+ All Categories
Home > Documents > Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and...

Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and...

Date post: 26-Jul-2020
Category:
Upload: others
View: 5 times
Download: 0 times
Share this document with a friend
169
Molecular engineering of photoinduced charge separation Citation for published version (APA): Marcos Ramos, A. (2003). Molecular engineering of photoinduced charge separation. Technische Universiteit Eindhoven. https://doi.org/10.6100/IR571230 DOI: 10.6100/IR571230 Document status and date: Published: 01/01/2003 Document Version: Publisher’s PDF, also known as Version of Record (includes final page, issue and volume numbers) Please check the document version of this publication: • A submitted manuscript is the version of the article upon submission and before peer-review. There can be important differences between the submitted version and the official published version of record. People interested in the research are advised to contact the author for the final version of the publication, or visit the DOI to the publisher's website. • The final author version and the galley proof are versions of the publication after peer review. • The final published version features the final layout of the paper including the volume, issue and page numbers. Link to publication General rights Copyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright owners and it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights. • Users may download and print one copy of any publication from the public portal for the purpose of private study or research. • You may not further distribute the material or use it for any profit-making activity or commercial gain • You may freely distribute the URL identifying the publication in the public portal. If the publication is distributed under the terms of Article 25fa of the Dutch Copyright Act, indicated by the “Taverne” license above, please follow below link for the End User Agreement: www.tue.nl/taverne Take down policy If you believe that this document breaches copyright please contact us at: [email protected] providing details and we will investigate your claim. Download date: 28. Sep. 2020
Transcript
Page 1: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Molecular engineering of photoinduced charge separation

Citation for published version (APA):Marcos Ramos, A. (2003). Molecular engineering of photoinduced charge separation. Technische UniversiteitEindhoven. https://doi.org/10.6100/IR571230

DOI:10.6100/IR571230

Document status and date:Published: 01/01/2003

Document Version:Publisher’s PDF, also known as Version of Record (includes final page, issue and volume numbers)

Please check the document version of this publication:

• A submitted manuscript is the version of the article upon submission and before peer-review. There can beimportant differences between the submitted version and the official published version of record. Peopleinterested in the research are advised to contact the author for the final version of the publication, or visit theDOI to the publisher's website.• The final author version and the galley proof are versions of the publication after peer review.• The final published version features the final layout of the paper including the volume, issue and pagenumbers.Link to publication

General rightsCopyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright ownersand it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights.

• Users may download and print one copy of any publication from the public portal for the purpose of private study or research. • You may not further distribute the material or use it for any profit-making activity or commercial gain • You may freely distribute the URL identifying the publication in the public portal.

If the publication is distributed under the terms of Article 25fa of the Dutch Copyright Act, indicated by the “Taverne” license above, pleasefollow below link for the End User Agreement:www.tue.nl/taverne

Take down policyIf you believe that this document breaches copyright please contact us at:[email protected] details and we will investigate your claim.

Download date: 28. Sep. 2020

Page 2: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Molecular Engineering of

Photoinduced Charge Separation

Page 3: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1
Page 4: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Molecular Engineering of

Photoinduced Charge Separation

PROEFSCHRIFT

ter verkrijging van de graad van doctor aan de Technische Universiteit

Eindhoven, op gezag van de Rector Magnificus, prof.dr. R.A. van Santen,

voor een commissie aangewezen door het College voor Promoties in het

openbaar te verdedigen op woensdag 29 oktober 2003 om 16.00 uur

door

Alicia Marcos Ramos

geboren te Reus, Spanje

Page 5: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Dit proefschrift is goedgekeurd door de promotoren:

prof.dr.ir. R.A.J. Janssen

en

prof.dr. J.C. Hummelen

This research has been financially supported by the Dutch Government through the

E.E.T. program (EETK97115), NOVEM (146.120.008.3 and 146.120.003.4), and by the

European Comission (Joule III contract No. JOR3CT9802026).

Omslagontwerp: Isabel Marcos Ramos en Jan-Willem Luiten.

Druk: Universiteitsdrukkerij, Technische Universiteit Eindhoven.

CIP-DATA LIBRARY TECHNISCHE UNIVERSITEIT EINDHOVEN

Marcos Ramos, Alicia Molecular Engineering of Photoinduced Charge Separation / by Alicia Marcos Ramos. – Eindhoven :

Technische Universiteit Eindhoven, 2003.

Proefschrift. – ISBN 90-386-2695-9

NUR 914

Trefwoorden: donor-acceptor systemen / π-geconjugeerde polymeren /supramoleculaire chemie /

electronenoverdracht / energie-overdracht

Subject headings: donor-acceptor systems / π-conjugated polymers /supramolecular chemistry /

electron transfer / energy transfer

Page 6: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

A mis padres

Máximo e Isabel

Page 7: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1
Page 8: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Table of Contents

Chapter 1

Covalently linked donor-acceptor materials 1

1.1 Organic solar cells 2

1.2 Donor-acceptor materials 5

1.2.1 Multichromophoric donor-acceptor materials 5

1.2.2 Donor-acceptor polymers 9

1.2.3 Supramolecular approach 11

1.3 Aim of the thesis 11

1.4 Outline of the thesis 12

1.5 References 13

Chapter 2

Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene

vinylene)-fullerene triad 17

2.1 Introduction 18

2.2 Synthesis 19

2.3 Electronic properties and energetic considerations 21

2.4 Photophysical processes in solution 27

2.4.1 Photoluminescence spectroscopy 27

2.4.2 Near steady state photoinduced absorption (PIA) spectroscopy 29

2.4.3 Femtosecond pump-probe spectroscopy in polar solvents 31

2.4.4 Kinetic considerations 36

2.5 Photophysical processes in the solid state 38

2.5.1 Near steady state PIA spectroscopy 38

2.5.2 Femtosecond pump-probe spectroscopy 39

2.6 Conclusions 40

2.7 Experimental section 42

2.8 References and notes 45

Chapter 3

Photoinduced multistep electron transfer in oligoaniline-oligo(p-phenylene vinylene)-perylene

arrays 51

3.1 Introduction 52

Page 9: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

3.2 Synthesis 53

3.3 Electronic properties and energetic considerations 56

3.4 Photophysical processes in solution 61

3.4.1 Photoluminescence spectroscopy 61

3.4.2 Near steady state photoinduced absorption (PIA) spectroscopy 63

3.4.3 Subpicosecond transient pump-probe spectroscopy 64

3.4.3.1 Pentad 1 65

3.4.3.2 Pentad 2 in THF 69

3.5 Kinetic considerations 69

3.6 Conclusions 71

3.7 Experimental section 72

3.8 References and notes 74

Chapter 4

Supramolecular control over donor-acceptor photoinduced charge separation 79

4.1 Introduction 80

4.2 Synthesis 82

4.3 Conformational states of the bridge 84

4.3.1 Folding in chloroform/heptane mixtures 84

4.3.2 Folding in other solvents 87

4.4 Electronic properties and energetic considerations 88

4.5 Photoinduced energy and electron transfer in different conformational states of the bridge 92

4.5.1 Bridge in a random coil conformation 92

4.5.2 Bridge in an folded conformation 95

4.5.2.1 Folded bridge in heptane/chloroform mixtures 95

4.5.2.2 Folded bridge in other solvents 101

4.5 Conclusions 101

4.6 Experimental section 103

4.7 References 104

Chapter 5

Photoinduced electron transfer of conjugated polymers with pendant fullerenes 107

5.1 Introduction 108

5.2 PPV-PPE polymers with pendant fullerenes 109

5.2.1 Synthesis and characterization 109

5.2.2 Photophysical properties 113

Page 10: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

5.2.3 Photovoltaic device 116

5.3 Low-bandgap π-conjugated polymers with pendant fullerenes 117

5.4 Conclusions 118

5.5 Experimental section 119

5.6 References 122

Chapter 6

Polyacetylenes with pendant donor-acceptor dyads 125

6.1 Introduction 126

6.2 Design 128

6.3 Synthesis and characterization 129

6.3.1 Ethynyl perylene bisimide 129

6.3.2 Polyacetylene with pendant OPV and PERY chromophores 130

6.3.3 Polyacetylene with pendant OPVE-PERY dyads 131

6.3.4 Polyacetylene with pendant OPE-PERY dyads 133

6.4 UV/Visible absorption and circular discroism spectroscopies 135

6.5 Photophysical properties 136

6.5.1 Solid state 136

6.5.2 Chloroform solution 138

6.6 Conclusions and outlook 140

6.7 Experimental section 141

6.8 References and notes 147

Summary

Resumen

Curriculum Vitae

Acknowledgment

Page 11: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1
Page 12: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 1

Covalently linked

donor-acceptor materials

Abstract

Blends of donor and acceptors have successfully been incorporated as the active layer in plastic solar

cells. Improving the performance of such devices implies optimizing the morphology of the bulk-

heterojunction. The covalent linkage between the donor and acceptor materials is the ultimate key to

obtain a well-defined spatial organization. In this chapter a concise overview on covalently linked

donor-acceptor multichromophoric arrays and polymers is given. Well-defined donor-acceptor dyads,

triads and larger arrays give valuable information on the photophysics occurring in plastic solar

cells. The reported donor-acceptor polymers provide an important guideline for design of new

polymers for photovoltaic applications.

Page 13: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 1

2

1.1 Organic solar cells

Photoinduced charge separation is the primary step in photosynthesis. In natural systems light

induced charge separation is achieved through a unique spatial arrangement of the pigments and

elements of the transport chain. Electronic excitations that reach the reaction centers are converted to

chemical energy in the form of charge separation across the photosynthetic membrane. In these

organized arrays electrons flow rapidly (< ms) over distances as great as 20 Å, with little loss of

energy.1,2 The photosynthesis in plants is a source of inspiration for scientists to engineer non-natural

systems that similarly convert light into chemical potential or electrical energy.

The opto-electronic and semiconducting properties of conjugated organic and polymer

materials has been one of the highlights in chemistry and physics in the past decade and has

developed into an important area of academic and applied research. The breakthrough that enabled

this development was the discovery of the conducting behavior of doped polyacetylene.3 Like for all

π-conjugated polymers (Figure 1.1), the structure of polyacetylene is characterized by the alternation

of single and double bonds that leads to extended π-orbitals. As a result, π-conjugated materials

exhibit a strong optical absorption in the visible region. Furthermore, charges created in these

materials are mobile and can be transported along and between the polymer chains. One of the

important advantages of π-conjugated polymers is that their properties can be widely tuned by

chemical modification. Chemists have been able to adjust the optical band gap, valence and

conduction band energies, charge transport characteristics as well as the solubility and structural

properties by modifying the nature of the polymer backbone or by changing the substituent groups.

**n

* *n

*

*n S* *n

polyacetylene poly-p-phenylene poly(p-phenylene vinylene) polythiophene

Figure 1.1. Chemical structures of some π-conjugated polymers.

The absorption of light and transport of charges are the two important elements that, in

principle, allow conjugated polymers to be used in the solar cells. However, in contrast to inorganic

semiconductors, photoexcitation of conjugated polymers does not (or with low efficiency) result in

the formation of free charge carriers that are required for a photovoltaic effect. Instead, excitons are

formed that represent a bound electron-hole pair that will not spontaneously dissociate into free

charge carriers. In order to be able to generate charge carriers in π-conjugated organic molecules and

polymers using photoexcitation, scientists have used the donor-acceptor concept conceived in natural

photosynthesis. By combining an electron donor (p-type) with an acceptor (n-type) material and

Page 14: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Covalently linked donor-acceptor materials

3

utilizing the different electron affinity and ionization potential of these materials, it is possible to

dissociate excitons created in either material.

The first organic p/n-heterojunction devices based on a donor and acceptor bilayer

configuration reached energy conversion efficiencies of about 1%.4 Such photovoltaic device typically

consists of a bilayer of the organic materials sandwiched between two different conducting contacts,

an optically transparent indium tin oxide (ITO) front electrode and a metal (Au, Al, Ca, Mg) back

electrode (Figure 1.2). One of the disadvantages of such donor-acceptor double layer device is the

limited interfacial area. More importantly, the organic and polymer semiconductors often have an

exciton diffusion range that is limited to 10 nm. As a consequence only the excitons created close to

the donor-acceptor interface contribute to the photocurrent, which strongly limits the performance. A

significant improvement in the performance of polymer photovoltaic devices was achieved by

blending donor and acceptor into a bulk- heterojunction.5,6 Heeger et al. used blends of a substituted

p-phenylene vinylene polymer (MEH-PPV) as the electron donor with a soluble fullerene derivative

(PCBM) as the electron acceptor,5 while Friend et al. used two polymers: MEH-PPV as the donor and

a cyano-substituted PPV (CN-PPV) as acceptor.6 In these bulk-heterojunctions donor and acceptor

materials form an interpenetrating network, resulting in a very high interfacial area between the two

materials and a continuous phase to connect the materials to the electrodes. The large interface is

advantageous for an efficient electron transfer, while the bi-continuity is required for efficient charge

transport. The polymeric nature of these materials gives them good mechanical properties and

processing advantages to fabricate solar cells in a commercially attractive manner.7

Figure 1.2. Schematic representation of an organic photovoltaic device.

One of the most successful bulk-heterojunction photovoltaic devices made up to date are

those based on the MDMO-PPV/PCBM system (Figure 1.3) that reached efficiencies up to 2.5%.8

Recently, these efficiencies have improved to above 3% by utilizing polythiophenes as donor material

in combination with PCBM derivatives9 and by using MDMO-PPV with a C70 PCBM derivative as an

acceptor.10 Perylene diimid dyes (PERY) constitute another promising class of acceptor materials. In

Glass

PEDOT:PSSActive layer

Al SMU+-

Illumination

ITOGlass

PEDOT:PSSActive layer

Al SMU+-

Illumination

ITO

Page 15: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 1

4

contrast to the fullerene molecules, the PERY acceptor has a strong absorption in the visible, and

plays an important role in the collection of light. Tang incorporated this acceptor in his double-layer

device4 and Friend et al. have shown that it is possible to use perylene in combination with π-

conjugated polymers in bulk-heterojunction devices by processing from solution.11

C12H25

C12H25

C12H25

C12H25

C12H25

C12H25

NN

O

O

O

O

O

O

OOMe

HBC-PhC12

PERY

+

n

+

MDMO-PPV PCBM

Figure 1.3. Components of the most successful organic solar cells up to date.

Because of their difference in chemical structure, donors and acceptors are commonly not

miscible and tend to macro-phase segregate. The resulting morphology of the donor/acceptor blend is

crucial because it affects several aspects of the photovoltaic performance. High interfacial areas are

beneficial for efficient charge separation and interconnection between domains of equal electron

affinity is essential for the transport of opposite charges to opposite electrodes. These factors need to

be optimized in the existing bulk-heterojunction devices, though remarkable improvements in

morphology have been achieved. In polymer/fullerene blends an increased efficiency has been

achieved by optimizing on the processing conditions.8-10 For perylene-based devices the highest

efficiency has been achieved by the use of components that can preorganize via π-π interactions and

via liquid crystallinity (hexaphenyl-substituted hexabenzocoronene, HBC-PhC12, Figure 1.3).12

The development of the plastic solar cell has revealed that it is possible to artificially convert

visible light into photoinduced charge separation and use this state to generate electrical energy.7 So

far the design of plastic solar cells has not been using the sophisticated supramolecular organization as

found in natural photosynthesis. Implementing design elements of the natural system such as multi-

chromophoric arrays and supramolecular organization into the design of donor/acceptor materials will

help in the understanding the photophysical process and might increase the performance of the solar

cells.

Page 16: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Covalently linked donor-acceptor materials

5

1.2 Donor-acceptor materials

Control over the morphology of donor/acceptor bulk-heterojunctions can be pursued at

different dimensional levels. The smallest dimension of control is the molecular scale. The discipline

of molecular engineering gives the opportunity to generate donor and acceptor systems with an almost

unlimited control over their size, shape, and intermolecular interactions on this molecular scale. By

design, synthesis and photophysical evaluation the processes occurring at the molecular level can be

understood and improved. Systems with the donor and acceptor elements connected, either covalently

or via a supramolecular interaction, represent the ultimate bottom-up approach to gain control over the

morphology of donor-acceptor bulkheterojunctions. As an example, the design and synthesis of

covalently linked donor–acceptor dyads, triads and multichromophoric arrays in general, has helped

to understand the photophysical phenomena occurring in photosynthesis.13 The design principles

derived from this approach can be incorporated into the design of polymers containing both

components in a well-defined arrangement. This should allow for the creation of bicontinuous

networks of donor and acceptor materials. Moreover, by judicious choice and use of supramolecular

interactions, donor-acceptor dyads and triads can be brought to preorganize in the solid state into

well-ordered architectures.

1.2.1 Multichromophoric donor-acceptor arrays

Molecular dyads. Donor-acceptor dyads constitute the most elemental molecularly

engineered multichromophoric arrays with a high interfacial area between donor and acceptor. They

can be used as model systems to understand the intricate photophysical processes occurring in

polymer donor/acceptor devices. For photovoltaic applications the donor-acceptor dyads have to be

designed to undergo a fast electron transfer and generate long-lived charge-separated states.

The progress in synthetic methods for the functionalization of fullerene14 and perylene dye15

electron acceptors has allowed the preparation and subsequent photophysical study of a number of

donor-acceptor materials. A large number of fullerene-porphyrin dyads in a variety of arrangements

(for example dyads 1 to 3 in Figure 1.4) have been synthesized in order to gain control over the

electronic coupling, geometrical overlap, and nature of the intervening spacer in the photophysical

processes between the donor and acceptor.16 The use of perylene in multichromophoric arrays has a

more recent history, and it is mainly found in the pioneering work of Wasielewski17 et al. and Lindsay

et al.18 As in the case of fullerene dyads, mostly porphyrins have been used as donor chromophore in

junction with perylene dyes (for example compounds 4 and 5 in Figure 1.4). Lindsay et al. have

shown that arrays containing perylenes and porphyrins can be designed to select for excited-state

energy or charge transfer by tuning the redox and photophysical properties of the components. In

particular, the use of perylene diimid dyes over perylene monoimid dyes results in a photoinduced

charge separation process that prevails over the energy transfer reaction. These comprehensive studies

Page 17: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 1

6

on fullerene/perylene and porphyrin based donor-acceptor dyads have laid down some of the

important basic rules for the design of novel donor-acceptor multichromophoric arrays with desired

photophysical properties.

Ar

NN

NNAr

Ar

N

O

H

M

O

OO

O

HO

ZnNN

NN

Ph

Ph

O

OO

OOO O

OMeMeO

O

O

NN

NNH

H

Ph

Ph

PhO

OO

O

1 2 3

M

4

NN

NN

Ar

Ar

N

OR

N

OR

O

O

M

N

N

O

O

O

OOAr

ArO

N

N

O

O

O

OOAr

ArO

N

N

O

O

O

O

OAr

ArO

N

N

O

OO

O

OAr

ArO

N

N N

N

5

Figure 1.4. Several examples of donor-acceptor dyads based on porphyrin as donor and fullerene or

perylene as acceptor.

The success of π-conjugated polymers in photovoltaic applications, together with the

development of strategies to synthesize functional well-defined oligomers,19 have motivated in the late

90’s the study of photoinduced charge generation in covalently linked dyads, consisting of conjugated

oligomers and fullerenes20-24 or perylenes.25 This type of donor-acceptor dyad is of high interest

because of the remarkable advantages of the oligomer approach. In contrast to π-conjugated polymers,

the oligomers are monodisperse and well defined. This allows the study of oligomer based donor-

acceptor dyads in an isolated form in solution with the possibility to discern the intramolecular

photophysical processes from the intermolecular ones occurring in the solid state. When of sufficient

length, the oligomers feature the essential electrical and optical properties of the corresponding

polymers. Because of this, oligomer-acceptor dyads are excellent models to understand the

photophysics in polymer/acceptor bulk-heterojunction devices. Another interesting characteristic,

derived from their well-defined nature, is the capability of oligomers to organize into crystalline

domains26 or pack efficiently into highly ordered nano-aggregates,27 properties that are crucial in the

use of self-organization of donor-acceptor materials via π-π interaction.

Page 18: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Covalently linked donor-acceptor materials

7

In the design of oligomer-acceptor dyads, the size and functionalization of the oligomer has to

be carefully considered. By going from polymers to shorter oligomers a reduction of the donor

capabilities occurs, to which the fullerene molecule is especially sensitive. In systems using short

donor oligomers the energy transfer process might compete with and even overrule the photoinduced

charge separation. The importance of oligomer length has been manifested in two homologous series

of OPV-C60 dyads, 6-8 and 9-12 (Figure 1.5), synthesized by Nierengarten et al.20 a,b and Janssen et

al., 20d respectively. For dyads 6-8 no significant charge separation has been observed in solution.

Although devices made with these compounds show photovoltaic effect, indicating that electron

transfer occurs in the solid state, the performance is limited by the competing energy transfer

processes. In contrast, the study of the analogous oligomers 9-12 reveals that the dyads undergo

electron transfer in polar solvents when the oligomer exceeds a critical length (three or more phenyl

rings). In this case the oligomers have been heavily decorated with alkoxy side chains. The side

chains not only ensure solubility of the ensembles but also confer a higher electron donating character

to the oligomers. An important conclusion of this study is that in the solid state the charge separation

is much faster than in solution, and that the photoinduced charges are much longer lived, both

implying that in bulk-heterojunctions made out of dyads, the intermolecular processes prevail over the

intramolecular ones.

N

OO

OO

n

N

C12H25

C12H25On

6 n = 07 n = 28 n = 3

9 n = 010 n = 111 n = 212 n = 3

Figure 1.5. Chemical structures of two analogous sets of OPV-C60 dyads.

Another important class of oligomer based donor-acceptor dyads that have attracted attention

are those based on oligothiophenes (nT) and C60. As in OPV-C60 dyads, the length and chemical

modification of the donor oligomer play an important role in determining the outcome of each

photoexcitation. Janssen et al. have studied a series of symmetrical C60-nT-C60 triads (compounds 13-

15, Figure 1.6).21f,g The longer triads 14 and 15 exhibit photoinduced electron transfer in polar

solution and in the solid state and for the shorter triad 13 the same photophysical process occurs,

however only in a small extend. Investigations on a set of analogous dyads with oligomer lengths

ranging from four to sixteen thiophene rings (compounds 16-19, Figure 1.6) show that charge

Page 19: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 1

8

separation occurs for all dyads, in polar solvents as in the solid state. The combined results of 13-19

reveal that four thiophene rings is the critical minimum size for efficient charge separation to occur in

this class of oligothiophene-C60 systems.21d,e,h,i Dyads 16 to 19 have been incorporated in photovoltaic

cells with a incident photon-to-current efficiency that systematically increases with oligothiophene

length.21k

NN

S S

S n

C12H25n

S NS S

S

H

C6H13

C6H13

13 n = 114 n = 215 n = 3

16 n = 117 n = 218 n = 319 n = 4

Figure 1.6. Chemical structures of two analogous sets of nT-C60 dyads and symmetrical triads.

Triads and larger arrays. Extending the lifetime of the charge-separated state and avoiding

the undesired back electron transfer can be achieved using a multistep electron transfer as occurring in

photosynthesis. For a sequential charge transfer a gradient of potentials within a multichromophoric

array is needed. In that regard, a number of multi-site covalently linked porphyrin-fullerene

combinations (triads, tetrads, and pentads) have been described.28,29 The design of molecular triads

and larger arrays that efficiently accomplish the purpose of multistep electron transfer is far from

trivial because of the many possible (and competitive) photophysical processes in these

multichromophoric arrays. In fact, for the first C60-based reported triad, made up of a carotenoid

polyene, a diaryl-porphyrin and a fullerene (C-P-C60), the overall quantum yield for the formation of

the C•+-P-C60•–

charge-separated state was only 0.14.29a In the mean time, a few examples have been

described that reach efficiencies for the charge separation of unity.29i,o The record for the longest lived

charge-separated state with 380 ms has been reported for a tetrad incorporating sequentially one

ferrocene (Fc), one zinc 5,10,15,20-tetraphenylporphyrin (ZnTPP), one free base porphyrin (TPP) and

a fullerene molecule (Fc-ZnTPP-TPP-C60, compound 20, Figure 1.7).29p Imahori et al. have used Fc-

H2TPP-C60 and Fc-ZnTPP-C60 triads (compounds 21 and 22, Figure 1.7), to create photoactive self-

assembled monolayers on Au(111) surfaces. These systems feature the highest efficiency for

photocurrent generation of monolayer-modified electrodes and show that the application of molecular

triads as such in photovoltaic applications is viable.29h

Page 20: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Covalently linked donor-acceptor materials

9

Fe

NO

NONN

NNN

O NOS

Fe

NO

NONN

NN

NN

NNZn N

O

HH

N

Au(111)M

20

21 M = H2

22 M = Zn

Figure 1.7. Multichromophoric donor-acceptor arrays.

1.2.2 Donor-acceptor polymers

Incorporation of donor and acceptor within one polymer chain using an appropriate structural

design is the most convenient way to obtain bicontinuous networks. In these polymers, donors and

acceptors are enforced to occupy predefined positions by the covalent connection. Two main designs

of donor-acceptor polymers have been employed up to date. One approach consists of making π-

conjugated polymers with pendant acceptors. 30 This idea has come to be known as the ‘double-cable’

approach. In the double-cable polymer the photoinduced generated charges are expected to diffuse

away from each other: the hole by an intrachain diffusion process and the electron by hopping from

acceptor to acceptor. The first double-cables ever made (for example polymer 23, Figure 1.8) met the

requirements for photoinduced charge separation, however they were insoluble and have not been

incorporated in photovoltaic devices. 31-33 The synthesis of processable double-cables has been

achieved by the use of multiple solubilizing side chains in the polymer backbone. The first of such

polymers that has been reported is a hybrid polymer of poly(p-phenylene vinylene) (PPV) and poly(p-

phenylene ethynylene) (PPE) with pendant methanofullerenes, which shows good photophysical

properties and was incorporated as the active layer in photovoltaic device. A more detailed description

of this polymer is given in chapter 5 of this thesis. Two following examples of soluble double-cables

are polymers 24 and 25 (Figure 1.8).34,35 Polymer 24 has been synthesized by random

copolymerization of thiophenes bearing the fullerene with thiophenes destined to ensure solubility,

and polymer 25 by grafting the fullerene molecule to an existing functionalized polymer backbone.

Polymer 24 has been successfully used in a photovoltaic device. Furthermore, the organization of the

polymer chain was not disturbed by the presence of side-bonded fullerenes. Apart from fullerenes,

Page 21: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 1

10

other acceptors such as tetracyanoanthraquinodimethane (TCAQ) have been used in double-cable

polymers.36 Following the same approach a polyfluorene with pendant perylenes has been

synthesized, though in this case the polymer is meant to be used in light emitting diodes (LEDs).37

(OCH2CH2)3ON

SS n

N

N

N

OC5H11

C5H11Ox y n

x = 0.2 y = 0.8

23 24 25

OC8H16N

n mSS

OC2H4OC2H4OCH3

C8H17O

Figure 1.8. Examples of double-cable polymers.

Another macromolecular strategy to gain order within the bulkheterojunctions is the

incorporation of donor and acceptors in blockcopolymers. In these class of polymers self-assembly of

the different components takes place through phase segregation at the nanometer scale. Hadziioannou

et al. have synthesized the first donor-acceptor diblock copolymer (polymer 26, figure 1.9), by

growing a flexible poly(styrene-stat-chloromethylstyrene) (PS) coil from an end-functionalized rigid

PPV block and subsequent functionalization of the flexible PS block with C60.38 In this polymer an

efficient electron transfer was inferred from photoluminescence quenching of the PPV block.

Morphology studies reveal that the film of the polymer exhibits a honeycomb structuring at the

micrometer scale. As the active layer in a device, the donor-acceptor diblock copolymer 26 shows

enhanced photovoltaic response, specially a much higher short-circuit current, relative to a blend of its

constituent polymers.39 Another example of donor-acceptor blockcopymer is that reported by Janssen

et al. consisting of alternating OPV and perylene diimid (PERY) segments connected via saturated

spacers, polymer 27 (Figure 1.9). 40 In this example π-π interactions are added as structuring elements

to the blockcopolymer approach. Photophysical studies of this polymer in solution reveal that the

photoinduced charge separated state is long-lived owing to the long saturated spacer between the

donor and acceptor. However, face-to-face orientations of OPV and PERY segments diminish the

lifetime of the charge-separated state in the film and limit the photovoltaic performance.

Page 22: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Covalently linked donor-acceptor materials

11

OOO

O

O

N

O

N

OO

OO

OO

O

O

n

RORO

OROR

ONO

7n m

H

26

27

Figure 1.9. Examples of donor-acceptor blockcopolymers.

1.2.3 Supramolecular approach

Proteins acquire their three-dimensional architecture (tertiary and quaternary structures) by

folding from an unordered chain using a variety of supramolecular interactions. By analogy, an

optimal morphology of the donor-acceptor materials in the solid state could be achieved through the

wise choice of such supramolecular interactions. These interactions can be used to bring donor and

acceptors in close proximity and to induce the self-organization of donor-acceptor multichromophoric

arrays or donor-acceptor blockcopolymers into highly ordered assemblies. Among the many possible

supramolecular interactions, π-π stacking and hydrogen bonding have already resulted in some

remarkable examples of organized donor-acceptor nano-aggregates. Wasieleswki et al.17 have shown

that a system, made of four perylene diimides connected to a central zinc 5,10,15,20-

tetraphenylporphyrin (ZnTPP) electron donor (compound 5, Figure 1.4), self-assembles into ordered

nanoparticles, both in solution and in the solid state, driven by the π-π stacking of the perylene dyes.

These nanoparticles exhibit photoinduced charge separation and charge transport, which demonstrates

the feasibility of the supramolecular approach. Another interesting example of self-organization is

given by Meijer et al. combining both hydrogen-bonding and π-π interactions to give well-defined

chiral self-assemblies having OPV donors and perylene dyes as acceptor chromophores. 41

1.3 Aim of the thesis

The aim of this thesis is to use molecular engineering to study fundamental issues of

photoinduced charge separation on novel well-defined multichromophoric systems and to develop

polymeric donor-acceptor materials with unprecedented architectures. The study of

multichromophoric donor-acceptor systems and of the role of supramolecular chemistry in directing

the interaction between donors and acceptors will assist in finding the optimal means to control

photoinduced charge separation. The generation of novel donor-acceptor polymers, based on

Page 23: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 1

12

knowledge gained from the single molecule models, may lead the way to processable materials with

optimal properties for photovoltaic devices.

As building blocks for the synthesis of donor-acceptor materials well-defined π-conjugated

oligomers as well as fullerenes and perylenes will be used. The well-defined nature of the oligomers

in combination with the acceptor molecules facilitates the identification of the different photophysical

intermediates. In addition, the planar structure of the oligomers in combination with that of the

perylene acceptor allows for the use of π-π interactions as a structural element that can aid in self-

assembly in the solid state.

1.4 Outline of the thesis

Two donor-donor-acceptor (D(1)-D(2)-A) multichromophoric arrays with a gradient of

potentials have been synthesized by connecting an oligoaniline (OAn), an oligo(p-phenylene

vinylene) (OPV) and a fullerene (C60) (OAn-OPV-C60, chapter 2) or a perylene diimid (PERY), (OAn-

OPV-PERY, chapter 3) in a linear array. In these ensembles photoexcitation of any of the

chromphores generates, after a sequential event that involves multiple energy and charge transfers, the

D(1) •+-D(2)-A•– charge separated state. The overall efficiency for the formation of this D(1) •+-D(2)-

A•– charge-separated state depends on the ratio of rates for forward electron transfer and charge

recombination to generate the ground state. In turn, this ratio can be controlled by changing the

polarity of the solvent. The D(1) •+-D(2)-A•– charge-separated state has a much longer lifetime than

that of the associated D(2) -A dyad, because of the weak electronic coupling between D(1) •+ and A•–.

The results of the photophysical studies on both triads are rationalized in terms of the Marcus theory

for electron transfer.

Chapter 4 deals with the use of supramolecular chemistry to modulate donor and acceptor

interactions. An OPV donor and a PERY acceptor are linked at the opposite ends of a long foldable

cross-conjugated oligomer. The conformation of the bridge between donor and acceptor can be varied

in a controlled way between random coil and helically folded states, by external stimuli such as

solvent polarity. This property of the bridge provides a tool to govern the interaction and, thus, the

photophysical processes occurring between the two redox centers. By going from the random coil

conformation to the collapsed state the distance between the two chromophores is drastically

decreased which allows for electron transfer to occur.

In the last part of the thesis, the structure and photophysical properties of well-defined dyads

are transferred to the polymeric level. In chapter 5, a new double-cable polymer is discussed. Its

synthesis is based on an A-B type Sonogashira copolymerization to ensure the alternation of both

monomers and to provide the opportunity to synthesize a well-defined donor-acceptor polymer. One

monomer bears the fullerene acceptor, the other monomer is a bifunctional OPV consisting of three

Page 24: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Covalently linked donor-acceptor materials

13

phenyl units, which confers donor capabilities to the polymer backbone. The photoinduced electron

transfer is studied by means of photoluminescence and photoinduced absorption spectroscopies and

the photovoltaic properties of the polymer have been tested. In chapter 6, the design and synthesis of a

new donor-acceptor polymer is explored: a π-conjugated polyacetylene with pendant donor-acceptor

dyads. The dyads consist of an OPV and a PERY connected via a saturated spacer. In this new design

the covalent and supramolecular approaches are combined in order to guarantee donor-donor and

acceptor-acceptor selective interaction patterns.

1.5 References

1 (a) The Photosynthetic Reaction Center, Deisenhofer, J., Norris, J. R., Eds.; Academic Press: New York,

1993. (b) Anoxygenic Photosynthetic Bacteria, Blankenschip, R. E., Madigan, M. T., Bauer, C. E., Eds.;

Kluwer Academic Publishers: Dordrecht, 1995.

2 Electron Transfer in Chemistry Vol. I-IV, Balzani, V., Wiley-VCH, Weinheim, 2001.

3 Shirakawa, H.; Louis, E. J.; MacDiarmid, A. G.; Chiang, C. K.; Heeger, A. J. J. Chem. Soc., Chem.

Comm. 1977, 578. b) Chiang, C. K.; Fincher, C. R.; Park, Y. W.; Heeger, A. J. Shirakawa, H.; Louis, E. J.

Gau, S. C.; MacDiarmid, A. G. Phys. Rev. Lett. 1977, 39, 1098.

4 Tang, C. W. Appl. Phys. Lett. 1986, 48, 183.

5 Yu, G.; Gao, J.; Hummelen, J. C.; Wudl, F.; Heeger, A.J Science, 1995, 270, 1789.

6 Halls, J. J. M.; Walsh, C. A.; Greenham, N. C.; Marseglia, E. A.; Friend, R. H.; Maratti, S. C.; Holmes,

A. B. Nature 1995, 376, 498.

7 a) Heeger, A. J. J. Phys. Chem. B 2001, 105, 8475. b) Brabec, C. J.; Sariciftci, N. S.; Hummelen, J. C.

Adv. Funct. Mater. 2001, 11, 15. c) Nelson, Current Opinion in Solid State & Materials Science 2002, 6,

87. d) Nunzi, J.-M. C. R. Physique 2002, 3, 523.

8 Shaheen, S. E.; Brabec, C. J.; Sariciftci, N. S.; Padinger, F.; Fromherz, T.; Hummelen, J. C. Appl. Phys.

Lett. 2001, 78, 841.

9 a) Schilinsky, P.; Waldauf, C.; Brabec, C. J. Appl. Phys. Lett. 2002, 81, 3885. b) Padinger, F.; Rittberger,

R. S.; Sariciftci, N.S. Adv. Funct. Mater. 2003, 13, 85.

10 Wienk, M. M.; Kroon, J. M.; Verhees, W. J. H.; Knol, J.: Hummelen, J. C.; van Hal, P. A.; Janssen, R. A.

J. Angew. Chem. Int. Ed. 2003¸ 42, 3371

11 Dittmer, J. J.; Marseglia, E. A.; Friend, R. H. Adv. Mater. 2000, 12, 1270.

12 Schmidt-Mende L.; Frechtenkötter, A.; Müllen, K.; Moons, E.; Friend, R. H.; MacKenzie, J. D. Science

2001, 293, 1119.

13 Wasielewski, M. R. Chem. Rev. 1992, 92, 435-461.

14 (a) Fullerenes: Chemistry, Physics, and Technology, Kadish, K. M.; Ruoff, R. S., Eds.; J. Wiley, New

York, 2000. (b) Fullerenes: From Synthesis to Optoelectronic Properties, Guldi, D. M.; Martín, N., Eds.

Kluwer Academic Publishers, Dordrech, 2002.

15 Langhals, H. Heterocycles 1995, 40, 477.

Page 25: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 1

14

16 For reviews see: (a) Imahori, H.; Sakata, Y. Adv. Mater. 1997, 119, 5744. (b) Imahori, H.; Sakata, Y. Eur.

J. Org. Chem. 1999, 2445. (a) Martín, N.; Sánchez, L.; Llescas, B.; Pérez, I. Chem. Rev. 1998, 98, 2527.

(d) Guldi, D. M. Chem. Comm. 2000, 321. (e) Guldi, D. M. Chem. Soc. Rev. 2002, 31, 22.

17 For example: Van der Boom, T.; Hayes, R. T.; Zhao, Y.; Bushard, P. J.; Weiss, E. A.; Wasielewski, M.

R. J. Am. Chem. Soc. 2002, 124, 9582.

18 For example: (a) Kirmaier, C.; Hindin, E.; Schwartz, J. K.; Sazanovich, I. V.; Diers, J. R.;

Muthukumaran, K.; Taniguchi, M.; Bocian, D. F.; Lindsey, J. S.; Holten, D. J. Phys. Chem. B 2003, 107,

3443. (b) Uthukumaran, K.; Loewe, R. S.; Kirmaier, C.; Hinidn, E.; Schwartz, J. K, Sazanovich, I. V.;

Diers, J. R.; Bocian, D. F.; Holten, D.; Lindsey, J. S. J. Phys. Chem. B 2003, 107, 3431.

19 (a) Tour, J. M. Chem. Rev. 1996, 96, 537. (b) Martin, R. E.; Diederich, F. Angew. Chem. Int. Ed. 1999,

38, 1350. (b) Segura, J. L.; Martín, N. J. Mater. Chem. 2000, 10, 2403.

20 Oligo(p-phenylene vinylene)-fullerene dyads: (a) Nierengarten, J. F.; Eckert, J. F.; Nicoud, J. F.; Ouali,

L.; Krasnikov, V.; Hadziioannou, G. Chem. Commun. 1999, 617. (b) Eckert, J. F.; Nicoud, J. F.;

Nierengarten, J. F.; Liu, S. G.; Echegoyen, L.; Barigelletti, F.; Armaroli, N.; Ouali, L.; Krasnikov, V. V.;

Hadziioannnou, G. J. Am. Chem. Soc. 2000, 122, 7467. (c) Armaroli, N.; Barigelletti, F.; Ceroni, P.;

Eckert, J.-F.; Nicoud, J.-F.; Nierengarten, J.-F. Chem. Commun. 2000, 599. (d) Peeters, E.; van Hal, P.

A.; Knol, J.; Brabec, C. J.; Sariciftci, N. S.; Hummelen, J. C.; Janssen, R. A. J. J. Phys. Chem. B 2000,

104, 10174. (e) Segura, J. L.; Gómez, R.; Martín, N.; Luo, C. P.; Swartz, A.; Guldi, D. M. Chem.

Commun. 2001, 707. (f) van Hal, P. A.; Janssen, R. A. J.; Lanzani, G.; Cerullo, G.; Zavelani-Rossi, M.;

De Silvestri, S. Phys. Rev. B 2001, 64, 075206. (g) Nierengarten, J.-F.; Armaroli, N.; Accorsi, G.; Rio, Y;

Eckert, J.-F. Chem. Eur. J. 2003, 9, 37.

21 Oligothiophene-fullerene dyads: (a) Benincori, T.; Brenna, E.; Sannicolo, F.; Trimarco, L.; Zotti, G.;

Sozzani, P. Angew. Chem., Int. Ed. Engl. 1996, 35, 648. (b) Effenberger; F.; Grube ,G. Synthesis 1998,

1372. (c) Knorr, S.; Grupp, A.; Mehring, M.; Grube, G.; Effenberger, F. J. Chem. Phys. 1999, 110, 3502.

(d) Yamashiro, T.; Aso, Y.; Otsubo, T.; Tang, H.; Harima, Y.; Yamashita, K. Chem. Lett. 1999, 443. (e)

Fujitsuka, M.; Ito, O.; Yamashiro, T.; Aso, Y.; Otsubo, Y. J. Phys. Chem. A 2000, 104, 4876. (f) van Hal,

P. A.; Knol, J.; Langeveld-Voss, B. M. W.; Meskers, S. C. J.; Hummelen, J. C.; Janssen, R. A. J. J. Phys.

Chem. A 2000, 104, 5974. (g) van Hal, P. A.; Janssen, R. A. J.; Lanzani, G.; Cerullo, G.; Zavelani-Rossi,

M.; De Silvestri, S. Chem. Phys. Lett. 2001, 345, 33. (h) Fujitsuka, M.; Masahura, A.; Kasai, H.; Oikawa,

H.; Nakanishi, H.; Ito, O.; Yamashiro, T.; Aso, Y.; Otsubo, T. J. Phys. Chem. B. 2001, 105, 9930. (i)

Fujitsuka, M.; Matsumoto, K.; Ito, O.; Yamashiro, T.; Aso, Y.; Otsubo, T. Res. Chem. Intermed. 2001,

27, 73. (j) Hirayama, D.; Takimiya, K.; Aso, Y.; Otsubo, T. Hasobe, T.; Yamada, H. Imahori, H.;

Fukuzumi, S.; Sakata, Y. J. Am. Chem. Soc. 2002, 124, 532. (k) Negishi, N.; Yamada, K.; Takimiya, K.;

Aso, Y.; Otsubo, T.; Harima, Y. Chem. Lett. 2003, 32, 404-405.

22 Oligo(thienylene vinylene)-fullerene dyads: (a) Liu, S.-G.; Shu, L.; Rivera, J.; Liu, H.; Raimundo, J.-M.;

Roncali, J.; Gorgues, A.; Echegoyen, L. J. Org. Chem. 1999, 64, 4884. (b) Liu, S.-G.; Martineau, C.;

Raimundo, J.-M.; Roncali, J.; Echegoyen, L. Chem. Commun. 2001, 913. (c) Martineau, C.; Blanchard,

Page 26: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Covalently linked donor-acceptor materials

15

P.; Rondeau, D.; Delaunay, J.; Roncali, J. Adv. Mater. 2002, 14, 283. (d) Apperloo, J. J.; Martineau, C.;

van Hal, P. A.; Roncali, J.; Janssen, R. A. J. J. Phys. Chem. A 2002, 106, 21.

23 Oligoene-fullerene dyads: (a) Imahori, H.; Cardoso, S.; Tatman, D.; Lin, S.; Noss, L.; Seely, G. R.;

Sereno, L.; Chessa de Silber, J.; Moore, T. A.; Moore, A. L.; Gust, D. Photochem. Photobiol. 1995, 62,

1009. (b) Yamazaki, M.; Araki, Y.; Fuijtsuha, M.; Ito, O. J. Phys. Chem. A. 2001, 105, 8615.

24 Miscellaneous: oligomer-fullerene dyads: (a) Segura, J. L.; Gómez, R.; Martín, N.; Luo, C.; Guldi, D. M.

Chem. Commun. 2000, 701. (b) Guldi, D. M.; Swartz, Luo, C.; Gómez, R.; Segura, J. L.; Martín, N. J.

Am. Chem. Soc. 2002, 124, 10875. (c) Guldi, D. M.; Luo, C.; Schwartz, A.; Gómez, R.; Segura, J. L.;

Martin, N.; Brabec, C. J.; Sariciftci, N. S. J. Org. Chem. 2002, 67, 1141. (d) Gu, T.; Tsamouras, D.;

Melzer, C.; Krasnikov, V.; Gisselbrecht, J.-P.; Gross, M.; Hadziioannou, G.; Nierengarten, J.-F. Chem.

Phys. Chem. 2002, 124;

25 (a) Peeters, E.; van Hal, P. A.; Meskers, S. C. J.; Janssen, R. A. J.; Meijer, E. W. Chem. Eur. J. 2002,

8(19), 4470. (b) Asha S., Schenning, A. P. H. J.; Meijer, E. W.,Chem. Eur. J. 2002, 8, 15, 3353-61.

26 Gill, R. E.; Meetsma, A.; Hadziioannou, G. Adv. Mater. 1996, 8, 212.

27 (a) Peeters, E.; Marcos Ramos, A.; Meskers, S. C. J.; Janssen, R. A. J. J. Chem. Phys. 2000, 112, 9445.

(b) Schenning, A. P. H. J.; Kilbinger, A. F. M.; Biscarini, F.; Cavallini, M.; Cooper, H. J.; Derrick, P. J.;

Feast, W. J.; Lazzaroni, R.; Leclere, Ph.; McDonell, L. A.; Meijer, E. W.; Meskers, S. C. J. J. Am. Chem.

Soc. 2002, 124, 1269.

28 For recent reviews see: (a) Gust, D.; Moore, T. A.; Moore, A. L. Acc. Chem. Res. 2001, 34, 40. (b) Guldi,

D. M. Chem. Soc. Rev. 2002, 31, 22.

29 (a) Liddell, P. A.; Kuciauskas, D.; Sumida, J. P.; Nash, B.; Nguyen, D.; Moore, A. L.; Moore, T. A.;

Gust, D. J. Amer. Chem. Soc. 1997, 119, 1400. (b) Imahori, H.; Yamada, K.; Hasegawa, M.; Taniguchi,

S.; Okada, T.; Sakata, Y. Angew. Chem., Int. Ed. Engl. 1997, 36, 2626. (c) Carbonera, D.; Di Valentin,

M.; Corvaja, C.; Agostini, G.; Giacometti, G.; Liddell, P. A.; Kuciauskas, D.; Moore, A. L.; Moore, T.

A.; Gust, D. J. Amer. Chem. Soc. 1998, 120, 4398. (d) Kuciauskas, D.; Liddell, P. A.; Moore, A. L.;

Moore, T. A.; Gust, D. J. Amer. Chem. Soc. 1998, 120, 10880. (e) Tamaki, K.; Imahori, H.; Sakata, Y.;

Nishimura, Y.; Yamazaki, I. Chem. Commun. 1999, 625. (f) Imahori, H.; Yamada, H.; Ozawa, S.; Sakata,

Y.; Ushida, K. Chem. Commun. 1999, 1165. (g) Fujitsuka, M.; Ito, O.; Imahori, H.; Yamada, K.;

Yamada, H.; Sakata, Y. Chem. Lett. 1999, 721. (h) Imahori, H.; Yamada, H.; Nishimura, Y.; Yamazaki,

I.; Sakata, Y. J. Phys. Chem. B 2000, 104, 2099. (i) Kuciauskas, D.; Liddell, P. A.; Lin, S.; Stone, S. G.;

Moore, A. L.; Moore, T. A.; Gust, D. J. Phys. Chem. B 2000, 104, 4307. (j) Luo, C.; Guldi, D. M.;

Imahori, H.; Tamaki, K.; Sakata, Y J. Amer. Chem. Soc. 2000, 122, 6535. (k) Imahori, H.; Tamaki, K.;

Yamada, H.; Yamada, K.; Sakata, Y.; Nishimura, Y.; Yamazaki, I.; Fujitsuka, M.; Ito, O. Carbon 2000,

38, 1599. (l) Bahr, J. L.; Kuciauskas, D.; Liddell, P. A.; Moore, A. L.; Moore, T. A.; Gust, D. Photochem.

Photobiol. 2000, 72, 598. (m) Imahori, H.; Norieda, H.; Yamada, H.; Nishimura, Y.; Yamazaki, I.;

Sakata, Y.; Fukuzumi, S. J. Amer. Chem. Soc. 2001, 123, 100. (n) Fukuzumi, S.; Imahori, H.; Yamada,

H.; El-Khouly, M. E.; Fujitsuka, M.; Ito, O.; Guldi, D. M. J. Amer. Chem. Soc. 2001, 123, 2571. (o)

Imahori, H.; Tamaki, K.; Guldi, D. M.; Luo, C.; Fujitsuka, M.; Ito, O.; Sakata, Y.; Fukuzumi, S. J. Amer.

Page 27: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 1

16

Chem. Soc. 2001, 123, 2607. (p) Imahori, H.; Guldi, D. M.; Tamaki, K.; Yoshida, Y.; Luo, C.; Sakata, Y.;

Fukuzumi, S. J. Amer. Chem. Soc. 2001, 123, 6617. (q) Fukuzumi, S.; Imahori, H.; Okamoto, K.;

Yamada, H.; Fujitsuka, M.; Ito, O.; Guldi, D. M. J. Phys. Chem. A 2002, 106, 1903. (r) Liddell, P. A.;

Kodis, G.; De la Garza, L.; Bahr, J. L.; Moore, A. L.; Moore, T. A.; Gust, D. Helv. Chim. Acta 2001, 84,

2765. (s) Ikemoto, J.; Takimiya, K.; Aso, Y.; Otsubo, T.; Fujitsuka, M.; Ito, O. Org. Lett. 2002, 4, 309. (t)

Imahori, H.; Tamaki, K.; Araki, Y.; Hasobe, T.; Ito, O.; Shimomura, A.; Kundu, S.; Okada, T.; Sakata,

Y.; Fukuzumi, S. J. Phys. Chem. A 2002, 106, 2803. (u) Imahori, H.; Tamaki, K.; Araki, Y.; Sekiguchi,

Y.; Ito, O.; Sakata, Y.; Fukuzumi, S. J. Amer. Chem. Soc. 2002, 124, 5165. (v) D'Souza, F.; Deviprasad,

G. R.; Zandler, M. E.; El-Khouly, M. E.; Fujitsuka, M.; Ito, O. J. Phys. Chem. B 2002, 106, 4952. (w)

Liddell, P. A.; Kodis, G.; Moore, A. L.; Moore, T. A.; Gust, D. J. Amer. Chem. Soc. 2002, 124, 7668. (x)

Kodis, G.; Liddell, P. A.; de la Garza, L.; Moore, A. L.; Moore, T. A.; Gust, D. J. Mater. Chem. 2002, 12,

2100. (y) Sánchez, L.; Pérez, I.; Martín, N.; Guldi, D. M. Chem. Eur. J. 2003, 9, 2457.

30 Cravino, A.; Sariciftci, N. S. J. Mater. Chem. 2002, 12, 1931.

31 Benincori, T.; Brenna, E.; Sannicoló, F.; Trimarco, L.; Zotti, G. Angew. Chem. 1996, 108, 718.

32 Ferraris, J. P.; Yassar, A.; Loveday, D.; Hmyene, M. Opt. Mat. 1998, 9, 34.

33 (a) Cravino, A.; Zerza, G.; Maggini, M.; Bucella, D.; Svensson, M.; Andersson, M. R.; Neugebauer, H.;

Sariciftci, N. S. Chem. Commun. 2000, 2487. (b) Cravino, A.; Zerza, G.; Neugebauer, H.; Maggini, M.;

Bucella, S.; Menna, E.; Svensson, M.; Andersson, M. R.; Brabec, C. J.; Sariciftci, N. S. J. Phys. Chem. B

2002, 106, 70.

34 Zhang, F.; Svensson, M.; Andersson, M. R.; Maggini, M.; Bucella, S.; Menna, E.; Inaganäs, O. Adv.

Mater. 2001, 13, 1871.

35 Xiao, S.; Wang, S.; Fang, H.; Li, Y.; Shi, Z.; Du, C.; Zhu, D. Macromol. Rapid. Commun. 2001, 22,

1313.

36 (a) Zerza, G.; Cravino, A.; Neugebauer, H.; Sariciftci, N. S.; Gómez, R.; Segura, J. L.; Martín, N.;

Svensson, M.; Anderson, M. R. J. Phys. Chem. A 2001, 105, 4172. (b) Giacalone, F.; Segura, J. L.;

Martín, N.; Catellani, M.; Luzzati, S.; Lupsac, N. Org. Lett. 2003, 5, 1669.

37 Ego, C.; Marsitzky, D.; Becker, S.; Zhang, J.; Crimsdale, A.; Müllen, K.; MacKenzie, J. D.; Silva, C.;

Friend, R. H. J. Am. Chem. Soc. 2003, 125, 437.

38 Stalmach, U.; de Boer, B.; Videlot, C.; van Hutten P. F.; Melzer, C.; Krasnikov V. V.; Hadziioannou, G.

J. Am. Chem. Soc. 2000, 122, 5464.

39 de Boer, B.; Stalmach, U.; van Hutten P. F.; Melzer, C.; Krasnikov V. V.; Hadziioannou, G. Polym..

2001, 42, 9097.

40 Neuteboom, E. E.; Meskers, S. C. J.; van Hal, P. A.; van Duren, J. K. J.; Meijer, E. W.; Dupin, H.;

Pourtois, G.; Cornil, J.; Lazzaroni, R.; Brédas, J.-L.; Beljonne, D.; Janssen, R. A. J. J. Am. Chem. Soc.,

2003, 125, 8625.

41 Schenning, A. P. H. J.; van Herrikhuyzen, J.; Jonkheijm, P.; Chen, Z.; Würthner, F.; Meijer, E. W. J. Am.

Chem. Soc. 2002, 124, 10252.

