• Aucun résultat trouvé

Modelling IR Spectral Dynamics upon Symmetry Breaking of a Photo-Excited Quadrupolar Dye

N/A
N/A
Protected

Academic year: 2022

Partager "Modelling IR Spectral Dynamics upon Symmetry Breaking of a Photo-Excited Quadrupolar Dye"

Copied!
43
0
0

Texte intégral

(1)

Article

Reference

Modeling Infrared Spectral Dynamics upon Symmetry Breaking of a Photo-Excited Quadrupolar Dye

NAZAROV, Alexey E., IVANOV, Anatoly I., VAUTHEY, Eric

Abstract

A significant number of quadrupolar dyes with a D-$pi$-A-$pi$-D or A-$pi$-D-$pi$-A structure, where D and A are electron donor and acceptor groups, were shown to undergo symmetry breaking (SB) upon optical excitation. During this process, the electronic excitation, originally distributed evenly over the molecule, concentrates on one D-$pi$-A branch, and the molecule becomes dipolar. This process can be monitored by time resolved infra-red (TRIR) spectroscopy and causes significant spectral dynamics. A theoretical model of excited-state SB developed earlier ( extit{J. Phys. Chem. C}, extbf{2018}, extit{132}, 29165) is extended to account for the temporal changes taking place in the IR spectrum upon SB. This model can reproduce the IR spectral dynamics observed in the $-mathrm{C}equiv mathrm{C}-$

stretching region with a D-$pi$-A-$pi$-D dye in two polar solvents using a single set of molecular parameters. This approach allows estimating the degree of asymmetry of the excited state in different solvents as well as its change during SB. Additionally, the relative contribution of the different mechanisms responsible for the [...]

NAZAROV, Alexey E., IVANOV, Anatoly I., VAUTHEY, Eric. Modeling Infrared Spectral

Dynamics upon Symmetry Breaking of a Photo-Excited Quadrupolar Dye. Journal of Physical Chemistry. C, 2020, vol. 124, no. 4, p. 2357-2369

DOI : 10.1021/acs.jpcc.9b10565

Available at:

http://archive-ouverte.unige.ch/unige:132980

Disclaimer: layout of this document may differ from the published version.

(2)

Modelling IR Spectral Dynamics upon Symmetry Breaking of a Photo-Excited Quadrupolar Dye

Alexey E. Nazarov,

Anatoly I. Ivanov,

∗,†

and Eric Vauthey

∗,‡

†Volgograd State University, University Avenue 100, Volgograd 400062, Russia

‡Department of Physical Chemistry, University of Geneva, 30 Quai Ernest-Ansermet, CH-1211 Geneva 4, Switzerland

E-mail: Anatoly.Ivanov@volsu.ru; Eric.Vauthey@unige.ch

(3)

Abstract

A significant number of quadrupolar dyes with a D-π-A-π-D or A-π-D-π-A struc- ture, where D and A are electron donor and acceptor groups, were shown to undergo symmetry breaking (SB) upon optical excitation. During this process, the electronic excitation, originally distributed evenly over the molecule, concentrates on one D-π-A branch, and the molecule becomes dipolar. This process can be monitored by time resolved infra-red (TRIR) spectroscopy and causes significant spectral dynamics. A theoretical model of excited-state SB developed earlier ( Ivanov, A. I.J. Phys. Chem.

C, 2018, 122, 29165-29172) is extended to account for the temporal changes taking place in the IR spectrum upon SB. This model can reproduce the IR spectral dynamics observed in the−C≡C−stretching region with a D-π-A-π-D dye in two polar solvents using a single set of molecular parameters. This approach allows estimating the degree of asymmetry of the excited state in different solvents as well as its change during SB.

Additionally, the relative contribution of the different mechanisms responsible for the splitting of the symmetric and antisymmetric −C≡ C− stretching bands, which are both IR active upon SB, can be determined.

Introduction

1

Over the past few years, there has been an increasing interest for symmetric molecular

2

systems, for which photoexcitation triggers a charge transfer process along one of several

3

energetically-equivalent pathways, causing a breaking of the symmetry.1 These systems in-

4

clude molecules containing two or more identical chromophores where the excited subunit can

5

act either as electron donor (D) or acceptor (A).2–13 Such symmetry-breaking (SB) charge

6

separation process could be advantageously exploited for applications in photovoltaics and ar-

7

tificial photosynthesis.9,14,15Another class of compounds comprises multibranched molecules

8

with a D(-π-A)n or A(-π-D)n (n = 2,3) motif, that are attracting considerable interest for

9

(4)

their promising two-photon absorption properties.16–23Here SB manifests itself by a large flu-

10

orescence solvatochromism, indicative of a dipolar excited state, despite a symmetric ground

11

state, as evidenced by a negligible absorption solvatochromism and by the relative intensity

12

of the S1 ←S0 and S2 ←S0 bands in the one- and two-photon absorption spectra.19,24–38

13

Such excited-state symmetry breaking (ES-SB) involves a transition from a quadrupolar or

14

octupolar Franck-Condon excited state to a dipolar relaxed excited state, and, thus corre-

15

sponds to an increase of the excitation density on one branch of the molecule at the expense

16

of the other(s).1,39–42

17

This ES-SB process could be visualised in real time in quadrupolar molecules using

18

femtosecond time-resolved infrared (TRIR) spectroscopy. This was achieved by monitoring

19

either −C≡N vibrations located on the acceptor ends of A-π-D-π-A molecules consisting of

20

a pyrrolopyrrole D core and two cyanophenyl acceptors,43–45 or−C≡C−vibrations in the

21

π linkers of D-π-A-π-D molecules.46,47 Upon ES-SB, the TRIR spectrum transforms from a

22

spectrum with a single −C≡N or −C ≡ C− stretching band, as expected for a symmetric

23

excited state, to a spectrum with two bands, pointing to an asymmetric excited state. In the

24

case of the A-π-D-π-A dyes, the splitting of the−C≡N bands was shown to increase with the

25

extent of SB. This process was found to be very sensitive to the solute-solvent interactions.

