Chemical Bonding
Chemical Bonding
This is the most important fact about chemical bonding that you should know, but it
is not of itself a workable theory of the chemical bond because it does not describe the
conditions under which bonding occurs, nor does it make useful predictions about
the properties of the bond.
What is a molecule?
All atoms attract one another at small distances; the universal attractive interactions
known as van der Waals forces exist between all matter, and play an important part in
determining the properties of liquids and solids. These attractions are extremely weak,
however, and they lack specificity: they do not lead to aggregates having any special
structure or composition.
A molecule is an Chemical bonding connotes the existence of an aggregate of
atoms that is sufficiently stable to possess a characteristic
aggregate of
structure and composition. The important thing to understand
atoms that pos-
about the definition written at the left is that it is essentially an
sesses distinctive operational one; as our ability to observe the characteristic
and distinguish- properties of loosely-bound aggregates of atoms increases, our
ing properties. ideas of what constitutes a molecule will change. This was illus-
trated quite vividly in the early 1980’s, when metal clusters—
stable arrangements of 5-20 metallic atoms— were first characterized. These had not
previously been recognized as molecules because no one know how to observe them.
More recently, advances in technology that allow chemists to observe chemical species
that can only exist for tiny fractions of a second have greatly extended the range of
what we can call “molecules”.
The internuclear distance at which the energy minimum occurs defines the bond
length. This is more correctly known as the equilibrium bond length, because thermal
motion causes the two atoms to vibrate about this distance. In general, the stronger
the bond, the smaller will be the bond length.
Potential energy and kinetic energy. Quantum theory tells us that an electron in
an atom possesses kinetic energy K as well as potential energy P, so the total energy E
is always the sum of the two: E = P + K. The relation between them is surprisingly
simple: K = –0.5 P. This means that when a chemical bond forms (an exothermic pro-
cess with ∆E < 0), the decrease in potential energy is accompanied by an increase in
the kinetic energy (embodied in the momentum of the bonding electrons), but the
magnitude of the latter change is only half as much, so the change in potential energy
always dominates. The bond energy –∆E has half the magnitude of the fall in potential
energy.
H C N O F Cl Br I Si
43 41 39 46 56 43 37 29 39
H
6 5 0 4 9 2 0 5 5
34 29 35 43 33 27 24 36
C
5 0 0 9 0 5 0 0
16 20 27 20 24
N
0 0 0 0 5
14 18 20 22 20 37
O
0 5 5 0 0 0
16 25 23 28 54
F
0 5 5 0 0
24 22 21 35
Cl
3 0 0 9
19 18 29
Br
0 0 0
15 21
I
0 0
23
Si
0
Table 1: Average energies of some single bonds (kJ/mol)
On the atomic scale in which all motions are quantized, a vibrating system can pos-
sess only certain allowed vibrational frequencies, or states. These are depicted by the
horizontal lines in the potential energy curve shown here. Notice that the very bottom
of the curve does not correspond to an allowed state, as this allows for no change in
the position of the atoms at all, and would therefore violate the uncertainty principle.
VSEPR model
The valence shell electron repulsion model is not so much a model of chemical bonding
as a scheme for explaining the shapes of molecules. It is based on the quantum
mechanical view that bonds represent electron clouds— physical regions of negative
electric charge that repel each other and thus try to stay as far apart as possible. We
will explore this concept in much greater detail beginning on Page 30.
1. For more information on the coulombic model, see the articles by Lawrence J. Sacks
in J. Chemical Education, 1986: 288-297, and 373-378.
Lewis’ idea that the electrons are shared in pairs stemmed from his observation that
most molecules possess an even number of electrons. This paired sharing also allows
the formulas of a large number of compounds to be rationalized and predicted— a fact
that led to the widespread acceptance of the Lewis model by the early 1920s.
For the lightest atoms the octet rule must be modified, since the noble-gas configura-
tion will be that of helium, which is simply s2 rather than s2p6. Thus we write LiH as
Li:H, where the electrons represented by the two dots come from the s orbitals of the
separate atoms.
The octet rule applies quite well to the first full row of the periodic table (Li through F),
but beyond this it is generally applicable only to the non-transition elements, and even
in many of these it cannot explain many of the bonding patterns that are observed. The
principal difficulty is that a central atom that is bonded to more than four peripheral
atoms must have more than eight electrons around it if each bond is assumed to con-
sist of an electron pair. In these cases, we hedge the rule a bit, and euphemistically
refer to the larger number of electrons as an “expanded octet”.
In spite of the octet rule’s many exceptions and limitations, the shared electron-pair
model is the one that chemists most frequently employ in their day-to-day thinking
about molecules. It continues to serve as a very useful guiding principle and can be a
good starting point for more sophisticated theories.
