Comparison between Theoretically and Experimentally Determined

Feb 3, 2015 - ... Experimentally Determined. Electronic Properties: Applications to Two-Photon Singlet Oxygen. Sensitizers. Christian Benedikt Orea Ni...
0 downloads 0 Views 1MB Size
Subscriber access provided by GEORGIAN COURT UNIVERSITY

Article

Comparison between Theoretically and Experimentally Determined Electronic Properties: Applications to Two-Photon Singlet Oxygen Sensitizers. Christian Benedikt Orea Nielsen, Henning Osholm Sørensen, and Jacob Kongsted J. Phys. Chem. A, Just Accepted Manuscript • DOI: 10.1021/jp5122849 • Publication Date (Web): 03 Feb 2015 Downloaded from http://pubs.acs.org on February 18, 2015

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

The Journal of Physical Chemistry A is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

February, 2015

Comparison

between

Theoretically

and

Experimentally Determined Electronic Properties: Applications

to

Two-Photon

Singlet

Oxygen

Sensitizers. Christian B. O. Nielsena,b,c Henning Osholm Sørensen,d and Jacob Kongstede a

Polymer Department, Risø National Laboratory, Frederiksborgvej 399, DK-4000 Roskilde,

Denmark b

Department of Chemistry, University of Aarhus, Langelandsgade 140, DK-8000 Aarhus C,

Denmark c

Present address: Sun Chemical A/S, Københavnsvej 112, DK-4600 Køge, Denmark

d

Nano-Science Center, Department of Chemistry, Universitetsparken 5, DK-2100 Copenhagen Ø,

Denmark

ACS Paragon Plus Environment

1

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 37

e

Department of Physics, Chemistry and Pharmacy, University of Southern Denmark, Campusvej

55, 5230 Odense M, Denmark Correspondence should be addressed to [email protected] Running title: Electronic structure of oligo-phenylene vinylene molecules.

Abstract A series of oligo-phenylene vinylenes are investigated with spectroscopic measurements and calculations with the aim to rationalize the correlation between the electronic structure of the molecules and their efficiency as two-photon singlet oxygen sensitizers. The band-shape functions of selected two-photon absorption processes are analyzed and the corresponding spectra are deconvoluted. This analysis allows for a comparison between calculated and measured twophoton absorption cross sections where a reasonable agreement is found.

Introduction Predictions of two-photon absorption (TPA) properties and in particular the TPA cross section of molecules have recently received significant interest in materials applications such as imaging and optical limiters,1-5 3D-microfabrication,6-8 lasing materials9-11 and two-photon sensitized production of singlet oxygen.12-19 Seminal work within the area of two-photon absorption processes originates from Göppert-Meyer20 who not only defined the TPA cross-section, but also derived selection rules for electronic transitions in a TPA process: Molecules possessing centrosymmetric symmetry obey selection rules for which excited electronic states can be populated. In a one-photon absorption process only electronic states with opposite parity to the initial state can

ACS Paragon Plus Environment

2

Page 3 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

be populated. This is in contrast to the TPA process where only electronic states of the same parity as the initial state can be populated. Experimentally observed TPA cross-sections can be rationalized according to the expression by Göppert-Meyer,20

2 e 4 2  2 2 2  c 0 h

 i

 f  i

i   g

i  

2

g  2 Eq.(1)

where e is the charge of the electron, c is the speed of light, ε0 is the vacuum permittivity, h is Planck’s constant, ω is the energy of the incident photons, g(2ω) is the line width function of the final state,  is the dipole operator,  f is the final electronic state, Y i is an intermediate virtual state, g is the ground state, and ωi is the energy separation between the ground state and the intermediate states. The TPA cross section has the unit of cm4·s but is often expressed in the Göppert-Meyer unit (GM): 1 GM = 10-50 cm4·s. It is important to emphasize that Eq. (1) is only valid when the two-photon excitation process occurs with two photons of the same wavelength. Otherwise, an additional term is needed that take into account the different wavelengths of the two photons.20 Another important aspect is the energies of the excited states which in a twophoton processes is the sum of energies of the two photons needed to populate the excited state. The implementation of the response methodology in for example the Dalton quantum chemistry program21,22 allows a direct calculation of the sum-over-states term in Eq. (1), but it is not possible to calculate the line width function using ab-initio methods. Often, the calculated TPA cross sections reported in the literature uses a fixed value for the band-shape function at the absorption maximum.23-24 One of the purposes of the present work is to quantify the band shape function for a variety of molecules in order to probe whether meaningful comparisons can be

ACS Paragon Plus Environment

3

The Journal of Physical Chemistry

made using a fixed value for the band-shape function at the absorption maximum for different molecules. TPA cross sections can be measured using a variety of experimental techniques such as Zscan,25 two-photon fluorescence26 and photoacoustic measurements.27,28 The two latter techniques use either fluorescence or an acoustic signal as the probe for the optical process. Another probe useful for measuring TPA cross sections is singlet oxygen formed in a cascade reaction upon relaxation of the excited state of the molecule studied (hereafter referred to as the singlet oxygen sensitizer), where energy is transferred between the singlet oxygen sensitizer and oxygen itself (Figure 1).

Sn Sm IC ISC

S1 virtual Fluorescence

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 37

T1 Oxygenation and/or Oxidation of M

IC

O 2(a1Δ g) ISC Energy Transfer

S0

Two One Photon Photon

a-X emission and physical deactivation

M

O 2(X 3Σ g- )

Figure 1. Illustration of one- and two-photon triplet-state photosensitized production of singlet oxygen, O2(a1g). Depending on the sensitizer, the simultaneous absorption of two photons may or may not populate the same state as that created upon the absorption of a single higher-energy photon. ISC denotes intersystem crossing and IC denotes internal conversion.

The label “a-X emission” refers to the 1270 nm

phosphorescence of singlet oxygen. M denotes the singlet oxygen sensitizer.

Obtaining high TPA cross sections on its own has been the focus of much work and has led to molecular design rules, briefly summarized in the following. For example, narrowing the

ACS Paragon Plus Environment

4

Page 5 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

excitation bandwidth of the line width function, g(2ω) in Eq. (1), will result in an increase of the line width function at its maximum value since this function should be normalized. Decreasing the width of the line width function is done by reducing the density of states function for the final electronic state, i.e. reducing the number of vibrational states in the vicinity of the excited state. Also, by increasing the conjugation length of a π-conjugated system an increase in the density of states for all the electronic states is obtained.29 Another design rule focusses on having a onephoton transition close to the two-photon laser frequency. This will create a singularity in Eq. (1) as a i- difference close to zero would be obtained. One way to achieve this is to have a large conjugation length. On the other hand this could impose solubility problems. The most common approach in the molecular design of molecules possessing large TPA cross sections is to introduce chemical functional groups in the molecules that give rise to significant charge relocalization upon excitation, i.e. significant changes in ground-to-excited state dipole moments.3033

A variety of molecular templates have been used along this line including distyrylbenzenes and

naphthalene derivatives.30-37 Most of these compounds are centrosymmetric and are designed for use as fluorescence probes. Significant changes of the dipole moment, however, is not always necessary to obtain molecules with large TPA cross sections. Polyfluorenes, for example, exhibit a large cross section of approximately 70,000 GM. This large value has been suggested to be the result of the high molecular weight and therefore larger conjugation compared to low molecular weight molecules.38 Similar findings have been obtained for certain dendrimers.39 Optimization of both a large TPA cross section and a large singlet oxygen quantum yield has been the focus of some recent works.12-19 The present work aims at quantifying the electronic properties that will lead to high TPA cross sections and also a high singlet oxygen quantum yield.

