+ All Categories
Home > Documents > Donor-Acceptor dyads for molecular rectifying...

Donor-Acceptor dyads for molecular rectifying...

Date post: 12-Sep-2018
Category:
Upload: trinhthuy
View: 235 times
Download: 1 times
Share this document with a friend
248
Donor-Acceptor dyads for molecular rectifying devices Mykola Kondratenko Department of Chemistry McGill University Montreal, Quebec Canada June 2011 A Thesis submitted to McGill University In partial fulfillment of requirements for the degree of Doctor of Philosophy Mykola Kondratenko 2011
Transcript

Donor-Acceptor dyads for molecular

rectifying devices

Mykola Kondratenko

Department of Chemistry

McGill University

Montreal, Quebec

Canada

June 2011

A Thesis submitted to McGill University

In partial fulfillment of requirements for the degree of Doctor of Philosophy

Mykola Kondratenko 2011

i

To my parents

ii

Abstract

The interest in molecular electronics began in the 1970s with the work of Aviram

and Ratner, who proposed that a donor-acceptor dyad, specifically TTF––TCNQ

molecule (TTF–tetrathiafulvalene, –nonconjugated bridge and TCNQ–

tetracyanoquinodimethane), can resemble the electric properties of a p-n junction, acting

as a unimolecular diode. The reason of such behaviour lies in asymmetrically distributed

electronic levels, and very low HOMO-LUMO gap (0.3 eV) that was imposed for the

model molecule. Up to date, numerous donor-acceptor dyads were investigated as

candidates for molecular rectifiers, which included some D––A dyads with weak or

moderate donor and acceptor moieties, numerous D––A and also molecules without

obvious asymmetry in the electronic structure. However, neither the original TTF––

TCNQ molecule nor any other molecule with similar HOMO–LUMO gap have been

studied in molecular electronics applications, which was due to synthetic unavailability

of such molecules.

In this thesis we present molecular design, synthesis as well as characterization

of series of Donor-Acceptor dyads with different combinations of well studied

electroactive moieties (TTF--fluorene, nEDOT-3CNQ, nEDOT-NDI). Herein we

describe our progress towards the main synthetic challenge in the field of molecular

rectifiers – coupling together strong donor and strong acceptor molecules. To achieve

this we employed different synthetic strategies, namely, use of intermediates with

moderated redox properties and highly reactive derivatives to avoid formation of charge-

transfer complexes between donor and acceptor as well as utilization of the donor-

acceptor complexation which results their covalent binding. The synthetic design

includes two types of approaches allowing binding the dyad molecules to electrode

surface: (1) amphiphilic structure which enables deposition of molecular monolayers via

Langmuir-Blodgett technique and (2) thiol/disulfide functionality suitable for covalent

grafting of molecules to metals. Characterization of such monolayers by different

spectroscopic and electrochemical techniques as well as analysis of the alignment and

packing of the molecules within the films and monolayers stability are discussed.

Finally, we describe fabrication of Electrode/Organic monolayer/Electrode junctions and

iii

discuss results of the charge transport measurements of the synthesized donor-acceptor

dyads.

iv

Résumé

L'intérêt pour l'électronique moléculaire a commencé dans les années 70, avec la

découverte d'Aviram et de Ratner. Ils ont proposé une dyade donneur-accepteur telle que

la molécule TTF––TCNQ (TTF–tetrathiafulvalene, liaisons isolées et TCNQ–

tetracyanoquinodimethane) qui pourrait fonctionner comme une jonction p-n, jouant le

rôle d'une diode unimoléculaire. Ce phénomène est dû à une distribution asymétrique

des niveaux électroniques, ainsi qu'au très faible écart HOMO-LUMO (0.3 eV) de cette

molécule. Jusqu'à présent, un grand nombre de dyades donneur-accepteur ont été

étudiées comme candidates pour la synthèse de redresseurs moléculaires. Ceux-ci

incluent certaines dyades D––A avec des groupes donneur et accepteur faibles ou

modérés, de nombreux dyades D––A, ainsi que des molécules ayant une forte

asymétrie dans leur structure électronique.

Dans cette thèse, nous présentons la conception moléculaire, la synthèse et la

caractérisation d'une série de donneur-accepteurs avec de différentes combinaisons de

groupes connus comme étant électroactifs (TTF--fluorene, nEDOT-3CNQ, nEDOT-

NDI). Nous décrivons les progrès que nous avons apportés au domaine complexe de

redresseurs moléculaires par l’entremise du couplage de puissant donneurs et accepteurs.

Pour réussir cela, nous avons employé différentes stratégies dont: l'utilisation

d'intermédiaires avec des propriétés oxido-réductives modérées et des dérivés très

réactifs pour empêcher la formation de complexes à transfert de charge ainsi que de se

servir de ces mêmes complexe pour obtenir des liaisons covalentes. La synthèse utilisée

explore deux approches qui permettent la liaison des dyades à la surface de l'électrode:

(1) l'utilisation de structures amphiphiles permettant la déposition de monocouches

moléculaires par la technique Langmuir-Blodgett et (2) l'utilisation de groupes de thiol

et dedisulfure permettant la liaison covalente des molécules avec des métaux. La

caractérisation de ces monocouches a été fait à l’aide de techniques spectroscopiques et

électrochimiques.L'analyse de la densité, l'ordre des molécules dans les films et leur

stabilité a aussi été étudié. Finalement, nous décrivons la fabrication de jonctions

électrode/monocouche organique/électrode et nous discutons les résultats des mesures de

transport de charges des dyades donneur-accepteur synthétises.

v

Acknowledgments

I heartily thank my supervisor, Prof. D. Perepichka, without who's

encouragement, guidance and support from the very beginning to the final editing

stages, this theses would not be possible. I would also like to thank him for his help in

fostering my understanding of the subject and his patience with me as I developed my

organic synthesis skills, through the characterization of the monolayers and a great many

experiments.

During my study I also have been lucky to work with some remarkable people at

Laval University, INRS-EMT and of course, in D. Perepichka’s group. My special

thanks to Karin Arseneault (group of M. Pezolet, Laval University) who introduced me

to Langmuir-Blodgett technique. I am also grateful to Jacky Brusso, Matt Morantz, Julia

Schneider, Andrey Moiseev and Afshin Dadvand for many helpful discussions and

proof-reading my manuscripts.

I would also like to thank the many people who have helped me during my stay

at McGill University. First of all, I’m grateful to Chantal Marotte for her guidance

during my graduate studies. I would like to recognize the inestimable help of Fred Morin

and Nadim Saade with the characterization of new materials; Fred Kluck, Jean-Philippe

Guay, Richard Rossi and Weihua Wang for their help in building, breaking, and fixing

the tools that made my research easier.

Finally I would like to thank my family for their support, understanding and

encouragement through all these years.

vi

Table of contents

Abstract ............................................................................................................................ ii

Résumé ............................................................................................................................ iv

Acknowledgments ........................................................................................................... v

List of abbreviations ...................................................................................................... ix

Introduction .................................................................................................................. - 1 -

Motivations and objectives ........................................................................................ - 3 -

Outline of the thesis .................................................................................................... - 4 -

Chapter I: Overview: Unimolecular organic rectifiers .......................................... - 6 -

1.1 Aviram-Ratner concept ................................................................................ - 6 -

1.2 Experimental works in the field of unimolecular rectifiers ............................. - 13 -

1.2.1 Donor--Acceptor molecules .................................................................... - 13 -

1.2.2 Donor-π-Acceptor molecules .................................................................... - 17 -

1.2.3. Non Donor-Acceptor rectifying systems ................................................. - 21 -

1.2.4 Synthetic strategies for donor-acceptor dyads .......................................... - 28 -

1.3 Molecular assemblies ....................................................................................... - 31 -

1.3.1 Langmuir-Blodgett deposition .................................................................. - 32 -

1.3.2 Self-assembly by chemisorption ............................................................... - 36 -

1.4 Characterization of organic monolayers .......................................................... - 39 -

1.4.1 Ellipsometry .............................................................................................. - 39 -

1.4.2 Contact angle ............................................................................................. - 40 -

1.4.3 Reflection-absorption and ATR infrared spectroscopy ............................. - 41 -

1.4.5 Electrochemical characterization of SAMs ............................................... - 45 -

1.5. Fabrication of molecular junctions ................................................................. - 47 -

1.5.1 Vacuum deposition of the metal on top of the organic layer .................... - 48 -

1.5.2 Liquid metals drop junctions ..................................................................... - 49 -

1.5.4 Electron Transfer in Metal-Molecule-Metal Junctions ............................. - 51 -

Conclusions ............................................................................................................ - 54 -

Chapter II. The first studies of a tetrathiafulvalene-σ-acceptor molecular

rectifiers .................................................................................................................... - 55 -

Introduction ............................................................................................................ - 55 -

vii

2.1. ―Amphiphilic design‖ ..................................................................................... - 57 -

2.2. Synthesis of TTF--fluorene dyads ................................................................ - 57 -

2.3. Geometry and electronic structure of the dyad 2.4 ......................................... - 59 -

2.4. Electrochemical characterization of the dyads in bulk ................................... - 61 -

2.5. Preparation of monolayer of 2.4 on air-water interface .................................. - 63 -

2.6. Deposition of the monolayers on solid substrates ........................................... - 65 -

2.7. Spectroscopic characterization of LB monolayers.......................................... - 66 -

2.8. Fabrication and electrical studies of n-Si/SiO2/2.4/Ti junction devices ......... - 71 -

2.9. Fabrication and electrical studies of Au/2.4/Hg junction devices .................. - 76 -

Conclusions ............................................................................................................ - 79 -

Experimental section .............................................................................................. - 80 -

Chapter III. Self-Assembled Monolayers of Strong Electron Acceptors:

Polynitrofluorenes on Gold ..................................................................................... - 83 -

Introduction ............................................................................................................ - 83 -

3.1 Synthesis .......................................................................................................... - 84 -

3.2 Formation of SAMs of the fluorene derivatives 3.4 and 3.5............................ - 87 -

3.3. Electrochemical and spectroscopic characterizations of 3.4 and 3.5 in solution- 87

-

3.4 Electrochemistry of SAMs ............................................................................... - 91 -

3.5 Reflectance-absorbance infrared spectroscopy (RAIRS) of SAMs ................. - 96 -

3.6 Ellipsometry and contact angle measurements ................................................ - 98 -

3.7. Rectification study of dyad 3.5 ....................................................................... - 99 -

Conclusions .......................................................................................................... - 103 -

Experimental Part ................................................................................................. - 104 -

Chapter IV. Synthesis and characterization of TTF--nitrofluorene dyads for self-

assembly on gold surface. ...................................................................................... - 107 -

Introduction .......................................................................................................... - 107 -

4.1. Synthesis ....................................................................................................... - 107 -

4.2. Characterization ............................................................................................ - 117 -

Conclusions .......................................................................................................... - 123 -

Experimental Part ................................................................................................. - 124 -

viii

Chapter V. Molecular rectification of hexyl-nEDOT-3CNQ dyads in Langmuir-

Blodgett film ........................................................................................................... - 130 -

Introduction .......................................................................................................... - 130 -

5.1. Synthesis ....................................................................................................... - 131 -

5.2. DFT Calculations .......................................................................................... - 136 -

5.3. Characterization ............................................................................................ - 138 -

5.4. Langmuir-Blodgett deposition of the monolayer of 5.12 on the solid

substrates .............................................................................................................. - 143 -

5.5. Rectification measurements of the LB films of the dyad 5.12 in mercury-drop

junctions ............................................................................................................... - 146 -

Conclusions .......................................................................................................... - 152 -

Experimental part ................................................................................................. - 153 -

Chapter VI. Stable nEDOT-NDI molecular rectifiers with self-assembly

capability. ................................................................................................................ - 156 -

Introduction .......................................................................................................... - 156 -

6.1. Synthesis of nEDOT-NDI dyads ................................................................... - 158 -

6.2. Calculations ................................................................................................... - 160 -

6.3. Absorption/Emission spectra ........................................................................ - 161 -

6.4. Electrochemistry ........................................................................................... - 167 -

6.5. SAM preparation and characterization ......................................................... - 172 -

6.6. Preliminary rectification study of the bis-EDOT-NDI dyads ....................... - 177 -

6.7. Potential for improvement of the acceptor properties ................................... - 181 -

Experimental part ................................................................................................. - 184 -

Conclusions ............................................................................................................. - 189 -

References ............................................................................................................... - 192 -

Appendix ................................................................................................................. - 229 -

Author’s contribution ............................................................................................ - 235 -

List of publications ................................................................................................. - 237 -

ix

List of abbreviations

A Acceptor

Ac Acetyl

AFM Atomic force microscope

AR Aviram-Ratner

ATR Attenuated total reflection

CTC Charge-transfer complex

CV Cyclic Voltammetry

D Donor

DCC N,N'-dicyclohehylcarbodiimide

DMAP 4-Dimethylaminopyridine

DTeF 9-dicyanomethylene-2,4,5,7-tetranitrofluorene

DFT Density functional theory

EDOT 3,4-Ethylenedioxythiophene

Et ethyl

FT-IR Fourier transform infrared spectroscopy

Fc Ferrocene

HOMO Highest occupied molecular orbital

HLG HOMO-LUMO gap

LB Langmuir-Blodgett

LUMO Lowest unoccupied molecular orbital

Me methyl

NDI 1,4,5,8-naphthalenetetracarboxylic diimide

NMR Nuclear magnetic resonance spectroscopy

Ph phenyl

PDI 3,4,9,10-Perylenetetracarboxylic diimide

x

PTCDA 3,4,9,10-Perylenetetracarboxylic dianhyride

RR Rectification ratio

SAM Self-assembled monolayer

STM Scanning tunneling microscopy

STS Scanning tunneling spectroscopy

TCNQ Tetracyanoquinodimethane

TLC Thin-layer chromatography

TNF 2,4,5,7-tetranitrofluorene-9-one

TTF Tetrathiafulvalene

UV Ultra-violet spectroscopy

- 1 -

Introduction

Molecular electronics can be defined as a field of research that studies electrical

processes in a single or at least in a limited number of molecules. This also involves the

study of a wide number of different molecular assemblies of any scale and organization

and the application of organic and biological molecules in electronic devices.

Richard Feynman was the first scientist to point out on the perspective future of

molecular-scale systems in his famous speech, ―There is Plenty of Room at the Bottom‖

in the 1950s. He brought attention to the point that physical laws do not limit the ability

to manipulate and study single molecules and even atoms. He correctly noted simply the

lack of instrumentation for doing so and predicted that in a near future it would be

possible to perform atomically precise manipulations [1].

Today, molecular electronics is an important multidisciplinary field in the

fundamental theoretical research of the physical and electrochemical properties of

organic materials and the application of such materials in novel electronic devices [2-4].

One of the reasons that molecular electronics has attracted so much attention was the

hope that single molecules could possibly become an alternative for the present silicon-

based integrated circuits. In 1965, Gordon Moore from Intel has quantitatively described

the trend of the computer’s power growth By making an observation, made in 1965, was

that the number of transistors per unit area on integrated circuits, or functionality per

chip, doubles every 2-3 years since the integrated circuit was invented [5]. He also

predicted that this trend would continue in the foreseeable future. This is achieved by

reducing the size of devices. Over the past decades, silicon based devices has continually

been scaled down in size. In late 2009, Intel began production of a process featuring a

32 nm feature size. But this minimization cannot go on forever and eventually

technology will face hard technical difficulties. The organic molecules, with size 1-3 nm

can possibly do similar tasks that current silicon-based devices are doing. Molecular

electronics involves a bottom-up technology that uses atoms to build nanometre-sized

molecules that could further self-assemble into a desired computational circuitry. This

bottom-up approach gives rise to the prospect of manufacturing electronic circuits in

rapid, cost-efficient, flow-through processes.

- 2 -

Two main focuses in the field of molecular electronics are: design and synthesis

of the molecular-scale systems with tailored electronic properties and the study of such

systems as electronic devices for processing electrical, optoelectronic and other signals.

Bulk organic materials are already widely utilized in thin-film electronics and successful

application of such materials is a rather developed field. Organic light-emitting diodes

[6-7], organic field effect transistors [8-12] and photovoltaic cells [13-14] are already on

the market.

There are, of course, a number of challenges related to the use of the single

organic molecules as electronic device.

Organic molecules are easier to synthesize in large quantities then it is to

manufacture the Si based devices, but they are more difficult to arrange on a surface or

in a three-dimensional array using existing technologies (e.g. Photolithography). It is

also difficult to ensure that the molecules stay in place.

The stability and life-time of the organic molecules can be an issue as

well. The heat dissipation becomes very important consideration for the electronic

devices with a million-fold increase in circuit density. Thus, extremely efficient cooling

systems would be needed to prevent decomposition of the molecule and damage of the

device.

To address these challenges the development of new technology together with

scientific understanding of the processes in single molecules could make progress

towards molecular electronics possible. Nowadays, synthesis of organic molecules is a

highly developed tool and by choosing different compositions and geometries it is easy

to vary a molecule’s charge transport, binding, electrical, and structural properties. The

size scale of organic molecules lies between 1 and 100 nm, a scale that permits

functional nanostructures with accompanying advantages in cost, efficiency, and power

dissipation. The advantages of specific intermolecular interactions allow formation of

nanoscale structures by self-assembly.

- 3 -

Motivations and objectives

In 1974 Ari Aviram and Mark Ratner proposed a theoretical concept of

unimolecular rectification and this work brought an attractive idea for development

systems that can potentially compete with today’s electronic devices. Although

numerous experimental strategies to achieve rectification in donor–acceptor molecules

have been attempted and various molecular structures, electrodes, and junction assembly

approaches have been tested, the precise mechanism for the rectification in molecular

junctions is still a subject of controversy.

The main objective of this thesis is molecular design and experimental study of

the unimolecular rectification of series of donor-acceptor molecules. For this purpose a

set of three donors and three acceptors were chosen to test testing the mechanism

proposed in the original Aviram-Ratner (AR) concept. Two approaches of depositing the

molecules on the electrode surface were employed in this work: Langmuir-Blodgett

deposition of the amphiphilic molecular structures and self-assembly of the thiol-

functionalized molecules. Different methods of assembly allow comparison of electrical

properties of molecules with different types of metal-molecule contacts. The electrode

surface, modified with the electroactive molecules were studied by different

spectroscopic and electrochemical methods was conducted during the course of this

work in order to get complete information about the chemical composition and structure

of the monolayers. Electron transfer through the monolayers of the dyads was studied in

metal-molecule-metal junctions made by thermal evaporation of the second electrode on

top of the organic film and by mercury drop technique. Comparison of the results may

illuminate details of the electron transport mechanisms of the single molecules.

- 4 -

Outline of the thesis

In the first Chapter we present an introduction to the field of molecular

rectification where we discuss general concepts, previously published experimental

results and current progress in the field. Also we present an attempt to systematize

important ―tools‖ and challenges in this field (synthesis, assembling of the molecules on

the electrode surface, characterization of the monolayers and fabrication of the

molecular junctions).

Chapter II discusses synthesis, characterization and rectification study of TTF--

nitrofluorene based dyad with amphiphilic structure. The charge transport of this dyad

was studied in LB films sandwiched between two metal electrodes. Confirmation of

molecular origin of rectification behaviour of this molecule in metal-dyad-metal

junctions was also presented.

In the Chapter III we discuss self-assembly of electroactive molecules on the

electrode surface via chemical grafting of organic molecule on metal. Therein we

present the synthesis and study of monolayer self-assembly study of the poly-

nitrofluorene acceptor on metal surfaces.

Chapter IV focuses on the synthesis of the TTF--nitrofluorene based dyads with

self-assembly functionality and discusses in details the challenges associated with it.

There we present design, synthesis and characterization of the series of new TTF--

nitrofluorene based dyads with different thiol terminated functional groups.

In the Chapter V we present different design of donor-acceptor dyad based on

nEDOT-3CNQ. Synthesis and characterization of the molecules in bulk and as LB

monolayers on the solid substrate as well as rectification study of LB films are

discussed.

Finally, in the Chapter VI we discuss synthesis, characterization and rectification

properties of series another nEDOT containing donor-acceptor dyads with NDI as and

acceptor moiety.

In Conclusion section we summarize experimental results obtained during the

course of this work.

- 5 -

Appendix includes preliminary theoretical and experimental results towards

improving the acceptor properties in the Donor-Acceptor dyads described in the Chapter

VI.

- 6 -

Chapter I: Overview: Unimolecular organic rectifiers

A rectifier is an electrical device that converts alternating current (AC) to direct

current (DC) by allowing the current to flow only in one direction. The process is called

rectification and is the main function of conventional diodes. Diodes were the first

semiconductor electronic devices and are currently the key components of integrated

circuits along with transistors, resistors, capacitors and other electronic components. A

conventional diode is made of a crystal of semiconducting material that has impurities

added to it to create a region on one side that contains negative charge carriers

(electrons), called an n-type semiconductor, and a region on the other side that contains

positive charge carriers (holes), called a p-type semiconductor. Today most diodes are

made of silicon but sometimes other semiconductors such as germanium are used,

however, this is not an exhaustive description of present types of diodes [15].

The main parameter that characterizes performance of diode-like devices is the

rectification ratio (RR), which is obtained from the ratio of the current at equal voltages

of opposite sign:

|

| (1.1)

The RR of silicon-based diodes is usually in the range of a few hundred [16-17].

1.1 Aviram-Ratner concept

For decades researchers have been studying electron transport within large

molecules. In the 1950s, Henry Taube proved that electron transfer across an organic

bridge between two dissimilar metal ions occurs more slowly across aliphatic bridges

than across conjugated aromatic bridges [18]. This launched extensive studies of

intramolecular electron transfer in molecules in solution by fluorescence and time-decay

spectroscopy. Theoretical understanding of electron transfer was developed by Rudolph

A. Marcus, Noel S. Hush, and others [19-21]. ―Marcus theory‖ explains the rates of

- 7 -

electron transfer reactions, – a process by which an electron can move by ―hopping‖

from one chemical species (called the electron donor) to another (called the electron

acceptor). The basic equation of Marcus theory is built on the classical Arrhenius

equation and is expressed as:

(

) (1.2)

where λ is the reorganization energy, Go is the total Gibbs free energy change for

electron transfer between A and B, and kb is the Boltzmann constant.

According to this equation, the electron-transfer rate will increase with increasing

the driving force (-∆G°) of the reaction (―normal case‖) up to a maximum when -∆G° =

λ (―optimal case‖), but decreases when -∆G° exceeds λ (―inverted case‖). Experimental

proof of this inverted region was obtained by Gerhard Closs, John R. Miller, and co-

workers by measuring the rate of electron transfer for a series of D--A dyads [22-24].

Marcus theory was used to describe a number of important processes in

chemistry and biology, including photosynthesis, corrosion, certain types of

chemiluminescence, charge separation in some types of solar cells and more.

The electron-transfer theory could be employed as a basis for understanding the

functioning of electronic devices of molecular size. Particularly, the one-way electron

transport within organic molecules is a subject of current interest in the field of

unimolecular diodes. The idea of unimolecular rectifiers was first proposed in 1974 by

Ari Aviram and Mark Ratner [25]. This publication was the first theoretical proposal

that started the field of molecular electronics. Aviram-Ratner proposed diode-like

behaviour from the junction based on a single D--A molecule (1.1) composed of

tetrathiafulvalene (TTF) and tetracyanoquinodimethane (TCNQ) – a good electron donor

[26] and acceptor [27], respectively – connected through an insulating saturated bridge

(-bridge). Electrical conduction within molecule 1.1 would be favoured from the

electron-poor moiety to the electron-rich moiety, but disfavoured in the reverse

direction. The purpose of the -bridge is to isolate the HOMO from the LUMO which

- 8 -

are mainly localized on the donor and acceptor moieties, respectively. Thus, the bridge

physically prevents the molecular orbitals of the donor and acceptor from mixing and

provides asymmetry of the electronic structure of the dyad.

The rectifying behaviour of the D-A dyad can be easier understood from the

energy band diagram (Fig. 1.1). Figure 1.1 (A) presents the molecular energy levels of

an ideal rectifying molecule, when it is placed between two electrodes with no external

bias applied. The LUMO of the acceptor should lies close to the Fermi energy of the

cathode (or slightly above) whereas the LUMO of the donor lies as high as possible

above the Fermi energy of the anode. As well, the HOMO of the donor should lie close

to the Fermi energy of the anode. For the current to flow, electrons must tunnel through

the potential barriers between the molecule and electrodes as well as through the -

bridge.

Application of external bias to the junction leads to overlap of the work functions

of the electrodes with the molecular orbitals of the donor and acceptor moieties. Under

forward bias (Fig. 1.1 (B)), an electron travels from the Fermi level of the cathode to the

empty LUMO of the acceptor. A similar process takes place on the other side of the

junction – an electron is transferred from the occupied HOMO of the donor to the anode.

Injection of electrons and holes into TTF and TCNQ from electrodes will create the first

exited state followed by intramolecular electron transfer to form the ground state via

inelastic tunneling through the –bridge. The efficiency of the tunnelling increases when

HOMO and LUMO orbitals are close enough energetically (i.e., small HOMO-LUMO

gap). This electron transfer is irreversible by nature of the molecule’s energy levels, as

- 9 -

the LUMO is higher in energy than the HOMO. Now, if we apply a reverse bias to the

same junction (Fig. 1.2), a much higher voltage is required to bring the Fermi level of

the cathode in resonance with the LUMO of the donor and the Fermi level of the anode

in resonance with the HOMO of the acceptor [21]. Comparing diagram B in Fig 1.1 with

diagram C in Fig 1.2 we can see that simply applying the same voltage in the reverse

direction cannot induce the resonance between the anode and the LUMO energy level of

the donor moiety. This is a classical interpretation of the behaviour of molecular diodes.

This electron transfer in the donor-acceptor molecule is well explained by

Marcus theory [19]. The rate of the electron transfer will increase as the energy

difference between the HOMO of donor and LUMO of acceptor is increasing, but only

until a certain point. Past that point, the electron transfer rate will actually decrease as

the energy difference continues to increase – this is called the ―Marcus inverted region".

The electron transfer at reverse bias and through large HOMO-LUMO gap is actually in

this inverted region, where the difference between LUMO of the donor and HOMO of

the acceptor is large.

Since molecule 1.1 has never been synthesised, there is no experimental evidence

for the rectification behaviour of this specific molecule. However, theoretical

calculations (INDO semi-empirical method) of the I-V characteristics for the 1.1

(EHOMO= 5.3 eV, ELUMO=5.0 eV) placed between two electrodes (=5.1 eV) [25] (Fig.

1.3) predict its rectification behaviour. The threshold potential for current passage,

determined as the voltage necessary for an overlap of the HOMO (D) and LUMO (A)

with corresponding Fermi levels of electrodes, is around 0.3 V. If the reverse bias

voltage is high enough (0.55 V for the system presented in [25]) to allow the overlap of

the LUMO of donor and HOMO of acceptor moieties with respective electrode, than the

current begins to flow in opposite direction.

It is important to note that the original concept also allows a second, potentially

possible, competing process called ―autoionization‖. In this case, the first step involves

internal tunneling of the electron from the HOMO of the donor onto the LUMO of the

acceptor resulting in formation of an exited state (i.e., a ―zwitterion‖). The electron

transfer then occurs in a second step across the molecule–electrode interface [25, 28].

According to this mechanism the electron would be transferred in the opposite direction,

- 10 -

Figure 1.1: The energy band diagrams of the Aviram-Ratner concept of molecular

rectifiers. The D--A molecule is placed between two electrodes with A) no external

bias and B) a small forward bias.

- 11 -

Figure 1.2: The energy band diagrams of Aviram-Ratner concept of molecular rectifier

at reverse bias.

Figure 1.3: Calculated I-V characteristics of the Aviram-Ratner molecular rectifier.

Copied from [25].

- 12 -

from the electrode close to the donor onto the electrode close to the acceptor. However,

this alternative mechanism also involves a threshold voltage necessary to bring the

HOMO of the donor and the LUMO of the acceptor to the same energy levels before

internal tunnelling may occur [25]. Thus, a molecular rectifier must be designed with the

knowledge that the pathway for which the threshold voltage is smallest will define the

preferential direction of conductance.

Herein we briefly summarize most important issues in studying unimolecular

rectification.

(1) Synthesis of the rectifying dyads requires coupling together a strong

oxidizing agent (acceptor) with a strong reducing agent (donor) in order to achieve small

HOMO-LUMO gap. This is not a trivial task as interaction between these species may

result in formation of a charge transfer complex rather than covalent bond formation.

The linker between two electroactive moieties should provide separation of their

molecular orbitals. The length of the spacer also controls the efficiency of the electron

tunnelling process during device operation. Coupling a strong acceptor and a strong

donor with a somewhat short bridge might be not enough separation for the HOMO and

LUMO energy levels. From the other hand, a too long and flexible bridge can give

additional conformational freedom and may result in creation of a horseshoe-shaped

molecule with intramolecular donor-acceptor overlap. Such molecule may not function

as a rectifier.

(2) Special consideration should also be given to the analysis and testing of the

rectifiers. The small size of the organic molecules was presented earlier as an advantage

of molecular scale electronics; however, manipulation and investigation of nano-meter

scale systems reproducibly and reliably remain challenging tasks. Using molecular

monolayers facilitates fabrication of electronic devices where the electron transport still

occurs through (and is controlled by) a single molecule. Different techniques for

assembling the molecules on the surface were already developed. Detailed

characterization of the molecular assemblies, including orientation of the molecules

within the films as well as actual formation of the molecular junctions which requires

deposition of the metal contact on top of the organic film or positioning the molecule

- 13 -

between two metal electrodes are important aspects for understanding structure-property

relations, electronic and charge transport behaviour of molecular systems.

(3) The nature and properties of the molecule-metal interface plays an important

role in the electron transfer process through the junction and can in many cases dominate

the overall device performance. Thus, one should also consider the electronic influence

of functional groups used for assembling the dyad molecules between metal electrodes.

Whether physisorption or chemisorption methods is selected, proper design of the donor

and acceptor synthons should be chosen prior to coupling them together as it becomes

extremely difficult to make any changes with the dyad afterwards. Additionally, choice

of the metal contacts should not be neglected since matching the HOMO and LUMO of

the molecules with Fermi energies of corresponding metal electrodes is important for

correct functioning of the system.

1.2 Experimental works in the field of unimolecular rectifiers

Since the original Aviram-Ratner proposal more than 30 years ago, numerous

attempts to synthesize and investigate molecular rectifiers based on single molecules

have been undertaken. However, only a few D–σ–A molecules have been shown to

rectify current. Many other studies have been focused on D–π–A dyads as potential

candidates for Aviram-Ratner rectifiers. As long as the π–bridge serves as an effective

barrier to prevent orbitals mixing, such systems may be an alternative model in the field

of unimolecular rectification. Herein we will summarize previous experimental work on

molecular rectifiers that was reported in the literature since the original Aviram-Ratner

publication [25].

1.2.1 Donor--Acceptor molecules

The first attempt of experimental study of rectification behaviour of a D-A dyad

in monolayer was reported by Aviram and co-workers in 1988 [29]. Using a scanning

tunneling microscopy (STM) tip as an electrode (Fig. 1.4), they observed asymmetric I-

- 14 -

V characteristics of deposited on gold surface molecule 1.2, which comprise catechol

and o-quinone rings.

Figure 1.4. The first experimental attempt to study a molecular rectifier. Adapted from

[29]

It was proposed that at negative bias (–0.2V) electrons flowed from the tip to the

quinone (acceptor) and from the catechol (donor) to the gold surface. However, it was

concluded that the rectifying behaviour may have been due to proton transfer from the

catechol to the quinone. This process produces a semiquinone structure, which would be

a conductor and so results in enhanced current flow through molecule. In any case, soon

after publishing these results were retracted [30]

Sambles et al [31] have studied the rectification properties of donor- acceptor

molecule 1.3, constructed of strong acceptor (TCNQ) and a weak donor (alkoxybenzene)

They observed rectification behaviour from LB films of a the 1.3, with highly

asymmetric I-V curves with the preferential current flow from donor to acceptor [31],

which is opposite to the direction proposed by Aviram-Ratner model. Further

investigation of this molecule revealed that a Schottky barrier due to magnesium oxide

layer created as a result of breaking the vacuum after evaporation of the Mg electrode

was the reason of the current asymmetry [32].

- 15 -

Recent reinvestigation of the I-V characteristics of the same molecule 1.3 by

scanning tunneling spectroscopy (STS) technique [33] showed that in contrast to

previous results, LB films on a gold substrate showed expected current rectification in

direction from the substrate to TCNQ acceptor and then to dodecyloxyphenyl donor.

Mikayama et al. [34] studied molecular rectification in D--A dyad based on a

dinitrobenzene acceptor and dihydrophenazine donor (1.4). The authors observed

rectification behaviour of the LB film of this dyad on the gold surface by STS

measurements. The rectification ratio was found to be ~7 at ±1 V under illumination and

lower in the dark, revealing the characteristics of a photodiode. The direction of the

rectification was from the acceptor to donor moiety, in accordance with Aviram-Ratner

concept.

Theoretical and experimental studies of D-σ-A molecule (1.5) constructed from

pyrene (a moderate donor) and dinitrobenzene (a weak acceptor), were performed by

Sambles and co-workers [35]. The junction was made by transferring the molecule as a

LB multilayer onto a silver electrode and contacted with magnesium pads evaporated on

top of the organic film. A five-layer film of 1.5 showed diode-like behaviour with a

rectification ratio RR ~130 at 2.5 V. The direction of the preferential current flow, in this

case, is from Mg to pyrene (D) and from dinitrobenzene (A) to Ag, i.e. opposite to that

predicted by the AR model. It was concluded that the electrical conductivity of the film

of the 1.5 involves both inter- and intramolecular charge-transfer. Application of the

external electrical field results in variations in the HOMO-LUMO gap of the molecule.

- 16 -

This lead to the increase of the accessibility of the exited state (D+--A

-) of the molecule

and increase of the probability of the resonant tunneling between A- and silver electrode,

which results in asymmetry in the current flow through the molecules.

Unimolecular rectification was also observed for bulky, fullerene based,

molecule 1.6 [36]. Interestingly, this donor-acceptor dyad can form Langmuir films on

the water surface despite the luck of hydrophobic hydrocarbon chains. Such structure

allows symmetric positioning of the D-A dyad between electrodes, which is important to

avoid artefacts in rectification behaviour [37]. The rectification of this molecule was

measured in the Au/Langmuir-Shaefer monolayer/Au junctions, where the top gold

electrode was evaporated on top of the organic layer. The junctions show current

asymmetry (RR ~16) in the direction from acceptor to donor (in agreement with AR

model) and the rectification ratio does not decay after several cycles (in contrast to other

literature results for LB films [38-39]).

1.6

A series of PDI-based rectifiers 1.7 were synthesized by Wescott and Mattern

[40] and studied in the Metzger group [41]. LB monolayers of 1.7b and 1.7c sandwiched

between two gold electrodes showed only weak asymmetry of the current flow with RR

~2 and 1.5 respectively. However, significantly larger RR (14 – 28 at ±1 V) which did

- 17 -

not decrease after 40 cycles was observed for monolayers of 1.7a. This was explained by

theoretical calculations which showed that for the dyads 1.7b and 1.7c energy of the

HOMO orbital is significantly below the work function of the electrode and cannot

participate in the electron transfer. Whereas for the 1.7a both HOMO and LUMO are

energetically close to the work function of the electrodes and show preferential charge

transport in the acceptor-to-donor direction.

1.2.2 Donor-π-Acceptor molecules

A family of zwitterionic molecules including C16H33Q-3CNQ (1.8) were studied

in 90’s by the group of Ashwell [42-43] and then continued by Metzger [38-39, 44-48]

confirming the unimolecular rectification of the molecule 1.8 and clarifying its

mechanism.

- 18 -

Measurements of the static dipole moment in the CH2Cl2 solution of 1.8 revealed

that this molecule has a clear zwitterionic ground state (D+-π-A

-) with dipole moments

43±8 D and a neutral (D0-π-A

0) first excited state with dipole moments 3-9 D. These

results were also supported by theoretical calculations, NMR, UV-Vis and IR data [44,

49]. The important difference of molecule 1.8 from the Aviram-Ratner model is a

conjugated π-bridge. However, the twist angle about 30°, caused by steric hindrance

between donor and acceptor moieties, provides some separation of the donor and

acceptor orbitals. This molecule forms monolayers at the air-water interface [44]. The

highly polar CN groups on the acceptor and the ―fatty‖ tail of the donor moiety facilitate

the upstroke transfer of LB film on the hydrophilic substrate as: glass, gold, aluminum

surfaces, with transfer ratio close to 100%.