26

It does not take place in apolar solvents, but occurs not only in dipolar solvents, but also in

27

quadrupolar as well as in halogen- and hydrogen-bonding solvents.45 Moreover, its timescale

28

is similar to that of solvent relaxation.

29

Theoretically, ES-SB could be successfully described in terms of an essential state model.39–42,48,49

30

It was assumed to be induced by the interaction of the solute molecule with the solvent po-

31

larization and by electronic interactions with antisymmetric vibrational modes. This model

32

could reproduce the effects of ES-SB on one- and two-photon absorption spectra in quadrupo-

33

lar and octupolar donor-acceptor chromophores.39–42,48,49 More recently, a two-level model

34

describing the effect of SB on the vibrational spectrum of a quadrupolar dye with an exact

35

solution has been derived.50 This model was able to successfully reproduce the IR absorp-

36

(5)

tion spectrum in the −C≡N stretching region of a A-π-D-π-A dye after ES-SB in dipolar

37

solvents of varying dielectric constant. The two-level model was recently shown to predict

38

results very similar to those obtained within the essential state model when the S1-S0 gap

39

is significantly larger than the S2-S1 energy splitting.51 The −C≡N stretching vibrations

40

were described in terms of two normal modes, an IR active antisymmetric mode and an IR

41

inactive symmetric mode, with an intrinsic frequency splitting. Upon ES-SB, the dye is no

42

longer centro-symmetric and, thus the symmetric stretching mode has a non-zero IR cross

43

section, and the frequency splitting changes. The increase of this frequency splitting with

44

increasing polarity could be perfectly accounted for by the model.

45

Here we present an approach to describe the effect of solvent relaxation on the SB dy-

46

namics, as well as a method to model the accompanying IR spectral dynamics. It will be

47

applied to recent results obtained with a quadrupolar D-π-A-π-D dye (1, Figure 1) consisting

48

in a central benzothiadiazole electron acceptor flanked by two alkoxyphenyls donors, linked

49

through −C ≡ C− bridges.47 In non polar solvents, the TRIR spectrum exhibits a single

50

−C≡C−stretching band, in agreement with a purely quadrupolar symmetric excited state.

51

In polar solvents, the TRIR spectra show complex dynamics including the appearance of a

52

second −C ≡ C− stretching band and a frequency splitting that decreases with time. We

53

will show that the TRIR spectra recorded at different times and in two different solvents

54

can be perfectly reproduced with this extended model using the same set of parameters to

55

describe the dye. This approach allows determining the degree of asymmetry in different

56

solvents and its change during ES-SB. It also provides rich information on the different

57

mechanisms responsible for the band splitting. Finally, it confirms that the appearance of a

58

second vibrational band in the TRIR spectrum is an unambiguous evidence of ES-SB.

59

(6)

Theory

60

Overview of the ES-SB model

61

The theoretical model of ES-SB based on a two-essential-states approach was presented in

62

ref. 52 and its extension that includes the effect of electron-vibration coupling can be found

63

in ref. 50. Here, we present a brief overview of this model and the most relevant results.

64

The wavefunction of the excited states of a quadrupolar dye like 1 is expressed as:

65

Ψ =aLϕL+aRϕR (1)

whereϕL,R are the wavefunctions of the ’essential’ zwitterionic states with the positive charge

66

localized entirely on left or right branch, respectively, andaL,R are the associated coefficients

67

witha2L+a2R = 1. This quantities determine the positive charges of the donorsδL,R =−a2L,Re

68

with e being the electron charge. The total Hamiltonian accounting for symmetry breaking

69

and accompanying variation in the vibrational frequencies is39,50

70

H =HM+Hs+Hc+Hv0+Uint (2)

where

71

HM=

 0 V V 0

 (3)

includes the coupling between the two zwitterionic states, V, which determines the split-

72

ting of the one- and two-photon absorption bands, Hs describes the dipolar solute-solvent

73

interactions, Hc accounts for the Coulombic interactions, Hv0 and Uint are intramolecular

74

vibration and electron-vibration interactions, respectively. Within the self-consistent field

75

approximation, the Hamiltonian of the interaction of a polar molecule with a polar medium

76

is39,52

77

Hs =−λDDˆ (4)

(7)

where the dissymmetry parameter, D,

78

D=a2L−a2R = 2a2L−1 (5)

quantifies the extent of ES-SB. It varies from 0 in the absence of SB to 1 when the excitation

79

is entirely localised on one branch, i.e. when the excited state is purely dipolar. The

80

dissymmetry operator is defined as

81

Dˆ = ˆPL−PˆR (6)

with the projection operators ˆPL and ˆPR

82

R(L)Ψ = ˆPR(L)(aLϕL+aRϕR) = aR(L)ϕR(L) (7)

The mean value of the dissymmetry operator is equal to the magnitude of the dissymmetry

83

parameter D = hΨ|D|Ψi. According to the Onsager model,ˆ 53 the interaction between a

84

molecule with a permanent dipole moment and a polar solvent is described as50

85

λ= µ20∆f

r3d1∆f (8)

whereµ0 is the solute dipole moment,rd the cavity radius, and ∆f =f(εs)−f(n2) with the

86

solvent polarity function f(x) = 2(x−1)/(2x+ 1), the static dielectric constant εs and the

87

refractive index n. Eq 4 assumes that a molecule in a state with a dissymmetry parameter

88

D has an average dipole moment µ=Dµ0. The Hamiltonian of the Coulombic interactions

89

is defined as follows39,50

90

Hc=−γCDD, γˆ C = e2

imRLR (9)

where RLR is the distance between the charges located on the two donor sub-units, and εim

