frontiers of molecular spectroscopy || vibrational potential energy surfaces in electronic excited...

70
CHAPTER 4 V IBRATIONAL P OTENTIAL E NERGY S URFACES IN E LECTRONIC E XCITED S TATES Jaan Laane Contents 1. Introduction 64 2. Theory 67 2.1. The quartic potential energy 67 2.2. Calculation of energy levels 70 2.3. Kinetic energy functions 71 3. Experimental Methods 73 4. Electronic Ground State 77 4.1. One-dimensional potential functions 77 4.2. Two-dimensional potential energy functions 79 5. Electronic Excited States 84 5.1. Cyclic ketones 84 5.2. Stilbenes 97 5.3. Bicyclic aromatics 104 5.4. Cavity ringdown spectroscopy of enones 124 6. Summary and Conclusions 129 Acknowledgments 130 References 130 Abstract Several spectroscopic methods, including infrared and ultraviolet absorption, Raman, jet-cooled laser-induced fluorescence (LIF), and cavity ringdown, have been utilized to map out the vibrational quantum states of molecules in their ground and excited electronic states. Data on the higher excited vibrational levels for large-amplitude vibrations such as ring-puckering, ring-twisting, ring-flapping, and internal rotation allow one- or two-dimensional potential energy surfaces (PESs) to be accurately deter- mined. In many cases, ab initio and/or DFT computations are utilized to complement the experimental work. Following a discussion of theory, experimental methods, and com- putational methods, the spectroscopic results and PESs for several types of molecules are presented. First, the PESs for the carbonyl wagging vibration of seven cyclic ketones in their S 1 (n,% ] ) excited states are reviewed. Except for 2-cyclopentenone (2-CP), which is conjugated and planar, the PESs have a barrier to planarity which increases with angle strain. PESs for the ring-bending and ring-twisting vibrations were also determined for these ketones in both their ground and excited states. The LIF study of trans-stilbene 63

Upload: jaan

Post on 23-Dec-2016

214 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

C H A P T E R 4

VIBRATIONAL POTENTIAL ENERGY SURFACES

IN ELECTRONIC EXCITED STATES

Jaan Laane

Contents1. Introduction 642. Theory 67

2.1. The quartic potential energy 672.2. Calculation of energy levels 702.3. Kinetic energy functions 71

3. Experimental Methods 734. Electronic Ground State 77

4.1. One-dimensional potential functions 774.2. Two-dimensional potential energy functions 79

5. Electronic Excited States 845.1. Cyclic ketones 845.2. Stilbenes 975.3. Bicyclic aromatics 1045.4. Cavity ringdown spectroscopy of enones 124

6. Summary and Conclusions 129Acknowledgments 130References 130

Abstract

Several spectroscopic methods, including infrared and ultraviolet absorption, Raman,jet-cooled laser-induced fluorescence (LIF), and cavity ringdown, have been utilized tomap out the vibrational quantum states of molecules in their ground and excitedelectronic states. Data on the higher excited vibrational levels for large-amplitudevibrations such as ring-puckering, ring-twisting, ring-flapping, and internal rotationallow one- or two-dimensional potential energy surfaces (PESs) to be accurately deter-mined. In many cases, ab initio and/or DFT computations are utilized to complement theexperimental work. Following a discussion of theory, experimental methods, and com-putational methods, the spectroscopic results and PESs for several types of moleculesare presented. First, the PESs for the carbonyl wagging vibration of seven cyclic ketonesin their S1(n,�] ) excited states are reviewed. Except for 2-cyclopentenone (2-CP), whichis conjugated and planar, the PESs have a barrier to planarity which increases with anglestrain. PESs for the ring-bending and ring-twisting vibrations were also determined forthese ketones in both their ground and excited states. The LIF study of trans-stilbene

63

Page 2: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

and two substituted stilbenes allowed two-dimensional PESs for the internal rotations ofthe phenyl groups to be calculated for both ground and S1(�,�] ) states. The torsionabout the CTC double bond in the S1(�,�] ) state was also investigated to understandthe photoisomerization. The ring-puckering and ring-flapping of four molecules in theindan (IND) family were investigated in both ground and S1(�,�] ) states, and the PESs,which differ substantially between the electronic states, were determined. Results for1,3-benzodioxole (13BZD) were particularly interesting in that it possesses the anomericeffect that results in a non-planar five-membered ring. In its ground state, the barrier toplanarity is relatively small due to suppression of the anomeric effect due to interactionswith the benzene ring. However, the suppression is considerably reduced in the S1(�,�] )state and the barrier almost doubles. Spectroscopic results and PESs for two dihydro-naphthalenes, tetralin (TET), and two benzodioxoles are also reviewed as is the work on2-indanol. The latter molecule possesses intramolecular hydrogen bonding that wasanalyzed in both ground and excited electronic states. Finally, the cavity ringdownspectra (CRDS) of 2-CP and 2-cyclohexenone (2-CHO) are presented along with theirring-puckering and ring-twisting PESs. This highly sensitive technique allows the T1(n,�] )triplet levels and the corresponding PESs to be determined for 2-CP. The flipped spin forthe triplet state results in a barrier to planarity, whereas the ground and S1(n,�] ) statesare planar.

Keywords: molecular vibrations; potential energy surfaces; molecular structure; infraredspectroscopy; Raman spectroscopy; laser-induced fluorescence; large-amplitudevibrations; electronic excited states; fluorescence spectra

1. INTRODUCTION

Many discussions of potential energy functions concentrate on the use of theharmonic oscillator. This is the case, for example, for representing the vibration of adiatomic molecule. Whenever greater accuracy is desired, perturbations to theharmonic oscillator are used and this is known as the anharmonicity. Similarly,the vibrations of polyatomic molecules can be considered to be nearly harmonic,and quadratic force fields have often been utilized. Although these kinds ofinvestigations can be very informative about bonding forces, direct structuralinformation can generally not be obtained. Potential functions that are not harmo-nic, however, can be much more informative, especially if their energy minima donot correspond to the coordinate origins. The much studied inversion vibration ofammonia provides an illustrative example. Dennison [1,2] first recognized in 1932that there would be inversion doubling for ammonia due to the fact that theinversion vibration has a double minimum potential energy function. Since then,this vibration has been studied many dozens of times [3–5]. Figure 1 presents ourown representation of the potential energy function for the ammonia inversionvibration along with the experimentally observed energy spacings. This is assumedto be a one-dimensional system where the only coordinate is that for the inversionmotion. Although ammonia has six vibrations, the inversion can be consideredindependently as all of the other vibrations are either of different symmetry

64 Jaan Laane

Page 3: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

(E symmetry species instead of A1 for the inversion) or are much higher infrequency (�1, the symmetric stretching mode). The function in Figure 1 is wellrepresented by a harmonic potential energy function with a Gaussian barrier

V = Ax2þBe�Cx2

; ð1Þwhere x represents the inversion coordinate and where x = 0 represents a planarammonia molecule. A, B, and C are potential energy parameters. The function inFigure 1 has a barrier of 2031 cm�1 (5.80 kcal/mol), and this value is the energydifference between the planar form of ammonia and its equilibrium structure. Theenergy minima correspond to the equilibrium values of the out-of-plane distance x.When used in the Schrodinger equation, this three-parameter function repro-duces the experimental data very accurately. As we showed many years ago [6], asimpler function also reproduces the observed frequencies almost as well. This hasthe form

V = Ax4þBx2; ð2Þwhere a negative B value is utilized to produce the barrier. As we shall see, this formof the potential energy function has proven to be extremely valuable for our workon the ring systems to be discussed below.

Another type of vibration that has a potential function very different from aharmonic oscillator was recognized by Bell in 1945 [7]. He postulated that a four-membered ring molecule such as cyclobutane should have a ring-puckering vibra-tion governed by a quartic potential energy function

V = Ax4: ð3Þ

4000

547

512.0

501.3

284.6

629.3

932.5968.3

2031

35.8

0.79

xmin0–xmin

3000

2000

1000

0

V (c

m–1

)

X

Figure 1 Potential energy function for the inversion of ammonia.

Vibrational Potential Energy Surface 65

Page 4: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

As it turns out, and as will be discussed below, the function in Equation (2) moreappropriately represents these types of ring-puckering vibrations, which are low infrequency and have large amplitudes of motion. The first far-infrared observation ofthe ring-puckering transitions was the study of trimethylene oxide (TMO) reportedfrom the R. C. Lord laboratory at MIT nearly half a century ago [8,9]. Later studiesthere and at Berkeley established that TMO has a tiny barrier to planarity and asmall negative B coefficient for Equation (2)[10–13]. Studies of cyclobutanone(CB) [13] and trimethylene sulfide [14] at Berkeley by the Gwinn and Straussresearch groups and of silacyclobutane (SCB) at MIT by Laane and Lord [15]further paved the way for the dozens of far-infrared investigations to follow. Laaneand Lord further recognized that molecules such as cyclopentene [16] and 1,4-cyclohexadiene [17] could be considered to be ‘‘pseudo-four-membered rings’’ andtheir ring-puckering spectra could be fit well with the one-dimensional function inEquation (2). As a result, the ring-puckering of many five- and six-membered ringswere studied in the following decades. Much of the work on four-, five-, and six-membered rings through 1979 was reviewed by Carreira et al. [18].

The situation with cyclopentane and other unsaturated five-membered rings,however, is more complicated. Pitzer and co-workers [19,20] initially showed thatthe two out-of-plane ring motions of cyclopentane could be described by a nearlyfree pseudorotation and a radial mode. They postulated that the pseudorotationcould be represented by a one-dimensional potential energy model with V = 0 andthat this would give rise to energy levels 2B apart, where B is the pseudorotationalconstant. In 1968, Durig and Wertz [21] observed a number of the pseudorota-tional transitions as combinations with a CH2 bending mode, and in 1988, Baumanand Laane [22] presented improved spectra, a comprehensive analysis, and a two-dimensional function for this molecule. One- and two-dimensional energy calcula-tions based on experimental data have followed for many other five-membered ringmolecules, and reviews have discussed these [23,24]. Two-dimensional studieswere also extended to six-membered ring molecules such as cyclohexene and tobicyclic molecules, and these will be discussed later.

In the 1970s, vapor-phase Raman spectroscopy also became available to com-plement the far-infrared studies. TMO was investigated by Kiefer and Bernstein[25] in 1970 and cyclopentene by Chao and Laane [26] in 1972 in early work.Many studies since then, including the study of cyclobutane [27], utilized vapor-phase Raman as an additional tool. In addition to out-of-plane ring vibrations andinversions, internal rotations (or torsions) have also been investigated by far-infraredand Raman spectroscopy. One- and two-dimensional periodic potential energyfunctions have been determined for many molecules, and a number of reviews areavailable [28,29]. The energy level calculations have generally been based onmethods developed in the Laane laboratories [30].

These types of potential energy studies took a major step forward in the 1980swith the development of supersonic jet cooling and with the availability of tunablelaser sources in the ultraviolet region. This allowed the vibronic energy states ofelectronically excited states to be determined directly and therefore made it possibleto determine the potential energy functions and structures for these states. Much ofthe discussion here will focus on these excited state investigations.

66 Jaan Laane

Page 5: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

2. THEORY

2.1. The quartic potential energy

Bell theorized [7] that cyclobutane should have a quartic potential energy functionfor its ring-puckering vibration, but he never detailed his derivation. This was donein 1991 by Laane [31]. Figure 2 shows a four-membered ring, such as the fourcarbon atoms of cyclobutane, with four equal bond angles �. The degree ofpuckering is measured by the puckering coordinate x, which reflects the distancethat each carbon atom has moved away from a planar structure. Hence, thedifference between the two ring diagonals is 2x. The structure of the ring is fullydefined by x and the carbon—carbon bond distance R, which is assumed to befixed as bond stretching force constants are much larger than angle bendingconstants. The angle strain at each angle � is assumed to be the sole contributorto the potential energy, and the angle bending forces are assumed to be harmonic.

The derivative of the resulting potential energy function is based on somegeometrical relationships:

sinð�=2Þ= A=R ð4Þ

A2þ4x2 = B2 ð5Þand

A2þB2 = R2: ð6ÞFrom Equations (5) and (6)

A =1

2R2�2x2

� �1=2

ð7Þ

and then

sinð�=2Þ= ð1=2� 2x2=R2 Þ1=2: ð8ÞDifferentiation gives

cosð�=2Þd�= ð�4x=R2Þð1=2� 2x2=R2 Þ �1=2dx: ð9Þ

A A

A

B

φ2

2x

Figure 2 The ring-puckering coordinate x and the geometrical parameters for thecyclobutane ring.

Vibrational Potential Energy Surface 67

Page 6: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

From Equations (6) and (7),

B =1

2R2þ2x2

� �1=2

ð10Þ

and since

cosð�=2Þ= B=R; ð11Þwe have

cosð�=2Þ= ð1=2þ 2x2=R2 Þ1=2: ð12Þ

Then from Equations (9) and (12)

d�=�8xdxðR4� 16x4 Þ �1=2: ð13Þ

As x = 0 for the planar structure, dx = x. Then D�= d� is given by

D �=�8x2 ðR4�16x4 Þ �1=2: ð14ÞThe bond distance R is typically an order of magnitude greater than the puckeringcoordinate x at the equilibrium position. Thus, R4ii16x4 and

D� ffi �8x2=R2: ð15ÞBell’s conclusion that the potential energy is quartic (Equation (3)) must clearly bebased on the assumption that the potential energy function for the ring-puckeringvibration arises from a harmonic dependence on each angle change D� for each ofthe four angles:

V = 4½1=2k� ðD�Þ2�= 128k�x4=R4: ð16Þ

Thus, a in Equation (3) is given by

a = 128k�=R4 ð17Þ

However, Equation (16) neglects initial strain present at each of the already strainedC—C—C angles in the four-membered ring. Assuming that each of the ring anglesprefers to have an ideal value of �ideal (probably tetrahedral), instead of Equation(16) we then have

Vang = 2k� ð�ideal��Þ2 ð18Þand

�=�0 þ D �=�0 � 8x2 ðR4�16x4 Þ �1=2 ffi �0 � 8x2=R2; ð19Þ

where �0 is the internal ring angle for the planar conformation (90�). Defining theinitial strain by

S0 =�ideal � �0 ð20Þ

68 Jaan Laane

Page 7: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

then yields the potential energy

Vang ffi 2k� ðS0þ 8x2=R2 Þ2 = 2k�S02þð32k�S0=R

2Þx2þð128k�=R4Þx4 ð21Þ

or

Vang = aangx4þbangx

2þV0 ð22Þ

where

aang = 128k�=R4; bang = 32k�S0=R

2; V0 = 2k�S2

0 : ð23Þ

In quantum mechanical calculations for the energy levels, the value of V0 can be setequal to zero so that V(0) = 0. It should also be noted that if D� is given by theexact expression in Equation (14) rather than Equation (15), then each R2 inEquation (23) should be replaced by ðR4�16x4 Þ1=2.