The rectification measurements were accomplished by ―sandwiching‖ the LB

film between two metal electrodes. Several different junctions where used to study

electronic properties of this molecule: Al/LB film/evaporated Al [50] (RR=26 at ±1.5V),

Au/LB film/evaporated Au [38, 47] (RR=27.5 at ±2.2 V), Al/LB multilayer/Mg [43],

graphite/LB film or multilayer/STM tip [44, 51] (RR=20 at ±1 V) and Au/LB

multilayer/Au [52]. Upon multiple scanning, the rectification ratio gradually decreases

after every scan. The reason behind such behaviour is the very large electrical field

applied across the monolayer. Under such field, dipolar molecules can flip over to

minimize the total energy [44].

In continuations of this project, the group of Prof. Ashwell used scanning probe

microscopy for investigation of self-assembled analogue of 1.8 covalently attached

molecules 1.9a and 1.9b to the gold surface (Fig. 1.5) [53] Due to the methyl substituent

the dyad 1.9a has a significant twist angle between donor and acceptor planes, while

dyad 1.9b is planar. Rectification measurements of the SAMs with a STM gold tip

showed a diode-like behaviour for the molecule 1.9a (RR is ~11 at 1 V) and no current

rectification for the dyad 1.9b. Significantly smaller out-of-plane rotation in the 1.9b,

compared to the 1.9a suggests that better conjugation in the former that leads to

delocalization of both HOMO and LUMO throughout the entire molecule, is detrimental

for the current rectification [54].

- 19 -

Figure 1.5: Au-S-CnH2n-Q3CNQ assemblies studied by Ashwell [55].

Two chevron-shaped molecules 1.10a and 1.10b were studied by Ashwell et al.

[56]. The molecules have a central cationic acceptor and two π-bridged donor groups

with an angle of ca. 120 between the charge-transfer axes of the chevron-shaped D-π-

(A+)-π-D unit. These molecules form stable LB films and, when placed between two

gold electrodes, exhibited rectifying behaviour with rectification ratio up to 90 at ±1 V.

The current asymmetry is enhanced at a forward bias of 0.5–1.0 V, as electrons flow

preferentially from the iodide ion to the pyridinium ion. It was suggested that such high

asymmetry of the I-V characteristics must be due to interionic rather than intramolecular

electron transfer [39].

A conjugated diblock co-oligomer system 1.11 consisting of two blocks with

opposite electronic demand was reported to behave as a molecular diode (Fig. 1.6). The

molecule consists of an electron-rich bithiophene segment and an electron-poor

bithiazole segment, which are efficient hole- and electron-transporting agents,

- 20 -

respectively. The geometry of the molecule suggests presence of a large dihedral angle

between two blocks of the oligomer due to the steric hindrance caused by methyl groups.

This provides separation of the molecular orbitals of the donor and acceptor moieties.

Electrical measurements, performed by scanning tunneling spectroscopy (STS) for

SAMs on gold surface, clearly reveal a moderate current rectification effect (RR ~6).

The proof for the molecular nature of the rectifying effect in this conjugated diblock

molecule was provided by control experiments with a structurally similar reference

oligomer, tetrathiophene, with no asymmetric charge polarization [57-58].

Figure 1.6: Illustration of possible orientation of di-block oligomers 1.11 attached on

the gold surface. Adapted from [57].

A current rectification was also shown for other non-symmetric diblock

dipyrimidinyldiphenyl molecule 1.12 (Fig. 1.7) by group of NJ Tao [59]. The molecule

with two thiol-based end groups was placed between two gold electrodes (a gold

substrate for self-assembly and gold coated STM tip). Important part of the work is that

the orientation of the molecule within the junction was controlled by selective

deprotection of the each thiol terminal group.

The rectification ratio for these junctions is 5 at forward bias (the current

preferentially flows from dipyrimidinyl to diphenyl moiety). The molecular origin of the

rectification was proved in the control experiments with symmetric tetraphenyl molecule

that showed symmetric I-V characteristics.

- 21 -

Figure 1.7: Non-symmetric dipyrimidinyl–diphenyl molecule 1.12 and its symmetric

equivalent. Adapted from [59].

1.2.3. Non Donor-Acceptor rectifying systems

Interest in molecular rectification properties is not limited only to the Aviram-

Ratner model. Some relatively recent work described rectification properties of

molecular junctions with no D-bridge-A structure. Particularly, it was suggested [60]

that rectification of the C16H33Q-3CNQ molecule is due to the asymmetric position of

the HOMO and LUMO with respect to the Fermi levels of the electrodes. The groups of

Whitesides and Rampi showed current rectification behaviour of disulfide-terminated

acceptor (TCNQC10S)2 (1.13). The SAMs of this molecule were deposited on the gold

- 22 -

or silver electrode and sandwiched between SAM of the alkanethiol of a mercury

electrode [61].

The I-V curves' asymmetry indicate the preferential currents flow at forward bias

(from gold/silver electrode onto the mercury electrode) with rectification ratio RR= 9±2

at 1 V. The rectification in this molecule, lacking an obvious D-A structure, was

attributed to the asymmetric position of the redox center in the metal/insulator/metal

junction.

Whitesides et al [62] have recently published a systematic study of the

conductivity of junctions with Ag bottom electrodes and liquid metal (Ga2O3/EGaIna)

top electrodes, based on SAMs with an electrically ―conductive‖ ferrocene (Fc) moiety

1.14 and insulating alkyl moiety (Figure 1.8) varying the proximity of the redox centre

to each electrode. It was shown that junctions with the Fc moiety placed symmetrically

(Fig. 1.8d) between the electrodes (by varying the length of the insulating alkyl section)

did not rectify, however, rectification was observed in the junctions where Fc moiety is

placed closer to one electrode (Fig. 1.8a, e and f).

a Eutectic indium-gallium alloy

- 23 -

Figure 1.8: Schematic representations of the tunneling junctions described in [62]

- 24 -

Table 1.1: Summary data to discussed molecular rectifiers.

Structure Junction RR Rectification direction Ref.

Donor--Acceptor

1.2

Au/SAM/STM tip <2 - [29]

1.3

Pt/LB/evaporate Mg

Au/LB/STM tip

5 DA (Schottky barrier due to MgO)

AD

[31-

33]

1.4

Au/LB/STM tip 7 AD (photoconductor) [34]

1.5

Ag/LB/evaporated Mg 130 DA [35]

- 25 -

1.6

Au/LB/evaporated Au 16

DA (Asymmetric rectifier)

[36]

1.7a

1.7b

1.7c

Au/LB/ evaporated Au

14-28

2

1.5

DA [40-

41]

Donor-π-Acceptor

1.8

Al/LB/evaporated Al [50],

Au/LB/evaporated Au [38, 47],

Al/LB/evaporated Mg [43],

graphite/LB/STM tip [44, 51]

Au/LB multilayer/Au [52].

26

27.5

20

zwitterionic ground state (D+-π-

A-) with dipole moments 43±8D

D+A

-

- 26 -

1.9a

1.9b

Au/SAM/STM tip 12

1

AD

[53-

54]

1.10a

1.10b

Au/LB/Au 90 Interionic charge-transfer [56]

1.11

Au/SAM/STM tip 6 DA [57-

58]

- 27 -

1.12

Au/SAM/STM tip 5 AD [59]

Non Donor-Acceptor rectifiers

1.13

Au/SAM/Hg 9 [61]

1.14

Au/SAM/EGaIn ~100 [62]

- 28 -

1.2.4 Synthetic strategies for donor-acceptor dyads

As was mentioned earlier in this Chapter, coupling a strong oxidizing agent with

a strong reducing agent is a challenge that requires special synthetic methods. This part

of the Chapter will cover synthetic approaches that have been employed to obtain

desired molecular systems and give important examples of the synthesised dyads. It also

discusses the criteria for the coupling reactions required to form a covalent link between

two electroactive species.

After the Aviram-Ratner proposal of the TTF--TCNQ dyad, many researchers

focused on the synthesis of such molecules. However, it was found that covalent

bonding of strong donors and acceptors is difficult and the formation of insoluble

charge-transfer salts followed by side reactions of radical ion species was likely a

competing process. The earliest attempt of coupling TTF and TCNQ was published by

Metzger et al .[63], reporting synthesis of the first monomeric example of the TTF--

TCNQ dyad (Fig. 1.9). However, while a molecular ion for the target dyad has been

observed in the mass-spectra, the isolation and purification of the compounds proved to

be hard to achieve.

Figure 1.9: First attempt of coupling strong electron donor (TTF) and strong electron

acceptor (TCNQ).

To overcome the problem of covalently linking strong electron acceptors and

strong electron donors, many synthetic strategies focused on the design of intermediates

with moderate or weak electroactive parts then converting them into strong moieties

(Figure 1.10). For example, use of benzoquinone (BQ) instead of TCNQ allows for

straightforward coupling with a strong donor (TTF). Many TTF-BQ dyads were

synthesized and characterized. However, all attempts to increase the electron affinity of

- 29 -

the acceptor by converting them into TCNQ failed due to incompatibility of TTF to

strong acidic conditions [64-66].

Figure 1.10: Series of donor–acceptor dyads with weak (BQ based) acceptors [66-67].

Interesting results were observed upon converting moderate acceptors into strong

acceptors, as shown for the polynitrofluoren-9-ones (Fig. 1.11). In contrast to

benzoquinones, condensation of fluorenone with malononitrile can be done under mild

conditions resulting in dicyanomethylene derivatives that have similar electron affinity

to TCNQ [68]. It was shown that TTF-fluorene-9-one dyad can be prepared by simple

coupling of carboxyl group on the nitrofluorenone acceptor with hydroxy or amino

group on the TTF [69-70].

Figure 1.11: Synthesis of TTF--fluorene dyads [69].

- 30 -

Another approach to reduce the formation of CTC was based on creating steric

hindrance that will reduce the tendency to form these complexes. Figure 1.12 shows a

series of dyads that are readily synthesized by direct coupling of donor and acceptor

(TCNAQ) fragments. The drawback of the TCNAQ is in the distortion of the π-electron

system which dramatically reduces its electron acceptor ability [71].

Figure 1.12: Donor-acceptor dyads with steric hindrance that prevents formation of π-π

stacking (From [67]).

Perepichka et al. [72] demonstrated that strong acceptor–strong donor dyads can

be synthesized by direct coupling of appropriate synthons. A highly reactive acid

chloride derivative of TCNQ and lithium alcoholate derivative of TTF were coupled

together at very low temperature (–100C). Under these conditions the rate of the

charge-transfer complex formation can be reduced resulting in the desired dyad in

acceptable yields (Figure 1.13).

The major problem of this structure is its flexible -bridge that allows two

conformations of the molecule (Fig. 1.13): ―extended‖ and ―head-to-tail folded‖ [72].

This ―folded‖ conformation is responsible for π-π interaction between donor and

acceptor and, thus, increases the HOMO-LUMO gap.

Many synthetic approaches were focused on the alternative D-π-A systems.

Usually, the π-bridge corresponds to a more rigid structure that prevents unwanted

structural conformations of the Donor-Acceptor system. In most cases the synthesis of

- 31 -

such dyads requires specific metal-catalyzed cross coupling reactions. These types of

catalytic reactions of organic electrophiles with organometallic reagents are powerful

tools for the formation of carbon-carbon bonds as well as carbon-nitrogen, carbon-

oxygen and carbon-sulphur bonds [73]. For example the Stille [74-76] reaction has been

successfully utilized for donor-acceptor coupling of a wide range of compounds

including conjugated polymers with donor-acceptor sequence [77-78] and molecular

rectifiers [58].

Figure 1.13: Synthesis of non-conjugated TTF-TCNQ dyad and structure of two

possible conformations [72]

1.3 Molecular assemblies

Implementation of molecular-scale electronics depends upon the ability to

address individual or small numbers of molecules. The key issue is to find a way to

assemble molecules in a repeatable fashion and develop methods to test these molecules.

To resolve individual molecules electronically, one has to position the molecule between

two electrodes. One way of establishing contact between the molecule and electrode is

self-assembly. Potentially, it allows to position molecules selectively on a surface with

sub-nanometer precision. This simple process, with its intrinsic error-correction

advantage, makes SAMs inherently manufacturable and thus technologically attractive.

In addition, SAMs can be designed and engineered to provide an extremely high density

- 32 -

of functional group on the surface. On the other hand, in order to perform highly

complex functions such as those of integrated circuits, a self-assembly strategy that

enables easy formation of complex patterns to "program" the structures and (electrical)

properties of materials at nanometer levels must be developed.

The molecular assemblies on the surface can be divided in different groups

depending on the type of the interactions between molecules and the surface. Herein we

describe two types of molecular self-assembly used as deposition methods:

physisorptions of the molecules with amphiphilic structure and covalent coupling to the

surface via specific ―anchor‖ group of the molecules.

1.3.1 Langmuir-Blodgett deposition

The first attempts of producing 2D assemblies of molecules were made by

Pockels and Rayleigh in late XIX century [79]. They observed formation of monolayers

of fatty acids on the surface of water. Their studies were continued by Langmuir, who

has developed equipment for studying films of amphiphilic molecules on the water

surface (Langmuir trough) [80]. He has discovered that molecules with such structure

could be aligned at the air-water interface, with the polar functional groups immersed in

the water and the non-polar chains sticking out in the air. Later, Katharine Blodgett was

able to transfer such monolayers onto a solid substrate [81], a process known nowadays

as Langmuir-Blodgett deposition (LB). These discoveries gave an opportunity for deeper

investigations of mono- and multilayer properties, initiating a variety of works to study

the spectroscopic, optical and electrical properties of organic thin films.

The Langmuir-Blodgett technique has been widely used for the fabrication of

molecular structures because it offers good control of order and alignment of the

molecules in the monolayer. All this makes the LB technique a very attractive method

for different fields of research. In spite of the possibility of producing films with high

precision, the LB method is not perfect. The main disadvantages, such as poor thermal

and mechanical stability, have led to a search for other methods of preparing molecular

films, which would be less sensitive to environmental conditions.

- 33 -

To make organic thin films by the LB technique, one needs molecules that can

form a monomolecular layer on an aqueous phase. Such molecules normally have

amphiphilic structure and consist of two parts: a hydrophilic, polar head group and a

hydrophobic, non-polar tail. Immediately after depositing on the surface, the molecules

form a loosely packed monolayer, in which the hydrophilic head-groups of the

molecules interact strongly with water (via its dipole-dipole or by hydrogen bonding

interactions) and the hydrophobic ends protrude from the water surface. The large

hydrophobic moiety prevents dissolution of the molecules in water. An important

indicator of the Langmuir monolayer is an isotherm of surface pressure as a function of

the area available for a single molecule. Surface pressure can be defined as changes in

the surface tension of water upon covering it with molecules, and it can be recorded

during compression of the film [82]. A typical isotherm for fatty acids is shown in

Figure 1.14.

Figure 1.14: Surface pressure/area isotherm of fatty acid. Adapted from [83].

The typical surface pressure/area isotherm presents three distinct regions. As the

surface area is reduced from its initial value (hydrophobic chains are near the water

surface), there is a gradual onset of the surface pressure until a horizontal region that

corresponds to the state where hydrophobic chains are being lifted away from the

surface (surface pressure <1 mN/m). This region corresponds to the 2D ―gas‖ phase

- 34 -

(Fig. 1.14a and 1.15a) and is not always resolved by the instrument. As the barriers

move, the next steeply sloping linear region appears corresponding to a partially

compressed monolayer – the ―liquid‖ phase (fig. 1.14b and 1.15 b). The abrupt increase

of the slope is indicative of the phase change and represents a transition to an ordered

solid-like arrangement of the two-dimensional array of molecules (Fig. 1.14c and 1.15

c). If this second linear region is extrapolated to zero surface pressure, the intercept

gives the area per molecule that would be expected for the ideal state of the

uncompressed close-packed layer.

Figure 1.15: Monolayer of amphiphilic molecules on a water surface: a) expanded; b)

partly compressed; c) close packed.

The Langmuir films floating on the water surface can be transferred on various

solid substrates to study the interaction of the molecules within the two-dimensional

system. The actual deposition process can be visualized as shown in Figure 1.16. It is a

delicate process which depends on many factors such as the direction and speed of the

substrate movement, the surface pressure, composition, temperature, and pH of the

subphase. The deposition process depends on the hydrophobic/hydrophilic properties of

the substrate. In the case of hydrophilic substrate it should first be immersed in the clean

water and then the molecules are spread on the surface. The film is then compressed to

the surface pressure which gives the best results for the transfer ratio, a value that can be

established empirically. Traditionally, the LB deposition is performed using films in the

―solid‖ phase, and deposition is carried out at a constant surface pressure to maintain the

film structure. For the hydrophilic substrate, deposition will follow scheme (a) on Fig.

1.16. The water wets the substrate surface and the meniscus turns up, as the slide is

withdrawn the meniscus is wiped over the surface and leaves the monolayer behind (―Z-

type‖a). The hydrophilic groups of the molecules are turned toward the hydrophilic

a We are aware that X-, Y- and Z-types of LB films (Fig. 1.16A) are used for classification of

multilayers. However here, for clarity, we used terms ―X-type‖ and ―Z-type‖ for monolayers to

distinguish the direction of transfer on a substrate (immersion and withdrawal, correspondingly).

- 35 -

surface of the substrate. If the substrate surface is hydrophobic, the meniscus will be

turned down and deposition should be started on the first immersion of the substrate into

the subphase through the organic film (―X-type‖) (Figure 1.16 (b)).

The Langmuir-Blodgett method of deposition has been used to construct highly

ordered films for different applications, including molecular electronics [43, 49, 84]. The

limitation of such compounds in this area is due to the increase in insulating properties

of the monolayers as a result of the long alkyl chains and fragility of the LB films

The pioneering work on semiconducting LB film was done on N-docosyl

pyridinium-TCNQ charge transfer salt deposited on a CaF2 substrate. The LB films

showed good conductivity after doping with iodine [85-86]. Another attempt to fabricate

LB films with semiconducting properties was done by Petty et al [87]. They deposited

alternating layers of alkyl-chain derivatives of TCNQ and TTF on glass substrates and

such multilayer films showed semiconducting properties. The Langmuir–Blodgett

technique has been most often used to study rectification behaviour of monolayers and

multilayers of Donor-Acceptor dyads (see above).

A

B

Figure 1.16: A) Different types of deposited LB multilayers; B) Scheme of the

deposition process of the monolayer on the solid substrate: a) Z-deposition; b) X-

deposition.

- 36 -

Many external factors can affect the quality of LB films. For example, the

presence of contamination on the water subphase can change the position of the

pressure-area isotherm, thus giving incorrect values for molecular area, and affect the

concentration of molecules constituting the film. Vibrations and larger contaminations

like dust particles may cause collapse of the monolayer and therefore change its average

thickness [88].

The main advantage of the Langmuir-Blodgett deposition is that coverage of the

surface can be measured and controlled directly during the deposition as a transfer ratio;

The limitations of this type of molecular assemblies are:

LB films do not have strong bonding to the surface causing

structure changes over time as the film tends towards a thermodynamic steady

state;

any contaminants, previously present on the electrode surface

(also on water surface, in the solvent or in the compound itself, will become a

part of the LB film, thus influencing the electronic properties of the final device.

1.3.2 Self-assembly by chemisorption

Surface self-assembly, which is defined as the spontaneous adsorption of organic

molecules on a solid surface, was first described by Zisman and co-workers in 1946

[89], where they studied the absorption of monolayers of polar organic molecules (such

as long alkyl-chain alcohols and amines) on polished metal surfaces. The wide interest

in self-assembly began with the work conducted by Nuzzo and Allara in 1983, in which

they studied chemisorption of organic thiols and disulfides on gold surfaces [90]. In this

process, the molecules form strong chemical bonds with the substrate via special

terminal anchor groups, thus providing stable and robust monolayers. The bonding can

be purely covalent (e.g. Si-O-Si on oxidized surfaces), covalent and slightly polar (e.g.

Au-S for alkanethiols on gold), or fully ionic. Due to the substrate–anchor group

interaction, molecules try to attach to every available binding site on the surface. Also,

van der Waals interactions between the methylene chains cause the molecules to pack

- 37 -

densely on the surface. In general, the longer the chain, the more ordered monolayer

structure is [91-92].

A number of compounds have been found be capable to form SAMs on various

substrates: chloro- and alkoxysilanes on various hydroxylated substrates (silicon

dioxide, aluminum oxide, quartz, glass, mica, zinc and germanium oxide) [93]; fatty

acids on metal oxides surface (aluminum oxide [94], silver oxide [95]). The most

extensively studied type of SAMs is alkanethiols (and their chemical equivalent,

disulfides) adsorbed onto various metal surfaces: gold [90-91, 96-102], silver [103-104],

copper [101], palladium [105-106], platinum [107] and mercury [108]. The applications

of the SAMs range from studies of the molecular and cellular interactions with specific

functional groups, surface energies, surface charge, or other interfacial properties to the

introduction of specific functionalities to study cell signalling, cell adhesion [109], and

protein interactions [110]. SAMs have also been used for constructing molecular

switches [111], biosensors [112] etc. The covalent self-assembly was widely adopted in

the field of molecular electronics and particularly molecular rectifiers [113-115].

The most common protocol for preparing SAMs of thiols on metal surfaces is

immersion of a freshly prepared or clean substrate into a dilute (1-10 mM) ethanol

solution of thiols for ca. 12-24 hours at room temperature (Figure 1.17). This procedure

allows the use of different solvents, variation in temperature and exposure time to

optimize the formation of SAMs. The self-assembly process is essentially an exchange

between organosulfur molecules with anchor groups and whatever materials were

adsorbed on the surface of the substrate before self-assembly. Thiols are able to displace

various impurities and contaminants that are already present on the surface. The

displacement will require desorption of impurities and this process will therefore affect

the kinetics of SAM formation. Different methods of substrate preparation or cleaning

(―piranha‖ solution, oxygen/air plasmas) are used to facilitate the SAMs formation.

Within the first minutes of self-assembly one can obtain a dense coverage on the

substrate with alkylthiol monolayer but then the slow process of reorganization will

require hours to maximize the density of the molecules and minimize the defects in

SAM [116].

- 38 -

Figure 1.17: Process of the growth of the SAMs: a) immersion of the substrate into the

dilute solution of the molecules; b) initiation of the self-assembly process; c) formation

of the densely-packed monolayer.

For monolayers containing closely packed alkanethiols, the spacing of the alkane

chains is 4.97 Å as determined by low-energy electron diffraction [117]. This spacing is

almost three times larger than the van der Waals radius of sulfur (1.85 Å) suggesting

minimal S-S interactions [118]. This distance is also greater than the distance of closest

approach of the alkyl chains (4.24 Å). This difference in spacing causes the axis of the

alkyl chains to tilt by 30° from the surface normal [97, 118-120]. The tilt angle is

virtually independent (within a few degrees) of the functionality of the head group, as

long as it is not larger than the spacing between the alkyl chains [118].

Numerous theoretical studies suggest that the reaction of the thiol with the gold

surface proceeds through as oxidative addition of the S-H bond to the Au, followed by

elimination of the hydrogen. Such chemical bonding corresponds to energy of ~40

kcal/mol [93]. Also the monolayer packs tightly due to van der Waals interactions (~1

kcal/mol per each methylene group in the chain [121]), thereby reducing its own free

energy [116, 122]. All this makes the SAMs stable in a wide range of temperature,

solvents and potentials. The thermal stability of alkanethiolate SAMs has been studied in

a number of papers. It has been reported that loss of sulfur from hexadecanethiolate

monolayer on gold occurred over the range of 170-230°C. A temperature-programmed

desorption of methanethiolate SAMs on gold reported a desorption maximum at 220°C

[123-124].

- 39 -

1.4 Characterization of organic monolayers

Analysis of the surface composition, structure and its physical properties as well

as alignment of molecules in monolayers is important for understanding their electrical

behaviour in molecular junctions. In contrast to inorganic thin films or organic

compounds in bulk, molecular monolayers are extremely fragile and soft, and thus

require non-destructive analytical tools. Among them, many spectroscopic methods such

as attenuated total reflection FT-IR, surface-enhanced infrared absorption, X-ray

photoelectron spectroscopy (XPS), near-edge X-ray absorption fine structure

spectroscopy (NEXAFS), time-of-flight secondary ion mass spectrometry (TOF-SIMS)

and surface plasmon resonance (SPR) have been widely used to obtain information

about thickness, structural disorder, chemical composition and presence of impurities in

the molecular films. Contact angle measurements provide additional information about

the changes in surface wettability after its modification. Finally, electrochemical

characterization of the SAMs on the electrode surface provides knowledge about the

conductivity of the film and the nature of the redox activity.

1.4.1 Ellipsometry

Measuring the thickness of the film can provide important information about the

geometric structure (monolayer or multilayer) of the film and alignment and order of the

molecules within the monolayer. A common technique to determine the thickness of the

films, for ~1 nm to several microns, is ellipsometry. Figure 1.18 shows the principle of

the ellipsometric measurement.

Ellipsometry was extensively used to study the physical and optical properties of

both LB films and SAMs. In addition to theoretical elaboration of ellipsometric models

for studying thin films, [125-127] a great deal of experimental research has also been

done in this field. For example, Porter et al. found by ellipsometry a noticeable decrease

of the thickness for monolayers formed from alkanethiols with chains shorter than 8

methylene groups. These results were interpreted as a decrease in monolayer order for

molecules with shorter chains [97]. The alignment of molecules in LB films can also be

- 40 -

studied by ellipsometry measurements [128]. Ellipsometry was shown to be very useful

for real-time in situ monitoring of the films formation and growth [129-131].

Figure 1.18: Schematic presentation of the ellipsometry setup.

1.4.2 Contact angle

The quality of the monolayer can often be estimated from wetting properties of

the surface. Wetting is the ability of a liquid to maintain contact with a solid surface,

resulting from intermolecular interactions when the two are brought together. The

degree of wetting (wettability) is determined by the force balance between the adhesive

and cohesive forces. The free energy of the surface will affect the free energy of the

droplet of water on that surface and, as a result, will determine the shape of the droplet.

Figure 1.19: Measurement of contact angle

- 41 -

The contact angle (θ), as seen in Figure 1.19, is the angle at which the liquid-

vapour interface meets the solid-liquid interface. The contact angle is determined by the

resultant between adhesive and cohesive forces and provides an inverse measure of

wettability. If the liquid is very strongly attracted to the solid surface (for example

oxidized surfaces that can form hydrogen bonding with water) the droplet will

completely spread out on the solid surface and the contact angle will be close to 0°. Less

hydrophilic solids will have higher contact angle. On many highly hydrophilic surfaces,

water droplets will exhibit contact angles of 0° to 30° (e.g. pure metals). If the solid

surface is hydrophobic, the contact angle will be larger than 90°. On highly hydrophobic

surfaces the surfaces have a water contact angles as high as ~120° on low energy

materials (e.g. polymers, such as PVC and Teflon). However some materials with highly

rough surface may have water contact angle greater than 150°. These are called super-

hydrophobic surfaces [132-133].

The relations between the contact angle and properties of the thin films and

monolayers have been extensively studied. For example, Bain et al. attributed the

difference in contact angles of mercaptocarboxylic acids with different alkyl chain

lengths to the order/disorder factor in monolayers (longer chain acids form more ordered

monolayers) [134]. The same group has also studied the effect of various terminal

groups (CH3, OH, etc.), chain length and other parameters on wettability of the thiol-

based monolayers [91, 119].

A good estimate of the SAMs uniformity can be obtained by monitoring the

difference between the advancing and receding contact angles. This difference, called

the contact angle hysteresis, is a measure of the physical and chemical nonuniformity of

the surface [135].

1.4.3 Reflection-absorption and ATR infrared spectroscopy

IR spectroscopy is one of the most surface-sensitive, non-destructive methods for

the characterization of chemical composition, structure, and orientation of molecules at

surfaces with monolayer resolution. The properties of the molecules in thin films

absorbed on a surface can be very different from those in bulk and these differences can

- 42 -

be characterized by IR spectroscopy. For the analysis of metal surfaces, the most

successfully used method is reflection-absorption infrared spectroscopy (RAIRS or

Grazing-angle IR). In addition to the techniques used for metals, semiconductors and

insulator surfaces are widely probed by attenuated internal reflections (ATR) IR

spectroscopy. Other infra-red techniques, such as photoacoustic spectroscopy (see for

example [136]) and photothermal displacement spectroscopy [137], are also infrequently

employed to test the absorbance of different materials on the surface. General

considerations about using one method or another usually depend on different factors

and are based on the relative sensitivity of the method to studying the monolayer and

surface.

Francis and Ellison first recognized that a large incident angle is necessary to

enhance sensitivity [138] and the pioneering work on the RAIRS for SAMs was done by

Allara and Swalen [139]. The first studies of alkyl thiols on gold surfaces showed that

both the width and the position of the CH2 stretching vibrations are sensitive to the

structure of the film.

The RAIRS method requires IR radiation to be incident at a particular grazing

angle to a metal surface, on which an organic layer has been deposited (Fig. 1.20).

Incident perpendicularly-polarized radiation undergoes a phase shift of 180 on

reflection from the metal surface and so the electric field components of the incident and

reflected radiation cancel at the metal/thin film interface. In contrast, the incident and

reflected components of parallel-polarized IR radiation differ by only about 90 at

grazing incidence. This is the origin of the surface selection rule, which results in the

very useful ability to distinguish vibrations that possess a transition dipole moment with

a large component perpendicular to the surface [140].

- 43 -

Figure 1.20: Schematic diagram of the grazing-angle FTIR spectroscopy from a thin

film-covered surface.

In attenuated total reflection spectroscopy (ATR), a beam of infrared light is

passed through the ATR crystal in such a way that it reflects off the internal surface.

Changing the angle of incidence may vary the number of reflections. The total internal

reflection occurs when the light strikes the medium boundary at the angle larger than a

particular critical angle with respect to the surface normal. This internal reflectance

creates an evanescent wave that extends beyond the surface of the crystal into the

sample held in contact with the crystal (Fig. 1.21). The penetration depth into the sample

is typically between 0.5 and 2 micrometers, with the exact value being determined by

the wavelength of light, the angle of incidence and the type of ATR crystal [141].

Figure 1.21: Formation of the evanescent wave at the internal reflection surface.

- 44 -

A large variety of molecules and monolayers have been studied using grazing

angle and ATR infrared spectroscopy For the purposes of this thesis, it is not possible to

completely review all the examples of IR methods used to quantitatively study

monolayers structures and the orientation of molecules in monolayer assemblies. We

will cite only those most related to the work in this thesis. The application of IR

spectroscopy to study the orientation of molecules at the surface has been well

developed [139, 142-145]. The most studied by IR spectroscopy are monolayers of alkyl

thiols on gold surfaces [146], fatty acids on metal oxide surfaces [97, 103, 147-149] and

LB films [150-152]. For these monolayers the specific spectral features of C-H

vibrations can be correlated with order/disorder of the film. In principle, the relative

intensity of the peaks and their position are associated with the orientation of the dipole

moment of the vibration modes for C-H bonds and, accordingly, provide information

about the molecular orientation on the surface. Important information can be obtained

from the difference between the IR spectra of the bulk and the monolayers. The

frequencies of the CH2 stretching are very sensitive to the conformational ordering of the

alkyl chain within the monolayer. The frequency shifts are not large (5-7 cm-1

);

nevertheless, it was found that when alkyl chains in the monolayer are all in trans

conformation and highly ordered, the absorption bands of asymmetric and symmetric

methylene vibrations appear at 2918 and 2850 cm-1

, respectively. However, with the

presence of gauche conformation, the bands shift bathochromically to 2926 and 2856

cm-1

, as studied for both LB films and SAMs [149, 153]. The bandwidths of the CH2

vibrational modes were also found to be dependent on the degree of mobility and

flexibility of the molecules within the monolayer and it increase with decreasing order.

This parameter can also be used to determine the structure of the films [154-156]. The

relative intensities of the stretching vibration when using polarized light provide

information about the tilt angle of the bond with respect to the surface normal [143,

157].

- 45 -

1.4.5 Electrochemical characterization of SAMs

Cyclic voltammetry (CV) is a powerful electrochemical tool to study the stability

of the films and electron-transfer in redox active species present on the surface of an

electrode. CV has also been used to investigate monolayer structure and packing of

molecules adsorbed on the electrode [97, 158]. The presence of defects and pinholes in

the monolayers can also be verified by electrochemistry. As the close-packed SAMs on

the electrode block diffusion of species to the electrode surface, evidence for pinholes or

defects can be obtained by analysing the blocking behaviour of additional layers

specifically at the bare metal surface [159-161].

The nature of the electrode surface is a critical factor in all of the electrochemical

processes that take place. When the electrode surface is coated with a layer of organic

molecules one can expect the electrochemical processes to be significantly affected.

Specifically, the film present on the electrode surface can retard or completely block the

electron transfer by increasing the separation between the electrode surface and redox-

active molecules. However, if the thickness of this film is small (for example a self-

assembled monolayer on the gold electrode) the electrons can still tunnel through it.

Also, if the electroactive centers are all close to the electrode surface, then diffusion

should not have any influence in the process. Accordingly, one should expect the

oxidation and reduction peaks be at the same potential for electrochemically reversible

redox reactions (Fig. 1.22) [162]. However, in some cases the separation between the

two potentials can still be observed. Often, this can be explained by slow electron

transfer at the electrode surface.

The most common example of electroactive SAM is formed by ferrocene

derivatives with thiol-terminated alkyl tail. Adsorption kinetics [163] and exchange of

the thiols between the SAM and solution [164-165] can be studied if one of the

components is electrochemically active. The relationship between the film structure and

CV response was studied in experiments with mixed SAMs composed of thiol-

terminated ferrocene and different alkyl thiols. For example, the location of the

ferrocene head-groups in the SAM and on the surface of the SAM causes the existence

of multiple formal potentials which results in broadening of the peaks [164].

- 46 -

Experiments with varying the alkyl chain length of the thiols, leading to the ferrocene

group ―buried‖ in the SAM, confirm the broadening of the peaks and shift to more

positive potentials [166-168].

Figure 1.22: Cyclic voltammogram of an Au electrode coated with a mixed monolayer

containing thiol terminated ferrocene [169].

For an electroactive species adsorbed on the surface of an electrode, cyclic

voltammetry can be used to determine the surface density. This is done by integration of

the area under the oxidative curve to quantify the total charge passed in the

electrochemical process, and is expressed by the following equation:

(1.3)

where Γ is the surface coverage in mole/cm2, Q is the charge passed to

oxidize/reduce the molecule, n is the number of electrons in the electron-transfer

process, F is Faraday’s constant, and A is the area of the electrode.

Stability of the SAMs is an important factor that contributes to the application of

the molecules in electronic devices. By analyzing the redox response during the multiple

CVs scanning, a quantitative characterization of the SAM stability can be obtained [170-

172].

- 47 -

1.5. Fabrication of molecular junctions

To study the rectification of a single molecule one would need to attach two

metallic probes on both sides of the molecule and examine the conductance of the

junction by studying its I-V characteristic at forward and reverse bias. However,

addressing an individual molecule presents a series of technical difficulties. Only a

recent progress in nanostructure characterization (STM/AFM) and nanofabrication tools

made such experiments possible.

The most simple and common approach for fabricating the metal-molecule-metal

junctions is based on ordered structures assembled on the first electrode, on which the

second electrode is then fabricated or positioned. This approach uses a metal electrode

supporting a SAM or LB film as one contact and a second contact generated on top of

the organic surface by different methods such as:

depositing a metal film by vacuum deposition or electrodeposition [44, 173];

transferring a metal film by flotation or nanocontact printing [174];

positioning a conducting probe (STM [175] or conducting AFM [176];

making a contact with a liquid metal contact (mercury) [177];

using conductive polymer mixtures (PEDOT:PSS) as a barrier between metals and

organic monolayer [178].

A second approach involves fabrication of junctions by positioning the

molecules across the nano-gap between the electrodes. The gap can be fabricated by

breaking a single wire mechanically or electrochemically [179] or by narrowing a gap

by electrodeposition of metal [173].

Each type of junction has certain advantages and limitations. In this work we

will discuss in detail three of these methods, thermal vacuum evaporation of the top

metal electrode, mercury-drop junctions and scanning probe microscopy, as the most

accessible tools that a synthetic chemist can use to study the electronic properties of the

molecules.

- 48 -

1.5.1 Vacuum deposition of the metal on top of the organic layer

The preparation of metal thin films by vacuum deposition is one of the first

techniques used for metallization of organic surfaces. The nature of the contact between

organic molecules and metals depends strongly on chemical reactions that can occur

between an evaporated metal and the terminating group of a molecule. Evaporated metal

atoms may also penetrate through the organic monolayer and short out the device or

even create meta-stable filaments that switch the conductance on and off during the

measurements [180]. Many studies of the thin films have shown that metals such as Ti

and Cr are highly reactive toward organic functional groups, while Au and Ag are

mainly inert [180]. The degree of penetration of the metal atoms increases with

decreasing reactivity, which makes Ti a contact metal of choice for many studies [179,

181]. Titanium is a unique metal for an evaporated top electrode. Due to its high

reactivity, it immediately cleaves terminal C–H bonds forming a thin titanium carbide

layer on the surface of the monolayer that prevents further penetration of the Ti atoms

inside the film [182]. In the case of less reactive metals, during the evaporation on top of

the organic film, a sample holder should be cooled by liquid nitrogen. This is usually

sufficient for Al deposition [44]. For gold deposition, the so-called ―cold gold‖ method

was developed to prevent organic film damage. In this technique of deposition the Au

atoms are forced to undergo multiple scattering before they reach the substrate by adding

Ar gas to the evaporation chamber and by positioning the substrate on the opposite side

of the holder (away from the crucible) [28, 38, 47].

