6.1 Overlap Integrals
Operations and representations are merely
theoretical constructs. What is actually observed are the
interactions. In quantum
mechanics, interactions are expressed as matrix elements of
operators in a function space. When the operator is the unit
operator, the matrix elements are just overlap integrals. These are
the simplest form of interactions.
We start our analysis by examining symmetry
selection rules for overlap integrals. Consider the overlap
integral between the ith
component of a function space which transforms according to the
irrep Γ, and the
kth component of another
function space transforming as Γ′. The overlap integral, S ki , is a scalar quantity and
thus must be invariant under the action of linear symmetry
operators acting on the functions.
An integral being an infinite sum, the operator can be brought
inside the bracket and then transform the bra and ket parts
directly.
By summing over all
and dividing by the group order one
obtains a form to which the GOT can be applied.
We now rewrite this result in terms of elements of the overlap
matrix
:
This simple derivation yields three important results:
These results clearly illustrate the importance of the GOT. It not
only provides a selection rule at the level of the irreps, but also
at the level of the components. Of course, the latter selection
rule will work only if we ensured that the symmetry adaptation of
the basis set has been carried out at the component level, as was
explained in Sect. 5.3.
![$$ S_{ki}= \bigl\langle\phi^{\varGamma'}_k\big| \psi^{\varGamma}_i\bigr\rangle=\hat{R} \bigl\langle \phi^{\varGamma'}_k\big|\psi^{\varGamma}_i\bigr\rangle $$](A303787_1_En_6_Chapter_Equ1.gif)
(6.1)
![$$ \hat{R} \bigl\langle\phi^{\varGamma'}_k\big|\psi^{\varGamma}_i \bigr\rangle= \bigl\langle\hat {R}\phi^{\varGamma'}_k\big|\hat{R} \psi^{\varGamma}_i\bigr\rangle= \sum_{jl} \bar {D}^{\varGamma'}_{lk}(R) D^{\varGamma}_{ji}(R) \bigl \langle\phi^{\varGamma'}_l\big|\psi^{\varGamma}_j\bigr \rangle $$](A303787_1_En_6_Chapter_Equ2.gif)
(6.2)
![$\hat{R} \in G$](A303787_1_En_6_Chapter_IEq1.gif)
![$$\begin{aligned} \bigl\langle\phi^{\varGamma'}_k\big|\psi^{\varGamma}_i \bigr\rangle =& \frac{1}{|G|} \sum_{R \in G} \hat{R} \bigl\langle\phi^{\varGamma'}_k\big|\psi ^{\varGamma}_i \bigr\rangle \\ =& \frac{1}{|G|}\sum_{jl} \biggl( \sum _R \bar{D}^{\varGamma'}_{lk}(R) D^{\varGamma}_{ji}(R) \biggr) \bigl\langle\phi^{\varGamma'}_l\big| \psi^{\varGamma}_j \bigr\rangle \\ =& \delta_{\varGamma' \varGamma} \delta_{ik} \frac{1}{\mathrm{dim}({\varGamma})} \sum _j \bigl\langle\phi^{\varGamma'}_j\big| \psi^{\varGamma}_j \bigr\rangle \end{aligned}$$](A303787_1_En_6_Chapter_Equ3.gif)
(6.3)
![$\mathbb{S}$](A303787_1_En_6_Chapter_IEq2.gif)
![$$ S_{ki} = \delta_{\varGamma' \varGamma} \delta_{ik} \frac{1}{\mathrm{dim}({\varGamma })} \sum_j S_{jj} = \delta_{\varGamma' \varGamma} \delta_{ik} \frac{1}{\mathrm{dim}({\varGamma })} \mathrm{Tr} (\mathbb{S}) $$](A303787_1_En_6_Chapter_Equ4.gif)
(6.4)
1.
Overlap integrals between functions which
transform according to different irreps are zero.
2.
Overlap integrals between functions which belong
to different components of the same irrep are zero.
3.
Overlap integrals between functions with the same
symmetry properties, i.e. transforming as the same component of the same irrep, are independent of the
component choice provided that both components are
normalized.
A further consequence is that SALCs on peripheral
atom sites can quite often easily be derived from central
symmetry-adapted orbitals. One simply has to make sure that the
SALCs have the same nodal characteristics as the central functions,
so as to guarantee maximal overlap. This is well illustrated in
Fig. 4.4.
6.2 The Coupling of Representations
Overlap integrals are scalar products of a bra
and a ket function. A general matrix element is an integral of
the outer product of a bra, an operator, and a ket, giving rise to
a triad of irreps. The evaluation of such elements is based on the
coupling of irreps. This
concept refers to the formation of a product space. The simplest
example is the formation of a two-electron wavefunction, obtained
by multiplying two one-electron functions. This section will be
devoted entirely to the formation of such product spaces.
Consider two sets of orbitals, transforming as
the irreps Γ
a and
Γ b respectively, each occupied by
one electron. A two-electron wavefunction with electron 1 in
the γ a component of the first set, and
electron 2 in the γ
b component of
the second set is written as a simple product function:
|Γ a γ a (1)〉|Γ b γ b (2)〉. Clearly, since the
one-electron function spaces are invariants of the group, their
product space is invariant, too. Now the question is to determine
the symmetry of this new space. The recipe to find this symmetry
can safely be based on the character theorem: first determine the
character string for the product basis, and then carry out the
reduction according to the character theorem. Symmetry operators
are all-electron operators affecting all particles together; hence,
the effect of a symmetry operation on a ket product is to transform
both kets simultaneously.
The transformation of the product functions is thus expressed by a
super matrix, each element of which is a product of two matrix
elements for the individual orbital transformations. The trace of
this super matrix is given by:
This is a gratifying result. The character of a product space is
simply the product of the characters of the factor spaces.
Accordingly, the symmetry of the product space is identified as the
direct product of the
orbital irreps, and is denoted as Γ a ×Γ b . If both irreps are
degenerate, the direct product will be reducible. Let c Γ be the number of times that the
irrep Γ occurs in the
direct product:
By straightforward application of the character theorem one
obtains:
Here we have, for the first time, a formula with a triad of irreps.
This will form the basis for the symmetry evaluation of general
matrix elements. The c
Γ coefficients
are obtained by performing product manipulations on the character
tables. As an example, Table 6.1 illustrates the
reduction of the E
g ×T 2g product in O h , as given in Eq. (6.9). Product tables are
given in Appendix E.
![$$ \hat{R}\bigl(\big|\varGamma_a\gamma_a (1) \big\rangle\big| \varGamma_b\gamma_b (2)\big\rangle\bigr) =\sum _{\gamma_a'}\sum_{\gamma_b'} D^{\varGamma_a}_{\gamma_a' \gamma_a} (R) D^{\varGamma_b}_{\gamma_b' \gamma_b} (R)\big | \varGamma_a\gamma_a' (1)\big\rangle \big| \varGamma_b\gamma_b' (2)\big\rangle $$](A303787_1_En_6_Chapter_Equ5.gif)
(6.5)
![$$\begin{aligned} \chi^{\varGamma_a \times\varGamma_b}(R) =& \sum_{\gamma_a \gamma_b} D^{\varGamma_a}_{\gamma_a \gamma_a}(R) D^{\varGamma_b}_{\gamma_b \gamma_b}(R) \\ =& \chi^{\varGamma_a}(R) \chi^{\varGamma_b}(R) \end{aligned}$$](A303787_1_En_6_Chapter_Equ6.gif)
(6.6)
![$$ \varGamma_a \times\varGamma_b = \sum _{\varGamma} c_{\varGamma} \varGamma $$](A303787_1_En_6_Chapter_Equ7.gif)
(6.7)
![$$\begin{aligned} c_{\varGamma} =& \frac{1}{|G|} \sum_R \bar{\chi}^{\varGamma} (R)\chi ^{\varGamma_a \times\varGamma_b}(R) \\ =& \frac{1}{|G|} \sum_R \bar{ \chi}^{\varGamma} (R)\chi^{\varGamma_a}(R) \chi^{\varGamma_b}(R) \end{aligned}$$](A303787_1_En_6_Chapter_Equ8.gif)
(6.8)
![$$ E_g \times T_{2g} = T_{1g} + T_{2g} $$](A303787_1_En_6_Chapter_Equ9.gif)
(6.9)
Table 6.1
Direct product of E g ×T 2g in O h symmetry
![A303787_1_En_6_Tab1_HTML.gif](A303787_1_En_6_Tab1_HTML.gif)
Let us now proceed with the two-electron problem
and address the next question, which is that, after having
determined which symmetry species are present, we should like to
know what the corresponding two-electron wavefunctions look like,
i.e. we should like to construct the SALCs. This construction does
not pose any new problems; the projection operators that were
introduced in Sect. 4.5 will do the job perfectly well.
Some notation is important here. The product function will be
written as:
The combination coefficient is itself identified as a matrix
element, by multiplying left and right with the one-electron bra
functions and using orthonormality of the basis orbitals.
This coefficient is known as a Clebsch–Gordan (CG) coupling
coefficient and denoted by the 3Γ bracket 〈Γ a γ a Γ b γ b |Γγ〉. It indicates how the orbital
irreps Γ a and Γ b have to be combined to yield a
product ket that transforms as |Γγ〉. The CG-coefficients can be
determined by using projection operators. The results are listed in
Appendix F. It is often possible to obtain these results by a
simpler procedure. We illustrate this for the components of the
T 1g two-electron state, obtained in
Eq. (6.9). The
z-component of this state
is the only component that is totally symmetric under the
splitting field. It is clear that this symmetry can be obtained
only by multiplying the |e
g ϵ〉 and |t 2g ζ〉 components, since these are both
antisymmetric and thus will form a symmetric product. From here on
we will adopt for the product functions the usual notation of small
letters for the orbitals and capital letters for the coupled
states. Hence:
The coupling coefficient 〈E
g ϵT 2g ζ|T 1g z〉 is thus equal to 1. The
x and y components may then immediately be
obtained by applying the cyclic
generator. As an example for the
x-component:
Thus:
.
The resulting coupling coefficients are shown in
Table 6.2.
![$$ \big|\varGamma\gamma(1,2)\big\rangle= \sum_{\gamma_a}\sum _{\gamma_b} c_{\gamma _a\gamma_b}^{\varGamma\gamma} \big| \varGamma_a\gamma_a (1)\big\rangle\big|\varGamma_b \gamma_b (2) \big\rangle $$](A303787_1_En_6_Chapter_Equ10.gif)
(6.10)
![$$\begin{aligned} c^{\varGamma}_{\gamma_a\gamma_b} =& \bigl\langle\varGamma_a \gamma_a (1) \varGamma_b\gamma_b (2)| \varGamma\gamma (1,2)\bigr\rangle \\ \equiv& \langle\varGamma_a\gamma_a \varGamma_b \gamma_b|\varGamma\gamma\rangle \end{aligned}$$](A303787_1_En_6_Chapter_Equ11.gif)
(6.11)
![$\hat{C}_{4}$](A303787_1_En_6_Chapter_IEq3.gif)
![$$ |T_{1g}z\rangle= |e_g \epsilon\rangle|t_{2g} \zeta\rangle $$](A303787_1_En_6_Chapter_Equ12.gif)
(6.12)
![$\hat{C}_{3}$](A303787_1_En_6_Chapter_IEq4.gif)
![$$\begin{aligned} |T_{1g}x\rangle =& \hat{C}_3 |T_{1g}z\rangle \\ =& \bigl( \hat{C}_3 |e_{g}\epsilon\rangle \bigr)\bigl ( \hat{C}_3 |t_{2g}\zeta\rangle \bigr) \\ =& \biggl( -\frac{\sqrt{3}}{2} |e_{g}\theta\rangle-\frac{1}{2} |e_{g}\epsilon\rangle \biggr) |t_{2g}\xi\rangle \\ =&-\frac{\sqrt{3}}{2} |e_{g}\theta\rangle|t_{2g}\xi\rangle- \frac {1}{2} |e_{g}\epsilon\rangle|t_{2g}\xi\rangle \end{aligned}$$](A303787_1_En_6_Chapter_Equ13.gif)
(6.13)
![$\langle\theta\xi| x\rangle= -\sqrt{3}/2; \langle\epsilon \xi|x\rangle= -1/2$](A303787_1_En_6_Chapter_IEq5.gif)
Table 6.2
The coupling coefficients for the direct
product E g ×T 2g in O h symmetry
![A303787_1_En_6_Tab2_HTML.gif](A303787_1_En_6_Tab2_HTML.gif)
6.3 Symmetry Properties of the Coupling Coefficients
The CG-coefficients in the finite point groups
stem from Wigner’s celebrated coupling coefficients for the
spherical symmetry group [1].
Wigner proposed reformulating these coefficients in terms of more
primitive 3j symbols, which
contain, in a uniform way, the permutational properties of the
spherical coupling coefficients. Several attempts have been made to
define similar 3Γ symbols
for the point group, but this requires the introduction of quite
detailed phase conventions, which limits the efficiency of this
formalism [2, 3]. We shall therefore not engage in a further
factorization of the coupling coefficients, but, express the
important symmetry properties of the couplings at the level of the
brackets. Two guidelines will thereby be used: when dealing with
coupling coefficients it is important to bear in mind that the
coupling is based on the formation of a product, as we have
illustrated in the preceding section, and, secondly, that we should
treat the coupling coefficients as far as possible as ordinary
brackets.
A direct consequence of the latter viewpoint is
that the rules for complex conjugation of brackets apply:
![$$ \overline{\langle\varGamma_a\gamma_a \varGamma_b \gamma_b|\varGamma\gamma \rangle} = \langle \varGamma\gamma|\varGamma_a\gamma_a \varGamma_b \gamma _b\rangle $$](A303787_1_En_6_Chapter_Equ14.gif)
(6.14)
Being expansion coefficients of SALCs, the
coupling coefficients also obey two orthogonality rules.
Column-wise orthonormality results from the orthonormal properties
of the coupled states.
In addition, the scalar products along rows are orthonormal,
because of the orthonormal properties of the basic kets. Note that
the summation runs over the irreps of the entire product space:
Γ∈Γ a ×Γ b .
![$$ \sum_{\gamma_a \gamma_b} \bigl\langle\varGamma' \gamma'|\varGamma_a\gamma_a \varGamma _b \gamma_b\bigr\rangle \langle\varGamma_a \gamma_a \varGamma_b \gamma_b|\varGamma \gamma\rangle = \delta_{\varGamma\varGamma'} \delta_{\gamma\gamma'} $$](A303787_1_En_6_Chapter_Equ15.gif)
(6.15)
![$$ \sum_{\varGamma\gamma} \bigl\langle\varGamma_a \gamma'_a \varGamma_b \gamma'_b|\varGamma \gamma\bigr\rangle \langle\varGamma \gamma|\varGamma_a\gamma_a \varGamma_b \gamma _b\rangle = \delta_{\gamma'_a \gamma_a} \delta_{\gamma'_b \gamma_b} $$](A303787_1_En_6_Chapter_Equ16.gif)
(6.16)
The permutational properties of the
CG-coefficients refer to interchange of the bra and ket irreps. If
Γ a and Γ b are not equivalent, their
ordering will not affect the symmetry of the coupled state, since
the factors in the direct product commute:
We can therefore define the coupling coefficients in such a way
that interchange of the coupled irreps leaves the coefficient
invariant:
![$$ \varGamma_a \times\varGamma_b = \varGamma_b \times\varGamma_a $$](A303787_1_En_6_Chapter_Equ17.gif)
(6.17)
![$$ \varGamma_a \neq\varGamma_b : \langle \varGamma_a\gamma_a \varGamma_b \gamma _b|\varGamma\gamma\rangle\equiv\langle\varGamma_b \gamma_b \varGamma_a \gamma _a|\varGamma \gamma\rangle $$](A303787_1_En_6_Chapter_Equ18.gif)
(6.18)
If the coupled electrons are equivalent, i.e.
belong to the same shell,
the situation is different. In this case, Γ a =Γ b , and the direct product
becomes a direct square.
For non-equivalent irreps, exchange of the components γ a and γ b was not possible, because they
refer to different irreps. However, when they are components of the
same irrep, this exchange is an important symmetry of the product
space. Indeed, the squared space can be split into two separate
blocks, one block which contains product functions that are
symmetric under exchange of the component labels and one block
which is antisymmetric. This implies that we can define two
separate sets of direct square coupling coefficients, which are
either symmetric or antisymmetric under exchange of the labels,
i.e.: 〈Γ a γ a Γ a γ b |Γγ〉=±〈Γ a γ b Γ a γ a |Γγ〉. The symmetrized part of the direct
square is denoted as [Γ
a ]2.
For n=dim(Γ a ), the dimension of this
subspace is equal to the number of symmetric combinations:
On the other hand, if the coupling coefficients are antisymmetric
under exchange of the labels, the coupled state belongs to the
antisymmetrized direct square, denoted as {Γ a }2. This product
space is restricted to combinations with γ a ≠γ b ; its dimension is equal to
n(n−1)/2. The characters for either part
of the square can be determined separately. For the character of
the {Γ a }2 part the
derivation runs as follows: one first applies a symmetry operator
to an arbitrary antisymmetric function. The ket product
|Γ a γ a (1)〉|Γ a γ b (2)〉 will be abbreviated here
as: γ a (1)γ b (2).
Taking the trace then yields:
The trace for the symmetrized product is then found by subtracting
the trace in Eq. (6.21) from the total trace for the direct
product.
