0% found this document useful (0 votes)
17 views14 pages

Chapter 1

The document provides a comprehensive overview of the basics of organic chemistry, including the structure and significance of atomic orbitals, the principles of chemical bonding, and the Octet Rule. It explains the differences between ionic and covalent bonds, the concept of electronegativity, and the VSEPR theory for predicting molecular geometry. Key concepts such as core and valence orbitals, sigma and pi bonds, and polar covalent bonds are also discussed to illustrate their roles in chemical reactivity and bonding behavior.

Uploaded by

Yagnika Variya
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
17 views14 pages

Chapter 1

The document provides a comprehensive overview of the basics of organic chemistry, including the structure and significance of atomic orbitals, the principles of chemical bonding, and the Octet Rule. It explains the differences between ionic and covalent bonds, the concept of electronegativity, and the VSEPR theory for predicting molecular geometry. Key concepts such as core and valence orbitals, sigma and pi bonds, and polar covalent bonds are also discussed to illustrate their roles in chemical reactivity and bonding behavior.

Uploaded by

Yagnika Variya
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 14

1.

1: Brief Refresher of the Basics


The Periodic Table represents one way of organizing the known elements based
on their properties. While a full Periodic Table contains many useful pieces of
information most of it is not relevant to a discussion of the basics of Organic
Chemistry. An abridged Periodic Table (Figure 1.1) with the atomic symbols, element
numbers, and Group numbers (column numbers) is suf cient in the rst half of this
text. Notice that, depending on which Periodic Table you use, the Group numbers may
be different. For all references to Group numbers in this text, please use these values.
Recall that electrons occupy regions of space called orbitals. Atomic orbitals have
prede ned shapes that can be determined from their corresponding quantum
numbers and the use of a wavefunction. These atomic orbitals represent the area
around the nucleus where the electrons in it are most likely to be (a probability
distribution).
The simplest atomic orbitals are s orbitals, which look like spheres (Figure 1.2). For
our purposes the presence of nodes in s orbitals can be ignored.
The next atomic orbitals are p orbitals, which are often compared to dumbbells
(Figure 1.3). There is a nodal plane that splits the orbitals into two differently phased
lobes on either side of the nucleus. “Phase” represents the sign of the function for that
area (+ or -) and is not physically meaningful, nor does it have any relationship to
charge. It does however play a crucial role in bonding.
As a result of the quantum numbers used to derive them, there are 3 equivalent p
orbitals for each level: px, py, and pz. These three orbitals are orthogonal
(perpendicular with respect to each other) and most easily viewed as aligned with the
coordinate axes.
Other types of atomic orbitals (d, f, g, h, i) can be de ned, but are not normally
considered/occupied in simple organic systems
As a result of their spin quantum number and the Pauli Exclusion Principle (no two
electrons can have the same four quantum numbers) each orbital may hold a
maximum of two spin-paired electrons. This means an s orbital may hold up to two
electrons, and each of the three p orbitals (px, py, and pz) may also hold up to two
electrons.
In any given system most things like to rest in the lowest possible energy state.
The Aufbau Principle describes the order in which electrons ll orbitals in order of
fi
fi
fi
fi
fi
increasing energy. An arrowed diagram (Figure 1.4) is the most common way to
depict the order the ascending energy levels take.
Based on the order of ascending orbital energies (Figure 1.4) electrons rst ll the
lowest energy orbital (1s), then the next lowest (2s), then the next lowest (2p), and so
on (Figure 1.5). It is important to remember that atomic orbitals of the same type also
have the same energy. For instance, the 2p orbitals are the 2px, 2py, and 2pz orbitals,
all of which have exactly the same energy level. A set of orbitals with exactly the same
energies is sometimes referred to as being degenerate.
Orbitals in systems like atoms or molecules can be broadly divided into two
categories. The fully occupied orbitals of the lowest energies are referred to as core
orbitals, and the electrons in them as core electrons. These orbitals and electrons are
low in energy (stable) and close to the nucleus. As a result, they are not normally
considered important for bonding or chemical reactivity and are typically ignored.
