Organic Chemistry
Organic Chemistry
CHEMISTRY
TABLE OF CONTENTS
The element carbon has unique properties making it a foundation for very large and wonderfully intricate molecules. The properties and
function of the more than 50 million unique substances the human race understands, most of which are organic compounds, can be
understood by exploring the composition, connectivity, shape, and sense of carbon-based molecules. Organic Chemistry is a challenging
and valuable discussion transferable to both the UC and CSU systems.
10: ORGANOHALIDES
10.1: Chapter Objectives
10.2: Introduction to Organohalides
10.3: Names and Properties of Alkyl Halides
10.4: Preparing Alkyl Halides from Alkanes - Radical Halogenation
10.5: Preparing Alkyl Halides from Alkenes - Allylic Bromination
10.6: Stability of the Allyl Radical - Resonance Revisited
10.7: Preparing Alkyl Halides from Alcohols
10.8: Reactions of Alkyl Halides - Grignard Reagents
KEY TERMS
Make certain that you can define, and use in context, the key term below.
organic chemistry
All living things on earth are formed mostly of carbon compounds. The prevalence of carbon compounds in living things has led to the
epithet “carbon-based” life. The truth is we know of no other kind of life. Early chemists regarded substances isolated from organisms
(plants and animals) as a different type of matter that could not be synthesized artificially, and these substances were thus known as
organic compounds. The widespread belief called vitalism held that organic compounds were formed by a vital force present only in
living organisms. The German chemist Friedrich Wöhler was one of the early chemists to refute this aspect of vitalism, when, in 1828,
he reported the synthesis of urea, a component of many body fluids, from nonliving materials. Since then, it has been recognized that
organic molecules obey the same natural laws as inorganic substances, and the category of organic compounds has evolved to include
both natural and synthetic compounds that contain carbon. Some carbon-containing compounds are not classified as organic, for
example, carbonates and cyanides, and simple oxides, such as CO and CO2. Although a single, precise definition has yet to be
identified by the chemistry community, most agree that a defining trait of organic molecules is the presence of carbon as the principal
element, bonded to hydrogen and other carbon atoms.
Figure 1.0.1: All organic compounds contain carbon and most are formed by living things, although they are also formed by
geological and artificial processes. (credit left: modification of work by Jon Sullivan; credit left middle: modification of work by Deb
Tremper; credit right middle: modification of work by “annszyp”/Wikimedia Commons; credit right: modification of work by George
Shuklin)
Today, organic compounds are key components of plastics, soaps, perfumes, sweeteners, fabrics, pharmaceuticals, and many other
substances that we use every day. The value to us of organic compounds ensures that organic chemistry is an important discipline
within the general field of chemistry. In this chapter, we discuss why the element carbon gives rise to a vast number and variety of
compounds, how those compounds are classified, and the role of organic compounds in representative biological and industrial
settings.
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]).
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
KEY TERMS
Make certain that you can define, and use in context, the key terms below.
atomic number
atomic weight
electron
mass number
neutron
proton
Almost all of the mass of an atom is contained within a tiny (and therefore extremely dense) nucleus which carries a positive
electric charge whose value identifies each element and is known as the atomic number of the element.
Almost all of the volume of an atom consists of empty space in which electrons, the fundamental carriers of negative electric
charge, reside. The extremely small mass of the electron (1/1840th the mass of the hydrogen nucleus) causes it to behave as a
quantum particle, which means that its location at any moment cannot be specified; the best we can do is describe its behavior in
terms of the probability of its manifesting itself at any point in space. It is common (but somewhat misleading) to describe the
volume of space in which the electrons of an atom have a significant probability of being found as the electron cloud. The latter has
no definite outer boundary, so neither does the atom. The radius of an atom must be defined arbitrarily, such as the boundary in
which the electron can be found with 95% probability. Atomic radii are typically 30-300 pm.
Because the electrons of an atom are in contact with the outside world, it is possible for
one or more electrons to be lost, or some new ones to be added. The resulting electrically-
charged atom is called an ion.
The other nuclear particle is the neutron. As its name implies, this particle carries no electrical charge. Its mass is almost the same as
that of the proton. Most nuclei contain roughly equal numbers of neutrons and protons, so we can say that these two particles together
account for almost all the mass of the atom.
Atomic numbers were first worked out in 1913 by Henry Moseley, a young member of Rutherford's research group in Manchester.
Moseley searched for a measurable property of each element that increases linearly with atomic number. He found this in a class of X-
rays emitted by an element when it is bombarded with electrons. The frequencies of these X-rays are unique to each element, and they
increase uniformly in successive elements. Moseley found that the square roots of these frequencies give a straight line when plotted
against Z; this enabled him to sort the elements in order of increasing atomic number.
You can think of the atomic number as a kind of serial number of an element, commencing at 1 for hydrogen and increasing by one for
each successive element. The chemical name of the element and its symbol are uniquely tied to the atomic number; thus the symbol
"Sr" stands for strontium, whose atoms all have Z = 38.
A = Z +N (1.2.1)
Two nuclides having the same atomic number but different mass numbers are known as isotopes. Most elements occur in nature as
mixtures of isotopes, but twenty-three of them (including beryllium and fluorine, shown in the table) are monoisotopic. For example,
there are three natural isotopes of magnesium: 24Mg (79% of all Mg atoms), 25Mg (10%), and 26Mg (11%); all three are present in all
compounds of magnesium in about these same proportions.
Approximately 290 isotopes occur in nature. The two heavy isotopes of hydrogen are especially important— so much so that they have
names and symbols of their own:
ATOMIC WEIGHTS
Atoms are of course far too small to be weighed directly; weight measurements can only be made on the massive (but unknown)
numbers of atoms that are observed in chemical reactions. The early combining-weight experiments of Dalton and others established
that hydrogen is the lightest of the atoms, but the crude nature of the measurements and uncertainties about the formulas of many
compounds made it difficult to develop a reliable scale of the relative weights of atoms. Even the most exacting weight measurements
we can make today are subject to experimental uncertainties that limit the precision to four significant figures at best.
CONTRIBUTORS
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
Prof. Steven Farmer (Sonoma State University)
William Reusch, Professor Emeritus (Michigan State U.), Virtual Textbook of Organic Chemistry
Organic Chemistry With a Biological Emphasis by Tim Soderberg (University of Minnesota, Morris)
KEY TERMS
Make certain that you can define, and use in context, the key terms below.
nodal plane
node
orbital
quantum mechanics
wave function
ATOMIC ORBITALS
An orbital is the quantum mechanical refinement of Bohr’s orbit. In contrast to his concept of a simple circular orbit with a fixed
radius, orbitals are mathematically derived regions of space with different probabilities of having an electron.
One way of representing electron probability distributions was illustrated in Figure 6.5.2 for the 1s orbital of hydrogen. Because Ψ2
gives the probability of finding an electron in a given volume of space (such as a cubic picometer), a plot of Ψ2 versus distance from
the nucleus (r) is a plot of the probability density. The 1s orbital is spherically symmetrical, so the probability of finding a 1s electron
at any given point depends only on its distance from the nucleus. The probability density is greatest at r = 0 (at the nucleus) and
decreases steadily with increasing distance. At very large values of r, the electron probability density is very small but not zero.
In contrast, we can calculate the radial probability (the probability of finding a 1s electron at a distance r from the nucleus) by adding
together the probabilities of an electron being at all points on a series of x spherical shells of radius r1, r2, r3,…, rx − 1, rx. In effect, we
are dividing the atom into very thin concentric shells, much like the layers of an onion (part (a) in Figure 1.2.1), and calculating the
probability of finding an electron on each spherical shell. Recall that the electron probability density is greatest at r = 0 (part (b) in
Figure 1.2.1), so the density of dots is greatest for the smallest spherical shells in part (a) in Figure 1.2.1. In contrast, the surface area of
each spherical shell is equal to 4πr2, which increases very rapidly with increasing r (part (c) in Figure 1.2.1). Because the surface area
of the spherical shells increases more rapidly with increasing r than the electron probability density decreases, the plot of radial
probability has a maximum at a particular distance (part (d) in Figure 1.2.1). Most important, when r is very small, the surface area of a
spherical shell is so small that the total probability of finding an electron close to the nucleus is very low; at the nucleus, the electron
probability vanishes (part (d) in Figure 1.2.1).
S ORBITALS
Three things happen to s orbitals as n increases (Figure 1.2.2):
1. They become larger, extending farther from the nucleus.
2. They contain more nodes. This is similar to a standing wave that has regions of significant amplitude separated by nodes, points
with zero amplitude.
3. For a given atom, the s orbitals also become higher in energy as n increases because of their increased distance from the nucleus.
Orbitals are generally drawn as three-dimensional surfaces that enclose 90% of the electron density, as was shown for the hydrogen 1s,
2s, and 3s orbitals in part (b) in Figure 1.2.2. Although such drawings show the relative sizes of the orbitals, they do not normally show
the spherical nodes in the 2s and 3s orbitals because the spherical nodes lie inside the 90% surface. Fortunately, the positions of the
spherical nodes are not important for chemical bonding.
P ORBITALS
Only s orbitals are spherically symmetrical. As the value of l increases, the number of orbitals in a given subshell increases, and the
shapes of the orbitals become more complex. Because the 2p subshell has l = 1, with three values of ml (−1, 0, and +1), there are three
2p orbitals.
ELECTRON CONFIGURATIONS
Before assigning the electrons of an atom into orbitals, one must become familiar with the basic concepts of electron configurations.
Every element on the periodic table consists of atoms, which are composed of protons, neutrons, and electrons. Electrons exhibit a
negative charge and are found around the nucleus of the atom in electron orbitals, defined as the volume of space in which the electron
can be found within 95% probability. The four different types of orbitals (s,p,d, and f) have different shapes, and one orbital can hold a
maximum of two electrons. The p, d, and f orbitals have different sublevels, thus can hold more electrons.
As stated, the electron configuration of each element is unique to its position on the periodic table. The energy level is determined by
the period and the number of electrons is given by the atomic number of the element. Orbitals on different energy levels are similar to
each other, but they occupy different areas in space. The 1s orbital and 2s orbital both have the characteristics of an s orbital (radial
nodes, spherical volume probabilities, can only hold two electrons, etc.) but, as they are found in different energy levels, they occupy
different spaces around the nucleus. Each orbital can be represented by specific blocks on the periodic table. The s-block is the region
Using the periodic table to determine the electron configurations of atoms is key, but also keep in mind that there are certain rules to
follow when assigning electrons to different orbitals. The periodic table is an incredibly helpful tool in writing electron configurations.
For more information on how electron configurations and the periodic table are linked, visit the Connecting Electrons to the Periodic
Table module.
Groups 3-12: transition metals 2* (The 4s shell is complete and cannot hold any more electrons)
* The general method for counting valence electrons is generally not useful for transition metals. Instead the modified d electron count
method is used.
** Except for helium, which has only two valence electrons.
Answer:
S1.2.1
1= 3px; 2= 3s ; 3= 2pz
CONTRIBUTORS
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
Prof. Steven Farmer (Sonoma State University)
William Reusch, Professor Emeritus (Michigan State U.), Virtual Textbook of Organic Chemistry
Organic Chemistry With a Biological Emphasis by Tim Soderberg (University of Minnesota, Morris)
KEY TERMS
Make certain that you can define, and use in context, the key terms below.
ground-state electronic configuration
Hund's rule
Pauli exclusion principle
aufbau principle
The electron configuration of an element is the arrangement of its electrons in its atomic orbitals. By knowing the electron
configuration of an element, we can predict and explain a great deal of its chemistry.
From the orbital diagram, we can write the electron configuration in an abbreviated form in which the occupied orbitals are identified
by their principal quantum number n and their value of l (s, p, d, or f), with the number of electrons in the subshell indicated by a
superscript. For hydrogen, the single electron is placed in the 1s orbital, which is the orbital lowest in energy (Figure 1.2.5), and the
electron configuration is written as 1s1 and read as “one-s-one.”
A neutral helium atom, with an atomic number of 2 (Z = 2), has two electrons. We place one electron in the orbital that is lowest in
energy, the 1s orbital. From the Pauli exclusion principle, we know that an orbital can contain two electrons with opposite spin, so we
place the second electron in the same orbital as the first but pointing down, so that the electrons are paired. The orbital diagram for the
helium atom is therefore
written as 1s2, where the superscript 2 implies the pairing of spins. Otherwise, our configuration would violate the Pauli principle.
The next element is lithium, with Z = 3 and three electrons in the neutral atom. We know that the 1s orbital can hold two of the
electrons with their spins paired. Figure 1.2.5 tells us that the next lowest energy orbital is 2s, so the orbital diagram for lithium is
At carbon, with Z = 6 and six electrons, we are faced with a choice. Should the sixth electron be placed in the same 2p orbital that
already has an electron, or should it go in one of the empty 2p orbitals? If it goes in an empty 2p orbital, will the sixth electron have its
spin aligned with or be opposite to the spin of the fifth? In short, which of the following three orbital diagrams is correct for carbon,
remembering that the 2p orbitals are degenerate?
Because of electron-electron repulsions, it is more favorable energetically for an electron to be in an unoccupied orbital than in one that
is already occupied; hence we can eliminate choice a. Similarly, experiments have shown that choice b is slightly higher in energy (less
stable) than choice c because electrons in degenerate orbitals prefer to line up with their spins parallel; thus, we can eliminate choice b.
Choice c illustrates Hund’s rule (named after the German physicist Friedrich H. Hund, 1896–1997), which says that the lowest-energy
electron configuration for an atom is the one that has the maximum number of electrons with parallel spins in degenerate orbitals. By
Hund’s rule, the electron configuration of carbon, which is 1s22s22p2, is understood to correspond to the orbital diagram shown in c.
Experimentally, it is found that the ground state of a neutral carbon atom does indeed contain two unpaired electrons.
When we get to nitrogen (Z = 7, with seven electrons), Hund’s rule tells us that the lowest-energy arrangement is
with three unpaired electrons. The electron configuration of nitrogen is thus 1s22s22p3.
At oxygen, with Z = 8 and eight electrons, we have no choice. One electron must be paired with another in one of the 2p orbitals,
which gives us two unpaired electrons and a 1s22s22p4 electron configuration. Because all the 2p orbitals are degenerate, it doesn’t
matter which one has the pair of electrons.
Notice that for neon, as for helium, all the orbitals through the 2p level are completely filled. This fact is very important in dictating
both the chemical reactivity and the bonding of helium and neon, as you will see.
VALENCE ELECTRONS
As we continue through the periodic table in this way, writing the electron configurations of larger and larger atoms, it becomes tedious
to keep copying the configurations of the filled inner subshells. In practice, chemists simplify the notation by using a bracketed noble
gas symbol to represent the configuration of the noble gas from the preceding row because all the orbitals in a noble gas are filled. For
example, [Ne] represents the 1s22s22p6 electron configuration of neon (Z = 10), so the electron configuration of sodium, with Z = 11,
which is 1s22s22p63s1, is written as [Ne]3s1:
Neon Z = 10 1s22s22p6
Sodium Z = 11 1s22s22p63s1 = [Ne]3s1
Because electrons in filled inner orbitals are closer to the nucleus and more tightly bound to it, they are rarely involved in chemical
reactions. This means that the chemistry of an atom depends mostly on the electrons in its outermost shell, which are called the valence
electrons. The simplified notation allows us to see the valence electron configuration more easily. Using this notation to compare the
electron configurations of sodium and lithium, we have:
Sodium 1s22s22p63s1 = [Ne]3s1
Lithium 1s22s1 = [He]2s1
It is readily apparent that both sodium and lithium have one s electron in their valence shell. We would therefore predict that sodium
and lithium have very similar chemistry, which is indeed the case.
As we continue to build the eight elements of period 3, the 3s and 3p orbitals are filled, one electron at a time. This row concludes with
the noble gas argon, which has the electron configuration [Ne]3s23p6, corresponding to a filled valence shell.
EXAMPLE 1.3.1
Draw an orbital diagram and use it to derive the electron configuration of phosphorus, Z = 15. What is its valence electron
configuration?
Given: atomic number
Asked for: orbital diagram and valence electron configuration for phosphorus
Strategy:
A. Locate the nearest noble gas preceding phosphorus in the periodic table. Then subtract its number of electrons from those in
phosphorus to obtain the number of valence electrons in phosphorus.
B. Referring to Figure 1.3.1, draw an orbital diagram to represent those valence orbitals. Following Hund’s rule, place the valence
electrons in the available orbitals, beginning with the orbital that is lowest in energy. Write the electron configuration from your
orbital diagram.
C. Ignore the inner orbitals (those that correspond to the electron configuration of the nearest noble gas) and write the valence
electron configuration for phosphorus.
Solution:
A Because phosphorus is in the third row of the periodic table, we know that it has a [Ne] closed shell with 10 electrons. We begin
by subtracting 10 electrons from the 15 in phosphorus.
B The additional five electrons are placed in the next available orbitals, which Figure 1.2.5 tells us are the 3s and 3p orbitals:
Hund’s rule tells us that the remaining three electrons will occupy the degenerate 3p orbitals separately but with their spins aligned:
EXERCISE 1.3.1
Draw an orbital diagram and use it to derive the electron configuration of chlorine, Z = 17. What is its valence electron
configuration?
Answer:
[Ne]3s23p5; 3s23p5
The general order in which orbitals are filled is depicted in Figure 1.3.1. Subshells corresponding to each value of n are written from
left to right on successive horizontal lines, where each row represents a row in the periodic table. The order in which the orbitals are
filled is indicated by the diagonal lines running from the upper right to the lower left. Accordingly, the 4s orbital is filled prior to the 3d
orbital because of shielding and penetration effects. Consequently, the electron configuration of potassium, which begins the fourth
period, is [Ar]4s1, and the configuration of calcium is [Ar]4s2. Five 3d orbitals are filled by the next 10 elements, the transition metals,
followed by three 4p orbitals. Notice that the last member of this row is the noble gas krypton (Z = 36), [Ar]4s23d104p6 = [Kr], which
has filled 4s, 3d, and 4p orbitals. The fifth row of the periodic table is essentially the same as the fourth, except that the 5s, 4d, and 5p
orbitals are filled sequentially.
Figure 1.3.1: Predicting the Order in Which Orbitals Are Filled in Multielectron Atoms. If you write the subshells for each value of the
principal quantum number on successive lines, the observed order in which they are filled is indicated by a series of diagonal lines
running from the upper right to the lower left.
The sixth row of the periodic table will be different from the preceding two because the 4f orbitals, which can hold 14 electrons, are
filled between the 6s and the 5d orbitals. The elements that contain 4f orbitals in their valence shell are the lanthanides. When the 6p
orbitals are finally filled, we have reached the next (and last known) noble gas, radon (Z = 86), [Xe]6s24f145d106p6 = [Rn]. In the last
row, the 5f orbitals are filled between the 7s and the 6d orbitals, which gives the 14 actinide elements. Because the large number of
protons makes their nuclei unstable, all the actinides are radioactive.
EXAMPLE 1.3.2
Write the electron configuration of mercury (Z = 80), showing all the inner orbitals.
SUMMARY
Based on the Pauli principle and a knowledge of orbital energies obtained using hydrogen-like orbitals, it is possible to construct the
periodic table by filling up the available orbitals beginning with the lowest-energy orbitals (the aufbau principle), which gives rise to
a particular arrangement of electrons for each element (its electron configuration). Hund’s rule says that the lowest-energy
arrangement of electrons is the one that places them in degenerate orbitals with their spins parallel. For chemical purposes, the most
important electrons are those in the outermost principal shell, the valence electrons.
EXERCISES
QUESTIONS
Q1.3.1
Give the electron configurations for Al, Br, Fe.
SOLUTIONS
S1.3.1
Al = 1s22s22p63s23p1
Br = 1s22s22p63s23p64s23d104p5
Fe = 1s22s22p63s23p64s23d6
CONTRIBUTORS
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
Prof. Steven Farmer (Sonoma State University)
William Reusch, Professor Emeritus (Michigan State U.), Virtual Textbook of Organic Chemistry
Organic Chemistry With a Biological Emphasis by Tim Soderberg (University of Minnesota, Morris)
STUDY NOTES
When drawing any organic structure, you must remember that a neutral carbon atom will almost always have four bonds. Similarly,
hydrogen always has one bond; neutral oxygen atoms have two bonds; and neutral nitrogen atoms have three bonds. By committing
these simple rules to memory, you can avoid making unnecessary mistakes later in the course.
The “wedge-and-broken-line” type of representation, which helps to convey the three-dimensional nature of organic compounds,
will be used throughout the course.
BONDING OVERVIEW
Why are some substances chemically bonded molecules and others are an association of ions? The answer to this question depends
upon the electronic structures of the atoms and nature of the chemical forces within the compounds. Although there are no sharply
defined boundaries, chemical bonds are typically classified into three main types: ionic bonds, covalent bonds, and metallic bonds. In
this chapter, each type of bond and the general properties found in typical substances in which the bond type occurs will be discussed.
1. Ionic bonds results from electrostatic forces that exist between ions of opposite charge. These bonds typically involve a metal with
a nonmetal
2. Covalent bonds result from the sharing of electrons between two atoms. The bonds typically involve one nonmetallic element with
another
3. Metallic bonds are found in solid metals (copper, iron, aluminum) with each metal atom bonded to several neighboring metal atoms
and the bonding electrons are free to move throughout the 3-dimensional structure.
Each bond classification is discussed in detail in subsequent sections of the chapter. Let's look at the preferred arrangements of
electrons in atoms when they form chemical compounds.
Figure 1.4.1: G. N. Lewis and the Octet Rule. (a) Lewis is working in the laboratory. (b) In Lewis’s original sketch for the octet rule,
he initially placed the electrons at the corners of a cube rather than placing them as we do now.
LEWIS SYMBOLS
At the beginning of the 20th century, the American chemist G. N. Lewis (1875–1946) devised a system of symbols—now called Lewis
electron dot symbols, often shortened to Lewis dot symbols—that can be used for predicting the number of bonds formed by most
elements in their compounds. Each Lewis dot symbol consists of the chemical symbol for an element surrounded by dots that represent
its valence electrons.
To write an element’s Lewis dot symbol, we place dots representing its valence electrons, one at a time, around the element’s chemical
symbol. Up to four dots are placed above, below, to the left, and to the right of the symbol (in any order, as long as elements with four
or fewer valence electrons have no more than one dot in each position). The next dots, for elements with more than four valence
electrons, are again distributed one at a time, each paired with one of the first four. For example, the electron configuration for atomic
sulfur is [Ne]3s23p4, thus there are six valence electrons. Its Lewis symbol would therefore be:
Fluorine, for example, with the electron configuration [He]2s22p5, has seven valence electrons, so its Lewis dot symbol is constructed
as follows:
The number of dots in the Lewis dot symbol is the same as the number of valence electrons, which is the same as the last digit of the
element’s group number in the periodic table. Lewis dot symbols for the elements in period 2 are given in Figure 1.4.2.
Lewis used the unpaired dots to predict the number of bonds that an element will form in a compound. Consider the symbol for
nitrogen in Figure 1.4.2. The Lewis dot symbol explains why nitrogen, with three unpaired valence electrons, tends to form compounds
in which it shares the unpaired electrons to form three bonds. Boron, which also has three unpaired valence electrons in its Lewis dot
symbol, also tends to form compounds with three bonds, whereas carbon, with four unpaired valence electrons in its Lewis dot symbol,
tends to share all of its unpaired valence electrons by forming compounds in which it has four bonds.
Atoms often gain, lose, or share electrons to achieve the same number of electrons as the
noble gas closest to them in the periodic table.
MOLECULAR SHAPE
A stick and wedge drawing of methane shows the tetrahedral angles...(The wedge is coming out of the paper and the dashed line is
going behind the paper. The solid lines are in the plane of the paper.)
The following examples make use of this notation, and also illustrate the importance of including non-bonding valence shell electron
pairs when viewing such configurations.
Bonding configurations are readily predicted by valence-shell electron-pair repulsion theory, commonly referred to as VSEPR in most
introductory chemistry texts. This simple model is based on the fact that electrons repel each other, and that it is reasonable to expect
that the bonds and non-bonding valence electron pairs associated with a given atom will prefer to be as far apart as possible. The
bonding configurations of carbon are easy to remember, since there are only three categories.
Configuration Bonding Partners Bond Angles Example
Tetrahedral 4 109.5º
Linear 2 180º
Figure 1.4.3 In the three examples shown above, the central atom (carbon) does not have any non-bonding valence electrons;
consequently the configuration may be estimated from the number of bonding partners alone. However, for molecules of water and
AX2: BEH2
1. The central atom, beryllium, contributes two valence electrons, and each hydrogen atom contributes one. The Lewis electron
structure is
AX2: CO2
1. The central atom, carbon, contributes four valence electrons, and each oxygen atom contributes six. The Lewis electron structure is
2. The carbon atom forms two double bonds. Each double bond is a group, so there are two electron groups around the central atom.
Like BeH2, the arrangement that minimizes repulsions places the groups 180° apart.
3. Once again, both groups around the central atom are bonding pairs (BP), so CO2 is designated as AX2.
4. VSEPR only recognizes groups around the central atom. Thus the lone pairs on the oxygen atoms do not influence the molecular
geometry. With two bonding pairs on the central atom and no lone pairs, the molecular geometry of CO2 is linear (Figure 1.4.3).
2. There are three electron groups around the central atom. To minimize repulsions, the groups are placed 120° apart (Figure 1.4.3).
3. All electron groups are bonding pairs (BP), so the structure is designated as AX3.
4. From Figure 1.4.3 we see that with three bonding pairs around the central atom, the molecular geometry of BCl3 is trigonal planar.
AX3: CO32−
1. The central atom, carbon, has four valence electrons, and each oxygen atom has six valence electrons. As you learned previously, the
Lewis electron structure of one of three resonance forms is represented as
In our next example we encounter the effects of lone pairs and multiple bonds on molecular geometry for the first time.
AX2E: SO2
1. The central atom, sulfur, has 6 valence electrons, as does each oxygen atom. With 18 valence electrons, the Lewis electron structure
is shown below.
2. There are three electron groups around the central atom, two double bonds and one lone pair. We initially place the groups in a
trigonal planar arrangement to minimize repulsions (Figure 1.4.3).
3. There are two bonding pairs and one lone pair, so the structure is designated as AX2E. This designation has a total of three electron
pairs, two X and one E. Because a lone pair is not shared by two nuclei, it occupies more space near the central atom than a bonding
pair (Figure 1.4.4). Thus bonding pairs and lone pairs repel each other electrostatically in the order BP–BP < LP–BP < LP–LP. In SO2,
we have one BP–BP interaction and two LP–BP interactions.
4. The molecular geometry is described only by the positions of the nuclei, not by the positions of the lone pairs. Thus with two nuclei
and one lone pair the shape is bent, or V shaped, which can be viewed as a trigonal planar arrangement with a missing vertex (Figures
1.4.2.1 and 1.4.3).
Figure 1.4.4: The Difference in the Space Occupied by a Lone Pair of Electrons and by a Bonding Pair
As with SO2, this composite model of electron distribution and negative electrostatic potential in ammonia shows that a lone pair of
electrons occupies a larger region of space around the nitrogen atom than does a bonding pair of electrons that is shared with a
AX4: CH4
1. The central atom, carbon, contributes four valence electrons, and each hydrogen atom has one valence electron, so the full Lewis
electron structure is
2. There are four electron groups around the central atom. As shown in Figure 1.4.2, repulsions are minimized by placing the groups in
the corners of a tetrahedron with bond angles of 109.5°.