The effect of torsional interactions is not included in Equation (23). Forcyclobutane the planar conformation has each of the four CH2 groups eclipsedby two adjacent groups, and the torsional contribution to the potential function isat a maximum at x = 0. A three-fold potential function can be used to approximatethe internal rotation about each bond. When this is converted to a dependence onx, we have a power series in x2 to represent this contribution

Vtors = btorsx2þatorsx

4þ � � � ð24Þ

Thus, when Equation (24) is added to Equation (22) to determine the overallpotential function, the function will be of the double-minimum type when btors isnegative and btorsj j > bang. It will have a single minimum when bang þ btors>0. Itcan also be seen that ators will contribute somewhat to the quartic coefficient in thetotal potential function, but aang in Equation (22) will dominate.

The derivation described above only applies directly for cyclobutane, but the sameprinciples apply for other four-membered ring molecules. Moreover, they can also beapplied to ‘‘pseudo-four-membered rings’’ such as cyclopentene, 1,4-cyclohexadiene,and their analogs. Figure 3 shows how the ring-puckering coordinate is defined forthese larger ring molecules.

2x

2x

2x

Figure 3 Definition of the ring-puckering coordinate for three types of ringmolecules.

Vibrational Potential Energy Surface 69

Page 8: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

2.2. Calculation of energy levels

Initial studies of ring-puckering potential energy functions assumed a fixed reducedmass � for the computations. In this case the one-dimensional wave equation is

��h2=2�� �

d2c=dx2þVc= Ec: ð25Þ

This can be transformed to the reduced form [6] using

Z = 2�=�h2� �1=6

a1=6x ð26ÞE = A� ð27Þ

A = �h2=2�� �2=3

a1=3 ð28Þand

B = 2�=�h2� �1=3

a�2=3b: ð29ÞThis results in

�d2c=dZ2þ Z4þBZ2� �

c=�c; ð30Þ

where Z is the dimensionless coordinate. B defines the quadratic contribution thatmay be positive or negative, and the � are the eigenvalues that are proportional tothe vibrational quantum states E. Each value of B defines a set of eigenvalues thatmay be computed and then scaled by the parameter A to best fit the experimentaldata. Hence, the effective potential energy function is

V = A Z4þBZ2� �

: ð31ÞFigure 4 shows the first 17 eigenvalues calculated for this function for different values ofB (with A = 1). This data were presented in tabular form in 1970 when the computa-tion of eigenvalues was slow and tedious and became known as Laane’s tables [6].Laane utilized numerical methods based on Milne to calculate the eigenvalues. Hislaboratory and others nowadays utilize matrix diagonalization methods based onharmonic oscillator functions to calculate energy levels. For B = 0, the function isthat of a pure quartic oscillator. For positive B, a mixed quartic/quadratic functionexists, and a pure harmonic oscillator is approached as B!1. When B is negative, thefunction represents a double minimum potential with a barrier of B2/4, and pairs ofenergy levels begin to merge below the barrier. When B!�1, the pairs of energylevels become equally spaced and also approach those of a harmonic oscillator. Figure 5shows the experimentally determined potential energy functions for the ring-puckeringvibrations of several different molecules with values of B ranging from B =þ4.68 to�9.33. 2,5-Dihydrothiophene [32] (DHT) has a mixed quartic/quadratic potential(B = 4.68) while 3-silacyclopent-1-ene [33] (SCP) is a nearly perfect quartic oscillatorwith B = –0.17. Both molecules are planar. TMO [8–13], 1,3-disilacyclobutane [34](DSCB), cyclopentene [16], and SCB [15] have increasingly negative B values (–1.47,–4.65, –6.17 and –9.33, respectively) and increasing barriers to planarity. Theirpotential energy functions show how the energy levels merge to produce (inversion

70 Jaan Laane

Page 9: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

doubling.) As the barrier increases, more and more levels can be seen to become nearlydoubly degenerate. This is also clear in Figure 4 where the lower � values merge as Bbecomes more and more negative. The eigenvalues at the specific B values shown inFigure 5 correlate to the energy levels on the potential energy curves when theappropriate scaling factor A is used.

2.3. Kinetic energy functions

Many of the original ring-puckering studies utilized the wave equation (Equation(25)) with the mixed quartic/quadratic potential energy (Equation (2)) and a fixedreduced mass to fit the experimental data. However, as this vibration has a largeamplitude and the reduced mass changes with the vibrational coordinate, thecalculations can be improved by utilizing a computed reduced mass function thatdepends on the coordinate. In one dimension, the Hamiltonian becomes

H xð Þ= ��h2=2� �

@=@x g44 xð Þð Þ@=@xþV xð Þ; ð32Þ

where g44 is the coordinate-dependent reciprocal-reduced mass function that canbe represented by a polynomial

g44 = gð0Þ

44 þ gð2Þ

44 x2 þ gð4Þ

44 x4 þ gð6Þ

44 x6; ð33Þwhere the g

ðiÞ44 are the parameters determined computationally to best fit the

coordinate dependence. The odd-powered terms in this polynomial representation

90

100

SCB CP DSCB TMO SCP DHT

16

B 2

4= Barrier ht.

1615

1413

12

11

10

9

8

7

6

5

4

3

2

1

n = 0

14.15

12.13

10.11

8.9

6.7

4.5

2.3

0.1

80

70

60

50

40Eig

en v

alue

s

30

20

10

0–20 –18 –16 –14 –12 –10 –8

B in V = Z 4

+ BZ 2

–6 –4 –2 0 2 4 6 8 10

Figure 4 Eigenvalues � as a function of B for the potential energy of the formV= z4þBz2.

Vibrational Potential Energy Surface 71

Page 10: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

are zero if puckering up and down are equivalent. Typically, only the terms upthrough the sixth power are included. The subscripts on g44 reflect the fact thevalues 1 to 3 are reserved for the molecular rotations. As shown by Malloy [35], thereduced mass for the puckering motion can be readily calculated at differentcoordinate values using vector methods. The difficult part is to correctly modelthe vibrational motions as derivatives of the type @r

*

i=@Qj must be computed. Theseare then utilized to set up a G matrix from which the g44 terms as a function of thecoordinate Qj can be determined. Laane and co-workers [36–41] have describedthe computation of the kinetic energy expressions for several different types ofvibrations using vector methods for both one- and two-dimensional cases. As anexample, Figure 6 shows the two-dimensional kinetic energy (reciprocal-reducedmass) function calculated for the ring-puckering of phthalan (PHT) [42] in terms ofits ring-puckering (x1) and ring-flapping (x2) coordinates. The coordinate depen-dence is clearly evident. In two dimensions, the Hamiltonian becomes

H x1 ; x2

� �=��h2=2

�@=@x1 g44 x1 ; x2

� �� �@=@x1 þ @=@x2 g55 x1 ; x2

� �� �@=@x2

þ @=@x1 g45 x1 ; x2

� �� �@=@x2Þ@=@x2 g45 x1 ; x2

� �� �@=@x1�

þ V x1 ; x2

� �; ð34Þ

2000

1600

1200

800

400

0

1000

800

600

400

200

0

11

11

12

12

13

14

10

10

10

11

SH2Si

O

15

14

13

12

12

12

17117.1

207.2

199.2

190.2

177.1

164.7

153.8124.2

13594.

0.3

0.003

113.2

110.5

106.4

101.0

98.5

92.1

86.2

79.2

74.749.935.4

133.5 141.8

157.8

16

15

14

13

11

11

1110

10

109

9

9

8

8

8

7

7

7

6

6

6

5

5

5

3

334

44

2

22

11

1

00

056.0 (2.99)

0.9131.6

61.0

66.0

70.6

74.7

78.5

82.0

85.3

88.2

91.2

93.9 132.6

126.1

119.4

113.3

107.5

99.8

92.0

76.6

83.125.2

127.149.554.5

9

9

9

8

8

8

7

7

7

6

6

6

5

5

5

44

43

3

32 2

21 110 00

39.253.4

89.9104.7

118.1128.8

138.9

147.3

155.0

161.6

168.2

174.9

180.4

185.5

190.1

53.4

61.0

67.1

72.0

76.3

80.1

83.4

86.6

89.3

91.8

94.2

86.5

95.2

101.5

106.6

110.8

114.5

117.7

120.6

123.1

124.9

125.3

V (c

m–1

)V

(cm

–1)

V (c

m–1

)

V (c

m–1

)

V (c

m–1

)

V (c

m–1

)

1250

1000

750

500

250

0

1200

1000

800

600

400

200

0

1200

800

400

00

1000

750

500

250

–0.2 0.0 0.2

X (Å)

–0.2 0.0 0.2

X (Å)

–0.2 0.0 0.2

X (Å)

–0.2 0.0 0.2

X (Å)–0.2 0.0 0.2

X (Å)

–0.2 0.0 0.2

X (Å)

SiH2

H2Si

SiH2

Figure 5 Experimentallydetermined potential energy functions for various differentBvalues.

72 Jaan Laane

Page 11: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

where V(x1,x2) typically has the two-dimensional form to be discussed later. Thegij(x1,x2) for i,j = 4,5 is calculated as a function of both coordinates, again usingvector methods. Utilization of these kinetic energy functions rather than a fixedreduced mass tends to affect the potential energy parameters by less than 5% butdoes reduce the deviation between observed and calculated frequencies by about afactor of 2 (typically from an average 1–2 cm�1 deviation for the fixed reduced massmodel to half of that).

3. EXPERIMENTAL METHODS

Figure 7 shows the types of spectroscopic transitions used in the experimentalwork for determining the vibrational energy levels for both the ground (S0) and thefirst excited S1 electronic states of the molecules to be discussed. The diagram showsfar-infrared absorption and Raman transitions that are used to determine thevibrational quantum states for the electronic ground state. These generally havequantum number changes of Dv = 1 and 2 for the principal transitions. Laser-induced fluorescence (LIF) of the jet-cooled molecules results from transitionswhich for the most part originate from the vibrational ground state in S0. However,if the molecular jet is warmed somewhat, or if the ground state is nearly degenerate,weak transitions from the first excited vibrational level of low-lying states can alsobe seen. For LIF, the laser system is tuned, and when the energy of the vibroniclevel is reached, the fluorescence signal is detected as fluorescence excitation spectra(FES). For molecules excited from the v = 1 level, fluorescence results when thefrequency matches the separation between v = 1 and the vibronic level. Figure 8shows a schematic diagram of the LIF system in our laboratories for FES and singlevibronic level fluorescence (SVLF) investigations. Sensitized phosphorescence exci-tation spectroscopy (SPES) can also be carried out. A Nd : YAG laser is used to

0.008g44

0.004

0.20.0

0.20.1

0.0

–0.1

–0.2–0.2x2 (Å)

x 1 (Å

)

–0.4

0.010

0.006

Figure 6 Kinetic energy function for the ring-bending (g44) of phthalan in terms of thebending (x1) and twisting (x2) coordinates.

Vibrational Potential Energy Surface 73

Page 12: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

Virtual state

Electronicexcited state

S1

Electronicground state

Far-infrared Raman SVLF(DF)

FES(LIF)

Electronicabsorption

4

3

2

1v ′ = 0

4

3

2

1v ″ = 0

S0

Figure 7 Spectroscopic transitions involving the ground and excited electronic states ofmolecules.

Laser pulse fromOPO or dye laser

ISA HR640Czerny–Turnermonochromator

Height andorientationmirrors

Supportoptics

Phosphorescencecold tip

Outline of chamberbodyPhosphorescence

excitation photomultiplier

Fluorescenceexcitation photomultiplier

Skimmer

Focusing lens

Pulsed nozzleAccelerationgrids

Microchannel platedetector

Figure 8 Laser-induced fluorescence (LIF) system in the Laane laboratories. (See color plate 5).

74 Jaan Laane

Page 13: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

drive an optical parametric oscillator (OPO). The molecules are cooled through asupersonic jet expansion using a pulsed valve. Detection is with a photomultipliertube (PMT) or by a time-of-flight mass spectrometer. Dispersed fluorescence orSVLF spectra can also be recorded using a single 0.64-m monochromator and PMTor CCD detection. More details can be found elsewhere [43–45].

The ultraviolet absorption spectra are recorded at room temperature and hencetransitions can originate from more than a dozen of the low-energy vibrationallevels in the S0 state. The symmetry of the molecules restricts which transitions areallowed, and generally only v = even to even or odd to odd transitions are observed.The jet-cooled LIF spectra are most valuable in that they clearly show whichtransitions originate from the vibrational ground state. However, the Franck–Condon factor limits the number of S1 vibronic states that can be accessed. Inthese cases, the ultraviolet absorption spectra are very helpful in that many of thehigher vibronic states can be reached only when the transitions originate fromupper vibrational levels in S0. The ultraviolet absorption spectra in our laboratoryare recorded on a Bomem 8.02 Fourier transform spectrometer, which is capable of0.02 cm�1 resolution and operates up to 50,000 cm�1. In some cases, as many as100,000 scans were averaged to achieve high signal to noise ratios.

The far-infrared (3–400 cm�1) and mid-infrared (400–4000 cm�1) spectra are alsorecorded on the Bomem instrument. Heatable long path multiple-reflection cells up to20 m are used to investigate weak signals. Before FT-IR spectroscopy became commonin the 1970s, the best far-infrared instrument was constructed by Jarrell Ash for the R. C.Lord laboratory [46] at M.I.T. at a cost of about one quarter million 1964 dollars. Thisinstrument is shown in Figure 9. It produced many excellent spectra including those ofSCB [15] and cyclopentene [16]. Its monochromator had a 5-m path length and its longpath cell, sticking out to the left, had a volume of more than 200 L, so it required about agram of sample to produce one torr of vapor pressure. Many grating and filter changes

Figure 9 Jarrell Ash far-infrared spectrometer atMIT in the1960s. (See color plate 6).

Vibrational Potential Energy Surface 75

Page 14: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

were required to investigate the 25–400 cm�1 region, and several days of scanning wererequired to do this. Figure 10 shows a small portion of the far-infrared spectrum of CP[16] recorded with this instrument. This shows the rotational fine structure of the nearlysymmetric top molecule as well as several ring-puckering bands.

Raman spectra are recorded on a JY U-1000 monochromator equipped with aCCD detector. Either a Coherent Innova 20 argon ion laser or a Coherent Verdi 10operating at 532 nm has been used as the excitation source. As many of the sampleshave high boiling points, a special cell had to be constructed [47] for heating samplesas high as 350� C to obtain substantial vapor pressures. This is shown in Figure 11.

60

Abs

orpt

ion

70 80

Wavenumber (cm–1)90

H2O

Figure 10 Aportion of the far-infrared spectrum of cyclopentene recorded on theMIT JarrellAsh instrument.

Figure 11 High-temperatureRaman cell for vapors. (See color plate 7).

76 Jaan Laane

Page 15: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

4. ELECTRONIC GROUND STATE

4.1. One-dimensional potential functions

Figure 5 already showed several experimentally determined one-dimensional potentialenergy functions. Here additional detail will be provided for other examples. Figure 12shows the far-infrared spectrum of 2,3-dihydrofuran [48] and Figure 13 shows itspotential energy function determined from the data for its ring-puckering vibration.The experimental data were fit very well using the simple two-parameter potential

cm–1160

11–1210–11

9–108–9

7–8

6–7

5–6 4–5

3–4

1–22–3

0

0.5

1.0

140 120 100 80

Abs

orba

nce

Figure 12 Far-infrared spectrum of 2,3-dihydrofuran.