In Table 6.3 these quantities are given for the direct
product H g ×H g in icosahedral symmetry. The
product resolution is as follows:
![$$ \mathrm{dim}\bigl([\varGamma_a]^2\bigr) = \sum _{\gamma_a} 1 + \sum_{\gamma_a < \gamma _b} 1 = n + n(n-1)/2 = n(n+1)/2 $$](A303787_1_En_6_Chapter_Equ19.gif)
(6.19)
![$$\begin{aligned} &\hat{R} \bigl( \gamma_a(1) \gamma_b(2) - \gamma_b(1) \gamma_a(2) \bigr) \\ &\quad= \sum_{\gamma_a' \gamma_b'} \bigl( \gamma_a' (1) \gamma_b' (2) - \gamma_b' (1) \gamma_a' (2) \bigr) D^{\varGamma_a}_{\gamma_a' \gamma _a}(R) D^{\varGamma_a}_{\gamma_b' \gamma_b}(R) \\ &\quad= \sum_{\gamma_a' \gamma_b'} \gamma_a' (1) \gamma_b' (2) \bigl( D^{\varGamma_a}_{\gamma_a' \gamma_a}(R) D^{\varGamma_a}_{\gamma_b' \gamma_b}(R) - D^{\varGamma_a}_{\gamma_a' \gamma_b}(R) D^{\varGamma_a}_{\gamma _b' \gamma_a}(R) \bigr) \\ &\quad= \frac{1}{2} \sum_{\gamma_a' \gamma_b'}\bigl( \gamma_a' (1) \gamma _b' (2) - \gamma_b' (1) \gamma_a' (2) \bigr) \bigl( D^{\varGamma_a}_{\gamma_a' \gamma_a}(R) D^{\varGamma_a}_{\gamma_b' \gamma_b}(R) -D^{\varGamma_a}_{\gamma_a' \gamma_b}(R) D^{\varGamma_a}_{\gamma_b' \gamma_a}(R) \bigr) \end{aligned}$$](A303787_1_En_6_Chapter_Equ20.gif)
(6.20)
![$$\begin{aligned} \chi^{\{\varGamma_a\}^2}(R) =& \frac{1}{2} \sum_{\gamma_a \gamma_b} \bigl( D^{\varGamma_a}_{\gamma_a \gamma_a}(R) D^{\varGamma_a}_{\gamma_b \gamma _b}(R) -D^{\varGamma_a}_{\gamma_a \gamma_b}(R) D^{\varGamma_a}_{\gamma_b \gamma_a}(R) \bigr) \\ =& \frac{1}{2} \biggl( \bigl(\chi^{\varGamma_a} (R) \bigr)^2 - \sum_{\gamma_a} D^{\varGamma_a}_{\gamma_a \gamma_a} \bigl(R^2\bigr) \biggr) \\ =& \frac{1}{2} \bigl( \bigl( \chi^{\varGamma_a} (R) \bigr)^2 - \chi ^{\varGamma_a}\bigl(R^2\bigr) \bigr) \end{aligned}$$](A303787_1_En_6_Chapter_Equ21.gif)
(6.21)
![$$\begin{aligned} \chi^{[\varGamma_a]^2}(R) =& \chi^{{\varGamma_a}^2}(R) - \chi^{\{ \varGamma_a\} ^2}(R) \\ =& \frac{1}{2} \bigl( \bigl( \chi^{\varGamma_a} (R) \bigr)^2 + \chi ^{\varGamma_a}\bigl(R^2\bigr) \bigr) \end{aligned}$$](A303787_1_En_6_Chapter_Equ22.gif)
(6.22)
![$$ H_g \times H_g = [ A_{g} + G_g + 2 H_{g} ] + \{ T_{1g} + T_{2g} + G_g \} $$](A303787_1_En_6_Chapter_Equ23.gif)
(6.23)
Table 6.3
Direct square H g ×H g in I h symmetry
![A303787_1_En_6_Tab3_HTML.gif](A303787_1_En_6_Tab3_HTML.gif)
Note that this product contains one
totally-symmetric irrep, notably in the symmetrized part. In
general, for irreps with real characters the totally-symmetric
irrep, Γ 0,
appears in a direct square only once. This can easily be derived
from Eq. (6.8). When Γ a is an irrep with real characters, one has:
In the case of irreps that can be represented by real transformation matrices, it is
possible to show that this totally-symmetric irrep will belong to
the symmetrized part. In order to apply the character theorem to
Eq. (6.22),
the following intermediate result is needed:
In order to arrive at this result we have made use of the GOT, on
the assumption that the
matrices are real. Combining the results of
Eqs. (6.24)
and (6.25)
with Eq. (6.22) then leads to the conclusion that the
unique totally-symmetric irrep belongs to the symmetrized part of
the direct square:
For irreps with real characters, but transformation matrices that
cannot be all real, the unique totally-symmetric product appears in
the antisymmetrized part. This is the case for spin
representations, which will be dealt with in
Chap. 7. In summary, as far as
complex-conjugation properties are concerned, we have three kinds
of irreps:
![$$\begin{aligned} \bigl\langle\chi^{\varGamma_0}|\chi^{\varGamma_a \times\varGamma_b}\bigr\rangle =& \sum _R \bar{\chi}^{\varGamma_0} (R)\chi^{\varGamma_a}(R) \chi^ {\varGamma_b}(R) \\ =&\sum_R \chi^{\varGamma_a}(R) \chi^ {\varGamma_b}(R) \\ =&|G| \delta_{\varGamma_a \varGamma_b} \end{aligned}$$](A303787_1_En_6_Chapter_Equ24.gif)
(6.24)
![$$\begin{aligned} \sum_R \chi^{\varGamma_a}\bigl(R^2 \bigr) =& \sum_{R} \sum _i \bigl[ \mathbb {D}^{\varGamma_a}(R) \times \mathbb{D}^{\varGamma_a}(R) \bigr]_{ii} \\ =& \sum_R \sum_i \sum_j D^{\varGamma_a}_{ij}(R) D^{\varGamma_a}_{ji}(R) \\ =& \frac{|G|}{\mathrm{dim}(\varGamma_a)} \sum_i \sum _j \delta_{ij} = |G| \end{aligned}$$](A303787_1_En_6_Chapter_Equ25.gif)
(6.25)
![$\mathbb{D}$](A303787_1_En_6_Chapter_IEq6.gif)
![$$ \frac{1}{|G|} \bigl\langle\chi^{\varGamma_0} | \chi^{[\varGamma_a]^2}\bigr \rangle= \frac{1}{2|G|} \sum_R \chi^{\varGamma_0}(R) \bigl[ \chi^{{\varGamma_a}^2}(R) + \chi^{\varGamma_a} \bigl(R^2\bigr) \bigr] =1 $$](A303787_1_En_6_Chapter_Equ26.gif)
(6.26)
1.
Irreps with real characters, and for which all
transformation matrices can be put in real form. In this case:
Γ
0∈[Γ
a
]2.
![$\mathbb{D}(R)$](A303787_1_En_6_Chapter_IEq7.gif)
2.
Irreps with real characters, but which cannot be
represented by transformation matrices that are all real. In this
case Γ
0∈{Γ
a
}2.
3.
Irreps with complex characters. In this case
there is always a complex-conjugate irrep, and
.
![$\varGamma_{0} \in\varGamma\times\bar {\varGamma}$](A303787_1_En_6_Chapter_IEq8.gif)
Equation (6.23) further exemplifies a case of
product multiplicity. This
is when an irrep occurs more than once in the decomposition of a
direct product. Both the H
g and
G g irreps appear twice in the
direct product H
g ×H g . In the point groups product
multiplicity is quite rare. It occurs only in the icosahedral group
for the products G×H and H×H, as well for spin representations in
cubic and icosahedral symmetries. Product multiplicity means that
there are different coupling schemes for arriving at the product
states. Each of these “channels” corresponds to a separate set of
CG coupling coefficients. There are several ways of obtaining
linearly-independent sets of coupling coefficients. For the
separation of the two G
g irreps in Eq.
(6.23)
symmetrization of the product space is sufficient, since the
symmetrized and antisymmetrized parts each contain one G g . This strategy does not work
for the two H
g irreps, which
are both the result of symmetrized coupling. In this case, more
elaborate splitting schemes have been constructed, based
inter alia on higher
symmetries [4, 5].
Last but not least, we should consider the
relationship between coupling coefficients where irreps from bra
and ket parts are interchanged. We shall limit the discussion here
to the simplified case in which all ingredients of the coupling are
taken to be real. A case with complex irreps will be treated
in Chap. 7. Consider two related couplings:
Γ a ×Γ b =Γ and Γ×Γ b =Γ a . The corresponding expansion
coefficients are scalar matrix elements and are thus invariant
under the group action. By importing the group action inside the
brackets, as we have frequently done before, we obtain a set of
equations in the CG-coefficients:
These equations form a system of homogeneous linear equations from
which the coupling coefficients can be obtained. At present, we
shall use this result only in the simplified case where all
components have been chosen to be real, so that Eqs. (6.27) and (6.28) form the same
system of equations.1
From this it follows that the corresponding coupling coefficients
will be proportional to each other, independent of the components;
hence:
The proportionality constant can be determined by summing the
square of the coefficients over all components and using the
normalization result from Eq. (6.15).
The permutation of irreps between bra and ket in the
CG-coefficients thus requires a uniform dimensional
renormalization:
The renormalization leaves a phase factor undetermined. This phase
factor is the same for the entire coupling table, and thus can be
chosen in arbitrarity. As an example, in the group O (see Appendix F), the
coefficients 〈T
2 ξT
1 x|Eθ〉 and 〈EθT 1 x|T 2 ξ〉 are related as follows:
Here the phase was chosen to be +1.
![$$\begin{aligned} \langle\varGamma_a \gamma_a \varGamma_b \gamma_b | \varGamma\gamma\rangle =& \sum _{\gamma_a' \gamma_b'\gamma} \biggl[\frac{1}{|G|}\sum _R \bar {D}_{\gamma_a' \gamma_a}^{\varGamma_a} (R) \bar{D}_{\gamma_b' \gamma _b}^{\varGamma_b} (R) {D}_{\gamma' \gamma}^{\varGamma} (R) \biggr] \bigl\langle\varGamma_a \gamma_a' \varGamma_b \gamma_b' | \varGamma \gamma' \bigr\rangle \\ \end{aligned}$$](A303787_1_En_6_Chapter_Equ27.gif)
(6.27)
![$$\begin{aligned} \langle\varGamma\gamma\varGamma_b \gamma_b | \varGamma_a \gamma_a \rangle =& \sum _{\gamma_a' \gamma_b'\gamma} \biggl[\frac{1}{|G|}\sum _R {D}_{\gamma_a' \gamma_a}^{\varGamma_a} (R) \bar{D}_{\gamma_b' \gamma _b}^{\varGamma_b} (R) \bar{D}_{\gamma' \gamma}^{\varGamma} (R) \biggr] \bigl\langle\varGamma\gamma' \varGamma_b \gamma_b' | \varGamma_a \gamma_a' \bigr\rangle \end{aligned}$$](A303787_1_En_6_Chapter_Equ28.gif)
(6.28)
![$$ \langle\varGamma_a \gamma_a \varGamma_b \gamma_b | \varGamma\gamma\rangle= x \langle\varGamma\gamma \varGamma_b \gamma_b | \varGamma_a \gamma_a \rangle $$](A303787_1_En_6_Chapter_Equ29.gif)
(6.29)
![$$\begin{aligned} \sum_{\gamma_a \gamma_b \gamma} \bigl \vert \langle \varGamma_a \gamma_a \varGamma _b \gamma_b | \varGamma\gamma\rangle\bigr \vert ^2 =& x^2 \sum_{\gamma_a \gamma_b \gamma} \bigl \vert \langle \varGamma\gamma\varGamma_b \gamma_b | \varGamma_a \gamma_a \rangle\bigr \vert ^2 \\ \mathrm{dim}(\varGamma) =& x^2 \mathrm{dim}(\varGamma_a) \end{aligned}$$](A303787_1_En_6_Chapter_Equ30.gif)
(6.30)
![$$ \bigl[ \mathrm{dim}(\varGamma) \bigr]^{-1/2} \langle \varGamma_a \gamma_a \varGamma _b \gamma_b | \varGamma\gamma\rangle= \pm \bigl[ \mathrm{dim}(\varGamma _a) \bigr]^{-1/2} \langle\varGamma\gamma \varGamma_b \gamma_b | \varGamma_a \gamma_a \rangle $$](A303787_1_En_6_Chapter_Equ31.gif)
(6.31)
![$$ \frac{1}{\sqrt{2}} \langle T_2{\xi}T_1x|E\theta\rangle = \frac{1}{\sqrt{3}} \langle E\theta T_1x|T_2\xi\rangle= - \frac{1}{2} $$](A303787_1_En_6_Chapter_Equ32.gif)
(6.32)
6.4 Product Symmetrization and the Pauli Exchange-Symmetry
In principle, the T 1g and T 2g coupled two-electron states,
which we obtained in Table 6.2 of the previous section, could apply to the
case of the (t
2g
)1(e
g )1
excited states of a d
2 transition-metal ion in an octahedral ligand field,
which splits the d orbitals
into a t 2g and an e g shell. However, these coupled
descriptions are not yet sufficient, since they make a distinction
between electron 1, which resides in the t 2g orbital, and electron 2,
which was promoted to the e
g level. The
fundamental symmetry requirement that electrons must be
indistinguishable is thus not fulfilled. The operator that permutes
the two electrons is represented as
:
The |Γγ(1,2)〉 and
|Γγ(2,1)〉 states will have
exactly the same symmetries, since the factors in the direct
product commute:
As a result,
commutes with the spatial symmetry operators, and we can symmetrize
the coupled states with respect to the electron permutation. The
permutation operator is the generator of the symmetric group,
S 2, which has
only two irreps, one symmetric and one antisymmetric,
corresponding, respectively, to the plus and minus combination in
Eq. (6.35).
These states have distinct permutation symmetries, and spatial
symmetry operators cannot mix + and − states. This is a very
general property of multi-particle states, to which no exceptions
are known.
![$\hat{P}_{12}$](A303787_1_En_6_Chapter_IEq9.gif)
![$$ \hat{P}_{12} \big|\varGamma\gamma(1,2)\big\rangle= \big|\varGamma\gamma(2,1) \big\rangle=\sum_{\gamma_a}\sum_{\gamma_b} \langle\varGamma_a\gamma_a \varGamma_b \gamma _b|\varGamma\gamma\rangle \big|\varGamma_a \gamma_a (2)\big\rangle\big|\varGamma_b\gamma_b (1) \big\rangle $$](A303787_1_En_6_Chapter_Equ33.gif)
(6.33)
![$$ \varGamma_a \times\varGamma_b = \varGamma_b \times\varGamma_a $$](A303787_1_En_6_Chapter_Equ34.gif)
(6.34)
![$\hat{P}_{12}$](A303787_1_En_6_Chapter_IEq10.gif)
![$$\begin{aligned} |\varGamma\gamma; \pm\rangle =& \frac{1}{\sqrt{2}}\bigl[\big |\varGamma\gamma(1,2) \big\rangle\pm\big|\varGamma\gamma(2,1)\big\rangle\bigr] \\ =&\frac{1}{\sqrt{2}}\sum_{\gamma_a}\sum _{\gamma_b}\langle\varGamma_a\gamma _a \varGamma_b \gamma_b|\varGamma\gamma\rangle \\ &{}\times \bigl[\big|\varGamma_a\gamma_a (1)\big\rangle\big| \varGamma_b\gamma _b (2)\big\rangle\pm \big| \varGamma_b\gamma_b (1)\big\rangle\big|\varGamma_a \gamma_a (2)\big\rangle\bigr] \end{aligned}$$](A303787_1_En_6_Chapter_Equ35.gif)
(6.35)
On the other hand the permutation symmetry of
multi-electron wavefunctions is restricted by the Pauli
principle.
Theorem 11
The total
wavefunction should be antisymmetric with respect to exchange of
any pair of electrons. Hence, in the symmetric group S 2, or, for an n-electron system, the symmetric group, S n , the total wavefunction should change sign
under odd permutations, i.e. under permutations that consist of an odd
number of transpositions of two elements, and should remain invariant under even
permutations.
Until now we have limited ourselves to the
spatial part of the wavefunction. So far, only the antisymmetrized
part obeys the Pauli principle. However, the principle places a
requirement only on the total wavefunction. This also involves
a spin part, which should be multiplied by the orbital part.
Anticipating the results of Chap. 7, we here provide the spin functions
for a two-electron system. Spin functions are characterized by a
spin quantum number, S, and
a component, M
S , in the range
{−S,−S+1,−S+2,…,S−1,S}. The total number of components,
hence the dimension of the spin-space for a given S, is equal to 2S+1. This number is called the spin
multiplicity. For a two-electron system, S can be 0 or 1; hence, there is one
singlet state, and there are three components belonging to a
triplet state. In a |SM
S 〉 notation
they are given by:
These functions also exhibit permutation symmetry: the triplet
functions are symmetric under exchange of the two particles, while
the singlet function is antisymmetric under such an exchange. The
total wavefunction can thus always be put in line with the Pauli
principle by combining the coupled orbital states with spin states
of opposite permutation symmetry. Altogether we can thus construct
four states: 1 T
1g ,1
T 2g ,3 T 1g ,3 T 2g . This set of four states,
totalling 24 wavefunctions, forms a manifold, representing all the coupled
states resulting from the (t 2g )1(e g )1 configuration.
The dimension of the manifold is equal to the product of the six
possible t
2g substates
(including spin), and the four possible e g substates. In this case, where
the coupling involves electrons belonging to different shells, the
Pauli principle does not restrict the total dimension of the
manifold, since all combinations remain possible. All states can be
written as linear combinations of Slater determinants. As an
example, for the |1 T 1g z〉 state, one writes:
![$$\begin{aligned} \begin{aligned} |0,0\rangle &= \frac{1}{\sqrt{2}} \bigl[ \alpha(1) \beta(2) - \beta(1) \alpha(2) \bigr] \\ |1,1\rangle &= \alpha(1) \alpha(2) \\ |1,0\rangle &= \frac{1}{\sqrt{2}} \bigl[ \alpha(1) \beta(2) + \beta(1) \alpha(2) \bigr] \\ |1,-1\rangle &= \beta(1) \beta(2) \end{aligned} \end{aligned}$$](A303787_1_En_6_Chapter_Equ36.gif)
(6.36)
![$$\begin{aligned} \big|{}^1T_{1g}z \big\rangle =& \frac{1}{\sqrt{2}} \bigl(\big| \epsilon(1)\big\rangle\big|\zeta (2)\big\rangle+ \big|\epsilon(2)\big\rangle\big|\zeta(1)\big\rangle \bigr) \frac{1}{\sqrt {2}} \bigl[ \alpha(1) \beta(2) - \beta(1) \alpha(2) \bigr] \\ =& \frac{1}{\sqrt{2}} \bigl \vert (\epsilon\alpha) (\zeta\beta) \bigr \vert - \frac{1}{\sqrt{2}} \bigl \vert (\epsilon\beta) (\zeta\alpha) \bigr \vert \end{aligned}$$](A303787_1_En_6_Chapter_Equ37.gif)
(6.37)
The situation is different when coupling two
equivalent electrons; these are electrons that belong to
the same shell. In this
case, the coupled states are already eigenfunctions of the exchange
operator as a result of the special symmetrization properties of
the coupling coefficients for direct squares. Equation
(6.10) will
take the following form:
Now, if the product representation belongs to the symmetrized
square, Γ∈[Γ a ]2, this result
simplifies to:
Hence, this function is symmetric under the
operator. It also obeys the normalization condition of Eq.
(6.15). It
should always be multiplied by an antisymmetric singlet
spin-function in order to obey the Pauli exclusion principle. On
the other hand, if the product representation belongs to the
antisymmetrized square, Γ∈{Γ a }2, the coupled
state is given by:
This function is antisymmetric under the
operator, and should be multiplied by a symmetric triplet
spin-function in order to obey the Pauli principle. As an example,
for the (e
g )2
configuration the allowed states are:
For equivalent electrons the Pauli principle thus really does
function as an exclusion principle, since the coupled states are
either triplets or singlets, depending on their symmetrization. The
dimension of the manifold is given by the binomial coefficient,
where q is the number of
equivalent substates (including spin), and n is the number of electrons:
For the (e
g )2
problem, one has n=2 and
q=4; there are thus six
two-electron states in this configuration (see Eq. (6.41)).
