Sunteți pe pagina 1din 29

Department of Chemistry

University of Cambridge

Part II Chemistry 2004-2005

High-Resolution Molecular Spectroscopy


Electronic Spectroscopy

Lecturer: Professor Jacek Klinowski


2
3

Electronic spectra of diatomic molecules

The Born-Oppenheimer approximation


A diatomic molecule may undergo electronic, vibrational and rotational transitions.
The Born-Oppenheimer approximation (“Since the energies of the various motions are
very different, motions of a diatomic molecule may be considered as independent”)
allows us to express the total molecular energy as:
Etotal ≈ Eelec + Evib + Erot
or equivalently using term values:
T ≈ Te + G(v) + F(J)
where Te is the electronic energy which depends only on the electronic configuration
and nuclear repulsion. A change in the total energy of a molecule (in wavenumbers) is:
Δεtotal = Δεelec + Δεvib + Δεrot cm-1
where the approximate orders of magnitude of the three quantities are:
Δεelec ≈ Δεvib x 103 ≈ Δεrot x 106
Vibrational changes thus produce a “coarse structure”, and rotational changes a “fine
structure” on the electronic transitions. Unlike in vibrational and rotational
spectroscopies, all molecules give electronic spectra.

Vibrational coarse structure


Ignoring rotational changes, we have
Δεtotal = Δεelec + Δεvib cm-1
or
εtotal = εelec + (v + 1/2)ϖe – xe(v + 1/2)2 ϖe cm-1 (v = 0, 1, 2,…)
where ϖe is the equilibrium vibrational frequency expressed in wavenumbers and xe is
the anharmonicity constant. The lower energy states are marked with a double prime,
and the higher states by a single prime. The figure below shows the vibrational coarse
structure in an electronic transition from the ground state (v’ = 0), to a higher state. By
convention, X is used to denote the electronic ground state, and A, B, C etc. the excited
states in order of increasing energy. The electronic energy is measured from the bottom
of the respective potential energy curves, while the lowest molecular energy level is
offset upwards by the zero point energy of that vibration. In general v” ≠ v’.
4

Vibrational transitions occur providing the electronic transition dipole moment and the
Franck-Condon factors are non-zero. In the infrared, the selection rules for v arise from
the symmetry of the different vibrational wavefunctions, and are strictly Δv = ±1 for the
simple harmonic oscillator. For an anharmonic oscillator, this strict rule begins to break
down and overtones are allowed, albeit with rapidly declining intensity. For vibrational
transitions between different electronic levels, this rule breaks down completely and
there is now no longer any restriction on Δv, so that every v’ ← v” transition has some
probability, giving rise to many spectral lines. However, absorption spectra from the
ground state are much simpler: virtually all the molecules are in their lowest vibrational
state (v” = 0), so that the only transitions we observe are (0, 1), (0, 2), (0, 3) etc.

Note that the upper state is given first within the brackets. Each set of transitions in a
band is called a v’ progression, since the value of v’ increases by unity for each line in
the set. The lines in a band are closer together at high frequencies because of the
5

anharmonicity of the upper state vibrations, which causes vibration energy levels to
converge. Note that, in general, the spacings of vibrational energy levels in the two
electronic states are different.
To quantify:
Δεtotal = Δεelec + Δεvib. = (ε’ – ε”) +
+ [(v’ + 1/2)ϖe’ – xe’(v’ + 1/2)2 ϖe’ – (v” + 1/2)ϖe” – xe”(v” + 1/2)2 ϖe”]
cm-1
If a sufficient number (at least five) lines can be resolved in the band, the values of xe’,
ϖe’, xe”, ϖe” and (ε’ – ε”) can be calculated. Information on the excited electronic states is
particularly valuable, since such states are extremely unstable, and therefore short-lived.

