You are on page 1of 46

orbitals with more p character will be

directed towards groups that are more


electronegative. By removing the
assumption that all hybrid orbitals are
equivalent spn orbitals, better predictions
and explanations of properties such as
molecular geometry and bond strength
can be obtained.[4] Bent's rule has been
proposed as an alternative to VSEPR
theory as an elementary explanation for
observed molecular geometries of simple
molecules with the advantages of being
more easily reconcilable with modern
theories of bonding and having stronger
experimental support.
Bent's rule can be generalized to d-block
elements as well. The hybridisation of a
metal center is arranged so that orbitals
with more s character are directed towards
ligands that form bonds with more
covalent character. Equivalently, orbitals
with more d character are directed
towards groups that form bonds of greater
ionic character.[1] The validity of Bent's rule
for 75 bond types between the main group
elements was examined recently.[5] For
bonds with the larger atoms from the
lower periods, trends in orbital
hybridization depend strongly on both
electronegativity and orbital size.
History
In the early 1930s, shortly after much of
the initial development of quantum
mechanics, those theories began to be
applied towards molecular structure by
Pauling,[6] Slater,[7] Coulson,[8] and others.
In particular, Pauling introduced the
concept of hybridisation, where atomic s
and p orbitals are combined to give hybrid
sp, sp2, and sp3 orbitals. Hybrid orbitals
proved powerful in explaining the
molecular geometries of simple molecules
like methane (tetrahedral with an sp3
carbon). However, slight deviations from
these ideal geometries became apparent
in the 1940s.[9] A particularly well known
example is water, where the angle
between hydrogens is 104.5°, far less than
the expected 109.5°. To explain such
discrepancies, it was proposed that
hybridisation can result in orbitals with
unequal s and p character. A. D. Walsh
described in 1947[9] a relationship between
the electronegativity of groups bonded to
carbon and the hybridisation of said
carbon. Finally, in 1961, Henry A. Bent
published a major review of the literature
that related molecular structure, central
atom hybridisation, and substituent
electronegativities [2] and it is for this work
that Bent's rule takes its name.
Justification

Structure of fluoromethane. The C-H and C-F bonds


are polar and so the electron density will be shifted
towards carbon in the C-H bonds and towards fluorine
in the C-F bonds. Directing orbitals with more s
character towards the hydrogens is stabilizing, while
directing orbitals of less s character towards the
fluorine is destabilizing but to a lesser extent.

Polar covalent bonds


An informal justification of Bent's rule
relies on s orbitals being lower in energy
than p orbitals.[2] Bonds between elements
of different electronegativities will be polar
and the electron density in such bonds will
be shifted towards the more
electronegative element. Applying this to
the molecule fluoromethane provides a
demonstration of Bent's rule. Because
carbon is more electronegative than
hydrogen, the electron density in the C-H
bonds will be closer to carbon. The energy
of those electrons will depend heavily on
the hybrid orbitals that carbon contributes
to these bonds because of the increased
electron density near the carbon. By
increasing the amount of s character in
those hybrid orbitals, the energy of those
electrons can be reduced because s
orbitals are lower in energy than p orbitals.

By the same logic and the fact that


fluorine is more electronegative than
carbon, the electron density in the C-F
bond will be closer to fluorine. The hybrid
orbital that carbon contributes to the C-F
bond will have relatively less electron
density in it than in the C-H case and so
the energy of that bond will be less
dependent on the carbon's hybridisation.
By directing hybrid orbitals of more p
character towards the fluorine, the energy
of that bond is not increased very much.

Instead of directing equivalent sp3 orbitals


towards all four substituents, shifting s
character towards the C-H bonds will
stabilize those bonds greatly because of
the increased electron density near the
carbon, while shifting s character away
from the C-F bond will increase its energy
by a lesser amount because that bond's
electron density is further from the carbon.
The atomic s character on the carbon
atom has been directed toward the more
electropositive hydrogen substituents and
away from the electronegative fluorine,
which is exactly what Bent's rule suggests.

