You are on page 1of 12

Chem@HC

Secondary 4 Chemistry Olympiad


Name: ______________________ Class: Sec 4_____ Date: ________

Concepts in Covalent Bonding


Formal Charges

In Lewis structures of molecules and polyatomic ions, we can associate with each atom an
apparent charge, under the assumption that the two electrons in each covalent bond are
shared equally among the two bonding atoms. This assigned charge is known as the formal
charge of an atom in a polyatomic species. Before we further explain the concept, we
highlight two key points:

1. We always speak of the formal charge of an atom in a polyatomic species. It does


not make sense to speak of the formal charge of a whole species, or a group of more
than one atom in a species. That is, it is wrong to say “the formal charge of the nitrate
anion”, and “the formal charge of the carboxylate group in the benzoate ion” (a
carboxylate group is a group of three atoms commonly encountered in organic
chemistry). It is only valid to speak of “the formal charge of the central nitrogen atom
in the nitrate ion”, for instance.
2. This assigned charge is not a definitive indicator of the distribution of electron density
in the species. This is because electronegativity differences are ignored in the
assignment of formal charges. Furthermore, the effect of resonance will also weigh
in, which will be discussed in a later section.

The idea of formal charges is perhaps best illustrated with a few examples. Take for instance
the familiar carbon dioxide molecule, CO2 , as shown below.

Suppose we are interested in the formal charge of the central carbon atom. Assume each
covalent bond assigns one electron to each of the bonding atoms. Since carbon has 4 bonds
around it, it gets assigned 4 electrons in the above structure. Since it has no lone pairs
around it, its final valence electron count is 4. We then compare this number to the number
of valence electrons in a lone carbon atom: since a lone carbon atom also has 4 valence
electrons, the numbers match and thus the carbon atom has a formal charge of 0.

If we are instead concerned with a peripheral oxygen atom, we get 2 electrons assigned to it
due to the covalent bonds, and another 4 electrons direct from the lone pairs. This totals to 6
valence electrons, which again matches the valence electron count in a lone oxygen atom.
Therefore, the oxygen atom has a formal charge of 0.

1 Chem@HC
Formal charges are (of course) not always 0. For a more interesting example, let us look at

the ammonium ion, NH4 . Here is one representation of the ammonium ion:

Let us consider the formal charge of the nitrogen centre. Since there are 4 covalent bonds
on the nitrogen atom and no lone pairs, its final valence electron count is 4. However, a lone
nitrogen atom has 5 valence electrons. When the two counts do not match, we always
subtract the electron count in the species from the electron count in the lone atom. In
our case, we take 5  4  1, and conclude that the nitrogen centre has a 1 formal charge.

If we repeat this procedure with each peripheral hydrogen atom, we will conclude that they
all have a formal charge of 0. Indeed, even the one that bears a positive charge also has a
formal charge of 0. With this, we can obtain another way of representing the Lewis structure
of the ammonium ion:

In this representation, the charge on each atom is its formal charge, and the “arrowed” dative
bond is absent. Indeed, one of the biggest advantages of presenting Lewis structures based
on formal charges is doing away with the dative bond notation. The strength of this is
perhaps best appreciated in a lesson on organic reaction mechanisms and “arrow-pushing”.

In summary, here are the steps to assigning formal charges to atoms:

1. Count the number of covalent bonds on the atom.


2. To the count in step 1, add the number of non-bonding electrons on the atom.
3. Subtract the total count in step 2 from the number of valence electrons in the lone
atom.

This formula summarizes the procedure above:

Formal charge  V  B  L,

where V is the number of valence electrons in the lone atom, B is the number of covalent
bonds and L is the number of lone electrons on the atom.

Finally, you should always remember that the sum of the formal charges on each atom in
the Lewis structure must be equal to the overall charge of the entire molecule.