![$$\begin{aligned} |\varGamma\gamma(1,2)\rangle =& \sum_{\gamma_a\gamma_b} \langle \gamma _a\gamma_b|\varGamma\gamma\rangle \big| \gamma_a (1)\big\rangle\big|\gamma_b (2)\big\rangle = \sum _{\gamma_a}\langle\gamma_a\gamma_a| \varGamma\gamma\rangle \big|\gamma_a (1) \big\rangle\big|\gamma_a (2)\big\rangle \\ &{}+ \sum_{\gamma_a < \gamma_b} \bigl[ \langle \gamma_a\gamma_b|\varGamma\gamma \rangle \big| \gamma_a (1)\big\rangle\big|\gamma_b (2)\big\rangle+\langle \gamma_b\gamma_a|\varGamma \gamma\rangle \big| \gamma_b (1)\big\rangle\big|\gamma_a (2)\big\rangle\bigr] \end{aligned}$$](A303787_1_En_6_Chapter_Equ38.gif)
(6.38)
![$$\begin{aligned} |\varGamma\gamma\rangle =& \sum _{\gamma_a} \langle\gamma_a\gamma_a| \varGamma\gamma\rangle\big|\gamma_a (1)\big\rangle\big|\gamma_a (2) \big\rangle \\ &{}+ \sum_{\gamma_a < \gamma_b} \langle\gamma_a \gamma_b|\varGamma\gamma\rangle \bigl(\big|\gamma_a (1) \rangle\big|\gamma_b (2)\rangle +\big|\gamma_b (1)\big\rangle \big| \gamma_a (2)\big\rangle\bigr) \end{aligned}$$](A303787_1_En_6_Chapter_Equ39.gif)
(6.39)
![$\hat{P}_{12}$](A303787_1_En_6_Chapter_IEq11.gif)
![$$ |\varGamma\gamma\rangle= \sum_{\gamma_a < \gamma_b} \langle \gamma_a\gamma_b|\varGamma\gamma\rangle \bigl(\big| \gamma_a (1)\big\rangle\big|\gamma_b (2)\big\rangle- \big|\gamma_b (1)\big\rangle \big|\gamma_a (2)\big\rangle \bigr) $$](A303787_1_En_6_Chapter_Equ40.gif)
(6.40)
![$\hat{P}_{12}$](A303787_1_En_6_Chapter_IEq12.gif)
![$$ e_g \times e_g = [A_{1g} + E_g] + \{A_{2g}\} \Rightarrow {}^1 A_{1g} + {} ^1 E_g + {} ^3 A_{2g} $$](A303787_1_En_6_Chapter_Equ41.gif)
(6.41)
![$$ \left ( \begin{array}{c} q \\ n \end{array} \right ) = \frac{q!}{ n! (q-n)!} $$](A303787_1_En_6_Chapter_Equ42.gif)
(6.42)
As a special result, we examine the symmetry of
the maximal spin-multiplicity ground state of a system with a
half-filled shell. The shell consists of the components
|f
1〉⋯|f
n 〉,
transforming according to the irrep Γ a , and each will be occupied by
one electron with α spin.
The ground state corresponds to a single determinant:
Here, σ∈S n is an element of the
permutation group of the n
electron labels, and sgn(σ)
is its parity.2
Equation (6.43) indicates that this permutation can
equally well be applied to the component labels, since the
determinant is invariant under matrix transposition. We can now
calculate the matrix element in the symmetry operator:
In this equation we have used the result from
Eq. (2.8), which identified matrix
elements over symmetry operators as elements of the representation
matrix. The double summation over permutations covers the
permutation group twice and could be reduced to a single sum. The
result indicates that the spatial symmetry of the half-filled shell
ground state transforms as the determinant of the irrep of the
shell. This is also called the determinantal representation. For the
e g shell the determinantal irrep
is A 2g . The shell ground state is
thus a 3 A
2g .
![$$\begin{aligned} |\varPsi\rangle= | f_1\alpha\cdots f_2\alpha| =& \frac{1}{\sqrt{n !}} \sum_{\sigma\in S_n} \mathrm{sgn}(\sigma) \bigl[ f_1\alpha(\sigma_1)\cdots f_n\alpha( \sigma_n) \bigr] \\ =& \frac{1}{\sqrt{n !}} \sum_{\sigma\in S_n} \mathrm{sgn}( \sigma) \bigl[ f_{\sigma_1}\alpha(1) \cdots f_{\sigma_n}\alpha(n) \bigr] \end{aligned}$$](A303787_1_En_6_Chapter_Equ43.gif)
(6.43)
![$$\begin{aligned} \langle\varPsi|\hat{R}|\varPsi\rangle =& \frac{1}{n!} \sum _{\sigma, \pi\in S_n} \mathrm{sgn}(\sigma) \mathrm {sgn}(\pi) \bigl[ \bigl \langle f_{\sigma_1}\alpha(1)|\hat{R}|f_{\pi_1}\alpha (1)\bigr\rangle \cdots \bigl\langle f_{\sigma_n}\alpha(n)|\hat{R}|f_{\pi_n}\alpha (n) \bigr\rangle \bigr] \\ =& \frac{1}{n!} \sum_{\sigma, \pi\in S_n} \mathrm{sgn}( \sigma) \mathrm {sgn}(\pi) \bigl[ D^{\varGamma_a}_{\sigma_1 \pi_1}(R) \cdots D^{\varGamma _a}_{\sigma_n \pi_n}(R) \bigr] \\ =& \sum_{\lambda\in S_n} \mathrm{sgn}(\lambda) \bigl[ D^{\varGamma_a}_{1 \lambda_1}(R) \cdots D^{\varGamma_a}_{n \lambda_n}(R) \bigr] \\ =& \mathrm{det}\bigl(\mathbb{D}^{\varGamma_a}\bigr) \end{aligned}$$](A303787_1_En_6_Chapter_Equ44.gif)
(6.44)
6.5 Matrix Elements and the Wigner–Eckart Theorem
A general interaction element is a bracket around
an operator. Each of the three ingredients, bra, ket, and operator,
can be put in symmetry-adapted form, so that it transforms
according to a given irrep. Moreover, provided that the symmetry
adaptation is done properly, not only the irrep itself but also the
subrepresentation is well defined. Altogether, the matrix element
will thus be characterized by six symmetry labels, as:
.
The labels imply that the symmetry behaviour of each of these
ingredients is fully known:
Note that the general form of the operator
refers to a component of an
irreducible set. Such a set of operators is usually referred to as
a tensor operator. Obvious
examples are the components of the electric or magnetic
dipole-moment operators. The Wigner–Eckart theorem introduces a
symmetry factorization, which simplifies the evaluation of matrix
elements.
![$\langle\psi^{\varOmega}_{\omega}|O^{\varLambda}_{\lambda}|\phi^{\varGamma }_{\gamma}\rangle$](A303787_1_En_6_Chapter_IEq13.gif)
![$$\begin{aligned} \begin{aligned} \hat{R} \big\langle\psi^{\varOmega}_{\omega}\big| &= \sum _{\omega'} \bar{D}^{\varOmega}_{\omega' \omega} (R) \bigl\langle\psi^{\varOmega}_{\omega'}\big| \\ \hat{R}\big|\phi^{\varGamma}_{\gamma}\bigr\rangle&= \sum _{\gamma'} D^{\varGamma }_{\gamma' \gamma} (R) \big| \phi^{\varGamma}_{\gamma'}\big\rangle \\ \hat{R} \big|O^{\varLambda}_{\lambda}\big|R^{-1} &= \sum _{\lambda'} D^{\varLambda }_{\lambda' \lambda} (R) \big|O^{\varLambda}_{\lambda'}\big| \end{aligned} \end{aligned}$$](A303787_1_En_6_Chapter_Equ45.gif)
(6.45)
![$|O^{\varLambda}_{\lambda}|$](A303787_1_En_6_Chapter_IEq14.gif)
Theorem 12
A matrix
element, involving a tensor
operator, may be factorized
into a product of an intrinsic scalar part and an
appropriate 3Γ
coupling coefficient.
The scalar constant is denoted by 〈ψ Ω ∥O Λ ∥ϕ Γ 〉, and is called the
reduced matrix element.
To prove this theorem, one first considers the coupling of two
ingredients of the matrix element, and then compares the result
with the third one. We thus first consider the coupling of the
operator and the ket. The transformation of their product does
indeed correspond to the super matrix which is due to the direct
product Λ×Γ.
This means that we couple the tensor operator and the ket to form
product entities:
We now invert this equation, using the unitary properties of the
coupling coefficients, to yield:
Then we combine this expression with the bra.
The matrix elements on the right-hand side are now in fact reduced
to an overlap integral where the direct product irreps are compared
with the irrep of the bra. Hence, the selection rules for the
overlap integrals apply:
The trace summation in this equation is identified as a scalar
interaction constant, which is represented by the reduced matrix
element.
![$$ \bigl\langle\psi^{\varOmega}_{\omega}\big|O^{\varLambda}_{\lambda}| \phi^{\varGamma }_{\gamma}\bigr\rangle= \bigl\langle\psi^{\varOmega} \parallel O^{\varLambda}\parallel \phi^{\varGamma }\bigr\rangle \langle\varOmega\omega| \varLambda\lambda\varGamma\gamma\rangle $$](A303787_1_En_6_Chapter_Equ46.gif)
(6.46)
![$$ \hat{R} \big|O^{\varLambda}_{\lambda}|\phi^{\varGamma}_{\gamma} \big\rangle= \hat{R} \big|O^{\varLambda}_{\lambda}|\hat{R}^{-1} \hat{R} |\phi^{\varGamma}_{\gamma }\big\rangle = \sum_{\lambda'} \sum_{\gamma'} D^{\varLambda}_{\lambda' \lambda} (R) D^{\varGamma}_{\gamma' \gamma} (R) |O^{\varLambda}_{\lambda'}\big| \phi^{\varGamma }_{\gamma'}\big\rangle $$](A303787_1_En_6_Chapter_Equ47.gif)
(6.47)
![$$ \big|(O\phi)^{\varPi}_{\pi}\big\rangle= \sum _{\lambda'}\sum_{\gamma'} \big|O^{\varLambda }_{\lambda'}|\phi^{\varGamma}_{\gamma'}\big\rangle \bigl\langle\varLambda\lambda' \varGamma\gamma'|\varPi \pi\bigr\rangle $$](A303787_1_En_6_Chapter_Equ48.gif)
(6.48)
![$$ \big|O^{\varLambda}_{\lambda}|\phi^{\varGamma}_{\gamma}\big\rangle= \sum_{\varPi'}\sum_{ \pi'} \big|(O \phi)^{\varPi'}_{\pi'}\big\rangle \bigl\langle\varPi' \pi' | \varLambda\lambda \varGamma\gamma\bigr\rangle $$](A303787_1_En_6_Chapter_Equ49.gif)
(6.49)
![$$ \bigl\langle\psi^{\varOmega}_{ \omega}|O^{\varLambda}_{\lambda}| \phi^{\varGamma }_{\gamma}\bigr\rangle= \sum _{\varPi'}\sum_{ \pi'} \bigl\langle \psi^{\varOmega}_ {\omega }|(O\phi)^{\varPi'}_{\pi'}\bigr \rangle \bigl\langle\varPi' \pi' | \varLambda\lambda \varGamma\gamma\bigr\rangle $$](A303787_1_En_6_Chapter_Equ50.gif)
(6.50)
![$$ \bigl\langle\varOmega\omega|(O\phi)^{\varPi'}_{\pi'}\bigr\rangle= \delta_{\varOmega, \varPi '}\delta_{\omega\pi'} \frac{1}{\mathrm{dim}({\varOmega})} \sum _{\omega'} \bigl\langle\varOmega\omega'\big|(O \phi)^{\varOmega'}_{\omega'}\bigr\rangle\equiv\bigl\langle \psi^{\varOmega}\parallel O^{\varLambda}\parallel \phi^{\varGamma}\bigr\rangle $$](A303787_1_En_6_Chapter_Equ51.gif)
(6.51)
Combination of Eqs. (6.50) and (6.51) then yields the
Wigner–Eckart theorem of Eq. (6.46), where the total interaction is the
product of a scalar interaction constant and a CG coupling
coefficient. The former refers to the interaction itself, the
latter extracts the transformation properties. In case of product
multiplicity, there will be one reduced matrix element for every
coupling channel, and the matrix element is decomposed into a sum
over the channels. The Wigner–Eckart theorem or matrix-element
theorem is at the heart of most chemical applications of group
theory. It provides an elegant method for separating interactions
into an intrinsic part and a part that depends only on the symmetry
of the problem under consideration.
An important consequence of the matrix element
theorem concerns the definition of selection rules. An interaction
will be forbidden if the corresponding coupling coefficient in the
Wigner–Eckart theorem is zero. The conditions that control the zero
values of the coupling coefficients are called triangular conditions, since they
involve the combination of three irreps. Two kinds of triangular
conditions must be taken into account:
1.
Selectivity on the representations: an
interaction element is forbidden if the coupling of the three
irreps involved is zero, i.e. if the direct product of the operator
and ket parts does not include the irrep of the bra.
The triad of the three irreps may also be seen as a triple direct
product,
, where
the bra irrep appears in its complex-conjugate form. Equation
(6.8) can now
also be read as the character overlap between the totally-symmetric
irrep and the triple product. Accordingly, the selection rule of
Eq. (6.52)
can also be reformulated as: an interaction will be forbidden if
the triple product of the irreps does not contain the
totally-symmetric irrep.
![$$ \varOmega\notin\varLambda\times\varGamma $$](A303787_1_En_6_Chapter_Equ52.gif)
(6.52)
![$\bar{\varOmega} \times\varLambda\times\varGamma$](A303787_1_En_6_Chapter_IEq15.gif)
![$$ \varGamma_0 \notin\bar{\varOmega} \times\varLambda\times\varGamma $$](A303787_1_En_6_Chapter_Equ53.gif)
(6.53)
2.
Selectivity on the subrepresentations:
subrepresentations that are defined in a splitting field must obey
the triangular conditions for the subduced irreps in the
corresponding subgroup.
6.6 Application: The Jahn–Teller Effect
In 1937 Jahn and Teller made the claim that
degenerate states of molecules are intrinsically unstable
[6, 7].
Theorem 13
Non-linear molecules in a
spatially-degenerate
electronic state are subject to spontaneous
symmetry-breaking forces
that distort the molecule to a geometry of lower symmetry,
where the degeneracy is
removed.
The theorem is based on a perturbation of the
Hamiltonian by small displacements of the nuclei.
A high-symmetry geometry is chosen as the origin, and the
nuclear displacements are described by normal modes which transform
as irreps of the point group. The nuclear positions are parameters
in the electronic Hamiltonian. One has, to second-order:
The partial derivatives with respect to the normal modes will
affect only the electrostatic V Ne term in the Hamiltonian. These
operators are thus electrostatic one-electron operators. At the
coordinate origin, the electronic state is degenerate, and is
described by a set of wavefunctions, |Γ a γ a 〉, where Γ a is a degenerate irrep. The
energies as functions of the coordinates are obtained by
diagonalizing the Hamiltonian matrix,
, with elements:
The matrix in
is
also called the Jahn–Teller (JT) matrix. The linear terms in this
matrix are of type:
We have used the fact that the integration in this matrix element
runs over electronic coordinates, and does not affect the nuclear
coordinates. The Wigner–Eckart theorem can be applied to derive the
selection rules. Since the Hamiltonian is invariant under the
elements of the symmetry group, the transformation properties of
the operator part in this matrix element will be determined by the
partial derivatives, ∂/∂Q Γγ . As we have seen in
Sect. 1.3, a partial derivative in a
variable has the same transformation properties as the variable
itself.3 The operator
part is thus given by:
The coupling coefficient on the right-hand side of Eq.
(6.57)
restricts the symmetry of the nuclear displacements to the direct
square of the irrep of the electronic wavefunction. This selection
rule is made even more stringent by time-reversal symmetry. The
Hamiltonian is based on displacement of nuclear charges, and not on
momenta, so as an operator it is time-even or real.4 For spatially-degenerate irreps,
which are of the first kind, i.e. can be represented by real
functions, JT matrix elements can thus be chosen to be entirely
real, which implies:
Combining this result with Eq. (6.57) implies that the coupling coefficients,
to first-order, should obey:
In view of Eq. (6.31) this condition can be rewritten as:
The JT distortion modes are thus restricted to the symmetrized
square of the degenerate irrep of the electronic state, minus the
totally-symmetric modes, since these cannot lower the symmetry:
Modes that obey this selection rule, are said to be JT active. The evaluation of the
second-order matrix elements requires two steps. One first couples
the two distortion modes to a composite tensor operator:
|Ωω|.
The second-order matrix element then becomes:
The second-order elements thus are related to a product of two
3Γ symbols.5 A special element arises when
Ω is totally symmetric. In
this case, the coupling coefficients are given by:
The second-order expressions then are reduced to a diagonal matrix
element:
K Γ in this equation is the
harmonic force-constant. It gives rise to a constant diagonal term
which provides an attractive potential around the minimum and keeps
the surface bound at larger distances from the origin. The general
expression for the potential-energy surface then becomes:
Here, ε k (Q) represents the kth root of the Hamiltonian matrix.
This equation describes a surface with multiple sheets, one for
each root, which cross in the high-symmetry origin. In its simplest
form the Hamiltonian can be restricted to the linear terms only. In
the second-order approximation non-totally-symmetric second-order
terms will also be included.
![$$\begin{aligned} \begin{aligned} \mathcal{H} &= \mathcal{H}_0 + \mathcal{H}' \\ \mathcal{H}' &= \sum_{Q_{\varGamma\gamma}} \biggl( \frac{\partial{\mathcal {H}}}{\partial{Q_{\varGamma\gamma}}} \biggr)_0 Q_{\varGamma\gamma} + \frac{1}{2}\sum _{Q_{\varGamma\gamma}} \sum_{Q_{\varGamma'\gamma'}} \biggl(\frac{\partial^2{\mathcal{H}}}{\partial {Q_{\varGamma\gamma}} \partial Q_{\varGamma' \gamma'}} \biggr)_0 Q_{\varGamma \gamma}Q_{\varGamma'\gamma'} \end{aligned} \end{aligned}$$](A303787_1_En_6_Chapter_Equ54.gif)
(6.54)
![$\mathbb{H}$](A303787_1_En_6_Chapter_IEq16.gif)
![$$ H_{\gamma_a \gamma_b} = \langle{\varGamma_a} {\gamma_a} | \mathcal{H}|{\varGamma _a} {\gamma_b}\rangle= E_0 \delta_{\gamma_a \gamma_b} + \bigl\langle{\varGamma _a} {\gamma_a} |\mathcal{H}'|{\varGamma_a} { \gamma_b}\bigr\rangle $$](A303787_1_En_6_Chapter_Equ55.gif)
(6.55)
![$\mathcal{H}'$](A303787_1_En_6_Chapter_IEq17.gif)
![$$ \biggl\langle{\varGamma_a} {\gamma_a} \biggl \vert \biggl(\frac{\partial{\mathcal {H}}}{\partial{Q_{\varGamma\gamma}}} \biggr)_0 Q_{\varGamma\gamma}\biggr \vert { \varGamma_a} { \gamma_b} \biggr\rangle= \biggl\langle{ \varGamma_a} {\gamma_a} \biggl \vert \frac{\partial{\mathcal {H}}}{\partial{Q_{\varGamma\gamma}}} \biggr \vert {\varGamma_a} { \gamma_b} \biggr \rangle_0 Q_{\varGamma\gamma} $$](A303787_1_En_6_Chapter_Equ56.gif)
(6.56)
![$$ \biggl\langle{\varGamma_a} {\gamma_a} \biggl \vert \frac{\partial{\mathcal {H}}}{\partial{Q_{\varGamma\gamma}}} \biggr \vert {\varGamma_a} { \gamma_b} \biggr\rangle_0 = \langle{ \varGamma_a} \parallel \varGamma\parallel \varGamma_a\rangle \langle \varGamma_a \gamma_a | \varGamma\gamma \varGamma_a \gamma_b\rangle $$](A303787_1_En_6_Chapter_Equ57.gif)
(6.57)
![$$ \bigl\langle{\varGamma_a} {\gamma_a} | \mathcal{H}'|{\varGamma_a} {\gamma_b}\bigr \rangle = \bigl\langle{\varGamma_a} {\gamma_b} | \mathcal{H}'|{\varGamma_a} {\gamma_a}\bigr \rangle $$](A303787_1_En_6_Chapter_Equ58.gif)
(6.58)
![$$ \langle\varGamma_a \gamma_a | \varGamma\gamma \varGamma_a \gamma_b\rangle =\langle \varGamma_a \gamma_b | \varGamma\gamma \varGamma_a \gamma_a\rangle $$](A303787_1_En_6_Chapter_Equ59.gif)
(6.59)
![$$ \langle\varGamma_a \gamma_a \varGamma_a \gamma_b |\varGamma\gamma\rangle =\langle\varGamma_a \gamma_b \varGamma_a \gamma_a |\varGamma \gamma\rangle $$](A303787_1_En_6_Chapter_Equ60.gif)
(6.60)
![$$ \varGamma\in\bigl( [\varGamma_a \times\varGamma_a ] - \varGamma_0 \bigr) $$](A303787_1_En_6_Chapter_Equ61.gif)
(6.61)
![$$ \biggl \vert \frac{\partial^2{\mathcal{H}}}{\partial{Q_{\varGamma\gamma}} \partial Q_{\varGamma' \gamma'}}\biggr \vert =\sum _{\varOmega\omega}|{\varOmega\omega }|\bigl\langle\varOmega\omega| \varGamma\gamma\varGamma' \gamma' \bigr\rangle $$](A303787_1_En_6_Chapter_Equ62.gif)
(6.62)
![$$\begin{aligned} & \biggl\langle\varGamma_a \gamma_a\biggl \vert \frac{\partial^2{\mathcal {H}}}{\partial{Q_{\varGamma\gamma}} \partial Q_{\varGamma' \gamma'}}\biggr \vert \varGamma_a \gamma_b \biggr\rangle_0 \\ &\quad = \sum_{\varOmega\omega} \langle\varGamma_a \parallel \varOmega\parallel \varGamma_a \rangle \bigl\langle\varOmega\omega| \varGamma\gamma\varGamma' \gamma' \bigr\rangle \langle \varGamma_a \gamma_a |\varOmega\omega \varGamma_a \gamma_b \rangle \end{aligned}$$](A303787_1_En_6_Chapter_Equ63.gif)
(6.63)
![$$\begin{aligned} \begin{aligned} \bigl\langle\varGamma_0 |\varGamma\gamma \varGamma' \gamma' \bigr\rangle&= \frac {1}{\sqrt{\mathrm{dim}(\varGamma)}} \delta_{\varGamma\varGamma'} \delta_{\gamma \gamma'} \\ \langle\varGamma_a \gamma_a | \varGamma_0 \varGamma_a \gamma_b \rangle&= \delta _{\gamma_a \gamma_b} \end{aligned} \end{aligned}$$](A303787_1_En_6_Chapter_Equ64.gif)
(6.64)
![$$\begin{aligned} & \langle\varGamma_a \parallel \varGamma_0 \parallel \varGamma_a \rangle \bigl\langle \varGamma_0 |\varGamma\gamma\varGamma' \gamma' \bigr\rangle \langle\varGamma_a \gamma_b | \varGamma_0 \varGamma_a \gamma_b \rangle \\ &\quad = \frac{1}{\sqrt {\mathrm{dim}(\varGamma)}} \langle\varGamma_a \parallel \varGamma_0 \parallel \varGamma_a \rangle \delta_{\varGamma\varGamma'} \delta_{\gamma\gamma'} \delta _{\gamma_a \gamma_b} = K_{\varGamma} \delta_{\varGamma\varGamma'} \delta_{\gamma\gamma'} \delta _{\gamma_a \gamma_b} \end{aligned}$$](A303787_1_En_6_Chapter_Equ65.gif)
(6.65)
![$$ E_k(Q) = E_0 + \sum_{\varGamma} \frac{1}{2} K_{\varGamma} \biggl( \sum_{\gamma} Q_{\varGamma\gamma}^2 \biggr) +\varepsilon_k(Q) $$](A303787_1_En_6_Chapter_Equ66.gif)
(6.66)
The prototype of the JT surface is the celebrated
Mexican hat potential,
which describes the effect of the twofold-degenerate cubic or
trigonal E state.