3. All electron groups are bonding pairs, so the structure is designated as AX4.
4. With four bonding pairs, the molecular geometry of methane is tetrahedral (Figure 1.4.3).
AX3E: NH3
1. In ammonia, the central atom, nitrogen, has five valence electrons and each hydrogen donates one valence electron, producing the
Lewis electron structure
2. There are four electron groups around nitrogen, three bonding pairs and one lone pair. Repulsions are minimized by directing each
hydrogen atom and the lone pair to the corners of a tetrahedron.
3. With three bonding pairs and one lone pair, the structure is designated as AX3E. This designation has a total of four electron pairs,
three X and one E. We expect the LP–BP interactions to cause the bonding pair angles to deviate significantly from the angles of a
perfect tetrahedron.
AX2E2: H2O
1. Oxygen has six valence electrons and each hydrogen has one valence electron, producing the Lewis electron structure
2. There are four groups around the central oxygen atom, two bonding pairs and two lone pairs. Repulsions are minimized by directing
the bonding pairs and the lone pairs to the corners of a tetrahedron Figure 1.4.3.
3. With two bonding pairs and two lone pairs, the structure is designated as AX2E2 with a total of four electron pairs. Due to LP–LP,
LP–BP, and BP–BP interactions, we expect a significant deviation from idealized tetrahedral angles.
4. With two hydrogen atoms and two lone pairs of electrons, the structure has significant lone pair interactions. There are two nuclei
about the central atom, so the molecular shape is bent, or V shaped, with an H–O–H angle that is even less than the H–N–H angles in
NH3, as we would expect because of the presence of two lone pairs of electrons on the central atom rather than one.. This molecular
shape is essentially a tetrahedron with two missing vertices.
EXERCISES
QUESTIONS
Q1.4.1
List the bond angles for each of the following compounds: BH3, CF4, H2O.
Q1.4.2
Why is sulfur dioxide a bent molecule (bond angle less than 180°)?
SOLUTIONS
S1.4.1
HBH = 120°
FCF = 109.5°
OHO = 104°
S1.4.2
This deviation is due to the lone pairs on the sulfur. These force the molecule to exhibit a
“bent” geometry and therefore a deviation from the 180°.
KEY TERMS
Make certain that you can define, and use in context, the key terms below.
bond strength
covalent bond
ionic bond
Lewis structure
lone-pair electron
non-bonding electron
sigma (σ) bond
π
pi ( ) bond
valence bond theory
STUDY NOTES
To draw Lewis structures successfully, you need to know the number of valence electrons present in each of the atoms involved.
Memorize the number of valence electrons possessed by each of the elements commonly encountered in organic chemistry: C, H,
O, N, S, P and the halogens.
IONIC BONDING
Ions are atoms or molecules which are electrically charged. Cations are positively charged and anions carry a negative charge. Ions
form when atoms gain or lose electrons. Since electrons are negatively charged, an atom that loses one or more electrons will become
positively charged; an atom that gains one or more electrons becomes negatively charged.
Ionic bonding is the attraction between positively- and negatively-charged ions. These oppositely charged ions attract each other to
form ionic networks (or lattices). Electrostatics explains why this happens: opposite charges attract and like charges repel. When many
ions attract each other, they form large, ordered, crystal lattices in which each ion is surrounded by ions of the opposite charge.
Generally, when metals react with non-metals, electrons are transferred from the metals to the non-metals. The metals form positively-
charged ions and the non-metals form negatively-charged ions.
Ionic bonds form when metals and non-metals chemically react. By definition, a metal is relatively stable if it loses electrons to form a
complete valence shell and becomes positively charged. Likewise, a non-metal becomes more stable by gaining electrons to complete
its valence shell and become negatively charged. When metals and non-metals react, the metals lose electrons by transferring them to
the non-metals, which gain them. Consequently, ions are formed, which instantly attract each other—ionic bonding.
For full video of making NaCl from sodium metal and chlorine gase, see https://www.youtube.com/watch?v=WVonuBjCrNo. These
ions are arranged in solid NaCl in a regular three-dimensional arrangement (or lattice):
The arrow indicates the transfer of the electron from sodium to chlorine to form the Na+ metal ion and the Cl- chloride ion. Each ion
now has an octet of electrons in its valence shell:
Na+: 2s22p6
Cl-: 3s23p6
CHEMICAL BONDS
Chemical bonds are the attractive forces that hold atoms together in the form of compounds. They are formed when electrons are
shared between two atoms. There are 3 types of bonds: covalent bonds, polar covalent bonds and ionic bonds. The simplest example of
bonding can be demonstrated by the H2 molecule. We can see from the periodic table that each hydrogen atom has a single electron. If
2 hydrogen atoms come together to form a bond, then each hydrogen atom effectively has a share in both electrons and thus each
resembles a noble gas and is more stable. The 2 electrons that are shared can be represented either by 2 dots or a single dash between
the atoms.
Valence bond theory describes a chemical bond as the overlap of atomic orbitals. In the case of the hydrogen molecule, the 1s orbital of
one hydrogen atom overlaps with the 1s orbital of the second hydrogen atom to form a molecular orbital called a sigma bond.
Attraction increases as the distance between the atoms gets closer but nuclear-nuclear repulsion becomes important if the atoms
approach too close.
Hydrogen represents is a special case, of course – a hydrogen atom cannot fulfill the octet rule; it needs only two electrons to have a
full shell (you could think of this as the ‘doublet rule’ for hydrogen).
One of the simplest organic molecules is methane with the molecular formula CH4. Methane is the ‘natural gas’ burned in home
furnaces and hot water heaters, as well as in electrical power generating plants. To illustrate the covalent bonding in methane using the
Lewis method, we first must recognize that, although a carbon atom has a total of six electrons, the two electrons in the inner 1s orbital
do not participate in bonding interactions. It is the other four- those in the 2s and 2p orbitals - that form covalent bonds with other
atoms. Only the partially occupied, highest energy shell of orbitals - in this case the 2s and 2p orbitals - can overlap with orbitals on
other atoms to form covalent bonds. Electrons in these orbitals are termed ‘valence electrons’.
A carbon atom, then, has four valence electrons with which to form covalent bonds. In order to fulfill the octet rule and increase the
occupancy of its second shell to eight electrons, it must participate in four electron-sharing interactions - in other words, it must form
four covalent bonds. In a methane molecule, the central carbon atom shares its four valence electrons with four hydrogen atoms, thus
forming four bonds and fulfilling the octet rule (for the carbon) and the ‘doublet rule’ (for each of the hydrogens).
The next relatively simple organic molecule to consider is ethane, which has the molecular formula C2H6. If we draw each atom with
its valence electron(s) separately, we can see that the octet/doublet rule can be fulfilled for all of them by forming one carbon-carbon
bond and six carbon-hydrogen bonds.
EXERCISE 1.5.1
Draw the Lewis structure for ammonia, NH3.
Answer:
What about multiple bonds? The molecular formula for ethene (also known as ethylene, a compound found in fruits,
such as apples, that signals them to ripen) is C2H4. Arranging the atoms and surrounding them with their valence
electrons, you can see that the octet/doublet rule can be fulfilled for all atoms only if the two carbons share two pairs
of electrons between them - in other words, only if a double bond is formed.
Because a hydrogen atom has only a 1s orbital to work with, it cannot form more than one single bond, otherwise it would exceed its
doublet rule.
Following this pattern, the triple bond in ethyne molecular formula C2H2, (also known as acetylene, the fuel used in welding torches),
is formed when the two carbon atoms share three pairs of electrons between them.
What about ions? The hydroxide ion, OH-, is drawn simply by showing the oxygen atom with its six valence electrons, then adding one
more electron to account for the negative charge. Now the oxygen has three non-bonding lone pairs, and can only form one bond to a
hydrogen. (Bear in mind that this is merely a description of the thought process going into drawing a Lewis structure, and is not meant
to describe any actual chemical process).
To draw a Lewis structure of the hydronium ion, H3O+, you again start with the oxygen atom with its six valence electrons, then take
one away to account for the positive charge (there is now one more proton than there are electrons). The oxygen now can form bonds
to three hydrogen atoms.
EXERCISE 1.5.2
Draw a Lewis structure for the ammonium ion, NH4+.
Answer:
Answer:
FORMAL CHARGES
Consider the Lewis structure of methanol, CH3OH (methanol is the so-called ‘wood alcohol’ that unscrupulous bootleggers sometimes
sold during the Prohibition days, often causing the people who drank it to go blind). Just like in a water molecule, the oxygen atom in a
methanol molecule has two non-bonding lone pairs of electrons. And just like a water molecule can be protonated to form the H3O+
cation, a methanol molecule can be protonated to form the CH3OH2+ cation.
This polyatomic cation, as you can see, has an overall charge of +1. But we can be more specific than that - we can also state that the
positive charge is located specifically on the oxygen atom, rather than on the carbon or any of the hydrogens. When a charge can be
located on a particular atom in a polyatomic ion, this atom is said to have a ‘formal charge’. Figuring out the formal charge on different
atoms of a polyatomic ion is a straightforward process - it’s simply a matter of adding up valence electrons. Remember that an oxygen
atom needs six valence electrons (in addition to the two electrons in the non-valence 1s orbital) to completely balance the charge of the
eight protons in its nucleus. Let’s figure out how many electrons the oxygen atom in our CH3OH2+ ion ‘owns’.
EXERCISE 1.5.4
Label all non-zero formal charges on the molecules/ions below. All atoms have a full octet of electrons (lone pairs are not shown).
Answer:
This valence shell repulsion model can be illustrated at home with a very fun experiment!
EXAMPLE 1.5.5: METHANE
In the case of methane, the three 2p orbitals of the carbon atom are combined with its 2s orbital to form four new orbitals called sp3
hybrid orbitals. The name is simply a tally of all the orbitals that were blended together to form these new hybrid orbitals. Four
hybrid orbitals were required since there are four atoms attached to the central carbon atom. These new orbitals will have an energy
slightly above the 2s orbital and below the 2p orbitals as shown in the following illustration. Notice that no change occurred with
the 1s orbital.
These hybrid orbitals have 75% p-character and 25% s-character which gives them a shape that is shorter and fatter than a p-orbital.
The new shape looks a little like...
A stick and wedge drawing of ammonia showing the non-bonding electrons in a probability area for the hybrid orbital...
A space-filling model of ammonia would look like...(Note the non-bonded electron pair is not shown in this model.)
A space-filling model of water would look like...(Note the non-bonded electron pairs are not shown in this model.)
In the following stick model, the empty p orbital is shown as the probability area...one end shaded blue and the other is white...there
are no electrons in this orbital!
In the following stick model, the empty p orbitals are shown as the probability areas...one green and one blue.
SUMMARY OF HYBRIDIZATION
In the following summary, groups are considered to be atoms and/or pairs of electrons and hybrid orbitals are the red lines and wedges.
When the octet of an element is exceeded, then hybridization will involve d-orbitals. Non-hybridized p-orbitals are shown as
probability areas in blue and green for sp hybridization and blue for sp2 hybridization. A single electron as found in a radical would
occupy an non-hybridized p-orbital.
Number of Groups Attached to a Description and 3-Dimensional Shape
Central Atom
Two Groups
sp
Three Groups
sp2
Four Groups
sp3
The 3-dimensional model of ethene is planar with H-C-H and H-C-C bond angles of 120º...the pi bond is not shown in this picture.
These p orbitals will undergo parallel overlap to form two pi bonds at right angles to each other.
The 3-dimensional model of acetylene is therefore linear...the pi bonds are not shown in this picture.
Table 1.5.1: Average Bond Energies (kJ/mol) for Commonly Encountered Bonds at 273 K
Single Bonds Multiple Bonds
H–H 432 C–C 346 N–N ≈167 O–O ≈142 F–F 155 C=C 602
H–C 411 C–Si 318 N–O 201 O–F 190 F–Cl 249 C≡C 835
H–Si 318 C–N 305 N–F 283 O–Cl 218 F–Br 249 C=N 615
H–N 386 C–O 358 N–Cl 313 O–Br 201 F–I 278 C≡N 887
H–P ≈322 C–S 272 N–Br 243 O–I 201 Cl–Cl 240 C=O 749
H–O 459 C–F 485 P–P 201 S–S 226 Cl–Br 216 C≡O 1072
H–S 363 C–Cl 327 S–F 284 Cl–I 208 N=N 418
H–F 565 C–Br 285 S–Cl 255 Br–Br 190 N≡N 942
H–Cl 428 C–I 213 S–Br 218 Br–I 175 N=O 607
Source: Data from J. E. Huheey, E. A. Keiter, and R. L. Keiter, Inorganic Chemistry, 4th ed. (1993).
1. Bonds between hydrogen and atoms in the same column of the periodic table decrease in strength as we go down the column. Thus
an H–F bond is stronger than an H–I bond, H–C is stronger than H–Si, H–N is stronger than H–P, H–O is stronger than H–S, and so
forth. The reason for this is that the region of space in which electrons are shared between two atoms becomes proportionally
smaller as one of the atoms becomes larger (Figure 1.5.1a)
2. Bonds between like atoms usually become weaker as we go down a column (important exceptions are noted later). For example,
the C–C single bond is stronger than the Si–Si single bond, which is stronger than the Ge–Ge bond, and so forth. As two bonded
atoms become larger, the region between them occupied by bonding electrons becomes proportionally smaller. (Figure 1.5.1b)
Noteworthy exceptions are single bonds between the period 2 atoms of groups 15, 16, and 17 (i.e., N, O, F), which are unusually
weak compared with single bonds between their larger congeners. It is likely that the N–N, O–O, and F–F single bonds are weaker
than might be expected due to strong repulsive interactions between lone pairs of electrons on adjacent atoms. The trend in bond
energies for the halogens is therefore
Cl– Cl > Br– Br > F – F > I – I (1.6.1)
Similar effects are also seen for the O–O versus S–S and for N–N versus P–P single bonds.
Bonds between hydrogen and atoms in a given column in the periodic table are weaker
down the column; bonds between like atoms usually become weaker down a column.
3. Because elements in periods 3 and 4 rarely form multiple bonds with themselves, their multiple bond energies are not accurately
known. Nonetheless, they are presumed to be significantly weaker than multiple bonds between lighter atoms of the same families.
Compounds containing an Si=Si double bond, for example, have only recently been prepared, whereas compounds containing C=C
double bonds are one of the best-studied and most important classes of organic compounds.
NOTE
Bond strengths increase as bond order increases, while bond distances decrease.
Table 1.5.2: Average bond energies:
Bond (kJ/mol)
C-F 485
C-Cl 328
C-Br 276
C-I 240
C-C 348
C-N 293
C-O 358
C-F 485
C-C 348
C=C 614
C=C 839
EXERCISES
QUESTIONS
Q1.5.1
Draw an energy diagram for energy vs. intermolecular distance for a fluorine molecule (F2)
and describe the regions of the graph.
SOLUTIONS
S1.5.1
CONTRIBUTORS
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
Richard Banks (Boise State University)
KEY TERMS
Make certain that you can define, and use in context, the key terms below.
bond angle
hybridization
sp3 hybrid
STUDY NOTES
The tetrahedral shape is a very important one in organic chemistry, as it is the basic shape of all compounds in which a carbon atom
is bonded to four other atoms. Note that the tetrahedral bond angle of H − C − H is 109.5°.
There is a serious mismatch between this structure and the modern electronic structure of carbon, 1s22s22px12py1. The modern
structure shows that there are only 2 unpaired electrons to share with hydrogens, instead of the 4 which the bonding picture requires.
You can see this more readily using the electrons-in-boxes notation. Only the 2nd level electrons are shown. The 1s2 electrons are too
deep inside the atom to be involved in bonding. The only electrons directly available for sharing are the 2p electrons. Why then isn't
methane CH2?
PROMOTION OF AN ELECTRON
When bonds are formed, energy is released and the system becomes more stable. If carbon forms 4 bonds rather than 2, twice as much
energy is released and so the resulting molecule becomes even more stable. There is only a small energy gap between the 2s and 2p
orbitals, and so it pays the carbon to provide a small amount of energy to promote an electron from the 2s to the empty 2p to give 4
unpaired electrons. The extra energy released when the bonds form more than compensates for the initial input.
The carbon atom is now said to be in an excited state. Now that we've got 4 unpaired electrons ready for bonding, another problem
arises. In methane all the carbon-hydrogen bonds are identical, but our electrons are in two different kinds of orbitals. You aren't going
to get four identical bonds unless you start from four identical orbitals.
HYBRIDIZATION
sp3 hybrid orbitals look a bit like half a p orbital, and they arrange themselves in space so that they are as far apart as possible. You can
picture the nucleus as being at the center of a tetrahedron (a triangularly based pyramid) with the orbitals pointing to the corners. For
clarity, the nucleus is drawn far larger than it really is.
WHAT HAPPENS WHEN THE BONDS ARE FORMED?
Remember that hydrogen's electron is in a 1s orbital - a spherically symmetric region of space surrounding the nucleus where there is
some fixed chance (say 95%) of finding the electron. When a covalent bond is formed, the atomic orbitals (the orbitals in the individual
atoms) merge to produce a new molecular orbital which contains the electron pair which creates the bond.
Four molecular orbitals are formed, looking rather like the original sp3 hybrids, but with a hydrogen nucleus embedded in each lobe.
Each orbital holds the 2 electrons that we've previously drawn as a dot and a cross. The principles involved - promotion of electrons if
necessary, then hybridization, followed by the formation of molecular orbitals - can be applied to any covalently-bound molecule.
CONTRIBUTORS
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
Prof. Steven Farmer (Sonoma State University)
Organic Chemistry With a Biological Emphasis by Tim Soderberg (University of Minnesota, Morris)
In the ethane molecule, the bonding picture according to valence orbital theory is very similar to that of methane. Both carbons are sp3-
hybridized, meaning that both have four bonds arranged with tetrahedral geometry. The carbon-carbon bond, with a bond length of
1.54 Å, is formed by overlap of one sp3 orbital from each of the carbons, while the six carbon-hydrogen bonds are formed from
overlaps between the remaining sp3 orbitals on the two carbons and the 1s orbitals of hydrogen atoms. All of these are sigma bonds.
Because they are formed from the end-on-end overlap of two orbitals, sigma bonds are free to rotate. This means, in the case of ethane
molecule, that the two methyl (CH3) groups can be pictured as two wheels on a hub, each one able to rotate freely with respect to the
other.
In chapter 3 we will learn more about the implications of rotational freedom in sigma bonds, when we discuss the ‘conformation’ of
organic molecules.
The sp3 bonding picture is also used to described the bonding in amines, including ammonia, the simplest amine. Just like the carbon
atom in methane, the central nitrogen in ammonia is sp3-hybridized. With nitrogen, however, there are five rather than four valence
electrons to account for, meaning that three of the four hybrid orbitals are half-filled and available for bonding, while the fourth is fully
occupied by a non-bonding pair of electrons.
C2H4, also known as ethylene or ethene, is a gaseous material created synthetically through steam cracking. In nature, it is released in
trace amounts by plants to signal their fruits to ripen. Ethene consists of two sp2-hybridized carbon atoms, which are sigma bonded to
each other and to two hydrogen atoms each. The remaining non-hybridized p orbitals on each carbon overlap to form a pi bond, which
gives ethene its reactivity.
EXERCISE
QUESTIONS
Q1.7.1
Draw pentane, CH3CH2CH2CH2CH3, predict the bond angles within this molecule.
SOLUTIONS
S1.7.1
KEY TERMS
Make certain that you can define, and use in context, the key terms below.
pi (π) bond
sp2 hybrid
BONDING IN ETHYLENE
A key component of using Valence Bond Theory correctly is being able to draw the Lewis dot diagram correctly. Ethylene, commonly
known as ethene, has a double bond between the carbons and single bonds between each hydrogen and carbon: each bond is
represented by a pair of dots, which represent electrons. Each carbon requires a full octet and each hydrogen requires a pair of
electrons. The correct Lewis structure for ethylene is shown below:
In the molecule ethene, both carbon atoms will be sp2 hybridized and have one unpaired electron in a non-hybridized p orbital.
These p-orbitals will undergo parallel overlap and form one sigma bond with bean-shaped probability areas above and below the plane
of the six atoms. This pair of bean-shaped probability areas constitutes one pi bond and the pair of electrons in this bond can be found
in either bean-shaped area.
The 3-dimensional model of ethene is therefore planar with H-C-H and H-C-C bond angles of 120º. Thepi bond is not shown in this
picture.
According to valence bond theory, two atoms form a covalent bond through the overlap of individual half-filled valence atomic
orbitals, each containing one unpaired electron. In ethene, each hydrogen atom has one unpaired electron and each carbon is sp2
hybridized with one electron each sp2 orbital. The fourth electron is in the p orbital that will form the pi bond. The bond order for
ethene is simply the number of bonds between each atom: the carbon-carbon bond has a bond order of two, and each carbon-hydrogen
bond has a bond order of one.
EXERCISE
QUESTIONS
Q1.8.1
Consider the following molecule:
At each atom, what is the hybridization and the bond angle? At atom A draw the molecular orbital.
SOLUTIONS
S1.8.1
A - sp2, 120°
B - sp3, 109°
C - sp2, 120° (with the lone pairs present)
D - sp3, 109°
CONTRIBUTORS
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
KEY TERMS
Make certain that you can define, and use in context, the key term below.
sp hybrid orbital
STUDY NOTES
The bond angles associated with sp3-, sp2- and sp‑hybridized carbon atoms are approximately 109.5°, 120° and 180°, respectively.
BONDING IN ACETYLENE
Finally, the hybrid orbital concept applies well to triple-bonded groups, such as alkynes and nitriles. Consider, for example, the
structure of ethyne (another common name is acetylene), the simplest alkyne.
This molecule is linear: all four atoms lie in a straight line. The carbon-carbon triple bond is only 1.20Å long. In the hybrid orbital
picture of acetylene, both carbons are sp-hybridized. In an sp-hybridized carbon, the 2s orbital combines with the 2px orbital to form
two sp hybrid orbitals that are oriented at an angle of 180°with respect to each other (eg. along the x axis). The 2py and 2pz orbitals
remain non-hybridized, and are oriented perpendicularly along the y and z axes, respectively.
The C-C sigma bond is formed by the overlap of one sp orbital from each of the carbons, while the two C-H sigma bonds are formed
by the overlap of the second sp orbital on each carbon with a 1s orbital on a hydrogen. Each carbon atom still has two half-filled 2py
and 2pz orbitals, which are perpendicular both to each other and to the line formed by the sigma bonds. These two perpendicular pairs
of p orbitals form two pi bonds between the carbons, resulting in a triple bond overall (one sigma bond plus two pi bonds).
In alkene B, however, the carbon-carbon single bond is the result of overlap between an sp2 orbital and an sp3 orbital, while in alkyne
C the carbon-carbon single bond is the result of overlap between an sp orbital and an sp3 orbital. These are all single bonds, but the
single bond in molecule C is shorter and stronger than the one in B, which is in turn shorter and stronger than the one in A.
The explanation here is relatively straightforward. An sp orbital is composed of one s orbital and one p orbital, and thus it has 50% s
character and 50% p character. sp2 orbitals, by comparison, have 33% s character and 67% p character, while sp3 orbitals have 25% s
character and 75% p character. Because of their spherical shape, 2s orbitals are smaller, and hold electrons closer and ‘tighter’ to the
nucleus, compared to 2p orbitals. Consequently, bonds involving sp + sp3 overlap (as in alkyne C) are shorter and stronger than bonds
involving sp2 + sp3 overlap (as in alkene B). Bonds involving sp3-sp3overlap (as in alkane A) are the longest and weakest of the group,
because of the 75% ‘p’ character of the hybrids.
Notice that as the bond order increases the bond length decreases and the bond strength increases.
EXERCISES
QUESTIONS
Q1.9.1
1-Cyclohexyne is a very strained molecule. By looking at the molecule explain why there is
such a intermolecular strain using the knowledge of hybridization and bond angles.
SOLUTIONS
S1.9.1
The alkyne is a sp hybridized orbital. By looking at a sp orbital, we can see that the bond angle is 180°, but in cyclohexane the regular
angles would be 109.5°. Therefore the molecule would be strained to force the 180° to be a 109°.
CONTRIBUTORS
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
Prof. Steven Farmer (Sonoma State University)
William Reusch, Professor Emeritus (Michigan State U.), Virtual Textbook of Organic Chemistry
KEY TERMS
Make certain that you can define, and use in context, the key term below.
lone pair electrons
STUDY NOTES
Nitrogen is frequently found in organic compounds. As with carbon atoms, nitrogen atoms can be sp3-, sp2- or sp‑hybridized.
Note that, in this course, the term “lone pair” is used to describe an unshared pair of electrons.
The valence-bond concept of orbital hybridization can be extrapolated to other atoms including nitrogen, oxygen, phosphorus, and
sulfur. In other compounds, covalent bonds that are formed can be described using hybrid orbitals.
METHYL AMINE
The nitrogen is sp3 hybridized which means that it has four sp3 hybrid orbitals. Two of the sp3 hybridized orbitals overlap with s
orbitals from hydrogens to form the two N-H sigma bonds. One of the sp3 hybridized orbitals overlap with an sp3 hybridized orbital
from carbon to form the C-N sigma bond. The lone pair electrons on the nitrogen are contained in the last sp3 hybridized orbital. Due
to the sp3 hybridization the nitrogen has a tetrahedral geometry. However, the H-N-H and H-N-C bonds angles are less than the typical
109.5o due to compression by the lone pair electrons.
Methylamine
METHANOL
The oxygen is sp3 hybridized which means that it has four sp3 hybrid orbitals. One of the sp3 hybridized orbitals overlap with s orbitals
from a hydrogen to form the O-H sigma bonds. One of the sp3 hybridized orbitals overlap with an sp3 hybridized orbital from carbon to
form the C-O sigma bond. Both the sets of lone pair electrons on the oxygen are contained in the remaining sp3 hybridized orbital. Due
to the sp3 hybridization the oxygen has a tetrahedral geometry. However, the H-O-C bond angles are less than the typical 109.5o due to
compression by the lone pair electrons.
Methanol
METHYL PHOSPHATE
Phosphorus can have have expanded octets because it is in the n = 3 row. Typically, phosphorus forms five covalent bonds. In
biological molecules, phosphorus is usually found in organophosphates. Organophosphates are made up of a phosphorus atom bonded
to four oxygens, with one of the oxygens also bonded to a carbon. In methyl phosphate, the phosphorus is sp3 hybridized and the O-P-
O bond angle varies from 110 ° to 112o.
Methanethiol
Dimethyl sulfide
EXERCISES
QUESTIONS
Q1.10.1
Identify geometry and lone pairs on each heteroatom of the molecules given.
SOLUTIONS
S1.10.1
Diethyl ether would have two lone pairs of electrons and would have a bent geometry around the oxygen.
Dimethyl amine would have one lone pair and would show a pyramidal geometry around the nitrogen.
CONTRIBUTORS
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
Prof. Steven Farmer (Sonoma State University)
KEY TERMS
Make certain that you can define, and use in context, the key terms below.
anti-bonding molecular orbital
bonding molecular orbital
molecular orbital (MO) theory
As we have seen, valence bond theory does a remarkably good job of explaining the bonding geometry and properties of many organic
compounds. There are some areas, however, where the valence bond theory falls short. It fails to adequately account, for example, for
some interesting properties of compounds that contain alternating double and single bonds. In order to understand these properties, we
need to think about chemical bonding in a new way, using the ideas of molecular orbital (MO) theory.
Following the aufbau ('building up') principle, we place the two electrons in the H2 molecule in the lowest energy molecular orbital,
which is the (bonding) sigma orbital.