Multiple bonds
If one pair of electrons shared between two atoms constitutes a
chemical bond, it seems logical that two or three pairs could be C O
shared to produce double and triple bonds. Such formulations
appear quite naturally when the octet rule is applied to elements H –C C–H
such as C, O, S, and N.
Since multiple bonds place more electron density between the two N N
nuclei, the latter are held toward each other more closely and tightly;
multiple bonds are therefore shorter and stronger than single bonds. –
C N
Expanded octets
As mentioned previously, the octet rule works best for the elements in the first period
(Li through F) of the periodic table. The reason for this is that electrons, whether
shared or not, must be contained in orbitals, and the energies of electrons in such
orbitals must be relatively low; otherwise, there would be no energetic advantage in
forming a bond in the first place-- the atoms would be better off by themselves.
For the first-period elements, the only available orbitals are the s2p6 set that, when
filled, comprise the octet. In order to place more than four other atoms around any one
of these elements, the additional shared electrons would have to go into the much-
Paramagnetic molecules
A somewhat different situation is exemplified by the common oxygen
molecule. It is easy to write a proper Lewis structure for O2 that places O O
an octet around each oxygen atom and a double bond between them.
However, it takes only a simple experiment to show that the electrons in
dioxygen cannot all be arranged in pairs: if you place a magnet near some liquid oxy-
gen, the liquid will be drawn to the magnet. This can only mean one thing: there are at
least two unpaired electrons in the O2 molecule. A more exact experiment shows that
this number is exactly two. Are they in the bond or are they non-bonding electrons?
You can decide this by sketching out a few possible structures.
The paramagnetism of oxygen is an anomaly in terms of the Lewis theory, although it is
predicted by a more comprehensive theory that we will look at later. There are, how-
ever, a few other molecules that we would expect to be paramagnetic simply because
they contain an odd number of valence electrons. The most well known example is
nitric oxide, NO. Since oxygen has four and nitrogen has five outer electrons, the total
number of valence electrons is nine, and magnetic measurements show that one of
these is unpaired.
These structures differ only in which oxygen atom is attached by the double bond.
Since there is no reason to prefer one over another, the NO3– ion is regarded as a
superposition, or hybrid of these three structures.
The term resonance has been used to describe this phenomenon, which is indicated in
the above structures by the double-ended arrows. The choice of this word was unfortu-
nate, because it connotes the existence of some kind of dynamic effect that has led to
the mistaken idea that the structure is continually alternating between the three possi-
bilities. The correct interpretation is simply that none of the hybrid structures com-
pletely represents the molecule; the “real” structure is a superposition of the individual
contributing structures.
5.1 Electronegativity
The electrons constituting a chemical bond are simultaneously attracted by the electro-
static fields of the nuclei of the two bonded atoms. In a homonuclear molecule such as
O2 the bonding electrons will be shared equally by the two atoms. In general, however,
differences in the sizes and nuclear charges of the atoms will cause one of them to exert
a greater attraction on the bonding pair, causing the electron cloud to be displaced
toward the more strongly-attracting atom.
The electronegativity of an atom denotes its relative electron-attracting power in a chem-
ical bond.
It is important to understand that electronegativity is not a measurable property of an
atom in the sense that ionization energies and electron affinities are, although it can
be correlated with both of these properties. The actual electron-attracting power of an
atom depends in part on its chemical environment (that is, on what other atoms are
bonded to it), so tabulated electronegativities should be regarded as high-precision
predictors of the behavior of electrons in more complicated molecules.
There are several ways of computing electronegativities, which are expressed on an
arbitrary scale. The concept of electronegativity was introduced by Linus Pauling and
his 0-4 scale continues to be the one most widely used.
.
Fig. 8: Electronegativities
in which the core charge is the electric charge the atom would have if all its valence
electrons were removed.
In simple cases, the formal charge can be worked out visually directly from the Lewis
structure, as is illustrated farther on.
Problem Example: Lewis structures for carbon monoxide can be written in which the
C–O bond is either double or triple. Calculate the formal charges on carbon and oxy-
gen in the two alternative structures.
Solution: For ::C::O: :
C: 4 – 4 – 2 = –2 O: 6 – 2 – 2 = +2
For C:::O:::
C: 4 – 0 – 3 = +1 6 – 4 – 3 = +1
The general rule for choosing between alternative structures is that the one involving
Fig. 10: Alternative Lewis structures for the thiocyanate ion SCN–
The electrons in the structures of the top row are the valence electrons for each atom; an additional
electron (purple) completes the nitrogen octet in this negative ion. The electrons in the bottom row
are divided equally between the bonded atoms; the difference between these numbers and those
above gives the formal charges.