ACS Paragon Plus Environment

5

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 37

These studies also include the particulars of the band-shape function for the two-photon transitions. Experimental and computational details Voltammetry. In the cyclic voltammetric experiments a three electrode setup was employed with a glassy carbon (o.d.=1 mm) serving as working electrode. The reference electrode applied was Ag/AgI ( C I- =0.1 M) but all the potentials were also referenced against the ferrocenium/ferrocene (Fc+/Fc) redox couple. A platinum coil was used as counter electrode. The ohmic drop was compensated with a positive feedback system incorporated in the home-built potentiostat. Fluorescence. Solutions were degassed for 15 min with argon prior to use. Data were recorded using an instrument comprised of a 450 W Xe lamp for steady-state measurements and a diode laser for lifetime measurements. The detection system was a single-photon-counting photomultiplier tube in a peltier-cooled housing. All spectra where measured in a perpendicular geometry using 1-cm quartz cuvettes. Steady-state measurements were obtained with 1.8 nm band pass filters and corrected for wavelength dependent intensity variation of the excitation light source. Quantum yields were determined using 9,10-diphenylanthracene in cyclohexane40 (Φf =1.00) as the fluorescence standard with refractive index and differential absorption corrections. In all cases, time-resolved fluorescence decay traces were single exponential and lifetimes were determined using least-squares analysis. All time-resolved measurements were conducted with 7.3 nm band pass filters. Singlet oxygen quantum yields. Singlet oxygen quantum yields, Φ, were determined using instruments and an approach that has been described previously.35,41 For these measurements, the

ACS Paragon Plus Environment

6

Page 7 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

samples were irradiated at 355 nm (third harmonic of a nanosecond Nd:YAG pulsed laser). The standard used for experiments performed in toluene was phenalenone (Φ=1.00±0.05).42

Crystallographic structure determinations. Crystals suitable for structure determination by single crystal X-ray diffraction were obtained of compounds 10, 11 and 17. The diffraction data were collected using a Nonius KappaCCD employing Mo Kα radiation (λ = 0.71073 Å) on crystals cooled to 122(1) K. The scan parameters were chosen following the recommendations in Ref. [43]. Data reductions were performed with EvalCCD44 for 10 and 11 and DENZO and SCALEPAK45 for 17. All reflections were corrected for background, Lorentz and polarization effects. Absorption correction was performed using the numerical Gaussian integration procedure46 for 17 only within the maXus suite.47 The structures were solved by Direct methods using SIR9748 for 10 and SHELXS9749 for 11 and 17. The structures were refined by full matrix leastsquares on |F|2 values against all reflections with SHELXL97.49 The non-hydrogen atoms were refined with anisotropic displacement parameters, except for the atoms in disorded solvent molecules found in 10 and 11. All hydrogen atoms were introduced using a riding model with idealised geometry. In the crystal structure 10 solvent molecule methylphenylether was cocrystallizated. The solvent molecule was refined as a disordered entity with two molecular positions. The occupancy of the two positions were almost equal, 0.486/0.514(4). Restraints were used for the anisotropic displacement parameters for the carbon atoms in the phenyl group such that the displacements are approximatel isotropic. Also compound 11 was cocrystallized with solvent molecules. The toluene molecule was found in a position over an inversion center, hence four molecules shared the position of which two were unique. The occupancy of the unique molecules were 0.179(5) and 0.321(5). Together with the inversion symmetry this makes 1 molecule in average on this position. The carbon atoms were refined with isotropic displacement

ACS Paragon Plus Environment

7

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 37

parameters which within each molecule were restrained to have approximately the same displacement parameter. The detailed crystal data are presented below.

Crystal data for 10. C48H40N2O2, C7H8O, Mw = 784.95 g mol−1, crystal dimensions 0.49×0.30×0.11 mm3, triclinic, space group P-1, a = 10.069(4), b = 13.876(3), c = 16.500(6) Å, α = 109.803(19)◦, β = 94.46(2)◦, γ = 103.692(20)◦, V = 2075.6(12) Å3, Z = 2, ρcalc = 1.256 g cm−3, μ(Mo Kα) = 0.077 mm−1, 1.33 < θ < 25.00◦, of 53270 measured reflections, 7293 were independent (Rint =0.0622) and 5585 observed with I > 2σ(I); R1 = 0.0733, wR2 = 0.1255, GOF = 1.055 for 587 parameters, Δρmax/min = 0.451/−0.246.

Crystal data for 11. C48H36N2O2, C7H8, Mw = 764.92 g mol−1, crystal dimensions 0.52×0.18×0.10 mm3, triclinic, space group P-1, a = 8.5037(11), b = 9.8109(13), c = 13.1211(7) Å, α = 92.898(10)◦, β = 97.615(9)◦, γ = 107.974(9)◦, V = 1027.2(2) Å3, Z = 1, ρcalc = 1.237 g cm−3, μ(Mo Kα) = 0.074 mm−1, 1.57 < θ < 25.01◦, of 24859 measured reflections, 3618 were independent (Rint =0.0668) and 2812 observed with I > 2σ(I); R1 = 0.0706, wR2 = 0.1283, GOF = 1.008 for 266 parameters, Δρmax/min = 0.336/−0.251.

Crystal data for 17. C46H34N4O4, Mw = 706.77 g mol−1, crystal dimensions 0.60×0.25×0.16 mm3, orthorhombic, space group Pbca, a = 10.981(2), b = 16.0090(19), c = 19.9870(18) Å, V = 3513.6(9) Å3, Z = 4, ρcalc = 1.336 g cm−3, μ(Mo Kα) = 0.086 mm−1, 2.04 < θ < 32.08◦, of 85028 measured reflections, 6122 were independent (Rint =0.0779) and 4549 observed with I > 2σ(I); R1 = 0.0904, wR2 = 0.1686, GOF = 1.065 for 244 parameters, Δρmax/min = 0.646/−0.244.

ACS Paragon Plus Environment

8

Page 9 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Computational details. Geometry optimizations for determining the molecular structures in the present work were performed at the B3LYP/6-31G(d) level using the Gaussian 03 suite of programs.50 Vibrational frequency calculations were carried out on the optimized structures in order to verify that the structures represent minima on the potential energy surface. These geometry optimizations and frequency calculations were performed for the molecules in isolation. The optimized structures were then used as input for response calculations at the CAM-B3LYP/631G(d) level (both in vacou and employing a solvent model) carried out with the DALTON program package.21 Results and discussion All the molecules discussed in the present work can be described with the general structure shown in Chart 1. 

Aryl



Aryl

Aryl





A





D







D

D



A



D

A



A



A

Chart 1: General structure of the sensitizers examined in the present work. The letters A and D refer to electron acceptors and donors, respectively, and illustrate methods by which substituents can be used to influence the electronic character of the chromophore. The π-moiety is either aryl (phenyl, toloyl), alkenyl, acrylonitrile or acrylaldehyde.

The molecules are all centrosymmetric and are built around a core aromatic ring with vinyl substituents in the 1- and 4-positions (phenyl) or the 2- and 6-positions (naphthalene). The enddonor groups are either amine, thio-methyl-ether or methoxy groups. The only end-acceptor

ACS Paragon Plus Environment

9

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 37

group applied was a substituted imidazole. The acceptor groups on the center aryl ring are bromine , aldehydes, ketones, amides, nitriles, sulfones, fluorine or nitro groups. In continuation of previous works,12,13,15,29,51,52 the success criterion for a good TPA singlet oxygen sensitizer is defined as a large value of the product of the TPA cross section () and the singlet oxygen quantum yield (ΦΔ). Some of the molecular design features for obtaining large TPA cross sections are outlined above, with the presence of a high degree of intramolecular charge transfer (CT) in the molecule as a key feature. However, the extent of CT, both within the sensitizer itself as well as in the sensitizer oxygen complex (intermolecular CT) can have large adverse effects on the efficiency in which singlet oxygen is generated.53-58 In general, CT character in a sensitizer and/or sensitizer-oxygen complex facilitates non-radiative deactivation of the excited state at the expense of energy transfer from the sensitizer to produce singlet oxygen. It is assumed that the extent of intermolecular CT in the sensitizer-oxygen complex will be larger for a sensitizer with a proclivity for intramolecular CT (i.e. intermolecular CT to yield a Sens-O2 state with Sens+··O2- character will more readily occur for a sensitizer with an electron-donating moiety). One particular design principle known to promote higher singlet oxygen quantum yields is the heavy atom effect, which facilitates the efficiency of S1->T1 intersystem (see Figure 1) crossing due to the increasing amount of spin-orbit coupling in the molecule.59,60 In addition, the El-Sayed selection rule59,60 states that intersystem crossing from the S1 to the T1 state becomes allowed if there is a change in the orbital angular momentum to balance the change in spin angular momentum. This is possible when the electronic transition includes excitation from or to an n- or orbital to or from a -orbital. Functional groups where such transitions are likely to occur include carbonyl groups.59,60 Quantification of some of the key photophysical parameters have been carried out in previous work51 whereas the present work is focused on a quantum