The bonding sigma orbital, which holds both electrons in the ground state of the molecule, is egg-shaped, encompassing the two nuclei,
and with the highest likelihood of electrons being in the area between the two nuclei. The high-energy, anti-bonding sigma-star orbital
can be visualized as a pair of droplets, with areas of higher electron density near each nucleus and a ‘node’, (area of zero electron
density) midway between the two nuclei.
Remember that we are thinking here about electron behavior as wave behavior. When two separate waves combine, they can do so
with what is called constructive interference, where the two amplitudes reinforce one another, or destructive interference, where the
two amplitudes cancel one another out. Bonding MO’s are the consequence of constructive interference between two atomic orbitals
which results in an attractive interaction and an increase in electron density between the nuclei. Anti-bonding MO’s are the
In the bonding Ψ1 orbital, the two shaded lobes of the 2pz orbitals interact constructively with each other, as do the two unshaded lobes
(remember, the shading choice represents mathematical (+) and (-) signs for the wavefunction). Therefore, there is increased electron
density between the nuclei in the molecular orbital – this is why it is a bonding orbital.
In the higher-energy anti-bonding Ψ2* orbital, the shaded lobe of one 2pz orbital interacts destructively with the unshaded lobe of the
second 2pz orbital, leading to a node between the two nuclei and overall repulsion.
By the aufbau principle, the two electrons from the two atomic orbitals will be paired in the lower-energy Ψ1orbital when the molecule
is in the ground state.
EXERCISES
1. Draw a simple molecular orbital diagram for each of the following molecules
a. nitrogen, N2.
b. oxygen, O2.
ANSWERS:
Answers
CONTRIBUTORS
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
Prof. Steven Farmer (Sonoma State University)
Organic Chemistry With a Biological Emphasis by Tim Soderberg (University of Minnesota, Morris)
STUDY NOTES
When drawing the structure of a neutral organic compound, you will find it helpful to remember that
each carbon atom has four bonds.
each nitrogen atom has three bonds.
each oxygen atom has two bonds.
each hydrogen atom has one bond.
It is necessary to draw structural formulas for organic compounds because in most cases a molecular formula does not uniquely
represent a single compound. Different compounds having the same molecular formula are called isomers, and the prevalence of
organic isomers reflects the extraordinary versatility of carbon in forming strong bonds to itself and to other elements. When the group
of atoms that make up the molecules of different isomers are bonded together in fundamentally different ways, we refer to such
compounds as constitutional isomers. There are seven constitutional isomers of C4H10O, and structural formulas for these are drawn
in the following table. These formulas represent all known and possible C4H10O compounds, and display a common structural feature.
There are no double or triple bonds and no rings in any of these structures.
Simplification of structural formulas may be achieved without any loss of the information they convey. In condensed structural
formulas the bonds to each carbon are omitted, but each distinct structural unit (group) is written with subscript numbers designating
multiple substituents, including the hydrogens. Shorthand (line) formulas omit the symbols for carbon and hydrogen entirely (unless
the hydrogen is bonded to an atom other than carbon). Each straight line segment represents a bond, the ends and intersections of the
lines are carbon atoms, and the correct number of hydrogens is calculated from the tetravalency of carbon. Non-bonding valence shell
electrons are omitted in these formulas.
Developing the ability to visualize a three-dimensional structure from two-dimensional formulas requires practice, and in most cases
the aid of molecular models. As noted earlier, many kinds of model kits are available to students and professional chemists, and the
beginning student is encouraged to obtain one.
KEKULÉ FORMULA
CONDENSED FORMULA
Condensed structural formulas show the order of atoms like a structural formula but are written in a single line to save space and make
it more convenient and faster to write out. Condensed structural formulas are also helpful when showing that a group of atoms is
connected to a single atom in a compound. When this happens, parenthesis are used around the group of atoms to show they are
together.
Ex. Condensed Structural Formula for Ethanol: CH3CH2OH (Molecular Formula for Ethanol C2H6O).
SHORTHAND FORMULA
Because organic compounds can be complex at times, line-angle formulas are used to write carbon and hydrogen atoms more
efficiently by replacing the letters with lines. A carbon atom is present wherever a line intersects another line. Hydrogen atoms are then
assumed to complete each of carbon's four bonds. All other atoms that are connected to carbon atoms are written out. Line angle
formulas help show structure and order of the atoms in a compound making the advantages and disadvantages similar to structural
formulas.
EXERCISES
Write down the molecular formula for each of the compounds shown here.
ANSWERS:
A. C7H7N
B. C5H10
C. C5H4O
D. C5H6Br2
QUESTIONS
Q1.12.1
Below is the molecule for caffeine. Give the molecular formula for it.
CONTRIBUTORS
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
Prof. Steven Farmer (Sonoma State University)
William Reusch, Professor Emeritus (Michigan State U.), Virtual Textbook of Organic Chemistry
Organic Chemistry With a Biological Emphasis by Tim Soderberg (University of Minnesota, Morris)
The four identical C-H single bonds in CH4 form as the result of sigma bond overlap between the sp3 hybrid orbitals of carbon and
the s orbital of each hydrogen.
1.7: sp Hybrid Orbitals and the Structure of Ethane
3
The C-C bond in C2H6 forms as the result of sigma bond overlap between a sp3 hybrid orbital on each carbon. and the s orbital of
each hydrogen. The six identical C-H single bonds in form as the result of sigma bond overlap between the sp3 hybrid orbitals of
carbon and the s orbital of each hydrogen.
1.8: sp Hybrid Orbitals and the Structure of Ethylene
2
The C=C bond in C2H4 forms as the result of both a sigma bond overlap between a sp2 hybrid orbital on each carbon and a pi bond
overlap of a p orbital on each carbon
1.9 sp Hybrid Orbitals and the Structure of Acetylene
SKILLS TO MASTER
Skill 1.1 Determine the number of protons, neutrons, and electrons in a nuclide.
Skill 1.2 Write the electron configuration and orbital diagram for an atom.
Skill 1.3 Determine the number of valence electrons in an atom.
Skill 1.4 Draw the molecular formula, Lewis Dot Structure, structural formula, condensed structural formula, shorthand formula and
wedge-dash structure of simple organic molecules.
Skill 1.5 Use Lewis Dot structures to predict molecular shape, bond angle, hybridization.
Skill 1.6 Calculate formal charge on an atom in a molecule.
Skill 1.7 Determine the number of sigma and pi bonds in organic molecules.
Skill 1.8 Determine relative bond energy and bond length based on atoms involved in the bond and bond type.
Skill 1.9 Describe and draw the orbital overlap and types of bonding in simple organic molecules like methane, ethane, ethylene and
acetylene.
Skill 1.10 Describe the bonding in organic molecules using both the Valence Bond Theory and Molecular Orbital Theory.
CONTRIBUTORS
Dr. Kelly Matthews (Professor of Chemistry, Harrisburg Area Community College)
2.4: RESONANCE
Resonance structures are a set of two or more Lewis Structures that collectively describe the electronic bonding a single polyatomic
species including fractional bonds and fractional charges. Resonance structure are capable of describing delocalized electrons that
cannot be expressed by a single Lewis formula with an integer number of covalent bonds.
KEY TERMS
Make certain that you can define, and use in context, the key terms below.
electronegativity inductive effect
polar covalent bond
STUDY NOTES
Students often wonder why it is important to be able to tell whether a given bond is polar or not, and why they need to know which
atoms carry a partial positive charge and which a partial negative charge. Consider the chloromethane (CH3Cl) molecule. The
carbon atom is shown as carrying a partial positive charge. Now, recall that opposite charges attract. Thus, it seems reasonable that
the slightly positive carbon atom in chloromethane should be susceptible to attack by a negatively charged species, such as the
hydroxide ion, OH−. This theory is borne out in practice: hydroxide ions react with chloromethane by attacking the slightly positive
carbon atom in the latter. It is often possible to rationalize chemical reactions in this manner, and you will find the knowledge of
bond polarity indispensible when you start to write reaction mechanisms.
Note: Because of the small difference in electronegativity between carbon and hydrogen, the C-H bond is normally assumed to be
nonpolar.
ELECTRONEGATIVITY
The elements with the highest ionization energies are generally those with the most negative electron affinities, which are located
toward the upper right corner of the periodic table. Conversely, the elements with the lowest ionization energies are generally those
with the least negative electron affinities and are located in the lower left corner of the periodic table.
Because the tendency of an element to gain or lose electrons is so important in determining its chemistry, various methods have been
developed to quantitatively describe this tendency. The most important method uses a measurement called electronegativity
(represented by the Greek letter chi, χ, pronounced “ky” as in “sky”), which is defined as the relative ability of an atom to attract
electrons to itself in a chemical compound. Elements with high electronegativities tend to acquire electrons in chemical reactions and
are found in the upper right corner of the periodic table. Elements with low electronegativities tend to lose electrons in chemical
reactions and are found in the lower left corner of the periodic table.
Unlike ionization energy or electron affinity, the electronegativity of an atom is not a simple, fixed property that can be directly
measured in a single experiment. In fact, an atom’s electronegativity should depend to some extent on its chemical environment
because the properties of an atom are influenced by the neighboring atoms in a chemical compound. Nevertheless, when different
methods for measuring the electronegativity of an atom are compared, they all tend to assign similar relative values to a given element.
For example, all scales predict that fluorine has the highest electronegativity and cesium the lowest of the stable elements, which
suggests that all the methods are measuring the same fundamental property.
To get a bond like this, A and B would usually have to be the same element, for example, H2 or Cl2 molecules. Note: It's important to
realize that this is an average picture. The electrons are actually in a molecular orbital, and are moving around all the time within that
That means that the B end of the bond will have a greater share of the electron density and so becomes slightly negative. At the same
time, the A end (rather short of electrons) becomes slightly positive. This unequal sharing of the bonding electrons is indicated in the
diagram above using the symbols "δ+" and "δ-". "δ" (read as "delta") means "slightly", so "δ+" means slightly positive, and "δ-"
means slightly negative. Because the electrons are shared unequally, the bond between A and B is polarized, and the bond is
called a polar covalent bond.
A polar covalent bond is a covalent bond in which there is a separation of charge between one end and the other - in other words one
end is slightly positive and the other slightly negative. Examples include most covalent bonds, like the hydrogen-chlorine bond in HCl
or the hydrogen-oxygen bonds in water.
If B is a lot more electronegative than A, then the electron pair is dragged completely over to B's end of the bond. To all intents and
purposes, A has lost control of its electron, and B has complete control over both electrons. Ions have been formed. Because the
electrons are no longer shared, the bond is an ionic bond rather than a covalent bond.
A "SPECTRUM" OF BONDS
There is no clear-cut division between covalent and ionic bonds. In a pure non-polar covalent bond, the electrons are held on average
exactly half way between the atoms. In a polar bond, the electrons have been dragged slightly towards one end. How far does this
dragging have to go before the bond counts as ionic? There is no real answer to that. Sodium chloride is typically considered an ionic
solid, but even here the sodium has not completely lost control of its electron. Because of the properties of sodium chloride, however,
we tend to count it as if it were purely ionic. Lithium iodide, on the other hand, would be described as being "ionic with some covalent
character". In this case, the pair of electrons has not moved entirely over to the iodine end of the bond. Lithium iodide, for example,
dissolves in organic solvents like ethanol - not something which ionic substances normally do. A general rule in organic chemistry is if
the bond is between metal and a non-metal atoms, then the bond should be considered ionic. Examples of this are the lithium - carbon
bond in methyl lithium and the potassium - oxygen bond in potassium tert-butoxide,
SUMMARY
No electronegativity difference between two atoms leads to a non-polar covalent bond.
A small electronegativity difference leads to a polar covalent bond.
A large electronegativity difference leads to an ionic bond.
EXERCISES
1. Identify the positive and negative ends of each of the bonds shown below.
Answer:
QUESTIONS
Q2.1.1
Rank the following from least polar to most polar using knowledge of electronegativity
CH3CH2-Li CH3CH2-K CH3CH2-F CH3CH2-OH
SOLUTIONS
S2.1.1
(least polar) OH < F < Li < K (most polar)
CONTRIBUTORS
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
Prof. Steven Farmer (Sonoma State University)
Organic Chemistry With a Biological Emphasis by Tim Soderberg (University of Minnesota, Morris)
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.
Jim Clark (Chemguide.co.uk)
KEY TERMS
Make certain that you can define, and use in context, the key term below.
dipole moment
STUDY NOTES
You must be able to combine your knowledge of molecular shapes and bond polarities to determine whether or not a given
compound will have a dipole moment. Conversely, the presence or absence of a dipole moment may also give an important clue to a
compound’s structure. BCl3, for example, has no dipole moment, while NH3 does. This suggests that in BCl3 the chlorines around
boron are in a trigonal planar arrangement, while the hydrogens around nitrogen in NH3 have a less symmetrical arrangement -
trigonal pyramidal.
Remember that the C−H bond is assumed to be non-polar.
which is slightly negative and one which is slightly positive. The whole of the outside of the molecule is somewhat negative, but
there is no overall separation of charge from top to bottom, or from left to right.
In contrast, CH Cl is a polar molecule (right panel in figure above). The hydrogen at the top of the molecule is less
3
electronegative than carbon and so is slightly positive. This means that the molecule now has a slightly positive "top" and a slightly
negative "bottom", and so is overall a polar molecule.
A polar molecule will need to be "lop-sided" in some way.
Mathematically, dipole moments are vectors; they possess both a magnitude and a direction. The dipole moment of a molecule is
therefore the vector sum of the dipole moments of the individual bonds in the molecule. If the individual bond dipole moments cancel
one another, there is no net dipole moment. Such is the case for CO2, a linear molecule (Figure 2.2.1a). Each C–O bond in CO2 is
polar, yet experiments show that the CO2 molecule has no dipole moment. Because the two C–O bond dipoles in CO2 are equal in
magnitude and oriented at 180° to each other, they cancel. As a result, the CO2 molecule has no net dipole moment even though it has a
substantial separation of charge. In contrast, the H2O molecule is not linear (Figure 2.2.1b); it is bent in three-dimensional space, so
the dipole moments do not cancel each other. Thus a molecule such as H2O has a net dipole moment. We expect the concentration of
negative charge to be on the oxygen, the more electronegative atom, and positive charge on the two hydrogens. This charge
polarization allows H2O to hydrogen-bond to other polarized or charged species, including other water molecules.
Figure 2.2.2: Molecules with Polar Bonds. Individual bond dipole moments are indicated in red. Due to their different three-
dimensional structures, some molecules with polar bonds have a net dipole moment (HCl, CH2O, NH3, and CHCl3), indicated in
purple, whereas others do not because the bond dipole moments cancel (BCl3, CCl4, PF5, and SF6).
EXAMPLE 2.2.1
Which molecule(s) has a net dipole moment?
a. H2S
b. NHF2
c. BF3
Given: three chemical compounds
Asked for: net dipole moment
Strategy:
For each three-dimensional molecular geometry, predict whether the bond dipoles cancel. If they do not, then the molecule has a net
dipole moment.
Solution:
a. The total number of electrons around the central atom, S, is eight, which gives four electron pairs. Two of these electron pairs
are bonding pairs and two are lone pairs, so the molecular geometry of H2S is bent. The bond dipoles cannot cancel one another,
so the molecule has a net dipole moment.
c. The molecular geometry of BF3 is trigonal planar. Because all the B–F bonds are equal and the molecule is highly symmetrical,
the dipoles cancel one another in three-dimensional space. Thus BF3 has a net dipole moment of zero:
EXERCISE 2.2.1
Which molecule(s) has a net dipole moment?
a. CH3Cl
b. SO3
c. XeO3
Answer: CH3Cl; XeO3
EXERCISES
1. Determine whether each of the compounds listed below possesses a dipole moment. For the polar compounds, indicate the direction
of the dipole moment.
a. O=C=O
b. I Cl
c. SO 2
d. CH −O−CH
3 3
e. CH C(=O)CH
3 3
Answers:
1. a.
b.
c.
d.
e.
QUESTIONS
Q2.2.1
The following molecule has no dipole moment in the molecule itself, explain.
Q2.2.2
Q2.2.3
Within reactions with carbonyls, such as a reduction reaction, the carbonyl is attacked from the carbon side and not the oxygen side.
Using knowledge of electronegativity explain why this happens.
SOLUTIONS
S2.2.1
The hydroxyl groups are oriented opposite of one another and therefore the dipole moments
would “cancel” one another out. Therefore having a zero net-dipole.
S2.2.2
1, 3, and 4 have a net dipoles.
S2.2.3
The oxygen is more electronegative than the carbon and therefore creates a dipole along the bond. This leads to having a partial
positive charge on the carbon and the reduction can take place.
CONTRIBUTORS
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
Prof. Steven Farmer (Sonoma State University)
Organic Chemistry With a Biological Emphasis by Tim Soderberg (University of Minnesota, Morris)
KEY TERMS
Make certain that you can define, and use in context, the key term below.
valence electrons
bonding and non-bonding electrons
formal charge
carbocations
STUDY NOTES
It is more important that students learn to easily identify atoms that have formal charges of zero, than it is to actually calculate the
formal charge of every atom in an organic compound. Students will benefit by memorizing the "normal" number of bonds and non-
bonding electrons around atoms whose formal charge is equal to zero.
A formal charge compares the number of electrons around a "neutral atom" (an atom not in a molecule) versus the number of
electrons around an atom in a molecule. Formal charge is assigned to an atom in a molecule by assuming that electrons in all chemical
bonds are shared equally between atoms, regardless of relative electronegativity. To calculate formal charges, we assign electrons in the
molecule to individual atoms according to these rules:
Non-bonding electrons are assigned to the atom on which they are located.
Bonding electrons are divided equally between the two bonded atoms, so one electron from each bond goes to each atom.
The formal charge of each atom in a molecule can be calculated using the following equation:
Formal charge = (# of valence e- in neutral atom) - [(# of non-bonding e-) + (# of bonds)] Eqn. 2.3.1
To illustrate this method, let’s calculate the formal charge on the atoms in ammonia (NH3) whose Lewis structure is as follows:
A neutral nitrogen atom has five valence electrons (it is in group 15). From the Lewis structure, the nitrogen atom in ammonia has one
lone pair and three bonds with hydrogen atoms. Substituting into Equation 2.3.1, we obtain
A neutral hydrogen atom has one valence electron. Each hydrogen atom in the molecule has no non-bonding electrons and one bond.
Using Equation 2.3.1 to calculate the formal charge on hydrogen, we obtain
The sum of the formal charges of each atom must be equal to the overall charge of the molecule or ion. In this example, the nitrogen
and each hydrogen has a formal charge of zero. When summed the overall charge is zero, which is consistent with the overall neutral
charge of the NH3 molecule.
EXAMPLE 2.3.1
Calculate the formal charges on each atom in the NH4+ ion.
Given: chemical species
Asked for: formal charges
Strategy:
Identify the number of valence electrons in each atom in the NH4+ ion. Use the Lewis electron structure of NH4+ to identify the
number of bonding and non-bonding electrons associated with each atom and then use Equation 2.3.1 to calculate the formal charge
on each atom.
Solution:
The Lewis electron structure for the NH4+ ion is as follows:
The nitrogen atom in ammonium has zero non-bonding electrons and 4 bonds. Using Equation 2.3.1, the formal charge on the
nitrogen atom is therefore
Each hydrogen atom in has one bond and zero non-bonding electrons. The formal charge on each hydrogen atom is therefore
The formal charges on the atoms in the NH4+ ion are thus
Adding together the formal charges on the atoms should give us the total charge on the molecule or ion. In this case, the sum of the
formal charges is 0 + 1 + 0 + 0 + 0 = +1, which is the same a s the overall charge of the ammonium polyatomic ion.
EXERCISE 2.3.1
Write the formal charges on all atoms in BH4−.
Answer:
And yet, organic chemists, and especially organic chemists dealing with biological molecules, are expected to draw the structure of
large molecules such as this on a regular basis. Clearly, you need to develop the ability to quickly and efficiently draw large structures
and determine formal charges. Fortunately, this only requires some practice with recognizing common bonding patterns.
CARBON
Carbon, the most important element for organic chemists. In the structures of methane, methanol, ethane, ethene, and ethyne, there are
four bonds to the carbon atom. And each carbon atom has a formal charge of zero. In other words, carbon is tetravalent, meaning that
it commonly forms four bonds.
Carbon is tetravalent in most organic molecules, but there are exceptions. Later in this chapter and throughout this book are examples
of organic ions called ‘carbocations’ and carbanions’, in which a carbon atom has a positive or negative formal charge, respectively.
Carbocations occur when a carbon has only three bonds and no lone pairs of electrons. Carbocations have only 6 valence electrons
and a formal charge of +1. Carbanions occur when the carbon atom has three bonds plus one lone pair of electrons. Carbanions have 8
valence electrons and a formal charge of -1.
Two other possibilities are carbpon radicals and carbenes, both of which have a formal charge of zero. A carbon radical has three
bonds and a single, unpaired electron. Carbon radicals have 7 valence electrons and a formal charge of zero. Carbenes are a highly
reactive species, in which a carbon atom has two bonds and one lone pair of electrons, giving it a formal charge of zero. You may
encounter carbenes in more advanced chemistry courses, but they will not be discussed any further in this book.
HYDROGEN
The common bonding pattern for hydrogen is easy: hydrogen atoms in organic molecules typically have only one bond, no unpaired
electrons and a formal charge of zero. The exceptions to this rule are the proton, H+, the hydride ion, H-, and the hydrogen radical, H..
The proton is a hydrogen with no bonds and no lone pairs and a formal charge of +1. The hydride ion is a is a hydrogen with no
bonds, a pair of electrons, and a formal charge of -1. The hydrogen radical is a hydrogen atom with no bonds, a single unpaired
electron and a formal charge of 0. Because this book concentrates on organic chemistry as applied to living things, however, we will
not be seeing ‘naked’ protons and hydrides as such, because they are too reactive to be present in that form in aqueous solution.
Nonetheless, the idea of a proton will be very important when we discuss acid-base chemistry, and the idea of a hydride ion will
become very important much later in the book when we discuss organic oxidation and reduction reactions. As a rule, though, all
hydrogen atoms in organic molecules have one bond, and no formal charge.
OXYGEN
The common arrangement of oxygen that has a formal charge of zero is when the oxygen atom has 2 bonds and 2 lone pairs. Other
arrangements are oxygen with 1 bond and 3 lone pairs, that has a -1 formal charge, and oxygen with 3 bonds and 1 lone pair that has a
formal charge of +1. All three patterns of oxygen fulfill the octet rule.
If it has two bonds and two lone pairs, as in water, it will have a formal charge of zero. If it has one bond and three lone pairs, as in
hydroxide ion, it will have a formal charge of-1. If it has three bonds and one lone pair, as in hydronium ion, it will have a formal
charge of +1.
When we get to our discussion of free radical chemistry in chapter 17, we will see other possibilities, such as where an oxygen atom
has one bond, one lone pair, and one unpaired (free radical) electron, giving it a formal charge of zero. For now, however, concentrate
on the three main non-radical examples, as these will account for virtually everything we see until chapter 17.
NITROGEN
Nitrogen has two major bonding patterns, both of which fulfill the octet rule:
If a nitrogen has three bonds and a lone pair, it has a formal charge of zero. If it has four bonds (and no lone pair), it has a formal
charge of +1. In a fairly uncommon bonding pattern, negatively charged nitrogen has two bonds and two lone pairs.
Two third row elements are commonly found in biological organic molecules: sulfur and phosphorus. Although both of these elements
have other bonding patterns that are relevant in laboratory chemistry, in a biological context sulfur almost always follows the same
bonding/formal charge pattern as oxygen, while phosphorus is present in the form of phosphate ion (PO43-), where it has five bonds
HALOGENS
The halogens (fluorine, chlorine, bromine, and iodine) are very important in laboratory and medicinal organic chemistry, but less
common in naturally occurring organic molecules. Halogens in organic compounds usually are seen with one bond, three lone pairs,
and a formal charge of zero. Sometimes, especially in the case of bromine, we will encounter reactive species in which the halogen has
two bonds (usually in a three-membered ring), two lone pairs, and a formal charge of +1.
These rules, if learned and internalized so that you don’t even need to think about them, will allow you to draw large organic structures,
complete with formal charges, quite quickly.
Once you have gotten the hang of drawing Lewis structures, it is not always necessary to draw lone pairs on heteroatoms, as you can
assume that the proper number of electrons are present around each atom to match the indicated formal charge (or lack thereof).
Occasionally, though, lone pairs are drawn if doing so helps to make an explanation more clear.
CO2
This structure has an octet of electrons around each O atom but only 4 electrons around the C atom.
Both Lewis electron structures give all three atoms an octet. How do we decide between these two possibilities? The formal charges for
the two Lewis electron structures of CO2 are as follows:
Both Lewis structures have a net formal charge of zero, but the structure on the right has a +1 charge on the more electronegative atom
(O). Thus the symmetrical Lewis structure on the left is predicted to be more stable, and it is, in fact, the structure observed
NOTE
The Lewis structure with the set of formal charges closest to zero is usually the most stable.
EXAMPLE 2.3.2
The thiocyanate ion (SCN−), which is used in printing and as a corrosion inhibitor against acidic gases, has at least two possible
Lewis electron structures. Draw two possible structures, assign formal charges on all atoms in both, and decide which is the
preferred arrangement of electrons.
Given: chemical species
Asked for: Lewis electron structures, formal charges, and preferred arrangement
Strategy:
A Use the step-by-step procedure to write two plausible Lewis electron structures for SCN−.
B Calculate the formal charge on each atom using Equation 2.3.1.
C Predict which structure is preferred based on the formal charge on each atom and its electronegativity relative to the other atoms
present.
Solution:
A Possible Lewis structures for the SCN− ion are as follows:
B We must calculate the formal charges on each atom to identify the more stable structure. If we begin with carbon, we notice that
the carbon atom in each of these structures shares four bonding pairs, the number of bonds typical for carbon, so it has a formal
charge of zero. Continuing with sulfur, we observe that in (a) the sulfur atom shares one bonding pair and has three lone pairs and
has a total of six valence electrons. The formal charge on the sulfur atom is therefore 6 - (6 + 2/2) = -1. In (b), the sulfur atom has a
formal charge of 0. In (c), the sulfur atom has a formal charge of +1. Continuing with the nitrogen, we observe that in (a) the
nitrogen atom shares three bonding pairs and has one lone pair and has a total of 5 valence electrons. The formal charge on the
nitrogen atom is therefore 5 - (2 + 6/2) = 0. In (b), the nitrogen atom has a formal charge of -1. In (c), the nitrogen atom has a
formal charge of -2.
C Which structure is preferred? Structure (b) is preferred because the negative charge is on the more electronegative atom (N), and
it has lower formal charges on each atom as compared to structure (c): 0, −1 versus +1, −2.
EXERCISE
Salts containing the fulminate ion (CNO−) are used in explosive detonators. Draw three Lewis electron structures for CNO− and use
formal charges to predict which is more stable. (Note: N is the central atom.)
Answer:
EXERCISES
1. Draw the Lewis structure of each of the molecules listed below.
CH3+, NH2−, CH3−, NH4+, BF4−
In each case, use the method of calculating formal charge described to satisfy yourself that the structures you have drawn do in fact
carry the charges shown.