91

is the ”intramolecular” dielectric constant. The vibrational Hamiltonian of a quadrupolar

92

(8)

molecule with inversion symmetry has the form50

93

Hv0 = 1 2

X

i

p2sisi2x2si

+X

j

p2ajaj2 x2aj

(10)

where psi, paj, xsi, xaj, ωsi, ωsj are the momenta, the coordinates and the frequencies of the

94

symmetric and antisymmetric vibrational modes, respectively. The electron-vibration inter-

95

action is introduced through its formal expansion in powers of the dissymmetry parameter

96

and vibrational displacements up to the second order50

97

Uint =Uint(1)+Uint(2) (11)

Uint(1) = ˆDX

j

ζjxaj (12)

Uint(2) = ˆDX

ij

δijxajxsj +DDˆ 2

X

ik

ikxaixakikxsixsk] (13)

where only the terms that are invariant under the inversion symmetry transformation are

98

kept; ζj is the linear electron-vibration interaction parameter, δij describes the coupling

99

between the antisymmetric and symmetric vibrational modes upon SB, whereas the param-

100

eters αik and βik quantify their frequency shifts due to SB. The last term, Uint(2), describes

101

the frequency variations of the normal molecular vibrations and the Duschinsky effect upon

102

SB. Since the frequency variations are small in real molecules, Uint(2) is small as well and its

103

effect on SB can be neglected. It should be noted that in the adiabatic approximation, the

104

linear electron-vibration interaction transforms into50

105

int(1) =−ζDD, ζˆ =X

j

ζj2

ω2aj. (14)

Eq. 14 describes the effect of the electron-vibration interaction on the electronic subsystem

106

of the molecule.

107

To determine the resulting stationary excited states and energies of the molecule inter-

108

(9)

Figure 1: Energy-level scheme of a quadrupolar D-π-A-π-D molecule predicted by the model for the case where ES-SB is operative (D = 1), together with the structure of molecule 1. The distribution of the excitation in each state is represented by the red shading. The different shading for the state at E highlights the antisymmetric nature of its associated wavefunction. OPA and TPA stand for one- and two-photon absorption, respectively.

acting with the solvent, the wavefunction eq 1 is inserted into the Schr¨odinger equation

109

HΨ = EΨ (15)

where H = HM+Hs +Hc+Hv0 + ¯Uint(1). It is important to note that eq 15 has two sets

110

of the solutions. The first set of solutions is associated with the nonpolarized medium and

111

describes the symmetric (aL =aR) and an antisymmetric (aL=−aR) ’excitonic’ states with

112

the energies, E+ and E, respectively (Figure 1). The energy splitting of these states is

113

equal to 2V, and is equivalent to the Davydov splitting. The second set corresponds to the

114

polarized medium and includes two degenerate asymmetric states with the charge localized

115

predominantly on either of the two arms (|aL| 6= |aR|). Neglecting the quadratic electron-

116

vibrational interaction Uint(2), whose effect on the IR spectra is treated perturbatively, the

117

energy of these states is given by a simple expression

118

Eas =−(λ+γ), γ =γc+ζ. (16)

(10)

These states only exist when Eas < E. This condition is fulfilled when V < λ+γ. In this

119

case, optical population of the antisymmetric electronic state upon one photon absorption

120

is followed by SB. The dipolar character of the asymmetric states is given by

121

D= s

1− V2

(λ+γ)2 (17)

One can see that the D is determined by a single dimensionless parameter V /(λ+γ).

122

Symmetry Breaking and IR Absorption Spectrum

123

To see how the vibrational spectrum of the excited molecule is affected by ES-SB, we consider

124

two equivalent vibrational modes, here the −C ≡ C− stretching vibrations, localised on

125

each of the two arms of the molecule with the coordinates xL and xR. The interaction

126

between these two local modes results into antisymmetric and symmetric normal modes, with

127

coordinatesxaandxsand frequenciesωaandωs. If the two−C≡C−groups are separated by

128

several atoms, as it is the case here, direct coupling between the two local modes is negligible,

129

and ωa and ωs can be expected to be very close to each other.54 However, anharmonic

130

interactions between the vibrational modes as well as electron-vibration interactions can

131

lead to a splitting of these frequencies.50Both mechanisms are fairly universal, so a splitting

132

is expected in most molecules, although its magnitude can vary widely. In the model, this

133

splitting, in the absence of any ES-SB (D= 0), is given by ∆ω0s0−ωa0orω2s0−ωa022,

134

whereωs0s(D= 0),ωa0a(D= 0), andκis connected to the linear electron-vibration

135

coupling parameter, κ22/V50 (Figure 2, top).

136

The experimental TRIR spectra reveal that the changes in the vibrational frequencies of

137

the −C ≡ C− groups upon SB is about 1%. This weak effect indicates that the electron-

138

vibration interaction of these modes is weak as well. Therefore, the effect of SB on these

139

vibrational modes can be described in terms of perturbation theory, and the effect of these

140

vibrational modes on the electronic state of the solute can be ignored. On the other hand, the

141

(11)

Figure 2: Schematic illustration of the frequency splitting of the symmetric and antisym- metric−C≡C−stretching vibrations of 1in an IR absorption spectrum. In the absence of SB (D= 0), the splitting, is predominantly due to electron-vibration coupling, but only the antisymmetric vibration is IR active (top). Upon SB (D >0), other mechanisms associated with the parameters α, β and δ come into play. Both vibrations are IR active (bottom).

Here, α and δ are taken as positive, whereas β < 0. The functions f(x) should not be confused with the polarity functions.

linear electron-vibration coupling with the many other vibrational modes of the solute that

142

are not experimentally monitored can be significant, and was thus included in the electronic

143

Hamiltonian.