Conversely, the highest energy orbitals (vacant, semi-occupied, or fully occupied) are
referred to as valence orbitals. Any electrons in these orbitals are called valence
electrons. These orbitals and electrons are higher in energy (less stable) and farther
from the nucleus. As a result, they are heavily involved in bonding and chemical
reactivity.
Core orbitals: fully occupied orbitals of lowest energy
Core electrons: electrons found in core orbitals
(NOT available for bonding / as lone pairs)
(We can ignore them)
Valence orbitals: occupied orbitals of highest energy(and any accompanying
empty orbitals of similar energy)
Valence electrons: electrons found in valence orbitals
(Available for bonding / as lone pairs)
Consider the ground-state electron con guration for carbon: 1s22s22p2 (Figure
1.6). The 1s orbital would be considered a core orbital, and the two electrons in it
core electrons. The 2s, 2px, 2py, and 2pz orbitals would be considered valence
orbitals and the four electrons in them valence electrons. Notice that carbon has 4
valence electrons and is in Group 4 in the Periodic Table. This pattern holds for all
atoms in the rst few rows of the table; the number of valence electrons for an atom is
equal to its Group number in the Periodic Table.
fi
fi
fi
fi
There are two primary forms of chemical bonding between atoms.
Ionic bonds typically occur when a valence electron is fully transferred from one
atom’s valence orbital to another’s. The loss of an electron makes that atom cationic
(positively charged). The gain of an electron makes the other atom anionic (negatively
charged). The two charged species are then attracted towards each other, forming a
tight bond. These types of compounds are commonly referred to as salts
Covalent bonds occur when valence electrons are shared between atoms. Two
semi-occupied valence orbitals overlap and combine to form a new fully occupied
molecular orbital (Figure 1.8). The bond can be thought of as an area between the
two nuclei where the two electrons are being shared between them.
1.2: The Octet Rule
The atoms that participate in covalent bonding share electrons in a way that
enables them to acquire a stable electronic con guration, or full valence shell.
Consider the ground-state electron con guration of carbon (Figure 1.5 above).
Carbon has 4 valence electrons already and needs another 4 electrons to ll its
valence orbitals, for a total of 8 valence electrons. This holds for the other elements in
that row, such as nitrogen and oxygen, as well as the halogens (F, Cl, Br, and I). We call
this the Octet Rule.
Octet Rule: Atoms in the rst row (C, N, O) and the halogens (F, Cl, Br, I) look to ll
their valence shells with EIGHT electrons – NEVER more, VERY RARELY less.
It is very important to remember that one commonly encountered element does
not follow this pattern: hydrogen. Hydrogen only has one valence orbital (the 1s
orbital) and as a result looks to ll its valence shell with TWO electro
ns, not eight.
Each covalent bond an atom makes with another atom functionally allows it to
gain an electron in its valence orbitals (this works for the other atom as well). As a
result, the number of bonds that an atom normally makes is directly equal to the
number of additional valence electrons it needs to ll its valence shell. For example,
silicon (Si) is in Group 4 like carbon and usually makes 4 bonds. Phosphorous (P) is in
Group 5 like nitrogen and often makes 3 bonds, though sometimes it makes more
(e.g. phosphoric acid, H3PO4). Sulfur (S) is in Group 6 like oxygen and often makes 2
bonds, though sometimes it makes more (e.g. sulfuric acid, H2SO4). Why elements
outside of the rst row do not always follow the Octet Rule is beyond the scope of this
fi
fi
fi
fi
fi
fi
fi
fi
text, but involves the smaller energy differences between available orbitals once they
are suf ciently high in energy.
1.3 Basics of Bonding
Covalent bonds are the result of constructively overlapping orbitals. Covalent
bonds may be subdivided into two main types depending on how the orbitals
involved overlap: sigma (σ) and pi (π) bonds.