Vibrational band analysis: the Deslandres table


The Deslandres table is a convenient means of verifying vibrational assignments. The
differences between rows and columns should be constant, corresponding to the energy
difference between a particular pair of vibrational levels in either the upper state (rows)
or the lower state (columns):

A portion of the Deslandres table for the 1Π ← 1Σ electronic transition of CO is given


below. Vibrational wavenumbers are shown in bold, with the differences in plain
characters.
6

v” = 0 v” = 1 – 0 1 v” = 2 – 1 2 v” = 3 – 2 3
v’ = 0 64748 2143 62605 2117 60488 2090 58398
v’ = 1 – 0 1480 1480 1480 1480
1 66228 2143 64085 2117 61968 2090 59878
v’ = 2 – 1 1440 1440 1440 1440
2 67668 2143 65525 2116 63409 2090 61319
v’ = 3 – 2 1402 1402 1402 1402
3 69070 2143 66927 2117 64810 2090 62720

Intensity of vibrational-electronic spectra: the Franck-Condon principle


The vibrational lines in a progression are not of the same intensity. Intensities of the
vibrational bands are determined by three factors:
˜ coefficient).
(1) The intrinsic strength of the transition (the B
(2) The populations of the levels involved.
(3) The overlap of the vibrational wavefunctions (the Franck-Condon factor).
€ “Since electronic transitions occur very rapidly
The Franck-Condon principle states:
(≈ 10-15 s), vibration and rotation of the molecule do not change the internuclear distance
appreciably during the transition”. The intensity of a transition is greatest for the
largest of the vibrational wavefunctions. The figure shows the probability distribution
for a diatomic molecule. The nuclei are most likely to be found at distances apart given
by the maxima of the curve for each vibrational level.

In view of the Franck-Condon principle, electronic transitions occur “vertically” on a


potential energy diagram. Consider three distinct situations.
7

(1) If the internuclear distances in the upper and lower states are equal (re” ≈ r e’), the
most probable transition is (0, 0), as indicated by the vertical line in the figure below.
However, there is a non-zero probability of (1, 0), (2, 0), (3, 0) etc. transitions. The
successive lines will therefore have rapidly diminishing intensities.
(2) If the excited electronic state has a slightly larger nuclear separation than the ground
state (re” > re’), the most probable (and thus most intense) transition is (2, 0). The
intensities of the neighbouring transitions are lower.
(3) When the excited electronic state has a considerably larger nuclear separation than
the ground state (re” >> re’), the vibrational state to which the transition takes place
has a high v’ value. Further, transitions can occur to levels where the molecule has
energy in excess of its dissociation energy. From such states the molecule will
dissociate without any vibrations and, since the fragments which are formed may
take up any value of kinetic energy, the transitions are not quantized and a
continuum results.
8

The Franck-Condon factor: quantum mechanical treatment


The Franck-Condon factor can be derived quantum mechanically from the fact that the
intensity of any transition is proportional to the square of the transition dipole moment:
Mfi = ∫ Ψ*f µ Ψi dτN
where µ is the dipole moment operator and both Ψf and Ψi are functions of the electronic
and vibrational states of the upper and lower levels, Ψ e’v’ and Ψ e”v” respectively. The
rotational component of the total wavefunction is ignored at this stage.
Separating the dipole moment operator into electronic and a nuclear components:
µ = µe + µ N
and writing Ψe’v’ as Ψ e’Ψv’ (the Born Oppenheimer approximation), the transition dipole
moment becomes:
Mfi = ∫ ∫ Ψ*e’Ψ*v’(µe + µN)Ψe”Ψv” dτedτN
= ∫ Ψ*e’ µeΨe”dτe ∫ Ψ*v’Ψv” dτN
+ ∫ Ψ*e’Ψe”dτe ∫ Ψ*v’ µN Ψv” dτN
Note that, neglecting anharmonicity, vibrational wavefunctions within one electronic
state are orthogonal, but between different electronic states they are not. Since Ψ*e’ and
Ψe” are distinct electronic states (which are orthogonal), the second term is zero, and the
transition dipole moment can be written as:
Mfi = Relec v' v"

Where the electronic dipole moment Relec = ∫ Ψ*e’ µeΨe”dτe


and

v' v" = ∫ Ψ*v’ Ψv” dτN
The intensity of the transition is proportional to Mfi2, and thus to  v' v" 2, which is
the Franck-Condon factor, consistent with the classical approach.