Page 28: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 2

Photoinduced multistep energy and

electron transfer in an oligoaniline –

oligo(p-phenylene vinylene) – fullerene

triad

Abstract

A donor-donor-acceptor triad, OAn-OPV-C60, with a redox gradient has been synthesized by

covalently linking an oligoaniline (OAn), an oligo(p-phenylene vinylene) (OPV), and a fullerene (C60)

in a non-conjugated linear array. Photoluminescence and femtosecond pump-probe spectroscopy

studies reveal that photoexcitation of any of the three chromophores of this triad in a polar solvent

results in formation the OAn-OPV•+-C60 •– charge-separated state, subsequent to an efficient ultrafast

(< 190 fs) singlet-energy transfer to the fullerene singlet-excited state. The initial OAn-OPV•+-C60•–

state can rearrange to the low-energy OAn•+-OPV-C60•– charge-separated state via an intramolecular

redox reaction. Because the competing charge recombination of the OAn-OPV•+-C60•– state to the

ground state is fast (≤ 65 ps) and increases with increasing polarity of the solvent, the quantum yield

for this charge shift is the highest (~0.4) in weakly polar solvents such as chlorobenzene. Once

formed, the OAn•+-OPV-C60•– state has a long lifetime (> 1 ns) due to weak electronic coupling

between the distant redox sites in the excited state. The stabilization gained is more than one order of

magnitude. The experimental results are found to be in qualitative agreement with Marcus theory. In

thin films, the OAn•+-OPV-C60•– state is formed at a higher rate and in higher quantum yield than in

solution.

Page 29: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 2

18

2.1 Introduction

Photoinduced energy and electron transfer reactions are the key steps in natural

photosynthesis and the elucidation of their mechanism continues to attract considerable interest.1

Similar processes occur in artificial photoactive and redoxactive molecular donor entities linked to

acceptors. These systems are considered as promising for the application in molecular and

supramolecular electronics, light harvesting, and photocatalysis.2 Molecular donor-acceptor

combinations also find application in organic and polymer photovoltaic cells to convert sunlight into

electrical energy.3,4 One of the potentially promising materials for photovoltaic cells is a blend of a

conjugated polymer as a donor and a methanofullerene derivative as acceptor.5 In these polymer/C60

bulkheterojunctions, the forward electron transfer is extremely fast (<< 1 ps),6 while the electron

recombination extends to the millisecond time domain.7 The large difference between the forward and

backward transfer rates ensures efficient charge generation and gives the opportunity to transport and

collect the photogenerated charges at the electrodes, and both processes have been studied extensively

in recent years.

In this context there is a considerable interest in the study of photoinduced charge generation

in covalently linked dyads, consisting of linear conjugated oligomers and fullerenes.8-12 Compared to

the polymer/C60 blends, these covalent molecular dyads are much more well-defined and allow to

study the charge transfer reactions in different media. Using this approach, it has been shown that

photoexcitation of the oligo(p-phenylene vinylene) unit of an OPV-C60 dyad (Figure 1) in a polar

organic solvent results in an ultrafast (< 190 fs) singlet-energy transfer (1OPV*-C60 → OPV-1C60*),

followed by a much slower (~10 ps) electron transfer reaction (OPV-1C60*→ OPV•+-C60•–) that

produces a charge-separated state with a lifetime of 50 ps.9e More recent studies revealed that the rate

of the forward electron transfer strongly depends on the relative orientation of the two moieties.13 For

an end-to-end substitution of donor and acceptor the electron transfer is much slower (~10 ps) than for

a face-to-face orientation (<< 1 ps), suggesting that the latter configuration explains the fast forward

reaction observed in polymer/C60 blends. With respect to charge recombination, however, there

remains a substantial discrepancy between the lifetimes of oligomer-C60 dyads in solution, which are

typically less than 100 ps, and the long-lived charges in polymer/C60 films. Nature solved the problem

of fast charge recombination in photosynthesis by creating a multi-step electron transfer to increase

the distance between the charges and slow down recombination.1 In mimicking natural

photosynthesis, a number of multi-site covalently linked porphyrin-fullerene combinations (triads,

tetrads, and pentads) have been described,14,15 showing that also in artificial fullerene systems multi-

step charge transfer results in an increase in lifetime of the charge-separated state. Hence, inspired by

nature, a tentative explanation for the long lifetime in the polymer/C60 films is the diffusion of charges

to different sites in the blend.

Page 30: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced multistep energy and electron transfer in a molecular triad

19

In this contribution, the complementary views that originate from mimicking natural

photosynthesis and polymer/C60 solar cells are connected by extending the OPV-C60 dyad to include a

p-oligoaniline (OAn) moiety as an additional donor, to create a donor-donor-acceptor triad (OAn-

OPV-C60, Figure 2.1). By using a meta-substituted phenylene ring in OAn-OPV-C60, the OAn and

OPV parts are electronically decoupled in the ground state and operate as isolated redox active

segments. In OAn-OPV-C60, the oxidation potential decreases from C60, via OPV, to OAn, while at

the same time the reduction potential increases (vide infra). By introducing this redox gradient it is

expected that the energetically most favorable charge-separated state corresponds to OAn•+-OPV-

C60•– and that the lifetime of this state is enhanced as a result of the larger distance between the centers

of positive and negative charge density. The detailed analysis of the photophysical processes in OAn-

OPV-C60 has been performed using photoluminescence and femtosecond pump-probe spectroscopy in

solvents of different polarity and in the solid state, and by comparing the results with those of the

model compounds OAn-OPV, OPV-C60, OPV, OAn, and MP-C60 (Figure 2.1) that have only one or

two chromophores.

N N

NO

OO

O

OO

NO

OO

O

OO

O

O

N N

OO

O

O

OO

D

N N

OO

O

O

OO

O

O

N

OAn-OPV-C60

OPV-C60

OPV

OAn-OPVMP-C60

OAn

Figure 2.1. Structure of OAn-OPV-C60 triad and reference compounds.

2.2 Synthesis

The synthesis of OAn,16 OPV, MP-C60, and OPV-C609d have been described elsewhere. The

synthesis of the triad OAn-OPV-C60 (7, Scheme 2.1) starts from OPV aldehyde 1, which has been

described previously.9d The aldehyde functionality of 1 was protected as a dimethyl acetal 2.

Aldehyde 4 was obtained after reacting amine 3 with 3-bromobenzaldehyde in a palladium-catalyzed

Page 31: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 2

20

reaction. Aldehyde 4 was subsequently converted into the Schiff base 5 after refluxing with aniline in

ethanol. N-phenylaldimine 5 was then reacted with the methyl group of 2 in a Siegriest reaction17

affording, after acidic work up, aldehyde 6. A chlorobenzene solution of aldehyde 6, N-

methylglycine, and C60 was stirred for 16 h in the dark at reflux temperature to yield a mixture of C60,

the desired monoadduct, and higher adducts. The triad 7 was isolated after extensive column

chromatography in a 43% yield. This last synthetic step was done in collaboration with Joop Knol

(University of Groningen).

OR*OR*

OR*

R*OR*O

R*OO

O

OR*OR*

OR*

R*OR*O

R*OO

a

HN N

bN N

O

N N

N

N NOR*

OR*OR*

R*OR*O

R*OO

N NOR*

OR*OR*

R*OR*O

R*ON

c

d

e

OR*=O

1

6

2

3 4 5

7

Scheme 2.1. Synthesis of OAn-OPV-C60 (7). a. Amberlite IR 120, trimethyl orthoformate, methanol,

70 °C, 2 h, 91%; b. 3-bromobenzaldehyde, Pd2(dba)3, BINAP, Cs2CO3, toluene, 100° C, 5 days, 62%;

c. aniline, ethanol, 85 °C, 4 h, 79%; d. 1. Compound 2, t-BuOK, DMF, 80° C, 3 h, 72%; 2. HCl; e. N-

methylglycine, C60, chlorobenzene, reflux, 18 h, 43%.

Page 32: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced multistep energy and electron transfer in a molecular triad

21

For the synthesis of the OAn-OPV dyad (10, Scheme 2.2), the bromine atom in 8 was

exchanged for a deuterium atom18 by lithiation and subsequent deuteration with D2O to yield 9. A

Siegriest reaction of 9 with N-phenylaldimine 5 afforded 10. All compounds used in the photophysical

investigations were fully characterized using 1H and 13C NMR spectroscopy, mass spectrometry, FT-

IR, and elemental analysis.

OR*OR*

OR*

R*OR*O

R*O Br

OR*OR*

OR*

R*OR*O

R*O D

N N

D

OR*OR*

OR*

R*OR*O

R*O

O

a

OR*=

8 9

10

b

Scheme 2.2. Synthesis of OAn-OPV (10). a. 1. n-BuLi, diethyl ether, -10 °C; 2. D2O, room

temperature, 41%; b. 1. Compound 5, t-BuOK, DMF, 80° C, 5 h, 50%; 2. HCl.

2.3 Electronic properties and energetic considerations

UV/Visible absorption. The absorption spectrum of OAn-OPV-C60 in toluene solution

(Figure 2.2) exhibits two strong absorption bands at 327 and 440 nm and a weak absorption at 705

nm. Whereas each of the three chromophores contributes to the absorption band at 327 nm, the

absorption at 440 nm is dominated by the π−π* transition of the OPV segment. The absorption at 705

nm is characteristic for fulleropyrrolidines.9d The absorption spectrum of OAn-OPV-C60 is a near

superposition of the absorption spectra of the different components of the triad; only at high energies

there is a slight deviation of the linear combination (Figure 2.2). This is likely due to the fact that the

OAn-OPV-C60 triad has one less alkoxy substituted phenylene group compared to combined

chromophores (OAn + OPV + MP-C60)

Page 33: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 2

22

650 700 750

0

500

300 400 500 600 700

0.0

2.5

5.0

7.5

10.0

ε (l/

mol

.cm

) x

10-4

Wavelength (nm)

Figure 2.2. UV/Visible absorption spectra of the OAn-OPV-C60 triad (solid line) and model

compounds OAn (dotted line), OPV (dashed line), and MP-C60 (dashed-dotted line) recorded in

toluene solution, and the summation of the spectra of all three reference compounds (squares). Inset:

Magnification of the 705 nm absorptions.

Electrochemistry. The OAn-OPV-C60 triad exhibits a reversible first reduction wave at –0.70

V, corresponding to the fullerene moiety, and three reversible oxidation waves due to the OAn (+0.55

V and +1.03 V) and the OPV (+0.83 V) moieties (potentials are given vs. SCE, calibrated against

Fc/Fc+, recorded in dichloromethane with 0.1 M TBAPF6) (Table 2.1). These oxidation potentials are

in close agreement with the values established for the reference compounds (Table 2.1). 9d The small

difference in oxidation potential between the OPV moiety in triad OAn-OPV-C60 (+0.83 V) and the

OPV chromophore (+0.78 V)9d is attributed to the smaller number of electron donating alkoxy

substituents of the former. The reduction potentials of OPV and OAn-OPV (-1.91 and –1.87 V) were

measured in tetrahydrofuran (THF) and are much more negative than that of the fullerene (Table 2.1).

Page 34: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced multistep energy and electron transfer in a molecular triad

23

Table 2.1. One-electron redox potentials (E0) of OAn, OPV, MP-C60, OAn-OPV, and OAn-OPV-C60

(vs. SCE) calibrated with Fc/Fc+ (in dichloromethane with 0.1 M TBAPF6).

Compound E0red (V) E0

ox (V)

OAn 0.53 / 1.02

OPV -1.91 a 0.78

MP-C60 -0.70

OAn-OPV -1.87 a 0.53

OAn-OPV-C60 -0.70 0.53 / 0.83 / 1.03

a Measured in THF

Absorption spectra of redox states. UV/visible/near-IR spectroscopy enables to monitor the

electronic transitions of the donor-acceptor arrays during a stepwise oxidation process. Quantitative

chemical oxidation of the OAn-OPV dyad in dichloromethane solution was achieved by the addition

of thianthrenium perchlorate19 (Figure 2.3). After the addition of one equivalent of this oxidizing

agent, the intensity of the absorption band at 3.79 eV decreases and two absorption bands emerge in

the spectrum, one at 1.44 eV, and the other overlapping with the absorption band of the OPV unit at

2.83 eV. Comparison with the electronic transitions of the N,N,N´,N´-tetraphenyl-1,4-benzenediamine

radical cation (1.44 and 3.05 eV),20 demonstrates that the absorption bands are associated with the

formation of an OAn•+ radical cation, while the disappearing band is that of the neutral OAn moiety.

When a second equivalent of thianthrenium perchlorate is added, the band of the neutral OPV unit, at

2.83 eV, decreases and two absorption bands with vibronic fine structure, at 0.74 and 1.64 eV, related

to an OPV•+ radical cation,21 appear in the spectrum. At the same time the absorption of OAn•+ at 1.44

eV remains and the second band of OAn•+ at 3.04 eV is now clearly observable. After two equivalents

of oxidizing agent the absorption spectrum exhibits the characteristics of both OAn•+ and OPV•+

radical cations and therefore corresponds to that of the OAn•+-OPV•+ dication diradical.

Page 35: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 2

24

1 2 3 4

0.00

0.25

0.50

0.75

Abs

orba

nce

(O. D

.)

Energy (eV)

0.00

0.25

0.50

0.75

1.00

b

a

OPV+OPV+

OPV

OAn+

OAn

OAn+

Figure 2.3. UV/visible/near-IR spectra recorded during the conversion of OAn-OPV by stepwise

oxidation using thianthrenium perchlorate19 in CH2Cl2: (a) before (solid line) and after (dashed line)

adding 1 equivalent; (b) after adding 1 equivalent (dashed line) and 2 equivalents (solid line).

Energetic considerations. In the two dyads (OAn-OPV, OPV-C60) and triad (OAn-OPV-C60)

numerous processes may occur after photoexcitation. Apart from the intrinsic decay of the singlet-

excited state of the individual chromphores (photoluminescence, intersystem crossing, and thermal

decay), energy and electron transfer reactions involving more than one chromophore (or redox active

group) are possible. Whether these reactions occur depends, amongst others, on whether these

processes are exergonic. The absorption spectra (Figure 2.2) reveal that the lowest singlet-excited

state is located on the fullerene at 1.76 eV, followed by the singlet state of the OPV at 2.48 eV, while

that of the OAn segment is positioned at approximately 3.40 eV.

The energies of the charge-separated states can be estimated by calculating the change in free

energy (∆G0) for charge separation using a continuum model 22:

( ) ( )( )

+−−−−=∆ −+

sref0

2

ccs0

2

00redox0 1111

84AD

εεπεεπε rr

e

R

eEEEeG (2.1)

Page 36: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced multistep energy and electron transfer in a molecular triad

25

In this equation, Eox(D) and Ered(A), are the oxidation and reduction potentials of the donor

and acceptor molecules or moieties measured in a solvent with relative permittivity εref, E00 is the

energy of the excited state from which the electron transfer occurs, and Rcc is the center-to-center

distance of the positive and negative charges in the charge separated state. The radii of the positive

and negative ions are given by r+ and r– and εs is the relative permittivity of the solvent, -e is the

elemental charge, and ε0 is the vacuum permittivity.

Table 2.2. Change in free energy (∆G0) with reference to the lowest singlet state, reorganization

energy (λ), and barrier (∆G‡) for charge separation (CS), charge recombination (CR1), charge shift

(CSH), and charge recombination (CR2) in toluene (TOL), chlorobenzene (CB), o-dichlorobenzene

(ODCB), and benzonitrile (BZN) as determined using Eqs 2.1 and 2.3.

reaction solvent ∆G0

(eV) λ

(eV) ∆G‡ (eV)

OAn-OPV-1C60* → OAn-OPV+-C60– TOL 0.21 0.35 0.224

CS CB -0.22 0.75 0.093 ODCB -0.36 0.86 0.073 BZN -0.46 0.99 0.070

OAn-OPV•+-C60•–→ OAn-OPV-C60 TOL -1.97 0.35 1.905

CR1 CB -1.54 0.75 0.204 ODCB -1.40 0.86 0.088 BZN -1.30 0.99 0.024

OAn-OPV•+-C60•–→ OAn•+-OPV-C60

•– TOL -0.08 0.35 0.051 CSH CB -0.21 0.80 0.106

ODCB -0.25 0.91 0.119 BZN -0.29 1.06 0.140

OAn•+-OPV-C60• –→ OAn-OPV-C60 TOL -1.89 0.36 1.618

CR2 CB -1.32 0.89 0.053 ODCB -1.15 1.03 0.004 BZN -1.01 1.20 0.007

OAn-1OPV* → OAn•+-OPV•– TOL 0.41 0.35 0.407 CS CB -0.07 0.80 0.165

ODCB -0.22 0.91 0.133 BZN -0.34 1.06 0.123

The change in free energy for charge separation was calculated using Eq. 2.1 for four solvents

of interest with increasing polarity: toluene (ε = 2.38), chlorobenzene (ε = 5.71), o-dichlorobenzene (ε

= 9.93), and benzonitrile (ε = 25.2) (Table 2.2). For the calculations, the Rcc distances were

determined assuming that the charges are located at the centers of the OAn, OPV and C60 moieties.23

These distances were 15 and 30 Å for the OAn-OPV•+-C60•– and OAn•+-OPV-C60

•– charge-separated

states respectively. The radius of the negative ion is reported for C60 to be r– = 5.6 Å24 and that of the

Page 37: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 2

26

ions of OPV to r+/ r– = 5.1 Å 9d The radius of the positive charge of OAn was set to r+ = 4.8 Å, as

calculated from molecular modeling. Small variations of the approximated Rcc and r+/ r– values did

not significantly alter the outcome of the calculations for the change in free energy. According to Eq.

2.1 charge separation (CS: OAn-OPV-1C60* → OAn-OPV•+-C60

•–) and the charge shift (CSH: OAn-

OPV•+-C60•– → OAn•+-OPV-C60

•–) are energetically feasible in the three polar solvents (Table 2.2).

The experimental and estimated energies of the various neutral and charge-separated states of

the OAn-OPV-C60 triad are depicted in Figure 2.4, assuming chlorobenzene as the medium. Figure 2.4

also shows the photophysical reactions that have been identified in the triad or its model compounds,

with the prevailing processes highlighted in black, vide infra.

1OAn*-OPV-C601OAn*-OPV-C60

OAn-1OPV*-C60

OAn-OPV-C60

OAn-1OPV*-C60OAn-1OPV*-C60

OAn-OPV-C60OAn-OPV-C60

OAn-OPV•+-C60• -

OAn•+-OPV-C60• -

OAn-OPV•+-C60• -OAn-OPV•+-C60• -

OAn•+-OPV-C60• -OAn•+-OPV-C60• -

OAn•+-OPV•--C60

OAn-OPV-1C60*OAn-OPV-1C60*

OAn-OPV-3C60*OAn-OPV-3C60*

kET

kCS

kCSH

kPLOPV

k´ET

kISCC60

OAn-3OPV*-C60

kISCOPV

kPLC60

Energy

k´CS

k´CR

kCR1

kCR2

Figure 2.4. Schematic energy levels (in chlorobenzene) and photoinduced processes in OAn-OPV-

C60. The rate constants are collected in Table 2.3. The fastest processes have been highlighted in

black (vide infra).

Page 38: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced multistep energy and electron transfer in a molecular triad

27

2.4 Photophysical processes in solution

2.4.1 Photoluminescence spectroscopy

Photoluminescence (PL) experiments on the triad, dyads, and model compounds, were used to

study the energy and electron transfer reactions that occur in these molecules. These transfer

relaxation pathways are expected to quench the PL of the chromophores, especially when their rate

constants are higher than that of the intrinsic decay.

500 600 700 8000

3

6

300 400 500 600 7000.0

0.5

1.0b

a

PL

Inte

nsity

(a.

u.)

Wavelength (nm)

Nor

mal

ized

PL

Inte

nsity

Wavelength (nm)

Figure 2.5. (a) PL spectra of OPV-C60 in toluene (dashed line) and of OAn-OPV-C60 in toluene (solid

line) and o-dichlorobenzene (dotted line). (b) Excitation spectra of the 710 nm emission of OPV-C60

(dashed line) and OAn-OPV-C60 (solid line) compared to the absorption spectrum of OAn-OPV-C60

(dashed line), all in toluene.

The PL at 499 nm of the OPV moiety of the OAn-OPV-C60 triad, dissolved in toluene, is

highly quenched (quenching factor Q > 4000) compared to the PL of the corresponding OPV

molecule. Apart from a residual emission at 499 nm,25 photoexcitation of the OPV moiety of OAn-

OPV-C60 results in a weak PL signal at 715 nm (Figure 2.5a), characteristic of the fluorescence

emission band of fulleropyrrolidines.9d In toluene the quantum yield of this emission is nearly

identical to that of the reference compound MP-C60. The same result is observed for the OPV-C60

dyad, although in this case the PL quenching of the OPV is somewhat less (Q ≈ 1500).25 The strong

Page 39: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 2

28

quenching of the 1OPV* singlet-excited state (S1) in the OPV-C60 dyad has previously been studied in

detail and was found to involve an ultrafast photoinduced intramolecular singlet-energy transfer (ET)

towards the fullerene moiety (1OPV*-C60 → OPV-1C60*), which occurs with a time constant of less

than 190 fs.9e,26 By analogy, the same ET process must occur in the OAn-OPV-C60 triad. In

accordance with this proposition, the excitation spectrum of the fullerene fluorescence of OAn-OPV-

C60 coincides with the corresponding absorption spectrum of OAn-OPV-C60 (Figure 2.5b). It is

interesting to note that the fullerene excitation spectra of OPV-C60 and OAn-OPV-C60 differ

appreciably at lower wavelengths, where the OAn moiety absorbs. The excellent agreement between

the absorption spectrum of the OAn-OPV-C60 triad and the fullerene excitation spectrum implies that

not only excitation of OPV, but also excitation of OAn is responsible for the emission of the fullerene.

This points to an efficient sequential singlet-energy transfer in the triad (Figure 2.4), that starts with

the formation of the 1OAn*-OPV-C60 singlet-excited state and ends at the OAn-OPV-1C60* fullerene

singlet-excited state.

In a more polar solvent, like o-dichlorobenzene (ε = 9.93), the PL of the OPV unit of the

OAn-OPV-C60 triad is quenched to a similar extent as in toluene (Figure 2.5a). In contrast, the PL of

the fullerene moiety at 715 nm is significantly quenched in this more polar solvent as compared to

toluene (Q ≈ 400, Figure 2.5a). The same result has previously been observed for the OPV-C60 dyad9d,e

and gives evidence of a photoinduced charge separation (CS) that occurs from the fulleropyrrolidine

singlet-excited state and produces the OPV•+-C60•– charge-separated state. For the OAn-OPV-C60 triad,

a photoinduced CS will initially produce a similar state (OAn-OPV•+-C60•–), which may then go

through a charge shift (CSH) to generate the energetically more favorable OAn•+-OPV-C60•– state

(Figure 2.4).

For a further comparison, we have studied the PL of the OAn-OPV dyad in solvents of

different polarity. Figure 2.6a shows that the emission of 1(OAn-OPV)*, dissolved in toluene, stems

exclusively from the OPV moiety, irrespective of the excitation wavelength (330 nm (OAn) or 444

nm (OPV)). This implies that an efficient singlet-energy transfer occurs from the excited 1OAn* state

to OPV (1OAn*-OPV → OAn-1OPV*). In toluene, the fluorescence quantum yield of OAn-OPV is

slightly (~10%) higher than that of OPV, but the PL is progressively quenched with increasing solvent

polarity (Figure 2.6a), providing quenching factors of Q ≈ 2, 9, and 22 for chlorobenzene, o-

dichlorobenzene, and benzonitrile, respectively. The quenching of OAn-OPV PL after excitation at

440 nm in more polar solvents is attributed to an intramolecular photoinduced electron transfer

reaction in the excited state to produce the OAn•+-OPV•– state. The excitation spectrum of the residual

emission of OAn-OPV recorded in chlorobenzene (Figure 2.6b) closely corresponds to the absorption

spectrum of OAn-OPV. This indicates that the 1OAn*-OPV → OAn-1OPV* singlet-energy transfer is

significantly faster than an electron transfer from the same state (1OAn*-OPV → OAn•+-OPV•–).

Page 40: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced multistep energy and electron transfer in a molecular triad

29

Hence, OAn•+-OPV•– is primarily formed via the OAn-1OPV* singlet state, irrespective of the

excitation wavelength.

400 500 600 700

0

20

40

60

80

100

120

300 400 500 600

0.0

0.5

1.0 b

a

Cou

nts

x 10

-4

Wavelength (nm)

Nor

mal

ized

inte

sity

Wavelength (nm)

Figure 2.6. (a) PL spectra of OAn-OPV in toluene (solid squares), chlorobenzene (open squares), o-

dichlorobenzene (closed circles), and benzonitrile (open circles) solutions with excitation at 444 and

330 nm (for toluene only, dashed line). (b) Excitation spectrum of the OPV emission of OAn-OPV in

chlorobenzene (dashed line) and compared to the absorption spectrum (solid line).

2.4.2 Near steady state photoinduced absorption (PIA) spectroscopy

Near steady state PIA spectroscopy in the microsecond and millisecond time domain is a very

sensitive technique (detection limit ∆T/T ~ 10-6) to probe small concentrations of long-lived

photoexcitations such as triplet states and intermolecular charge-separated states. The PIA spectrum

of OAn-OPV-C60 in toluene solution, recorded with excitation at 458 nm, exhibits a band at 1.78 eV

with a shoulder at 1.52 eV, characteristic of the long-lived (~40 µs) triplet state of the

fulleropyrrolidine moiety (OAn-OPV-3C60*) formed via intersystem crossing from the fullerene

singlet-excited state. 9d

The photoinduced electron transfer in OAn-OPV-C60 in o-dichlorobenzene and subsequent

charge shift or charge recombination will likely occur in the picosecond to nanosecond time regime

and cannot be resolved with the near-steady state PIA technique. However, absorptions of the charged

Page 41: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 2

30

states can be observed in mixtures of the different model compounds of the triad in o-dichlorobenzene

solution, by using the redox activity of the corresponding triplet states. Photoexcitation of OPV (at

458 nm) or MP-C60 (at 528 nm) will result in the formation of the corresponding excited triplet states.

These triplet states (3OPV* and MP-3C60*) can undergo an electron transfer to one of the other redox

active chromophores present in solution to produce an intermolecularly charge-separated state, which

is long-lived because the formed cation and anion radicals diffuse away in solution.

0.5 1.0 1.5 2.00.0

0.5

1.0

75 Hz 2500 Hz

Energy (eV)

0.0

2.5

c

-∆T

/T x

104

b

a

0.0

0.5

1.0

1.5

Figure 2.7. (a) Normalized photoinduced absorption spectra of the mixtures OPV/MP-C60 (1:1) (solid

line) and OAn/MP-C60 (1:1) (dashed-line) in o-dichlorobenzene (excitation at 528 nm with 25 mW

and modulation frequency of 275 Hz). (b) Photoinduced absorption spectra of the mixtures OAn-

OPV/MP-C60 (1:1) (solid line) or OAn/OPV/MP-C60 (1:1:1) (dashed line) in o-dichlorobenzene

(excitation at 528 nm with 25 mW and modulation frequency of 275 Hz). (c) Normalized photoinduced

absorption spectra of the mixture OAn/OPV/MP-C60 (1:4:4) in o-dichlorobenzene recorded at

modulation frequencies of 75 Hz (solid line) and 2500 Hz (dashed line) (excitation at 528 nm with 25

mW).

Page 42: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced multistep energy and electron transfer in a molecular triad

31

Accordingly, selective photoexcitation of MP-C60 at 528 nm in mixtures with OPV or OAn in

o-dichlorobenzene solution produces the charge-separated states OPV•+/MP-C60•– and OAn•+/MP-C60

•–

respectively, which are characterized by the absorptions of OPV•+ (at 0.68 and 1.52 eV), OAn•+ (at

1.40 eV), and MP-C60•– (at 1.24 eV) (Figure 2.7a). Likewise, photoexcitation of MP-C60 in a mixture

with OAn-OPV in o-dichlorobenzene results in a band at 1.40 eV (Figure 2.7b) attributed to OAn•+-

OPV radical cation. Although in this mixture the OAn-OPV•+ radical cation can also be formed, it will

probably rearrange by an intramolecular redox reaction to the OAn•+-OPV state within the timescale

of the experiment. In an equimolar mixture of OAn, OPV, and MP-C60, charge transfer from the

positively charged OPV+ onto the OAn becomes intermolecular, i.e. diffusion limited and slower.

This allows for the detection of a weak residual absorption band characteristic of the OPV•+ radical

cation in the PIA spectrum at 0.68 eV (Figure 2.7b).

Even in a mixture where both OPV and MP-C60 are present in fourfold excess with respect to

OAn, the PIA spectrum is still dominated by the absorption of the OAn•+ radical cation (Figure 2.7c).

At high modulation frequency (2500 Hz) the relative intensity of the OPV•+ radical cation band at

0.68 eV increases compared to low modulation frequency (75 Hz) (when normalized at 1.24 eV),

indicating that in this mixture OPV•+ has a shorter lifetime than the OAn•+ radical cation, consistent

with the proposed OPV•+ + OAn → OPV + OAn•+ redox reaction.

2.4.3 Femtosecond pump-probe spectroscopy in polar solvents

The formation and decay of the transient charged species generated after photoexcitation of

the OPV moiety in OAn-OPV-C60 and the reference compounds OPV-C60 and OAn-OPV have been

investigated with sub-picosecond pump-probe spectroscopy in solvents of different polarity.

Upon excitation at 450 nm of OPV-C60 or OAn-OPV-C60, the transient absorption at 1450 nm

(0.85 eV) that is associated with the OPV•+ radical cation, i.e. the OPV•+-C60•– and OAn-OPV•+-C60

•–

states, exhibits a rise and a decay component (Figure 2.8). It is important to note that the charge

separation occurs subsequent to the ultrafast singlet-energy transfer: OAn-1OPV*-C60 → OAn-OPV-1C60* (kET ≥ 5.3 × 1012

s-1).9e Fitting of the temporal differential absorption data at 1450 nm to a

biexponential function27 provides the rate constants for forward and backward electron transfer

reactions (Table 2.3). As can be seen in Figure 2.8 (and Table 2.3), the rate constants for the charge

separation (kCS) are equal in the dyad and triad within experimental error. The rates for charge

separation in the dyad and triad increase significantly with solvent polarity from kCS = 4.3 × 1010 s

-1 in

chlorobenzene to kCS = 2.0 × 1011 s-1 in benzonitrile. A similar increase is found for the charge

recombination. This effect of the medium on the rate constant has often been observed in donor-

acceptor dyads and is in agreement with Marcus theory, vide infra.28

Page 43: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 2

32

-60

-40

-20

0

0 50 100 150 200 250

-60

-40

-20

0

-60

-40

-20

0

a

b

c

∆T/T

(a.

u.)

Time delay (ps)

Figure 2.8. Differential transmission dynamics of the OPV•+ transition at 1450 nm for OAn-OPV•+-

C60•–

(solid line) and OPV•+-C60•– (dashed line) in (a) chlorobenzene, (b) o-dichlorobenzene, and (c)

benzonitrile after excitation at 450 nm.

In comparison to the dyad, the charge shift (CSH) from OAn-OPV•+-C60•– to OAn•+-OPV-

C60•– provides an extra pathway in the triad for the decay of the OPV•+ radical other than charge

recombination (CR1, Figure 2.4) to the ground state. Hence, the OPV•+ radical cation absorption may

disappear faster in the triad than in the dyad. The experimental data in Figure 2.8 show, however, that

this difference is only discernable in chlorobenzene, i.e. the least polar solvent. This is a direct

consequence of the increased rate for charge recombination in more polar solvents (Figure 2.8), which

implies that the probability for the competing (slower) charge shift is strongly reduced. Non-linear

least squares analysis of the experimental OPV•+ traces (Figure 2.8) does not give a clear indication of

the magnitude of the rate of charge shift; it yields values of kCR1 and (kCR1 + kCSH) that are essentially

the same (see table 3). This may be explained in part by a difference in kCR1 for the dyad and the triad

and in part by the difficulties in the data analysis itself. Due to the fact that the rates for formation and

decay of the OPV•+ species are very similar, it is difficult to estimate the rate. Non-linear-least-

Page 44: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced multistep energy and electron transfer in a molecular triad

33

squares-fit procedures invariably yield rate parameters for formation and decay that are strongly

correlated.

Table 2.3. Rate constants for intrinsic decay of the lowest-energy singlet-excited chromophore (k0OPV,

k0C60), singlet-energy transfer (kET), charge separation (kCS), charge recombination (kCR1), and charge

shift (kCSH) in the studied compounds in toluene (TOL), chlorobenzene (CB), o-dichlorobenzene

(ODCB), benzonitrile (BZN).

Compound TOL

k (ns-1)

CB

k (ns-1)

ODCB

k (ns-1)

BZN

k (ns-1) OPV k0

OPV 0.83 0.79 0.74 0.69

MP-C60 k0C60 0.68 0.72 0.75 0.68

OPV-C60 kET ≥ 5300 n.d. ≥ 5300 n.d.

kCS - 40±2 50±9 250±40

kCR1 - 12±1 19±3 37±8

OAn-OPV-C60 kET ≥ 5300 n.d. ≥ 5300 n.d.

kCS - 48±5 71±12 201±54

kCR1+ kCSH - 11±1 16±3 61±19

OAn-OPV k´CS - 16.4 15.6 40.8

k´CR - 1.0 1.4 4.1

Additional information on the CSH reaction can be obtained by measuring the photoinduced

absorption at 1030 nm (1.20 eV). At this probe wavelength the S1 states 1OPV* and 1MP-C60*, as well

as the radical ions MP-C60•–, and OAn•+ contribute to the induced absorption (Figure 2.7). The

differential transmission at 1030 nm of OPV-C60 and OAn-OPV-C60 in chlorobenzene undergoes a

strong rise and drop within 2 ps (Figure 2.9). This transient signal is associated with the formation and

decay of the 1OPV* singlet-excited state and involves the Sn←S1 transition of the OPV unit.9e The

short lifetime of the 1OPV* state is due to the ultrafast ET onto the fullerene moiety as described

above. After the ET, the 1030 nm signal for the OAn-OPV-C60 triad remains constant over the

timescale of the experiment (1 ns), while for the dyad the signal decays to zero within 200 ps (Figure

2.9). It is proposed that the remaining signal is due to absorption by C60•– and OAn•+ radical ions, and

hence characteristic of the OAn•+-OPV-C60•– charge separated state. Using the following extinction

coefficients for these species (7.7 × 104 M-1cm-1 (OPV(S1), estimate), 7 × 103 M-1cm-1 (MP-C60(S1) 29),

7 × 103 M-1cm-1 (MP-C60•– 29), 8.2 × 103 M-1cm-1 (OAn•+ 20)) the time profile for the absorption at 1030

nm (Figure 2.9) has been modeled for both the OPV-C60 dyad and the OAn-OPV-C60 triad. Rate

Page 45: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 2

34

constants were taken from the transient absorption measurements at 1450 nm probe wavelength

(Table 2.3) and the rate for charge shift (kCSH) is used as an adjustable parameter.30 The results show a

large induced absorption due to the OPV(S1) species. The duration of this transient is mainly

determined by the width of the laser pulses used. For the modeling a cross-correlation of 500 fs

(FWHM) for the pump and probe pulse was assumed. For the dyad, a smooth trace is observed after

the initial contribution of the excited OPV moiety. This is consistent with the fact that the C60(S1) and

the C60•– groups have almost equal extinction coefficients. For the triad, a long-lived signal is

observed which can be modeled taking kCSH = 5 × 109 s-1 and kCR1 = 6 × 109 s-1. Taking a higher

(lower) value for kCSH results in a curve that bends downward (upward) after 60 ps. Using kCSH = 5 ×

109 s-1 results in a calculated yield of formation for the OAn•+-OPV-C60•– species of about 0.4 per

absorbed photon. This result can be compared with the following crude estimate. At 30 ps mainly the

Oan-OPV•+-C60•– state is present of which only the C60

•– makes a major contribution to the absorption

at 1030 nm. At t= 200 ns only the OAn•+-OPV-C60•– state is present and OAn•+ and C60

•–contribute

almost equally to the transient absorption. The experimental observation that the induced absorption

at 1030 nm hardly changes between 30 and 200 ps thus indicates that the probability of formation of

OAn•+-OPV-C60•– out of OAn-OPV•+-C60

•– is roughly one half, in good agreement with the estimate

above. Figure 2.9 also shows the induced absorption signal at 1450 nm as observed for the triad. This

latter signal has been modeled using the same parameters as mentioned above.

0.1 1 10 100 1000

-100

-50

0

∆T/T

(a.

u.)

Time delay -1 (ps)

Figure 2.9. Differential transmission dynamics of OAn-OPV-C60 (open circles) and OPV-C60 (closed

squares) in chlorobenzene monitored at 1030 nm with excitation at 450 nm. The induced absorption

signal at 1450 nm as observed for the triad is shown with open squares. The lines correspond to a

numerical simulation based on the model described in the text. For the simulation the rate constants

from Table 3 are used and the rate for charge shift is used as an adjustable parameter. The time delay

has been shifted by 1 ps to show the signals on a logarithmic plot.

Page 46: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced multistep energy and electron transfer in a molecular triad

35

We also studied the reference compound OAn-OPV with pump-probe spectroscopy in

solution. In particular, it is of interest to investigate whether an OAn-1OPV*-C60 → OAn•+-OPV•–-C60

electron transfer reaction can compete with the OAn-1OPV*-C60 → OAn-OPV-1C60* energy transfer.

As a result of the electron-hole symmetry in conjugated oligomers, the optical spectra of OPV

radical cations and OPV radical anions are expected to be very similar and, hence, we probe the

formation of OPV•– at 1450 nm.31 The transient differential transmission recorded for OAn-OPV,

dissolved in solvents with increasing polarity, at 1450 nm after excitation at 455 nm of the OPV

chromophore is shown in Figure 2.10. In full agreement with the PL quenching of the OAn-OPV

fluorescence (Figure 2.6), the OAn•+-OPV•– charge-separated state is only formed in the three polar

solvents (chlorobenzene, o-dichlorobenzene, and benzonitrile) but not in (less polar) toluene (Figure

2.10). The rate constants for charge separation and charge recombination (k´CS and k´CR, Table 2.3)

increase with increasing polarity. Table 2.3 reveals that the rate for the OAn-1OPV* → OAn•+-OPV•–

charge separation (k´CS = 1.6 - 4.1 × 1010 s

-1) is significantly less than the rate of the 1OPV*-C60 →

OPV-1C60* energy transfer (kET ≥ 5.3 × 1012 s

-1) and as a consequence the OAn•+-OPV•–-C60 state is

unlikely be formed in a significant yield from OAn-1OPV*-C60. Hence, excitation of OAn-OPV-C60 in

solution will always provide OAn-OPV-1C60* as an intermediate state, irrespective of the excitation

wavelength and solvent polarity.

0 200 400 600 800 1000-0.3

-0.2

-0.1

0.0

∆T/T

(a.

u.)

Time delay (ps)

Figure 2.10. Differential transmission dynamics of OAn•+-OPV•– in toluene (closed squares),

chlorobenzene (closed circles), o-dichlorobenzene (open circles), and benzonitrile (open squares)

monitored at 1450 nm with excitation at 455 nm.

Charge recombination in OAn•+-OPV•– is almost an order of magnitude slower than in

OPV•+-C60•– and occurs in the nanosecond regime. In this respect, a remarkable phenomenon was

observed when the fluorescence lifetimes of OAn-OPV were recorded. In the most polar solvents o-

Page 47: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 2

36

dichlorobenzene and benzonitrile the lifetime of the OAn-OPV emission is significantly reduced to

~140 and ~100 ps, compared to the OPV model compound (1.36 and 1.44 ns, respectively). This is in

accordance with the proposed OAn-1OPV* → OAn•+-OPV•– electron transfer reaction, which reduces

the lifetime of the 1OPV* state. The lifetimes of 1OPV* and OAn-1OPV* in toluene are very similar

(1.20 and 1.38 ns, Figure 2.9) as expected because electron transfer does not occur here. However,

and surprisingly, a significant increase of fluorescence lifetime was observed for OAn-1OPV* in

chlorobenzene (2.65 ns) compared to OPV (1.27 ns) (Figure 2.11). At first glance this increase in

fluorescence lifetime seems to contradict the observed OAn-1OPV* → OAn•+-OPV•– charge

separation reaction, which is approximately 20 times faster than the intrinsic decay (Figure 2.10,

Table 2.3). Instead of an increase in fluorescence lifetime, a decrease would be expected. Note that

except for a loss in intensity (only a factor of 2), the fluorescence spectra of OPV and OAn-OPV are

identical (Figure 2.6), and that the emission is thus from the OAn-1OPV* state. These experimental

observations lead to the conclusion that in chlorobenzene the singlet OAn-1OPV* state and the

OAn•+-OPV•– charge-separated state are nearly degenerate and that back electron transfer from

OAn•+-OPV•– reproduces the OAn-1OPV* singlet state. The estimated change in free energy for OAn-1OPV* → OAn•+-OPV•– of only ∆G0 = – 0.07 eV (Table 2.2) supports this suggestion.

0 5 10

10

100

1000

10000

OAn-OPV: TOL

OAn-OPV: CB

OPV: CB

OPV: TOL

Cou

nts

Time (ns)

Figure 2.11. Time-resolved fluorescence of OPV and OAn-OPV in toluene (TOL) and chlorobenzene

(CB) recorded with excitation at 400 nm.

2.4.4 Kinetic considerations

The final outcome of a photoexcitation not only depends on the energetics of the reaction, but

also on the kinetics. Marcus theory provides an estimate for the free energy barrier (∆G‡) for electron

transfer reactions based on the change in free energy (∆G0) and the reorganization energy (λ) via:

Page 48: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced multistep energy and electron transfer in a molecular triad

37

( )

λλ

4

20‡ +∆=∆ G

G (2.2)

The reorganization energy consists of an internal contribution (λ i) and a solvent term (λs),

which can be approximated via the Born-Hush approach to give after summation:

++=+= −+

s2

0

2

isi11111

2

1

4 επελλλλ

nRrr

e

cc

(2.3)

The rate constants for the different processes, are not only a function of the energy barrier

∆G‡, but also of the reorganization energy (λ) and the electronic coupling (V) between donor and

acceptor in the excited state according to the equation:

( )

+∆−

=

Tk

GV

Tkhk

B

202

21

B2

2

4exp

4

λλ

λπ

(2.4)

The values of ∆G0, λ, and ∆G‡ calculated on the basis of Eqs. 2.1 to 2.4 collected in Table

2.2, show that the initial charge separation (kCS) is in the Marcus normal region (–∆G0 < λ). As the

polarity of the solvent increases, the OAn-OPV•+-C60•– charge-separated state is stabilized with a

concomitant increase of the reorganization energy. The combination of these trends results in

reduction of the barrier for charge separation in more polar solvents (Table 2.2). As a consequence,

the rate for charge separation (kCS) is expected to increase with polarity as has been found

experimentally (Figure 2.8, Table 2.3). Charge recombination in OAn-OPV•+-C60•– is in the Marcus

inverted region (–∆G0 > λ). The use of Eq. (2.4) in the inverted region often underestimates the true

rate constant because of nuclear tunneling,32 but will be used here for qualitative comparison. We find

that the barrier for charge recombination is strongly reduced in more polar solvents, consistent with

the higher recombination rate (kCR1, Table 2.3). Apart from recombination to the ground state, the

OAn-OPV•+-C60•– state may undergo a charge shift to form OAn•+-OPV-C60

•–. The charge shift occurs

in the normal region (–∆G0 < λ) and is energetically less favorable than the recombination (Table 2.3).

However, in contrast to the recombination, the barrier for charge shift is reduced with decreasing

polarity. Hence, the balance between charge recombination and the competing charge shift, will move

towards the latter with decreasing polarity. As a result, the barrier for the charge shift in OAn-OPV•+-

C60•– is lower than that for charge recombination in chlorobenzene, but not in o-dichlorobenzene and

benzonitrile. This is in full agreement with the experimental result that the charge shift was more

easily observed in chlorobenzene (Figure 2.8 and 2.9).

Page 49: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 2

38

The energy barriers for relaxation to the ground state in OAn•+-OPV-C60•– are small.

However, besides the energy barrier, the rate constant (Eq. 2.4) is also determined by the electronic

coupling V. This electronic coupling depends exponentially on the center-to-center distance between

the donor and acceptor via ))(exp()( 0cc02

02 RRRVV −−= β , with R0 the contact distance. Hence, V

is orders of magnitude less for OAn•+-OPV-C60•– (Rcc = 30.0 Å) than for OAn-OPV•+-C60

•– (Rcc = 15.4

Å).33 The reduction of the electronic coupling V, caused by the longer distance between the centers of

positive and negative charge density in OAn•+-OPV-C60•–, is the origin of the increase in lifetime for

OAn•+-OPV-C60•– compared to OAn-OPV•+-C60

•–.

The rate constants of the various photoinduced processes in solvents of different polarity have

been calculated using Eq. 2.4 relative to the corresponding rate constant in chlorobenzene by

assuming that V is independent on the solvent, and are collected in Table 2.4.

Table 2.4. Rate constants for the charge separation (CS), charge recombination (CR1), charge shift

(CSH), and charge recombination (CR2) in o-dichlorobenzene (ODCB) and benzonitrile (BZN) with

relative to the corresponding rate constant in chlorobenzene (CB), calculated using Eq. 2.4.

rate constant CS CR1 CSH CR2

k(CB) 1 1 1 1

k(ODCB) 2.00 89 0.58 6.25

k(BZN) 2.16 980 0.23 5.06

2.5 Photophysical processes in the solid state

2.5.1 Near steady state PIA spectroscopy

Near steady state PIA spectra of thin films of OPV-C60 and OAn-OPV-C60 were recorded at

80 K with excitation at 458 nm. The PIA spectrum of the OPV-C60 film (Figure 2.12) exhibits the

signals of OPV•+ at 0.68 and 1.52 eV and that of C60•– at 1.24 eV, characteristic of a charge-separated

state. The PIA spectrum of a film of OAn-OPV-C60 (Figure 2.12) lacks the bands at 0.68 and 1.52 eV

and exhibits only one broad band that peaks at 1.25 eV. This signal is attributed to the overlapping

transitions of the OAn•+ and C60•– radical ions and gives evidence for the formation of an

intramolecular (OAn-OPV•+-C60•–) or intermolecular (OAn•+-OPV-C60 / OAn-OPV-C60

•–) charge-

separated state. The charge-separated state in the OPV-C60 and OAn-OPV-C60 films, measured with

this PIA technique, extend into the millisecond time domain. The PIA band at 1.25 eV increases with

the pump intensity (I) following a square-root power law (–∆T/T ~ I0.5). This suggests a non-geminate

bimolecular recombination of the photoinduced charges. We propose that the long-lived (ms domain)

Page 50: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced multistep energy and electron transfer in a molecular triad

39

charges observed with near steady state PIA in films of OPV-C60 and OAn-OPV-C60 at 80 K, are

associated with a small fraction of positive and negative charges that have escaped from geminate

recombination by charge migration to other sites in the film where they became trapped and are thus

associated with different molecules in the film (i.e. OAn•+-OPV-C60 and OAn-OPV-C60•–).

0.5 1.0 1.5 2.0

0.0

0.5

1.0

1.5

2.0

∆T

/T x

104

Energy (eV)

Figure 2.12. Photoinduced absorption spectra of OAn-OPV-C60 (solid line) and OPV-C60 (dotted

line) in thin films. Recorded at 80 K with excitation at 458 nm (25 mW) and a modulation frequency

of 275 Hz.

2.5.2 Femtosecond pump-probe spectroscopy

Femtosecond spectroscopy shows that in films of OPV-C60 and OAn-OPV-C60 at room

temperature a charge-separated state is formed within 0.5 ps, as evidenced by the instantaneous rise of

the 1450 nm differential transmission (Figure 2.13a) associated with OPV•+ radical cations. Although

a lack of higher time-resolution precludes an unambiguous conclusion, we have no evidence that a

singlet-energy transfer precedes the electron transfer reaction in the solid state. Possibly, the

photoinduced electron transfer in the solid state is predominantly intermolecular. The OPV•+-C60•–

state is much longer lived in the film than in solution (Table 2.3). This phenomenon has also been

observed in other donor-acceptor dyads and has been attributed to the migration of opposite charges to

different sites in the film.8h,9d,34 In contrast, the OPV•+ signal in the film of the triad rapidly decays,

with a time constant of 20 ps over the first 100 ps. This is interpreted to result from an intermolecular

or intramolecular charge shift (CSH) from OPV•+ to OAn•+. The time profile of the 1030 nm

differential transmission (Figure 2.13b) is consistent with the formation of an OAn•+-OPV-C60•– state

in the triad and explains the short lifetime of the OAn-OPV•+-C60•– state (Figure 2.13a). This time

profile cannot be fitted to monoexponential decay, and suggests several lifetimes. This fact can be

rationalized by a combination of a direct charge recombination and an indirect charge recombination

Page 51: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 2

40

after migration of the charges in the film.8h,9d It should be noticed however, that contrary to what is

observed for the OPV-C60 dyad, the OAn•+-OPV-C60•– state seems to be longer lived in solution than

the OAn•+-OPV-C60 / OAn-OPV-C60•– state in the film. As a tentative explanation for this difference

we propose that in solution the triads are isolated from each other such that the weak electronic

coupling between the OAn and the C60 decelerates the intramolecular charge recombination in OAn•+-

OPV-C60•–. In the film the triads are in intimate contact with each other, and intermolecular charge

recombination between OAn•+-OPV-C60 and OAn-OPV-C60•– can occur.

0 100 200 300 400 500 600-120

-80

-40

0

0 200 400 600 800 1000-100

-80

-60

-40

-20

0b

a

Time delay (ps)

∆T/T

(a.

u.)

∆T/T

(a.

u.)

Time delay (ps)

Figure 2.13. (a) Differential transmission dynamics at 1450 nm of thin films of OAn-OPV-C60 (open

circles) and OPV-C60 (closed circles) at 298 K with excitation at 450 nm. (b) Differential transmission

dynamics at 1030 nm of a thin film of OAn-OPV-C60 at 298 K with excitation at 450 nm.

2.6 Conclusions

A molecular triad, OAn-OPV-C60, with a redox gradient in a linear array has been

synthesized. The photophysical processes that may occur in this system are schematically depicted in

Figure 2.4 and have been investigated in solution and in thin films with photoluminescence and

transient absorption spectroscopy.

Page 52: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced multistep energy and electron transfer in a molecular triad

41

In solution, photoexcitation of either chromophore of the OAn-OPV-C60 triad results in an

ultrafast (sequence of) singlet-energy transfer (ET, Figure 2.4) and provides a singlet state on the

fulleropyrrolidine unit (OAn-OPV-1C60*), irrespective of the polarity of the solvent. The competitive

process of charge separation from the primary 1OAn*-OPV-C60 or OAn-1OPV*-C60 states are more

than one order of magnitude slower and were not observed in the triad. In toluene, the singlet OAn-

OPV-1C60* state decays via intersystem crossing (ISC) to the OAn-OPV-3C60* triplet state and via PL

to the ground state. In more polar solvents, the singlet OAn-OPV-1C60* state gives rise to an

intramolecular charge separation reaction (CS) that generates the OAn-OPV•+-C60•– state. The rate for

this forward electron transfer reaction increases with the polarity of the solvent from kCS = 4.8 × 1010 s

-1

in chlorobenzene to kCS = 2.0 × 1011 s

-1 in benzonitrile, in qualitative agreement with Marcus theory.

Because the oxidation potential of the OAn segment is below that of the OPV unit, the primary OAn-

OPV•+-C60•– charge-separated state may undergo an intramolecular redox reaction, or charge shift

(CSH), to form OAn•+-OPV-C60•–. The charge recombination in OAn-OPV•+-C60

•– (CR1), however,

competes with the charge shift. Because the charge recombination is slowed down in less polar

solvents, the quantum yield for the charge shift is the highest (~0.4) in the least polar solvent in which

electron transfer occurs, i.e. chlorobenzene. The charge recombination in the secondary OAn•+-OPV-

C60•– charge-separated state (CR2) is significantly slower (kCR2 < 1 × 109

s-1) than that of the primary

OAn-OPV•+-C60•– state (kCR1 = 1.1 × 1010

s-1). The observed trends in the various rate constants (kCS,

kCSH, kCR1, kCR2, k´SC and k´CR) with changing solvent polarity are in qualitative agreement with Marcus

theory when the free energies of the charge-separated states are determined using a continuum model

(Eq. 2.1).

In thin films, charge generation on the OPV unit of OAn-OPV-C60 is much faster (kCS ≥ 3.0 ×

1012 s

-1) and likely predominantly intermolecular. In the films a subsequent charge-shift occurs from

the primary OPV•+ radical cation to an OAn•+ radical cation with a rate close to kCSH = 5.0 × 1010 s

-1.

Because we consider it likely that in the film the primary charge-separated state involves two

molecules, also the charge shift probably involves an intermolecular OAn-OPV•+-C60 → OAn•+-OPV-

C60 reaction, with the negative charge located on the fullerene unit of a third molecule (OAn-OPV-

C60•–). The lifetime of the charges formed in the film (OAn•+-OPV-C60 / OAn-OPV-C60

•–) is somewhat

less as compared to the solution.

The experiments on OAn-OPV-C60 demonstrate that in solution the OAn-OPV•+-C60•– →

OAn•+-OPV-C60•– charge shift and the resulting spatial extension of the charges increase the lifetime

of the charge-separated state compared to OAn-OPV•+-C60•–, because the electronic coupling between

the redox active groups is strongly reduced. This provides a rationale to explain the long-lifetime of

the charge-separated state in conjugated polymer:C60 blends in terms of charge migration. The major

differences in the kinetics of the electron transfer reactions observed after photoexcitation of OAn-

Page 53: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 2

42

OPV-C60 in solution or in thin films, further demonstrate that intermolecular interactions are of crucial

significance in this respect. Creating and investigating well-defined multichromophoric

supramolecular donor-acceptor assemblies, consisting of many judiciously positioned chromophores,

will enable a more detailed understanding of photoinduced charge-separation processes in natural and

artificial systems.