Answer:
QUESTIONS
Q2.3.1
Give the formal charges for all non-hydrogen atoms in the following moelcules:
BH4−, H2O, CH3O−
SOLUTIONS
S2.3.1
BH4− (B = −1)
H2O (O = 0)
CH3O− (C = 0, O = −1)
CONTRIBUTORS
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
Prof. Steven Farmer (Sonoma State University)
William Reusch, Professor Emeritus (Michigan State U.), Virtual Textbook of Organic Chemistry
Organic Chemistry With a Biological Emphasis by Tim Soderberg (University of Minnesota, Morris)
KEY TERMS
Make certain that you can define, and use in context, the key term below.
resonance form
delocalization
Sometimes, even when formal charges are considered, the bonding in some molecules or ions cannot be described by a single Lewis
structure. Resonance is a way of describing delocalized electrons within certain molecules or polyatomic ions where the bonding
cannot be expressed by a single Lewis formula. A molecule or ion with such delocalized electrons is represented by several
contributing structures (also called resonance structures or canonical forms). Such is the case for ozone (O3), an allotrope of oxygen
with a V-shaped structure and an O–O–O angle of 117.5°.
O3
1. We know that ozone has a V-shaped structure, so one O atom is central:
5. At this point, both terminal oxygen atoms have octets of electrons. We therefore place the last 2 electrons on the central atom:
6. The central oxygen has only 6 electrons. We must convert one lone pair on a terminal oxygen atom to a bonding pair of electrons—
but which one? Depending on which one we choose, we obtain either
Before the development of quantum chemistry it was thought that the double-headed arrow indicates that the actual electronic structure
is an average of those shown, or that the molecule oscillates between the two structures. Today we know that the electrons involved in
the double bonds occupy an orbital that extends over all three oxygen molecules, combining p orbitals on all three (Figure 2.4.1).
Figure 2.4.1: The resonance structure of ozone involves a molecular orbital extending all three oxygen atoms.In ozone, a molecular
orbital extending over all three oxygen atoms is formed from three atom centered pz orbitals. Similar molecular orbitals are found in
every resonance structure.
We will discuss the formation of these molecular orbitals in the next chapter but it is important to understand that resonance structures
are based on molecular orbitals not averages of different bonds between atoms. We describe the electrons in such molecular orbitals as
being delocalized, that is they cannot be assigned to a specific bond between two atoms.
When it is possible to write more than one equivalent resonance structure for a molecule or ion, the actual structure involves a
molecular orbital which is a linear combination of atomic orbitals from each of the atoms.
CO32−
Like ozone, the electronic structure of the carbonate ion cannot be described by a single Lewis electron structure. Unlike O3, though,
the Lewis structures describing CO32− has three equivalent representations.
1. Because carbon is the least electronegative element, we place it in the central position:
2. Carbon has 4 valence electrons, each oxygen has 6 valence electrons, and there are 2 more for the −2 charge. This gives 4 + (3 × 6)
+ 2 = 24 valence electrons.
3. Six electrons are used to form three bonding pairs between the oxygen atoms and the carbon:
4. We divide the remaining 18 electrons equally among the three oxygen atoms by placing three lone pairs on each and indicating the
−2 charge:
As with ozone, none of these structures describes the bonding exactly. Each predicts one carbon–oxygen double bond and two carbon–
oxygen single bonds, but experimentally all C–O bond lengths are identical. We can write resonance structures (in this case, three of
them) for the carbonate ion:
As the case for ozone, the actual structure involves the formation of a molecular orbital from pz orbitals centered on each atom and
sitting above and below the plane of the CO32− ion.
EXAMPLE 2.4.1
Benzene is a common organic solvent that was previously used in gasoline; it is no longer used for this purpose, however, because it
is now known to be a carcinogen. The benzene molecule (C6H6) consists of a regular hexagon of carbon atoms, each of which is
also bonded to a hydrogen atom. Use resonance structures to describe the bonding in benzene.
Given: molecular formula and molecular geometry
Asked for: resonance structures
Strategy:
A Draw a structure for benzene illustrating the bonded atoms. Then calculate the number of valence electrons used in this drawing.
B Subtract this number from the total number of valence electrons in benzene and then locate the remaining electrons such that each
atom in the structure reaches an octet.
C Draw the resonance structures for benzene.
Solution:
A Each hydrogen atom contributes 1 valence electron, and each carbon atom contributes 4 valence electrons, for a total of (6 × 1) +
(6 × 4) = 30 valence electrons. If we place a single bonding electron pair between each pair of carbon atoms and between each
carbon and a hydrogen atom, we obtain the following:
Each carbon atom in this structure has only 6 electrons and has a formal charge of +1, but we have used only 24 of the 30 valence
electrons.
B If the 6 remaining electrons are uniformly distributed pair-wise on alternate carbon atoms, we obtain the following:
Each structure has alternating double and single bonds, but experimentation shows that each carbon–carbon bond in benzene is
identical, with bond lengths (139.9 pm) intermediate between those typically found for a C–C single bond (154 pm) and a C=C
double bond (134 pm). We can describe the bonding in benzene using the two resonance structures, but the actual electronic
structure is an average of the two. The existence of multiple resonance structures for aromatic hydrocarbons like benzene is often
indicated by drawing either a circle or dashed lines inside the hexagon:
This combination of p orbitals for benzene can be visualized as a ring with a node in the plane of the carbon atoms.
EXERCISE 2.4.1
The sodium salt of nitrite is used to relieve muscle spasms. Draw two resonance structures for the nitrite ion (NO2−).
Answer:
Resonance structures are particularly common in oxyanions of the p-block elements, such as sulfate and phosphate, and in aromatic
hydrocarbons, such as benzene and naphthalene.
EXERCISES
QUESTIONS
Q2.4.1
Draw the resonance structures for the following molecule:
SOLUTIONS
S2.4.1
CONTRIBUTORS
The above resonance structures show that the electrons are delocalized within the molecule and through this process the molecule gains
extra stability. Ozone with both of its opposite formal charges creates a neutral molecule and through resonance it is a stable molecule.
The extra electron that created the negative charge one terminal oxygen can be delocalized by resonance through the other terminal
oxygen.
Benzene is an extremely stable molecule due to its geometry and molecular orbital interactions, but most importantly, due to its
resonance structures. The delocalized electrons in the benzene ring make the molecule very stable and with its characteristics of a
nucleophile, it will react with a strong electrophile only and after the first reactivity, the substituted benzene will depend on its
resonance to direct the next position for the reaction to add a second substituent.
The Hybrid Resonance forms show the different Lewis structures with the electron been delocalized. This is very important for the
reactivity of chloro-benzene because in the presence of an electrophile it will react and the formation of another bond will be
directed and determine by resonance. The lone pair of electrons delocalized in the aromatic substituted ring is where it can
potentially form a new bond with an electrophile, as it is shown there are three possible places that reactivity can take place, the
first to react will take place at the para position with respect to the chloro substituent and then to either ortho position.
EXERCISES
QUESTIONS
Q2.5.1
Are all the bond lengths the same in the following molecule?
SOLUTIONS
S2.5.1
CONTRIBUTORS
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
Prof. Steven Farmer (Sonoma State University)
William Reusch, Professor Emeritus (Michigan State U.), Virtual Textbook of Organic Chemistry
KEY WORDS
resonance structure
resonance hybrid
Resonance is a mental exercise and method within the Valence Bond Theory of bonding that describes the delocalization of electrons
within molecules. It compares and contrasts two or more possible Lewis structures that can represent a particular molecule. Resonance
structures are used when one Lewis structure for a single molecule cannot fully describe the bonding that takes place between
neighboring atoms relative to the empirical data for the actual bond lengths between those atoms. The net sum of valid resonance
structures is defined as a resonance hybrid, which represents the overall delocalization of electrons within the molecule. A molecule
that has several resonance structures is more stable than one with fewer. Some resonance structures are more favorable than others.
INTRODUCTION
Electrons have no fixed position in atoms, compounds and molecules (see image below) but have probabilities of being found in
certain spaces (orbitals). Resonance forms illustrate areas of higher probabilities (electron densities). This is like holding your hat in
either your right hand or your left. The term Resonance is applied when there are two or more possibilities available. Resonance
structures do not change the relative positions of the atoms like your arms in the metaphor. The skeleton of the Lewis Structure remains
the same, only the electron locations change. A double headed arrow on both ends of the arrow ( ↔ ) between Lewis structures is used
to show their inter-connectivity. It is different from the double harpoons ( ⇌ ) used for designating equilibria. A double headed arrow
on only one end ( → ) is used to indicate the movement of two electrons in a single resonance structure.
Figure: This is an animation of how one can do a resonance with ozone by moving electrons.
FORMAL CHARGE
Even though the structures look the same, the formal charge (FC) may not be. Formal charges are charges that are assigned to a
specific atom in a molecule. If computed correctly, the overall formal charge of the molecule should be the same as the oxidation
charge of the molecule (the charge when you write out the empirical and molecular formula) We want to choose the resonance
structure with the least formal charges that add up to zero or the charge of the overall molecule.
The equation for finding Formal Charge is:
Formal Charge = (number of valence electrons in free orbital) - (number of lone-pair electrons) - ( number bond pair
1
electrons)
The formal charge has to equal the molecule's overall charge.
Ex.) CNS- has an overall charge of -1, so the Lewis structure's formal charge has to equal -1. See Lewis Structure for more information.
2. Resonance: All elements want an octet, and we can do that in multiple ways by moving the terminal atom's electrons around
(bonds too).
2
number bond pair electrons)
Remember to determine the number of valence electron each atom has before assigning Formal Charges
C = 4 valence e-, N = 5 valence e-, S = 6 valence e-, also add an extra electron for the (-1) charge. The total of valence electrons is
16.
4. Find the most ideal resonance structure. (Note: It is the one with the least formal charges that adds up to zero or to the molecule's
overall charge.)
Step 2: Combine the resonance structures by adding (dotted) bonds where other resonance bonds can be formed.
Step 3: Add only the lone pairs found on ALL resonance structures.
EXAMPLE 2.6.4
The above resonance structures show that the electrons are delocalized within the molecule and through this process the molecule
gains extra stability. Ozone with both of its opposite charges creates a neutral molecule and through resonance it is a stable
molecule. The extra electron that created the negative charge on either terminal oxygen can be delocalized by resonance through the
terminal oxygens.
Benzene is an extremely stable molecule and it is accounted for its geometry and molecular orbital interaction, but most importantly
it's due to its resonance structures. The delocalized electrons in the benzene ring make the molecule very stable and with its
characteristics of a nucleophile, it will react with a strong electrophile only and after the first reactivity, the substituted benzene will
depend on its resonance to direct the next position for the reaction to add a second substituent.
The next molecule, the Amide, is a very stable molecule that is present in most biological systems, mainly in proteins. By studies of
NMR spectroscopy and X-Ray crystallography it is confirmed that the stability of the amide is due to resonance which through
molecular orbital interaction creates almost a double bond between the Nitrogen and the carbon.
The Hybrid Resonance forms show the different Lewis structures with the electron been delocalized. This is very important for the
reactivity of chloro-benzene because in the presence of an electrophile it will react and the formation of another bond will be
directed and determine by resonance. The lone pair of electrons delocalized in the aromatic substituted ring is where it can
potentially form a new bond with an electrophile, as it is shown there are three possible places that reactivity can take place, the
first to react will take place at the para position with respect to the chloro substituent and then to either ortho position.
REFERENCES
1. Petrucci, Ralph H., et al. General Chemistry: Principles and Modern Applications. New Jersey: Pearson Prentice Hall, 2007. Print.
2. Ahmad, Wan-Yaacob and Zakaria, Mat B. "Drawing Lewis Structures from Lewis Symbols: A Direct Electron Pairing Approach."
Journal of Chemical Education: Journal 77.3: n. pag. Web. March 2000. Link to this journal:
http://pkukmweb.ukm.my/~mbz/c_penerb...83%29/p329.pdf
OUTSIDE LINKS
1. http://en.wikipedia.org/wiki/Resonance_(chemistry)
2. http://www.absoluteastronomy.com/topics/Resonance_(chemistry)#encyclopedia
3. http://www.nku.edu/~russellk/tutorial/reson/resonance.html
4. http://en.wikipedia.org/wiki/Formal_charge
5. http://commons.wikimedia.org/wiki/Main_Page (for the electronegatvity chart)
6. http://misterguch.brinkster.net/PRA037.pdf (for problem 5)
7. http://commons.wikimedia.org/wiki/File:Phosphite-ion-resonance-structures-2D.png (for the (HPO32-) problem 4 answer)
8. http://commons.wikimedia.org/wiki/File:Sulfate-resonance-2D.png (for the Sulfate answer)
9. http://www.mpcfaculty.net/mark_bishop/resonance.htm
10. http://www.chem.ucla.edu/harding/tutorials/resonance/draw_res_str.html
PROBLEMS
1. True or False, The picture below is a resonance structure?
ANSWERS
1. False, because the electrons were not moved around, only the atoms (this violates the Resonance Structure Rules).
2. Below are the all Lewis dot structure with formal charges (in red) for Sulfate (SO42-). There isn't a most favorable resonance of the
Sulfate ion because they are all identical in charge and there is no change in Electronegativity between the Oxygen atoms.
3. Below is the resonance for CH3COO-, formal charges are displayed in red. The Lewis Structure with the most formal charges is not
desirable, because we want the Lewis Structure with the least formal charge.
4. The resonance for HPO32-, and the formal charges (in red).
CONTRIBUTORS
Sharon Wei (UCD), Liza Chu (UCD)
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
Prof. Steven Farmer (Sonoma State University)
KEY TERMS
Make certain that you can define, and use in context, the key terms below.
Brønsted-Lowry acid
Brønsted-Lowry base
conjugate acid
conjugate base
STUDY NOTES
You should already be familiar with the Brønsted-Lowry concept of acidity and the differences between strong and weak acids. You
may wish to review this topic before proceeding.
In 1923, chemists Johannes Brønsted and Martin Lowry independently developed definitions of acids and bases based on compounds
abilities to either donate or accept protons (H+ ions). Here, acids are defined as being able to donate protons in the form of hydrogen
ions; whereas bases are defined as being able to accept protons. This took the Arrhenius definition one step further as water is no
longer required to be present in the solution for acid and base reactions to occur.
BRØNSTED-LOWERY DEFINITION
J.N. Brønsted and T.M. Lowry independently developed the theory of proton donors and proton acceptors in acid-base reactions,
coincidentally in the same region and during the same year. The Arrhenius theory where acids and bases are defined by whether the
molecule produces hydrogen ion or hydroxide ion when dissolved in water was too limiting, because not all chemical reactions,
especially organic reactions, occur in water. The Brønsted-Lowry Theory defines an acid a proton donor, while a base is a proton
acceptor. This is illustrated in the following reactions:
+ −
H Cl + H OH → H3 O + Cl (2.7.1)
+ −
H OH + N H3 → N H + OH (2.7.2)
4
Acid Base
Donates hydrogen ions Accepts hydrogen ions.
The determination of a substance as a Brønsted-Lowry acid or base can only be done by examining the reaction, since many chemicals
can be either an acid or a base. For example, HOH is a base in the first reaction and an acid in the second reaction.
Similarly, in the reaction of acetic acid with water, acetic acid donates a proton to water, which acts as the base. In the reverse reaction,
H3O+ is the acid that donates a proton to the acetate ion, which acts as the base. Once again, we have two conjugate acid–base pairs:
the parent acid and its conjugate base (CH3CO2H/CH3CO2-) and the parent base and its conjugate acid (H3O+/H2O).
Some common conjugate acid–base pairs are shown in Figure 2.7.1. The strongest acids are at the bottom left, and the strongest bases
are at the top right. The conjugate base of a strong acid is a very weak base, and, conversely, the conjugate acid of a strong base is a
very weak acid.
Figure 2.7.1 The Relative Strengths of Some Common Conjugate Acid–Base Pairs
EXERCISES
1. Identify the Brønsted-Lowry acids and bases in the reactions given below.
a. $\ce{\sf{CH3CH2O^- + H2O <=> CH3CH2OH + OH^- }}$
b. $\ce{\sf{CH3CH2OH + H2SO4 <=> CH3CH2OH2+ + HSO4- }}$
Answer:
1. a.
QUESTIONS
Q2.7.1
Is the following molecule a Brønsted acid or base?
HSO4−
SOLUTIONS
S2.7.1
It can be both, consider the following schemes:
CONTRIBUTORS
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
Organic Chemistry With a Biological Emphasis by Tim Soderberg (University of Minnesota, Morris)
Prof. Steven Farmer (Sonoma State University)
KEY TERMS
Make certain that you can define, and use in context, the key terms below.
acidity constant, Ka
equilibrium constant, Keq
STUDY NOTES
Calculations and expressions involving Ka and pKa were covered in detail in your first-year general chemistry course. Note that
acidity constant is also known as the acid dissociation constant.
You are no doubt aware that some acids are stronger than others. Sulfuric acid is strong enough to be used as a drain cleaner, as it will
rapidly dissolve clogs of hair and other organic material.
Not surprisingly, concentrated sulfuric acid will also cause painful burns if it touches your skin, and permanent damage if it gets in
your eyes (there’s a good reason for those safety goggles you wear in chemistry lab!). Acetic acid (vinegar), will also burn your skin
and eyes, but is not nearly strong enough to make an effective drain cleaner. Water, which we know can act as a proton donor, is
obviously not a very strong acid. Even hydroxide ion could theoretically act as an acid – it has, after all, a proton to donate – but this is
not a reaction that we would normally consider to be relevant in anything but the most extreme conditions.
The relative acidity of different compounds or functional groups – in other words, their relative capacity to donate a proton to a
common base under identical conditions – is quantified by a number called the acid dissociation constant, abbreviated Ka. The
common base chosen for comparison is water.
We will consider acetic acid as our first example. When a small amount of acetic acid is added to water, a proton-transfer event (acid-
base reaction) occurs to some extent.
Notice the phrase ‘to some extent’ – this reaction does not run to completion, with all of the acetic acid converted to acetate, its
conjugate base. Rather, a dynamic equilibrium is reached, with proton transfer going in both directions (thus the two-way arrows) and
finite concentrations of all four species in play. The nature of this equilibrium situation, as you recall from General Chemistry, is
expressed by an equilibrium constant, K.
The equilibrium constant is actually a ratio of activities (represented by the symbol a), but activities are rarely used in courses other
than analytical or physical chemistry. To simplify the discussion for general chemistry and organic chemistry courses, the activities of
all of the solutes are replaced with molarities, and the activity of the solvent (usually water) is defined as having the value of 1.
Because dividing by 1 does not change the value of the constant, the "1" is usually not written, and Ka is written as:
− +
[CH3 COO ][ H3 O ]
−5
Keq = Ka = = 1.75 × 10 (2.8.2)
[CH3 COOH ]
In more general terms, the dissociation constant for a given acid is expressed as:
− +
[A ][ H3 O ]
Ka = (2.8.3)
[H A]
or
+
[A][ H3 O ]
Ka = (2.8.4)
+
[H A ]
Equation 2.8.3 applies to a neutral acid such as like HCl or acetic acid, while Equation 2.8.4 applies to a cationic acid like ammonium
(NH4+).
The value of Ka = 1.75 x 10-5 for acetic acid is very small - this means that very little dissociation actually takes place, and there is
much more acetic acid in solution at equilibrium than there is acetate ion. Acetic acid is a relatively weak acid, at least when compared
to sulfuric acid (Ka = 109) or hydrochloric acid (Ka = 107), both of which undergo essentially complete dissociation in water.
A number like 1.75 x 10- 5 is not very easy either to say or to remember. Chemists often use pKa values as a more convenient term to
express relative acidity. pKa is related to Ka by the following equation
p Ka = − log Ka (2.8.5)
Doing the math, we find that the pKa of acetic acid is 4.8. The use of pKa values allows us to express the acidity of common
compounds and functional groups on a numerical scale of about –10 (very strong acid) to 50 (not acidic at all). Table 2.8.1 at the end
of the text lists exact or approximate pKa values for different types of protons that you are likely to encounter in your study of organic
and biological chemistry. Looking at Table 2.8.1, you see that the pKa of carboxylic acids are in the 4-5 range, the pKa of sulfuric acid
is –10, and the pKa of water is 14. Alkenes and alkanes, which are not acidic at all, have pKa values above 30. The lower the pKa value,
the stronger the acid.
Table 2.8.1: Representative acid constants
Any particular acid will always have the same pKa (assuming that we are talking about an aqueous solution at room temperature) but
different aqueous solutions of the acid could have different pH values, depending on how much acid is added to how much water.
Our table of pKa values will also allow us to compare the strengths of different bases by comparing the pKavalues of their conjugate
acids. The key idea to remember is this: the stronger the conjugate acid, the weaker the conjugate base. Sulfuric acid is the strongest
acid on our list with a pKa value of –10, so HSO4- is the weakest conjugate base. You can see that hydroxide ion is a stronger base than
ammonia (NH3), because ammonium (NH4+, pKa = 9.2) is a stronger acid than water (pKa = 14.0).
The stronger the conjugate acid, the weaker the conjugate base.
While Table 2.8.1 provides the pKa values of only a limited number of compounds, it can be very useful as a starting point for
estimating the acidity or basicity of just about any organic molecule. Here is where your familiarity with organic functional groups will
come in very handy. What, for example, is the pKaof cyclohexanol? It is not on the table, but as it is an alcohol it is probably
somewhere near that of ethanol (pKa = 16). Likewise, we can use Table 2.8.1 to predict that para-hydroxyphenyl acetaldehyde, an
intermediate compound in the biosynthesis of morphine, has a pKa in the neighborhood of 10, close to that of our reference compound,
phenol.
Notice in this example that we need to evaluate the potential acidity at four different locations on the molecule.
Answer:
a. The most acidic group is the protonated amine, pKa ~ 5-9
b. Alpha proton by the C=O group, pKa ~ 18-20
c. Thiol, pKa ~ 10
d. Carboxylic acid, pKa ~ 5
e. Carboxylic acid, pKa ~ 5
EXERCISES
1. Write down an expression for the acidity constant of acetic acid, CH3COOH.
ANSWERS
− + − +
[CH3 CO ][ H ] [CH3 CO ][ H3 O ]
2 2
1. Ka = or K
a =
[CH3 CO2 H ] [CH3 CO2 H ]
2. p K a = − log10 Ka = 4.74
−5
Thus, log10 Ka = − 4.72 and K a = anti-log(−4.72)= 1.9 × 10
4. Benzoic acid is stronger than acetic acid. [Benzoic acid has a higher Ka and a lower pKa.]
CONTRIBUTORS
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
Prof. Steven Farmer (Sonoma State University)
Organic Chemistry With a Biological Emphasis by Tim Soderberg (University of Minnesota, Morris)
Tom Neils (Grand Rapids Community College)
KEY TERMS
Make certain that you can define, and use in context, the key term below.
pKa
First, we need to identify the acid species on either side of the equation. On the left side, the acid is of course acetic acid, while on the
right side the acid is methyl ammonium. The specific pKa values for these acids are not on our very generalized pKa table, but are
given in the figure above. Without performing any calculations, you should be able to see that this equilibrium lies far to the right-hand
side: acetic acid has a lower pKa, is a stronger acid, and thus it wants to give up its proton more than methyl ammonium does. Doing
the math, we see that
ΔpKa 10.6–4.8 5.8 5
Keq = 10 = 10 = 10 = 6.3 × 10 (2.9.1)
So K is a very large number (much greater than 1) and the equilibrium lies far to the right-hand side of the equation, just as we had
eq
predicted.
If you had just wanted to approximate an answer without bothering to look for a calculator, you could have noted that the difference
in pKa values is approximately 6, so the equilibrium constant should be somewhere in the order of 106, or one million. Using the
pKa table in this way, and making functional group-based pKa approximations for molecules for which we don’t have exact values,
we can easily estimate the extent to which a given acid-base reaction will proceed.
EXAMPLE 2.9.1
Show the products of the following acid-base reactions, and estimate the value of Keq. Use the pKa table from Section 2.8 and/or
from the Reference Tables.
EXERCISES
QUESTIONS
Q2.9.1
In the following reactions give the resulting products and label the conjugate acid and bases.
CONTRIBUTORS
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
Prof. Steven Farmer (Sonoma State University)
Organic Chemistry With a Biological Emphasis by Tim Soderberg (University of Minnesota, Morris)
This page explains the acidity of simple organic acids and looks at the factors which affect their relative strengths.
A hydronium ion is formed together with the anion (negative ion) from the acid. This equilibrium is sometimes simplified by leaving
out the water to emphasize the ionization of the acid.
− +
AH(aq) ⇌ A +H (2.10.2)
(aq) (aq)
If you write it like this, you must include the state symbols - "(aq)". Writing H+(aq) implies that the hydrogen ion is attached to a water
molecule as H3O+. Hydrogen ions are always attached to something during chemical reactions.
The organic acids are weak in the sense that this ionization is very incomplete. At any one time, most of the acid will be present in the
solution as un-ionized molecules. For example, in the case of dilute ethanoic acid, the solution contains about 99% of ethanoic acid
molecules - at any instant, only about 1% have actually ionised. The position of equilibrium therefore lies well to the left.
Remember - the smaller the number the stronger the acid. Comparing the other two to ethanoic acid, you will see that phenol is very
much weaker with a pKa of 10.00, and ethanol is so weak with a pKa of about 16 that it hardly counts as acidic at all!
NOTE
The smaller the pK , the stronger the acid
a
The acidic hydrogen is the one attached to the oxygen. When ethanoic acid ionises it forms the ethanoate ion, CH3COO-.
You might reasonably suppose that the structure of the ethanoate ion was as below, but measurements of bond lengths show that the
two carbon-oxygen bonds are identical and somewhere in length between a single and a double bond.
To understand why this is, you have to look in some detail at the bonding in the ethanoate ion. Like any other double bond, a
carbon-oxygen double bond is made up of two different parts. One electron pair is found on the line between the two nuclei - this is
known as a sigma bond. The other electron pair is found above and below the plane of the molecule in a pi bond. Pi bonds are made
by sideways overlap between p orbitals on the carbon and the oxygen.
In an ethanoate ion, one of the lone pairs on the negative oxygen ends up almost parallel to these p orbitals, and overlaps with them
This leads to a delocalised pi system over the whole of the -COO- group, rather like that in benzene.
All the oxygen lone pairs have been left out of this diagram to avoid confusion. Because the oxygens are more electronegative than
the carbon, the delocalised system is heavily distorted so that the electrons spend much more time in the region of the oxygen
atoms.
The dotted line represents the delocalisation. The negative charge is written centrally on that end of the molecule to show that it
isn't localised on one of the oxygen atoms. The more you can spread charge around, the more stable an ion becomes. In this case, if
you delocalise the negative charge over several atoms, it is going to be much less attractive to hydrogen ions - and so you are less
likely to re-form the ethanoic acid.
When the hydrogen-oxygen bond in phenol breaks, you get a phenoxide ion, C6H5O-. Delocalization also occurs in this ion. This
time, one of the lone pairs on the oxygen atom overlaps with the delocalised electrons on the benzene ring.
This overlap leads to a delocalisation which extends from the ring out over the oxygen atom. As a result, the negative charge is no
longer entirely localised on the oxygen, but is spread out around the whole ion.
But the delocalisation spreads this charge over the whole of the COO group. Because oxygen is more electronegative than carbon,
you can think of most of the charge being shared between the two oxygens (shown by the heavy red shading in this diagram).
If there wasn't any delocalisation, one of the oxygens would have a full charge which would be very attractive towards hydrogen
ions. With delocalisation, that charge is spread over two oxygen atoms, and neither will be as attractive to a hydrogen ion as if one
of the oxygens carried the whole charge.
That means that the ethanoate ion won't take up a hydrogen ion as easily as it would if there wasn't any delocalisation. Because
some of it stays ionised, the formation of the hydrogen ions means that it is acidic.
In the phenoxide ion, the single oxygen atom is still the most electronegative thing present, and the delocalised system will be
heavily distorted towards it. That still leaves the oxygen atom with most of its negative charge.