Formal charge can also help answer the question “where is the charge located?” that is
frequently asked about polyatomic ions. Thus by writing out the Lewis structure for the
ammonium ion NH4+, you should be able to convince yourself that the nitrogen atom
has a formal charge of +1 and each of the hydrogens has 0, so we can say that the pos-
itive charge is localized on the central atom.
Oxidation number is another arbitrary way of characterizing atoms in molecules. In
contrast to formal charge, in which the electrons in a bond are assumed to be shared
equally, oxidation number is the electric charge an atom would have if the bonding elec-
trons were assigned exclusively to the more electronegative atom. Oxidation number
serves mainly as a tool for keeping track of electrons in reaction in which they are
exhanged between reactants, and for characterizing the “combining power” of an atom
in a molecule or ion.
Fig. 11: Comparison of electron assignments for formal charge and oxidation number.
The resulting substance is sometimes said to contain an ionic bond. Indeed, the proper-
ties of a number of compounds can be adequately explained using the ionic model. But
does this mean that there are really two kinds of chemical bonds, ionic and covalent?
It is obvious that the electron-pair bond brings about this situation, and this is the rea-
son for the stability of the covalent bond. What is not so obvious (until you look at the
numbers such as were quoted for LiF above) is that the “ionic” bond results in the same
condition; even in the most highly ionic compounds, both electrons are close to both
nuclei, and the resulting mutual attractions bind the nuclei together. This being the
case, is there really any fundamental difference between the ionic and covalent bond?
The answer, according to modern chemical thinking1 is probably “no”; in fact, there is
some question as to whether it is realistic to consider that these solids consist of “ions”
in the usual sense. The preferred picture that seems to be emerging is one in which the
electron orbitals of adjacent atom pairs are simply skewed so as to place more electron
density around the “negative” element than around the “positive” one.
This being said, it must be reiterated that the ionic model of bonding is a useful one for
many purposes, and there is nothing wrong with using the term “ionic bond” to
describe the interactions between the atoms in “ionic solids” such as LiF and NaCl.
Polar covalence
If there is no such thing as a “completely ionic” bond, can we have one that is com-
pletely covalent? The answer is yes, if the two nuclei have equal electron attracting
powers. This situation is guaranteed to be the case with homonuclear diatomic mole-
cules-- molecules consisting of two identical atoms. Thus in Cl2, O2, and H2, electron
sharing between the two identical atoms must be exactly even; in such molecules, the
center of positive charge corresponds exactly to the center of negative charge: halfway
between the two nuclei.
1. See R.T. Sanderson: Chemical bonds and bond energy, Chapter 9 (Academic Press)
In BF3, one the 2p orbitals is unoccupied and can accommodate the lone pair on the
nitrogen atom of ammonia. The electron acceptor, BF3, acts as a Lewis acid here, and
NH3 is the Lewis base.
Bonds of this type (sometimes known as coordinate covalent or dative bonds) tend to be
rather weak (usually 50-200kJ/mol); in many cases the two joined units retain suffi-
cient individuality to justify writing the formula as a molecular complex or adduct.
Plane-trigonal molecules
In the molecule BF3, there are three regions of elec-
tron density extending out from the central boron
atom. The repulsion between these will be at a mini-
mum when the angle between any two is 120°. This
requires that all four atoms be in the same plane;
the resulting shape is called trigonal planar, or sim-
ply trigonal.
Ethene, CH2CH2
Ethane, CH3CH3
Both of these structures can be extended indefinitely into very long hydrocarbon
chains.
The O-H bond angle in water is 104.5°. This is somewhat less than the tetrahedral
angle of 109.5°, and it means that the four orbitals are not entirely equivalent. This is
not at all surprising, considering that two of them are nonbonding. Because a non-
bonding orbital has no atom at its far end to draw the electron cloud to it, the charge in
such an orbital will be concentrated closer to the central atom. As a consequence, non-
bonding orbitals exert more repulsion on other orbitals than do bonding orbitals. Thus
in H2O, the two nonbonding orbitals push the bonding orbitals closer together, closing
the angle somewhat.
The bent shape of the water molecule can be a source of confusion if you are not care-
ful: the oxygen atom in water is tetrahedrally coordinated, but the molecule itself has
a shape defined by its atoms rather than its orbitals. The two hydrogen atoms are sit-
uated near the corners of a tetrahedron that is centered on the oxygen atom, but
these three points define only a bent shape, not a complete tetrahedron.
The only way that we can obtain two unpaired electrons for bonding in beryllium is to
promote one of the 2s electrons to the 2p level. However, the energy required to carry
out this promotion would be sufficiently great to discourage bond formation. It is
observed that Be does form reasonably stable bonds with other atoms. Moreover, the
two bonds in BeH2 and similar molecules are completely equivalent; this would not be
the case if the electrons in the two bonds shared Be orbitals of different types.