Sigma (σ) bonds result from direct (head-on) overlap between two orbitals. The
simplest case to consider is the hydrogen molecule, H2. When we say that the two
hydrogen nuclei share their electrons to form a covalent bond, what we mean is that
the two spherical 1s orbitals overlap and contain two electrons with opposite spin
These two electrons are now attracted to the positive charge of both of the
hydrogen nuclei, with the result that they serve as a sort of ‘chemical glue’ holding the
two nuclei together.
The combination of an s orbital with another s orbital described above is the
simplest to visualize (Figure 1.10). The two orbitals directly overlap with each other,
forming an area between the two nuclei where the electrons are shared.However, a σ
bond may arise from any direct overlap between two properly aligned orbitals of the
same phase. The combination of an s orbital with an aligned p orbital (Figure 1.11a) is
slightly more complicated but follows the same pattern.
The idea of constructive overlap is important here. That is, the phase of the two
orbitals being combined must be the same to form a covalent bond. In this case phase
is being represented by colour, with the ‘blue’ s orbital constructively overlapping
with the ‘blue’ portion of the p orbital.
Imagine the p orbital above were rotated 90° clockwise on the page (equivalent to
using a pz instead of a px orbital; Figure 1.11b). Now, as the two orbitals come
together any constructive overlap between ‘blue’ portions at the top would be
cancelled out by a deconstructive overlap between the ‘blue’ portion of the s orbital
and the ‘orange’ portion at the bottom of the p orbital. This means a covalent bond
would not be possible with this alignment of orbitals, with every constructive overlap
matched by an equal deconstructive one.
Imagine the p orbital above were instead rotated 180° clockwise on the page
(equivalent to switching the ‘blue’ and ‘orange’ portions; Figure 1.11c). Now, as the
two orbitals come together there is instead a deconstructive overlap between the
fi
‘blue’ portion of the s orbital and the ‘orange’ portion of the p orbital. Instead of a
stabilizing bond being formed, a destabilizing anti-bonding orbital is generated (σ*).
The involvement of anti-bonding orbitals in chemical bonds is complex and well
beyond the scope of this text; at an introductory level assume anti-bonding orbitals
are not important for a discussion of bonding, but play a crucial role in many kinds of
chemical reactions.
To avoid confusion later it is important to clarify that σ bonds do not require an s
orbital in their formation. For example, two properly aligned p orbitals may overlap to
form a σ bond (Figure 1.12). In other words, the only important requirement is that the
two portions overlapping combine constructively and in a direct overlap. This speci c
combination (two p orbitals forming a sigma bond) is uncommon in organic
chemistry but plays important roles in inorganic chemistry.
Pi (π) bonds result from indirect (side-on) overlap between two orbitals. Unlike σ
bonds, the two orbitals are not able to directly overlap, instead simply coming closer
together and slightly ‘bending’ to constructively overlap.
In simple organic systems all π bonds result from the indirect overlap of two p
orbitals (Figure 1.13). As a result there are two areas, one above and one below the
nuclei, which form the bond. The two lobes together make one π bond.
The simplest case to consider is ethylene, which contains a carbon double
bonded to another carbon (Figure 1.14). The rst bond between them is a σ bond,
with one of the lines between them representing this bond. The second bond
between them is a π bond formed by the combination of a p orbital from each of the
carbons. The other line represents this bond. Notice that even though there are three
areas of electron density (lobes), there are only two bonds because the π bond is
made from overlap of both lobes on the p orbitals.
In more complex systems, including those involving heavy metal elements, more
complex orbitals (commonly d orbitals) may also be involved. These do not play a
role in any systems encountered in this text.
In covalent bonds electrons are shared between the two atoms involved. When
the two atoms involved are the same element (and barring other factors) the
electrons are shared equally. However, not all elements have the same af nity for
electron density around them. We can quantify a measurement called
electronegativity, which is de ned as the relative ability of an atom to attract electrons
fi
fi
fi
fi
to itself in a chemical compound. Electronegativity of an atom is not a simple, xed
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 in uenced 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. The most common values used are those in the Pauling scale
When two atoms with different electronegativities are covalently bonded together
an unequal sharing of electrons occurs. This is referred to as a polar covalent bond.