1600 167.9 480.812

11

10

9

8

7

6

5

4

3

21

0

460.6

[441.2]

422.1

398.9

372.7

343.2

301.8

275.9

191.2

93.4

160.4

152.7

147.4

141.1

133.3

124.9

114.9

103.8

83.5

[18.9]88.8

1200

800

400

0

–0.2 –0.1 0 0.1 0.2

X(Å)

V(c

m–1

)

Figure 13 Ring-puckering potential energy function for 2,3-dihydrofuran.

Vibrational Potential Energy Surface 77

Page 16: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

function of Equation (2). This molecule has only a small barrier to planarity of 93 cm�1,but this is sufficient to pucker it with a dihedral angle of 22�. The barrier to planarityarises from the torsional interaction between the two CH2 groups.

Another especially interesting molecule is 1,3-dioxole in that it might be expectedto be planar as it has no CH2—CH2 torsional interactions. However, it shows theanomeric effect that results in a puckered molecule. Figures 14 and 15 show its far-infrared and Raman spectra of the vapor, and Figure 16 presents the potential energyfunction that has a substantial barrier of 325 cm�1 [49]. The double quantum jumps

Water

O

300 200

Wavenumber (cm–1)100 40

O

Tra

nsm

ittan

ce

Figure 14 Far-infrared spectrum of 1,3-dioxole vapor.

100 200

1–3

3–5

4–6

5–7 6–8 0–4

0–22–4

Wavenumber (cm–1)300

Figure 15 Raman spectrum of 1,3-dioxole vapor.

78 Jaan Laane

Page 17: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

of the puckering observed in the Raman spectra were very important for comple-menting the far-infrared data and helping to make the assignments for the spectralbands. The anomeric effect, which is present in molecules with the —O—CH2—O—configuration, is responsible for twisting the molecule out-of-plane. This will beexamined in more detail when the 13BZD molecule is discussed later.

Another study worth noting is that of 1,3-cyclohexadiene as this was achievedusing high-temperature vapor-phase Raman spectroscopy [50]. Figures 17 and 18show the Raman spectra and potential energy function, respectively. The numberof Raman transitions observed is remarkable because they extend above the highbarrier to planarity (1132 cm�1) for this molecule.

The experimentally determined potential energy functions and barriers of thefour-membered ring and pseudo-four-membered ring molecules for which Equa-tion (2) was utilized have been presented elsewhere [18] and in many of thereferences herein. As discussed above, the a and b potential energy parameters forthe most part reflect angle strain and torsional forces, respectively.

4.2. Two-dimensional potential energy functions

The one-dimensional approximation described above can only be applied when asingle vibration such as the ring-puckering can be investigated independently from allthe other vibrations. When a second low-frequency vibration is present, two-dimen-sional potential energy surfaces (PESs) can be utilized. These generally have the form

V = ax41þbx2

1þcx42þdx2

2þex21x

22: ð35Þ

Two-dimensional studies have been carried out for molecules in the cyclopentaneand cyclohexene families, both of which have ring-bending and ring-twisting

–0.2

0

500

1000

1500185.8

180.0

172.0

165.9

155.0

144.4

133.3

115.9

113.147.9

158.6

319.7

208.6 161.3

277.0

206.3161.3

229.0

249.0

278.0

321.8

0.1

234

5

6

7

8

9

10

11

12

299.0

–0.1 0 0.1

X (Å)

V (c

m–1

)

0.2

Figure 16 Ring-puckering potential energy function of 1,3-dioxole.

Vibrational Potential Energy Surface 79

Page 18: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

vibrations, and for bicyclic molecules similar to IND, which has ring-bending andring-flapping vibrations.

The cyclopentane molecule is a particularly interesting case as it has 10 equiva-lent bent forms and 10 equivalent twisted forms [19,20]. Moreover, the energydifference between bent and twisted forms in less than a few cm�1 [51]. Figure 19shows how these 20 structures readily interconvert into one another resulting in thepseudorotational process. Figure 20 shows the combination bands of the

0–2

Ram

an in

tens

ity

2–4

4–6

200 180 160

cm–1140 120

6–8

9–118–10

13–1511–13

10–12

Figure 17 Vapor-phase Raman spectra of 1,3-cyclohexadiene at150�C.

0,1

4,5

2,3

6,7

8,9

10111213

1514

16

146 145

142

155

169.6

181.6

191.3

1132

153

117138116

155 157.9

169.6

181.5

191.1

198.7 199.8200

600

1000

1400

V (c

m–1

)

Observed Calculated

0 0.5–0.5

τ (radians)

Figure 18 Potential energy function for the ring-twisting vibration of 1,3-cyclohexadiene.

80 Jaan Laane

Page 19: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

0.30

0.25

0.15

0.20

0.10

0.05

1

2

34

5

0.00

φ

B1+T4+

T2+

B2–

T5–

B3+

T1+

B4–

T2–

B5+

T3+B1–

T4–

B2+

T5+

B3–

T1–

B4+

B5–

T3–

0.95

0.90

0.85

0.80

0.75

0.70

0.65

0.60

0.55 0.500.45

0.40

0.35

Figure 19 Pseudorotation scheme for cyclopentane. B andTrefer to bent and twist structures,respectively.

1400

0

50

100

% T

1450 1500

C5H10

cm–1

Figure 20 Cyclopentane pseudorotational bands in combinationwith a CH2 bendingmode.

Vibrational Potential Energy Surface 81

Page 20: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

pseudorotation with a CH2 bending mode [52]. These are characteristic of nearlyfree pseudorotation where the energy levels are separated by 2Bpseudo, the pseudor-otational constant. In this case, Bpseudo ffi 2:7 cm�1. Figure 21 (left) shows the two-dimensional PES determined for cyclopentane by Bauman and Laane [52], whichalso fit the data for the radial mode near 272 cm�1. A similar result was firstpresented by Carreira, Mills, and Person [53]. This function has a barrier toplanarity of 1808 cm�1, and the bent and twisted structures all have the sameenergy. What this PES shows is that all 20 conformations of cyclopentane caninterconvert essentially without encountering any energy barriers, and the moleculecan totally invert in this manner without passing through the planar structure.When the cyclopentane ring is substituted with a heteroatom, a barrier to pseudor-otation results. This is the case for 1,3-oxathiole, and its two-dimensional PES isalso presented in Figure 21 (right) [54,55]. As can be seen, the molecule is twisted,and the bent structures correspond to saddle points that lie 570 cm�1 above theenergy minima. The anomeric effect helps to reduce the energy of the bent form.The planar structure is 2289 cm�1 higher in energy. Although there is a barrier topseudorotation at the bent conformation, the molecule can still invert more readilyalong this pathway than by passing through the high energy planar structure.

Cyclohexene [56,57] and similar molecules containing heteroatoms have alsobeen investigated using far-infrared spectroscopy and two-dimensional PESs[58,59]. These molecules in general are twisted with high barriers to planarities.The lowest frequency spectra arise from the ring-bending motion, but twistingbands and twisting-bending combinations were also utilized to determine the PESs.

We have also reported results for numerous molecules in the indan family wherethese bicyclic systems have interacting low-frequency ring-bending and ring-flappingmodes. The definitions of these two coordinates are shown in Figure 22. As anexample, Figure 23 shows the far-infrared spectrum of phthalan [60]. Figure 24shows its PES in terms of these two coordinates [42]. The molecule is quasi-planarwith only a tiny barrier to planarity of 35 cm�1. A one-dimensional potential functionwas not able to fit the experimental data well, but the two-dimensional PES together

2500

200015001000500

–0.8

–0.8 –0.40

0.40.8

Ring-twisting coordinate

–0.8–0.4 0

0.4 0.8

Ring-twisting coordinateRing-bending coordinate

Ring-bending coordinate

–0.40

0.40.8

–0.4–0.2

0

0.20.4

0

25002000

150010005000

2500

2000

15001000

Potential energy

Potential energy

5000

2500

20001500

1000500

0

Figure 21 Two-dimensional potential energy surface (PES) for the ring-bending (x1) andring-twisting (x2) of cyclopentane (left) and1,3-oxathiolane (right).

82 Jaan Laane

Page 21: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

2x1

x2

2 2τ

1, 3

1, 3

4, 5 4, 5

Figure 22 Ring-bending (x1) and ring-flapping (x2 or t) coordinates for indan and relatedmolecules.

100

20

40

80

60

100

%T

80 60 40

cm–1

Figure 23 Far-infrared spectrum of phthalan in the 25^105 cm�1 region.

0

200

400

600

0

0.25

–0.25–0.2

–0.10

0.1

X 1 (A

)X2 (A)

0.2

0

200

400

600

Potential energy (cm

–1)

O

Figure 24 Potential energy surface (PES) for phthalan in terms of its ring-bending (x1) andring-flapping (x2) modes.

Vibrational Potential Energy Surface 83

Page 22: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

with the calculated kinetic energy expansion similar to Equation (34) very nicely fitthe data. The energy level pattern was highly unusual due to the kinetic energyinteractions. Far-infrared data and PESs for the related molecules coumaran (COU)[61,62], 13BZD [63], and IND [64] have also been published, and these will bediscussed later.

5. ELECTRONIC EXCITED STATES

5.1. Cyclic ketones

When a ketone molecule undergoes an n ! �] transition, the oxygen atom of thecarbonyl group typically bends out of the plane of the adjoining carbon atoms.Similarly, the formaldehyde molecule has long been known to be non-planar in itsS1(n,�] ) excited state [65]. In the 1990s, we carried out a number of investigationsusing FES of cyclic ketones including 2-cyclopenten-1-one (2CP) [66], 3-cyclopen-ten-1-one (3CP) [67], CP [68,69], CB [70], bicyclo[3.1.0]hexan-3-one [71] (BCH),tetrahydrofuran-3-one (THFO), and tetrahydrothiophen-3-one (THTP) [72].

In each case, we analyzed the vibronic bands resulting from the carbonyl waggingmotion in the S1(n,�] ) state and determined the one-dimensional potential energyfunction governing this vibration. The bending and twisting vibrations associatedwith the conformational changes of the rings were also examined in detail. Except for2CP, where sufficient conjugation is retained after the n! �] transition to keep thecarbonyl oxygen in the plane of the ring, the other cyclic ketones each have double-minimum carbonyl wagging potential energy functions for the S1(n,�] ) state demon-strating that the carbonyl groups are bent out of the ring planes. In the followingsections, the results on these molecules will be presented.

5.1.1. 2-Cyclopenten-1-one (2CP)Figure 25 shows the survey FES of 2CP [66]. The spectra were recorded in a regionfrom below the electronic origin (27,210 cm�1) to about 1800 cm�1 beyond it. Incomparison to 3CP and CP, which will be considered later, the 2CP electronic bandorigin is considerably lower. This is the expected result from the conjugationbetween the CTO and the CTC groups, which results in a lower energy �] orbital.The electronic origin for 2CP is also extremely intense in contrast to the S1(n,�] )origins of similar molecules. This is the result of a planar excited state structure and ahigh Franck–Condon factor. The frequencies of some of the fundamental vibrationsdetermined for the electronic ground state and S1 excited state are given in Table 1.In addition to these, more than 50 other bands were seen in the FES, with 34 of theseinvolving �30, the ring-puckering, and 20 involving �29, the ring-twisting (several areassociated with both). Combinations with the CTO in-plane (�19) and out-of-plane

84 Jaan Laane

Page 23: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

(�28) wags, which occur at 348 and 422 cm�1, are also common. These modes are bothconsiderably lower in frequency than in the electronic ground state reflecting thedecrease in � character of the CTO bond. The intense bands at 1357 and1418 cm�1 for 2CP are due to the CTO and CTC stretches, respectively. Thesetwo vibrations are Franck–Condon active due to the increased bond lengths for bothbonds. Intense combination bands for each of these stretches were observed with thecarbonyl in-plane wag, 191

0. All of the other ring mode transitions were also observed,and many of them were also found to be associated with combination bands.

The ring-puckering potential functions for the S0 and S1(n,�] ) states, whichwill be compared later to that in the T1(n,�] ) state, were determined based on thefar-infrared [73] and FES [66] data, both of which show series of band progressions.

345

1600 1200 800 400 0

350 355 360

Frequency shift (cm–1)

365

Wavelength (nm)

Inte

nsity

Figure 25 Fluorescence excitation spectrum of jet-cooled 2-cyclopentenone (2CP).

Table 1 Vibrational frequencies (cm�1) for the ground andexcited S1(n,�] ) states of 2CP

Approx. description Ground Excited

�5 CTO stretch 1748 1357�6 CTC stretch 1599 1418�13 Ring mode 1094 1037�14 Ring mode 999 974�15 Ring mode 912 906�16 Ring mode 822 849�17 Ring mode 753 746�18 Ring mode 630 587�19 CTO def (ll) 464 348�26 a-CH bend 750 768�28 CTO def (?) 537 422�29 CTC twist 287 274�30 Ring-puckering 94 67

Vibrational Potential Energy Surface 85

Page 24: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

Equation (2) fits the data very well for both the ground and the excited states. Themolecule is planar in both states but becomes much less rigid in the excited statedue to the decreased conjugation resulting from the transition to the antibondingorbital.

5.1.2. 3-Cyclopenten-1-one (3CP)Figure 26 shows the jet-cooled fluorescence excitation spectrum of 3CP [67]. Theband origin is observed at 30,238 cm�1. For the molecule lying in the xz plane,each v = 0 (in the electronic ground state) to v = n (in the A2 electronic excitedstate) transition of the CTO wag has B2 vibrational symmetry for n = odd but hasA1 vibrational symmetry for n = even. Only transitions to the n = odd states can beobserved. These show up as intense Type B bands arising from A2 x B2 = B1

symmetry. The first five of these transitions are shown in Figure 26. Because thev = 0 and v = 1 levels in the S1(n,�] ) state are near-degenerate, the band origin liesvery close to the 0! 1 frequency. The other bands in the spectrum include manycombinations of the CTO wag with the ring-puckering vibration and also com-binations of these with other fundamentals including the CTO stretch.

For 3CP, the reduced mass and the carbonyl wagging potential energy para-meters that best fit the observed frequency separations are listed in Table 2. Theexperimentally determined potential energy function is shown in Figure 27 alongwith both the observed and the calculated frequency separations. The minimumenergy corresponds to wagging angles of +26�, and the barrier to inversion is926 cm�1 (2.65 kcal/mol). For 3CP, the ring-puckering frequency of 127 cm�1 inthe S1 state is considerably higher than the value of 83 cm�1 in the ground stateindicating a stiffer and asymmetric puckering potential function.

30,900

600

295

400 200 0

Inte

nsity

30,700 30,500 30,300 30,100Wavenumber (cm–1)

0 2920 293

0

2910

3010

2930 302

0 2910 302

0

3010

3030

Figure 26 Fluorescence excitation spectrum of 3-cyclopenten-1-one.