2 Chem@HC
Question 1:

Draw the Lewis structures based on formal charges for the following species:
2
(a) SO 4

(b) BF4
(c) CO

(d) NO 3
(e) O3

(f) Cl O 2

Question 2:

State if the following statements are true or false. Provide a brief explanation or justification
to your answer.

a. Formal charges take the values 1, 0 and 1 only.


b. Negative formal charges are always assigned to the elements with higher
electronegativity.
c. There exists a species in which an atom with a positive formal charge to be directly
bonded to another atom with a negative formal charge.
d. There exists a species in which an atom with a positive formal charge to be directly
bonded to another atom with a positive formal charge.
e. There exists a species in which an atom with a negative formal charge to be directly
bonded to another atom with a negative formal charge.
f. Isoelectronic species have the same formal charges on their corresponding atoms.

Resonance
2
Consider the carbonate ion, CO3 , whose Lewis structure is shown below (two of the oxygen
atoms are deliberately coloured):

We notice that there are:

 two oxygen atoms with a 1 formal charge singly bonded to the carbon centre, and
 one oxygen atom with a 0 formal charge doubly bonded to the carbon centre.

The choice of which oxygen atom is doubly bonded is in some sense arbitrary. For instance,
the following are also valid Lewis structures for the ion:

3 Chem@HC
The above are not rotations of the first structure by 120; the oxygen atoms did not move
and there is only a rearrangement of the electrons in the Lewis structure.

Which of these three Lewis structures correctly depicts the electron distribution in the
carbonate ion? None! Any one of the three structures above incorrectly suggest that one of
the bonds is different from the other two. However, experiments show that the three bonds
are equivalent, and the length of each bond is intermediate between those of a single bond
and a double bond. The true structure of the ion is a “blend” of all the three feasible Lewis
structures above.

This is called resonance: there is more than one feasible Lewis structure, and the actual
electron distribution is a hybrid of all such structures. Resonance is indicated by a double
headed arrow, as below:

The true structure (i.e. the “blend” of all the possible Lewis structures) is termed the
resonance hybrid. Each individual Lewis structure is called a resonance structure (or a
canonical structure). The resonance hybrid cannot be easily represented in a single
structure, and therefore a diagram, like the one above, is often used to indicate resonance.

However, drawing such diagrams every single time is tedious, especially since too many
species exhibit resonance. Often, only one resonance structure is displayed and the
resonance is implied, especially when it is not an important part of the discussion. For
example, it will be a chore to always draw the following diagram for every benzene ring in
organic chemistry:

As such, resonance often must be detected by the alert reader.

It should also be emphasized that resonance should be visualized as a “blending” of the


individual Lewis structures; the species does not flicker in alternation among all the feasible

4 Chem@HC
resonance structures. A common analogy for this is that of a mule. A mule does not change
back and forth between a horse and a donkey; it is always half-horse and half-donkey.

If you have learnt some quantum mechanics, the resonance hybrid is a superposition of the
individual wavefunctions of the resonance structures. Quantum mechanical computations
will then show that the resonance hybrid is lower in energy than any of its individual
resonance structures. In other words, we have the slogan: resonance leads to stability.

Question 3:

Draw a diagram to indicate resonance in the following species, displaying all major
resonance structures:

(a) O3
2
(b) SO 4

(c) CH3 CO2

(d) OCN

Question 4:

The following two species exhibit resonance, and one resonance structure is given below.
Provide at least two other resonance structures for each species.

H H
H H
N

C C
H C H
O C O
N N

H
H H

5 Chem@HC
Valence Bond Theory

The quantum mechanical model of the atom was a huge breakthrough in chemistry, but we
then need a theory of chemical bonding compatible with it. Two bonding theories prevail, one
of which is the valence bond (VB). The use of Lewis structures is already part of VB theory;
this section and the next describe the orbital picture in the theory.

In VB theory, when two or more atoms form a molecule, their atomic orbitals overlap to form
bonds. There are (at least) three different types of bonds. The first type is the σ bond (read
sigma bond), where two orbitals overlap “head-on:

The orbitals must be oriented correctly for overlap to occur. By convention, the p orbital that
forms σ bonds is the p z orbital. For d orbitals, only the dz orbital forms σ bonds.
2

The second type is the π bond (read pi bond), when two orbitals overlap sideways:

Note that s orbitals always form σ bonds; only p and d orbitals can form π bonds.