The atom that takes extra electron density (the more electronegative atom) is denoted
with a symbol, δ–, while the atom that gains less electron density is denoted with the
symbol δ+. These symbols emphasize that there are partial negative and positive
charges on those atoms. Polar covalent bonds can be thought of as being similar to
ionic bonds, but where the electrons are still ‘shared’ between the two atoms
In general, an electronegativity difference of >0.4 is required for us to consider a
bond polar. For instance, a carbon-hydrogen bond is not polar (0.4 difference), but a
carbon-nitrogen bond is (0.5 difference) and a carbon- uorine bond (1.5 difference)
is even more so
It is occasionally useful to denote the direction that a bond or molecule is polar in.
Convention uses a crossed arrow pointing in the direction of the more negative
portion. Be aware that in other areas of chemistry, particularly inorganic chemistry, this
symbol may be used to mean other things. The arrow convention is more useful when
describing the overall direction of polarity in a molecule (a sum of vectors; a net
dipole), such as when there are multiple polar bonds oriented in different directions.
Determining the net dipole (sometimes called dipole moment) from a molecular
structure can be complicated when there is more than one type of polar bond (e.g. a
C-Cl bond and a C-O bond) because they do not affect the net polarity to the same
degree. At an introductory level these differences are typically ignored and complex
examples are avoided. Each polar bond is considered an equivalent ‘pull’ in the
direction of the more electronegative atom and the overall direction being pulled
towards is determined. When present three-dimensional structure (see Section 1.4)
must be considered to accurately determine the overall direction of the net dipole.
With practice this process can be streamlined and done internally.
fl
fl
fi
1.4 VSEPR Theory
Valence Shell Electron Pair Repulsion (VSEPR) theory states that the most stable
three-dimensional structure for a molecule is the one in which valence electron pairs
are as far apart as possible. This is intuitive; electrons are small negative charges, and
negative charges repel rather than attract each other. In practical terms VSEPR theory
amounts to keeping groups (attached atoms or lone pairs) as far apart from each
other as possible.
If there are two groups then maximizing the distance between them makes a
Linear geometry (Figure 1.18). The two groups are on opposite sides of the central
point (atom) to keep as far from each other as possible. The typical angle between
them is thus 180°
If there are three groups then maximizing the distance between them makes a
Trigonal Planar geometry (Figure 1.19). The three groups are arranged in a triangle
around the central point (atom) to keep as far from each other as possible. The typical
angle between them is thus around 120° (this varies slightly depending on what the
groups are, but the variation is not normally important).
If there are four groups then maximizing the distance between them makes a
Tetrahedral geometry (Figure 1.20). This is the rst truly three-dimensional geometry.
If the four groups were instead arranged two-dimensionally (i.e. as a square around
the central point) then the angle between them would be approximately 90°. By
instead arranging the four groups three-dimensionally as a tetrahedron the angle,
and by extension the distance, between them increases to around 109°. This again
varies slightly depending on the exact groups, but the idea of maximal distance is
upheld. For those unfamiliar with the shape, a tetrahedron is equivalent to a four-
sided pyramid where each of the three sides and the bottom are equilateral triangles.
Additional complex geometries are possible and may be encountered in other
subjects, such as inorganic chemistry, but are not relevant in simple organic systems.
Some older textbooks denote a difference between lone pairs and atoms when
talking about ‘groups’ to describe geometry (Figure 1.21). These are usually referred
to as electron and molecular geometries.So-called molecular geometries have no
practical (or theoretical) uses and may be ignored. In the examples above the correct
description for the geometry of all four nitrogen atoms would simply be tetrahedral.
1.5: Hybrid Orbitals
fi
Covalent bonds involve the sharing of spin-paired electrons. This creates localized
bonds, con ned to the region between the bonded atoms, from an overlap of
orbitals. The better the overlap between the orbitals, the stronger the bond (e.g. σ vs.
π bonds discussed above).