2.7 Experimental section

All reagents and solvents were used as received or purified using standard procedures. C60 was

purchased from BuckyUSA. NMR spectra were recorded on a Varian Unity Inova and a Varian Unity Plus at frequencies of 500 and 125 MHz for 1H and 13C nuclei; a Varian Mercury Vx at frequencies of 400 and 100 MHz for 1H and 13C nuclei or Varian Gemini 2000 at frequencies of 300 and 75 MHz for 1H and 13C nuclei, respectively. Tetramethylsilane (TMS) was used as an internal standard for 1H NMR and CDCl3, CD3COCD3 or CS2 for 13C NMR. Infrared (FT-IR) spectra were recorded on a Perkin-Elmer Spectrum One UATR FT-IR. Elemental analyses were preformed on a Perkin Elmer 2400 series II CHN Analyzer. Matrix assisted laser desorption ionization time-of-flight mass spectrometry (MALDI-TOF MS) was performed on a Perseptive DE PRO Voyager MALDI-TOF mass spectrometer using a dithranol matrix. All HPLC analyses were performed on a Hewlett Packard HP LC-Chemstation 3D (HP 1100 Series) with DAD detection using an Inertsil® 5 Si column (250x3 mm). A Shimadzu LC-10AT system combined with a Polymer Laboratories MIXED-D column (Particle size: 5µm; Length/I.D. (mm): 300 × 7.5) and UV detection was employed for size exclusion chromatography (SEC), using CHCl3 as an eluent (1 mL/min). (E,E)-4-{4-(4-methyl-2,5-bis[(S)-2-methylbutoxy]styryl)-2,5-bis[(S)-2-methylbutoxy]styryl}-2,5-bis[(S)-2-methylbutoxy}-5-benzaldehyde-dimethylacetal (2). Amberlite IR 120 (1.5 g), trimethyl orthoformate (20 mL) and 1 (1.2 g, 1.78 mmol) where added to 100 mL methanol. The suspension was stirred under an argon atmosphere at 70 °C for 2 h. The reaction mixture was cooled to room temperature and 1.5 g of Na2CO3 was added. The suspension was filtered and the solvent was removed in vacuo to yield 1.38 g (91%) of 2, which was used without further purification. 1H NMR (CDCl3, 400 MHz): δ (ppm) 7.50 (d, 1H),7.49 (s, 2H), 7.44 (d, 1H), 7.18 (s, 1H), 7.17 (s, 1H), 7.16 (s, 1H), 7.10 (s, 1H), 7.08 (s, 1H), 6.73 (s, 1H), 3.92-3.74 (m, 12H), 3.42 (s, 6H), 2.24 (s, 3H), 1.98-1.88 (m, 6H), 1.69-1.54 (m, 6H), 1.39-1.26 (m, 6H), 1.10-0.96 (m, 36 H); 13C NMR (CDCl3, 100 MHz): δ (ppm) 151.69, 151.10, 150.96, 150.67, 150.45, 128.17, 127.71, 127.57, 126.98, 126.56, 125.21, 123.19, 123.06, 122.54, 121.71, 116.32, 111.80, 110.14, 109.94, 109.26, 108.39, 99.75, 74.69, 74.35, 74.28, 74.23, 73.73, 73.38, 54.34, 35.11, 35.08, 35.05, 34.96, 34.89, 26.36, 26.26, 26.21, 16.79, 16.70, 16.40, 11.46, 11.37, 11.33. N-(4-Diphenylaminophenyl)-N-phenyl-3-aminobenzaldehyde (4). To a tube fitted with a magnetic stirrer was added 3 (0.8 g, 2.38 mmol), 3-bromobenzaldehyde (1.32 g, 7.14 mmol), Pd2(dba)3 (0.022g, 0.024 mmol), BINAP (0.044 g, 0.071 mmol) and Cs2CO3 (1.16 g, 3.57 mmol). After purging with argon, freshly distilled toluene (11.9 mL) was added. The reaction mixture was heated at 100 °C under Ar atmosphere. After 48 h Pd2(dba)3 (0.022g, 0.024 mmol), BINAP (0.044 g, 0.071 mmol) and Cs2CO3 (1.16 g, 3.57 mmol) were added and the mixture was heated for another 72 h. After cooling to room temperature, the reaction mixture was filtered over Celite 545 and concentrated in vacuo. Column chromatography (SiO2, heptane/CH2Cl2 1:1, Rf = 0.4) and evaporation to dryness from heptane yielded 0.643 g (62 %) of a yellow powder. IR (UATR) ν (cm-1) 3034, 2923, 2851, 1698, 1583, 1485, 1277, 1263, 749, 690, 626. 1H NMR (CDCl3, 400 MHz): δ (ppm) 9.92 (s, 1H), 7.56 (t, 1H), 7.43(dt, 1H), 7.37 (t, 1H), 7.34-7.23 (m, 7H), 7.13-7.11 (m, 6H), 7.08-6.96 (m, 7H); 13C NMR (CDCl3, 100 MHz): δ (ppm) 192.24, 149.14, 148.00, 147.45, 144.17, 142.11, 137.91, 129.98, 129.70, 129.45, 128.18, 126.37, 125.41, 124.62, 124.20, 123.67, 123.07, 122.91, 122.66. Anal. Cald for C31H24N2O: C, 84.5; H, 5.5; N, 6.4. Found: C, 84.1; H, 5.1; N, 6.1. (E)-N,N’-(Diphenyl)-N’-(4-diphenylaminophenyl)-3-aminobenzaldimine (5). To a suspension of 4 (0.55 g, 1.24 mmol) in ethanol (50 mL), was added aniline (0.14g, 1.50 mmol). The reaction mixture was heated at 85 °C for 4 h. After cooling to room temperature the product precipitated slowly

Page 54: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced multistep energy and electron transfer in a molecular triad

43

from the ethanol. The product was obtained as 0.508 g (79%) of a yellow powder after washing with ethanol. IR (UATR) ν (cm-1) 3034, 1628, 1589, 1486, 1267, 751, 693. 1H NMR (CDCl3, 400 MHz): δ (ppm) 8.34 (s, 1H), 7.61 (t, 1H), 7.53 (dt, 1H), 7.38-7.31(m, 3H), 7.28-7.10 (m, 16H), 7.03-6.97 (m, 7H); 13C NMR (CDCl3, 100 MHz): δ (ppm) 160.31, 151.99, 148.47, 147.78, 147.56, 143.29, 142.34, 137.40, 129.62, 129.33, 129.19, 129.09, 126.32, 125.91, 125.66, 125.26, 123.95, 123.88, 123.53, 122.82, 122.52, 122.44, 120.85. Anal. Cald for C37H29N3: C, 86.2; H, 5.7; N, 8.1. Found: C, 85.8; H, 5.3; N, 8.0. MALDI-TOF MS (Mw = 515.65) m/z = 515.23 [M]+. (E,E,E)-4-[4-{4-[N-(4-Diphenylaminophenyl)-N-(phenyl)-3-aminostyryl]2,5-bis[(S)-2-methylbutoxy]styryl}-2,5-bis[(S)-2-methylbutoxy]styryl}-2,5-bis[(S)-2-methylbutoxy]benzaldehyde (6). Schiff base 5 (0.450 g, 0.873 mmol) and acetal 2 (0.775 g, 0.873 mmol) were dissolved in DMF (5 mL). The mixture was heated to 80 °C under an argon atmosphere and potassium tert-butoxide (0.352 g, 3.142 mmol) was added. The reaction mixture was stirred for 3 h. After cooling to room temperature, the reaction mixture was poured on ice, washed with HCl 3N and brine, dried over MgSO4 and concentrated in vacuo. Column chromatography (SiO2, toluene/cyclohexane 7:3 Rf = 0.4, and heptane/CH2Cl2 6:2, Rf = 0.2) and evaporation to dryness from heptane yielded 0.795 g (72%) of an orange powder. IR (UATR) ν (cm-1) 3062, 2960, 2917, 2873, 1675, 1589, 1504, 1490, 1422, 1262, 1200, 1039, 968, 744, 693. 1H NMR (CDCl3, 500 MHz): δ (ppm) 10.43(s, 1H), 7.64 (d, 1H), 7.54 (d, 1H), 7.51 (d, 1H), 7.50 (d, 1H), 7.40 (d, 1H), 7.33 (s, 1H), 7.28-6.98 (m, 29H), 3.79 3.98 (m, 12H), 1.98-1.91 (m, 6H), 1.70-1.57 (m, 6H), 1.39-1.31 (m, 6H), 1.13-0.94 (m, 36H); 13C NMR (CDCl3, 100 MHz): δ (ppm) 189.06, 156.43, 151.44, 151.26, 151.04, 150.64, 148.21, 147.84, 142.91, 142.70, 139.13, 135.22, 129.45, 129.19, 129.18, 128.50, 128.45, 127.45, 126.74, 126.48, 126.42, 125.46, 125.30, 123.97, 123.79, 123.74, 123.56, 123.19, 122.98, 122.56, 122.43, 122.40, 122.21, 121.90, 120.29, 110.64, 110.29, 110.17, 109.82, 109.74, 109.64, 74.39, 74.25, 74.11, 74.05, 73.91, 73.69, 35.13, 35.01, 34.93, 34.88, 34.84, 26.38, 26.34, 26.33, 26.18, 16.88, 16.86, 16.81, 16.76, 16.63, 11.52, 11.46, 11.38, 11.32. Analysis Calcd for C85H102N2O7 C, 80.8; H, 8.1; N, 2.2. Found: C, 80.7; H, 7.6; N, 2.2. MALDI-TOF MS (Mw = 1263.75) m/z = 1263.71 [M]+. N-Methyl-2<4-[4-{4-[N’-(4-Diphenylaminophenyl)-N’-(phenyl)]3-aminostyryl]-2,5-bis[(S)-2-methylbutoxy]styryl}-2,5-bis[(S)-2-methylbutoxy]styryl]-2,5-bis[(S)-2-methylbutoxy]>-3,4-fulleropyrrolidine (7). A solution of aldehyde 6 (0.2 g, 0.158 mmol), finely ground N-methylglycine (84.4 mg, 0.949 mmol) and C60 (227 mg, 0.316 mmol) in chlorobenzene (60 mL) was stirred and refluxed in the dark under an atmosphere of dry nitrogen for 18 h. After cooling to room temperature the solvent was removed in vacuo and the remaining residue was purified by column chromatography on silica gel (toluene/cyclohexane: 2/1, Rf = 0.4) to afford triad 7 as a ~1/1 mixture of diastereomers according to HPLC (eluent: toluene/cyclohexane 50/50 v/v; flow: 1 mL/min.; peaks at tr = 9.1 and 10.3 min.). Traces of impurities were effectively removed after a 2nd chromatographic purification on silica gel (CS2/toluene: 1/0 to 7/3 Rf = 0.3) to afford analytically pure material (assay >99.5%) according to HPLC and GPC analysis. The product was precipitated from a concentrated toluene solution with methanol (100 mL) and the resulting solid was washed with methanol (2 × 100 mL) and finally dried in vacuo at 55 ºC. The triad was obtained as a light brown powder (137 mg, 43%). IR (UATR) ν (cm-1) 2957, 2914, 2872, 1589, 1502, 1491, 1264, 1190, 965, 750, 694, 526. 1H NMR (CS2, 500 MHz): δ (ppm) 7.61-6.96 (m, 35H), 5.62 (s, 1H), 5.05 (d, 1H), 4.41 (d, 1H), 4.15-3.70 (m, 12H), 2.92 (s, 3H), 2.16-1.26 (m, 18H), 1.02-1.26 (m, 36H). 13C NMR (CS2, 125 MHz): δ (ppm) 156.50, 154.91, 154.84, 154.18, 154.15, 153.53, 151.61, 150.93, 150.76, 150.71, 150.67, 150.63,147.79, 147.40, 147.06, 146.60, 146.57, 146.52, 146.13, 146.07, 146.03, 146.00, 145.93, 145.89, 145.85, 145.75, 145.50, 145.47, 145.37, 145.26, 145.19, 145.07, 145.04, 145.01, 144.98, 144.91, 144.50, 144.37, 144.31, 144.15, 142.92, 142.86, 142.51, 142.46, 142.43, 142.40, 142.35, 142.12, 142.10, 142.06, 141.99, 141.98, 141.95, 141.92, 141.79, 141.63, 141.53, 141.51, 140.06, 139.98, 139.69, 139.51, 139.07, 136.35, 135.99, 135.88, 135.86, 134.49, 134.44, 129.42, 129.20, 129.17, 127.99, 127.68, 127.66, 127.40, 127.27, 126.88, 126.30, 125.24, 125.14, 124.85, 124.80, 123.79, 123.70, 123.61, 123.29, 122.69, 122.56, 122.52, 122.47, 122.29, 121.96, 120.39, 114.29, 114.16, 110.04, 109.63, 109.40, 109.22, 108.71, 108.61, 76.53, 75.43, 75.40, 73.97, 73.91, 73.69, 73.59, 73.52, 73.44, 72.95, 72.91, 69.70, 68.98, 40.02, 35.41, 35.31, 35.29, 35.25, 35.20, 35.18, 35.17, 26.86, 26.84, 26.80, 26.49, 17.14, 17.09, 17.06, 17.02, 17.00, 16.80, 12.04, 12.03, 12.00, 11.94, 11.89, 11.85. Analysis Calcd for C147H107N3O6 C, 87.7; H, 5.4; N, 2.1. Found: C, 87.8; H, 5.0; N, 2.1. MALDI-TOF MS (Mw = 2011.44) m/z = 2011.18 [M]+. (E,E)-2-{4-(4-methyl-2,5-bis[(S)-2-methylbutoxy]styryl)-2,5-bis[(S)-2-methylbutoxy]styryl}-1,4-bis[(S)-2-methylbutoxy]-5-deuteriobenzene (9). n-BuLi 1.6 M (0.13 mL, 0.21 mmol) was added dropwise to a solution of 8 (0.15 g, 0.16 mmol) in freshly distilled diethyl ether (3 mL) at –10° C. The reaction mixture was stirred for 5 minutes. After the addition of

Page 55: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 2

44

D2O (0.5 mL) the cooling bath was removed and the reaction mixture was stirred for 2 h at room temperature. The mixture was estracted with diethyl ether dried over MgSO4 and concentrated in vacuo. Column (SiO2; heptane/CH2Cl2: 1/1, Rf = 0.3) and recrystallization from ethanol yielded 54 mg (41%) of the product as yellow crystals. 1H NMR (CDCl3, 400 MHz): δ (ppm) 7.49 (d, 1H), 7.48 (s, 2H), 7.44 (d, 1H), 7.19(s, 1H), 7.18 (s, 1H), 7.17 (s, 1H), 7.10 (s, 1H), 6.82 (s, 1H), 6.72(s, 1H), 3.92-3.71 (m, 12H), 1.99-1.82 (m, 6H), 1.68-1.54 (m, 6H), 1.38-1.23 (m, 6H), 1.20-0.86 (m, 36H); 13C NMR (CDCl3, 100 MHz): δ (ppm) 151.68, 151.14, 150.94, 150.90, 150.44, 128.21, 127.61, 127.54, 127.10, 125.23, 123.24, 122.98, 122.82, 121.74, 116.31, 113.96, 111.58, 110.36, 109.85, 108.34, 74.68, 74.46, 74.30, 74.26, 73.59, 73.38, 35.11, 35.06, 35.04, 34.97, 34.80, 26.36, 26.34, 26.26, 26.19, 16.80, 16.70, 16.56, 16.40, 11.46, 11.40, 11.37, 11.31. Anal. Cald for C53H79DO6: C, 77.8; H, 10.2. Found: C, 77.5; H, 9.1. MALDI-TOF MS (Mw = 813.96) m/z = 813.53[M]+.

(E,E,E)-2-[4-{4-[N-(4-Diphenylaminophenyl)-N-(phenyl)-3-aminostyryl]-2,5-bis[(S)-2-methylbutoxy]styryl}-2,5-bis[(S)-2-methylbutoxy]styryl]-1,4-bis[(S)-2-methylbutoxy]-5-deuteriobenzene (10). Schiff base 5 (0.025 g, 0.049 mmol) and 9 (0.40 g, 0.049 mmol) were dissolved in DMF (5 mL). The mixture was heated at 80 °C under an argon atmosphere and potassium tert-butoxide (0.020 g, 0.176 mmol) was added. The reaction mixture was stirred for 5 h at 80 °C. After cooling to room temperature, the reaction mixture was poured onto ice and washed with HCl 3N and brine. The organic layer was dried over MgSO4 and concentrated in vacuo. Column chromatography (SiO2, toluene/cyclohexane 7:3, Rf = 0.6) and recyrstallization from hexane, heptane and a few drops of CH2Cl2 yielded 30 mg of 10 (50%) as a yellow powder. IR (UATR) ν (cm-1) 2960, 2917, 2873, 1588, 1504, 1489, 1255, 1205, 1042, 970, 744, 693. 1H NMR (CD3COCD3, 400 MHz): δ (ppm) 7.61 (s, 2H), 7.60 (d, 1H), 7.55 (d, 1H), 7.47(d, 1H), 7.35-6.95 (m, 30H), 4.01-3.76 (m, 12H), 2.09-1.83 (m, 6H), 1.74-1.55 (m, 6H), 1.45-1.26 (m, 6H), 0.95-1.15 (m, 36H); 13C NMR (100 MHz): δ (ppm) 155.25, 152.89, 152.84, 152.72, 150.10, 149.58, 149.54, 144.87, 144.58, 141.06, 131.32, 131.03, 130.98, 130.19, 129.47, 129.13, 128.96, 128.91, 128.15, 127.30, 127.04, 125.39, 125.35, 125.16, 124.93, 124.87, 124.73, 124.68, 124.37, 124.34, 124.30, 123.17, 122.16, 115.71, 113.14, 112.34, 111.95, 111.58, 111.39, 75.61, 75.56, 75.53, 75.36, 74.72, 36.76, 36.68, 36.63, 36.44, 27.85, 27.81, 27.60, 18.04, 17.98, 17.95, 17.92, 17.60, 12.63, 12.60, 12.58, 12.52, 12.39. Anal. Cald for C85H102N2O7: C,81.6; H, 8.2; N, 2.3. Found: C, 81.2; H, 7.7; N, 2.2. MALDI-TOF MS (Mw = 1236.74) m/z = 1235.62 [M]+.

Electrochemistry. Cyclic voltammograms were measured in 0.1 M tetrabutylammonium hexafluorophosphate (TBAPF6) as a supporting electrolyte in dichloromethane (or THF) using a Potentioscan Wenking POS73 potentiostat. The working electrode was a Pt disk (0.2 cm2), the counter electrode was a Pt plate (0.5 cm2), and a saturated calomel electrode (SCE) was used as reference electrode, calibrated against Fc/Fc+ (+0.43 V).

Absorption and Photoluminescence. UV/visible/near-IR absorption spectra were recorded on a Perkin Elmer Lambda 900 spectrophotometer. Fluorescence spectra were recorded on an Edinburgh Instruments FS920 double-monochromator spectrometer and a Peltier-cooled red-sensitive photomultiplier. Time-correlated single photon counting. Time-correlated single photon counting fluorescence studies were performed using an Edinburgh Instruments LifeSpec-PS spectrometer. The LifeSpec-PS comprises a 400 nm picosecond laser (PicoQuant PDL 800B) operated at 2.5 MHz and a Peltier-cooled Hamamatsu micro-channel plate photomultiplier (R3809U-50). Lifetimes were determined from the data using the Edinburgh Instruments software package.

Near steady state PIA. Solutions were prepared in a nitrogen-filled glove box in order to exclude interference of oxygen during measurements. The PIA spectra were recorded between 0.5 and 3.0 eV by exciting with a mechanically modulated cw Ar ion laser (λ = 458 or 528 nm, 275 Hz) pump beam and monitoring the resulting change in transmission of a tungsten-halogen probe light through the sample (∆T) with a phase-sensitive lock-in amplifier after dispersion by a grating monochromator and detection, using Si, InGaAs, and cooled InSb detectors. The pump power incident on the sample was typically 25 mW with a beam diameter of 2 mm. The PIA (-∆T/T ≈ ∆αd) was directly calculated from the change in transmission after correction for the PL, which was recorded in a separate experiment. PIA and PL spectra were recorded with the pump beam in a direction almost parallel to the direction of the probe beam. The solutions were studied in a 1 mm near-IR grade quartz cell at room temperature. Solvents for PIA measurements were distilled under nitrogen before use. The solid-state measurements were performed on films, drop cast from chloroform solution, on quartz substrate and held at 80 K in an Oxford continuous flow cryostat.

Page 56: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced multistep energy and electron transfer in a molecular triad

45

Transient subpicosecond photoinduced absorption. The femtosecond laser system used for pump-probe experiments consists of an amplified Ti/sapphire laser (Spectra Physics Hurricane). The single pulses from a cw mode-locked Ti/sapphire laser were amplified by a Nd-YLF laser using chirped pulse amplification, providing 150 fs pulses at 800 nm with an energy of 750 µJ and a repetition rate of 1 kHz. The pump pulses at 450 nm were created via optical parametric amplification (OPA) of the 800 nm pulse by a BBO crystal into infrared pulses which were then two times frequency doubled via BBO crystals. The probe beam was generated in a separate optical parametric amplification set-up in which 1030 and 1450 nm pulses were created. The pump beam was focused to a spot size of about 1 mm2 with an excitation flux of 1 mJ cm-2 per pulse. For the 1030 and 1450 nm pulses a RG 850 nm cut-off filter was used to avoid contributions of residual probe light (800 nm) from the OPA. The probe beam was reduced in intensity compared to the pump beam by using neutral density filters of OD = 2. The pump beam was linearly polarized at the magic angle of 54.7° with respect to the probe, to cancel out orientation effects in the measured dynamics. The temporal evolution of the differential transmission was recorded using Si or an InGaAs detector by a standard lock-in technique at 500 Hz. Solutions in the order of 2-5 × 10-4 M were excited at 450 nm, i.e. providing primarily excitation of the OPV part within the molecules. 2.8 References and notes 1 (a) The Photosynthetic Reaction Center, Deisenhofer, J., Norris, J. R., Eds.; Academic Press: New York,

1993. (b) Anoxygenic Photosynthetic Bacteria, Blankenschip, R. E., Madigan, M. T., Bauer, C. E., Eds.;

Kluwer Academic Publishers: Dordrecht, 1995.

2 (a) Molecular Electronics, Jortner, J., Ratner, M. Eds.; Blackwell: London, 1997. (b) Lehn, J.-M.,

Supramolecular Chemistry; VCH, Weinheim, 1995. (c) Electron Transfer in Chemistry Vol. I-IV,

Balzani, V., Wiley-VCH, Weinheim, 2001,

3 Tang, C. W. Appl. Phys. Lett. 1986, 48, 183.

4 (a) Nunzi, J.-M.C. R. Physique 2002, 3, 523. (b) Nelson, J. Curr. Opn. Solid State Mater. Sci. 2002, 6, 87.

5 (a) Sariciftci, N. S.; Smilowitz, L.; Heeger, A. J.; Wudl, F. Science 1992, 258, 1474. (b) Yu, G.; Gao, J.;

Hummelen, J. C.; Wudl, F.; Heeger, A. J. Science 1995, 270, 1789.

6 (a) Kraabel, B.; McBranch, D.; Sariciftci, N. S.; Moses, D.; Heeger, A. J. Phys. Rev. B 1994, 50, 18543.

(b) Kraabel, B.; Hummelen, J. C.; Vacar, D.; Moses, D.; Sariciftci, N. S.; Heeger, A. J. J. Chem. Phys.

1996, 104, 4267. (c) Brabec, C. J.; Zerza, G.; Cerullo, G.; De Silvestri, S.; Luzatti, S.; Hummelen, J. C.;

Sariciftci, N. S. Chem. Phys. Lett. 2001, 340, 232.

7 (a) Smilowitz, L.; Sariciftci, N. S.; Wu, R.; Gettinger, C.; Heeger, A. J.; Wudl, F. Phys. Rev. B. 1993, 47,

13835. (b) Meskers, S. C. J.; van Hal, P. A.; Spiering, A. J. H.; van der Meer, A. F. G.; Hummelen, J. C.;

Janssen, R. A. J. Phys. Rev. B. 2000, 61, 9917. (c) Montanari, I.; Nogueira, A. F.; Nelson, J.; Durrant, J.;

Loi, M.-A.; Winder, C.; Sariciftci, N. S.; Brabec, C. J. Appl. Phys. Lett. 2002, 81, 3001.

8 Oligothiophene-fullerene dyads: (a) Benincori, T.; Brenna, E.; Sannicolo, F.; Trimarco, L.; Zotti, G.;

Sozzani, P. Angew. Chem., Int. Ed. Engl. 1996, 35, 648. (b) Effenberger; F.; Grube ,G. Synthesis 1998,

1372. (c) Knorr, S.; Grupp, A.; Mehring, M.; Grube, G.; Effenberger, F. J. Chem. Phys. 1999, 110, 3502.

(d) Yamashiro, T.; Aso, Y.; Otsubo, T.; Tang, H.; Harima, Y.; Yamashita, K. Chem. Lett. 1999, 443. (e)

Fujitsuka, M.; Ito, O.; Yamashiro, T.; Aso, Y.; Otsubo, Y. J. Phys. Chem. A 2000, 104, 4876. (f) van Hal,

P. A.; Knol, J.; Langeveld-Voss, B. M. W.; Meskers, S. C. J.; Hummelen, J. C.; Janssen, R. A. J. J. Phys.

Chem. A 2000, 104, 5974. (g) van Hal, P. A.; Janssen, R. A. J.; Lanzani, G.; Cerullo, G.; Zavelani-Rossi,

M.; De Silvestri, S. Chem. Phys. Lett. 2001, 345, 33. (h) Fujitsuka, M.; Masahura, A.; Kasai, H.; Oikawa,

Page 57: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 2

46

H.; Nakanishi, H.; Ito, O.; Yamashiro, T.; Aso, Y.; Otsubo, T. J. Phys. Chem. B. 2001, 105, 9930. (i)

Fujitsuka, M.; Matsumoto, K.; Ito, O.; Yamashiro, T.; Aso, Y.; Otsubo, T. Res. Chem. Intermed. 2001,

27, 73. (j) Hirayama, D.; Takimiya, K.; Aso, Y.; Otsubo, T. Hasobe, T.; Yamada, H. Imahori, H.;

Fukuzumi, S.; Sakata, Y. J. Am. Chem. Soc. 2002, 124, 532.

9 Oligo(p-phenylene vinylene)-fullerene dyads: (a) Nierengarten, J. F.; Eckert, J. F.; Nicoud, J. F.; Ouali,

L.; Krasnikov, V.; Hadziioannou, G. Chem. Commun. 1999, 617. (b) Eckert, J. F.; Nicoud, J. F.;

Nierengarten, J. F.; Liu, S. G.; Echegoyen, L.; Barigelletti, F.; Armaroli, N.; Ouali, L.; Krasnikov, V. V.;

Hadziioannnou, G. J. Am. Chem. Soc. 2000, 122, 7467. (c) Armaroli, N.; Barigelletti, F.; Ceroni, P.;

Eckert, J.-F.; Nicoud, J.-F.; Nierengarten, J.-F. Chem. Commun. 2000, 599. (d) Peeters, E.; van Hal, P.

A.; Knol, J.; Brabec, C. J.; Sariciftci, N. S.; Hummelen, J. C.; Janssen, R. A. J. J. Phys. Chem. B 2000,

104, 10174. (e) van Hal, P. A.; Janssen, R. A. J.; Lanzani, G.; Cerullo, G.; Zavelani-Rossi, M.; De

Silvestri, S. Phys. Rev. B 2001, 64, 075206. (f) Nierengarten, J.-F.; Armaroli, N.; Accorsi, G.; Rio, Y;

Eckert, J.-F. Chem. Eur. J. 2003, 9, 37.

10 Oligo(thienylene vinylene)-fullerene dyads: (a) Liu, S.-G.; Shu, L.; Rivera, J.; Liu, H.; Raimundo, J.-M.;

Roncali, J.; Gorgues, A.; Echegoyen, L. J. Org. Chem. 1999, 64, 4884. (b) Liu, S.-G.; Martineau, C.;

Raimundo, J.-M.; Roncali, J.; Echegoyen, L. Chem. Commun. 2001, 913. (c) Martineau, C.; Blanchard,

P.; Rondeau, D.; Delaunay, J.; Roncali, J. Adv. Mater. 2002, 14, 283. (d) Apperloo, J. J.; Martineau, C.;

van Hal, P. A.; Roncali, J.; Janssen, R. A. J. J. Phys. Chem. A 2002, 106, 21.

11 Oligoene-fullerene dyads: (a) Imahori, H.; Cardoso, S.; Tatman, D.; Lin, S.; Noss, L.; Seely, G. R.;

Sereno, L.; Chessa de Silber, J.; Moore, T. A.; Moore, A. L.; Gust, D. Photochem. Photobiol. 1995, 62,

1009. (b) Yamazaki, M.; Araki, Y.; Fuijtsuha, M.; Ito, O. J. Phys. Chem. A. 2001, 105, 8615.

12 Miscellaneous: oligomer-fullerene dyads: (a) Segura, J. L.; Gómez, R.; Martín, N.; Luo, C.; Guldi, D. M.

Chem. Commun. 2000, 701. (b) Guldi, D. M.; Swartz, Luo, C.; Gómez, R.; Segura, J. L.; Martín, N. J.

Am. Chem. Soc. 2002, 124, 10875. (c) Guldi, D. M.; Luo, C.; Schwartz, A.; Gómez, R.; Segura, J. L.;

Martin, N.; Brabec, C. J.; Sariciftci, N. S. J. Org. Chem. 2002, 67, 1141. (d) Gu, T.; Tsamouras, D.;

Melzer, C.; Krasnikov, V.; Gisselbrecht, J.-P.; Gross, M.; Hadziioannou, G.; Nierengarten, J.-F. Chem.

Phys. Chem. 2002, 124;

13 van Hal, P. A.; Beckers, E. H. A.; Meskers, S. C. J.; Janssen, R. A. J.; Jousselme, B.; Blanchard, P.;

Roncali, J. Chem. Eur. J. 2002, 8, 5415.

14 For recent reviews see: (a) Gust, D.; Moore, T. A.; Moore, A. L. Acc. Chem. Res. 2001, 34, 40. (b) Guldi,

D. M. Chem. Soc. Rev. 2002, 31, 22.

15 (a) Liddell, P. A.; Kuciauskas, D.; Sumida, J. P.; Nash, B.; Nguyen, D.; Moore, A. L.; Moore, T. A.;

Gust, D. J. Amer. Chem. Soc. 1997, 119, 1400. (b) Imahori, H.; Yamada, K.; Hasegawa, M.; Taniguchi,

S.; Okada, T.; Sakata, Y. Angew. Chem., Int. Ed. Engl. 1997, 36, 2626. (c) Carbonera, D.; Di Valentin,

M.; Corvaja, C.; Agostini, G.; Giacometti, G.; Liddell, P. A.; Kuciauskas, D.; Moore, A. L.; Moore, T.

A.; Gust, D. J. Amer. Chem. Soc. 1998, 120, 4398. (d) Kuciauskas, D.; Liddell, P. A.; Moore, A. L.;

Moore, T. A.; Gust, D. J. Amer. Chem. Soc. 1998, 120, 10880. (e) Tamaki, K.; Imahori, H.; Sakata, Y.;

Nishimura, Y.; Yamazaki, I. Chem. Commun. 1999, 625. (f) Imahori, H.; Yamada, H.; Ozawa, S.; Sakata,

Y.; Ushida, K. Chem. Commun. 1999, 1165. (g) Fujitsuka, M.; Ito, O.; Imahori, H.; Yamada, K.;

Page 58: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced multistep energy and electron transfer in a molecular triad

47

Yamada, H.; Sakata, Y. Chem. Lett. 1999, 721. (h) Imahori, H.; Yamada, H.; Nishimura, Y.; Yamazaki,

I.; Sakata, Y. J. Phys. Chem. B 2000, 104, 2099. (i) Kuciauskas, D.; Liddell, P. A.; Lin, S.; Stone, S. G.;

Moore, A. L.; Moore, T. A.; Gust, D. J. Phys. Chem. B 2000, 104, 4307. (j) Luo, C.; Guldi, D. M.;

Imahori, H.; Tamaki, K.; Sakata, Y J. Amer. Chem. Soc. 2000, 122, 6535. (k) Imahori, H.; Tamaki, K.;

Yamada, H.; Yamada, K.; Sakata, Y.; Nishimura, Y.; Yamazaki, I.; Fujitsuka, M.; Ito, O. Carbon 2000,

38, 1599. (l) Bahr, J. L.; Kuciauskas, D.; Liddell, P. A.; Moore, A. L.; Moore, T. A.; Gust, D. Photochem.

Photobiol. 2000, 72, 598. (m) Imahori, H.; Norieda, H.; Yamada, H.; Nishimura, Y.; Yamazaki, I.;

Sakata, Y.; Fukuzumi, S. J. Amer. Chem. Soc. 2001, 123, 100. (n) Fukuzumi, S.; Imahori, H.; Yamada,

H.; El-Khouly, M. E.; Fujitsuka, M.; Ito, O.; Guldi, D. M. J. Amer. Chem. Soc. 2001, 123, 2571. (o)

Imahori, H.; Tamaki, K.; Guldi, D. M.; Luo, C.; Fujitsuka, M.; Ito, O.; Sakata, Y.; Fukuzumi, S. J. Amer.

Chem. Soc. 2001, 123, 2607. (p) Imahori, H.; Guldi, D. M.; Tamaki, K.; Yoshida, Y.; Luo, C.; Sakata, Y.;

Fukuzumi, S. J. Amer. Chem. Soc. 2001, 123, 6617. (q) Fukuzumi, S.; Imahori, H.; Okamoto, K.;

Yamada, H.; Fujitsuka, M.; Ito, O.; Guldi, D. M. J. Phys. Chem. A 2002, 106, 1903. (r) Liddell, P. A.;

Kodis, G.; De la Garza, L.; Bahr, J. L.; Moore, A. L.; Moore, T. A.; Gust, D. Helv. Chim. Acta 2001, 84,

2765. (s) Ikemoto, J.; Takimiya, K.; Aso, Y.; Otsubo, T.; Fujitsuka, M.; Ito, O. Org. Lett. 2002, 4, 309. (t)

Imahori, H.; Tamaki, K.; Araki, Y.; Hasobe, T.; Ito, O.; Shimomura, A.; Kundu, S.; Okada, T.; Sakata,

Y.; Fukuzumi, S. J. Phys. Chem. A 2002, 106, 2803. (u) Imahori, H.; Tamaki, K.; Araki, Y.; Sekiguchi,

Y.; Ito, O.; Sakata, Y.; Fukuzumi, S. J. Amer. Chem. Soc. 2002, 124, 5165. (v) D'Souza, F.; Deviprasad,

G. R.; Zandler, M. E.; El-Khouly, M. E.; Fujitsuka, M.; Ito, O. J. Phys. Chem. B 2002, 106, 4952. (w)

Liddell, Paul A.; Kodis, Gerdenis; Moore, Ana L.; Moore, Thomas A.; Gust, Devens. J. Amer. Chem.

Soc. 2002, 124, 7668. (x) Kodis, G.; Liddell, P. A.; de la Garza, L.; Moore, A. L.; Moore, T. A.; Gust, D.

J. Mater. Chem. 2002, 12, 2100. (y) Sánchez, L.; Pérez, I.; Martín, N.; Guldi, D. M. Chem. Eur. J. 2003,

9, 2457.

16 Strohrieghl, P.; Jesberger, G.; Heinze, J.; Moll, T. Makromol. Chem. 1992, 193, 909.

17 (a) Siegrist, A. E. Helv. Chim. Acta 1967, 50, 906. (b) Siegrist, A. E.; Meyer, H. R.; Weber, K. Helv.

Chim, Acta 1969, 52, 2521.

18 The bromine has been replaced by a deuterium atom in order to simplify the 1H-NMR spectrum of the

molecule.

19 Murata, Y.; Shine, H. J. J. Org. Chem. 1969, 34, 3368.

20 van Meurs, P. J. High-Spin Molecules of p-Phenylenediamnine Radical Cations, PhD Thesis, Eindhoven

University of Technology, 2002, ISBN 90-386-2574-X.

21 van Hal, P. A.; Beckers, E. H. A.; Peeters, E.; Apperloo, J. J.; Janssen, R. A. J. Chem. Phys. Lett. 2000,

328, 403.

22 Weller, A. Z. Phys. Chem. Neue Folge 1982, 133, 93.

23 The assumption that charges are located at the centers, is of course a simplification of the actual situation

in which charges are delocalized. Especially for C60, the charge is not expected to be at the center, but

rather at the outer surface.

24 Williams, R. M.; Zwier, J. M.; Verhoeven, J. W. J. Am. Chem. Soc. 1995, 117, 4093.

Page 59: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 2

48

25 The PL lifetime of the residual emission at 510 nm contained a contribution of long lifetime (~ 1 ns) and

is in part attributed to residual PL emission by minor impurities in the sample.

26 In toluene the energy level of initial charge separated state [OAn-OPV•+-C60•–] is higher than that of

OAn-OPV-1C60* due to the destabilization in the nonpolar solvent. In such case one can also expect that

initial electron transfer from 1OPV* to C60 occurs to yield OAn-OPV•+-C60•–, which subsequently

recombines to generate OAn-OPV-1C60*. The stepwise pathway would compete with the direct energy

transfer pathway from 1OPV* to C60 to produce the same state. However, in view of the high rate of the

energy transfer (190 fs) it is very unlikely that this process occurs and the same rate of 190 fs has been

found in ODCB, where the OAn-OPV•+-C60•– state is below the OAn-OPV-1C60* state (Ref 9e).

27 The function f(t)= a0(exp(-a1(t-a3) - exp(-a2(t-a3)) + a4) was used to fit the data and all 5 parameters (a0 –

a4) were optimized.

28 (a) Wasielewski, M. R.; Niemczyk, M.P.; Svec, W. A.: Pewitt, E. B. J. Am. Chem. Soc. 1985, 107, 1080.

(b) Asahi, T.; Ohkohchi, M.; Matsusaka, R.; Mataga, N.; Zhang, R. P.; Osuka, A.; Maruyama, K. J. Am.

Chem. Soc. 1993, 115, 5665. (c) Macpherson, A. N.; Liddell, P. A.; Lin, S.; Noss, L.; Seely, G. R.;

DeGraziano, J. M.; Moore, A. L.; Moore, T. A.; Gust, D. J. Am. Chem. Soc. 1995, 117, 7202. (d) Kroon,

K.; Verhoeven, J. W.; Paddon-Row, M. N.; Oliver, A. M. Angew. Chem Int. Ed. Engl. 1991, 30, 1358. (e)

Imahori, H.; El-Khouly, M. E.; Fujitsuka, M.; Ito, O.; Sakata, Y.; Fukuzumi S. J. Phys. Chem. A 2001,

105, 325.

29 Guldi, D. M.; Prato, M. Acc. Chem. Res. 2000, 33, 695.

30 The following kinetic scheme has been used to model the photoinduced absorption at 1030 nm. In this

scheme, the intramolecular recombination of charges in OAn+-OPV-C60– and also charge generation from

electronic states other than the singlet-excited state of the C60 moiety have been neglected.

CRCOPV

kkk

kkk

CSHCSET

↓↓↓

−− →−−→→ −•+•−•+•

6000

60601601 COPVOAnCOPVOAn)S(C)OPV(S

For this model the following set of coupled differential equations can be written describing the change in

the concentration of each of the four species involved over time.

−−−−

+−+−+

+−

=

−−−−

−•+•

−•+•

−•+•

−•+•

]COPV[OAn

]COPV[OAn

)]S([C

)][OPV(S

000

0)(0

00)(

000)(

]COPV[OAn

]COPV[OAn

)]S([C

)][OPV(S

60

60

160

1

1

1

0

0

60

60

160

1

60

CR

CSHCRCS

CSC

ET

ETOPV

k

kkk

kkk

kk

dt

d

The singlet-excited state of the OPV is assumed to be generated by photon absorption at time t = 0 and

thus the relevant boundary value condition is:

Page 60: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced multistep energy and electron transfer in a molecular triad

49

==

−−−−

−•+•

−•+•

0

0

0

)](0)[OPV(S

)0(

]COPV[OAn

]COPV[OAn

)]S([C

)][OPV(S 1

60

60

160

1

t

A solution to the above set of equations is:

))(exp()0]()OPV(S[))](OPV(S[ 011 tkkt ETOPV +−=

[ ]))(exp())(exp()0]()OPV(S[)]()(SC[ 60

6000

001160 tkktkk

kkkk

kt C

CSETOPV

CCSET

OPVET +−++−−

−−+=

[ ]

[ ]

[ ]

+−

−−+−−+

++−−−+−−+

+

+−

−−+−−+

=−− −•+•

))(exp())((

))(exp())((

))(exp())((

)0)](OPV(S[

)](COPVOAn[

1

1010

0

1000

0

1000

1

60

60

60

6060

60

tkkkkkkkkkk

kk

tkkkkkkkkkk

kk

tkkkkkkkkkk

kk

t

CSHCR

CSHCRETOPV

CSHCRC

CS

CSET

CCS

CSHCRC

CSC

CSETOPV

CSET

ETOPV

CSHCRETOPVC

CSETOPV

CSET

[ ]

[ ]

[ ]

+−−+−−+

++−−+−−+

+

+−−+−−+

++−

+−−+−−+−

++−+−−+−−+

+

+−

+−−+−−+−

×=−− −•+•

))()((

))()((

))()((

))(exp())()((

)(exp())()((

))(exp())()((

)0)](OPV(S[)](COPVOAn[

11000

01000

11000

1

11000

0

01000

0

1000

160

60

606060

60

60

60

606060

60

CSHCRCSHCRETOPVC

CSETOPV

CSHCSET

CCSCSHCR

CCS

CCSET

OPVCSHCSET

CSHCRCSHCRETOPVC

CSETOPV

CSHCSET

CSHCR

CSHCRCSHCRETOPVC

CSETOPV

CSHCSET

CCSC

CSCSHCRC

CSC

CSETOPV

CSHCSET

ETOPV

CSHCRCSHCRETOPVC

CSETOPV

CSHCSET

kkkkkkkkkk

kkk

kkkkkkkkkk

kkk

kkkkkkkkkk

kkk

tkkkkkkkkkkkk

kkk

tkkkkkkkkkkkk

kkk

tkkkkkkkkkkkk

kkk

t

In these equations it can be verified quite easily that they satisfy the boundary condition at t = 0.

Furthermore, setting kCS = 0, the yield of C60(S1) (abbrv. φ(C60(S1))) is given by kET/( kET + k0OPV).

Similarly, setting kCR1 = 0, the yield of OAn-OPV•+-C60•– is φ(C60(S1)) kCS /( kCS + k0

C60). Finally, the yield

of OAn• +-OPV-C60•– is: φ (C60(S1)) φ(OAn-OPV•+-C60

•–) kCSH /( kCSH + kCR1).

Page 61: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 2

50

31 Deussen, M; Bässler, H. Chem. Phys. 1992, 164, 247.

32 Jortner, J. J. Chem. Phys. 1976, 64, 4860.

33 Imahori et. al. reported that the electronic coupling in a triad (0.019 cm-1) is much less compared to that

in a dyad (3.9 cm-1), see ref. 15o.

34 (a) Thomas, K. G.; Biju, V.; Guldi, D. M.; Kamat, P. V. George, M. V. J. Phys. Chem. B 1999, 103,

8864. (b) Imahori, H.; Hasobe, T.; Yamada, H.; Kamat, P. V.; Barazzouk, S.; Fujitsuka, M.; Ito, O.;

Fukuzumi, S. Chem. Lett. 2001, 784.

Page 62: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 3

Photoinduced multistep electron transfer in

oligoaniline - oligo(p-phenylene vinylene) -

perylene arrays

Abstract

Multistep energy and electron transfer have been studied in solutions containing two different OAn-

OPV-PERY-OPV-OAn pentads, corresponding to a donor(1)-donor(2)-acceptor-donor(2)-donor(1)

symmetrical arrangement. The pentads have been synthesized by double coupling of an oligoaniline-

oligo(p-phenylene vinylene) (OAn-OPV) dyad to a central perylene diimide (PERY) segment, directly

connected (pentad 1) and via a saturated spacer (pentad 2). Photoexcitation of any of the

chromophores yields in first instance the OAn-OPV•+-PERY•—-OPV-OAn charge separated state. This

state can relax to the ground state via charge recombination or generate the OAn•+-OPV-PERY•–-

OPV-OAn state via a charge shift. In pentad 1, the intramolecular charge separation and charge

recombination are extremely fast. The same processes are slowed down in pentad 2, because the

saturated spacer reduces the electronic coupling between the OPV donor and PERY acceptor. For

pentad 1 in toluene, the OAn-OPV•+-PERY•–-OPV-OAn and the OAn•+-OPV-PERY•–-OPV-OAn states

are nearly isoenergetic and thus there is no driving force for the charge shift process in this apolar

solvent. In more polar media, the charge shift competes with the fast charge recombination to the

ground state with an estimated efficiency of 0.22. The slower charge recombination of OAn-OPV•+-

PERY•–-OPV-OAn to the ground state in pentad 2, with respect to that in pentad 1, is beneficial for

the charge shift process that occurs with an efficiency of 0.28 in polar media. The low electronic

coupling between the OAn and PERY chromophores accounts for the remarkably long-lived OAn•+-

OPV-PERY•–-OPV-OAn geminate ion pair (τ > 1 ns).

Page 63: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 3

52

3. 1 Introduction

Fuelled by intriguing applications in molecular electronics, light harvesting, photocatalysis,

and artificial photosynthesis, photoinduced energy and electron transfer in covalently linked donor-

acceptor molecules attract enormous attention in recent years.1 For practical applications it is often of

interest to combine a high rate for charge separation with a low rate for charge recombination. For this

purpose, multichromophoric arrays have been designed in which the initial photoinduced charge

separation reaction is followed by charge migration reactions along a well-defined redox gradient that

ultimately provides the spatial separation of the photogenerated charges, which is essential to lower

the rate for charge recombination. Various elegantly designed triads, tetrads, pentads, etc. have been

synthesized and investigated in detail.2,3

Multistep electron transfer and associated charge migration is possibly also the origin of the

longevity of charges in solid state blends of donor and acceptor materials used in organic and polymer

solar cells.4 In these cells, excitons, created by the absorption of light, dissociate at the donor-acceptor

interface forming an electron-hole pair. When this electron and hole escape from geminate

recombination and diffuse away from the interface, they can be collected at electrodes as a

photocurrent.

Recent studies on covalently linked oligo(p-phenylene vinylene) (OPV) donors and perylene

diimide (PERY) acceptors in liquid crystalline OPV-PERY-OPV triads,5,6 π-stacks of hydrogen-

bonded OPV-PERY-OPV trimers,7 and alternating (OPV-PERY)n copolymers,8 have revealed that

photoinduced electron transfer can be extremely fast in solution (< 1 ps) and that the lifetime is

generally rather short (<1 ns). Because of their high absorption coefficients in the visible region and

their charge transport properties, OPV9-14 and PERY15-19 chromophores have also attracted interest for

organic solar cells.

With the aim to extend the lifetime of the charge-separated state, two new multichromophoric

molecular arrays incorporating a central OPV-PERY-OPV triad augmented with two p-oligoaniline

(OAn) moieties as additional donor, have been designed (1 and 2 in Figure 3.1). In both donor(1)-

donor(2)-acceptor-donor(2)-donor(1) pentads, the OAn and OPV donors are electronically decoupled

in the ground state by an m-phenylene linkage and operate essentially as isolated redox active

segments. The two pentads differ in the connectivity between the central PERY segment and the OPV

donor. In pentad 1 the OPV and PERY units are directly connected providing a close proximity of the

chromophores and possibly some overlap of the π-electronic system, even though the two

chromophores are not coplanar. In 2 the OPV-PERY distances are increased and their π-overlap is

fully interrupted by saturated bonds. The redox potentials of the OAn, OPV, and PERY segments

favor the OAn•+-OPV-PERY•–-OPV-OAn state as the lowest-energy charge-separated state after

photoexcitation. To elucidate the mechanism of multistep photoinduced electron transfer in these

pentads and to assess the lifetime of the charge separated state, the photophysical properties of 1 and 2

Page 64: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced multistep electron transfer in multichromophoric arrays

53

have been studied using photoluminescence and transient photoinduced absorption spectroscopy in

solvents of different polarity in comparison to those of the isolated chromophores and dyad

combinations. The rates for charge separation, migration, and recombination are rationalized using

Marcus theory and a continuum model for describing the free energy.

OO

O

OO

OOO

O

OO

ONNN N

NN

O

O

O

O

1

OO

O

OO

O

NN

OO

O

OO

ON N

NO

ONN

N O

O

O

O

O

O

2

N N NN

O

O

O

O

OO

O

OO

ON N

O

O

O

O

OO

O

OO

O

OAn OPV PERY

OAn-OPV

Figure 3.1 Chemical structures of pentads 1 and 2 and reference compounds.

3.2 Synthesis

The synthesis of pentads 1 and 2 and the reference compound OAn-OPV are depicted in

schemes 3.1 to 3.3. The synthesis of pentad 1 (Scheme 3.1) starts with the coupling of aldehyde 3,

whose synthesis has been described in chapter 2, to diethyl 4-nitrobenzyl phosphonate 4 via a Wittig-

Horner reaction, affording nitro compound 5. The reduction of the nitro group to the terminal amine 6

was achieved using stannous dichloride and ethanol in ethyl acetate. Condensation of 6 with 3,4,9,10-

perylenetetracarboxy-dianhydride in imidazole and catalytic amounts of Zn(OAc)2 afforded pentad

OAn-OPV-PERY-OPV-OAn 1. Purification of 1 was achieved via extensive column chromatography

and preparative size exclusion chromatography, which resulted in a low yield for this last step. For the

Page 65: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 3

54

synthesis of reference dyad OAn-OPV 8, aldehyde 3 was reacted with diethyl benzylphosphonate 7 in

a Wittig-Horner reaction (Scheme 3.2).

N NOR*

OR*OR*

R*OR*O

R*OO

N NOR*

OR*OR*

R*OR*O

R*OR

N

N NOR*

OR*OR*

R*OR*O

R*OO

O

NO2(EtO)2P

O

O

6. R = NH2

5. R = NO2

1

a

b

c

2

OR* =

43

+

Scheme 3.1. Synthesis of OAn-OPV-PERY-OPV-OAn (1). a. t-BuOK, DMF/THF (2/1), r.t., 2 h, 86%;

b. SnCl2·H2O, EtOH, EtOAc, 95° C, 5 h, 70%; c. Imidazole, Zn(OAc)2, 160 °C, 4 h, 15%.

N NOR*

OR*OR*

R*OR*O

R*OO

N NOR*

OR*OR*

R*OR*O

R*O

(OEt)2P

O

O

3

8

a

OR* =

7

+

Scheme 3.2. Synthesis of OAn-OPV (8). a. t-BuOK, DMF/THF (2/1), r.t., 2.5 h, 73%.

Page 66: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced multistep electron transfer in multichromophoric arrays

55

In pentad 2 (Scheme 3.3) a saturated spacer has been intercalated between the OAn-OPV

segments and the perylene diimide central unit. This saturated spacer was first introduced on the

perylene unit by a double condensation of (S)-(+)-leucinol with 3,4,9,10-

tetracarboxyperyleneanhydride, affording the dihydroxy terminated perylene diimide 10. The amino

function in OAn-OPV 6 was reacted with phosgene in toluene to yield isocyanate 9. Pentad 2 was

synthesized by double reaction of isocyanate 9 with the hydroxyl functions of 10 using dibutyltin

dilaurate as catalyst.

N NOR*

OR*OR*

R*OR*O

R*ONCO

OHNN

OH

O

O

O

O

N NOR*

OR*OR*

R*OR*O

R*ONH2

O

N

O

O

O

N NOR*

OR*OR*

R*OR*O

R*ON

H

O

6

9 10

2

a

b

OR* =

+

2

Scheme 3.3. Synthesis of OAn-OPV-PERY-OPV-OAn (10). a. Phosgene, toluene, 95 °C, 16 h, 100%;

b. Dibutylin dilaurate, CH2Cl2, reflux, 20 h, 47%.

All final compounds were characterized using 1H-NMR spectroscopy, mass spectrometry, and

size exclusion chromatography.

In analogy with a previously studied analogous molecular triad based on OPV-PERY-OPV,6

the pentads and reference dyad were studied with 1H-NMR spectroscopy. Pentad 1 and reference dyad

8 feature the spectral characteristics of molecularly dissolved species when measured in deuterated

dichloromethane. However, pentad 2 features four clearly separated and well-resolved signals for the

perylene protons, whereas in a molecular dissolved state only two different signals are expected

(Figure 3.2a). This difference between the aromatic protons points to the formation of aggregates via

Page 67: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 3

56

intermolecular hydrogen bonding as has been established for the analogous triads.6 The spectrum of

pentad 2 recorded in THF (Figure 3.2b) shows the signals typical for molecular dissolved species and

indicates that intermolecular hydrogen bonding is much less present in THF.

7.88.28.69.09.49.8

N-H

N-H

a

b

1 2 3 4

1-4

7.88.28.69.09.49.8

N-H

N-H

a

b

1 2 3 4

1-4O RORNN

O

O

O

ONH O

NO H

2 1

13

3

4

4

2

R = OPV-OAn

Figure 3.2. 1H-NMR spectra of pentad 2 in dichloromethane (a) and THF (b).

The 1H-NMR results indicate that pentad 1 and reference dyad 8 are molecular dissolved in

most organic solvents of intermediate polarity. Pentad 2, however, will dimerize or aggregate in a

variety of solvents due to intermolecular hydrogen bonding of the urethane functionalities. This has

implications for the opto-electronic behavior of pentad 2 and therefore the solvent for 2 must be

carefully selected. Tetrahydrofuran seems to be a solvent that allows for the study of pentad 2 in its

molecularly dissolved state.