86 Jaan Laane

Page 25: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

5.1.3. Cyclopentenone (CP)The FES spectrum of cyclopentenone [68,69] is shown in Figure 28. The bandorigin is at 30,276 cm�1. In the ground state, the molecule is twisted, and inthe C2v approximation, the vibrational ground state is nearly doubly degeneratewith symmetry species A1 and A2. The twisting conformation (and degeneracy)carries through to the electronic excited state. The purely electronic transition is1A2 1A1, which is forbidden in the C2v approximation. However, combinations

Table 2 Potential energy parameters and reduced masses for CTO wagging vibrations in theS1(n,�] ) electronic state

Molecule m (au) V=ax4þ bx2 Barrier (cm�1) �min

a(cm^1/A‡4) b(cm^1/A‡2)

THTP 3.572a 13.4� 103 �5.94� 103 659 20�CP 5.569b 10.49� 103 �5.34� 103 680 22�BCH 6.49c 7.84� 103 �5.18� 103 873 23�3CP 5.260d 8.11� 103 �5.48� 103 926 26�THFO 5.203e 8.51� 103 �6.26� 103 1152 26�CB 4.244f 2.47� 103 �4.38� 103 1940 41�

a g44 = 0.2799� 0.1872x2þ 0.0887x4� 0.0165x6.b g44 = 0.17957� 0.049144x2þ 0.014227x4� 0.002181x6.c g44 = 01541þ 0.1483x� 0.0428x2þ 0.0143x3� 0.04199x4þ 0.0586x5þ 0.0239x6; V includes �0.097� 103x3.d g44 = 0.19012� 0.054853x2þ 0.016335x4� 0.002554x6.e g44 = 0.1920� 0.0546x2þ 0.0152x4� 0.0021x6.f g44 = 0.23565� 0.076454x2þ 0.024096x4� 0.003922x6.

–40°

0

Pot

entia

l ene

rgy

(cm

–1)

500

1000

1500

2000

–20° 0° 20° 40°

Wagging angle

336

304

309

325

386

20

939

0,1

2,3

45

67

8

9

10

O11

Figure 27 Vibrational potential energy function for the CTO out-of-plane waggingvibration of 3-cyclopenten-1-one.

Vibrational Potential Energy Surface 87

Page 26: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

with odd quantum transitions of the CTO wagging with B2 symmetry results inType B bands from B1 symmetry. If either the ground or excited electronic state isalso in combination with the near-degenerate A2 twisting state, the even quantaCTO wagging transitions can also be observed as Type A (A1)bands A2�A2� B2ð Þn = A1 for n = even½ �. As can be seen in Figure 29, the transi-tions for both even and odd quantum states of the CTO wag in the S1 state arereadily observed.

Figure 30 shows the CTO wagging potential energy function for cyclopenta-none. The barrier is 680 cm�1 and the energy minima are at +22�. The kinetic andpotential energy terms are given in Table 2. The comparison between the S0 and S1

states for the fundamental vibrational frequencies of several other modes is given inTable 3.

The observed data for the ring-twisting and ring-bending motions of CP in theS0 and S1 states have also been analyzed [69]. The fundamental frequencies for thesetwo modes are changed little in the two states as the ring has a similar twistedconformation for each state. However, the two-dimensional PESs for these twomodes have been determined for both states and they are quite different. In S0, thebarrier to planarity is 1408 cm�1 and that for pseudorotation is 1358 cm�1. Thetwist angle is 29�. These barrier values become 1445 and 596 cm�1, respectively,for the excited state. Figure 31 shows the PES for the S1(n,�] ) state. The lowerpseudorotation barrier shows that the bent conformation has become considerablylower in energy.

32,000

Inte

nsity

259

2000

O

1500 1000 500 0

31,000Wavenumber (cm–1)

0

1810 258

0

1810 256

0

1810 252

0

1810 254

0

2570

2550

2530

2510

000

Figure 28 Fluorescence excitation spectrum of cyclopentanone.

88 Jaan Laane

Page 27: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

5.1.4. Cyclobutanone (CB)The FES spectrum of CB is shown in Figure 32 [70]. The �26 carbonyl bendingbands are labeled, and transitions up to the 11th quantum state can be seen. Inaddition to the carbonyl wagging, puckering data for the S1(n,�] ) state can also bededuced from the low-frequency data shown in Figure 33. The potential energyfunction for the carbonyl wagging is shown in Figure 34, and this highly strained

510410500400

228

140

040B1B2A1A2

A1

A2

A1

A1

A2

A2

A1A2

A1A2

A1

A2

A1

A2

A1

A2

B2

B2

B1

B1

B2

B1

B2

B1

B2

B1

B2

B1

B2

B2

B1

B1

S1

S0

B1B2A1A2

A1

A2

A1

A2

B1B2A1A2

130030120020

110

010

309

209

315

113

88

81

88

89

91

93

95

91

013103003112012102002

111011101001

238

236

Ring twisting C O wagging (out of plane) Ring bending

446

105

005

104

004

103

003

102

002

101

001

400

E – ν0

(cm–1)

ν18 ν25 ν26

ν18ν25ν26

E (cm–1)

200

0

400

200

0

310210300200

110010100000

500

400

300

200

100

000

Figure 29 Energy level diagram for cyclopentanone.

Vibrational Potential Energy Surface 89

Page 28: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

four-membered ring can be seen to have a wagging barrier (2149 cm�1) muchhigher than the other ketones. At its energy minimum, the wagging angle is alarge 39�. The presence of the wagging barrier also complicates the ring-puckeringvibration. As shown in Figure 35, both the puckering and the carbonyl inversioncan change the conformation of the molecule and result in two puckered forms ofdifferent energy. Hence, the puckering potential function is asymmetric in theS1(n,�] ) state. Figure 36 shows several similar one-dimensional ring-puckeringpotential energy functions which reproduce the experimental data very well.Figure 37 shows the two-dimensional PES in terms of the carbonyl wagging andring-puckering modes.

5.1.5. Bicyclo[3.1.0]hexan-3-oneThe bicyclic BCHO molecule is similar to CP except that it has an attached three-membered ring that induces asymmetry for both the carbonyl wagging and thering-puckering modes. The FES spectrum shows about three dozen bands within1000 cm�1 of the 00

0 band at 30,262 cm�1 [71]. Many of these are associated withthe ring-puckering (�22) and/or the carbonyl wagging (�21), and these wereanalyzed to determine the one-dimensional ring-puckering potential energy func-tion in Figure 38 and the function for the wagging in Figure 39. The molecule isnot puckered but shows the expected asymmetry. The barrier to carbonyl inversionis 873 cm�1, and the asymmetry resulting from the energy difference between thetwo inversion directions is approximately very small, 36 cm�1.

1600

1200V

(cm

–1)

800

400

0

–40° –20° 0°Wagging angle

201

187

164

156

88

209

309 315680

0,1

23

45

6

7

8

9

20° 40°

Figure 30 CTO wagging potential energy function and observed energy spacings forcyclopentanone.

90 Jaan Laane

Page 29: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

Table 3 Comparison of frequencies of several vibrations of cyclic ketones in the ground and S1(n,�] ) electronic excited states

Vibration 2CP THTP CP BHO 3CP THFO CB

S0 S1 S0 S1 S0 S1 S0 S1 S0 S1 S0 S1 S0 S1

Ring-puckering 94 67 67 58 95 91 86 134 83 127 59 82 36 106Ring-twisting 287 274 170 — 38 238 — — 378 377 228 224 — —CTO wag o.p. 537 422 427 326 446 309 — 315 450 336 463 344 395 355CTO wag i.p. 464 348 483 329 467 342 — — 458 339 463 365 454 392CTO stretch 1748 135 1760 1240 1770 1230 — — 1773 1227 1755 1232 1816 1251

i.p., in-plane; o.p., out-of-plane.

Vib

rationalPotentialEnergyS

urface9

1

Page 30: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

0

1000

20002000

1000

0

S1(n,π∗)

Potential energy (cm

–1)

0.2

0

–0.2

Ring-bending coordinate

–0.4–0.8

–0.4

Ring-twisting coordinate

0

0.4

Figure 31 Vibrational potential energy surface for cyclopentanone in the S1(n,�]) electronic

excited state.The contour lines are150 cm�1apart.

32,000

1600

O

1200 800 400 0

Inte

nsity

31,500 31,000

Wavenumber (cm–1)

30,500

2690

26110

2670 265

0

2630

2610

000

Figure 32 Fluorescence excitation spectrum of cyclobutanone.

92 Jaan Laane

Page 31: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

30,800

400

O

200 0

30,600 30,400

Wavenumber (cm–1)

Inte

nsity

2630

910 273

0 2010

2620 270

1

2610 272

0 2710

2610

2701

000

Figure 33 Low-frequency region of the cyclobutanone fluorescence excitation spectrum.

0

400

–40° –20°

309 315

209

88

156

164

187

201

d0

680

0,1

23

4

5

6

7

8

9

Wagging angle

V (

cm–1

)

0° 20°

800

1200

1600

Figure 34 Carbonyl wagging potential energy functions and observed energy spacings forcyclobutanone.

Vibrational Potential Energy Surface 93

Page 32: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

5.1.6. Tetrahydrofuran-3-one (THFO) and tetrahydrothiophen-3-one (THTP)The FES spectra of THFO and THTP were investigated to determine theircarbonyl wagging potential energy functions in their S1(n,�] ) states [72]. Bothspectra are unusual in that the band intensities die out by about 1500 cm�1. TheTHFO spectrum is shown in Figure 40. The carbonyl wagging bands (�28) are

o

o

o

o

Puckering

Puckering

InversionInversion

φ

φ

φ

φ

Figure 35 Conformations of cyclobutanone in the S1(n,�]) state resulting from the ring-

puckering or CTO wagging (inversion) vibrations. For the (x1,x2) notation, the (þ,þ) and(�,�) conformations are equivalent as are the (þ,�) and (�,þ) conformations.

–0.1

0106

0

1

2

3

c (×105):

166

185

200

400

600

V – V0(cm–1)

–0.0 –0.1

x (Å)

8.304.980.00

Figure 36 One-dimensional potential energy curves of the form V= ax24þ bx2

2þ cx23 for

ring-puckering of cyclobutanone in the S1(n,�]) state.

94 Jaan Laane

Page 33: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

–40°

0.1

0.0

–0.1

0° 40°

Rin

g-pu

cker

ing

(Å)

C O wagging angle

Figure 37 Two-dimensional potential energy function for the carbonyl wagging and ring-puckering coordinates of cyclobutanone.The dots mark the energy minima.The first contourline lies 50 cm�1above each minimumpoint.The others are100 cm�1apart.

–0.1

–20 0 20

S1

S0

40

0

500

1000

1500

0.0 0.1

121

Dihedral angle

123

127

129

133

134

X(Å)

V(cm–1)

0.2

Figure 38 Ring-puckering potential energy function forbicyclo[3.1.0]hexan-3-one (BCHO)in the S1(n,�

]) electronic state. The function has been translated so that x= 0 represents theplanar structure for the five-membered ring.This is compared to the function in the S0 groundstate (dashed curve).

Vibrational Potential Energy Surface 95

Page 34: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

labeled, and these were used to determine the potential function of Figure 41. Thebending modes (�30) for the hindered pseudorotation are also labeled. The THTPspectra, which can be found elsewhere [72], were also used to determine itswagging potential function which is shown in Figure 42. As can be seen, theTHTP barrier of 659 cm�1 is considerably less than that for THFO (1152 cm�1).

5.1.7. SummaryThe studies of these cyclic ketones have demonstrated a number of interestingproperties. One of these is that the inversion barrier for the carbonyl groupincreases with the angle strain. Molecular mechanics (MM3) calculations wereused to estimate the CCC angle at the carbonyl atom, and Figure 43 plots the

–40°

0

500

1000

1500

–20° 0°

Obs.

873 315

260

144

79

22

36

0 1

2 3

4 5 6 7

20°

Wagging angle

V (c

m–1

)

40°

Figure 39 Carbonyl wagging potential energy function for bicyclo[3.1.0]hexan-3-one(BCHO).

32,000

2000 1500 1000 500 0

31,500 31,000 30,500

3 2 1

30,000

Wavenumber (cm–1)

2810

Inte

nsity

0 2890 288

0

2870 286

0

284,50

282,30 0280,1

30v0

Figure 40 Fluorescence excitation spectrum of tetrahydrofuran-3-one (THFO).

96 Jaan Laane

Page 35: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

barrier height versus this angle [72]. As this angle becomes more strained, thebarrier increases. The cyclopentanone molecule appears to have too low a barrier,but this may result from its lower CCC angle resulting from the effect of torsionalforces.

Table 2 summarizes the results for the carbonyl inversion potential energyfunctions, and Table 3 summarizes data for several of the other relevant vibrationsof these cyclic ketones.

5.2. Stilbenes

Both trans- and cis-stilbene are stable molecules with the conformation of the transform slightly lower in energy by about 4.6 kcal/mol and with the barrier to internalrotation (isomerization) of 48.3 kcal/mol [74]. This precludes any isomerizationabout the double bond occurring in the ground state, but the trans to cis photo-isomerization in the S1(�,�] ) state has been often investigated [74,75]. In thisexcited state, the lowest energy form of stilbene has a twisted configuration andonly a small barrier to internal rotation exists between the trans and the twist forms.A dynamical study in 1992 concluded that this barrier is about 1200 cm�1 [76].

In 1995, Chiang and Laane [75] studied the LIF spectra of trans-stilbene. TheFES is shown in Figure 44 and the SVLF in Figure 45. The analysis of the spectra

–40°

0

600

1200

160 170

139 144

56 170 162 54

252 242

320 326

1152

365 358

2,3

0,1

4,5

6

7

8

9

1800

–20° 0° 20° 40°

Wagging angle

Observed Calculated

Pot

entia

l ene

rgy

(cm

–1)

Figure 41 Two similar potential energy functions for the carbonyl wagging oftetrahydrofuran-3-one (THFO).

Vibrational Potential Energy Surface 97

Page 36: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

–40° –20°

1800

250 256

234 237

214

175

220

170

20943

216 41

343

659

346

0,1

2

3

4

5

6

7

8

1200

600

0

0° 20° 40°

Wagging angle

Pot

entia

l ene

rgy

(cm

–1)

Observed Calculated

Figure 42 Potential energy function for the carbonyl wagging of tetrahydrothiophen-3-one(THTP).The solid curve corresponds toV1and the dashed curve toV2.

2000

O

OO

OO

O

O

s

1500

Bar

rier

(cm

–1)

1000

500

090° 100°

C—C(O)—C angle

110°

Figure 43 Correlation of inversion barrier with � (CCC angle at the carbonyl carbon atom).The dashed line indicates the‘‘strain-free’’angle according to theMM3 calculation.

98 Jaan Laane

Page 37: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

is greatly complicated by the fact that trans-stilbene has eight low-frequencyvibrations and these needed to be assigned before the internal rotation vibrationscould be analyzed. This assignment was greatly aided by the vapor-phase Ramanspectra of this molecule [77], which clearly identified the Raman active modes.From the Raman and the LIF spectra, the energy diagram in Figure 46 wasdetermined, and the analysis of the three torsional (internal rotational) modesbecame feasible.

Figure 47 defines the �1 and �2 coordinates that represent the phenyl torsionsand � that represents the internal rotation about the CTC bond. The phenyltorsions were analyzed two dimensionally, and a computer program was written to

0 100

32,200 32,000 31,800

Wavenumber (cm–1)

31,600

Inte

nsity

019

4576

118

238

400

410

430

454

623

609

216

202

200 300 400 500 600

H

HC C

700

Figure 44 Dispersed fluorescence spectra of the 000 band of trans-stilbene.