An interesting modification of this technique was developed by Reed and co-

workers [183]. In this method a pore in the SiN with a 30–50 nm aperture was fabricated

using combination of different microfabrication methods (Fig. 1.23). The pore was then

filled with gold by thermal evaporation and the whole device was immersed in a solution

of the studied molecule (4-thioacetylbiphenyl) to form a SAM. After deposition, the

second electrode was formed by evaporating gold onto the sample, at 77 K to minimize

damage to the SAM [184]. The significant disadvantage of this method is that once the

structure is sealed by the evaporated electrodes, it is impossible to study the structure,

order or orientation of the SAM even before the top electrode is evaporated in place.

- 49 -

Figure 1.23: ―Nanopore‖ junction [183].

1.5.2 Liquid metals drop junctions

Extraction of reliable data out of molecular junctions requires statistical analysis of a

large number of measurements. This creates a demand for a simple and soft method to

form a mechanical contact to an organic film supported on electrode surface (such as

SAM on the gold). Mercury has been used for a long time in the formation of tunneling

junctions. Mann and Kuhn [185] used mercury to form various metal/LB film/metal

tunneling junctions. Later, the groups of Whitesides and Rampi [113-115] and Majda

[186-187] independently developed the method to create metal/SAM//SAM/metal

junctions using mercury drop as one of the electrodes or both electrodes. Thiol-

terminated molecules can form SAMs on the mercury surface, similarly to gold.

Bringing the modified mercury electrode into contact with the studied SAM on the

second electrode results in formation of metal-SAM-SAM-metal junctions (Fig.1.24).

Such junctions are fast and easy to make, allow for different combinations of SAMs and

metals to be used, and provide control over the size of the contact. One difference

compared to other molecular junction methods is a new SAM-SAM interface. The

organic monolayer on the top mercury electrode is necessary for protection of two metal

- 50 -

electrodes from direct contact, which would lead to amalgam formation. The electron

transport in these junctions can be monitored as a function of the interaction between

these two SAMs. It was found that covalent, hydrogen bond and van-der-Waals

interactions could change the conductivity of the junctions by more than four orders of

magnitude [188]. Similar systems with only one SAM (e.g., contact made between bare

Hg electrode and SAM-modified gold substrate) are usually not stable and easily form

an amalgam due to presence of defects in the organic layer. However, use of mercury-

SAM-semiconductor (ex. Si substrate) interface avoids this problem [189]. The

disadvantages of mercury drop based junctions are in significant influence of the surface

topography and the structure of the SAMs on reproducibility of the measurements [190].

Recently, Whitesides and Nijhuis successfully used another liquid electrode

(eutectic alloy of gallium and indium, Fig. 1.8). The Ga2O3/EGaIn electrode is much less

likely able to form short-circuits with the bottom electrode and thus eliminates the

necessity of using an alkylthiol protective layer on the second electrode (like in case of

mercury) [191] and significantly increases the yield of successful measurements [192].

Figure 1.24: Mercury drop junction technique for measuring electronic properties of the

molecules [114]).

- 51 -

1.5.4 Electron Transfer in Metal-Molecule-Metal Junctions

Above we described the methods and techniques to build the systems for

studying the molecular conductivity. Interpreting the results though, require

understanding of the processes by which an electron moves across the metal-molecule-

metal junction, and is generally very complicated. Several important components of the

system determine how electrons traverse metal–molecule–metal junctions [193]:

energy and position of the molecular orbitals;

the type of bonding in the metal–molecule junctions (molecule-metal

interface ) which determine the interaction;

energy alignment of molecular orbital levels with the Fermi levels in the

metal.

The basic mechanism for electron transport in all types of systems generally

involves electron tunnelling – a process of crossing a finite potential barrier by a

wavelike particle. Its probability depends on the barrier width and availability of the

unoccupied states (LUMO or conduction band) on the other side of the barrier. The

tunnelling current shows an exponential dependence on the length of the barrier

(molecule):

(1.4)

where β is a structure-dependent tunnelling attenuation factor, l is a width of the

barrier (length of the molecule).

The electronic structure of the molecule plays a significant role in the electrical

behaviour of molecular devices. If the molecular orbitals are fully delocalized along the

entire molecule then the electron can traverse the molecule in a resonance process and

the molecular junctions will exhibit high conductivity, as for example the π-conjugated

oligo(phenylene ethynylene) [173, 194]. In contrast, if the molecular orbitals are

localized on a specific part of the molecule, then the electron will have to tunnel through

the non-conductive part of the system. In most cases, the HOMO and LUMO orbitals of

the molecule are not aligned relative to the electrode Fermi levels. Applying an external

- 52 -

bias shifts the Fermi levels of the metal electrodes: the negative voltage will raise the

Fermi level and positive voltage will lower it. The electronic structure of molecules also

substantially changes upon making the electrical contact with metal electrodes and as a

function of applied bias voltage. Generally, this process lowers the potential barrier

between the electrode and the molecular orbitals and makes tunnelling possible.

Another important characteristic of the junction is the nature of the contact

between the molecule (anchor functionality) and the electrode. In the ideal case, a low-

barrier ohmic contact between the metal and organic molecules will allow to study the

―pure‖ molecular electronic behaviour. However in real junctions it is not the case and

most interfaces used to study the conductivity of the molecules and molecular

assemblies have significant barriers for the electrons or holes injection, which may even

dominate the whole junction behaviour. The contact, actually, controls the energy and

mixing between molecular orbitals and the electronic states of the metals [195]. It was

demonstrated that a conjugated molecular wire, chemically bound to one gold electrode

by the thiol linker and in only physically contact with the second gold electrode shows

current rectification [196]. The effect of the different contacts was shown in series of

experiments by keeping the contact on one end of the molecule constant (S-Au) and

varying the contacts on the other end (Fig. 1.25) [194, 196]. Since one end of the

junctions is always the Au-S the observed rectification ratio can be related to the amount

of electronic coupling between the molecule and the metal at the second contact. The

more effective interaction between molecule and the metal leads to the symmetric I-V

characteristics. However, poor orbitals mixing at metal-molecule interface results in

strong current rectification. Thus, it shows the importance of metal/molecule contact in

charge transport of the entire junction.

- 53 -

Figure 1.25: Rectification ratio as a function of applied bias for several

Au/molecule/Au junctions with different contacts at two terminals [194].

The complete picture of the charge transport across the junctions is complicated

and depends on many different factors. In addition to the search for the low-barrier

contacts between electrode and the molecule we should consider that the alignment may

also determine the charge transport barrier. Furthermore, the position of HOMO and

LUMO energies varies with the applied electrical field [197-198]. Even more

complication arises in multicomponent molecular junctions (ex. Metal-SAM/SAM-

Metal). The interaction between the molecules in such junctions may result in

electrostatic barrier, orbitals mixing and energy splitting [199]. Finally, simple defects in

the junctions sometimes may mislead the interpretation of the junction behaviour.

- 54 -

Conclusions

The properties of the molecules play a significant role in the behaviour of the

molecular devices. A careful selection of the components and design of future molecular

rectifiers is therefore very important for successful projects. Such design rules include a

proper choice of the Donor and Acceptor moieties that will result in a small and

adjustable HOMO-LUMO gap of the dyad and position of molecular orbitals vs. Fermi

levels of the electrodes. The linker between the electroactive moieties should provide a

separation of the molecular orbitals of the donor and the acceptor. Finally, the molecular

blueprint should include a design of appropriate functional groups which will assist in

the assembly of the molecules between two electrodes.

The nature of the metal-molecule junction can have substantial influence on the

performance of the device. One of the important issues in regards to analysis of the

results are the difficulties associated with understanding the chemical nature and

structure of the junctions and contacts that are being measured. Use of wide range of

current analytical techniques can provide exhaustive information about chemical

composition of the bulk materials and thin films, alignment of the molecules within the

monolayer and properties. There are also many approaches to fabricate the electrode-

molecule-electrode junctions, such as vapour deposition of the top electrode on the

monolayer and liquid electrode junctions that allow measurement of conductivity of a

small number of molecules.

Although molecular rectification was shown and confirmed in numerous papers

for different molecular systems, there are still many open questions and challenges that

retain the interest of the researchers in this field. Particularly, the molecular rectifier with

the structure of the original Aviram-Ratner model has not been tested experimentally

yet. Several donor-acceptor dyads with low HOMO-LUMO gap were previously

reported in literature; however, their rectification properties were not demonstrated.

Creating such molecular systems with very strong electronic asymmetry could be a

potential way to eliminate the influence of the external factors (contacts, defects, etc.) on

the rectification behaviour of the molecular junctions.

- 55 -

Chapter II. The first studies of a tetrathiafulvalene-σ-acceptor

molecular rectifier

(Part of this Chapter was adapted with permission from: G. Ho, J. Heath, M.

Kondratenko, D. F. Perepichka, K. Arseneault, M. Pezolet, M. R. Bryce, The first

studies of a tetrathiafulvalene–σ–acceptor molecular rectifier, Chem. Eur. J. 2005, 11,

2914–2922. Copyrights 2005 Wiley-VCH Verlag GmbH&Co. KGaA, Weinheim)

Introduction

Shortly after discovery of metallic conductivity in the purely organic donor–

acceptor (D–A) complex of TTF and TCNQ [200], Aviram and Ratner proposed the

concept of a molecular rectifier [25]. Their work was based on theoretical study of a

covalent donor-bridge-acceptor (D--A) molecule, such as TTF--TCNQ (2.1) in which

is a saturated aliphatic linker.

Since then, numerous attempts of the synthesis of molecules with a small

HOMO-LUMO gap have been done [28, 201-202]. Still, the design and synthesis of

such systems presents considerable challenges. And since, the hypothetical TTF--

TCNQ molecule 2.1 with a HOMO-LUMO gap of 0.3 eV was never synthesized, a

special synthetic interest exists for coupling together these moieties.

The closest analogue to the Aviram-Ratner rectifier 2.1, a compound 2.2 was

successfully synthesised in analytically pure form from TTF lithium alcoholate and a

TCNQ acid chloride at –100 C; this favoured covalent coupling via ester formation

over electron-transfer processes [72]. The molecule has two reversible oxidations and

two reversible reductions in cyclic voltammetry (CV) experiments that are characteristic

- 56 -

of the TTF and TCNQ moieties, respectively, and the difference between the E0

1ox and

E0

1red values suggests an HLG of only 0.17 eV. However, such an extremely low HLG

leads to an easy electron transfer in the solution at thermal excitation. Also, ―head-to-

tail‖ geometry of the molecule limits the use of the molecule for rectification

measurements. The coupling of different acceptors, fluorene derivatives

(dicyanomethylene derivative of tetranitrofluorene has very strong electron affinity

[203], very similar to TCNQ), with TTF-CH2OH was first attempted by Perepichka et

al. [69, 204]. Following standard conditions for pyridine catalyzed ester formation they

successfully synthesised compounds 2.3. Important feature of the nitrofluorene as an

acceptor is that a condensation of it with malononitrile takes place under mild

conditions, which results in facile conversion of a moderate acceptor into a strong one,

thus, reducing probability of the charge transfer salt formation during the coupling

reaction [68]. The flexible linker in the molecule 2.3 is long enough to allow an

unwanted head-to-tail conformation, which leads to a formation of the intramolecular

charge-transfer complex. Using an acceptor derivative like fluorene-4-carbonyl chloride

would provide a shorter bridge with donor moiety and, thus, prevent formation of the

intramolecular complex [204]. TTF-fluorene dyad 2.4 was synthesized by coupling

fluorene acid chloride with the lithium alkoxide derivative of TTF-CH2OH at –100 C.

Increased reactivity of donor synthon and a lower reaction temperature in this coupling

reaction minimize charge-transfer complex formation. A very low HOMO–LUMO gap

of approximately 0.3 eV was found for molecule 2.4 [69] which is close to the value of

the original A-R molecule (2.1). Thus, the dyad 2.4 appears as a suitable candidate for

testing as a molecular rectifier.

- 57 -

Herein we present an improved synthesis of TTF--nitrofluorene dyad 2.4, its

characterization by spectroscopic and electrochemical methods, preparation of its

Langmuir-Blodgett monolayer on solid substrates and its analysis by ATR IR

spectroscopy. Finally, we discuss our results on the rectification study of the dyad in

Si/LB-film/Ti and Au/LB-film//C16S-Hg junctions.

2.1. “Amphiphilic design”

The general design of the D–bridge-A dyad 2.4 includes a strong electron donor

(TTF) and a strong electron acceptor (nitrofluorene) separated by a saturated bridge, and

an amphiphilic structure enabling fabrication of a Langmuir–Blodgett films [204]. Such

design makes this molecule a suitable candidate for molecular electronic devices,

particularly in the frame of the original Aviram–Ratner rectification concept. A 4,5-

dipentyl-4’-methyl-TTF 2.5 was considered to be an appropriate TTF building block as

alkyl substitution is known to lower significantly the oxidation potential of TTF (i.e.

raise the HOMO). Pentyl chains would also enhance solubility in organic solvents and,

moreover, serve as hydrophobic elements in the amphiphilic structure of the dyad. The

linker between donor and acceptor in 2.4 is short enough (three atoms) to prevent an

intramolecular head-to-tail interaction between the donor and acceptor fragments of the

dyad.

2.2. Synthesis of TTF--fluorene dyads

Modification of the donor moiety was done following literature procedure [204]

and is presented in Scheme 2.1.

- 58 -

Scheme 2.1: Synthesis of TTF-donor synthon.

Lithiation of the TTF derivative 2.5 [205] with LDA, followed by the reaction of

the lithium salt with N-methylformanilide afforded an aldehyde 2.6 in 66% yield. The

aldehyde was reduced with NaBH4 to give hydroxymethyl-TTF 2.7 in 88% yield.

The covalent linkage of the TTF fragment with fluorene acceptor was achieved

through the formation of an ester bond. The carbonyl group on the acceptor (fluorene)

component 2.8 was converted into fluorene-4-carbonyl chloride 2.9 as presented in

Scheme 2.2.

Scheme 2.2: Synthesis of acid chloride derivative of the acceptor synthon.

The dyad 2.11 was successfully synthesized in 78% yield by coupling 2.9 with

the lithium alkoxide derivative of 2.7 (Scheme 2.3) at –100 C, resulting in desired dyad

as a dark crystalline compound.

- 59 -

Scheme 2.3: Coupling together donor and acceptor synthons. i) 2.7 + BuLi, THF, –

100C 20C, 1h, then + 2.9, –100C –15C, 3h, then –15C, 12h.

The acceptor ability of the fluorene moiety of 2.11 was increased by conversion

to the dicyanomethylene derivatives 2.4 (with 95% yield) by treatment with

malononitrile in DMF solution (Scheme 2.5).

Scheme 2.5: Conversion of the fluorenone 2.11 into dicyanomethylene derivative 2.4.

As the separation of the dyads 2.11 and 2.4 was difficult, the completion of the

conversion was achieved by large excess of the malononitrile. It was reported that under

these conditions a blue colour by-product can be obtained as a result of substitution of

the cyano group of 2.4 with a second molecule of malononitrile [204]. However, the

purification of the final molecule 2.4 from this by-product is easy by filtering the crude

mixture through silica gel layer.

2.3. Geometry and electronic structure of the dyad 2.4

To evaluate the molecular properties of the dyad 2.4, we calculated the geometry

and the orbital energies of a simplified molecule (lacking long alkyl substituents, Fig.

- 60 -

2.1) by using density functional theory (DFT) at the B3LYP level, with a 6–31G(d) basis

set. This method is reliable for describing geometry and orbital energies of organic

molecules [206], and was successfully used to predict the HOMO–LUMO gap of a

related low-gap TTF--TCNQ dyad [72]. In accordance with previously reported

experimental observationsa [204], our calculations show no possibility for intramolecular

π-π complexation between the TTF and fluorene fragments. The conformational analysis

reveals the presence of several stable conformations, differing in energy by 1–2.5

kcal/mol.

Figure 2.1: The calculated [B3LYP/6–31G (d)] geometry of the dyad 2.4 in the

minimum energy conformation and the plot of HOMO (left) and LUMO (right) orbitals.

The calculated structural features of dyad 2.4 are similar to those previously

found by X-ray crystallographic analysis in the related compound 2.12 [204]; this

confirms the applicability of the chosen theoretical model. The variation of the

calculated HOMO–LUMO gap (0.30–0.35 eV) in different conformations is low, which

is in contrast to the dyads with a long linker [72]. The calculated HOMO-LUMO gap

is also very close to the experimentally observed electrochemical gap of compound 2.4

(0.29 V) obtained from cyclic voltammetry experiments. The HOMO–LUMO gap of 2.4

a Change in the solution UV-Vis absorbance as a function of concentration showed no linear trend and

almost no absorption at concentrations below 10–4

M, which clearly indicate that only the intermolecular

CT complex, is responsible for the long-wavelength absorption in dyad 2.4.

- 61 -

fits almost exactly the original Aviram–Ratner model, in which the asymmetric I–V

curve was calculated assuming a 0.3 eV gap [25]. At the same time, the electronic state

of 2.4 (and, therefore, the rectification behaviour) in the tunnelling junction might be

difficult to predict. Although the ground state of individual molecules is neutral, the

electron transfer is very facile in these systems, as manifested by a relatively weak ESR

signal in solution. When an LB film of 2.4 is sandwiched between conducting

electrodes, the energy levels may be shifted significantly (due to intermolecular

interactions, electrode interface effects [207], and the applied electric field [35]). Such

shifts may even cause the gap to close, so that the zwitterionic biradical ground state of

2.4 (D+

--A–

) in the junction is a possibility.

2.4. Electrochemical characterization of the dyads in bulk

The cyclic voltammetry (CV) of the dyads 2.4 and 2.11 in CH2Cl2 presents clear

characteristic multiredox amphoteric behaviour (Figure 2.2) consisting of two reversible

single-electron oxidation waves corresponding to a radical cation and dication state of

the TTF moiety, and three reversible single-electron reductions of the fluorene fragment,

which afford a radical anion, dianion and radical trianion species:

The electrochemical potentials E0 of dyads 2.4 and 2.11 are given in Table 2.1,

together with data for donor synthons 2.5 – 2.7 to show the mutual influence of the TTF

and fluorene fragments.

As expected [208], the conversion of the keto group into the dicyanomethylene

group (2.112.4) results in a significant positive shift in the reduction potentials with

larger shifts being observed for first (E1red0=350 mV) and third (E3red

0=180 mV)

reduction peaks than for second ((E2red0=90 mV). For the TTF moiety, a decrease,

followed by increase in its donor ability was found upon the attachment of the electron

withdrawing CHO group (2.52.6) and conversion of it into electron releasing CH2OH

group (2.62.7).

- 62 -

Table 2.1: Electrochemistry data (0.2 M Bu4NPF6 in CH2Cl2 vs. Fc/Fc+).

Compound E1ox0,V E2ox

0,V E1red

0,V E2red

0,V E3red

0,V E,eV

2.5a –0.15 0.32 – – – –

2.6a 0.00 0.52 – – – –

2.7a –0.23 0.24 – – – –

2.11 –0.10 0.40 –0.74 –1.02 –1.81 0.64

2.4 –0.10 0.41 –0.39 –0.93 –1.63 0.29

a) data from [204]

-1.5 -1.0 -0.5 0.0 0.5

2.4

2.11

E, volts (vs. Fc/Fc+

)

Figure 2.2: CV of dyads 2.4 and 2.12 measured in 0.2M Bu4NPF6 in CH2Cl2, vs.

Fc/Fc+.

As intramolecular charge-transfer interaction is not possible in the short-bridge

dyads 2.4 and 2.11, the donor ability of the TTF fragment in these compounds is not

perturbed (within 10 mV) by the ketone dicyanomethylene transformation. In

- 63 -

accordance with the fact that this charge-transfer interaction should disappear with

reduction of the acceptor fragment to the radical anion, the presence of the TTF moiety

has no influence on the second and third reduction waves of the fluorene nuclei. The

difference between the oxidation and reduction potentials (E1ox0 – E1red

0) for compound

2.4 is as small as 290 mV, which represents one of the smallest solution HOMO –

LUMO gaps (the smallest reported was 0.17 eV for dyad 2.2 [205]) so far reached for

closed-shell organic compounds [209].

2.5. Preparation of monolayer of 2.4 on air-water interface

Deposition of the dyads on the electrode surface was performed by the

Langmuir-Blodgett technique. The initial procedure for the dyads 2.4 and 2.11 was

elaborated in the group of Prof. Heath. Following their method, LB films were prepared

by spreading a fresh solution (10-3

M) of the molecule 2.4 in chloroform on a dilute

aqueous buffer subphase (5×10-4

M Na2CO3/NaHCO3)a. After the solvent evaporated the

more polar dicyanomethylene fluorene fragments of 2.4 were presumably exposed to the

polar water phase, whereas the hydrophobic trialkyl-TTF moieties are stretched into the

air. The surface area was decreased with Teflon barriers, which created a change in

surface pressure that can be recorded by microbalances to yield an isotherm curve (Fig.

2.3). When the surface area was decreased below about 60 Å2 per molecule it resulted in

a sharp increase of the pressure. Further decrease up to an area about 30–35 Å2 per

molecule leads to monolayer collapse. The non-uniform shape of the isotherm with a

shoulder at about 50 Å2

per molecule may be explained by the conformational flexibility

of 2.4, resulting in multiple molecular orientations, each with their own characteristic

molecular area.

Brewster angle microscopy (BAM) of the Langmuir monolayers revealed small

spots (~100 microns in diameter) on the trough surface that appear immediately upon the

dropwise addition of a solution of 2.4 in CHCl3 to the surface (Figure 2.3). These islands

a For the control experiment with eicosanoic acid 3.045×10

-4M CdCl2/6.415×10

-5 NaOH aqueous

subphase was used.

- 64 -

presumably correspond to molecular aggregates. The spots, however, spread out within

seconds to form a microscopically uniform surface. During the monolayer compression

with a speed of 2 cm2/min, the surface pressure was periodically held steady to check the

stability of the films. At a fixed pressure of 12 mN m-1

, the area dropped by less than

0.4% over 2 minutes, indicating a very stable monolayer. At higher pressures of 16 and

25 mN m-1

, there was an increased rate of monolayer relaxation (the trough area dropped

by 1.6 and 2.9 %, respectively, over a period 2 min). This is consistent with the

statement on conformational changes to 2.5 at the surface pressure 16-25 mN m-1

(or

surface area around 50 Å2).

Figure 2.3: Surface pressure isotherm of 2.4.

The few small spots (presumably dust particles), visible on the images (Fig. 2.4),

captured by the BAM remained steady on the surface for the entire two minutes,

indicating that the formed LB films were 2D solids. As stated above, the more polar

dicyanomethylene fluorene fragments of 2.4 were presumably exposed to the polar

water phase, whereas the hydrophobic trialkyl-TTF moieties are stretched into the air.

- 65 -

Figure 2.4: BAM images of a) clean Langmuir trough; b) same, immediately after

dropwise addition of a solution of 2.4 in CHCl3; c) LB film of 2.4 at 12 mN m-1

; d) LB

film of 2.4 at 26 mN m-1

(the picture width is 480 mm).

2.6. Deposition of the monolayers on solid substrates

The LB films, compressed to the desired surface pressure were transferred onto

three different substrates required for characterization and study experiments.

LB films transferred onto hydrophilic (oxygen-terminated) polycrystalline n-

doped silicon substrates, and

freshly cleaned hydrophilica[210] gold substrates with contact angle 42±4 (for

rectification study);

freshly cleaned Ge ATR crystal (for infrared characterization).

Deposition on the Si wafer was done at three values of surface pressure (12, 16

and 25 mN m-1

). The transfer direction (slides start out dipped into the water before the

addition of the molecules, and are then lifted up through the LB film, ―Z-type‖)

necessarily resulted in the acceptor moiety being exposed to the slide surface. The

transfer on the gold slides was done in two opposite direction: 1) ―Z-type‖ as described

a We should note that gold surface has two surface energy stated: with bulk atomic properties

(unreconstructed) and reconstructed surface with shorter atomic lengths. This could be a possible

reason for duality of the gold surface presenting hydrophilic and hydrophobic behavior

respectively [210].

- 66 -

above for the Si; and 2) ―X-type‖ resulting in the hydrophobic alkyl tails of the donor

being in contact with the gold surface.

Deposition was performed at constant rate (5 mm/min) with continuous control

of the surface pressure to insure uniformity of the deposited films. The transfer ratio for

the films deposited on the Ge crystal was always close to 1, while the transfer ratio for

the gold samples varied from sample to sample in the range from 0.4 to 1. This is

possible due to the non-uniformity of the slides’ surface. Normally one side on the slides

was covered with gold while other remains uncoated.

2.7. Spectroscopic characterization of LB monolayers

Infrared spectroscopy is very useful and highly sensitive technique to analyze the

nitrile stretching frequency depending on the charge accumulated on the molecule, and

can be utilized to monitor charge transfer interactions involving cyano-substituted

acceptors. For example, for TCNQ, CN = 2225 and 2180 cm-1

for the neutral and anion

radical species, respectively [211]. The CN of the dyad 2.4 (ca. 2203 cm-1

in KBr

pellets, fig 2.5) is substantially lower than that of similar fluorene acceptors without a

TTF moiety (2235 cm-1

) [204] and it could be attributed to significant charge transfer in

the solid state of 2.4. According to correlations of nitrile stretching frequencies with

degree of charge transfer (Z values) for TCNQ salts, Z-value for 2.4 equals

approximately 0.6 [211]. This value is very close to that of TTFTCNQ salt which

indicates the possibly of charge transfer in the solid state due to the π-π interaction

between the molecules.

To establish the preservation of chemical structure of 2.4 in transferred LB

monolayers, we have performed an infrared spectroscopic characterization of films

transferred (―Z-type‖) onto the surface of a Ge crystal by the attenuated total reflectance

(ATR) technique. We also investigated films transferred onto a gold substrate by

grazing-angle reflection–absorption infrared spectroscopy (RAIRS). Comparison with

the spectrum of bulk 2.4 (powdered in KBr, Figure 2.5), and of spectra of compounds

containing separate TTF and fluorene fragments, reveals the presence of all main

absorption peaks, suggesting the structural integrity of the transferred molecules (see

- 67 -

Table 2.2 and Fig. 2.5). In spite of certain differences (such as much broader lines in the

LB film and somewhat different relative intensities) expected for different molecular

orientations and intermolecular interactions in the LB film and in the crystal, one can

clearly see the presence of a potentially vulnerable ester group (C=O, C–O), the

electron-acceptor fluorene fragment (cyano and nitro-groups as well as aromatic C=C

bonds), and the electron-donor trialkyl-TTF fragment (CH2, CH3).

3500 3000 2500 2000 1500 1000

KBr

wavenumber (cm-1

)

LB film

Figure 2.5: Infrared spectroscopy of the dyad 2.4 in bulk (KBr pellets) and LB film on

gold substrate.

The position of the significantly broadened CN band (2205 cm-1

, similar to that in the

bulk sample) is between those for a completely neutral (2225 cm-1

) and radical-anion

species (2180 cm-1

) [204], suggesting a partial charge transfer (either from the TTF

fragment or from the Ge surface). The lack of strong characteristic absorption bands of

the TTF core precludes a detailed analysis of this fragment, however, its presence is

observed from the strong aliphatic C–H stretching at ~2900 cm-1

. Also, the absence of a

sulfoxide bond (at ca. 970–990 cm-1

) [212], expected for S-oxidized species, suggests

- 68 -

that no irreversible oxidative decomposition of this fragment took place (although one

cannot exclude a reversible formation of a TTF radical cation).

Table 2.2: Assignment of the major IR absorption peaks of 2.4 in LB film and in bulk,

their relative intensity (IX/INO2) as well as the dichroic ratio (RATR), order parameter

<P2> and average tilt angle () obtained from polarization experiments.

Peaks Bulk, cm–

1

LB, cm–1

IX /INO2

Bulk

IX /INO2

LB

RATR=As/Ap <P2> []

CHar 3106,

3093

3096 0.03 0.08

a CH3

a, CH2

s CH2

2956

2929

2859

2959

2928

2856

0.10

0.12

0.06

0.39

0.41

0.23

1.130.01

1.060.01

1.030.01

–0.010.03

–0.160.02

–0.210.03

55

61

65

CN 2203 2205 0.27 0.09

C=O 1735,

1718

1728 0.19 0.55 1.110.00 –0.060.00 57

C=Cring 1601,1579 1607 1591

1577

0.12 0.48 1.100.02 –0.050.03 57

a NO2 1534 1536 0.43 0.87 1.080.01 –0.120.02 60

C=C 1512 – 0.24

a CH3 1456 1457 0.19 0.45 1.070.01 –0.130.03 60

C-O 1422 – 0.48

1403 – 0.44

s CH3 1379 – 0.72

s NO2 1336 1341 1.00 1.00 1.090.01 –0.090.03 58

COC 1281 (1288) 0.53 0.27

COC 1255 (1262) 0.39 0.21

COC 1235 (1242) (1222) 0.23 0.19

a COC 1207 – 0.31

a COC 1185 (1190) 0.53 0.18

ra CH3 1151 1155 0.51 0.27

ra CH3 1102 (1102) 0.19 0.16

s COC 1089 (1081) 0.15 0.15

1075 (1058) 0.11 0.13

C-CH3 1034 (1025) 0.14 0.11

To study the structure of LB film of 2.4, an orientation analysis of the films

transferred on a Ge crystal was performed in collaboration with Prof. M. Pezolet (Fig.

2.6).

- 69 -

Figure 2.6: Geometric representation of the determination of the molecules orientation

by ATR spectroscopy. S (transition moment) located at angle from c (molecular axis)

which is inclined at angle from Z (normal to the crystal surface); is an angle between

the transition moment S and Z [213].

ATR-FTIR spectra were thus recorded by using polarized infrared radiation, and

the dichroic ratio (RATR, eq. 2.1) was calculated from the absorbance of characteristic

bands for 2.4, obtained with the infrared radiation polarized parallel (Ap) and

perpendicular (As) to the plane of incidence.

2.1

The associated order parameters of the transition moment of a given vibration

with respect to the film normal <P2> can be calculated by using mean-square electric

field amplitudes (E2

x,y,z) obtained from the Harrick thin-film equations [214].

⟨ ⟩

2.2

The order parameter of the molecular axis, <P2(cos)>, can be calculated from

the order parameter of the transition moment, <P2> using the Legendre addition

theorem:

- 70 -

⟨ ⟩ ⟨ ⟩

⟨ ⟩ ⟨ ⟩

2.3

where is the angle between the molecular axis and the normal to the ATR

crystal, and is the angle between the transition moment and the molecular axis.

Assuming that orientation distribution of the molecular axes is infinitely narrow,

the tilt angle can be calculated from the eq. 2.3, as:

√ ⟨ ⟩

2.4

The values of RATR, <P2> and the average angle γ between the transition

moment and the surface normal for the films transferred at the highest pressure (25

mNm-1

) are given in Table 2.2 for the major bands. It is important to remember here that

<P2> should be equal to zero for an isotropic sample (γ=54.9, the magic angle), and to

one and –0.5 for perfect orientation of the transition moments (S) parallel (γ =0) or

perpendicular (γ =90) to the surface normal, respectively [213, 215]. Table 2.2 shows

that the order parameter differs significantly from zero for several bands. For example, a

<P2> of about –0.2 is observed for the two methylene C-H stretching modes, revealing

that the CH2 groups are preferentially oriented in the plane of the LB film. Even though

the high wavenumber position of the maximum of the 2859 and 2929 cm-1

bands shows

that the alkyl chains are significantly disordered [149], as expected from the molecular

model of the films (Figure 2.7), the polarized ATR results indicate that they are

preferentially oriented along the surface normal with an average tilt angle of

approximately 30. On the other hand, the dichroic behaviour of the 2960 cm-1

band

indicates that the methyl groups are unoriented. The values for the angle of about 60

found for peaks corresponding to C=C and C=O double bonds and to the symmetric

stretching vibration of the NO2 groups at 1341 cm-1

(for which the transition moment is

bisector of the NO2 angle) shows unequivocally that the fluorene moiety is not lying flat

on the surface, although a more precise determination of its orientation is precluded by

the complexity of the structure (the presence of several similar bonds with different

orientations). It is worth noting that reproducibility of the orientation measurements

- 71 -

(presented by the absolute error in Table 2.2 calculated for three different films), while

being quite acceptable for the films transferred at 25 mN m-1

was persistently very low

for the films transferred at lower pressures, as expected for poorly aligned low-density

films.

Figure 2.7: The top (left) and side (right) views of a model of an LB film of 2.4,

corresponding to a molecular area of 50 Å2 (the TTF-fluorene core is the DFT optimized

minimum energy conformer; the alkyl chains are added and optimized with the

molecular mechanics force field MM+; the molecules are manually placed in the closest

position, respecting van-der-Waals distances; the hydrogen atoms omitted for clarity).

2.8. Fabrication and electrical studies of n-Si/SiO2/2.4/Ti junction devices

Figure 2.8 shows energy levels for the studied junctions of the dyad 2.4.

According to the Aviram-Ratner proposal the preferred direction of the electron flow

should be from acceptor to donor (M2 to M1). The bias necessary to reach resonance

between work functions of the electrodes M2 and the LUMO of the acceptor is

significantly smaller than that necessary for the resonance between M1 and LUMO of

the donor. Notably, as the work function of the mercury electrode lies between energies

of HOMO and LUMO of the molecule and the difference between them is very small,

such system can be ideal for the future study of electronic properties of this dyad in Hg-

molecule-Hg junctions.

The molecular rectification of the dyad 2.4 initially was studied in n-Si/dyad/Ti

junctions in Prof. Heath’s group (Fig. 2.9). The monolayer of 2.4 was deposited on

degeneratively n-doped Si slide by LB technique (see section 2.6). Then, a second

- 72 -

electrode (10 nm Ti followed by 4 m of Al) was deposited by evaporation on top of the

transferred monolayers to complete the fabrication of molecular tunnel junction devices.

Figure 2.8: Scheme of the molecular rectifier based on the dyad 2.4. The energy levels

of the molecule are calculated using DFT. Work functions for the electrodes are taken

from the literature.

The emerging picture of the deposition of metallic thin films, such as Ti and Au,

on molecular monolayers, is that the thickness and stoichiometry of the metallic film, as

well as the structure of the molecular monolayer, all play critical roles in determining the

extent to which the molecules are modified or damaged by the deposition [182, 216-

218]. Titanium is a unique metal for an evaporated top electrode. Due to its high

reactivity, it immediately cleaves terminal C–H bonds forming a thin titanium carbide

layer on the surface of the monolayer and may prevent further penetration of the Ti

atoms inside the film [182], as often observed for gold [217-220]. Preservation of the

molecular features buried inside the aliphatic chain protected monolayers has been

demonstrated by both X-ray photoelectron spectroscopy [216], and more recently by

RAIRS [182], which showed disappearance of only terminal CH3 vibrations, whereas all

- 73 -

other infrared spectral features were unperturbed. At the same time, evaporation of a Ti

layer of >30 Å on self-assembled monolayers (SAMs) of predictably more reactive

conjugated compounds such as oligothiophenes [218] or oligo(phenylethynylene)s [217]

can result in complete destruction of the molecules. SAMs are also lower-density

molecular monolayers than compressed LB films, which would result in higher tilt

angles and exposure of larger part of the molecules to the incident Ti flux. During the

evaporation of the Ti we have tried to account for these facts by putting protecting alkyl

chains on the TTF moiety against the Ti and by using highly compressed LB films.

Figure 2.9: Scheme of the n-Si/2.4/Ti junction studied in Heath’s group.

We are also well aware of the criticism toward the claims of molecular

rectification from junctions based on oxidizable metal contacts. The titanium oxide

induced rectification was first pointed out by Ashwell et al. in 1980 [221] and a number

of rectifying Ti-based junctions for molecules lacking an evident ―diode‖ structure have

been reported to date [182, 222-223]. As was shown in these reports, such rectification

depends upon the level of oxidation at the molecule/Ti interface and can be controlled

by the level of vacuum used during the deposition of the Ti. Using a sufficiently high

quality e-beam deposition system (providing a vacuum of 5×10-7

Torr), we are able to

routinely control the vacuum, both to increase and to decrease the rectification. As

shown below in control experiments, under the correct Ti deposition conditions, and for

a degeneratively doped poly-Si bottom electrode, such rectification can be effectively

suppressed and experimentally separated from the molecular features.