This has nothing at all going for it. There is no way of delocalizing the negative charge, which remains firmly on the oxygen atom.
That intense negative charge will be highly attractive towards hydrogen ions, and so the ethanol will instantly re-form. Since
ethanol is very poor at losing hydrogen ions, it is hardly acidic at all.
NOTE
Remember that the higher the value for pKa, the weaker the acid is.
Why is ethanoic acid weaker than methanoic acid? It again depends on the stability of the anions formed - on how much it is possible
to delocalise the negative charge. The less the charge is delocalised, the less stable the ion, and the weaker the acid.
The methanoate ion (from methanoic acid) is:
The only difference between this and the ethanoate ion is the presence of the CH3 group in the ethanoate. But that's important! Alkyl
groups have a tendency to "push" electrons away from themselves. That means that there will be a small amount of extra negative
charge built up on the -COO- group. Any build-up of charge will make the ion less stable, and more attractive to hydrogen ions.
Ethanoic acid is therefore weaker than methanoic acid, because it will re-form more easily from its ions.
The other alkyl groups have "electron-pushing" effects very similar to the methyl group, and so the strengths of propanoic acid and
butanoic acid are very similar to ethanoic acid. The acids can be strengthened by pulling charge away from the -COO- end. You can do
this by attaching electronegative atoms like chlorine to the chain.
As the next table shows, the more chlorines you can attach the better:
The chlorine is effective at withdrawing charge when it is next-door to the -COO- group, and much less so as it gets even one carbon
further away.
This page explains why simple organic bases are basic and looks at the factors which affect their relative strengths. For A'level
purposes, all the bases we are concerned with are primary amines - compounds in which one of the hydrogens in an ammonia
molecule, NH3, is replaced either by an alkyl group or a benzene ring.
An ammonium ion is formed together with hydroxide ions. Because the ammonia is only a weak base, it doesn't hang on to the extra
hydrogen ion very effectively and so the reaction is reversible. At any one time, about 99% of the ammonia is present as unreacted
molecules. The position of equilibrium lies well to the left.
The ammonia reacts as a base because of the active lone pair on the nitrogen. Nitrogen is more electronegative than hydrogen and so
attracts the bonding electrons in the ammonia molecule towards itself. That means that in addition to the lone pair, there is a build-up
of negative charge around the nitrogen atom. That combination of extra negativity and active lone pair attracts the new hydrogen from
the water.
The only difference between this and ammonia is the presence of the CH3 group in the methylamine. But that's important! Alkyl
groups have a tendency to "push" electrons away from themselves. That means that there will be a small amount of extra negative
charge built up on the nitrogen atom. That extra negativity around the nitrogen makes the lone pair even more attractive towards
hydrogen ions.
Making the nitrogen more negative helps the lone pair to pick up a hydrogen ion. What about the effect on the positive
methylammonium ion formed? Is this more stable than a simple ammonium ion? Compare the methylammonium ion with an
ammonium ion:
In the methylammonium ion, the positive charge is spread around the ion by the "electron-pushing" effect of the methyl group. The
more you can spread charge around, the more stable an ion becomes. In the ammonium ion there is not any way of spreading the
charge.
WHY ARE AROMATIC PRIMARY AMINES MUCH WEAKER BASES THAN AMMONIA?
An aromatic primary amine is one in which the -NH2 group is attached directly to a benzene ring. The only one you are likely to come
across is phenylamine. Phenylamine has the structure:
The lone pair on the nitrogen touches the delocalized ring electrons . . .
That means that the lone pair is no longer fully available to combine with hydrogen ions. The nitrogen is still the most electronegative
atom in the molecule, and so the delocalized electrons will be attracted towards it, but the intensity of charge around the nitrogen is
nothing like what it is in, say, an ammonia molecule.
The other problem is that if the lone pair is used to join to a hydrogen ion, it is no longer available to contribute to the delocalisation.
That means that the delocalization would have to be disrupted if the phenylamine acts as a base. Delocalization makes molecules more
stable, and so disrupting the delocalization costs energy and will not happen easily.
Taken together - the lack of intense charge around the nitrogen, and the need to break some delocalization - this means that
phenylamine is a very weak base indeed.
EXERCISES
QUESTIONS
Q2.10.1
Determine which of the one of the molecules is an acid or a base.
CONTRIBUTORS
Jim Clark (Chemguide.co.uk)
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
KEY TERMS
Make certain that you can define, and use in context, the key terms below.
Lewis acid
Lewis base
STUDY NOTES
The Lewis concept of acidity and basicity will be of great use to you when you study reaction mechanisms. The realization that an
ion such as
is electron deficient, and is therefore a Lewis acid, should help you understand why this ion reacts with substances which are Lewis
bases (e.g., H2O).
The Brønsted-Lowry picture of acids and bases as proton donors and acceptors is not the only definition in common use. A broader
definition is provided by the Lewis theory of acids and bases, in which a Lewis acid is an electron-pair acceptor and a Lewis base is an
electron-pair donor. This definition covers Brønsted-Lowry proton transfer reactions, but also includes reactions in which no proton
transfer is involved. The interaction between a magnesium cation (Mg+2) and a carbonyl oxygen is a common example of a Lewis
acid-base reaction. The carbonyl oxygen (the Lewis base) donates a pair of electrons to the magnesium cation (the Lewis acid).
As we will see in Chapter 19 when we begin the study of reactions involving carbonyl groups, this interaction has the very important
effect of increasing the polarity of the carbon-oxygen double bond. The Brønsted-Lowry equivalent of the reaction above is simply
protonation of the carbonyl group. This, too, has the effect of increasing the polarity of the carbonyl double bond.
While it is important to be familiar with the Lewis definition, the focus throughout the remainder of this chapter will be on acid-base
reactions of the Brønsted-Lowry type, where an actual proton transfer event takes place.
LEWIS ACIDS
Borane is unusual because it is a compound without an octet. The central boron atom has only six valence electrons. It needs one more
pair of electrons to obtain an octet. The boron is a Lewis acid.
Figure 2.11.2: Boron, aluminum and indium are from the same column of the periodic table. All three are often Lewis acidic; they can
accept electrons from donors.
Boron, aluminum and indium compounds are often Lewis acids.
The eight-electron rule does not hold throughout the periodic table. In order to obtain noble gas configurations, some atoms may need
eighteen electrons in their valence shell. For example, transition metals such as titanium often follow an eighteen-electron rule.
Titanium has four valence electrons and can form four bonds in compounds such as titanium tetrakis (isopropoxide), below, or titanium
tetrachloride, TiCl4. However, the titanium atom in that compound has only eight valence electrons, not eighteen. It can easily accept
electrons from donors.
Figure 2.11.3: Although titanium has eight electrons in this molecule, titanium tetrakis(isopropoxide), it can accommodate up to
eighteen. It is Lewis acidic. The cerium atom in cerium tris(dimethylamide) comes from a similar part of the periodic table and is also
Lewis acidic.
Transition metals such as titanium, iron and nickel may have up to eighteen electrons and can frequently accept electron pairs from
Lewis bases. Transition metals are often Lewis acids.
Lanthanides such as cerium and samarium could conceivably have up to thirty-two electrons in their valence shells! They never do.
However, they are usually strong Lewis acids.
Positive ions are often Lewis acids because they have an electrostatic attraction for electron donors. Examples include alkali and
alkaline earth metals in the group IA and IIA columns. K+, Mg2+ and Ca2+ are sometimes seen as Lewis acidic sites in biology, for
example. These ions are very stable forms of these elements because of their low electron ionization potentials. However, their positive
charges do attract electron donors.
Figure 2.11.5: A few alkali, alkaline earth and transition metals that are commonly found as cations.
Many cations such as Ca2+ or Sc3+ are good Lewis acids. Their positive charges attract electrons.
LEWIS BASES
What makes a molecule (or an atom or ion) a Lewis base? It must have a pair of electrons available to share with another atom to form
a bond. The most readily available electrons are those that are not already in bonds. Bonding electrons are low in energy. Non-bonding
electrons are higher in energy and may be stabilized when they are delocalized in a new bond.
Lewis bases usually have non-bonding electrons or lone pairs.
Lewis bases may be anionic or neutral. The basic requirement is that they have a pair of electrons to donate. Examples of Lewis bases
include halide ions such as bromide or chloride. To the right of the halides in the periodic table are Noble gases such as neon. Noble
gases do have lone pairs, but are stable enough that they do not usually react. They are not very good Lewis bases. To the left of the
halides, however, are other examples in oxygen and nitrogen compounds. Water also has lone pairs and is a common Lewis base, and
so is hydroxide ion, HO-.
Figure 2.11.2: Some examples of Lewis basic ions and molecules. Note that neon, although it has nonbonding electron pairs or lone
pairs, does not usually act as a Lewis base.
Halides, water, ammonia and hydroxide ion are examples of Lewis bases.
One column further to the left in the periodic table from nitrogen is carbon. Carbon does not normally have a lone pair. For example,
methane, CH4, has all of its valence electrons in bonding pairs. These bonding pairs are too stable to donate under normal conditions.
Methane is not a Lewis base.
Figure 2.11.3. Carbon and boron "hydrides". Neither of these compounds has a lone pair, and neither is a good Lewis base.
Even further to the left is boron. A simple boron compound is borane, BH3. Borane has no lone pairs; all its valence electrons are in
bonds. Boron is not a good Lewis base.
PROBLEM
Which of the following compounds appear to be Lewis bases?
a) SiH4 b) AlH3 c) PH3 d) SH2 e) -SH
LEWIS ACID-BASE COMPLEXES
What happens when a Lewis base donates a pair of electrons to a Lewis acid? The arrow formulism we have been using to illustrate the
behaviour of Lewis acids and Lewis bases is meant to show the direction of electron movement from the donor to the acceptor.
However, given that a bond can be thought of as a pair of electrons that are shared between two atoms (in this case, between the donor
and the acceptor), these arrows also show where bonds are forming.
Figure AB4.2: Formation of a Lewis acid-base complex from ammonia and boron trifluoride.
When the nitrogen donates a pair of electrons to share with the boron, the bond that forms is sometimes called a coordinate bond.
Another term for this kind of bond is a dative bond. A coordinate or dative bond is any covalent bond that arose because one atom
brought a pair of its electrons and donated them with another.
There is another piece of terminology you should get used to here. Sometimes, the electron donor is called a nucleophile and the
electron acceptor is called an electrophile. Ammonia is a nucleophile and boron trifluoride is an electrophile.
Because Lewis bases are attracted to electron-deficient atoms, and because positive charge is generally associated with the nucleus
of an atom, Lewis bases are sometimes refered to as "nucleophiles". Nucleophile means nucleus-loving.
Because Lewis acids attract electron pairs, Lewis acids are sometimes called "electrophiles". Electrophile meanse electron-loving.
Lewis acid-base complexes frequently have very different properties from the separate compounds from which they were formed. For
example, titanium tetrachloride is a yellow liquid at room temperature. It is so Lewis acidic that it reacts with moisture in the air,
undergoing a reaction that generates HCl gas in the form of white smoke. Tetrahydrofuran (or THF), a mild Lewis base, is a colourless
liquid. When THF and TiCl4 are combined, a Lewis acid-base complex is formed, TiCl4(THF)2. TiCl4(THF)2 is a yellow solid at room
temperature. Although it still reacts with the air, it does so very slowly, and shows no visible change when exposed to the air for
several minutes.
Figure AB4.5: A Lewis acid-base complex between tetrahydrofuran (THF) and titanium tetrachloride.
PROBLEM 1
Show, using arrow notation, the reaction between THF and titanium tetrachloride to form the Lewis acid-base complex, TiCl4(THF)2.
Also show the structures of the complexes formed.
PROBLEM 2
A similar Lewis acid-base complex is formed between THF and borane, BH3.
1. Which compound is the Lewis acid? Which one is the Lewis base?
2. Which atom in the Lewis acid is the acidic site? Why?
3. Which atom in the Lewis base is the basic site? Why?
4. How many donors would be needed to satisfy the acidic site?
5. Show, using arrow notation, the reaction to form a Lewis acid-base complex.
6. Borane is highly pyrophoric; it reacts violently with air, bursting into flames. Show, using arrow notation, what might be happening
when borane contacts the air.
7. Borane-THF complex is much less pyrophoric than borane. Why do you suppose that is so?
PROBLEM 3
When a neutral Lewis acid combines with an anionic Lewis base, the product is called a complex ion. The same is true if a cationic
Lewis acid combines with a neutral Lewis base.
Show the formation of the following polyatomic anions from the Lewis acid-base pairs that were combined in each case.
a) BF4- b) PF6- c) AlCl4- d) AlH4- e) Ag(NH3)2+
EXERCISES
QUESTIONS
SOLUTIONS
S2.11.1
Acetone is a Lewis base and a Brønsted base. Ammonium cation is both a Lewis acid and a weak Brønsted acid.
CONTRIBUTORS
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
Prof. Steven Farmer (Sonoma State University)
Chris P Schaller, Ph.D., (College of Saint Benedict / Saint John's University)
Organic Chemistry With a Biological Emphasis by Tim Soderberg (University of Minnesota, Morris)
STUDY NOTES
You will have noticed that we have given two names for most of the compounds discussed up to this point. In general we shall be
using systematic (i.e., IUPAC—International Union of Pure and Applied Chemistry) names throughout the course. However, simple
compounds are often known principally by their common names, which may be more familiar to you than their IUPAC
counterparts. We shall address the subject of nomenclature (naming) in Chapter 3.
ethanol,
formaldehyde (methanal),
acetone (propanone),
EXERCISES
1. Construct a molecular model of each of the compounds listed below.
a. CH 3
−CN
b. CH 3
−N=C=O
c. CH 3
−CH −O−CH
2 3
Hint: Use the curved sticks to form the multiple bonds and the straight sticks for single bonds.
ANSWERS:
A.
B.
C.
CONTRIBUTORS
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
CONTRIBUTORS
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
KEY TERMS
Make certain that you can define, and use in context, the key term below.
functional group
STUDY NOTES
The concept of functional groups is a very important one. We expect that you will need to refer back to tables at the end of Section
3.1 quite frequently at first, as it is not really feasible to learn the names and structures of all the functional groups and compound
types at one sitting. Gradually they will become familiar, and eventually you will recognize them automatically.
Functional groups are atoms or small groups of atoms (two to four) that exhibit a characteristic reactivity. A particular functional
group will almost always display its characteristic chemical behavior when it is present in a compound. Because of their importance in
understanding organic chemistry, functional groups have characteristic names that often carry over in the naming of individual
compounds incorporating specific groups
As we progress in our study of organic chemistry, it will become extremely important to be able to quickly recognize the most common
functional groups, because they are the key structural elements that define how organic molecules react. For now, we will only worry
about drawing and recognizing each functional group, as depicted by Lewis and line structures. Much of the remainder of your study of
organic chemistry will be taken up with learning about how the different functional groups tend to behave in organic reactions.
The carbon-carbon triple bond in ethyne is the simplest example of an alkyne function group.
What about ethane? All we see in this molecule is carbon-hydrogen and carbon-carbon single bonds, so in a sense we can think of
ethane as lacking a functional group entirely. However, we do have a general name for this ‘default’ carbon bonding pattern: molecules
or parts of molecules containing only carbon-hydrogen and carbon-carbon single bonds are referred to as alkanes. If the carbon of an
alkane is bonded to a halogen, the group is now referred to as a haloalkane (fluoroalkane, chloroalkane, etc.). Chloroform, CHCl3, is
an example of a simple haloalkane.
ALCOHOLS AND THIOLS
We have already seen the simplest possible example of an alcohol functional group in methanol. In the alcohol functional group, a
carbon is single-bonded to an OH group (this OH group, by itself, is referred to as a hydroxyl). If the central carbon in an alcohol is
bonded to only one other carbon, we call the group a primary alcohol. In secondary alcohols and tertiary alcohols, the central carbon is
bonded to two and three carbons, respectively. Methanol, of course, is in class by itself in this respect.
In an ether functional group, a central oxygen is bonded to two carbons. Below are the line and Lewis structures of diethyl ether, a
common laboratory solvent and also one of the first medical anaesthesia agents.
In sulfides, the oxygen atom of an ether has been replaced by a sulfur atom.
One of the most important properties of amines is that they are basic, and are readily protonated to form ammonium cations.
Phosphorus is a very important element in biological organic chemistry, and is found as the central atom in the phosphate group. Many
biological organic molecules contain phosphate, diphosphate, and triphosphate groups, which are linked to a carbon atom by the
phosphate ester functionality.
Because phosphates are so abundant in biological organic chemistry, it is convenient to depict them with the abbreviation 'P'. Notice
that this 'P' abbreviation includes the oxygen atoms and negative charges associated with the phosphate groups.
Molecules with carbon-nitrogen double bonds are called imines, or Schiff bases.
As the name implies, carboxylic acids are acidic, meaning that they are readily deprotonated to form the conjugate base form, called a
carboxylate (much more about carboxylic acids in the acid-base chapter!).
In amides, the carbonyl carbon is bonded to a nitrogen. The nitrogen in an amide can be bonded either to hydrogens, to carbons, or to
both. Another way of thinking of an amide is that it is a carbonyl bonded to an amine.
In esters, the carbonyl carbon is bonded to an oxygen which is itself bonded to another carbon. Another way of thinking of an ester is
that it is a carbonyl bonded to an alcohol. Thioesters are similar to esters, except a sulfur is in place of the oxygen.
Finally, in a nitrile group, a carbon is triple-bonded to a nitrogen. Nitriles are also often referred to as cyano groups.
A single compound often contains several functional groups. The six-carbon sugar molecules glucose and fructose, for example,
contain aldehyde and ketone groups, respectively, and both contain five alcohol groups (a compound with several alcohol groups is
often referred to as a ‘polyol’).
Capsaicin, the compound responsible for the heat in hot peppers, contains phenol, ether, amide, and alkene functional groups.
The male sex hormone testosterone contains ketone, alkene, and secondary alcohol groups, while acetylsalicylic acid (aspirin) contains
aromatic, carboxylic acid, and ester groups.
While not in any way a complete list, this section has covered most of the important functional groups that we will encounter in
biological and laboratory organic chemistry. The table on the inside back cover provides a summary of all of the groups listed in this
section, plus a few more that will be introduced later in the text.
PROBLEMS
1: Identify the functional groups in the following organic compounds. State whether alcohols and amines are primary, secondary, or
tertiary.
EXERCISES
QUESTIONS
Q3.1.1
The following is the molecule for ATP, or the molecule responsible for energy in human cells. Identify the functional groups for ATP.
SOLUTIONS
S3.1.1
KEY TERMS
Make certain that you can define, and use in context, the key terms below.
branched-chain alkane
constitutional or structural isomer
homologous series
isomer
saturated hydrocarbon
straight-chain alkane (or normal alkane)
STUDY NOTES
A series of compounds in which successive members differ from one another by a CH2 unit is called a homologous series. Thus, the
series CH4, C2H6, C3H8 . . . CnH2n+2, is an example of a homologous series.
It is important that you commit to memory the names of the first 10 straight-chain alkanes (i.e., from CH4 to C10H22). You will use
these names repeatedly when you begin to learn how to derive the systematic names of a large variety of organic compounds. You
need not remember the number of isomers possible for alkanes containing more than seven carbon atoms. Such information is
available in reference books when it is needed. When drawing isomers, be careful not to deceive yourself into thinking that you can
draw more isomers than you are supposed to be able to. Remember that it is possible to draw each isomer in several different ways
and you may inadvertently count the same isomer more than once.
Alkanes are organic compounds that consist entirely of single-bonded carbon and hydrogen atoms and lack any other functional
groups. Alkanes have the general formula C H n and can be subdivided into the following three groups: the linear straight-chain
2n+2
alkanes, branched alkanes, and cycloalkanes. Alkanes are also saturated hydrocarbons.
Cycloalkanes are cyclic hydrocarbons, meaning that the carbons of the molecule are arranged in the form of a ring. Cycloalkanes are
also saturated, meaning that all of the carbons atoms that make up the ring are single bonded to other atoms (no double or triple bonds).
There are also polycyclic alkanes, which are molecules that contain two or more cycloalkanes that are joined, forming multiple rings.
This is an introductory page about alkanes, such as methane, ethane, propane, butane and the remainder of the common alkanes. This
page addresses their formulae and isomerism, their physical properties, and an introduction to their chemical reactivity.
MOLECULAR FORMULAS
Alkanes are the simplest family of hydrocarbons - compounds containing carbon and hydrogen only. Alkanes only contain carbon-
hydrogen bonds and carbon-carbon single bonds. The first six alkanes are as follows:
methane CH4
ethane C2H6
propane C3H8
butane C4H10
pentane C5H12
hexane C6H14
You can work out the formula of any of the alkanes using the general formula CnH2n+2
ISOMERISM
There are also endless other possible ways that this molecule could twist itself. There is completely free rotation around all the carbon-
carbon single bonds. If you had a model of a molecule in front of you, you would have to take it to pieces and rebuild it if you wanted
to make an isomer of that molecule. If you can make an apparently different molecule just by rotating single bonds, it's not different -
it's still the same molecule.
In structural isomerism, the atoms are arranged in a completely different order. This is easier to see with specific examples. What
follows looks at some of the ways that structural isomers can arise. The names of the various forms of structural isomerism probably
do not matter all that much, but you must be aware of the different possibilities when you come to draw isomers.
CHAIN ISOMERISM
These isomers arise because of the possibility of branching in carbon chains. For example, there are two isomers of butane, C4H10. In
one of them, the carbon atoms lie in a "straight chain" whereas in the other the chain is branched.
Be careful not to draw "false" isomers which are just twisted versions of the original molecule. For example, this structure is just the
straight chain version of butane rotated about the central carbon-carbon bond.
You could easily see this with a model. This is the example we've already used at the top of this page.
CONTRIBUTORS
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
Prof. Steven Farmer (Sonoma State University)
Organic Chemistry With a Biological Emphasis by Tim Soderberg (University of Minnesota, Morris)
KEY TERMS
Make certain that you can define, and use in context, the key terms below.
alkyl group
methyl group
primary carbon
quaternary carbon
secondary carbon
tertiary carbon
STUDY NOTES
The differences among primary, secondary, tertiary and quaternary carbon atoms are explained in the following discussion. A
convenient way of memorizing this classification scheme is to remember that a primary carbon atom is attached directly to only one
other carbon atom, a secondary carbon atom is attached directly to two carbon atoms, and so on.
The IUPAC system requires first that we have names for simple unbranched chains, as noted above, and second that we have names
for simple alkyl groups that may be attached to the chains. Examples of some common alkyl groups are given in the following table.
Note that the "ane" suffix is replaced by "yl" in naming groups. The symbol R is used to designate a generic (unspecified) alkyl group.
Table 3.4.1: Alkyl Group names
Group CH3– C2H5– CH3CH2CH2– (CH3)2CH– CH3CH2CH2CH2– (CH3)2CHCH2– CH3CH2CH(CH3)– (CH3)3C– R–
Name Methyl Ethyl Propyl Isopropyl Butyl Isobutyl sec-Butyl tert-Butyl Alkyl
ALKYL GROUPS
Alkanes can be described by the general formula CnH2n+2. An alkyl group is formed by removing one hydrogen from the alkane chain
and is described by the formula CnH2n+1. The removal of this hydrogen results in a stem change from -ane to -yl. Take a look at the
following examples.
The same concept can be applied to any of the straight chain alkane names provided in Table 3.4.2.
EXAMPLE 3.4.1
You will find that hydrogen atoms are also classified in this manner. A hydrogen atom attached to a primary carbon atom is called a
primary hydrogen; thus, isobutane, has nine primary hydrogens and one tertiary hydrogen.
EXAMPLE 3.4.2
EXERCISES
QUESTIONS
Q3.3.1
Consider the following molecule. How many carbons are in the longest chain? Find a primary and quaternary carbon, and label an
ethyl group.
SOLUTIONS
S3.3.1
A = 4° Carbon
B = Ethyl Group
C = 1° Carbon
The longest chain is 10 carbons long
KEY TERMS
Make certain that you can define, and use in context, the key term below.
IUPAC system
STUDY NOTES
The IUPAC system of nomenclature aims to ensure
1. that every organic compound has a unique, unambiguous name.
2. that the IUPAC name of any compound conveys the structure of that compound to a person familiar with the system.
One way of checking whether the name you have given to an alkane is reasonable is to count the number of carbon atoms implied by the chosen name. For example, if you named a compound 3‑ethyl-
4‑methylheptane, you have indicated that the compound contains a total of 10 carbon atoms—seven carbon atoms in the main chain, two carbon atoms in an ethyl group, and one carbon atom in a
methyl group. If you were to check the given structure and find 11 carbon atoms, you would know that you had made a mistake. Perhaps the name you should have written was 3‑ethyl-
4,4‑dimethylheptane!
When naming alkanes, a common error of beginning students is a failure to pick out the longest carbon chain. For example, the correct name for the compound shown below is 3‑methylheptane, not
2‑ethylhexane.
Remember that every substituent must have a number, and do not forget the prefixes: di, tri, tetra, etc.
You must use commas to separate numbers, and hyphens to separate numbers and substituents. Notice that 3‑methylhexane is one word.
Hydrocarbons having no double or triple bond functional groups are classified as alkanes or cycloalkanes, depending on whether the carbon atoms of the molecule are arranged only in chains or also in
rings. Although these hydrocarbons have no functional groups, they constitute the framework on which functional groups are located in other classes of compounds, and provide an ideal starting point for
studying and naming organic compounds. The alkanes and cycloalkanes are also members of a larger class of compounds referred to as aliphatic. Simply put, aliphatic compounds are compounds that do
not incorporate any aromatic rings in their molecular structure.
The following table lists the IUPAC names assigned to simple continuous-chain alkanes from C-1 to C-10. A common "ane" suffix identifies these compounds as alkanes. Longer chain alkanes are well
known, and their names may be found in many reference and text books. The names methane through decane should be memorized, since they constitute the root of many IUPAC names. Fortunately,
common numerical prefixes are used in naming chains of five or more carbon atoms.
Table 3.5.1: Simple Unbranched Alkanes
Na Molecular Structural Isomers Name Molecular Structural Iso
me Formula Formula Formula Formula me
rs
me CH4 CH4 1 hexane C6H14 CH3(CH2)4CH3 5
tha
ne
eth C2H6 CH3CH3 1 heptane C7H16 CH3(CH2)5CH3 9
an
e
pro C3H8 CH3CH2CH3 1 octane C8H18 CH3(CH2)6CH3 18
pa
ne
but C4H10 CH3CH2CH2CH3 2 nonane C9H20 CH3(CH2)7CH3 35
an
e
pe C5H12 CH3(CH2)3CH3 3 decane C10H22 CH3(CH2)8CH3 75
nta
ne
Beginning with butane (C4H10), and becoming more numerous with larger alkanes, we note the existence of alkane isomers. For example, there are five C6H14 isomers, shown below as abbreviated line
formulas (A through E):
Although these distinct compounds all have the same molecular formula, only one (A) can be called hexane. How then are we to name the others?
The IUPAC system requires first that we have names for simple unbranched chains, as noted above, and second that we have names for simple alkyl groups that may be attached to the chains. Examples of
some common alkyl groups are given in the following table. Note that the "ane" suffix is replaced by "yl" in naming groups. The symbol R is used to designate a generic (unspecified) alkyl group.