Although fluoromethane is a special case,


the above argument can be applied to any
structure with a central atom and 2 or
more substituents. The key is that
concentrating atomic s character in
orbitals directed towards electropositive
substituents by depleting it in orbitals
directed towards electronegative
substituents results in an overall lowering
of the energy of the system. This
stabilizing trade off is responsible for
Bent's rule.
Nonbonding orbitals

Bent's rule can be extended to rationalize


the hybridization of nonbonding orbitals as
well. On the one hand, a lone pair (an
occupied nonbonding orbital) can be
thought of as the limiting case of an
electropositive substituent, with electron
density completely polarized towards the
central atom. Bent's rule predicts that, in
order to stabilize the unshared, closely
held nonbonding electrons, lone pair
orbitals should take on high s character. On
the other hand, an unoccupied nonbonding
orbital can be thought of as the limiting
case of an electronegative substituent,
with electron density completely polarized
towards the ligand. Bent's rule predicts
that, in order to leave as much s character
as possible for the remaining occupied
orbitals, unoccupied nonbonding orbitals
should maximize p character.

Experimentally, the first conclusion is in


line with the reduced bond angles of
molecules with lone pairs like water or
ammonia compared to methane, while the
second conclusion accords with the planar
structure of molecules with unoccupied
nonbonding orbitals, like monomeric
borane and carbenium ions.
Examples
Bent’s rule can be used to explain trends in
both molecular structure and reactivity.
After determining how the hybridisation of
the central atom should affect a particular
property, the electronegativity of
substituents can be examined to see if
Bent’s rule holds.

Bond angles

Knowing the angles between bonds is a


crucial component in determining a
molecular structure. In valence bond
theory, covalent bonds are assumed to
consist of two electrons lying in
overlapping, usually hybridised, atomic
orbitals from bonding atoms. Orbital
hybridisation explains why methane is
tetrahedral and ethylene is planar for
instance. However, there are deviations
from the ideal geometries of spn
hybridisation such as in water and
ammonia. The bond angles in those
molecules are 104.5° and 107°
respectively, which are below the expected
tetrahedral angle of 109.5°. The traditional
approach to explain those differences is
VSEPR theory. In that framework, valence
electrons are assumed to lie in localized
regions and lone pairs are assumed to
repel each other to a greater extent than
bonding pairs.

Bent’s rule provides an alternative


explanation as to why some bond angles
differ from the ideal geometry. First, a
trend between central atom hybridisation
and bond angle can be determined by
using the model compounds methane,
ethylene, and acetylene. In order, the
carbon atoms are directing sp3, sp2, and
sp orbitals towards the hydrogen
substituents. The bond angles between
substituents are 109.5°, ~120°, and 180°.
This simple system demonstrates that
hybridised atomic orbitals with higher p
character will have a smaller angle
between them. This result can be made
rigorous and quantitative as Coulson's
theorem (see Formal theory section
below).

Now that the connection between


hybridisation and bond angles has been
made, Bent’s rule can be applied to
specific examples. The following were
used in Bent’s original paper, which
considers the group electronegativity of
the methyl group to be less than that of
the hydrogen atom because methyl
substitution reduces the acid dissociation
constants of formic acid and of acetic
acid.[2]

Molecule Bond angle between substituents

111°
Dimethyl ether

 
107-109°
Methanol

104.5°
Water

103.8°
Oxygen difluoride

As one moves down the table, the


substituents become more electronegative
and the bond angle between them
decreases. According to Bent's rule, as the
substituent electronegativies increase,
orbitals of greater p character will be
directed towards those groups. By the
above discussion, this will decrease the
bond angle. This agrees with the
experimental results. Comparing this
explanation with VSEPR theory, VSEPR
cannot explain why the angle in dimethyl
ether is greater than 109.5°

In predicting the bond angle of water,


Bent’s rule suggests that hybrid orbitals
with more s character should be directed
towards the lone pairs, while that leaves
orbitals with more p character directed
towards the hydrogens, resulting in
deviation from idealized O(sp3) hybrid
orbitals with 25% s character and 75% p
character. In the case of water, with its
104.5° HOH angle, the OH bonding orbitals
are constructed from O(~sp4.0) orbitals
(~20% s, ~80% p), while the lone pairs
consist of O(~sp2.3) orbitals (~30% s,
~70% p). As discussed in the justification
above, the lone pairs behave as very
electropositive substituents and have
excess s character. As a result, the
bonding electrons have increased p
character. This increased p character in
those orbitals decreases the bond angle
between them to less than the tetrahedral
109.5°. The same logic can be applied to
ammonia (107.0° HNH bond angle, with
three N(~sp3.4 or 23% s) bonding orbitals
and one N(~sp2.1 or 32% s) lone pair), the
other canonical example of this
phenomenon.