Acceptor

Linker

Donor

e

Ti

Si

- 74 -

The work functions of n-doped Si (–4.85 eV) and Ti (–4.33 eV) are similar to

each other (to minimize the rectification due to p-n junction), and also fit reasonably

well with the HOMO/LUMO levels of 2.5. As described below, any rectification arising

from the dissimilar poly-Si and Ti electrode materials or formation of an oxide layer

[221] can be experimentally separated from that which arises from the molecular

component.

Current-voltage curves obtained for these devices are depicted in Figure 2.10.

Notably, as the transfer pressure goes up, the magnitude of the current through the

junction decreases, implying an increasing distance between the top and bottom

electrodes as the monolayer aligns. The most dramatic effect, however, is that the

rectification ratio (RR) sharply increases with increased transfer pressure, from 1.5 for

56 Å2 molecule

–1 to 5 for 52 Å

2 molecule

–1 and 18 for 43 Å

2 molecule

–1. These

observations are in agreement with the alignment of molecules of 2.4 during

compression to form a well-packed monolayer with the D-A / surface angle being close

to normal.

Figure 2.10: The current–voltage curves obtained from the n-Si/2.4/Ti molecular tunnel

junction devices made from the three films indicated in the isotherm measurements.

Note that as the area per molecule decreases, the rectification ratio increases (by a factor

of 10), but the current decreases. Both the increased current rectification and the

decrease in current magnitude are indicative of an increased alignment of the molecular

monolayer.

-5

-4

-3

-2

-1

0

-1 -0.5 0 0.5 1

Applied Bias (V)

56 Å2

52 Å2 (x8)

43 Å2 (x35)

Cu

rre

nt

(mic

roA

mp

s)

0

20

40

60

10 30 50 70

Area per Molecule (Å2)

Su

rfac

e P

res

su

re(m

N·m

-2)

collapse

Transfer points

-5

-4

-3

-2

-1

0

-1 -0.5 0 0.5 1

Applied Bias (V)

56 Å2

52 Å2 (x8)

43 Å2 (x35)

Cu

rre

nt

(mic

roA

mp

s)

-5

-4

-3

-2

-1

0

-1 -0.5 0 0.5 1-1 -0.5 0 0.5 1

Applied Bias (V)

56 Å2

52 Å2 (x8)

43 Å2 (x35)

Cu

rre

nt

(mic

roA

mp

s)

0

20

40

60

10 30 50 70

Area per Molecule (Å2)

Su

rfac

e P

res

su

re(m

N·m

-2)

collapse

Transfer points

- 75 -

Note that the only thing that is changing for the devices in Figure 2.10 is the area

per molecule, which translates into molecular orientation. Ellipsometry measurements of

a film transferred at 17 mN m–1

(~50 Å2 molecule

–1) suggested a thickness of 15-20 Å,

consistent with the formation of a monolayer of the most stable conformer (shown in

Figure 2.7). This conformer has a calculated thickness and molecular area of 21 Å and

50 Å2, respectively (Figure 2.7). Further compression at pressures above ~17 mN m

–1

would require conformational changes, and the film transferred at 26 mN m–1

most likely

has both fluorene and TTF fragments perpendicular to the surface. Thus, any current-

voltage asymmetry that might arise from the dissimilar electrodes, titanium oxide

formation, etc. is effectively a constant through this series of devices. Nevertheless, we

checked this conclusion by preparing similar devices containing an eicosanoic acid LB

monolayer in place of 2.4. In a number of experiments, these devices yielded a RR close

to 1, and never more than 1:2 – 1:3. Furthermore, no dependence of the rectification

ratio on the transfer pressure was found for eicosanoic acid, although the total current

was also observed to decrease for films transferred at increasing values of area/molecule.

Therefore, we assign the dominant contribution to the observed rectification (1:18) in the

monolayer of 2.4 as a molecular feature.

The maximum RR for a n-Si/2.4/Ti device is achieved at the relatively low

potential of 0.9 V. Above 1.0 V the rectification ratio decreases (Figure 2.11). This is

expected, because at sufficiently high applied bias, direct tunneling of charge carriers

between the two electrodes becomes an increasingly important (and eventually

dominant) mechanism of charge transport. In other words, at sufficiently high bias, the

specific details of the molecule become less important. The RR does not decrease (it

actually slightly increases) after 10 scans (up to 1.75 V). This is in contrast to D––A

systems in which the reorientation of highly polar molecules reduced the RR by a factor

of two every second scan [47, 55]. The direction of the observed rectification indicates

that the preferred electron current is from fluorene acceptor to TTF donor (from Si on

Ti).

- 76 -

Figure 2.11: The dependence of the rectification ratio in the n-Si/2.4/Ti device on the

applied bias for a series of voltage cycles.

2.9. Fabrication and electrical studies of Au/2.4/Hg junction devices

The previously discussed rectification study of the dyad 2.4 in n-

Si/SiO2/molecule/Ti junctions [224] has few point of criticism. Briefly they are: (i) the

rectification is known for the junctions based on oxidizable metal contacts [28]; (ii) the

titanium oxide induced rectification was first pointed out by Ashwell et al. in 1980

[221], and a number of rectifying Ti-based junctions for molecules lacking an evident

―diode‖ structure have been reported to date [113-114, 182]; (iii) it was shown before

[113-114, 221], such rectification depends upon the level of oxidation at the molecule/Ti

interface and can be controlled by the level of vacuum used during the deposition of the

Ti. Although the evaporation of Ti was performed in high vacuum (<10-7

mbar), and the

pressure-dependence of RR shows the importance of molecular alignment, we cannot

completely rule out the possibility of small levels of oxidation, during the fabrication or

measurements of the dyad.

To further verify the molecular nature of the rectification behaviour displayed by

2.4, we performed an electrical measurement of the LB monolayer of 2.4 between

higher-work-function gold and mercury electrodes (5.3 eV and 4.49 eV, respectively).

- 77 -

Such junctions are different from the n-Si/SiO2/molecule/Ti junctions because no oxides

are present on the metal surface (Au) and because the junctions themselves are relatively

simpler. The mercury electrode was covered with protecting alkyl thiol monolayer to

prevent direct contact of two metal surfaces through defects in the organic film (Fig.

2.12).

The LB film of 2.4 was transferred onto a gold substrate in the previously

described fashion (―Z-type‖, Au/fluorene--TTF interface, Fig. 2.12 a) [224], and the

electrical junction was established by micromanipulator-controlled contacting with a

hexadecanethiol-protected hanging mercury drop electrode. The dense defect-free Hg-

SC16H33 monolayer prevents electrical shorts (due to possible defects in the LB layer)

and formation of radical-ion salts on the mercury surface. Similar to Si/2.4/Ti devices,

the current-voltage response of the Au/2.4/C16H32S-Hg junction is highly asymmetric

(Fig. 2.13) with higher current at forward bias, which corresponds to the electron flow

from Hg to Au (i.e. from donor to acceptor), i.e. opposite to that predicted by AR model.

The unimolecular origin of the rectification was confirmed when the junction

consisted of the LB film deposited with opposite orientation: Au/TTF--fluorene/Hg

(Fig. 2.12b). Also, in a number of experiments performed, the devices containing an

eicosanoic acid LB monolayer in place of 2.4 yielded a RR of approximately 1.5–2, and

never more than 3. The more important fact is that the direction of the electrons flow has

changed to the opposite direction (from Au to Hg, at forward bias). This confirms that

the direction of the current flow strongly depends on the molecular orientation within the

LB film. The difference in the rectification direction for Si/Ti and Au/Hg (AD and

DA, respectively) could lie in the extremely low HOMO–LUMO gap of 2.4, which in

specific junction devices may change the ground state from a neutral TTF-σ-fluorene to

a zwitterionic TTF+

--Fluorene– (in which the TTF

+ becomes an acceptor and

Fluorene–

becomes a donor). Whatever the case, the results do highlight the important

role that the electrodes play in determining the current–voltage response of a molecular

electronic device. Although the exact mechanism of the different rectification directions

in these two studied junctions is certainly disputable, we believe that the molecular

- 78 -

origin of such behaviour is adequately proved by the above molecular reorientation and

alignment studies as well as control experiments with eicosanoic acid.

Figure 2.12: Mercury drop junction of the dyad 2.4 a) junction with ―Z-type‖ deposited

LB monolayer on gold substrate; b) junction with ―X-type‖ deposited LB film; c) photo

of the mercury-drop junction device.

- 79 -

Figure 2.13: I–V characteristics of Au/2.4/C16H32S-Hg junction devices for different

molecular orientations.

Conclusions

To summarize the work discussed in this Chapter, we have prepared and

characterized the first molecular tunnel junctions based on a TTF--fluorene dyad with

an extremely low HOMO–LUMO gap (0.29 eV). We discussed results of complete

study and characterization of the structure of LB film transferred on ATR crystals and

gold surface as well as orientation of the molecules in the monolayer by IR and other

spectroscopic methods. The rectification behaviour was studied in two types of

junctions: n-Si/dyad/Ti and Au/dyad/C16SH-Hg. The rectification ratio was found to

increase rapidly to 1:18 upon alignment of the molecules in compressed Langmuir–

Blodgett monolayers. An opposite rectification direction was found for n-Si/dyad/Ti and

Au/dyad/C16SH-Hg junctions experiments. However, the molecular origin of the

rectification was confirmed in Au/molecule/Hg tunnel junction by changing the

orientation of the molecule (from D--A to A--D) and by control experiments with

eicosanoic acid in case of n-Si/dyad/Ti.

- 80 -

Experimental section

Preparation of Langmuir–Blodgett (LB) films: Single monolayers were prepared at

20C on an aqueous (18.2 MOhm H2O) subphase by using either a 600 cm2 Nima 611D

(Nima Technology, Coventry, UK) or a 400 cm2 KSV 3000 (KSV Instruments, Helsinki,

Finland) Langmuir–Blodgett (LB) trough. Images of the Langmuir films were recorded

with a Nanofilm Surface Analysis Brewster angle microscope (BAM) (Gӧttingen,

Germany). For compound 2.4, a dilute buffer (5×10–4

M Na2CO3/NaHCO3) was

employed as a subphase in the trough to prevent acid-catalyzed oxidation of the TTF

units. For the eicosanoic acid controls, a 3.045×10–4

M CdCl2/6.415×10–5

M NaOH

aqueous subphase was used. The molecules were first dissolved in slightly basic, freshly

distilled chloroform (~0.5 gL–1

) and then immediately spread to the subphase to form the

monolayer. After an equilibrating period of 30 min allowing solvent evaporation, the

monolayer was compressed at constant speed of 10 mmmin_1 and transferred at

constant surface pressure onto the surface of interest (n-Si or Au electrodes for I–V

experiments or Ge crystal for ATR experiments).

Characterization of 2.4 in monolayers: Contact angle and ellipsometry measurements

for the monolayer were taken by depositing the monolayer on a bare Si<111> wafer and

transferring at 17.0 mN m–1

. Contact angle measurements were obtained by using a

Ramé Hart goniometer 100–00. Ellipsometry measurements were obtained using a

Gaertner L116B Ellipsometer equipped with a He-Ne laser at 632.8 nm; a refractive

index of 3.842 and an extinction coefficient of –0.016 were used for the silicon

substrate. The refractive index of the monolayer was assumed to be 1.46 with an

extinction coefficient of 0.00.

To perform ATR infrared measurements, single monolayers of 2.4 were transferred at a

constant speed of 5 mm min–1

onto Ge parallelograms (angle of incidence 45C) of

50×20×2 mm, allowing 24 internal reflections. Before deposition, the substrates were

cleaned with chloroform and methanol, immersed in chloroform in a Branson 1510

ultrasonic bath (Branson Ultrasonics Corporation, Danbury, CT) for 5 min, and put in a

plasma cleaner (Harrick Scientific, Ossining, NY) for 2 min. Finally, dust was removed

with a nitrogen gas flow. The germanium crystals were placed in a vertical ATR

accessory (Harrick Scientific, Ossining, NY) and the spectra were recorded using Magna

550 FTIR spectrometer (Thermo-Nicolet, Madison, WI) equipped with a liquid-N2

cooled MCT detector. A motorized rotating ZnSe wire-grid polarizer (Specac,

Orpington, UK) was positioned in front of the sample to obtain parallel- and

perpendicular-polarized spectra without breaking the purge of the spectrometer. A total

of 500 scans at 4 cm–1

were sufficient to achieve a high signal-to-noise ratio. RAIRS

spectroscopy was performed for monolayers of 2.4 transferred onto gold substrates by

using a Nexus 670 FTIR spectrometer (Thermo-Nicolet, Madison, WI) equipped with a

liquid-N2 cooled MCT-II detector and grazing angle (80C) Smart-SAGA accessory.

The measurements were done in an atmosphere of dried, CO2-free air, and an identical

gold-covered slide (prepared in the same Au-evaporation run), freshly cleaned by

soaking in HPLC-grade dichloromethane and drying in vacuo, was used to record a

background spectrum.

Fabrication and studies of Si/2.4/Ti junctions: The process for the fabrication of the

solid-state molecular diode tunnel junctions follows the methods described previously

[225-226]. For the bottom electrodes, a layer of n-type polycrystalline (poly-Si) was

- 81 -

formed by means of direct chemical vapour deposition growth onto <100> Si wafers

coated with oxide. The poly-Si was then etched into 5 μm wide electrodes by using

standard optical lithography techniques. The LB monolayer of 2.4 was transferred onto

the silicon substrate by X-type deposition (the substrate was lifted up from the

subphase). A top electrode of 10 Å titanium, followed by 4000 Å aluminum, was then

deposited on top of the monolayer by electron-beam evaporation at a residual pressure of

~5×10–7

Torr. Current–voltage characteristics were taken in air at room temperature by

using a shielded probe station with coaxial probes. For Si/2.4/Ti junctions, bias voltages

were applied to the polysilicon electrode, and the top metal electrode was connected to

ground through a Stanford Research Systems SR570 low-noise current preamplifier.

Fabrication and studies of Au/2.4/C16H33S-Hg junctions: The junction was assembled

in a procedure, similar to the described before [222-223]. A gold layer (~200 nm) was

thermally evaporated on <100> Si wafers or freshly cleaved mica slides. Before the LB

film transfer, the gold surface was cleaned by 10 min exposure to O2-plasma followed

by immersing in HPLC grade ethanol to decompose the formed oxides [227]. The LB

film of 2.5 was transferred (always at 25 mN m–1

) on thus prepared substrate in X or Z

deposition mode, resulting in formation of Au/fluorene--TTF or Au/TTF--fluorene

sandwiches, respectively. The gold substrate was put under deionized water (to improve

the stability of the junction; ion-exchange purification followed by distillation was

employed to reduce the water conductivity, whereas no other solvent could be used due

to LB film instability). A hanging drop of mercury (from a microsyringe, ~500μm in

diameter), covered with a monolayer of hexadecylthiolate by 15 min exposure to a

solution of C16H33SH in ethanol and rinsed with fresh ethanol, was brought into contact

with the monolayer of 2.4 under the water, by using a micromanipulator. The substrate

was grounded, the bias voltages were applied to the mercury electrode (two-electrode

scheme), and the I-V characteristics were recorded with a potentiostat EG&G PAR273A

(sensitivity 0.1 nA) at a scan rate of 1000 mVs–1

and sampling rate of 20 mV per point.

Calculations: The geometry optimization was performed at the DFT (RB3LYP) level of

theory, by using the 6–31G(d) basis set, as implemented in Gaussian 03 [228]. The

calculated RB3LYP wavefunction was found to be stable according to the Gaussian

stability test. The frequency check for all conformations was used to confirm that they

are true minima.

5-Methyl-4,5-dipentyltetrathiafulvalene-4-carbaldehyde (2.6): LDA (0.77 mL of a

1.8 M solution in THF/heptane, 1.38 mmol) was added at -78 °C to a solution of TTF

derivative 2.5 [205] (0.4 g, 1.1 mmol) in anhydrous diethyl ether (20 mL), the reaction

mixture was stirred for 2 h at -78 °C followed by addition of N-methylformanilide (0.17

mL, 1.38 mmol). The reaction mixture was then warmed to 20 °C overnight and

quenched with ice-water (acidified with AcOH). The organic layer was separated,

washed with water and brine, dried over MgSO4 and, after evaporation, was purified by

chromatography on silica gel, eluting with ethyl acetate/hexanes (1:3 v/v). The red

fraction was evaporated and dried in vacuo to give aldehyde 2.6 (0.24 g, 58 %) as a dark

red solid. M.p. 54-57 °C; 1H NMR (300 MHz, acetone-d6): δ=9.77 (s, 1H), 2.54 (s, 3H),

2.44 (t, J=7.5 Hz, 4H), 1.58-1.45 (m, 4H), 1.40-1.26 (m, 8H), 0.89 (t, J=6 Hz, 6H); 13

C

NMR (75 MHz; acetone-d6): δ=180.4, 153.9, 134.3, 130.1, 129.7, 113.1, 103.7, 31.9,

30.20, 30.18, 29.2, 23.1, 14.27, 14.19.

- 82 -

4-Hydroxymethyl-5-methyl-4,5-dipentyltetrathiafulvalene (2.7): Sodium

borohydride (30 mg, 0.78 mmol) was added to a solution of aldehyde 2.6 (0.24 g, 0.62

mmol) in anhydrous EtOH (15 ml) and the reaction mixture was stirred at 20 °C for 1 h

(the red color of 2.8 vanished within 10 min). Then ethyl acetate (20 mL) was added, the

organic layer was washed with water and brine, dried over MgSO4, and filtered through

a 1 cm pad of silica gel. Evaporation of the filtrate gave alcohol 2.7 (0.21 g, 89 %) as an

amorphous solid. M.p. 59-62 °C; 1H NMR (200 MHz; acetone-d6): δ=4.42-4.30 (m, 3H),

2.41 (t, J=7.5 Hz, 4H), 1.99 (s, 3 H), 1.60-1.42 (m, 4H), 1.42-1.24 (m, 8H), 0.90 (t, J=6

Hz, 6H); 13

C NMR (50 MHz; acetone-d6): δ=131.2, 129.75, 129.68, 125.7, 108.1, 107.7,

58.1, 31.9, 30.2, 29.2, 23.1, 14.3, 13.6.

2,5,7-Trinitro-4-chlorocarbonylfluorene-9-one (2.9) was obtained as described in

[229] from 2.8. 1H NMR (200 MHz; CDCl3): δ=9.20 (d, J=2 Hz, 1H), 8.81 (d, J=2 Hz,

1H), 8.65 (d, J=2 Hz, 1H), 8.54 (dd, J=8, 2 Hz, 1H), 8.42 (d, J=8 Hz, 1H).

Dyad 2.11: BuLi (1.6 M; 0.36 mL, 0.57 mmol) was added at –78°C to a solution of

compound 2.7 (210 mg, 0.543 mmol) in dry THF (10 mL). The reaction mixture was

then cooled to -100 °C, a solution of acid chloride 2.9 (250 mg, 0.0.68 mmol) in dry

THF (5 mL) was added dropwise and reaction mixture was stirred for 1 h at -100 °C,

then allowed to warm up overnight in a freezer (-15 °C). The solvent was removed in

vacuum and the residue was dissolved in ethyl acetate. The organic layer was

subsequently washed with water, NaHCO3 solution and brine, dried over MgSO4 and

evaporated. Flash chromatography on silica gel eluting with hexane/ethyl acetate (3:1

v/v) gave a green fraction. After the solvent was evaporated the, the residue was

recrystallized from acetone resulting in 2.12 (18 mg, 50 %). M.p. 140 °C; 1H NMR (300

MHz, acetone-d6): δ=9.02 (d, J=2 Hz, 1 H), 8.84 (d, J=2 Hz, 1H), 8.80 (d, J=2 Hz, 1H),

8.71 (d, J=2 Hz, 1H), 5.13 (s, 2H), 2.40 (m, 4H), 2.19 (s, 3H), 1.50 (m, 4H), 1.32 (m,

8H), 0.89 (m, 6H); 13

C NMR (75 MHz; acetone-d6, 50 °C): =186.0, 165.1, 150.9, 150.8,

147.4, 144.4, 140.4, 140.2, 139.4, 133.8, 132.3, 131.3, 130.0, 126.5, 123.1, 122.7, 122.4,

61.6, 32.05, 32.03, 30.2, 23.1, 14.3, 14.2; IR (KBr): ν=1731 (C=O), 1614, 1593, 1537,

1465, 1342, 1174, 739 cm-1

.

Dyad 2.4: Fluorenone 2.11 (100 mg, 0.136 mmol) and malononitrile (12 mg, 0.19

mmol) were dissolved in DMF (3 mL) and stirred at 20 °C for 8 h. Then the solvent was

distilled off in vacuo and the residue was dissolved in acetonitrile (3 ml) and diluted

with methanol (20 ml). The brown precipitate was filtered off and washed with methanol

giving compound 2.4 as a black solid (77 mg, 75 %). M.p. 160 °C; 1H NMR (300 MHz;

acetone-d6): δ=9.9-9.3 (br, 2 H, fluorene-H), 5.15 (br s, 2H), 2.40 (br s, 4H), 2.19 (br s,

3H), 1.50 (m, 4H), 1.31 (m, 8H), 0.89 (m, 6H); IR (KBr) ν=2204 (C≡N), 1718 (C=O),

1605, 1535, 1340cm-1

; MS (FAB): m/z (%): 777 (35).

- 83 -

Chapter III. Self-Assembled Monolayers of Strong Electron Acceptors:

Polynitrofluorenes on Gold

(Part of this Chapter was adapted with permission from: D.F. Perepichka, M.

Kondratenko, M.R. Bryce, Self-Assembled Monolayers of Strong Electron Acceptors:

Polynitrofluorenes on Gold and Platinum, Langmuir 2005, 21, 8824–8831. Copyrights

2005 American Chemical Society)

Introduction

Strong chemical binding of π-functional organic molecules to metal surfaces

resulting in the formation of self-assembled monolayers has been a major focus of recent

research in connection with molecular electronic devices (molecular diodes, switches,

wires, memories, etc.) [184, 230-231], electrochemical sensors [232-236], electrode

modification for OLEDs [237], and photovoltaics [238-240].

A number of moderate and strong electron donor molecules, including ferrocene

[165, 241-242], TTF [233, 235, 243-248], and its π-extended analogue [249],

oligothiophenes [250-253], N-alkylcarbazole [254], pyrene [255], tetraalkylphenylene-p-

diamine [256], and porphyrin derivatives [255, 257-260] have been attached to gold

surfaces. The ability of these monolayers to release electrons forming stable cationic

states is detrimental for such applications as photocurrent generation, memory devices,

switches, cation sensors, etc. (as all these processes include formation of ion radicals in

the device operation process). Significantly less is known about complementary SAMs

consisting of electron acceptor molecules. Most of the reported self-assembled π-

electron acceptors possess only moderate electron affinity (Ered < –0.6V vs. Fc/Fc+) viz.

fullerene [252, 261-265], p-benzoquinone [266-267], naphthalene-1,4,5,8-

tetracarboxylic diimides [268], phthalocyanine [269], and viologen derivatives [230,

270-272]. Before this study was started, the only example of a strong electron acceptor

used in preparation of SAMs is a (TCNQ-C10S)2 derivative, reported by Frisbie et al

[256]. This monolayer has been used for direct determination of charge-transfer

complexation (using a tetraalkylphenylenediamine-covered AFM tip) [273]. Also a

molecular junction was made of (TCNQ-C10S)2 by sandwiching SAM between Ag and

alkylthiol-covered Hg drop electrodes [37]. Although the mechanism for current

- 84 -

rectification in this system is controversial molecular layers with such a low LUMO

energy are certainly of interest for a number of applications in molecular electronics and

surface science. One can also envisage the SAM-forming (TCNQ-C10S)2 as an

intermediate toward analogues of the original Aviram-Ratner molecular rectifier TTF-

TCNQ. However, the relatively low stability of (TCNQ-C10S)2 and no easy way for

further synthetic modification have prevented its widespread use in molecular

electronics and related applications.

Previously, Perepichka et al demonstrated that the polynitrofluorene electron

acceptor moiety, by virtue of its very high electron affinity and synthetic versatility, is a

convenient building block for the construction of donor-acceptor dyads with extremely

low HOMO-LUMO gaps [204]. In Chapter 2, we have established molecular

rectification in Langmuir-Blodgett monolayers of nitrofluorene-TTF dyad physisorbed

on metal surfaces [274]. Here, we describe the immobilization of nitrofluorene electron

acceptors on gold and platinum electrodes by means of covalent bond formation with the

surface and characterization of the corresponding SAMs. A rectification study of the

junctions formed by sandwiching the SAMs of the nitrofluorene derivatives between

gold electrode and thiol coated mercury drop is also presented.

3.1 Synthesis

Tuning acceptor properties of the fluorene derivatives with different electron-

withdrawing substituents was a subject of interest over past decades [229]. Our

particular interest is in electron-withdrawing substituents that allow for further

modifications of the fluorene moiety (such as coupling with other functional groups) and

do not dramatically reduce the acceptor properties of the molecule. Substitution of nitro

groups in the nitrofluorenes with different nucleophilic reagents and particularly with

alkyl mercaptans was developed by I. F. Perepichka et al. [275] resulting in the

- 85 -

introduction of electron-withdrawing sulfanyl groups into the aromatic core. Placing

different functional groups on the alkanethiol reagents (for example hydroxyl and

carboxyl groups) potentially allows further coupling of the molecule with another

electroactive moiety or surface ―anchor‖ groups.

The first step of TNF (3.1) functionalization was achieved by regioselective

nucleophilic substitution of a nitro group with commercially available 3-

mercaptopropanol in presence of NaHCO3 affording sulfide 3.2 (Scheme 3.1). The

product 3.2 has bright red color as a result of intramolecular charge transfer from sulfur

atom (donor) onto acceptor fluorene moiety.

The effect of an electron-donating alkylthio substituent in 3.2 (which reduces the

total electron affinity of the molecule) can be partially eliminated by oxidizing the

sulfide group into the electron-withdrawing sulfonic group in the presence of hydrogen

peroxide (Scheme 3.1) resulting in compound 3.3 as a yellow solid.

Scheme 3.1: Substitution of nitro group in TNF with sulfone.

The anchor disulfide functionality can be easily introduced in the molecule by

DCC (N,N'-Dicyclohexylcarbodiimide) promoted esterification of the terminal hydroxyl

group with thioctic acid resulting in the corresponding thioctic ester 3.4 (Scheme 3.2).

- 86 -

Scheme 3.2: Coupling of acceptor synthon with thioctic anchor group.

As was shown previously [204, 274], a unique feature of fluorene acceptors is

the possibility of converting chemically stable fluorene-9-one derivatives into the

stronger electron accepting fluorene-9-dicyanomethylene derivatives under very mild

conditions. Thus, treating a solution of 3.4 in DMF with malononitrile gave acceptor 3.5

in good yield, 80 % (Scheme 3.3).

Scheme 3.3: Conversion of fluorene-9-one into dicyanomethylene derivative.

A donor TTF moiety with a disulfide functional group (as a model compound to

study relative stability of the SAMs of donor and acceptor and their influence on each

other within the same monolayer) was also synthesized from hydroxyl functionalized

TTF synthon 3.6 and thioctic acid [276], following DCC coupling protocol (Scheme

3.4).

Scheme 3.4: Synthesis of self-assembly functionalized TTF derivative.

- 87 -

3.2 Formation of SAMs of the fluorene derivatives 3.4 and 3.5

The disulfide functionality in fluorenes 3.4 and 3.5 enables their covalent

attachment to a gold metal surface. Two types of metal substrates were used in these

studies: polished polycrystalline Au disk electrodes (pre-cleaned by immersing into hot

1:2 H2O2/H2SO4) were used for electrochemical studies, and large area Au substrates

(freshly prepared by thermal vacuum evaporation of gold onto microscope slides

covered with a Cr or Ti adlayer) were used for spectroscopy, ellipsometry and contact

angle studies. Self-assembled monolayers (SAMs) have been fabricated by immersing

the above substrates into ca. 10–3

M MeCN solutions of acceptors 3.4, 3.5 (in the dark)

for 24-72 hours. After this period the metal substrates were thoroughly washed by

rinsing with CH2Cl2 and soaking (followed by sonication for a few seconds) in MeCN

and CH2Cl2 solvents (HPLC grade), dried in vacuo and stored under Ar. It should be

noted that formation of SAMs by this fluorene acceptor is more hindered in comparison

with other thioctic esters containing a tetrathiafulvalene redox unit (as determined by

competitive absorption; see Electrochemistry of SAMs section). On the other hand,

prolonged exposure, particularly when combined with increased solution concentration

(due to evaporation) results in the formation of multilayers of the corresponding

compounds, which is manifested in increased film thicknesses and very strong

electrochemical reduction signals (see below). In the case of self-assembly of 3.5 on a

polycrystalline gold substrate for 4 weeks (in saturated concentration and with partial

light exposure) the formation of a visible film of the material, whose electrochemical

properties resembled that of the product of electro-oxidative polymerization 3.5, was

observed (see Section 3.4).

3.3. Electrochemical and spectroscopic characterizations of 3.4 and 3.5 in solution

The electrochemical behaviour of new nitrofluorene derivatives 3.2–3.5 in

solution, studied by cyclic voltammetry, reveals from two to four reversible single-

electron reduction waves, depending on substituents present in the molecules (Figure

3.1, Table 3.1). The strongest acceptor of the fluorene series (DTeF) presents electron

- 88 -

acceptor properties very similar to those of TCNQ [204]. As seen from Table 3.1,

substitution of a nitro group in fluorene 3.2 with an alkylsulfanyl moiety results in a

significant decrease of the electron affinity by >250 mV (See E1red for 3.1 and 3.2). An

oxidation of 3.2 to the sulfonyl derivative 3.3 partially restores the electron acceptor

properties, so replacing a nitro group with an alkylsulfonyl group causes an overall

decrease of the electron affinity of about 70-180 mV (see Table 3.1 3.1 3.4). The

introduction of a dicyanomethylene fragment shifts the first reduction wave to less

negative potentials by >400 mV, rendering 3.5 almost as strong an electron acceptor as

DTeF. For the oxidation process, one partially reversible peak was observed for

compounds 3.4 and 3.5, revealing the electron donor properties of the dithiolane moiety

[277]. Accordingly, a donor-acceptor interaction between the 1,2-dithiolane and

nitrofluorene moieties (presumably through-space) is manifested in a weak charge-

transfer band in the UV-Vis absorption spectra (Fig. 3.2). The intramolecular character

of this band was corroborated by its linear concentration dependence. The significant

bathochromic shift of this band for compound 3.5 (~550 nm) compared to that for 3.4

(~450 nm) is in agreement with the stronger acceptor properties of the fluorene nucleus

in the former.

-2.5 -2.0 -1.5 -1.0 -0.5 0.0 0.5 1.0

3.4

3.5

Fc/Fc+

20A

2A

E / V vs. Fc/Fc+

Figure 3.1: Cyclic voltammograms of the compounds 3.4 and 3.5.

- 89 -

400 500 600 700 8000

200

400

600

800

1000

3.5

3.4

/

M-1 c

m-1

/ nm

Figure 3.2: Electronic absorption spectra of compounds 3.4 and 3.5 in acetone solution.

Notably, multiple scanning of solutions of the ―ethylene‖ analogue of 3.5,

namely compound 3.8 [278], to positive potentials beyond the oxidation potential (at a

scan rate of 20 mV/s) resulted in deposition of an insoluble white creamy film on the

glassy carbon electrode (GCE) surface. The CV of the film, carefully washed with the

solvent, in pure electrolyte, retains the well-defined reduction waves of the fluorene-

dicyanomethylene moiety of 3.8 (Figure 3.3). The significant potential shift during the

first CV scan can be attributed to penetration of ions into a relatively thick film, and

highly reproducible electrochemical behaviour was observed during the subsequent

scans. Although no definitive evidence of the structure of this product was obtained, we

speculate that oxidation of the dithiolane fragment results in ring-opening

polymerization, giving the insoluble poly(propylenedisulfide) derivative of 3.8. Indeed,

ring-opening polymerization of the dithiolane fragment seems to be a common feature of

the thioctic anchor [279]; the signs of such polymerization were observed occasionally

during the self-assembly process.

- 90 -

Table 3.1: Redox potentials (0.1M Bu4NPF6 in MeCN, vs. Fc/Fc+) in solution of

synthesized electron acceptors.

Figure 3.3: The electrochemical polymerization of 3.8 on GCE electrode in 0.2 M

Bu4NPF6/CH2Cl2 (left) and the CV of the film of poly-3.8 in fresh electrolyte solution

(right).

-1500 -1000 -500 0 500

E / mV vs. Fc/Fc+

Compound E1red, V E2red, V E3red, V E4red, V E1ox, V

DTeF –0.19 –0.81 –1.47 –2.09

3.1 –0.57 –0.83 –1.73

3.2 –0.84 –1.11

3.3 –0.71 –1.12 –2.03

3.4 –0.64 –0.95 –1.79 0.69

3.5 –0.29 –0.83 –1.53 0.66

- 91 -

3.4 Electrochemistry of SAMs

Multi-redox behaviour has been also observed in the SAMs: two reversible single-

electron reduction waves for compound 3.4 and even three single-electron reductions for

the dicyanomethylene-fluorene (3.5) (Table 3.2, Fig. 3.4). It should be noted that the

formation of multiply charged redox species should be hindered in monolayers, as

compared to solution, due to higher coulomb repulsion energy. High negative potentials

required for multielectron reduction may result in reductive desorption of the molecule

(as RS–) [280]. Thus, for SAMs of (TCNQC10S)2 derivative only the first reduction wave

was reversible [256]. Therefore, the sequential and reversible accepting of three

electrons by a monolayer of 3.5 within a readily-achievable potential window is

remarkable. To the best of our knowledge, it presents the first observation of a radical

trianion species in SAMs. Three single-electron reductions in Langmuir-Blodgett

monolayers have, also, been observed for a fullerene derivative [281]. The reduction

potentials of the SAMs and their dependence on structural variations are very similar to

those obtained in solution (Table 3.1 and 3.2). The anodic-to-cathodic peak separation in

the CVs of the SAMs at lower scan rates (≤100 mV/s) were less than 10 mV, and the

peak current increased linearly with the scan rate (Fig. 3.5), thus revealing the space-

confined nature of the process.

Table 3.2: Redox potentials (0.2M 0.1M Bu4NPF6 in THF, vs. Fc/Fc+) of studied

electron acceptors in SAMs.

Compound Media/RE E1red E2red E3red ref

(TCNQC10S)2 MeCN –0.35 –0.74p.a. [256]

3.4 THF –0.65 –0.98

3.5 THF –0.27 –0.94 –1.61

The surface coverage (), calculated from the CV peak area, varied significantly

from sample to sample, with no clear dependence on the exposure time. The highest ,

obtained after 1–2 days of self-assembly was ~3.5×10–10

mol cm–2

with typical values

- 92 -

being in the range (1–2)×10–10

mol cm–2

, which correspond to molecular areas of ~ 0.5

nm2 and 0.8–1.5 nm

2, respectively. These values are comparable to those obtained for

SAMs from other electrochemically active molecules (e.g., 3–3.510–10

mol/cm2 for

(TCNQC10S)2 [256], 2.110–10

mol/cm2 for TTF-thioctic ester [243]). A molecular area

of ~0.5 nm2 should be expected for dense-packed ―stand-up‖ monolayers of molecules

of this size. Generally, it was more difficult to form dense monolayers with

dicyanomethylene derivatives 3.5, than with fluorenone 3.4. Self-assembly was also

observed on Pt electrodes, although the typical surface coverage was several times lower

on Pt. Prolonged exposure (for a week or more) often (but not always) resulted in

significantly increased coverage of ~510–10

– 510–9

mol/cm2 and higher. This

coverage is not compatible with a monolayer model considering the size of the

molecule, and a multilayer structure with disulfide bridges was assumed (see also

section 3.6).

The SAMs possess good electrochemical stability as judged by a gradual

decrease in the current. Less than 10% desorption (decomposition) was observed after

100 scans over the range 0 and –0.75 V (formation of the radical anion, fig. 3.6).

Predictably, cycling to more negative potentials (formation of the dianion, and even

more, the trianion species) resulted in more rapid desorption of the SAMs: ~8% for 20

cycles between 0 and –1.1 V at 600 mV/sec for 3.4; ~24% for 30 cycles between +0.2 V

and –1.05 V at 200 mV/sec for 3.5.