Table 3.5.2: Alkyl Groups Names
Group CH3– C2H5– CH3CH2CH2– (CH3)2CH– CH3CH2CH
(CH
2CH
3)22CHCH
– 2– CH3CH2CH(CH3)– (CH3)3C–
Name Methyl Ethyl Propyl Isopropyl B Isobutyl sec-Butyl tert-Butyl
u
t
y
l
ALKYL GROUPS
Alkanes can be described by the general formula CnH2n+2. An alkyl group is formed by removing one hydrogen from the alkane chain and is described by the formula CnH2n+1. The removal of this
hydrogen results in a stem change from -ane to -yl. Take a look at the following examples.
The same concept can be applied to any of the straight chain alkane names provided in the table above.
Name Molecular Formula Condensed Structural Formula
Methane CH4 CH4
EXAMPLE 3.5.2
What is the name of the follow molecule?
SOLUTION
Rule #1: Choose the longest, most substituted carbon chain containing a functional group. This example does not contain any functional groups, so we only need to be concerned with choosing the
longest, most substituted carbon chain. The longest carbon chain has been highlighted in red and consists of eight carbons.
Rule #2: Carbons bonded to a functional group must have the lowest possible carbon number. If there are no functional groups, then any substitute present must have the lowest possible number.
Because this example does not contain any functional groups, we only need to be concerned with the two substitutes present, that is, the two methyl groups. If we begin numbering the chain from the
left, the methyls would be assigned the numbers 4 and 7, respectively. If we begin numbering the chain from the right, the methyls would be assigned the numbers 2 and 5. Therefore, to satisfy the
second rule, numbering begins on the right side of the carbon chain as shown below. This gives the methyl groups the lowest possible numbering.
EXAMPLE 3.5.3
What is the name of the follow molecule?
SOLUTION
Rule #1: Choose the longest, most substituted carbon chain containing a functional group. This example contains two functional groups, bromine and chlorine. The longest carbon chain has been
highlighted in red and consists of seven carbons.
Rule #2: Carbons bonded to a functional group must have the lowest possible carbon number. If there are no functional groups, then any substitute present must have the lowest possible number. In this
example, numbering the chain from the left or the right would satisfy this rule. If we number the chain from the left, bromine and chlorine would be assigned the second and sixth carbon positions,
respectively. If we number the chain from the right, chlorine would be assigned the second position and bromine would be assigned the sixth position. In other words, whether we choose to number from
the left or right, the functional groups occupy the second and sixth positions in the chain. To select the correct numbering scheme, we need to utilize the third rule.
Rule #3: After applying the first two rules, take the alphabetical order into consideration. Alphabetically, bromine comes before chlorine. Therefore, bromine is assigned the second carbon position, and
chlorine is assigned the sixth carbon position.
EXAMPLE 3.5.4
What is the name of the follow molecule?
SOLUTION
Rule #1: Choose the longest, most substituted carbon chain containing a functional group. This example contains two functional groups, bromine and chlorine, and one substitute, the methyl group. The
longest carbon chain has been highlighted in red and consists of seven carbons.
Rule #2: Carbons bonded to a functional group must have the lowest possible carbon number. After taking functional groups into consideration, any substitutes present must have the lowest possible
carbon number. This particular example illustrates the point of difference principle. If we number the chain from the left, bromine, the methyl group and chlorine would occupy the second, fifth and
sixth positions, respectively. This concept is illustrated in the second drawing below. If we number the chain from the right, chlorine, the methyl group and bromine would occupy the second, third and
sixth positions, respectively, which is illustrated in the first drawing below. The position of the methyl, therefore, becomes a point of difference. In the first drawing, the methyl occupies the third
position. In the second drawing, the methyl occupies the fifth position. To satisfy the second rule, we want to choose the numbering scheme that provides the lowest possible numbering of this substitute.
Therefore, the first of the two carbon chains shown below is correct.
Once you have determined the correct numbering of the carbons, it is often useful to make a list, including the functional groups, substitutes, and the name of the parent chain.
Rule #3: After applying the first two rules, take the alphabetical order into consideration. Alphabetically, bromine comes before chlorine. Therefore, bromine is assigned the second carbon position, and
chlorine is assigned the sixth carbon position.
Parent chain: heptane 2-Chloro 3-Methyl 6-Bromo
EXERCISES
3.4 EXERCISES
QUESTIONS
Q3.4.1
Are the following structures properly named, and if they are not, what is the correct naming?
Q3.4.2
Give the name of the following molecules:
SOLUTIONS
S3.4.1
They are both labeled incorrectly:
3-bromo-2-hydroxypentane
2, 3-dimethylpentane
S3.4.2
1 = 3,4-Dimethyl hexane
2 = 2-methyl pentane
3 = 2,2,4-trimethyl pentane
CONTRIBUTORS
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
Prof. Steven Farmer (Sonoma State University)
Organic Chemistry With a Biological Emphasis by Tim Soderberg (University of Minnesota, Morris)
KEY TERMS
Make certain that you can define, and use in context, the key term below.
van der Waals force
Alkanes are not very reactive and have little biological activity; all alkanes are colorless and odorless.
BOILING POINTS
The boiling points shown are for the "straight chain" isomers of which there is more than one. The first four alkanes are gases at room
temperature, and solids do not begin to appear until about C H , but this is imprecise because different isomers typically have
17 36
different melting and boiling points. By the time you get 17 carbons into an alkane, there are unbelievable numbers of isomers!
Cycloalkanes have boiling points that are approximately 20 K higher than the corresponding straight chain alkane.
There is not a significant electronegativity difference between carbon and hydrogen, thus, there is not any significant bond polarity.
The molecules themselves also have very little polarity. A totally symmetrical molecule like methane is completely non-polar, meaning
that the only attractions between one molecule and its neighbors will be Van der Waals dispersion forces. These forces will be very
small for a molecule like methane but will increase as the molecules get bigger. Therefore, the boiling points of the alkanes increase
with molecular size.
Where you have isomers, the more branched the chain, the lower the boiling point tends to be. Van der Waals dispersion forces are
smaller for shorter molecules and only operate over very short distances between one molecule and its neighbors. It is more difficult
for short, fat molecules (with lots of branching) to lie as close together as long, thin molecules.
EXERCISE 3.5.1
For each of the following pairs of compounds, select the substance which you expect to have the higher boiling point:
a. octane and nonane.
b. octane and 2,2,3,3-tetramethylbutane.
NOTE
This is a simplification because entropic effects are important when things dissolve.
EXERCISES
1. For each of the following pairs of compounds, select the substance you expect to have the higher boiling point.
a. octane and nonane.
b. octane and 2,2,3,3‑tetramethylbutane.
ANSWERS:
a. Nonane will have a higher boiling point than octane, because it has a longer carbon chain than octane.
b. Octane will have a higher boiling point than 2,2,3,3‑tetramethylbutane, because it branches less than 2,2,3,3‑tetramethylbutane, and
therefore has a larger “surface area” and more van der Waals forces.
Note: The actual boiling points are
nonane, 150.8°C
octane, 125.7°C
2,2,3,3‑tetramethylbutane, 106.5°C
CONTRIBUTORS
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
Prof. Steven Farmer (Sonoma State University)
Jim Clark (Chemguide.co.uk)
KEY TERMS
Make certain that you can define, and use in context, the key terms below.
conformation (conformer, conformational isomer)
eclipsed conformation
Newman projection
sawhorse representation
staggered conformation
strain energy
torsional strain (eclipsing strain)
STUDY NOTES
You should be prepared to sketch various conformers using both sawhorse representations and Newman projections. Each method
has its own advantages, depending upon the circumstances. Notice that when drawing the Newman projection of the eclipsed
conformation of ethane, you cannot draw the rear hydrogens exactly behind the front ones. This is an inherent limitation associated
with representing a 3-D structure in two dimensions.
Conformational isomerism involves rotation about sigma bonds, and does not involve any differences in the connectivity or geometry
of bonding. Two or more structures that are categorized as conformational isomers, or conformers, are really just two of the exact
same molecule that differ only in terms of the angle about one or more sigma bonds.
ETHANE CONFORMATIONS
Although there are seven sigma bonds in the ethane molecule, rotation about the six carbon-hydrogen bonds does not result in any
change in the shape of the molecule because the hydrogen atoms are essentially spherical. Rotation about the carbon-carbon bond,
however, results in many different possible molecular conformations.
In order to better visualize these different conformations, it is convenient to use a drawing convention called the Newman projection.
In a Newman projection, we look lengthwise down a specific bond of interest – in this case, the carbon-carbon bond in ethane. We
depict the ‘front’ atom as a dot, and the ‘back’ atom as a larger circle.
The six carbon-hydrogen bonds are shown as solid lines protruding from the two carbons at 120°angles, which is what the actual
tetrahedral geometry looks like when viewed from this perspective and flattened into two dimensions.
The lowest energy conformation of ethane, shown in the figure above, is called the ‘staggered’ conformation, in which all of the C-H
bonds on the front carbon are positioned at dihedral angles of 60°relative to the C-H bonds on the back carbon. In this conformation,
the distance between the bonds (and the electrons in them) is maximized.
This is the highest energy conformation because of unfavorable interactions between the electrons in the front and back C-H bonds.
The energy of the eclipsed conformation is approximately 3 kcal/mol higher than that of the staggered conformation. Another
60°rotation returns the molecule to a second eclipsed conformation. This process can be continued all around the 360°circle, with three
possible eclipsed conformations and three staggered conformations, in addition to an infinite number of variations in between.
Figure 3.6.X: The potential energy associated with the various conformations of ethane varies with the dihedral angle of the bonds.
Newman projections of butane conformations & their relative energy differences (not total energies). Conformations form when butane
rotates about one of its single covalent bond. Torsional/dihedral angle is shown on x-axis. Conformations (according to IUPAC): A:
antiperiplanar, anti or trans B: synclinal or gauche C: anticlinal or eclipsed D: synperiplanar or cis Valleys of the pink graph are
conformations lowest in energy (shown as A & B). Peaks are conformations highest in energy (shown as C & D). Energies are
A<B<C<D with D highest & A lowest in energy. A is thus the most stable conformation &, of all the other conformations, occurs most
often in room temperature. Valleys A & B are local energy minima & A is global minima. A & B can thus be classified as rotamers (a
class of conformers). Peaks are not rotamers, & are caused by repulsive forces of the hydrogens & methyls (-CH3). Source for
conformation names & conformer classification: Pure & Appl. Chem., Vol. 68, No. 12, pp. 2193-2222, 1996 Image used with
permission (Public Domain; Keministi).
Although the conformers of ethane are in rapid equilibrium with each other, the 3 kcal/mol energy difference leads to a substantial
preponderance of staggered conformers (> 99.9%) at any given time. The animation below illustrates the relationship between ethane's
potential energy and its dihedral angle
EXERCISES
QUESTIONS
Q3.6.1
What is the most stable rotational conformation of ethane and explain why it is preferred over the other conformation?
SOLUTIONS
S3.6.1
Staggered, as there is less repulsion between the hydrogen atoms.
CONTRIBUTORS
William Reusch, Professor Emeritus (Michigan State U.), Virtual Textbook of Organic Chemistry
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
KEY TERMS
Make certain that you can define, and use in context, the key terms below.
anti conformation
gauche conformation
eclipsed conformation
steric repulsion
In butane, there are now three rotating carbon-carbon bonds to consider, but we will focus on the middle bond between C2 and C3.
Below are two representations of butane in a conformation which puts the two CH3 groups (C1 and C4) in the eclipsed position.
This is the highest energy conformation for butane, due to what is called ‘van der Waals repulsion’, or ‘steric repulsion’, between
the two rather bulky methyl groups.
What is van der Waals repulsion? Didn’t we just learn in Chapter 2 that the van der Waals force between two nonpolar groups is an
attractive force? Consider this: you probably like to be near your friends, but no matter how close you are you probably don’t want to
share a one-room apartment with five of them. When the two methyl groups are brought too close together, the overall resulting
noncovalent interaction is repulsive rather than attractive. The result is that their respective electron densities repel one another.
If we rotate the front, (blue) carbon by 60°clockwise, the butane molecule is now in a staggered conformation.
This is more specifically referred to as the ‘gauche’ conformation of butane. Notice that although they are staggered, the two methyl
groups are not as far apart as they could possibly be. There is still significant steric repulsion between the two bulky groups. A further
rotation of 60°gives us a second eclipsed conformation (B) in which both methyl groups are lined up with hydrogen atoms.
This is the lowest energy conformation for butane. The diagram below summarizes the relative energies for the various eclipsed,
staggered, and gauche conformations.
At room temperature, butane is most likely to be in the lowest-energy anti conformation at any given moment in time, although the
energy barrier between the anti and eclipsed conformations is not high enough to prevent constant rotation except at very low
temperatures. For this reason (and also simply for ease of drawing), it is conventional to draw straight-chain alkanes in a zigzag form,
which implies anti conformation at all carbon-carbon bonds.
EXAMPLE 3.8.1
Draw Newman projections of the eclipsed and staggered conformations of propane.
Answer:
Answer:
The following diagram illustrates the change in potential energy that occurs with rotation about the C2–C3 bond.
Figure 3.8.2: Potential curve vs dihedral angle of the C2-C3 bond of butane.
EXERCISES
QUESTIONS
Q3.7.1
Draw the energy diagram for the rotation of the bond highlighted in pentane.
SOLUTIONS
S3.7.1
KEY TERMS
Make certain that you can define, and use in context, the key terms below.
catalytic cracking
catalytic reforming
fractional distillation
octane number (octane rating)
STUDY NOTES
The refining of petroleum into usable fractions is a very important industrial process. In the laboratory component of this course,
you will have the opportunity to compare this industrial process to the procedure as it is performed in the laboratory.
PETROLEUM
The petroleum that is pumped out of the ground at locations around the world is a complex mixture of several thousand organic
compounds, including straight-chain alkanes, cycloalkanes, alkenes, and aromatic hydrocarbons with four to several hundred carbon
atoms. The identities and relative abundances of the components vary depending on the source. So Texas crude oil is somewhat
different from Saudi Arabian crude oil. In fact, the analysis of petroleum from different deposits can produce a “fingerprint” of each,
which is useful in tracking down the sources of spilled crude oil. For example, Texas crude oil is “sweet,” meaning that it contains a
small amount of sulfur-containing molecules, whereas Saudi Arabian crude oil is “sour,” meaning that it contains a relatively large
amount of sulfur-containing molecules.
GASOLINE
Petroleum is converted to useful products such as gasoline in three steps: distillation, cracking, and reforming. Recall from Chapter 1
"Introduction to Chemistry" that distillation separates compounds on the basis of their relative volatility, which is usually inversely
proportional to their boiling points. Part (a) in Figure 3.8.1 shows a cutaway drawing of a column used in the petroleum industry for
separating the components of crude oil. The petroleum is heated to approximately 400°C (750°F), at which temperature it has become
a mixture of liquid and vapor. This mixture, called the feedstock, is introduced into the refining tower. The most volatile components
(those with the lowest boiling points) condense at the top of the column where it is cooler, while the less volatile components condense
nearer the bottom. Some materials are so nonvolatile that they collect at the bottom without evaporating at all. Thus the composition of
the liquid condensing at each level is different. These different fractions, each of which usually consists of a mixture of compounds
with similar numbers of carbon atoms, are drawn off separately. Part (b) in Figure 3.8.1 shows the typical fractions collected at
refineries, the number of carbon atoms they contain, their boiling points, and their ultimate uses. These products range from gases used
in natural and bottled gas to liquids used in fuels and lubricants to gummy solids used as tar on roads and roofs.
OCTANE RATINGS
The quality of a fuel is indicated by its octane rating, which is a measure of its ability to burn in a combustion engine without knocking
or pinging. Knocking and pinging signal premature combustion (Figure 3.8.2), which can be caused either by an engine malfunction or
by a fuel that burns too fast. In either case, the gasoline-air mixture detonates at the wrong point in the engine cycle, which reduces the
power output and can damage valves, pistons, bearings, and other engine components. The various gasoline formulations are designed
to provide the mix of hydrocarbons least likely to cause knocking or pinging in a given type of engine performing at a particular level.
As shown in Figure 3.8.3, many compounds that are now available have octane ratings greater than 100, which means they are better
fuels than pure isooctane. In addition, antiknock agents, also called octane enhancers, have been developed. One of the most widely
used for many years was tetraethyllead [(C2H5)4Pb], which at approximately 3 g/gal gives a 10–15-point increase in octane rating.
Since 1975, however, lead compounds have been phased out as gasoline additives because they are highly toxic. Other enhancers, such
as methyl t-butyl ether (MTBE), have been developed to take their place. They combine a high octane rating with minimal corrosion to
engine and fuel system parts. Unfortunately, when gasoline containing MTBE leaks from underground storage tanks, the result has
been contamination of the groundwater in some locations, resulting in limitations or outright bans on the use of MTBE in certain areas.
As a result, the use of alternative octane enhancers such as ethanol, which can be obtained from renewable resources such as corn,
sugar cane, and, eventually, corn stalks and grasses, is increasing.
Figure 3.8.3: The Octane Ratings of Some Hydrocarbons and Common Additives
CONTRIBUTORS
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
CONTRIBUTORS
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
STUDY NOTES
Provided that you have mastered the IUPAC system for naming alkanes, you should find that the nomenclature of cycloalkanes does not
present any particular difficulties. Concentrate on the examples in which the substituent or substituents is or are an alkyl group, a
halogen, or both.
Cycloalkanes are cyclic hydrocarbons, meaning that the carbons of the molecule are arranged in the form of a ring. Cycloalkanes are also
saturated, meaning that all of the carbons atoms that make up the ring are single bonded to other atoms (no double or triple bonds). There
are also polycyclic alkanes, which are molecules that contain two or more cycloalkanes that are joined, forming multiple rings.
INTRODUCTION
Many organic compounds found in nature or created in a laboratory contain rings of carbon atoms with distinguishing chemical properties;
these compounds are known as cycloalkanes. Cycloalkanes only contain carbon-hydrogen bonds and carbon-carbon single bonds, but in
cycloalkanes, the carbon atoms are joined in a ring. The smallest cycloalkane is cyclopropane.
Figure 4.1.1:
If you count the carbons and hydrogens, you will see that they no longer fit the general formula C H
n . By joining the carbon atoms in
2n+2
a ring,two hydrogen atoms have been lost. The general formula for a cycloalkane is C H . Cyclic compounds are not all flat molecules.
n 2n
All of the cycloalkanes, from cyclopentane upwards, exist as "puckered rings". Cyclohexane, for example, has a ring structure that looks
like this:
Figure 4.1.2: This is known as the "chair" form of cyclohexane from its shape, which vaguely resembles a chair. Note: The cyclohexane
molecule is constantly changing, with the atom on the left, which is currently pointing down, flipping up, and the atom on the right flipping
down. During this process, another (slightly less stable) form of cyclohexane is formed known as the "boat" form. In this arrangement, both
of these atoms are either pointing up or down at the same time
In addition to being saturated cyclic hydrocarbons, cycloalkanes may have multiple substituents or functional groups that further determine
their unique chemical properties. The most common and useful cycloalkanes in organic chemistry are cyclopentane and cyclohexane,
although other cycloalkanes varying in the number of carbons can be synthesized. Understanding cycloalkanes and their properties are
crucial in that many of the biological processes that occur in most living things have cycloalkane-like structures.
Although polycyclic compounds are important, they are highly complex and typically have common names accepted by IUPAC. However,
the common names do not generally follow the basic IUPAC nomenclature rules. The general formula of the cycloalkanes is C H where
n 2n
n is the number of carbons. The naming of cycloalkanes follows a simple set of rules that are built upon the same basic steps in naming
CONTENTS
For simplicity, cycloalkane molecules can be drawn in the form of skeletal structures in which each intersection between two lines is
assumed to have a carbon atom with its corresponding number of hydrogens.
same as same as
Cycloalkane Molecular Formula Basic Structure
Cyclopropane
C3H6
Cyclobutane
C4H8
Cyclopentane
C5H10
Cyclohexane
C6H12
Cycloheptane
C7H14
Cyclooctane
C8H16
Cyclononane
C9H18
Cyclodecane
C10H20
cyclopropane cyclopropyl
cyclobutane cyclobutyl
cyclopentane cyclopentyl
cyclohexane cyclohexyl
cycloheptane cycloheptyl
cyclooctane cyclooctyl
cyclononane cyclononanyl
cyclodecane cyclodecanyl
EXAMPLE 4.1.1
The longest straight chain contains 10 carbons, compared with cyclopropane, which only contains 3 carbons. Because cyclopropane is a
substituent, it would be named a cyclopropyl-substituted alkane.
(ex: 2-bromo-1-chloro-3-methylcyclopentane)
6) Indicate the carbon number with the functional group with the highest priority according to alphabetical order. A dash"-" must be placed
between the numbers and the name of the substituent. After the carbon number and the dash, the name of the substituent can follow. When
there is only one substituent on the parent chain, indicating the number of the carbon atoms with the substituent is not necessary.
EXAMPLE 4.1.2
Notice that "f" of fluoro alphabetically precedes the "m" of methyl. Although "di" alphabetically precedes "f", it is not used in determining
the alphabetical order.
EXAMPLE 4.1.3
cis-1-chloro-2-methylcyclopentane
9) After all the functional groups and substituents have been mentioned with their corresponding numbers, the name of the cycloalkane can
follow.
REACTIVITY
Cycloalkanes are very similar to the alkanes in reactivity, except for the very small ones, especially cyclopropane. Cyclopropane is
significantly more reactive than what is expected because of the bond angles in the ring. Normally, when carbon forms four single bonds,
the bond angles are approximately 109.5°. In cyclopropane, the bond angles are 60°.
With the electron pairs this close together, there is a significant amount of repulsion between the bonding pairs joining the carbon atoms,
making the bonds easier to break.
EXAMPLE 4.1.4
The alcohol substituent is given the lowest number even though the two methyl groups are on the same carbon atom and labeling 1 on
that carbon atom would give the lowest possible numbers. Numbering the location of the alcohol substituent is unnecessary because the
ending "-ol" indicates the presence of one alcohol group on carbon atom number 1.
EXAMPLE 4.1.5
EXAMPLE 4.1.6
trans-cyclohexane-1,2-diol
alkyne -yne
alcohol -ol
ether -ether
nitrile -nitrile
amine -amine
aldehyde -al
ketone -one
ester -oate
amide -amide
Although alkynes determine the name ending of a molecule, alkyne as a substituent on a cycloalkane is not possible because alkynes are
planar and would require that the carbon that is part of the ring form 5 bonds, giving the carbon atom a negative charge.
EXAMPLE 4.1.7
ethynylcyclooctane
EXAMPLE 4.1.8
1-propylcyclohexane
SUMMARY
1. Determine the parent chain: the parent chain contains the most carbon atoms.
2. Number the substituents of the chain so that the sum of the numbers is the lowest possible.
3. Name the substituents and place them in alphabetical order.
4. If stereochemistry of the compound is shown, indicate the orientation as part of the nomenclature.
5. Cyclic hydrocarbons have the prefix "cyclo-" and have an "-alkane" ending unless there is an alcohol substituent present. When an
alcohol substituent is present, the molecule has an "-ol" ending.
GLOSSARY
alcohol: An oxygen and hydrogenOH hydroxyl group that is bonded to a substituted alkyl group.
alkyl: A structure that is formed when a hydrogen atom is removed from an alkane.
cyclic: Chemical compounds arranged in the form of a ring or a closed chain form.
cycloalkanes: Cyclic saturated hydrocarbons with a general formula of CnH(2n). Cycloalkanes are alkanes with carbon atoms attached
in the form of a closed ring.
functional groups: An atom or groups of atoms that substitute for a hydrogen atom in an organic compound, giving the compound
unique chemical properties and determining its reactivity.
hydrocarbon: A chemical compound containing only carbon and hydrogen atoms.
saturated: All of the atoms that make up a compound are single bonded to the other atoms, with no double or triple bonds.
skeletal structure: A simplified structure in which each intersection between two lines is assumed to have a carbon atom with its
corresponding number of hydrogens.
PROBLEMS
Name the following structures. (Note: The structures are complex for practice purposes and may not be found in nature.)
1) 2) 3) 4) 5) 6)
7)
18) 19)
INSIDE LINKS
Nomenclature of Alcohols
Nomenclature of Ethers
Nomenclature of Esters
Nomenclature of Alkenes
Nomenclature of Ketones and Aldehydes
Nomenclature of Alkynes
OUTSIDE LINKS
More Practice Problems on Nomenclature of Cycloalkanes
Vollhardt, Schore. Organic Chemistry. 5th ed.
Wikipedia: Cycloalkanes
http://www.cem.msu.edu/~reusch/VirtualText/nomen1.htm
http://www.chemguide.co.uk/organicprops/alkanes/background.html
http://www.cem.msu.edu/~reusch/VirtualText/nomen1.htm
http://science.csustan.edu/nhuy/chem...IVNamecyal.htm
http://en.wikibooks.org/wiki/Organic...s/Cycloalkanes
REFERENCES
1. ACD/ChemSketch Freeware, version 11.0, Advanced Chemistry Development, Inc., Toronto, ON, Canada, www.acdlabs.com, 2008.
2. Bruice, Paula Yurkanis. Oragnic Chemistry. 5th. CA. Prentice Hall, 2006.
3. Fryhle, C.B. and G. Solomons. Organic Chemistry. 9th ed. Danvers, MA: Wiley, 2008.
4. McMurry, John. Organic Chemistry. 7th ed. Belmont, California: Thomson Higher Education, 2008.
5. Sadava, Heller, Orians, Purves, Hillis. Life The Science of Biology. 8th ed. Sunderland, MA: W.H. Freeman, 2008.
6. Vollhardt, K. Peter C., and Neil E. Schore. Organic Chemistry. 5th ed. New York: W.H. Freeman, 2007.
EXERCISES
Q4.1.2
Draw the following structures
1 = 1,2-dimethylcyclohexane
2 = 2-cyclopropyl butane
3 = 1,2,3-trimethyl cyclopentane
SOLUTIONS
S4.1.1
1 = 1-cyclopropyl cyclopropane
2 = 1-isopropyl cyclohexane
3 = 2-propenyl cyclopentane
CONTRIBUTORS
Pwint Zin
Jim Clark (ChemGuide)
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
Prof. Steven Farmer (Sonoma State University)
Organic Chemistry With a Biological Emphasis by Tim Soderberg (University of Minnesota, Morris)
KEY TERMS
Make certain that you can define, and use in context, the key terms below.
cis-trans isomers
cycloalkane
stereoisomer
STUDY NOTES
Compounds in which the carbon atoms are joined in a ring, and in which no multiple bonds or other functional groups are present,
are called cycloalkanes. A cycloalkane containing only one ring will correspond to the general formula CnH2n, but alkenes
containing only one double bond also correspond to this formula. Cycloalkanes containing more than one ring are not considered in
this section.
Stereoisomers are also observed in certain disubstituted (and higher substituted) cyclic compounds. Unlike the relatively flat molecules
of alkenes, substituted cycloalkanes must be viewed as three-dimensional configurations in order to appreciate the spatial orientations
of the substituents. By agreement, chemists use heavy, wedge-shaped bonds to indicate a substituent located above the average plane of
the ring, and a hatched line for bonds to atoms or groups located below the ring. As in the case of the 2-butene stereoisomers,
disubstituted cycloalkane stereoisomers may be designated by nomenclature prefixes such as cis and trans. The stereoisomeric 1,2-
dibromocyclopentanes below are an example.
In general, if any two sp3 carbons in a ring have two different substituent groups (not counting other ring atoms) stereoisomerism is
possible. This is similar to the substitution pattern that gives rise to stereoisomers in alkenes; indeed, one might view a double bond as
a two-membered ring. Four other examples of this kind of stereoisomerism in cyclic compounds are shown below.