The same trend holds for nitrogen


containing compounds. Against the
expectations of VSEPR theory but
consistent with Bent's rule, the bond
angles of ammonia (NH3) and nitrogen
trifluoride (NF3) are 107° and 102°,
respectively.

Unlike VSEPR theory, whose theoretical


foundations now appear shaky, Bent's rule
is still considered to be an important
principle in modern treatments of
bonding.[10] For instance, a modification of
this analysis is still viable, even if the lone
pairs of H2O are considered to be
inequivalent by virtue of their symmetry
(i.e., only s, and in-plane px and py oxygen
AOs are hybridized to form the two O-H
bonding orbitals σO-H and lone pair nO(σ),
while pz becomes an inequivalent pure p-
character lone pair nO(π)), as in the case of
lone pairs emerging from natural bond
orbital methods.

Bond lengths
Similarly to bond angles, the hybridisation
of an atom can be related to the lengths of
the bonds it forms.[2] As bonding orbitals
increase in s character, the σ bond length
decreases.

Molecule Average carbon–carbon bond length

 
1.54 Å

1.50 Å

1.46 Å

By adding electronegative substituents


and changing the hybridisation of the
central atoms, bond lengths can be
manipulated. If a molecule contains a
structure X-A--Y, replacement of the
substituent X by a more electronegative
atom changes the hybridization of central
atom A and shortens the adjacent A--Y
bond.

Molecule Average carbon–fluorine bond length

1.388 Å

Fluoromethane

1.358 Å

Difluoromethane

1.329 Å

Trifluoromethane

1.323 Å

Tetrafluoromethane

Because fluorine is so much more


electronegative than hydrogen, in
fluoromethane the carbon will direct hybrid
orbitals higher in s character towards the
three hydrogens than towards the fluorine.
In difluoromethane, there are only two
hydrogens so less s character in total is
directed towards them and more is
directed towards the two fluorines, which
shortens the C—F bond lengths relative to
fluoromethane. This trend holds all the
way to tetrafluoromethane whose C-F
bonds have the highest s character (25%)
and the shortest bond lengths in the
series.

The same trend also holds for the


chlorinated analogs of methane, although
the effect is less dramatic because
chlorine is less electronegative than
fluorine.[2]

Molecule Average carbon–chlorine bond length

1.783 Å

Chloromethane

1.772 Å

Dichloromethane

1.767 Å

Trichloromethane

1.766 Å

Tetrachloromethane

The above cases seem to demonstrate


that the size of the chlorine is less
important than its electronegativity. A
prediction based on sterics alone would
lead to the opposite trend, as the large
chlorine substituents would be more
favorable far apart. As the steric
explanation contradicts the experimental
result, Bent’s rule is likely playing a primary
role in structure determination.

JCH Coupling constants

Perhaps the most direct measurement of s


character in a bonding orbital between
hydrogen and carbon is via the 1H−13C
coupling constants determined from NMR
spectra. Theory predicts that JCH values
will be much higher in bonds with more s
character.[11][12] In particular, the one bond
13C-1H coupling constant 1J13C-1H is
related to the fractional s character of the
carbon hybrid orbital used to form the
bond through the empirical relationship
. (For
instance the pure sp3 hybrid atomic orbital
found in the C-H bond of methane would
have 25% s character resulting in a
expected coupling constant of 500 Hz ×
0.25 = 125 Hz, in excellent agreement with
the experimentally determined value.)
Molecule JCH (of the methyl protons)

125 Hz

Methane

127 Hz

Acetaldehyde

134 Hz
1,1,1–Trichloroethane

 
141 Hz
Methanol

149 Hz

Fluoromethane

As the electronegativity of the substituent


increases, the amount of p character
directed towards the substituent increases
as well. This leaves more s character in
the bonds to the methyl protons, which
leads to increased JCH coupling constants.
Inductive effect

The inductive effect can be explained with


Bent’s rule.[13] The inductive effect is the
transmission of charge through covalent
bonds and Bent’s rule provides a
mechanism for such results via
differences in hybridisation. In the table
below,[14] as the groups bonded to the
central carbon become more
electronegative, the central carbon
becomes more electron-withdrawing as
measured by the polar substituent
constant. The polar substituent constants
are similar in principle to σ values from the
Hammett equation, as an increasing value
corresponds to a greater electron-
withdrawing ability. Bent's rule suggests
that as the electronegativity of the groups
increase, more p character is diverted
towards those groups, which leaves more
s character in the bond between the
central carbon and the R group. As s
orbitals have greater electron density
closer to the nucleus than p orbitals, the
electron density in the C−R bond will more
shift towards the carbon as the s character
increases. This will make the central
carbon more electron-withdrawing to the R
group.[9] Thus, the electron-withdrawing
ability of the substituents has been
transferred to the adjacent carbon, exactly
what the inductive effect predicts.