ACS Paragon Plus Environment

10

Page 11 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

mechanical approach to interpret the measured singlet oxygen quantum yields and the TPA cross sections. 1. Changes in the central moiety of the Oligo-phenylene vinylene motif. In the oligophenylene vinylene (OPV) motif the central moiety was systematically changed in order to address the points made above. One-photon absorption properties and fluorescence, fluorescence quantum yields and lifetimes and singlet oxygen quantum yields are presented in Table 1. The E0,0-energies are obtained from the points where the corresponding absorption and emission profiles (normalized intensities) crosses and is thus interpreted as the S1-energies.

Table 1:

One-photon optical characterization (absorption, fluorescence and singlet oxygen

quantum yields measurements) of sensitizers with different aromatic cores substituted with diphenylamino-phenylenevinyl.a  abs max [nm] is the absorption maximum, log ε is the log of the extinction coefficient (ε),  em max [nm] is the emission maximum, Φf is the fluorescence quantum yield,  [ns] is the fluorescence lifetime and E0,0 [eV] is the energy of the first excited state. Ph2N Aryl NPh2

Compound

F

2

 [ns]

 em max [nm]

428/4.90 304/4.61

486 517

0.8

1.1

409/4.90 300/4.62

459 481

0.8

1.0

ΦΔ

E0,0 [eV]

TPA [GM]

0.09

N.D.

1970

0.08

N.D.

2030

F

1 F

Φf

 abs max [nm]/log ε

F

ACS Paragon Plus Environment

11

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 37

Br

0.2

0.4

0.13

N.D.

1410

452 479

0.7

1.0

0.06

N.D.

N.D.

375/4.75 300/4.63

426

0.5

1.6

0.05

N.D.

N.D.

382/4.92 302/4.77

525

0.7

3.3

N.D.

N.D.

N.D.

7

425/2.92 303/4.09

492 524

0.3

0.7

0.46

2.63

1310

8

410/3.02 306/4.05

458 586

0.9

1.1

0.08

2.79

1150

9

472/2.67 304/4.08

531

0.87

1.5

0.13

2.45

2815

10

428/2.90 306/4.05

484 513

0.8

1.1

0.11

2.64

1350

11

480/2.58 394/3.15 347/3.57 303/4.09

601

0.4

2.6

0.15

2.25

2280

12

429/2.89 305/4.07

516

0.7

1.6

0.17

2.57

1970

3

423/4.83 306/4.63

480 507

4

399/4.98 303/4.61

5

Br

F

C8F17

6 C8F17

F

ACS Paragon Plus Environment

12

Page 13 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

13

416/2.98 302/4.11

475 504

0.8

1.1

0.09

2.70

1460

14

431/2.88 302/4.11

556

0.05

0.2

0.02

2.51

1800

15

419/2.96 303/4.09

481 511

0.1

0.4

0.04

2.68

1040

16

454/2.73 306/4.05

534

0.6

1.5

0.06

2.46

4110

-

0.0

-

0.00

-

-

17

a

Except for 4,5 and 6 all data for

419/2.96 303/4.09

em  abs max [nm]/log ε,  max [nm], Φf, ΦΔ and TPA [GM] are taken from Ref. [51]

As discussed above, the product of the TPA cross section and the singlet oxygen quantum yield truly classifies the efficiency of the two-photon singlet oxygen sensitizer. In order to study this property as a function of intramolecular CT a suitable measure of CT needs to be defined. In previous work describing the TPA properties of 7-16 the 1H-chemical shift of the protons on the central ring in the OPV motif was used as such a measure.51 Indeed a correlation between the CT character of the molecule and the TPA cross section was seen and is reproduced in Figure 2 along the with the curve describing the dependence of · ΦΔ on the intramolecular CT. The OPV analog substituted with a NO2-group (17) is not included, as it does not produce singlet oxygen in a measureable amount.

ACS Paragon Plus Environment

13

The Journal of Physical Chemistry

a

b

4500

500

4000 3500

Br

400

3000 300

* 

2500

GM]

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 37

2000

200

1500 100

1000

COMe

500 0 7.2

0 7.4

7.6

7.8

8.0

8.2

8.4

8.6

7.2

7.4

7.6

7.8

8.0

8.2

8.4

8.6

Chemical shift [ppm]

Figure 2: (a) TPA cross section as a function of the intramolecular CT for 7-16 (reproduced from Ref. [51]). (b) Dependence between the intramolecular CT and · ΦΔ. The analogs 7 (Br) and 14 (COMe) seem to be outliers. The solid lines are added to guide the eye.

The molecules 2-6 do not fit directly into the series 7-17 but they do give some insight into the photophysical processes taking place in the OPV motif. For example comparing 2 and 3 one would expect that 3 would have a much higher quantum yield than 2 (see for example 7 and 8). However, the quantum yields are not significantly different. Though there is a decade in difference between the non-radiative rate constants, which is the sum of the rate constants for internal conversion (kic) and intersystem crossing (kisc): For 2 the non-radiative rate constant is ~1.3·109 s-1, whereas for 3 the rate constant is 1.3·1010 s-1. It has previously been demonstrated that the energy level of the triplet state can otherwise lower the singlet oxygen quantum yield

ACS Paragon Plus Environment

14

Page 15 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

despite expectations of a higher quantum yield.51 This will be further elaborated with calculations below. The “configuration” of the “building blocks” i.e. the connectivity was probed on compound 4 and 5, where only a negligible difference in ΦΔ was observed despite differences in fluorescence quantum yields and lifetimes. The non-radiative decay lifetimes amounts to approximately the same value; ~1.4·109 s-1 for 4 and ~1.3·109 s-1 for 5. The bromine (7) and methyl-ketone (14) analogs do not seem to follow the trend of the other molecules in the series 7-16 as observed in Figure 2: The bromine analog has a · ΦΔ value larger than what would be expected from the amount of intramolecular CT present in the molecule, whereas the opposite is seen for the methyl-ketone. However, 7 cannot be directly compared to the rest of the series due to the heavy atom effect, which increases the rate for intersystem crossing because of the mass of the bromine atoms. It is particularly interesting that 14 has a low singlet oxygen quantum yield (0.02) compared to the other molecules. Likewise, the tert-butylketone has a similar low quantum yield (0.04). It can be speculated that the methyl-ketone analog has considerably lower two-photon singlet oxygen production efficiency than expected because of the possible presence of a keto-enol tautomer reaction. Such a reaction can act as an energy sink. Thus a populated Sn state can isomerize to an enol-tautomer with a lower intersystem crossing efficiency than the keto-tautomer. The oxygen lone pair orbitals in the enol tautomer will no longer be in plane. This will most like result in a decreased orbital overlap and eventually cause a lower n->π* transition matrix element. The bromine analog on the other hand has higher twophoton singlet oxygen production efficiency than expected. This can possibly be ascribed to the heavy atom effect. Electronic structure calculations. In Ref. [51] the singlet oxygen quantum yield of 7 and an analog, which includes two carbonyl groups, showed an unexpected order. Additional carbonyl

ACS Paragon Plus Environment

15

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 37

groups did not give an expected higher singlet quantum yield. Instead the singlet oxygen quantum yield was lowered. This finding was ascribed to the differences in triplet energy levels of the two compounds. In order to pursue the triplet energy levels as an explanation for the differences in singlet oxygen quantum yields, singlet and triplet excitation energies were calculated for 7-17 along with the corresponding (singlet) oscillator strengths (f) (CAM-B3LYP/6-31G(d) level). This exchange-correlation functional is especially promising for transitions showing a large degree of charge-transfer; see the discussion below. Furthermore, the basis set 6-31G(d) represents a reasonable balance between computational cost and accuracy. The results of these calculations are presented in Table 2 which refers to calculations performed in vacuum The calculations were validated by also calculating the T1-energy for the analog of 7 containing two carbonyl groups51 obtaining a value of 2.00 eV, which is to be compared to the experimentally determined value of 1.66 eV. For 7 a calculated T1-energy is found to be 1.63 eV and the experimentally determined value is 1.53 eV. Even though there is a difference of 0.34 eV and 0.10 eV, respectively, for the calculated and experimentally determined values the order of the T1-leves are reproduced.