In fact, there is little reason to believe that s, p, and d orbitals really do exist in the
outer shells of many bonded atoms. Remember that these different orbitals arise in the
first place from the interaction of the electron with the central electrostatic force field
associated with the positive nucleus. An outer -shell electron in a bonded atom will be
under the influence of a force field, emanating from two positive nuclei rather than one,
so we would expect the orbitals in the bonded atoms to have a somewhat dif ferent
character from those in free atoms. We can, in fact, throw out the concept of atomic
orbital altogether and reassign the electrons to a new set of molecular orbitals that are
characteristic of each molecular configuration. This approach is indeed valid, but we
will defer a discussion of it until later. For now, we will look at a less-radical model that
1. Linus Pauling (1901-1994) was the most famous American chemists of the 20th cen-
tury and the author of the classic The nature of the chemical bond. His early work pio-
neered the application of X-ray diffraction to determine the structure of complex
molecules; he then went on to apply quantum theory to explain these observations and
predict the bonding patterns and energies of new molecules. Pauling, who spent most of
his career at Cal Tech, won the Nobel Prize for Chemistry in 1954 and the Peace Prize in
1962.
atomic orbitals hybrid orbitals First, recall that the electron, being a
quantum particle, cannot have a dis-
2p tinct location; the most we can do is
2s sp
define the region of space around the
nucleus in which the probability of find-
ing the electron exceeds some arbi-
trary value, such as 90% or 99%. This
region of space is the orbital. Because
in-phase combination of the wavelike character of matter, the
orbital corresponds to a standing wave
pattern in 3-dimensional space which
2s we can often represent more clearly in
2-dimensional cross section. The
p quantity that is varying (“waveing”) is a
number denoted by ψ whose value
varies from point to point according to
the wave function for that particular
orbital.
Sigma bonds
In the ground state of the free carbon atom, there are two unpaired electrons in sepa-
rate 2p orbitals. In order to form four bonds (tetravalence), need four unpaired elec-
trons in four separate but equivalent orbitals. We assume that the single 2s, and the
three 2p orbitals of carbon mix into four sp3 hybrid orbitals which are chemically and
geometrically identical; the latter condition implies that the four hybrid orbitals extend
toward the corners of a tetrahedron centered on the carbon atom.
The tetrachloro zinc ion is another structure derived from zinc and chlorine. As we
might expect, this ion is tetrahedral; there are four chloride ions surrounding the cen-
tral zinc ion. The zinc ion has a charge of +2, and each chloride ion is -1, so the net
charge of the complex ion is -2.
7.5 Hybrid types and multiple bonds
Up to now we have dealt with molecules in which the orbitals on a single central atom
are considered to be hybridized. Let us now see how these various models of hybridiza-
tion can help us understand bonding between pairs of such atoms. Because of its
This alternative hybridization scheme explains how carbon can combine with four
atoms in some of its compounds and with three other atoms in other compounds. You
may be aware of the conventional way of depicting carbon as being tetravalent in all its
compounds; it is often stated that carbon always forms four bonds, but that some-
times, as in the case of ethylene, one of these may be a double bond. This concept of
the multiple bond preserves the idea of tetravalent carbon while admitting the exist-
ence of molecules in which carbon is clearly combined with fewer than four other
atoms.
As shown above, ethylene can be imagined to form when two -CH2 fragments link
together through overlap of the half-filled sp2 hybrid orbitals on each. Since sp2 hybrid
orbitals are always in the same plane, the entire ethylene molecule is planar. However,
there remains on each carbon atom an electron in an unhybridized atomic pz orbital
that is perpendicular to the molecular plane. These two parallel pz orbitals will interact
with each other; the two orbitals merge, forming a sausage-like charge cloud (the p
bond) that extends both above and below the plane of the molecule. It is the pair of
electrons that occupy this new extended orbital that constitutes the “fourth” bond to
each carbon, and thus the “other half” of the double bond in the molecule.
Pi orbitals, on the other hand, require the presence of two atomic p orbitals on adja-
cent atoms. Most important, the charge density in the π orbital is concentrated above
and below the molecular plane; it is almost zero along the line-of-centers between the
two atoms.It is this perpendicular orientation with respect to the molecular plane (and
the consequent lack of cylindrical symmetry) that defines the π orbital. The combina-
tion of a σ bond and a π bond extending between the same pair of atoms constitutes the
double bond in molecules such as ethylene.
Triple bonds
We have not yet completed our overview of multiple bonding, however. Carbon and
hydrogen can form yet another compound, acetylene, in which each carbon is con-
nected to only two other atoms: a carbon and a hydrogen. This can be regarded as an
example of divalent carbon, but is usually rationalized by writing a triple bond between
the two carbon atoms.