3.3 Electronic properties and energetic considerations

Absorption spectroscopy. The absorption spectrum of pentad 1 in toluene solution (Figure

3.3 left) exhibits two strong absorption bands, one centered at 327 nm and the other featuring vibronic

fine structure with maxima at 464, 494 and 532 nm. For comparison, the absorption spectra of the

corresponding individual chromophores OAn, OPV and PERY as well as that of dyad OAn-OPV are

also plotted in Figure 3.3. Whereas all chromophores contribute to the absorption of the UV region of

the pentad, the absorption in the visible region is dominated by the π-π* transitions of the OPV and

PERY chromophores. The vibronic fine structure typical for the PERY chromophore is still present in

the visible region.

Page 68: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced multistep electron transfer in multichromophoric arrays

57

300 400 500 6000.0

0.5

1.0

1.5N

orm

aliz

ed A

bsor

banc

es (

O.D

.)

Wavelength (nm)300 400 500 600

0.0

0.3

0.6

Abs

orba

nce

(O. D

.)

Wavelength (nm)

Figure 3.3. Left: UV/Visible absorption spectra of the OAn-OPV-PERY-OPV-OAn pentad 1 (solid

line) and model compounds OAn-OPV (solid squares), OAn (open squares), OPV (open triangles),

and PERY (open circles) recorded in toluene solution. Right: UV-Visible absorption spectra of the

OAn-OPV-PERY-OPV-OAn pentad 2 in tetrahydrofuran (solid line), dichloromethane (dashed line)

and toluene (dotted line) solutions.

The UV-visible absorption spectrum of pentad 2 (Figure 3.3 right) recorded in

tetrahydrofuran, consists of the overlapping absorptions of the individual chromophores OAn, OPV

and PERY. However, the low energy absorption band of the PERY at 532 nm exhibits a lower

intensity in pentad 2 with respect to pentad 1, suggesting aggregation, most probably via π-π

interactions. In toluene and dichloromethane the absorption spectra show a further reduction of the

intensity of the 532 nm peak and a weak contribution above 550 nm, which indicates aggregation of

the pentads. In these solvents, a combination of intermolecular hydrogen bonding and π-π interactions

is probably responsible for the aggregation.6

Electrochemistry. The electrochemical properties of pentads 1 and 2 and of the reference

compounds OAn, OPV, and PERY were investigated using cyclic voltammetry. The redox potentials

are collected in Table 3.1. The cyclic voltammogram of pentad 1 shows two reversible reduction

waves at –0.53 and –0.76 V, corresponding to the reduction of the PERY acceptor, and four reversible

oxidation waves of the OAn (+0.53 V and +1.03 V) and the OPV (+0.76 and 0.96 V) donor moieties

(potentials are given vs. SCE, calibrated against Fc/Fc+, recorded in dichloromethane with 0.1 M

TBAPF6) (Table 3.1). In dichloromethane pentad 2 also exhibits four oxidation waves at +0.53 and

+1.07 V associated with the OAn segment and at +0.74 and +0.90 V corresponding to the OPV

segment. In this solvent no reduction waves could be observed for pentad 2. The measurement of the

reduction potentials in dichloromethane was most likely hampered by the aggregation via hydrogen

Page 69: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 3

58

bonding. In THF, however, the cyclic voltammogram of pentad 2 exhibits the two reduction waves of

the PERY moiety at -0.60 and –0.89 V.

Table 3.1 reveals that the redox potentials of the OAn and OPV donors are hardly affected by

their linkage to the other units. In contrast, the reduction potentials of the PERY acceptor depend on

the substitution of the imide functionalities. The reduction potentials of 1 and 2 are shifted by +0.12

and +0.05 V respectively with respect to that of the PERY reference compound (-0.65 V). For pentad

2 the shift is less, owing to the similar alkyl substitution as in the PERY reference. In pentad 1, the

overlap of the π-electronic clouds of the OPV and PERY chromophores is possible, although the first

phenyl ring of the OPV moiety will not be coplanar with the PERY unit.

Table 3.1. One-electron redox potentials (E0) of OAn, OPV, PERY, OAn-OPV, and OAn-OPV-PERY-

OPV-OAn (vs. SCE) calibrated with Fc/Fc+ (in dichloromethane with 0.1 M TBAPF6).

Compound E0red (V) E0

ox (V)

OAn 0.53 / 1.02

OPV 0.73/ 0.80

PERY -0.65/-0.85

Pentad 1 -0.53/-0.76 0.53 / 0.76/ 0.96/ 1.07

Pentad 2 -0.60/-0.89a 0.53 / 0.74/ 0.90/ 1.07

a Measured in THF.

Energetic considerations. The scheme in Figure 3.4 illustrates the various photophysical

processes that may occur in the pentads upon illumination. Excitation of one of the redox-active

chromophores generates locally excited states, e.g. the OAn-OPV-1PERY*-OPV-OAn singlet excited

state. Exergonic processes such as energy transfer (ET) and charge separation (CS) might compete

with intrinsic decay processes generating cor OAn•+-OPV•–-PERY-OPV-OAn charge-separated states.

These charge-separated states either decay to the ground state via charge recombination (CR) or

evolve via charge shift (CSH) to the more stable OAn•+-OPV-PERY•–-OPV-OAn state.

Page 70: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced multistep electron transfer in multichromophoric arrays

59

1OAn*-OPV-PERY1OAn*-OPV-PERY

OAn-1OPV*-PERYOAn-1OPV*-PERY

OAn-OPV-PERYOAn-OPV-PERY

OAn-OPV•+-PERY• -

OAn•+-OPV-PERY• -

OAn•+-OPV•--PERYOAn-OPV-1PERY*OAn-OPV-1PERY*kET2

kCS1

kCR1

kCR2

kCSH1

kPLOPV

kET1

kISCOPV

kPLPERY

Energy

kCS3

kCR3

kCS2

OAn-3OPV*-PERY

kCSH2

kCS4

kPLOAn

Figure 3.4. Schematic energy levels (in THF) and photoinduced processes in OAn-OPV-PERY-OPV-

OAn pentads 1 and 2. (In the scheme the pentads are defined as OAn-OPV-PERY for simplification)

The rate constants are collected in Table 3.2. The fastest processes have been highlighted in black.

The energies of the singlet-excited states are at 3.40, 2.39 and 2.33 eV for the OAn, OPV and

PERY chromophores respectively, as extracted from absorption spectroscopy. In contrast to the

singlet- excited states, the energy of the charged separated states is strongly influenced by the polarity

of the medium. The Weller equation (Eq. 3.1) provides a means of calculating the energy of the

charged states in function of the polarity of the solvent and distance between chromophores. 20

( ) ( )( )

+−−−−=∆ −+

sref0

2

ccs0

2

00redox0 1111

84AD

εεπεεπε rr

e

R

eEEEeG (3.1)

In this equation, Eox(D) and Ered(A) are the oxidation and reduction potentials of the donor

and acceptor molecules or moieties measured in a solvent with relative permittivity εref, E00 is the

Page 71: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 3

60

energy of the excited state from which the electron transfer occurs, and Rcc is the center-to-center

distance of the positive and negative charges in the charge separated state. The radii of the positive

and negative ions are given by r+ and r– and εs is the relative permittivity of the solvent, -e is the

elemental charge, and ε0 is the vacuum permittivity.

Using equation 3.1 the change in free energy of the different processes of pentads 1 and 2 in

solutions of low to high polarity, i.e. toluene (ε = 2.38), chlorobenzene (ε = 5.72), THF (ε = 7.51), and

o-dichlorobenzene (ε = 9.93), have been calculated (Table 3.2). For the calculation the radius of the

negative perylene radical anion, PERY•–, was set to r– = 4.7 Å8 and that of the radical cations of OPV

and OAn to r+ = 5.4521 and 4.8 Å (Chapter 2) respectively. The Rcc distances were determined,

assuming that the charges are located at the centers of the OAn, OPV, and PERY moieties.22 The Rcc

distances were, in pentad 1, 21 and 41 Å for the OAn-OPV•+-PERY•–-OPV-OAn and OAn•+-OPV-

PERY•–-OPV-OAn charge-separated states respectively. In pentad 2 the relative position between the

OAn-OPV segment and the PERY central unit is not constant due to the relatively flexible character

of the spacer separating them. The upper limit for the Rcc distances in pentad 2 (27 and 44 Å for OAn-

OPV•+-PERY•–-OPV-OAn and OAn•+-OPV-PERY•–-OPV-OAn respectively) have been determined

in a conformation in which all chromophores are coplanar.

The experimental and estimated energies of the various neutral and charge-separated states of

the pentads are depicted in Figure 3.4, assuming THF as the medium. In the energetic scheme, the

dominant photophysical processes are depicted in black to highlight them from the slower and

therefore less probable processes, which are depicted in gray. The slow processes are the intrinsic

decay to the ground state of the individual chromophores (kPLOAn, kPL

OPV and kPLPERY), the intersystem

crossing to the triplet state for the OPV chromophore (kISCOPV), and the formation of the OAn•+-OPV•–

-PERY-OPV-OAn charge-separated state either from the 1OAn*-OPV-PERY-OPV-OAn or OAn-1OPV*-PERY-OPV-OAn singlet excited states.23 The change in free energy for the fast processes is

collected in Table 3.2. Here, the charge separation (CS1) process refers to the reaction taking place

from the lowest singlet excited state, i.e. OAn-OPV-1PERY*-OPV-OAn → OAn-OPV•+-PERY•–-

OPV-OAn. This separated state can also originate directly from the OPV singlet excited state (kCS2),

however, energy transfer to generate the PERY singlet excited (kET2) state is likely to compete with

this process.

The charge separation process, charge shift and charge recombination (CS1, CSH1, CR1, and

CR2) are exergonic in all solvents, with the exception that the driving force for charge shift is zero in

toluene. Similar energy values are obtained for pentads 1 and 2 because the only difference taken in

consideration was the Rcc parameter.

Page 72: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced multistep electron transfer in multichromophoric arrays

61

Table 3.2. Change in free energy (∆G0) with reference to the lowest singlet excited state,

reorganization energy (λ), and barrier (∆G‡) for charge separation (CS1, OAn-OPV-1PERY*-OPV-

OAn → OAn-OPV•+-PERY•–-OPV-OAn), charge recombination (CR1, OAn-OPV•+-PERY•–-OPV-

OAn→ OAn-OPV-PERY-OPV-OAn), charge shift (CSH1, OAn-OPV•+-PERY•–-OPV-OAn→ OAn•+-

OPV-PERY•–-OPV-OAn), and charge recombination (CR2, OAn•+-OPV-PERY•–-OPV-OAn→ OAn-

OPV-PERY-OPV-OAn) in toluene (TOL), chlorobenzene (CB), tetrahydrofuran (THF), and o-

dichlorobenzene (ODCB) as determined using Eq. 3.1 and 3.3.

Pentad 1 Pentad 2

Reaction Solvent ∆G0

(eV) λ

(eV) ∆G‡

(eV) ∆G0

(eV)λ

(eV) ∆G‡

(eV) CS1 TOL -0.45 0.36 0.006 -0.34 0.36 0.000

CB -0.98 0.85 0.005 -0.90 0.89 0.000 THF -1.07 0.94 0.004 -1.00 0.99 0.000 ODCB -1.14 0.98 0.007 -1.08 1.03 0.001

CR1 TOL -1.88 0.36 1.625 -1.99 0.36 1.846 CB -1.35 0.85 0.072 -1.43 0.89 0.080 THF -1.26 0.94 0.026 -1.33 0.99 0.029 ODCB -1.19 0.98 0.011 -1.25 1.03 0.013

CSH1 TOL 0.00 0.35 0.073 0.00 0.35 0.060 CB -0.16 0.83 0.135 -0.16 0.83 0.135 THF -0.18 0.92 0.148 -0.18 0.83 0.128 ODCB -0.20 0.95 0.149 -0.19 0.95 0.152

CR2 TOL -1.85 0.37 1.474 -1.93 0.37 1.636 CB -1.19 0.98 0.011 -1.26 0.99 0.019 THF -1.08 1.10 0.000 -1.15 1.11 0.000 ODCB -0.99 1.18 0.008 -1.06 0.99 0.001

3.4 Photophysical processes in solution

3.4.1 Photoluminescence spectroscopy

The quenching of the photoluminescence (PL) of any of the chromophores indicates that

faster processes are occurring than the intrinsic decay to the ground state from the singlet-excited

state. As shown in Figure 3.4, the competitive processes are energy and charge transfer. Whereas the

quenching of the PL of OAn and OPV chromophores can be caused by energy and electron transfer

(ET1, CS4, ET2, CS3, and CS2, Figure 3.4), the quenching of the PERY acceptor emission gives

direct evidence of charge separation taking place (CS1), because the PERY S1 state is the lowest-

energy singlet-excited state.

In pentad 1, photoexcitation of either the OAn or OPV chromophore in toluene solution

results in a very weak emission that is associated with the OPV chromophore (Figure 3.5, right).

When compared to the reference OPV compound, the PL of the OPV in the pentad is quenched by a

Page 73: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 3

62

factor of Q = 50. Selective photoexcitation of the PERY chromophore at 530 nm in pentad 1 gives

also a very weak emission, that corresponds to the PERY chromophore and that is quenched by a

factor of Q = 3000 with respect to the PERY reference (Figure 3.5, right). The excitation spectra of

the observed residual photoluminescence at 519 and 540 nm do not superimpose with the absorption

spectrum of the pentad but instead coincide with the absorption spectrum of the OAn-OPV dyad

(Figure 3.5, left). The features corresponding to the PERY chromophore are absent in the excitation

spectra meaning that the contribution of acceptor excitation to the residual fluorescence is negligible.

The fact that the excitation spectrum of the residual OPV emission corresponds to the OAn-OPV

segments gives evidence that energy transfer occurs from the OAn to the OPV segment (ET1, Figure

3.4). Similar quenching factors have been measured in solvents of higher polarity like o-

dichlorobenzene. The photoluminescence of the PERY chromophore is nearly completely quenched

in pentad 1, meaning that the quantum yield for charge separation to generate the OAn-OPV•+-

PERY•–-OPV-OAn charge-separated state is close to unity and that this photophysical process is

extremely fast in all solvents. The OAn-OPV•+-PERY•–-OPV-OAn in itself may subsequently go

through a charge shift (CSH) to generate the energetically more favorable OAn•+-OPV-PERY•–-OPV-

OAn state.

300 400 500

0

1

Nor

mal

ized

inte

nsity

Wavelength (nm)500 600

0

100

200

300

x 200x 20

Inte

nsity

(a.

u.)

/ 10

4

Wavelength (nm)

Figure 3.5. Left: Absorption (solid line) and excitation spectra of the 519 (open circles) and 540 nm

emissions (open squares) in toluene solution. Right: PL spectra of the OPV (solid circles) and PERY

(solid squares) reference compounds and of pentad 1 after selective excitation of the OPV (open

circles) and the PERY (open squares) chromophores at 400 and 530 nm, in toluene solution.

For pentad 2 the PL of the OPV and PERY chromophores are also strongly reduced, with

quenching factors of Q = 66 and 300 respectively (Figure 3.6, right). Photoexcitation of OAn or OPV

results in a weak fluorescence that can be attributed to the sum of the OPV and PERY emissions. This

result is an indication that energy transfer is occurring from the OAn, to the PERY via the OPV

chromophore (ET1 and ET2, Figure 3.4). In pentad 2 the emission of the PERY acceptor is 10 times

Page 74: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced multistep electron transfer in multichromophoric arrays

63

more intense than in pentad 1, indicating that the charge separation process (CS1, Figure 3.4) must be

10 times slower for pentad 2 than for pentad 1, due to the saturated spacer that separate donor and

acceptor.8 The excitation spectra of both the OPV donor and PERY acceptor coincide with the

absorption spectrum of pentad 2 (Figure 3.6, left), consistent with the multistep energy transfer

mentioned above (processes ET1 and ET2, Figure 3.4).

300 400 500 6000

1

Nor

mal

ized

inte

nsity

Wavelength (nm)

500 6000

100

200

x 20

Inte

nsity

(co

unts

) / 1

04

Wavelength (nm)

Figure 3.6. Left: Absorption (solid line) and excitation of the 534 nm emission (open circles) spectra

of pentad 2 in THF solution. Right: PL spectra of the OPV (solid circles) and PERY (solid squares)

reference compounds and of pentad 2 after selective excitation of the OPV (open circles) and the

PERY (open squares) chromophores at 400 and 530 nm, in THF solution.

3.4.2. Near steady state photoinduced absorption (PIA) spectroscopy

By measuring the PIA spectrum of a 1:1 mixture of the OAn-OPV dyad and the PERY

reference compound in a o-dichlorobenzene solution, the feasibility of a charge shift occurring in the

triad (CSH1, OAn-OPV•+-PERY•–-OPV-OAn → OAn•+-OPV-PERY•–-OPV-OAn) can be assessed.

Illumination of this mixture with monochromatic light of 458 nm results in the almost selective

excitation of the OPV chromophore.24 The OPV singlet excited state can evolve into the OAn•+-OPV•–

charged separated state (Figure 3.7, path a) or via intersystem crossing to the 3OPV* triplet excited

state (Figure 3.7, path b). While the charge-state decays to the ground state within a few nanoseconds

with no further consequences,25 the triplet state is characterized by a very long lifetime (τ = 25 µs).21

This long-lived triplet state can encounter a PERY chromophore in solution by diffusion and undergo

intermolecular charge transfer to generate the OAn-OPV•+ and PERY•– charged molecules (Figure

3.7, path d). These charged species will diffuse apart resulting in a long lifetime. During its lifetime,

the photogenerated positive charge of OAn-OPV•+ will most likely shift to the OAn segment to yield

OAn•+-OPV, because OAn has a lower oxidation potential than OPV (Figure 3.7, path e). The

experimental PIA spectrum (Figure 3.8) is in agreement with this sequence of reactions (paths b to e

Page 75: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 3

64

in Figure 3.7). It consists of a structured absorption band with maxima at 1.28, 1.54 and 1.72 eV that

are characteristic for the PERY•– radical anion.26 The absorption of the OAn•+ radical cation cannot be

distinguished because the electronic transitions of this charged species coincide in energy with those

of the PERY radical anion and are less intense (ε (1.44 eV) =15 × 103 M-1cm-1 for OAn•+ vs. ε (1.72

eV) ∼ 100 × 103 M-1cm-1 for PERY•–).26 Nevertheless, the absence of the typical polaronic absorptions

corresponding to the OPV radical cation, at 0.59 and 1.43 eV,21 in particular the one at low energy,

gives evidence that the positive charge has shifted from the OPV to the OAn segment.

OAn - 1 OPV * + PERY

OAn •+

+ - OPV •−

- + PERY

OAn - 3 OPV * + PERY OAn - OPV •+ + PERY •−-

a

b

c

d OAn •+

+ - OPV + PERY •− e

OAn - OPV + PERY

Figure 3.7. Probable photophysical reactions in a mixture of OAn-OPV PERY 1:1 after excitation at

458 nm.

Figure 3.8. Photoinduced absorption spectra of the mixture OAn-OPV/PERY (1:1) (solid line) in o-

dichlorobenzene solution (excitation at 458 nm with 25 mW and modulation frequency of 275 Hz).

3.4.3. Subpicosecond transient pump-probe spectroscopy

To unravel the different photophysical processes occurring on the picosecond to nanosecond

time domain, subpicosecond transient pump-probe spectroscopy has been performed at room

temperature on solutions of pentad 1 in solvents of different polarity. For pentad 2 the study has been

restricted to THF, because in this solvent aggregation is less pronounced. The experiments have been

performed by selective excitation of the OPV or PERY chromophores at 455 or 520 nm, respectively,

and by monitoring the transient absorptions at 1450, 900, and 700 nm. The absorption at 1450 nm is

0.5 1.0 1.5 2.0 2.5-1

0

1

2

3

4

-∆T

/T x

104

Energy (eV)

Page 76: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced multistep electron transfer in multichromophoric arrays

65

caused by the low-energy transition of the OPV•+ radical cation21 and, hence, this transient signal

gives direct information on the rates of formation (kCS = kCS1+kCS2) and decay (kdecay) of the OAn-

OPV•+-PERY•–-OPV-OAn charge separated state. The decay of this charge-separated state is the sum

of the rate constants for charge recombination to the ground state and charge shift to the OAn•+-OPV-

PERY•–-OPV-OAn state (kdecay = kCR1 + kCSH1). The radical ions of all chromophores absorb at 700

nm, with estimated molar absorption coefficients of 7 × 103, 15 × 103 and ~80 × 103 M-1cm-1 for

OAn•+, OPV•+,27 and PERY•–26 respectively. The 1OPV* and 1PERY* states dominate the transient

absorption at 900 nm. Also at this wavelength all radical ions absorb, though with much lower molar

absorption coefficients for OPV•+ and PERY•– than at 700 nm. On the basis of the molar absorption

coefficients of the radical ions a OAn-OPV•+-PERY•–-OPV-OAn to a OAn•+-OPV-PERY•–-OPV-OAn

charge shift can be identified by comparing the intensities of transient absorptions at 700 nm

(dominated by the PERY•– radical anion), 900 nm (dominated by the S1 states of the OPV or PERY

chromophores), and 1450 nm (dominated by the OPV•+ radical cation).

3.4.3.1 Pentad 1

Apolar medium (toluene solution). The 1450 nm transient absorption of pentad 1 in toluene

solution after excitation at 450 nm exhibits a fast rise and a slow decrease of the (negative) intensity

(Figure 3.9). The rise occurs within 1 ps, which corresponds to a rate constant for charge separation

(kCS) of > 1000 ns-1. This fast formation of the OAn-OPV•+-PERY•–-OPV-OAn charge-separated state

is in agreement with the complete quenching of the PERY emission (Figure 3.5, right). The slow

decay (3×109 s-1) points to a long-lived charge-separated state (~ 366 ps) in toluene.

The transient absorption recorded at 700 nm essentially superimposes with the absorption at

1450 nm (Figure 3.9). This strongly suggests that both absorptions originate from the same charged

species, i.e. the OAn-OPV•+-PERY•–-OPV-OAn and, hence, no charge shift is taking place in toluene

solution.

The 900 nm differential transmission, after photoexcitation at 450 nm (OPV segment),

exhibits in the first picosecond an abrupt rise and decay of the intensity corresponding to the

formation and decay of the OPV singlet excited state (Figure 3.9). A similar feature is observed if the

PERY chromophore is excited instead, although in this case the short-lived absorption corresponds to

the PERY singlet excited state. Both S1 states evolve within 1 ps into the OAn-OPV•+-PERY•–-OPV-

OAn charge-separated state, as evidenced by the 1450 and 700 nm transient absorptions. Apart from

this initial transient feature, the absorption at 900 nm exhibits the same temporal evolution as the 1450

and 700 nm differential transmissions, which is associated with the OAn-OPV•+-PERY•–-OPV-OAn

charge-separated state.

Page 77: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 3

66

0 100 200 300 400 500 600

-6

-5

-4

-3

-2

-1

0

-4 -2 0 2 4 6

-5

0

∆T (

a. u

.)

Time (ps)

∆T (

a. u

.)

Time (ps)

Figure 3.9. Differential transmission dynamics of pentad 1 in toluene monitored at 1450 (open

circles), 900 nm (closed squares) and 700 (open triangles) with excitation at 450 nm. The inset shows

the differential transmissions on shorter time scale.

Polar media. The 1450 nm transient absorption reveals that the formation of the OAn-OPV•+-

PERY•–-OPV-OAn charge-separated state in pentad 1 in THF solution is as fast as in toluene, but that

the lifetime of the charged state is dramatically shorter (~ 12 vs. 366 ps) (Figure 3.10). Again, the

same kinetics are observed regardless which chromophore is excited, OPV or PERY.

In the differential transmission at 700 nm of pentad 1 in THF two different regimes can be

distinguished (Figure 3.10). The first regime consists of a fast rise and decay of the signal occurring in

the initial 30 ps. This initial feature is associated with the formation and decay of the OAn-OPV•+-

PERY•–-OPV-OAn charge-separated and, thus, the OPV•+ and PERY•– radical ions account for the

absorption in this first regime. After 30 ps most of the absorption has decayed but a less intense long-

lived signal remains, which constitutes the second regime. Because the 1450 nm absorption of the

OAn-OPV•+-PERY•–-OPV-OAn state has disappeared after 30 ps, the remaining signal at 700 nm

corresponds to the OAn•+-OPV-PERY•–-OPV-OAn charge separated state. In the first picosecond

after photoexcitation, the differential transmission at 900 nm features the formation and decay of the

singlet excited state of the OPV and PERY chromophores (depending on which has been excited). As

in toluene, this initial singlet excited state is immediately quenched by the formation of the OAn-

OPV•+-PERY•–-OPV-OAn charge-separated state. After 20 ps an almost constant signal remains that

corresponds to the OAn•+-OPV-PERY•–-OPV-OAn charge separated state.

Page 78: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced multistep electron transfer in multichromophoric arrays

67

0 200 400 600 800 1000

-1

0

∆T (

a. u

.)∆T

(a.

u.)

∆T (

a. u

.)

Time (ps)

0 20 40 60 80 100

-1

0

Time (ps)

0 2 4 6 8

-1

0

Time (ps)

Figure 3.10. Differential transmission dynamics of pentad 1 in THF monitored at 1450 (open circles),

900 (solid squares) and 700 nm (open triangles) with excitation at 455 nm, measured on different time

scales.

Even though the rate for charge shift cannot be extracted from the data, an efficiency of 22%

for the process in THF can be estimated from the 700 nm transient absorption, using the molar

absorption coefficients of the different radical ions at this wavelength and assuming that the

maximum intensity at 3 ps corresponds to the OAn-OPV•+-PERY•– charge separated state, while at 50

ps the remaining absorption corresponds to the OAn•+-OPV-PERY•– charged state only.

Page 79: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 3

68

In chlorobenzene and o-dichlorobenzene, solvents of lower and higher polarity than THF, the

differential transmissions at 1450, 900 and 700 nm of pentad 1 exhibit similar time profiles as

observed in THF. Also the efficiencies for charge shift are similar in all the polar solvents.

0 200 400 600 800 1000

-1

0

∆T (

a. u

.)

Time (ps)

0 100 200 300 400

-1

0

∆T (

a. u

.)

Time (ps)

0 2 4 6 8 10 12

-1

0

∆T (

a. u

.)

Time (ps)

Figure 3.11. Differential transmission dynamics of pentad 2 in THF monitored at 1450 (open circles),

900 (solid squares) and 700 nm (open triangles) with excitation at 455 nm, measured on different time

scales.

Page 80: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced multistep electron transfer in multichromophoric arrays

69

3.4.3.2. Pentad 2 in THF

The differential transmission at 1450 nm of pentad 2 dissolved in THF (Figure 3.11) reveals

that the charge separation process to the OAn-OPV•+-PERY•–-OPV-OAn charge-separated state is

slowed down to kCS1 = 285 ns-1, compared to 1000 ns–1 for the same process in pentad 1. The reduced

rate of charge formation is consistent with the lower photoluminescence quenching of the PERY

chromophore in pentad 2 compared to 1. Moreover, the lifetime of the charge-separated state is

influenced by the longer and non-conjugated connection between donor and acceptor. As can be

expected, the rate for charge recombination has decreased resulting in a longer lifetime of the primary

charge-separated state (OAn-OPV•+-PERY•–-OPV-OAn) in pentad 2 (100 ps) than in pentad 1 (12 ps).

The efficiency of the charge shift process to generate the OAn•+-OPV-PERY•–-OPV-OAn charge-

separated state has improved to 28%, judging from the 700 nm transient absorption.

3.5. Kinetic considerations

Marcus theory provides an estimate for the free energy barrier (∆G‡) for electron transfer

reactions based on the change in free energy (∆G0) and the reorganization energy (λ) via:

( )

λλ

4

20‡ +∆=∆ G

G (3.2)

The reorganization energy consists of an internal contribution (λ i) and a solvent term (λs),

which can be approximated via the Born-Hush approach to give after summation, with n the index of

refraction:

++=+= −+

s2

0

2

isi11111

2

1

4 επελλλλ

nRrr

e

cc

(3.3)

The rate constants for the different processes, are not only a function of the energy barrier

∆G‡, but also of the reorganization energy (λ) and the electronic coupling (V) between donor and

acceptor in the excited state according to:

( )

+∆−

=

Tk

GV

Tkhk

B

202

21

B2

2

4exp

4

λλ

λπ

(3.4)

Page 81: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 3

70

The values of ∆G0, λ, and ∆G‡ calculated on the basis of Eqs. 3.1 to 3.4 are collected in Table

3.2. For the initial charge separation (kCS1) the values of free energy variation and reorganization

energies are comparable (–∆G0 ~ λ), close to the Marcus optimal region. In such cases, the reaction

rate is governed mainly by the electronic coupling V between the donor and acceptor and the polarity

of the solvent becomes less important. The electronic coupling depends on the nature of the spacer

and on the separation of donor and acceptor via ))(exp()( 0cc02

02 RRRVV −−= β , with R0 the

contact distance. The electronic coupling in pentad 1 is probably high due to the semi-conjugated

fashion by which the OPV and the PERY chromophores are linked. This explains the extraordinarily

high rate constants for the process observed for pentad 1 in all solvents (kCS > 1000 ns-1). Charge

recombination in OAn-OPV•+-PERY•–-OPV-OAn is in the Marcus inverted region (–∆G0 > λ) for all

solvents (Table 3.2). The use of Eq. 3.4 in the inverted region often underestimates the true rate

constant because of nuclear tunneling,28 but will be used here for qualitative comparison. In general

the barrier for charge recombination is reduced upon increasing the polarity of the medium. In

toluene, the most apolar solvent, the energy barrier is remarkably high, due to the large difference

between the free and reorganization energies. This is in agreement with the rather long-lived (366 ps)

OAn-OPV•+-PERY•–-OPV-OAn charge-separated stated observed for pentad 1 in toluene. In the more

polar solvents, the energy barrier is up to two orders of magnitude less than in toluene and, hence

charge recombination is expected to be much faster than in toluene. This explains, to some extend, the

12 ps lifetime of the OAn-OPV•+-PERY•–-OPV-OAn charge-separated state in THF. The charge shift

occurs in the normal region (–∆G0 < λ) and is energetically less favorable than the recombination

process (Table 3.2). In addition, the energy barrier of this process is higher than that of charge

recombination. These two facts imply that the charge shift is probably a slower process than the

charge recombination, although the influence of V is not accounted for. The study on the analogous

OAn-OPV-C60 molecular triad reported in chapter 2, has shown that increasing the lifetime of the

OAn-OPV•+-PERY•–-OPV-OAn charge-separated state is beneficial for the slower charge shift

process to compete. In principle this condition is nicely fulfilled for pentad 1, in toluene solution.

However, the driving force for the charge shift is close to zero in toluene and, hence, the charge shift

does not occur. However, in the polar solvents, where charge recombination is much faster, the charge

shift can be observed. The efficiency of the charge shift is similar in all polar solvents, consistent with

the small variation in energies reported in Table 3.2.

For pentad 2 the rate for charge separation (kCS) is less than for pentad 1, even though the

energy values depicted in Table 3.2 predict a similarly low barrier. Whereas in pentad 1, the electron

can be transferred from the OPV donor, through a ‘conjugated’ pathway to the PERY acceptor, the

electron transfer in pentad 2 occurs through space or through the saturated spacer by superexchange

mechanism. Moreover, the Rcc distance is higher in pentad 2. Both effects result in a lower electronic

Page 82: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced multistep electron transfer in multichromophoric arrays

71

coupling in the excited state between the OPV donor and the PERY acceptor in pentad 2, and thus a

slower charge separation process. The same explanation holds for the charge recombination process

and explains the longer lifetime of the OAn-OPV•+-PERY•–-OPV-OAn state of pentad 2 in THF

solution (100 ps) compared to that of pentad 1 (12 ps). As a result of the slower charge recombination

to the ground state a slightly higher efficiency for charge shift is found for pentad 2 than for pentad 1.

The electronic coupling between the OAn and the PERY moieties is very weak.

Consequently, the charge recombination process is very slow and a long lifetime is observed for the

OAn•+-OPV-PERY•–-OPV-OAn charge-separated state.

3.6 Conclusions

Two multichromophoric arrays have been synthesized by covalently linking OAn, OPV and

PERY chromophores in an OAn-OPV-PERY-OPV-OAn symmetrical arrangement. In pentad 2 all

chromophores are decoupled in the ground state while in pentad 1 some π-overlap is possible between

the OPV and the PERY chromophores. Because of its urethane functionality, pentad 2 dimerizes or

aggregates in a variety of solvents (except THF) due to intermolecular hydrogen bonding.

Upon illumination of the pentad a number of sequential photophysical events occur. These

processes are mainly multistep energy and electron transfers. Even though no unambiguous evidence

has been established for an ultrafast multistep singlet-energy transfer process (ET1 & ET2, Figure

3.4) resulting in the OAn-OPV-1PERY*-OPV-OAn state prior to the charge transfer and irrespective

of the excitation wavelength, some facts seem to indicate that such sequential event indeed occurs.

First, the excitation spectrum of the residual PERY emission in pentad 2, superimposes with the

absorption spectrum of the pentad (Figure 3.6, left). Second, the rate for intramolecular charge

separation that generates the OAn-OPV•+-PERY•–-OPV-OAn state is independent of the chromophore

that has been excited, strongly suggesting that they originate from the same state.

The sequential electron transfer starts by the formation of OAn-OPV•+-PERY•–-OPV-OAn.

This state can relax to the ground state via charge recombination or generate via a charge shift the

long-lived OAn•+-OPV-PERY•–-OPV-OAn charge-separated state (CR1 and CSH1, Figure 3.4). In

pentad 1, intramolecular charge separation to generate the OAn-OPV•+-PERY• –-OPV-OAn state is

extremely fast (kCS > 1000 ns-1), regardless the polarity of the solvent, because it occurs under optimal

conditions (–∆G0 ~ λ). The charge recombination of OAn-OPV•+-PERY•–-OPV-OAn to the ground

state occurs in the Marcus inverted region (–∆G0 > λ) and is therefore very fast in the polar solvents

and remarkably slow in toluene. The charge separation and recombination are slowed down in pentad

2, because the saturated spacer reduces the electronic coupling between the OPV donor and PERY

acceptor. In pentad 1, the charge shift process that generates OAn•+-OPV-PERY•–-OPV-OAn

competes with the faster charge recombination with an efficiency of 0.22. The slower charge

Page 83: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 3

72

recombination of OAn-OPV•+-PERY•–-OPV-OAn to the ground state in pentad 2, with respect to that

in pentad 1, is beneficial for the charge shift process that occurs with an efficiency of 0.28 in polar

media. Thus placing the OPV and PERY chromophores at longer distances seems to be the key to

improve the quantum yield for charge shift. As a matter of fact, the steady state PIA experiment on

the OAn-OPV and PERY mixture is an extreme situation of such a large distance, resulting in a high

efficiency because for every OAn-OPV•+/PERY•– the OAn•+-OPV/PERY•– is formed.

Pentads 1 and 2 show a few differences with respect to the OAn-OPV-C60 molecular triad

presented in chapter 2. In contrast to the fullerene-based triad, charge separation is possible in apolar

media for the pentads owing to the lower reduction potential of the perylene diimid acceptor.

Furthermore, in the pentads the various photophysical processes are less sensitive to the solvent

polarity (within the polar solvents) than in the OAn-OPV-C60 triad. This means that even though

similar trends are followed (for example, faster charge recombination in polar solvents), these changes

are almost negligible in the pentads. The most striking difference between the two multichromophoric

arrays originates from the fact that the charge shift process to the OAn•+-OPV-PERY•–-OPV-OAn is

able to compete with the charge recombination to the ground state in the pentads, even though the

latter process seems to be even much faster than in the OAn-OPV-C60 triad (Chapter 2).

In this chapter, the photophysical phenomena following photoexcitation have been

investigated in pentads 1 and 2 in an isolated form (molecularly dissolved). These molecules have the

potential to form ordered molecular aggregates, because of the strong and directing π−π interactions

of the perylene unit and additionally for pentad 2 the possibility of hydrogen bonding via the urethane

linkage. Future studies on the aggregated and solid states of the pentads might reveal promising bulk

photophysical properties.

3.7 Experimental section

For general methods and materials the reader is referred to chapter 2 of this thesis. Diethyl(4-nitrobenzyl) phosphonate (4). Triethyl phosphite (2.30 g, 13.88 mmol) and 4-nitrobenzyl bromide (2g, 9.25 mmol) were heated to 160 °C and stirred for 2 h. Subsequently, the mixture was cooled to 70 °C and the formed ethyl bromide and the excess of triethyl phosphite were distilled under reduced pressure. The residue was dissolved in ethyl acetate and filtrated over silica gel. The solvent was removed in vacuo to yield 2.2 g of diethyl 4-nitrobenzyl phosphonate. 1H NMR (CDCl3, 300 MHz): δ 8.18 (d, 2H), 7.5 (d, 2H), 4.07 (m, 4H), 3.27 (d, 2H), 1.28 (t, 6H); 13C NMR (CDCl3, 75 MHz): δ 146.78, 139.62 (d), 130.45(d), 123.44 (d), 62.20 (d), 34.61, 32.79, 16.15 (d). (E,E,E,E)-4-<4-[4-{4-[N-(4-Diphenylaminophenyl)-N-(phenyl)-3-aminostyryl]-2,5-bis[(S)-2-methylbutoxy]styryl]-2,5-bis[(S)-2-methylbutoxy]styryl}-2,5-bis[(S)-2-methylbutoxy]styryl>nitrobenzene (5).

Page 84: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced multistep electron transfer in multichromophoric arrays

73

Diethyl(4-nitrobenzyl) phosphonate (73 mg, 0.27 mmol), was dissolved in anhydrous DMF (6 mL) under an argon atmosphere and KtBuO (35 mg, 0.32 mmol) was added to the solution at room temperature. After 15 minutes a solution of aldehyde compound 3 (307 mg, 0.243 mmol) in 3 mL DMF/THF (2/1) was added dropwise to the reaction mixture. Twenty minutes after the addition was completed, the product starts precipitating from solution. The reaction mixture was stirred for 2 h more. The product was isolated by filtration and washed with ethanol and methanol as a red powder (0.290 g, 86%). 1H NMR (CDCl3, 300 MHz): δ 8.22 (d, 2H), 7.65 (d, 1H), 7.63(d, 2H), 7.57 (d, 1H), 7.52 (s, 2H), 7.51 (d, 1H), 7.40 (d, 1H), 7.31-6.96 (m, 31H), 4.05-3.72 (m, 12H), 2.10-1.85(m, 6H), 1.70-1.45 (m, 6H ), 1.45-1.2 (m, 6H), 1.2-0.85 (m, 36H); 13C NMR (CDCl3, 75 MHz): δ 151.81, 151.25, 151.01, 148.21, 147.84, 146.43, 144.73, 142.90, 142.70, 139.15, 129.45, 129.18, 128.42, 128.31, 127.75, 127.59, 127.07, 126.63, 126.04, 125.45, 125.30, 125.11, 124.18, 123.78, 122.96, 122.79, 122.68, 122.42, 122.24, 120.29, 111.15, 110.67, 110.01, 109.89, 109.70, 109.53, 74.47, 74.25, 74.16, 74.04, 73.94, 35.13, 35.05, 34.94, 26.37, 16.85, 11.52, 11.38. MALDI-TOF MS (Mw =1382.87) m/z = 1382.0 [M]+; Anal. Calc for C92H107N3O8: C, 79.9; H, 7.8; N, 3.0. Found: C, 79.7; H, 7.8; N, 2.7. (E,E,E,E)-4-<4-[4-{4-[N-(4-Diphenylaminophenyl)-N-(phenyl)-3-aminostyryl]-2,5-bis[(S)-2-methylbutoxy]styryl]-2,5-bis[(S)-2-methylbutoxy]styryl}-2,5-bis[(S)-2-methylbutoxy]styryl>aniline (6). Under an Ar atmosphere nitro compound 5 (0.28 g, 0.2 mmol) was suspended in EtOAc (15 mL). SnCl2.H2O (0.36 g, 1.62 mmol) and EtOH (2 mL) were added and the mixture was heated to 85 °C. The reaction mixture was stirred for 5 h at 95 °C and subsequently cooled to room temperature. After cooling to room temperature the reaction mixture was poured into crushed ice. The aqueous phase was slightly basified by the addition of NaOH 0.1 M and was subsequently extracted three times with diethyl ether. The collected organic fractions were dried over MgSO4, and the solvent removed in vacuo. Purification by column chromatography (silica gel, pentane/CH2Cl2 1:1, Rf = 0.4), yielded 6 (0.19 g, 70%) as an orange solid. 1H NMR (CDCl3, 400 MHz): δ 7.52 (s, 4H), 7.42 (d, 1H), 7.36 (d, 2H), 7.31 (d, 1H), 7.27-6.97 (m, 31H), 6.69 (d, 2H), 3.93-3.89 (m, 12H), 1.70-1.45 (m, 6H), 1.45-1.2 (m, 6H), 1.2-0.85 (m, 36H); 13C NMR (CDCl3, 75 MHz): δ 151.23, 151.08, 151.02, 150.95, 148.17, 147.81, 145.94, 142.86, 142.69, 139.15, 129.43, 129.17, 128.77, 128.67, 128.30, 127.72, 127.53, 127.36, 127.21, 126.79, 126.43, 125.44, 125.30, 123.77, 123.71, 123.59, 122.68, 122.41, 122.25, 120.26, 119.88, 115.22, 110.62, 110.41, 109.85, 109.65, 74.44, 74.28, 74.06, 73.99, 35.13, 35.04, 34.98, 34.92, 26.37, 26.29, 16.89, 16.87, 16.82, 16.79, 11.54, 11.48, 11.41, 11.37. MALDI-TOF MS (Mw =1352.89) m/z = 1352.92 [M]+.

N,N’-Bis[(E,E,E,E)-4-<4-[4-{4-[N-(4-Diphenylaminophenyl)-N-(phenyl)-3-aminostyryl]-2,5-bis[(S)-2-methylbutoxy]styryl]-2,5-bis[(S)-2-methylbutoxy]styryl}-2,5-bis[(S)-2-methylbutoxy]styryl>phenyl]-3,4,9,10-perylene bis(dicarboximide) (1). Amine compound 6 (55 mg, 0.04 mmol), 3,4,9,10-perylenetetracarboxylic dianhydride (8 mg, 0.02 mmol), imidazole (0.5g, 7.3 mmol), and a catalytic amount of Zn(AcO)2 were mixed and heated till 160 °C. After stirring for 4 h, the reaction mixture was cooled to room temperature. The residue was purified by column chromatography (silica gel, ethyl acetate/ CHCl3 1:0 to 1:4, Rf = 0 to 0.4) and repetitive preparative size exclusion chromatography (Bio Beads S-X1 and S-X3, CH2Cl2) to yield 9 mg (15 %) of 1 as a dark red solid. 1H NMR (CD2Cl2, 400 MHz): δ 8.61-8.52 (m, 8H), 7.61 (d, 4H), 7.51-6.89 (m, 77H), 3.90-3.70 (m, 24H), 1.94-1.82 (m, 12H), 1.65-1.52 (m, 12H), 1.36-1.25 (m, 12H), 1.07-0.90 (m, 64H). MALDI-TOF MS (Mw =3062.08) m/z = 3060.64 [M]+. Diethyl benzylphosphonate (7). Triethyl phosphite (1.45 g, 8.77 mmol) and benzyl bromide (1 g, 5.84 mmol) were heated to 160 °C and stirred for 3 h. Subsequently, the mixture was cooled to 90 °C and the formed ethyl bromide and the excess of triethyl phosphite were distilled under reduced pressure. The residue was dissolved in ethyl acetate and filtrated over silica gel. The solvent was removed in vacuo to yield 1.05 g (79 %)of diethyl benzylphosphonate. 1H NMR (CDCl3, 300 MHz): δ 7.34 (m, 5H), 4.06-3.95 (m, 4H), 3.15 (dd, 2H), 1.23 (td, 6H); 13C NMR (CDCl3, 75 MHz): δ 131.55 (d), 129.70 (d), 128.45 (d), 126.78 (d), 62.02 (d), 34.63, 32.80, 16.29 (d). (E,E,E,E)-4-<4-[4-{4-[N-(4-Diphenylaminophenyl)-N-(phenyl)-3-aminostyryl]-2,5-bis[(S)-2-methylbutoxy]styryl]-2,5-bis[(S)-2-methylbutoxy]styryl}-2,5-bis[(S)-2-methylbutoxy]styryl>benzene (8). Phosphonate 7 (17 mg, 0.074 mmol) was dissolved in anhydrous DMF (1 mL) under an argon atmosphere and KtBuO (18 mg, 0.016 mmol) was added to the solution at room temperature. After 15 minutes, a solution of aldehyde compound 3 (54 mg, 0.042 mmol) in 2.2 mL DMF/THF (2/1) was added dropwise to the reaction mixture. The reaction mixture was stirred for 2.5 h. Evaporation of the solvent and precipitation in methanol yielded 7 (42 mg, 73 %) as an orange solid. 1H NMR (CD2Cl2, 400 MHz): δ 7.56-6.98 (m, 42 H), 3.98-3.81 (m,

Page 85: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 3

74

12H), 2.04-1.92 (m, 6H), 1.73-1.59 (m, 6H), 1.43-1.32 (m, 6H), 1.14-0.96 (m, 36H); 13C NMR (CDCl3, 100 MHz): δ 151.65, 151.50, 151.41, 148.73, 148.27, 143.46, 143.21, 139.54, 138.39, 129.84, 129.55, 129.06, 128.93, 128.71, 127.86, 127.80, 127.65, 126.91, 126.77, 126.02, 125.81, 124.11, 123.78, 123.68, 123.07, 122.92, 122.83, 122.25, 120.61, 110.96, 110.78, 110.15, 109.93, 74.76, 74.56, 74.42, 35.56, 35.49, 35.43, 26.77, 17.00, 11.70, 11.56. MALDI-TOF MS (Mw =1337.90) m/z = 1337.74 [M]+. (E,E,E,E)-4-<4-[4-{4-[N-(4-Diphenylaminophenyl)-N-(phenyl)-3-aminostyryl]-2,5-bis[(S)-2-methylbutoxy]styryl]-2,5-bis[(S)-2-methylbutoxy]styryl}-2,5-bis[(S)-2-methylbutoxy]styryl>phenylisocyanate (9). Amine compound 6 (50 mg, 0.04 mmol) was suspended in 20% phosgene solution in toluene (2.4 mL) and stirred at 95 °C for 16 h under an inert atmosphere. The reaction mixture was cooled to room temperature and the solvent removed in vacuo. The complete conversion of the amine to the isocyanate was monitored by IR spectroscopy by observing the disappearance of the amine peak at 3364 cm-1 and the formation of the isocyanate peak at 2264 cm-1 Compound 9 was used without further purification. MALDI-TOF MS (Mw =1378.91) m/z = 1378.64 [M]+. N,N’-Di-((S)-1-isobutyl-2-hydroxyethyl)-3,4,9,10-perylene bis(dicarboximide) (10). (S)-(+)-Leucinol (0.455 g, 3.88 mmol), 3,4,9,10-perylenetetracarboxylic dianhydride (0.63 g, 1.61 mmol), imidazole (10 g), and catalytic amounts of Zn(OAc)2 were heated to 160 °C and stirred for 1.5 h. After cooling to room temperature the reaction mixture was dissolved in methylene chloride and washed 2 times with HCl 1M, brine, and dried over MgSO4. The solvent was removed in vacuo to yield 0.72g (75%) of a dark red solid. 1H NMR (CD2Cl2, 400 MHz): δ 8.40-7.52 (m, 8H), 5.54-5.44 (m, broad signal, 2H), 4.60-4.48 (m, broad signal, 2H), 3.98-3.94 (m, 2H), 2.10-2.01 (m, 2H), 1.80-1.46 (m, broad signal, 4H), 1.10-0.92 (m, 12). MALDI-TOF MS (Mw = 590.2) m/z = 590.1 [M]+. N,N’-Bis[(S)-1-Isobutyl-2-[(E,E,E,E)-4-<4-[4-{4-[N-(4-Diphenylaminophenyl)-N-(phenyl)-3-aminostyryl]-2,5-bis[(S)-2-methylbutoxy]styryl]-2,5-bis[(S)-2-methylbutoxy]styryl}-2,5-bis[(S)-2-methylbutoxy]styryl>phenyl carbamicacid]ethyl] 3,4,9,10-perylene bis(dicarboximide) (2). A solution of the isocyanate (38 mg, 0.027 mmol) in dry methylene chloride was added to a solution of 6 (7.4 mg, 0.012 mmol) in dry methylene chloride. Dibutylin dilaurate (3.22 mg) was added as a catalyst, and the reaction mixture was heated at reflux for 20 h under an argon atmosphere. The crude mixture was cooled to room temperature, the solvent was removed in vacuo. The residue was purified by silica gel chromatography (CH2Cl2/pentane; 4:1, Rf = 0.15) and preparative size exclusion chromatography (Bio Beads S-X3 , CH2Cl2) to yield 20 mg (47 %) of 7 as a red solid. 1H NMR (CD2Cl2, 400 MHz): δ 9.38 (br s, 2H), 8.57, 8.21, 8.12, 7.79 (4 x d, 8H), 7.58-7.06 (m, 82H), 5.94-5.78 (br m, 2H), 5.35-5.18 (br m, 2H), 4.76-4.57 (br. m, 2H), 3.98-3.83 (m, 24H), 2.37-2.22 (m, 2H), 2.02-1.89 (m, 14H), 1.75-1.60 (m, 16H), 1.52-1.32 (m, 12H), 1.16-0.99 (m, 84H); 13C NMR (THF, 100 MHz): δ 165.00 (broad signal), 154.83, 153.07, 152.95, 150.14, 149.76, 144.83, 144.73, 141.33, 140.59, 136.06, 134.32, 132.50, 131.21, 131.09, 130.82, 130.60, 129.89, 129.79, 129.36, 129.18, 129.04, 128.89, 128.65, 128.34, 128.26, 127.74, 127.09, 126.99, 126.79, 126.79, 125.45, 125.06, 124.49, 124.37, 124.22, 124.12, 123.97, 123.28, 123.09, 121.83, 119.90, 112.15, 112.02, 111.47, 111.20, 111.13, 80.28, 75.64, 75.46, 53.07, 40.05, 37.07, 36.97, 36.89, 28.11, 27.26, 24.44, 23.65, 18.08, 12.73, 12.61.MALDI-TOF MS (Mw =3348.49) m/z = 3347.85 [M]+. 3.8 References and notes 1 (a) Molecular Electronics, Jortner, J., Ratner, M. Eds.; Blackwell: London, 1997. (b) Lehn, J.-M.,

Supramolecular Chemistry; VCH, Weinheim, 1995. (c) Electron Transfer in Chemistry Vol. I-IV,

Balzani, V., Wiley-VCH, Weinheim, 2001.

2 For recent reviews see: (a) Gust, D.; Moore, T. A.; Moore, A. L. Acc. Chem. Res. 2001, 34, 40. (b) Guldi,

D. M. Chem. Soc. Rev. 2002, 31, 22.

Page 86: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced multistep electron transfer in multichromophoric arrays

75

3 (a) Liddell, P. A.; Kuciauskas, D.; Sumida, J. P.; Nash, B.; Nguyen, D.; Moore, A. L.; Moore, T. A.;

Gust, D. J. Amer. Chem. Soc. 1997, 119, 1400. (b) Imahori, H.; Yamada, K.; Hasegawa, M.; Taniguchi,

S.; Okada, T.; Sakata, Y. Angew. Chem., Int. Ed. Engl. 1997, 36, 2626. (c) Carbonera, D.; Di Valentin,

M.; Corvaja, C.; Agostini, G.; Giacometti, G.; Liddell, P. A.; Kuciauskas, D.; Moore, A. L.; Moore, T.

A.; Gust, D. J. Amer. Chem. Soc. 1998, 120, 4398. (d) Kuciauskas, D.; Liddell, P. A.; Moore, A. L.;

Moore, T. A.; Gust, D. J. Amer. Chem. Soc. 1998, 120, 10880. (e) Tamaki, K.; Imahori, H.; Sakata, Y.;

Nishimura, Y.; Yamazaki, I. Chem. Commun. 1999, 625. (f) Imahori, H.; Yamada, H.; Ozawa, S.; Sakata,

Y.; Ushida, K. Chem. Commun. 1999, 1165. (g) Fujitsuka, M.; Ito, O.; Imahori, H.; Yamada, K.;

Yamada, H.; Sakata, Y. Chem. Lett. 1999, 721. (h) Imahori, H.; Yamada, H.; Nishimura, Y.; Yamazaki,

I.; Sakata, Y. J. Phys. Chem. B 2000, 104, 2099. (i) Kuciauskas, D.; Liddell, P. A.; Lin, S.; Stone, S. G.;

Moore, A. L.; Moore, T. A.; Gust, D. J. Phys. Chem. B 2000, 104, 4307. (j) Luo, C.; Guldi, D. M.;

Imahori, H.; Tamaki, K.; Sakata, Y J. Amer. Chem. Soc. 2000, 122, 6535. (k) Imahori, H.; Tamaki, K.;

Yamada, H.; Yamada, K.; Sakata, Y.; Nishimura, Y.; Yamazaki, I.; Fujitsuka, M.; Ito, O. Carbon 2000,

38, 1599. (l) Bahr, J. L.; Kuciauskas, D.; Liddell, P. A.; Moore, A. L.; Moore, T. A.; Gust, D. Photochem.

Photobiol. 2000, 72, 598. (m) Imahori, H.; Norieda, H.; Yamada, H.; Nishimura, Y.; Yamazaki, I.;

Sakata, Y.; Fukuzumi, S. J. Amer. Chem. Soc. 2001, 123, 100. (n) Fukuzumi, S.; Imahori, H.; Yamada,

H.; El-Khouly, M. E.; Fujitsuka, M.; Ito, O.; Guldi, D. M. J. Amer. Chem. Soc. 2001, 123, 2571. (o)

Imahori, H.; Tamaki, K.; Guldi, D. M.; Luo, C.; Fujitsuka, M.; Ito, O.; Sakata, Y.; Fukuzumi, S. J. Amer.