ACS Paragon Plus Environment

16

Page 17 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Table 2: Calculated singlet and triplet excitation energies for compounds 7-17 at the CAMB3LYP/6-31G(d) level. Excitation energies (and differences) are presented in eV. f denotes onephoton oscillator strengths. Excited state

7

8

9

10

11

12

13

14

15

16

17

Br

H

CN

OMe

CHO

CO2Et

CONEt2

COMe

COtBu

SO2Me

NO2

Singlet states (E0,0÷1-1Au)

0.50

0.42

0.48

0.42

0.74

0.48

0.58

0.45

0.59

0.63

-

1-1Au

3.13

3.21

2.93

3.06

2.99

3.05

3.28

2.96

3.27

3.09

2.66

1-1Ag

3.73

3.87

3.54

3.84

3.56

3.68

3.83

3.55

3.80

3.62

3.21

2-1Au

4.34

4.39

3.68

4.07

3.58

3.82

4.40

3.66

3.94

4.15

3.34

2-1Ag

4.40

4.39

4.30

4.38

3.66

4.35

4.41

3.78

4.01

4.41

3.80

3-1Au

4.40

4.50

4.36

3.38

3.67

4.40

4.42

3.72

4.06

4.42

3.78

3-1Ag

4.57

4.54

4.42

4.52

4.10

4.42

4.54

4.21

4.41

4.50

3.84

f(1-1Au)

2.77

2.88

2.62

2.61

1.64

2.21

2.68

1.81

2.60

2.57

1.22

f(2-1Au)

0.01

0.05

0.17

0.17

0.99

0.46

0.04

0.17

0.05

0.08

1.46

f(3-1Au)

0.08

0.01

0.04

0.04

0.09

0.05

0.01

0.65

0.04

0.05

0.02

Triplet states 1-3Au

1.63

1.64

1.51

1.54

1.73

1.67

1.80

1.67

1.81

1.76

1.61

1-3Ag

2.26

2.30

2.21

2.29

2.26

2.26

2.32

2.25

2.30

2.26

2.19

2-3Au

3.03

3.07

2.89

2.99

2.75

2.94

3.05

2.83

3.01

3.01

2.47

2-3Ag

3.16

3.14

3.18

3.13

3.05

3.14

3.14

3.15

3.15

3.17

2.47

3-3Au

3.16

3.14

3.10

3.13

2.94

3.17

3.14

3.10

3.34

3.17

2.60

3-3Ag

3.35

3.36

3.34

3.36

3.35

3.35

3.36

3.22

3.36

3.36

3.18

S1÷T1)

1.50

1.57

1.42

1.52

1.26

1.38

1.48

1.29

1.46

1.33

1.05

S1÷T2)

0.87

0.91

0.72

0.77

0.73

0.79

0.96

0.71

0.97

0.83

0.47

S1÷T3)

0.10

0.14

0.04

0.07

0.24

0.11

0.34

0.13

0.26

0.08

0.19

ACS Paragon Plus Environment

17

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 37

The calculated 1-1Au energies are compared to the experimentally determined E0,0-energies, differences between the two energies are from 0.42 to 0.74 eV. These differences are relatively high and even though the results are quite scattered, there is a tendency for molecules with a large amount of intramolecular CT (e.g. 11 and 16) to have a higher energy difference, than for compounds like 8 and 10, where less intramolecular CT is expected. Several studies where the CAM-B3LYP functional is used have shown a consistent underestimation of excitation energies.61 However, there is a correlation between the E0,0 energies and the excitation energies calculated with the CAM-B3LYP functional in the present study. Previously, a good correlation between measured and calculated optical properties for large aromatic systems has been achieved using the CAM-B3LYP functional29 which is the main motivation for using this level of theory. As described above the amount of CT character in the sensitizer-oxygen complex can have an adverse effect of generating singlet oxygen. The extent of CT in the sensitizer-oxygen complex can be evaluated by the Rehm-Weller equation,62 ECT = F(EMOX ÷ EO2red) + C, where ECT (in eV) is the amount of CT, F is Faraday’s constant, EMOX (in V) is the oxidation potential for the sensitizer (see Table 3), EO2red (in V) is the reduction potential of oxygen and C is a constant that depends on the electrostatic interaction energy which is inversely proportional to the static relative permittivity εr of the solvent and on the differences in solvation energies of the separate ions and the ion pairs. It is assumed that there is little variation of C for the ECT-energies between the sensitizer-oxygen complexes for the sensitizers listed in Table 1. In Figure 3 the singlet oxygen quantum yields are plotted as a function of the oxidation potential of the sensitizer.

ACS Paragon Plus Environment

18

Page 19 of 37

0.5 Br

0.4 0.3 

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

0.2 OMe

0.1 0.0 1.50 1.55 1.60 1.65 1.70 1.75 1.80 1.85 Ox

red

F(EM -EO ) [eV] 2

Figure 3: Singlet oxygen quantum yields as a function of the difference in the oxidation potential for the OPV sensitizer and reduction potentials for oxygen.

The bromine and methoxy analogs are outliers in this series. If these two points are excluded from the analysis, Figure 3 reveals that there seems to be an optimum value for the amount of CT in the sensitizer-oxygen complex. The justification for excluding the methoxy-analog is that the methoxy group is an electron donating group whereas all the other analogs contain electron accepting groups. Furthermore, the bromine analog can be excluded since the mechanism for forming singlet oxygen is a spin-orbit coupling mechanism, which is different from the remaining sensitizers. Thus, by excluding these two compounds we find that there exists an optimum amount of intramolecular CT needed for achieving the highest value of· ΦΔ.

ACS Paragon Plus Environment

19

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 37

Figure 4: Crystal structures of 11 (CHO) (left), 17 (NO2) (middle) and 10 (OMe) determined by X-ray diffraction.

The difference between the methoxy analog and the other compounds is clearly confirmed by crystal structures of 10, 11 and 17. In the methoxy analogue, 10, the configuration around the N atom suggests that the lone pair of the N atom cannot be conjugated with the distilbene moiety. This is different from both the nitro, 11, and aldehyde, 17, analogues where the configuration around the N atom enables the lone pairs to be in conjugation with the distilbene moiety.

Table 3: Cyclic voltammetry data (in V) for the OPV’s (oxidations carried out in CH2Cl2 and reductions in DMF). All potentials are referenced against the Fc/Fc+·-couple. Compound

Reduction [V]

Oxidation [V]

1 (F4)

-2.318

-1.943

0.472

7 (Br)

-2.122

-1.964

0.472

8 (H)

-2.577

-2.333

0.369

9 (CN)

-2.211

-1.715

0.498

10 (OMe)

-2.499

-2.247

0.285

11 (CHO)

-2.073

-1.598

0.447

12 (CO2Et)

-2.340

-1.942

0.406

13 (CONEt2)

-2.466

-2.164

0.402

14 (COMe)

-2.241

-1.863

0.565

ACS Paragon Plus Environment

20

Page 21 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

15 (COtBu)

0.327

16 (SO2Me)

-2.216

-1.765

0.524

17 (NO2)

-1.304

-1.096

0.532

TPA transition probabilities were calculated in different solvents using the response formalism coupled to the polarizable continuum model (PCM) as described in Ref. [63]. Comprehensive descriptions of the relation between calculated and experimentally determined TPA cross sections have been given in the literature.64 In short, the experimental TPA cross section, as discussed above, can be interpreted through the equation,

   

2 2 2 2    g 2    , 15

where  is the fine structure constant, and    is a sum of transition moments available through response calculations (the sum-over-states expression in Eq. 1). The calculated transition moments and excitation energies at the CAM-B3LYP/6-31G(d) level of theory in various solvents for 7-17 are presented in Table 4 for the 1Ag state.