We assume here that since two geometrically equivalent bonds are formed by each car-
bon, this atom must be sp-hybridized in acetylene. On each carbon, one sp hybrid
bonds to a hydrogen and the other bonds to the other carbon atom, forming the σ bond
skeleton of the molecule.
In addition to the sp hybrids, each carbon atom has two half-occupied p orbitals, ori-
ented at right angles to each other, and to the interatomic axis. These are two sets of
parallel and adjacent p orbitals, and they can thus merge into two sets of π orbitals.
The triple bond in acetylene is seen to consist of one σ bond joining the line-of-centers
between the two carbon atoms, and two π bonds whose lobes of electron density are in
mutually-perpendicular planes. The acetylene molecule is of course linear, since the
angle between the two sp hybrid orbitals that produce the σ skeleton of the molecule is
Fig. 22:
A conjugated carbon
chain
When single and double
bonds alternate in a chain of
carbon atoms, the two π
bonding arrangements
shown at the top are equiva-
lent. The resulting structure
is said to be conjugated and
tend to be especially stable.
The C–C bond order is 1.5.
Benzene
The classic example of π bond delocalization is found in the cyclic molecule benzene
(C6H6), which consists of six carbon atoms bound together in a hexagonal arrange-
ment. Each carbon has a single hydrogen atom attached to it. Earlier, you learned to
represent the benzene structure as a composite of two resonance forms.
Modern hybridization theory offers a somewhat less contrived view: each carbon atom
is sp hybridized, producing trigonal bond geometry about each node in the hexagon.
Each carbon atom also has a half-occupied unhybridized p orbital; two of these
together, on adjacent carbon atoms, can form two equivalent sets of π bonds as shown
in the center section. The net effect is depicted at the right. The π bonding system is
drawn out into two donut shaped electron clouds, one on either side of the molecular
plane. The “double bonds” in benzene are delocalized so that they extend around the
entire hexagonal skeleton.
Nitrogen has three half-occupied p orbitals available for bonding, all perpendicular to
one another. Since the nitrate ion is known to be planar, we are forced to assume that
the nitrogen outer electrons are sp2 hybridized. The addition of an extra electron fills
all three hybrid orbitals completely. Each of these filled sp2 orbitals forms a σ bond by
overlap with an empty oxygen 2pz orbital; this, you will recall, is an example of coordi-
nate covalent bonding, in which one of the atoms contributes both of the bonding elec-
trons.
The empty oxygen 2p orbital is made available when the oxygen electrons themselves
become sp hybridized; we get three filled sp hybrid orbitals, and an empty 2p atomic
orbital, just as in the case of nitrogen.
The π bonding system arises from the interaction of one of the occupied oxygen sp
orbitals with the unoccupied 2pz orbital of the nitrogen. Notice that this, again, is a
coordinate covalent sharing, except that in this instance, it is the oxygen atom that
Some of the most important and commonly encountered compounds which involve the
d orbitals in bonding are the transition metal complex. The term “complex” in this con-
text means that the molecule is composed of two or more kinds of species, each of
which can have an independent existence. For example, the ions Pt 2+ and Cl– can form
the ion [PtCl4]2– which is found experimentally to be square planar. To understand the
hybridization scheme, it helps to start with the neutral Pt atom, then imagine it losing
two electrons to become an ion, followed by grouping of the two unpaired 5d electrons
into a single d orbital, leaving one vacant.This vacant orbital, along with the 6s and two
of the 6p orbitals, can then accept an electron pair from four chlorines.
Octahedral coordination
This same type of hybridization can be invoked to explain the structures of many tran-
sition metal cations; the hexamminezinc(II) cation depicted below is typical.
All of the four-coordinated molecules we have discussed so far have tetrahedral geome-
try around the central atom. Methane, CH4, is the most well known example. It may
come as something as a surprise, then, to discover that the tetrachlorplatinum (II)
ion [PtCl4]2– has an essentially two-dimensional square-planar configuration. This
In sp3d2 hybridization the bonding orbitals are derived by mixing atomic orbitals hav-
ing the same principal quantum number (n = 4 in the preceding example). A slightly
different arrangement, known as d2sp3 hybridization, involves d orbitals of lower prin-
cipal quantum number. This is possible because of the rather small energy differences
between the d orbitals in one “shell” with the s and p orbitals of the next higher one—
hence the term “inner orbital” complex which is sometimes used to describe ions such
as hexaminecobalt(III), shown below. Both arrangements produce octahedral coordi-
nation geometries.