6 Chem@HC
The third type is far less common, and is not known to occur in main group species. It is the
δ bond (read delta bond), where two d orbitals overlap “face-on”.

The δ bond can be found when d block elements form quadruple bonds, such as in the
2
[Re2Cl8 ] ion.

In general,

 A single bond is a σ bond.


 A double bond consists of one σ bond and one π bond.
 A triple bond consists of one σ bond and two π bonds.
 A quadruple bond consists of one σ bond, two π bonds and one δ bond.

Question 5:

Count the number of σ and π bonds in the following molecule.

CH2
O
H
CH2
C
NH H2C O
H
C C
H2C C O

H2C O
C

[NOTE: some of the bonds are not explicitly shown.]

Hybridisation

We have discussed how orbitals overlap in VB theory, but atomic orbitals point in directions
seemingly unrelated to the geometry predicted by the Valence Shell Electron Pair
Repulsion (VSEPR) theory. To reconcile this, we introduce a new notion, best illustrated
using a simple example.

Consider the methane molecule, CH4. VSEPR theory predicts the molecule to be tetrahedral,
which is indeed the case. However, the atomic orbitals in carbon do not point in the correct
directions. The three p orbitals are mutually orthogonal, far from the tetrahedral geometry.
To resolve this, VB theory asserts that the valence orbitals (in this case the 2s and 2p
orbitals of carbon) will “mix” and give rise to new orbitals that are degenerate and of the
7 Chem@HC
correct geometry. This process is called hybridisation, and the new orbitals are called
hybrid orbitals. For methane, the 2s and all three 2p orbitals of carbon will hybridise to give
3
four hybrid orbitals. Because of the constituent orbitals, the hybrid orbitals are called sp
orbitals.

A hybrid orbital consists of two lobes of different sizes, and the larger lobe points outwards
for bonding with other atoms. As shown in the diagram that follows, the s and p orbitals
3 3
together form four sp orbitals. All four of the sp orbitals are degenerate (i.e. they have the
same energy) and point in the tetrahedral direction.

s px py pz

3
Combine to generate four sp orbitals

Which are represented as the set

3
Figure 1. Mixing of the s and p orbitals to give sp hybrid orbitals. Figure taken from General Chemistry:
Principles and Modern Applications (10e) by Petrucci et. al.

The electron-in-box representation of the hybridised carbon atom in methane is shown below:

↿ ↿ ↿ ↿
3
sp
Since these orbitals are now all singly occupied, they form σ bonds with the 1s orbital of the
peripheral hydrogen atoms, completing the VB picture of this molecule.
8 Chem@HC
3
In general, we will use sp hybridisation to produce tetrahedral geometries in molecules.
Therefore, if VSEPR theory predicts that a molecule is tetrahedral, VB theory will then assert
3
that the central atom undergoes sp hybridisation. In fact, it includes any molecule with 4
electron pairs (in the VSEPR sense, so multiple bonds count has one electron pair) around
3
the central atom. Therefore, the oxygen atom in water undergoes sp hybridisation. The six
valence electrons get distributed in the hybrid orbitals as such:

↿⇂ ↿⇂ ↿ ↿
3
sp

Since two of the orbitals are now singly occupied, they form σ bonds with the 1s orbital of the
hydrogen atoms, giving the bent geometry as expected. A similar procedure can be applied
to ammonia, and this is left as an exercise.
3
Since sp hybridisation only produces tetrahedral geometries, we must employ other
hybridisation schemes to produce the other shapes that VSEPR theory predicts. One other
2
very important hybridisation scheme is sp hybridisation, which corresponds to trigonal
planar geometry. Let us illustrate this using a representative trigonal planar species, say
2
boron trifluoride, BF3 . In sp hybridisation, one s orbital and two p orbitals mix to give three
degenerate hybrid orbitals pointing in a trigonal planar fashion. These hybrid orbitals are
2
correspondingly named the sp orbitals. There is one valence p orbital that is left
unhybridised.

s px py

2
Combine to generate three sp orbitals
120

Which are
represented
as the set

2
Figure 2. Mixing of the s and p orbitals to give sp hybrid orbitals. Figure taken from General Chemistry:
Principles and Modern Applications (10e) by Petrucci et. al.