Dissociation energy €
The figure below shows two ways in which electronic excitation can lead to dissociation.
The figure on the left describes the case for re” >> r e’. The dashed lines of the Morse
curve represent the dissociation of the normal and excited molecule, the dissociation
energies being Do’ and Do”. The total energy of the dissociation products from the upper
state is greater by an amount Eex than that of the products of dissociation in the lower
state. This is the excitation energy of the products of dissociation. The lower
wavenumber limit of the spectral continuum must represent just sufficient energy to
9

cause no more than dissociation. The dissociation products separate with zero kinetic
energy, and we have:
v˜ continuum limit = Do” + Eex cm-1

and we can therefore determine D o” if we know Eex. Thermochemical measurements


€ lead to an approximate value of D ”, and hence, since D ” + E is accurately
often o o ex

measured spectroscopically, an approximate value of Eex is obtained. Secondly, if more


than one spectroscopic dissociation limit is found (corresponding to dissociation into
two or more different states of products), the separation between the excitation energies
often corresponds with the separations between only one set of excited states of the
atoms observed spectroscopically. This gives the nature and the energies of the excited
products.

When the internuclear distances in both states are such that transitions near to the
dissociation limit are of negligible probability, no continua appear in the spectra.
However, it is still possible to determine the dissociation energy by noting how the
vibrational lines converge. The separation between the neighbouring vibrational levels:
Δε = εv+1 – εv = ϖe[1 – 2xe(v + 1)] cm-1
This separation decreases with increasing v, and the dissociation limit is reached when
Δε tends to zero. The maximum value of v is vmax, where
ϖe[1 – 2xe(vmax + 1)] = 0
so that
vmax = (1/2xe) – 1
Since xe is of the order of 10-2, vmax is about 50.
10

The Birge-Sponer extrapolation


The expression for the separation of vibrational levels given above does not apply
exactly at high values of v because higher order anharmonicity terms are then
significant. However, it is still possible to obtain the dissociation energy by using the
Birge-Sponer extrapolation, involving plotting Δε versus the vibrational quantum
number v. The plot is not a straight line, but by extrapolating it to Δε = 0 we obtain vmax,
while the area under the plot gives the dissociation energy directly.

To summarize, provided that transitions are assigned correctly, the two methods
available for the determination of the dissociation energies Do” and Do’ are:
(1) The Birge-Sponer extrapolation
(2) If the Morse potential is assumed to be correct, the equilibrium vibration
frequencies νe’ and ν e” and the anharmonicity constants v˜ exe’ and v˜ exe” can be
determined for the two levels using:
v˜ (v” = 1 ← v” = 0) = v˜ e” – 2 v˜ exe”
€ €
v˜ (v” = 2 ← v” = 1) = v˜ e” – 4 v˜ exe”
v˜e2 €
€ De = €
4 v˜e x e €
€ €
A minimum of five entries in the Deslandres table (the bordered region) is required for
the determination of the dissociation energies of both electronic states.

Rotational fine structure of electronic-vibrational transitions
If we ignore centrifugal distortion and use the Born-Oppenheimer approximation, the
total energy of a diatomic molecule is:
˜ J(J + 1)
εtotal = εelec + εvib + B cm-1
while changes in the total energy are


11

˜ J(J + 1)] cm-1


Δεtotal = Δ(εelec + εvib) + Δ[ B
and the wavenumber of a spectroscopic line corresponding to such a change is:
˜ J(J + 1)]
v˜ spect. = v˜ v’,v” + Δ[ B cm-1

where v˜ v’,v” represents the wavenumber of electronic-vibrational transitions.
The selection rule for J depends on the type of electronic transition which the molecule
€ €
undergoes. If both€ the lower and upper electronic states are 1Σ states (with no

electronic angular momentum about the internuclear axis, S = 0), the selection rule is:
ΔJ = ±1 only for 1Σ ↔ 1Σ transitions
whereas for all other transitions (provided that at least one of the states has S ≠ 0) the
rule is:
ΔJ = 0 or ±1
In the latter case there is an added restriction that a state with J = 0 cannot undergo a
transition to another J = 0 state: J = 0 ↔ J = 0 is not allowed. Once more conservation of
angular momentum can be used to explain these selection rules. However, for
electronic transitions bond orders are expected to change and thus bond lengths are
likely to be markedly different between different electronic states (e.g. CO). The
spreading out or bunching up of transitions which occurs in vibrational-rotational
spectra may thus be more marked for electronic transitions, and “band heads” now
become evident at low J. Frequently these are more visible in spectra than the band
origin, resembling Q branches in IR spectra.
Thus for transitions between 1Σ states, only P and R branches will appear, while for
other transitions Q branches will appear in addition. We have for the difference of the
two electronic states:

ν˜ spect = v˜v',v" + B˜ ’J’(J’ + 1) – B˜ ”J”(J” + 1) cm-1


giving the P, R and Q branches as follows:
(1) P branch: ΔJ = –1, J” = J’ + 1
€ € ν˜ = v€
˜ – ( ˜
B ’ +

˜ ”)(J’ + 1) + ( B
B ˜’– B
˜ ”) (J’ + 1)2 cm-1
P v',v"

where J’ = 0, 1, 2,…
(2) R branch: ΔJ = +1, J’ = J” + 1
€ € € € € €
ν˜ R = v˜v',v" + ( B˜ ’ + B˜ ”)(J” + 1) + ( B˜ ’ – B˜ ”) (J’ + 1)2 cm-1
where J” = 0, 1, 2,…
These two equations can be combined into:
€ € € € €
€ ν˜ P,R = v˜v',v" + ( B˜ ’ + B˜ ”) m + ( B˜ ’ – B˜ ”) m2 cm-1
where m = ±1, ±2,… with positive m values for the R branch (ΔJ = +1) and negative
values for the P branch (ΔJ = –1). Note that m cannot be zero (as this would correspond
€ € € € € €
12

to J’ = –1 for the P branch), so that no line from the P and R branches appears at the
band origin, v˜v',v" . The figure shows the resulting spectra for ˜’< B
B ˜ ” and a 10%
˜ ’ and
difference in the magnitude of B ˜ ”.
B P branch lines appear at the low
wavenumber side of the band origin, and the spacing between the lines increases with
€ €
m. €
The R branch appears on the high wavenumber side, and the line spacing decreases
€ at which
rapidly with m. The point € the R branch separation decreases to zero is known
as the band head.

(3) Q branch: ΔJ = 0, J’ = J” ≠ 0 (since J = 0 ↔ J = 0 is not allowed)


ν˜ Q = v˜v',v" + ( B˜ ’ – B˜ ”)J” + ( B˜ ’ – B˜ ”) J”2 cm-1
The Q branch lines lie on the low wavenumber side of the origin.

The
€ Fortrat € €
€ diagram € €
Rewrite the expressions for the P, R and Q lines with continuously variable parameters p
and q:
ν˜ P,R = v˜v',v" + ( B˜ ’ + B˜ ”) p + ( B˜ ’ – B˜ ”) p2
ν˜ Q = v˜v',v" + ( B˜ ’ – B˜ ”) q + ( B˜ ’ – B˜ ”) q2
Each equation represents a Fortrat parabola, where p takes both positive and negative
€ € € € € €
values, while q is always positive. The band head is clearly at the vertex of the P, R
parabola. € € the position
€ € We calculate € € of the vertex by differentiation:
˜’+ B
d ν˜ P,R /dp = B ˜ ” + (B
˜’– B
˜ ”) p = 0
13

or
˜’+ B
p = – (B ˜ ”)/2( B
˜’– B
˜ ”) for band head
˜’< B
If B ˜ ”, the band head occurs at positive p values (i.e. in the R branch). Conversely,
˜’> B
if B ˜ ”, the band head is found at negative p values (i.e. in the P branch). A 10%
€ € € €
˜ ’ and B
difference between B ˜ ” gives p = 11.
€ €
€ €
€ €

Predissociation
A phenomenon called predissociation arises when the Morse curves of a molecule in
two different excited states intersect, with one of the states being stable (with a
minimum on the energy curve), and the other continuous. Suppose a transition takes
place from some lower state into the vibrational levels shown bracketed on the left.

Now if a transition takes place into the levels labelled a, b or c, a normal vibrational-
electronic spectrum occurs, complete with rotational fine structure. Two such bands
14

appear on the left of the spectrum below. However, for a transition into the levels
labelled d, e or f, the molecule may “cross over” onto the continuous curve and
dissociate. Such transition (known as “radiationless transfer”) is faster than the time
taken by the molecule to rotate (ca. 10-10 s), but usually slower than the vibrational time
(ca. 10-13 s). Thus predissociation will occur before the molecule rotates (destroying the
rotational fine structure), while the vibrational structure is preserved. If the cross-over
is faster than the vibrational time, a complete continuum will be observed.