In some cases, the same central atom can form either inner or outer complexes
depending on the particular ligand and the manner in which its electrostatic field
affects the relative energies of the different orbitals.Thus the hexacyanoiron(II) ion uti-
lizes the iron 3d orbitals, whereas hexaaquoiron(II) achieves a lower energy by accept-
ing two H2O molecules in its 4d orbitals.
As two H nuclei move toward each other, the 1s atomic orbitals of the isolated atoms
gradually merge into a new molecular orbital in which the greatest electron density
falls between the two nuclei. Since this is just the location in which electrons can exert
the most attractive force on the two nuclei simultaneously, this arrangement consti-
tutes a bonding molecular orbital. Regarding it as a three- dimensional region of space,
we see that it is symmetrical about the line of centers between the nuclei; in accord
with our usual nomenclature, we refer to this as a (sigma) orbital.
Fig. 31: Bonding and antibonding molecular orbitals formed by in- and out-
of-phase combinations of atomic orbitals
When the two 1s wave functions combine out-of-phase, the regions of high electron
probability do not merge. In fact, the orbitals act as if they actually repel each other.
Notice particularly that there is a region of space exactly equidistant between the nuclei
at which the probability of finding the electron is zero. This region is called a nodal sur-
face, and is characteristic of antibonding orbitals. It should be clear that any electrons
that find themselves in an antibonding orbital cannot possibly contribute to bond for -
mation; in fact, they will actively oppose it.
We see, then, that whenever two orbitals, originally on separate atoms, begin to interact
as we push the two nuclei toward each other, these two atomic orbitals will gradually
merge into a pair of molecular orbitals, one of which will have bonding character, while
the other will be antibonding. In a more advanced treatment, it would be fairly easy to
show that this result follows quite naturally from the wave-like nature of the combining
orbitals.
What is the difference between these two kinds of orbitals, as far as their potential
energies are concerned? More precisely, which kind of orbital would enable an electron
to be at a lower potential energy? Clearly, the potential energy decreases as the electron
moves into a region that enables it to “see” the maximum amount of positive charge. In
Dihydrogen
If one electron in the bonding orbital is conducive to bond for-
mation, might two electrons be even better? We can arrange
this by combining two hydrogen atoms-- two nuclei, and two
electrons. Both electrons will enter the bonding orbital, as
depicted in the Figure. We recall that one electron lowered the
potential energy of the two nuclei by 270 kJ/mole, so we
might expect two electrons to produce twice this much stabili-
zation, or 540 kJ/mole.
Experimentally, one finds that it takes only 452 kJ to break apart a mole of hydrogen
molecules. The reason the potential energy was not lowered by the full amount is that
the presence of two electrons in the same orbital gives rise to a repulsion that acts
against the stabilization. This is exactly the same effect we saw in comparing the ion-
ization energies of the hydrogen and helium atoms.
Dilithium
For example, when lithium, whose configuration is
1s22s1,bonds with itself to form Li2, we can forget about the
1s atomic orbitals and consider only the σ bonding and anti-
bonding orbitals shown at the right. Since there are not
enough electrons to populate the antibonding orbital, the
attractive forces win out and we have a stable molecule. The
bond energy of dilithium is 104.6 kJ/mole; notice that this
value is far less than the 270 kJ bond energy in dihydrogen,
which also has two electrons in a bonding orbital.
The reason, of course, is that the 2s orbital of Li is much farther from its nucleus than
is the 1s orbital of H, and this is equally true for the corresponding molecular orbitals.
It is a general rule, then, that the larger the parent atom, the less stable will be the cor-
responding diatomic molecule.
Fig. 32: Formation of sigma bonding- and antibonding molecular orbitals from end-to-end
atomic p orbitals.
These directional differences lead to the formation of two different classes of molecular
orbitals. The above figure shows how two px atomic orbitals interact. In many ways the
resulting molecular orbitals are similar to what we got when s atomic orbitals com-
bined; the bonding orbital has a large electron density in the region between the two
nuclei, and thus corresponds to the lower potential energy. In the out-of-phase combi-
nation, most of the electron density is away from the internuclear region, and as
before, there is a surface exactly halfway between the nuclei that corresponds to zero
electron density. This is clearly an antibonding orbital-- again, in general shape, very
much like the kind we saw in hydrogen and similar molecules. Like the ones derived
from s-atomic orbitals, these molecular orbitals are orbitals.
Sigma orbitals are cylindrically symmetric with respect to the line of centers of the
nuclei; this means that if you could look down this line of centers, the electron density
Fig. 33: Formation of pi bonding and antibonding orbitals from two parallel
atomic-p orbitals.