A typical example is the 2 E g ground state of octahedral
Cu2+ complexes, with (t 2g )6(e g )3 configuration.
The JT-active mode in this case is restricted to an e g mode, corresponding to the
symmetrized square.
This distortion mode consists of the tetragonal and orthorhombic
stretchings, which we already encountered as vibrational modes of
UF6, and are depicted in Fig. 6.1. By use of the
appropriate 〈Ei|EjEk〉
coupling coefficients the JT matrix can easily be derived. The
force element is defined as:
The matrix then becomes:
To diagonalize this Hamiltonian, it is convenient to transform to
cylindrical coordinates {ρ,φ}:
Then the secular equation of the force element matrix in Eq.
(6.69)
becomes:
Two roots are found, which are independent of the angular
coordinate. The corresponding eigenfunctions are:
The surface consists of two sheets and exhibits rotational
symmetry.
![$$ [ E_g \times E_g ] - A_{1g}= E_g $$](A303787_1_En_6_Chapter_Equ67.gif)
(6.67)
![$$ F_E = \biggl\langle E \theta\bigg|\frac{\partial\mathcal{H}}{\partial Q_{\theta }}\bigg|E \theta\biggr \rangle $$](A303787_1_En_6_Chapter_Equ68.gif)
(6.68)
![$$ \mathbb{H} = \biggl( E_0 +\frac{1}{2} K_E \bigl(Q_{\theta}^2 + Q_{\epsilon}^2 \bigr) \biggr) \left [{ \begin{array}{c@{\quad}c} 1 & 0 \\ 0 & 1 \end{array} } \right ] + F_E \left [ { \begin{array}{c@{\quad}c} Q_{\theta} & Q_{\epsilon} \\ Q_{\epsilon} & -Q_{\theta} \end{array} } \right ] $$](A303787_1_En_6_Chapter_Equ69.gif)
(6.69)
![$$\begin{aligned} \begin{aligned} Q_{\theta} &= \rho\cos{\varphi} \\ Q_{\epsilon} &= \rho\sin{\varphi} \end{aligned} \end{aligned}$$](A303787_1_En_6_Chapter_Equ70.gif)
(6.70)
![$$ \varepsilon_k^2 - F_E^2 \rho^2\cos^2{\varphi} - F_E^2 \rho^2\sin ^2{\varphi} = 0 $$](A303787_1_En_6_Chapter_Equ71.gif)
(6.71)
![$$\begin{aligned} \begin{aligned} \epsilon_1 = F_E \rho &\longrightarrow | \psi_1\rangle= \cos\frac {\varphi}{2} |E\theta\rangle+ \sin \frac{\varphi}{2}|E\epsilon\rangle \\ \epsilon_2 = -F_E \rho & \longrightarrow | \psi_2\rangle= -\sin \frac{\varphi}{2} |E\theta\rangle+ \cos \frac{\varphi}{2} |E\epsilon \rangle \end{aligned} \end{aligned}$$](A303787_1_En_6_Chapter_Equ72.gif)
(6.72)
![$$\begin{aligned} E_{\pm} =& E_0 +\frac{1}{2} K_E \rho^2 \pm F_E \rho \\ =& E_0 + \frac{1}{2} K_E \bigl(Q_{\theta}^2 + Q_{\epsilon}^2 \bigr) \pm F_E \sqrt {Q_{\theta}^2 + Q_{\epsilon}^2} \end{aligned}$$](A303787_1_En_6_Chapter_Equ73.gif)
(6.73)
![A303787_1_En_6_Fig1_HTML.gif](A303787_1_En_6_Fig1_HTML.gif)
Fig. 6.1
The Mexican hat potential-energy surface of
the E×e linear JT problem. The nuclear
displacement coordinates are the tetragonal elongation,
Q θ , and the orthorhombic in-plane
distortion, Q
ϵ
A cross section of this surface looks like a
two-well potential, with two displaced parabolæ. The depth of the
well is called the JT stabilization energy:
In the 2D space of the active modes these parabolæ revolve around
the centre, giving rise to the Mexican hat appearance. At the
origin this surface has the shape of a conical intersection,
indicating that the high-symmetry point is unstable, and will
spontaneously relax to the circular trough surrounding the
degeneracy [8]. The distorted
system in the trough orbits around the origin. This motion is a
pseudo-rotation, i.e. it is
not a rotation of the molecular frame, but a gradual redistribution
of the distortions between the Cartesian directions. We illustrate
this in Fig. 6.2. The starting point is at φ=0, in the direction of the
Q θ mode. In this mode the
z-axis is elongated, and
the xy-plane is contracted.
A counterclockwise rotation activates the orthorhombic
Q ϵ mode, while the tetragonal
Q θ mode is receding. This
introduces a difference between the x- and y-axes: the distortion in the
x-direction becomes more
pronounced, while the y-axis contracts further. At the same
time the elongation of the z-axis diminishes. At an angle of
60∘ the x- and
z-axes are both elongated
to an equal extent, giving rise to a weak xz-plane and a strong y-axis. At the 90∘ point the
Q θ contribution vanishes, and the
distortion is orthorhombic, with a short y-axis, an elongated x-axis, and an undistorted z-axis. At an angle of 120∘
we reach the point where the x-axis is weak, and the perpendicular
yz-plane is strong. We thus
have regained an elongated tetragonal configuration, but the
elongation has been rotated from the z-axis to the x-axis. Continuing now at
240∘ we shall have travelled another third of the trough
and reoriented the tetragonal axis along the y-direction. We can also follow the
wavevector along the trough. If it is assumed that F E <0, the lower eigenfunction
will be the |ψ
1〉 eigenfunction of Eq. (6.72). In the starting
elongated tetragonal configuration the ground state coincides with
the |Eθ〉 basis function. By
the time we have reached the orthorhombic configuration at
φ=90∘, the
ground state has rotated by only half that angle and equals
.
At 180∘, we reach a structure which is tetragonally
compressed along the z-direction. Accordingly, the
Q θ mode has changed sign, in
contrast to the eigenfunction, where |Eθ〉 is replaced by |Eϵ〉. The observation of rotational
symmetry is an unexpected feature, which is not related to the
point group, but which stems from the limitation of the JT
Hamiltonian to linear terms. To describe this symmetry we first
reformulate the force element Hamiltonian in Eq. (6.69) in Dirac notation:
![$$ E_{JT} = -\frac{F_E^2}{2K_E} $$](A303787_1_En_6_Chapter_Equ74.gif)
(6.74)
![${1}/{\sqrt{2}} ( |E\theta \rangle+|E\epsilon\rangle )$](A303787_1_En_6_Chapter_IEq18.gif)
![$$ \mathcal{H}' = F_E \bigl( Q_{\theta} \bigl[ |E \theta\rangle \langle E\theta| - |E\epsilon\rangle \langle E\epsilon| \bigr] + Q_{\epsilon} \bigl[ |E\theta\rangle \langle E\epsilon| + |E\epsilon \rangle \langle E\theta| \bigr] \bigr) $$](A303787_1_En_6_Chapter_Equ75.gif)
(6.75)
![A303787_1_En_6_Fig2_HTML.gif](A303787_1_En_6_Fig2_HTML.gif)
Fig. 6.2
Rotation of the distortion in the trough of
the Mexican hat. Along the Q θ coordinate the complex is
elongated along its z-axis.
Rotation around the centre in the direction of Q ϵ will shorten the z-axis and increase the x-axis. At an angle of revolution of
120∘ a tetragonally elongated structure is again found,
but this time with the elongation along the x-direction, and similarly at
240∘, with the elongation along the y-axis
The angular momentum operator, corresponding to a
rotation in coordinate space, is given by:
The partial derivatives were obtained from Eq. (6.70). The commutator of
this operator with the Hamiltonian is:
Surprisingly, this commutator does not vanish. This is an important
observation, which directly points to the vibrational-electronic or
vibronic coupling between
the distortion modes and the electronic wavevector. When the system
rotates around the origin in coordinate space, not only are the
coordinates changing, but the wavevector is also rotating
simultaneously, so we must also provide an angular momentum
operator for a rotation in the function space (see [9]). We can construct this by analogy with Eq.
(6.76), but
with an important amendment: as we have argued while discussing
Fig. 6.2,
the coordinates rotate twice as quickly as the wavevector, and
hence a prefactor of 1/2 is required!
Only in this case does the total momentum operator
commute with
the Hamiltonian:
As the reader will have noticed, we have made use of the standard
spectroscopic symbols for orbital angular momentum, spin momentum,
and total momentum. Vibronic coupling is indeed analogous to
coupling of spin and orbit momenta in cylindrical molecules. To
form the vibronic wavefunction, describing the dynamics of the
Mexican hat system, the electronic state has to be combined with
nuclear wavefunctions. If the JT effect is pronounced, the vibronic
levels take the form of a radial oscillator, describing transverse
oscillations in the bottom of the through, and pseudo-rotational
levels, describing the longitudinal motion along the bottom of the
trough. The total vibronic wavefunction should of course be
single-valued after a full turn around the trough, which takes the
system back to the starting point. Hence, since the electronic part
changes sign after a full turn, the vibrational part should also
show a compensating sign change. This is indeed the case: the
pseudo-rotational levels are characterized by half-integral angular
momentum [10].
![$$ \mathcal{L} = \frac{\partial}{\partial\varphi} = \frac{\partial Q_{\epsilon}}{\partial\varphi} \frac{\partial}{\partial Q_{\epsilon}} + \frac{\partial Q_{\theta}}{\partial\varphi} \frac{\partial}{\partial Q_{\theta}} = Q_{\theta} \frac{\partial}{\partial Q_{\epsilon}} - Q_{\epsilon} \frac {\partial}{\partial Q_{\theta}} $$](A303787_1_En_6_Chapter_Equ76.gif)
(6.76)
![$$ \bigl[\mathcal{L},\mathcal{H}'\bigr] = F_E \bigl( - Q_{\epsilon} \bigl[ |E\theta\rangle \langle E \theta| - |E\epsilon\rangle \langle E \epsilon|\bigr] + Q_{\theta} \bigl[ |E \theta\rangle \langle E \epsilon| + |E\epsilon \rangle \langle E\theta| \bigr] \bigr) $$](A303787_1_En_6_Chapter_Equ77.gif)
(6.77)
![$$ \mathcal{S} =\frac{1}{2} \bigl( |E\theta\rangle \langle E\epsilon| - |E \epsilon\rangle \langle E \theta| \bigr) $$](A303787_1_En_6_Chapter_Equ78.gif)
(6.78)
![$\mathcal{J} = \mathcal{L} + \mathcal{S}$](A303787_1_En_6_Chapter_IEq19.gif)
![$$ \bigl[\mathcal{J}, \mathcal{H}'\bigr] =\bigl[\mathcal{L}, \mathcal{H}'\bigr]+\bigl[\mathcal{S}, \mathcal{H}' \bigr]= 0 $$](A303787_1_En_6_Chapter_Equ79.gif)
(6.79)
6.7 Application: Pseudo-Jahn–Teller interactions
Pseudo-JT interactions (PJT) refer to the
second-order vibronic coupling between electronic states which are
separated by a gap [11]. In this
section we describe the case of a non-degenerate ground state,
|ψ Σσ 〉, which is coupled to an
excited manifold. The Hamiltonian is identical to the expression in
Eq. (6.54).
Application of the selection rules shows that the diagonal
contribution,
,
is limited to the totally-symmetric operator associated with the
harmonic restoring-force, as in Eq. (6.66). Perturbation
theory further provides interactions between the ground and excited
states. These interactions are usually limited to first-order
contributions, which give rise to a quadratic coordinate
dependence. Hence one has, to second-order in the displacements:
where we have used the property that the Hamiltonian matrix is
hermitian. The selection rule in this process resides with the
matrix elements in the enumerator of the bilinear term. The
vibronic operator must couple ground and excited states; hence, it
is required that their triple direct product contains the
totally-symmetric irrep:
Applying the Wigner–Eckart theorem to the matrix element yields:
The sum over the λ
components of the excited state, transforming as the Λ irrep, can be simplified by using the
orthonormality property of the coupling coefficients from Eq.
(6.16).
Note that in case of a non-degenerate ground state the product
Γ×Σ yields only one irrep, since the norm
of the product character string equals the order of the group.
The summation over λ∈Λ thus covers the entire product space
of Γ×Σ. Combining the sum rule of Eq.
(6.83) with
the total expression for the PJT, one finds:
Hence, when the ground state is non-degenerate, the first-order
dependence of the energy on symmetry-lowering displacement
vanishes, and the second-order term contains two contributions: the
diagonal harmonic force constant, which is always positive, and the
bilinear relaxation term, which is always negative. If the excited
states are close in energy to the ground state, and if the vibronic
coupling is strong, the relaxation term may be dominant, and a
second-order symmetry-breaking effect will result. This is known as
the pseudo-JT effect. There are two main applications of this
effect: in geometry optimization, and in reaction dynamics.
![$\langle\psi_{\varSigma\sigma}|\mathcal{H}|\psi_{\varSigma \sigma}\rangle$](A303787_1_En_6_Chapter_IEq20.gif)
![$$\begin{aligned} E(Q) =& E_0 + \sum _{\varGamma} \frac{1}{2} K_{\varGamma} \biggl( \sum _{\gamma} Q_{\varGamma\gamma}^2 \biggr) \\ &{}+\sum_{\varLambda\lambda} \sum_{\varGamma\gamma} \frac{\vert \langle\psi _{\varLambda\lambda} \vert \frac{\partial{\mathcal{H}}}{\partial {Q_{\varGamma\gamma}}} \vert \psi_{\varSigma\sigma} \rangle_0 \vert ^2}{E_0-E_{\varLambda}} Q_{\varGamma\gamma}^2 \end{aligned}$$](A303787_1_En_6_Chapter_Equ80.gif)
(6.80)
![$$ \varGamma_0 \in\bar{\varGamma}_{\varLambda} \times\varGamma\times \varSigma $$](A303787_1_En_6_Chapter_Equ81.gif)
(6.81)
![$$ \biggl\langle\psi_{\varLambda\lambda}\biggl \vert \frac{\partial{\mathcal {H}}}{\partial{Q_{\varGamma\gamma}}} \biggr \vert \psi_{\varSigma\sigma}\biggr\rangle _0 = \langle\varLambda\lambda| \varGamma\gamma\varSigma\sigma\rangle \langle\varLambda\parallel \varGamma\parallel \varSigma \rangle $$](A303787_1_En_6_Chapter_Equ82.gif)
(6.82)
![$$\begin{aligned} \sum_{\lambda\in\varLambda} \frac{\vert \langle\psi_{\varLambda\lambda} \vert \frac{\partial{\mathcal{H}}}{\partial{Q_{\varGamma\gamma}}} \vert \psi_{\varSigma\sigma}\rangle_0 \vert ^2}{ E_0-E_{\varLambda}} =& \sum _{\lambda\in\varLambda} \frac{\vert \langle\varLambda\lambda| \varGamma\gamma\varSigma\sigma\rangle\langle\varLambda\parallel \varGamma\parallel \varSigma \rangle \vert ^2}{ E_0-E_\varLambda} \\ =& \frac{\vert \langle\varLambda\parallel \varGamma\parallel \varSigma\rangle \vert ^2}{E_0 - E_\varLambda} \end{aligned}$$](A303787_1_En_6_Chapter_Equ83.gif)
(6.83)
![$$ \bigl\langle\chi^{\varGamma\times\varSigma}| \chi^{\varGamma\times\varSigma}\bigr\rangle= \bigl\langle \chi^{\varGamma} \chi^{\varSigma}| \chi^{\varGamma} \chi^{\varSigma} \bigr\rangle = \bigl\langle\chi^{\varGamma}| \chi^{\varGamma}\bigr\rangle= |G| $$](A303787_1_En_6_Chapter_Equ84.gif)
(6.84)
![$$ E(Q) = E_0 + \sum_{\varGamma} \biggl\{ \frac{1}{2} K_{\varGamma} + \sum_{\varLambda} \frac{\vert \langle\varLambda\parallel \varGamma\parallel \varSigma\rangle \vert ^2}{E_0 - E_\varLambda} \biggr\} \biggl( \sum_{\gamma} Q_{\varGamma\gamma}^2 \biggr) $$](A303787_1_En_6_Chapter_Equ85.gif)
(6.85)
In reaction dynamics the PJT may be responsible
for stereoselectivity, because of the selection rules for vibronic
coupling matrix elements. Via these relaxation matrix elements the
Wigner–Eckart theorem is at the basis of the Woodward–Hoffmann
rules [12]. We shall not discuss
these rules in general, but consider some simple illustrations,
related to electrocyclic reactions.6 Take as a simple example the ring
closure of cis-butadiene,
as illustrated in Fig. 6.3. The relevant occupied orbitals are the
π-bonds in the reagent, and
the remaining π- and newly
formed σ-bonds in the
product. As the diagram shows, in the common C 2v point group there is a mismatch
between the symmetries. In order for the reaction to occur, the
reaction coordinate has to reduce the symmetry so that the
a 2-orbital can
interchange with an a
1-orbital. This interchange is taking place via a PJT
mechanism which couples the a 2 occupied orbital to an
a 1 virtual
orbital in the reagent. As the reaction coordinate proceeds, this
coupling is intensified and leads to an interchange of both. The
relevant matrix element is thus an orbital vibronic coupling
element:
Hence, a distortion coordinate is required which transforms as
a 2×a 1=a 2. The coordinate with
this symmetry is the one that destroys the symmetry planes but
keeps the twofold axis. This is typically a conrotatory reaction, where the
extremal carbon atoms rotate simultaneously in the same sense, to
form the σ-bond. Ring
closure of substituted butadienes thus follows a conrotatory
reaction stereochemistry, at least if the reaction is concerted.