This last part creates a problem for organic chemists. Most organic molecules are
primarily made from hydrogen, carbon, oxygen, and nitrogen. The valence orbitals
involved in bonding for these elements are s and p orbitals
s Orbitals have a spherical shape and can be easily used to rationalize bonding
from any direction. However, p orbitals exist as px, py, and pz and are orthogonal
(90°) from each other. Based on VSEPR theory, typical bond angles should be ~109°
(tetrahedral geometries), ~120° (trigonal planar geometries), and 180° (linear
geometries). We are thus left with a question of how to reconcile the strong σ bonds
observed in these molecules with the apparent poor overlap of p orbitals at these
geometries.
Orbital hybridization is one approach used to explain this discrepancy. Hybrid
orbitals are a mathematical construct (a model) arising from the combination of 2s
and 2p orbitals. Orbital hybridization is not a physical process. However, the model is
much simpler to understand and discuss than the more rigorous (and accurate)
approaches developed since.
The exact math behind how individual hybrid orbitals may be generated from
atomic orbitals is interesting but not important for understanding. However, two
important consequences of the approach should be noted.
First, the number of hybrid orbitals generated is exactly equal to the number of
atomic orbitals used to make them. This means that in all cases we will start with four
valence orbitals (2s, 2px, 2py, and 2pz) and end with four valence orbitals (the new
hybrid orbitals and any unused atomic orbitals). The new hybrid orbitals are used in
forming σ bonds or contain lone pairs. Any p orbitals that remain are used in forming
π bonds; p orbitals are not involved in σ bonds in hybridization models.
Second, in all cases the hybrid orbitals arise from the combination of an s orbital
with one or more p orbitals. As a result, all of the hybrid orbitals look the same (Figure
1.23). In each case they resemble a ‘biased’ p orbital where one lobe is larger than
before and the other is smaller than before. This is because one lobe combined
constructively with the s orbital (both were the same phase) and one lobe combined
fi
deconstructively with the s orbital (the two were opposite phases). The only difference
between hybrid orbitals is the directionality of the lobes
If we combine the 2s orbital with the 2px, 2py, and 2pz orbitals we get four sp3
orbitals (Figure 1.24). The four valence orbitals combine mathematically to form four
equivalent hybrid orbitals, which are named sp3 orbitals because they are formed
from mixing one s and three p orbitals.
As a result of how the orbitals were combined, they each point in different
directions. The angle between each orbital is now ~109°, which allows construction of
tetrahedral geometries
We can describe σ bonds made from the atom as arising from the overlap of one
of its sp3 orbitals with an orbital from the adjacent atom. For example, the C-H σ
bonds in methanol form from the overlap of a sp3 orbital on carbon with the s orbital
on hydrogen (Figure 1.26). Conversely, the C-O σ bond forms from the overlap of an
sp3 orbital from carbon with an sp3 orbital from oxygen (the oxygen has four groups
around it, and so is also sp3 hybridized). Hybrid orbitals are not always involved in
bonding and can instead localize lone pairs.
If we combine the 2s orbital with the 2px and 2py orbitals (and leave the 2pz
orbital untouched) we get three sp2 orbitals (Figure 1.27). The three valence orbitals
combine mathematically to form three equivalent hybrid orbitals, which are named
sp2 orbitals because they are formed from mixing one s and two p orbitals. Note that
the new hybrid orbitals exist only in the x/y plane.
As a result of how the orbitals were combined, they each point in different
directions. The angle between each orbital is now ~120°, which allows construction of
trigonal planar geometries (Figure 1.28). The remaining 2pz orbital is perpendicular
to the plane occupied by the sp2 orbitals.
We can describe σ bonds made from the atom as arising from the overlap of one
of its sp2 orbitals with an orbital from the adjacent atom. For example, the C-H σ
bonds in formaldehyde form from the overlap of a sp2 orbital on carbon with the s
orbital on hydrogen (Figure 1.29). Conversely, the C-O σ bond forms from the overlap
of a sp2 orbital from carbon with a sp2 orbital from oxygen (the oxygen has three
groups around it, and so is also sp2 hybridized). Hybrid orbitals are not always
involved in bonding and can instead localize lone pairs. Finally, the C-O π bond forms
from the overlap of the untouched 2pz orbital on carbon with the corresponding 2pz
orbital on oxygen.