In the case of boron trifluoride, this is the electron-in-a-box diagram of the hybridised boron
centre:

↿ ↿ ↿
2
sp p

9 Chem@HC
The hybrid orbitals are of a lower energy than the unhybridised p orbital, because they
possess some s orbital character. Therefore, the three valence electrons of boron will fill the
hybrid orbitals singly, leaving the unhybridised orbital empty. The hybrid orbitals then form σ
bonds with the fluorine orbitals to give the molecule.
2
The unhybridised p orbital is orthogonal to the sp plane, and this has important
consequences in both inorganic and organic chemistry. In the case of boron trifluoride, the
lone pairs on fluorine can form π bonds with this p orbital, giving the boron-fluorine bond
partial double bond character. This can also be illustrated with resonance structures:

As a final example, and to illustrate both multiple bonding and resonance, consider the
carbonate ion. We have encountered this ion previously, and the three major resonance
structures are reproduced here:

The ion is trigonal planar, in line with VSEPR theory. Therefore, we know that the carbon
2
centre undergoes sp hybridisation. The 4 valence electrons are distributed as follows:

↿ ↿ ↿ ↿
2
sp p
Even though the hybrid orbital has a lower energy than the p orbital, the fourth electron must
go into the p orbital for the bonding to take place. Then, the hybrid orbitals will each form a σ
bond with an oxygen, while the unhybridised p orbital will form a π bond with one oxygen
atom, completing the VB picture for one resonance structure.

However, as we have learnt in the previous section on resonance, all the three bonds in the
ion are equivalent. In fact, the carbon p orbital will overlap with a p orbital from all three
oxygen atoms to form an extended π system. The two electrons in the π system is
delocalized throughout the whole system, and not localized between any two atoms. On the
contrary, the electrons in a regular bond are localized in between the two atoms which
contribute the orbitals.

In sp hybridisation, the 2s and one 2p orbital give rise two two sp orbitals, while the other
two p orbitals remain unhybridised. This gives rise to linear geometry.

10 Chem@HC
s px

Combine to generate two sp orbitals

180

Which are
represented
as the set

Figure 3. Mixing of the s and p orbitals to give sp hybrid orbitals. Figure taken from General Chemistry: Principles
and Modern Applications (10e) by Petrucci et. al.

In the linear carbon dioxide molecule, the four valence electrons of carbon are distributed
into the two sp orbitals and the two p orbitals. In this case, the two sp orbitals are involved in
σ bonding with oxygen, while the unhybridised carbon p orbitals are involved in π bonding.
This forms the two double bonds between the carbon and oxygen atoms.

↿ ↿ ↿ ↿
sp p
Other geometries correspond to other hybridisation schemes. The table below lists the five
common schemes you should familiarize yourself with:

Geometry Hybridisation scheme


Linear sp
2
Trigonal planar sp
3
Tetrahedral sp
3
Trigonal bipyramidal sp d
3 2
Octahedral sp d

It must be emphasized that hybridisation is not a tool to predict molecular geometry:


hybrid orbitals are generated to agree with the VSEPR-predicted shape. Furthermore, when
you learn about d metal complexes, you will also not be able to apply VB theory very well. In
short, VB theory, especially hybridisation, should only be applied primarily to main
group species. Also, as a final note, we only speak of σ , π and δ bonds in VB theory; we
never say σ , π and δ orbitals. The latter belongs to the language of the other prevalent
bonding theory which will not be covered here.

11 Chem@HC
Question 6:

State the hybridisation scheme of the central atoms in the following species:

(a) NO2
(b) SF4
(c) HCN

(d) N3

(e) Cl O 4

Question 7:

Amides are carboxylic acid derivatives, commonly encountered in organic chemistry. Amides
are found everywhere, from the proteins in your body to the synthetic material nylon (in fact,
both are examples of polyamides). The primary amides are structurally the simplest, and
their general structure can be represented by the following Lewis structure, where R is an
organic group:

Primary amides (and generally all amides) are trigonal planar about the nitrogen atom. Does
this conflict with VSEPR theory? How do we reconcile this conflict?

12 Chem@HC

You might also like