- 93 -

-1.6 -1.2 -0.8 -0.4 0.0 0.4

-50

-40

-30

-20

-10

0

10

20

30

J, A

/cm

2

E, Volt (vs. Fc/Fc+

)

Figure 3.4: Cyclic voltammograms of a SAM of 3.5 (electrolyte 0.2 M Bu4NPF6 in

THF).

As was mentioned above, the formation of SAMs of the fluorene acceptors is

more hindered in comparison with thioctic acid containing TTF unit. Figure 3.7 presents

a CV experiment of competitive absorption of two molecules containing acceptor and

donor redox moieties (3.5 and 3.7, respectively). Samples were prepared by immersing a

gold substrate into mixed solutions of 3.5 (5.8×10–3

M) and TTF-thioctate 3.7 (2.0×10–3

M) (in 3:1 molar ratio) for 3 days to achieve thermodynamic equilibrium. The CV

experiment (Fig. 3.7) clearly shows that relative intensities of the oxidation and

reduction peaks of donor and acceptor molecules are disproportional to the molar ratio

of 3.5 and 3.7 in the mixed solution (the current density of reduction is 1.5 times lower

then oxidation). Another important information we can see from this experiment is a ca.

200mV shift of the redox potentials of 3.5 in the mixed SAM compared to the SAM of

pure 3.5 (E1red –1.03 and V –0.44 V respectively). This suggests strong π-π interaction

between the donor and acceptor moieties within the monolayer resulting in raising the

LUMO energy in the acceptor 3.5 [72, 204-205].

- 94 -

-1.4 -1.2 -1.0 -0.8 -0.6 -0.4 -0.2 0.0 0.2 0.4

-1.0µ

-500.0n

0.0

500.0n

1.0µ 800 mV/sec

400 mV/sec

200 mV/sec

100 mV/secJ,

nA

/cm

2

Potential, V vs. Fc/Fc+

Figure 3.5: Cyclic voltammograms of a SAM of 3.4 at different scan rates (top).

Dependence of the first reduction peak current on the scan rate, for SAM of 3.4 (circles)

and 3.5 (triangles).

0

0.5

1

1.5

2

2.5

0 500 1000 1500

Scan rate (mV/sec)

I (

A)

- 95 -

-0.7 -0.6 -0.5 -0.4 -0.3

-10.0

-7.5

-5.0

-2.5

0.0

2.5

5.0

7.5

10.0

I,

Am

p

E, volt

Figure 3.6: Cyclic voltammogram of 3.4 in SAM, 100 scans at 200 mV/sec (Bu4NPF6,

MeCN). Large potential gap between cathodic and anodic peaks was due to

uncompensated solution resistance.

-1.5 -1.0 -0.5 0.0 0.5 1.0

3.5

3.5 + 3.7

Potential, V

Figure 3.7: Cyclic voltammogram of SAM prepared by immersing gold electrode in

mixed solution of 3.5 and 3.7 (3.5 (5.8×10–3

M) and 3.7 (2.0×10–3

M), 3 days, 0.2M

Bu4NPF6 in MeCN, 100 mV/s, vs. Fc/Fc+).

- 96 -

3.5 Reflectance-absorbance infrared spectroscopy (RAIRS) of SAMs

The strong infrared absorption features of compounds 3.4 and 3.5 have allowed

facile analysis of their SAMs even at low coverage. Grazing angle FTIR spectra of the

monolayers on gold were in agreement with the preservation of their molecular

structures during the self-assembly process (Fig. 3.8). The characteristic C=O vibration

was observed in all the SAMs as a strong band near 1740 cm–1

(Table 3.2). This band

includes vibrations of the carboxylic (ester) group and, for compounds 3.4, also

fluorenone C=O vibrations. The ester C–O single bond vibration is found around 1100

cm–1

. The fluorene moiety is evidenced by the strong nitro group features around 1553

(asymmetric) and 1344 (symmetric) cm–1

as well as by the characteristic sulfone group

vibrations at 1363 (asymmetric) and 1143 cm–1

(symmetric). Interestingly, the vibrations

of the cyano groups, typically at ~2200 cm–1

, are not seen in the spectra of SAMs of 3.5,

although their presence in the structure is clearly supported by the electrochemical data.

This vibration is revealed as a very weak band at 2234 cm–1

in powder samples of these

compounds (Figure 3.8, top), and also is not observable in solution. The C–H vibrations

in all the studied SAMs are quite weak. The observed peaks at ca. 3080 cm–1

are

attributed to aromatic C–H vibrations, whereas the peaks at 2850–2937 cm–1

are

symmetric/asymmetric vibrations within the methylene groups.

Table 3.3. Assignment of the major IR absorption peaks of compounds 3.4 and 3.5 in

SAMs on gold.

Peak assignment 3.4, cm-1

3.5, cm-1

CH aromatic 3079 3078

as/sym (CH2) 2922/2851 2926/2856

(C=O) 1740 1738

Aromatic (C=C) 1618 1617

as NO2/symNO2 1552/1344 1552/1344

as/sym (S=O) 1364/1161 1362/1166

(C–O) 1092 1105

- 97 -

3000 2500 2000 1500 10000.00

0.1

a.u

.

3.5

3.4

wavenumber, cm-1

3000 2500 2000 1500 1000

0.0

01 a

.u.

3.4

3.5

wavenumber, cm-1

Figure 3.8: FTIR-spectra of compounds 3.4 and 3.5 in bulk (in KBr, top) and in SAMs

on gold (grazing angle spectra, bottom).

- 98 -

3.6 Ellipsometry and contact angle measurements

To shed light on the structure of monolayers we undertook ellipsometry studies

and contact angle measurements. After 24 h soaking of evaporated gold substrates in

solutions of 3.4 and 3.5 in MeCN their reflectivity suggested an overlayer of organic

molecules with thicknesses of 1.30.2 and 1.30.1 nm, respectively. This thickness is

significantly lower than the calculated molecular length (2.5 nm for 3.4/3.5, in fully

extended conformations, Figure 3.9) and suggests a highly tilted and likely quite

disordered orientation of the molecules. The static contact angle measurements indicate

relatively hydrophobic surfaces for all the monolayers: 3.4 (671) and 3.5 (702).

As we have already mentioned, prolonged self-assembly of fluorene-thioctic

esters on gold results in the formation of multilayers: the film thickness of 3.90.1 nm

for 3.4 and 3.5, which is ca. twice the value expected for a monolayer, was found by

ellipsometry after 1 week of self-assembly. The increased film thickness is accompanied

by an increase of the contact angle (by approximately 10). Most likely, the molecules

are bound in the multilayer by a polymeric disulfide bond.

Figure 3.9: Molecular model of 3.4, calculated by DFT B3LYP method with 6-31 G (d)

basis set.

- 99 -

3.7. Rectification study of dyad 3.5

In Chapter I we have already discussed electrical junctions using SAMs of

redox-active molecules that are not of donor-acceptor type [37, 62]. These results show

clear rectification behaviour of such junctions, suggesting that a single electrochemically

active center (Donor or Acceptor) asymmetrically placed in the junctions can

demonstrate diode-like behaviour initially proposed for the donor-acceptor dyadsa. Thus,

it was interesting to test if the molecule 3.5 will show similar electrical properties.

For the rectification study we used junctions formed by physical contact between

SAMs of the 3.5 on the gold surface and a liquid mercury drop electrode coated with

hexadecane thiolb protecting monolayer (Au-S-3.5/C16S-Hg). The technique has been

previously described by Whitesides, Rampi et al. [115] SAMs of 3.5 were formed as

described above by soaking gold substrates in the MeCN solution of 3.5. SAMs on

mercury drop were formed immediately before each measurement by immersing an

electrode into a 10-3

M ethanol solution of the thiol for 30 minc followed by careful

washing the drop in pure solvent. All measurements were done in hexadecane which

provides additional mechanical stability for the junctions.

All junctions showed clear and reproducible current rectification in the direction

from gold electrode to mercury. Typical asymmetric I-V curve is shown in Figure 3.10

(top). The average rectification ratio (RR) was found to be 12 at 1 V, which is

comparable with previous results for TCNQ (RR=9±2 at 2 V) monolayers. However, the

range of the RR values is large: from 1 to 76 (Fig. 3.11) and it depends on the applied

bias: the RR increased with increasing the potential up to 0.7 V and then decreases. The

direction of the rectification corresponds to enhanced current flow from the gold

electrode with SAM of 3.5 to the mercury electrode (same direction of the rectification

a The rectification in this molecule, lacking an obvious D−A structure, was attributed to the

asymmetric position of the redox center in the metal/insulator/metal junction. However,

assembly of the junction under a solution of alkylthiol could result in substitution of a cyano

group in the TCNQ with a donor alkylsulfide group, affording a covalent D−π−A structure.

Furthermore, using a disulfide binding on one electrode and a thiol binding on the other can also

result in asymmetric conductance b Using shorter alkyl thiols (C8H17SH and C12H25SH) did not yield stable junctions.

c Longer exposure time usually leads to the detachment of the mercury drop from the column of

mercury in the electrode due to the solution moving up by capillary forces.

- 100 -

was reported for the TCNQ based SAM [37]). In contrast, junctions formed between two

hexadecane thiol SAMs supported on both, gold and mercury, electrodes presents more

symmetric I-V curves with small RR 2.3±0.3 at 1 V (in the same as for the junctions

with 3.5). Thus, we can speculate that the current asymmetry of the junctions with 3.5

also includes such non molecular based rectification.

Upon multiple scanning in the range ±1 V the rectification ratio drops

approximately 2-3 times after 4 cycles (Fig. 3.10, bottom). For comparison, the TCNQ-

based SAMs, reported by Whitesides [37], characterized by nearly indistinguishable I-V

curves upon cycling the potential from +1 to –1 V. The possible reason of such drop of

the RR for the SAMs of 3.5 could be in low stability of the monolayer. Upon applying

the electrical field and increase of the number of the defects, the direct tunneling

between two metal electrodes becomes a preferential process and covers the electrical

conductivity of the junction related to the molecular structure.

- 101 -

-1.0 -0.5 0.0 0.5 1.0

-15.0n

-10.0n

-5.0n

0.0

5.0n

Curr

ent, n

A

Voltage, V

-1.0 -0.5 0.0 0.5 1.0

-10.0

-9.5

-9.0

-8.5

-8.0

-7.5

1st scan

2nd scan

3rd scan

4th scan

log

(A

)

Voltage, V

Figure 3.10: (top) I-V characteristics of the Au-S-3.5/C16S-Hg junction (RR=4.5);

(bottom) decrease of the rectification during multiple scanning.

- 102 -

0 10 20 30 40 50 60 70 800

5

10

15

20

Co

un

t

RR at 0.7V

-1.0 -0.5 0.0 0.5 1.0 1.5 2.0 2.5 3.00

5

10

15

Co

un

t

logRR

Figure 3.11: (top) Statistical analysis of rectification ratio RR and logRR of the

Au/3.5/C16S/Hg junctions; (bottom) Dependence of the RR on the applied bias voltage

for the Au/3.5/C16S/Hg junction.

0.0 0.2 0.4 0.6 0.8 1.0-2

0

2

4

6

8

10

12

RR

Bias voltage, V

- 103 -

Conclusions

Nitrofluorene derivatives have been successfully employed for the preparation of

SAMs for the first time. By exploiting the synthetic versatility of tetranitrofluoren-one, a

series of derivatives has been obtained which retain strong electron acceptor ability and

can be covalently attached to a gold electrode. The resulting SAMs possess unique

electrochemical characteristics. The radical anion, dianion and radical trianion redox

states can be reversibly formed in these monolayers. The good electrochemical stability

of the SAMs, especially to repeated cycling in the potential range between 0 and –0.75

V, augurs well for applications of nitrofluorene-derived acceptors in molecular

electronics. At the same time, the observed tendency of the thioctic anchor to form

multilayers via disulfide links could limit the development of these specific compounds

(as well as other previously reported molecules with thioctic functionality) in molecular

electronics. We also demonstrated that junctions, constructed from SAMs of the 3.5 on

the gold electrode and alkylthiol protected Hg electrode can rectify electrical current.

- 104 -

Experimental Part

Gold substrates. Two different types of gold substrates were used in these experiments:

a) a 1.6 mm BAS gold electrode; this was polished with 0.05 m alumina, sonicated

in water, then immersed in warm (60 C) ―piranha‖ solution for 30 sec and thoroughly

washed with mili-Q water; b) an ITO glass slide with ~50–200 nm gold layer deposited

by vacuum (10–6

Torr) evaporation techniques. Both substrates were rinsed with

methanol and immediately immersed in a self-assembly solution.

Cyclic voltammetry. CV experiments were performed on a Princeton EG&G PAR273A

potentiostat under nitrogen, with a three-electrode cell in CH2Cl2, MeCN or THF as

solvent using 0.2 M Bu4NPF6 as an electrolyte, at different scan rates (20–1500 mV/s).

Platinum wire and Ag/Ag+ electrodes were used as the counter and reference electrodes,

respectively. The oxidation of ferrocene under our conditions occurs at E1/2

ox = +0.20 V

vs. Ag/Ag+ (in CH2Cl2), E

1/2ox = +0.135 V vs. Ag/Ag

+ (in THF) and E

1/2ox = +0.08 V vs.

Ag/Ag+ (in MeCN). A glassy carbon electrode (BAS, d = 2.5 mm) was used as a

working electrode for studying the solution electrochemistry and gold disk electrodes

(BAS, d = 1.6 mm and home-made, d = 6.0 mm) with self-assembled monolayers were

used to record the CVs of SAMs. The thiol monolayer was self-assembled on gold

electrodes as described above.

FTIR spectroscopy. FTIR spectra were recorded with a Nexus 670 FTIR

spectrometer (Thermo-Nicolet, Madison, WI) equipped with a liquid-N2 cooled MCT-II

detector with spectral resolution of 4 cm–1

. Transmission mode was used for bulk

samples (in KBr pellets) and grazing angle (80) reflectance-absorbance mode (RAIRS),

using grazing angle Smart-SAGA accessory was employed for monolayers on gold

substrates. The measurements were done in an atmosphere of dried, CO2-free air, and an

identical gold-covered slide (prepared in the same Au-evaporation run) freshly cleaned

by soaking in HPLC-grade dichloromethane and dried in vacuo, was used to record a

background spectrum.

Ellipsometry. SAM thicknesses were measured on a Sentech SE 400 ellipsometer

equipped with a He–Ne laser (λ=632.8 nm) at an incidence angle of 70º with respect to

the surface normal. Optical constants of the gold-coated substrates were measured using

a bare gold slide as described earlier [251]. Both reference sample and SAMs were

cleaned by soaking in HPLC dichloromethane and dried immediately before the

measurements. All the layer thicknesses reported were calculated after averaging over

10–20 measurements. The refractive index of the monolayer was assumed to be 1.45.

Contact angle measurements. The advancing, receding and static contact angles of

deionized water (> 18 MΩ cm) were measured on a home-made contact angle

goniometer equipped with a video camera, and averaged over 3–5 spots. The advancing

angles are produced as fluid is added to the drop and the receding angles as fluid is

withdrawn. The ―clean‖ Au surface produced a static contact angle of 733 which is

believed to be due to hydrocarbon impurities absorbed from air [227, 282].

Rectification measurements in mercury drop junctions: The junctions were

assembled in a procedure, similar to the described in the Chapter II. A gold layer (~200

nm) was thermally evaporated on Si wafers or glass slides pre-coated with adhesion

layer of Ti. The gold substrate was immersed in to the solution of 3.5 (10-3

M, CH2Cl2)

for 12h A hanging drop of mercury (from a microsyringe, ~500μm in diameter), covered

- 105 -

with a monolayer of hexadecylthiolate by 15-20 min exposure to a solution of C16H33SH

in ethanol and rinsed with fresh ethanol, was brought into contact with the SAM of 3.5

under hexadecane, by using a micromanipulator. The substrate was grounded, the bias

voltages were applied to the mercury electrode (two-electrode scheme), and the I-V

characteristics were recorded with a potentiostat EG&G PAR273A (sensitivity 0.1 nA)

at a scan rate of 1000 mVs–1

and sampling rate of 20 mV per point.

2-(2-Hydroxypropylsulfanyl)-4,5,7-trinitrofluorene-9-one (3.2). 3-Mercaptopropanol

(1.5 ml, 17.5 mmol) was added to a solution of fluorenone 3.1 (5.0 g, 14 mmol) in

MeCN (150 ml) followed by well-ground NaHCO3 (3.5 g, 42 mmol), which resulted in a

brown colorization. The reaction mixture was stirred at 20°C for 12 h, and the inorganic

salts were filtered off. The filtrate was concentrated in vacuum to 10 mL, and hot 2-

propanol (100 mL) was added. The red precipitate which formed on cooling was filtered

off and washed with 2-propanol and methanol to give sulfide 3.2 (4.81 g, 86%): mp 157-

159 °C; 1H NMR (300 MHz; acetone-d6) δ= 8.93 (1H, d, J=2.1 Hz), 8.69 (1H, d, J =2.1

Hz), 8.11 (1H, d, J =1.86 Hz), 8.05 (1H, d, J =1.83 Hz), 3.88 (1H, t, J =5.5 Hz, OH),

3.74 (3H), 3.42 (3H). Anal. Calcd for C16H11N3O8S: C, 47.41; H, 2.74; N, 10.37. Found:

C, 47.39; H, 2.70; N, 10.49.

2-(2-Hydroxypropylsulfonyl)-4,5,7-trinitrofluorene-9-one (3.3). Hydrogen peroxide

(10 ml, excess; 33 wt % aqueous solution) was added to a hot solution of sulfide 3.2

(4.35 g, 10.74 mmol) in AcOH, and the reaction solution was stirred at 50-60°C for 3 h,

which resulted in a change from deep-red to yellow color. The yellow precipitate which

formed on cooling was filtered off and washed with 2-propanol and methanol, affording

sulfone 3.3 (3 g, 70%): mp 196-199 °C; 1H NMR (300 MHz; acetone-d6) δ= 9.08 (1H,

d, J=2 Hz), 8.88 (1H, d, J=2 Hz), 8.77 (1H, d, J=2 Hz), 8.65 (1H, d, J=2 Hz), 3.85 (1H,

t, J=5.5 Hz, OH), 3.6-3.7 (6H, m). Anal. Calcd for C16H11N3O10S: C, 43.94; H, 2.54; N,

9.61. Found: C, 43.98; H, 2.54; N, 9.49.

2-(2-Hydroxypropylsulfonyl)-4,5,7-trinitrofluorene-9-one thioctic ester (3.4). A

solution of DCC (230 mg, 1.15 mmol) in dry CH2Cl2 was added to a solution of thioctic

acid (177 mg, 0.86 mmol) in CH2Cl2 (5 ml) at 0 °C, and the reaction mixture was stirred

at 20 °C for 1 h after which period 4-(dimethylamino)- pyridine (4.3 mg) and sulfone 3.3

(300 mg, 0.69 mmol) were added in one portion. The reaction mixture was stirred for 60

h at 20°C, then filtered from the formed dicyclohexylurea, evaporated, and

chromatographed on silica (eluting with CH2Cl2-ethylacetate 5:1 v/v). The brown-violet

fraction was evaporated; the oily residue was triturated with MeOH, decanted, and dried

in vacuum to give compound 3.4 (250 mg, 58%) as a brown powder: mp 68-71°C; 1H

NMR (400 MHz; CDCl3) δ= 9.05 (1H, d, J=2 Hz), 8.86 (1H, d, J=2 Hz), 8.75 (1H, d,

J=2 Hz), 8.63 (1H, d, J=2 Hz), 4.18 (2H, t, J=6.28 Hz), 3.69 (2H, t, J=7.88 Hz), 3.57

(1H), 3.23-3.05 (2H), 2.46 (1H), 2.3 (2H, t, J=7.34 Hz), 2.14 (2H), 1.92-1.84 (1H, m),

1.65-1.53 (4H, m); IR (KBr) ν= 1736 (C=O), 1617, 1541(NO2), 1364 (S=O), 1135 cm-1

.

Anal. Calcd for C24H23N3O11S3: C, 46.07; H, 3.71; N, 6.72. Found: C, 46.32; H, 3.73; N,

6.52.

2-(2-Hydroxypropylsulfonyl)-4,5,7-trinitrofluorenylidene-9-dicyanomethylene

thioctic ester (3.5). Fluorenone 3.4 (50 mg, 0.08 mmol) and malononitrile (6.86 mg, 0.1

- 106 -

mmol) were stirred in DMF (2 ml) at 20 °C for 6 h, then the resulting precipitate was

filtered off and washed with MeOH, yielding dicyanomethylene 3.5 (40 mg, 80%) as a

dark-green powder: mp 123-126 °C; IR (KBr) ν= 2234 (C≡N), 1736 (C=O), 1611, 1553

(NO2), 1362 (S=O), 1137 cm-1

; 1H NMR (300 MHz; acetone-d6) δ= 9.72 (1H, s), 9.44

(1H, s), 9.08 (1H, s), 8.79 (1H, s), 4.2 (2H br s), 3.68 (2H, br s), 2.84 (3H, m), 2.31 (3H,

m), 1.58 (7H, m),1.43 (3H,m).

4-Hydroxymethyl-5,4’,5’-thimethyltrathiafulvalene (3.6) synthesised as described in

Chapter 2 for the compound 2.7 [283].

Compound 3.7. Synthesised from 4-hydroxymethyl-3,3’,4’-trimethyl-TTF (3.6) (0.05 g,

0.18 mmol) and thioctic acid (0.052 g, 0.25 mmol), DCC (0.074 g, 0.36 mmol) and

DMAP (0.002 g, 0.01 mmol) ) in dry CH2Cl2 (10 mL). Product 3.7 was isolated as a

salmon pink solid (0.07 g, 83%). Mp 88–90 °C. 1H NMR [CO(CD3)2–TMS]: 4.85 (s, 2

H), 3.62 (m, 1 H), 3.16 (m, 2 H), 2.48 (m, 1 H), 2.38 (t, 2 H), 2.10 (s, 3 H), 1.96 (s, 6 H),

1.90 (m, 1 H), 1.65 (m, 4 H) 1.50 (m, 2 H); 13

C NMR (CO(CD3)2–TMS): 173.15,

131.23, 124.43, 123.78, 123.72 109.11, 106.33, 58.67, 57.13, 40.88, 39.08, 35.29, 34.17,

25.41, 13.90, 13.60, 13.58; MS (EI) m/z = 464 (M+). Calc. for C18H24O2S6: C, 46.55; H,

5.17. Found: C, 46.58; H, 5.20%.

- 107 -

Chapter IV. Synthesis and characterization of TTF--nitrofluorene

dyads for self-assembly on gold surface.

Introduction

The TTF-nitrofluorene based donor-acceptor dyads were studied for different

applications [69, 204, 278, 284]. Previously we described the design of an amphiphilic

TTF-nitrofluorene dyad 2.4 ([274] and Chapter II) and presented its rectification

behaviour in n-Si/LB film/Ti and Au//LB films//Hg junctions. Despite of the limitation

of the Langmuir-Blodgett films for application in real devices such molecular design

provides facile synthetic accessibility of the target molecules without necessity to

introduce additional functional groups. On the other hand, attractive advantages of the

covalent self-assembly of the organic molecules on metal substrates brought us to the

idea of the design of TTF-nitrofluorene dyad ―equipped‖ with an anchor group. Towards

achieving this goal, we have demonstrated in Chapter III synthesis and self-assembly of

the nitrofluorene derivatives on metal surfaces using disulfide (thioctic) anchor group

[278]. Herein, we describe synthesis of new self-assembly capable donor-acceptor dyads

with thiol-based anchor groups. The particular focus of this work was on asymmetric

modification of the fluorene moiety for its subsequent functionalization with the

appropriate ―anchor‖ group and electron-donor functionality.

4.1. Synthesis

Our synthetic target was a molecule in which donor and acceptor moieties are

coupled together by a saturated -bridge and one of the molecule’s ends is

functionalized with an anchor group capable of binding to the metal surface. 2,4,5,7-

Tetranitrofluorene-9-one (TNF) is a convenient starting material for this purpose as the

nitro groups in positions 2 and 7 can be selectively and sequentially substituted with

different ―arms‖. It allows us to create an asymmetrically functionalized acceptor

synthon with linkers available for further selective coupling with required building

blocks, such as TTF-donor and anchor group for self-assembly. Hence, modification of

- 108 -

the TNF was done by regioselective nucleophilic substitution of two nitro group in

positions 2- and 7- with different alkanethiol reagents. The substitution of the first nitro

group with 3-mercaptopropanole was described in the Chapter III (Scheme 3.1,

compound 3.3).

In the next step, a second ―arm‖ necessary for coupling with donor synthon was

installed by substitution of the second nitro group in the position 7- of fluorene 3.3 with

tert-butyl-3-mercaptopropionate 4.1. This reagent was synthesised from the commercial

tert-butyl-3-bromopropionate in 65% yield following a literature procedure [285].

Follow-up oxidation of the sulfide 4.2 resulted in sulfone 4.3 in 80% yield (Scheme 4.2).

Scheme 4.1: Synthesis of bifunctionalized fluorene 4.3.

As a result, the constructed acceptor synthon 4.3 carries a hydroxyl group on one

side and the protected carboxylic group on the other. The necessity of the t-Bu ester

protection is dictated by further synthetic route: (1) to allow selective deprotection of

one arm for further functionalization, (2) to maintain good solubility of the molecule,

and (3) to prevent polymerization in a self-esterification reaction. The subsequent DCC

promoted esterification of the terminal hydroxyl with thioctic acid yields the

corresponding thioctic ester 4.4 with anchoring disulfide group (Scheme 4.2).

Scheme 4.2: Coupling of the acceptor synthon with thioctic acid.

- 109 -

The tert-butyl ester group of 4.4 was selectively hydrolyzed with catalytic

amount of CF3COOH to afford acid 4.5 (Scheme 4.5). However, the acid 4.5 appeared to

be unstable, presumably due to acid-promoted polymerization of the dithiolidene cycle

of the thioctic group. The evidences of polymerization of thioctic ester of fluorene in

SAMs were also shown earlier in Chapter III of this thesis.

Scheme 4.3: Deprotection of the carboxylic group in the synthon 4.4.

Since the above attempt to use thioctic ester as a functional group for self-

assembly was problematic and also because of its tendency to form multilayers on the

electrode surface, we focused on finding appropriated anchor functionality for the dyads.

As we already discussed in the Chapter III the thiol group (widely used in designing of

SAMs) is not an appropriate ―anchor‖ as it readily reacts with polynitrofluorene. Thus,

in our synthetic strategy we are limited to the sulfur containing ―anchor‖ groups such as

disulfides and protected thiols. Below we discuss our results of synthesis of series of

acceptor molecules and TTF--nitrofluorene dyad with a self-assembly functional

group.

Our first option was to use a dialkyl disulfide with an appropriate head functional

group for easy and strong coupling with the fluorene synthon. The absorbance of

disulfides on gold surface is a well-understood process that leads to formation of stable

and dense monolayers identical to those of thiols [124, 286]. The first potential

candidate that met our requirements was cystamine (Scheme 4.4). This commercially

available disulfide has two amino groups that can be coupled with carboxylic groups

forming a stable and robust amide bond which provides stable synthon for the next

reaction steps and may reinforce the self-assembly through hydrogen bonding. Thus,

- 110 -

synthesis of the acceptor synthon was modified in order to have two carboxyl-terminated

―arms‖. To achieve this we choose two ester terminated thiols for substitution of nitro

groups in TNF: n-butyl-3-mercaptopropionatea and tert-butyl-3-mercaptopropionate.

This allows sequential deprotection of the carboxylic groups for successive attachment

of an amine-terminated anchor group and a hydroxyl-terminated TTF synthon.

Modification of the acceptor moiety was performed similarly to synthetic route

described above (Scheme 4.4). Starting from TNF, one nitro group was substituted with

butyl-3-mercaptopropionate in the presence of NaHCO3, resulting in sulfide 4.6 with

81% yield. Following oxidation of the 4.6 in the next step resulted in sulfone 4.7 in 92%

yields. The second nitro-group was substituted with tert-butyl 3-mercaptopropionate

following by oxidation of the sulfide 4.8, in to the acceptor synthon 4.9 (83 and 88%

yields, respectively).

The tert-butyl group was selectively hydrolyzed in presence of CF3COOH in

CH2Cl2 solution resulting in acid 4.10 which was then converted into the corresponding

acid chloride 4.11 by refluxing in pure oxalylchloride. The coupling of the 4.11 with

cystamine was done in the presence of pyridineb resulting in compound 4.12a in good

yield (87%). However, the very low solubility of the acid 4.13, a product of hydrolysis

of 4.12a, limited further synthetic applicability of this molecule.

In an effort to overcome the solubility issue we have prepared a longer ―anchor‖

group – bis-(6-aminohexyl) disulfide 4.16 (Scheme 4.7). It was prepared according to

literature procedure [287] starting with 1,6-dibromohexane and potassium phthalimide.

The 1-bromohexyl-6-phthalimide 4.14 was obtained in 87% yield and converted into the

disulfide 4.15 in 67% yield by treatment with Na2S2O3 followed by oxidation with I2.

The final bis-(1-hexylamine)disulfide 4.16 was obtained in 60–80% yield by treatment

of the 4.15 with hydrazine monohydrate. The diamine 4.16 was reacted with acid

chloride 4.11 in a pyridine catalyzed coupling reaction. However, the solubility of the

product 4.12b was the limiting factor in this case as well. So far the explanation of such

a Our initial attempt intended to use a methyl ester, however the solubility of the product was

now enough to continue the synthesis. b Limited solubility of the acid 4.13 prevented us from using the DCC coupling protocol.

- 111 -

low solubility may be due to the large size of the molecule, which carries two acceptor

moieties.

In our next attempt to introduce the self-assembly functional group we decided to

reduce the size of the molecule and use a protected alkyl thiol as an anchor group. As a

protection group for the thiol we have chosen a tert-butyl group, widely used in the

peptide synthesis [285]. Such an anchor group allows us to decrease the size of the

acceptor synthon (as compared to 4.12 a, b), and further improve its solubility

throughout further synthesis. The t-BuS group has fairly good stability to the basic and

acidic conditions and can tolerate subsequent reactions. There are also various ways to

remove it or exchange with other, easy removable, protective groups [288-289]. The

most common transprotection procedure consists in regenerating a free thiol by acidic

hydrolysis (CF3COOH, BBr3, etc.) followed by its immediate re-protection with acetyl

halogen providing a more labile acetyl protecting group.

- 112 -

Scheme 4.4: Synthesis of the disulfide terminated acceptor synthons.

- 113 -

Scheme 4.5: Synthesis of bis-(6-aminohexyl) disulfide.

To follow this route, 6-tert-butylsulfanylhexylamine 4.18 was synthesized from

the bromide 4.14 (Scheme 4.6). First, by substitution of the bromine with 2-methyl-2-

propanethiol, the starting compound was converted into phthalimide derivative 4.17 in

91% yield, which was then cleaved with hydrazine monohydrate, resulting in final

product 4.18 in 93% yield.

Scheme 4.6: Synthesis of tert-butyl protected anchor moiety. i) 2-methyl-2-

mercaptopropane, K2CO3, DMF; ii) hydrazine monohydrate, ethanol

- 114 -

The amine terminated ―anchor‖ 4.18 was coupled with acid chloride 4.11

(Scheme 4.9) in presence of pyridine resulting in the acceptor synthon with anchor

functionality 4.18 in 34% yield. Unfortunately, after hydrolysis of the n-butyl ester of

4.19 (Scheme 4.7) the acid 4.20 still showed very low solubility to continue with the

next synthetic steps. We suspected that hydrogen-bonded aggregation due to the

presence of primary amide may have limited the solubility of the compounds of this

series.

Scheme 4.7: Synthesis of the acceptor synthon 4.20

To test this hypothesis, we came to the molecular design in which we tried to

combine tert-butyl protected thiol as a soluble intermediate of the anchor group and ester

linkers between the dyad’s moieties. As the anchor moiety in our design we used 3-tert-

butylsulfanylpropionic acid 4.21, which was synthesized in 67% yield [290] from

commercially available 3-bromopropionic acid and 2-methyl-2-propanethiol (Scheme

4.8).

- 115 -

Scheme 4.8: Synthesis of the anchor functionality 4.21

The synthesis of tert-butylthiol-functionalized dyad 4.27 is presented in Scheme

4.9. We have used the previously synthesised acceptor synthon 4.3 with two ―arms‖

containing a hydroxy and a protected carboxylic group. Through the first, hydroxy-

terminated ―arm‖, the fluorene 4.3 (Scheme 4.1) was coupled with anchor moiety 4.21 in

a DCC promoted esterification reaction, to give ester 4.22 in 62% yield. Two tert-butyl

terminal groups on the 4.22 provide very good solubility to the compound, which is

maintained for the monoacid 4.23 after selective hydrolysis of the tert-butyl ester in the

presence of CF3COOH. The dyad 4.24 was obtained in 24% yield by DCC-promoted

esterification of carboxy-terminated ―arm‖ of the acceptor synthon 4.23 with hydroxy-

functionalized donor synthon 4.28 (Scheme 4.11).

Synthesis of the donor synthon 4.28 (Scheme 4.10) was done following the

synthetic procedure, described in the Chapter II for dipentyl-substituted TTF (Scheme

2.1). Lithiation of the trimethyl-TTF 4.26 with LDA, followed by the reaction of the

corresponding lithium salt N-methylformanilide resulted in an aldehyde 4.27 in 51%

yield. The aldehyde was reduced with NaBH4 to give hydroxymethyl-TTF derivative

4.28 in 64% yield.

Finally, the acceptor ability of the dyad 4.24 was improved by converting it in to

the dicyanomethylene derivative 4.25. (See Chapters II and III).

Transprotection of the tert-butyl group with a more labile acetyl group was

attempted by acid catalyzed (BBr3 or CF3COOH) conversion procedure [288]. However,

we found that these conditions are harsh enough to decompose the donor moiety, which

was observed by absence of the corresponding NMR signals in the crude product. Other

methods of deprotection such as bromine catalyzed conversion [289] were also

unsuccessful showing no signs of the desired product. As a future direction in this

project, deprotection of tert-Bu group can be attempted on stage of the compound 4.23

- 116 -

or more labile triphenylmethyl group as protection can be used [291]. Nonetheless, the

dyads 4.24 and 4.25 may be used to study their rectification behaviour as the tert-BuS

terminal groups can be physisorbed on the electrode surface done for other molecular

wires [292].

Scheme 4.9: Synthesis of the TTF-fluorene dyad.

Scheme 4.10: Synthesis of TTF-alcohol 4.31.

- 117 -

4.2. Characterization

The electrochemical behaviour of new nitrofluorene derivatives 4.4, 4.6-4.9 and

donor-acceptor dyads 4.24 and 4.25 in solution was studied by cyclic voltammetry. In

CV experiments, acceptor synthons 4.4, 4.6-4.9 reveal similar multiredox characteristics

(Table 4.1) as was discussed in the Chapter III for mono-functionalized fluorene

derivatives. As expected, replacing the second nitro group with a sulfonyl group leads to

a larger decrease (130mV) of the total electron affinity of the fluorene moiety. Due to

the presence of TTF moiety, with characteristic oxidation waves the dyads 4.24 and 4.25

showed clear amphoteric behaviour (see Figure 4.1 and Table 4.1). As expected, the

conversion of the keto group of the fluorene into the dicyanomethylene derivative shifts

the reduction potentials. This decreases of the HOMO-LUMO gap from 0.62 to 0.34 eV.

Interestingly, the electron affinity of the acceptor moiety in 4.28 is the same as for the

dyad 2.4 with trinitrocarboxy fluorene acceptor moiety. Coupling together the fluorene

and TTF moieties has no significant influence on the electrochemical properties of the

acceptor and donor in dyad 4.27, whereas, its conversion into dicyanomethylene

derivative 4.28 results in 100 mV positive shift of the first oxidation potential due to the

stronger π-π interaction between donor and acceptor which leads to the perturbation of

donor ability of TTF.

- 118 -

-2.0 -1.5 -1.0 -0.5 0.0 0.5 1.0

4.27

4.28

Voltage, V vs. Fc

0.34V

0.62V

Figure 4.1: CV of the dyads 4.27 and 4.28 (0.2M Bu4NPF6 in CH2Cl2, vs. Fc/Fc+.)

Table 4.1: Redox potentials (vs. Fc/Fc+) in solution of synthesized fluorene derivatives

and donor-acceptor dyads.