144

To calculate the IR absorption spectrum in a state with broken symmetry, we use the clas-

145

sical theory of the electromagnetic wave absorption. The interaction energy of the vibrations

146

with the electric field of an electromagnetic wave is written as

147

HintEM =−µviba E(t)−µvibs E(t) (18)

(12)

where E(t) is the projection of the electric component of the electromagnetic wave on the

148

direction of the oscillator dipole moment, µviba and µvibs are the dipole moments associated

149

with the antisymmetric and symmetric vibrations, respectively, which are assumed to be

150

parallel. In the electronic antisymmetric state, µvibs = 0 because the symmetric vibration

151

is not IR active. However, the dipole moments of the −C ≡ C− groups that link the

152

donor and acceptor sub-units of the molecule depends on D. This results in a difference

153

in the absorption coefficients of these two groups in the asymmetric electronic state. This

154

can be formally described by expanding the interaction energy eq 18 in a power series of

155

the asymmetry parameter D up to the second order. Considering the invariance of the

156

interaction energy eq 18 with respect to the inversion transformation and the fact that the

157

electric field E(t) is an odd function of this transformation, we obtain

158

µviba E(t)'F0cos(ωt)xa(1−ξaD2), µvibs E(t)'F0cos(ωt)xsξsD (19)

Here we have supposed that E(t) = E0cos(ωt) and included the amplitude of the electrical

159

field E0 in the quantity F0. It is shown below that the quantities ξa and ξs are measures of

160

the difference in dipole moment of the−C≡C−groups in the two branches of the molecule.

161

162

Furthermore, we assume that the electronic,τe, vibrational,τv, and medium,τm, timescales

163

satisfy the conditionτeτvτm. If the medium controls SB, then the dissymmetry param-

164

eterDis a much slower function of time than the high frequency vibrations, and consequently

165

the intramolecular high-frequency vibration modes follow adiabatically the variation of D.

166

Under these assumptions, the equations of motion of−C≡C−stretching vibrations in the

167

presence of an IR optical field at frequencyω have the following form

168

¨

xa =−∂Hv

∂xa −ηaa +Facos(ωt) (20)

¨

xs =−∂Hv

∂xs −ηss+Fscos(ωt) (21)

(13)

where Hv =Hv0+Uint, the dots above the vibrational coordinates denote time derivatives,

169

ηa and ηs are the damping frequencies of the antisymmetric and symmetric normal modes

170

and determine the homogeneous width of the IR absorption bands. The amplitudes of the

171

interaction between the vibrations and the IR field are

172

Fa =F0(1−ξaD2), Fs =F0ξsD (22)

Equation 22 is a simplest approximation of the dependence of the dipole moment of the

173

−C ≡ C− groups on the dissymmetry parameter. The comparison of the simulated and

174

experimental spectra described below evidences that the dipole moment of the −C ≡ C−

175

groups varies considerably upon changing the dissymmetry parameter.

176

A particular solution to the system of linear differential eqs 20–21 can be found in the

177

form of a sum complex conjugate solutions xa(s)(t) +xa(s)(t) where xa(s)(t) = ¯xa(s)eiωt. The

178

complex amplitudes ¯xa and ¯xs are given by a system of linear algebraic equations

179

a2−ω2+iηa)¯xa+Dδ¯xs+D2α¯xa =Fa (23) (ωs2−ω2+iηs)¯xs+Dδx¯a+D2βx¯s =Fs (24)

Here, the equations are written without the linear electron-vibrational interaction, since it

180

does not affect the form of the bands. The solutions of eqs 23 and 24 are

181

¯

xa = Fas−Fs

as−D2δ2, x¯s = Fsa−Fa

as−D2δ2 (25)

with

182

a = ¯ωa2−ω2+iηaω, ∆s = ¯ωs2−ω2 +iηsω (26)

183

¯

ωa220−κ2+D2α, ω¯2s02+D2β (27) where ω0s0.

184

(14)

According to the classical theory of absorption, the amount of energy absorbed per unit

185

time is equal to the work of the external force during the same time interval. It can be

186

calculated as

187

I(ω) =A(ω) = 1 T

Z T 0

Fadxa

dt +Fsdxs dt

cos(ωt)dt (28)

where

188

xa(s)(t) = 1

2 x¯a(s)eiωt+ ¯xa(s)e−iωt

(29) and T = 2π/ω is the period of the electromagnetic wave. This results in the IR spectral

189

intensity50

190

I(ω) =−1

2ω[FaIm¯xa+FsIm¯xs] (30) In the symmetric state, D = 0, eq 30 predicts a spectrum with a single band, which

191

corresponds to the absorption of the antisymmetric vibrational mode, the other one being

192

Raman active only (Figure 2, top). However, in a state with broken symmetry, D 6= 0,

193

the calculated spectrum exhibits two bands at the frequencies of the antisymmetric and

194

symmetric vibrations (Figure 2, bottom). In the case of narrow bands, ηa, ηs ω0, the

195

magnitude of the band splitting can be well estimated using eq 31:

196

∆ω= q

[(α−β)D2−κ2]2+ 4D2δ2

0 (31)

The model distinguishes four frequency splitting mechanisms illustrated in Figure 2, three

197

of which being due to SB. The first splitting mechanism is also operative in the symmetric

198

state of the quadrupolar molecule and its magnitude is related to the parameter κ (blue).

199

It originates from weak linear electron-vibration interaction50 and vibrational anharmonic

200

coupling. The second and third mechanisms are the frequency shifts of the antisymmetric and

201

symmetric vibrations due to SB (red). Their magnitudes are determined by the parameters

202

α and β. The fourth mechanism is associated with the interaction between the symmetric

203

and antisymmetric vibrations in an asymmetric electronic state and is described by the

204

(15)

parameter δ (green). Although, the molecule is no longer centro-symmetric upon SB, we

205

still refer to the vibrations as symmetric and antisymmetric to emphasize their origin.