EXERCISES
QUESTIONS
Q4.2.1
Draw the following molecules:
trans-1,3-dimethylcyclohexane
trans-1,2-dibromocyclopentane
cis-1,3-dichlorocyclobutane
SOLUTIONS
S4.2.1
KEY TERMS
Make certain that you can define, and use in context, the key terms below.
angle strain
heat of combustion
steric strain
torsional strain
Cycloalkanes have one or more rings of carbon atoms. The simplest examples of this class consist of a single, unsubstituted carbon ring, and these form a homologous series similar to the unbranched
alkanes. The IUPAC names of the first five members of this series are given in the following table. The last (yellow shaded) column gives the general formula for a cycloalkane of any size. If a simple
unbranched alkane is converted to a cycloalkane two hydrogen atoms, one from each end of the chain, must be lost. Hence the general formula for a cycloalkane composed of n carbons is CnH2n. Although
a cycloalkane has two fewer hydrogens than the equivalent alkane, each carbon is bonded to four other atoms so such compounds are still considered to be saturated with hydrogen.
Table 4.4.1: Examples of Simple Cycloalkanes
Name Cyclopropane Cyclobutane Cyclopentane Cyclohexane Cycloheptane Cycloalkane
Molecular
C3H6 C4H8 C5H10 C6H12 C7H14 CnH2n
Formula
Structural
(CH2)n
Formula
Line
Formula
One of the most interesting developments in ‘the stereochemistry of organic compounds in recent years has been the demonstration that transcyclooctene (but not the cis isomer) can be resolved into stable
chiral isomers (enantiomers, Section 5-IB). In general, a trans-cycloalkene would not be expected to be resolvable because of the possibility for formation of achiral conformations with a plane of
symmetry. Any conformation with all of the carbons in a plane is such an achiral conformation (Figure 12-20a). However, when the chain connecting the ends of the double bond is short, as in trans-
cyclooctene, steric hindrance and steric strain prevent easy.
Below is a chart of cycloalkanes and their respective heats of combustion (ΔHcomb). The ΔHcomb value increases as the number of carbons in the cycloalkane increases (higher membered ring), and the
ΔHcomb/CH2 ratio decreases. The increase in ΔHcomb can be attributed to the greater amount of London Dispersion forces. However, the decrease in ΔHcomb/CH2can be attributed to a decrease in the ring
strain.
Certain cycloalkanes, such as cyclohexane, deal with ring strain by forming conformers. A conformer is a stereoisomer in which molecules of the same connectivity and formula exist as different isomers,
in this case, to reduce ring strain. The ring strain is reduced in conformers due to the rotations around the sigma bonds. More about cyclohexane and its conformers can be seen here.
DIFFERENT TYPES OF STRAIN
There are many different types of strain that occur with cycloalkanes. In addition to ring strain, there is also transannular strain, eclipsing, or torsional strain and bond angle strain. Transannular strain exists
when there is steric repulsion between atoms. Eclipsing (torsional) strain exists when a cycloalkane is unable to adopt a staggered conformation around a C-C bond, and bond angle strain is the energy
needed to distort the tetrahedral carbons enough to close the ring. The presence of angle strain in a molecule indicates that there are bond angles in that particular molecule that deviate from the ideal bond
angles required (i.e., that molecule has conformers).
EXERCISES
QUESTIONS
Q4.3.1
trans-1,2-Dimethylcyclobutane is more stable than cis-1,2-dimethylcyclobutane. Explain this observation.
SOLUTIONS
S4.3.1
The trans form does not have eclipsing methyl groups, therefore lowering the energy within the molecule. It does however have hydrogen-methyl interactions, but are not as high in energy than methyl-
methyl interactions.
STUDY NOTES
Notice that in both cyclobutane and cyclopentane, torsional strain is reduced at the cost of increasing angular (angle) strain.
Although the customary line drawings of simple cycloalkanes are geometrical polygons, the actual shape of these compounds in most
cases is very different.
Cyclopropane is necessarily planar (flat), with the carbon atoms at the corners of an equilateral triangle. The 60º bond angles are much
smaller than the optimum 109.5º angles of a normal tetrahedral carbon atom, and the resulting angle strain dramatically influences the
chemical behavior of this cycloalkane. Cyclopropane also suffers substantial eclipsing strain, since all the carbon-carbon bonds are
fully eclipsed. Cyclobutane reduces some bond-eclipsing strain by folding (the out-of-plane dihedral angle is about 25º), but the total
eclipsing and angle strain remains high. Cyclopentane has very little angle strain (the angles of a pentagon are 108º), but its eclipsing
strain would be large (about 10 kcal/mol) if it remained planar. Consequently, the five-membered ring adopts non-planar puckered
conformations whenever possible.
Rings larger than cyclopentane would have angle strain if they were planar. However, this strain, together with the eclipsing strain
inherent in a planar structure, can be relieved by puckering the ring. Cyclohexane is a good example of a carbocyclic system that
virtually eliminates eclipsing and angle strain by adopting non-planar conformations. Cycloheptane and cyclooctane have greater strain
than cyclohexane, in large part due to transannular crowding (steric hindrance by groups on opposite sides of the ring).
Cyclic systems are a little different from open-chain systems. In an open chain, any bond can be rotated 360 degrees, going through
many different conformations. That complete rotation isn't possible in a cyclic system, because the parts that you would be trying to
twist away from each other would still be connected together. Cyclic systems have fewer "degrees of freedom" than aliphatic systems;
they have "restricted rotation".
Because of the restricted rotation of cyclic systems, most of them have much more well-defined shapes than their aliphatic
counterparts. Let's take a look at the basic shapes of some common rings.
Many biologically important compounds are built around structures containing rings, so it's important that we become familiar with
them.
In nature, three- to six-membered rings are frequently encountered, so we'll focus on those.
CYCLOPROPANE
A three membered ring has no rotational freedom whatsoever. A plane is defined by three points, so the three carbon atoms in
cyclopropane are all constrained to lie in the same plane.
Furthermore, if you look at a model you will find that the neighboring C-H bonds (C-C bonds, too) are all held in eclipsed
conformations.
However, the ring isn't big enough to introduce any steric strain, which does not become a factor until we reach six membered rings.
Until that point, rings are not flexible enough for two atoms to reach around and bump into each other.
The really big problem with cyclopropane is that the C-C-C bond angles are all too small.
All the carbon atoms in cyclopropane appear to be tetrahedral.
These bond angles ought to be 109 degrees.
The angles in an equilateral triangle are actually 60 degrees, about half as large as the optimum angle.
This factor introduces a huge amount of strain in the molecule, called ring strain.
CYCLOBUTANE
Cyclobutane is a four membered ring. In two dimensions, it is a square, with 90 degree angles at each corner.
However, in three dimensions, cyclobutane is flexible enough to buckle into a "butterfly" shape, relieving torsional strain a little bit.
When it does that, the bond angles get a little worse, going from 90 degrees to 88 degrees.
In a line drawing, this butterfly shape is usually shown from the side, with the near edges drawn using darker lines.
With bond angles of 88 rather than 109 degrees, cyclobutane has a lot of ring strain, but less than in cyclopropane.
Torsional strain is still present, but the neighbouring bonds are not exactly eclipsed in the butterfly.
Cyclobutane is still not large enough that the molecule can reach around to cause crowding. Steric strain is very low.
Cyclobutanes are a little more stable than cyclopropanes and are also a little more common in nature.
CYCLOPENTANE
Cyclopentanes are even more stable than cyclobutanes, and they are the second-most common paraffinic ring in nature, after
cyclohexanes. In two dimensions, a cyclopentane appears to be a regular pentagon.
In three dimensions, there is enough freedom of rotation to allow a slight twist out of this planar shape. In a line drawing, this three-
dimensional shape is drawn from an oblique view, just like cyclobutane.
The ideal angle in a regular pentagon is about 107 degrees, very close to a tetrahedral bond angle.
Cyclopentane distorts only very slightly into an "envelope" shape in which one corner of the pentagon is lifted up above the plane
of the other four, and as a result, ring strain is entirely removed.
The envelope removes torsional strain along the sides and flap of the envelope. However, the neighbouring carbons are eclipsed
along the "bottom" of the envelope, away from the flap. There is still some torsional strain in cyclopentane.
Again, there is no steric strain in this system.
SOLUTIONS
S4.4.1
There are 8 eclipsing interactions (two per C-C bond). The extra strain on this molecule would be 32 kJ/mol (4 kJ/mol x 8).
S4.4.2
The first conformation is more stable. Even though the methyl groups are cis in the model on the left, they are eclipsing due the
conformation, therefore increasing the strain within the molecule.
CONTRIBUTORS
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
Prof. Steven Farmer (Sonoma State University)
Chris P Schaller, Ph.D., (College of Saint Benedict / Saint John's University)
William Reusch, Professor Emeritus (Michigan State U.), Virtual Textbook of Organic Chemistry
KEY TERMS
Make certain that you can define, and use in context, the key terms below.
chair conformation
twist-boat conformation
Rings larger than cyclopentane would have angle strain if they were planar. However, this strain, together with the eclipsing strain
inherent in a planar structure, can be relieved by puckering the ring. Cyclohexane is a good example of a carbocyclic system that
virtually eliminates eclipsing and angle strain by adopting non-planar conformations. Cycloheptane and cyclooctane have greater strain
than cyclohexane, in large part due to transannular crowding (steric hindrance by groups on opposite sides of the ring).
CONFORMATIONS OF CYCLOHEXANE
A planar structure for cyclohexane is clearly improbable. The bond angles would necessarily be 120º, 10.5º larger than the ideal
tetrahedral angle. Also, every carbon-carbon bond in such a structure would be eclipsed. The resulting angle and eclipsing strains
would severely destabilize this structure. If two carbon atoms on opposite sides of the six-membered ring are lifted out of the plane of
the ring, much of the angle strain can be eliminated.
This boat structure still has two eclipsed bonds and severe steric crowding of two hydrogen atoms on the "bow" and "stern" of the boat.
This steric crowding is often called steric hindrance. By twisting the boat conformation, the steric hindrance can be partially relieved,
but the twist-boat conformer still retains some of the strains that characterize the boat conformer. Finally, by lifting one carbon above
the ring plane and the other below the plane, a relatively strain-free 'chair' conformer is formed. This is the predominant structure
adopted by molecules of cyclohexane.
EXERCISES
QUESTIONS
Q4.5.1
Consider the conformations of cyclohexane, chair, boat, twist boat. Order them in increasing strain in the molecule.
SOLUTIONS
S4.5.1
Chair < Twist Boat < Boat (most strain)
CONTRIBUTORS
Robert Bruner (http://bbruner.org)
William Reusch, Professor Emeritus (Michigan State U.), Virtual Textbook of Organic Chemistry
KEY TERMS
Make certain that you can define, and use in context, the key terms below.
axial position
equatorial position
ring flip
On careful examination of a chair conformation of cyclohexane, we find that the twelve hydrogens are not structurally equivalent. Six
of them are located about the periphery of the carbon ring, and are termed equatorial. The other six are oriented above and below the
approximate plane of the ring (three in each location), and are termed axial because they are aligned parallel to the symmetry axis of
the ring.
In the figure above, the equatorial hydrogens are colored blue, and the axial hydrogens are in bold. Since there are two equivalent chair
conformations of cyclohexane in rapid equilibrium, all twelve hydrogens have 50% equatorial and 50% axial character. The figure
below illustrates how to convert a molecular model of cyclohexane between two different chair conformations - this is something that
you should practice with models. Notice that a 'ring flip' causes equatorial hydrogens to become axial, and vice-versa.
In ISIS/Draw, the "up wedge" and "down bond" that I used, along with other variations, are
available from a tool button that may be labeled with any of them, depending on most
recent use. It is located directly below the tool button for ordinary C-C bonds.
When the methyl group in the structure above occupies an axial position it suffers steric crowding by the two axial hydrogens located
on the same side of the ring.
EXERCISES
QUESTIONS
Q4.6.1
Draw two conformations of cyclohexyl amine (C6H11NH2). Indicate axial and equatorial positions.
Q4.6.2
Draw the two isomers of 1,4-dihydroxylcyclohexane, identify which are equatorial and axial.
Q4.6.3
In the following molecule, label which are equatorial and which are axial, then draw the chair flip (showing labels 1,2,3).
SOLUTIONS
S4.6.1
S4.6.2
S4.6.3
Original conformation: 1 = axial, 2 = equatorial, 3 = axial
Flipped chair now looks like this.
KEY TERMS
Make certain that you can define, and use in context, the key term below.
1,3‑diaxial interaction
STUDY NOTES
1,3-Diaxial interactions are steric interactions between an axial substituent located on carbon atom 1 of a cyclohexane ring and the
hydrogen atoms (or other substituents) located on carbon atoms 3 and 5.
Be prepared to draw Newman-type projections for cyclohexane derivatives as the one shown for methylcyclohexane. Note that this
is similar to the Newman projections from chapter 3 such as n-butane.
Because axial bonds are parallel to each other, substituents larger than hydrogen generally suffer greater steric crowding when they are
oriented axial rather than equatorial. Consequently, substituted cyclohexanes will preferentially adopt conformations in which the
larger substituents assume equatorial orientation.
When the methyl group in the structure above occupies an axial position it suffers steric crowding by the two axial hydrogens located
on the same side of the ring.
The conformation in which the methyl group is equatorial is more stable, and thus the equilibrium lies in this direction.
The relative steric hindrance experienced by different substituent groups oriented in an axial versus equatorial location on cyclohexane
may be determined by the conformational equilibrium of the compound. The corresponding equilibrium constant is related to the
energy difference between the conformers, and collecting such data allows us to evaluate the relative tendency of substituents to exist
in an equatorial or axial location.A table of these free energy values (sometimes referred to as A values) may be examined by clicking
here.
CH −
3
1.7 O N−
2
1.1
CH H −
2 5
1.8 N≡C− 0.2
(CH ) CH−
3 2
2.2 CH O−
3
0.5
(CH ) C−
3 3
≥ 5.0 0.7
F− 0.3 1.3
Cl− 0.5 C H −
6 5
3.0
Br− 0.5
I− 0.5
CHLOROCYCLOHEXANE
This is an example of the next level of complexity, a mono-substituted cycloalkane. See Fig 3.
Chlorocyclohexane: simple hexagon and chair structures, showing hydrogen atoms.
Figure 4.7.3:
So what is new here? Not much, with the hexagon formula, Fig 3A. That type of formula shows the basic "connectivity" of the atoms -
- who is connected to whom. This chemical has one Cl on the ring, and it does not matter where we show it. There is now only one H
on that C, but since we are not showing H explicitly here, that is not an issue in drawing the structure. (It is an issue when you look at it
and want to count H.)
With the chair formula (Fig 3B), which shows information not only about connectivity but also about conformation, there is important
new information here. In a chair, there are two "types" of substituents: those pointing up or down, and called axial, and those pointing
"outward", and called equatorial. I have shown the chlorine atom in an equatorial position. Why? Two reasons: it is what we would
predict, and it is what is found. Why do we predict that the Cl is equatorial? Because it is bigger than H, and there is more room in the
equatorial positions.
HELPFUL HINTS
If possible, examine a physical model of cyclohexane and chlorocyclohexane, so that you can see the axial and equatorial positions.
Common ball and stick models are fine for this. It should be easy to see that the three axial H on one side can get very near each
other.
EXERCISES
QUESTIONS
Q4.7.1
In the molecule, cyclohexyl ethyne there is little steric strain, why?
SOLUTIONS
S4.7.1
The ethyne group is linear and therefore does not affect the hydrogens in the 1,3 positions to say to the extent as a bulkier or a bent
group (e.g. ethene group) would. This leads to less of a strain on the molecule.
CONTRIBUTORS
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
Prof. Steven Farmer (Sonoma State University)
Robert Bruner (http://bbruner.org)
Organic Chemistry With a Biological Emphasis by Tim Soderberg (University of Minnesota, Morris)
KEY TERMS
Make certain that you can define, and use in context, the key term below.
conformational analysis
STUDY NOTES
When faced with the problem of trying to decide which of two conformers of a given disubstituted cyclohexane is the more stable,
you may find the following generalizations helpful.
1. A diequatorial conformation will always be more stable than a diaxial one.
2. When one substituent is axial and the other is equatorial, the most stable conformation will be the one with the bulkiest
substituent in the equatorial position. Steric bulk decreases in the order
tert-butyl > isopropyl > ethyl > methyl > hydroxyl > halogens
MONOSUBSTITUTED CYCLOHEXANES
The conformation in which the methyl group is equatorial is more stable, and thus the equilibrium lies in this direction.
The relative steric hindrance experienced by different substituent groups oriented in an axial versus equatorial location on cyclohexane
may be determined by the conformational equilibrium of the compound. The corresponding equilibrium constant is related to the
energy difference between the conformers, and collecting such data allows us to evaluate the relative tendency of substituents to exist
in an equatorial or axial location. Table 4.9.1 show some of these free energy values (sometimes referred to as A values).
DICHLOROCYCLOHEXANES
The next level of complexity are di-substituted cycloalkanes and we will use dichlorocyclohexane as the model system for discussing
these compounds. The first question we must ask is which C the two chlorine substituents are on. For now, I want to discuss 1,3-
dichlorocyclohexane. This introduces another issue: are the two Cl on the same side of the ring, or on opposite sides? We call these
"cis" (same side) and "trans" (opposite sides). We focus on one of these, cis-1,3-dichlorocyclohexane. And for now, we will just look at
hexagon structural formulas, leaving the question of conformation for later. Let's go through this one step at a time.
1,3-DICHLOROCYCLOHEXANE
Our first attempt to draw 1,3-dichlorocyclohexane might look something like Figure 4.9.1
CIS-1,3-DICHLOROCYCLOHEXANE
Figure 4.8.2 shows two common ways to show how the substituents are oriented relative to the plane of the ring. The compound shown
here is cis-1,3-dichlorocyclohexane.
Figure 4.9.2
The basic idea in both of these is that we can imagine the ring to be planar, and then show the groups above or below the plane of the
ring. How we show this is different in the two parts of Figure 4.9.2. In Figure 4.9.1A, we look at the ring "edge-on". The thick line for
the bottom bond is intended to convey the edge-on view (or side view); this is sometimes omitted, especially with hand-drawn
structures, but be careful then that the meaning is clear. Once we understand that we are now looking at the ring edge-on, it is clear that
the two Cl atoms are both above the ring, hence cis. In Figure 4.9.2B, we view the ring "face-on" (or top view), and use special bond
symbols -- "stereo bonds" -- to convey up and down: the heavy wedge -- an "up wedge" -- points upward, toward you, and the dashed
bond -- a "down bond" -- points downward, away from you. Again, both Cl are "up", hence cis.
NOTE
For some notes on how to draw the stereo bonds, see the section E.1. Note: How to draw stereo bonds ("up" and "down" bonds).
In discussing Figure 4.9.2, I started by saying that we imagine the ring to be planar. Emphasize that cycloalkane rings are not really
planar (except for cyclopropane rings). As so often, the structural formula represents the general layout of the atoms, but not the
actual molecular geometry.
Figure 4.9.3
The structure shows cis-1,3-dichlorocyclohexane: a 6-ring; 2 Cl atoms, at positions 1 and 3; and cis, with both Cl on the same side of --
above -- the H that is on the same C.
I showed the 2 Cl atoms at corner positions, and I showed the H at the key positions explicitly. These points follow from some of the
Helpful Hints discussed earlier. It is not required that you do these things, but they can make things easier for you -- and for anyone
reading your structures.
Now, what is new here? The conformation. We start with the notion that the conformation of cyclohexane derivatives is based on the
"chair". At each position, one substituent is axial (loosely, perpendicular to the ring), and one is equatorial (loosely, in the plane of the
ring). There is more room in the equatorial positions (not easily seen with these simple drawings, but ordinary ball and stick models do
help with this point). Thus we try to put the larger substituents in the equatorial positions. In this case, we put the Cl equatorial and the
H axial at each position 1 and position 3.
We are now done with this compound, cis-1,3-dichlorocyclohexane. However, we have missed one very important concern -- because
it is not an issue in this case. So let's look at another compound.
Figure 4.9.15
Figures 4.9.4 and 4.9.5 introduce no new ideas or complications; these two figures should be straightforward. So, what is the preferred
conformation of cis-1,2-dichlorocyclohexane? This requires careful consideration; an important lesson from this exercise is to realize
that we cannot propose a good conformation based simply on what we have learned so far. The two guidelines we have so far for
conformation of 6-rings are:
The carbon ring is in a chair.
Larger substituents are in equatorial positions.
Let's explore the difficulty here by looking at some things people might naively draw. Figure 4.9.6 shows an attempt to draw a chair
conformation of cis-1,2-dichlorocyclohexane. It satisfies both of the guidelines listed above. But it is wrong.
Figure 4.9.6: An attempt to draw a chair conformation of cis-1,2-dichlorocyclohexane. But this is the wrong chemical.
Why is Figure 4.9.6 wrong? It is the wrong chemical. The structure shown in Figure 4.9.6 is trans, not cis. Look carefully at the 1 and
2 positions. At one of them, the H is above the Cl; at the other, the Cl is above the H. Trans. Wrong chemical. The structure shown in
Figure 4.9.6 is not the requested chemical. Figure 4.9.7 attempts to fix the problem with Figure 4.9.6. However, it is also wrong.
Figure 4.9.7
Why is Figure 4.9.8 wrong? After all, it seems to address the criteria presented. It contains both of the larger atoms (Cl) equatorial, and
they are cis as desired. However, in Figure 4.8.7, the two axial groups on carbons # 1 and 2 (the two H that are shown) are both
pointing up. This is impossible. In a valid chair, the axial groups alternate up/down as one goes around the ring
Figure 4.9.8
EXERCISES
QUESTIONS
Q4.8.1
For the following molecules draw the most stable chair conformation and explain why you chose this as an answer
1 = trans-1,2-dimethylcyclohexane
2 = cis-1,3-dimethylcyclohexane
SOLUTIONS
S4.8.1
1 – The most stable conformation would be to have the methyl groups axial reducing steric interaction
2 – The most stable conformation would be to have the groups equatorial this would reduce the strain if they were axial
KEY TERMS
Make certain that you can define, and use in context, the key terms below.
bridgehead carbon atom
polycyclic molecule
STUDY NOTES
A bridgehead carbon atom is a carbon atom which is shared by at least two rings. The hydrogen atom which is attached to a
bridgehead carbon may be referred to as a bridgehead hydrogen.
Note that bicyclo[2.2.1]heptane is the systematic name of norborane. You need not be concerned over the IUPAC name of
norbornane. The nomenclature of compounds of this type is beyond the scope of this course.
A bridged bicycloalkane is a bicycloalkane whose molecule has two carbon atoms shared by all three rings identifiable in the molecule.
The two carbon atoms shared by the three rings are called bridgehead carbon atoms. A bond or a chain of bonds connecting the
bridgehead carbon atoms is called a bridge.
eg. 1: Decalin
eg. 2: bicyclo[2.2.1]heptane
STEROIDS
Steroids include such well known compounds as cholesterol, sex hormones, birth control pills, cortisone, and anabolic steroids.
CHOLESTEROL
The best known and most abundant steroid in the body is cholesterol. Cholesterol is formed in brain tissue, nerve tissue, and the blood
stream. It is the major compound found in gallstones and bile salts. Cholesterol also contributes to the formation of deposits on the
inner walls of blood vessels. These deposits harden and obstruct the flow of blood. This condition, known as atherosclerosis, results in
various heart diseases, strokes, and high blood pressure.
Much research is currently underway to determine if a correlation exists between cholesterol levels in the blood and diet. Not only does
cholesterol come from the diet, but cholesterol is synthesized in the body from carbohydrates and proteins as well as fat. Therefore, the
elimination of cholesterol rich foods from the diet does not necessarily lower blood cholesterol levels. Some studies have found that if
certain unsaturated fats and oils are substituted for saturated fats, the blood cholesterol level decreases. The research is incomplete on
this problem.
EXERCISES
QUESTIONS
Q4.9.1
Someone stated that trans-decalin is more stable than cis-decalin. Explain why this is incorrect.
SOLUTIONS
S4.9.1
Cis-decalin has fewer steric interactions than trans-decalin.
CONTRIBUTORS
Gamini Gunawardena from the OChemPal site (Utah Valley University)
John D. Robert and Marjorie C. Caserio (1977) Basic Principles of Organic Chemistry, second edition. W. A. Benjamin, Inc. ,
Menlo Park, CA. ISBN 0-8053-8329-8. This content is copyrighted under the following conditions, "You are granted permission
for individual, educational, research and non-commercial reproduction, distribution, display and performance of this work in any
format."
CONTRIBUTORS
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
KEY TERMS
Make certain that you can define, and use in context, the key term below.
enantiomer
STUDY NOTES
Stereoisomers are isomers that differ in spatial arrangement of atoms, rather than order of atomic connectivity. One of their most
interesting type of isomer is the mirror-image stereoisomers, a non-superimposable set of two molecules that are mirror image of
one another. The existance of these molecules are determined by concept known as chirality. The word “chiral” was derived from
the Greek word for hand, because our hands display a good example of chirality since they are non-superimposable mirror images
of each other.
INTRODUCTION
The opposite of chiral is achiral. Achiral objects are superimposable with their mirror images. For example, two pieces of paper are
achiral. In contrast, chiral molecules, like our hands, are non superimposable mirror images of each other. Try to line up your left hand
perfectly with your right hand, so that the palms are both facing in the same directions. Spend about a minute doing this. Do you see
that they cannot line up exactly? The same thing applies to some molecules
A Chiral molecule has a mirror image that cannot line up with it perfectly- the mirror images are non superimposable. The mirror
images are called enantiomers. But why are chiral molecules so interesting? A chiral molecule and its enantiomer have the same
chemical and physical properties(boiling point, melting point,polarity, density etc...). It turns out that many of our biological molecules
such as our DNA, amino acids and sugars, are chiral molecules.
It is pretty interesting that our hands seem to serve the same purpose but most people are only able to use one of their hands to write.
Similarily this is true with chiral biological molecules and interactions. Just like your left hand will not fit properly in your right glove,
one of the enantiomers of a molecule may not work the same way in your body.
This must mean that enantiomers have properties that make them unique to their mirror images. One of these properties is that they
cannot have a plane of symmetry or an internal mirror plane. So, a chiral molecule cannot be divided in two mirror image halves.
Another property of chiral molecules is optical activity.
Organic compounds, molecules created around a chain of carbon atom (more commonly known as carbon backbone), play an essential
role in the chemistry of life. These molecules derive their importance from the energy they carry, mainly in a form of potential energy
between atomic molecules. Since such potential force can be widely affected due to changes in atomic placement, it is important to
understand the concept of an isomer, a molecule sharing same atomic make up as another but differing in structural arrangements. This
article will be devoted to a specific isomers called stereoisomers and its property of chirality (Figure 5.1.1).
SPATIAL ARRANGEMENT
First and foremost, one must understand the concept of spatial arrangement in order to understand stereoisomerism and chirality.
Spatial arrangement of atoms concern how different atomic particles and molecules are situated about in the space around the organic
compound, namely its carbon chain. In this sense, spatial arrangement of an organic molecule are different another if an atom is shifted
in any three-dimensional direction by even one degree. This opens up a very broad possibility of different molecules, each with their
unique placement of atoms in three-dimensional space .
STEREOISOMERS
Stereoisomers are, as mentioned above, contain different types of isomers within itself, each with distinct characteristics that further
separate each other as different chemical entities having different properties. Type called entaniomer are the previously-mentioned
mirror-image stereoisomers, and will be explained in detail in this article. Another type, diastereomer, has different properties and will
be introduced afterwards.