![$$ \biggl\langle a_2 \biggl \vert \frac{\partial{\mathcal{H}}}{\partial{Q_{\varGamma \gamma}}} \biggr \vert a_1 \biggr\rangle\neq0 $$](A303787_1_En_6_Chapter_Equ86.gif)
(6.86)
![A303787_1_En_6_Fig3_HTML.gif](A303787_1_En_6_Fig3_HTML.gif)
Fig. 6.3
Ring closure of cis-butadiene to cyclo-butene. In C 2v symmetry there is a symmetry
mismatch between the a
2 and a
1 occupied orbitals. Vibronic orbital coupling requires
a concerted mechanism, based on a conrotatory ring closure, which
conserves only the
axis
![$\hat{C}_{2}$](A303787_1_En_6_Chapter_IEq21.gif)
This ring-closure selection rule is further
confirmed by the closure reaction for the cis-1,3,5 hexatriene to
1,3-cyclohexadiene, as illustrated in Fig. 6.4. Here,
a b
1-orbital has to interchange with a virtual orbital of
a 1 symmetry.
The selection takes thus place at the level of the orbital matrix
element:
Clearly, the distortion coordinate should now be of b 1×a 1=b 1 symmetry, and this
corresponds to the disrotatory mode, which destroys the
axis
but keeps the vertical reflection plane.
![$$ \biggl\langle b_1 \biggl \vert \frac{\partial{\mathcal{H}}}{\partial{Q_{\varGamma \gamma}}} \biggr \vert a_1 \biggr\rangle\neq0 $$](A303787_1_En_6_Chapter_Equ87.gif)
(6.87)
![$\hat{C}_{2}$](A303787_1_En_6_Chapter_IEq22.gif)
![A303787_1_En_6_Fig4_HTML.gif](A303787_1_En_6_Fig4_HTML.gif)
Fig. 6.4
Ring closure of cis-1,3,5-hexatriene to cyclo-hexadiene. In C 2v symmetry there is a symmetry
mismatch between the b
1 and a
1 occupied orbitals. Vibronic orbital coupling requires
a concerted mechanism, based on a disrotatory ring closure, which
conserves only the
plane
![$\hat{\sigma}_{1}$](A303787_1_En_6_Chapter_IEq23.gif)
6.8 Application: Linear and Circular Dichroism
Selection rules are of primary importance in
spectroscopy, where they provide direct evidence concerning the
nature of excited states. As an application, we study the linear
and circular dichroism of tris-chelate transition-metal complexes
[14]. The prototype is a divalent
ruthenium complex with three 2,2′-bipyridyl ligands, which is an
important chromophore for energy conversion. In this section we
shall describe the charge transfer and intra-ligand transitions of
this type of complex. The linear dichroism (LD) spectrum measures
the absorption of the chromophore under plane-polarized incident
light for different orientations of the polarization with respect
to the molecular frame. This requires that the molecules should be
embedded in an oriented phase, such as a crystalline host. Circular
dichroism (CD) measures the difference in absorption between left
and right circularly polarized light. Since this is based on the
intrinsic helicity of the molecule, it can be performed in
non-oriented medium, such as a solution.
As always, we start the treatment by making a
simple sketch of the structure. Two sets of Cartesian axes are
relevant. In the usual octahedral coordinate system the
x,y,z-axes coincide with the metal-ligand
bond directions, assuming that the ligator atoms form a perfect
octahedron. In addition, in Fig. 3.6 of Chap. 3 a primed x′,y′,z′-coordinate system was introduced,
which is adapted to the tris-chelate geometry. The z′-axis is along the threefold
direction, and the x′ axis
is oriented along a twofold axis, coinciding with the bisector of
the positive x and the
negative y axes. Next, we
determine the point group, which in the present case is
D 3. This is a
rotational group, which implies that the molecule is chiral. The
figure shows the Δ-enantiomer.7 Thirdly, we define the functional
basis. The relevant orbitals are the metal t 2g orbitals, which are fully
occupied in the Ru2+ ground state, and the frontier
orbitals on the ligand. For conjugated bidentate ligands, such as
2,2′-bipyridyl (bipy) or 2,4-pentanedionate (also named
“acetylacetonate”, acac−), the frontier orbitals are of
π-type. The essential parts
of these orbitals are the contributions on the ligator atoms. These
are either symmetric or antisymmetric with respect to the twofold
axis through the bidentate ligand, as shown in
Fig. 6.5.
Following Orgel, we denote them as χ- or ψ-type, respectively [15]. The standard techniques of characters and
projection operators yield SALCs for all these basis sets. The
results are shown in Table 6.4. For the e-irrep the components are labelled as
e θ and e ϵ , following the standard
canonical format. As a splitting field we use the twofold axis
along x′. Finally, we
also include in the table the symmetries of the transition dipoles,
which are the operators for the optical transitions. This completes
the groundwork for the symmetry analysis.
![A303787_1_En_6_Fig5_HTML.gif](A303787_1_En_6_Fig5_HTML.gif)
Fig. 6.5
Δ-enantiomer of tris-chelate octahedral
complex of D 3
symmetry. The xyz-coordinate system passes through
the ligator atoms; the primed coordinate system has the
z′-direction along the
threefold axis, and x′ on a
twofold axis through ligand A. The ligand orbital shown on the
right is of ψ-type: it is
antisymmetric under a rotation about the twofold axis through the
ligator bridge. The ψ
orbital on ligand A
interacts with the
-combination on
the metal
![$\frac{1}{\sqrt{2}} ( d_{xz} - d_{yz} )$](A303787_1_En_6_Chapter_IEq24.gif)
Table 6.4
Symmetry-adapted zeroth-order metal and
ligand orbital functions
t
2g
-orbitals
|
dipole moments
|
---|---|
![]() |
a
2:μ
z′
|
![]() |
e
θ :μ x′
|
![]() |
e
ϵ :μ y′
|
ψ-orbitals
|
χ-orbitals
|
---|---|
![]() |
![]() |
![]() |
![]() |
![]() |
![]() |
Linear Dichroism
The linear dichroism is associated with the
metal-to-ligand charge-transfer (CT) transitions [16]. Dipole-allowed transitions between the
orbitals are governed by the appropriate D 3 coupling coefficients.
However, since both donor and acceptor orbitals, as well as the
transition operators, each involve two irreps, several
symmetry-independent coupling channels are possible. As is often
the case in transition-metal spectroscopy, it is not sufficient to
identify the reduced matrix elements; for a deeper understanding a
further development of the model is often required to compare the
reduced matrix elements. In the case of the CT bands the model of
Day and Sanders offers just that little extra [17]. According to this simple model, a
charge-transfer (CT) transition between metal and ligand gains
intensity when the relevant metal and ligand orbitals
interact.
We first calculate the interaction terms between
the metal and isolated ligand orbitals. The bipy ligand has
low-lying unoccupied levels of ψ-character, which form π-acceptor interactions with the metal
t 2g orbitals. Let H π represent the elementary
interaction between a ligand ψ orbital and a metal t 2g orbital, directed towards one
ligator. The allowed interactions are then obtained by cyclic
permutation:
In order to apply the model of Day and Sanders, we now consider the
CT transition between the ligand orbital on A and the t 2g combination that interacts with
it. As shown in Fig. 6.5, the ψ A -acceptor orbital is
antisymmetric with respect to the
axis and antisymmetric in the
xy-plane. The only matching
t 2g combination on the metal is the
|e ϵ (t 2g )〉 component (see
Table 6.4). In the local C 2v symmetry, |ψ A 〉 and |e ϵ (t 2g )〉 both transform as
b 2 (taking the
horizontal plane as the local
). Their interaction element is
expressed as:
We now consider the transition dipole moment between these orbitals
along the x′ direction,
with μ x′=−ex′. In C 2v symmetry this component
transforms as a
1, while μ
y′ and
μ z′ are antisymmetric with respect
to the
axis. According to the Wigner–Eckart theorem, a transition dipole
between two b 2
orbitals must transform as the direct product b 2×b 2=a 1; hence, only the
x′- component will be
dipole-allowed. In a perturbative approach, which takes into
account the symmetry-allowed interaction between the metal and
ligand orbitals, one has:
In this expression the first term is the contact term between the zeroth-order
orbitals. The second term is the transfer term, arising from the
interaction between the donor and acceptor orbitals. In the
simplified model of Day and Sanders this term is the dominant
contribution. The transfer-dipole matrix element in Eq.
(6.90) is
approximated as the dipole length of the transferred charge, which
we will represent as μ
A .
where R A is the radius vector from the
origin to the centre of ligand A, with length ρ. Since the three ligands are
equivalent, we further write:
The three vectors of the ligand positions can be expressed in a row
notation for the primed x′,y′,z′ coordinate system as:
The transfer term then becomes:
where the parameter κ is an
overlap factor which indicates what fraction of the charge is
actually transferred:
Note that the transfer term is always polarized in the direction of
the transferred charge.
![$$\begin{aligned} H_{\pi} =& \bigl\langle d_{xz}|\mathcal{H}| \psi^A\bigr\rangle= - \bigl\langle d_{yz}|\mathcal{H}| \psi^A \bigr\rangle \\ =& \bigl\langle d_{xy}|\mathcal{H}|\psi^B\bigr\rangle= - \bigl\langle d_{xz}|\mathcal {H}|\psi^B \bigr\rangle \\ =& \bigl\langle d_{yz}|\mathcal{H}|\psi^C\bigr\rangle= - \bigl\langle d_{xy}|\mathcal {H}|\psi^C \bigr\rangle \end{aligned}$$](A303787_1_En_6_Chapter_Equ88.gif)
(6.88)
![$\hat {C}_{2}^{x'}$](A303787_1_En_6_Chapter_IEq34.gif)
![$\hat{\sigma}_{1}$](A303787_1_En_6_Chapter_IEq35.gif)
![$$ \bigl\langle e_{\epsilon}(t_{2g})|\mathcal{H}|\psi^A \bigr\rangle= \frac{1}{\sqrt{2}} \bigl\langle d_{xz} -d_{yz}|\mathcal{H}|\psi^A\bigr\rangle= \sqrt{2} H_{\pi} $$](A303787_1_En_6_Chapter_Equ89.gif)
(6.89)
![$\hat{C}_{2}^{x'}$](A303787_1_En_6_Chapter_IEq36.gif)
![$$ \mu\bigl(e_\epsilon(t_{2g}) \rightarrow\psi^A\bigr) = \bigl\langle e_{\epsilon }(t_{2g})|\mu_{x'}| \psi^A\bigr\rangle - \frac{\langle e_{\epsilon }(t_{2g})|\mathcal{H}|\psi^A\rangle}{E_{\psi} - E_{t_{2g}}} \bigl\langle \psi^A|\mu_{x'}|\psi^A \bigr\rangle $$](A303787_1_En_6_Chapter_Equ90.gif)
(6.90)
![$$ \bigl\langle\psi^A|\mu_{x'}|\psi^A \bigr \rangle= - \mathrm{e} \bigl\langle\psi ^A|{x'}| \psi^A \bigr\rangle \approx-\mathrm{e} |\mathbf{R}_A|=- \mathrm{e} \rho\equiv\mu_A $$](A303787_1_En_6_Chapter_Equ91.gif)
(6.91)
![$$ \mu_A = \mu_B = \mu_C \equiv \mu^{\perp} $$](A303787_1_En_6_Chapter_Equ92.gif)
(6.92)
![$$\begin{aligned} \begin{aligned} \boldsymbol{R}_A &= \rho ( 1,0,0 ) \\ \boldsymbol{R}_B &= \rho \biggl( -\frac{1}{2}, \frac{\sqrt{3}}{2}, 0 \biggr) \\ \boldsymbol{R}_C &= \rho \biggl( -\frac{1}{2},- \frac{\sqrt{3}}{2}, 0 \biggr) \end{aligned} \end{aligned}$$](A303787_1_En_6_Chapter_Equ93.gif)
(6.93)
![$$ \boldsymbol{\mu} \bigl(e_\epsilon(t_{2g}) \rightarrow \psi^A\bigr) = -\mathrm{e} \kappa\mathbf{R}_A = \kappa \boldsymbol{\mu}_A $$](A303787_1_En_6_Chapter_Equ94.gif)
(6.94)
![$$ \kappa= -\frac{\langle e_\epsilon(t_{2g})|\mathcal{H}|\psi^A\rangle }{E_{\psi} - E_{t_{2g}}} =-\frac{\sqrt{2} H_{\pi}}{E_{\psi} - E_{t_{2g}}} $$](A303787_1_En_6_Chapter_Equ95.gif)
(6.95)
This parametrization can now be used to calculate
the transfer term for the relevant trigonal transitions. The
Hamiltonian operator is of course totally symmetric, so allowed
interactions can take place only between orbitals with the same
symmetry, and are independent of the component; hence:
![$$\begin{aligned} \begin{aligned} \bigl\langle e_{\epsilon}(t_{2g})| \mathcal{H}|e_{\epsilon}(\psi)\bigr\rangle &= \sqrt{3} H_{\pi} \\ \bigl\langle e_{\theta}(t_{2g})|\mathcal{H}|e_{\theta}( \psi)\bigr\rangle &= \sqrt{3} H_{\pi} \end{aligned} \end{aligned}$$](A303787_1_En_6_Chapter_Equ96.gif)
(6.96)
Symmetry prevents interaction between the
a 1(t 2g ) and a 2(ψ) orbitals. The metal-ligand
π acceptor interaction will
thus stabilize the e-component of the t 2g shell, while leaving the
a 1-orbital in
place, as shown in the simple orbital-energy diagram in the left
panel of Fig. 6.6. We can now calculate the transfer term for
the e→e and e→a 2 orbital transitions. In
each case only one component needs to be calculated. The
interaction element in this case is obtained from Eq. (6.96) and the transfer
fraction reads:
The transfer-dipole element is given by:
Here, we have made use of the fact that the sum of the three dipole
vectors vanishes. The effective transfer term thus becomes:
In the Wigner–Eckart formalism, this matrix element is written as:
The coupling coefficient in this equation is equal
to
. We
can thus identify the reduced matrix element as:
All other e→e transfer terms can then be obtained
by simply varying the coupling coefficients. We give one more
example of a transition that requires an operator which is
μ y′ polarized:
The vector μ
B −μ C in this expression is directed
in the μ y′ direction, as required by the
selection rule. Moreover, the length of this vector is
:
Hence, the transfer-dipole length for this y′-polarized transition also measures
, which is exactly the same as for the
x′-polarized transition,
given in Eq. (6.99). This is expected since the corresponding
coupling coefficients, 〈Eϵ|EθEϵ〉 and 〈Eθ|EϵEϵ〉, are equal.
![$$ -\frac{\sqrt{3} H_{\pi}}{E_{\psi} - E_{t_{2g}}} = \sqrt{\frac {3}{2}}\kappa $$](A303787_1_En_6_Chapter_Equ97.gif)
(6.97)
![$$ \frac{1}{6} {\bigl\langle} 2 \psi^A - \psi^B - \psi^C {|}\boldsymbol {\mu} {|}2 \psi^A - \psi^B - \psi^C {\bigr\rangle} = \frac{1}{6} [ 4 \boldsymbol{\mu}_A + \boldsymbol{\mu}_B + \boldsymbol{ \mu}_C ] = \frac{1}{2} \boldsymbol{\mu}_A $$](A303787_1_En_6_Chapter_Equ98.gif)
(6.98)
![$$ \boldsymbol{\mu} \bigl(e_\epsilon(t_{2g}) \rightarrow e_{\epsilon} (\psi)\bigr) = \sqrt{\frac{3}{8}} \kappa\boldsymbol{ \mu}_A $$](A303787_1_En_6_Chapter_Equ99.gif)
(6.99)
![$$ \bigl\langle e_{\epsilon}(t_{2g})|\mu_{x'}|e_{\epsilon}( \psi)\bigr\rangle = \langle E \epsilon| E \theta E\epsilon\rangle \bigl\langle e(t_{2g})\parallel e(\mu)\parallel e(\psi) \bigr\rangle $$](A303787_1_En_6_Chapter_Equ100.gif)
(6.100)
![$1/\sqrt{2}$](A303787_1_En_6_Chapter_IEq37.gif)
![$$ \bigl\langle e(t_{2g})\parallel e(\mu)\parallel e(\psi) \bigr\rangle= \frac{\sqrt{3}}{2} \kappa{\mu}_A $$](A303787_1_En_6_Chapter_Equ101.gif)
(6.101)
![$$\begin{aligned} \bigl\langle e_{\theta}(t_{2g}) |\boldsymbol{\mu} | e_{\epsilon}(\psi )\bigr\rangle =& \frac{1}{\sqrt{8}}\kappa {\bigl\langle} \psi^C - \psi^B {|}\boldsymbol{\mu} {|}2 \psi^A - \psi^B - \psi^C {\bigr\rangle} \\ =& \frac{1}{\sqrt{8}}\kappa ( \boldsymbol{\mu}_B - \boldsymbol{\mu }_C ) \end{aligned}$$](A303787_1_En_6_Chapter_Equ102.gif)
(6.102)
![$\sqrt{3} \mu^{\perp}$](A303787_1_En_6_Chapter_IEq38.gif)
![$$ ( \boldsymbol{\mu}_B - \boldsymbol{\mu}_C ) \cdot ( \boldsymbol{\mu}_B - \boldsymbol{\mu}_C ) = 2 \bigl( \mu^{\perp } \bigr)^2 - 2 \boldsymbol{\mu}_B \cdot \boldsymbol{\mu}_C = 3 \bigl(\mu^{\perp} \bigr)^2 $$](A303787_1_En_6_Chapter_Equ103.gif)
(6.103)
![$\sqrt{3/8}\kappa$](A303787_1_En_6_Chapter_IEq39.gif)
![A303787_1_En_6_Fig6_HTML.gif](A303787_1_En_6_Fig6_HTML.gif)
Fig. 6.6
Allowed CT transitions from the
t 2g shell to ψ- or χ-type ligand acceptor orbitals for
tris-chelate complexes with D 3 symmetry
Using the transfer model, we can also express the
reduced matrix elements for the e→a 2 channel. Even though
there is no overlap between these orbitals, they do give rise to a
transfer-term intensity. Orbital interaction does indeed delocalize
the e(t 2g ) orbitals over the ligands.