If we combine the 2s orbital with the 2px orbital (and leave the 2py and 2pz
orbitals untouched) we get two sp orbitals (Figure 1.30). The two valence orbitals
combine mathematically to form two equivalent hybrid orbitals, which are named sp
orbitals because they are formed from mixing one s and one p orbitals.
As a result of how the orbitals were combined, they point in different directions.
The angle between the orbitals is now ~180°, which allows construction of linear
geometries (Figure 1.31). The remaining 2py and 2pz orbitals are perpendicular to
the axis occupied by the sp orbitals and to each other.
We can describe σ bonds made from the atom as arising from the overlap of one
of its sp orbitals with an orbital from the adjacent atom. For example, the C-C σ bond
in acetonitrile forms from the overlap of a sp orbital on the central carbon with the
sp3 orbital on the other carbon (the other carbon has four groups around it, and so is
sp3 hybridized; Figure 1.32). Conversely, the C-N σ bond forms from the overlap of a
sp orbital from the central carbon with a sp orbital from the nitrogen (the nitrogen has
two groups around it, and so is also sp hybridized). Hybrid orbitals are not always
involved in bonding and can instead localize lone pairs. Finally, the C-N π bonds form
from the overlap of the untouched p orbitals on carbon and nitrogen: the 2py orbitals
on each combine to make one π bond, and the 2pz orbitals on each combine to make
the other π bond. Visualizing the three-dimensional orientation of localized bonds
using two dimensional graphics can be challenging. It is important to remember that
the two π bonds exist simultaneously in orthogonal planes (i.e. one is in a y plane and
one is in a z plane).
It is important to remember that orbital hybridization is a model used to rationalize
the geometries we observe. For example, saying that a carbon atom is sp3 hybridized
is a synonym for saying the geometry around that carbon is tetrahedral. There is thus a
simple trick to quickly determine the hybridization state of any atom: count the
number of connected groups. In this case ‘groups’ refers to any connected atom,
regardless of how many bonds connect the two, and any lone pairs. The number of
groups around it is equal to the sum of exponents in the hybridization name
1.6: Lewis Structure
There are a variety of ways to depict chemical structures. The classical approach is
through Lewis Structures.
A Lewis Structure displays the atoms of the molecule in the order they are
bonded. It also depicts how the atoms are bonded to one another, for example as
single, double, and triple covalent bonds. Covalent bonds are shown using lines. The
number of lines indicates whether the connection is a single, double, or triple
covalent bond. All atom labels are shown and all lone pairs are shown.
Notably, Lewis Structures do not attempt to show geometry; all angles are
normally shown as 90° and no three-dimensional features are indicated.
Some atoms carry formal charges (they are cations or anions). This typically
happens when the atom has either an incomplete octet or a different number of
bonds than usual (see Section 1.2). As such, formal charge may be calculated using
the following formula:
Formal Charge = (# of valence electrons) – (# of bonds) – (# of non-bonded
electrons)
The formal charge is then indicated by placing a circled positive or negative
symbol adjacent to the atom with the formal charge
Drawing Lewis Structures from molecular formulae can be complicated by the fact
that there is often more than one structure (molecule) possible. It is much easier to
convert so-called condensed formulae, which show the order of atoms like a Lewis
Structure but are written in a single line to save space. The order of the atoms
suggests the connectivity in the molecule. This provides hints for how to construct the
Lewis Structure of the speci c molecule, instead of a related compound
Example: CH3CH2CONH2
• Using the condensed formula as a hint, arrange the atoms.
• Add valence electrons around the atoms. If there is a formal charge, add an
electron for anionic or remove an electron for cationic
• Form all obvious σ bonds by combining pairs of electrons from adjacent
elements. Follow the Octet Rule
• Form all obvious π bonds by combining pairs of electrons from adjacent
elements. Follow the Octet Rule
• Con rm that each element in the Lewis Structure follows the Octet Rule. Add
formal charges and create additional bonds as needed
fi
fi
• The last step is crucial in ensuring success. Some attempts may generate
structures that are not easily resolved
In these cases the structure will never be able to satisfy the Octet Rule for all atoms
and contain the proper number of formal charges, so it may be ruled out as a
candidate structure.