Chem. Soc. 2001, 123, 2607. (p) Imahori, H.; Guldi, D. M.; Tamaki, K.; Yoshida, Y.; Luo, C.; Sakata, Y.;

Fukuzumi, S. J. Amer. Chem. Soc. 2001, 123, 6617. (q) Fukuzumi, S.; Imahori, H.; Okamoto, K.;

Yamada, H.; Fujitsuka, M.; Ito, O.; Guldi, D. M. J. Phys. Chem. A 2002, 106, 1903. (r) Liddell, P. A.;

Kodis, G.; De la Garza, L.; Bahr, J. L.; Moore, A. L.; Moore, T. A.; Gust, D. Helv. Chim. Acta 2001, 84,

2765. (s) Ikemoto, J.; Takimiya, K.; Aso, Y.; Otsubo, T.; Fujitsuka, M.; Ito, O. Org. Lett. 2002, 4, 309. (t)

Imahori, H.; Tamaki, K.; Araki, Y.; Hasobe, T.; Ito, O.; Shimomura, A.; Kundu, S.; Okada, T.; Sakata,

Y.; Fukuzumi, S. J. Phys. Chem. A 2002, 106, 2803. (u) Imahori, H.; Tamaki, K.; Araki, Y.; Sekiguchi,

Y.; Ito, O.; Sakata, Y.; Fukuzumi, S. J. Amer. Chem. Soc. 2002, 124, 5165. (v) D'Souza, F.; Deviprasad,

G. R.; Zandler, M. E.; El-Khouly, M. E.; Fujitsuka, M.; Ito, O. J. Phys. Chem. B 2002, 106, 4952. (w)

Liddell, Paul A.; Kodis, Gerdenis; Moore, A. L.; Moore, Thomas A.; Gust, Devens. J. Amer. Chem. Soc.

2002, 124, 7668. (x) Kodis, G.; Liddell, P. A.; de la Garza, L.; Moore, A. L.; Moore, T. A.; Gust, D. J.

Mater. Chem. 2002, 12, 2100. y) Sánchez, L.; Pérez, I.; Martín, N.; Guldi, D. M. Chem. Eur. J. 2003, 9,

2457.

4 (a) Kraabel, B.; McBranch, D.; Sariciftci, N. S.; Moses, D.; Heeger, A. J. Phys. Rev. B 1994, 50, 18543.

(b) Kraabel, B.; Hummelen, J. C.; Vacar, D.; Moses, D.; Sariciftci, N. S.; Heeger, A. J. J. Chem. Phys.

1996, 104, 4267. (c) Meskers, S. C.J.; van Hal, P. A.; Spiering, A. J. H.; Hummelen, J. C.; van der Meer,

A. F.G.; Janssen, R. A.J. Phys. Rev. B 2000, 61, 9917. (d) Nogueira, A. F.; Montanari, I.; Nogueira, A. F.;

Nelson, J.; Durrant, J. R.; Winder, C.; Loi, M. A.; Sariciftci, N. S.; Brabec, C. J. Phys. Chem. B 2003,

107, 1567

5 Peeters, E.; van Hal, P. A.; Meskers, S. C. J.; Janssen R. A. J.; Meijer, E. W. Chem. Eur. J., 2002, 8, 4470

Page 87: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 3

76

6 Asha, S.; Schenning, A. P. H. J.; Meijer, E. W.,Chem. Eur. J. 2002, 8, 3353.

7 Schenning, A. P. H. J.; van Herrikhuyzen, J.; Jonkheijm, P.; Chen, Z.; Würthner F.; Meijer, E. W. J. Am.

Chem. Soc., 2002, 124, 10252.

8 Neuteboom, E. E.; Meskers, S. C. J.; van Hal, P. A.; van Duren, J. K. J.; Meijer, E. W.; Dupin, H.;

Pourtois, G.; Cornil, J.; Lazzaroni, R.; Brédas, J.-L.; Beljonne, D.; Janssen, R. A. J. J. Am. Chem. Soc.,

2003, 125, 8625

9 Nierengarten, J.-F.; Eckert, J.-F.; Nicoud, J.-F.; Ouali, L.; Krasnikov, V.; Hadziioannou, G. Chem.

Commun. 1999, 617.

10 Eckert, J.-F.; Nicoud, J.-F.; Nierengarten, J.-F.; Liu, S.-G.; Echegoyen, L.; Barigelletti, F.; Armaroli, N.;

Ouali, L.; Krasnikov, V.; Hadziioannou, G. J. Am. Chem. Soc. 2000, 122, 7467

11 Peeters, E.; van Hal, P. A.; Knol, J.; Brabec, C. J.; Sariciftci, N. S.; Hummelen, J. C.; Janssen, R. A. J. J.

Phys. Chem. B 2000, 104, 10174.

12 El-ghayoury, A.; Schenning, A. P. J. H.; van Hal, P. A.; Van Duren, J. K. J.; Janssen, R. A. J.; Meijer, E.

W. Angew. Chem. Int. Ed. 2001, 40, 3660.

13 Marcos Ramos, A.; Rispens, M.T.; van Duren, J. K. J.; Hummelen, J. C.; Janssen, R. A. J. J. Am. Chem.

Soc. 2001, 123, 6714.

14 De Boer, B.; Stalmach, U.; Van Hutten, P. F.; Melzer, C.; Krasnikov, V. V.; Hadziioannou, G. Polymer

2001, 42, 9097.

15 (a) Wohrle, D.; Meissner, D. Adv. Mater. 1991, 3, 129. (b) Schlettwein, D.; Wohrle, D.; Karmann, E.;

Melville, U. Chem. Mater., 1994, 6, 3. (c) Ferrere, S.; Zaban, A.; Gregg, B. A. J. Phys. Chem. B, 1997,

101, 4490.

16 Schmidt-Mende, L.; Fechtenkötter, A.; Müllen, K.; Moons, E.; Friend, R. H.; MacKenzie, J. D. Science

2001, 293, 1119.

17 Dittmer, J. J.; Marseglia, E. A.; Friend, R. H. Adv. Mater. 2000, 12, 1270.

18 Dittmer, J. J.; Petritsch, K.; Marseglia, E. A.; Friend, R. H.; Rost, H.; Holmes, A. B. Synth. Met. 1999,

102, 879.

19 Angadi, M. A.; Gosztola, D.; Wasielewski, M. R. J. Appl. Phys.1998, 83, 6187.

20 Weller, A. Z. Phys. Chem. Neue Folge 1982, 133, 93.

21 Van Hal, P. A.; Beckers, E. H. A.; Peeters, E.; Apperloo, J. J., Janssen, R. A. J. Chem. Phys. Lett. 2000,

328, 403.

22 The assumption that charges are located at the centers of the chromophores is of course a simplification

of the actual situation in which charges are likely to be strongly delocalized.

23 This assumption is based on the studies of the analogous OAn-OPV-C60 triad, described in chapter 2.

24 The small amount of PERY chromophore that might be excited will decay to the ground state by

fluorescence before encountering any other chromophore.

25 As reported in chapter 2, charge recombination in an analogous dyad, OAn+-OPV– with the OPV

featuring four phenyl rings instead of five, occurs in the nanosecond regime.

26 Salbeck, J. J. Electroanal. Chem. 1992, 340, 169.

Page 88: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced multistep electron transfer in multichromophoric arrays

77

27 The molar absorption coefficient of the OPV chromophore at 700 nm has been extracted from the

photoinduced absorption spectrum of a OPV/MP-C60 (1:1) mixture in o-dichlorobenzene reported in

reference 21, using the extinction coefficients of the MP-C60 reported in Guldi, D. M; Prato, M. Acc.

Chem. Res. 2000, 33, 695.

28 Jortner, J. J. Chem. Phys. 1976, 64, 4860.

Page 89: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1
Page 90: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 4

Supramolecular control over

donor-acceptor photoinduced charge

separation

Abstract

A novel donor-bridge-acceptor system has been synthesized by covalently linking a p-phenylene

vinylene oligomer (OPV) and a perylene diimid (PERY) at opposite ends of a m-phenylene ethynylene

oligomer (FOLD) of twelve phenyl rings, containing non-polar, (S)-3,7-dimethyl-1-octanoxy side

chains. For comparison, also model compounds have been prepared in which either the donor or

acceptor is absent. In chloroform solution, the oligomeric bridge is in a random coil conformation.

Upon addition of a poor solvent (heptane) the oligomeric bridge first folds into a helical stack and

subsequently intermolecular aggregation of the stacks into columnar architectures occurs. In almost

pure heptane, the compounds bearing PERY chromophores exhibit higher degrees of aggregation. In

a random coil conformation, the small interaction between the donor and acceptor chromophores

allows for energy transfer from the OPV singlet-excited state to yield the PERY singlet-excited state.

In the folded and aggregated states of the bridge, donor and acceptor are in a favorable orientation

to generate the OPV•+-FOLD-PERY•– charge transfer product upon photoexcitation. The quantum

yield for charge separation depends on the degree of folding and aggregation of the bridge between

donor and acceptor and therefore on the apolar nature of the medium. As a consequence, and

contrary to conventional photoinduced charge separation processes, the formation of the OPV•+-

FOLD-PERY•– charge-separated state is more favored in apolar media.

Page 91: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 4

80

4.1 Introduction

Distance and orientation between covalently linked electron donors and acceptors are key

factors that determine the nature and kinetics of the photophysical processes that occur upon

illumination. Photoinduced charge separation is a reaction that is much more sensitive to the distance

between the chromophores than long-range energy transfer. In particular, for oligo(p-phenylene

vinylene)-perylene diimide (OPV-PERY) donor(D)-acceptor(A) dyads, direct connection of the redox

centers leads to a very fast charge separation, but also very fast charge recombination.1 When a

saturated spacer is interposed between the chromophores both processes are slowed down2 and if the

spacer is rigid and long, the OPV and PERY only exchange energy.3 Much effort has been given to

study the effects of distance and orientation in donor-acceptor dyads that are either directly linked4 or

connected via rigid, or flexible spacers.5

On/off switchability of intramolecular photophysical reactions can be achieved by the use of

supramolecular connections between donor and acceptor, such as hydrogen bonding3,6 and metal

coordination,7 or, in covalent D-bridge-A systems, by photoisomerization of the bridge.8 It is also

possible to modify the distance of covalently connected D-A dyads by modifying the conformation of

the flexible bridge between them. Two examples of that are complexation of the bridge with metal

cations9 and denaturation of the helical structure of peptide bridges.10

In the present work an OPV (donor) and a PERY (acceptor) have been linked at opposite ends

of a m-phenylene ethynylene oligomer made of twelve phenyl rings and containing non-polar (S)-3,7-

dimethyl-1-octanoxy side chains (FOLD) (Figure 4.1).

The type of oligomer that has been used as a bridge in OPV-FOLD-PERY, can go through

conformational changes in solution by varying the polarity of the media as well as the temperature,

when it is of sufficient length (n > 10, n being the number of phenyl rings).11 This behavior arises

from the difference in polarity between the aromatic hydrocarbon backbone and the side chains. In a

good solvent the oligomer adopts a random coil conformation and when a poor solvent (for the

backbone) is added, the oligomer folds into a helical stack, keeping the side chains exposed to the

solution. As a second step, aggregation of the helical stacks into larger stacks can also occur (Figure

4.2). A powerful technique to monitor the conformational changes is UV/Visible absorption

spectroscopy, as the different conformational states feature different absorption spectra. This has been

shown for oligomers featuring polar12 and apolar13 side chains. In both cases, the side chains contain a

chiral center, and this small perturbation is enough to bias the helical twist and as a result solutions of

the folded state show circular dichroism (CD) in the electronic absorption of light by the π system of

the backbone. For apolar side chain oligomers, the twist sense bias is only expressed when

intermolecular aggregation of the stacks into columnar architectures takes place.

Page 92: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Supramolecular control over donor-acceptor photoinduced charge separation

81

R*'O

R*'O

R*'O OR*'

OR*'

OR*'

N

OOR*

R*O OR*

OOR*

OR*

OR*

R*O

N

O

O

O

O

O

OR*O

O

OR*

OO

O

OOR*

O

OR*

OR*O

O

O

O

R*O =

R*'O =

Figure 4.1. Chemical structure of the OPV-FOLD-PERY dyad.

Figure 4.2. Schematic representations of the possible conformations of the FOLD oligomer upon

increasing the volume percentage of poor solvent in good solvent (from left to right).

The control over the conformation of these oligomers provides an unprecedented means to

influence the relative spatial orientation between donor and acceptor when FOLD is used as a bridge

in a dyad (OPV-FOLD-PERY, Figure 4.1). In a random coil conformation the OPV and PERY

chromophores are presumably too far away to interact and in the folded conformation or stacked state

they are expected to neighbor each other and to lead to high photophysical activity. In this chapter, the

synthesis of the OPV-FOLD-PERY dyad, with FOLD featuring apolar side chains, as well as the

synthesis of model compounds, in which either the donor or acceptor are missing, i.e. OPV-FOLD

and FOLD-PERY, are described. The conformational changes that the FOLD bridge undergoes by the

folding experiments are monitored with absorption and CD spectroscopy. Photoluminescence studies

of the different conformational states provide a measure of the interaction between donor and acceptor

in the excited state.

Page 93: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 4

82

4.2 Synthesis

The OPV-FOLD-PERY dyad 10 and its model compounds FOLD-PERY 7 and OPV-FOLD 9

have been synthesized following the synthetic approach previously developed for the m-phenylene

ethynylene oligomers. This synthetic strategy is based on the palladium-catalyzed Sonogashira

coupling, which allows for an orthogonal protecting group procedure and thus enables a divergent

modification of the folding oligomer. The monofunctional model compounds were obtained by

coupling either a PERY or an OPV moiety to one of the bridge oligomers, and the dyad was obtained

by selectively coupling of PERY and OPV at either side of the bridge. The synthesis of the bridged

donor-acceptor system and the model compounds is depicted in Scheme 4.1. To allow for the donor

and acceptor chromophores to be linked to the foldamer via the palladium-catalyzed cross-coupling

reaction, the OPV and the PERY chromophores had to be functionalized with ethynyl and iodine,

respectively. The building blocks for these functionalized chromophores are perylene monoanhydride

monoimide 114 and OPV 3,15 whose syntheses have been described elsewhere. Iodine-functionalized

perylene diimide 2 was obtained after reaction of 4-iodoaniline with the perylene monoanhydride

monoimide 1. Coupling of trimethylsilylacetylene with bromine 3 and subsequent deprotection with

tetrabutylammonium fluoride (TBAF) afforded the ethynyl end-capped OPV 4. Model compound 7

(FOLD-PERY) was obtained by coupling of iodoperylene diimide 2 with the trimethylsilylethynyl

end-capped dodecamer 6, in a Sonogashira reaction with in situ deprotection of the ethynyl functional

group.16 Reaction of 2 with previously deprotected dodecamer 6 resulted only in low yields due to the

occurrence of a considerable amount of homocoupled 6. Model compound 9 (OPV-FOLD) was

obtained via reaction of ethynyl OPV 5 with the iodine functionalized foldamer 8. This OPV-FOLD

was further reacted with 2 under similar conditions as used to obtain PERY-FOLD 7, utilizing the

ethynyl functionality at the other end of the chain, affording the OPV-FOLD-PERY bridged dyad 10

in low (9%) yield.

All compounds used in the photophysical investigations were characterized using 1H-NMR

spectroscopy, mass spectrometry, and size exclusion chromatography. Mass spectrometry established

the correct mass of the compounds and showed additional masses corresponding to extra methylene

groups. These additional signals arise due to the fact that the (S)-3,7-dimethyl-1-octanol used for the

m-phenylene ethynylene bridge is not completely pure but contained some alcohols of higher mass.

For each of the compounds a single, symmetrical peak was obtained by size exclusion

chromatography, revealing the absence of starting material or undesired longer adducts and that the

compounds are highly monodisperse.

Page 94: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Supramolecular control over donor-acceptor photoinduced charge separation

83

N

N

O

O

O

O

OR*O

OR*'

OR*'

OR*'

OR*'

OR*'

OR*'

OR*'

OR*'

OR*'

OR*'

OR*'

OR*'

Br

OR*'

OR*'

OR*'

OR*'

OR*'

OR*'

X

NN

O

O O

O

ION

O

O O

O

NN

O

O O

OOR*

O

H

OR*O

tms

H

OR*O

I

tms

OR*

O

tms

OR*'

OR*'

OR*'

OR*'

OR*'

OR*'

OR* = O

OR*' = O

12

10

f

3

b

c4. X = TMS

5. X = H

1 2

a

d

67

1212

12

12

e

98

Scheme 4.1. a. 4-Iodoaniline, imidazole, Zn(OAc)2, 160 °C, 1 h, 92%; b. Pd(PPh3)2Cl2, PPh3, CuI,

Et3N, 80 °C, 16 h, 45%; c. TBAF, THF, 1 min, r.t., 95%; d. 2, Pd(PPh3)4, KOAc, DMF, toluene, 100

°C, 16 h, 22%; e. 5, Pd(PPh3)2Cl2, PPh3, CuI, Et3N, 70 °C, 16 h, 48%; f. 2, 9, Pd(PPh3)4, KOAc,

DMF, toluene, 100 °C, 5 h, 9%.

Page 95: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 4

84

4.3 Conformational states of the bridge

4.3.1 Folding in chloroform/heptane mixtures

The UV/Visible absorption spectrum of OPV-FOLD-PERY in chloroform solution (Figure

4.3) shows the superposition of the individual electronic transitions of the FOLD11 bridge (288 and

304 nm), the OPV donor17 (430 nm), and the PERY acceptor1 (454, 490 and 526 nm) chromophores.

For OPV-FOLD and FOLD-PERY the absorption spectra display the related chromophore transitions.

300 400 500 6000.0

0.2

0.4

0.6

0.8

1.0

A

bsor

banc

e (O

. D.)

Wavelength (nm)

Figure 4.3. UV/Visible absorption spectra of OPV-FOLD-PERY (solid line), OPV-FOLD (dashed

line), and FOLD-PERY (dotted line) in chloroform solution.

The absorption spectrum of the FOLD bridge exhibits a shoulder at 304 and a peak at 288 nm

in chloroform solution. The ratio of the absorbance of the shoulder (As) and the peak (Ap) is

characteristic of the conformational state of the FOLD bridge.13 Transoidal conformations result in a

high As/Ap ratio, while cisoidal conformations lower the As/Ap ratio. The random-coil conformation,

in which transoidal conformations persist, is therefore characterized by a high As/Ap ratio, whereas

cisoidal conformations account for the formation of helices and a significant lower As/Ap ratio.

The high As/Ap absorption ratio observed in chloroform solution indicates that in the FOLD

bridge transoidal conformations dominate for OPV-FOLD, FOLD-PERY, and OPV-FOLD-PERY

and that the bridge is in a random coil conformation (Figure 4.3).13

In order to establish the conditions at which the bridge folds, solutions of OPV-FOLD-PERY

in chloroform/heptane mixtures of different composition were prepared and studied by means of

UV/Visible absorption and CD spectroscopy. With increasing amount of heptane in chloroform a

decrease of the As/Ap ratio of OPV-FOLD-PERY (Figure 4.4) is observed, characteristic of a helical

folding of the bridge.13 In addition, the shoulder at 304 nm and the peak at 288 nm observed in

chloroform, undergo small shifts with increasing the heptane content. Starting at 80% heptane, a

Cotton effect is observed in the UV region of the CD spectrum (Figure 4.4, inset). The appearance of

optical activity is accompanied by a small blue shift of the FOLD absorption band. In general, a CD

Page 96: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Supramolecular control over donor-acceptor photoinduced charge separation

85

effect in these m-phenylene ethynylene oligomers is only observed when intermolecular aggregation

of the molecular helices into columnar architectures occurs.13 As the relative amount of heptane

increases further, the overall FOLD absorption becomes less intense (hypochromicity) while the CD

effect grows. Above 95% heptane, an unexpected inversion of the chirality in the region of the FOLD

bridge occurs. In the absorption spectrum, the FOLD features a red–shifted broader absorption band

and a sudden small increase of the As/Ap ratio is observed.

300 400 500 6000,0

0,4

0,8

1,2

300 400 500 600

-60

-30

0

30

60

0%

80% 93%

96%

99%

CD

(mde

g)

Wavelength (nm)

Abs

orba

nce

(O. D

.)

Wavelength (nm)

Figure 4.4. UV/Visible spectra of OPV-FOLD-PERY in 0% (solid star), 60% (diamond), 80%

(triangle), 96% (circle) and 99% (solid square) volume heptane in chloroform. Inset: CD spectra of

OPV-FOLD-PERY in heptane/chloroform solutions of different composition (the numbers denote

volume percentage heptane in chloroform).

The absorption bands associated with the PERY chromophore also undergo some changes

upon the addition of heptane, which are a direct consequence of the conformational changes occurring

in the FOLD bridge. Figure 4.5 shows that the optical density of the lowest energy PERY absorption

band decreases considerably above 80% heptane content for both the OPV-FOLD-PERY dyad and the

FOLD-PERY model compound. This indicates aggregation of the PERY chromophores under these

conditions.

Page 97: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 4

86

0 20 40 60 80 100

0.0

0.2

0.4

0.6

0.8

1.0

1.2

Nor

mal

ized

abs

orba

nces

Volume % heptane in chloroform

Figure 4.5. The absorbance due to the S1 ← S0 transition the PERY at 528 nm of OPV-FOLD-PERY

(solid squares) and FOLD-PERY (open squares) in chloroform/heptane solutions of different

composition.

The As/Ap ratios of the OPV-FOLD-PERY dyad, the OPV-FOLD and the FOLD-PERY

model compounds and the unsubstituted FOLD bridge are plotted against the volume percentage of

chloroform in heptane in Figure 4.6. This plot shows that all compounds possess a similar folding and

aggregation behavior with decreasing polarity of the medium, up to a heptane content of 95%, where

the compounds with a PERY chromophore start to deviate because of aggregation.

0 20 40 60 80 100

0.6

0.7

0.8

0.9

As/A

p

Volume % heptane in chloroform

Figure 4.6. Ap/As ratios of OPV-FOLD-PERY (solid star), FOLD-PERY (open triangle), OPV-FOLD

(open circle) and FOLD (open square) in chloroform/heptane solutions of different composition.

The inversion of chirality together with the loss of the isodichroic points observed for the

compounds bearing PERY, i.e. OPV-FOLD-PERY and FOLD-PERY, at high percentage of heptane,

Page 98: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Supramolecular control over donor-acceptor photoinduced charge separation

87

(Figure 4.4, inset) suggests that several columns could be interacting with each other laterally

resulting in multi-columnar architectures with a different overall chirality. The formation of this

highly collapsed state is driven by the high tendency of perylene diimid chromophores to aggregate.18

In summary, absorption and CD spectroscopy show that OPV-FOLD-PERY adopts different

conformational states depending on the volume percentage of heptane (Cartoon 4.1). In pure

chloroform the OPV-FOLD-PERY dyad is in a random coil conformation (state A) and upon the

addition of heptane it folds into the well-known helical conformation (state B). Above 70% heptane,

intermolecular aggregation of the helical stacks into columnar architectures (state C) occurs. Above

95% heptane, several columns may interact with each other laterally resulting in a highly compact

conformational state (state D) that resembles the solid state but still is soluble. In accordance with

previous studies on the oligomeric bridge, it is assumed that the bridge chain is in one of the proposed

states and the observed UV/Vis and CD spectra are linear combinations of the spectra of these four

conformations.11-13

A B C D

Donor AcceptorBridge

A B C DA B C D

Donor AcceptorBridgeDonor AcceptorBridge

Cartoon 4.1. Schematic drawing of the different conformational states of OPV-FOLD-PERY with

decreasing polarity of the medium (from left to right). A: random coil; B: folded (helical)

conformation; C: columnar stacks; D: aggregated state.

4.3.2 Folding in other solvents

The number of solvents in which the bridge of OPV-FOLD-PERY can fold is rather high.19

Different As/Ap ratios are observed in different solvents (Table 4.1) and indicate that OPV-FOLD-

PERY adopts a varying degree of intramolecular folding and intermolecular aggregation. In general,

the As/Ap ratio is lower for less polar solvents. The As/Ap ratio increases with Reichardt’s ENT index.20

The ENT index has found acceptance in the literature as a reliable and convenient measurement of

solvochromatic effects of solvents of different polarity.

Page 99: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 4

88

Table 4.1. As/Ap absorption ratios observed for OPV-FOLD-PERY dissolved in solvents of different

ENT.

solvent ENT As/Ap

heptane - 0.65

cyclohexane 0.006 0.61

carbon tetrachloride 0.052 0.76

dioxane 0.164 0.74

tetrahydrofuran 0.207 0.83

chloroform 0.259 0.90

The lowest As/Ap ratio is observed for the most apolar solvents, heptane and cyclohexane.

Only in these two solvents a Cotton effect is observed (Figure 4.7). The absence of optical activity

together with a low As/Ap absorption ratio in solvents like tetrahydrofuran, dioxane, and carbon

tetrachloride, indicates that the conformation of the bridge in solvents of medium polarity is restricted

to random coil or helical conformations of the bridge, and that no aggregation occurs (only

conformational states A and B are present).

250 300 350 400 450 500 550

-80

-60

-40

-20

0

20

40

60

CD

(m

deg)

Wavelength (nm)

THF Dioxane CCl

4

Cyclohexane Heptane

Figure 4.7. CD spectra of OPV-FOLD-PERY in different solvents.

4.4 Electronic properties and energetic considerations

To establish whether energy or electron transfer reactions between FOLD, OPV, and PERY

chromophores can be expected after photoexcitation, it is of interest to determine the energy of the

Page 100: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Supramolecular control over donor-acceptor photoinduced charge separation

89

various electronic states. Absorption and fluorescence spectroscopy in combination with cyclic

voltammetry have been used to obtain these parameters.

Absorption and photoluminescence spectroscopy. The UV/Visible absorption spectrum of

OPV-FOLD-PERY in chloroform solution (Figure 4.3) is a near superposition of the spectra of the

individual chormophores. The para connectivity between the OPV and the FOLD bridge results in the

extension of the conjugation length of the OPV moiety to at least the first phenyl ring of the bridge.

The meta linkage of the next phenyl ring disrupts further conjugation. Thus, although the OPV only

contains two vinylene bonds, the observed absorption resembles that of an OPV with three vinylene

bonds.17 The PERY chromophore preserves its spectroscopic identity after being connected to the

bridge and the observed absorption spectra corresponds to the individual perylene diimid

chromophore.1,2

The photoluminescence spectra of the donor and acceptor moieties have been measured in

chloroform solution after selective excitation of the OPV and PERY chromophores in the OPV-FOLD

and FOLD-PERY model compounds respectively (Figure 4.8). The emission of the OPV

chromophore maximizes at 502 nm. The photoluminescence of the PERY consists of an emission

band with vibronic fine structure and has a maximum intensity at 535 nm.

Hence, absorption and photoluminescence spectroscopy reveal that in OPV-FOLD-PERY the

optical bandgap, and hence the energy of the excited singlet state, decreases from the FOLD bridge

(4.55 eV), via OPV (2.47 eV), to the PERY (2.32 eV) chromophore. As a consequence singlet-energy

transfer may occur in the series: 1FOLD* → 1OPV* → 1PERY* (Figure 4.9).

450 500 550 600 650 7000

50

100

150

Inte

nsity

(co

unts

/104 )

Wavelength (nm)

Figure 4.8. Fluorescence spectra of OPV-FOLD (dashed line) and FOLD-PERY (dotted line) after

photoexcitation of the OPV and PERY chromophores respectively.

Page 101: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 4

90

Cyclic voltammetry. The oxidation and reduction potentials of the redox centers have been

measured for the donor OPV-FOLD and acceptor FOLD-PERY reference compounds and are

assumed to be identical in the OPV-FOLD-PERY dyad. The cyclic voltammogram of OPV-FOLD

exhibits two reversible oxidation waves at 0.83 and 1.05 V, while for FOLD-PERY two reversible

reduction waves are found at –0.60 and –0.82 V (potentials are given vs. SCE, recorded in

dichloromethane with 0.1 M TBAPF6). In the –1.0 to +1.0 V range no redox reactions occur that are

associated with the FOLD bridge.

Energetic considerations. The most probable photophysical processes after photoexcitation

of either donor or acceptor in dyad OPV-FOLD-PERY are schematically represented in Figure 4.9.

Photoinduced electron transfer from the OPV donor to the PERY acceptor can only occur

when the OPV•+-FOLD-PERY•– charge-separated state is lower in energy than the PERY singlet-

excited state (OPV-FOLD-1PERY*) and competes with the intrinsic relaxation processes (mainly

fluorescence) of the 1PERY* state. Contributions of the FOLD bridge in charge transfer processes,

either in the random coil or helical stacked conformations, are not likely because of the high redox

potential of the FOLD segment.

OPV-FOLD-PERY

kET

kCR

kPLOPV

kPLPERY

Energy

1OPV*-FOLD-PERY

OPV-FOLD-1PERY*

OPV+-FOLD-PERY-

kCSPERY

kCSOPV

OPV-1FOLD*-PERY

kPLFOLD

kET

kET

OPV-FOLD-PERY

kET

kCR

kPLOPV

kPLPERY

EnergyEnergy

1OPV*-FOLD-PERY

OPV-FOLD-1PERY*

OPV+-FOLD-PERY-

kCSPERY

kCSOPV

OPV-1FOLD*-PERY

kPLFOLD

kET

kET

Figure 4.9. Schematic diagram describing the energy levels of the singlet and charge-separated states

of OPV-FOLD-PERY.

Page 102: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Supramolecular control over donor-acceptor photoinduced charge separation

91

Most of the folding experiments of the bridge in OPV-FOLD-PERY presented in this chapter

are performed by the addition of heptane to a chloroform solution of the dyad. Two parameters with

opposite effects on the charge separation reaction change as the amount of heptane is increased. On

one hand, at high volume percentages of heptane, folding of the bridge occurs which reduces the

distance between donor and acceptor and, thereby, favors the charge transfer. On the other hand, the

low polarity of heptane compared to chloroform, makes this solvent less effective in screening the

photogenerated charges and thereby will increase the energy of the OPV•+-FOLD-PERY•– charge-

separated state.

An estimate of the change in Gibbs free energy for photoinduced electron transfer in solution

for the OPV-FOLD-PERY dyad can be obtained from the Weller equation: 21

( ) ( )( )

+−−−−=∆ −+

sref0

2

0

2

00redox0 1111

84AD

εεπεεπε rr

e

R

eEEEeG

ccsCS (4.1)

In this equation, Eox(D) and Ered(A), are the oxidation and reduction potentials of the donor

and acceptor molecules or moieties respectively measured in a solvent with relative permittivity εref,

E00 is the energy of the excited state from which the electron transfer occurs (PERY singlet-excited

state), and Rcc is the center-to-center distance of the positive and negative charges in the charge-

separated state. The radii of the positive and negative ions are given by r+ and r- and εs is the relative

permittivity of the solvent, -e is the elemental charge, and ε0 is the vacuum permittivity.

The limiting distance at which charge separation becomes no longer exergonic in heptane is

21 Å (Table 4.2), as calculated from equation 4.1 taking Eox(D) = 0.83 V, Ered(A) = -0.60 V and

setting the ionic radii to r– = 4.7 Å2 and r+ = 5.6 Å.15 Larger distances will result in an OPV•+-FOLD-

PERY•– charge-separated state that is less stable than the PERY singlet excited state.

Table 4.2. Change in Gibbs free energy change for intramolecular and intermolecular electron

transfer in OPV-FOLD-PERY dyad in chloroform and heptane calculated from equation 4.1.

Solvent Rcc

(Å) CSG∆ (PERY (S1))

(eV)

CSG∆ (OPV (S1))

(eV)

Heptane 106 0.28 0.11

21 0.00 -0.17

Chloroform 106 -0.61 -0.78

21 -0.73 -0.90

Page 103: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 4

92

In the more polar solvent chloroform, charge separation is in principle feasible ( CSG∆ <0)

even when the molecule is in a completely extended configuration (distance ~106 Å). However,

kinetic factors will impede the intramolecular charge transfer at such large distances, because the

electronic coupling between donor and acceptor in the excited state is negligible at this distance.

4.5 Photoinduced energy and electron transfer in different conformational states of the bridge

The interaction between donor and acceptor in the excited state can be investigated with

photoluminescence and photoinduced absorption spectroscopies. Quenching of either donor or

acceptor fluorescence indicates an energy or charge transfer process. Photoinduced absorption

spectroscopy (PIA) can be used to monitor the formation and decay of the OPV•+-FOLD-PERY•–

charged separated state. In the following paragraphs the donor acceptor interaction in the excited state

is first described for a random coil conformation followed by the results for folded conformations and

aggregated states.

4.5.1 Bridge in a random coil conformation

In chloroform, the bridge is in a random-coil conformation. Hence, the relative orientation of

donor and acceptor chromophores for OPV-FOLD-PERY is not well defined and probably numerous

conformers exist.

500 600 7000

50

100

150

OPV-FOLD-PERY OPV-FOLD

Inte

nsity

(co

unts

/104 )

Wavelength (nm)

500 600 7000

50

100

150

OPV-FOLD-PERY FOLD-PERY

Inte

nsity

(co

unts

/104 )

Wavelength (nm)

Figure 4.10. Left: Photoluminescence spectra of OPV-FOLD-PERY and OPV-FOLD in chloroform

solution after selective photoexcitation of the OPV moiety at 380 nm. Right: Photoluminescence

spectra of OPV-FOLD-PERY and FOLD-PERY in chloroform solution after selective photoexcitation

of the PERY moiety at 512 nm.

Photoluminescence spectroscopy reveals that, in chloroform, the two chromophores interact

in the excited state. Selective photoexcitation of the OPV unit in OPV-FOLD-PERY at 380 nm results

in emission of the PERY part (λem, max = 537 nm), concomitant with a reduction of the fluorescence

Page 104: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Supramolecular control over donor-acceptor photoinduced charge separation

93

intensity of the OPV chromophore (λem, max = 504 nm) compared with the fluorescence of the OPV-

FOLD compound (Figure 4.10, left).

In contrast, the quantum yield of the PERY emission, after selective photoexcitation of the

PERY moiety in OPV-FOLD-PERY at 512 nm, approaches that of the FOLD-PERY reference

compound (Figure 4.10, right). According to the same diagram in Figure 4.9, a quenching of the

PERY emission would indicate the formation of the OPV•+-FOLD-PERY•– charge-separated state.

Because the PERY emission is not significantly quenched, photoinduced charge separation is

negligible in chloroform. In a random coil conformation, the OPV-FOLD-PERY dyad resembles D-

bridge-A systems with a saturated hydrocarbon chain as a bridge, albeit with less conformational

freedom. Out of all possible conformations in chloroform, barely a few will provide sufficient

closeness between the OPV donor and the PERY acceptor to allow for charge transfer to occur.

The excitation spectrum of the PERY emission nearly coincides with the absorption spectrum

of OPV-FOLD-PERY (Figure 4.11). The overlap is particularly good in the region of the PERY and

OPV absorption. This implies that singlet-energy transfer to the PERY chromophore occurs after

photoexcitation of OPV chromophore, and to a considerable extent also after exciting the FOLD

(Figure 4.9). However, there is a small mismatch between the absorption and excitation spectra in the

region of the FOLD absorption, indicating that energy transfer from the bridge is not complete. The

excitation spectrum of the OPV emission also reveals that energy transfer occurs from the FOLD

bridge onto the OPV chromophore. The difference between the absorption and excitation spectra is

larger for the OPV moiety. This indicates that some, but not all, of the photoexcitations of the FOLD

oligomer transfer their excited state energy first to the OPV chromophore, before ending up at the

PERY moiety.

300 400 500 6000

1

2

3

4

Inte

nsity

(a.

u.)

Wavelength (nm)

Figure 4.11. Excitation spectra of the 496 nm (solid triangles) and 535 nm (solid squares) emission

and absorption spectrum of OPV-FOLD-PERY, in chloroform.

Page 105: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 4

94

Because the excitation spectrum of OPV-FOLD-PERY coincides with the absorption

spectrum above 350 nm, the quenching of the OPV fluorescence is mainly due to a intramolecular

photoinduced singlet-energy transfer from the 1OPV* towards the more stable PERY the singlet-

excited state (Figure 4.9). The quenching of the OPV emission can be used to estimate an average

distance between the donor and acceptor using the Förster model for energy transfer.22 This model,

which is based on a dipole-dipole interaction mechanism, provides an expression for the rate for the

energy transfer, FETk , via:

6

1

=

d

Rk cF

ET τ (4.2)

In this equation τ is the lifetime of the donor chromophore, cR is the critical transfer radius

and d the center-to-center distance between the redox centers, the parameter that we wish to

determine. Using the OPV fluorescence quenching (Q = 4.16) and the fluorescence lifetime of the

OPV segment in OPV-FOLD-PERY (τ = 1.41 ns) the experimental rate constant for energy transfer

of kET = 2 ns-1 can be obtained via the expression:

τ

1−= maz

ETQ

k (4.3)

cR can be calculated using the Förster equations:

45

6

1283

2)10ln(9000

nN

JR

F

c⋅⋅⋅

⋅⋅⋅⋅=

π

φ (4.4)

∫=

dvvf

dvv

vvf

J F )(

))(

).((4

ε

(4.5)

In expressions 4.4 and 4.5, φ is the luminescence quantum yield of OPV(S1) (φ = 0.84, for

OPV-FOLD), N the Avogadro constant, and n the refractive index of the solvent (1.444). FJ is the

overlap integral of the luminescence spectrum of the donor ( )(vf of OPV) on an energy scale (cm-1)

and the absorption spectrum of the acceptor ( )(vε of PERY). For OPV-FOLD-PERY the overlap

integral is FJ = 2.5 × 10-13 cm6/mol. The average distance between the chromophores is then d = 5.5

nm. This implies that the foldamer does not adopt an extended configuration, in which the center-to-

center distance is of roughly 10 nm, but as expected, a random conformation (Cartoon 4.2).

Page 106: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Supramolecular control over donor-acceptor photoinduced charge separation

95

5.5 nm

Cartoon 4.2. Chemical structure of one of the many possible random coil conformations of OPV-

FOLD-PERY in chloroform solution.

4.5.2 Bridge in a folded conformation

4.5.2.1 Folded bridge in heptane/chloroform mixtures

Steady-state and time-resolved photoluminescence spectroscopy. To study the influence

of the decreasing polarity and the associated folding and intermolecular aggregation of the FOLD on

the emission of the individual chromophores, experiments were performed on the OPV-FOLD and

FOLD-PERY model compounds in solvents with different heptane/chloroform content.

The fluorescence intensity of OPV-FOLD increases upon addition of heptane (Figure 4.12).

The absence of any quenching in OPV-FOLD with decreasing polarity excludes the possibility of

OPV-OPV aggregates in any heptane/chloroform mixture.

500 600 7000

20

40

60

80

100

120 Heptane

CHCl3

Inte

nsity

(co

unts

/104 )

Wavelength (nm)

Figure 4.12. Steady-state photoluminescence spectra of OPV-FOLD in chloroform/heptane mixtures

of different composition.

Page 107: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 4

96

0 20 40 60 80 100

50

100

150

500 550 600 650 700 7500

50

100

150

Inte

nsity

(co

unts

/104 )

Wavelength (nm)

Inte

nsity

(co

unts

/104 )

Volume % heptane in chloroform

0 50 100 1501

10

100

1000

10000

99% Heptane85% Heptane

80% Heptane70% Heptane

CHCl3

Inte

nsity

(co

unts

)

Time (ns)

Figure 4.13. Left: Steady-state photoluminescence spectra of FOLD-PERY model compound in

solutions of chloroform/heptane mixtures of different composition. Inset: Maximum fluorescence

intensity of FOLD-PERY in chloroform/heptane mixtures of different composition (The fluorescence

intensities have been corrected for small deviations from O.D. = 0.1 at the excitation wavelength).

Right: Time profiles of the fluorescence at 600 nm of FOLD-PERY model compound in solutions of

chloroform/heptane mixtures of different composition.

Quite on the contrary, the photoluminescence of the acceptor chromophore in FOLD-PERY

undergoes a dramatic quenching starting from 80% of heptane in chloroform (Figure 4.13, left and

inset therein). This quenching coincides with a decreased intensity of the lowest energy PERY

absorption band (Figure 4.5) and is consistent with aggregation of the PERY chromophores and with

intermolecular aggregation of the helixes (state C) above 80% heptane. Moreover, at high heptane

contents, the photoluminescence spectra of FOLD-PERY feature an additional emission centered at

approximately 630 nm, associated with perylene excimers.23 Photoluminescence lifetime

measurements of the PERY emission at 600 nm reveal that the perylene aggregates (with an excited-

state lifetime of τ = 30 ns vs. ~ 4 ns of molecularly dissolved PERY chromophores) start to appear

already at 70 volume percent heptane and become the predominant species in almost pure heptane

(Figure 4.13, right).

For OPV-FOLD-PERY, an increase of the percentage of heptane results in a quenching of the

PERY emission, regardless which of the two chromophores is excited (Figure 4.14, left and right). In

such apolar environment, the folding of the bridge (state B) and the formation of intermolecular stacks

(states C and D) decreases the distance between donor and acceptor. When the two chromophores are

in close proximity, charge separation may occur in the excited state. The lack of excimer emission at

630 nm for OPV-FOLD-PERY at all chloroform/heptane ratios, indicates the absence of perylene

diimid excimers and leaves charge separation as the most likely fluorescence quenching process

Page 108: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Supramolecular control over donor-acceptor photoinduced charge separation

97

(Figure 4.9). The absence of a red-shifted emission also implies that donor and acceptor do not form

an emissive exciplex prior to electron transfer.

450 500 550 600 6500

40

80

120

99% Heptane

CHCl3

Inte

nsity

(co

unts

/104 )

Wavelength (nm)550 600 650 700

0

40

80

120

99% Heptane

CHCl3

Inte

nsity

(co

unts

/104 )

Wavelength (nm)

Figure 4.14. Steady-state photoluminescence spectra of OPV-FOLD-PERY in chloroform/heptane

solutions of different composition after selective excitation of the OPV at 380 nm (left) and the PERY

at 510 nm (right).

Time-resolved fluorescence of OPV-FOLD-PERY measured at 600 nm (Figure 4.15) reveals

that the lifetime of the PERY emissions is ~ 4 ns up to 80% heptane, while above 80% slightly shorter

lifetimes are observed. The absence of a 30 ns lifetime component in the fluorescence decay curves

upon the addition of heptane supports the absence of perylene diimid excimers in the aggregated

states in OPV-FOLD-PERY (states B to D, Cartoon 4.1). On the other hand, the absence of a

reduction of the lifetime above 80% heptane is surprising, because a reduction of the PERY emission

is observed in steady-state fluorescence (Figure 4.14, right). This suggests that the electron transfer

reaction that causes the PL quenching is so fast in state B that that the corresponding lifetime is

shorter than the time resolution of the experimental set up and is not detected by time-resolved

photoluminescence. Such a fast formation of the charge-separated state is rarely observed in solution,

unless donor and acceptor are directly connected1 or in a face-to-face orientation,5h,k,o especially when

taking into account that the decreasing polarity of the medium usually slows down the rate for the

charge separation. Hence, within the folded state (state B) only in conformations where donor and

acceptor are placed in a face-to-face orientation, charge transfer may take place as a predominant

decay route, while for less favorably folded conformers radiative decay of 1PERY* will occur

(Cartoon 4.3).

Page 109: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 4

98

0 20 40 60 80 1001

10

100

1000

10000 100 % Heptane 80% Heptane 70% Heptane 60% Heptane 30% Heptane 0% Heptane

Inte

nsity

(co

unts

)

Time (ns)

Figure 4.15. Time-resolved photoluminescence spectra of OPV-FOLD-PERY in chloroform/heptane

solutions of different composition.

-+-+-+

Cartoon 4.3. Conformational states B of OPV-FOLD-PERY. Only those molecules with appropriate

orientation of donor and acceptor yield the OPV+-FOLD-PERY– charge-separated state after

photoexcitation. In other conformations, relaxation to the ground state takes place via PERY

emission.

In the region above 80% heptane, where intermolecular and inter-columnar aggregation

occurs (states C and D, Cartoon 4.1), more donor and acceptors are brought in close proximity and

intermolecular charge separation can occur in addition to the intramolecular process.

A comparison of the steady-state photoluminescence of PERY-FOLD and OPV-FOLD-PERY

gives an estimate of the extent of charge transfer in the dyad. The significant quenching of the PERY

photoluminescence upon addition of heptane to chloroform solutions of OPV-FOLD-PERY strongly

suggests that charge separation occurs to yield the OPV•+-FOLD-PERY•– charge-separated state. For

low amounts of heptane the charge separation takes place only to a small extent and upon folding and

intermolecular aggregation in OPV-FOLD-PERY the quantum yield for charge transfer increases.

Above 80% heptane the comparison of the two intensities is no longer valid, because the excimer

Page 110: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Supramolecular control over donor-acceptor photoinduced charge separation

99

formation of FOLD-PERY results in a quenching of the fluorescence and that is not associated with

charge separation. The observed photoluminescence quenching supports the proposed conformational

exchanges in OPV-FOLD-PERY inferred from absorption and CD spectroscopy.

0 20 40 60 80 100

0

4

8

τ (n

s)

Volume % heptane in chloroform

50

100

150

CS

Inte

nsity

(co

unts

/104 )

Figure 4.16. Photoluminescence intensities (above) and lifetimes (below) of the PERY emission in

OPV-FOLD-PERY (solid squares) and FOLD-PERY (open squares) in chloroform/heptane mixtures

of different composition. The extent of charge separation can be estimated from the difference

between the two emission intensities. The arrows denote the amount of the charge separation (CS).

Femtosecond pump-probe spectroscopy. The formation and decay of the photoinduced

charges have been followed by means of transient photoinduced absorption spectroscopy. In this

experiment, the OPV chromophore is excited at 450 nm and the absorption at 1450 nm associated

with the OPV•+ radical cation is monitored in time.2

At early stages of intermolecular aggregation of the bridge in OPV-FOLD-PERY (60%

volume heptane in chloroform, state B), the differential transmission at 1450 nm is not very intense

(Figure 4.17, left). The small signal is in agreement with the low quenching of the PERY emission

observed at 60% heptane (Q ≈ 2, Figure 4.16). The time profile of the 1450 nm transient absorption

indicates that the charges are formed within 1 ps. This rapid formation of the OPV•+-FOLD-PERY•–

charge-separated state is consistent with the absence of short lifetimes in the time-resolved

photoluminescence of the PERY chromophore and is attributed to occur for conformations of OPV-

FOLD-PERY with a face-to-face orientation of donor and acceptor moieties (Cartoon 4.3). The decay

of the signal cannot be fitted to a single exponential function, suggesting that several charged species

are involved in the measurement. A feasible explanation for this result could be the conformational

Page 111: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 4

100

heterogeneity of the system: in the folded conformation of the bridge (state B), the position between

donor and acceptor is dynamic and not all conformers provide the face-to-face orientation between

donor and acceptor required for immediate charge transfer. The fact that the signal is low suggests

that few systems have the required face-to-face orientation, and the fact that the signal does not grow

with time implies that the conformational changes are slow on the time scale of the experiment.

-20 0 20 40 60 80 100-2.5

-2.0

-1.5

-1.0

-0.5

0.0

0.5

1.0

∆T

(a.

u.)

Time (ps)

0 100 200 300 400 500-120

-100

-80

-60

-40

-20

0

∆T (

a. u

.)

Time (ps)

Figure 4.17. Differential transmission dynamics for OPV-FOLD-PERY in mixtures of

chloroform/heptane of 60:40 (left) and 99:1 (right) at the radical cation absorption of OPV at 1450

nm with excitation of primarily OPV at 450 nm from –20 to 100 ps (left) and –100 to 500 ps (right).

At almost 100% volume heptane, the transient absorption is very intense (Figure 4.17, right).

This is consistent with the much higher PL quenching and attributed to intermolecular photoinduced

electron transfer in the aggregated state. The helical conformations have folded into columnar

architectures, and even several columns may be interacting with each other (state D, Cartoon 4.1).

More donor and acceptors are brought in close proximity and apart from intramolecular processes,

also intermolecular charge separation takes place. Under these conditions, the formation of the

charges is immediate, even though the heptane content is the highest and the medium is strongly

apolar. In strongly aggregated states the polarity of the solvent is less relevant, because charge transfer

takes place as it would do in a solid-state blend of the donor and acceptor. Paddon-Row et al.

observed that in a face-to-face orientated donor-acceptor dyad, photoinduced electron transfer took

place within 1 ps, irrespective of the polarity of the medium.5b Again the decay of the 1450 nm signal

cannot be fitted to a single exponential function. However, the charged-separated states are much

longer lived at high percentage of heptane, than in 60% heptane. These results can be rationalized by

migration of the charges to different parts of the aggregates.

Page 112: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Supramolecular control over donor-acceptor photoinduced charge separation

101

4.5.2.2 Folded bridge in other solvents

Analogous to the behavior in different heptane/chloroform mixtures, a decrease of the solvent

polarity results in an increasing quenching of the PERY photoluminescence in OPV-FOLD-PERY

(Figure 4.18, left). The quenching in these solvents is also attributed to photoinduced charge

separation from the OPV to the PERY chromophore caused by the folding of the bridge. There seems

to be a direct relation between the As/Ap ratio and fluorescence quenching in different solvents of

different polarity (Figure 4.18, right). For example, the observed As/Ap ratio and the quenching of the

PERY emission are comparable in carbon tetrachloride and a mixture of heptane/chloroform (60:40)

(Figure 4.16 and 4.18 left).

500 600 7000

50

100

150

200

Inte

nsity

(co

unts

/104 )

Wavelength (nm)

CHCl3

THF Dioxane CCl

4

Cyclohexane Heptane

-5 0 5 10 15 20 25 30 350.5

0.6

0.7

0.8

0.9

1.0

AS/A

P

Quenching

Figure 4.18. CD and PL spectra of OPV-FOLD-PERY in different solvents (left) and quenching ratio

vs. the AS/AP ratio in chloroform/heptane mixtures of different composition (open squares) and in pure

solvents of different polarity (solid squares) (right).

These results show that the conformational folding of OPV-FOLD-PERY and the resulting

enhancement of the photoinduced charge transfer is a general phenomenon that can be achieved and

controlled by modifying the polarity of the solvent, either in solvent mixtures of by varying the nature

of the pure solvent.

4.5 Conclusions

A novel donor-bridge-acceptor system has been synthesized by connecting OPV donor and

PERY acceptor chromophores at the opposite ends of a long foldable cross-conjugated oligomer,

FOLD. The conformation of this bridge can be controlled between random coil and helically folded

states, by changing the polarity of the medium.

In chloroform solution the bridge is in a random coil conformation. As a consequence most

donor and acceptor chromophores are too far apart to undergo a photoinduced charge transfer

reaction, and only energy transfer from the 1OPV* singlet excited state to the PERY occurs (Figure

Page 113: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 4

102

4.9), as inferred from photoluminescence spectroscopy (Figure 4.10). The average distance between

OPV and PERY chromophores in chloroform, as estimated by the Förster equations for energy

transfer, is 5.5 nm. This distance supports the idea that the FOLD oligomer in chloroform is not in a

completely extended conformation (~ 10 nm) but adopts a random coil conformation.

Upon addition of heptane to chloroform solutions of OPV-FOLD-PERY, the bridge folds up

into a helix (state B, Cartoon 4.1) and, with increasing amount of heptane, these helices aggregate into

columnar architectures (state C, Cartoon 4.1). In almost pure heptane (starting from 80% heptane in

chloroform), an additional higher degree of aggregation has been inferred from further changes in

absorption and CD spectra (Figure 4.4). The observed inversion of the chirality suggests that several

columns might interact with each other laterally, resulting in multi-columnar architectures with a

different overall chirality (state D, Cartoon 4.1). A comparison of the spectra of the OPV-FOLD-

PERY dyad with the model compounds OPV-FOLD, FOLD-PERY, and the unsubstituted bridge

FOLD shows that state D only forms in compounds containing the PERY chromophore. Possibly, the

multi-columnar architectures in state D originate from the tendency of the PERY chromophore to

aggregate. In the collapsed states of OPV-FOLD-PERY, OPV and PERY are prone to be involved in

a photoinduced charge transfer, because of a decreased distance. In state B, intramolecular

photoinduced charge transfer occurs within 1 ps. The high speed suggests that electron transfer only

occurs in (helical) conformers that provide a face-to-face orientation of donor and acceptor (Cartoon

4.3). In states C and D, two phenomena favor photoinduced electron transfer. On one hand, more

chromophores are placed in close proximity. On the other hand, under aggregated conditions, the

chromphores experience a ‘solid state environment’ and, hence, the solvent is less important with

respect to screening of charges and solvating ions.

The folding of the bridge and the associated changes in photoinduced reactions of OPV-

FOLD-PERY not only occurs for high heptane/chloroform ratios, but also is a general phenomenon in

less polar solvents.

In conclusion, it has been shown that the bridge in OPV-FOLD-PERY provides a unique

means to change the distance and orientation, and thereby the photophysical interaction between

donor and acceptor. Full control over the hierarchical build-up of a supramolecular D-A system has

been obtained, starting from two semi-isolated chromophores showing only energy transfer, via

intermediates capable of rendering electron transfer, into an aggregated state featuring, in addition to

electron transfer, charge migration. This donor-foldamer-acceptor dyad can be considered as a first

step towards the supramolecular construction of synthetic architectures that are able to perform

complicated photophysical and photochemical processes, similar to those observed in the

photosynthetic reaction center. Such appealing synthetic analogues may find future application in

photocatalytic reactions and solar light energy conversion.