206

The IR spectrum calculated with eq 30 only depends on the dissymmetry parameter

207

through the αD2, βD2, δD, ξaD2, and ξsD products. Therefore, it is an invariant of the

208

transformation:

209

α0 =αg2, β0 =βg2, δ0 =δg, ξa0ag2, ξs0sg, D0 =D/g (32)

with an arbitrary factor g. As a result, only relative D values can be determined from the

210

fit of eq 30 to the experimental spectra. It should be emphasized thatg is not an adjustable

211

parameter of the theory, since the simulated spectra does absolutely not depend on it.

212

The theory leading to eq 30 operates with normal vibrational coordinates of the quadrupo-

213

lar molecule in the symmetric electronic state. These coordinates have to be symmetric or

214

antisymmetric with respect to inversion. SB changes the electronic state of the molecule and

215

should result in a variation of the vibrational frequencies and to the Duschinsky effect. As

216

a result, normal vibrations in a state with broken symmetry are neither symmetric nor anti-

217

symmetric. To describe these variations, the vibrational Hamiltonian is formally expanded

218

in a power series of D up to the second order.50 This implies that the variations in the

219

natural frequencies should be small. However, this does not require the dissymmetry param-

220

eter, D, to be small also, the theory is applicable when the conditions including additional

221

multipliers:

222

αD202, βD220, δD/ω02, ξaD2, ξsD1 (33) are fulfilled. This is because, in fact, the power series contains the terms listed in eq 33.

223

Symmetry Breaking Dynamics

224

The TRIR spectra recorded with the D-π-A-π-D molecule 1 in polar solvents show pro-

225

nounced dynamics upon ES-SB that can be assigned to the increase of D with time.47 The

226

(16)

theoretical model can be generalized to describe the time dependence ofDreflected by these

227

spectral dynamics.50,52 A comparison of the theoretical model with the experimental data

228

enables the estimation of important parameters related to the solute and to its interaction

229

with polar solvents.

230

The magnitude of the dissymmetry parameter at equilibrium, Deq is given by eq 17.50,52

231

To include the time evolution of D in the theory, we start by noting that the operator of

232

the interaction energy of the solute with the solvent polarization described by eq 4 includes

233

the productDD. In this expression, the multiplier ˆˆ Daccounts for the dipole moment of the

234

solute, whileDis proportional to the magnitude of the solvent polarization. WhereasDand

235

the mean value of ˆD are equal at equilibrium, they can differ in a nonequilibrium state. For

236

an arbitrary nonequilibrium state, this interaction energy can be written as

237

Wint =hΨ|Hs|Ψi=−λDDm (34)

where Dm is a new quantity characterizing both the equilibrium and nonequilibrium polar-

238

ization of the solvent.

239

To justify eq 34, we note that, in the Onsager model, the interaction energy between a

240

solute dipole moment, ~µ, and a polar solvent is Wint =−~µ ~E, where E~ =~µ∆f /rd3 is the so-

241

called equilibrium solvent reaction field.53 By principle, the nonequilibrium solvent reaction

242

fieldE~m < ~E and corresponds to the field generated by an electric dipoleµ~m< ~µ. The energy

243

of the solvent reaction field Wm can be calculated as follows. In the linear approximation

244

of the solvent response, Wm has to be a quadratic function of Dm. The exact expression of

245

Wm is determined by the condition that the energy Wint+Wm has a minimum at Dm=D.

246

This gives Wm=λDm2/2.

247

The quantity Dm is thus directly associated with the solvent reaction field. For example,

248

Dm = 0 in an unpolarized solvent. As long as the solvent polarization is not equilibrated

249

with the field of the solute dipole moment, Dm is determined by the equality µm = µ0Dm.

250

(17)

Dm plays the same role in the theory of SB electron transfer as the reaction coordinate in

251

the electron transfer theory.

252

Solvent relaxation is usually expressed in terms of a relaxation function X(t).55 By defi-

253

nition, this function decreases from unity to zero. It reflects the decay of the solvent polari-

254

sation from its initial value in the presence an external field to zero after the external field

255

has been switched off. Consequently, the temporal evolution of Dm can be approximated as

256

Dm(t) = Deq(1−X(t)). This equation assumes a linear solvent response. The solvent po-

257

larisation, henceDm, is initially zero and then increases to its equilibrium value. As solvent

258

relaxation is assumed to be slow relative to electronic and intramolecular vibrational motion,

259

the electrons and nuclei of the solute follow adiabatically the changes in solvent polarization.

260

Within this approximation, the interaction energyWint =−λ(Deq)2(1−X(t)) increases with

261

time from zero to its equilibrium value. Formally, this equation can be obtained from eq 34

262

with the substitution

263

λ→λ(1−X(t)) (35)

Applying this substitution to eq 17 results in eq 36, which accounts for time dependence of

264

the dissymmetry parameter due to the varying solvent response:

265

D(t) = s

1− V2

[γ+λ(1−X(t))]2 (36) Eq. 36, designated from now on approach 1, is a relatively crude estimation of the time

266

dependence ofD. A better description (approach 2) can be obtained by solving the stationary

267

Schr¨odinger equation with a fixed value ofDm (adiabatic approximation). Eqs 18 – 20 from

268

ref 50 can be recast in the form

269

(λDm+γxy)x−(V −E)y = 0

(V +E)x+ (λDm+γxy)y = 0 (37)

x2+y2 = 2

(18)

wherex=aL−aR, andy=aL+aR. Forγ = 0, a simple equation relating the dissymmetry

270

parameter and solvent polarization is obtained:

271

D= λDm

pV22Dm2 (38) Here D = xy is the equilibrium value of the dissymmetry parameter of the molecule in

272

a frozen solvent with a given polarization Dm. When γ 6= 0, the functional dependence

273

D = D(Dm) can be calculated by solving eqs 37 numerically. The time dependence of the

274

dissymmetry parameter can then be calculated assumingDm(t) = Deq(1−X(t)). To describe

275

the temporal evolution of SB completely self-consistently, it is necessary to solve a dynamic

276

problem. We plan to do this in the future.