ACS Paragon Plus Environment

21

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 37

Table 4: Excitation energies and transition moments calculated at the CAM-B3LYP/6-31G(d) level for the 1Ag state in various solvents (cyclohexane, toluene, THF, PhCN and DMSO). To be consistent with the notation above, the excitation energies are described as twice the energy needed to populate the state in a two-photon process.

   [10-31 cm4/eV2]

2·ω[eV] vac

C6H12

PhCH3

THF

PhCN

DMSO

Vac

C6H12

PhCH3

THF

PhCN

DMSO

7 (Br)

3.73

3.61

3.60

3.60

3.61

3.59

3.01

3.89

3.97

4.00

3.94

4.09

8 (H)

3.87

3.76

3.74

3.76

3.77

3.71

1.83

2.18

2.21

2.23

2.21

2.27

9 (CN)

3.54

3.42

3.40

3.42

3.43

3.41

5.73

7.93

8.18

8.04

7.77

8.14

10 (OMe)

3.84

3.73

3.71

3.72

3.72

3.70

1.93

2.40

2.45

2.61

2.66

2.71

11 (CHO)

3.56

3.45

3.44

3.43

3.43

3.42

2.55

4.52

4.70

4.80

4.71

4.95

12 (CO2Et)

3.67

3.56

3.54

3.53

3.53

3.51

3.30

4.43

4.57

4.72

4.70

4.92

13 (CONEt2)

3.83

3.72

3.71

3.70

3.70

3.68

1.76

2.22

2.31

2.46

2.45

2.56

14 (COMe)

3.55

3.45

3.44

3.43

3.44

3.41

3.74

5.10

5.26

5.28

5.19

5.44

15 (COtBu)

3.80

3.69

3.68

3.67

3.67

3.65

1.81

2.43

2.51

2.65

2.66

2.77

16 (SO2Me)

3.62

3.50

3.48

3.48

3.49

3.47

3.69

5.16

5.34

5.42

5.26

5.53

17 (NO2)

3.21

3.08

3.06

3.03

3.03

3.01

5.30

7.91

8.23

7.82

7.42

8.02

It is particularly noteworthy that there is hardly any change in excitation energies when going from one solvent to another. This finding is confirmed experimentally in Ref. [65], where UVVIS spectra of 7 (Br), 9 (CN) and 10 (OMe) were recorded in toluene, cyclohexane, THF, acetonitrile and benzonitrile. For example, the experimentally determined energy difference between the S0 and S1 states for 9 (CN) are 2.45 eV (toluene), 2.52 eV (cyclohexane), 2.42 eV (THF) and 2.34 eV (benzonitrile) as found from the crossing of normalized absorption and fluorescence spectra. Absorption maxima are observed at 2.63 eV (toluene), 2.70 eV

ACS Paragon Plus Environment

22

Page 23 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(cyclohexane), 2.68 eV (THF) and 2.63 eV (benzonitrile) showing an even less pronounced solvent effect. Only a few studies of the influence of the solvent on the TPA properties23,65-67 have been carried out, in which solvent effects are discussed. Usually, solvent effects are identified in correlations between the molecular property in question and the dielectric constant of the solvents or various reaction field parameters. However, no apparent correlation is observed between the solvent type (polarity) and TPA transition moments against either the optical (εo), static dielectric constants or reaction field parameters. An observation previously made by inspection of solvent effects in calculated transition moments is the apparent grouping of the values in two categories; one category for solvents with high static dielectric constants and one for solvents with low static dielectric constants (see Ref. 68 for further discussion). In Figure 5, the term  2   in Eq. (1) calculated for 11 (CHO) is plotted as a function of the optical dielectric constant for each of the solvents considered in the present work. As described above, the energy ω is half the energy of the excited state in a two-photon process.

ACS Paragon Plus Environment

23

The Journal of Physical Chemistry

1.44 DMSO

-30

4

cm ]

1.42

2

  [10

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 37

THF

1.40 1.38

Toluene Benzoenitrile

1.36

Cyclohexane

1.34 1.8

1.9

2.0

2.1

2.2

0

Figure 5: Dependence of  2   (calculated for 11 (CHO)) on the optical dielectric constant.

When trying to elucidate solvent effects in experimental data, it is likely that different band shape functions complicate the comparison of TPA cross sections in different solvents. It is outlined in the following how to deconvolve  2   from experimentally determined TPA spectra and these values are then compared to calculated ab initio results.

Well determined TPA cross sections exist for 7 (Br) and 9 (CN) over a broad spectral range as shown in Figure 6 (the values for the TPA cross sections as a function of wavelength were taken from Ref. 28 and 51).

ACS Paragon Plus Environment

24

Page 25 of 37

a

b

3000 2500 2000

 [GM]

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

1500 1000 500 0

1.9

2.0

2.1

2.2

1.3

1.5

1.7

1.8

2.0

 [eV]

Figure 6: TPA spectra in frequency space of 7 (Br) (a) and 9 (CN) (b). Band shape functions consisting of a sum of Gaussian functions (two for 7 (Br) and three for 9 (CN)) fitted to the measurements are shown as solid lines. Dotted lines indicate the regions in which Gaussian functions were fitted.

Band shape functions consisting of a sum of Gaussian functions (two for 7 (Br) and three for 9 (CN)) were fitted to the experimental cross section (Figure 6). These band shape functions were subsequently normalized.69,70 The same analysis was also made for the carbonyl analogue in Ref. 51 (Ac) and the data from this analysis is also presented below. In Table 5, the values of the band shape functions are presented at absorption maximum along with the values for  2   derived by Eq. (1).

ACS Paragon Plus Environment

25

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 37

Table 5: Experimentally and theoretically derived TPA cross sections. The solvent used in the PCM calculations is toluene. Experiment g(2  ) [s-1]

 /eV  ( )

Theory

[cm4·eV-2]

 2  

[cm4]



[eV]

 ( )

[cm4·eV-2]

 2  

[cm4]

7 (Br) 3.47·10-14

2.05

1.28·10-30

5.38·10-30

1.80

0.40·10-30

1.29·10-30

9 (CN) 4.47·10-14

1.47

4.18·10-30

8.99·10-30

1.70

0.82·10-30

2.36·10-30

Ac1

1.47

2.09·10-30

4.50·10-30

1.77

0.10·10-30

0.31·10-30

1

5.52·10-14

Calculations do not include a PCM model. Data from Ref. 51 were used to derive the parameters

presented.

The calculated excitation energies, ω, in Table 5 differ by 0.25 eV (12 %) and 0.30 eV (20 %) from the experimental values. The deviation in the  ( ) values are somewhat larger, however, the relative order is predicted correctly and given the theoretical level in terms of solvent model, the deviation is still impressively small. Many theoretical rationalizations of two-photon properties employs a fixed value for the band shape function (e.g. 2.77·10-14 s-1), which is not completely off when comparing with the values in Table 5. However, there is a significant difference between the two numbers presented and more measurements should be conducted to supplement these values. 2. Changes in the end-groups of the OPV motif. In the OPV motif the central moiety was systematically changed in order to evaluate the effect of the electron-donating effect of the endgroups on the singlet oxygen quantum yields. The singlet oxygen quantum yields along with photophysical parameters are presented in Table 6.