300 0100200

32,500 32,400 32,300

Wavenumber (cm–1)

32,200

Inte

nsity

3704

4802

3604

2401

3502

2360

000

H

H

C C

3742

3722 37

Figure 45 Fluorescence excitation spectra of jet-cooled trans-stilbene.

Vibrational Potential Energy Surface 99

Page 38: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

calculate the coordinate-dependent kinetic energy function in terms of �1 and �2.This was then used in the potential energy computation with

V ð�1; �2Þ= 12 V2ð2þcos 2�2ÞþV12cos 2�1cos 2�2þV 012 sin 2�1 sin 2�2:�

ð36Þ

Ene

rgy

(cm

–1)

99 48

4

3

48 37

35110

120

118

109 S1

S0

108

35

35

47.5

47.5

99

2

2

1

0 0, 1

2, 3

4, 5

6, 7

8, 9

6, 7

4, 5

2, 3

0, 1

4, 5

2, 3

0, 1

1

0

2

1

1

0

101

10158

17141412109

12, 1310, 118, 96, 74, 52, 30, 1

ν35

C Ctorsion

Ring flap Phenyltorsion

Phenyltorsion

ν36 ν37 ν48

0

Figure 46 Vibrational energy spacings for four low-frequency vibrations of trans-stilbene inits ground and excited states.

θ

φ 2

φ 1

Figure 47 Torsional coordinates for trans-stilbene.

100 Jaan Laane

Page 39: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

A separate computer program was written to determine the energy levels for thisPES and to optimize the potential energy parameters V2;V12; and V 012. Figure 48shows the PES for the S1(�,�] ) excited state. The coordinates are defined so that at�1 =�2 = 90�; both phenyl groups are in the —C—CTC—C— plane. Thepotential energy has its maxima of 3000 cm�1 (2V2) at �1 =�2 = 0�. There arefour equivalent energy minima at �1 = –90� and �2 = –90�, where the entireskeleton of the molecule lies in a plane. The barrier to rotating a single phenylring is 1670 cm�1. The S0 PES is qualitatively similar with a barrier of 3100 cm�1.The barrier to rotating a single phenyl group is only 875 cm�1. The S0 and S1

surfaces each require only three potential energy parameters, and these do anexcellent job of fitting all of the observed energy spacings.

For the internal rotation about the CTC double bond, a computer programwas written to calculate the coordinate-dependent kinetic energy function. For theS0 ground state, energy level calculation was used together with the one-dimen-sional potential function

V ð�Þ= 12 V1ð1�cos �Þþ 1

2 V2ð1�cos 2�Þþ 12 V4ð1�cos 4�Þ:=

��ð37Þ

Here, �T 0� and 180� correspond to the trans and cis isomers, respectively, V1 isthe energy difference between the trans and the cis forms, V2 almost entirelydetermines the barrier, and V4 is a shaping parameter. When the literature values[66] V1 = 1605 cm�1 (4.6 kcal/mol) and V2 = 16,892 cm�1 (48.3 kcal/mol) areused, excellent frequency agreement between the experimental and the calculateddata is obtained using V4 = –900 cm�1. The data is confined to the bottom regionof a potential energy function that has a very high barrier, so the calculation of V2

by extrapolation does not give a very accurate value (15,000+ 3000 cm�1), but thisis consistent with the literature values. The lower half of Figure 49 shows thepotential energy curve for the S0 state.

4000

3000

2000

1000

0

4000

3000

2000

1000

0

–90

S1

0

90

0

0

00 3000

–90

90

0

Ene

rgy

(cm

–1)

φ 2φ 1

Figure 48 Potential energy surface for the phenyl torsions of trans-stilbene in its S1(�,�]) state.

Vibrational Potential Energy Surface 101

Page 40: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

The CTC torsion data for the S1(�,�] ) state were only observed for the transconformation due to the Franck–Condon principle. Figure 49 (top section) showsthe qualitative potential energy curve estimated in the literature. The figure alsoshows the vibronic energy levels observed within the trans potential well. No FEScould be observed going to levels in the twist well. In addition, the cis well isapparently too shallow to allow transitions originating from the S0 state of cis-stilbene. Consequently, the fluorescence data allow only the shape of the trans wellto be calculated, and this is shown as the dotted line in Figure 49. This can berepresented by [75]

V ð�Þ= 12 V1ð1�cos �Þþ 1

2 V2ð1�cos 2�Þ ð38Þ=�þ1

2 V4ð1�cos 4�Þ þ 12 V8ð1�cos 8�Þ:==

4

–2

–4

20

15

10

0Obs.

16892

1605

cis

Twist

Twist

Obs.

Trans

Lit

cis

0 90θ (degrees)

180 270

5

2

0E

nerg

y (×

103 c

m–1

)E

nerg

y (×

103 c

m–1

)

S1

S0

Figure 49 Potential energy function for the internal rotation of trans-stilbene about the CTCbond. �=0� corresponds to the trans isomer.

102 Jaan Laane

Page 41: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

V1 represents the trans/cis energy difference, and V2 primarily determines the depthof the twist well. V4 primarily determines the trans ! twist barrier while V8 is aminor shaping term. Analysis of the experimental data, which extend 1230 cm�1

above the trans well minimum, shows that the trans ! twist barrier is somewhathigher than the 1200 cm�1 value estimated from the dynamics data [76].

Investigations have also been carried out for 4,40-dimethoxy- and 4,4-dimethyl-trans-stilbene [78,79], and similar results have been obtained for the internal rota-tions. In addition, the data for the methyl torsions of the dimethyl compound werealso analyzed and the quantum states for these resembled those of m-xylene ratherthan p-xylene. Figure 50 shows the energy diagram for the methyl internal rotationsfor both S0 and S1 states, and these were fit with the one-dimensional periodicfunctions shown in Figure 51. The S0 state has a tiny six-fold barrier, whereas theS1(�,�]) state has a small but higher three-fold barrier.

(2,3)(3,2)

(1,4)(4,1)

(1,4)(4,1)(0,4)(4,0)

(0,4)(4,0)

(1,3)(3,1)

(1,3)(3,1)

72

77.4

46.8

51.3

33.8

–7.5

–3.7

30.2

(0,3)(3,0)

(0,3)(3,0)

(1,2)(2,1)

(1,2)(2,1)

(0,2)(2,0)

(0,2)(2,0)

(1,1)

(1,1)

(0,0)

S0

0

20

40

60

E (

cm–1

)

80

S1

0(0,0)

(0,1)(1,0)

(0,1)(1,0)

(2,2)

Figure 50 4,4-dimenthyl-trans-stilbene (DMS) methyl torsion quantum states and transitionsfor S0 and S1electronic states.The transitions to the (0,3) and (1,3) are not completely shown.

Vibrational Potential Energy Surface 103

Page 42: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

5.3. Bicyclic aromatics

5.3.1. Indan and related moleculesIndan (IND), phthalan (PHT), coumaran (COU) and 1,3-benzodioxole (13BZD),shown below, are ‘‘pseudo-four-membered-ring’’ molecules because their ring-puck-ering vibrations resemble those of four-membered rings as the two atoms of the five-membered ring joined to the benzene ring tend to move together as a single unit.

These molecules also have low-frequency ring-flapping vibrations (also calledbutterfly motions) of the same symmetry species as the ring-puckering, and thesecan interact strongly with the puckering motions. Figure 22 showed the definitionof these two low-frequency vibrations. The puckering (x1 or �) is basically the out-

–180

0

20

40

60

80

100

0

20

40

60

80

100

–120 –60 0 60

φ1 (degrees)

V (c

m–1

)V

(cm

–1)

S1

4

3

2

1

0

S0

4

5

3

2

1

0

120 180

–180 –120 –60 0 60

φ1 (degrees)

120 180

Figure 51 One-dimensional potential energy functions for a single methyl torsion in DMS inits S0 and S1 electronic states.

104 Jaan Laane

Page 43: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

of-plane motion of the apex CH2 group or oxygen atom relative to all the otheratoms in the two rings. The flapping (x2) is the motion of the entire five-memberedring relative to the benzene ring. As with the other molecules discussed, vectormethods have been developed to represent the motion of all the atoms relative tothe two coordinates so that the kinetic energy (reciprocal-reduced mass) expansionsfor the molecules can be calculated as a function of the coordinates. These expan-sions have been presented elsewhere [59–64].

These bicyclic molecules have been studied in both their ground and excitedelectronic states. The discussion here will concentrate on the S1(�,�]) excited states.Figure 52 shows a molecular orbital diagram correlating the benzene � orbitals tothose of PHT (and the other three molecules in this group). The reduction of thebenzene D6h symmetry to C2v in PHT produces the splitting of the degenerate E1g

and E2u orbitals into A2, B2 pairs. The transition resulting in the S1(�,�]) state is alsoshown in the figure.

5.3.1.1. PhthalanThe far-infrared spectrum of PHT in Figure 23 shows not only single quantumtransitions but also weaker double and triple quantum jumps [59,60]. The primaryseries is in the 30–105 cm�1 region, and this also shows side bands arising fromtransitions in the flapping excited state. Figure 53 shows the energy map andobserved transitions for the S0 ground state for the ring-puckering (vP) and ring-flapping (vF) states. The primary puckering sequence is irregular and cannot be fitwell with a one-dimensional function [59]. However, when the ring-flappingcoordinate is included for a two-dimensional calculation that includes a cross-

B1g

E2u

E1gA2

B2

A2

A2– –+ +

++

++ +

+

+

+ –+ –

––

+

++ +

+

++

+

+

B2

B2

C2vD6h

E

A2u

O

Figure 52 Molecular orbital diagram correlating to �orbitals of benzene to those of phthalan.The skeletal atoms of phthalan are assumed to be in the xz plane.

Vibrational Potential Energy Surface 105

Page 44: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

kinetic-energy term, an excellent agreement with the experimental data is obtainedusing a PES with a tiny barrier to planarity of 35 cm�1 [60].

The LIF of jet-cooled PHT and its ultraviolet absorption spectrum were used todetermine the puckering and flapping quantum states for the S1(�,�] ) excited state[80]. Figure 54 shows the region of the UV spectrum near the band origin at37,034.2 cm�1. The puckering transitions to the vibronic levels in S1 are labeled. Atwo-dimensional PES was again required to reproduce the experimental data.However, in this case, the PES has no barrier and the surface is considerably stifferalong the puckering coordinate. Figure 55 shows a cut of the PES along the ring-puckering coordinate. The puckering component is nearly pure quartic while theflapping is quadratic. The excited state energy level separations are also shown. Abinitio calculations also predict a stiffer excited state PES for PHT.

5.3.1.2. CoumaranThe original study of the far-infrared spectrum of COU [61] identified an inversionsplitting of 3.1 cm�1, and this was assigned to the v = 2�3 separation. However,Ottavani and Caminati [81] later utilized millimeter wave spectroscopy and showedthat the v = 0�1 splitting was 3.12 cm�1, and this assignment was then utilized todetermine the one-dimensional ring-puckering potential energy function for COU

8

110.3

232.0

102.8

215.2

214.5

213.3

211.1

73.7

35.0

76.1

O

94.1

105.3

75.2

72.5

33.0

(230.1)

229.2

221.9

220.9

218.5

215.9

7

105.8

100.6

93.2

98.5

74.3

200

400

600

E (cm–1)

800

0

70.4

30.9

vF = 1

vF = 2

vF = 0vP = 0

4

3

2

1

5

6

Figure 53 Energy level diagram for the ring-puckering vibration (�P) in different ring-flapping (�F) states of phthalan.

106 Jaan Laane

Page 45: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

[62]. The inversion barrier was found to be 154 cm�1, and the puckering angle atthe energy minima is 25�. A triple zeta ab initio calculation predicted a barrier of238 cm�1 and a dihedral angle of 26.5�.

The LIF spectra of COU were first reported by Watkins and co-workers [82]but were not correctly assigned due in part to the incorrect far-infrared assign-ments [61]. Yang and co-workers [83] in the Laane laboratory later reported theFES, SVLF, and UV absorption spectra that not only confirmed the electronic

2–4

0–31–35–6 2–1'

1'–3

2'–0'

1'–2

2'–1'

4'–3'

3'–2'

1'–1'

2'–2'2–2

1–1 0–0

0'–1

0–0' 2'–34–0'

2–36–6

4–4

O

cm–1

100200

Tra

nsm

ittan

ce

0 –100

3–3

3–1

2–0 6–4

5–5

0–2

Figure 54 Ultraviolet absorption spectra of phthalan vapor near the electronic band origin.The ring-puckering transitions are labeled. Primes refer to the �F =1 state.

800

600

400

200

0

–0.2 –0.1

70.8

88.6

95.1

103.7

110.3

115.6

o

0.0

S0

X (Å)

V (

cm–1

)

S1

0.1 0.2

Figure 55 Ring-puckering potential energy curves for the phthalan S1(�,�]) excited state

compared to the S0 ground state.The flapping coordinate x2 is set equal to zero.

Vibrational Potential Energy Surface 107

Page 46: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

ground state assignments but also provided the data for determining the energymap for both S0 and S1(�,�] ) states. Figures 56 and 57 show the FES and SVLFspectra of COU, and Figure 58 shows the energy diagram. As was the case forthe electronic ground state, the ring-puckering data can be fit very nicely with aone-dimensional potential energy function for S1. In the excited state, however,the barrier drops to 34 cm�1 and the puckering dihedral angle drops to 14�.

1000 900 800

FES

UV absorption(a)

cm–1

700 600 500

O

4204401 1

4204404511 1 1

4204501 1

4404501 1

420451FES

400 300

cm–1

200 100 0500

UV absorption

1 2

4502

4511

000

4404511 0430451

1 0

4513

4404511 2

4204511 0

4204501 3

440450

290

1

1

3

(b)

Figure 56 Fluorescence excitation spectra of jet-cooled coumaran and the correspondingultraviolet absorption spectra at ambient temperatures.

108 Jaan Laane

Page 47: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

Figure 59 compares the two ring-puckering potential energy functions to eachother. The barrier in each case is due to the single CH2—CH2 torsionalinteraction. For the excited electronic state, the barrier is reduced because ofthe increased angle strain in the five-membered ring arising from the reduced �character of the benzene ring.

4520

4540

2510

2410 211

0

2610

2710

Excitation

Excitation

cm–1–800–600–400–200 –10000

UV absorption(a)

4502

000

4414510 2

4522

4502

4542

2810291

0

000

44145010

4511

4531

4551

29145110

27145110

21145110

29245110

–200 –400

cm–1

–600 –800 –10000

UV absorption(b)

Excitation4511

Figure 57 Single vibronic level fluorescence (SVLF) spectra of coumaran from the 000

band at 34,965.9 cm�1 and the 4502 and 451

1 bands which are 110.8 and 31.9 cm�1 higher,respectively.