Electronic structure of diatomic molecules

Molecular orbital theory


An atomic orbital is the space within which an electron belonging to the orbital spends
95% of its time. The orbital ψ is described by three quantum numbers: n, l and m (or lz).
The energy of the orbital also depends on electron spin. In molecular orbital theory
orbitals embrace two or more nuclei. The same rules: (1) lowest energy first; (2)
maximum of two paired electrons per orbital; (3) parallel spins in degenerate orbitals,
apply to the filling of molecular orbitals. In the LCAO theory, the approximate shape of
molecular orbitals is made up by taking sums and differences of the atomic orbitals. For
the hydrogen molecule we have the sum:
ψ H 2 = ψ1s + ψ1s
This bonding orbital (called 1sσ, as it is produced from two s atomic orbitals) is a simple
symmetrical ellipsoid. It does not change sign upon inversion about the centre of

symmetry, which is marked by the subscript g (German gerade = even). The orbital is
thus known as 1sσg.
The difference:
ψ H 2 = ψ1s − ψ1s
In this antibonding orbital (called 1sσ∗) the charge is concentrated outside the nuclei,
which repel one another. This orbital does change sign upon inversion, which is

marked by the subscript u (German ungerade = odd). It is thus known as 1sσu*.
15

The energies of the two s orbitals depend on the internuclear distance as shown below.

Using the same procedure we obtain the two 2pσ orbitals by combining atomic 2p z
orbitals:
16

which results in bonding (2pπu) and antibonding (2pπg*) orbitals.


When two 2py orbitals combine, the result is as follows:

The figure below shows the various orbitals in order of increasing energy. For light
diatomic molecules, such as Li2, there is considerable overlap and interaction between
2pπu, 2pσg, and 2sσu* orbitals. For H2 this overlap does not occur, so that the two cases
are considered separately.
17

Consider some specific homonuclear diatomic molecules:


Hydrogen, H2. In the ground state there are two 1s electrons occupying the molecular
1sσg orbital with opposing spins. Their energy is lower than in two separate atoms, so
that the molecule is stable.
Helium, He2. There is a total of four electrons. Only two can go to 1sσ g, the others
would have to go to 1sσ u*. However, since more energy would be absorbed by the
former than evolved by the latter, the molecule is unstable.
Nitrogen, N2. Given a total of 14 electrons, omitting the asterisks, the configuration is:
1sσg21sσu22sσg22sσu22pσg22pπu4
1sσg2 and 1sσu2 as well as 2sσg2 and 2sσu2 contributions cancel each other, and we are left
with six electrons, in bonding 2pσg and 2pπu orbitals. This results in a triple bond.
Oxygen, O2. There are two more electrons than in the nitrogen molecule. The lowest
available orbital is the antibonding 2pπg*. Two electrons in this orbital cancel the
contribution from two electrons in 2pπu, and we are left with four electrons, i.e. a double
bond. Further, since the antibonding 2pπg* orbitals are degenerate, the electrons occupy
one each to satisfy electron repulsion. However, according to Hund’s rule, their spins
must be parallel. The unpaired electrons make the oxygen molecule paramagnetic.

Electronic angular momentum in diatomic molecules


The total energy of an electron depends mainly on its distance from the nucleus
(represented by the quantum number n) and the orbital and spin angular momenta
(quantum numbers l and s), and on the way these are coupled together (quantum
number j).
The same applies to electrons in molecules. However, in order to discuss the
components of the vector l we need to specify some reference direction called the z
direction. The angular momentum of a diatomic molecule is referenced to the
internuclear axis. Thus, l can be thought of as precessing about the internuclear axis,
but with a well defined component (in units of h) along it. The direction of the angular
momentum along the internuclear axis does not affect the overall energy, so that for
ml > 0 there is a degeneracy of two.
When referring to molecules, the equivalent Greek symbols are used, lower case for
individual electrons, upper case for the sum over the molecule. Thus, the angular
momentum of electron i along the internuclear axis is λi (i.e. σ electrons have λ = 0, π
electrons λ = 1 etc.), and the total molecular orbital angular momentum is:

Λ = ∑ λi (equivalent to the single atom case where L = ∑ li )


i i

€ €
18

The same argument applies to spin angular momentum, with the projection of spin
angular momentum along the inter-nuclear axis denoted Σ (equivalent to S in an atom).
This Σ is not to be confused with a Σ state, for which Λ = 0. Thus for a diatomic
molecule, the total angular momentum J (neglecting nuclear spin) is now the vector sum
of the orbital (Λ), spin (Σ) and rotational (R) angular momentum:

Note that the rotational angular momentum is perpendicular to the inter-nuclear axis
(since Iz = 0). The total angular momentum thus has a projection Ω (in units of h) along
the inter-nuclear axis. There are always 2S+1 spin components, where 2S+1 is the spin
multiplicity.
Each electronic state can thus be described by a term symbol of the form:

(spin multiplicity)
[Component of L](net angular momentum component)

2S+1
or ΛΩ
NO, which in its ground state has an electronic configuration …(2pσ)2(2pπ)4(2pπ∗), has a
single unpaired electron. Its spin multiplicity is thus 2 (a doublet state), and it has total
orbital angular momentum of 1 (an unpaired π electron). Its term symbol is thus 2Π, or
more completely 2Π1/2 or 2Π3/2, depending on whether the spin and orbital angular
momentum components are parallel or anti-parallel. Clearly NO may also possess
19

rotational angular momentum, but this does not affect the term symbol which refers to
the electrons alone.
Two additional subscripts and superscripts are used in term symbols. The parity
indicates whether a wavefunction has the same sign (g) or the opposite sign (u) under
the process of inversion. Heteronuclear molecules do not have centres of inversion and
thus g and u are not applicable. The second parameter indicates the symmetry of the
electronic wavefunction with respect to reflection in a plane containing the nuclei, and
is applied only to Σ wavefunctions. If an electronic orbital wavefunction is symmetric
with respect to reflection (Σ+) it must, by the Pauli principle, be associated with an anti-
symmetric electron spin wavefunction, and vice-versa.
For example, the ground state electronic configuration of O2 is …(2pσ g)2(2pπu)4(2pπ∗g)2.
Hund’s rules indicate that the lowest energy state is the triplet (which is symmetric in

spin), and the term symbol is thus 3 Σ g .

Labelling of electronic states in atoms and molecules


€ orbital momentum spin total
momentum momentum
atoms
single electron l (symbol s, p, d for l = 0, 1, 2,…) s j
single electron
(z-component) lz sz jz
several electrons L (symbol S, P, D for L = 0, 1, 2,…) S J
several electrons
(z-component) Lz Sz Jz
molecules
single electron l s ja
single electron λ (symbol σ, π, δ for λ = 0, 1, 2,…) σ ω
(axial component)
several electrons L S Ja
several electrons Λ (symbol Σ, Π, Δ for Λ = 0, 1, 2,…) Σ Ω
(axial component)
20

The spectrum of molecular hydrogen


There are two electrons. We thus expect to find singlet and triplet states, depending on
whether the electron spins are paired or parallel. In the ground state both electrons are
in the same 1sσg orbital, and must form a singlet state. We have λ1 = λ 2 = 0 and Λ = 0:
the state is thus 1Σ. We also specify the orbital geometry: since both electrons are in the
same orbital, we write 1Σg. Finally, since the wavefunction of the electron is unchanged
upon reflection in the plane of symmetry, the full term symbol for the ground state of H2
is (1sσg)2 1Σg+.
We consider the three possible excited states: (1sσg2sσg), (1sσg2pσg) and (1sσg2pπu).

(1sσg2sσg). Since both electrons are σ electrons, Λ = λ1 + λ2 = 0. Since we are considering


singlet states, S = 0. Both constituting orbitals are even and symmetrical, so we have
(1sσg2sσg) 1Σg+.

(1sσg2pσg). We have again a 1Σ state, since both electrons are σ, but the overall state is
now odd (a combination of an electron arising from an even 1s state with an electron
from an odd p state gives and overall odd state). Thus (1sσg2pσg) 1Σu+.

(1sσg2pπu). Now Λ = λ1 + λ2 = 1, since one electron is in a π state and one originates from
a 2p orbital, the overall state is (1sσg2pπu) 1Πu.