The dipole operators, centred on the complex origin, will then
couple the e(ψ) and a 2(ψ) ligand-centred orbitals. Hence, we
write:
The dipole matrix element in this expression can easily be
evaluated:
The total transfer term is obtained by combining Eqs. (6.104) and (6.105):
A final task is to calculate the transition-moments between the
corresponding multi-electronic states based on the
orbital-transition moments obtained. In the tris-chelate complex
under consideration, a 1 A 1→1
E state transition can be
associated with each allowed orbital-transition. The 1
A 1 corresponds
to the closed-shell ground state, based on the (t 2g )6 configuration.
Both the e→a 2 and e→e transitions will give rise to a
twofold-degenerate 1 E state. As an example, the
θ states are written in
determinantal notation as follows, where we write only the orbitals
that are singly occupied:
The resulting state transition-moments are then expressed in terms
of orbital transition-moments as:
On the other hand, the a
1(t
2g ) orbital
does not delocalize over the ligands. As a result, there can be no
transfer term associated with transitions from this orbital. One
expects only a weak contact term. The lowest transition corresponds
to a
1(t
2g
)→a
2(ψ). The only
non-zero coupling coefficient for this transition is 〈A 1|A 2 A 2〉. This transition will
thus be dipole allowed under μ z′. Polarized absorption spectra
are in line with this analysis: the spectral onset of the CT region
is characterized by a weak absorption band in parallel
polarization, followed by two strong absorption bands in
perpendicular polarization. This assignment is based on the
assumption that the vertical Franck–Condon excitations reach
delocalized charge-transfer states. At least in the case of
, this is supported by
detailed spectral measurements [18]. An entirely similar analysis can be
performed in the case when the ligand orbital is of χ-type. The transition-moments are
collected in Table 6.5. In this case, the ligand and metal part
both transform as a
1+e (see
Table 6.4). As a result, three transitions are found
to carry transfer-term intensity, as indicated in
Fig. 6.6.
![$$\begin{aligned} \boldsymbol{\mu} \bigl(e_\epsilon(t_{2g}) \rightarrow a_{2} (\psi)\bigr) =& - \frac{\langle e_{\epsilon}(t_{2g})|\mathcal{H}|e_{\epsilon}(\psi )\rangle}{E_{\psi} - E_{t_{2g}}} \bigl\langle e_{\epsilon}(\psi)|\boldsymbol{\mu}|a_{2}(\psi) \bigr\rangle \\ =& \sqrt{\frac{3}{2}} \kappa\bigl\langle e_{\epsilon}(\psi)| \boldsymbol{\mu }|a_{2}(\psi) \bigr\rangle \end{aligned}$$](A303787_1_En_6_Chapter_Equ104.gif)
(6.104)
![$$\begin{aligned} \bigl\langle e_{\epsilon}(\psi)|\boldsymbol{\mu}|a_{2}(\psi) \bigr\rangle =& \frac{1}{3\sqrt{2}} {\bigl\langle} 2 \psi^A - \psi^B - \psi^C {|}\boldsymbol{\mu} {|} \psi^A + \psi^B + \psi^C {\bigr\rangle} \\ =& \frac{1}{3\sqrt{2}} ( 2 \boldsymbol{\mu}_A - \boldsymbol{\mu }_B - \boldsymbol{\mu}_C ) \\ =& \frac{1}{\sqrt{2}} \boldsymbol{\mu}_A \end{aligned}$$](A303787_1_En_6_Chapter_Equ105.gif)
(6.105)
![$$ \boldsymbol{\mu} \bigl(e_\epsilon(t_{2g}) \rightarrow a_{2} (\psi)\bigr) = \frac{\sqrt{3}}{2}\kappa\boldsymbol{ \mu}_A $$](A303787_1_En_6_Chapter_Equ106.gif)
(6.106)
![$$\begin{aligned} \begin{aligned}\big |{}^1E_{\theta}(e\rightarrow a_2) \big\rangle&= \frac{1}{\sqrt{2}} \bigl[ \big|\bigl(e_\epsilon(t_{2g}) \alpha\bigr) \bigl(a_2 (\psi) \beta\bigr)\big| -\big|\bigl(e_\epsilon(t_{2g}) \beta\bigr) \bigl(a_2 (\psi) \alpha\bigr)\big| \bigr] \\ \big|{}^1E_{\theta}(e\rightarrow e)\big\rangle&= \frac{1}{2} \bigl[ \big|\bigl(e_\theta(t_{2g}) \alpha\bigr) \bigl(e_{\theta} (\psi) \beta\bigr)\big| -\big|\bigl(e_\theta (t_{2g}) \beta\bigr) \bigl(e_{\theta}(\psi) \alpha\bigr)\big| \\ & \quad \ \ {}- \big|\bigl(e_\epsilon(t_{2g}) \alpha\bigr) \bigl(e_{\epsilon} (\psi) \beta\bigr)\big| +\big|\bigl(e_\epsilon(t_{2g}) \beta\bigr) \bigl(e_{\epsilon}(\psi) \alpha\bigr)\big| \bigr] \end{aligned} \end{aligned}$$](A303787_1_En_6_Chapter_Equ107.gif)
(6.107)
![$$\begin{aligned} \begin{aligned} \bigl\langle{{}^1A}_1 |\boldsymbol{\mu}| {{}^1E}_{\theta}(e\rightarrow a_2)\bigr\rangle&= \sqrt{2} \bigl\langle e_{\epsilon} (t_2g)|\boldsymbol{\mu} | a_2(\psi )\bigr\rangle= \sqrt{3/2} \kappa \\ \bigl\langle{{}^1A}_1 |\boldsymbol{\mu}| {{}^1E}_{\theta}(e\rightarrow e)\bigr\rangle&= 2 \bigl\langle e_{\theta} (t_2g)|\boldsymbol{\mu} | e_\theta(\psi) \bigr\rangle= \sqrt{3/2} \kappa \end{aligned} \end{aligned}$$](A303787_1_En_6_Chapter_Equ108.gif)
(6.108)
![$\mathrm {Ru(bipy)}_{3}^{2+}$](A303787_1_En_6_Chapter_IEq40.gif)
Table 6.5
Transfer-term contributions to 1
A 1→1
E CT transitions, with
ψ and χ acceptor orbitals
ψ
ligand orbitals
|
χ
ligand orbitals
|
||
---|---|---|---|
1 E(a 1→e(ψ))
|
0
|
1 E(a 1→e(χ))
|
![]() |
1 E(e→a 2(ψ))
|
![]() |
1 E(e→a 1(χ))
|
![]() |
1 E(e→e(ψ))
|
![]() |
1 E(e→e(χ))
|
![]() |
Circular Dichroism
The tris-chelate compounds are chiral compounds,
with an apparent helical structure, which can easily be related to
their circular-dichroic properties by use of symmetry selection
rules. The CT transitions that we have just discussed cannot be
responsible for the primary CD strength, since they are in-plane
polarized, and thus do not carry intrinsic helicity. Instead, the
prominent peaks in the CD spectrum are observed at higher energies,
and are associated with the intra-ligand ππ ∗-transitions. These
transitions take place between occupied and virtual ligand-centred
orbitals which are of opposite signature, and hence are of type
ψ→χ or vice-versa. Such transitions are
long-axis polarized, i.e. the transition dipole moment is oriented
along the ligand bridge as shown in Fig. 6.7.
![A303787_1_En_6_Fig7_HTML.gif](A303787_1_En_6_Fig7_HTML.gif)
Fig. 6.7
Allowed intra-ligand transitions from
χ- to ψ-type ligand orbitals for tris-chelate
complexes with D
3 symmetry. The circular dichroism has a lower
right-circularly polarized (rcp) band and an upper left-circularly
polarized (lcp) band. This gives the CD spectrum the appearance of
the first derivative of a Gaussian curve, with a negative part at
longer wavelength and a positive part at shorter wavelength
We designate these dipole moments as
.
These vectors can be expressed in a row notation for the primed
x′,y′,z′ coordinate system as follows:
The scalar products between these orientations are equal to 1/2,
which corresponds to angles of 60∘. Each of the three
transitions gives rise to an excited state. In D 3 symmetry these states
transform as A
2+E. The
composition of these exciton states8 is as follows:
Here, the notation refers to a singlet orbital transition, which
can be written in determinantal form as:
To first approximation, the metal centre is not taking part in the
electronic properties, but merely serves as a structural template
which keeps the ligands in place. Distant interactions between the
three transitions can be described by a simple exciton-coupling
model. In this model, the interaction between transitions is
approximated by the electrostatic interaction potential between the
corresponding transition dipoles. This potential is given by:
where R ij is the distance between the
dipoles, and R
ij =R j −R i . The length of the distance
vector is thus
.
The energies of the exciton states are then given by:
The 1 A
2 state thus goes up in energy twice as much as the
1 E state goes
down, thus keeping the barycentre energy at the zeroth-order
position. Now, in order to determine the CD strength, we need for
the two states both the electric and the magnetic transition
dipoles from the ground state. The electric dipoles are easily
obtained by combining the state vectors:
The calculation of the magnetic transition dipoles requires a
preamble. The magnetic moment was already defined in
Eq. (4.128) of Chap. 4. By explicitly writing the angular
momentum operator in terms of the linear momentum operator as
r×p one obtains:
The commutator of the one-electron Hamiltonian with the position
operator is given by:
Here, we used the Heisenberg commutator relation between the
conjugate position and momentum operators: [x,p x ]=iħ. The magnetic moment matrix element
of the intra-ligand transition with respect to the common origin of
the coordinate system is given by:
where it was assumed that the chromophore has no intrinsic magnetic
transition-moment. The momentum matrix element in this equation can
now be evaluated with the help of Eq. (6.116):
Here, ν is the frequency of
the intra-ligand transition. The combination of this result with
Eq. (6.117)
yields:
As we indicated the above formalism applies to chromophores that
have no intrinsic magnetic moment.
A transition will be characterized by a helical displacement of the
electron if the magnetic and electric transition dipoles are
aligned. This is reflected in the Rosenfeld equation for the CD
intensity or rotatory
strength,
, for a transition from a
ground state a to an
excited state j in a
collection of randomly-oriented molecules:
Straightforward application to the exciton bands yields:
The out-of-plane polarized transition to the 1
A 2 state, which
lies at higher energy, has a positive CD signal, while the in-plane
polarized transition to the lower 1 E state has a negative CD signal. The
latter transition consists of two components along the two in-plane
directions. Summing over the three components in Eq. (6.122), shows that the
total rotatory strength, for randomly-oriented molecules, is
exactly zero. This is a general sum rule for CD spectra. If one now
takes the spectrum of the chiral antipode, the Λ tris-chelate complex, the spectra are
exactly the same but the signs are reversed. Mirror image in actual
geometry thus becomes reflection symmetry in the spectrum.
![$\boldsymbol{\mu}_{A}^{\parallel}, \boldsymbol{\mu}_{B}^{\parallel}, \boldsymbol{\mu}_{C}^{\parallel}$](A303787_1_En_6_Chapter_IEq46.gif)
![$$\begin{aligned} \begin{aligned} \boldsymbol{\mu}_A^{\parallel} &= \mu^{\parallel} \biggl( 0,\frac {1}{\sqrt{3}} ,\frac{\sqrt{2}}{\sqrt{3}} \biggr) \\ \boldsymbol{\mu}_B^{\parallel} &= \mu^{\parallel} \biggl( - \frac {1}{2},-\frac{1}{2\sqrt{3}}, \frac{\sqrt{2}}{\sqrt{3}} \biggr) \\ \boldsymbol{\mu}_C^{\parallel} &= \mu^{\parallel} \biggl( \frac {1}{2},-\frac{1}{2\sqrt{3}}, \frac{\sqrt{2}}{\sqrt{3}} \biggr) \end{aligned} \end{aligned}$$](A303787_1_En_6_Chapter_Equ109.gif)
(6.109)
![$$\begin{aligned} \begin{aligned} \big|{{}^1A}_2\big\rangle&= \frac{1}{\sqrt{3}} \bigl[ (\chi_A \rightarrow\psi_A)^1 +( \chi_B \rightarrow\psi_B)^1 + ( \chi_C \rightarrow\psi_C)^1 \bigr] \\ \big|{{}^1E}_{\theta}\big\rangle&= \frac{1}{\sqrt{2}} \bigl[ -( \chi_B \rightarrow\psi_B)^1 + ( \chi_C \rightarrow\psi_C)^1 \bigr] \\ \big|{{}^1E}_{\epsilon}\big\rangle&= \frac{1}{\sqrt{6}} \bigl[2 ( \chi_A \rightarrow\psi_A)^1 -( \chi_B \rightarrow\psi_B)^1 - ( \chi_C \rightarrow\psi_C)^1 \bigr] \end{aligned} \end{aligned}$$](A303787_1_En_6_Chapter_Equ110.gif)
(6.110)
![$$ (\chi_A \rightarrow\psi_A)^1 = \frac{1}{\sqrt{2}} \bigl[ \big|(\chi_A \alpha) (\psi_A \beta)\big| - \big|(\chi_A \beta) (\psi_A \alpha)\big| \bigr] $$](A303787_1_En_6_Chapter_Equ111.gif)
(6.111)
![$$ V_{ij} =\frac{1}{4\pi\epsilon_0} \biggl(\frac{\boldsymbol{\mu}^i \cdot \boldsymbol{\mu}^j}{R^3_{ij}} - \frac{3 ( \boldsymbol{\mu}^i \cdot \mathbf{R}_{ij} ) ( \boldsymbol{\mu}^j \cdot\mathbf {R}_{ij} )}{R^5_{ij}} \biggr) $$](A303787_1_En_6_Chapter_Equ112.gif)
(6.112)
![$\sqrt{3}\rho$](A303787_1_En_6_Chapter_IEq47.gif)
![$$\begin{aligned} \begin{aligned} \bigl\langle{}^1A_{2} |V| {}^1A_{2} \bigr\rangle&= \frac{ (\mu^{\parallel } )^2}{4 \pi\epsilon_0 \rho^3} \frac{1}{6 \sqrt{3}} \\ \bigl\langle{}^1E |V| {}^1E \bigr\rangle&= - \frac{ (\mu^{\parallel} )^2}{4 \pi\epsilon_0 \sqrt{3} \rho^3} \frac{1}{12 \sqrt{3}} \end{aligned} \end{aligned}$$](A303787_1_En_6_Chapter_Equ113.gif)
(6.113)
![$$\begin{aligned} \begin{aligned} \boldsymbol{\mu}\bigl({}^1A_1 \rightarrow{}^1A_2\bigr) &= \frac{1}{\sqrt{3}} \bigl( \boldsymbol{\mu}^\parallel_A + \boldsymbol{\mu}^\parallel_B + \boldsymbol{\mu}^\parallel_C \bigr) =\sqrt{2} \mu^{\parallel} (0,0,1) \\ \boldsymbol{\mu}\bigl({}^1A_1 \rightarrow{}^1E_{\epsilon} \bigr) &= \frac{1}{\sqrt{6}} \bigl(2 \boldsymbol{\mu}^\parallel_A - \boldsymbol{\mu}^\parallel_B - \boldsymbol{ \mu}^\parallel_C \bigr) = \frac{1}{\sqrt{2}} \mu^\parallel (0,1,0) \\ \boldsymbol{\mu}\bigl({}^1A_1 \rightarrow{}^1E_{\theta} \bigr) &= \frac{1}{\sqrt{2}} \bigl( - \boldsymbol{\mu}^\parallel_B + \boldsymbol{\mu}^\parallel_C \bigr) = \frac{1}{\sqrt{2}} \mu^\parallel(1,0,0) \end{aligned} \end{aligned}$$](A303787_1_En_6_Chapter_Equ114.gif)
(6.114)
![$$ \boldsymbol{m} = -\frac{\mathrm{e}}{2m}\boldsymbol{l} = - \frac{\mathrm {e}}{2m} \boldsymbol{r}\times\boldsymbol{p} $$](A303787_1_En_6_Chapter_Equ115.gif)
(6.115)
![$$ [\mathcal{H},\boldsymbol{r}] = \biggl[ \biggl(\frac{\boldsymbol{p} \cdot \boldsymbol{p} }{2m}+ V( \boldsymbol{r}) \biggr), \boldsymbol{r} \biggr] = -\frac{i \hbar}{m} \boldsymbol{p} $$](A303787_1_En_6_Chapter_Equ116.gif)
(6.116)
![$$ \mathbf{m}_A = \langle\psi_A | \boldsymbol{m} | \chi_A\rangle = - \frac{\mathrm{e}}{2m}\bigl\langle\psi_A | ( \mathbf{R}_A + \boldsymbol{r}) \times\boldsymbol{p} | \chi_A\bigr\rangle= - \frac{\mathrm{e}}{2m}\mathbf {R}_A \times\langle\psi_A | \boldsymbol{p} | \chi_A\rangle $$](A303787_1_En_6_Chapter_Equ117.gif)
(6.117)
![$$\begin{aligned} \langle\psi_A | \boldsymbol{p} | \chi_A\rangle =& \frac{i m}{\hbar} \langle\psi_A |\mathcal{H}\boldsymbol{r}- \boldsymbol{r}\mathcal {H}|\chi_A\rangle \\ =& \frac{ i m}{\hbar} \bigl( \langle\mathcal{H} \psi_A |\boldsymbol {r}|\chi_A\rangle - \langle\psi_A |\boldsymbol{r} \mathcal{H}|\chi_A\rangle \bigr) \\ =& \frac{ i m }{\hbar} (E_{\psi} - E_{\chi} ) \langle\psi _A|\boldsymbol{r}|\chi_A\rangle \\ =& 2 \pi i m \nu\langle\psi_A|\boldsymbol{r}|\chi_A \rangle \end{aligned}$$](A303787_1_En_6_Chapter_Equ118.gif)
(6.118)
![$$\begin{aligned} \begin{aligned} \mathbf{m}_A &= i\pi\nu \bigl( \mathbf{R}_A \times\boldsymbol{\mu }_A^{\parallel} \bigr) = i\pi\nu\rho \mu^{\parallel} \biggl( 0, -\frac {\sqrt{2}}{\sqrt{3}},\frac{1}{\sqrt{3}} \biggr) \\ \mathbf{m}_B &= i\pi\nu \bigl( \mathbf{R}_B \times \boldsymbol{\mu }_B^{\parallel} \bigr) = i\pi\nu\rho \mu^{\parallel} \biggl( \frac {1}{\sqrt{2}}, \frac{1}{\sqrt{6}}, \frac{1}{\sqrt{3}} \biggr) \\ \mathbf{m}_C &= i\pi\nu \bigl( \mathbf{R}_C \times \boldsymbol{\mu }_C^{\parallel} \bigr) = i\pi\nu\rho \mu^{\parallel} \biggl( -\frac {1}{\sqrt{2}}, \frac{1}{\sqrt{6}}, \frac{1}{\sqrt{3}} \biggr) \end{aligned} \end{aligned}$$](A303787_1_En_6_Chapter_Equ119.gif)
(6.119)
![$$\begin{aligned} \begin{aligned} \mathbf{m}\bigl(^1A_1 \rightarrow^1A_2\bigr) &= \frac{1}{\sqrt{3}} \bigl( \mathbf{m}^\parallel_A + \mathbf{m}^\parallel_B + \mathbf{m}^\parallel_C \bigr) = i\pi\nu\rho \mu^{\parallel }(0,0,1) \\ \mathbf{m}\bigl(^1A_1 \rightarrow^1E_{\epsilon} \bigr) &= \frac{1}{\sqrt{6}} \bigl(2 \mathbf{m}^\parallel_A - \mathbf{m}^\parallel_B - \mathbf{m}^\parallel_C \bigr) = - i\pi\nu\rho\mu^{\parallel }(0,1,0) \\ \mathbf{m}\bigl(^1A_1 \rightarrow^1E_{\theta} \bigr) &= \frac{1}{\sqrt{2}} \bigl( - \mathbf{m}^\parallel_B + \mathbf{m}^\parallel_C \bigr) = - i\pi\nu\rho \mu^{\parallel}(1,0,0) \end{aligned} \end{aligned}$$](A303787_1_En_6_Chapter_Equ120.gif)
(6.120)
![$\mathcal{R}_{a\rightarrow j}$](A303787_1_En_6_Chapter_IEq48.gif)
![$$ \mathcal{R}_{a\rightarrow j} = \mathrm{Im} \bigl\{ \langle a| \boldsymbol {\mu}| j\rangle \cdot\langle j| \boldsymbol{m}| a \rangle \bigr\} $$](A303787_1_En_6_Chapter_Equ121.gif)
(6.121)
![$$\begin{aligned} \begin{aligned} \mathcal{R}\bigl({{}^1A_1 \rightarrow{{}^1A}_2}\bigr) &= \sqrt{2} \pi\nu\rho \bigl( \mu^{\parallel} \bigr)^2 \\ \mathcal{R}\bigl({{}^1A_1 \rightarrow{{}^1E}_{\epsilon}} \bigr) &= -\frac{1}{\sqrt {2}} \pi\nu\rho \bigl( \mu^{\parallel} \bigr)^2 \\ \mathcal{R}\bigl({{}^1A_1 \rightarrow{{}^1E}_{\theta}} \bigr) &= -\frac{1}{\sqrt{2}} \pi\nu\rho \bigl( \mu^{\parallel} \bigr)^2 \end{aligned} \end{aligned}$$](A303787_1_En_6_Chapter_Equ122.gif)
(6.122)
6.9 Induction Revisited: The Fibre Bundle
In Chap. 4 we left induction after the proof
of the Frobenius reciprocity theorem. In that proof the important
concept of the positional representation was introduced. This
described the permutation of the sites under the action of the
group elements. Further, we defined local functions on the sites
which transformed as irreps of the site symmetry. As an example, if
we want to describe the displacement of a cluster atom in a
polyhedron, two local functions are required: a totally-symmetric
one for the radial displacement and a twofold-degenerate one for
the tangential displacements. In cylindrical symmetry, these are
labelled σ and π, respectively. The mechanical representation, i.e. the
representation of the cluster displacements, is then the sum of the
two induced representations:
As an example using the induction tables in Sect. C.2 for an
octahedron, we have:
This is precisely the set of fluorine displacements that we
constructed in Sect. 4.8 in order to describe the
vibrational modes of UF6. One remarkable result of
induction theory is that the mechanical representation can also be
obtained as the direct product of the positional representation and
the translational representation, T 1u ; this is the representation of
the three displacements of the centre of the cluster.