Page 114: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Supramolecular control over donor-acceptor photoinduced charge separation

103

4.6 Experimental section

For general methods and materials the reader is referred to chapter 2 of this thesis. N-(1-Ethylpropyl)-N’-(4-iodophenyl) perylene diimide (2). Perylene monoanhydride monoimide 1 (38.6 mg, 0.08 mmol), 4-iodoaniline (182 mg, 0.83 mmol), imidazole (0.53g, 7.8 mmol), and catalytic amounts of Zn(AcO)2 were mixed and stirred for 1 h at 160 °C. After cooling to room temperature the solid reaction mixture was extensively washed with methanol to yield 49 mg of 2 (92%) as a dark red powder. 1H NMR (CDCl3, 300 MHz): δ 8.66 (t, 4H), 8.56 (d, 4H), 7.90 (d, 2H), 7.12 (d, 2H), 5.11-5.00 (m, 1H), 2.35-2.17 (m, 2H), 2.03-1.89 (m, 2H), 0.95 (t, 6H); 13C NMR (CDCl3, 100 MHz): δ 138.65, 135.17, 134.77, 134.09, 131.81, 131.38 (broad signal), 130.60, 129.72, 129.46, 126.57, 126.29, 123.32, 123.00, 94.66, 57.81, 25.01, 11.38. MALDI-TOF MS (Mw =662.49) m/z=662.19 [M]+. (E,E)-4-{4-(4-Methyl-2,5-bis[(S)-2-methylbutoxy]styryl)-2,5-bis[(S)-2-methylbutoxy]styryl}-2,5-bis[(S)-2-methylbutoxy]-4-trimethylsylilethynylbenzene (4). Bromide 3 (0.2 g, 0.22 mmol), PdCl2 (3.97 mg, 0.02 mmol), PPh3 (17.6 mg, 0.07 mmol) and Cu(AcO)2 (4.47 mg, 0.02 mmol) were dissolved in 5 mL anhydrous triethylamine. Argon was purged trough the solution for 30 minutes after which (trimethylsilyl)acetylene (44 mg, 0.44 mmol) was added. The reaction mixture was heated at 80 °C for 16 h. The solvent was removed in vacuo, the residue dissolved in methylene chloride and washed with NH4Cl, water and dried over MgSO4. Purification by column chromatography (chloroform:pentane, 1:1, Rf= 0.2) afforded 92 mg (45%) of compound 4. 1H NMR (CDCl3, 300 MHz): δ 7.52 (d, 1H), 7.50 (d, 1H), 7.45 (d, 1H). 7.44 (d, 1H), 7.18 (s, 1H), 7.16 (s, 1H); 7.12 (s, 1H), 7.10 (s, 1H), 6.94 (s, 1H), 6.72 (s, 1H), 3.92-3.67 (m, 12H), 2.22 (s, 3H), 2.00-1.84 (m, 6H), 1.70-1.55 (m, 6H), 1.40-1.25 (m, 6H). 1.11-1.02 (m, 18H), 1.02-0.95 (m, 18H), 0.27 (s, 9H); 13C NMR (CDCl3, 100 MHz): δ 154.91, 151.67, 151.20, 150.93, 150.46, 150.20, 129.03, 127.95, 127.62, 126.76, 125.15, 123.97, 123.20, 122.16, 121.64, 117.38, 116.28, 112.06, 110.14, 109.92, 109.60, 108.38, 101.79, 98.91, 74.66, 74.25, 74.14, 73.38, 35.10, 35.04, 35.01, 34.98, 34.95, 26.36, 26.32, 26.25, 26.14, 16.78, 16.70, 16.52, 16.40, 11.44, 11.36, 0.01. MALDI-TOF MS (Mw = 909.37) m/z=908.35 [M]+. (E,E)-4-{4-(4-Methyl-2,5-bis[(S)-2-methylbutoxy]styryl)-2,5-bis[(S)-2-methylbutoxy]styryl}-2,5-bis[(S)-2-methylbutoxy]-4-ethynylbenzene (5). To a solution of 4 (36.7 mg, 0.04 mmol) in dry THF was added 1 M tetrabutylammonium fluoride in THF (0.044 mL). The reaction mixture was stirred for 1 min and subsequently filtered over silica gel using chloroform as eluent. The solvent was removed in vacuo yielding 32 mg (95%) of a yellow solid which was used without further purification. 1H NMR (CDCl3, 300 MHz): δ 7.53 (d, 1H), 7.51 (d, 1H), 7.45 (d, 1H), 7.44 (d, 1H), 7.18 (s, 1H), 7.16 (s, 1H); 7.15 (s, 1H), 7.10 (s, 1H), 6.97 (s, 1H), 6.72 (s, 1H), 3.92-3.67 (m, 12H), 3.31 (s, 1H), 2.22 (s, 3H), 2.00-1.84 (m, 6H), 1.70-1.55 (m, 6H), 1.40-1.25 (m, 6H). 1.11-1.02 (m, 18H), 1.02-0.95 (m, 18H). MALDI-TOF MS (Mw = 837.25) m/z = 836.51 [M]+. PERY-FOLD (7). A mixture of 2 (11 mg, 0.016 mmol), 5 (39.8mg, 0.0014 mmol), Pd(PPh3)4 (1.06 mg, 0.0009 mmol) and potassium acetate (2 mg, 0.020 mmol) in dry toluene (1 mL) and N,N-dimethylformamide (1 mL) was heated to 100 °C for 24 hours, and subsequently the solvent was evaporated in vacuo. The residue was triturated with water and extracted with methylene chloride. The methylene chloride extract was worked up to give crude products, which were purified by flash column chromatography (CH2Cl2/pentane to 9:1, Rf = 0.4) to yield 10 mg (22 %) of 4 as a light red solid. 1H NMR (CDCl3, 400 MHz): δ 8.77-8.67 (m, 8H), 8.20-8.17 (m, 22 H), 8.03 (ddd, 1H), 7.92-7.87 (m, 11H), 7.75 (d, 2H), 7.70 (ddd, 1H), 7.45 (t, 1H), 7.38 (d, 2H), 5.10-5.02 (m, 1H), 4.50-4.32 (m, 24 H), 2.33-2.22 (m, 2H), 2.00-1.90 (m, 2H), 1.90-1.78, (m, 12 H), 1.70-1.47 (m, 36H), 1.40-1.12(m, 72 H), 1.00-0.90 (m, 42 H), 0.88-0.84 (m, 72H); 13C NMR (CDCl3, 100 MHz): δ 165.23, 165.18, 165.12, 163.43, 138.24, 135.70, 135.34, 134.31, 132.74, 132.67, 132.42, 131.97, 131.44, 131.37, 130.95, 129.90, 129.71, 129.58, 128.93, 128.56, 126.77, 126.46, 124.06, 123.89, 123.58, 123.49, 123.42, 123.12, 122.99, 90.34, 89.18, 89.11, 89.02, 88.92, 64.30, 63.96, 57.76, 50.89, 39.18, 37.10, 35.51, 29.94, 29.70, 27.94, 25.01, 24.59, 22.69, 22.59, 19.66, 11.35. MALDI-TOF MS (Mw =3949.19) m/z= 3948.18 [M]-. OPV-FOLD (9). A mixture of 8 (38.7 mg, 0.01 mmol), 5 (11.7 mg, 0.012 mmol), Pd3(dba)2 (2.4 mg, 0.003 mmol), PPh3 (7mg, 0.03 mmol), CuI (1mg, 0.005 mmol) dissolved in dry triethylamine (1 mL) was purged with Ar for 15 min and subsequently stirred and heated at 70 °C for 16 h. The solution was filtered over celite to remove the precipitate

Page 115: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 4

104

and concentrated in vacuo. The crude product was purified by pressure column chromatography (CH2Cl2/pentane 1:1, Rf = 0.1 and CHCl3/pentane 3:2 to 4:1, Rf = 0.1) to yield 23 mg (48%) of a yellow waxy solid. 1H NMR (CDCl3, 400 MHz): δ 8.22 -8.14 (m, 21 H), 8.38-8.18 (m, 2H), 8.08 (t, 1H), 7.89-7.87 (m, 12 H), 7.86-7.84 (m, 2H), 7.81(t, 1H), 7.42-7.59 (m, 4H), 7.19 (s, 2H), 7.17 (s, 1H), 7.10 (s, 1H), 7.02 (s, 1H), 6.73 (s, 1H), 4.48-4.32 (m, 24H), 4.00-3.74 (m, 12H), 2.25 (s, 3H), 2.05-1.78 (m, 18H), 1.70-1.47 (m, 42H), 1.40-0.9 (m,150H), 0.90-0.81 (m, 72H), 0.26 (s, 9H); 13C NMR (CDCl3, 75 MHz): δ 165.30, 165.09, 154.58, 151.67, 151.24, 150.93, 150.47, 150.36, 138.68, 138.23, 138.01, 132.67, 131.44, 131.26, 131.18, 129.46, 128.08, 127.64, 126.65, 125.13, 124.85, 124.22, 124.10, 123.59, 123.27, 122.04, 121.62, 117.07, 116.27, 111.49, 110.18, 109.87, 109.65, 108.36, 92.09, 89.38, 89.03, 88.77, 88.62, 74.64, 74.28, 73.37, 64.28, 39.18, 37.10, 35.51, 35.11, 34.99, 29.93, 27.93, 26.34, 26.28, 24.59, 22.68, 22.59, 19.65, 16.80, 16.69, 11.47, 11.37, -0.22. MALDI-TOF MS (Mw = 4321.78) m/ z= 4320.17 [M]+. OPV-FOLD-PERY (10) Similar to the procedure for the synthesis of PERY-FOLD a mixture of 2 (6 mg, 0.009 mmol), 9 (31 mg, 0.0072mmol), Pd(PPh3)4 (1 mg, 0.0008 mmol) and potassium acetate (1 mg, 0.01 mmol) in toluene (1 mL) and N,N-dimethylformamide (1 mL) was heated to 100 °C for 5 hours, the solvent was evaporated in vacuo. The residue was triturated with water and extracted with methylene chloride. The methylene chloride extract was worked up to give crude products, which were purified by pressure column chromatography (CH2Cl2/pentane to 9:1, Rf = 0.4) to yield 3 mg (9 %) of 10 as a light red solid. 1H NMR (CDCl3, 400 MHz): δ 8.76-8.65 (m, 8H), 8.20-8.15 (m, 23H), 8.12 (t, 1H), 7.91 (t, 1H), 7.90-7.87 (m, 10H), 7.84 (t, 1H), 7.74 (d, 2H), 7.56-7.41 (m, 4H), 7.38 (d, 2H), 7.18 (s, 2H), 7.16 (s, 1H), 7.09 (s, 1H), 7.01 (s, 1H), 6.72 (s, 1H), 5.11-5.02 (m, 1H), 4.45-4.35 (m, 24H), 3.97-3.74 (m, 12H), 2.31-2.24 (m, 5H), 2.05-1.78 (m, 20H), 1.70-1.47 (m, 42H), 1.40-0.95 (m, 156H), 0.88-0.84 (m, 72H).MALDI-TOF MS (Mw = 4784.90) m/ z= 4783.38 [M]+. Transient subpicosecond photoinduced absorption. Solutions in the order of 10-5 M were excited at 450 nm, i.e. providing primarily excitation of the OPV part within the molecules. 4.7 References

1 Peeters, E.;Van Hal, P. A.; Meskers, S. C. J.; Janssen, R. A. J.; Meijer, E. W. Chem. Eur. J. 2002, 8,

4470.

2 Neuteboom, E. E.; Meskers, S. C. J.; van Hal, P. A.; van Duren, J. K. J.; Meijer, E. W.; Dupin, H.;

Pourtois, G.; Cornil, J.; Lazzaroni, R.; Brédas, J.-L.; Beljonne, D.; Janssen, R. A. J. J. Am. Chem. Soc.

2003, 125, 8625.

3 Neuteboom, E. E.; Beckers, E. H. A.; Meskers, S. C. J.; Janssen, R. . A. J. Org. Biomol. Chem. 2003,

1,198.

4 Hauke, F.; Hirsch, A.; Liu, S.-G.; Echegoyen, L.; Swartz, A.; Luo, C.; Guldi, D. M. Chem. Phys. Chem.

2002, 3, 195.

5 (a) Williams, R. M.; Koeberg, M.; Lawson, J. M.; An, Y.-Z.; Rubin, Y.; Paddon-Row, M. N.; Verhoeven,

J. W. J. Org. Chem. 1996, 61, 5055. (b) Bell, T. D. M.; Smith, T. A.; Ghiggino, K. P.; Ranasinghe, M. G.;

Shephard, M. J.; Paddon-Row, M.. Chem. Phys. Lett. 1997, 268, 223. (c) Guldi, D. M.; Maggini, M.;

Scorrano, G.; Prato, M. J. Am. Chem. Soc. 1997, 119, 974. (d) Tkachenko, N. V.; Tauber, A. Y.;

Grandell, D.; Hynninen, P. H.; Lemmetyinen, H. J. Phys. Chem. A 1999, 103, 3646. (e) Tkachenko, N.

V.; Rantala, L.; Tauber, A. Y.; Helaja, J.; Hynninen, P. H.; Lemmetyinen, H.; J. Am. Chem. Soc. 1999,

121, 9378. (f) Thomas, k. George; Biju, V.; Guldi, D. M.; Kamat, Prashant V.; George, M. V. J. Phys.

Chem. A 1999, 103, 10755. (g) Armaroli, N.; Marconi, G.; Echegoyen, L.; Bourgeois, J.-P.; Diederich, F..

Chem. Eur. J. 2000, 6, 1629. (h) Koeberg, M.; deGroot, M.; Verhoeven, J. W.; Lokan, N. R.; Shephard,

Page 116: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Supramolecular control over donor-acceptor photoinduced charge separation

105

M. K.; Paddon-Row, M. N. J. Phys. Chem. A 2001, 105, 3069. (i) Vehmane, V.; Tkachenko, N. V.;

Tauber, A. Y.; Hynninen, P. H.; Lemmetyinen, H. Chem. Phys. Lett. 2001, 345, 213. (j) Daub, J.; Engl,

R.; Kurzawa, K.; Miller, S. E.; Schneider, S.; Stockmann, A.; Wasielewski, M. R., J. Phys. Chem. A

2001, 105, 5655. (k) Guldi, D. M.; Luo, C.; Prato, M.; Troisi, A.; Zerbetto, F.; Scheloske, M.; Dietel, E.;

Bauer, W.; Hirsch, A.. J. Am. Chem. Soc. 2001, 12, 9166. (l) Kreher, D.; Cariou, M.; Liu, S.-G.;

Levillain, E.; Veciana, J.; Rovira, C.; Gorgues, A.; Hudhomme, P.. J. Mat. Chem. 2002, 12, 2137. (m)

Schuster, D. I.; Jarowski, P. D.; Kirschner, A. N.; Wilson, S. R. J. Mater. Chem. 2002, 12, 2041. (n)

Schuster, D. I.; Jarowski, P. D.; Kirschner, A. N.; Wilson, S. R. J. Mater. Chem. 2002, 12, 2041. (o) Van

Hal, P. A.; Beckers, E. H. A.; Meskers, S. C. J.; Janssen, R. A. J.; Jousselme, P.; Roncali, J. Chem. Eur. J.

2002, 8, 5415.

6 a) Schenning, A. P. H. J.; v. Herrikhuyzen, J.; Jonkheijm, P.; Chen, Z.; Wurthner, F.; Meijer, E. W. J.

Am. Chem. Soc. 2002; 124(35); 10252. b) Beckers, E. H. A.; van Hal, P. A.; Schenning, A. P. H. J.; El-

Ghayoury, A.; Peeters, E.; Rispens, M. T.; Hummelen, Jan C.; Meijer, E. W.; Janssen, R. A. J. J. Mater.

Chem 2002, 12(7), 2054.

7 (a) Ballardani, R.; Balzani, V.; Clemente-León, M.; Credi, A.; Gandolfi, M. T.; Ishow, E.; Perkins, J.;

Stoddart, J. F.; Tseng, H.-R.; Wenger, S.. J. Am. Chem. Soc. 2002, 124, 12786. (b) Kercher, M; König, B;

Zieg, H.; De Cola, L. J. Am. Chem. Soc. 2002, 124, 11541. (c) Baran, P. S.; Monaco, R. R.; Khan, A. U.;

Schuster, D. I., Wilson, S. R. J. Am. Chem. Soc. 1997, 119, 8363. (d) Konishi, T.; Ikeda, A.; Kishida, T.;

Rasmussen, B. S., Fujitsuka, M.; Ito, O.; Shinkai, S. J. Phys. Chem. A. 2002, 106, 10254. (e) Ayabe, M.;

Ikeda, A.; Shinkai, S.; Sakamoto, S.; Yamaguchi, K. Chem. Comm. 2002, 1032. (f) D’Souza, F.;

Deviprasad, G. R.; Zandler, M. E.; El-Khouly, M. E.; Fujitsuka, M.; Ito, O. J. Phys. Chem. B 2002,

16(19), 4952. (g) Hauke, F; Swartz, A.; Guldi, D. M.; Hirsch, A. J. Mat. Chem. 2002, 12, 2088. (h)

D’Souza, F.; Gadde, S.; Zandler, M. E.; Arkady, K.; El-Khouly, M. E.; Fujitsuka, M.; Ito, O. J. Phys.

Chem. A 2002, 106, 12393.

8 (a) Endtner, J. M.; Effenberger, F.; Hartschuh, A.; Port, H. J. Am. Chem. Soc. 2000, 122, 3037. (b) Bahr,

J. L.; Kodis, G.; de la Garza, L.; Lin, S.; Moore, A. L.; Moore, T. A., Gust, D. J. Am. Chem. Soc. 2001,

123; 7124.

9 Baran, P. S.; Monaco, R. R.; Khan, A. U.; Schuster, D. I.; Wilson, S. R. J. Am. Chem. Soc. 1997, 119,

8363.

10 Polese, A.; Mondini, S.; Bianco, A.; Toniolo, C.; Scorrano, G.; Guldi, D. M.; Maggini, M. J. Am. Chem.

Soc. 1999, 121, 3446.

11 Nelson, J. C.; Saven, J. G.; Moore, J. S.; Wolynes, P. G. Science 1997, 277, 1793.

12 (a) Gin, M. S.; Yokozawa, T.; Prince, R. B.; Moore, J. S. J. Am. Chem. Soc. 1999, 121, 2643. (b) Prince,

R. B.; Saven, J. G.; Wolynes, P. G.; Moore, J. S.. J. Am. Chem. Soc. 1999, 121, 3114. (c) Prince, R. B.;

Brunsveld, L.; Meijer, E. W.; Moore, J. S.. Angew. Chem., Int. Edit., 2000, 39, 228. (d) Yang, W. Y.;

Prince, R. B.; Sabelko, Jo.; Moore, J. S.; Gruebele, M. J. Am. Chem. Soc. 2000, 122, 3248. (e) Brunsveld,

L.; Meijer, E. W.; Prince, R. B.; Moore, J. S. J. Am. Chem. Soc. 2001, 123, 7978.

Page 117: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 4

106

13 Brunsveld, L.; Prince, R.B.; Meijer, E.W; Moore, J.S. Org. Lett. 2000, 2, 1525.

14 Nagao, Y.; Naito, T.; Abe, Y.; Misono, T. Dyes and Pigments 1996, 32, 71.

15 Peeters, E.; van Hal, P. A.; Knol, J.; Brabec, C. J.; Sariciftci, N. S.; Hummelen, J. C.; Janssen, R. A. J. J.

Phys. Chem. B 2000, 104, 10174.

16 Akita, Y.; Kanekawa, H.; Kawasaki, T.; Shiratori, I.; Ohta, A. J. Heterocyclic Chem. 1988, 25, 975.

17 Peeters, E.; Marcos Ramos, A.; Meskers, S. C. J.; Janssen, R. A. J. J. Chem Phys. 2000, 112, 9445.

18 Würthner, F.; Thalacker, C.; Diele, S.; Tschierske, C. Chem. Eur. J. 2001, 7, 2245.

19 Hill, D. J.; Moore, J. S. Proc. Nat. Acad. Sci., 2002, 99, 5053.

20 Reichart, C.Solvents and Solvent Effects in Organic Chemistry (VCH, New York), 1990.

21 Weller, A. Z. Phys. Chem. Neue. Folge. 1982, 133, 93.

22 Förster. T. Discuss. Faraday Soc. 1959, 27, 7.

23 (a) Ford, W. E.; Kamat, P. V.; J. Phys. Chem. 1987, 91, 6373. (b) Gómez, U.; Leonhardt, M.; Port, H.;

Wolf, H. C. Chem. Phys. Lett. 1997, 268, 1. (c) Puech, K.; Fröb, H.; Leo, K. J. Lumin. 1997, 72-74, 524.

Page 118: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 5

Photoinduced electron transfer of

conjugated polymers with pendant

fullerenes

Abstract

A processable π-conjugated polymer bearing covalently linked methano-fullerenes has been

synthesized using the Sonogashira polycondensation reaction between a diiodoaryl functionalized

fullerene and a bisethynylterminated p-phenylene vinylene oligomer. The photophysical properties of

this donor-acceptor polymer have been studied and compared to an analogous polymer lacking the

pendant fullerenes. Photoexcitation of the donor-acceptor polymer results in a photoinduced charge

transfer reaction from the conjugated backbone to the pendant C60 moieties. This novel polymer was

applied via spin coating to form the active layer of the first polymer solar cell based on a covalently

linked donor-acceptor bulk-herojunction. The synthesis of analogous polymers with low-bandgap

characteristics was investigated using the Suzuki polycondensation.

Page 119: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 5

108

5.1 Introduction

Photoinduced electron transfer from a donor to an acceptor is widely studied to mimic the

natural photosynthetic reaction center and to investigate the prospects of molecular materials in

photovoltaic energy conversion.1 Promising energy conversion efficiencies have been obtained in so-

called bulk-heterojunction solar cells in which the active layer is a composite film of a conjugated

donor polymer and an acceptor polymer or a fullerene derivative.2,3 In these blends charges are

preferentially formed at the donor-acceptor interface and intimate mixing of donor and acceptor is

therefore beneficial for charge generation. For efficient transport of the positive charge carriers

through the donor phase and of electrons via the acceptor phase to the electrodes, a phase-segregated

bicontinuous network is required. A convenient route to obtain a predefined nanoscopic phase-

segregated network is linking donor and acceptor via a covalent bond. As a first step towards the

making covalently linked donor-acceptor materials, small fragments of π-conjugated polymers have

been connected to fullerene moieties in well-defined donor-acceptor molecular dyads.4,5 The same

concept can be extended to macromolecules by synthesizing π-conjugated polymers with pendant

fullerenes (Figure 5.1).

ITOMetal

electrode

e-e-

h+

h+

ITOMetal

electrode

e-e-

h+

h+

Figure 5.1. Cartoon of the working principle of a π-conjugated polymer with pendant fullerene

moieties in a solar cell.

The preparation of well-defined polymers incorporating fullerenes has remained a challenge

over the years and only few synthetic routes have ensured full structural homogeneity of the final

polymer.6 Conjugated polymers incorporating fullerenes have previously been prepared by

electrochemical polymerization of oligothiophene-fullerene dyads or by grafting C60 on precursor

polymers but these materials have not been incorporated in electro-optical devices because of

insolubility of the resulting polymers.7 Processable polymers with pendant fullerenes have been

prepared by random copolymerization of thiophenes bearing fulleropyrrolidine moieties with

thiophenes carrying solubilising agents using oxidative coupling with FeCl3.8 These polymers, that

Page 120: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced electron transfer of conjugated polymers with pendant fullerenes

109

were only soluble if the percentage of fullerene monomer in the feed was not very high, have

successfully been used as the active layer in photovoltaic devices. Examples of soluble polymers

having a backbone containing p-phenylene vinylene and carbazole units, and also triphenylamine

moieties have been reported.9

In this chapter, the synthesis of a novel processable π-conjugated polymer with covalently

linked methanofullerenes is described. The π-conjugated backbone is based on short p-phenylene

vinylene segments that are linked via p-phenylene ethynyl units. The polymer has been synthesized

using the Sonogashira palladium-catalyzed cross-coupling reaction. This polymerization can be

performed under mild conditions and is one of the reactions developed in recent years for the

synthesis of electronic polymer materials, which ensures the alternation of the two monomers. A

similar polymer that lacks the pendant methanofullerenes has also been synthesized and is used as a

reference. The photophysical processes occurring after photoexcitation of the polymer backbone have

been studied by means of photoluminescence and photoinduced absorption spectroscopies. The

polymer has been tested as the active layer of a polymer solar cell. Moreover, the synthesis of an

analogous polymer with low-bandgap properties has been explored using the Suzuki

polycondensation reaction.

5.2 PPV-PPE polymers with pendant fullerenes 5.2.1 Synthesis and characterization

Polymer 3 was synthesized using the Sonogashira palladium catalyzed cross-coupling

reaction between a diiodinated aryl compound 1 and an oligo(p-phenylene vinylene) with terminal

ethynyls 2 (Scheme 5.2). The pendant fullerenes were incorporated into the polymer via the

diiodinated monomer 1. Because of the well-known poor solubility of fullerene molecules, both

monomers have been decorated with numerous branched side chains in order to ensure solubility of

the resulting polymer, required for processability from solution.

Monomer 1, bearing the fullerene, was synthesized by Minze Rispens (University of

Groningen).10 Diethynylene monomer 2 was readily synthesized starting from methylhydroquinone

(6) (Scheme 5.1). Etherification of 6 with 2-ethylhexyl-p-toluenesulfonate, followed by radical

bromination using NBS in the presence of AIBN and ionic bromination with NBS, gave 8.

Phosphonate 9 was obtained by treatment of 8 with triethylphosphite. For the central unit of 1

etherification of 4-methoxyphenol with 3,7-dimethyloctyl-p-toluenesulfonate gave 11, which was then

brominated to give 12, followed by bis-formylation using butyllithium and N,N-dimethylformamide to

yield dialdehyde 13. A double Wittig Horner coupling of 9 and 13 gave 14 which was reacted with

Page 121: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 5

110

(trimethylsilyl)acetylene using a palladium-catalyzed coupling to afford monomer 2 after

deprotection.

OR

RO

OR

ROBr

Br

OR

ROP

Br

O

OEt

OEt

OR'

MeO

OR'

MeO

Br Br

OR'

MeO

O

O

OR'

MeO

X

X

OR

OR

RO

RO

O

O

O

H

H

O

O

O

6: R = H7: R = CH2CH(C2H5)(CH2)3CH3

8 9

10: R' = H11: R' = CH2CH2CH(CH3)(CH2)3CH(CH3)2

12 13

14: X = Br 15: X = C≡C-TMS

a

b

d

e f g

h

i

2

R = CH2CH(C2H5)(CH2)3CH3 R' = CH2CH2CH(CH3)(CH2)3CH(CH3)2

c

Scheme 5.1. a. CH3(CH2)3CH(C2H5)CH2OTs, K2CO3, TBAC, MEK, 93%; b. 1. NBS, AIBN, CCl4; 2.

NBS, THF, 24%; c. P(OEt)3, 160 °C, 1.5 h. 100%; d. (CH3)2CH(CH2)3CH(CH3)CH2CH2OTs; e. Br2,

HOAc, 65-116 °C, 2 h. 75%; f. 1. BuLi, DMF, Et2O, -10 °C, 56%; g. 9, KtBuO, DMF, 35%; p. TMS-

CCH, NEt3, PdCl2, PPh3, Cu(OAc)2, 50%; i. TBAF, THF, 100%

Polymer 3 was synthesized according to Scheme 5.2 under inert conditions using Pd(PPh3)4

and CuI as catalysts, in a mixture of 1,2-dichlorobenzene/triethylamine (7:3 v/v). The polymerization

reaction was followed by absorption spectroscopy, as a red shift in time of the π-π* transition of the

polymer with respect to that of monomer 2 (λmax= 428 nm) (Figure 5.1, left). Polymer 5, which is

similar to 3 but lacks the pendant methanofullerenes was synthesized by reaction of 2 and 4 under the

same conditions used for 3. The three hexyloxy chains ensure that 1 is a highly soluble C60 derivative,

but its reactivity towards 2 under these conditions is less than that of 1,4-diiodo-2,5-bis(2-

ethylhexyloxy)benzene (4). After 24 h of polymerization, the effective conjugation length of 3, as

inferred from λmax = 468 nm, approaches that of 5 (λmax= 474 nm) (Figure 5.1, left). The polymers

Page 122: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced electron transfer of conjugated polymers with pendant fullerenes

111

were isolated by precipitation into methanol. Polymer 3 consisted of a soluble brown powder and an

insoluble fraction. The partial insolubility is probably due to high molecular weight chains.8

O O

O

O

O

O

O

O

O

O n

O

O

O

O

O

O

H

H

O O

O

O

II

O

O

O

O

O

O

O

O n

O

O

IIPd(PPh3)4

CuI, NEt3ODCB

2

3

1

4

5

Scheme 5.2. Synthesis of polymers 3 and 5.

The molecular weight of the polymers determined by size-exclusion chromatography (SEC,

Figure 5.1, right) shows that the soluble fraction of 3 (Mw =16.2 kg/mol, PDI = 2.82) has a lower

degree of polymerization than 5 (Mw =35.8 kg/mol, PDI = 2.32), consistent with the small 6 nm

hypsochromic shift (Figure 5.1, left). The difference in SEC molecular weights might not reflect the

actual situation accurately. For monomer 2 the molecular weight determined by SEC corresponds to

the actual value (SEC: Mw = 1172 g/mol; MALDI-TOF Mw = 1029.46 g/mol) but there is a significant

underestimation for 1 by a factor of three (SEC: Mw = 555 g/mol; MALDI-TOF Mw = 1510.6 g/mol)

(Figure 5.1 right).

Page 123: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 5

112

300 400 500 600 7000

1

2 3 5

Nor

mal

ized

Abs

orba

nce

Wavelength (nm)

300 400 500 600

1235

Nor

mal

ized

Abs

orba

nce

(a.u

.)

Time (s)

Figure 5.1. Left: UV-Vis spectra of 2, 3 and 5 in chloroform. Right: SEC traces of monomers 1 and 2,

and polymers 3 and 5 in chloroform.

The IR-spectrum of 3 clearly shows an absorption at 526 cm-1, characteristic of the fullerene

moiety. This peak is absent in the IR-spectrum of the reference 5.

The 1H-NMR spectrum of 3 features the characteristic signals of the polymer backbone and

additional absorptions corresponding to the pendant moieties (Figure 5.2). The polymeric nature of

the sample broadens the signals. The broad multiplet at 3 ppm corresponds to the -CH2- closest to the

fullerene. A comparison of the integrals of this multiplet with that of the peaks at around 4 ppm,

which are characteristic for the –OCH2- and –OCH3 protons, gives an estimate of the relative content

of the monomers in the polymer. According to this quantitative analysis the ratio of the monomers 1

and 2 is 0.7 to 1. The ratio of the monomer content can be explained by polymer chains that are

mainly terminated with monomer 2. Actually, it is also possible to observe a small signal

corresponding to ethynyl terminated chains at 3.5 ppm.

Page 124: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced electron transfer of conjugated polymers with pendant fullerenes

113

C≡CH end group

CHCl3

C≡

Ha

0.51.01.02.02.02.53.03.03.54.04.04.55.05.05.56.06.06.57.07.07.58.08.0

O

O

OO

O

O

O

OO

O

Ha Ha

-O-CH2-and –O-CH3

C≡CH end group

CHCl3

C≡

Ha

0.51.01.02.02.02.53.03.03.54.04.04.55.05.05.56.06.06.57.07.07.58.08.0

O

O

OO

O

O

O

OO

O

Ha Ha

-O-CH2-and –O-CH3

Figure 5.2. 1H-NMR spectrum of polymer 3 in deuterated chloroform solution.

5.2.2 Photophysical properties

UV/Visible absorption spectroscopy. The absorption spectrum of polymer 3 features the

absorption of the π-conjugated backbone at 468 nm plus an additional absorption in the UV region,

corresponding to the pendant fullerene moiety (Figure 5.1 left). The optical bandgap of polymers 3

and 5 is approximately 2.35 eV, as estimated from the onset of the absorption spectra. The presence of

both double and triple bonds in the backbone make 3 and 5 hybrid polymers of poly(p-phenylene

vinylene) (PPV) and poly(p-phenylene ethynylene) (PPE), similar to those recently reported.11 As a

consequence, the bandgap of 3 and 5 is in between the typical values of the bandgaps for alkoxy

substituted PPV and PPE, which are 2.1 and 2.5 eV respectively.

Photoluminescence spectroscopy in solution. Excitation of 3 in dilute toluene solution at

486 nm reveals that the photoluminescence (PL) of 3 is quenched by two orders of magnitude in

comparison with the emission of 5 (Figure 5.3). Also a weak PL signal at 720 nm is observed after

photoexcitation of the polymer backbone, characteristic of the fluorescence emission band of

fulleropyrrolidines (Figure 5.3 right). The same result has been observed in well-defined molecular

donor-acceptor dyads based on fullerene as acceptor after photoexcitation of the donor, and attributed

Page 125: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 5

114

to a singlet-singlet energy transfer from the π-conjugated segment to the fullerene moiety in apolar

medium.5,12,13 In Figure 5.3 the photoluminescence of PCBM is also plotted. PCBM is a molecule that

has a structure close to the pendant methanofullerenes in polymer 3. The emission of the fullerene in

polymer 3 is slightly less intense than the emission of the corresponding reference PCBM. This small

quenching can be attributed to photoinduced charge separation to generate the polymer•+-C60•– charge-

separated state. This charge-separated state is only energetically feasible in such an apolar enviroment

for C60 donor-acceptor systems when donor and acceptor are in a face-to-face orientation.13 This can

happen in this polymer via folding of the pendant C60 onto the polymer chain, enabled by the flexible

nature of the spacer between the polymer backbone and the fullerene moiety. In more polar solvents

like benzonitrile, the additional fullerene emission is completely quenched implying a much more

efficient charge separation.

PL quenching in solution confirms unambiguously the covalent linkage of the C60 moieties to

the polymer backbone in 3.

500 600 700 8000

100

200

300

Inte

nsity

(a.

u.)

Wavelength (nm)700 750 800

0

1

2

Inte

nsity

(a.

u.)

Wavelength (nm)

Figure 5.3. Left: PL spectra of polymers 3 (solid circles) and 5 (open circles) in toluene solution.

Right: Fullerene emission of PCBM in chloroform solution (solid line) and of polymer 3 in toluene

(open squares) and benzonitrile (open circles) solutions.

Photoinduced absorption spectroscopy of the film. The PIA spectrum of a thin film of 3

exhibits a characteristic band at 1.20 eV of the methanofullerene radical anion and the two distinctive

strong bands of cation radicals (polarons) generated on the conjugated polymer at 0.62 and 1.53 eV

together with a bleaching band at 2.45 eV (Figure 5.4). The PIA spectrum of a film of reference

polymer 5 exhibits a single band at 1.55 eV of the triplet excited state. The near coincidence of the

triplet absorption of 5 and the high-energy radical cation absorption of 3 at 1.53 eV, is often

encountered in π-conjugated polymers.14 The low-energy radical cation band at 0.62 eV and the

methanofullerene anion band at 1.20 eV for 3, and their absence in the PIA spectrum of 5, give direct

Page 126: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced electron transfer of conjugated polymers with pendant fullerenes

115

spectral evidence of a photoinduced electron transfer in 3 between the conjugated chain and the

pendant methanofullerene.

0.5 1.0 1.5 2.0 2.5-1.5

-1.0

-0.5

0.0

0.5

1.0

Nor

mal

ized

-∆T

/T

Energy (eV)

Figure 5.4. Photoinduced absorption spectra of thin films of polymer 3 (solid line) and 5 (dashed

line) on quartz recorded at 80 K.

All PIA bands of 3 increase with the pump intensity (I) following a square-root power law (–

∆T~I0.5) (Figure 5.5, left). This suggests non-geminate bimolecular recombination of the photoinduced

charges, which indicates the migration of opposite charges to different sites in the film.

0 25 50 75 1000

25

50

75

100

125

0.62 eV 1.20 eV 1.53 eV

-∆T

(µV

)

Laser power (mW)100 1000

1

Polymer 3 0.62 eV 1.20 eV 1.53 eV

Polymer 5 1.56 eV

Nor

mal

ized

-∆T

Modulation frequency (Hz)

Figure 5.5. Intensity (left) and frequency (right) dependence of the PIA absorption bands at 0.62,

1.20 and 1.53 eV of polymer 3 and at 1.56 eV of polymer 5.

The lifetime of the photoinduced generated species can be estimated from the frequency

depence of the intensity of the photoinduced absorption bands (Figure 5.5, right). For the triplet state

in 5 the lifetime is 5.4 ms. The lifetime of the absorptions at 0.62 and 1.20 eV, associated with the

photoinduced generated charges in 3, also extend into the millisecond time domain. Figure 5.5 (right)

Page 127: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 5

116

reveals that the high energy polaronic absorption band of 3 peaking at 1.53 eV has a frequency

dependence that is a combination of the frequency dependence of the triplet state of 5 and that of the

charged species at 3. This implies that some triplet state is formed together with charge separation in 3

after photoexcitation of the backbone.

5.2.3 Photovoltaic device

Photovoltaic cells were prepared by spin coating 3 from chloroform onto a transparent ITO

front electrode covered with a conducting layer of polyethylenedioxythiophene polystyrenesulfonate

(PEDOT:PSS) and depositing an aluminum back electrode in vacuum. The film thickness was around

30 nm with a surface roughness of less than 5 nm. The dark current and photocurrent of the device

under white-light illumination (100 mW/cm2, AM1.5) reveal promising characteristics (Figure 5.6,

left). A short circuit current of Isc = 0.42 mA/cm2, an open circuit voltage of Voc = 0.83 V, and a fill

factor of 0.29 characterize the depicted cell. The rather low rectification ratio of 36 of the cell at ±2 V

in the dark is related to the low film thickness.

The incident monochromatic photon-to-current efficiency (IPCE) has an onset at 550 nm and

exhibits a maximum of 6% at 480 nm (Figure 5.6, right).

-2 -1 0 1 21E-5

1E-3

0.1

10

I (m

A/c

m2 )

Bias (V)400 500 600 700

0

2

4

6

8

10

IPC

E (

%)

Wavelength (nm)

Figure 5.6. Left: I/V curves of an ITO/PEDOT:PSS/3/Al device in the dark (dashed line) and under

white-light illumination (solid line). Right: The incident photo-to-current efficiency (IPCE, solid

squares) and absorption spectrum (solid line) of an ITO/PEDOT:PSS/3/Al device.

Page 128: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced electron transfer of conjugated polymers with pendant fullerenes

117

5.3. Low-bandgap π-conjugated polymers with pendant fullerenes

With the solar spectrum peaking at 700 nm, a mismatch occurs between the absorption

spectrum of polymer 3 and the solar emission. In order to improve the absorption of photons to create

charges, a backbone with low bandgap properties would be beneficial. Low-bandgap polymers are

typically obtained by alternating electron rich and electron deficient conjugated units along the

polymer chain. Another aspect of polymer 3 that needs to be addressed is its solubility. A less rigid

backbone might aid in solubilizing the donor-acceptor polymer.

BO

O

BO

OS S

NS

N

RR

R R

OC6H13

OC6H13

OC6H13

O*

S S

NS

N* n

RR

R R

+ 1

16. R = OC8H17

17. R = OC8H17

Scheme 5.3. Synthesis of low-bandgap polymers with pendant fullerenes.

These issues are taken in consideration in polymer 17, prepared by copolymerization of

boronic ester 16 with diiodo functionalized fullerene 1 using the Suzuki polycondensation reaction

(Scheme 5.3). 15 The alternation between thiophenes and benzothiodiazoles in monomers 16 confers

the low-bandgap character to the corresponding polymer, together with a less rigid structure.

The optical bandgap of polymer 17 is 1.7 eV, lower than that of polymer 3. The polymer is

also highly soluble in most common solvents and, according to SEC analysis, of rather high molecular

weight (Mw = 25.2 kg/mol, PDI =2.3). IR and UV-Visible absorption spectroscopies exhibit the

typical features of the C60 fragments that are evidence of the incorporation of the fullerene molecules

in the polymer chains. However, several characterization techniques seem to indicate that the polymer

does not correspond with the proposed structure. The SEC-trace of polymers 17 shows a bimodal

distribution, which indicates that likely more than one reaction mechanism occurs in the

polymerization. The bimodality is not observed in a reference polymer without the fullerene moieties

that has been made by reaction of 16 with 4. The resolution of the 1H-NMR spectrum is extremely

low due to peak broadening and the characteristic multiplet at 3 ppm of the –CH2– closest to the

fullerene moiety is not distinguishable.

Page 129: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 5

118

Photoinduced absorption spectroscopy of 17 reveals that charge separation occurs after

photoexcitation of the backbone. The PIA spectrum of 17 shows two bands at 0.48 and 1.08 eV of the

backbone radical cation with a residual triplet signal (Figure 5.7, left). The signal of the C60 radical

anion at 1.24 eV cannot be distinguished because of the radical cation absorptions present in the same

spectral region. As expected, the PIA spectrum of the corresponding reference polymer exhibits a

single photoinduced absorption attributed to the formation of the triplet state.

0.5 1.0 1.5 2.0-10

-5

0

5

10

15

Nor

mal

ized

-∆T

/T

Energy (eV)

-2 -1 0 1 21E-5

1E-3

0.1

10

I (m

A/c

m2 )

Bias (V)

Figure 5.7. Left: Photoinduced absorption spectra of polymer 17 (solid line) and its corresponding

reference polymer (dashed line). Right: I/V curves of an ITO/PEDOT:PSS/17/Al device in the dark

(dashed line) and under white-light illumination (solid line).

Following a procedure similar as for polymer 3, a photovoltaic device of 17 has been prepared

spincoating from a toluene solution. The semilogaritmic plot of the I/V characteristics in the dark and

under white-light illumination of the device are shown in Figure 5.7 (right). The I/V characteristics of

the cell show a diode behavior with a rectification ratio (RR) of 54.7 at ± 2 V. This rather low value is

related to low film thickness (~ 40 nm). Under white-light illumination (100 mW/cm2, AM1.5) an

open-circuit voltage of Voc= 0.75 V and a short circuit current of Isc = 0.31 mA/cm2 are obtained. The

fill-factor (FF) is 0.23. This low fill-factor can be attributed to a small parallel resistance caused by

the low film thickness.

5.4. Conclusions

Two novel π-conjugated polymers with pendant methanofullerenes have been synthesized

using palladium catalyzed cross-coupling reactions.

The first one, a hybrid polymer between p-phenylene vinylenes and p-phenylene ethynylenes

bearing fullerenes, 3, has been synthesized using the Sonogashira coupling between dihalogenated

and bisethynyl terminated aryl compounds 1 and 2. SEC reveals that polymer 3 has a lower molecular

Page 130: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced electron transfer of conjugated polymers with pendant fullerenes

119

weight than its reference polymer 5. Higher molecular weights result in insolubility of the polymer

chains. Photoluminescence and PIA spectroscopies reveal that photoexcitation of 3 results in a

electron transfer from the polymer backbone to the pendant fullerenes. A device with the structure

ITO/PEDOT:PSS/3/Al has been prepared by spincoating from a chloroform solution. The values of Isc

and Voc have improved compared to previously reported π-conjugated polymer/fullerene solar cells.16

Moreover, this performance is given by a polymer with a fullerene loading of 31 wt %, a value

considerably smaller than that used commonly in ‘bulk-heterojunction’ solar cells (about 75%). The

performance of the device of 3 could be improved by increasing the solubility of the polymer, which

would enable the preparation of thicker active layers. Also, there is a mismatch between the solar

spectrum and the absorption spectrum of 3, meaning that the solar spectrum is not used in an optimal

way to generate photoinduced charges.

In an attempt to increase the processability and the absorption of light of polymer 3, a

covalently linked donor-acceptor polymer with low-bangap characteristics, 17, has been synthesized

using the Suzuki polycondensation. Characterization of these polymers reveals that although the

polymers contain fullerene molecules, the polymer does not correspond to a perfect alternation of the

comonomers. This is attributed to a significant amount of side reactions taking place, apart from the

Suzuki cross-coupling reactions. Although this polymer shows some interesting photophysical and

photovoltaic properties, the Suzuki polycondensation is not an ideal alternative to improve this type of

polymers.

Overall, the results indicate that a bicontinuous network of donor and acceptors, confined to a

molecular scale, is an attractive approach to new materials for photovoltaic applications, although

much has to be improved to bring the performance of these novel polymers to that of the best bulk-

heterojunction cells.

5.5. Experimental section

For general methods and materials the reader is referred to chapter 2 of this thesis. 1,4-Bis(2-ethylhexyloxy)-2-methylbenzene (7). Under an argon atmosphere methylhydroquinone (10 g, 80.5 mmol), 2-ethylhexyl-p-toluenesulfonate (48 g, 169 mmol) and tetrabutylammonium chloride (2.66 g, 9.6 mmol) were added to a suspension of K2CO3 (66.3 g, 480 mmol) in dry 2-butanone (160 mL). The reaction mixture was stirred for 16 h at reflux temperature. After cooling, the suspension was filtered and the solvent was removed in vacuo. The resulting crude product was purified by column chromatography (silica gel, hexane/CHCl3 2:1). Evaporation of the solvent yielded 28.3 g (93%) of 7 as a pure colorless oil: 1H NMR (CDCl3) δ 6.79 (d, 1H), 6.77 (d, 1H), 6.71 (dd, 1H), 3.83 (m, 4H), 2.25 (s, 3H), 1.7 (m, 2H), 1.48 (m, 16H), 0.98 (m, 12H); 13C NMR (CDCl3) δ 152.95, 151.50, 127.99, 117.62, 111.66, 111.35, 70.94, 70.79, 39.62, 39.45, 30.68, 30.52, 29.11, 29.07, 24.05, 23.83, 23.08, 16.38, 14.09, 11.21, 11.09; GC-MS (Mw = 348.56) m/z = 348 [M]+. 1-Bromo-2,5-bis(2-ethylhexyloxy)-4-bromomethylbenzene (8). Under an argon atmosphere, NBS (6.12 g, 34.4 mmol) and AIBN (1.72 g, 10.3 mmol) were added to a solution of 7 (10 g, 28 mmol) in dry CCl4 (28 mL). After stirring for 1h under reflux, the reaction mixture was cooled to

Page 131: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 5

120

room temperature. The mixture was filtered and the solvent evaporated. To remove the last traces of AIBN and NBS, hexane was added to the residue, followed by filtration and evaporation of the solvent. Subsequently, dry THF (28 mL) and NBS (6.63 g, 37 mmol) were added and the reaction mixture was stirred at reflux temperature for 1h. After evaporation of the solvent, hexane was added. The solution was filtered and the solvent removed in vacuo. After column chromatography (silica gel, hexane) and evaporation of the solvents, 3.5 g of 8 (24%) was obtained as a colorless oil. 1H NMR (CDCl3) δ 7.06 (s, 1H), 6.80 (s, 1H), 4.48 (s, 2H), 3.84 (m, 4H), 1.76 (m, 2H), 1.4 (m, 16H), 0.92 (m, 12H); 13C NMR (CDCl3) δ 151.20, 149.55, 125.87, 116.98, 115.61, 113.15, 72.35, 71.04, 39.50, 30.57, 30.46, 29.05, 23.97, 23.86, 23.01, 14.05, 11.16. Diethyl[2,5-bis(2-ethylhexyloxy)-4-bromo-benzyl]phosphonate (9). Triethyl phosphite (1.55 g, 9.33 mmol) and 8 (3.15 g, 6.22 mmol) were stirred at 160 oC for 1.5 h. The reaction mixture was cooled to 75 oC and the ethyl bromide, formed during the reaction, and the excess of triethyl phosphite were distilled under reduced pressure. The product 9 was a light yellow oil. Yield 3.30 g (100%). 1H NMR (CDCl3) δ 7.02 (d, 1H), 6.86 (d, 1H), 4.03 (m, 4H), 3.79 (m, 4H), 3.16 (d, 2H), 1.70 (m, 2H), 1.36 (m, 22H), 0.90 (m, 12H); 13C NMR (CDCl3) δ 150.94 (d), 149.34 (d), 120.08 (d), 116.55 (d), 116.25 (d), 110.50 (d), 72.17, 71.27, 61.76 (d), 39.39 (d), 30.38 (d), 28.93 (d), 23.74 (d), 22.88, 16.22 (d), 13.91, 11.01. 1,4-Dibromo-2-(3,7-dimethyloctyloxy)-5-methoxybenzene (12). A solution of Br2 (11.78 g, 73.74 mmol) in glacial acetic acid (30 mL) was added dropwise to a solution of 1-(3,7-methyloctoxy)-4-methoxybenzene (11) (10 g, 37.82 mmol) in glacial acetic acid (45 mL) at 65 oC. After stirring during 45 min at 65 oC, the temperature was raised to reflux during 1h. The solution was then cooled to room temperature and subsequently poured on water (560 mL) and made alkaline with 2N NaOH (660 mL). The aqueous phase was extracted with CH2Cl2 (3 x 200 mL). The combined organic layers were washed with brine and dried over MgSO4. The resultant crude product was purified by column chromatography (silica gel, hexane/CHCl3 2:1) Evaporation of the solvent yielded 12 g (75%) of 12 as a colorless oil: 1H NMR (CDCl3) δ 7.09 (s, 1 2H), 1.52 (m, 1H), 1.32 (m, 2H), 1.68 (m, 4H), 0.94 (d, 3H), 0.87 (d, 6H); 13C NMR (CDCl3) δ 150.34, 150.09, 118.38, 116.91, 111.11, 110.13, 68.57, 56.90, 39.13, 37.15, 35.97, 29.70, 27.92, 24.61, 22.66, 22.56, 19.64; GC-MS (Mw = 422.20) m/z= 422 [M]+. 2-(3,7-Dimethyloctyloxy)-5-methoxybezene-1,4-dialdehyde (13). Dibromide 12 (7 g, 16.58 mmol) was dissolved in dry diethyl ether (135 mL). The solution was cooled to –10 oC and 1.6 M n-buthyllithium hexane solution (24.87 mL) was added slowly. The reaction mixture was stirred for 5 min., then, the cooling bath was removed and dry DMF (3.23 mL) was added dropwise. The mixture was stirred for another hour at room temperature. After addition of 6 M HCl (30 mL), the organic layer was washed with water (2 x 100 ml) a saturated NaHCO3 solution (100 mL) and again water (100 mL). The organic layer was dried over MgSO4 and the solvent was evaporated. The resultant crude product was purified by column chromatography (silica gel, hexane/toluene 2:1, Rf = 0.2). Evaporation of the solvent yielded 3.02 g (56%) of 13 as a light yellow solid. 1H NMR (CDCl3) δ10.54 (s, 1H), 10.49 (s, 1H), 7.44 (s, 2H), 4.13 (m, 2H), 3.94 (s, 3H), 1.88 (m, 1H), 1.66 (m, 2H), 1.53 (m, 1H), 1.25 (m, 6H), 0.96 (d, 3H), 0.87 (d, 6H); 13C NMR (CDCl3) δ189.20, 155.47, 155.28, 129.08, 111.70, 110.58, 67.53, 56.10, 39.11, 37.13, 35.89, 29.83, 27.88, 24.60, 22.61, 22.51, 19.57. (E,E)-1,4-Bis[4-bromo-2,5-bis(2-ethylhexyloxy)styryl]-2-(3,7-dimethyloctyloxy)-5-methoxybezene (14). Phosphonate 9 (1.82 g, 3.23 mmol) was dissolved in dry DMF (10 mL) under an argon atmosphere and 0.43 g (3.9 mmol) of KtBuO were added to the solution. After 15 min, a solution of dialdehyde 13 (0.5 g, 1.56 mmol) in dry DMF (12 mL) was added dropwise and the reaction mixture was stirred for 3 h. The solution was poured on crushed ice and of 6 M HCl (200 mL) was added. The aqueous phase was extracted twice with diethyl ether and the combined organic layers were subsequently washed with 3 M HCl, water, a saturated aqueous solution of Na2CO3 and dried over MgSO4. The solvent was removed in vauo. Column chromatography (silica gel, hexane/CHCl3 4:1, Rf = 0.6) and posterior evaporation of the solvent afforded 0.7g (35%) of 14 as a greenish oil: 1H NMR (CDCl3) δ 7.5 (d, 1H), 7.49 (d, 1H), 7.45(d, 1H), 7.44 (d, 1H), 7.18 (s, 2H), 7.17 (s, 2H), 7.10 (s, 1H), 7.10 (s, 1H), 4.10(t, 2H), 3.92 (m, 11H), 1.9-0.8 (m, 79H); 13C NMR (CDCl3) δ 151.37, 151.08, 150.98, 150.90, 149.92, 149.88, 127.10, 127.02, 126.90, 126.84, 123.41, 123.35, 123.22, 122.63, 117.55, 117.41, 111.59, 111.55, 111.08, 110.72, 110.35, 108.45, 72.34, 72.25, 71.79, 71.50, 67.17, 55.91, 39.61, 39.53, 39.21, 37.29, 36.46, 30.72, 30.73, 30.50, 30.21, 29.10, 29.06, 27.92, 24.70, 24.15, 23.89, 23.87, 23.05, 22.63, 22.54, 19.88, 14.11, 14.08, 14.06, 11.32, 11.26, 11.23, 11.20; MALDI-TOF MS (Mw = 1139.32) m/z = 1138.59 [M]+.