# E1red,

V

E2red,

V

E3red,

V

E1ox,

V

E2ox,

V

HOMOa/LUMOb

(exp)

HOMO/LUMO

(calc.) E,

eV

E, eV

(calc)

TNF –0.62 –0.96 –1.82

4.5 –0.72 –1.11 – 0.78

4.6 –0.87 –1.17 –1.92

4.7 –0.66 –1.03 –1.87

4.8 –0.99 –1.34 –2.09

4.9 –0.75 –1.18 –1.99

4.24 –0.77 –1.13 – –0.15 0.33 –4.65/

–4.03

0.62

4.25 –0.39 –0.95 – –0.05 0.37 –4.75/

–4.41

–4.81/

–4.44

0.34 0.37

a) determined from first oxidation peak (vs. Fc/Fc+) as HOMO= –4.8 – E

11/2ox

b) determined from first reduction peak (vs. Fc/Fc+) as LUMO= –4.8 – E

11/2red

The electronic structure of the dyad 4.25 was estimated from the quantum

mechanical calculations by DFT method (B3LYP level, with a 6–31G(d) basis set).

According to calculations, there are two possible stable conformers: with extended and

with folded geometry, differing in energy by 4.7 kcal/mol (the folded conformer is more

stable) (Fig. 4.2). Higher stability of the folded conformer is likely due to electrostatic

- 119 -

attraction between donor and acceptor in such CTC. However, this effect should be

diminished in the presence of solvent. Remarkably, our calculations show complete

localization of the HOMO and LUMO orbitals on the donor and acceptor fragments

respectively in the extended conformer. However, in the case of the folded conformer an

intramolecular π-π complexation between TTF and fluorene moieties is possible due to

their close proximity, which results in partial mixing of the HOMO and LUMO orbitals

(Fig. 4.2, bottom). The energy levels of HOMO and LUMO as well as HOMO-LUMO

gap for extended conformer are in good agreement with values obtained in CV

experiment (Table 4.1), while the folded conformer is characterized by decreased

HOMO and increased LUMO energies as well as larger value for HLG (1.19 eV) due to

the through space intramolecular charge transfer.

The presence of intramolecular charge transfer (ICT) was observed by broad

absorbance band in the electronic spectra of the solution of 4.28 with maximum at 1200

nm (Fig. 4.3). The optical HOMO-LUMO gap, calculated from UV-Vis absorbance

peaks are reasonably close (1.41 eV) to the calculated value for the folded conformer.

The intramolecular origin of this charge transfer was established by a linear

concentration dependence of the absorbance at 1200 nm in wide range from 10-4

to 10-6

M (figure 4.3, inset).

ICT complexation in the dyad 4.28 also results in strong EPR signal both in

solution and solid state (Figure 4.4). Solution EPR (in CH2Cl2, corrected with DPPH)

shows a broad singlet with g = 2.008, whereas frozen solution revealed rhombic spectra

(g1=2.017, g2=2.009 and g3=2.004) centred at g=2.009 which is in agreement with the

isotropic value in the solutiona. Such an EPR signal agrees well with the literature results

for a similar TTF-fluorene dyad [204] and corresponds to the TTF radical cation [293].

The presence of the second radical (fluorene radical anion) was not observed, possibly

due to the overlap with the TTF signal or quenching the radical by formation of the

dimer in head-to-tail conformation.

a g-values were calibrated according to the g-value of the 2,2-diphenyl-1-picrylhydrazyl (DPPH)

standard (g=2.0036)

- 120 -

Figure 4.2: Optimized geometries and calculated HOMO and LUMO orbitals for two

conformations of 4.25; (top) extended and (bottom) folded ―head-to-tail‖ structures.

- 121 -

600 900 1200 1500 1800 2100

0.0

0.5

1.0

-5.5 -5.0 -4.5 -4.0 -3.5-3.5

-3.0

-2.5

-2.0

-1.5

-1.0

log

A

log[C]

R2=0.989

y=2.69+1.08x

x

10

-3

cm

-1

wavelength, nm

Figure 4.3: UV-Vis-near IR absorbance spectra of dyads 4.24 (blue) and 4.25 (red). In

the inset a linear dependence of the absorbance on the concentration of 4.25 is presented.

2.04 2.03 2.02 2.01 2.00 1.99 1.98

289K

133K

g-value

Figure 4.4: EPR spectra of the dyad 4.28 in CH2Cl2 solution at room temperature and in

CH2Cl2 frozen matrix at 133K.

- 122 -

IR spectroscopy provides evidence of the partial charge transfer that occurs in

the molecule. We have discussed in Chapter II the influence of charge transfer on the

CN stretching frequency. As we can see from Figure 4.5, the position of the CN band

(νCN=2202 cm-1

) in 4.25 is significantly lower than in similar dicyanomethylene

derivative of fluorene 3.5 without TTF-donor moiety (ν(CN)=2230 cm-1

). This is a sign

of a partial charge transfer in the solid state (degree of charge transfer is approx. 0.6, See

Chapter II and [211]). This supports our previous observations from the electronic

absorption and EPR spectra.

3500 3000 2500 2000 1500 1000 500

4.27

4.28

wavenumber, cm-1

2202 cm-1

Figure 4.5: Infra-red spectroscopy of the dyads 4.27 and 4.28 in bulk (ATR crystal).

- 123 -

Conclusions

A donor--acceptor dyad based on the TTF-fluorene couple with a protected

thiol functionality was successfully synthesized for the first time. However, the tert-

butyl protection for the thiol moiety employed as the stable and soluble intermediate in

our synthetic strategy is not a suitable group due to incomparability of the donor moiety

with the used deprotection methods. We suggest that use of thiol anchor functionality is

not the best choice in our molecular design. In spite of the fact that thiol and disulfide

groups are widely used in SAMs it is extremely difficult to introduce such functionalities

throughout the synthesis of the donor-acceptor dyads with low HOMO-LUMO gap.

Other anchor groups, such as, for example, pyridine based anchor groups [294] might be

a better alternative. At the same time, we also note the tendency of the donor-acceptor

molecules to rapidly lose solubility as the size of the molecule increases. This

emphasizes the need for careful planning the synthetic part, considering the

compatibility of every component of the molecule with each other as well as their

appropriate functionalization conditions.

Synthesized TTF-fluorene dyads present very promising properties for

applications as electronic materials. Remarkably low HOMO-LUMO gap of the dyad

4.25 (~0.34eV) brings it to the same level with Donor-Acceptor dyad described by

Aviram and Ratner [25]. Significant charge transfer between donor and acceptor

moieties in the dyad 4.25 is manifested in strong ICT band in the UV-Vis-NIR

absorption spectra, strong EPR signal and lowering of the CN stretching in the IR

spectra. We suggest that this charge transfer complexation arises from the possible

intramolecular ―head-to-tail‖ conformation of the dyad due to the flexibility of the linker

((CH2)n where n=2) between the donor and acceptor. However, we expect that in a self-

assembled monolayer the dyad 4.25 will preferentially have extended conformation, in

order to maximize the packing efficiency.

- 124 -

Experimental Part

Cyclic voltammetry. CV experiments were performed on a CHI 760C potentiostat

under nitrogen, with a three-electrode cell in CH2Cl2, or THF as solvent using 0.2 M

Bu4NPF6 as an electrolyte, at scan rates 100 mV/s. Platinum disk electrode and Ag/AgCl

electrodes were used as the counter and reference electrodes, respectively.

Electronic spectroscopy. UV-Vis and near IR spectra were recorded with Jasco V-670

spectrophotometer.

EPR spectroscopy. EPR analysis was done on Bruker ElexysE580 spectrometer

operating at X-Band (9.8 GHz). The g-values were corrected to the signal of DPPH.

FTIR spectroscopy. FTIR spectra were recorded with a Nexus 670 FTIR spectrometer

(Thermo-Nicolet, Madison, WI), equipped with SMART-Orbit ATR accessory. The

measurements were done in an atmosphere of dried, CO2-free air.

tert-Butyl 2-mercaptopropionate (4.1) [285]. To a suspension of potassium xanthate

(12.8 g, 0.08 mol) in dry acetone tert-butyl-2-bromopropionate was added at 0 °C.

Reaction mixture was stirred overnight at room temperature. After filtering off KCl and

evaporation of the solvent residue was dissolved in ether and organic phase was washed

with 2%-solution of NaHCO3, water and brine, dried with MgSO4. Product was

dissolved in ethanolamine (5 ml) and stirred at room temperature for 2 hours. After

addition of EtOAc, organic phase was washed with 2%-solution of HCl, quickly washed

with water and brine, dried with MgSO4. Purification was done by fractional distillation

at low pressure resulting in desired product 4.6 as colorless liquid (9 g, 76%): 1H NMR

(300 MHz, CDCl3) δ= 2.70 (2H, q, J=6 Hz), 2.53 (2H, t, J=7 Hz), 1.59 (1H, t, J=9 Hz),

1.44 (9H, s),

2-(3-Hydroxypropylsulfonyl)-7-(tert-butyloxycarbonylethylsulfanyl)-4,5-

dinitrofluorene-9-one (4.2). This was obtained similarly to 3.2 (Chapter III) from tert-

butyl 2-mercaptoacetate (0.30 g), fluorenone 3.3 (319 mg), and NaHCO3 (326 mg):

yield 69%, mp 195 °C (dec; phase transition at ca. 135 °C); 1H NMR (300 MHz; CDCl3)

δ= 8.61 (1H, d, J=2 Hz), 8.47 (1H, d, J=2 Hz), 8.01 (1H, d, J=2 Hz), 7.96 (1H, d, J=2

Hz), 4.15 (2H, d, J=5.5 Hz), 3.78 (2H, s), 3.49 (2H, t, J=5.5 Hz), 1.48 (9H, s); MS

(CI)m/z 542(MNH4+, 7%), 524 (M+, 2%), 134 (100%). Anal. Calcd for C21H20N2O10S2:

C, 48.09; H, 3.84; N, 5.34. Found: C, 47.80; H, 3.77; N, 5.32.

2-(3-Hydroxypropylsulfonyl)-7-(tert-butyloxycarbonylethylsulfonyl)-4,5-

dinitrofluorene-9-one (4.3). This was obtained similarly to 3.3 (Chapter III) from

sulfide 4.2 (210 mg) and H2O2 (2 mL): yield 81%, mp 310 °C (dec); 1H NMR (300

MHz; acetone-d6) δ= 8.75 (1H, d, J=2 Hz), 8.74 (1H, d, J=2 Hz), 8.62 (1H, d, J=2 Hz),

8.61 (1H, d, J=2 Hz), 4.69 (2H, s), 4.20 (1H, t, J=5.5 Hz, OH), 4.06 (2H, q, J=5.5 Hz),

3.77 (2H, t, J=5.5 Hz), 1.48 (9H, s). Anal. Calcd for C21H20N2O12S2: C, 45.32; H, 3.62;

N, 5.03. Found: C, 44.93; H, 3.54; N, 4.95.

2-(3-Hydroxypropylsulfonyl)-7-(tert-butyloxycarbonylethylsulfonyl)-4,5-

dinitrofluorene-9-one thioctic ester (4.4). This was obtained similarly to 3.4 (Chapter

III). To a solution of thioctic acid (31 mg) in CH2Cl2 (15ml) DCC (34 mg) as added and

reaction mixture was stirred at room temperature of 1.5 hour. Then 4-(dimethylamino)-

pyridine (3 mg) and sulfone 4.3 (82 mg) were added and reaction mixture stirred for 12

- 125 -

hours. After the coupling was complete (followed by TLC, CH2Cl2:hexane 1:3) the

solution was filtered from urea, diluted with CH2Cl2 and washed with water, brine and

dried over Mg2SO4. Purification by column chromatography (CH2Cl2:hexanes, 1:3)

resulted in desired product 4.7 (yield 61%), mp 165 °C (dec; phase transition at ca. 135-

140 °C); 1H NMR (300 MHz; CDCl3) δ= 8.72 (1H, d, J) 2 Hz), 8.71 (1H, d, J=2 Hz),

8.624 (1H, d, J=2 Hz), 8.618 (1H, d, J=2 Hz), 4.56 (2H, t, J=5.5 Hz), 4.21 (2H, s), 3.63

(2H, t, J=5.5 Hz), 3.55-3.38 (1H, m), 3.21-2.99 (2H, m), 2.48-2.36 (1H, m), 2.17 (2H, t,

J=7.5 Hz), 1.92-1.79 (1H, m), 1.71-1.42 (4H, m), 1.48 (9H, s), 1.42-1.23 (2H, m); MS

(FAB) m/z 744 (65%). Anal. Calcd for C29H32N2O13S4: C, 46.76; H, 4.33; N, 3.76.

Found: C, 46.81; H, 4.35; N, 3.76.

2-(3-Hydroxypropylsulfonyl)-7-(carboxyethylsulfonyl)-4,5-dinitrofluorene-9-one

thioctic ester (4.5). Ester 4.4 (44 mg, 0.059 mmol) was dissolved in dry CH2Cl2 (0.3

mL), and trifluoroacetic acid (0.1 mL) was added at 0°C at stirring. The reaction mixture

was stirred for 1 h at 0 °C and left overnight at room temperature. The TLC analysis

showed completeness of the hydrolysis, and the product was isolated by precipitation

with ether (2 mL): yield 40 mg (98%); 1H NMR (300MHz, acetone-d6) δ= 8.78 (2H,

m), 8.66 (2H, m), 4.82 (2H, s), 4.54 (2H, t), 3.99 (2H, t), 3.12 (2H, m), 2.41 (1H, m),

2.13 (2H, t), 1.92 (2H, m), 1.65-1.1 (6H, m).

2-(2-Butoxycarbonylethylsulfanyl)-4,5,7-trinitrofluorene-9-one (4.6). n-Butyl-3-

mercaptopropionate (13.8 ml, 85 mmol) was added to a solution of

TNF (26.5 g, 73.5 mmol) in MeCN (750 ml) followed by addition of well-ground

NaHCO3 (20 g, 240 mmol).The resulting brown-orange reaction mixture was stirred at

20°C for 12 h, and the inorganic salts were filtered off. The filtrate was concentrated in

vacuum to 50 ml, and hot 2-propanol (200 ml) was added. The orange precipitate which

formed on cooling was filtered off and washed with 2-propanol to give sulfide 4.6

(27.87 g, 81%): mp 116-118 oC;

1H NMR (500 MHz; CDCl3) 8.93 (d, 1H, J=2.0 Hz),

8.75 (d, 1H, J=1.5 Hz), 7.94 (dd, 2H, J=2.0, 11.5 Hz), 4.15 (t, 2H, J=7.0 Hz), 3.41 (t, 2H,

J=7.5 Hz), 2.78 (t, 2H, J=7.0 Hz), 1.65-1.63 (m, 2H), 1.54-1.38 (m, 2H), 0.94 (t, 3H,

J=8.0 Hz); 13

C NMR (500 MHz; CDCl3) 185.8, 170.7, 148.6, 147.2, 139.5, 137.6, 137.6,

128.5, 127.0, 125.6, 125.5, 122.4, 65.2, 33.3, 30.6, 27.5, 16.1, 13.6; HR-MS (ESI)

calculated for C20H17O9N3SNa 498.0578 found 498.0574.

2-(2-Butoxycarbonylethylsulfonyl)-4,5,7-trinitrofluorene-9-one (4.7). Hydrogen

peroxide (100 ml, excess; 33 wt % aqueous solution) was added to a hot solution of

sulfide 4.6 (27.8 g, 58.6 mmol) in AcOH (500ml), and the reaction solution was stirred

at 65 °C for 6 h, which resulted in a change from deep-orange to pale yellow color.

Then, hot water (100 ml) was added, and the pale yellow precipitate which formed on

cooling was filtered off and washed with water, affording sulfone 4.7 (27.3 g, 92%). mp

150-152oC;

1H NMR (500 MHz; CDCl3) 9.03 (d, J=2.0 Hz, 1H), 8.89 (d, J=1.5 Hz, 1H),

8.68 (d, J=1.5 Hz, 1H), 8.60 (d, J=1.5 Hz, 1H), 4.05 (t, J=7.0 Hz, 2H), 3.60 (t, J=7.0 Hz,

2H), 2.90 (t, J=7.0 Hz, 2H), 1.61-1.58 (m, 2H), 1.38-1.33 (m, 2H), 0.92 (t, J=7.5 Hz,

3H); 13

C NMR (500 MHz; CDCl3) 183.9, 169.6, 146.8, 144.7, 138.3, 138.2, 137.7,

137.2, 130.3, 127.9, 127.6, 122.9, 65.8, 51.7, 30.4, 27.5, 19.0, 13.6; HR-MS (ESI)

calculated for C20H17O11N3SNa 530.0476 found 530.0474

2-(2-Butoxycarbonylethylsulfonyl)-7-(2-tert-butyloxycarbonylethylsulfanyl)-4,5-

dinitrofluorene-9-one (4.8). Mercaptane 4.1 (0.319 g, 1.97mmol) was added to a

solution of 4.7 (0.933 g, 1.79 mmol) in acetonitrile (200 mL), followed by addition of

sodium bicarbonate (0.70 g, 8.33 mmol). The mixture was stirred at room temperature

- 126 -

for two nights. The sodium bicarbonate was filtered off, and the clear red solution was

concentrated to 15-20ml. Hot isopropanol (200 mL) was added and the bright yellow

precipitate, formed upon cooling, was filtered, collected, and dried under vacuum to give

4.8 (0.950g, 83%): mp 132-134oC;

1H NMR (500 MHz; CDCl3) 8.56 (d, J=1.5 Hz, 1H),

8.43 (d, J=2.0 Hz, 1H), 7.91 (dd, J=2.0, 13.5 Hz, 2H), 4.06 (t, J=6.5 Hz, 2H), 3.56 (t,

J=7.5 Hz, 2H), 3.36 (t, J=7.5 Hz, 2H), 2.85 (t, J=7.0 Hz, 2H), 2.69 (t, J=7.0 Hz, 2H),

1.61-1.58 (m, 2H), 1.48 (s, 9H), 1.37-1.33 (m, 2H), 0.92 (t, J=7.5 Hz, 3H); 13

C NMR

(500 MHz; CDCl3) 186.3, 169.8, 169.6, 147.1, 146.9, 145.7, 142.0, 139.2, 137.4, 137.3,

130.1, 128.6, 127.0, 126.9, 125.5, 81.9, 65.7, 51.6, 34.4, 30.4, 28.1, 27.64, 27.58, 19.0,

13.6. HR-MS (ESI) calculated for C27H30O11N2S2Na 645.1183 found 645.1177.

2-(2-Butoxycarbonylethylsulfonyl)-7-(2-tert-butyloxycarbonylethylsulfonyl)-4,5-

dinitrofluorene-9-one (4.9). Hydrogen peroxide (4 mL, excess, 33 wt% aqueous

solution) was added to a solution of 4.8 (0.950g, 1.53 mmol) in acetic acid (100mL).

The solution was stirred at 80oC for three hours. Water was added to the reaction

mixture and the pale yellow powder was filtered out, collected, and dried under vacuum

yielding 4.9 (0.882g, 88%): mp 258-260oC;

1H NMR (500 MHz; CDCl3) 8.67 (d, J=1.5

Hz, 2H, 8.58 (d, J=1.5 Hz, 2H, 4.05 (t , 2H), 3.60 (t, 2H), 3.56 (t, J=7.5 Hz, 2H), 2.89 (t,

J=7.0 Hz, 2H), 2.80 (t, J=7.0 Hz, 2H), 2.58 (m, 2H), 1.40 (s, 9H), 1.36 (m, 2H), 0.92 (t,

J=7.5 Hz, 3H); 13

C NMR (500 MHz; CDCl3) 14.4, 169.5, 168.5, 146.7, 144.7, 144.5,

137.9, 137.9, 137.5, 137.4, 130.3, 130.2, 127.8, 127.7, 82.6, 65.8, 51.8, 51.7, 30.4, 28.6,

27.9, 27.5, 19.0, 13.6; HR-MS (ESI) calculated for C27H30O13N2S2Na 677.1081 found

677.1074.

2-(2-Butyloxycarbonylethylsulfonyl)-7-(2-carboxyethylsulfonyl)-4,5-

dinitrofluorene-9-one (4.10) To a solution of 4.9 (0.882g, 1.32 mmol) in acetic acid

(100 ml) was added trifluoroacetic acid (50 mL) under nitrogen. The reaction was stirred

at room temperature for one hour. Cold water was added to the pale yellow solution and

the precipitate was filtered out, collected, and dried yielding 4.10 (0.763g, 94%): 1H

NMR (400 MHz; acetone-d6) 8.75 (dd, J=1.6, 6.0 Hz, 2H), 8.63 (d, J=2.8 Hz, 2H), 4.03

(t , J=6.8 Hz, 2H), 3.89-3.83 (m, 4H), 2.88-2.83 (m, 4H), 1.60-1.53 (m, 2H), 1.37-1.30

(m, 2H), 0.89 (t, J=7.2 Hz, 3H). HR-MS (ESI) calculated for C23H21O13N2S2 597.0479

found 597.0485.

2-(2-Butoxycarbonylethylsulfonyl)-7-(2-chlorocarbonylethylsulfonyl)-4,5-

dinitrofluorene9-one (4.11). An acid 4.10 (1g) was refluxed for 5-6 hours in 3-5 ml of

oxalylchloride with presence of catalytic amount of dry DMF (2-3 μl). After removing

of oxalylchloride, by careful distillation under reduced pressure residue was dried in

vacuum and used for next step without further purification: quantitative yield, 1H NMR

(300 MHz; acetone-d6) 8.78 (1H, d, J=1.8 Hz), 8.74. (1H, d, J=1.8 Hz), 8.65 (1H, d,

J=1.8 Hz), 8.62 (1H, d, J=1.8 Hz), 4.03 (4H, t, J=6.9 Hz), 3.86 (2H, t, J=7.5 Hz), 3.63

(2H, t, J=7.5 Hz), 2.86 (2H, t, J=7.5 Hz), 1.57-1.54 (2H, m), 1.36-1.33 (2H, m), 0.89

(3H, t, J=7.2 Hz).

Disulfide 4.12a. To a solution of acid chloride 4.11 (0.5g, 0.84mmol) in dry THF (50

ml) at –30C a solution of a mixture of cystamine (0.11g, 0.71mmol) and pyridine (1ml)

in dry THF (5ml) was slowly added. The reaction mixture was stirred at –20-30C

overnight. Then water was added to the reaction mixture, precipitate was collected by

filtration and washed with water and methanol. Crude material was re-crystallized from

dioxane:butanol mixture resulting in desired product (0.406g, 87% yield). 1H NMR (300

- 127 -

MHz, DMSO-d6) 8.64 (d, 2H, J=4.5), 8.54 (d, 2H, J=1.5), 8.23 (t, br, 1H), 3.97 (t, 2H,

J=6.6), 3.91 (t, 2H, J=7.2), 3.81 (t, 2H, J=6.6), 3.18 (m, 2H), 2.72 (t, 2H, J=7.2), 2.63 (t,

2H, J=6.6), 2.52 (t, 2H, J=7.5), 1.50 (m, 2H), 1.28 (m, 2H), 0.84 (t, 3H, J=7.5). MS

(M+Na+) (ESI) calculated for C50H52N6 O24S6Na 1335.1255 found 1334.95.

Disulfide 4.13. An ester 4.12a (0.06g, 0.045mmol) was dissolved in hot trifluoroacetic

acid (3ml) and 3 drops of water. Reaction mixture was stirred at reflux for 48 h. After

the reaction was quenched with water and cooled down to room temperature the

precipitate was collected by filtration. 1H NMR (300 MHz, DMSO-d6) 8.65 (d, 1H,

J=1.2), 8.62 (d, 1H, J=1.8), 8.54 (m, 2H), 8.23 (t, br, 1H), 3.83 (m, 4H), 3.17 (m, 2H),

2.63 (t, 4H, J=6.9), 2.52 (t, 2H, J=7.2). MS (M–H) (ESI) calculated for C42H36N6O24S6

1200.0105 found 1198.88.

N–(6-Bromohexyl)phthalimide (4.14) [287]. To a solution of 1,6-dibromohexane (25g,

103 mmol) in dry acetone (100ml) at reflux potassium phthalimide (4.76g, 25.7mmol)

was added in portions. Reaction mixture was stirred at reflux for 4-6h. Then KBr was

filtered off and solvent was evaporated under reduced pressure giving colorless liquid.

Crude mixture of starting dibromohexane, mono and di-substituted imide was separated

by distillation at low pressure (0.1-0.15 mm of Hg) giving final product 4.14 as white

solid (6.9g, 87% yield). Unreacted starting dibromohexane was reused for another

synthesis. 1H NMR (300 MHz CDCl3) 7.83 (m, 2H), 7.69 (m, 2H), 3.68 (t, 2H, J=7.2),

3.39 (t, 2H, J=6.6), 1.85 (q, 2H, J=6.9), 1.69 (q, 2H, J=7.2), 1.5-1.3 (m, 4H).

N,N’-(Dithiodihexane–6,1-diyl)bisphthalimide (4.15) [287]. To a solution 4.14 (24g,

78 mmol) in methanol/water (300ml, 1:1) sodium thiosulfate (12g, 78 mmol) was added

and reaction mixture was refluxed for 4-5 h. Then iodine was added to hot reaction

mixture until brown color of solution remained. Iodine was neutralized with sodium

metabisulfite. Resulting yellow solution was left in the fridge overnight and then

precipitate was filtered off, dried and dissolved in ether and precipitated by addition of

methanol. Precipitate was filtered off, washed with methanol and dried under vacuum

giving 4.15 as a white solid (13.6g, yield 67%). 1H NMR (300 MHz, acetone-d6), 7.84

(m, 4H), 3.65 (t, 2H, J=7.2), 2.70 (t, 2H, J=7.2), 1.68 (q, 4H, J=7.2), 1.5-1.3 (m, 4H).

6,6’–Dithiobis(hexane-1-amine) (4.16) [287]. To a suspension of compound 4.15 (7.7g,

14.68 mmol) in ethanol (150ml) hydrazine hydrate (3g, excess) was added and the

reaction mixture was stirred at reflux for 1h. Then ethanol was evaporated and residue

was refluxed in 1M HCl for 1h. After evaporation of HCl solution under reduced

pressure, residue was redissolved in ethanol (150ml) and insoluble residue was filtered

off. Filtrate, then, was diluted with ether/ethyl acetate mixture (1:1, 100ml) and

precipitate was collected and re-crystallized from ethanol/ether/ethyl acetate (2:1:1)

mixture resulting 6,6’–dithiobis(hexane-1-amine) hydrochloride as a white solid.

Neutralization of hydrochloride salt was done by dissolving it in distilled water and

lowering pH of the solution with K2CO3 (pH 12). The final product was extracted from

aqueous solution with ether. Organic phase was washed with water and brine. After the

solvent was evaporated under reduced pressure a desired product 4.16 was obtained as

light yellow oil (3.0g, 77% yield). 1H NMR (300 MHz, CDCl3) 2.68 (m, 8H), 1.86 (br,

4H), 1.68 (m, 4H), 1.46-1.33 (m, 12H).

N-(6-tert-butylsulfanylhexyl-1)phthalimide (4.17). To a solution of 540 mg 4.14 and

tert-butylthiol (170 mg) in DMF (20 ml) a finely powdered K2CO3 was added and

reaction mixture was stirred at room temperature for 12 h. Product 4.17 was extracted

with EtOAc in 91% yield (500 mg). 1HNMR (300 MHz, CDCl3), 7.84-7.82 (m, 2H),

- 128 -

7.72-7.69 (m, 2H), 3.70-3.65 (m, 2H), 2.5 (t, 2H, J=7.2), 1.67 (t, 2H, J=6.9), 1.61-1.3

(m, 6H), 1.3 (s, 9H).

tert-Butylsulfanyl-6-hexylamine (4.18). To a suspension of compound 4.17 (1.10g,

3.44mmol) in methanol (125 ml) hydrazine (0.5 ml, 9.55mmol) was added and reaction

mixture was stirred at reflux for 1h. Then solvent was evaporated, and residue was

dissolved in CH2Cl2. Organic layer was washed with 10% KOH solution and brine, dried

with magnesium sulfate. The CH2Cl2 was evaporated under reduced pressure affording

amine 4.21 as a clear yellow oil (0.605g, 93%). 1H NMR (300 MHz, CDCl3), 2.67 (t,

2H, J=6.6), 2.51 (t, 2H, J=7.2), 1.62-1.52 (m, 2H), 1.50-1.26 (m, 6H), 1.30 (s, 9H).

2-(2-Butoxycarbonylethylsulfonyl)-7-(3-(N-(6-tert-

butylsulfanylhexyl)aminocarbonyl)propylsulfonyl)-4,5- dinitrofluorene9-one (4.19).

To a solution of 4.11 (0.16 g, 0.25mmol) in dry THF (50 ml) at -30 °C solution of 4.18

(100 mg, 0.53 mmol) and pyridine (100 mg) in THF (10 ml) was added. Reaction

mixture was stirred at room temperature for 48h. After THF was evaporated, residue was

washed with methanol and dried. Purification by column chromatography on silica

(CH2Cl2:EtOAc, gradient) resulted desired product as yellow solid (65 mg, 34 %), m.p.

223-224°C. 1HNMR (300MHz, CDCl3), 8.66 (d, 1H), 8.64 (d, 1H), 8.57 (dd, 2H), 5.62

(t, 1H, J=), 4.06 (t, 2H, J=6.6), 3.54-3.51 (m, 4H), 3.15 (q, 2H, J=6.6), 2.89 (t, 2H,

J=6.9), 2.75 (t, 2H), 2.5 (t, 2H, J=7.2), 1.7-1.2 (m, 12H), 1.3 (s, 9H), 0.92 (t, 3H, J=7.2). 13

C NMR (500 MHz; CDCl3) 184.5, 169.5, 167.4, 146.7, 146.5, 144.8, 144.4, 138.0,

137.9, 137.5, 137.3, 130.2, 130.1, 127.8, 127.7, 65.8, 51.9, 51.7, 41.8, 39.9, 34.4, 31.0,

30.4, 29.6, 29.2, 29.1, 28.7, 28.1, 27.5, 26.4, 19.0, 13.6. MS(EI) m/z =769 (100%)

2-(2-Hydroxycarbonylethylsulfonyl)-7-(3-(N-(6-tert-

butylsulfanylhexyl)aminocarbonyl)propylsulfonyl)-4,5- dinitrofluorene9-one (4.20).

A solution of ester 4.19 (0.134 g, 0.17 mmol) in trifluoroacetic acid (8 ml) and water (8

ml) was refluxed for 12 h. After the reaction mixture was diluted with water and cooled

to 0°C, precipitate was filtered off and washed with water resulting in acid 4.20 (94 mg)

in 77% yield: 1H NMR (400 MHz; acetone-d6) 8.76 (s, 1H), 8.70 (s, 1H), 8.63 (s, 1H),

8.58 (s, 1H), 5.63 (t, br, 1H), 3.87-3.78 (m, 4H), 3.06-3.03 (m, 2H), 2.88-2.78 (m, 4H),

2.50 (t, J=7.2 Hz, 2H), 1.40-1.35 (m, 2H), 1.27 (s, 9H); MS(EI): m/z=712 (100%).

2-(2-tert-butoxycarbonylethylsulfonyl)-7-(3-(3-tert-

butylsulfanylpropanoyloxy)propylsulfonyl)-4,5-dinitrofluorene-9-one (4.22). To a

solution of 4.21 (0.17g, 1mmol) in dry CH2Cl2 (50ml) at 0C, DCC (300 mg) was added

and the reaction mixture stirred for 2 h. During this time a precipitate was formed. Then

catalytic amount of DMAP (10mg) was added to the reaction mixture followed by

addition of 4.3 (0.5g, 0.86mmol). Reaction mixture was stirred for 4h at room

temperature. After all starting 4.3 was consumed (TLC monitoring, eluent

CH2Cl2:EtOAc), the reaction mixture was filtered off from precipitate and purified by

column chromatography (CH2Cl2:EtOAc, gradient) resulting in 4.22 as yellow solid

(207 mg, 62% yield). 1H NMR (400MHz, CDCl3) 8.695 (s, 1H), 8.692 (s, 1H), 4.25 (t,

2H, J=6Hz), 3.56 (t, 2H, J=7.2), 3.38 (t, 2H, J=7.6), 2.82-2.75 (m, 4H), 2.58 (t, 2H,

J=7.2), 2.23-2.18 (m, 2H), 1.39 (s, 9H), 1.31 (s, 9H).

2-(2-Hydrocabonyethylsulfonyl)-7-(3-(3-tert-

butylsulfanylpropanoyloxy)propylsulfonyl)-4,5-dinitrofluorene-9-one (4.23). To a

solution of ether 4.22 (0.18g, 2.6mmol) in CH2Cl2 (50ml) a trifluoroacetic acid (0.5ml)

was added and reaction mixture was stirred at room temperature until no ester left

(followed by TLC). Then hexane was added to the solution and precipitate was collected

- 129 -

by filtration and washed with water resulting in desired product as brownish solid

(0.10g, 60%). 1H NMR (400MHz, acetone-d6) 8.76 (d, 1H, J=1.6), 8.73 (d, 1H, J=1.2),

8.66 (d, 1H, J=1.6), 8.60 (d, 1H, J=1.2), 4.20 (t, 2H, J=6.4), 3.85 (t, 2H, J=7.2), 3.73-

3.68 (m, 2H), 2.86 (t, 2H, J=7.6), 2.76 (t, 2H, J=7.2), 2.56 (t, 2H, J=7.6), 2.15 (m, 2H),

1.28 (s, 9H). 13

C NMR (300MHz, acetone-d6) 185.1, 171.2, 170.4, 146.6, 144.2, 144.1,

138.54, 138.46, 137.1, 130.7, 130.4, 127.5, 127.3, 61.8, 52.0, 50.9, 41.7, 34.6, 30.2,

26.9, 23.1, 22.2, 14.7. MS (ESI): m/z = 681.01 (M+ + Na), m/z = 659.03 (M

+ +1).

Dyad 4.24. To a solution of 4.23 (60 mg, 0.09mmol) in a mixture of dry THF and

CH2Cl2 (1:1, 10 ml) DCC (100mg) was added at 0C and reaction mixture was stirred

for 2 h. Then catalytic amount of DMAP was added followed by 4.28 (35mg,

0.12mmol). The flask with reaction mixture was slowly warmed to room temperature

and stirred for 48 h. After addition of methanol the precipitate was filtered and washed

with methanol. Recrystallization from CH2Cl2:Methanol mixture resulting in desired

product 4.24 in 29% yield (25mg) as a green solid. M.p. 192-194 C, 1H NMR

(400MHz, CDCl3) 8.66 (m, 2H), 8.54 (d, 1H, J=1.6), 8.51 (d, 1H, J=1.6), 4.71 (s, 2H),

4.24 (t, 2H, J=6.4), 3.66 (t, 2H, J=7.2), 3.60 (m, 2H), 2.94 (t, 2H, J=7.2), 2.76 (t, 2H,

J=7.2), 2.57 (t, 2H, J=7.2), 2.19 (m, 2H), 1.98 (s, 3H), 1.93 (s, 6H), 1.31 (s, 9H). 1H

NMR (500MHz, CDCl3) 183.8, 171.7, 169.0, 146.7, 144.2, 144.0, 138.0, 137.7, 137.1,

130.4, 129.9, 127.6, 127.5, 123.0, 121.6, 104.9, 61.7, 61.5, 59.6, 53.3, 51.2, 42.6, 34.7,

30.8, 27.6, 23.3, 22.2, 13.9, 13.7. HR-MS (APCI) calculated for

C36H39O13N2S7 931.0492 found 931.0516.

Dyad 4.25. To mixture of 4.24 (10 mg, 0.011 mmol) and malononitrile (5 mg, (0.07

mmol) in DMF (5 ml) was stirred at room temperature for 4 h. After DMF was

evaporated at reduced pressure the residue was washed with methanol, dissolved in

CH2Cl2 and filtered through silica. Evaporation of the solvent resulted in desired product

in 58% yield (6 mg) as dark solid. 1H NMR (400MHz, CDCl3) 9.09 (s, 1H), 8.51 (s, 1H),

8.48 (s, 1H), 8.31 (s, 1H), 4.63 (s, 1H) 4.22 (t, 2H, J=5.6), 3.80 (t, 2H, J=7.2), 3.33 (t,

2H, J=7.6), 3.04 (t, 2H, J=6.4), 2.75 (t, 2H, J=7.2), 2.58 (t, 2H, J=7.2), 2.21 (s, 3H), 2.16

(s, 6H), 1.30 (s, 9H). HR-MS (APCI) calculated for C39H39O12H4S7 979.0604 found

979.0629.

5,4’,5’-trimethyltetrathiafulvalene-4-carbaldehyde (4.27) [283]. To a solution

trimetyl-TTF 4.26 (0.5g, 2mmol) in dry ether (100 ml) at –78C a solution of freshly

prepared LDA (1.2 eq.) in Et2O was added dropwise and reaction mixture was stirred at

the same temperature for 2 hours. Then N-methylformanilide (0.43 ml, 2.0 mmol) was

added and the reaction mixture was allowed to warm up overnight. After quenching with

water an organic phase was extracted with EtOAc, washed with water and brine.