The energies of the three states are 1Σu+ < 1Πu < 1Σg+.
Similar states are obtained by excitation to 3s and 3p states, 4s and 4p states, etc. In
addition, for n = 3, 4,… electrons can also be excited to the nd orbital. In this case we
have for the energies: (1sσndδ) 1Σg+ < (1sσndπ) 1Πg < (1sσndδ) 1Δg.
21

Transitions between the various energy levels follow the selection rules:

1. ΔΛ = 0, ±1
This is rationalised in terms of conservation of angular momentum of molecule plus
photon which has a spin angular momentum of h. For ΔΛ = 0, rotational angular
momentum changes by one (ΔJ = ± 1). Thus transitions Σ ↔ Σ, Σ ↔ Π and Π ↔ Π,
etc. are allowed, but Σ ↔ Δ, for example, is not allowed.
2. ΔS = 0.
Transitions which change the multiplicity are weak for molecules formed of light atoms,
where spin-orbit coupling is small. For heavier atoms, spin-orbit coupling effects are
larger, and transitions for which ΔS ≠ 0 can be seen.
3. ΔΩ = 0, ±1.
Σ+ states can go only to into other Σ+ states (or to Π states), while Σ– can go only to Σ– (or
Π). Thus Σ+ ↔ Σ+ and Σ– ↔ Σ–
g ↔ u (where applicable).

Consider now some triplet states of molecular hydrogen (S = 1). Both electrons occupy
the same orbital, so that the state of the lowest energy will be either (1sσ g2sσg),
(1sσg2pσg) or (1sσ g2pπu). The first two are 3Σ states, and the third is 3Π. Following the
above selection rules, the order of the energies will be:
22

(1sσg2pσg) 3Σu+ < (1sσg2pπu) 3Πu < (1sσg2sσg) 3Σg+.


Note that, because of the ΔS = 0 selection rule, transitions between singlet and triplet
states are not allowed.
For CO, the allowed electronic electric dipole transitions are:

The O2 molecule
23

The only allowed electronic electric dipole transition from the ground state is B ← X.
As O2 is homonuclear, no vibration-rotation spectrum would be expected within the X
state. Thus O2 would not be expected to absorb at wavelengths longer than ~ 200 nm.
Given the shapes of the two potential energy curves, this absorption would be expected
to be very temperature dependent (stronger at higher temperatures).
Hückel molecular orbital theory has been used to obtain a semi-quantitative picture of
the π-electronic structure of a number of organic molecules. Radiative transitions may
occur between these different electronic states.

The electronic structure of CH2O

The non-bonding state, n, corresponds to the lone pairs of electrons on the oxygen atom.
In its electronic ground state the HOMO (the X state) is the n-state. Promotion of
electrons leads to A and a states etc. The lowest energy electron promotion is to the π*
state, leading to a π* ← n transition, which occurs at ~290 nm. The other transitions
shown occur at shorter wavelengths (higher energies), e.g. the σ* ← n occurs at 190 nm.

Change of molecular shape on absorption


Consider a hydride H2A, where A is a polyvalent atom and the H–A bonds are at 90o to
one another. Two of the p orbitals of A are involved, leaving the third (“non-bonding”)
orbital unaffected. We label the bonding orbitals as a1 and a2. In the linear molecule we
encounter “orbital hybridization”, whereby atom A mixes its s orbital with one of its p
orbitals, forming two strong-bonding sp hybrid orbitals. In this configuration there are
24

now two non-bonding p orbitals, N 1 and N 2, and two bonding orbitals formed by
overlap between sp hybrids and hydrogen 1s, called b1 and b2. Given that the energy of
a non-bonding orbital is higher than that of a bonding orbital and that the sp orbital is
lower in energy than a p bonding orbital, we plot the energy changes for a smooth
transition from 90o to 180o bonding.

We see that (1) the energy of N1 remains constant; (2) the energy of the bonding orbital
a1 decreases as it passes over into the stronger b1 orbital; (3) bonding orbital a2 becomes
non-bonding N2, thus increasing in energy; (4) bonding orbitals b1 and b2 are formed by
absorption of the non-bonding s into a1; (5) 90o to 180o represent maxima and minima on
the energy curves.
Consider BeH2. Be has the electronic ground state configuration 1s22s2, with two outer
bonding electrons. Each hydrogen atom contributes a further electron, so that BeH2 has
four electrons to place into molecular orbitals. The most stable state will be for two
electrons to go into b1 and two into b 2, thus producing a linear molecule. However,
when the molecule is electronically excited, the next available orbitals for the excited
electron are N 2 or N1. With a configuration b12b21N21 the most stable state will be at a
bond angle α. Thus the excited state is bent.
Linear molecules such as CO2 and HC ≡ CH become bent on excitation.
25