Vibrational Potential Energy Surface 109

Page 48: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

500120.5

5 A"

3 A"

2 A'

1 A"0 A'

6 A'

4 A'

3' A'

1' A'

5' A'

4' A'' 5'' A'

6'' A''

4'' A''

3'' A'2'' A''

1'' A'

7'' A'

1' A'vp = 0' A''

vF = 0,v T = 0

ν45 (A") ν43 (A") ν44 (A") ν29 (A') ν42 (A")

vF = 1,v T = 0 vF = 0,v T = 1 vF = 0,v T = 1 vF = 0,v T = 0

vp = 0'' A''

vp = 0 A''vp = 0 A''

vp = 0' A''

3' *2' *

0' A"

3" A'

1 A"

1" A'

4 A''

3 A'

3 A''

1 A'0 A''

4 A'

3 A''3 A'2 A''

1 A'

2 A'

1 A''

0 A'

0" A"

2" A"

2' *

8 A'

9 A"

6 A'

4 A'

2 A'

7 A"

5 A"

3 A"

1 A"

95.5

75.7

35.6

180.9

91.7

35.2

364.9

102.9

91.1

75.2

35.2

245.9

75.4

36.1

122.3

110.1

101.0

92.1

75.8

35.0

106.7

101.477.8

180107

90.9

37.2

90.1

416.5389.2

76.2

71.4

36.1

92.6

191.8

68.639.4

90.9

251.4

95.0

87.9

76.7

70.1

36.7

91.7{ 3.1

{ 3.8

{ 3.7{ 3.4

{ 2.7

400

300S1

S0

E (

cm–1

)

200

100

0

500

600

400

300

200

100

0

Figure 58 Energy level diagram for the ring-puckering levels in excited states of other low-frequency modes of coumaran in its S0 and S1(�,�

]) states.

600

500

400

300

V (

cm–1

)

X (Å)

200

–0.2 –0.1

14° 25°

34 cm–1

154 cm–1

S0

S1(π,π*)

0.0 0.1 0.2

100

0

Figure 59 Comparison of the ring-puckering potential energy functions of coumaran in its S0and S1(�,�

]) states.

110 Jaan Laane

Page 49: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

5.3.1.3. 1,3-Benzodioxole13BZD is one of the most interesting molecules we have studied in our laboratory.Because the molecule has no CH2—CH2 interactions, which would be expected topucker the five-membered ring, it would seem likely that the molecule would beplanar. However, like 1,3-dioxole discussed earlier, 13BZD possesses the anomericeffect due to the —O—CH2—O— configuration, and this provides an impetus forthe five-membered ring to pucker. Duckett et al. [84] initially reported the far-infrared spectra of 13BZD but incorrectly assigned the spectra based on a planarstructure. Later, far-infrared and Raman spectra by Sakurai and co-workers [85],however, showed that the molecule has a barrier to planarity of 164 cm�1 andpuckering angles of +24�. A one-dimensional potential energy function was ableto fit the data moderately well, but all of the experimental ring-puckering and ring-flapping data were much better fit with a two-dimensional PES. The 164 cm�1

barrier provided evidence for the anomeric effect, but its magnitude was only abouthalf of what had been found for 1,3-dioxole. This indicated that the anomeric effectwas suppressed by the presence of the benzene ring.

The FES and UV absorption spectra of 13BZD were also investigated by Laaneand co-workers [86] to study the S1(�,�] ) excited state. Figure 60 shows the jet-cooled FES and UV absorption spectra for this molecule. This is a beautifulexample of how the two types of spectra complement each other extremely well.The FES shows only transitions from the vibrational ground state plus weakertransitions from the nearly degenerate v = 1 state, which is 9.6 cm�1 higher. TheUV spectrum, on the contrary, shows a very large number of bands arising fromthe many populated S0 vibrational levels at room temperature. Figure 61 shows the

500 400 300 200cm–1

FES

UV absorptiono

o

100 0

Figure 60 Fluorescence excitation spectrum (FES) and ultraviolet absorption spectrum of1,3-benzodioxole.The band origin is at 34,789.8 cm�1.

Vibrational Potential Energy Surface 111

Page 50: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

kind of very valuable detail available from the UV absorption spectra. Figure 62shows the ring-puckering and ring-flapping energy map generated from the spectraldata. As can be seen from the significant differences between the puckering levels inthe flapping vF = 0 and vF = 1 states, there is a large amount of interaction betweenthese two modes. This is not unexpected as they are both of low frequency andhave the same symmetry species (B2 for C2v symmetry). The S1(�,�] ) data werealso utilized for a two-dimensional PES calculation, and Figure 63 shows theresulting surface that has a barrier of 264 cm�1. To compare the inversion barriersfor the S0 and S1(�,�] ) states, Figure 64 shows the one-dimensional cut of the twoPESs along the puckering coordinate. As is obvious, the barrier for the S1(�,�] )state nearly doubles.

The experimental result that the 13BZD has a higher barrier and hence anincreased anomeric effect in the electronic excited state supports the view that thiseffect is suppressed in the S0 ground state. Apparently, as shown in Figure 65, thereis competition between the anomeric effect and the interaction of the oxygen non-bonded p orbital with the benzene ring. The anomeric effect also utilizes this porbital as it interacts with the ] orbital of the C—O bond involving the otheroxygen atom. In the S0 ground state, this p(O)-benzene interaction is higher andcompetes with the anomeric effect. However, in the S1(�,�] ) state, the � system ofthe benzene ring has been disturbed and thus the p(O)-benzene interaction isdiminished allowing the anomeric effect to dominate. The S1(�,�] ) barrier heightof 264 cm�1 is similar to that of 1,3-dioxole [48], where there is no suppression ofthis effect.

300 200

1–32–4

0–2

0–4

0–1'

2–1'

5–1"

4–3'

2–3'

1–2'

1–0'

0'–5

2'–2'

1'–1'

2–0

1–1

0–0

1–2

o

o

100

cm–1

% T

–1000

Figure 61 Electronic absorption spectra and assignments for 1,3-benzodioxole near the bandorigin.

112 Jaan Laane

Page 51: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

5.3.1.4. IndanThe far-infrared spectra of IND were first reported by Smithson et al. [87] in1984 and these are of high quality showing detail for both the ring-puckeringand ring-flapping levels. They analyzed the puckering spectra with a one-dimensional potential energy function and reported a barrier of 1900 cm�1 forthe S0 ground state. As this value appeared to be inconsistent with the 232 cm�1

barrier of CP [16], Arp and co-workers [64] reinvestigated the far-infrared workin 2002 using a two-dimensional PES. This PES had a barrier of 488 cm�1,minima at puckering angles of +30�, and reproduced the experimental data

6

5

4

3

2

1VP = 0

VP = 0

VF = 0

VF = 0

VF = 1

VF = 2

VF = 1

VF = 2

010.9

508.4

415.2

325.2

236.888.4

90.0

93.2

157.358.2

99.1

89.59.60

1

2

3

4

5

6

7

S0

S1

79.5

90.9

101.860.7

162.5

111.0

273.5

126.2

109.9

509.6 4' 2"

1"0"

504

101.4

402.6383.1

3'

2'

4'

3'

2'

1'0'

563.5 1"0"

472.6

383.7

297.8267.2

88.9

85.9

30.6

90.9

1'14.9

19.5

0' 189.5204.4

93.3

297.7

400.1

102.4383.4

Figure 62 Energy diagram for the ring-puckering (�P), ring-flapping (�F), and �37 vibrationsof 1,3-benzodioxole in its ground (S0) and excited (S1) electronic states. Single and doubleprimes on the ring-puckering (�P) quantum numbers indicate the �P =1 and 2 states,respectively.

Vibrational Potential Energy Surface 113

Page 52: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

1200

800

400

0

1200

800

400

0

X2 (Å) X 1

(Å)

Potential energy (cm

–1)

O

O

–0.1–0.50

0.5 0.10

Figure 63 Two-dimensional potential energy surface (cm�1) of 1,3-benzodioxole in itsS1(�,�

]) state: x1, ring-puckering; x2, ring-flapping.

600

109.9

111.0

60.7

90.9

26410.9}

126.2S0

O

O

S1

400

200

–0.2 –0.1 0.0

X1 (Å)

V (

cm–1

)

0.1 0.2

0

Figure 64 Comparison of the S0 and S1(�,�]) vibrational potential energy surfaces along the

ring-puckering coordinate (x1).The flapping coordinate x2 = 0.

114 Jaan Laane

Page 53: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

very well. In this case, the barrier arises from the two —CH2—CH2— torsionalinteractions, which tend to pucker the molecule so that the methylene groupsdo not eclipse each other.

Hollas et al. [88] reported the first UV absorption spectra of IND in 1977 andthen the LIF in 1991 [89]. They relied on the Smithson and co-workers [87]assignments and reported barriers of 1979 and 1800 cm�1 for the S0 and S1(�,�] )states, respectively. The Laane laboratory reinvestigated the jet-cooled LIF of INDand reported new data and a two-dimensional PES analysis in 2002 [64]. Figure 66shows the FES and the UV absorption spectra, and Figure 67 shows the vibroniclevels for the ring-puckering, ring-flapping, and ring-twisting states in S1(�,�] ).The two-dimensional PES analysis showed the barrier to be 441 cm�1, slightlyless than that in the ground state and also somewhat less than the ab initio valueof 528 cm�1.

Figure 65 Depiction of the interaction between the benzene � system and the oxygen atoms.Only one oxygen non-bonded p orbital is shown along with a ] (CçO) orbital with which theanomeric effect is achieved.

UV absorption

FES

37,400 37,200

cm–1

37,000 36,800

Figure 66 Laser-induced fluorescence spectrum (jet-cooled) and ultraviolet absorptionspectrum of indan (25�C).

Vibrational Potential Energy Surface 115

Page 54: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

5.3.1.5. SummaryThe investigation of these four members of the IND family in both their groundand excited electronic states has provided some valuable insights into these mole-cules. Some of the quantitative results for the experimental and calculated barriersto planarity are shown in Table 4. The PHT molecule has no CH2—CH2 torsionalinteractions, so angle strain is expected to keep the molecule planar in bothelectronic states. This is almost true, but the electronic ground state does in facthave a small 35 cm�1 barrier, and the ab initio calculation also predicts a smallbarrier. This may be because of a weak interaction involving the oxygen non-bonded orbital. The barrier disappears in the S1(�,�] ) state. COU does have asingle CH2—CH2 interaction, which is responsible for the puckering of the five-membered ring in both electronic states. In the S0 state, the barrier is 154 cm�1, amagnitude that is fairly typical of a single methylene�methylene interaction. For2,3-dihydrofuran, the barrier is 93 cm�1 [47]. In its S1(�,�] ) excited state, thebarrier almost disappears, and this reflects increased angle strain in the five-mem-bered ring. IND has two methylene–methylene torsional interactions, and, asexpected, it has higher and similar barriers to planarity in both S0 and S1(�,�] )states. There is little effect on the barrier as a result of the electronic transition. Themost interesting case is that of 13BZD, which possesses the anomeric effect becauseof its two oxygen atoms in a 1,3-configuration. As was the case for 1,3-dioxole[48], the five-membered ring puckers to increase the overlap between the non-bonded oxygen p orbital and the ] orbital of the adjacent C—O bond. In theelectronic ground state, this effect is suppressed as the benzene � system competesfor interaction with the oxygen p orbital. Hence, the barrier is only about half ofthat determined for 1,3-dioxole. When the benzene � system is perturbed by the�! �] excitation, however, its suppression of the anomeric effect is substantiallydiminished, and the barrier to planarity is almost doubled.

0

100

200

300

116

01

3

54

76

2

13710

1761'0'

2903'2'

3511"0"

401391

5'

4'

253 266270273 (ν49)325 (ν41)

32

0VF = 0 VF = 1 or 2 VT = 1

Otherlevels

230228

339344

Ene

rgy

(cm

–1)

400

2νT}

Figure 67 Energy level diagram for indan in its S1(�,�]) excited state.

116 Jaan Laane

Page 55: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

Table 4 Comparison of experimental and ab initio results for molecules in the indan family

Ground state S1(�,�]) excited state

Molecule Barrier Dihedral angles (cm�1)a Barrier Dihedral angles v00(cm�1)

Exp. Ab initio Exp. Ab initio Exp. Ab initio Exp. Ab initio Exp. Ab initio

Phthalan 35 91 0 23� 0 11 0 0 37,034 46,6351,3-Benzodioxole 164 171 24�, 3� 25�, 3� 264 369 24� 29� 34,790 36,986Coumaran 154 258 25� 27� 34 21 14� — 34,870 38,965Indan 488 662 30� 32� 441 528 39� — 36,904 —

a Puckering angle except for 1,3-benzodioxole for which the flapping angle is also given.

Vib

rationalPotentialEnergyS

urface117

Page 56: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

5.3.2. Other bicyclic aromatics5.3.2.1. DihydronaphthalenesThe ground and S1(�,�] ) excited states of both 1,2-dihydronaphthalene (12DHN)and 14DHN have been studied using LIF. Autrey and co-workers [90]

reported the LIF and UV absorption spectra along with ab initio calculationsof 12DHN in 2003. In addition to the determination of the PESs, vibrational

assignments were reported for both states showing the effect of the electronicexcitation. 12DHN is analogous to 1,3-cyclohexadiene [50] discussed earlier, butit possesses one extra out-of-plane ring mode, the ring-flapping, as shown inFigure 68. Figure 69 shows the FES and UV absorption spectra of 12DHN. Ascan be seen, the FES is very rich, but the UV shows much less detail than what isthe case for many of the other systems discussed in this work. Figure 70 shows theenergy diagram for both the ground and the excited electronic states. Vibrations51–54 are the out-of-plane ring modes while 35 is an in-plane ring-bendingmode. Numerous combinations between these vibrations were observed in theFES. Analysis of the data show the molecule to be twisted, and a one-dimensionalpotential energy function showed the barrier to inversion in the ground state to be1363+100 cm�1, with the uncertainty coming from the need to extrapolate thefunction above the observed experimental data. This agrees well with the triple-zeta ab initio value of 1524 cm�1. In the S1(�,�] ) state, the experimental datashow that the barrier increases substantially, probably to about 3000 cm�1, butan accurate value could not be determined. The increased barrier in the excitedstate reflects the decreased conjugation in the � system, and this leads to lowerangle strain.

++

ν19(A2)199 cm–1

131 cm–1 148 cm–1 374 cm–1 266 cm–1

ν36(B2)292 cm–1

ν18(A2)506 cm–1

Twist(out-of-phase)

Twist(in-phase)

FlapBend

+

++

+ –

+

–+

– –

+

+

+

+

++

++–

+

+

–+

––

+

+

––

+

Figure 68 Low-frequency out-of-plane vibrations of 1,3-cyclohexadiene and 1,2-dihydronaphthalene.

118 Jaan Laane

Page 57: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

14DHN can be classified as a ‘‘pseudo-four-membered ring.’’ Its LIF wasoriginally studied by Chakraborty and co-workers [91], but improved FES andUV data [92] in the Laane laboratories have led to a more reliable analysis. Figure 71shows its FES and UV absorption spectra, which were used to generate thequantum energy map for the low-frequency modes, including �54, the ring-puckering. In the S0 state, a regular sequence of energy separations beginning at33.8 cm�1 correspond to a very non-rigid ring system. In the S1(�,�] ) state, thesame type of sequence beginning at 77.4 cm�1 arises from a more rigid ring system.Figure 72 compares these one-dimensional ring-puckering potential energy func-tions to each other and also to 1,4-cyclohexadiene [17]. The stiffer potential energyfunction for the excited state again reflects increased angle strain [92].

800 600 400 200 0

FES

UV absorption(a)

000

cm–1

200

SVLF

(b)

0 –200 –400 –600 –800

ν00= 34,095 cm–1

UV absorption

cm–1

000

Figure 69 Fluorescence excitation spectrum (FES) and ultraviolet absorption spectrum of1,2-dihydronaphthalene (12DHN).