277

The physics underlying the above description is as follows. Immediately after photoex-

278

citation into the antisymmetric electronic excited state (Figure 1), the solute is purely

279

quadrupolar, D = 0, and is surrounded by non-polarized solvent. Therefore, the dipolar

280

solute-solvent interaction considered here is zero, corresponding formally to Dm= 0. Ther-

281

mal fluctuations of the solvent polarization rapidly induces a dipole moment on the solute,

282

which in turn polarises the solvent. Such a concurrent increase in both solvent polarization

283

and dipolar character of the solute proceeds until equilibrium is reached. This type of re-

284

laxation could be expected to be slower then the solvent response to a quasi instantaneous

285

change of permanent dipole moment of the solute, as determined for example from time-

286

resolved fluorescence measurements with push-pull molecules.56,57 The relaxation function

287

X(t) deduced from such measurements is usually approximated by eq 39:55

288

X(t) = X

i

aiexp(−t/τi), X

ai = 1 (39)

Figure 3 illustrates the influence of the parameters γ and λ on the time dependence of

289

D, assuming a single exponential solvent relaxation function X(t) = exp(−t/τ) with τ = 1

290

ps. A few important conclusions can be drawn from the analysis of eq 36 (approach 1) and

291

(19)

0 2 4 6 0 . 0

0 . 2 0 . 4 0 . 6 0 . 8

Dissymmetry parameter

t i m e / p s

γ/ V = 0 . 7 , λ/ V = 0 . 5 γ/ V = 0 . 9 , λ/ V = 0 . 5 γ/ V = 0 . 9 8 , λ/ V = 0 . 5 γ/ V = 1 . 2 , λ/ V = 0 . 5

Figure 3: Time dependence of the dissymmetry parameter D for different values of γ and λ, calculated with eq 36 (solid lines) and by solving eqs 37 (dashed lines), and assuming monoexponential solvent relaxation with a 1 ps time constant.

the solution of eqs 37 (approach 2). As mentioned above, SB is not possible at any time

292

when (γ +λ)/V < 1, and D(t) ≡ 0 (not shown in Figure 3). When γ/V is significantly

293

smaller than one and (γ +λ)/V slightly larger than unity, approaches 1 and 2 predict

294

strongly different D(t) (compare the red solid and dashed lines). Since approach 2 is more

295

consistent, the first one can be considered as inapplicable for such parameters. At the same

296

time, when γ/V is close to but larger than unity, both approaches lead to similar results

297

(compare the green solid and dashed lines). In this case, the much simpler approach 1 can be

298

applied. It should be noted that whenγ/V >1, the dissymmetry parameter differs from zero

299

already at time zero. This is the case when the intramolecular interactions, assumed here

300

to have an instantaneous response, suffice to break symmetry without solvent polarization.

301

Both approaches show that, immediately after the onset of SB, D rises very rapidly before

302

reaching a plateau. This fast increase ofDoccurs on a shorter timescale than that of solvent

303

relaxation due to the presence of a threshold in the dependence ofD on solvent polarization.

304

(20)

Results and Discussion

305

Analysis of TRIR Spectra

306

The above model is now put to the test with TRIR spectra recorded in the−C≡C−stretch-

307

ing region upon S1 ←S0excitation of the D-π-A-π-D molecule1in the non polar cyclohexane

308

(CHX), in the medium polar tetrahydrofuran (THF) and in the highly polar dimethylfor-

309

mamide (DMF).47 In CHX, the spectra show a single excited-state absorption band, whose

310

shape and position remain essentially independent of time. Two excited-state absorption

311

bands are observed in the polar solvents, with positions and relative intensities changing

312

significantly with time (Figure 4).47 These spectra have been recorded upon red-edge exci-

313

tation of the S1 ←S0 transition. Therefore, any spectral dynamics due to the relaxation of

314

vibrational excess energy can be safely excluded.

315

The experimental TRIR spectra of the D-π-A-π-D molecule 1 show that the bandwidth

316

of the −C≡C− stretching vibrations increases noticeably with the solvent polarity. In gen-

317

eral, such trend originates from the direct interaction of the polar bond vibrations of the

318

dye with solvent polarization fluctuations. However, for molecules undergoing symmetry

319

breaking, at least two specific band-broadening mechanisms should additionally be consid-

320

ered. Both are due to the interaction of the solvent polarization fluctuations with the overall

321

dipole moment of the solute that results in fluctuations of the dissymmetry parameter D.

322

These fluctuations of D lead to (i) a modulation of the intramolecular vibration frequencies

323

and (ii) a modulation of the dipole moment of the intramolecular polar bonds (see eq 19).

324

Both types of modulations affect the bandwidth and their contribution is expected to be

325

significant. Theoretical analysis of this phenomenon is very complex due to the nonlinear

326

coupling of the dissymmetry parameter fluctuations with the solvent fluctuations. Retar-

327

dation effects, i.e. the non-instantaneous response of intramolecular vibrations to changes

328

in solvent polarization, can also be important. This complex problem is not analysed here.