When we examine the results of the in- and out-of-phase combination of py and pz
orbitals, we get the bonding and antibonding pairs that we would expect, but the
resulting molecular orbitals have a different symmetry: rather than being rotationally
symmetric about the line of centers, these orbitals extend in both perpendicular direc-
tions from this line of centers.Orbitals having this more complicated symmetry are
called π (pi) orbitals. There are two of them, py and pz differing only in orientation, but
otherwise completely equivalent.
The different geometric properties of the π
σ* and orbitals causes the sigma orbitals to
split more than the pi orbitals, so that the
π* antibonding orbitals sigma antibonding orbital always has the
highest energy. The sigma bonding orbital
can be either higher or lower than the pi
bonding orbitals, depending on the partic-
px py pz px py pz ular atom.
atomic orbitals π atomic orbitals
If we combine the splitting schemes for
the 2s and 2p orbitals, we can predict
σ bonding orbitals bond order in all of the diatomic molecules
and ions composed of elements in the first
complete row of the periodic table.
Fig. 34: Splitting scheme for p orbitals
Remember that only the valence orbitals
of the atoms need be considered; as we
saw in the cases of lithium hydride and dilithium, the inner orbitals remain tightly
bound and retain their localized atomic character.
Dicarbon
Carbon has four outer-shell electrons, two 2s
and two 2p. For two carbon atoms, we therefore
have a total of eight electrons, which can be
accommodated in the first four molecular orbit-
als. The lowest two are the 2s -derived bonding
and antibonding pair, so the “first” four electrons
make no net contribution to bonding. The other
four electrons go into the pair of pi bonding orbit-
als, and there are no more electrons for the anti-
bonding orbitals— so we would expect the
dicarbon molecule to be stable, and it is.
You will recall that one pair of electrons shared
between two atoms constitutes a “single” chemi-
cal bond; this is Lewis’ original definition of the
covalent bond. In C2 there are two paris of elec-
tron in the pi bonding orbitals, so we have what
amounts to a double bond here. Actually, the
preferred nomenclature is “bond order”; the bond order in dicarbon is two.
Studies of this kind are carried out by placing a sample consisting of a solution of the
complex between the poles of an electromagnet. The sample is suspended from the
arm of a sensitive balance, and the change in weight is measured with the magnet on
and off. An increase in the weight when the magnet is turned on indicates that the
sample is attracted to the magnet (paramagnetism) and must therefore possess one
or more unpaired electrons. The precise number can be determined by calibrating the
system with a substance whose electron configuration is known.
d-orbital splitting
Although the five d orbitals of the central atom all have the same energy in a spheri-
cally symmetric field, their energies will not all be the same in the octahedral field
imposed by the presence of the ligands. The reason for this is apparent when we con-
sider the different geometrical properties of the five d orbitals. Two of the d orbitals,
designated dx2 and dx2-y2, have their electron clouds pointing directly toward ligand
atoms. We would expect that any electrons that occupy these orbitals would be subject
to repulsion by the electron pairs that bind the ligands that are situated at correspond-
ing corners of the octahedron. As a consequence, the energies of these two d orbitals
will be raised in relation to the three other d orbitals whose lobes are not directed
toward the octahedral positions.
The number of electrons in the d
orbital of the central atom is easily
determined from the location of the
element in the periodic table, taking
in account, of course, of the number
of electrons removed in order to form
the positive ion.
.
The effect of the octahedral ligand field due to the
ligand electron pairs is to split the d orbitals into
two sets whose energies differ by a quantity denoted
by ∆ which is known as the d orbital splitting. Note
that both sets of central-ion d orbitals are repelled
by the ligands and are both raised in energy; the
upper set is simply raised by a greater amount.
Both the total energy shift and ∆ are strongly depen-
dent on the particular ligands.
In contrast to this, the cyanide ion acts as a strong-field ligand; the d orbital splitting is
so great that it is energetically more favorable for the electrons to pair up in the lower
group of d orbitals rather than to enter the upper group with unpaired spins. Thus
hexacyanoiron(II) is a “low spin” complex— actually zero spin, in this particular case.
Different d orbital splitting occur in square planar and tetrahedral coordination geome-
tries, so a very large number of arrangements are possible. In most complexes the
value of ∆ corresponds to the absorption of visible light, accounting for the colored
nature of many such compounds in solution and in solids such as CuSO 4·5H2O.
Chlorophyll
Chlorophyll is the light-harvesting pigment present in green plants. Its name comes
from the Greek word chloros, meaning “green”— the same root from which chlorine gets
its name. Chlorophyll consists of a ring-shaped tetradentate ligand known as a porphin
coordinated to a central magnesium ion. A histidine residue from one of several types
of associated proteins forms a fifth coordinate bond to the Mg atom.