Page 132: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced electron transfer of conjugated polymers with pendant fullerenes

121

(E,E)-1,4-Bis[4-trimethylsilylethynyl-2,5-bis(2-ethylhexyloxy)styryl]-2-(3,7-dimethyloctyloxy)-5-methoxybezene (15). (Trimethylsilyl)acetylene (0.114 g, 1.17 mmol) and dibromide 14 (0.5 g, 0.39 mmol) were dissolved in anhydrous triethylamine (8 mL). Argon was purged trough the solution for 15 min and the temperature was raised to 80 oC. Then PdCl2 (6.98 mg, 0.04 mmol), triphenylphosphine (31 mg, 0.11 mmol) and copper(II) acetate (7.8 mg, 0.04 mmol) were added to the solution. The reaction mixture was stirred for 16 h. The solvent was removed in vacuo, the crude solid was dissolved in diethyl ether. The organic layer was washed with a saturated aqueous solution of NH4Cl and brine. After the organic layer was dried over MgSO4, the solvent was removed in vacuo. The residue was purified by column chromatography (silica gel, hexane/CHCl3 85:15, Rf = 0.4) and recrystalization from ethanol yielded 230 mg (50%) of 15 as yellow crystals: 1H NMR (CDCl3) δ7.48 (s, 2H), 7.47 (s, 2H), 7.15 (s, 2H), 7.11 (s, 1H), 7.10 (s, 1H), 6.94 (s, 2H), 4.07 (t, 2H), 3.90 (m, 11H), 1.89 (m, 1H), 1.77 (m, 5H), 1.51 (m, 16), 1.34 (m, 20H), 1.14 (m, 4H), 0.92 (m, 33H), 0.26 (s, 18H); 13C NMR (CDCl3) δ 154.90, 151.49, 151.01,150.40, 150.29, 128.72, 128.58, 127.26, 127.06, 123.94, 123.84, 123.48, 122.83, 117.07, 116.93, 112.17, 110.47, 110.05, 109,62, 108.47, 101.84, 98,86, 71.81, 71.70, 71.64, 71.35, 67.80, 55.94, 39.73, 39.94, 39.63, 39.24, 37.30, 36.47, 30.79, 30.76, 30.54, 30.24, 29.19, 29.15, 27.93, 24.69, 24.21, 23.92, 23.91, 23.09, 23.07, 22.64, 22.55, 19.90, 14.12, 14.08, 14.06, 11.35, 11.32, 11.28, 0.02; MALDI-TOF MS (Mw =1173.94) m/z = 1173.84 [M]+; Anal. Cald for C75H120O6Si2: C, 76.7; H, 10.3. Found: C, 76.64; H, 10.33. (E,E)-1,4-bis[4-ethynyl-2,5-bis(2-ethylhexyloxy)styryl]-2-(3,7-dimethyloctyloxy)-5-methoxybezene (2). To a solution of 15 (27 mg, 0.02 mmol) in dry THF was added 1 M tetrabuthylammonium fluoride in THF (0.023 mL). The reaction mixture was stirred for 1 min and subsequently filtrated over silica gel using chloroform as eluent. The solvent was removed in vacuo yielding 22.6 mg (100%) of a yellow solid which was used without further purification: 1H NMR (CDCl3) δ7.49 (s, 1H), 7.49 (s, 1H), 7.48 (s, 2H), 7.15 (s, 2H), 7.14 (s, 1H), 7.13 (s, 1H), 6.97 (s, 1H), 6.97 (s, 1H), 4.08 (t, 2H), 3.90 (m, 11H), 3.29 (s, 2H), 0.91 (m, 33H); 13C NMR (CDCl3) δ 154.87, 151.46, 150.99, 150.39, 150.28, 128.97, 128.81, 127.18, 127.01, 124.11, 124.03, 123.33, 122.73, 117.43, 117.27, 111.13, 110.38, 110.28, 109.88, 108.47, 81.41, 80.53, 80.50, 72.15, 72.05, 71.62, 71.34, 67.73, 55.93, 39.63, 38.57, 39.43, 39.22, 37.29, 36.45, 30.93, 30.76, 30.74, 30.52, 30.22, 29.68, 29.10, 29.06, 27.92, 24.69, 24.18, 23.89, 23.87, 23.05, 22.64, 22.54, 19.88, 14.10, 14.06, 11.31, 11.26, 11.23, 11.20. MALDI-TOF MS (Mw =1029.58) m/z = 1029.45 [M]+. Reference polymer (5). To a sealed tube fitted with a magnetic stirrer was added diiodo monomer 417 (14.85 mg, 0.025 mmol), diethylnyl monomer 2 (25 mg, 0.024 mmol), Pd(PPh3)4 (1.16 mg, 0.001 mmol), CuI (0.19 mg, 0.001 mmol), dry Et3N (0.3 mL) and dry ortodichlorobenzene (0.7 mL). The reaction mixture was degassed using freeze-pump-thaw cycles and heated at 75 °C under Ar atmosphere for 24 h. After cooling it to room temperature, the reaction mixture was added dropwise to rapidly stirred EtOH (30 mL). After stirring for 2 h, the precipitate was collected and dried under vacuum overnight. Polymer 5 was obtained as 25 mg (75%) of an orange solid. SEC (chloroform): Mw =35.8 kg/mol, PDI = 2.82. Polymer 3. This polymer was prepared by a procedure identical polymer 5 using diethylnyl monomer 2 (22 mg, 0.02 mmol), diiodo monomer 1 (27mg, 0.018mmol), Pd(PPh3)4 (0.92 mg, 0.0008 mmol), CuI (0.15 mg, 0.0008mmol), dry Et3N (0.3 mL) and dry orthodichlorobenzene (0.7 mL ). Polymer 3 was obtained as 34 mg (76%) of a brown solid. SEC (chloroform): Mw =16.2 kg/mol, PDI = 2.32. Photoinduced absorption. PIA measurements were performed between 0.25 eV and 3.0 eV by exciting thin films on quartz with a mechanically modulated (275 Hz, 25 mW 2 mm diameter, 488 nm) beam from a continuous wave argon ion laser (Spectra physics 2025). The change in transmission of probe light (∆T) was monitored with a phase-sensitive lock-in amplifier using Si, InGaAs, and cooled InSb detectors after dispersion by a triple grating monochromator. The photoinduced absorption, -∆T/T ≅ ∆αd, is directly calculated from the change in transmission after correction for fluorescence, which is recorded in a separate experiment. The lifetime of the photoexcitations has been determined by recording the intensity of the PIA bands as a function of the modulation frequency (ω) in the range of 30-4000 Hz. Photovoltaic cells. For photovoltaic cells, polyethylenedioxythiophene polystyrenesulfonate (PEDOT:PSS, Bayer AG) (90 nm) was spin coated on UV-ozone cleaned glass substrates covered with indium tin oxide (ITO) (140 nm), followed by spin coating a solution of 3 in chloroform to form the active layer of 30 nm as determined with a Tencor P-10 surface profiler. Finally, an aluminum back electrode (100 nm) was deposited in

Page 133: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 5

122

vacuum to give an active area of 4 mm2. I/V characteristics were measured under 100 mW/cm2 AM1.5 white-light illumination from a Steuernagel Solarconstant 1200 solar simulator with a Keithley 2400 Source Meter in inert nitrogen atmosphere at room temperature. Surface roughness was determined by AFM. IPCE measurements. Spectrally resolved photocurrent measurements were performed by illuminating the device with 1 mW/cm2 monochromatized light with a FWHM of ~6 nm from a Xe arc lamp. No noticeable degradation of the devices was observed during the measurement cycles. Light intensities were measured by a calibrated Si photodiode. The incident photon to current efficiency (IPCE) or external quantum efficiency was calculated from the photocurrent and light intensity as:

IPCE [%] =[ A / cm

[nm] [W / m ]

2

2

1240 ⋅

I

I

sc µλ

]

5.6 References 1 (a) Tang, C. W. Appl. Phys. Lett. 1986, 48, 183. (b) Granström, M.; Petrisch, K.; Arias, A. C.; Lux, A.;

Lux, M.; Andersson. M. R.; Friend, R. H. Nature 1998, 395, 257.

2 (a) Halls J. J. M.; Walsh, C. A.; Greenham, N. C.; Marseglia, E. A.; Friend, R. H.; Moratti, S. C.;

Holmes, A. B. Nature 1995, 376, 498. (b) Yu, G.; Gao, Y.; Hummelen, J. C.; Wudl, F.; Heeger, A. J.

Science 1995, 270, 1789.

3 Shaheen, S. E.; Brabec, J. C.; Padinger, F.; Fromherz, T.; Hummelen, J. C.; Sariciftci, N. S. Appl. Phys.

Lett. 2001, 78, 841.

4 (a) Nierengarten, J.-F.; Eckert, J.-F.; Nicoud, J.-F.; Ouali, L.; Krasnikov, V. V.; Hadziioannou, G. Chem.

Commun. 1999, 617. (b) Stalmach, U.; de Boer, B.; Videlot, C.; van Hutten, P. F.; Hadziioannou, G. J.

Am. Chem. Soc. 2000, 122, 5464. (c) Eckert, J.-F.; Nicoud, J.-F.; Nierengarten, J.-F.; Liu, S.-G.;

Echegoyen, L.; Barigelletti, F.; Armaroli, N.; Ouali, L.; Krasnikov, V.; Hadziioannou, G. J. Am. Chem.

Soc. 2000, 122, 7467.

5 Peeters, E.; van Hal, P. A.; Knol, J.; Brabec, C. J.; Sariciftci, N. S.; Hummelen, J. C.; Janssen. R. A. J. J.

Phys. Chem. B 2000, 104, 10174.

6 (a) Shi, S.; Khemani, K. C.; Li, C.; Wudl, F. J. Am. Chem. Soc. 1992, 114, 10656. (b) Sun, Y.-P.; Liu, B.;

Moton, D. K. Chem. Commun. 1996, 2699. (c) Zhang, N.; Schricker, S. R.; Wudl, F.; Prato, M.; Maggini,

M.; Scorrano, G. Chem. Mater. 1995, 7, 441. (d) Gügel, A.; Belik, P.; Walter, M.; Kraus, A.; Harth, E.;

Wagner, M.; Spickermann, J.; Müllen, K. Tetrahedron 1996, 52, 5007. (e) Kraus, A.; Müllen, K.

Macromolecules, 1999, 32, 4241. (f) Ilhan, F.; Rotello, V. M. J. Org. Chem. 1999, 64, 1455. (g) Xiao, L.;

Shimotani, H.; Ozawa, M.; Li, J.; Dragoe, N.; Saigo, K.; Kitazawa, K. J. Polym. Sci. A 1999, 37, 3632.

(h) Sun, Y.-P.; Lawson, G. E.; Huang, W.; Wright, A. D.; Moton, D. K. Macromolecules, 1999, 32, 8747.

(i) Okamura, H.; Miyazono, K.; Minoda, M.; Komatsu, K.; Fukuda, T.; Miyamoto, T. J. Polym. Sci. A

2000, 38, 3578.

7 (a) Benincori, T.; Brenna, E.; Sannicoló,F.; Trimarco, L.; Zotti, G.; Sozzani, P. Angew. Chem., Int. Ed.

Eng. 1996, 35, 648. (b) Ferraris, J. P.; Yassar, A.; Loveday, D. C.; Hmyene, M. Opt. Mater. 1998, 9, 34.

Page 134: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Photoinduced electron transfer of conjugated polymers with pendant fullerenes

123

(c) Cravino, A.; Zerza, G.; Maggini, M.; Bucella, S.; Svensson, M.; Andersson, M. R.; Neugebauer, H.;

Sariciftci, N. S. Chem. Commun. 2000, 2487.

8 Zhang, F., Svensson, M; Andersson, M. R.; Maggini, M.; Bucella, S.; Menna, E.; Inganäs, O. Adv. Mater.

2001, 13, 171.

9 (a) Xiao, S; Wang, S.; Fang, H.; Li, Y.; Shi, Z.; Du, C.; Zhu, D. Macromol. Rapid. Commun. 2001, 22,

1313. (b) Wang, S.; Xiao, S.; Li, Y.; Shi, Z.; Du, C.; Fang, H.; Zhu, D. Polymer, 2002, 43, 2049.

10 Marcos Ramos, A.; Rispens, M. T.; van Duren, J. K. J.; Hummelen, J. C.; Janssen, R. A. J. J. Am. Chem.

Soc., 2001, 123, 6714.

11 Brizius, G.; Pschirer, N. G.; Steffen, W.; Stitzer, K.; Zur Loye, H.-C.; Bunz, U. H. F. J. Am. Chem. Soc.,

2000, 122, 12435.

12 Van Hal, P. A.; Knol, J.; Langeveld-Voss, B. M. W.; Meskers, S. C. J.; Hummelen, J. C.; Janssen, R. A.

J. J. Phys. Chem. A 2000, 104, 5964.

13 van Hal, P. A.; Beckers, E. H. A.; Meskers, S. C. J.; Janssen, R. A. J.; Jousselme, B.; Blanchard, P.;

Roncali, J. Chem. Eur. J. 2002, 8, 5415.

14 Lane, P. A.; Wei, X.; Vardeny, Z. V. Phys. Rev. B 1997, 56, 4626.

15 For a full account on the synthesis of polymer 17 see: Hamelinck, P. Novel Double Cable Materials for

Polymer Solar Cells, 2003, research report Eindhoven University of Technology.

16 (a) Roman, L. S.; Andersson, M. R.; Yohannes, T.; Inganäs, O. Adv. Mater. 1997, 9, 1164. (b) Mattoussi,

H.; Rubner, M. F.; Zhou, F.; Kumar, J.; Tripathy, S. K.; Chiang, L. Y. Appl. Phys. Lett. 2000, 77, 1540.

17 Weder, C.; Wrighton, M. S. Macromolecules 1996, 29, 5157.

Page 135: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1
Page 136: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 6

Polyacetylenes with pendant

donor-acceptor dyads

Abstract

The synthesis of polyacetylenes functionalized with donors and acceptors in a variety of arrangements

has been investigated. The acceptor is a perylene diimide chromophore (PERY) and the donors are an

oligo(p-phenylene vinylene) (OPV), an oligo(p-phenylene ethynylene) (OPE) and a hybrid of OPV

and OPE, an OPVE. The polymers were prepared either by the copolymerization of OPV and PERY

functionalized acetylenes or by the polymerization of homologous OPVE-PERY and OPE-PERY

dyads. Absorption spectroscopy reveals that preorganization of the donor and acceptor might take

place in solution. Photoexcitation of the donor or acceptor results in charge separation in all

polymers. Even though the copolymerization approach represents the fastest and most facile method

to obtain donor-acceptor functionalized polyacetylenes, the alternating arrangement of donors and

acceptors results in very short-lived charged-separated states. In the polymers resulting from the

polymerization of the dyads, the lifetime of the photoinduced charge separated state is increased by

more than one order of magnitude with respect to the copolymer. The improved photophysical

properties are the result of a more refined architecture.

Page 137: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 6

126

6.1 Introduction

Promising power conversion efficiencies have been reported for spontaneously formed

interpenetrating networks of electron- and hole-accepting organic semiconductors.1 However, the poor

miscibility of donor and acceptor and the disordered nature of the resulting networks are thought to

limit the performance of these devices. In chapter 5, the advantage of having donors and acceptors

within one polymer chain has already been addressed. A large interfacial area is beneficial for

efficient charge generation and with an appropriate organization of the chromophores transport of the

charges to opposite electrodes can be expected. Here, in pursue of additional ordering and

consequently enhanced charge mobility, a new design for a donor-acceptor polymer is proposed. This

consists of well-defined donor-acceptor dyads that are covalently linked to a polyacetylene backbone

(Figure 6.1).

Donor Acceptor

Poly acetylene backbone

Donor Acceptor

Poly acetylene backbone

Donor Acceptor

Poly acetylene backbone

Donor Acceptor

Poly acetylene backbone

Figure 6.1. Schematic representation of a polyacetylene with pendant donor-acceptor dyads.

An important aspect of polyacetylenes to take into consideration is the stereoregularity of the

polymer backbone. Four conformers are possible, i.e., cis-cisoidal, cis-transoidal, trans-cisoidal, and

trans-transoidal. When the backbone mainly consists of the trans-transoidal conformer, it will be

stretched into a linear arrangement. If the backbone mainly consists of the cis-transoidal conformer, it

generally exhibits a helical arrangement. The handedness of this helix can be controlled via chiral

centers on the side-chains or via external stimuli.2 Furthermore, the stiffness of the polyacetylene

backbone and the tightness of the helical conformation highly depend on the side-chains. Cis-

transoidal polyacetylenes can be prepared in a selective manner by using [Rh(nbd)Cl]2

(nbd=norbornadiene) as a catalyst. This rhodium catalyst has been reported to be effective in the

polymerization of mono-substituted acetylenes, especially phenyl-substituted acetylenes.

There are a few examples of polyacetylenes with pendant redox-active components, yet none

of them contains both donor and acceptor elements. Schenning et al. have previously shown a

polyacetylene functionalized with oligo(p-phenylene vinylene) (OPV)3 and Vohlidal et al. have

prepared analogous polymers with pendant oligo(p-phenylene ethynylene) (OPEs).4 Yashima et al.

have incorporated acceptor fullerene moieties in a polyphenyl acetylene polymer.5

Page 138: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Polyacetylenes with pendant donor-acceptor dyads

127

Here, a perylene diimide chromophore (PERY) is used as an acceptor. This type of molecule

is potentially a better candidate for the acceptor role in plastic solar cells than fullerene molecules,

because of their lower reduction potential (~ -0.55 vs. ~ -0.7 V) and their strongly enhanced

absorption in the visible region. Furthermore, perylene diimide dyes readily stack via π-π interactions

and have the ability to form crystals that exhibit remarkable exciton diffusion ranges and charge

mobility.6 In the proposed design (Figure 6.1) the strong π-π interactions characteristic of perylene

chromophores are used in order to gain dimensional control over the phase separation. By having the

perylene at the periphery and the polyacetylene backbone connecting the dyads, donor-donor and

acceptor-acceptor interactions might be favored.

In this chapter the synthesis of different polyacetylenes functionalized with donors and

acceptors (polymers 1, 2 and 3, Figure 6.2) is described. Whereas the acceptor moiety is constant for

all polymers, the donor ranges from an oligo(p-phenylene vynylene) (OPV) in 1, to a oligo(p-

phenylene ethynylene) (OPE) in 3. In 2 the donor is a mixed OPV and OPE (OPVE). The

photophysical properties of the polymers are investigated in chloroform solution and in the solid state

by means of photoluminescence and photoinduced absorption spectroscopy.

NNO

O

O

O

O

O

H

n

O

OOO

O O ONN

O

O

O

O

OH

n

O

O

O

O

O

O

NNO

O

O

O

O

H

H

n

m

1

2

3

n

Figure 6.2. Chemical structure of polyacetylenes 1, 2 and 3.

Page 139: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 6

128

6.2 Design

Several design criteria have been formulated in order to screen for the optimal synthesis of the

polymers and to gain knowledge on how to obtain the most favorable photophysical properties.

A first consideration concerns the catalytic system to be employed for the polymerization.

The rhodium chlorine bridge catalyst [Rh(nbd)Cl]2 has extensively proven to be effective for the

polymerization of mono-substituted phenylacetylenes. The catalyst accounts for high molecular

weights and stereoregular control, yielding cis-transoidal polyphenylacetylenes.7 This rhodium

catalyst polymerizes selectively terminal ethynyl groups, allowing for the use of internal triple bonds

in the monomers as conjugating elements.4 The solvent system consisting of toluene and Et3N as a

base ensures sufficient solubility of the conjugated monomers used in the reaction.

A disadvantage of the rhodium catalyst is its tendency to generate cyclic trimers, consisting of

a benzene core made up out of three acetylenes. This side reaction may overrule the polymerization,

especially in cases where the terminal phenyl ring bearing the acetylene also carries alkoxy

substituents. Thus, in order to favor polymerization over the formation of cyclic trimers, the absence

of substituents on the terminal phenyl ring is required for all monomers.

An intrinsic feature of conjugated polymers is their poor solubility in conventional organic

solvents, due to their rigid backbone and tendency to aggregate. The designed polymers feature an

additional perylene function that is known to disfavor solubility. These features, lowering the

solubility, need to be compensated by flexible, preferably branched, side chains. The alkoxy side

chains on the donor segments and alkyl chains on the PERY have proven to greatly enhance

solubility. An extra requirement is the need for branching at the first carbon of the alkyl chains on the

PERY.8

An element that strongly influences the lifetime of the donor(+)-acceptor(-) charged separated

state is the connection of the donor and acceptor. Previous studies have shown that a direct connection

of these two chromophores will result in a very fast recombination of the created charges, whereas a

connection via a saturated spacer enhances the lifetime.9 As a long-lived charge separated state is

beneficial for charge transport through the system, the use of a flexible saturated spacer has been

included in the design of the polymers.

The synthesis of the monomers for polymers 2 and 3 requires the coupling of the perylene

functionality to only one side of the symmetrical donor functionality. Two approaches can be

envisaged to achieve these asymmetric monomers. The first approach, though not being very elegant,

is a fast one and consists of the unselective reaction of the phenyl acetylene modified perylene with a

symmetrical donor diiodide. This procedure will yield a mixture of non-reacted, mono-functionalized

and difunctionalized donor, thus limiting the best possible yield to 50 %. The second approach

consists of the total synthesis of a selectively addressable donor compound, featuring one halogenated

function and one protected acetylene function. The advantage of this approach is the intrinsic high

Page 140: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Polyacetylenes with pendant donor-acceptor dyads

129

yield of the coupling reaction at a crucial stage in the synthesis. The major drawback of this approach

is the extended synthetic pathway to the bifunctional donor, compared to that for the symmetrical

donor functionality.

The detailed synthesis of all monomers and polymers are described in the following section.

6.3 Synthesis and characterization

6.3.1 Ethynyl perylene bisimide (10)

Perylene diimid 10 (scheme 6.1) is a common element in all three polymers. It is one of the

comonomers of polymer 1 and a building block of the dyads in polymers 2 and 3. As 10 features a

terminal acetylene, it was also subjected to polymerization.

The synthesis of 10 starts with the protection of the amine function in (S)-(+)-leucinol.9b The

BOC-protected (S)-(+)-leucinol 4 was reacted with 4-iodophenol using Mitsonobu conditions,

affording, after deprotection of the amine with TFA, compound 6. Iodophenyl functionalized perylene

diimide 8 was obtained after reaction of amine 6 with the perylene monoanhydride monoimide 7.

Subsequent coupling of trimethylsilylacetylene (TMA), using a palladium-catalyzed coupling, and

desilylation with TBAF yielded monomer 10.

N O

O

OO

O

NH

O

OOH

NH

OO

OI

NH2

O I

N NO I

O

O

O

O

N NO

O

O

O

OR

NNO

O

O

O

O

*

*n

c

d

4 65

a b

7

89. R = TMS10. R = H

11

e

f

Scheme 6.1. a. 4-Iodophenol, PPh3, DEAD, toluene, r.t., 16 h, 40%; b. TFA, CH2Cl2, r.t., 16 h, 86%;

c. Imidazole, Zn(OAc)2,1 h, 160 °C, 67%; d. TMA, Pd(PPh3)2Cl2, CuI, Et3N, toluene, 50°C, 4 h, 84%;

e. TBAF, THF, 1 min, 81%; f. [Rh(nbd)Cl]2, Et3N, toluene, r.t., 24 h.

Page 141: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 6

130

Polymerization of monomer 10 results in a red solid, polymer 11, which is scarcely soluble in

any solvent. The SEC trace of the soluble fraction of 11 in chloroform reveals the presence of some

starting monomer, the formation of cyclotrimer, soluble polymer and mainly aggregated polymer. The

low solubility of this polymer is caused by the high content of perylene. Soluble acetylenes with

pendant perylene chromophores may be achievable by the copolymerization of 10 with a non-

perylene based comonomer.5

6.3.2 Polyacetylene with pendant OPV and PERY chromophores (polymer 1).

As a first facile synthesis of a polyacetylene incorporating pendant donor and acceptors,

perylene 10 and OPV 12,10 both featuring a terminal acetylene, were copolymerized to afford

copolymer 1 (Scheme 6.2) as a very soluble red solid. After work up, the SEC analysis of 1 revealed

the presence of cyclic trimers, which were removed from the polymer by preparative size exclusion

chromatography.

H

N NO

O

O

O

OH

R*O

OR*

R*OH

OR*

O

R*O

OR*

R*O

OR*

N NO

O

O

O

O

H

10

12

+

OR* = 1

n

ma

Scheme 6.2. Synthesis of copolymer 1: a.[Rh(nbd)Cl]2, Et3N, toluene, r.t. 16 h, 36%.

The size exclusion chromatography (SEC) trace of the polymer in chloroform solution does

not correspond to a perfect Gaussian distribution, which suggests aggregation of the polymer in

chloroform solution. Addition of 1% methanol does not alter the shape of the SEC trace (Figure 6.3).

The molecular weight of the polymer is Mw= 14.6 kg/mol with PDI = 1.6. The molecular weight has

been determined using polystyrene standards, which gives a rough indication, but is not quantitative

due to the different shape of the polymer with regard to polystyrene. In contrast to polystyrene, which

is in a random coil conformation under these conditions, the polyacetylene is most probably in a rigid

rod conformation. The IR spectrum of polymer 1 does not feature the characteristic terminal acetylene

absorption, present in the starting monomers at around 3300 cm-1. This indicates that the

polymerization reaction has proceeded at the terminal ethynyl position and that no starting monomer

is left in the polymers. The 1H-NMR spectrum of polymer 1 shows very broad resonances. Generally,

Page 142: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Polyacetylenes with pendant donor-acceptor dyads

131

polyphenylacetylenes prepared using [Rh(nbd)Cl]2 exhibit sharp line widths, because of the high

stereoregularity of these polymers.7 However, polyacetylene 1 bears donor and acceptor

chromophores that are highly prone to aggregate. The mobility of the polymer backbone is most

probably restricted by this aggregation and peak broadening of the proton ressonances occurs.11

Consequently, the characteristic vinylic proton corresponding to the acetylene backbone could not be

distinguished and therefore, the regularity of the polymer backbone could not be determined. The

good solubility of copolymer 1 with respect to polymer 11 shows however that random incorporation

of non-perylene oligomers, disrupts the perylene aggregation by which the processability is greatly

enhanced.

300 400 500 600 700

0

1

Nor

mal

ized

res

pons

e

Time (sec)

Figure 6.3. SEC traces of polymers 1 (solid line), 2 (open squares) and 3 (open triangles) measured

in chloroform with 1% methanol. Inset: SEC traces of polymer 2 recorded in chloroform (solid

squares) and in chloroform with 1% methanol (open squares).

6.3.3 Polyacetylene with pendant OPVE-PERY dyads (polymer 2)

Polymer 2 has been synthesized by polymerization of the terminal acetylene OPVE-PERY

dyad 20 (Scheme 6.3). This dyad has been assembled by the non-selective reaction of diiodinated

OPV 17 with ethynyl perylene 10.

The synthesis of the OPV chromophore started by the radical bromination of 4-iodotoluene

giving 14. Treatment of 14 with triethylphosphite afforded phosphonate 15. Diiodinated OPV 17 was

obtained after a double Wittig-Horner reaction of phosphonate 15 with dialdehyde 16.19 The coupling

between the OPV and the PERY was done using the palladium catalyzed Sonogashira cross-coupling

of ethynyl compound 10 with diiodinated OPV 17 affording 19. An excess of 17 was used in order to

minimize the double coupling reaction, and the dyad was obtained as the main product of a statistical

Page 143: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 6

132

mixture. Due to the different nature of the two alkoxy side chains on the central ring, 17 is not

symmetrical and 19 therefore consists of a mixture of two regioisomers (only one of which has been

represented in Scheme 6.3). Coupling of 19 with TMS-acetylene and subsequent desilylation gave

functional monomer 20.

Monomer 20 was polymerized for 16 hours. Already after 20 minutes of polymerization a red

solid starts precipitating out of solution. The polymer was worked up by precipitation from the

reaction mixture with methanol. Removal of the cyclic trimers was achieved by reprecipitation from

THF. The polymer is slightly soluble in chloroform and o-dichlorobenzene, but the solubility strongly

increases with a small percentage of ethanol in the chloroform.

IR

OO O

O O

O

I

I

NNO

O

O

O

O

O

O

R

tmsH

NNO

O

O

O

O

O

O

*

*

n

16 17

+

2

18. R = I19. R = 20. R =

13. R = H14. R = Br15. R = P(OEt)2O

a

b

c

d

e

f

g

Scheme 6.3. a. NBS, benzoyl peroxide, CCl4, reflux, 3.5 h, 25%; b. Triethylphosphite, 160 °C, 1.5 h,

100 %; c. t-BuOK, DMF, r.t., 3 h, 64 %; d. 10, Pd(PPh3)2Cl2, CuI, Et3N, toluene, 50 °C, 16 h, 47%; e.

TMA, Pd(PPh3)2Cl2, CuI, Et3N, toluene, 50 °C, 16 h, 80%; f. TBAF/Silica, THF, r.t., 30 min, 97%; g.

[Rh(nbd)Cl]2, Et3N, toluene, r.t. 16h, 60%.

In contrast to polymer 1, the SEC trace of 2 is enormously influenced by the addition of a

small percentage of methanol (Figure 6.3, and inset). The SEC trace in chloroform points to very

Page 144: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Polyacetylenes with pendant donor-acceptor dyads

133

strong aggregation, compared to polymer 1, but this is largely overcome by the addition of methanol.

But even methanol does not allow for complete disruption of the aggregation, which makes it difficult

to estimate an accurate molecular weight against polystyrene standards (Mw= 45.8 kg/mol with PDI

=1.9, using 1% methanol in chloroform as eluent). As for polymer 1, the 1H-NMR spectrum of

polymer 2 shows extremely broad resonances.

6.3.4 Polyacetylene with pendant OPE-PERY dyads (polymer 3)

For polymer 3 an acetylene terminated OPE-PERY dyad was synthesized. In this case, the

OPE donor segment has first been end-capped with two different functionalities (a TMS-ethynylene

group and a bromine) to facilitate a selective monocoupling to the phenylethynyl terminated perylene

acceptor 10.

For the synthesis of the bromine-ethynyl bifunctional donor OPE (Scheme 6.4) a repetitive

strategy developed by Ziener et al. was followed.12 The synthesis started by bisbromination and

subsequent bisalkylation of hydroquinone that gave dibromide 23. The exchange of one bromine for

one iodine to yield 24 was achieved by reaction of 23 with n-BuLi followed by treatment with 1,2-

diiodoethane, while working at temperatures below –80 °C. Selective coupling of TMS-acetylene

with 24 at the iodo-phenyl position yielded 25. In contrast to literature,12 iodo-selectivity was only

achieved by keeping the temperature of the reaction at 0 °C and by using exact amounts of TMS-

acetylene. One part of 25 was treated with TBAF to give acetylene 26. The other part was converted

into the iodo compound 27. The yield of the halogen exchange was only 77%, and several attempts of

separating 27 from 25 using column chromatography were unsuccessful. Nevertheless, the synthetic

procedure was continued and removal of 25 was achieved in the following synthetic step. An iodo-

bromo-selective cross-coupling of 26 with 27 gave dimer 28. Tetramer 31 was derived from 28 after

another cycle of desilylation, iodine-bromine exchange and a palladium catalyzed cross-coupling.

Although tetramer 31 was obtained in reasonable overall yield, the synthesis of the donor fragment is

hampered by difficult and tedious purification of some of the intermediates.

Coupling of the donor tetramer 31 with ethynyl terminated perylene 10, followed by treatment

with TBAF gave acetylene 33. An attempt to polymerize 33, resulted selectively in cyclic trimers due

to the alkoxy substituents on the first phenyl ring. Therefore 33 was extended by coupling it to 4-

iodotrimethylsilylethynylbenzene to afford monomer 35 after deprotection of the ethynyl function.

Polymer 3 was prepared using the usual polymerization conditions for 16 hours. The reaction

mixture was precipitated into methanol and THF to yield a red solid, which is highly soluble in

solvents of intermediate polarity such as chloroform, dichloromethane and o-dichlorobenzene.

Page 145: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 6

134

OH

OH

OH

OH

BrBr

OR

RO

BrBr

OR

RO

IBr

OR

RO

Br tms

OR

RO

Br H

OR

RO

tmsI

OR

RO

Br tms

OR

RO

Br H

OR

RO

I tms

OR

RO

Br tms

ONN

O

O O

O

OR

RO

R'

ONN

O

O O

O

OR

RO

R'

OR = O

ONN

O

O O

O

OR

RO

H

*

*n

2

2

2

4

a b c

de

f

g

h

i

j

4

4

l

m

32. R' = tms33. R' = H

n

o

34. R' = tms35. R' = H

21 22 23 24

25

26

27

28

29

30

31

4

3

k

Scheme 6.4. a. HBr, Br2, 100 °C, 24 h, 49%; b. ROTs, KOH, EtOH, reflux, 24 h, 57%; `c. n-BuLi,

1,2-diodoethane, THF, -60 °C, 72%; d. TMA, Pd(PPh3)2Cl2, CuI, Et3N, 0 °C, 1 h, 82%; e. TBAF,

THF, r.t.,1 min, 100%; f. n-BuLi, 1,2-diodoethane, THF, -60 °C, 77%; g. Pd(PPh3)2Cl2, CuI, Et3N, 0

°C, 3.5 h, 67%; h. TBAF, THF, r.t., 1 min, 100%; i. n-BuLi, 1,2-diodoethane, THF, -60 °C, 60%; j.

Pd(PPh3)2Cl2, CuI, Et3N, 0 °C, 3.5 h, 56%; k. 10, Pd(PPh3)2Cl2, CuI, Et3N, toluene, 50 °C, 16 h, 34%

; l. TBAF/Silica, THF, r.t., 30 min, 100% ; m. 4-Iodotrimethylsylilphenylacetylene Pd(PPh3)2Cl2, CuI,

Et3N, toluene, 50 °C, 16 h, 40% ; n. BAF/Silica, THF, r.t., 30 min, 99% ; o. [Rh(nbd)Cl]2, Et3N,

toluene, r.t. 16h, 30%.

Page 146: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Polyacetylenes with pendant donor-acceptor dyads

135

The molecular weight of the polymer according to SEC is Mw = 53.5 kg/mol with PDI= 1.5

(Figure 6.3). The high molecular weight, especially with respect to polymer 1, is partly due to the high

molecular weight of the monomers, but most probably to the good solubility of polymer 3.

Additionally, the different hydrodynamic volume caused by the large pendant dyads might contribute

to the high molecular weight. In the 1H-NMR spectrum of polymer 3, peak broadening is somewhat

less prominent than for polymers 1 and 2. The OPE segment is heavily substituted with alkoxy chains,

which may hamper aggregation and allow for a better resolution in the NMR spectrum.

6.4 UV/Visible absorption and circular dicroism spectroscopies

The absorption spectra of polymers 1, 2 and 3 in chloroform solution exhibit the electronic

transitions associated with their donor and acceptor chromophores (Figure 6.4). All spectra feature the

well-resolved vibronic structure absorption characteristic of the PERY chromophore with maxima at

490 and 530 nm, approximately. The donor moieties dominate the rest of the absorption spectrum.

Monosubstituted polyacetylenes mainly absorb in the UV region and have very low extinction

coefficients. 13 Therefore the contribution of the backbone to the absorption spectra can be considered

negligible.

In Figure 6.4 the absorption spectra of all monomers in chloroform have been plotted together

with those of the respective polymers. The polymers exhibit a red-shifted onset with respect to their

corresponding monomers. Also the normalized spectra of the monomers and the related polymers do

not superimpose. These differences in the spectra suggest that at least the perylene chromophore is

aggregating when incorporated in the polymeric chain. In the solid state the onset is even more red-

shifted and the absorption bands even broader. By comparing the absorption spectra of the polymers

in chloroform solution with those of the monomeric species and with that in the solid state it can be

concluded that already in chloroform solution the pendant chromophores and dyads are interacting

with each other and are preorganizing the polymer into its solid state conformation. Of all, 2 is the

polymer that experiences the highest aggregation in solution as the absorption spectra in solution

highly resembles the one in solid state.

None of the polymers showed a Cotton effect in chloroform solution, even though they are

decorated with stereocenters. The absence of optical activity might stem either from aggregation

phenomena that disrupt the helical arrangement or from the presence of racemic side-chains or

mixtures of side-chains with different chiral centers, resulting in non-preferential twist senses of the

helices.

Page 147: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 6

136

300 400 500 6000

1

Abs

orba

nce

(O.D

.)

Wavelength (nm)

0

1

0

1

b

c

a

Figure 6.4. Absorption spectra of the different polyacetylenes and monomers.(a) Polymer 1 in

solution (solid line), and solid state (doted line) and monomer 10 (dashed-dotted line) and 12 (dashed

line). (b) Polymer 2 in solution (solid line) and thin film (dotted line) and monomer 20 (dashed line).

(c) Polymer 3 in solution (solid line) and in solid state (dotted line) and monomer 35 (dashed line).

6.5 Photophysical properties

6.5.1 Solid state

Near steady-state photoinduced absorption (PIA) spectroscopy of thin films of polymers 1, 2

and 3 recorded at 80 K revealed the occurrence of charge separation upon illumination (Figure 6.5 left

and right). All PIA spectra exhibit the absorptions characteristic of the perylene bisimide radical anion

at 1.28, 1.54 and 1.73 eV.14 For polymer 3 two additional absorptions centered at 0.5 and at 2.0 eV are

observed. A similar PIA spectrum is obtained from a blend of a p-phenylene ethynylene polymer

(PPE)15 with a soluble perylene dye PERY (Figure 6.5 right), meaning that the additional absorption

bands observed in the PIA spectrum of polymer 3 correspond to the OPE•+ radical cations. The

Page 148: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Polyacetylenes with pendant donor-acceptor dyads

137

absorptions of the positively charged donor moieties in polymers 1 and 2 are expected at similar

energies as observed for the OPE segments.16 However, the PIA spectrum of polymers 1 and 2 display

very weak ill-defined absorptions at low energies. Similar low absorption of the OPV•+ radical cation

have been previously observed in the PIA spectrum of a molecular triad based on comparable

chromophores as in polymers 1 and 2, OPV3-PERY-OPV3 (Figure 6.5 left).17

0.5 1.0 1.5 2.0 2.5

-2

0

2

4

6

∆T/T

x 1

04

Energy (eV)

0.5 1.0 1.5 2.0 2.5

0

2

4

∆T/T

x 1

04

Energy (eV)

Figure 6.5. PIA spectra of thin films of the polyacetylene polymers. Left: polymers 1(solid line), 2

(dashed line) and OPV3-PERY-OPV3 (solid squares). Right: polymer 3 (solid line) and PPE/PERY

solid state blend (solid squares).

NN

O

O

O

O

* *

O

O

n

OO

OO

OO

OONN

O

O

O

O

OPV3-PERY-OPV3

PERY

PPE

Figure 6.6. Chemical structure of reference compounds OPV3-PERY-OPV3, PPE and PERY.

The rates of formation and decay of the charge-separated states in thin films of 2 and 3 have

been inferred from sub-picosecond pump-probe spectroscopy (Figure 6.7 and Table 6.1). The time

profile of the 1450 nm absorption, corresponding to the donor radical cation, has been measured after

exciting the polymers at 450 nm. Fitting of the temporal transient absorption to exponential functions

reveals that the charges are formed within 1 ps and that their lifetimes range from 300 to 400 ps. For

polymer 1 only extremely weak signals could be measured.

Page 149: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 6

138

0 200 400 600 800 1000-70

-60

-50

-40

-30

-20

-10

0

Time (ps)

∆T (

a. u

.)

Time (ps)

-3 0 3-60

-30

0

Figure 6.7. Differential transmission dynamics of the 1450 nm absorption of polymers 2 (solid

squares) and 3 (open squares) in a thin film, after excitation at 450 nm. The inset shows the 1450 nm

signals on short timescales.

Table 6.1. Lifetimes and rates for the charge separation (CS) and charge recombination (CR) of

polymers 1, 2 and 3 in the solid state and in chloroform solution.

Solid state CHCl3 solution

CS CR CS CR

τ (ps) k (s-1) τ (ps) k (s-1) τ (ps) k (s-1) τ (ps) k (s-1)

Polymer 1 n.d. n.d. <1 >1012 13 8·1010

Polymer 2 <1 >1012 340 3·109 2.2 5·1011 634 2·109

Polymer 3 <1 >1012 425 2·109 6 2·1011 400 2·109

6.5.2.Chloroform solution

The photoluminescence spectra of polymers 1, 2 and 3, recorded in chloroform solution after

excitation of either donor or acceptor, exhibit very weak emission intensities. In particular, the

fluorescence emission of the OPV chromophore is 60 times lower than that of the starting monomer

12. The PL of the PERY chromophore is also quenched in polymer 1, by a factor of 20 with respect to

the PL of the PERY model compound (Figure 6.6). For polymers 2 and 3 the PL quenching factor of

the PERY chromophore scales to 450 times. The modest PL quenching in copolymer 1, with respect

to 2 and 3, can be attributed to a less efficient charge transfer in the copolymer caused by the presence

of some domains of only OPV or PERY chromophores within the polymer chains.

Page 150: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Polyacetylenes with pendant donor-acceptor dyads

139

500 600 7000

50

Inte

nsity

(a.

u.)

Wavelength (nm)

Figure 6.8. PL spectra of monomers 10 (solid squares) and 12 (solid circles) and polymer 1 in

chloroform solution, after selective excitation of the OPV (open circles) and of the PERY (open

squares) chromophores at 400 and 520 nm respectively.

According to sub-picosecond pump-probe spectroscopy, charge separation occurs

immediately in copolymer 1 (Figure 6.9 left and Table 6.1), which is consistent with a face-to-face

orientation of donor and acceptor, the most probable orientation of two neighboring pendant

chromophores in the copolymer. The face-to-face orientation of the chromophores is also responsible

for a very fast recombination of the charges. Additional to the fast decay of the majority of the

charges, a weak signal remains that only decays very slowly. This residual signal most probably

corresponds to trapped charges in small domains of donor or acceptor units. The presence of small

domains of chromophore units with equal signature has also been inferred from fluorescence

spectroscopy.

In polymers 2 and 3 the charges are generated more slowly than in polymer 1, but what is

more important, their lifetimes have been increased by more than one order of magnitude (Figure 6.9

and Table 6.1). This is the result of the different design of the last two polymers. The spacer placed

between the donor and the acceptor moieties slows down the charge separation and recombination

processes by intercalating a distance between the redox centers and, because of its semirigid nature,

by preventing the folding of the PERY donor on top of the donor segment to a face-to-face

orientation.9b

Both for polymer 2 and for polymer 3 the formation of the charges in the solid state is faster

than in solution (Table 6.1). However, for polymer 2 the formation of the charges is increased by only

a factor of 2 upon going from solution to the solid state, whereas for polymer 3 this increases by a

factor of 6. This observation is in line with the results obtained by absorption spectroscopy, which

show that the conformation and aggregational state of polymer 2 in solution closely resembles that in

Page 151: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 6

140

the solid state. For polymer 3 this relationship was not manifested as strongly as for polymer 2.

Polymer 3 was found to be much better dissolved in chloroform and hence features a much larger

conformational flexibility. The decrease in conformational flexibility for polymer 3 upon going to the

solid state accounts for the chromophores lining up in close proximity, thus very rapidly generating

the charges.

0 200 400 600 800 1000

-5

0

∆T (

a. u

.)

Time (ps)

∆T (

a. u

.)

Time (ps)

-3 0 3

-5

0

0 200 400 600 800 1000

-100

-50

0

Time (ps)

∆T (

a. u

.)

Time (ps)

-5 0 5 10 15-100

-50

0

Figure 6.9. Differential transmission dynamics of the 1450 nm absorption of polymers 1 (left, solid

line), 2 (right, solid squares) and 3 (right, hollow squares) in chloroform solution, after excitation at

450 nm. The insets show the 1450 nm signals on short timescales.

Similar as for the charge separation in polymer 2 also the lifetime of the charges is decreased

by a factor of 2 upon going from solution to the solid state. This reduction of the lifetime results

probably from the combination of denser packing of the chromophores in the solid state and increased

interchain interactions, both speeding up the charge recombination in polymer 2 in the solid state.

Surprisingly, for polymer 3 a different behavior of the lifetime of the charges is observed. The

lifetime of the charges is similar in solution and in the solid state. This phenomenon might be the

result of the alkoxy substitution of the OPE element. Although a denser packing might also occur for

polymer 3 upon going to the solid state, the alkoxy side chains might hamper lateral interaction of the

OPE chromophores with acceptors belonging to other polymer chains, preventing ‘interchain

recombination’, and thus allowing longer lifetimes. The alkoxy side chains might play the same role

as the spacer between donor and acceptor in a single dyad.

6.6 Conclusions and outlook

Several polyacetylenes substituted with a OPV or OPE donor and a PERY acceptor have been

synthesized for the first time. By designing different architectures, the strengths and weaknesses of

the placement of the functional groups on processability and photophysical properties have been

explored.

Page 152: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Polyacetylenes with pendant donor-acceptor dyads

141

The polymers were prepared either by the copolymerization of OPV and PERY terminal

acetylenes or by the polymerization of homologous OPVE-PERY and OPE-PERY dyads. SEC

analyses exhibit the typical distribution corresponding to polymeric systems. The molecular weights

of the polymers are in the range of 15-55 kDa giving especially polymers 2 and 3 good processing

properties, due to their higher molecular weight and good solubility in common organic solvents.

Nevertheless, the low solubility of especially polymers 1 and 2 in toluene, used in the polymeriztion,

limits the chain length. Although the polydispersity of the polymers is good, the degree of

polymerization might eventually be optimized by a different choice of the polymerization solvent.

Even though the copolymerization approach as applied for polymer 1 offers the fastest

synthetic access to donor-acceptor functionalized polyacetylenes, the alternating arrangement of

donors and acceptors results in very short-lived charged-separated states. It remains to be seen if this

can be overcome by another design of the two monomers. On the contrary, polymers 2 and 3 show

promising photophysical properties owing to their more refined architecture, avoiding donor-acceptor

stacking and favoring donor-donor and acceptor-acceptor interactions. In these polymers the lifetime

of the photoinduced charge separated state is increased by at least more than one order of magnitude

with respect to polymer 1.

These results, together with the optimization of the polymerization conditions, make these

polyacetylenes an attractive system to explore in future experiments.

6.7 Experimental Section

For general methods the reader is referred to chapter 2 of this thesis. [(S)-1-Isobutyl-2-(4-iodo-phenyloxy)ethyl]carbamic acid tert-butyl ester (5). Iodophenol (4.34g, 19 mmol), 49 (4.25 g, 19 mmol) and triphenylphosphine (7.77 g, 29 mmol) were dissolved in toluene (60 mL). Diethyl azodicarboxylate (4.66 ml, 29 mmol) in toluene (30 mL) was slowly added while keeping the temperature below 35 °C. The reaction mixture was stirred for 16 h at room temperature and subsequently washed with 1N aqueous HCl. The organic phase was dried over MgSO4 and evaporated in vacuo. Column chromatography (n-heptane/EtOAc, 4/1, Rf = 0.3) and crystallization from hexane afforded the product (3.2 g, 40%) as white crystals.1HNMR (CDCl3, 300 MHz): δ 7.54 (d, 2H), 6.67 (d, 2H), 4.75-4.65 (m, 1H), 4.10-3.8 (m, 3H), 1.80-1.65 (m, 1H), 1.60-1.40 (m, 11 H), 1.00-0.88 (m, 6H); 13C NMR (CDCl3, 75 MHz): δ .158.50, 155.18, 138.04, 116.80, 82.91, 79.36, 70.25, 48.15, 41.00, 28.48, 24.94, 23.12, 22.38; Anal. Calc for C17H26INO3: C, 48.7; H, 6.2; N, 3.3. Found: C, 48.9; H, 6.1; N, 3.28. (S)-1-Isobutyl-2-(4-iodo-phenyloxy)-ethylamine (6). TFA (12 mL) was added to a solution of 5 (3 g, 7.3 mmol) in methylene chloride (12 mL). After stirring the reaction mixture for 16 h at room temperature under an argon atmosphere, NaHCO3 was added. The reaction mixture was subsequently washed with water, NaOH 1 N, water, brine and dried over Na2SO4.The product (2 g, 86%) was obtained after evaporation of the solvent as a colorless oil.1H NMR (CDCl3, 300 MHz): δ 7.54 (d, 2H), 6.68 (d, 2H), 3.86 (dd, 1H), 3.64 (dd, 1H), 3.22 (s, 1 H), 1.81-1.74 (m, 1 H), 1.25 (t, 2H), 0.97-0.92 (m, 6H); 13C NMR (CDCl3, 75 MHz): δ .158.64, 137.99, 116.76, 82.66, 73.69, 48.27, 42.98, 24.46, 23.27, 21.91.