Polar substituent constant


Substituent (larger values imply greater
electron-withdrawing ability)

−0.30

t–Butyl

0.00

Methyl

1.05

Chloromethyl

1.94

Dichloromethyl

2.65

Trichloromethyl

Formal theory
Bent's rule provides an additional level of
accuracy to valence bond theory. Valence
bond theory proposes that covalent bonds
consist of two electrons lying in
overlapping, usually hybridised, atomic
orbitals from two bonding atoms. The
assumption that a covalent bond is a
linear combination of atomic orbitals of
just the two bonding atoms is an
approximation (see molecular orbital
theory), but valence bond theory is
accurate enough that it has had and
continues to have a major impact on how
bonding is understood.[1]
In valence bond theory, two atoms each
contribute an atomic orbital and the
electrons in the orbital overlap form a
covalent bond. Atoms do not usually
contribute a pure hydrogen-like orbital to
bonds.[6] If atoms could only contribute
hydrogen-like orbitals, then the
experimentally confirmed tetrahedral
structure of methane would not be
possible as the 2s and 2p orbitals of
carbon do not have that geometry. That
and other contradictions led to the
proposing of orbital hybridisation. In that
framework, atomic orbitals are allowed to
mix to produce an equivalent number of
orbitals of differing shapes and energies.
In the aforementioned case of methane,
the 2s and three 2p orbitals of carbon are
hybridized to yield four equivalent sp3
orbitals, which resolves the structure
discrepancy. Orbital hybridisation allowed
valence bond theory to successfully
explain the geometry and properties of a
vast number of molecules.

In traditional hybridisation theory, the


hybrid orbitals are all equivalent.[15]
Namely the atomic s and p orbital(s) are
combined to give four spi3 = 1⁄√4 (s + √3pi)
orbitals, three spi2 = 1⁄√3 (s + √2pi) orbitals,
or two spi = 1⁄√2 (s + pi) orbitals. These
combinations are chosen to satisfy two
conditions. First, the total amount of s and
p orbital contributions must be equivalent
before and after hybridisation. Second, the
hybrid orbitals must be orthogonal to each
other.[15] If two hybrid orbitals were not
orthogonal, by definition they would have
nonzero orbital overlap. Electrons in those
orbitals would interact and if one of those
orbitals were involved in a covalent bond,
the other orbital would also have a
nonzero interaction with that bond,
violating the two electron per bond tenet
of valence bond theory.

To construct hybrid s and p orbitals, let the


first hybrid orbital be given by s + √λi pi,
where pi is directed towards a bonding
group and λi determines the amount of p
character this hybrid orbital has. This is a
weighted sum of the wavefunctions. Now
choose a second hybrid orbital s + √λj pj,
where pj is directed in some way and λj is
the amount of p character in this second
orbital. The value of λj and direction of pj
must be determined so that the resulting
orbital can be normalized and so that it is
orthogonal to the first hybrid orbital. The
hybrid can certainly be normalized, as it is
the sum of two normalized wavefunctions.
Orthogonality must be established so that
the two hybrid orbitals can be involved in
separate covalent bonds. The inner
product of orthogonal orbitals must be
zero and computing the inner product of
the constructed hybrids gives the
following calculation.

Note that the s orbital is normalized and


so the inner product . Also, the
s orbital is orthogonal to the pi and pj
orbitals, which leads to two terms in the
above equaling zero. Finally, the last term
is the inner product of two normalized
functions that are at an angle of ωij to
each other, which gives cos ωij by
definition. However, the orthogonality of
bonding orbitals demands that

, so we get

Coulson's theorem as a result:[15]