It is as if the displacements of the central point of the
octahedron were relocated to every ligand site. The elementary
function space of the displacements of the central atom, which
transforms as the translational irrep, T 1u , is called the standard fibre. This fibre is attached
to every site of the cluster, and the set of these fibres is the
fibre bundle. The action of
the group permutes fibres of the bundle. The following induction
theorem holds:
![$$ \varGamma_{\mathrm{mech}} = \varGamma(\sigma H\uparrow G) + \varGamma(\pi H \uparrow G) $$](A303787_1_En_6_Chapter_Equ123.gif)
(6.123)
![$$ \varGamma_{\mathrm{mech}} = ( A_{1g} + E_g + T_{1u} ) + (T_{1g} + T_{2g} + T_{1u} + T_{2u}) $$](A303787_1_En_6_Chapter_Equ124.gif)
(6.124)
![$$\begin{aligned} \varGamma_{\mathrm{mech}} =& T_{1u} \times ( A_{1g} + E_g + T_{1u} ) \\ =& T_{1u} + ( T_{1u} + T_{2u} ) + ( A_{1g} + E_g + T_{1g} + T_{2g} ) \end{aligned}$$](A303787_1_En_6_Chapter_Equ125.gif)
(6.125)
Theorem 14
Consider a
standard fibre, consisting
of a function space that is invariant under the action of the
group. In a cluster of
equivalent sites, we can
form a fibre bundle by associating this standard fibre with every
site position. The induced
representation of the fibre bundle is then the direct product of
the irrep of the standard fibre with the positional
representation.
For
being the representation of the standard
fibre, T 1u in our example, and
the
positional representation of the set of equivalent sites in the
molecule, one has for the induced representation:
For a proof of this theorem, we refer to the literature
[19, 20]. The theorem is not only applicable to
molecular vibrations but is also directly in line with the LCAO
method in molecular quantum chemistry. In this method the molecular
orbitals (MOs) are constructed from atomic basis sets that are
defined on the constituent atoms. An atomic basis set, such as
3d or 4f, corresponds to a fibre, emanating,
as it were, from the atomic centre. Usually, such basis sets obey
spherical symmetry, since they are defined for the isolated atoms.
As such, they are also invariant under the molecular point group
[21]. As an example, a set of
4f polarisation functions
on a chlorine ligand in a
complex is itself adapted to
octahedral symmetry as a
2u +t 1u +t 2u . This representation thus
corresponds to
. In
the C 4v site symmetry these irreps
subduce: a
1+b
1+b
2+2e. According
to the theorem, the LCAOs based on the 4f orbitals thus will transform as:
In this LCAO space several irreps occur multiple times, but they
can all be distinguished by the specific direct product from which
they originated.
![$\mathcal{V}$](A303787_1_En_6_Chapter_IEq49.gif)
![$\mathcal{P}$](A303787_1_En_6_Chapter_IEq50.gif)
![$$ \varGamma\bigl(\{\mathcal{V}\}H\uparrow G\bigr) = \mathcal{V} \times\mathcal {P}(H\uparrow G) $$](A303787_1_En_6_Chapter_Equ126.gif)
(6.126)
![$\mathrm{RhCl}_{6}^{3-}$](A303787_1_En_6_Chapter_IEq51.gif)
![$\mathcal{V}$](A303787_1_En_6_Chapter_IEq52.gif)
![$$\begin{aligned} &\varGamma\bigl( \{a_1 + b_1 + b_2 + 2e \} C_{4v}\uparrow O_h\bigr) \\ &\quad= (a_{2u} + t_{1u} + t_{2u} ) \times(a_{1g}+ e_g + t_{1u}) \\ &\quad= a_{1g} + a_{2g} + 2e_g + 2t_{1g} + 3 t_{2g} + a_{2u} + e_u + 3t_{1u} + 3t_{2u} \end{aligned}$$](A303787_1_En_6_Chapter_Equ127.gif)
(6.127)
6.10 Application: Bonding Schemes for Polyhedra
Leonhard Euler dominated the mathematics of the
18th century. One of his famous discoveries was the polyhedral
theorem, which marks the beginning of topology. A polyhedron
has three structural elements: vertices, edges, and faces.9 The numbers of these will
be represented as v,
e, and f, respectively. Then, for a
polyhedron, the following theorem holds:
Theorem 15
In a convex
polyhedron the alternating sum of the numbers of vertices,
edges, and faces is always equal to 2.
![$$ v-e+f=2 $$](A303787_1_En_6_Chapter_Equ128.gif)
(6.128)
As an example, in a cube one has v=8,e=12,f=6, and hence 8−12+6=2. The 2 in the
right-hand side of Eq. (6.128) is called the Euler invariant. It is a
topological characteristic.
Topology draws attention to properties of surfaces, which are not
affected when surfaces are stretched or deformed, as one can do
with objects made of rubber or clay. Topology is thus not concerned
with regular shapes, and in this sense seems to be completely
outside our subject of symmetry; yet, as we intend to show in this
section, there is in fact a deep connection, which also carries
over to molecular properties. The surface to which the 2 in the
theorem refers is the surface of a sphere. A convex polyhedron
is indeed a polyhedron which can be embedded or mapped on the
surface of a sphere. Group theory, and in particular the induction
of representations, provides the tools to understand this
invariant. To this end, each of the terms in the Euler equation is
replaced by an induced representation, which is based on the
particular nature of the corresponding structural element. In
Fig. 6.8
we illustrate the results for the case of the tetrahedron.
The following theorem [22]
applies:
-
The vertices, being zero-dimensional points, form a set of nodes, {〈u〉}, which are permuted under the symmetry operations of the polyhedron. The representation of this set is the positional representation, Γ σ (v). The σ here refers to the fact that the sites themselves transform as totally-symmetric objects in the site group. If the cluster contains several orbits, the induced representation is of course the sum of the individual positional representations. In Fig. 6.8 the vertex representation is A 1+T 2. In Sect. 4.7 we have already encountered these irreps, when discussing the sp 3 hybridization of carbon.(6.129)Fig. 6.8Face, edge and vertex SALCs for a tetrahedron. The δ symbol denotes taking the boundary, from faces to edges, and from edges to vertices (see text). The two topological invariants are the A 2 face term and the A 1 vertex term
-
The edges are one-dimensional lines. They form a set of ordered pairs, {〈u,v〉}. Each of these can be thought of as an arrow, directed along the edge. The symmetry operations will interchange these arrows, but may also change their sense. The corresponding representation is labelled as Γ ∥(e). This symbol indicates that the basic objects on the edge sites are not symmetric points but directed arrows. The site group through the centre of an edge has maximal symmetry C 2v and in this site group the arrows transform as the b 1 irrep, which is symmetric under reflection in a plane containing the edge and antisymmetric under the symmetry plane perpendicular to the edge. For a tetrahedron there are six edge vectors, transforming as T 1+T 2.(6.130)
-
The faces may be represented as closed chains of nodes, which are bordering a polyhedral face, {〈u,v,w,…〉}. The sequence forms a circulation around the face, in a particular sense (going from 〈u〉 to 〈w〉 over 〈v〉, etc.). The set of face rotations forms the basis for the face representation, which is denoted as Γ ↺(f). In a polyhedron the maximal site group of a face is C nv , and in this site group the face rotation transforms as the rotation around the
axis, i.e. it is symmetric under the axis and antisymmetric with respect to the vertical
planes, which invert the sense of rotation. For the tetrahedron, the face circulations transform as A 2+T 2, as shown in Fig. 6.8.
(6.131)
Theorem 16
The alternating
sum of induced representations of the vertex nodes,
edge arrows, and face rotations, is equal to the sum of the
totally-symmetric
representation, Γ
0, and the
pseudo-scalar
representation, Γ
ϵ . The latter representation is symmetric under
proper symmetry elements and antisymmetric under improper symmetry
elements.
![$$ \varGamma_{\sigma}(v) - \varGamma_{\parallel}(e) + \varGamma_{\circlearrowleft}(f)= \varGamma_0 + \varGamma_{\epsilon} $$](A303787_1_En_6_Chapter_Equ132.gif)
(6.132)
The Euler theorem may be considered as the
dimensional form of this theorem, which states that the alternating
sum of the characters of the induced representations under the unit
element,
, is
equal to 2, but the present theorem extends this character
equality to all the operations of the group. The theorem silently
implies that irreps can be added and subtracted. In the example of
the tetrahedron, the theorem is expressed as:
A straightforward interpretation of the theorem is possible in
terms of fluid flow on the surface of a polyhedron.10 Suppose observers are positioned
on the vertices, edge centres and face centres, and register the
local fluid flow. When the incoming and outgoing currents at a node
are not in balance, the observers located on these nodes will
report piling up or depletion of the local fluid level. This is the
scalar property represented by the vertex term. The corresponding
connection between edge flow and vertex density is expressed by the
boundary operation,
indicated by δ in
Fig. 6.8.
Taking the boundary of an edge arrow means replacing the arrow by
the difference of two vertex-localized scalars: a positive one
(indicated by a white circle in the figure) at the node to which
the arrow’s head is pointing, and a negative one (indicated by a
black circle) at the node facing the arrow’s tail. This projection
from edge to vertex will not change the symmetry. Hence, in this
way, the boundary of the T
2 edge irrep is the T 2 vertex SALC, as
illustrated in the figure. Similarly, observers in face centres
will notice the net current that is circulating around the face.
Such a circular current through the edges does not give rise to
changes at the nodes (indeed the incoming flow at a node is also
leaving again), but is observable from the centre of the face
around which the current is circulating. The boundaries of circular
currents around face centres are thus chains of arrows on the
edges, which again conserve the symmetry. In Fig. 6.8 the boundary of the
T 1 face term is
thus the T 1
edge term. Clearly, the sum of the vertex and face observations
should account for all currents going through the edges, except for
two additional terms which escape edge observations. These are the
two Euler invariants: the totally-symmetric Γ 0 component corresponds to
a uniform change of fluid amplitude at all vertex basins. This does
not give rise to edge currents, since it creates no gradients over
the edges. The other is the Γ ϵ component. It corresponds to a
simultaneous rotation around all faces in the same sense. Again,
such rotor flows do not create net flows through the edges, because
two opposite currents are flowing through every edge. The Euler
invariant thus points to two invariant characteristic modes of the
sphere. They are not boundaries of a mode at a higher level, nor
are they bounded by a mode at a lower level. The phenomena, that
these two terms describe, might also be referred to in a
topological context as the electric and magnetic monopoles.
![$\hat{E}$](A303787_1_En_6_Chapter_IEq55.gif)
![$$ \varGamma_{\sigma}(v) - \varGamma_{\parallel}(e) + \varGamma_{\circlearrowleft}(f)= (A_1 + T_2) - (T_1 + T_2) + (A_2 + T_2) =A_{1}+A_{2} $$](A303787_1_En_6_Chapter_Equ133.gif)
(6.133)
Because of this connection to density and
current, this theorem may be applied in various ways to describe
chemical bonding, frontier orbital structure, and vibrational
properties. The applications of this theorem can be greatly
extended by introducing fibre representations, as is shown
below.
Taking the Dual
To take the dual of a polyhedron is to replace
vertices by faces and vice-versa, as was already mentioned in
Sect. 3.7 in relation to the Platonic
solids. The dual has the same number of edges as the original, but
every edge is rotated 90∘. Hence the relations between
v D ,e D ,f D for the dual and v,e,f for the original are:
![$$\begin{aligned} v^D =& f \\ e^D =& e \\ f^D =& v \end{aligned}$$](A303787_1_En_6_Chapter_Equ134.gif)
(6.134)
As a result the Euler formula is invariant under
the dual operation.
A similar invariance holds for the symmetry extension, but in this
case “to take the dual” corresponds to multiplying all terms by the
pseudo-scalar irrep Γ
ϵ . The terms
are then changed as follows:
Hence, if the theorem holds for the original, it also holds for the
dual.
![$$ v-e+f = v^D -e ^D + f^D = 2 $$](A303787_1_En_6_Chapter_Equ135.gif)
(6.135)
![$$\begin{aligned} \begin{aligned} \varGamma_{\sigma}(v) \times\varGamma_{\epsilon} &= \varGamma_{\circlearrowleft}(v) =\varGamma_{\circlearrowleft} \bigl(f^D \bigr) \\ \varGamma_{\parallel}(e) \times\varGamma_{\epsilon} &= \varGamma_{\perp}(e) = \varGamma_{\parallel}\bigl(e^D\bigr) \\ \varGamma_{\circlearrowleft}(f) \times\varGamma_{\epsilon} &= \varGamma_{\sigma}(f)= \varGamma_{\sigma}\bigl(v^D \bigr) \\ ( \varGamma_0 + \varGamma_{\epsilon} ) \times \varGamma_{\epsilon} &= \varGamma_0 + \varGamma_{\epsilon} \end{aligned} \end{aligned}$$](A303787_1_En_6_Chapter_Equ136.gif)
(6.136)
![$$ \bigl(\varGamma_{\sigma}(v) - \varGamma_{\parallel}(e) + \varGamma_{\circlearrowleft}(f) \bigr) \times\varGamma_\epsilon= \varGamma_{\sigma}\bigl(v^D\bigr) - \varGamma_{\parallel} \bigl(e^D\bigr) +\varGamma_{\circlearrowleft }\bigl(f^D \bigr)= \varGamma_0 + \varGamma_{\epsilon} $$](A303787_1_En_6_Chapter_Equ137.gif)
(6.137)
Note especially the fibre modification of the
edge term. The maximal local symmetry of an edge is C 2v . The arrow along the edge
transforms as b
1, while the pseudo-scalar irrep in C 2v is a 2. The product
b 1×a 2 produces b 2, which is precisely the
symmetry of an arrow, tangent to the surface of the polyhedron, but
directed perpendicular to the edge. Multiplication with the
pseudo-scalar irrep thus has the effect of rotating the edges
through 90∘. In Eq. (6.136) the resulting representation is denoted
as Γ
⊥(e).
Deltahedra
Deltahedra are polyhedra that consist entirely of
triangular faces. Three of the Platonic solids are deltahedra: the
tetrahedron, the octahedron and the icosahedron. In a convex
deltahedron the bond stretches (i.e. stretchings of the edges) span
precisely the representation of the internal vibrations. In other
words, a convex deltahedron cannot vibrate if it is made of rigid
rods. This is the Cauchy theorem:
Theorem 17
Convex polyhedra
in three dimensions with congruent corresponding faces must be
congruent to each other. In
consequence, if a
polyhedron is made up of triangles with rigid rods,
the angles between the triangular
faces are fixed.
This result can be cast in the language of
induced representations. The stretchings of the edges correspond to
scalar changes of edge lengths and transform as σ-type objects, and hence will
correspond to Γ
σ (e). On the other hand, the internal
vibrations span the mechanical representation, which can be written
as a bundle of the translation, minus the spurious modes of
translation and rotation. The symmetries of these will be denoted
as Γ T and Γ R , respectively. One thus has:
![$$ \mbox{Deltahedron:}\quad \varGamma_{\sigma}(v) \times \varGamma_T - \varGamma_T - \varGamma_R = \varGamma_{\sigma}(e) $$](A303787_1_En_6_Chapter_Equ138.gif)
(6.138)
Trivalent Polyhedra
The dual of a deltahedron is a trivalent
polyhedron, meaning that every vertex is connected to three nearest
neighbours. The fullerene networks of carbon are usually trivalent
polyhedra. This reflects the sp 2 hybridization of
carbon, which can form three σ-bonds. Also in this case several
specialized forms of the Euler symmetry theorem can be formulated.
We may start from Eq. (6.138) and replace vertices by faces. The edge
terms remain the same since they are totally symmetric under the
local symmetries of the edges. Rotations of the edges by
90∘ will thus not affect these terms.