ACS Paragon Plus Environment

26

Page 27 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Table 6:

One-photon optical characterization (absorption, fluorescence and singlet oxygen

quantum yields measurements) of sensitizers with different end-groups on the vinyl-moieties of a 1,4-dibromo-2,5-divinylbenzene core. Br end-group end-group Br

End-group

Φf

 [ns]

0.3

0.7

0.46

0.5

0.6

N. D.

450 477

0.2

0.4

0.26

356/4.69

404 428

0.1

0.3

0.16

362/4.76

412 437

0.2

0.3

0.23

373/4.68

427 453

0.2

0.3

0.23

 abs max [nm]/log ε

 em max [nm]

425/4.80

492

303/4.60

524

N

397/4.50

475

N

304/4.13

504

383/4.67 343/4.50 330/4.43 293/4.58

20

21

7

Ph2N

18

19

22

N

MeO

ACS Paragon Plus Environment

ΦΔ

27

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

23

416/4.72

Me2N

Page 28 of 37

490 519

0.2

0.4

0.39

MeO

24

0.23

MeO MeO

25

O

398/4.60

N

486

0.2

0.5

The most striking result presented in Table 6 is that the strongest electron donor (diphenylamino, 7) also results in the highest singlet oxygen quantum yield. In particular, when comparing the singlet oxygen quantum yields, ΦΔ, for 23 (0.39), 19 (0.26) and 7 (0.46) it becomes apparent that the electron-donating properties of the amino-functional group are beneficial for achieving a high quantum yield. The reason for this is that the lone-pairs on the N atoms are, to a larger extent, delocalized in the pendent fused phenyl ring of the carbazole as opposed to the delocalization into the less conformational restrained phenyl groups of the diphenylamino group. Finally, no delocalization of the N lone-pairs other than into the distyrylbenzene skeleton is possible for 23. It also appears that more energy is lost to vibrational relaxation in for example 20, 21 and 22 in comparison to 23 and 7.

Conclusion A series of phenylene vinylenes have been characterized with quantum mechanical calculations, electrochemical measurements and their ability to produce singlet oxygen in a two-photon absorption scheme. It has been shown that even though charge transfer will result in higher two-

ACS Paragon Plus Environment

28

Page 29 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

photon absorption cross sections it is at the expense of the singlet oxygen quantum yield. Thus, it was demonstrated that in order to tune a molecule for optimum two-photon singlet oxygen sensitization specific charge transfer properties are essential. Quantum mechanical calculations at the DFT level provided insight into the band-shape function for two-photon processes and it was demonstrated that there is a reasonable agreement between experimental and calculated numbers. However, the maximum value for the band-shape function varied by a factor of two among the three molecules investigated. This value is essential for calculation of two-photon absorption cross sections, thus computational studies that do not take this value into account (i.e. uses the same value for a series of molecules) are not very accurate.

Acknowledgement. The authors would like to than Mikkel Jørgensen, Frederik C. Krebs and Peter R. Ogilby for valuable discussions.

Supporting Information Available: Synthetic procedures and characterization of compounds not described previously in Ref. [51]. Crystallographic cif files of compound 10, 11 and 17. This material is available free of charge via the Internet at http://pubs.acs.org.

ACS Paragon Plus Environment

29

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 30 of 37

References (1) Belfield, K. D.; Liu, Y.; Negres, R. A.; Fan, M.; Pan, G.; Hagan, D. J.; Hernandez, F. E. Twophoton Photochromism of an Organic Material for Holographic Recording Chem. Mater. 2002, 14, 3663-3667. (2) Denk, W.; Strickler, J. H.; Webb, W. W. 2-Photon Laser Scanning Fluorescence Microscopy Science 1990, 248, 73-76. (3) Williams, R. M.; Piston, D. W.; Webb, W. W. 2-Photon Molecular-Excitation Provides Intrinsic 3-Dimensional Resolution for Laser-Based Microscopy and Microphotochemistry FASEB J. 1994, 8, 804-813. (4) He, G. S.; Markowicz, P. P.; Lin, T. C.; Prasad, P. N. Observation of Stimulated Emission by Direct Three-Photon Excitation Nature 2002, 415, 767-770. (5) Reinhardt, B. A.; Brott, L. L.; Clarson, S. J.; Dillard, A. G.; Bhatt, J. C.; Kannan, R.; Yuan, L. X.; He, G. S.; Prasad, P. N. Highly Active Two-Photon Dyes: Design, Synthesis, and Characterization Toward Application Chem. Mater. 1998, 10, 1863-1874. (6) Watanabe, T.; Akiyama, M.; Totani, K.; Kuebler, S. M.; Stellacci, F.; Wenseleers, W.; Braun, K.; Marder, S. R.; Perry, J. W. Photoresponsive Hydrogel Microstructure Fabricated by TwoPhoton Initiated Polymerization Adv. Funct. Mater. 2002, 12, 611-614. (7) Belfield, K. D.; Ren, X.; Hagan, D. J.; Van Stryland, E. W.; Dubikovsky, V.; Miesak, E. J. Microfabrication via Two-Photon Photoinitiated Polymerization Abstr. Pap. -Am. Chem. Soc. 1999, 218, U629-U630. (8) Cumpston, B. H.; Ananthavel, S. P.; Barlow, S.; Dyer, D. L.; Ehrlich, J. E.; Erskine, L. L.; Heikal, A. A.; Kuebler, S. M.; Lee, I. Y. S.; McCord-Maughon, D.; et al. Two-Photon Polymerization Initiators for Three-Dimensional Optical Data Storage and Microfabrication Nature 1999, 398, 51-54. (9) He, G. S.; Prasad, P. N. Phase-Conjugation Properties of Two-Photon-Pumped BackwardStimulated Emission. I. Experimental Studies J. Opt. Soc. Am. B 1998, 15, 1078-1085. (10) He, G. S.; Cheng, N.; Prasad, P. N.; Liu, D.; Liu, S. H. Phase-Conjugation Properties of Two-Photon-Pumped Backward-Stimulated Emission. II. Theoretical Studies J. Opt. Soc. Am. B 1998, 15, 1086 1095. (11) Narang, U.; Zhao, C. F.; Bhawalkar, J. D.; Bright, F. V.; Prasad, P. N. Characterization of a New Solvent-Sensitive Two-Photon-Induced Fluorescent (Aminostyryl)pyridinium Salt Dye J. Phys. Chem. 1996, 100, 4521-4525.