Hemoglobin
Hemoglobin performs the essential task of transporting dioxygen molecules from the
lungs to the tissues in which it is used to oxidize glucose, this oxidation serving as the
source of free energy required for cellular metabolic processes. Like chlorophyll, hemo-
globin consists of a tetradentate ligand, heme, combined with a polypeptide (mini-pro-
tein) chain which together coordinate with an iron atom in five positions. The sixth
position in the octahedrally-coordinated iron is taken up either by an oxygen molecule
or by a water molecule, depending on whether the hemoglobin is in its oxygenated state
(in arteries) or deoxygenated state (in veins.)
Other ligands, notably cyanide ion and carbon monoxide, are able to bind to hemoglo-
bin much more strongly than does iron, thereby displacing it and rendering hemoglo-
bin unable to transport oxygen. Air containing as little as 1 percent CO will convert
hemoglobin to carboxyhemoglobin in a few hours, leading to loss of consciousness and
death. Even small amounts of carbon monoxide can lead to substantial reductions in
the availability of oxygen. The 400-ppm concentration of CO in cigarette smoke will tie
up about 6% of the hemoglobin in heavy smokers; the increased stress this places on
the heart as it works harder to compensate for the oxygen deficit is believed to be one
reason why smokers are at higher risk for heart attacks.
of positive ions immersed in a “sea of electrons” which can freely migrate throughout
the solid. In effect the electropositive nature of the metallic atoms allows their valence
electrons to exist as a mobile fluid which can be displaced by an applied electric field,
hence giving rise to their high electrical conductivities. Because each ion is surrounded
by the electron fluid in all directions, the bonding has no directional properties; this
accounts for the high malleability and ductility of metals.
This view is really an oversimplification that fails to explain metals in a quantitative
way, nor can it account for the differences in the properties of individual metals. A
more detailed treatment, known as the bond theory of metals, applies the idea of reso-
nance hybrids to metallic lattices. In the case of an alkali metal, for example, this
would involve a large number of hybrid structures in which a given Na atom shares its
electron with its various neighbors.
In Li4 we begin to
see a trend in which
the lower (mostly
bonding) MO’s are
filled and the upper
(mostly antibonding)
ones are empty.
Thermal conductivity Everyone knows that touching a metallic surface at room temper-
ature produces a colder sensation than touching a piece of wood or plastic at the same
temperature. The very high thermal conductivity of metals allows them to draw heat
out of our bodies very efficiently if they are below body temperature. In the same way, a
metallic surface that is above body temperature will feel much warmer than one made
of some other material. The high thermal conductivity of metals is attributed to vibra-
In most metals there will be bands derived from the outermost s-, p-, and d atomic lev-
els, leading to a system of bands, some of which will overlap as described above. Where
overlap does not occur, the almost continuous energy levels of the bands are separated
by a forbidden zone, or band gap. Only the outermost atomic orbitals form bands; the
inner orbitals remain localized on the individual atoms and are not involved in bond-
ing.
An insulator is characterized by a large band gap between the highest filled band and
an even higher empty band. The band gap is sufficiently great to prevent any signifi-
cant population of the upper band by thermal excitation of electrons from the lower
one. The presence of a very intense electric field may be able to supply the required
energy, in which case the insulator undergoes dielectric breakdown. Most molecular
crystals are insulators, as are covalent crystals such as diamond.
If the band gap is sufficiently small to allow electrons in the filled band below it to jump
into the upper empty band by thermal excitation, the solid is known as a semiconduc-
tor. In contrast to metals, whose electrical conductivity decreases with temperature
(the more intense lattice vibrations interfere with the transfer of momentum by the
electron fluid), the conductivity of semiconductors increases with temperature. In
many cases the excitation energy can be provided by absorption of light, so most semi-
conductors are also photoconductors. Examples of semiconducting elements are Se,
Te, Bi, Ge, Si, and graphite.
The presence of an impurity in a semiconductor can introduce a new band into the sys-
tem. If this new band is situated within the forbidden region, it creates a new and
smaller band gap that will increase the conductivity. The huge semiconductor industry
is based on the ability to tailor the band gap to fit the desired application by introduc-
ing an appropriate impurity atom (dopant) into the semiconductor lattice. The dopant
elements are normally atoms whose valance shells contain one electron more or less
than the atoms of the host crystal. For example, a phosphorus atom introduced as an
impurity into a silicon lattice possesses one more valence electron than Si. This elec-
tron is delocalized within the impurity band and serves as the charge carrier in what is
known as an N-type semiconductor. In a semiconductor of the P-type, the dopant might
be arsenic, which has only three valence electrons. This creates what amounts to an
electron deficiency, or hole in the electron fabric of the crystal, although the solid
remains electrically neutral overall. As this vacancy is filled by the electrons from sili-
con atoms the vacancy hops to another location, so the charge carrier is effectively a
positively charged hole, hence the P-type designation.