Page 153: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 6

142

N-[(S)-1-Isobutyl-2-(4-iodo-phenyloxy)-ethyl]-N’-(1-ethylpropyl)-3,4,9,10-perylenebis(dicarboximide) (8). N-(1-ethylpropyl)-3,4,9,10-perylenetetracarboxylic monoanhydride monoimide 518 (1.91 g, 4.17 mmol), amine 4 (2 g, 6.26 mmol), imidazol (30 g, mmol), and catalytic amounts of Zn(AcO)2 were mixed and stirred for 2.5 h at 160 °C. After cooling to room temperature the solid reaction mixture was dissolved in CH2Cl2 and washed with HCl 1N, water, brine and dried over MgSO4. Purification by column chromatography (flash silica, CH2Cl2, Rf = 0.36) yielded 1.97 g (67%) of as a dark red solid.1HNMR (CDCl3, 300 MHz): δ 8.76-8.59 (m, 8H), 7.46 (d, 2H), 6.65(d, 2H), 5.77-5.69 (m, 1H), 5.11-5.03 (m, 1H), 4.72 (dd, 1H), 4.33-4.30 (dd, 1H), 2.33-2.21 (m, 3H), 2.00-1.90 (m, 2H), 1.81- 1.75 (m, 1H), 1.67-1.59 (m, 1H) 1.02 (d, 3H), 098 (d, 3H), 0.94 (t, 6H); 13C NMR (CDCl3, 100 MHz): δ 163.16, 158.20, 137.87, 133.58, 133.22, 131.15, 130.61, 128.72, 125.35, 123.18, 122.32, 122.12, 117.02, 82.98, 68.78, 57.77, 54.41, 38.63, 25.64, 25.04, 23.14, 22.60, 11.62. Anal. Calc for C41H35IN2O5: C, 64.6; H, 4.6; N, 3.7. Found: C, 64.3; H, 4.4; N, 3.5. MALDI-TOF MS (Mw = 762.65) m/z= 762.08 [M]+. N-[(S)-1-Isobutyl-2-(4-trimethylsylylethynyl-phenyloxy)-ethyl]-N’-(1-ethylpropyl)-3,4,9,10-perylenebis(dicarboximide) (9). A mixture of 8 (1.42 g, 1.86 mmol), Pd(PPh3)2Cl2 (0.104 g, 0.148 mmol), CuI ( 0.021 g, 0.107 mmol) was dissolved in anhydrous toluene/triethylamine (320 mL, 5:1). After argon was purged trough the solution for 15 min trimethylsylylacetylene (0.32 mL, 2.23 mmol) was added. The reaction mixture was heated at 50 oC for 4 h. The crude reaction mixture was filtered through Celite, evaporated to dryness, and chromatographied (flash silica, CH2Cl2, Rf = 0.35) to afford a red solid (1.2 g, 84 %). 1H NMR (CDCl3, 300 MHz): δ 8.56-8.36 (m, 8H), 7.34 (d, 2H), 6.84 (d, 2H), 5.76-5.70 (m, 1H), 5.10-5.03 (m, 1H), 4.80 (dd, 1H), 4.39-4.34 (dd, 1H), 2.33-2.20 (m, 3H), 2.04-1.90 (m, 2H), 1.87-1.75 (m, 1H), 1.68-1.61 (m, 1H), 1.07-0.94 (m, 12H), 0.18 (s, 9H); 13C NMR (CDCl3, 100 MHz): δ 163.80 (broad signal), 158.78, 134.36, 134.01, 133.37, 131.20, 129.33, 126.08, 123.47, 122.88, 122.71, 115.41, 114.61, 105.10, 92.43, 68.64, 57.72, 53.40, 51.40, 38.53, 25.55, 24.98, 23.10, 22.42, 11.38, -0.03. Anal. Cald for C41H35IN2O5: C, 64.6; H, 4.6, N, 3.4. Found: C, 64.3; H, 4.4, N, 3.4. MALDI-TOF MS (Mw = 732.96) m/z= 732.30 [M]+. N-[(S)-1-Isobutyl-2-(4-ethynyl-phenyloxy)-ethyl]-N’-(1-ethylpropyl)-3,4,9,10-perylenebis(dicarboximide) (10). To a solution of 9 (0.4 g, 0.56 mmol) and tetrabutylammonium fluoride on silica (0.177 g, 0.68 mmol) in 70 mL THF was stirred for 0.5 h under argon. The reaction mixture was filtered through Celite and evaporated to dryness. The product was obtained as a red powder (0.3 g, 81%) and was used without further purification. 1H NMR (CDCl3, 300 MHz): δ 8.63-8.50 (m, 8H), 7.35 (d, 2H), 6.84 (d, 2H), 5.76-5.70 (m, 1H), 5.11-5.02 (m, 1H), 4.80 (dd, 1H), 4.39-4.34 (dd, 1H), 2.94 (s, 1H), 2.33-2.20 (m, 3H), 2.04-1.90 (m, 2H), 1.87-1.75 (m, 1H), 1.68-1.61 (m, 1H), 1.07-0.94 (m, 12H); 13C NMR (CD2Cl2, 100 MHz): δ.163.81, 159.55, 134.15, 133.89, 133.75, 131.56, 131.01, 129.25, 125.87, 123.72, 122.90, 122.70, 115.16, 114.65, 83.76, 76.14, 69.31, 58.40, 51.64, 38.99, 25.93, 25.36, 23.23, 22.63, 11.71. MALDI-TOF MS (Mw =660.78) m/z = 660.38 [M]+. 1-Bromomethyl-4-iodobenzene (14). A mixture of 4-iodotoluene (10 g, 458.8 mmol), N-bromosuccinimide (9.79 g, 55 mmol) and benzoyl peroxide (0.44 g, 1.8 mmol) in dry carbon tetrachloride (14 mL) was stirred and heated to reflux for 3.5 h and then cooled and filtered. The red filtrate was washed with saturated sodium thiosulfate solution (10 mL), dried and filtered. The solvent was removed in vacuo and the solid residue was further purified by recrystalization in hexane. The product was obtained as an off-white solid (3.5 g, 25%). 1H NMR (CDCl3, 300 MHz): δ 7.67 (d, 2H), 7.15(d, 2H), 4.41 (s, 1H); 13C NMR (CDCl3, 100 MHz): δ 137.94, 137.40, 130.82, 94.12, 32.45. Diethyl(4-iodo-benzyl)phosphonate (15). Triethyl phosphite (2.28 g, 13 mmol) and 14 (3.40 g, 11 mmol) were stirred at 160 oC for 1.5 h under an argon atmosphere. The reaction mixture was cooled to 75 oC and the ethyl bromide, formed during the reaction, and the excess of triethyl phosphite were distilled under reduced pressure. The product was obtained as a light yellow oil. Yield 3.89 g (100%). 1H NMR (CDCl3) δ 7.54(d, 2H), 6.97 (dd, 2H), 4.00- 3.84(m, 4H), 2.98 (d, 2H), 1.17(t, 6H); 13C NMR (CDCl3) δ 137.32 (d), 131.48 (d), 131.20 (d), 92.09 (d), 61.93 (d), 34.00, 32.17, 16.16 (d). (E,E)-1,4-Bis(4-iodo-styryl)-2-(3,7-dimethyloctyloxy)-5-methoxybenzene (17). Phosphonate 15 (0.8 g, 2.25 mmol) was dissolved in dry DMF (9 mL) under an argon atmosphere and 0.32 g (2.8 mmol) of KtBuO were added to the solution. After 15 min, a solution of dialdehyde 1619 (0.29 g, 0.9 mmol) in dry DMF (6 mL) was added dropwise and the reaction mixture was stirred for 3.5 h. The solution was poured on crushed ice and of 6 M HCl (200 mL) was added. The aqueous phase was extracted twice with diethyl ether

Page 154: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Polyacetylenes with pendant donor-acceptor dyads

143

and the combined organic layers were subsequently washed with 3 M HCl, water, a saturated aqueous solution of Na2CO3 and dried over MgSO4. The solvent was removed in vauo. Column chromatography (flush silica gel, CH2Cl2/pentane, 4:1, Rf = 0.3) and evaporation of the solvent afforded 0.42 g (64%) of the product as a yellow solid: 1H NMR (CDCl3; 300 MHz): δ 7.67 (d, 2H), 7.67 (d, 2H), 7.47(d, 1H), 7.46 (d, 1H), 7.27 (d, 2H), 7.26 (d, 2H), 7.11 (s, 1H), 7.09 (s, 1H), 7.06 (d, 1H), 7.02 (d, 1H), 4.09 (t, 2H), 3.91 (s, 3H), 1.97-1.86 (m, 1H), 1.82-1.45 (m, 4H), 1.45-1.05 (m, 6H), 0.99 (d, 3H), 0.87 (d, 3H). 13C NMR (CD2Cl3, 100 MHz) δ 151.54, 151.19, 137.76, 137.54, 137.48, 128.25, 128.20, 127.73, 127.69, 126.58, 126.35, 124.20, 123.98, 110.35, 109.09, 92.46, 92.42, 67.75, 56.19, 39.27, 37.41, 36.49, 30.14, 28.03, 24.85, 22.49, 22.39, 19.57. MALDI-TOF MS (Mw =720.48) m/z= 719.95[M]+. N-[(S)-1-Isobutyl-2-<4-{(E,E)-4-[4-(4-iodostyryl)-2-(3,7-dimethyloctyloxy)-5-methoxy)styryl]phenylethynyl}phenyloxyl>ethyl]-N’-(1-ethylpropyl)-3,4,9,10-perylenebis(dicarboximide) (18). A degassed solution of 17 (0.19 g, 0.26 mmol), 10 (0.12 g, 0.18 mmol), Pd(PPh3)2Cl2 (0.01 g, 0.014 mmol), CuI (0.002 g, 0.01 mmol) in anhydrous toluene/triethylamine (30 mL, 5:1) was heated at 50 oC for 16 h. The crude reaction mixture was filtered through Celite, evaporated to dryness, and chromatographied (flash silica, CHCl3/pentane, 9:1, Rf = 0.5) to afford a red solid (0.102 g, 47 %). 1H NMR (CDCl3, 400 MHz): δ 8.56-8.36 (m, 8H), 7.64 (d, 1H), 7.62 (d, 1H), 7.44-7.38 (m, 8H), 7.22 (d, 1H), 7.21 (d, 1H), 7.06-6.97 (m, 4H), 7.01 (d, 1H), 7.00(d, 1H), 6.99 (d, 1H), 6.93-6.90 (m, 2 H), 5.81-5.74 (m, 1H), 5.11-5.03 (m, 1H), 4.82 (dd, 1H), 4.41-4.38 (dd, 1H), 4.08-3.99 (m, 2H), 3.87 (s, 3H), 2.34-2.22 (m, 3H), 2.03-1.81 (m, 4H), 1.78-1.13 (m, 4 H),1.58-1.16 (m, 6H), 1.06 (d, 3H), 1.02 (d, 3H), 0.98 (d, 3H), 0.96 (t, 6H), 0.86 (d, 3H), 0.85 (d, 3H); 13C NMR (CDCl3, 100 MHz): δ 163.80 (broad signal), 158.66, 151.35, 151.37, 151.04, 137.65, 137.58, 137.36, 137.30, 134.39, 134.04, 132.93, 131.62, 131.56, 131.18, 129.34, 129.30, 128.24, 128.18, 128.11, 127.64, 127.61, 126.77, 126.52, 126.41, 126.34, 126.29, 126.14, 126.08, 124.17, 123.99, 123.91, 123.72, 122.87, 122.74, 122.31, 115.58, 114.85, 110.19, 108.93, 92.54, 92.49, 90.28, 88.40, 68.69, 67.63, 57.70, 56.13, 51.45, 39.19, 38.58, 37.36, 36.43, 30.06, 29.66, 27.93, 25.58, 24.98, 24.80, 24.78, 23.12, 22.67, 22.58, 22.48, 19.74, 11.41. MALDI-TOF MS (Mw =1253.34) m/z= 1252.23[M]+. N-[(S)-1-Isobutyl-2-<4-{(E,E)-4-[4-(4-trimethylsylylethynyl-styryl)-2-(3,7-dimethyloctyloxy)-5-methoxy)styryl]phenylethynyl}phenyloxy>ethyl]-N’-(1-ethylpropyl)-3,4,9,10-perylenebis(dicarboximide) (19). A mixture of 18 (0.073 g, 0.058 mmol), Pd(PPh3)2Cl2 (0.005 g, 0.007 mmol), CuI ( 0.001 g, 0.005 mmol) was dissolved in anhydrous toluene/triethylamine (18 mL, 5:1). After argon was purged trough the solution for 15 min trimethylsylylacetylene (0.1 mL, 0.707 mmol) was added. The reaction mixture was heated at 50 oC for 16 h. The crude reaction mixture was filtered through Celite, evaporated to dryness, and chromatographied (flash silica, CH2Cl2, Rf = 0.5) to afford a red solid (57 mg, 80%). 1H NMR (CDCl3, 300 MHz): δ 8.56-8.37 (m, 8H), 7.45-7.37 (m, 12H), 7.06 (d, 1H), 7.05, 7.04, 7.03, 7.03 (4xs, 2H), 7.02 (d, 1H), 6.92-6.89 (m, 2H), 5.80-5.73 (m, 1H), 5.09-5.03 (m, 1H), 4.82 (dd, 1H), 4.41-4.37 (dd, 1H), 4.05-4.01 (m, 2H), 3.82 (s, 3H2.34-2.22 (m, 3H), 2.03-1.81 (m, 4H), 1.78-1.13 (m, 4 H), 1.58-1.16 (m, 6H), 1.06 (d, 3H), 1.02 (d, 3H), 0.98 (d, 3H), 0.96 (t, 6H), 0.86 (d, 3H), 0.85 (d, 3H) ), 0.18 (s, 9H); 13C NMR (CDCl3, 100 MHz): δ 163.77 (broad signal), 158.66, 151.38, 151.07, 137.99, 137.41, 137.33, 134.39, 134.04, 132.91, 132.21, 131.59, 131.55, 131.13, 130.72, 129.30, 128.19, 128.06, 126.72, 126.56, 126.47, 126.33, 126.27, 126.18, 126.08, 124.33, 124.06, 123.75, 123.38, 122.87, 122.71, 122.28, 121.80, 115.58, 114.84, 110.21, 108.92, 105.26, 94.99, 90.25, 88.39, 68.68, 67.65, 57.69, 56.11, 53.39, 51.45, 39.19, 38.56, 37.34, 36.41, 30.05, 27.93, 25.57, 27.98, 24.77, 23.09, 22.65, 22.55, 22.45, 19.71, 11.38, -0.054. MALDI-TOF MS (Mw =1223.65) m/z = 1222.36 [M]+. N-[(S)-1-Isobutyl-2-<4-{(E,E)-4-[4-(4-Ethynyl-styryl)-2-(3,7-dimethyloctyloxy)-5-(methoxy)styryl]phenylethynyl}phenyloxy>ethyl]-N’-(1-ethylpropyl)-3,4,9,10-perylenebis(dicarboximide) (20). A solution of 19 (50 mg, 0.041 mmol) and tetrabutylammonium fluoride on silica (0.05 mL of a 1M solution, 0.05 mmols) in 5 mL THF was stirred for 0.5 h under argon. The reaction mixture was filtered through Celite and evaporated to dryness. The product was obtained as a red powder (46 mg, 97 %) and was used without further purification. The complete deprotection of the ethynyl bond was confirmed by 1H-NMR spectroscopy as the disappearance of the singlet at 0.18 ppm corresponding to the TMS end group and the appearance of the signal for the ethynylinic proton at 3.10 ppm. 2,5-Dibromo-hydroquinone (22). p-Benzoquinone (15 gr, 136 mmol) was dissolved in concentrated HBr (320 mL). Bromine (22.5 gr, 140 mmol) was added slowly to the solution while stirring. After 7 h the reaction mixture was heated to 100 °C and stirred

Page 155: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 6

144

for another 12 h. An additional amount of bromine (7 g, 43 mmol) was added and reacted at 100 °C for another 3 h. The reaction mixture was cooled to room temperature and the precipitated solid washed with water. The product was purified by repetitive recrystalization in water with addition of charcoal and obtained as white crystals (17.98 g, 49%). 1H NMR (CD3CO CD3, 400 MHz): δ 8.58 (s, 2H), 7.16(s, 2H); 13C NMR (CDCl3, 100 MHz): δ 147.98, 120.05, 108.90. 1,4-Dibromo-2,5-bis(2-ethylhexyloxy)benzene (23). Dibromohydroquinone 22 (26.27 g, 98 mmol) was dissolved in a solution of potassium hydroxide (21.99 g, 392 mmol) in ethanol (1250 mL) and heated to reflux, after what 2-ethylhexyl-p-toluenesulfonate (97.56 g, 343 mmol) was added. The reaction mixture was cooled to room temperature after 16 h. The precipitate formed was filtered off and washed with methanol. The solvent of the liquid phase was removed in vacuo. The resulting crude product was purified by column chromatography ( silica gel, hexane, Rf = 0.35). Evaporation of the solvent yielded 27.81 g (57 %) of a colorless oil. 1H NMR (CD3CO CD3, 300 MHz): δ 7.08 (s, 2H), 3.83 (d, 4H), 1.80-1.68 (m, 2H), 1.60-1.26 (m, 18 H), 0.96-0.90 (m, 12H); 13C NMR (CDCl3, 75 MHz): δ 150.28, 118.29, 111.15, 72.64, 39.52, 30.53, 29.11, 23.96, 23.10, 14.14, 11.23. 2,5-Bis(2-ethylhexyloxy)-4-bromo-iodobenzene (24). A solution of 23 (28 g, 56 mmol), in THF (225 mL) was cooled to –90 °C. n-BuLi in hexane (22.74 mL, 2.5 M) was added at such a rate that the internal temperature did not exceed –80 °C. This cold solution was added via a cannula to 1,2-diiodoethane (17.05 g, 60.5 mmol) in THF (113 mL) at –60 °C. After addition, the cooling bath was removed, and the reaction mixture was stirred for 30 min. The color changed from yellowish to brown. The reaction mixture was washed with saturated aqueous Na2S2O5 untill it lost the color. The water phase was extracted three times with diethyl ether. The combined organic phases were dried over MgSO4 and the solvent removed in vacuo. The crude product was purified by flash chromatography (silica gel, hexane, Rf = 0.5). Evaporation of the solvent yielded 22.10 g (72%) of a colorless oil. 1H NMR (CDCl3; 400 MHz): δ 7.27 (s, 1H), 6.98 (s, 1H), 3.85-3.81 (m, 4H), 1.80-1.70 (m, 2H), 1.60-1.48 (m, 16H), 0.96-0.90 (m, 12H); 13C NMR (CDCl3, 100 MHz): δ 152.04, 150.06, 123.52, 116.25, 112.01, 84.22, 72.24, 72.01, 39.22, 39.20, 30.30, 30.24, 28.84, 28.90, 23.75, 23.68, 22.84, 13.93, 13.90, 11.03, 11.01. 2,5-Bis(2-ethylhexyloxy)-4-trimethylsilylethynyl-bromobenzene (25). To a degassed solution of 24 (22 g, 40.77 mmol) in Et3N (123 mL), Pd(PPh3)2Cl2 (0.28 g, 0.4 mmol) and CuI (0.15 g, 0.8 mmol) were added, while cooling in an ice bath. (Trimethylsilyl)acetylene (5.76 mL, 40.97 mmol) was added and the mixture was stirred at 0 °C. After 50 min the reaction was quantitative according to GC-MS. The solvent was removed in vacuo and the residue was dissolved in water and diethyl ether. The water phase was extracted three times with diethyl ether. The combined organic phases were washed with saturated aqueous NH4Cl and dried over MgSO4. The solvent was removed and the residue was flash chromatographed (silica gel, hexane, Rf = 0.25). The product was obtained as a colorless oil (17.11 g, 82%). 1H NMR (CDCl3; 300 MHz): δ 7.03 (s, 1H), 6.93 (s, 1H), 3.88-3.77 (m, 4H), 1.77-1.69 (m, 2H), 1.61-1.25 (m, 16H), 0.96-0.90 (m, 12H), 0.24 (s, 9H); 13C NMR (CDCl3, 75 MHz): δ 154.89, 149.40, 117.62, 117.55, 113.52, 112.26, 100.68, 98.99, 72.35, 71.94, 39.61, 39.47, 30.47, 29.10, 23.89, 23.05, 14.06, 11.25, 11.14,-0.07. 2,5-Bis(2-ethylhexyloxy)-4-ethynyl-bromobenzene (26). To a solution of 25 (5.49 g, 10.7 mmol) in dry THF was added 1 M tetrabuthylammonium fluoride in THF (10.79 mL). The reaction mixture was stirred for 1 min and subsequently filtrated over silica gel using chloroform as eluent. The solvent was removed in vacuo yielding 4.68 g (100%) of a slightly colored oil which was used without further purification: 1H NMR (CDCl3; 300 MHz): δ 7.07 (s, 1H), 6.96 (s, 1H), 3.85-3.82 (m, 4H), 3.26 (s, 1H), 1.77-1.69 (m, 2H), 1.61-1.25 (m, 16H), 0.96-0.90 (m, 12H). 2,5-Bis(2-ethylhexyloxy)-4-trimethylsilylethynyl-iodobenzene (27). Following a similar procedure as for the synthesis of compound 24, the reaction of 25 (8.5 g, 16.6 mmol) in THF (100 mL) with 2.5 M BuLi in hexane (6.7 mL) and 1,2-diiodoethane (5.52 g, 19.6 mmol) in THF (50 mL) gave, after chromatography (flash silica gel, hexane, Rf = 0.23), 27 (7.2 g, 77%, purity~ 90%). 1H NMR (CDCl3; 400 MHz): δ 7.24 (s, 1H), 6.93 (s, 1H), 3.77-3.50 (m, 4H), 1.77-1.69 (m, 2H), 1.61-1.25 (m, 16H), 0.96-0.90 (m, 12H), 0.24 (s, 9H); 13C NMR (CDCl3, 75 MHz): δ 155.03, 151.70, 123.38, 115.88, 113.23, 100.85, 99.21, 87.65, 72.12, 71.95, 39.62, 39.46, 30.53, 3047, 29.10, 29.06, 23.96, 23.50, 14.09, 11.27, 11.18, -0.08.

Page 156: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Polyacetylenes with pendant donor-acceptor dyads

145

4-[4-Trimethylsilylethynyl-2,5-bis(2-ethylhexyloxy)phenylethynyl]-2,5-bis(2-ethylhexyloxy)bromobenzene (28). To a degassed solution of 26 (4.71 g, 10.78 mmol) and 27 (5.5 g, 10.78 mmol) in Et3N (48 mL), Pd(PPh3)2Cl2 (76 mg, 0.1 mmol) and CuI (41 mg, 0.21 mmol) were added, while cooling in an ice bath. The mixture was stirred at 0 °C for 3.5 h. The solvent was removed in vacuo and the residue was dissolved in water and diethyl ether. The water phase was extracted three times with diethyl ether. The combined organic phases were washed with saturated aqueous NH4Cl, dried over MgSO4 and the solvent evaporated in vacuo. The residue was extensively purified by flash column chromatography (silica gel, pentane/diethyl ether, 100:0 to 99:1 Rf = 0.1-0.5). The product was obtained as a light yellow oil (5.9 g, 67). 1H NMR (CDCl3; 300 MHz): δ 7.09 (s, 1H), 6.98 (s, 1H), 6.93 (s, 1H) 6.92 (s, 1H), 3.92-3.78 (m, 8H), 1.79-1.72 (m, 4H), 1.63-1.37 (m, 16H), 1.32-1.30 (m, 16H), 0.97-0.84 (m, 24H), 0.26 (s, 9H); 13C NMR (CDCl3, 100 MHz): δ 154.35, 154.20, 153.44, 149.57, 117.82, 117.20, 117.02, 116.33, 114.47, 113.40, 113.12, 112.90, 101.22, 99.84, 90.81, 90.47, 72.28, 72.20, 71.88, 71.64, 39.63, 39.56, 39.54, 39.46, 30.60, 30.56, 30.50, 30.47, 29.09, 29.08, 29.02, 24.07, 24.03, 23.87, 23.07, 23.05, 14.09, 14.06, 11.30, 11.25, 11.11, -0.05. MALDI-TOF MS (Mw = 866.20) m/z = 866.38 [M]+. 4-[4-Ethynyl-2,5-bis(2-ethylhexyloxy)phenylethynyl]-2,5-bis(2-ethylhexyloxy)bromobenzene (29). To a solution of 28 (1.76 g, 2 mmol) in dry THF was added 1 M tetrabuthylammonium fluoride in THF (2 mL). The reaction mixture was stirred for 1 min and subsequently filtrated over silica gel using chloroform as eluent. The solvent was removed in vacuo. The product was purified by column chromatogragphy (flush silica gel, hexane/toluene, 4:1, Rf = 0.2). The product (1.55 g, 95%) is light yellow oil: 1H NMR (CDCl3; 300 MHz): δ 7.09 (s, 1H), 6.98 (s, 1H), 6.96 (s, 2H), 3.90-3.82 (m, 8H), 1.79-1.73 (m, 4H), 1.60-1.26 (m, 32H), 0.98-0.84 (m, 24H); 13C NMR (CDCl3, 75 MHz): δ 154.44, 154.24, 153.50, 149.61, 117.86, 117.54, 117.26, 116.70, 114.90, 113.22, 112.85, 112.48, 90.87, 90.26, 82.14, 80.05, 72.32, 72.20, 72.07, 71.97, 39.54, 39.41, 30.58, 30.48, 29.08, 29.03, 24.04, 23.87, 23.03, 14.03, 11.28, 11.10. 4-[4-Trimethylsilylethynyl-2,5-bis(2-ethylhexyloxy)phenylethynyl]-2,5-bis(2-ethylhexyloxy)iodobenzene (30). Following a similar procedure as for the synthesis of compound 24, the reaction of 28 (3.14 g, 3.55 mmol) in THF (100 mL) with 2.5 M BuLi in hexane (1.56 mL) and 1,2-diiodoethane (1.10 g, 3.91 mmol) in THF (50 mL) gave, after chromatography (flash silica gel, toluene/pentane, 1:4, Rf = 0.42), 30 (2.1 g, 63%, purity~85%). 1H NMR (CDCl3; 400 MHz): δ 7.29 (s, 1H), 6.93 (s, 1H), 6.92 (s, 1H), 6.87 (s, 1H), 3.90-3.81 (m, 8H), 1.79-1.71 (m, 4H), 1.54-1.26 (m, 32H), 0.97-0.85 (m, 24H), 0.26 (s, 9H); 13C NMR (CDCl3, 75 MHz): δ 154.34, 153.44, 151.89, 123.68, 117.01, 116.34, 115.50, 114.48, 113.90, 113.45, 101.22, 99.81, 91.00, 90.67, 87.18, 72.22, 72.05, 71.85, 71.62, 39.63, 39.56, 39.47, 30.57, 30.51, 30.49, 29.08, 29.03, 24.03, 23.93, 23.88, 23.03, 14.06, 11.30, 11.24, 11.14, -0.07. MALDI-TOF MS (Mw = 913.20) m/z = 912.47 [M]+. 4-<4-{4-[4-Trimethylsilylethynyl-2,5-bis(2-ethylhexyloxy)phenylethynyl]-2,5-bis(2-ethylhexyloxy) phenylethynyl}-2,5-bis(2-ethylhexyloxy)phenylethynyl >-2,5- bis(2-ethylhexyloxy)bromobenzene (31). To a degassed solution of 29 (1.20 g, 1.5 mmol) and 30 (1.4 g, 1.5 mmol) in Et3N (15 mL), Pd(PPh3)2Cl2 (52 mg, 0.075 mmol) and CuI (14 mg, 0.075 mmol) were added, while cooling in an ice bath. The mixture was stirred at 0 °C for 2.5 h. The solvent was removed in vacuo and the residue was dissolved in water and diethyl ether. The water phase was extracted three times with diethyl ether. The combined organic phases were washed with saturated aqueous NH4Cl, dried over MgSO4and the solvent evaporated in vacuo. The residue was purified with flash column chromatography (silica gel, toluene/hexane, 3:7 Rf = 0.25). The product was obtained as a greenish fluorescent oil (1.4 g, 60%). 1H NMR (CDCl3; 400 MHz): δ 7.12 (s, 1H), 7.01 (s,3H), 7.01 (s, 2H) 6.96 (s, 1H), 3.94-3.83 (m, 16H), 1.88-1.72 (m, 8H), 1.70-1.40 (m, 32H), 1.40-1.25 (m, 32H), 1.01-0.88 (m, 48H), 0.27 (s, 9H); 13C NMR (CDCl3, 100 MHz): δ154.33, 154.18, 153.67, 153.64, 1253.45, 149.56, 117.79, 117.17, 117.00, 116.62, 116.33, 114.57, 114.22, 114.15, 114.10, 113.39, 113.07, 112.94, 101.23, 99.81, 91.52, 91.41, 90.83, 90.60, 72.22, 72.14, 71.87, 71.81, 71.59, 39.62, 39.57, 39.53, 39.44, 30.59, 30.48, 30.45, 29.07, 28.99, 24.04, 23.86, 23.05, 23.00, 14.06, 14.02, 11.27, 11.22, 11.08, -0.09. MALDI-TOF MS (Mw = 1579.31) m/z = 1579.12 [M]+. N-[(S)-1-Isobutyl-2-<4-[4-{4-[4-Trimethylsilylethynyl-2,5-bis(2-ethylhexyloxy)phenylethynyl]-2,5-bis(2-ethylhexyloxy)phenylethynyl}-2,5-bis(2-ethylhexyloxy)phenylethynyl>-2,5-bis(2-ethylhexyloxy)phenylethynyl]phenyloxy>-ethyl]-N’-(1-ethylpropyl)- 3,4,9,10-perylenebis(dicarboximide) (32). A degassed solution of 31 (0.3 g, 0.19 mmol), 10 (0.12 g, 0.19 mmol), Pd(PPh3)2Cl2 (0.01 g, 0.014 mmol), CuI (0.002 g, 0.01 mmol) in anhydrous toluene/triethylamine (30 mL, 5:1) was heated at 50 oC for 16 h. The crude

Page 157: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 6

146

reaction mixture was filtered through Celite, evaporated to dryness, and chromatographied (flash silica, CH2Cl2/pentane, 3:2 to 4:1, Rf = 0.23-0.4) to afford a red solid (0.144 g, 34 %). 1H NMR (CDCl3, 400 MHz): δ 8.64-8.45 (m, 8H), 7.41 (d, 2H), 6.99 (s, 4H), 6.96 (s, 1H), 6.94 (s, 3H), 6.91 (d, 2H), 5.81-5.73 (m, 1H), 5.10-5.03 (m, 1H), 4.82 (dd, 1H), 4.40 (dd, 1H), 3.94-3.80 (m, 16H), 2.32-2.45 (m, 3H), 2.01-1.94 (m, 2H), 1.86-1.25 (m, 66H), 1.06-0.82 (m, 60H), 0.26 (s, 9H); 13C NMR (CDCl3, 100 MHz): δ 163.51, 158.66, 153.65, 151.32, 153.45, 131.48, 131.11, 132.88, 131.21, 129.43, 126.18, 122.95, 116.99, 116.62, 115.90, 114.78, 114.57, 114.24, 114.10, 113.84, 113.34, 101.23, 99.82, 94.82, 91.54, 91.37, 84.49, 71.86, 71.59, 68.69, 57.71, 51.45, 39.55, 38.57, 30.74, 30.57, 30.47, 29.06, 25.57, 24.98, 24.03, 23.94, 2.85, 23.04, 22.45, 14.034, 11.36, 11.27, -0.069 MALDI-TOF MS (Mw = 2159.17) m/z = 2158.31[M]+. N-[(S)-1-Isobutyl-2-<4-[4-{4-[4-Ethynyl-2,5-bis(2-ethylhexyloxy)phenylethynyl]-2,5-bis(2-ethylhexyloxy)phenylethynyl}-2,5-bis(2-ethylhexyloxy)phenylethynyl>-2,5-bis(2-ethylhexyloxy)phenylethynyl]phenyloxy>-ethyl]-N’-(1-ethylpropyl)-3,4,9,10-perylenebis(dicarboximide) (33). To a solution of 32 (76 mg, 0.035 mmol) and tetrabutylammonium fluoride impregnated on silica (0.042 mL of a 1M solution, 0.042 mmol) in 10 mL THF was stirred for 0.5 h under argon. The reaction mixture was filtered through Celite and evaporated to dryness. The product was obtained as a red powder (70 mg, 96%) and was used without further purification. The complete deprotection of the ethynyl bond was confirmed by 1H-NMR spectroscopy as the disappearance of the singlet at 0.2 ppm and the appearance of a singlet for the ethynylinic proton at 3.2 ppm. N-[(S)-1-Isobutyl-2-<4-[4-{4-[4-(4-Trimethylsilylethynyl-phenylethynyl)-2,5-bis(2-ethylhexyloxy)phenylethynyl]-2,5-bis(2-ethylhexyloxy)phenylethynyl}-2,5-bis(2-ethylhexyloxy)phenylethynyl]-2,5-bis(2-ethylhexyloxy)phenylethynyl>phenyloxy>-ethyl]-N’-(1-ethylpropyl)-3,4,9,10-perylenebis(dicarboximide) (34). A degassed solution of 33 (70 mg, 0.035 mmol), 4-iodotrimethylsylilphenylacetylene (34 mg, 0.123 mmol), Pd(PPh3)2Cl2 (2 mg, 0.003 mmol), CuI ( 0.3 mg, 0.002 mmol) in anhydrous toluene/triethylamine (6 mL, 5:1) was heated at 50 oC for 16 h. The crude reaction mixture was filtered through Celite, evaporated to dryness, and chromatographied (flash silica, CH2Cl2/pentane, 1:1 to 2:1, Rf = 0.1-0.3) to afford a red solid (30 mg, 40 %). 1H NMR (CDCl3, 400 MHz):δ 8.72-8.58 (m, 8H), 7.44 (s, 4H), 7.39 (d, 2H), 6.99 (s, 4H), 6.98 (s, 2H), 6.96 (s, 1H), 6.95 (s, 1H), 6.87 (d, 2H), 5.90-5.70 (m, 1H), 5.11-5.01 (m, 1H), 4.78 (dd, 1H), 4.38 (dd, 1H), 3.96-3.82 (m, 16H), 2.32-2.23 (m, 3H), 1.99-1.92 (m, 2H), 1.86-1.73 (m, 8H), 1.68-1.25 (m, 66H), 1.05-0.84 (m, 60H), 0.26 (s, 9H); 13C NMR (CDCl3, 100 MHz): δ 163.51 (broad signal), 158.66, 153.87, 153.67, 153.60, 134.72, 134.34, 132.88, 132.04, 131.85, 131.24, 129.61, 129.55, 126.43, 126.38, 123.64, 123.12, 122.97, 122.74, 116.63, 116.59, 116.54, 116.51, 116.47, 115.88, 114.78, 114.65, 114.31, 114.26, 114.13, 114.09, 113.86, 113.37, 104.71, 96.22, 94.83, 94.38, 91.74, 91.62, 91.58, 91.52, 91.42, 91.37, 88.16, 84.84, 71.96, 71.88, 68.65, 57.71, 51.45, 39.58, 38.55, 30.65, 30.59, 29.08, 25.57, 24.99, 54.03, 23.97, 23.95, 23.14, 23.06, 22.44, 14.06, 11.34, 11.31, 11.28, 11.23, 11.21, -0.10. MALDI-TOF MS (Mw = 2259.29) m/z = 2258.69[M]+. N-[(S)-1-Isobutyl-2-<4-[4-{4-[4-(4-Ethynyl-phenylethynyl)-2,5-bis(2-ethylhexyloxy)phenylethynyl]-2,5-bis(2-ethylhexyloxy)phenylethynyl}-2,5-bis(2-ethylhexyloxy)phenylethynyl]-2,5-bis(2-ethylhexyloxy)phenylethynyl]phenyloxy>-ethyl]-N’-(1-ethylpropyl)-3,4,9,10-perylenebis(dicarboximide) (35). To a solution of 34 (28 mg, 0.012 mmol) and tetrabutylammonium fluoride impregnated on silica (0.013 mL of a 1M solution, 0.013 mmol) in 5 mL THF was stirred for 0.5 h under argon. The reaction mixture was filtered through Celite and evaporated to dryness. The product was obtained as a red powder (26 mg, 99%) and was used without further purification. The complete deprotection of the ethynyl bond was confirmed by 1H-NMR spectroscopy as the disappearance of the singlet at 0.26 ppm and the arising of the signal for the ethynylinic proton at 3.2 ppm. Synthesis of the polymers: Polyphenylacetylene with pendant perylenes (11). To a mixture of the ethynyl terminated compound 10 (37 mg, 0.06mmol) and triethylamine (0.8mL) in toluene (20 mL) at room temperature [Rh(nbd)Cl]2 (2.6 mg, 0.006 mmol) was added under an argon atmosphere. The reaction mixture was stirred for 24 h and the polymer was isolated by precipitation into methanol as a red polymer. The polymer was insoluble in any common solvent.

Page 158: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Polyacetylenes with pendant donor-acceptor dyads

147

Polyphenylacetylene with pendant OPVs and perylenes (1). To a mixture of monomers 10 (29 mg, 0.044 mmol) and 12 (30 mg, 0.045 mmol) in triethylamine (0.8 mL) and toluene (20 mL), [Rh(nbd)Cl]2 (2.6 mg, 0.006 mmol) was added under an argon atmosphere. The reaction was stirred for 24 h and subsequently poured into stirring methanol. The polymer was isolated from the cyclic trimers side products using preparative size exclusion chromatography (Biobeads, SXIII, CH2Cl2) and obtained as a red solid (21 mg, 36%). SEC (chloroform/methanol, 99/1): Mw =14.6 kg/mol, PDI = 1.6. Polyphenylacetylene with pendant OPV-PERY dyads (2). To a solution of ethynyl monomer 20 (44 mg, 0.038 mmol) dissolved in toluene (13 mL) and triethylamine (0.6 mL), [Rh(nbd)Cl]2 (2 mg, 0.004 mmol) was added. The reaction was stirred for 16 h and subsequently poured into stirring methanol. The cyclic trimers were removed by precipitation in THF. The polymer was obtained as a red solid (26 mg, 60 %). SEC (chloroform/methanol, 99/1): Mw =45.8 kg/mol, PDI = 1.9. Polyphenylacetylenes with pendant OPE-PERY dyads (3). To a solution of ethynyl monomer 35 (26 mg, 0.012 mmo) dissolved in toluene (4 mL) and triethylamine (0.15 mL), [Rh(nbd)Cl]2 (0.55 mg, 0.001 mmol) was added. The reaction was stirred for 16 h and subsequently poured into stirring methanol. The cyclic trimers were removed by precipitation in THF. The polymer was obtained as a red solid (7 mg, 30 %). SEC (chloroform/methanol, 99/1): Mw = 53.5 kg/mol, PDI = 1.5. Absorption and Photoluminescence. UV/visible/near-IR absorption spectra were recorded on a Perkin Elmer Lambda 900 spectrophotometer. Fluorescence spectra were recorded on a Perkin Elmer LS 50B spectrometer, using a 4 nm bandwidth and optical densities of the solutions of 0.1 at the excitation wavelength. Transient subpicosecond photoinduced absorption. Solutions in the order of 10-5 M were excited at 450 nm, i.e. providing primarily excitation of the donor part within the molecules. 6.8 References and notes 1 (a) Yu, G.; Gao, J.; Hummelen, J. C.; Wudl, F.; Heeger, A. J. Science 1995, 270, 1789; (b) Halls, J. J. M.;

Walsh, C. A.; Greenham, N. C.; Marseglia, E. A.; Friend, R. H., Moratti, S. C.; Holmes, A. B. Nature

1995, 376, 498; (c) Shaheen, S. E.; Brabec, C. J.; Padinger, F.; Fromherz, T.; Hummelen, J. C.; Sariciftci,

N. S. Appl. Phys. Lett. 2001, 78, 841; (d) Schmidt-Mende, L.; Fechtenkötter, A.; Müllen, K.; Moons, E.;

Friend, R. H.; MacKenzie, J. D. Science, 2001, 293, 1119.

2 (a) Shinohara, K.; Yasuda, S.; Kato, G.; Fujita, M.; Shigekawa, H. J. Am. Chem. Soc. 2001, 123, 3619.

(b) Yashima, E.; Katsuhiro, M.; Okamoto, Y. Nature 1999, 399, 449.

3 Schenning, A. P. H. J.; Fransen, M.; van Duren, J. K. L.; van Hal, P. A., Janssen, R. A. J.; Meijer, E. W.

Macromol. Rapid Commun. 2002, 23, 271.

4 Vohlidal, J.; Sedlacek, J.; Patev, N.; Lavastre, O.; Dixneuf, P. H.; Cabioch, S.; Balcar, H.; Pfleger, J.;

Blechta, V. Macromolecules 1999, 32, 6439.

5 Nishimura, T.; Takatani, K.; Sakurai, S.; Maeda, K.; Yashima, E. Angew. Chem. Int. Ed. 2002, 41, 3602.

6 Dittmer, J. J.; Marseglia, E. A.; Friend, R. H. Adv. Mater. 2000, 12, 1270 and references therein.

7 (a) Masuda, T.; Higashimura, T. Adv. Polym. Sci. 1987, 81, 121. (b)Yashima, E.; Maeda, Y.;

Matsushima, T.; Okamato, Y. Chirality 1997, 9, 593. (c)Tabata, M.; Sone, T.; Sadahiro, Y. Macromol.

Chem. Phys. 1999, 200, 265.

8 Demmig, S.; Langhals, H. Chem. Ber. 1988, 121, 225.

9 (a) Peeters, E.; van Hal, P. A.; Meskers, S. C. J.; Janssen, R. A. J.; Meijer, E. W. Chem. Eur. J. 2002, 8,

4470. (b) Neuteboom, E. E.; Meskers, S. C. J.; van Hal, P. A.; van Duren, J. K. J.; Meijer, E. W.; Dupin,

Page 159: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Chapter 6

148

H.; Pourtois, G.; Cornil, J.; Lazzaroni, R.; Brédas, J.-L.; Beljonne, D.; Janssen, R. A. J. J. Am. Chem. Soc.

2003, 125, 8625.

10 Albert Schenning, unpublished results.

11 Peak broadening has been reported by Okamoto in polyphenylacetylenes which mobility was hampered

by complexation of the pendant carboxylic acid groups with optically active aminoalcohols. Yashima, E.;

Matsushima, T.; Okamoto, Y. J. Am. Chem. Soc. 1997, 119, 6345.

12 Ziener, U.; Godt, A. J. Org. Chem. 1997, 62, 6137.

13 Fujita, Y.; Misumi, Y.; Tabata, M.; Masuda, T. J. Pol. Sci. A: Pol. Chem. 1998, 36, 3157.

14 Salbeck, J. J. Electroanal. Chem. 1992, 340, 169.

15 Polymer PPE was synthesized following a method as reported in Weder, C.; Wrighton, M. S.

Macromolecules 1996, 29, 5157.

16 Peeters, E.; van Hal, P. A.; Knol, J.; Brabec, C. J.; Sariciftci, N. S.; Hummelen, J. C.; Janssen, R. A. J. J.

Phys. Chem. B 2000, 104, 10174.

17 Peeters, E. Mesoscopic Order in π-conjugated Materials, Ph.D. Thesis, Eindhoven University of

Technology, 2000, ISBN 90-386-2951-6.

18 Nagao, Y.; Naito, T.; Abe, Y.; Misono, T. Dyes and Pigments 1996, 32, 71.

19 Marcos Ramos, A.; Rispens, M. T.; van Duren, J. K. J.; Hummelen, J. C.; Janssen, R. A. J. J. Am. Chem.

Soc., 2001, 123, 6714-6715.

Page 160: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Summary

Combinations of π-conjugated materials with electron donor and electron acceptor properties

have been used successfully as the active layer in the so-called plastic solar cells. The efficiency of

these devices, however, still needs to be improved and much remains to be learned about the

photophysical processes occurring in the photoactive materials. This multidisciplinary problem can be

tackled from many points of view. Molecular engineering provides a unique means for determining

the role of the individual components in π-conjugated systems consisting of many more components

at the molecular to multi-molecular level. This approach might aid in understanding and optimising

the sequence of events that result in solar energy conversion.

One of the major goals of molecular engineering is to increase the lifetime of the charge-

separated states, as it allows the generated charges to diffuse away from each other, before the non-

desired charge recombination can occur. This can be achieved establishing a redox gradient within a

multichromophoric array, similar as how it occurs in photosynthesis. A number of multichromophoric

arrays with donor(1)-donor(2)-acceptor arrangements were synthesized, studied, and compared to

reference compounds made of the combination of the individual components. The synthesized

systems were the molecular triad oligoaniline-oligo(p-phenylene vinylene)-fullerene (OAn-OPV-C60)

and two symmetrical molecular pentads OAn-OPV-PERY-OPV-OAn (PERY meaning perylene

diimid) with different connection between the OPV and PERY chromophores. The photophysical

studies of these multichromophoric arrays in solution revealed that sequential intramolecular energy

and electron transfer leads to the formation of the donor(1)•+-donor(2)-acceptor•– charge-separated

state. The lifetime of this charged state extended in all cases to the nanosecond time-regime. The

quantum yield for the formation of this state in solution is a result of the interplay between the redox

potential of the different chromophores, nature of the linkage between them and polarity of the

medium. Although, the results could be rationalized in terms of the Marcus theory for charge transfer,

the theory does not suffice to predict or design a system with an efficient sequential electron transfer.

By going to the solid state, intermolecular interactions that might influence the processes occurring at

the molecular level have to be taken into consideration. As observed for the triad Oan-OPV-C60, in the

solid state the formation of donor(1)•+-donor(2)-acceptor•– charge separated state is favored with

respect to the solution, but also an increase of the recombination rate of the charges occurs. In the

bulk, control over these processes is difficult, unless structural tools such as supramolecular chemistry

are used to control the arrangement of the chromophores.

The additional use of supramolecular chemistry in the molecular engineering approach was

achieved with a donor-bridge-acceptor system in which secondary interactions were employed to

control the distance and orientation of the chromophores and thus to govern their photophysical

Page 161: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

interaction. Such a system has been created by covalently linking an OPV and a PERY at opposite

ends of a long foldable m-phenylene ethynylene oligomer (FOLD), OPV-FOLD-PERY. The

conformation of the bridge can be varied from a random coil to a folded helical state by decreasing

the polarity of the medium. This conformational change affects the distance between the

chromophores, and thus allows for tuning the system to undergo energy or electron transfer. In very

apolar media, additional intermolecular interactions arise bringing more than one acceptor and donor

in close proximity. This creates a ‘microsolid’ state situation, increasing the overall yield of charge

generation.

Judicious engineering of polymers provides another means of obtaining spatial organization.

The most convenient way to obtain the highest contact between donors and acceptors, as obtained in

solid films of donor-acceptor dyads, but in a less chaotic manner, can be achieved by connecting

molecular dyads with each other into a polymeric chain. Pursuing this idea, a possible design consists

of only linking the donors elements to each other in a linear array. This results in a π-conjugated

polymer with dangling acceptors. Such a polymer has been made with the conjugated backbone being

a hybrid between poly(p-phenylene vinylene) and poly(p-phenylene ethynylene) (PPVE) and

methanofullerenes as the pendant acceptors. The photophysical studies on this donor-acceptor

polymer reveal that photoexcitation of the polymer results in a photoinduced electron transfer reaction

from the conjugated backbone to the pendant C60 moieties. This novel polymer has been applied via

spin coating to form the active layer of the first polymer solar cell based on a covalently linked donor-

acceptor polymer. This first approach has yielded promising photovoltaic properties, with an

exceptional low acceptor loading.

A second type of donor-acceptor polymer was engineered based on OPV and/or OPE donors

and PERY acceptors. A polyacetylene backbone was decorated with dangling donor-acceptor dyads

made up out of these elements. With this last structural design both the covalent and supramolecular

approach have been implemented, pursuing not only intimate contact between donors and acceptors

but also spatial organization. As a direct result of the higher degree of order, in these polymers,

photoexcitation of the donor or acceptor generates a long-lived charge separation in solution as well

as in the solid state. This long-lived charge-separated state makes these polymers ideal candidates for

the active layer in photovoltaic devices.

The systems studied show that molecular engineering is a powerful tool for optimising the

photophysical processes in donor-acceptor systems and for the creation of promising materials for

photovoltaic devices. The ability to control the spatial organization of the donor and acceptor

materials and to direct both the intra- and intermolecular interactions provides a level of control not

achievable by other techniques. The more refined architectures obtained via molecular engineering

not only help in understanding the photophysical processes but are also the key to the photovoltaic

devices of tomorrow.

Page 162: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Resumen

Combinaciones de materiales conjugados con propiedades de dador y aceptor de electrones

han sido utilizadas con éxito en la elaboración de las llamadas ‘células solares de plástico’. No

obstante, la eficacia de dichos dispositivos no es todavía óptima y poco se conoce sobre los procesos

fotofísicos que ocurren en los materials activos que los componen. Éste es un problema

multidisciplinar que se puede abordar desde muchos puntos de vista. La ingeniería molecular

proporciona una herramienta única mediante la cual se puede determinar el papel que juega cada uno

de los componentes activos en sistemas conjugados complejos a nivel molecular o multimolecular.

Esta aproximación al problema puede ayudar a entender y optimizar la secuencia de sucesos que

resulta en la conversión de la energía solar.

Uno de los mayores objetivos de la ingeniería molecular es incrementar la vida media de los

estados de separación de carga para permitir la difusión de las cargas generadas, evitando así la

recombinación no deseada de las mismas. Esto se puede lograr estableciendo un gradiente de

potenciales redox en un sistema consistente en varios cromóforos enlazados, al igual que ocurre en la

fotosíntesis. Diversos sistemas multicromofóricos que se ajustan al patrón dador(1)-dador(2)-aceptor

han sido sintetizados, estudiados y comparados con compuestos de referencia resultantes de la

combinación de los cromóforos individuales. Los sistemas preparados son la tríada molecular

oligoanilina-oligo(p-fenilenvinileno)-fulereno (OAn-OPV- C60) y dos péntadas moleculares simétricas

OAn-OPV-PERY-OPV-OAn (PERY significando perylendiimida), en las cuales únicamente se varió

el enlace entre los cromóforos OPV y PERY. Los estudios fotofísicos de estos sistemas

multicromóforos realizados en disolución revelaron que una transferencia secuencial de energía y

carga conduce a la formación del estado cargado de separación dador(1)•+-dador(2)-aceptor•–. La vida

media de dicho estado es, en todos los casos, superior a 1 nanosegundo debido a la considerable

distancia que separa las cargas. El rendimiento de separación de carga en disolución depende de los

potenciales redox de los diferentes cromóforos, de la naturaleza del enlace que los une y de la

polaridad del medio. En estado sólido se tienen que considerar interacciones intermoleculares que

pueden influir en los procesos que occurren a nivel molecular. Para la triada OAn-OPV-C60 en estado

sólido la generación del estado de separación de carga OAn •+-OPV-C60•– resulta, por una parte, más

favorable que en solución, pero, por otra, su vida media es más corta. En estado sólido el control

sobre todos estos procesos es sumamente complicado. Herramientas estructurales como la química

supramolecular pueden asistir en regular la disposición relativa de los diferentes cromóforos.

Page 163: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

El uso de la química supramolecular como elemento en la estrategia de la ingeniería

molecular se consiguió con un sistema dador-puente-aceptor en el que interacciones secundarias

fueron empleadas para controlar la distancia y orientación de los cromóforos, siendo así posible

gobernar su interacción fotofísica. Dicho sistema fue creado mediante la unión de un OPV y un PERY

a ambos extremos de un oligómero m-fenilenetinileno (FOLD) de longitud considerable el cual se

puede plegar sobre si mismo adoptando una conformación helicoidal, OPV-FOLD-PERY. El cambio

conformacional entre ‘ovillo estadístico’ y ‘hélice’ ocurre en el puente FOLD cuando la polaridad del

medio decrece. Esta transformación resulta en un acortamiento de la distancia entre los cromóforos

OPV y PERY proporcionando la manera de modular la actividad fotofísica del sistema. En un medio

extremadamente apolar, interacciones intermoleculares se suman a las intramoleculares, acercando

más de un dador y un aceptor. Se crea así una situación de ‘estado microsólido’ en el que el

rendimiento total de la transferencia de carga se ve incrementado.

Otro medio con el que lograr organización de los cromóforos en el espacio es mediante un

juicioso diseño del polímero. La manera más conveniente de obtener el máximo contacto entre

dadores y aceptores en estado sólido, similar al obtenido con las díadas dador-aceptor, aunque de

manera menos caótica, es mediante la incorporación de dichas díadas en una misma cadena

polimérica. Un diseño fiel a esta idea consiste en unir sólo los elementos dadores en una disposición

lineal, resultando en un polímero conjugado dador con aceptores pendiendo de él. Un sistema así,

consistiendo en un polímero híbrido de poli(p-fenilenvinileno) y poli(p-fenilenetinileno) (PPVE) con

metanofulerenos colgando de la cadena principal, se ha sintetizado. Estudios fotofísicos mostraron

que la fotoexcitación de dicho polímero resulta en una transferencia de electrón fotoinducida de la

cadena principal al los fulerenos laterales. Este polímero ha sido usado como la parte activa de la

primera célula solar basada en polímeros de tipo dador-aceptor. Este primer diseño se caracteriza por

unas propiedades fotovoltaicas prometedoras, si bien con un bajo contenido en material aceptor.

Un segundo tipo de polímero dador-aceptor fue diseñado usando OPV y/o OPE (oligo(p-

fenilenetinileno)) como dadores y PERY como aceptores. Un poliacetileno fue decorado con díadas

dador-aceptor basadas en dichos cromóforos. En este último diseño estructural, interacciones

covalentes y supremoleculares han sido combinadas con el fin de conseguir no sólo un contacto

íntimo entre dadores y aceptores, sino también una mayor organización espacial. Como resultado

directo del orden en estos polímeros la fotoexcitación de cualquiera de los cromóforos genera un

estado de separación de carga de larga vida media, tanto en disolución como en estado sólido. La

longevidad de las cargas en dicho estado cargado hace de estos polímeros candidatos ideales como la

parte activa de células solares.

Page 164: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

El trabajo descrito en esta tesis demuestra que la ingeniería molecular es una potente

herramienta con la que se pueden optimizar los procesos fotofísicos en sistemas dador-aceptor y con

la que se puede generar materiales prometedores para la fabricación de células solares. La habilidad

de dominar la organización de los materials dadores y aceptores y de gobernar las interacciones intra-

e intermoleculares proporciona un nivel de control no asequible con ninguna otra técnica. Las

refinadas arquitecturas obtenidas mediante la ingeniería molecular no sólo ayudan a entender los

complejos procesos fotofísicos que puedan ocurrir entre diferentes cromóforos, sino que también son

la clave para las células solares del futuro.

Page 165: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1
Page 166: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Curriculum Vitae

Alicia Marcos Ramos was born on December 11th 1973 in Reus (Spain). Her secondary

school studies took place at the I. B. “Gaudí” in Reus. She studied Chemistry at the ‘Rovira i Virgili”

University in Tarragona, Spain. During her studies she visited the laboratory of Marcromolecular and

Organic Chemistry at the Eindhoven University of Technology, the Netherlands (Prof. Dr. E.W.

Meijer) with an Erasmus grant. After working as a research assistant in the same group she started as

a Ph.D. student in the same laboratories under the supervision of Professor Dr. R.A.J. Janssen. The

most important results of the investigations are described in this thesis. As of September 2003 she will

be working as a post-doctoral research fellow at TNO-Industrial Technology in Eindhoven, the

Netherlands.

Page 167: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1
Page 168: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

Acknowledgments

The Spanish popular saying ‘nunca te acostarás sin saber una cosa más’ has been more than

ever before true during my Ph.D. period. Four years as a Ph.D. student in Eindhoven have enriched

me not only at the scientific and but also at the personal level, sometimes step by step, often with

gigantic leaps. My promotor Professor René Janssen has been responsible for initiating most of these

steps. He has provided me with freedom to embark on my scientific journey and supported this with

stimulating ideas and criticism. For this ideal working atmosphere and his contagious optimism, I

would like to thank him. Working from day to day in a stimulating environment has truly been an

‘onbetaalbare’ experience. Concerning this I also would like to thank Professor Bert Meijer. Dr.

Stefan Meskers has been a beacon in the dark on many problems of physical organic origin and

furthermore it has always been a pleasure to listen to what he had to say in matters of very different

nature.

I would like to thank Paul van Hal, Edwin Beckers, and Jeroen van Duren for their important

contribution to this work. Working together with them not only gave nice results but also taught me a

lot about the disciplines of photophysics and photovoltaics. Dr. Jef Vekemans and dr. Albert

Schenning were always enthusiastic in discussing synthetic and conceptual issues, my sincere

appreciation for that. For the fullerene chemistry presented in chapters 2 and 5, I would like to thank

Professor Kees Hummelen, dr. Minze Rispens, and dr. Joop Knol. They have also been magnificent

hosts during the week I spent in the ‘extremely cold’ Groningen. I would like to thank Professor

Jeffrey Moore, dr. David Hill and dr. Ryan Prince for our cooperation on chapter 4. I want to thank

Paul Hamelinck for his hard work and enthusiasm in the difficult subject described in chapter 5.

Quisiera agradecer al Profesor Nazario Martín por corregir esta memoria y por formar parte del

tribunal de mi defensa. También quisiera agradecerle su grata hospitalidad durante el encuentro que

tuvo lugar en Madrid

During my stay in the SMO group I have had the pleasure to discuss with many people about

science and about daily things. I would like to thank everybody who has contributed in such a way to

this thesis and to make of it an enjoyable four years. I would like to thank the members of the

‘Janssen-group’ for their many suggestions and critics during the lunch-meetings. For help and advice

with characterization I would like to thank Joost van Dongen, Xianwen Lou, Ralf Bovee, and Henk

Eding. For facilitating matters outside chemistry I would like to thank Hans Damen, Hanneke

Veldhoen, Ingrid Dirkx, Emma Eltink, Carine van der Vaart, Joke Rediker, and Hanny van der Lee. I

want to thank Koen Pieterse for his help with computers and for being there in desperate situations.

My former roommates Jan-Willem, Jack, Fiorella, and Hinke, I would like to thank them for

the entertaining conversations, resulting in laughter that sometimes even managed to shield us from

Page 169: Molecular engineering of photoinduced charge separation · Photoinduced multistep energy and electron transfer in an oligoaniline-oligo(p-phenylene vinylene)-fullerene triad 17 2.1

the music coming out of lab 3. Fiorella, voglio ringraziarti anche per avermi insignato l’arte delle

buone maniere e la gentilezza. Agradezco a Cristina, Elena, Marcela, Joaquín, Laia, Jorge y Marga,

los españoles a los que he tenido el placer de conocer aquí, porque reirse en castellano no es lo mismo

que reirse en holandés o inglés. For the good moments I also want to thank Ky, Francesca, Michel,

Edda, Mitsutoshi, Daniela, Dodo, Corinne…

Quisiera agradecer a mis padres y hermanos por su cariño y apoyo todo este tiempo desde la

distancia. Y a Luc, por hacer de la realidad un sueño maravilloso.

Alicia.


Recommended