ACS Paragon Plus Environment

30

Page 31 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(12) Nielsen, C. B.; Johnsen, M.; Arnbjerg, J.; Pittelkow, M.; McIlroy, S. P.; Ogilby, P. R.; Jørgensen, M. Synthesis and Characterization of Water-Soluble Phenylene-Vinylene-Based Singlet Oxygen Sensitizers for Two-Photon Excitation J. Org. Chem. 2005, 70, 7065–7079. (13) Arnbjerg, J.; Paterson, M. J.; Nielsen, C. B.; Jørgensen, M.; Christiansen, O.; Ogilby, P. R. One- and Two-Photon Photosensitized Singlet Oxygen Production: Characterization of Aromatic Ketones as Sensitizer Standards J. Phys. Chem. A 2007, 111, 5756–5767. (14) Arnbjerg, J.; Jiménez-Banzo, A.; Paterson, M. J.; Nonell, S.; Borrell, J. I.; Christiansen, O.; Ogilby, P. R. Two-Photon Absorption in Tetraphenylporphycenes: Are Porphycenes Better Candidates than Porphyrins for Providing Optimal Optical Properties for Two-Photon Photodynamic Therapy? J. Am. Chem. Soc. 2007, 129, 5188–5199. (15) McIlroy, S. P.; Cló, E.; Nikolajsen, L.; Frederiksen, P. K.; Nielsen, C. B.; Mikkelsen, K. V.; Gothelf, K. V.; Ogilby, P. R. Two-Photon Photosensitized Production of Singlet Oxygen: Sensitizers with Phenylene-Ethynylene-Based Chromophores J. Org. Chem. 2005, 70, 1134– 1146. (16) Oar, M. A.; Serin, J. M.; Dichtel, W. R.; Fréchet, J. M. J.; Ohulchanskyy, T. Y.; Prasad, P. N. Light-Harvesting Chromophores with Metalated Porphyrin Cores for Tuned Photosensitization of Singlet Oxygen via Two-Photon Excited FRET Chem. Mater. 2005, 17, 2267–2275. (17) Drobizhev, M.; Stepanenko, Y.; Dzenis, Y.; Karotki, A.; Rebane, A.; Taylor, P. N.; Anderson, H. L. Extremely Strong Near-IR Two-Photon Absorption in Conjugated Porphyrin Dimers: Quantitative Description with Three-Essential-States Model J. Phys. Chem. B 2005, 109, 7223–7236. (18) Andrasik, S. J.; Belfield, K. D.; Bondar, M. V.; Hernandez, F. E.; Morales, A. R.; Przhonska, O. V.; Yao, S. One- and Two-photon Singlet Oxygen Generation with new Fluorene-Based Photosensitizers ChemPhysChem 2007, 8, 399–404. (19) Belfield, K. D.; Bondar, M. V.; Hernandez, F. E.; Masunov, A. E.; Mikhailov, I. A.; Morales, A. R.; Przhonska, O. V.; Yao, S. Two-Photon Absorption Properties of New FluoreneBased Singlet Oxygen Photosensitizers J. Phys. Chem. C 2009, 113, 4706–4711. (20) Göppert-Mayer, M. Elementary File with Two Quantum Fissures Ann. Phys. 1931, 9, 273294. (21) Aidas, K.; Angeli, C.; Bak, K. L.; Bakken, V.; Bast, R.; Boman, L.; Christiansen, O.; Cimiraglia, R.; Coriani, S.; Dahle, P.; at al. The Dalton Quantum Chemistry Program System Wiley Inter. Rev.: Comput. Mol. Sci. 2014, 4, 269−284. (22) Olsen, J.; Jørgensen, P. Linear and Nonlinear Response Functions for an Exact State and for an MCSCF State J. Chem. Phys. 1985, 82, 3235-3264.

ACS Paragon Plus Environment

31

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 32 of 37

(23) Zhao, K.; Ferrighi, L.; Frediani, L.; Wang, C-K.; Luo, Y. Solvent Effects on Two-Photon Absorption of Dialkylamino Substituted Distyrylbenzene Chromophore J. Chem. Phys. 2007, 126, 204509. (24) Frediani, L.; Rinkevicius, Z.; Ågren, H. Two-Photon Absorption in Solution by Means of Time-Dependent Density-Functional Theory and the Polarizable Continuum Model J. Chem. Phys. 2005, 122, 244104. (25) Sheik-Bahae, M.; Said, A. A.; Wei, T.-H.; Hagan, D. J.; Van Stryland, E. W. Sensitive Measurement of Optical Nonlinearities using a Single Beam IEEE J. Quantum Electron. 1990, 26, 760-769. (26) Xu, C.; Webb, W. W. Nonlinear and Two-Photon-Induced Fluorescence. In Topics in Fluorescence Spectroscopy; Lakowicz, J., Ed.; Plenum Press: New York, 1997; Vol. 5, pp 471 540. (27) Braslavsky, S. E.; Heibel, G. E. Time-Resolved Photothermal and Photoacoustic Methods Applied to Photoinduced Processes in Solution Chem. Rev. 1992, 92, 1381-1410. (28) Arnbjerg, J.; Johnsen, M.; Frederiksen, P. K.; Braslavsky, S. E.; Ogilby, P. R. Two-Photon Photosensitized Production of Singlet Oxygen: Optical and Optoacoustic Characterization of Absolute Two-Photon Absorption Cross Sections for Standard Sensitizers in Different Solvents J. Phys. Chem. A, 2006, 110, 7375-7385. (29) Johnsen, M.; Paterson, M. J.; Arnbjerg, J.; Christiansen, O.; Nielsen, C. B.; Jørgensen, M.; Ogilby, P. R. Effects of Conjugation Length and Resonance Enhancement on Two-Photon Absorption in Phenylene-Vinylene Oligomers Phys. Chem. Chem. Phys., 2008, 10, 1177-1191. (30) Albota, M.; Beljonne, D.; Brédas, J. L.; Ehrlich, J. E.; Fu, J. Y.; Heikal, A. A.; Hess, S. E.; Kogej, T.; Levin, M. D.; Marder, S. R.; et al. Design of Organic Molecules with Large TwoPhoton Absorption Cross Sections Science 1998, 281, 1653-1656. (31) Marder, S.; Perry, J. Two Photon or Higher-Order Absorbing Optical Materials and Methods of Use. [US 6,267,913], 2001. (32) Pond, S. J. K.; Rumi, M.; Levin, M. D.; Parker, T. C.; Beljonne, D.; Day, M. W.; Brédas, J. L.; Marder, S. R.; Perry, J. W. One- and Two-Photon Spectroscopy of Donor-Acceptor-Donor Distyrylbenzene Derivatives: Effect of Cyano Substitution and Distortion from Planarity J. Phys. Chem. A 2002, 106, 11470-11480. (33) Rumi, M.; Ehrlich, J. E.; Heikal, A. A.; Perry, J. W.; Barlow, S.; Hu, Z. Y.; McCord Maughon, D.; Parker, T. C.; Rockel, H.; Thayumanavan, S.; et al. Structure-Property Relationships for Two-Photon Absorbing Chromophores: Bis-Donor Diphenylpolyene and Bis(styryl)benzene Derivatives J. Am. Chem. Soc. 2000, 122, 9500-9510.

ACS Paragon Plus Environment

32

Page 33 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(34) Poulsen, T. D.; Frederiksen, P. K.; Jørgensen, M.; Mikkelsen, K. V.; Ogilby, P. R. TwoPhoton Singlet Oxygen Sensitizers: Quantifying, Modeling, and Optimizing the Two-Photon Absorption Cross Section J. Phys. Chem. A 2001, 105, 11488-11495 (35) Frederiksen, P. K.; Jørgensen, M.; Ogilby, P. R. Two-Photon Photosensitized Production of Singlet Oxygen J. Am. Chem. Soc. 2001, 123, 1215-1221. (36) Strehmel, B.; Sarker, A. M.; Detert, H. The Influence of Sigma and Pi Acceptors on TwoPhoton Absorption and Solvatochromism of Dipolar and Quadrupolar Unsaturated Organic Compounds ChemPhysChem 2003, 4, 249-259. (37) Zojer, E.; Beljonne, D.; Kogej, T.; Vogel, H.; Marder, S. R.; Perry, J. W.; Brédas, J. L. Tuning the Two-Photon Absorption Response of Quadrupolar Organic Molecules J. Chem. Phys. 2002, 116, 3646-3658. (38) Najechalski, P.; Morel, Y.; Stephan, O.; Baldeck, P. L. Two-Photon Absorption Spectrum of Poly(fluorene) Chem. Phys. Lett. 2001, 343, 44-48. (39) Drobizhev, M.; Karotki, A.; Rebane, A.; Spangler, C. W. Dendrimer Molecules with Record Large Two-Photon Absorption Cross Section Opt. Lett. 2001, 26, 1081-1083. (40) Ware, W. R.; Rothman, W. Relative Fluorescence Quantum Yields using an Integrating Sphere - Quantum Yield of 9,10-Diphenylanthracene in Cyclohexane Chem. Phys. Lett. 1976, 39, 449-453. (41) Scurlock, R. D.; Nonell, S.; Braslavsky, S. E.; Ogilby, P. R. Effect of Solvent on the Radiative Decay of Singlet Molecular-Oxygen (A(1)Delta(G)) J. Phys. Chem. 1995, 99, 3521 3526. (42) Schmidt, R.; Tanielian, C.; Dunsbach, R.; Wolff, C. Phenalenone, a Universal Rererence Compound for the Determination of Quantum Yields of Singlet Oxygen O2(1-Delta-G) Sensitization J. Photochem. Photobiol., A 1994, 79, 11-17. (43) Sørensen, H. O.; Larsen, S. Measurement of High Quality Diffraction Data with a Nonius KappaCCD Diffractometer; Finding the Optimal Experimental Parameters. J. Appl. Crystallogr., 2003, 36, 931-939. (44) Duisenberg, A. J. M.; Kroon-Batenburg, L. M. J.; Schreurs, A. M. M. An Intensity Evaluation Method: EVAL-14. J. Appl. Crystallogr., 2003, 36, 220–229.