329

Instead, we refer to previous transient 2D-IR spectroscopy experiments, which showed that

330

(21)

the IR bands of a similar dye measured in a series of polar solvents can well reproduced by

331

a Voigt function with a homogeneous width of about 10 cm−1, similar to the value assumed

332

here (see Supporting Information in ref 58). Theory also shows that the IR band of solutes

333

in solvents with two different relaxation timescales can be well approximated with a Voigt

334

function.59

335

Equation 30 predicts the IR absorption bands to be close to Lorentzian since only ho-

336

mogeneous line broadening is considered. This is true as long as the widths ηa and ηs are

337

similar, as it is the case here. In the limit where ηa ηs, the weaker band has a strongly

338

asymmetric shape typical of a Fano resonance.60 However, comparison with experimental

339

spectra should take inhomogeneous broadening into account. For this, the TRIR spectra

340

with two clearly distinguishable bands were fitted with the sum of two Voigt functionsf(ω):

341

f(ω) =Z−1ω[V(ω−ω1, γ1, σ1) +V(ω−ω2, γ2, σ2)] (40) V(ω, σ, γ) =

Z

−∞

G(ω0, σ)L(ω−ω0, γ)dω0 (41) G(ω, σ) = 1

σ√

2πe−ω2/(2σ2) (42)

L(ω, γ) = γ

π(ω22) (43)

whereZ−1is a normalization factor. The functionV(ω−ωi, γi, σi) in eq 40 is the convolution

342

of a Gaussian inhomogenous distribution, G(ω), of width σi, with a Lorentzian function,

343

L(ω), of width γi and centred atωi. A fitting off(ω) to experimental TRIR spectra returns

344

the deconvoluted spectrum fD(ω)

345

fD(ω) =Ze−1ω[L(ω−ω1, γ1) +L(ω−ω2, γ2)] (44)

which does not include the inhomogeneous broadening. Here Ze−1 is another normalization

346

factor and determines the ratio of the heights of the two Lorentz bands. The simulated

347

spectrum eq 30 is then fitted to the function fD(ω).

348

(22)

Fitting the sum of two Voigt functions with arbitrary inhomogeneous (σi) and homo-

349

geneous widths (γi) to the TRIR spectra measured at different time delays in polar sol-

350

vents leads to a rather large scattering in the magnitude of the homogeneous width. Since

351

such result is not meaningful, the analysis was repeated with a fixed homogeneous width,

352

γ1 = γ2 = 10 cm−1. The quality of all fits is excellent in the area of both band maxima

353

as illustrated in Figure 4. The best-fit parameters are listed in Table 1. However, the

354

low-frequency band is strongly asymmetric and its low-frequency shoulder is not so well re-

355

produced. This discrepancy is minor in THF but increases in the more polar DMF. This

356

reflects the evolution of the distribution of frequencies in a highly asymmetric potential

357

E +Wm (see eq 37). As expected, this deviation decreases with time. This band-shape

358

asymmetry demonstrates the important role of the specific band-broadening mechanisms.

359

However, this discrepancy has a minor effect on the positions of the band maxima and can

360

lead to an overestimation of the relative amplitude of the high-frequency band. The error

361

on the parameter specified in Table 1 is about 15% at early time and decreases at longer

362

time.

363

This fit of eq 40 can be considered as a procedure to remove the inhomogeneous broad-

364

ening from the experimental TRIR data. The resulting deconvolved spectra, depicted with

365

solid coloured lines in Figure 4, can now be fitted with eq 30 to determine the parameters

366

associated with both the solute (ω0,κ,α,β,ξas) and the solvent response, which are indi-

367

rectly reflected in the magnitude of the dissymmetry parameter {Di} at a given time delay

368

{ti}. This fit involves a large number of adjustable parameters. To improve the convergence,

369

the optimization procedure was split into three stages. (i) In the first stage, the {Di} values

370

at different time delays were considered as independent adjustable parameters, whereas the

371

solute parameters were fixed. The {Di} values were then found using the golden section

372

search method in the 0 to 1 range. (ii) In the second stage, the solute parameters α, β, κ,

373

ξa, andξs were optimized using the particle swarm optimization method. Here, ω0 was kept

374

constant, since it strongly influences the spectral behavior. Note that to reduce computation

375

(23)

2020 2080 2140

ω / cm

1

f ( ω ) /ω , f

D

( ω ) /ω , a .u .

t=0.2 t=0.4 t=1.0 t=2.0 t=4.0 t=8.0 t=20

THF

2000 2060 2120 2180

ω / cm

1

DMF

Figure 4: TRIR spectra in the −C≡C− stretching region recorded at various times after excitation of 1in THF and DMF (dots) and best fits of the sum of two Voigt functions (eq 40, black dashed lines). The deconvolved spectra (eq 44) including only the homogeneous broadening are represented with solid coloured lines.

time, the particle swarm optimization method can be used as an initial approximation for

376

the Nelder-Mead optimization method. (iii) In the third and final stage, the value of ω0 was

377

determined using again the golden section search method, with the interval set within the

378

limits of the width of the weaker band. Thus, we formulated a triple embedded optimiza-

379

tion problem. To improve the accuracy in the determination of the solute parameters, it is

380

necessary to use sets of experimental data in at least two different solvents.

381

The best-fit of eq 30 to the deconvolved TRIR spectra in THF and DMF are presented

382

(24)

Table 1: Parameters obtained from the fit of eq 40, assuming fixed homogeneous widths γ1 = γ2 = 10 cm−1, to the TRIR spectra of 1 recorded at various time delays in THF and DMF. The band widths, σi, and positions, ωi, are in cm−1.

solvent time / ps ω1 σ1 ω2 σ2

CHX 20 0.000 2072.3 3.33 – –

THF 0.2 0.038 2083.5 6.79 2145.5 0.00 0.4 0.054 2083.5 7.32 2144.4 0.00 1 0.081 2086.0 8.22 2143.1 0.00 2 0.106 2088.0 8.68 2142.2 1.14 4 0.129 2089.4 9.26 2141.8 0.81 8 0.143 2090.7 9.19 2141.5 0.01 20 0.143 2091.4 9.34 2141.5 0.00 DMF 0.2 0.150 2090.1 14.85 2142.1 4.23 0.4 0.203 2092.6 14.70 2141.4 2.41 1 0.287 2095.5 15.16 2139.8 0.00 2 0.419 2098.4 15.57 2139.2 0.00 4 0.603 2101.3 15.41 2138.8 0.00 8 0.829 2103.2 14.84 2138.6 0.00 20 1.000 2103.8 15.08 2138.4 0.00

in Figure 5 with the parameters listed in Tables 2. Due to the invariance of the spectrum