ENANTIOMERS
This type of stereoisomer is the essential mirror-image, non-superimposable type of stereoisomer introduced in the beginning of the
article. Figure 3 provides a perfect example; note that the gray plane in the middle demotes the mirror plane.
Figure 5.1.2.
Note that even if one were to flip over the left molecule over to the right, the atomic spatial arrangement will not be equal. This is
equivalent to the left hand - right hand relationship, and is aptly referred to as 'handedness' in molecules. This can be somewhat
counter-intuitive, so this article recommends the reader try the 'hand' example. Place both palm facing up, and hands next to each other.
Now flip either side over to the other. One hand should be showing the back of the hand, while the other one is showing the palm. They
are not same and non-superimposable. This is where the concept of chirality comes in as one of the most essential and defining idea of
stereoisomerism.
CHIRALITY
Figure 5.1.3.
In this case, the molecule is considered 'achiral'. In other words, to distinguish chiral molecule from an achiral molecule, one must
search for the existence of the bisecting plane in a molecule. All chiral molecules are deprive of bisecting plane, whether simple or
complex. As a universal rule, no molecule with different surrounding atoms are achiral. Chirality is a simple but essential idea to
support the concept of stereoisomerism, being used to explain one type of its kind. The chemical properties of the chiral molecule
differs from its mirror image, and in this lies the significance of chilarity in relation to modern organic chemistry.
CONTRIBUTORS
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
Prof. Steven Farmer (Sonoma State University)
William Reusch, Professor Emeritus (Michigan State U.), Virtual Textbook of Organic Chemistry
Organic Chemistry With a Biological Emphasis by Tim Soderberg (University of Minnesota, Morris)
Jim Clark (Chemguide.co.uk)
KEY TERMS
Make certain that you can define, and use in context, the key terms below.
achiral
chiral
chiral (stereogenic) centre
plane of symmetry
A consideration of the chirality of molecular configurations explains the curious stereoisomerism observed for lactic acid, carvone and
a multitude of other organic compounds. Tetravalent carbons have a tetrahedral configuration. If all four substituent groups are the
same, as in methane or tetrachloromethane, the configuration is that of a highly symmetric "regular tetrahedron". A regular tetrahedron
several planes of symmetry and is achiral.
A carbon atom that is bonded to four different atoms or groups loses all symmetry, and is often referred to as an asymmetric carbon.
The configuration of such a molecular unit is chiral, and the structure may exist in either a right-handed configuration or a left-handed
configuration (one the mirror image of the other). This type of configurational stereoisomerism is termed enantiomorphism, and the
non-identical, mirror-image pair of stereoisomers that result are called enantiomers. In the general figure below, A and B are
nonsuperposable mirror images of one another, and thus are a pair of enantiomers.
The structural formulas of lactic acid and carvone are drawn on the right with the asymmetric carbon colored red. Consequently, we
find that these compounds exist as pairs of enantiomers. The presence of a single asymmetrically substituted carbon atom in a molecule
is sufficient to render the whole configuration chiral, and modern terminology refers to such groupings as chiral centers. Most of the
chiral centers we shall discuss are asymmetric carbon atoms, but it should be recognized that other tetrahedral or pyramidal atoms may
become chiral centers if appropriately substituted. When more than one chiral center is present in a molecular structure, care must be
taken to analyze their relationship before concluding that a specific molecular configuration is chiral or achiral. This aspect of
stereoisomerism will be treated later.
A useful first step in examining structural formulas to determine whether stereoisomers may exist is to identify all stereogenic
elements. A stereogenic element is a center, axis or plane that is a focus of stereoisomerism, such that an interchange of two groups
attached to this feature leads to a stereoisomer. Stereogenic elements may be chiral or achiral. An asymmetric carbon is often a chiral
stereogenic center, since interchanging any two substituent groups converts one enantiomer to the other. Alkenes having two different
groups on each double bond carbon constitute an achiral stereogenic element, since interchanging substituents at one of the carbons
changes the cis/trans configuration of the double bond.
Thalidomide had previously been used in other countries as an antidepressant, and was believed to be safe and effective for both
purposes. The drug was not approved for use in the U.S.A. It was not long, however, before doctors realized that something had gone
horribly wrong: many babies born to women who had taken thalidomide during pregnancy suffered from severe birth defects.
Researchers later realized the that problem lay in the fact that thalidomide was being provided as a mixture of two different isomeric
forms.
One of the isomers is an effective medication, the other caused the side effects. Both isomeric forms have the same molecular formula
and the same atom-to-atom connectivity, so they are not constitutional isomers. Where they differ is in the arrangement in three-
dimensional space about one tetrahedral, sp3-hybridized carbon. These two forms of thalidomide are stereoisomers.
Note that the carbon in question has four different substituents (two of these just happen to be connected by a ring structure).
Tetrahedral carbons with four different substituent groups are called stereocenters.
EXAMPLE 5.2.1
Locate all of the carbon stereocenters in the molecules below.
Looking at the structures of what we are referring to as the two isomers of thalidomide, you may not be entirely convinced that they are
actually two different molecules. In order to convince ourselves that they are indeed different, let’s create a generalized picture of a
tetrahedral carbon stereocenter, with the four substituents designated R1-R4. The two stereoisomers of our simplified model look like
this:
If you look carefully at the figure above, you will notice that molecule A and molecule B are mirror images of each other (the line
labeled 's' represents a mirror plane). Furthermore, they are not superimposable: if we pick up molecule A, flip it around, and place it
next to molecule B, we see that the two structures cannot be superimposed on each other. They are different molecules!
If you make models of the two stereoisomers of thalidomide and do the same thing, you will see that they too are mirror images, and
cannot be superimposed (it well help to look at a color version of the figure below).
Here are some examples of molecules that are achiral (not chiral). Notice that none of these molecules has a stereocenter.
It is difficult to illustrate on the two dimensional page, but you will see if you build models of these achiral molecules that, in each
case, there is at least one plane of symmetry, where one side of the plane is the mirror image of the other. Chirality is tied
conceptually to the idea of asymmetry, and any molecule that has a plane of symmetry cannot be chiral. When looking for a plane of
symmetry, however, we must consider all possible conformations that a molecule could adopt. Even a very simple molecule like
ethane, for example, is asymmetric in many of its countless potential conformations – but it has obvious symmetry in both the eclipsed
and staggered conformations, and for this reason it is achiral.
We will see in chapter 10 how researchers, in order to investigate the stereochemistry of reactions at the phosphate center, incorporated
sulfur and/or 17O and 18O isotopes of oxygen (the ‘normal’ isotope is 16O) to create chiral phosphate groups. Phosphate triesters are
chiral if the three substituent groups are different.
Asymmetric quaternary ammonium groups are also chiral. Amines, however, are not chiral, because they rapidly invert, or turn ‘inside
out’, at room temperature.
EXAMPLE 5.2.2
Label the molecules below as chiral or achiral, and locate all stereocenters.
Answer:
SOLUTIONS
S5.2.1
CONTRIBUTORS
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
Prof. Steven Farmer (Sonoma State University)
William Reusch, Professor Emeritus (Michigan State U.), Virtual Textbook of Organic Chemistry
Organic Chemistry With a Biological Emphasis by Tim Soderberg (University of Minnesota, Morris)
Jim Clark (Chemguide.co.uk)
KEY TERMS
Make certain that you can define, and use in context, the key terms below.
analyzer
dextrorotatory
levorotatory
optically active
plane-polarized light
polarimeter
polarizer
specific rotation, [α]D
STUDY NOTES
A polarizer is a device through which only light waves oscillating in a single plane may pass. A polarimeter is an instrument used
to determine the angle through which plane-polarized light has been rotated by a given sample. You will have the opportunity to use
a polarimeter in the laboratory component of the course. An analyzer is the component of a polarimeter that allows the angle of
rotation of plane-polarized light to be determined.
20
Specific rotations are normally measured at 20°C, and this property may be indicated by the symbol [α]D . Sometimes the solvent
is specified in parentheses behind the specific rotation value, for example,
20
[α]
D
= + 12 ° (chloroform)
For liquids, the specific rotation may be obtained using the neat liquid rather than a solution; in such cases the formula is
temp
α
[α] (neat) =
D
l × d
where α is the observed rotation, l is the path length of the cell (measured in decimetres, dm), and d is the density of the liquid.
Identifying and distinguishing enantiomers is inherently difficult, since their physical and chemical properties are largely identical.
Fortunately, a nearly two hundred year old discovery by the French physicist Jean-Baptiste Biot has made this task much easier. This
discovery disclosed that the right- and left-handed enantiomers of a chiral compound perturb plane-polarized light in opposite ways.
This perturbation is unique to chiral molecules, and has been termed optical activity.
POLARIMETRY
Plane-polarized light is created by passing ordinary light through a polarizing device, which may be as simple as a lens taken from
polarizing sun-glasses. Such devices transmit selectively only that component of a light beam having electrical and magnetic field
vectors oscillating in a single plane. The plane of polarization can be determined by an instrument called a polarimeter, shown in the
diagram below.
where
[α]D is the specific rotation
l is the cell length in dm
D designates that the light used is the 589 line from a sodium lamp
Compounds that rotate the plane of polarized light are termed optically active. Each enantiomer of a stereoisomeric pair is optically
active and has an equal but opposite-in-sign specific rotation. Specific rotations are useful in that they are experimentally determined
constants that characterize and identify pure enantiomers. For example, the lactic acid and carvone enantiomers discussed earlier have
the following specific rotations.
Carvone from caraway: [α]D = +62.5º this isomer may be referred to as (+)-carvone or d-carvone
Carvone from spearmint: [α]D = –62.5º this isomer may be referred to as (–)-carvone or l-carvone
Lactic acid from muscle tissue: [α]D = +2.5º this isomer may be referred to as (+)-lactic acid or d-lactic acid
Lactic acid from sour milk: [α]D = –2.5º this isomer may be referred to as (–)-lactic acid or l-lactic acid
A 50:50 mixture of enantiomers has no observable optical activity. Such mixtures are called racemates or racemic modifications, and
are designated (±). When chiral compounds are created from achiral compounds, the products are racemic unless a single enantiomer
of a chiral co-reactant or catalyst is involved in the reaction. The addition of HBr to either cis- or trans-2-butene is an example of
racemic product formation (the chiral center is colored red in the following equation).
CH3CH=CHCH3 + HBr (±) CH3CH2CHBrCH3
ENANTIOMERIC EXCESS
The "optical purity" is a comparison of the optical rotation of a pure sample of unknown stereochemistry versus the optical rotation of
a sample of pure enantiomer. It is expressed as a percentage. If the sample only rotates plane-polarized light half as much as expected,
the optical purity is 50%.
Because R and S enantiomers have equal but opposite optical activity, it naturally follows that a 50:50 racemic mixture of two
enantiomers will have no observable optical activity. If we know the specific rotation for a chiral molecule, however, we can easily
calculate the ratio of enantiomers present in a mixture of two enantiomers, based on its measured optical activity. When a mixture
contains more of one enantiomer than the other, chemists often use the concept of enantiomeric excess (ee) to quantify the difference.
Enantiomeric excess can be expressed as:
For example, a mixture containing 60% R enantiomer (and 40% S enantiomer) has a 20% enantiomeric excess of R: ((60-50) x 100) /
50 = 20 %.
Answer:
The observed rotation of the mixture is levorotary (negative, counter-clockwise), and the specific rotation of the pure S
enantiomer is given as dextrorotary (positive, clockwise), meaning that the pure R enantiomer must be levorotary, and the
mixture must contain more of the R enantiomer than of the S enantiomer.
Rotation (R/S Mix) = [Fraction(S) × Rotation (S)] + [Fraction(R) × Rotation (R)]
Let Fraction (S) = x, therefore Fraction (R) = 1 – x
Rotation (R/S Mix) = x[Rotation (S)] + (1 – x)[Rotation (R)]
–23 = x(+61) + (1 – x)(–61)
Solve for x: x = 0.3114 and (1 – x) = 0.6885
Therefore the percentages of (R)- and (S)-carvone in the sample are 68.9% and 31.1%, respectively.
ee = [(% more abundant enantiomer – 50) × 100]/50
= [68.9 – 50) × 100]/50 = 37.8%
Chiral molecules are often labeled according to the sign of their specific rotation, as in (S)-(+)-carvone and (R)-(-)-carvone, or (±)-
carvone for the racemic mixture. However, there is no relationship whatsoever between a molecule's R/S designation and the sign of its
specific rotation. Without performing a polarimetry experiment or looking in the literature, we would have no idea that (-)-carvone has
the R configuration and (+)-carvone has the S configuration.
EXERCISES
1. For each of the structures indicated below, mark the chiral centres with an asterisk.
Answer:
1.
QUESTIONS
Q5.3.1
A sample with a concentration of 0.3 g/mL was placed in a cell with a length of 5 cm. The resulting rotation at the sodium D line was
+1.52°. What is the [α]D?
SOLUTIONS
S5.3.1
5 cm = 0.5 dm
[α]D = α/(c x l) = +1.52/(0.3 x 0.5) = +10.1°
Because enantiomers have identical physical and chemical properties in achiral environments, separation of the stereoisomeric
components of a racemic mixture or racemate is normally not possible by the conventional techniques of distillation and
crystallization. In some cases, however, the crystal habits of solid enantiomers and racemates permit the chemist (acting as a chiral
resolving agent) to discriminate enantiomeric components of a mixture. As background for the following example, it is recommended
that the section on crystal properties be reviewed.
Tartaric acid, its potassium salt known in antiquity as "tartar", has served as the locus of several landmark events in the history of
stereochemistry. In 1832 the French chemist Jean Baptiste Biot observed that tartaric acid obtained from tartar
was optically active, rotating the plane of polarized light clockwise (dextrorotatory). An optically inactive, higher
melting, form of tartaric acid, called racemic acid was also known. A little more than a decade later, young Louis
Pasteur conducted a careful study of the crystalline forms assumed by various salts of these acids. He noticed that
under certain conditions, the sodium ammonium mixed salt of the racemic acid formed a mixture of
enantiomorphic hemihedral crystals; a drawing of such a pair is shown on the right. Pasteur reasoned that the
dissymmetry of the crystals might reflect the optical activity and dissymmetry of its component molecules. After picking the different
crystals apart with a tweezer, he found that one group yielded the known dextrorotatory tartaric acid measured by Biot; the second led
to a previously unknown levorotatory tartaric acid, having the same melting point as the dextrorotatory acid. Today we recognize that
Pasteur had achieved the first resolution of a racemic mixture, and laid the foundation of what we now call stereochemistry.
Optical activity was first observed by the French physicist Jean-Baptiste Biot. He concluded that the change in direction of plane-
polarized light when it passed through certain substances was actually a rotation of light, and that it had a molecular basis. His work
was supported by the experimentation of Louis Pasteur. Pasteur observed the existence of two crystals that were mirror images in
tartaric acid, an acid found in wine. Through meticulous experimentation, he found that one set of molecules rotated polarized light
clockwise while the other rotated light counterclockwise to the same extent. He also observed that a mixture of both, a racemic mixture
(or racemic modification), did not rotate light because the optical activity of one molecule canceled the effects of the other molecule.
Pasteur was the first to show the existence of chiral molecules.
CONTRIBUTORS
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
Prof. Steven Farmer (Sonoma State University)
William Reusch, Professor Emeritus (Michigan State U.), Virtual Textbook of Organic Chemistry
KEY TERMS
Make certain that you can define, and use in context, the key terms below.
absolute configuration
R configuration
S configuration
STUDY NOTES
When designating a structure as R or S, you must ensure that the atom or group with the lowest priority is pointing away from you,
the observer. The easiest way to show this is to use the wedge-and-broken-line representation. You can then immediately determine
whether you are observing an R configuration or an S configuration.
To name the enantiomers of a compound unambiguously, their names must include the "handedness" of the molecule. The method for
this is formally known as R/S nomenclature.
INTRODUCTION
The method of unambiguously assigning the handedness of molecules was originated by three chemists: R.S. Cahn, C. Ingold, and V.
Prelog and, as such, is also often called the Cahn-Ingold-Prelog rules. In addition to the Cahn-Ingold system, there are two ways of
experimentally determining the absolute configuration of an enantiomer:
1. X-ray diffraction analysis. Note that there is no correlation between the sign of rotation and the structure of a particular enantiomer.
2. Chemical correlation with a molecule whose structure has already been determined via X-ray diffraction.
However, for non-laboratory purposes, it is beneficial to focus on the R/S system. The sign of optical rotation, although different for
the two enantiomers of a chiral molecule,at the same temperature, cannot be used to establish the absolute configuration of an
enantiomer; this is because the sign of optical rotation for a particular enantiomer may change when the temperature changes.
Consider the first picture: a curved arrow is drawn from the highest priority (1) substituent to the lowest priority (4) substituent. If the
arrow points in a counterclockwise direction (left when leaving the 12 o' clock position), the configuration at stereocenter is considered
S ("Sinister" → Latin= "left"). If, however, the arrow points clockwise,(Right when leaving the 12 o' clock position) then the
stereocenter is labeled R ("Rectus" → Latin= "right"). The R or S is then added as a prefix, in parenthesis, to the name of the
enantiomer of interest.
EXAMPLE 5.5.1
(R)-2-Bromobutane
RULE 1
First, examine at the atoms directly attached to the stereocenter of the compound. A substituent with a higher atomic number takes
precedence over a substituent with a lower atomic number. Hydrogen is the lowest possible priority substituent, because it has the
lowest atomic number.
1. When dealing with isotopes, the atom with the higher atomic mass receives higher priority.
2. When visualizing the molecule, the lowest priority substituent should always point away from the viewer (a dashed line indicates
this). To understand how this works or looks, imagine that a clock and a pole. Attach the pole to the back of the clock, so that when
when looking at the face of the clock the pole points away from the viewer in the same way the lowest priority substituent should
point away.
3. Then, draw an arrow from the highest priority atom to the 2nd highest priority atom to the 3rd highest priority atom. Because the
4th highest priority atom is placed in the back, the arrow should appear like it is going across the face of a clock. If it is going
clockwise, then it is an R-enantiomer; If it is going counterclockwise, it is an S-enantiomer.
When looking at a problem with wedges and dashes, if the lowest priority atom is not on the dashed line pointing away, the molecule
must be rotated.
Remember that
Wedges indicate coming towards the viewer.
Dashes indicate pointing away from the viewer.
RULE 2
If there are two substituents with equal rank, proceed along the two substituent chains until there is a point of difference. First,
determine which of the chains has the first connection to an atom with the highest priority (the highest atomic number). That chain has
the higher priority.
If the chains are similar, proceed down the chain, until a point of difference.
For example: an ethyl substituent takes priority over a methyl substituent. At the connectivity of the stereocenter, both have a carbon
atom, which are equal in rank. Going down the chains, a methyl has only has hydrogen atoms attached to it, whereas the ethyl has
another carbon atom. The carbon atom on the ethyl is the first point of difference and has a higher atomic number than hydrogen;
therefore the ethyl takes priority over the methyl.
RULE 3
If a chain is connected to the same kind of atom twice or three times, check to see if the atom it is connected to has a greater atomic
number than any of the atoms that the competing chain is connected to.
If none of the atoms connected to the competing chain(s) at the same point has a greater atomic number: the chain bonded to the
same atom multiple times has the greater priority
If however, one of the atoms connected to the competing chain has a higher atomic number: that chain has the higher priority.
EXAMPLE 5.5.2
A 1-methylethyl substituent takes precedence over an ethyl substituent. Connected to the first carbon atom, ethyl only has one other
carbon, whereas the 1-methylethyl has two carbon atoms attached to the first; this is the first point of difference. Therefore, 1-
However:
Remember that being double or triple bonded to an atom means that the atom is connected to the same atom twice. In such a case,
follow the same method as above.
Caution!!
Keep in mind that priority is determined by the first point of difference along the two similar substituent chains. After the first point
of difference, the rest of the chain is irrelevant.
When looking for the first point of difference on similar substituent chains, one may encounter branching. If there is branching,
choose the branch that is higher in priority. If the two substituents have similar branches, rank the elements within the branches
until a point of difference.
USE YOUR MODELING KIT: Models assist in visualizing the structure. When using a model, make sure the lowest priority is
pointing away from you. Then determine the direction from the highest priority substituent to the lowest: clockwise (R) or
counterclockwise (S).
IF YOU DO NOT HAVE A MODELING KIT: remember that the dashes mean the bond is going into the screen and the wedges
means that bond is coming out of the screen. If the lowest priority bond is not pointing to the back, mentally rotate it so that it is.
However, it is very useful when learning organic chemistry to use models.
PROBLEMS
Are the following R or S?
SOLUTIONS
1. S: I > Br > F > H. The lowest priority substituent, H, is already going towards the back. It turns left going from I to Br to F, so it's a
S.
2. R: Br > Cl > CH3 > H. You have to switch the H and Br in order to place the H, the lowest priority, in the back. Then, going from
Br to Cl, CH3 is turning to the right, giving you a R.
3. Neither R or S: This molecule is achiral. Only chiral molecules can be named R or S.
4. R: OH > CN > CH2NH2 > H. The H, the lowest priority, has to be switched to the back. Then, going from OH to CN to CH2NH2,
you are turning right, giving you a R.
5. (5) S: −COOH > −CH OH > C≡CH > H. Then, going from −COOH to −CH OH to −C≡CH you are turning left, giving
2 2
you a S configuration.
OUTSIDE LINKS
http://www.youtube.com/watch?v=EphUiPiQiCo
http://en.wikipedia.org/wiki/Absolute_configuration
EXERCISES
QUESTIONS
Q5.5.1
Orient the following so that the least priority (4) atom is paced behind, then assign stereochemistry (R or S).
Q5.5.2
Draw (R)-2-bromobutan-2-ol.
Q5.5.3
Assign R/S to the following molecule.
SOLUTIONS
S5.5.1
A = S; B = R
S5.5.2
S5.5.3
The stereo center is R.
CONTRIBUTORS
Ekta Patel (UCD), Ifemayowa Aworanti (University of Maryland Baltimore County)
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
Prof. Steven Farmer (Sonoma State University)
KEY TERMS
Make certain that you can define, and use in context, the key term below.
diastereomer
Diastereomers are stereoisomers that are not related as object and mirror image and are not enantiomers. Unlike enatiomers which are
mirror images of each other and non-sumperimposable, diastereomers are not mirror images of each other and non-
superimposable. Diastereomers can have different physical properties and reactivity. They have different melting points and boiling
points and different densities. They have two or more stereocenters.
INTRODUCTION
It is easy to mistake between diasteromers and enantiomers. For example, we have four steroisomers of 3-bromo-2-butanol. The four
possible combination are SS, RR, SR and RS (Figure 5.6.1). One of the molecule is the enantiomer of its mirror image molecule and
diasteromer of each of the other two molecule (SS is enantiomer of RR and diasteromer of RS and SR). SS's mirror image is RR and
they are not superimposable, so they are enantiomers. RS and SR are not mirror image of SS and are not superimposable to each other,
so they are diasteromers.
Figure 5.6.1
Figure 5.6.2
We turn our attention next to molecules which have more than one stereocenter. We will start with a common four-carbon sugar called
D-erythrose.
A note on sugar nomenclature: biochemists use a special system to refer to the stereochemistry of sugar molecules, employing names
of historical origin in addition to the designators 'D' and 'L'. You will learn about this system if you take a biochemistry class. We will
use the D/L designations here to refer to different sugars, but we won't worry about learning the system.
As you can see, D-erythrose is a chiral molecule: C2 and C3 are stereocenters, both of which have the R configuration. In addition, you
should make a model to convince yourself that it is impossible to find a plane of symmetry through the molecule, regardless of the
conformation. Does D-erythrose have an enantiomer? Of course it does – if it is a chiral molecule, it must. The enantiomer of erythrose
is its mirror image, and is named L-erythrose (once again, you should use models to convince yourself that these mirror images of
erythrose are not superimposable).
Notice that both chiral centers in L-erythrose both have the S configuration.
NOTE
In a pair of enantiomers, all of the chiral centers are of the opposite configuration.
What happens if we draw a stereoisomer of erythrose in which the configuration is S at C2 and R at C3? This stereoisomer, which is a
sugar called D-threose, is not a mirror image of erythrose. D-threose is a diastereomer of both D-erythrose and L-erythrose.
The definition of diastereomers is simple: if two molecules are stereoisomers (same molecular formula, same connectivity, different
arrangement of atoms in space) but are not enantiomers, then they are diastereomers by default. In practical terms, this means that at
In L-glucose, all of the stereocenters are inverted relative to D-glucose. That leaves 14 diastereomers of D-glucose: these are molecules
in which at least one, but not all, of the stereocenters are inverted relative to D-glucose. One of these 14 diastereomers, a sugar called
D-galactose, is shown above: in D-galactose, one of four stereocenters is inverted relative to D-glucose. Diastereomers which differ in
only one stereocenter (out of two or more) are called epimers. D-glucose and D-galactose can therefore be refered to as epimers as
well as diastereomers.
EXAMPLE 5.6.1
Draw the structure of L-galactose, the enantiomer of D-galactose.
Draw the structure of two more diastereomers of D-glucose. One should be an epimer.
Answer:
Erythronolide B, a precursor to the 'macrocyclic' antibiotic erythromycin, has 10 stereocenters. It’s enantiomer is that molecule in
which all 10 stereocenters are inverted.
In total, there are 210 = 1024 stereoisomers in the erythronolide B family: 1022 of these are diastereomers of the structure above, one is
the enantiomer of the structure above, and the last is the structure above.
We know that enantiomers have identical physical properties and equal but opposite degrees of specific rotation. Diastereomers, in
theory at least, have different physical properties – we stipulate ‘in theory’ because sometimes the physical properties of two or more
diastereomers are so similar that it is very difficult to separate them. In addition, the specific rotations of diastereomers are unrelated –
they could be the same sign or opposite signs, and similar in magnitude or very dissimilar.
EXERCISES
QUESTIONS
Q5.6.1
SOLUTIONS
S5.6.1
CONTRIBUTORS
Organic Chemistry With a Biological Emphasis by Tim Soderberg (University of Minnesota, Morris)
Dr. Dietmar Kennepohl FCIC (Professor of Chemistry, Athabasca University)
KEY TERMS
Make certain that you can define, and use in context, the key term below.
meso compound
STUDY NOTES
You may be confused by the two sets of structures showing “rotations.” Of course in each case the two structures shown are
identical, they represent the same molecule looked at from two different perspectives. In the first case, there is a 120° rotation
around the single carbon-carbon bond. In the second, the whole molecule is rotated 180° top to bottom.
Meso Compounds
A meso compound is an achiral compound that has chiral centers. It is superimposed on its mirror image and is optically inactive
although it contains two or more stereocenters.
INTRODUCTION
In general, a meso compound should contain two or more identical substituted stereocenters. Also, it has an internal symmetry plane
that divides the compound in half. These two halves reflect each other by the internal mirror. The stereochemistry of stereocenters
should "cancel out". What it means here is that when we have an internal plane that splits the compound into two symmetrical sides,
the stereochemistry of both left and right side should be opposite to each other, and therefore, result in optically inactive. Cyclic
compounds may also be meso.
IDENTIFICATION
If A is a meso compound, it should have two or more stereocenters, an internal plane, and the stereochemistry should be R and S.
1. Look for an internal plane, or internal mirror, that lies in between the compound.
2. The stereochemistry (e.g. R or S) is very crucial in determining whether it is a meso compound or not. As mentioned above, a meso
compound is optically inactive, so their stereochemistry should cancel out. For instance, R cancels S out in a meso compound with
two stereocenters.
trans-1,2-dichloro-1,2-ethanediol