Purification by column chromatography (EtOAc: CH2Cl2, gradient) resulted in aldehyde

4.26 as an orange solid (0.28g, 51%). M.p. 216-218C. 1H NMR (300MHz, acetone-d6):

9.77 (s, 1H), 2.54 (s, 3H), 1.96 (t, 6H, J=1.8Hz).

4-Hydroxymethyl-5,4’,5’-thimethyltrathiafulvalene (4.28) [283]. To a solution of

TTF-aldehyde 4.27 (0.15g, 0.50mmol) in ethanol (50ml), NaBH4 (0.2g, excess) was

added and the reaction mixture was stirred at room temperature for 3 hours. Extraction

with EtOAc, followed by column chromatography resulted in TTF-alcohol derivative

4.28 as a yellow solid (0.12g, 79%). m.p. 203-205 C. 1H NMR (400 MHz, acetone-d6)

4.33 (s, 2H), 1.99 (s, 3H), 1.94 (t, 6H, J=3.6 Hz)

- 130 -

Chapter V. Molecular rectification of hexyl-nEDOT-3CNQ dyads in

Langmuir-Blodgett film

Introduction

Since the original Aviram-Rather concept of unimolecular rectifier [25] based on

a donor--acceptor molecule, a large number of studies have been conducted for various

donor-acceptor dyads, in monolayers as well as in single-molecule junctions [28, 35, 38,

43-44, 48, 54-55, 224, 295-297]. Among them, the first confirmed unimolecular

rectification behaviour was found for γ-hexadecylquinolinium tricyanoquinodimethane

(C16H33Q-3CNQ) placed between two metal electrodes [38, 43-44]. This is a donor-

acceptor dyad with conjugated π-bridge between electroactive moieties where

separations of HOMO and LUMO orbitals was realized through a twist angle between

donor and acceptor rings (Figure 5.1). However, it is not clear if one can apply the

original Aviram-Ratner theory to explain the mechanism of the rectification in this case.

The C16H33Q-3CNQ molecule has a zwitterionic ground state in solution resulting in

controversial explanations which electronic state of the molecule is dominating in the

LB film or SAMs and, thus, making the analysis of the rectification behaviour more

complicated [44, 298]. Particular question is which part of the molecule act as a donor

and which acts as an acceptor. Rectification behaviour of C16H33Q-3CNQ dyad,

however, was confirmed on numerous occasions [38, 43-44, 47, 50-52, 299]. Finally, a

large dipole moment of this molecule, associated with the zwitterionic structure of the

dyad (Fig. 5.1) reduces the stability of the monolayer under applied electric field, thus,

hindering the applications of such structures in molecular electronics.

Figure 5.1: The first confirmed molecular rectifier (C16H33Q-3CNQ).

Herein we describe the synthesis and detailed study of new Donor-π-Acceptor

dyads based on the tricyanoquinodimethane (3CNQ) acceptor and 3,4-

- 131 -

ethylenedioxythiophene (EDOT) donor moieties (Fig. 5.2). TCNQ has been known for a

long time as a strong one-electron acceptor. Due to its high electron affinity

(EA=2.84eV), TCNQ forms several types of stable charge transfer salts. Its salts show

very interesting electronic properties, including metallic conductivity and magnetism

[300]. EDOT is one of the most popular electron-rich building blocks for the

construction of functional conjugated materials [301-302]. The homopolymer of EDOT,

PEDOT, is one of the most stable and most widely used organic electrical conductors

[303].

Figure 5.2: Molecular design of the EDOT-3CNQ dyad.

In this work we used both of these building blocks to design the donor-acceptor

dyads (Fig. 5.2) where the quinodimethane moiety of the TCNQ is directly attached to

the thiophene ring of EDOT. The molecular design of the dyad includes an amphiphilic

structure (hydrophilic cyano groups and hydrophobic hexyl chains on the acceptor and

donor parts of the molecule, respectively) necessary for assembling Langmuir

monolayers at the air-water interface.

5.1. Synthesis

Since the first synthesis of TCNQ in 1960s [27] the chemistry of this acceptor

has been extensively studied. In general, ring substituted TCNQ derivatives are

synthesised by conversion of the corresponding precursors, like p-xylylene dihalides

[304], 1,4-diidobenzenes [305] and quinones [306] to TCNQ. The substitution reaction

- 132 -

of cyano group in TCNQ by nucleophiles, originally described by Hertler et al. [307],

results in products in which one or two CN groups are displaced by nucleophiles. In the

reactions with primary and secondary amines it was shown that the amino group reacts

with TCNQ via 1,6-addition, which is followed by elimination of HCN resulting in

mono- or bis-amino substitution products [307]. Also, upon mixing TCNQ with

electron-rich aromatic molecules (indole, pyrrole, phenol and aromatic amines) in a

solution [308], similar addition reactions mediated by formation of a deep blue charge-

transfer complex (I), can take place. Subsequent irradiation of the product (II) with UV

light leads to the release of HCN and formation of product III. Several approaches to

synthesise donor-acceptor dyads (Figure 5.3) employing the ability of TCNQ to form

charge-transfer complexes with variety of donors have appeared in the literature.

Synthesis of the dyad 5.2 was done by stirring equimolar amounts of TCNQ and

bisEDOT in PhCl at 80C and subsequent irradiation of the CH2Cl2 solution of 5.1 by

UV light (Scheme 5.1). The compound 5.2 is characterized by deep blue color as a result

of intramolecular charge transfer between the donor and the acceptor. In the mass-

spectra we also observed traces of an acceptor-donor-acceptor triad as a side product of

the coupling of bisEDOT with the second TCNQ. Thus, our initial experiments showed

a direct covalent linking of TCNQ acceptor to bisEDOT electron donor and which then

requires of the protection of the other reactive site of the EDOT.

- 133 -

Figure 5.3: Coupling reactions of TCNQ with C-nucleophiles: a) [308], b) [44], c)

[309], d) [310].

To prevent formation of the side product (as was shown on the Scheme 5.1,

compound 5.2) and to introduce the amphiphilic character (for monolayer formation)

and solubility to the dyad, we decided to add an alkyl tail to the far side of the EDOT

moiety.

- 134 -

Scheme 5.1: Coupling of bisEDOT and TCNQ.

Asymmetric modification of the donor moieties, EDOT and bisEDOT, was done

following literature procedure [311] by converting them into monoalkyl derivatives. The

dimer of the commercially available EDOT, bisEDOT 5.5, was obtained by oxidative

coupling of monolithiated derivatives of EDOT with CuCl2 [312-315]. Treatment of

starting EDOT or bisEDOT with n-BuLi, followed by addition of iodohexane resulted in

desired donor synthons 5.3 and 5.6, respectively. Monoalkylated EDOT derivative 5.3

was easily separated from the starting EDOT and dialkylated side product by distillation

at reduced pressure yielding the desired product in 30-40%. Monohexyl bisEDOT

derivative 5.6 was purified by column chromatography in average 43% yield. The

formation of disubstituted side products 5.4 and 5.7 was due to a proton exchange

between the monoalkylated product and the lithiated intermediate of nEDOT, followed

by a second substitution with the alkyliodide. It is known that the exchange reaction

occur at –50 °C, which is close to the on-set temperature of the alkylation reaction [311].

- 135 -

Scheme 5.2: Modification of nEDOT donor moieties.

Coupling of the donor synthons 5.3 and 5.6 with TCNQ was performed in

acetonitrile,a similarly to the procedure described above. It resulted in products 5.8 and

5.10 as yellow solids in 81 and 97% yield, respectively. Irradiation of the intermediates

5.8 and 5.10 in thoroughly degassed acetonitrile solution with the UV light (254nm) for

10-20 min in quartz photoreactor, resulted in color change of the solution from almost

colorless to deep blue. The dyad 5.9 has good solubility in acetonitrile and its

purification was done by column chromatography (CH2Cl2:EtOAc, 3:1), whereas dyad

5.11 precipitated from the solution. It was collected by simple filtration and washed with

acetonitrile to remove the starting material and by-products.

a In contrast to the bisEDOT the hexyl-bisEDOT is quite soluble in MeCN, which allowed us to

avoid using more toxic PhCl.

- 136 -

Scheme 5.3: Synthesis of (EDOT)n-TriCNQ dyads.

5.2. DFT Calculations

In order to investigate the electronic structure of the dyads 5.9 and 5.11 we

performed DFT calculations at B3LYP/6-31G (d) level of theory for the simplified

models of these dyads (R=CH3). As discussed above, the molecular structure of 5.9 and

5.11 contains no spacer between donor and acceptor moieties. However, it was shown

previously for other Ar-3CNQ [316] molecules that a dihedral angle between the donor

and acceptor rings can still separate HOMO and LUMO orbitals [28]. Geometry

optimization for our models of the dyads 5.9 and 5.11, shows two stable conformers for

both dyads with the energy difference between them less than 1 kcal/mol. The dihedral

angles between EDOT and 3CNQ ring for dyads 5.9 and 5.11 is in the range 24-36. Our

particular interest was focused on analysis of the location and energies of HOMO and

LUMO orbitals within the dyad. Calculations showed significant mixing of HOMO and

LUMO orbitals. Both HOMO and LUMO densities are spread over the entire molecule.

The calculated dipole moment for the models of the dyads 5.9 and 5.11 is 14-15 D and

can be attributed to a partial charge separated ground state of the molecules. NBO

- 137 -

charge analysis of the model of the dyad 5.11 indicates the accumulation of the positive

charge on the donor moiety and overall negative charge on the 3CNQ acceptor ring with

value of charge-transfer equal 0.241. Additionally, TD-DFT calculation for the dyad

5.11 show electronic absorbance band with the maximum at 625 nm which is close to

the experimental value (See Characterization section 5.3). The energies of HOMO and

LUMO orbitals of the dyads clearly indicate on moderate increase of the electron-

donating properties of dyad 5.11 (–5.33 eV) compared to 5.9 (–5.85 eV) and, as a result,

decrease of the HOMO-LUMO gap from 2.12 to 1.77 eV. Further increase of the

number of the EDOT units leads to only minor decrease of the HOMO-LUMO gap (1.53

eV for trisEDOT-3CNQ, calculated by DFT, B3LYP/6-31G(d)).

Figure 5.4: Geometry optimization (DFT, B3LYP/ 6-31G(d)) of the dyads 5.9 and 5.11,

calculated HOMO/LUMO energies and dihedral angles between donor and acceptor

planes.

- 138 -

5.3. Characterization

The photochemical conversion of the compound 5.8 into 5.9 and 5.10 into 5.11

(Scheme 5.3) is accompanied by the appearance of an intense blue color of the solution.

In the electronic spectra, this reaction is manifested by appearance of the long-wave

absorbance band with maxima at 555 and 775 nm for dyads 5.9 and 5.11, respectively

(Figure 5.5a). We identify these bands with intramolecular charge transfer between

donor and acceptor in the dyads. The reaction was followed spectroscopically and

showed a clear isosbestic point around 360 nm thus indicating on clear conversion of the

nEDOT-TCNQ into nEDOT-3CNQ dyad (Fig. 5.5 b). Study of the absorbance of the

dyad 5.11 in different solvents shows moderate positive solvatochromic effect (Fig. 5.6)

which indicated a further polarization in the excited state.

CV characterization of the dyad 5.11 in the solution (Fig. 5.7) revealed a

reversible first reduction wave and partially reversible second reduction that correspond

to the formation of radical anion and dianion of 3CNQ moiety and one quasi-reversible

oxidation wave of nEDOT moiety. Electrochemically irreversible (yet chemically

reversible) nature of the EDOT oxidation was a subject of several studies and can be

attributed to the formation of the dimerization of radical cations [317-318] (see Chapter

6 for details). The reduction potential and energies of HOMO and LUMO orbitals for

both dyads are presented in the Table 5.1. As shown, no substantial difference in

reduction potentials of dyads 5.9 and 5.11 was observed (E0

1red vs. Fc/Fc+ are –0.76 and

–0.71 V, respectively). However, the oxidation potential is significantly lower for the

dyad 5.11 (Table 5.1) which is a result of better donor ability of bisEDOT moiety

compared to EDOT. As expected, raising the HOMO energy results in decreasing

HOMO-LUMO gap in the dyad 5.11 (1.1 eV) with respect to the dyad 5.9 (2.24 eV).

A preliminary spectroelectrochemial experiment shows changes in UV-Vis

absorbance of the solution of the dyad 5.11 during the electrochemical oxidation of the

donor moiety (Fig. 5.8). As we can see, the ICT peak at 790 nm decreases with applying

the oxidation potential from 0 to 1.4 V and new bands rise at 490 and 580 nm. TD-DFT

calculations for radical cation of dyad 5.11, which is expected to be formed during the

oxidation, predict the absorption band at 1300 nm, while we observed a blue shift which

- 139 -

200 400 600 800 1000 1200

Ab

so

rba

nce

wavelengh, nm

5.115.9

5.10 a5.8

300 400 500 600 700 8000.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

Ab

so

rba

nce

wavelengh, nm

Figure 5.5: a): UV-Vis spectra of dyad 5.8, 5.9, 5.10 and 5.11 in MeCN solution. b)

Dynamics of photochemical conversion of 5.1 in to 5.2 in CHCl3 by irradiation with UV

light (254 nm).

- 140 -

500 600 700 800 900 1000 1100 1200

0.0

0.2

0.4

0.6

0.8

1.0

DMF

Toluene

DCMA

bs,

no

rma

lize

d

wavelength, nm

Figure 5.6: Solvatochromism in the solution of dyad 5.11.

-1.5 -1.0 -0.5 0.0 0.53.0µ

2.0µ

1.0µ

0.0

-1.0µ

-2.0µ

Curr

ent,

A

Potential, V vs. Fc/Fc+

Figure 5.7: CV of dyads 5.11 (0.1 M Bu4NPF6 in CH2Cl2).

- 141 -

clearly indicates a disappearance of the charge transfer in the dyad upon oxidation of the

donor moiety. These new bands in the electronic spectra of the dyad 5.11 might be

attributed to the dimerization of the radical cation of EDOT moiety as observed for other

EDOT containing molecules by UV and electrochemical spectroscopy [318].

Table 5.1: Electrochemistry data for synthesised dyads (0.1 M Bu4NPF6 in CH2Cl2;

potentials vs. Fc/Fc+).

Dyad E10

red ,V E20

red,V E10

ox,V HOMOa,

eV

LUMOb,

eV

(HOMO-LUMO) eV

5.9 –0.76 – 1.48

p.a.

–6.3 –4.0 2.24

5.11 –0.71 –1.37 0.39

p.a.

–5.2 –4.1 1.1

a) determined from anodic oxidation peak (vs. Fc/Fc+) using the equation (HOMO= –4.8–+E

aox );

b) determined from the first reduction peak (vs. Fc/Fc+) using the equation (LUMO= –4.8–E1/21 red).

The analysis of the IR spectra of the dyads 5.9 and 5.11 (Fig. 5.9) shows that the

position of the νCN stretching of 3CNQ moiety (2200 and 2198 cm-1

for 5.9 and 5.11

respectively) in bulk corresponds to a significant ionic character of the acceptor.

Comparing to the IR of the TCNQ, where the νCN for the neutral species is at 2230 and

that for radical-anion species is at 2180 cm-1

[211], we can conclude that the degrees of

charge transfer in the dyads 5.9 and 5.11 are approximately 0.6 and 0.8, respectively.

Thus, IR spectroscopy confirms our earlier observations of charge transfer in the dyads.

As was expected, the degree of the charge transfer in the dyad 5.9 is slightly lower than

in the dyad 5.11, as the 5.9 is weaker donor.

- 142 -

400 500 600 700 800

0.28

0.29

0.30

0.31

0.32

0.33

0.34

0.35

Absorb

ance

wavelengh, nm

0V

1.4V

-0.1V

Figure 5.8: Spectroelectrochemistry of the dyad 5.11 (0.2M Bu4NPF6 in CH2Cl2). Black

line – initial state, red – oxidized at 1.2V, blue - reduced back at –0.1V.

The analysis of the IR spectra of the dyads 5.9 and 5.11 (Fig. 5.9) shows that the

position of the νCN stretching of 3CNQ moiety (2200 and 2198 cm-1

for 5.9 and 5.11

respectively) in bulk corresponds to a significant charge transfer to the acceptor moiety.

Comparing these results to the IR of the TCNQ and its salts, where the νCN for the

neutral species is at 2230 cm-1

and for radical-anion species at 2180 cm-1

[211] we can

conclude that the degrees of charge transfer in the dyads 5.9 and 5.11 are approximately

0.6 and 0.63 respectively. Thus, IR spectroscopy confirms our earlier observations of

charge transfer in the dyads. As was expected, the degree of the charge transfer in the

dyad 5.9 is slightly lower than in the dyad 5.11, as the 5.9 is a weaker donor. Another

interesting observation was made when comparing IR spectra of dyads 5.8/5.10 (TCNQ

intermediate) and 3CNQ dyads 5.9/5.11. For the TCNQ derivatives 5.8 and 5.10, no CN

vibrations were observed while dyads 5.9 and 5.11 show strong characteristic signal

from CN vibrations. To understand the absence of the νCN peaks in the dyads 5.8 and

5.10 we performed frequency calculations by DFT (B3LYP/6-31G (d) level).

- 143 -

500 1000 1500 2000 2500 3000

wavenumber, cm-1

dyad 5.10

experimental

calculated

500 1000 1500 2000 2500 3000

wavenumber, cm-1

dyad 5.11

experimental

calculated

Figure 5.9: ATR-FTIR of the dyads 5.10 (left) and 5.11 (right) and corresponding

calculated spectra (scaling factor for frequencies calculated by DFT B3LYP, 6-31G(d))

is 0.9603 [319]). All spectra are normalized.

The calculations predict strong νCN absorbance peak at 2232 cm-1

for the dyad

5.11 (Oscillator strength =1200) and a very weak peak at 2268 cm-1

for the dyad 5.10 (

=14) (Figure 5.9). Such weak νCN peak for 5.10 can be explained by high symmetry of

the vibrational modes of the CN group (Fig 5.10).

Figure 5.10: Vibrations of the CN groups in the dyad 5.10 (DFT, B3LYP/6-31G(d)

calculation).

5.4. Langmuir-Blodgett deposition of the monolayer of 5.12 on the solid substrates

The amphiphilic structure of the dyads 5.9 and 5.11 allowed us to form Langmuir

films of the molecules at the air-water interface and subsequently transfer them to a solid

surface as LB monolayers. Two types of the substrate were prepared for this experiment:

- 144 -

glass slides or rectangular cut Si wafer with thermally evaporated 400nm gold layera for

―Z-type‖ deposition (used for spectroscopic characterization of the monolayer) and p-

doped, low resistance, Si wafers [320-321] with the covalently attached monolayer of 1-

dodecene for ―Z-type‖ deposition (used for rectification study). We chose degeneratively

p-doped Si because it was shown that for this type of substrate the Schottky barrier is

low and Si/alkyl-chain/Hg junctions shows near symmetric charge transport

characteristics [321].

Preparation of the hydrophobic surface on the Si substrate was done following

the literature procedure [322] by functionalization of H-terminated Si surface with 1-

dodecene. The reaction was carried out until the water droplet contact angle reaches

110±5 and the monolayer thickness of 10-15Åb is achieved.

Deposition of the donor-acceptor molecules at the air-water interface was done

by spreading diluted solution of dyad 5.12 in chloroform (0.7 mg/ml) on the surface of

water followed by compression of the film with the barriers. The pressure-area isotherm

of the dyad 5.12 is presented in Figure 5.9. The isotherm is characterized by increase of

the surface pressure starting at the molecular area ca. 45-50 Å2

(limiting areac is 37 Å

2).

The theoretical area of 5.12 was calculated from the surface area of the smallest

rectangle in which the projections of the optimized model of the molecule can fit. The

value of the estimated molecular area (39.4 Å2) matches well with the experimental

results. The monolayer collapses at surface pressure above 40 mN m-1

.

a Adhesion layer of Ti (3-5nm) was deposited on glass slides before the gold evaporation

b Calculated length of the 1-dodecene is 13 Å.

c Limiting area was obtained from extrapolation of the linear part of the isotherm to the X axis.

- 145 -

10 20 30 40 50 60 700

10

20

30

40

50

Su

rfa

ce

pre

ssu

re (

mN

m-1

)

Area per molecule, A2

Figure 5.9: Surface pressure isotherm of the dyad 5.12.

The films of the dyad 5.12, compressed to 35 mNm-1

, were transferred on the

solid substrate as Z- and Z-types. The transfer of the Langmuir films on the solid

substrate was done at constant surface pressure and slow deposition speed (2mm/min) to

ensure uniformity of the film. The ratio of the upstroke transfer of the hydrophilic

substrate (gold) of the LB film was 0.9-1.0, whereas downstroke transfer on the

hydrophobic substrate (p-Si) was characterized by transfer ratio around 0.6-0.7.

However, this did not necessarily indicate a poor quality film and might be explained by

higher sensitivity of the deposition process to the roughness of the surface (one side of

the Si substrate was polished while other not). The experimental thickness of the LB

films on the gold substrates, measured by ellipsometry, is 20-25 Å and matches very

well with calculated length of the molecule (22.8 Å). This indicates that molecules 5.12

in the film are vertically oriented on the surface.

The LB monolayer of 5.12 on the gold surface was also studied by IR

spectroscopy and reveals the same features in IR spectra as was observed in bulk (Fig.

5.10). The position of the CN peak shows no significant change indicating same charge-

- 146 -

transfer character of the dyad in monolayers as in bulk phase. The position of the

asymmetric and symmetric methylene vibration modes for the LB film on the gold

substrate are 2928 and 2860 cm-1

respectively, and are slightly shifted, in comparison

with crystalline sample of the dyad 5.12, which indicates some disorder in the alkyl tails

of the dyads (compare to 2918 and 2850 cm-1

for highly ordered film structure of SAMs

of

3500 3000 2500 2000 1500

99.0

99.5

100.0

T, %

wavenumber, cm-1

2198 cm-1

3200 3100 3000 2900 2800 2700

wavenumber, cm-1

2860 cm-1

2856 cm-1

2924cm-1

2928 cm-1

bulk

LB film

Figure 5.10: Grazing angle FTIR of the LB monolayer of the 5.12 on the gold slide

(left). Comparison of the CH2 vibration bands of the LB film and bulk (right).

alkyl thiols on gold and LB films of fatty acids at the air-water interface [149, 151]).

This indicates that floppy hexyl tails might have enough space above the donor-acceptor

core and do not form an ordered structure. However, the results of the ellipsometry study

and molecular area in the LB film allow us to conclude that the donor-acceptor cores of

the dyads are oriented vertically on the surface which is important for rectification

measurements.

5.5. Rectification measurements of the LB films of the dyad 5.12 in mercury-drop

junctions

To study the rectification ratio of the nEDOT-3CNQ dyads we have used

mercury-drop junction setup described in Chapter II. The LB film of the dyad 5.12 was

sandwiched between gold or p-doped silicon electrodes and a mercury drop electrode

- 147 -

pre-coated with an alkanethiols monolayer. As was already discussed in previous

Chapters, the SAM on mercury electrode is necessary to minimize the chance of direct

contact between the two metal electrodes. However, the stability of the junctions where

LB film of 5.10 was transferred on the gold electrodes was still very low, which

precluded corresponding conductivity measurements. The need for a thiolate layer on

Hg electrode is less critical in the case of p-Si electrode as it cannot form amalgam with

mercury, however, the alkanethiol SAM on the mercury surface is still necessary for

symmetric positioning of the bisEDOT-3CNQ core between electrodes Thus, the

rectification study was done using the junctions (p-Si-C12/C6-bisEDOT-3CNQ/C16-Hg)

as presented on Figure 5.11.

Figure 5.11: Scheme of the p-Si-C12/C6-bisEDOT-3CNQ/C16-Hg junctions.

The LB monolayer of 5.11 was transferred on the surface of the dodecene-

functionalized p-Si (with hexyl chains towards the solid surface) and the electrical

junction was established by contacting thus prepared surface with mercury electrode

- 148 -

functionalized with C16H33SH SAM. Rectification study of the dyad 5.11 in this type of

junctions showed definite current asymmetry with the higher current in the negative

quadrant of the I-V plot (Fig. 5.12). The statistical analysis shows close to exponential

distribution of the rectification ratio, with the average log(RR)=0.97±0.04 at ±0.4 V

(Fig. 5.13). The maximum RR (12) was observed at ±0.4 V (Fig. 5.14) which followed

by decrease at higher voltage. This suggests that at higher potential the direct tunneling

of the current between two electrodes becomes a dominating process and overshadows

the specific details of the molecular structure.

-1.0 -0.5 0.0 0.5 1.0

-80.0n

-60.0n

-40.0n

-20.0n

0.0

20.0n

Curr

ent, n

AVoltage, V

Figure 5.13: A) Current-voltage curve for the LB monolayer of dyad 5.11 between

mercury and p-Si electrodes, RR=12 at ±0.4 V.

- 149 -

0.8 0.9 1.0 1.1 1.20

2

4

6

8

10

Cou

nt

log(RR)

0.0 0.2 0.4 0.6 0.8 1.0

0

2

4

6

8

10

12

14

RR

Potential, V

Figure 5.14: (A) Statistical distribution of the log(RR) for p-Si/5.11/Hg junctions. (B)

The dependence of the rectification ratio in the p-Si-C12/C6-bisEDOT-3CNQ/C16-Hg

junction on the applied bias.

- 150 -

To prove the molecular origin of the rectification we compared the electron

transport properties of the junctions without the dyad 5.12. As we can see from the

Figure 5.15 these junctions show a more symmetric conductivity with slightly higher

current at positive bias (logRR=0.12±0.58). Importantly, since for the junctions with the

LB film of the dyad 5.12 the directions of the rectification was the opposite (higher

current at negative bias), we can conclude that the rectification observed in such

junctions is not due to intrinsic asymmetry of the contacts.

Thus, we can see that the asymmetry of the I-V characteristics of the junctions is

somewhat smaller to usually reported for other LB monolayer rectifiers (see Table 1.1).

Nevertheless, the direction of the current from the acceptor (TCNQ) to the donor

(bisEDOT) is in agreement with the Aviram-Ratner mechanism.

- 151 -

-0.2 -0.1 0.0 0.1 0.2-11

-10

-9

-8

-7

-6

log

(Cu

rre

nt)

Voltage, V

-1.0 -0.5 0.0 0.5 1.0 1.50

3

6

9

12

15

Co

un

t

log(RR)

Figure 5.15: Current-voltage characteristics of the p-Si-C12//C16-Hg junctions (top), and

Statistical distribution of log(RR) for p-Si-C12//C16-Hg junctions (bottom).

- 152 -

Conclusions

In this work we synthesised new donor-acceptor dyads based on nEDOT as the

electron-donating part and 3CNQ as the electron-withdrawing part. The new dyads

present relatively low HOMO-LUMO gap (1.05 eV for the 5.11) and good air stability.

The molecule was designed with amphiphilic structure for deposition of the electrode

surface via Langmuir-Blodgett technique. The rectification properties of the dyads were

studied in the LB films sandwiched between p-doped, oxide free hydrophobic Si

substrates from one side and SAM-protected liquid mercury electrode from the other.

Despite the significant HOMO-LUMO delocalization, characterized by presence of

strong ICT band, we observed small but clear asymmetry (log(RR)=0.97±0.04) in the

electrical conductivity of the LB films. This rectification behaviour was attributed to the

molecular origin, as was shown by control measurements without the LB films in the

junctions. The direction of the enhanced current flow is in agreement with the Aviram-

Ratner model and opposite to the C16H33Q-3CNQ dyad.

- 153 -

Experimental part

Cyclic voltammetry measurements were done using a CHI- 760C potentiostat under

nitrogen in a CH2Cl2 solution of electrolyte (0.1 M Bu4NPF6) with a Ag/AgCl

reference electrode and platinum disk (d=1.6 mm) as a working electrode. Fc/Fc+

(0.50 V vs. Ag/AgCl in these conditions) was used as an internal reference.

Calculations of geometry and electronic structure of the dyads were done using

density functional theory (DFT) with hybrid B3LYP functional and 6–31G(d) basis

set, as implemented in Gaussian W03 [228]. The alkyl substituent on donor moiety

was modeled using methyl group.

Absorbance/Emission spectroscopy. Absorption spectra were recorded with a Jasco

V-670 spectrophotometer in CH2Cl2 solution.

Preparation of Langmuir–Blodgett (LB) films: Single monolayers were prepared

at 20C on an aqueous (18.2 MOhm H2O) subphase by using 400 cm2 KSV 3000

(KSV Instruments, Helsinki, Finland) Langmuir–Blodgett (LB) trough. The

molecules were first dissolved HPLC grade chloroform (~0.7 mg/ml) and then

immediately spread to the subphase to form the monolayer. After an equilibrating

period of 20-30 min allowing solvent evaporation, the monolayer was compressed at

constant speed of 10 mm/min and transferred at constant surface pressure onto the

surface of interest (p-Si or Au electrodes for I–V experiments and spectroscopy

studies).

FTIR Spectroscopy. See Experimental part in Chapter III for details.

Ellipsometry. Monolayer thickness was measured on Gaertner ellipsometer

equipped with a He-Ne laser (λ= 632.8 nm) at an incidence angle of 70° with respect

to the surface normal. Optical constants of the gold-coated substrates were measured

using a bare gold slide (Ns=0.25, Ks= –3.46). Reference sample was cleaned by

soaking in HPLC-grade CH2Cl2 and exposed to air plasma immediately before the

measurements. The layer thickness was calculated by averaging over 10

measurements. The refractive index of the monolayer was assumed to be 1.46.

Spectroelectrochemistry experiments were performed in thin layer

spectroelectrochemial cell CHI140A from CH Instruments equipped with platinum

grid as a working, platinum wire as a counter and Ag/AgCl as a reference electrode.

UV-Vis spectra were recorded with a Jasco V-670 spectrophotometer and a BASi

Epsilon potentiostat and the static potential mode was used for oxidation/reduction of

the molecules. Potentials were applied in 50-100 mV steps and equilibrated by

allowing the current to drop until a negligible current change was achieved (less than

1% of initial current per minute).

Rectification measurements in mercury drop junctions. For rectification

measurements experiments the LB of 5.11was transferred on p-doped Si substrate.

The preparation of the p-Si substrate was done as follow. The p-doped Si wafer with

native oxide layer (resistance 0.01 Ω/cm2) was cleaned with freshly prepared

―piranha‖ solution (H2SO4:H2O2, 3:1) from organic contaminants and then immersed

in the 40% aqueous solution of NH4F for 30 min to remove the oxide layer and

produce H-terminated Si surface. The time of the immersing in the NH4F solution

was elaborated experimentally by ellipsometry measurements of the SiO2 layer

thickness before and after etching. After etching procedure the substrate was quickly

- 154 -

washed with water and dried under nitrogen flow. Then Si slides were placed in the

Schlenk tube containing neat 1-dodecene under inert atmosphere and heated to 120C

for 12h. The time of the deposition was elaborated experimentally by testing the

quality of the deposited SAM with contact angle and ellipsometry measurements.

Mercury electrode was covered with monolayer of hexadecanethiolate (10-3

M,

ethanol, 15-20 min).The junctions were assembled in a procedure, similar to the

described in the Chapter II.

bisEDOT-TCNQ (5.1). A mixture of TCNQ (0.1g, 0.5mmol) and bisEDOT (0.127 g,

0.50 mmol) in PhCl (20 ml) was stirred at 80C for 24h under inert atmosphere of N2

and 100mW light source. Then the solvent was evaporated under reduced pressure and

the residue was purified by the column chromatography (CH2Cl2:EtOAc 3:1) resulting in

colorless solid (0.21 g, 92%).1H NMR (300 MHz, CDCl3) 7.86 (s, 1H), 7.84 (s, 1H),

7.69 (s, 1H), 7.67 (s, 1H), 6.34 (s, 1H), 5.16 (s, 1H), 4.37-4.21 (m, 8H). UV-Vis (CH2Cl)

max (nm) 331, 349.

bisEDOT-3CNQ (5.2). A solution of 5.1 in CD2Cl2 was irradiated by UV light

(254nm). The reaction was performed in UV cell and monitored by UV-Vis

spectrometer until no further increase of the absorbance at 719 nm was observed. The

resulting blue solution was characterized by NMR and Mass-spectrometry. 1H NMR

(300 MHz, CDCl3) 7.86 (s, 1H), 7.84 (s, 1H), 7.69 (s, 1H), 7.67 (s, 1H), 6.34 (s, 1H),

4.37-4.21 (m, 8H). MS (EI) 459. UV-Vis (CH2Cl) max (nm) 719.

2-hexyl-EDOT (5.3) [323]. To a solution of EDOT (1 g, 7mmol) in dry THF (100ml) at

–78C a solution of nBuLi (2.5M, 3ml) was added dropwise. The reaction mixture was

slowly warmed to 0C, stirred for 1 hour at this temperature and then cooled to –78C.

Then 1-iodohexane (1.17 g, 7.1 mmol) was added and mixture was stirred at –50 –40C

overnight. The crude product was extracted with EtOAc. After the solvent was

evaporated, the organic phase was washed with water, brine and dried over MgSO4.

Purification was performed by column chromatography on silica using hexanes:EtOAc

(5:1) as eluent resulting in final product 5.3 as a colorless oil (0.56 g, 35%). 1H NMR

(300 MHz, CDCl3) 6.10 (s, 1H), 4.17 (m, 4H), 2.63 (t, 2H, J 7.6Hz), 1.6-1.61 (m, 4H),

1.4-1.36 (m, 4H), 0.88 (t, 3H, J 7.2 Hz).

5-Hexyl-bisEDOT (5.6) [323]. To a solution of bisEDOT [315] 5.5 (2.0 g, 7.1mmol) in

THF (200ml) at –78C a solution of nBuLi (2.5M, 3ml) was added dropwise. The

reaction mixture was slowly warmed to 0C, stirred for 1 hour at this temperature and

then cooled to –78C. Then 1-iodohexane (1.17 g, 7.1 mmol) was added and mixture

was stirred at –50...–40C overnight. The crude product was extracted with EtOAc.

After the solvent was evaporated, the organic phase was washed with water, brine and

dried over MgSO4. Purification was performed by column chromatography on silica

using hexanes:EtOAc (3:1) as eluent resulting in final product 5.6 as a yellow powder

(0.95 g, 37 %) and dihexylated by-product 5.7 as a white solid (26%). 5.6: M.p. 118-120

C (lit[323] 120 C); 1H NMR (400 MHz, CDCl3) 6.22 (s, 1H), 4.3-4.19 (m, 8H), 2.63

(t, 2H, J 7.6Hz), 1.6-1.61 (m, 4H), 1.4-1.36 (m, 4H), 0.88 (t, 3H, J 7.2 Hz). 5.7: δH (400

MHz, CDCl3) 4.3-4.19 (m, 8H), 2.63 (t, 2H, J 7.6Hz), 1.6-1.61 (m, 4H), 1.4-1.36 (m,

4H), 0.88 (t, 3H, J 7.2 Hz).

2-Hexyl-EDOT-TCNQ (5.8). Mixture of 5.4 (0.10 g, 0.44 mmol) and TCNQ (0.09 g,

0.44 mmol) in dry MeCN (10ml) was stirred at reflux for 12 h under atmosphere of N2.

- 155 -

The reaction mixture immediately turns dark green after mixing indicating formation of

CTC. Then solvent was evaporated and residue was purified by chromatography

(EtOAc:hexanes 1:1) resulting in yellow solid (0.153 g, 81 %). m.p. 211-212oC.

1H

NMR (500 MHz, CDCl3): 7.84 (s, 1H), 7.83 (s, 1H), 7.68 (s, 1H), 7.67 (s, 1H), 5.15 (s,

1H), 4.31-4.21 (m, 4), 2.58 (t, 2H, J=7.5Hz), 1.54 (m, 2H), 1.32-1.28 (m, 8H), 0.88 (t,

3H, J=6), 13

C HMR (125 MHz, CDCl3): 131.2, 128.9, 128.5, 128.4, 128.3, 112.7, 110.9,

65.2, 64.3, 31.4, 30.1, 28.7, 27.8, 25.7, 22.5, 14.0. HR-MS (APCI): calculated for

C24H23N4O2S (M+1) 431.1536, found 431.1528. UV-Vis (MeCN) max (nm) 400.

2-Hexyl-EDOT-3CNQ (5.9). A solution of 5.9 (0.050 g, 0.12 mmol) in dry and

degassed MeCN (100 ml) was irradiated with UV lamp at 254 nm for 10 min, until no

starting material was observed by TLC. The initial light green solution changed color to

deep blue. Then the solvent was evaporated and purified by column chromatography

(EtOAc:hexanes 1:3, then only EtOAc) resulting in purple solid (0.042 g, 89%). m.p.

227-228oC.