Vibrational Potential Energy Surface 119

Page 58: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

5.3.2.2. Tetralin and benzodioxansTetralin (TET), 1,4-benzodioxan (14BZD), and 1,3-benzodioxan (13BZN) areall analogous to the cyclohexene family of molecules and hence have twistedstructures with high barriers to planarity. Because of their low vapor pressures,they have not been studied by far-infrared spectroscopy, but their S0 vibrationaldata have been obtained using SVLF spectra of the jet-cooled molecules and

UV

FES

250 200 150 100

Wavenumber (cm–1)50 0

5302

2802

5402

5444

5433

5422 540

1

5411

000

Figure 71 Fluorescence excitation (bottom) and ultraviolet absorption (top) spectra of 1,4-dihydronaphthalene (14DHN).The wavenumbers are relative to the 00

0 band at 36,788.6 cm�1.

500

400318.7

472A' + 3B'

351A' + 2B'

229.1A' + B'

244.32B'

122.3B'

3663B'

487.54B' 440

2B' + C' 422A' + B' + C'

304.1A' + C'

412.82A' + C'

455.0B' + D'

292.6E'

318.7B' + C'

196.7C'

332D'3A'

216.62A'107.4A'

5104A

384.93A

259.22A

280.6A + B

4312A + B

5622A + 2B

296.82B

416B + C

5342C

267.0C

400A + C 347

D

385E

4493B

131.3A

A

B

B C D E

149.6

ν54 ν53 ν52 ν35 ν51

300

200

100

E(cm–1)0

400

300

200

100

0

500

Figure 70 Energydiagram for the low-frequency vibrations in their S0 and S1(�,�]) electronic

states of 1,2-dihydronaphthalene (12DHN).

120 Jaan Laane

Page 59: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

high-temperature vapor-phase Raman spectra. The FES data were also recorded.Figures 73 and 74 show the FES and SVLF spectra of 14BZD, respectively, andthis data were used to produce the quantum energy map of the low-frequencymodes shown in Figure 75 [93]. Because the inversion barrier is so high and theexperimental data extend to only about 700 cm�1 above the energy minima, anaccurate PES could not be reliably obtained. However, Figure 76 shows thecalculated relative energies of the twisted, bent, and planar conformations, and it

1500

1250

1000

750cm–1

500

250

0

–1.0 –0.8 –0.6 –0.4 –0.2 0

τ (radians)0.2 0.4 0.6

14DHN (S0)14DHN (S1)14CHD

0.8 1.0

Figure 72 Comparison of the 1,4-dihydronaphthalene (14DHN) S0 and S1 potential energyfunctions with the1,4-cyclohexadiene (14CHD) function.

UV absorption

00 = 35,563.1 cm–10

1401

2502

FES

800 600 400

cm–1200 0

4804 250

1 4804

4804

2502

2502 480

1

2501 480

2480

2

2501

000

2502 480

2

4801

1601

O

O

Figure 73 Fluorescence excitation spectra of jet-cooled 1,4-benzodioxan and ultravioletabsorption spectra at ambient temperature in the 0^900 cm�1 region. The band origin 00

0 isat 35,563.1cm�1.

Vibrational Potential Energy Surface 121

Page 60: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

also shows the direction and frequency values for the bending and twisting modes.A two-dimensional calculation produced a PES with a barrier of 3906 cm�1 inreasonable agreement with the ab initio value of 4095 cm�1. For the S1 excitedstate, the bending and twisting frequencies drop considerably to 79.8 and139.6 cm�1 reflecting a lower barrier to planarity. The two-dimensional PEScomputed from the data yields a barrier of 1744 cm�1.

The FES and SVLF data for 13BZN have also been recorded and analyzed [94].This molecule undergoes the anomeric effect due to the presence of the O—CH2—Oconfiguration. 13BZN is also twisted with a high barrier to planarity. Its bending andtwisting frequencies are 108.4 and 158.4 cm�1 for the ground state and drop to 96.3and 101.7 cm�1 in S1, again reflecting a decrease in barrier in the excited state.

The complete vibrational spectra of TET have been assigned for the S0 state, and 17of its 60 vibrations were determined for the S1(�,�] ) state including the out-of-planering modes [95]. Figure 77 shows its FES spectrum along with some of the assignments.As was done for the other molecules, a quantum energy map was generated andutilized to analyze the ring-bending and ring-twisting modes in particular. In the S0

ground state, the bending and twisting frequencies are 94.3 and 141.7 cm�1, respec-tively, and these decrease to 85.1 and 94.5 in the excited state. The barriers to planarityare close to 5000 cm�1 but cannot be determined with great accuracy.

5.3.2.3. 2-Indanol2-Indanol is a fascinating molecule in that it can exist in four different conforma-tions, one of which has intramolecular hydrogen bonding between the OH groupand the benzene ring. Both LIF and detailed ab initio calculations were carried outfor this molecule [96]. The four structures are shown in Figure 78. Conformation Apossesses the intramolecular hydrogen bond and is calculated to be about 450 cm�1

lower in energy than the other three conformations, all of which have similar

UV absorption

000

000

1410

1210

2520

2510 253

010×

1510

4610 481

0

10×

SVLFExcitation

0 –200 –400 –600 –800 –1000

cm–1

1410 251

0

O

O

Figure 74 Singlevibroniclevel fluorescencespectraofjet-cooled1,4-benzodioxanwithexcitationof the00

0bandat35,563.1cm�1andtheultravioletabsorptionspectraatambienttemperature.

122 Jaan Laane

Page 61: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

energies. The planar ring structure is calculated to be 1308 cm�1 higher in energy.The four conformers can interconvert into one another either through the ring-puckering or the OH internal rotation vibrations. These motions encounter bar-riers from about 400 to 1700 cm�1. Figure 79 shows the computed theoretical PESfor this molecule in terms of the ring-puckering angle and the OH internal rotationangle. The lowest energy form with the hydrogen bond corresponds to a puckeringangle of 35.4� and an internal rotation angle of 180�. The other forms are at anglesof –35.1� and 180� (b), 31.1� and 66.9� (c), and –35.0� and 54.5� (d). Figure 80shows the jet-cooled FES spectrum of 2-indanol, which strongly supports thetheoretical calculations. Bands from each of the four conformers are present withthe 00

0 band of A, the hydrogen bonded structure, by far the strongest. The bandintensities indicate a distribution of 82% A, 3% B, 13% C, and 9% D, and thesevalues are similar to those computed from the relative energies of the conformers.As is evident in Figure 80, the ring-puckering (�29) bands can also be seen for eachconformer.

500

600

400

79.8A'

613.26A

512.95A

411.94A

310.23A

207.72A

104.3A

ν48(A) ν25(B) ν47(C) ν24(D)

165.6B

159.32A'

238.63A'

218.5A'+ B'

574.24A + B 538

2A + 2B

697A + 2C

(434.6)A + 2B

297.1C

400.0A + C

594.12C

318.5D

472.93A + B

371.12A + B

268.7A + B

657.74B

494.33B

330.32B

2972A' + B'

434.22A' + 2B'

357.8A' + 2B'

3952A' + 2C'

237.22C'

4554A' + B'

475.96A'

396.95A'

416.13B'

278.32B'

139.6B'

306.5D'

553.04B'

317.84A'300S1

S0

200

100

0

400

300

200

100

0

500

600

700

Figure 75 Energy level diagram for the low-frequencymodes of 1,4-benzodioxan in its S0 andS1(�,�

]) states.A is the ring-bending,B andD are ring-twistingmodes, andC is ring-flapping.

Vibrational Potential Energy Surface 123

Page 62: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

5.4. Cavity ringdown spectroscopy of enones

In a collaboration between the Drucker and Laane laboratories, the cavity ringdownspectra (CRDS) of 2-cyclopenenone (2CP) and 2CH have been recorded andanalyzed. The CRDS system in the Drucker laboratory is capable of very high

Twist

0 cm–1

2722 cm–1 4095 cm–1

ν T

= 166 cm–1

ν B = 104 cm–1

B B BendP

T

T

Figure 76 Representation of the two-dimensional potential energy for the ring-twisting andring-bending vibrations of 1,4-benzodioxan in the S0 state. P, planar; T, twisted; B, bentstructure.

UV absorption

00 = 36,789.3 cm–10

000

1701

1701 310

1 1901 310

1

2801

4×180

1 1901

2901 300

1

3103 310

2

3101

FES

800 600 400 200 0

cm–1

Figure 77 Fluorescence excitation spectra of jet-cooled tetralin and the ultravioletabsorption spectra at ambient temperature in the 0^900 cm�1 region.

124 Jaan Laane

Page 63: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

sensitivity, and its details have been described elsewhere [97]. The high sensitivity,which arises from thousands of reflections between two highly reflecting mirrors, hasallowed very weak absorption signals to be recorded in the UV region for these twomolecules. Both 2CP and 2CH are conjugated possessing adjoining CTC andCTO double bonds. As already discussed, 2-CPO is planar due to this conjugation

(a)

(c)

(b)

(d)

Figure 78 Four stable conformations of 2-indanol. The intramolecular hydrogen bondingpresent in the most stable conformer (structureA) is represented by a dotted line.

2500

2000

1000

500

060 40 20 –20 –40 –60

180120 OH internal

rotation angle

600

0Puckering angle

V (

cm–1

)

Figure 79 Calculated potential energy surface of 2-indanol in terms of its ring-puckeringangle (degrees) on OH internal rotation (degrees relative to 180� at theA conformation). (Seecolor plate 8).

Vibrational Potential Energy Surface 125

Page 64: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

in the electronic ground state. The lowest frequency electronic transition for thesemolecules is the n!� transition, and this results in weaker conjugation in theelectronic excited state. As already shown in Table 1 for 2CP, the CTO andCTC frequencies are decreased in the S1(n,�] ) state. However, the single bondbetween these two double bonds is strengthened, and one of the other ring-stretch-ing modes increases in frequency. Figure 81 presents the survey CRDS spectrum of2CP showing the absorption by both transitions to the singlet S1(n,�] ) and the tripletT1(n,�] ) states [98]. Transitions to triplet states are spin forbidden, so observation ofsuch transitions is quite unusual. The S1 S0 transitions confirm those previouslydiscussed in the FES spectrum [66], but the T1 S0 had not been previously seen.The weaker absorption bands near the triplet band origin arise from transitionsbetween the ring-puckering (�30) levels of the S0 and T1 states. Namely,301

0; 3020; 303

0; 3034; 301

1; 3021; 303

1; 3022; 303

3 and many other transitions wereobserved. Figure 82 shows the energy level spacings for three isotopomers of 2CPfor both the singlet ground state and the triplet excited state. Figure 83 shows theone-dimensional potential energy function with the form of Equation (2) for thetriplet state. This is the first determination of a ring-puckering potential function for atriplet state. The function has a barrier of 42 cm�1, which is totally produced by theflipped electron spin as the S0 and S1 states are both planar. Figure 84 compares theS0, S1, and T1 ring-puckering functions. In S0, 2CP is planar and rigid, in the S1 stateit becomes floppy because of the loss of conjugation, and in the triplet state themolecule remains floppy, but also takes on a non-planar structure. This structuralchange and the presence of the barrier produced by the spin flip is highly unusual.

A294

300 200 100

cm–10

0

A2710

D5310

C5310

A293

OH

0 C2710

A2810

A2920

D2920

A2910

C2920

C2810

D2910

B2910

C2910

D000

B000 C00

0

A000

Figure 80 Laser-induced fluorescence excitation spectrum of 2-indanol.

126 Jaan Laane

Page 65: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

26,8000

2000

4000

6000

8000

10,000

26,600 26,400

Frequency (cm–1)

Wavelength (nm)

ε (M

–1 c

m–1

)

Fra

ctio

nal l

oss

per

pass

(pp

m)

26,200 26,000 25,800

374 376 378

191(S1 ← S0)0

00 (T1 ← S0)0

380 382 384 386

0.0

0.2

0.4

0.6

0.8

1.0

1.2

Figure 81 Room-temperatureCRDspectrumof 2-cyclopentenone (2CP) vapor showing theT1 S0 origin band near 385 nm, as well as the onset of S1 S0 transitions near 375 nm.

94

2CP – d0

S0

T1

Ene

rgy

(cm

–1)

0

100

200

300

0

100

200

300

400

500

2CP – d1 2CP – d2

100

104

89

94

98

25,95925,964

37119

214

325

446

199

10933 30102

185

281

386

25,956

85

90

93

Figure 82 Energy-level diagram, determined from the present vibronic assignments (tripletstate) and previous far-infrared results (ground state), for the ring-bending vibrational modeof 2CP-d0, 2CP-d1, and 2CP-d2.

Vibrational Potential Energy Surface 127

Page 66: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

To more fully understand the changes in 2CP, theoretical DFT and ab initiocalculations were carried out for the molecule in its S0, S1(n,�] ), T1(n,�] ), andT2(�,�] ) states [99]. These supported the experimental results and confirmed that theT1 state arises from the n! �] transition. In addition, the calculations provided datafor many of the other vibronic levels. The ab initio calculations predicted a barrier of8 cm�1 for the T1(n,�] ) state compared to the experimental value of 43 cm�1. Thebarrier for the T2(�,�] ) state was calculated to be much higher at 999 cm�1.

The CRDS of 2CHO have also been analyzed [100,101] and the potential energyfunctions for the inversion vibration have been determined for both the ground andS1(n,�] ) excited states. Earlier Raman [102] and infrared work [103] were used topostulate potential energy functions with barriers of 935 and 3379 cm�1, respectively,but both of these proved to be erroneous. Figure 85 shows the CRDS of 2CHOwhich has the S1(n,�] ) band origin at 26,081.3 cm�1. As can be seen, bands in both the

–0.1

0

200

400

600

0.0 0.1

Puckering coordinate (Å)

Ene

rgy

(cm

–1) S1

S0

T1

Figure 84 Comparison of experimentally determined ring-bending potential energyfunctions for the S0,T1(n,�

]), and S1(n,�]) states of 2-cyclopentenone (2CP).

–0.1

37 360

1

2

3

4

5

446

324

214

120119

214

325

446

Observed Calculated

0

200

400

600

Ene

rgy

(cm

–1)

0.0 0.1

Puckering coordinate (Å)

Figure 83 Ring-bending potential-energy function for theT1(n,�]) state of 2-cyclopentenone

(2CP).

128 Jaan Laane

Page 67: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

390n and 39m

0 series were readily observed where n is the S0 state quantum number forthe inversion vibration (�39) and m is that for the S1(n,�] ) state. The 390

n series togetherwith the far-infrared spectrum [103] allowed the inversion and twisting quantum statesfor the S0 electronic ground state to be correctly assigned, whereas the 39m

0 seriesprovided the data for the S1(n,�] ) state. A one-dimensional potential function for theground state determined the barrier to be 1900+ 300 cm�1, in good agreement withthe DFT value of 2090 cm�1 [100]. The excited state data extended to about 800 cm�1

above the energy minima and were used to determine a one-dimensional potentialfunction with a barrier of 3550+ 500 cm�1 for the S1 state. Ab initio calculationspredicted a value of only 2265 cm�1. The increased barrier height for the electronicexcited state results from the loss of conjugation and decreased angle strain allowing themolecule to distort more from a planar structure.