Electronic spectra of polyatomic molecules


Electronic spectra of polyatomic molecules are very complex and difficult to interpret.
However, the vibrational frequencies of a particular atomic grouping within a molecule,
such as CH3, C=O or C=C, are fairly insensitive to the nature of the rest of the molecule,
as are the bond length and dissociation energy. We may therefore, as an approximation,
discuss the spectrum of each bond in isolation. Such bonds are said to be “localized”. If
we recognize the lines from such a bond in the spectrum, we may assume that its
vibrational frequency gives a good estimate of the dissociation energy of the molecule.
We note that molecules can be investigated spectroscopically in their very short-lived
excited states.
Conjugated molecules are those in which there is an alternation of single and double
bonds along a chain of carbon atoms. The π molecular orbital energy level diagrams of
conjugated molecules can be constructed using a set of approximations suggested by
Hückel in 1931. In this approach, the π orbitals are treated separately from the σ
orbitals, and the latter form a rigid framework which determines the general shape of
the molecule. All the carbon atoms are treated identically.
The effect of bond conjugation is to shift the π* ← n absorption to lower energies. When
two isolated double bonds are brought into conjugation, both levels are shifted to give
bonding and anti-bonding orbitals:

The effect is that the LUMO π* orbital is shifted downwards, and the π* ← n absorption
shifts to lower energy. Repeated conjugation increases this effect. When sufficient
conjugation is present, the energy of the LUMO π* orbital is lowered to the point where
ν2 corresponds to the visible region of the spectrum (~500 nm rather than 290 nm), a
feature which is exploited in tuneable dye lasers.
26

Absorption by a C=C double bond excites a π electron into an antibonding π* orbital.


The chromophore activity is due to an π* ← π transition, corresponding to absorption at
180 nm. When the double bond is part of a conjugated chain, the energies of the
molecular orbitals lie closer together, and the π* ← π transition moves to longer
wavelengths. It may even lie in the visible region if the conjugated system is long
enough.

An important example is the photochemical mechanism of vision. The retina of the eye
contains “visual purple”, a combination of a protein with 11-cis-retinal, which acts as a
chromophore and is a receptor of photons entering the eye. 11-cis-retinal itself absorbs
at 380 nm, but in combination with the protein the absorption maximum shifts to ca. 500
nm and tails into the blue. The conjugated double bonds are responsible for the ability
of the molecule to absorb over the entire visible region. They play another important
role: in its electronically excited state the conjugated chain can isomerize by twisting
about an excited C=C bond, forming 11-trans-retinal. The primary step in vision is
photon absorption followed by isomerization, which triggers a nerve impulse to the
brain.
27

Re-emission of energy by an excited molecule


Following the absorption of a photon, the processes which may occur are:

Dissociation. The excited molecule breaks into two fragments.


Re-emission. This is simply a reverse of absorption.
Fluorescence. The spontaneously emitted fluorescent radiation ceases immediately
after the exciting radiation is switched off. The initial radiation takes the molecule into
an excited electronic state. The excited molecule gives up its energy non-radiatively via
collisions with the surrounding molecules, thus going down the vibrational levels to the
lowest vibrational level of the electronically excited state. At that point, the
surrounding molecules may be unable to accept the larger energy difference needed to
lower the molecule to the ground electronic state. They may survive long enough to
28

undergo spontaneous emission at a larger wavelength (because of the energy lost


during the collisions). The fluorescence spectrum has a vibrational structure
characteristic of the lower electronic state.

Phosphorescence. The initial radiation takes the molecule into an excited electronic
state. Because the energy curves of the singlet and triplet state cross one another, the
molecule may undergo intersystem crossing and become a triplet state. This happens
when the molecule contains a heavier atom, such as S, because the spin-orbit coupling is
then large. The molecule deposits its energy via collisions with the surrounding
molecules, but becomes trapped at the lowest vibrational energy level of the triplet. It
cannot radiate its energy, because return to the ground state is spin-forbidden.
However, the radiative transition is not totally spin-forbidden because of the spin-orbit
coupling, so that the molecule is able to radiate weakly.
29

Phosphorescence proceeds on a much longer time scale than fluorescence.

S-ar putea să vă placă și