383

calculated with eq 30 relative to the transformation eq 32, the value of one of the parameters,

384

δ/ω02 = 0.01, was fixed. Such choice only determines the relative values of D, and does not

385

affect the quality of the fit. According to eq 32, the D values obtained from the fit are not

386

calibrated and can, thus, exceed 1. Figure 5 reveals that the fit of eq 30 to the experimental

387

TRIR spectra at different time delays in THF and DMF using a single set of solute parameters

388

is excellent. The TRIR spectra in CHX are practically independent of time and comprise

389

a single band that points to an absence of SB in this solvent. The spectrum calculated for

390

D = 0 with the solute parameters obtained from the fit in polar solvents coincides very

391

well with the experimental spectra of 1 in CHX, as shown in Figure S1. Moreover, the

392

value of ω0 obtained from the fit leads to a frequency of the antisymmetric vibrations of

393

ωa00 −∆ω0 = 2073.6 cm−1, in perfect agreement with the band measured in CHX. It

394

is well known that the fit of a Voigt function to IR spectra does not allow determining the

395

Lorentz width with high precision. However, uncertainties in the homogeneous widths of the

396

(25)

2060 2100 2140

ω / cm

1

I ( ω ) /ω , a .u .

t=0.2 t=0.4 t=1.0 t=2.0 t=4.0 t=8.0 t=20

THF

2060 2100 2140

ω / cm

1

DMF

Figure 5: Best fits of eq 30 (black dashed lines) to the experimental TRIR data in different solvents after removal of the inhomogeneous broadening using eq 40 (coloured lines).

deconvoluted TRIR spectra do not have a significant effect on the solute parameters obtained

397

from the fit of eq 30. If the homogeneous bandwidth,γi, obtained from the fit overestimates

398

the true value by a factorh, the band-height is underestimated by the same factor. Therefore,

399

this imprecision affects neither the band positions nor the solute parameters.

400

The low-frequency band in the TRIR spectra of1in polar solvents is considerably broader

401

than the high-frequency band (Table 1). One of the reasons for this difference is that |α|

402

is significantly larger than |β|. Therefore, the position of the low-frequency band (anti-

403

symmetric mode) is much more sensitive to the degree of asymmetry than the position of

404

(26)

the high-frequency band (symmetric mode). In this case, it is natural to expect that fluc-

405

tuations in the dissymmetry parameter, themselves due to solvent fluctuations, lead to a

406

stronger inhomogeneous broadening of the low-frequency band.

407

Table 2 indicates that all parameters are within the limits of applicability of the theory

408

given by eq 33, except forξaD. The latter implies that the dipole moment of the −C≡C−

409

group varies strongly up SB. In reality, this dependence can significantly deviate from the

410

linear approximation assumed in eq 22.

411

Table 2: Solute parameters obtained from the fit of eq 30 to the deconvolved TRIR spectra measured with 1 in THF and DMF. The values of , 2γ1a, and 2γ2s are taken from Table 1.

parameter time, ps D(CHX) D(THF) D(DMF)

ω0, cm−1 2147.2 0.2 0.58 0.73

κ, cm−1 562.0 0.4 0.58 0.79

∆ω0, cm−1 73.5 1 0.64 0.84

δ/ω20 0.010 2 0.69 0.90

α/ω20 0.032 4 0.72 0.95

β/ω02 -0.012 8 0.74 0.98

ξa 0.604 20 0 0.76 1.0

ξs 0.190

Frequency Splitting and Dissymmetry Parameter

412

The best-fit parameters listed in Table 2 indicate that, in the absence of SB (D = 0), the

413

symmetric and antisymmetric −C≡C− stretching vibrations of 1 are split by ∆ω0 = 73.5

414

cm−1. In this case, only the antisymmetric vibration contributes to the IR spectrum (Figure

415

2). As SB takes place, i.e. as D increases, the ’symmetric’ vibration becomes also visible

416

in the IR spectrum. Figure 5 reveals that the splitting of the two bands decreases with

417

increasing D. This behaviour contrasts with that observed previously with an A-π-D-π-

418

A molecule consisting of a pyrrolopyrrole D core and two cyanophenyl acceptors where

419

the splitting was increasing with D. This opposite behaviour is due a different sign of the

420

parameterβ, which itself depends on the nature of the vibrational mode considered,−C≡C−

421

Références

Documents relatifs

Since the 2D model does not include a dissipation term and because the energy transferred into transverse motions is very small, almost all the energy of the initial impulse stays

Here we perform non-equilibrium molecular dynamics simulations of a defect-free graphite nanoplatelet suspended in a shear flow and measure the critical shear rate ˙γ c needed for

The proof is given that in a symmetric double well model of the symmetric H-bond with damped stretching mode, there is no asymptotic-time symmetry breaking owing to a full..

The two columns on the right show the surface elevation and free surface modal amplitude of a slightly larger wave packet where the breaking parameter is calculated as B x = 0.85091

Son implémentation dans un code spectral complet de 3 ème génération tel que SWAN fournit un outil opérationnel pour affiner la modélisation du déferlement

The whole set of measured line positions, intensities, as well as self- and N 2 -broadening coefficients for the ν 6 band of CH 3 Br is given in Table 2, in which the

Design, synthesis, and reaction of π-extended coumarin-based new caged compounds with two-photon absorption character in the near-IR region.. Chemistry Letters, Chemical Society

In a microbeam study, Lindau-Webb and Ladefoged (1989) showed that the x and y coordinates of two pellets located along the tongue midline may suffice to reconstruct