ACS Paragon Plus Environment

33

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 34 of 37

(45) Otwinowski, Z.; Minor,W. (1997) Macromolecular Crystallography, Part A, edited by C. W. J. Carter & R. M. Sweet, Methods in Enzymology, Vol. 276, pp. 307-326. New York: Academic Press. (46) P. Coppens, in Crystallographic Computing, ed. F. R. Ahmed, S. R. Hall and C. P. Huber, Munksgaard, Copenhagen, 1970, 255–270. (47) Mackay, S.; Gilmore, C. J.; Edwards, C.; Tremayne, M.; Stuart, N.; Shankland, K. maXus: A Computer Program for the Solution and Refinement of Crystal Structures From Diffraction Data; University of Glasgow, Scotland, UK, Nonius BV, Delft, The Netherlands and MacScience Co. Ltd., Yokohama, Japan, 1998. (48) Altomare, A.; Burla, M. C.; Camalli, M.; Cascarano, G.; Giacovazzo, C.; Guagliardi, A.; Moliterni, A. G. G.; Polidori, G.; Spagna, R. SIR97: A New Tool for Crystal Structure Determination and Refinement. J. Appl. Crystallogr., 1999, 32, 115–119. (49) Sheldrick, G. M. A Short History of SHELX. Acta Cryst. A, 2008, 64, 112–122. (50) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Montgomery, J. A., Jr.; Vreven, T.; Kudin, K. N.; Burant, J. C.; et al. Gaussian 03, revision C.02; Gaussian Inc.: Wallingford, CT, 2004. (51) Nielsen, C. B.; Arnbjerg, J.; Johnsen, M.; Jørgensen, M.; Ogilby, P. R. Molecular Tuning of Phenylene-Vinylene Derivatives for Two-Photon Photosensitized Singlet Oxygen Production J. Org. Chem., 2009, 74, 9094–9104. (52) Frederiksen, P. K.; McIlroy, S. P.; Nielsen, C. B.; Nikolajsen, L.; Skovsen, E.; Jørgensen, M.; Mikkelsen, K. V.; Ogilby, P. R. Two-Photon Photosensitized Production of Singlet Oxygen in Water J. Am. Chem. Soc., 2005, 127,255–269. (53) Schweitzer, C.; Schmidt, R. Physical Mechanisms of Generation and Deactivation of Singlet Oxygen Chem. Rev. 2003, 103, 1685–1757. (54) Kristiansen, M.; Scurlock, R. D.; Iu, K.-K.; Ogilby, P. R. Charge-Transfer State and Singlet Oxygen (1-Delta-G O2) Production in Photoexcited Organic-Molecule Molecular-Oxygen Complexes J. Phys. Chem. 1991, 95, 5190–5197.

ACS Paragon Plus Environment

34

Page 35 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(55) McGarvey, D. J.; Wilkinson, F.; Worrall, D. R.; Hobley, J.; Shaikh, W. Picosecond Absorption Studies on the Role of Charge-Transfer Interactions in the Mechanism of Quenching of Triplet-States by Molecular-Oxygen Chem. Phys. Lett. 1993, 202, 528–534. (56) Flors, C.; Ogilby, P. R.; Luis, J. G.; Grillo, T. A.; Izquierdo, L. R.; Gentili, P.-L.; Bussotti, L.; Nonell, S. Phototoxic Phytoalexins. Processes that Compete with the Photosensitized Production of Singlet Oxygen by 9-Phenylphenalenones Photochem. Photobiol. 2006, 82, 95– 103. (57) Arnbjerg, J.; Johnsen, M.; Nielsen, C. B.; Jørgensen, M.; Ogilby, P. R. Effect of Sensitizer Protonation on Singlet Oxygen Production in Aqueous and Nonaqueous media J. Phys. Chem. A 2007, 111, 4573–4583. (58) Jensen, P.-G.; Arnbjerg, J.; Tolbod, L. P.; Toftegaard, R.; Ogilby, P. R. Influence of an Intermolecular Charge-Transfer State on Excited-State Relaxation Dynamics: Solvent Effect on the Methylnaphthalene-Oxygen System and its Significance for Singlet Oxygen Production J. Phys. Chem. A 2009, 113, 9965–9973. (59) Klessinger, M.; Michl, J. Excited States and Photochemistry of Organic Molecules; VCH Publishers, Inc.: New York, 1995. (60) Turro, N. J. Modern Molecular Photochemistry; University Science Books: Sausalito, 1991. (61) Leang, S. S.; Zahariev, F.; Gordon, M. S. Benchmarking the Performance of TimeDependent Density Functional Methods J. Chem. Phys., 2012, 136, 104101. (62) Abdel-Shafi, A. A.; Wilkinson, F. Charge Transfer Effects on the Efficiency of Singlet Oxygen Production Following Oxygen Quenching of Excited Singlet and Triplet States of Aromatic Hydrocarbons in Acetonitrile J. Phys. Chem. A, 2000, 104, 5747-5757. (63) Frediani, L.; Aagren, H.; Ferrighi, L.; Ruud, K. Second-Harmonic Generation of Solvated Molecules using Multiconfigurational Self-Consistent-Field Quadratic Response Theory and the Polarizable Continuum Model J. Chem. Phys., 2005, 123, 144117. (64) He, G. S.; Tan, L.-S.; Zheng, Q.; Prasad, P. N. Multiphoton Absorbing materials: Molecular Designs, Characterizations, and Applications Chem. Rev., 2008, 108, 1245. (65) Johnsen, M.; Ogilby, P. R. Effect of Solvent on Two-Photon Absorption by Vinyl Benzene Derivatives J. Phys. Chem. A, 2008, 112, 7831-7839. (66) Woo, H. Y.; Liu, B.; Kohler, B.; Korystov, D.; Mikhailovsky, A.; Bazan, G. C. Solvent Effects on the Two-Photon Absorption of Distyrylbenzene Chromophores J. Am. Chem. Soc., 2005, 127, 14721-14729.

ACS Paragon Plus Environment

35

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 36 of 37

(67) Leng, W.; Bazan, G. C.; Kelley, A. M. Solvent Effects on Resonance Raman and HyperRaman Scatterings for a Centrosymmetric Distyrylbenzene and Relationship to Two-Photon Absorption J. Chem. Phys., 2009, 130, 044501 (68) Nielsen, C. B.; Sauer S. P. A.; Mikkelsen, K. V. Response Theory in the Multipole Reaction Field Model for Equilibrium and Nonequilibrium Solvation: Exact Theory and the Second Order Polarization Propagator Approximation J. Chem. Phys., 2003, 119, 3849. (69) Birge, R. R.; Bennett, J. A.; Pierce, B. M.; Thomas, T. M. 2-Photon Spesctroscopy of Visual Chromophores - Evidence for a Lowest Excited Ag-1-Like Pi Pi-Star State in All-Trans-Retinol (Vtiamin-A) J. Am. Chem. Soc., 1978, 100, 1533. (70) Birge, R. R.; Pierce, B. M. Theoretical-Analysis of the 2-Photon Properties of Linear Polyenes and the Visual Chromophores J. Chem. Phys., 1979, 70, 165.

ACS Paragon Plus Environment

36

Page 37 of 37

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

269x208mm (300 x 300 DPI)

ACS Paragon Plus Environment