For both 2CPO and 2CHO the absorption spectra are very weak and have notbeen observed by conventional methods. These two studies demonstrate the powerof CRDS to get these results. Direct observation of the triplet state spectra for2CPO was especially gratifying.

6. SUMMARY AND CONCLUSIONS

As demonstrated by the many studies discussed here, one- and two-dimensional PESs can be most invaluable for determining the structures of moleculesin both their ground and excited electronic states. They also help to elucidate theforces responsible for the structures and also provide quantitative detail on the energy

0

2

4

6

8

10

12

300

0.1 Torr

0.5 Torr3701

3601

3801

3901

000 391

0 3920 393

0(?)

3902390

3

400 200

Frequency shift (cm–1)

Wavelength (nm)

Fra

ctio

nal p

hoto

n lo

ss(p

arts

per

thou

sand

)

100 0 –100 –200 –300 –400

380378 382 384 386 388

Figure 85 Cavity ringdown spectra of 2CHO relative to the S1 000 band at 26,089.1 cm�1 in

the difference band region.

Vibrational Potential Energy Surface 129

Page 68: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

differences between various molecular structures. The results for electronic excitedstates, including triplet states, are particularly valuable in that photochemical pro-cesses proceed through such states after the molecules have been structurally distortedbecause of electronic transitions. It should also be emphasized that these investiga-tions validate the use of the Schrodinger equation along with the approximation thatthe vibrations of interest can be studied independently from all the other vibrations inthese molecules. This is based on the fact that the low-frequency vibrations do notinteract with the higher frequency modes or with those of different symmetry. Whiletables comparing observed and calculated frequencies for the PESs, often based onEquation (2), have not been shown here, it should be noted that in virtually all ofthese cases the agreement is fantastic. For SCB [15], for example, 34 frequencies arefit with the two-parameter potential function of Equation (2) with an averageaccuracy of better than 1 cm�1. Examples such as this should be added to physicalchemistry texts, which typically only discuss the rather boring harmonic oscillator.

ACKNOWLEDGMENTS

The author expresses his sincere gratitude to the more than 40 students, post-docs, and visiting faculty members who have contributed to this work. He alsothanks the National Science Foundation and the Robert A. Welch Foundation formany years of financial support. He is also grateful to Ms. Linda Redd for assistancewith the preparation of this manuscript.

REFERENCES

[1] D. M. Dennison and G. E. Uhlenbeck, Phys. Rev., 41: 313–321, 1932.[2] D. M. Dennison, Rev. Mod. Phys., 12: 175–321, 1940.[3] L. Coudert, A. Valentin, and L. Henry, J. Mol. Spectrosc., 120: 185–204, 1986.[4] V. Spirko and W. P. Kraemer, J. Mol. Spectrosc., 133: 331–344, 1989.[5] D. J. Rush and K. B. Wiberg, J. Phys. Chem. A, 101: 3143–3151, 1997.[6] J. Laane, Appl. Spectrosc., 24: 73–80, 1970.[7] R. P. Bell, Proc. R. Soc., A183: 328–337, 1945.[8] A. Danti and R. C. Lord, J. Chem. Phys., 30: 1310–1313, 1959.[9] A. Danti, W. J. Lafferty, and R. C. Lord, J. Chem. Phys., 33: 294–295, 1960.

[10] S. I. Chan, J. Zinn, and W. D. Gwinn, J. Chem. Phys., 33: 295–296, 1960.[11] S. I. Chan, J. Zinn, and W. D. Gwinn, J. Chem. Phys., 33: 1643–1655, 1960.[12] S. I. Chan, J. Zinn, and W. D. Gwinn, J. Chem. Phys., 34: 1319–1329, 1961.[13] S. I. Chan, T. R. Borgers, J. W. Russell, H. L. Strauss, and W. D. Gwinn, J. Chem. Phys., 44:

1103–1111, 1966.[14] T. R. Borgers and H. L. Strauss, J. Chem. Phys., 45: 947–955, 1966.[15] J. Laane and R. C. Lord, J. Chem. Phys., 48: 1508–1513, 1968.[16] J. Laane and R. C. Lord, J. Chem. Phys., 47: 4941–4945, 1967.[17] J. Laane and R. C. Lord, J. Mol. Spectrosc., 39: 340–344, 1971.[18] L. A. Carreira, R. C. Lord, and T. B. Malloy, Top. Curr. Chem., 82: 1–95, 1979.[19] J. E. Kilpatrick, K. S. Pitzer, and R. Spitzer, J. Am. Chem. Soc., 69: 2483–2488, 1947.[20] K. S. Pitzer and W. E. Donath, J. Am. Chem. Soc., 81: 3213–3218, 1959.[21] J. R. Durig and D. W. Wertz, J. Chem. Phys., 49: 2118–2121, 1968.

130 Jaan Laane

Page 69: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

[22] L. E. Bauman and J. Laane, J. Phys. Chem., 92: 1040–1051, 1988.[23] J. Laane, Vib. Spectra Struct., 1: 25–50, 1978.[24] H. Strauss, Ann. Rev. Phys. Chem., 34: 301–328, 1983.[25] W. Kiefer, H. J. Bernstein, M. Danyluk, and H. Wieser, Chem. Phys. Lett., 12: 605–609,

1972.[26] T. H. Chao and J. Laane, Chem. Phys. Lett., 14: 595–597, 1972.[27] F. A. Miller and R. J. Capwell, Spectrochim. Acta, 28A: 603–618, 1972.[28] Internal Rotation and Inversion(Lister, MacDonald, and Owen, eds), Academic Press, London.[29] P. Groner, J. F. Sullivan, and J. R. Durig, Vib. Spectra Struct., 9: 405–496, 1981.[30] J. D. Lewis, T. B. Malloy, Jr., T. H. Chao, and J. Laane, J. Mol. Struct., 12: 427–449, 1972.[31] J. Laane, J. Phys. Chem., 95: 9246–9249, 1991.[32] T. Klots, S. N. Lee, and J. Laane, J. Phys. Chem., 103: 833–837, 1999.[33] J. Laane, J. Chem. Phys., 50: 776–782, 1969.[34] R. M. Irwin, J. M. Cooke, and J. Laane, J. Am. Chem. Soc., 99: 3273–3278, 1977.[35] T. B. Malloy, J. Mol. Spectrosc., 44: 504–535, 1972.[36] J. Laane, M. A. Harthcock, P. M. Killough, L. E. Bauman, and J. M. Cooke, J. Mol. Spectrosc.,

91: 286–299, 1982.[37] M. A. Harthcock and J. Laane, J. Mol. Spectrosc., 91: 300–324, 1982.[38] R. W. Schmude, M. A. Harthcock, M. B. Kelly, and J. Laane, J. Mol. Spectrosc., 124:

369–378, 1987.[39] M. M. Tecklenburg and J. Laane, J. Mol. Spectrosc., 137: 65–81, 1989.[40] M. M. Strube and J. Laane, J. Mol. Spectrosc., 129: 126–139, 1988.[41] J. Yang and J. Laane, J. Mol. Struct., 798: 27–33, 2006.[42] S. Sakurai, N. Meinander, and J. Laane, J. Chem. Phys., 108: 3537–3542, 1998.[43] C. M. Cheatham, M. Huang, N. Meinander, M. B. Kelly, K. Haller, W.-Y. Chiang, and J.

Laane, J. Mol. Struct., 377: 81–92, 1996.[44] C. M. Cheatham, M.-H. Huang, and J. Laane, J. Mol. Struct., 377: 93–99, 1996.[45] K. Morris, Instrumentation for the Study of Jet-Cooled Molecules and Spectroscopic

Investigation of Low-Frequency Vibrations, Ph.D. Thesis, Texas A&M University, 1998.[46] T. M. Hard and R. C. Lord, Appl. Opt., 7: 589–598, 1968.[47] J. Laane, K. Haller, S. Sakurai, K. Morris, D. Autrey, Z. Arp, W.-Y. Chiang, and A. Combs, J.

Mol. Struct., 650: 57–68, 2003.[48] D. Autrey and J. Laane, J. Phys. Chem., 105: 6894–6899, 2001.[49] E. Cortez, R. Verastegui, J. R. Villarreal, and J. Laane, J. Am. Chem. Soc., 115: 12132–12136,

1993.[50] D. Autrey, J. Choo, and J. Laane, J. Phys. Chem., 105: 10230–10236, 2001.[51] T. H. Chao and J. Laane, J. Mol. Spectrosc., 70: 357–360, 1978.[52] L. E. Bauman and J. Laane, J. Phys. Chem., 92: 1040–1051, 1988.[53] L. A. Carreira, I. M. Mills, W. B. Person, J. Chem. Phys., 56: 1444–1448, 1972.[54] S. J. Leibowitz, J. R. Villarreal, and J. Laane, J. Chem. Phys., 96: 7298–7305, 1992.[55] S. J. Leibowitz and J. Laane, J. Chem. Phys., 101: 2740–2745, 1994.[56] V. E. Gaines, S. J. Leibowitz, and J. Laane, J. Am. Chem. Soc., 113: 9735–9742, 1991.[57] J. Laane and J. Choo, J. Am. Chem. Soc., 116: 3889–3891, 1994.[58] M. M. Tecklenburg and J. Laane, J. Am. Chem. Soc., 111: 6920–6926, 1989.[59] M. M. Tecklenburg, J. Villarreal, and J. Laane, J. Chem. Phys., 91: 2771–2775, 1989.[60] T. Klots, S. Sakurai, and J. Laane, J. Chem. Phys., 108: 3531–3536, 1998.[61] T. Klots, E. Bondoc, and J. Laane, J. Phys. Chem., 104: 275–279, 2000.[62] J. Yang, K. Okuyama, K. Morris, Z. Arp, and J. Laane, J. Phys. Chem. A, 109: 8290–8292,

2005.[63] S. Sakurai, N. Meinander, K. Morris, and J. Laane, J. Am. Chem. Soc., 121: 5056–5062, 1999.[64] Z. Arp, N. Meinander, J. Choo, and J. Laane, J. Chem. Phys., 116: 6648–6655, 2002.[65] J. C. D. Brand, J. Chem. Soc. 858–872, 1956..[66] C. M. Cheatham and J. Laane, J. Chem. Phys., 94: 7734–7743, 1991.[67] P. Sagear and J. Laane, J. Chem. Phys., 102: 7789–7797, 1995.[68] J. Zhang, W.-Y. Chiang, and J. Laane, J. Chem. Phys., 98: 6129–6137, 1993.

Vibrational Potential Energy Surface 131

Page 70: Frontiers of Molecular Spectroscopy || Vibrational Potential Energy Surfaces in Electronic Excited States

[69] J. Choo and J. Laane, J. Chem. Phys., 101: 2772–2778, 1994.[70] J. Zhang, W.-Y. Chiang, and J. Laane, J. Chem. Phys., 100: 3455–3462, 1994.[71] W.-Y. Chiang and J. Laane, J. Phys. Chem., 99: 11640–11643, 1995.[72] P. Sagear, S. N. Lee, and J. Laane, J. Chem. Phys., 106: 3876–3883, 1997.[73] C. M. Cheatham and J. Laane, J. Chem. Phys., 94: 5394–5401, 1991.[74] D. H. Waldeck, Chem. Rev., 91: 415–436, and references therein, 1991.[75] W.-Y. Chiang and J. Laane, J. Chem. Phys., 101: 8755–8767, and references therein, 1994.[76] L. Banares, A. A. Heikal, and A. H. Zewail, J. Chem. Phys., 96: 4127–4130, 1992.[77] K. Haller, W.-Y. Chiang, A. del Rosario, and J. Laane, J. Mol. Struct., 379: 19–23, 1996.[78] W.-Y. Chiang and J. Laane, J. Phys. Chem., 99: 11823–11829, 1995.[79] Z. Arp, J. Laane, A. Sakamoto, and M. Tasumi, J. Phys. Chem., 106: 3479–3484, 2002.[80] E. Bondoc, S. Sakurai, K. Morris, W.-Y. Chiang, and J. Laane, J. Chem. Phys., 112:

6700–6706, 2000.[81] P. Ottavani and W. Caminati, Chem. Phys. Lett., 405: 68–72, 2005.[82] M. J. Watkins, D. E. Belcher, and M. C. R. Cockett, J. Chem. Phys., 116: 7855–7867, 2002.[83] J. Yang, M. Wagner, K. Okuyama, K. Morris, Z. Arp, J. Choo, N. Meinander, and J. Laane, J.

Chem. Phys., 125: 034308/1–034308/9, 2006.[84] J. A. Duckett, T. L. Smithson, and H. Wieser, Chem. Phys. Lett., 64: 261–265, 1979.[85] S. Sakurai, N. Meinander, K. Morris, and J. Laane, J. Am. Chem. Soc., 121: 5056–5062, 1999.[86] S. Sakurai, E. Bondoc, J. Laane, K. Morris, N. Meinander, and J. Choo, J. Am. Chem. Soc.,

122: 2628–2634, 2000.[87] T. L. Smithson, J. A. Duckett, and H. Wieser, J. Phys. Chem., 88: 1102–1109, 1984.[88] J. M. Hollas, J. Mol. Spectrosc., 66: 452–464, 1977.[89] K. M. Hassan and J. M. Hollas, J. Mol. Spectrosc., 147: 100–113, 1991.[90] D. Autrey, Z. Arp, J. Choo, and J. Laane, J. Chem. Phys., 119: 2557–2568, 2003.[91] T. Chakraborty, J. E. Del Bene, and E. C. Lim, J. Chem. Phys., 98: 8–13, 1993.[92] M. Z. M. Rishard, M. Wagner, J. Yang, and J. Laane, Chem. Phys. Lett., 442: 182–186, 2007.[93] J. Yang, M. Wagner, and J. Laane, J. Phys. Chem. A, 110: 9805–9815, 2006.[94] K. McCann, M. Wagner, and J. Laane, to be published.[95] J. Yang, M. Wagner, and J. Laane, J. Phys. Chem. A, 111: 8429–8438, 2007.[96] A. A. Al-Saadi, M. Wagner, and J. Laane, J. Phys. Chem. A., 110: 12292–12297, 2006.[97] S. Drucker, J. L. Van Zanten, N. D. Gagnon, E. J. Gilles, and N. R. Pillsbury, J. Mol. Struct.,

692: 1–16, 2004.[98] N. Pillsbury, J. Choo, J. Laane, and S. Drucker, J. Phys. Chem. A, 107: 10648–10654, 2003.[99] J. Choo, S. Kim, S. Drucker, and J. Laane, J. Phys. Chem. A, 107: 10655–10659, 2003.

[100] E. J. Gilles, J. Choo, D. Autrey, M. Rishard, S. Drucker, and J. Laane, Can. J. Chem., 82:867–872, 2004.

[101] M. Z. M. Rishard, E. A. Brown, L. A. Ausman, S. Drucker, J. Choo, and J. Laane, J. Phys.Chem. A, 112, 38–44, (2007).

[102] L. A. Carreira, T. G. Towns, and T. B. Malloy, Jr., J. Chem. Phys., 70: 2273–2275, 1979.[103] T. L. Smithson and H. Wieser, J. Chem. Phys., 72: 2340–2346, 1980.

132 Jaan Laane