This means that the four s and p atomic


orbitals can be hybridised in arbitrary
directions provided that all of the
coefficients λ satisfy the above condition
pairwise to guarantee the resulting orbitals
are orthogonal.
Bent's rule, that centrals atoms direct
orbitals of greater p character towards
more electronegative substituents, is
easily applicable to the above by noting
that an increase in the λi coefficient
increases the p character of the s + √λi pi
hybrid orbital. Thus, if a central atom A is
bonded to two groups X and Y and Y is
more electronegative than X, then A will
hybridise so that λX < λY. More
sophisticated theoretical and computation
techniques beyond Bent’s rule are needed
to accurately predict molecular geometries
from first principles, but Bent’s rule
provides an excellent heuristic in
explaining molecular structures.
See also
Valence bond theory
Molecular orbital theory
Orbital hybridisation
VSEPR theory
Molecular geometry
Linear combination of atomic orbitals

References
1. Weinhold, F.; Landis, C. L. (2005), Valency
and Bonding: A Natural Donor-Acceptor
Perspective (1st ed.), Cambridge:
Cambridge University Press, ISBN 0-521-
83128-8
2. Bent, H. A. (1961), "An appraisal of
valence-bond structures and hybridization
in compounds of the first-row elements",
Chem. Rev., 61 (3): 275–311,
doi:10.1021/cr60211a005
3. Foster, J. P.; Weinhold, F. (1980), "Natural
hybrid orbitals", J. Am. Chem. Soc., 102
(24): 7211–7218,
doi:10.1021/ja00544a007
4. Orbital Hybridization: a Key Electronic
Factor in Control of Structure and
Reactivity. Alabugin, I. V.; Bresch S.; Gomes,
G. P. J. Phys. Org. Chem., 2015, 28, 147-
162.
http://onlinelibrary.wiley.com/doi/10.1002/
poc.3382/abstract
5. Hybridization Trends for Main Group
Elements and Expanding the Bent’s Rule
Beyond Carbon: More than
Electronegativity". Alabugin, I. V.; Bresch S.;
Manoharan, M. J. Phys. Chem. A 2014, 118,
3663– 3677.
http://pubs.acs.org/doi/abs/10.1021/jp502
472u
6. Pauling, L. (1931), "The nature of the
chemical bond. Application of results
obtained from the quantum mechanics and
from a theory of paramagnetic
susceptibility to the structure of
molecules", J. Am. Chem. Soc., 53 (4):
1367–1400, doi:10.1021/ja01355a027
7. Slater, J. C. (1931), "Directed Valence in
Polyatomic Molecules", Phys. Rev., 37 (5):
481–489, Bibcode:1931PhRv...37..481S ,
doi:10.1103/PhysRev.37.481
8. Coulson, C. A. (1961), Valence (2nd ed.),
Oxford: Clarendon Press
9. Walsh, A. D. (1947), "The properties of
bonds involving carbon", Discuss. Faraday
Soc., 2: 18–25,
doi:10.1039/DF9470200018
10. Weinhold, F.; Landis, Clark R. (2012).
Discovering Chemistry with Natural Bond
Orbitals. Hoboken, N.J.: Wiley. pp. 67–68.
ISBN 9781118119969.
11. Muller, N.; Pritchard, D. E. (1959), "C13 in
Proton Magnetic Resonance Spectra. I.
Hydrocarbons", J. Chem. Phys., 31 (3):
768–771, Bibcode:1959JChPh..31..768M ,
doi:10.1063/1.1730460
12. Muller, N.; Pritchard, D. E. (1959), "C13 in
Proton Magnetic Resonance Spectra. II.
Bonding in Substituted Methanes", J.
Chem. Phys., 31 (6): 1471–1476,
Bibcode:1959JChPh..31.1471M ,
doi:10.1063/1.1730638
13. Bent, H. A. (1960), "Distribution of
atomic s character in molecules and its
chemical implications", J. Chem. Educ., 37
(12): 616–624,
Bibcode:1960JChEd..37..616B ,
doi:10.1021/ed037p616
14. Taft Jr., R. W. (1957), "Concerning the
Electron—Withdrawing Power and
Electronegativity of Groups", J. Chem.
Phys., 26 (1): 93–96,
Bibcode:1957JChPh..26...93T ,
doi:10.1063/1.1743270
15. Coulson, C. A. (1961), Valence (2nd
ed.), Oxford: Clarendon Press, pp. 203–5
Non-equivalent hybrids

Retrieved from
"https://en.wikipedia.org/w/index.php?
title=Bent%27s_rule&oldid=860086029"

Last edited 28 days ago by Alsosaid…


Content is available under CC BY-SA 3.0 unless
otherwise noted.

You might also like