1.7: Hashed and Wedged Notations
More advanced ways of depicting molecules (see Section 1.8) attempt to show
three-dimensional structure in their construction. There are a variety of reasons this is
useful, but the primary reason involves the accurate portrayal of stereoisomers (see
Chapter 4).
The most common notation for showing three-dimensional orientation is the
hashed/wedged system. In this approach simple lines are used to represent bonds
that occupy the plane of the page (screen, board, etc.). Hashed bonds are drawn to
indicate that the attached atom is oriented away from the viewer (i.e. behind the
plane of the page). Wedged bonds are drawn to indicate that the attached atom is
oriented towards the viewer (i.e. in front of the plane of the page). This allows us to
show three-dimensional geometries, such as the tetrahedral geometry of methane
To be drawn properly the hashed/wedged bonds should start narrow at the atom
that lies in the plane of the page and widen as they approach the atom behind/in
front of the plane. Some older depictions lack this (Figure 1.37). This is ne for simple
molecules but can be confusing or ambiguous in complicated molecules. As such, it
should be avoided.
To be drawn properly the hashed/wedged bonds should both be on the same
side and should be made to resemble a tetrahedron
The left-to-right order of hashed vs. wedged bonds is not important; the hashed/
wedged system is an imperfect way of indicating the two groups are directly in front
of each other from the viewer’s perspective
1.8 Other Representation of Molecules
A condensed formula or structure is made up of the elemental symbols.
Condensed structural formulas show the order of atoms like a Lewis Structure but are
written in a single line to save space and make it more convenient and faster to write
out. The order of the atoms suggests the connectivity in the molecule. Condensed
structural formulas are also helpful when showing that a group of atoms is connected
fi
to a single atom in a molecule. When this happens, parentheses are used around the
group of atoms to show they are together. Also, if more than one of the same
substituent is attached to a given atom, it is shown with a subscript number. An
example is CH4, which represents four hydrogens attached to the same carbon.
Condensed formulas can be read from either direction and H3C is the same as CH3,
although the latter is more common (Figure 1.40). Condensed structures work best
for small, simple molecules where there is little ambiguity between possible
connection points
This style is largely a holdover from the typewriter era, when it was much easier to
include a line of text than to add a drawing for every molecule being discussed.
Nonetheless they continue to be used for many simple compounds.
Line-angle formulas are used to write carbon and hydrogen atoms more ef ciently
by replacing the letter “C” with lines. A carbon is present wherever a line intersects
another line and at the ends of lines (Figure 1.41). Hydrogens are omitted but are
assumed to be present to complete each of carbon’s four bonds. Hydrogens that are
attached to elements other than carbon are shown. Atom labels for all other elements
are shown.
Lone pair electrons are sometimes omitted, but they can be included without
issue (Figure 1.42a). Normally the angle between lines is 120° to approximate
tetrahedral and trigonal planar geometry, though angles of 180° are used when linear
geometries are present (Figure 1.42b). If the three-dimensional arrangement is
known (absolute con guration, see Chapter 4) it may be included using hashed/
wedged bonds, though hydrogens attached to carbons are still omitted
There are numerous other ways of depicting molecules in three-dimensions. In
organic chemistry a common approach is to use differently coloured ellipsoids
(spheres) to represent different atoms (Figure 1.43). Bonds are typically represented
using lines or cylinders connecting the spheres. These models are sometimes called
Ball-and-Stick and are meant to approximate seeing a molecular model. This
approach is very useful for showing three-dimensional features. These images are
generated using computer software and are not meant for students to recreate, but it
is important to be able to recognize what is being depicted. Most representations use
the same colours to refer to the same elements. For example, the computer-
generated image coloured the oxygen atom as red, and a standard molecular model
fi
fi
kit will have red balls for oxygen atoms. However, these are only conventions and not
rules; the computer-generated image coloured the carbon atoms grey, but a standard
molecular model kit will have black balls for carbon atoms.

You might also like