Furthermore, by multiplying the vertices in a trivalent polyhedron
by three, we have accounted for all the edges twice, since each
edge is linked to two vertices, hence:
The 3v in this formula
suggests once again taking the fibre representation Γ σ (v)×Γ T . In doing so we have
considered on each vertex one σ and two π objects. Hence, this is not only the
mechanical representation with three displacements on each vertex,
but it is equally well the symmetry of a set of sp 2 hybrids on every
vertex, directed along the three edges. Along each edge the hybrids
at either end can be combined in a local bonding and anti-bonding
combination. The corresponding induced representations are
respectively: Γ
σ (e) and Γ ∥(e); hence, the symmetry extension of
Eq. (6.140)
reads:
![$$ \mbox{Trivalent:} \quad \varGamma_{\sigma}(f) \times\varGamma_T - \varGamma_T - \varGamma_R = \varGamma_{\sigma}(e) $$](A303787_1_En_6_Chapter_Equ139.gif)
(6.139)
![$$ \mbox{Trivalent:} \quad 3v=2e $$](A303787_1_En_6_Chapter_Equ140.gif)
(6.140)
![$$ \mbox{Trivalent:} \quad\varGamma_{\sigma}(v) \times\varGamma_T = \varGamma_{\sigma}(e) +\varGamma_{\parallel}(e) $$](A303787_1_En_6_Chapter_Equ141.gif)
(6.141)
Edge Bonding in Trivalent Polyhedra
The understanding of the bonding schemes in
polyhedra is based on the correct identification of the local
hybridization scheme on the constituent fragments. Trivalent
polyhedra are often electron-precise: this means that the
fragment has three electrons in three orbitals, which are available
for cluster bonding and give rise to edge-localized σ-bonds. Such is the case for the
methyne fragment, CH, forming polyhedranes, but equally well for the
isolobal [24] organo-transition-metal fragments such as
, where M is a d 9 metal such as Co,Rh or
Ir. Figure 6.9 shows the bonding pattern based on such
electron-precise fragments. As indicated before, the orbital basis
corresponds to the fibre representation Γ σ (v)×Γ T , and contains 3n orbitals. Local interactions along
the edges will split this orbital basis into an occupied
σ-bonding half and a
virtual σ-anti-bonding
counterpart, transforming as Γ σ (e) and Γ ∥(e), respectively. This is precisely the
result of Eq. (6.141). Now, for each half, a more detailed
pattern can be discerned [25]. For
the anti-bonding orbitals, the general theorem, Eq. (6.132), can be applied
directly. The result is illustrated in Fig. 6.9. By this theorem, the
3n/2 edge anti-bonds are
split into two subsets containing (1+n/2) and (n−1) orbitals. The former, higher
lying, subset transforms as Γ ↺−Γ ϵ . These terms correspond to
circulations around the faces, which means that these levels will
be highly anti-bonding. In fact, they are always at the top of the
skeletal spectrum. Note that the pseudo-scalar term, Γ ϵ , does not take part. This is
because a uniform circulation around all faces in the same sense
has no contribution on the edges. Below this is a subset of weakly
anti-bonding orbitals, transforming as Γ σ (v)−Γ 0. These orbitals are more
localized on the vertices. The Γ 0 term is not included
since this is the totally-symmetric molecular orbital which is
completely bonding, and thus will appear in the lower half of the
diagram.
![$\mathrm{M(CO)}_{3}$](A303787_1_En_6_Chapter_IEq56.gif)
![A303787_1_En_6_Fig9_HTML.gif](A303787_1_En_6_Fig9_HTML.gif)
Fig. 6.9
Edge bonding in electron-precise trivalent
cages. The valence shell splits into an occupied set of localized
edge-bonds, and a matching virtual set of edge-anti-bonds. The sets
may be further differentiated by use of the symmetry theorems
Furthermore, the edge-bonding half can be
analysed with the help of Eq. (6.139). The 3n/2 edge bonds split into two subsets
of dimension (n−2) and
(2+n/2). This analysis
involves the fibre representation Γ σ (f)×Γ T , which can be decomposed into
a radial σ- and tangential
π-part. The σ-part corresponds to
cylindrically-symmetric bonds around the faces, and will thus be
strongly bonding. For the π-part the face terms contain a nodal
plane through the faces, and thus will be less bonding.
Frontier Orbitals in Leapfrog Fullerenes
Fullerenes are trivalent polyhedra of carbon,
consisting of hexagons and pentagons. The following relations hold:
The first two relationships are from Eqs. (6.128) and (6.140). The third
expresses that the total number of the faces is the sum of the
number of pentagons (f
5), and hexagons (f 6). The final equation
indicates that by counting the hexagons six times, and the
pentagons five times, we have counted all vertices three times,
since every vertex is at the junction of three faces. Even though
there are fewer equations, here, than unknowns, it can easily be
seen by manipulation of Eq. (6.142) that the only value that the number of
pentagons, f 5,
can take on is 12. Hence, the smallest fullerene is the
dodecahedron C20, which only consists of pentagons. Also
note that the number of atoms in a fullerene must be even, since
3v must be divisible by 2,
as e is an integer. Taking
the leapfrog, L, of a primitive fullerene,
P, is an operation of cage
expansion, which yields a fullerene with three times as many atoms
[26]. This procedure is described
by the following rule:
It involves two operations, which are carried out consecutively, as
illustrated in Fig. 6.10. One first places an extra capping atom on all pentagons and
hexagons. This leads to a cage which consists only of triangles,
and this is a deltahedron. By taking the dual one restores a
trivalent cage. As can be seen, all vertices of the primitive have
been turned into hexagons, while the original pentagons and
hexagons are recovered, but in a rotational stagger. The edges of
the primitive are also recovered, but rotated 90∘. In
summary, the leapfrog operation inserts 6 vertices in the hexagons
of P, and 5 vertices in the
pentagons, which, according to the final expression in Eq.
(6.142),
multiplies the number of atoms by 3. The first and best known
leapfrog is Buckminsterfullerene, C60, which is the
leapfrog of the dodecahedron itself. Each carbon atom contributes,
besides the sp 2
orbitals, which build the σ-frame, one radial p r -orbital. These orbitals form
π-bonds which control the
frontier orbitals of fullerenes. In the case of the leapfrog, this
frontier MO region is always characterized by six low-lying almost
non-bonding orbitals, which, moreover, always transform as
Γ T +Γ R . This can be explained with
the help of the Euler rules [27].
![$$\begin{aligned} \begin{aligned} v-e+f & = 2 \\ 3v & = 2e \\ f_5 + f_6 & = f \\ 5f_5 + 6 f_6 & = 3v \end{aligned} \end{aligned}$$](A303787_1_En_6_Chapter_Equ142.gif)
(6.142)
![$$ L = \mathrm{Dual} ( \mathrm{Omnicap } P ) $$](A303787_1_En_6_Chapter_Equ143.gif)
(6.143)
![A303787_1_En_6_Fig10_HTML.gif](A303787_1_En_6_Fig10_HTML.gif)
Fig. 6.10
The leapfrog extension consists of two
operations: first, place an extra atom in the centres of all the
polygons (middle panel),
then, take the dual. The result is indicated by the solid lines in the right panel
As in the case of C60, all leapfrogs
can be considered to be truncations of the primitive fullerenes, in
the sense that all the faces of the primitive have become isolated
islands, surrounded by
rings of hexagons. With every bond of the primitive is associated a
perpendicular bond, which always forms a bridge between these islands. Based on
this neat bond separation, two canonical valence-bond frames can be
constructed for the leapfrog, which in a sense are the extremes of
a correlation diagram, with the actual bonding somewhere in
between. These bond schemes are known as the Fries and the Clar
structures. The Fries structure is an extreme case where all
bridges are isolated
π-bonds. The induced
representation for these bonds corresponds to Γ σ (e P ). On the other hand, in the
Clar structures the bonding is completely redistributed to aromatic
sextets on the hexagonal and pentagonal islands. The corresponding
representation is the fibre bundle Γ σ (f P )×Γ T . We now compare the
representations of both bonding schemes, using the symmetry
theorems. We start with the main theorem, applied to the primitive
P, and multiply left and
right with Γ
R .
Since Γ R =Γ T ×Γ ϵ , we could already simplify the
face term to a form which precisely corresponds to the Clar
representation:
The vertex term can be expressed with the help of Eq. (6.141):
where the pseudo-scalar irrep turns a σ-object into a circular current, and
rotates the parallel edge current over 90∘. Note that
this is applied in the primitive cage, to the edges of P only. To complete the derivation one
final fibre bundle is needed, which applies to all convex
polyhedra:
![$$ \varGamma_{\sigma}\bigl(v^P\bigr) \times \varGamma_R - \varGamma_{\parallel}\bigl(e^P\bigr) \times\varGamma_R +\varGamma_{\circlearrowleft}\bigl(f^P \bigr)\times\varGamma_R= \varGamma_T + \varGamma_R $$](A303787_1_En_6_Chapter_Equ144.gif)
(6.144)
![$$ \varGamma_{\circlearrowleft}\bigl(f^P\bigr)\times\varGamma_R= \varGamma_{\sigma }\bigl(f^P\bigr)\times\varGamma_T= \varGamma_{\mathrm{Clar}} $$](A303787_1_En_6_Chapter_Equ145.gif)
(6.145)
![$$\begin{aligned} \mbox{Trivalent:} \quad \varGamma_{\sigma}\bigl(v^P\bigr) \times\varGamma_R = \varGamma _{\sigma}\bigl(e^P \bigr)\times{\varGamma_{\epsilon}} +\varGamma_{\parallel} \bigl(e^P\bigr)\times\varGamma_\epsilon= \varGamma_{\circlearrowleft } \bigl(e^P\bigr) + \varGamma_{\perp}\bigl(e^P \bigr) \end{aligned}$$](A303787_1_En_6_Chapter_Equ146.gif)
(6.146)
![$$ \varGamma_{\sigma}(e) \times\varGamma_T = \varGamma_{\sigma}(e) + \varGamma_{\parallel }(e) + \varGamma_{\perp}(e) $$](A303787_1_En_6_Chapter_Equ147.gif)
(6.147)
This result is based on the C 2v site-symmetry of an edge. The
translation in this site has a radial σ-component of a 1 symmetry, and two
tangential π-components of
b 1+b 2 symmetry. The fibre
bundle will thus correspond to the induction of a 1+b 1+b 2, which is precisely the
meaning of the three terms on the right-hand side of Eq.
(6.147).
This expression may be transformed in two steps to the term which
is required in the derivation. One first changes the substrate of
the fibre from Γ
σ (e) to Γ ∥(e). This associates the edges with
b 1 objects, and
combination with a
1+b
1+b 2
will thus yield b
1+a
1+a
2, or:
Finally, multiply this result by Γ ϵ :
We now combine Eqs. (6.146) and (6.149), and find:
This is precisely the representation of the Fries bonds. We can
thus compare the Fries and Clar structures in a general leapfrog,
and find from Eq. (6.144):
The Clar structure thus has six extra bonding orbitals as compared
with the Fries structure. When both bonding schemes are correlated,
as illustrated in Fig. 6.11, this sextet must correlate with the
anti-bonding half of the Fries structure. It will thus be placed on
top of the Clar band, and actually be nearly non-bonding, forming
six low-lying virtual orbitals, which explains the electron
deficiency of the leapfrog fullerenes. Moreover, as the derivation
shows, they transform exactly as rotations and translations.
![$$ \varGamma_{\parallel}(e) \times\varGamma_T = \varGamma_{\parallel}(e)+\varGamma_{\sigma}(e) + \varGamma_{\circlearrowleft}(e) $$](A303787_1_En_6_Chapter_Equ148.gif)
(6.148)
![$$ \varGamma_{\parallel}(e) \times\varGamma_R = \varGamma_{\perp}(e)+\varGamma _{\circlearrowleft}(e) + \varGamma_{\sigma}(e) $$](A303787_1_En_6_Chapter_Equ149.gif)
(6.149)
![$$ \mbox{Trivalent:} \quad \varGamma_{\parallel}\bigl(e^P\bigr) \times\varGamma_R -\varGamma_{\sigma }\bigl(v^P \bigr) \times\varGamma_R= \varGamma_{\sigma} \bigl(e^P\bigr) = \varGamma_{\mathrm{Fries}} $$](A303787_1_En_6_Chapter_Equ150.gif)
(6.150)
![$$ \varGamma_{\mathrm{Clar}} - \varGamma_{\mathrm{Fries}}= \varGamma_T + \varGamma_R $$](A303787_1_En_6_Chapter_Equ151.gif)
(6.151)
![A303787_1_En_6_Fig11_HTML.gif](A303787_1_En_6_Fig11_HTML.gif)
Fig. 6.11
Correlation diagram for C60. The
Fries and Clar structures are bonding extremes, where double bonds
are either localized on the 30 bonds between the pentagons (Fries),
or form isolated aromatic sextets on the twelve pentagons. The true
conjugation scheme is found in between, and is characterized by six
unoccupied levels, which are anti-bonding in the Fries structure
and bonding in the Clar structure, and which transform as rotations
and translations. Buckminsterfullerene has low-lying LUMO and
LUMO+1 levels of t
1u (Γ T ) and t 1g (Γ R ) symmetry
6.11 Problems
6.1
A three-electron wavefunction in an octahedron is
given by:
The vertical bars denote a Slater determinant. Determine the
symmetry of this function, starting from parent two-electron
coupled states, to which the third electron is coupled. Make use of
the coupling coefficients in Appendix F.
![$$ \varPsi=\big|(t_{1u}x\alpha) (t_{1u}y\alpha) (t_{1u}z\alpha)\big| $$](A303787_1_En_6_Chapter_Equ152.gif)
(6.152)
6.2
Write the Jahn-Teller matrix for a threefold
degenerate T
1u level in an
icosahedral molecule. How many reduced matrix elements are
needed?
6.3
Do you expect octahedral e g orbitals to show a magnetic
dipole moment?
6.4
Binaphthyl consists of two linked naphthalene
molecules. The dihedral angle between the two naphthyl planes is
around 70∘, and can be stabilized by bulky substituents
on the naphthyl units, as indicated below for the case of
2,2′-di-biphenylphosphine-1,1′-binaphthyl. A circular
dichroism signal is detected in the UV region, corresponding to the
long-axis polarized transitions of the naphthyl units (indicated by
the arrows in the figure). Construct the appropriate exciton states
and determine the CD profile of the two enantiomers of binaphthyl.
![A303787_1_En_6_Figa_HTML.gif](A303787_1_En_6_Figa_HTML.gif)
6.5
A diradical is a molecule with two open orbitals, each containing one
electron. Consider as an example twisted ethylene (D 2d symmetry, see
Fig. 3.9). The HOMO is a degenerate
e-orbital, occupied by two
electrons. Construct the e
2 diradical states for this molecule, and determine
their symmetries.
6.6
Planar trimethylenemethane (TMM), C 4 H 6, is a diradical with
trigonal symmetry. Determine the Hückel spectrum for the four
carbon p z -orbitals perpendicular to the
plane of the molecule. The HOMO in D 3h has e″ symmetry and is also occupied by two
electrons. Determine the corresponding diradical states, and
compare with the results for twisted ethylene. How would you
describe the valence bond structure of this molecule?
![A303787_1_En_6_Figb_HTML.gif](A303787_1_En_6_Figb_HTML.gif)
References
1.
Wigner, E.P.: Group Theory.
Academic Press, New York (1959)
2.
Griffith, J.S.: The
Irreducible Tensor Method for Molecular Symmetry Groups. Prentice
Hall, Englewood-Cliffs (1962)
3.
Butler, P.H.: Point Group
Symmetry Applications, Methods and Tables. Plenum Press, New York
(1981) CrossRef
4.
Ceulemans, A., Beyens, D.:
Monomial representations of point-group symmetries. Phys. Rev. A
27, 621 (1983) CrossRef
5.
Lijnen, E., Ceulemans, A.:
The permutational symmetry of the icosahedral orbital quintuplet
and its implication for vibronic interactions. Europhys. Lett.
80, 67006 (2007) CrossRef
6.
Jahn, H.A., Teller, E.:
Stability of polyatomic molecules in degenerate electronic states.
I. Orbital degeneracy. Proc. R. Soc. A 161, 220 (1937) CrossRef
7.
Bersuker, I.B.: The
Jahn–Teller Effect. Cambridge University Press, Cambridge
(2006)
8.
Domcke, W., Yarkony, D.R.,
Köppel, H.: Conical Intersections, Electronic Structure, Dynamics
and Spectroscopy. Advanced Series in Physical Chemistry, vol. 15.
World Scientific, Singapore (2004)
9.
Shapere, A., Wilczek, F.:
Geometric Phases in Physics. Advanced Series in Mathematical
Physics, vol. 5. World Scientific, Singapore (1989)
10.
Judd, B.R.: The theory of
the Jahn–Teller effect. In: Flint, C.D. (ed.) Vibronic Processes in
Inorganic Chemistry. NATO Advanced Study Institutes Series, vol.
C288, p. 79. Kluwer, Dordrecht (1988)
11.
Bersuker, I.B., Balabanov,
N.B., Pekker, D., Boggs, J.E.: Pseudo Jahn–Teller origin of
instability of molecular high-symmetry configurations: novel
numerical method and results. J. Chem. Phys. 117, 10478 (2002) CrossRef
12.
Hoffmann, R., Woodward,
R.B.: Orbital symmetry control of chemical reactions. Science
167, 825 (1970) CrossRef
13.
Halevi, E.A.: Orbital
symmetry and reaction mechanism, the OCAMS view. Springer, Berlin
(1992) CrossRef
14.
Rodger, A., Nordén, B.:
Circular Dichroism and Linear Dichroism. Oxford Chemistry Masters.
Oxford University Press, Oxford (1997)
15.
Orgel, L.E.: Double bonding
in chelated metal complexes. J. Chem. Soc. 3683 (1961).
doi:10.1039/JR9610003683
16.
Ceulemans, A.,
Vanquickenborne, L.G.: On the charge-transfer spectra of iron(II)-
and ruthenium(II)-tris(2,2′-bipyridyl) complexes. J. Am. Chem. Soc.
103, 2238 (1981) CrossRef
17.
Day, P., Sanders, N.:
Spectra of complexes of conjugated ligands. 2. Charge-transfer in
substituted phenantroline complexes: intensities. J. Chem. Soc. A
10, 1536 (1967) CrossRef
18.
Yersin, H., Braun, D.:
Localization in excited states of molecules. Application to
. Coord. Chem. Rev.
111, 39 (1991) CrossRef
![$\mbox{Ru(bipy)}_{3}^{2+}$](A303787_1_En_6_Chapter_IEq58.gif)
19.
Melvin, M.A.: Simplification
in finding symmetry-adapted eigenfunctions. Rev. Mod. Phys.
28, 18 (1956) CrossRef
20.
Ceulemans, A.: The
construction of symmetric orbitals for molecular clusters. Mol.
Phys. 54, 161 (1985)
CrossRef
21.
Biel, J., Chatterjee, R.,
Lulek, T.: Fiber structure of space of molecular-orbitals within
LCAO method. J. Chem. Phys. 86, 4531 (1987) CrossRef
22.
Ceulemans, A., Fowler, P.W.:
Extension of Euler theorem to the symmetry properties of polyhedra.
Nature 353, 52 (1991)
CrossRef
23.
Hilton, P.J., Wylie, S.:
Homology Theory. Cambridge University Press, Cambridge (1967)
24.
Hoffmann, R.: Building
bridges between inorganic and organic chemistry (Nobel lecture).
Angew. Chem., Int. Ed. Engl. 21, 711 (1982) CrossRef
25.
Mingos, D.M.P., Wales, D.J.:
Introduction to Cluster Chemistry. Prentice Hall, Englewood-Cliffs
(1990)
26.
Fowler, P.W., Manolopoulos,
D.E.: An Atlas of Fullerenes. Clarendon, Oxford (1995)
27.
Fowler, P.W., Ceulemans, A.:
Electron deficiency of the fullerenes. J. Phys. Chem. A
99, 508 (1995) CrossRef
Footnotes
1
The general case with complex irreps is
exemplified for the coupling of spin representations in
Sect. 7.4.
2
Any permutation can be expressed as a sequence of
transpositions of two elements. If the total number of
transpositions is even, sgn(σ)=+1; if it is odd, sgn(σ)=−1. See also Sect. 3.3.
5
Such combinations can be cast in a higher-order
symbol, known as 6Γ symbol,
by analogy with the 6j
coupling coefficients in atomic spectroscopy.
7
In tris-chelate complexes Δ refers to a
right-handed (dextro)
helix. A left-handed helix (lævo) is denoted as Λ.
8
The excitation creates an electron-hole pair,
which can move from one ligand to another. This is called an
exciton.