1H NMR (300MHz, acetone-d6): 7.25 (s, 1H), 7.22 (s, 1H), 6.98 (s, 1H),

6.95 (s, 1H), 4.40-4.27 (m, 4H), 2.58 (t, 2H, J=6Hz), 1.53 (m, 2H), 1.39-1.19 (m, 8H),

0.86 (t, 3H, J=6.3Hz). HR-MS (ESI): calculated for C23H22N3O2S (M+1) 404.1427,

found 404.1414. UV-Vis (MeCN) max (nm) 555.

5-Hexyl-bisEDOT-TCNQ (5.10). Mixture of 5.7 (0.10 g, 0.27 mmol) and TCNQ

(0.06g, 0.3mmol) in MeCN (5ml) was stirred at reflux for 12h. After the reaction

mixture was cooled to room temperature the precipitate was collected by filtration and

washed with hexanes. The crude material was recrystallized from EtOAc:Hexanes

mixture resulting is pure yellow solid (0.152 g, 97%). m.p. 168-150oC.

1H NMR (500

MHz, CDCl3): 7.86 (d, 1H, J=2), 7.85 (d, 1H, J=2.5), 7.68 (s, 1H), 7.66 (s, 1H), 5.15 (s,

1H), 4.36-4.18 (m, 8), 2.63 (t, 2H, J=7.5Hz), 1.54 (m, 2H), 1.40-1.27 (m, 8H), 0.88 (t,

3H, J=7), 13

C HMR (125 MHz, CDCl3): 140.7, 138.0, 136.9, 135.9, 134.6, 128.54,

128.45, 128.37, 118.9, 113.2, 112.6, 111.0, 103.9, 101.2, 65.3, 65.2, 64.8, 64.4, 31.5,

30.3, 28.8, 27.8, 25.8, 22.5, 14.1. HR-MS (APCI): calculated for C30H27N4O4S2 (M+1)

571.1467, found 571.1461. UV-Vis (MeCN) max (nm) 341, 355.

5-hexyl-bisEDOT-3CNQ (5.11). A solution of 5.11 (0.05g, 0.08mmol) in dry and

degassed MeCN (100 ml) was irradiated with UV light at 254nm for 10min. The

precipitate that was formed upon irradiation was collected by filtration and mother

liquor was irradiated with UV light for additional 10 min. The combined portions of

precipitates were washed with MeCN resulting in desired product as dark blue solid

(0.043g, 91%). M.p. 219-222 o

C. 1H NMR (300MHz, acetone-d6): 7.95-7.93 (m, 4H),

4.45-4.43 (m, 4H), 4.30-4.25 (m, 4H), 2.65 (t, 2H, J=7.8 Hz), 1.59 (m, 2H), 1.40-1.23

(m, 8H), 0.89 (t, br, 3H). HR-MS (APCI): calculated for C29H26N3O4S2 (M+1)

544.1359, found 544.1349. UV-Vis (MeCN) max (nm) 775.

- 156 -

Chapter VI. Stable nEDOT-NDI molecular rectifiers with self-

assembly capability.

(Part of this Chapter was adapted with permission from: M. Kondratenko, A. Moiseev,

D.F. Perepichka, New stable donor-acceptor Dyads for molecular electronics, J. Mater.

Chem. 2011, 21, 1470–1478. Copyrights 2011 The Royal Society of Chemistry)

Introduction

Among different problems related to the design of molecular rectifiers, the

stability and lack of rigidity are very important issues. For example, several reported

donor-acceptor dyads with the -bridge can have several energetically comparable

conformers, one of which allows through-space charge transfer complexation between

donor and acceptor [69, 72]. Intending to create simple, rigid and stable systems that

could function as unimolecular rectifiers, we turned our attention to two well-studied

molecular building blocks: 1,4,5,8-naphthalenetetracarboxydiimide (NDI) and 3,4,9,10-

perylenetetracarboxylic diimide (PDI). NDI and PDI are very interesting planar electron-

deficient molecular systems with exceptional chemical and thermal stability. These

molecules have been widely employed as acceptors in model donor-acceptor dyads

(used to study the fundamentals of electron transfer [324-328] and spin dynamics [329-

330]), as an n-channel semiconductor in organic field-effect transistors (OFETs) [331-

335] and photovoltaics [336-337], etc. Of particular relevance to this study, hybrid NDI-

thiophene oligomers [338] and polymers [339] showed ambipolar charge-transport

properties in OFETs, along with remarkable air-stability. As a donor fragment for this

project we used 3,4-ethylenedioxythiophene (EDOT) and bis-3,4-

ethylenedioxythiophene (bis-EDOT) that were also employed as electron donor moieties

in the donor-acceptor dyads described in Chapter V of this thesis. Our idea was that

coupling these donor and acceptor fragments together would provide dyads with very

high stability and desirable electronic properties for applications in molecular

electronics. As discussed in Chapter I, PDI was previously used as an acceptor moiety

for molecular rectifiers as a part of D--A dyad [340], however, due to very low

solubility of the starting 3,4,9,10-perylenetetracarboxylic dianhydride this building block

is difficult for asymmetric functionalization. Another advantage of these acceptor

- 157 -

building blocks is in the possibility for core-substitution in naphthalene [324] and

perylene [341-342] leading to improvement of the acceptor properties of the molecules.

Herein we discuss synthesis and detailed characterization of the electronic

properties for a series of new donor-acceptor dyads NDI-EDOT (6.7), NDI-bisEDOT

(6.7), RS-NDI-bisEDOT (6.8) and NDI-bisEDOT-SR (6.9) that carry NDI acceptor and

EDOT donor moieties linked together through a phenyl bridge. The protected thiol

functionality is introduced to provide covalent attachment of the molecules to gold

electrodes, either on the acceptor side or on the donor side. Corresponding self-

assembled monolayers (SAMs) on gold were prepared from solution and their

spectroscopic properties and electrochemical behaviour were examined.

Figure 6.1: Series of synthesized Donor-Acceptor dyads consisting of nEDOT and NDI

moieties.

- 158 -

6.1. Synthesis of nEDOT-NDI dyads

To obtain donor-acceptor dyads with desired solubility and self-assembly functionalities

the NDI acceptor and nEDOT donor moieties were asymmetrically modified with linear

or brancheda alkyl chains, or with a thiol-containing group 4.21 (Scheme 5.1).

Scheme 6.1. Synthesis of acceptor synthons.

Bromine-functionalized acceptor synthons 6.4 and 6.5 were prepared from a

commercially available 1,4,5,8-naphthalenetetracarboxylic dianhydride 6.1 through a

sequential reaction with corresponding primary amines to yield monoimides 6.2 and 6.3b

followed by condensation with 4-bromoaniline (Scheme 6.1). A symmetric diimide side

product 6.2a was isolated from the synthesis of 6.2 and used as a model acceptor

molecule for comparative electrochemical and spectroscopic studies.

a Although branched chains could prevent efficient packing of molecules in the monolayers, their

presence was found necessary to achieve sufficient solubility in these dyad molecules b Synthesis of tert-butylsulfanylhexylamine was described in Chapter IV as compound 4.21

- 159 -

Tin-functionalized donor synthons 6.12 and 6.13 were prepared by

monoalkylation of a lithium salt of bisEDOT with an appropriate alkyl iodide (1-

iodohexane or 6.16) to give the monoalkylated bisEDOT derivatives 6.10 and 6.11

(Scheme 6.2), followed by a second lithiation and coupling with tributyltin chloride.

Dihexyl-bisEDOT side product 6.10a was isolated from the synthesis of 6.10 and used

as a model donor compound. Similarly, alkylation of unsubstituted EDOT produced a

model donor compound dihexylEDOT 6.14.

The final assembly of dyads 6.6, 6.7, 6.8a and 6.9a was achieved through Pd-

catalyzed Stille coupling of acceptor synthons 6.4 or 6.5 with donor synthons 6.12, 6.13

or 6.15 (Scheme 6.3).

Scheme 6.2. Synthesis of the asymmetrically functionalized donor synthons.

- 160 -

Scheme 6.3. Synthesis of the donor-acceptor dyads.

The tert-butyl protecting group in molecules 6.8a and 6.9a is necessary for

successful synthesis, purification and prolonged storage of these thiol-functionalized

dyads. In order to achieve grafting of the molecules on gold electrodes, the tert-butyl

group was removed [288] at low temperature, by treatment with a strong Lewis acid

(BBr3). Then, addition of acetyl bromide to the reaction mixture led to the formation of

corresponding acetylsulfanyl derivatives 6.8b and 6.9b that act as more stable

equivalents of corresponding thiols 6.8 and 6.9 and can be attached to gold electrodes in

the presence of a catalytic amount of aqueous ammonia.

Thermogravimetric analysis (TGA) showed excellent thermal stability of the dyad 6.7

Tdec of 390C (5% loss) was measured for the dyad 6.7 under nitrogen. (Appendix, Fig.

A1).

6.2. Calculations

In order to investigate the electronic structure of the prepared dyads, model

molecules NDI-EDOT, NDI-bis-EDOT and NDI-tris-EDOT with methyl substituents

- 161 -

were calculated with Density Functional Theory (DFT) at B3LYP/6-31G(d) level of

theory (Fig. 6.2). We were particularly interested in the energy and distribution of the

HOMO and LUMO orbitals within the dyads. Optimized molecular geometries predict a

large (72º) dihedral angle between the phenyl bridge and the acceptor, preventing

conjugation between the donor and the acceptor. On the other hand, a moderate (18–20º)

dihedral angle between the EDOT and the phenyl rings should allow substantial electron

delocalization between the donor moiety and the bridge. Indeed, the orbital topology

shows that the LUMO orbital is fully localized on the NDI moiety and the HOMO is

mostly localized on the EDOT (bis-EDOT, tris-EDOT) fragment but partially extends on

phenyl ring (Fig. 6.2). The contribution of the phenyl bridge to the HOMO decreases

with an increasing number of EDOT units in the donor fragment. Overall, calculations

predict asymmetric distribution and spatial separation of the HOMO and LUMO orbitals

and confirm high rigidity of the molecules with no possibility for intramolecular

complexation between the donor and the acceptor moieties. The calculated HOMO-

LUMO gap is reduced dramatically, from 2.2 eV to 1.45 eV upon introduction of a

second EDOT ring into the donor moiety. However, adding the third EDOT ring leads to

only a moderate further decrease of the gap (1.10 eV for NDI-tris-EDOT). This trend

rationalizes our choice of bis-EDOT based dyads 6.7, 6.8 and 6.9 as the synthetic targets

of this study.

The calculations also predict a relatively low polarity of the designed dyads. The

dipole moment of the NDI-bisEDOT molecule is only 2.7 D, which is drastically lower

than that of the first and most extensively studied molecular rectifier C16H33Q-3CNQ (25

D).[343] and of dyad 5.11 described in Chapter V (15 D). We note that repulsive dipole-

dipole interactions of the latter have been previously identified as one of the reasons for

the low cycling stability of molecular rectifiers made from C16H33Q-3CNQ [44].

6.3. Absorption/Emission spectra

First, we have analyzed the UV-Vis spectra of CH2Cl2 solutions of separate

donor and acceptor molecules. The model donor compounds dihexyl-EDOT (6.14) and

dihexyl-bisEDOT (6.10a) exhibit absorption in the UV-Vis region with maxima at 290

- 162 -

nm and 330 nm, respectively (Fig. 5.3). This 50 nm red shift is due to the extended

conjugation of the π-conjugated EDOT-EDOT system [344]. Vibronically structured

absorption of 6.10a (peaks at 317, 329 and 345 nm) is indicative of the rigid structure of

bisEDOT moiety. Note that conjugation of the bisEDOT moiety with the phenyl bridge,

as in dyad 6.7 should cause a further bathochromic shift as was observed for diphenyl-

bisEDOT that displays a absorption band with vibronic peaks at 375, 400, and 427 nm

[345].

Figure 6.2: Calculated molecular orbitals for nEDOT-NDI dyads.

- 163 -

250 300 350 400

6.2a

6.6

6.7

6.10a

6.14

No

rmaliz

ed a

bsorb

an

ce

Wavelength /nm

Figure 6.3: UV-Vis spectra of DBA dyads 6.6 and 6.7 and separate model donor (6.10a,

6.14) and acceptor model molecule (6.2a) in MeCN.

The absorption spectrum of acceptor 6.2a is dominated by a typically (for NDI)

strong π→π* transition at max 383 nm, also with clear vibronic structure. A conjugation

of the donor moieties with the phenyl bridge leads to a bathochromic shift of their

absorption vs. the models 6.10a and 6.14. For comparison, diphenyl-EDOT absorbs at

max = 345 nm, diphenyl-bisEDOT absorbs at max = 387 nm [345]. Thus, the absorption

spectra of the dyad molecules are essentially a superposition of absorptions of the donor

and acceptor moieties, showing no evidence for intramolecular charge-transfer in the

ground state (Fig. 6.3). Furthermore, lack of long-wavelength ―charge-transfer‖

absorbance and no change in the absorbance spectra observed in

- 164 -

400 600 800 1000 12000.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

1.1

Absorb

ance/

a.u

.

Wavelength/nm

600 800 1000

0.02

0.03

Absorb

ance/ a.u

.

Wavelenght/ nm

Figure 6.4: Absorption spectra of a spin-coated film of the dyad 6.7 on glass (black, the

inset shows a magnification of the 500 – 1500 nm region) and SAM of the dyad 6.8b

(red) on gold (normalized).

-6.5 -6.0 -5.5 -5.0 -4.5 -4.0 -3.5 -3.0

-2.0

-1.5

-1.0

-0.5

0.0

0.5

1.0

log

(lA

) a

t 4

00

nm

log(concentration)

y=3.54+0.85x, R2=0.999

Figure 6.5: Linear dependence of the absorbance of the 6.6a in CH2Cl2 vs.

concentration.

- 165 -

the wide concentration range of 6.7 (510–7

–510–3

M in CH2Cl2) (Fig. 6.5) suggests

that intermolecular charge-transfer complexation is not significant between these

molecules in solution. This is in line with the relatively weak charge-transfer

complexing (CTC) ability of NDI (e.g. association constants for CTC of N,N´-dihexyl-

NDI with pyrene is 20 M–1

) [346]. UV-Vis spectra of thin films of the dyad 6.8b on the

glass slide (Fig. 6.4, black solid line) showed essentially the same absorbance as in

solution, with max = 380 nm, apart from a new extremely weak absorption band at max

= 695 nm. This weak absorption can be attributed to intermolecular charge transfer in

the solid state (Fig. 6.4, inset). Similar long wavelength absorption was observed for

CTCs of N,N’-dipyridyl-NDI with other π-donors in the solid state [347].

Study of fluorescence in solution reveals a rather weak photoluminescence (PL)

of the model acceptor compound 6.2a (PL = 0.34% in CH2Cl2), which is slightly

lowered to PL = 0.27% upon attachment of the nEDOT donor moiety in dyads 6.6 and

6.7. It was suggested earlier that fast intersystem crossing (τPL=16.4 ps [348]) is

responsible for the low fluorescence quantum yield. The emission spectrum of 6.7 is

bathochromically shifted compared to emission of the dyad 6.6, and is broader than that

of the model 6.2a containing only the NDI moiety (Fig. 5.6a). Its position is consistent

with an expected emission of the phenyl-substituted bisEDOT structure [349]. This can

likely be attributed to the resonance energy transfer between acceptor and donor

fragments in the dyad 6.7 since the absorbance of the bisEDOT moiety overlaps with the

emission of the and NDI moiety. Such energy transfer was not observed in the dyad 6.6,

where the absorbance of the EDOT moiety occurs at higher energy.

Bathochromic shift of the NDI emission is also possible due to the formation of

the excimer [350], but it can usually occur in polar solvents like toluene and was

observed at 500 nm for 6.7 (Fig. 6.6 b).

- 166 -

350 400 450 500 550 600

6.2a

6.6

6.7

6.10a

Flu

ore

scence inte

nsity /a.u

.

Wavelength /nm

400 450 500 550 600

in PhMe

in MeCN

Flu

ore

sce

nce

inte

nsity /

a.u

.

Wavelenth / nm

b

Figure 6.6: a) normalized emission spectra of the dyads 6.6 and 6.7, NDI acceptor 6.2a

and bisEDOT donor 6.10a in MeCN (excitation at 340 nm); b) normalized emission of

dyad 6.7 in MeCN and in PhMe with clear evidence of excimer formation.

- 167 -

6.4. Electrochemistry

The electrochemical behaviour of the synthesized molecules was studied using

cyclic voltammetry (Fig. 6.7, Table 6.1). All NDI derivatives showed two well-

separated, reversible, one-electron reduction waves corresponding to the formation of

the radical-anion and dianion. The reduction potentials of dyads 6.6 and 6.7 were

slightly less negative than those of the model acceptor 6.2a and did not depend on the

nature of the donor moiety. This can be attributed to the slightly electron withdrawing

effect of the phenyl bridge, in comparison with the electron donating nature of the alkyl

group. The one-electron oxidation corresponding to formation of a radical-cation on the

nEDOT fragment was electrochemically irreversible. The corresponding oxidation

potential decreases by ~600 mV upon addition of the second EDOT unit to the donor

fragment (6.6 6.7), whereas the reduction potentials stay practically unaffected. The

HOMO and LUMO values deduced from electrochemistry (Table 6.1) are in reasonable

agreement with values calculated by DFT (Fig. 6.1). For dyads 6.6 and 6.7, the

difference between calculated HOMO-LUMO gap and the gap measured in

electrochemical experiments is 0.2 and 0.05 eV, respectively (Table 6.1).

-2.0 -1.5 -1.0 -0.5 0.0 0.5 1.0

Potential/ V

15 A 6.6

6.730 A

Figure 6.7: Cyclic voltammogram of 6.6 and 6.7 in CH2Cl2 vs. Fc/Fc+.

- 168 -

Table 6.1: Electrochemical data for the synthesized dyads and model donor and

acceptor compounds.

E01red,

V

E02red, V E

pa1ox, V HOMO,

a eV LUMO,

b eV Gap, eV Gap,,

eV(calc.)

6.2a –1.13 –1.59 –

6.6 –1.06 –1.51 0.98 –5.8 –3.7 2.0 2.20

6.7 –1.07 –1.52 0.31 –5.1 –3.7 1.4 1.45

6.10a 0.36

6.14 – – 0.97

a) determined from anodic oxidation peak (vs. Fc/Fc+) using the equation (HOMO= –4.8–+E

aox ); b)

determined from the first reduction peak (vs. Fc/Fc+) using the equation (LUMO= –4.8–E1/21 red).

The electrochemically-irreversible nature of the oxidation requires a special

discussion in light of our aim to create highly stable donor-acceptor dyads. Such

behaviour was also observed for model dihexyl-EDOT 6.14 and dihexyl-bisEDOT 6.10a

and was previously speculated to be due to dimerization of the radical-cations formed

[344]. To prove the nature of the oxidized species we have performed bulk electrolysis

of the model donor 6.10a. The experiment was done in the electrochemical cell for bulk

electrolysis in degassed MeCN solution by applying constant potential at 1.0 V for 20

min. The analysis of EPR spectra of the oxidized species showed no radical species

present in solution. UV-Vis spectroelectrochemical studies in dry, degassed MeCN

reveal that upon gradual increase of the redox potential, the absorption of the neutral

molecule in the 300–360 nm region is gradually replaced by a new absorption in the

360–500 nm region, with clear isosbestic points at 258 and 355 nm (Fig. 6.8). The

absorption of oxidized species grows rapidly (the equilibrium is reached within ~1-2

min at each new potential value). During the backward reduction of the product,

disappearance of the absorbance band at 360–500 nm and recovery of the original

spectrum occurs at much slower rate (15–20 min required to reach the equilibrium after

each change of the potential). Such slow reduction explains the irreversible CV.

Nevertheless, the neutral molecule can be fully recovered upon reversal of the potential.

A small additional shoulder at ~520 nm which is visible on the forward (oxidative)

direction but which quickly disappears and is not observed during the reduction sweep is

- 169 -

likely due to absorption of the radical-cation transient. Based on the above observation,

the overall process appears as depicted in Scheme 6.5. An electron transfer from 6.10a

onto the electrode forming the radical cation is followed by a fast dimerization process

to give the bisEDOT dimer dication (6.10a)22+

. Two-electron reduction of this dication,

which leads to recovery of the neutral monomeric 6.10a, however, is a slow process that

leads to an electrochemically-irreversible CV signature.

300 400 500 6000.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8

2.0

Absorb

ance

Wavelenth/ nm

Figure 6.8: Spectroelectrochemistry of the bisEDOT 6.10a in MeCN; red – oxidation,

blue – reduction.

- 170 -

Scheme 6.5. Proposed scheme of the reversible oxidative dimerization of bisEDOT

derivatives.

Similar results were obtained in a spectroelectrochemical study of the dyad 6.7

(Fig. 6.9). Neutral 6.7 exhibits absorption at 325-420 nm region which is an overlap of

the donor and acceptor absorption bands. Upon gradual increase of the oxidation

potential, from 0 to 1.3 V, a new absorption band of the oxidized species appears at max

= 505 nm. The shoulder at ~400 nm corresponding to the bisEDOT moiety attenuates

but the vibronically-split band of NDI at 380 nm persists. Electrochemical reduction

back to the neutral state leads to complete disappearance of the oxidized species and

restores the pristine absorption of the neutral 6.7. We speculate that such behaviour

could give rise to bistable switching characteristics in the transport properties of the

molecular junctions based on dyads 6.7, although the kinetics of dimer dissociation

appear too slow for practical applications.

- 171 -

350 400 450 500 550 600 650

0.8

0.9

1.0

1.1

Ab

so

rba

nce

Wavelength/ nm

0 V

1.3 V

Figure 6.9: Spectroelectrochemistry of the dyad 6.7 in MeCN (blue line shows

restoration of the original spectrum after reduction of the oxidized solution at –0.2 V).

300 400 500 600 700 800

0.18

0.20

0.22

0.24

0.26

0.28

0.30

0.32

0.34

0.36

0.38

Absorb

ance,

a.u

.

Wavelengh, nm

0V

-1.5V

Figure 6.10: Reductive spectroelectrochemistry of the dyad 6.7 (reduction of the NDI

moiety) in DMF (0.1 M Bu4NPF6).

- 172 -

We have also studied cathodic (reductive) electrochemistry of the dyad 6.7 (Fig.

6.10). Reduction of the compound at negative potential reveals characteristic absorbance

bands at max = 460, 610, 700 and 780 nm that have been earlier reported for the radical

anion of dialkyl-NDI [324]. A broad absorbance band of the donor moiety at

approximately 380 nm remains.

To summarize the characterization studies conducted in solution we will point out

the most important features of the nEDOT-NDI dyads that we need to note for further

interpretation of the rectification measurements:

According to the calculation results a good asymmetric localization of the

HOMO and LUMO orbitals was observed. Mainly due to a torsion angle (72)

between NDI and phenyl bridge.

Fluorescence spectroscopy studies showed possibility of energy transfer between

the acceptor and the donor moieties.

A reasonably low HOMO-LUMO gap (1.4 eV) for bisEDOT-NDI dyads

observed by electrochemical study together with remarkable overall stability of

these dyads suggest robust device performance.

6.5. SAM preparation and characterization

The thiol functionality on either the donor or acceptor part of dyads 6.8 and 6.9

allows covalent attachment of the molecules to the gold metal surface. Monolayers of

Au-S-NDI-bisEDOT and Au-S-bisEDOT-NDI were prepared via self-assembly of

acetyl-protected molecules 6.8b and 6.9b, respectively, onto evaporated gold substrates

from THF solution, in the presence of a catalytic amount of NH4OH. Freshly prepared

SAMs were characterized by grazing angle FT-IR and UV-Vis spectroscopy,

ellipsometry, contact angle and electrochemical measurements.

Ellipsometry indicates that the thickness of the SAMs of 6.8 and 6.9 are 31±2 Å

and 26±2 Å, respectively. This agrees well with the calculated length of the molecules

(34 Å) and suggests an essentially up-right (with a small tilt) orientation on the surface.

Static contact angle measurements (70±2 for dyad 6.8 and 75±2 for dyad 6.9)

indicate that the surface of the SAMs is relatively hydrophobic. Compared to the 110

- 173 -

contact angle for highly ordered and very dense monolayers of alkyl thiols, we can

conclude that terminal alkyl tails of 6.8 and 6.9 in neat SAMs are loosely packed. This is

not unexpected, considering the twisted geometry of the core. A slightly more

hydrophobic surface for dyad 6.8 can be attributed to the branched 2-ethylhexyl tail that

fully covers a polar NDI fragment.

Grazing incidence angle FT-IR spectra of the SAMs studied on planar gold

mirror substrates shows the same features as those of the bulk, proving the preservation

of the molecular structure in the monolayer (Fig. 6.11). The characteristic C=O vibration

of the two imide groups appears at 1670 cm-1

and 1709 cm-1

in the spectra of the SAMs.

The frequencies of CH2 stretching modes at 2930 and 2858 cm-1

are higher than those of

densely packed SAMs of normal long-chain alkanethiols [vas (CH2) = 2917 cm–1

,

vs(CH2) = 2849 cm–1

], indicating the significant disorder of alkyl chains of 6.8 and 6.9 in

SAMs [149, 351].

UV-Vis absorption of SAMs of dyad 6.8 was studied on thin, semi-transparent,

gold-coated (~50 nm) microscope glass slides. The spectrum of a freshly prepared SAM

of 6.8 presents a characteristic peak at 380 nm, fully resembling that of the 6.7 in spin-

coated films (Fig. 4.3). An additional weak shoulder at ~500 nm can be attributed to a

plasmonic band of gold nanoislands in thin, vacuum-coated gold films [352]. Its

interference precludes us from assessing a possible weak charge-transfer band is the

SAMs.

- 174 -

3000 2500 2000 1500 1000

Tra

nsm

isio

n (

norm

aliz

ed)

Wavenumber /cm-1

Figure 6.11: ATR FTIR of the bulk (black) and GA-FTIR of the monolayer on gold

(red) of 6.8 (all spectra are normalized).

Cyclic voltammetry of the SAMs of the dyads 6.8 and 6.9 resembles that of

solution experiments (Fig. 5.12). Two fully reversible reduction waves, characteristic of

the NDI fragment, appear at E0

red 1.08 and 1.55 V vs. Fc/Fc+, respectively. A partially

reversible oxidation peak due to the bisEDOT fragment is observed at Epa

ox = 0.46 V vs.

Fc/Fc+. Multiple scanning of the SAMs through the first reduction wave (formation of

the radical anion) shows moderate stability with a 20–30% drop of the current after 50

cycles in the range of 0 to –0.8 V (see Fig 6.13, top). This can be attributed to desorption

of the molecules from the surface due to repulsion of the positively charged molecules.

The peak current scales linearly (Fig. 6.13, bottom) with the scan rate, indicating a

surface-confined nature of the process.

- 175 -

-2.0 -1.5 -1.0 -0.5 0.0 0.5 1.030

20

10

0

-10

-20

J, A

/cm

2

Potential/ V vs Fc/Fc+

Figure 6.12: Cyclic voltammogram of a SAM of 6.8; electrolyte 0.2M Bu4NPF6 in

CH2Cl2.

The surface coverage () was calculated from the CV peak area. An average

value for SAMs calculated from anodic peaks is 4.210–10

mol/cm2

for dyad 6.8 and

1.610-10

mol/cm2 for dyad 6.9, which corresponds to an average molecular area of 40

Å2 and 110 Å

2 respectively. Comparing these values with the results reported for SAMs

of TCNQ-alkanethiol (3–3.510–10

mol/cm2) [256], TTF-thioctic ester (2.110

–10

mol/cm-10

) [243], fluorene-thioctic ester (3.510–10

mol/cm2) [278] and considering the

twisted geometry of the molecules we can conclude that the dyad 6.8 forms well packed

- 176 -

-0.8 -0.6 -0.4 -0.2 0.0 0.21.0x10

-6

8.0x10-7

6.0x10-7

4.0x10-7

2.0x10-7

0.0

-2.0x10-7

-4.0x10-7

-6.0x10-7

Curr

ent,

A

Potential, V

0 100 200 300 400 500 600

0.0

1.0x10-4

2.0x10-4

3.0x10-4

4.0x10-4

5.0x10-4

6.0x10-4

7.0x10-4

8.0x10-4

Cu

rre

nt,

A

Scan Rate, mV/s

y=1.36*10-6+ 1.25*10

-6x, R

2=0.997

Figure 6.13: Multiple scanning of the SAM of the dyad 6.8 (top). Linear dependence of

the first reduction peak of the SAM of dyad 6.8 on the scan rate (on Au electrode)

(bottom).

- 177 -

monolayers of ―stand-up‖ molecules. The larger molecular area for dyad 6.9 can be

explained by a more bulky 2-ethylhexyl tail on the NDI fragment that leads to less dense

packing, and is consistent with the lower thickness of SAMs of 6.9 observed by

ellipsometry (see above).

6.6. Preliminary rectification study of the bis-EDOT-NDI dyads

To perform rectification measurements the SAMs of the dyads 6.8 and 6.9 were

sandwiched between a gold slide and a mercury drop electrode. First, SAMs of the

corresponding dyads were deposited on the freshly prepared gold. Each gold slide with

the monolayer was used for 5-10 measurements at different spots for statistical analysis.

The mercury electrode was coated with the SAM of dodecanethiol before each

measurement. After both electrodes were covered with SAMs, they were gently brought

into contact inside of the hexadecane bath using a micromanipulator. The molecules of

hexadecane can fill the defects in the monolayer, thus, protecting the metals from direct

contact.

The rectification measurements of the 6.8, deposited on the gold substrate, in the

Au-6.8//C12S-Hg junction (Fig. 6.14a) showed significant asymmetry of the I-V curves

(Fig. 6.15) with higher current at forward bias (the current preferentially flows from gold

electrode to mercury). The direction of the current corresponds to the electron transfer

from the acceptor to the donor, which is in agreement with that predicted by Aviram-

Ratner model [25]. For statistical analysis we used all successful junctions (52) which

are ~50% of all our attempted contacts. The rectification ratio RR varies from junction

to junction in the quite wide range: from 3 to 1400 at 2 V which fit in Gaussian

distribution with log(RR)=2.1±0.6 (Fig. 6.15, inset). The rectification direction is

reproducible among all junctions. The current density at ±1 V is smaller than 1×10–13

A/cm2, thus, they show very low conductivity, whereas, the current density at the ±2 V

is in order of 1×10-5

Amp/cm2. A threshold voltage, determined as tangent to the I-V

curve, lies in the range between 1.6 to 1.8 V depending on the sample. A control

experiment with dodecanethiol SAM on the gold electrode instead of 6.8 (Au-

- 178 -

SC12//C12S-Hg) showed nearly symmetrical I–V curves (log(RR)=0.09±0.37). Figure

6.16 illustrates energy diagram of the junction Au/NDI-bisEDOT/Hg.

Interestingly, the voltage of 1.4V necessary to bring LUMO of the dyad (–3.7

eV) in resonance with the work functions of the gold electrode (–5.1 eV) corresponds

well to the observed threshold voltage. The hysteresis observed for all I-V curves is

probably a result of charge trapping in the monolayer which may also promote

degradation of the molecules.

Figure 6.14: Structure of the junctions studied with mercury drop electrode.

-2 -1 0 1 2

-3.0x10-7

-2.0x10-7

-1.0x10-7

0.0

1.0x10-7

0 1 2 3 40

5

10

15

20

# o

f co

unt

lg(RR)

Curr

ent,

A

Voltage, V

Figure 6.15: The rectification of the SAM of the 6.8 sandwiched between gold and

mercury electrode. Inset: statistical analysis of log(RR)).

- 179 -

Figure 6.16: Energy diagram for the junction Au/NDI-bisEDOT/Hg.

The successive cycling of the voltage through the junctions from 0V to –2 V and

then to +2 V leads to the decrease of the rectification ratio from 114 to approximately 70

after the first cycle and then continues decreasing, resulting in almost symmetric I-V

characteristics (RR=~1.5) after 7 cycles (Fig. 6.17a). The stability of the junctions is

lower if applying higher bias potential and successive cycling of the voltage in the range

±2.5 V results in a sharp drop of the rectification (Fig. 6.17b). Such a drop of the RR

may be a result of reorientation of the molecules in the junction or even complete

desorption from the electrode surface at certain voltage, which creates defects within a

junction that could be filled with solvents. As the number of the defects increased a

portion of the direct tunneling between two electrodes becomes more pronounced which

results in symmetric I-V characteristic of the junction. Sweeping of the bias voltage in

opposite direction (from 0 V to 2 V, and then to –2 V) in most cases results in junctions

breakdown during the first half of the cycle (0 V2 V). The overall asymmetry for the

successful junctions was the same as for the junctions that were scanned in direction

- 180 -

0 V –2 V +2 V 0 V. To test if the observed rectification behaviour follows the

Aviram-Ratner mechanism, we have prepared junctions where donor and acceptor

moieties of the dyad are positioned in inverse orientation with respect to the electrodes

(i.e., Au//bisEDOT-NDI//Hg). This was achieved through design of the dyad 6.9 which

has self-assembly functional group on the donor (bisEDOT) moiety. The SAMs of the

dyad 6.9 on gold substrate contacted with the mercury drop electrode, is expected to

form a junction, where the donor is close to the gold electrode and the acceptor is close

to the mercury electrode. The electrical conductivity measurements of corresponding

mercury-drop junctions, showed high asymmetry of the current (the average

log(RR)=1.5±0.7), similar to that of the dyad 6.8. Against our expectations, the direction

of rectification (from Au to Hg) has not changed and now corresponded to the electron

tunneling from donor to the acceptor moiety.

0 1 2 3 4 5 6 7 8

0

40

80

120

RR

Number of cycles

A

1 2 3 4 5 6 7

0

400

800

1200

1600

RR

Number of cycles

B

Figure 6.17: The rectification of an SAM monolayer of 6.8 junctions decreases with

successive cycling: A) (0 V –2 V +2 V 0 V); B) (0 V –2.5 V +2.5 V 0

V).

- 181 -

-2 -1 0 1 2

-200.0n

-100.0n

0.0

Curr

ent, n

A

Voltage, V

Figure 6.18: I-V characteristics of the junctions Au/6.9//C12-S-Hg.

Despite the fact that the contribution of the donor-acceptor structure of the

bisEDOT-NDI dyads in the rectification behaviour of the SAMs was not confirmed, the

importance of molecular structure is obvious from the control experiments with

junctions containing simple alkanethiols on both electrodes. One of the possible

explanations for the independence of the rectification direction on the orientation of the

dyad (i.e., 6.8 vs. 6.9) is in asymmetric position of their electroactive core between two

electrodes (cf. [59]). The structure of the molecule and the junction defines the spacing

between the redox-active center (bisEDOT or NDI cores) and gold electrode equal to six

methylene groups, while the spacing between mercury electrode is 18 methylene groups

(12 for dodecanethiol SAM on mercury electrode and 6 for the alkyl tail on the dyad) for

both types of the studied junctions.

6.7. Potential for improvement of the acceptor properties

As was mentioned in the introduction for this chapter, the advantage of the

nEDOT-NDI model is in the potential synthetic tuning of its molecular orbitals energies.

This includes the optimization of the HLG, the positions of HOMO and LUMO orbitals

with respect to the Fermi energies of the electrodes, etc. Above we already presented

- 182 -

how the donor properties can be tuned by increasing of the number of the EDOT unites

from one to two, which leads to significant rising of the HOMO and, thus, decrease of

the HLG. Further increase of the number of EDOT units does not provide significant

improvement of the donor ability of the dyad. As shown on the energy diagram (Fig.

6.16), it is more important lower the LUMO brining it closer to the Fermi level of the

electrodes.

In the Appendix we present computational and synthetic exploration of potential

approaches for further tailoring the electron acceptor characteristics on the dyads. Our

initial idea for the design of the Donor-Acceptor dyad was to use the PDI as an acceptor

moiety (See Appendix). However, because very poor solubility of the starting PTCDA,

which drastically hinders asymmetric functionalization, we turned to the NDI synthon.

Despite the acceptor properties of the NDI are lower than of PDI (LUMONDI= –3.41 eV,

LUMOPDI= –3.46 eV) they can be dramatically improved by introducing the electron-

withdrawing substituents in the aromatic core [334, 353]. Many different substituents

were already introduced in the NDI core, such as halogens, nitrile, alkylsulfones and

other [354]. Synthetic efforts to lower the LUMO of the NDI by introducing electron-

withdrawing core-substituents were reported in the literature [332, 334]. The

introduction of two cyano groups in the NDI core decreased LUMO energy levels [332,

354-355]. Good improvement of the acceptor ability of the NDI could be obtained from

introducing the sulfide groups in the core and subsequent oxidation of them into sulfone

groups [353]. Our DFT (B3LYP, 6-31G(d)) calculation for the bisEDOT-NDI

tetrasulfone based dyads (Appendix, Fig. A3) shows a dramatic decrease of the HLG

(0.84eV) in the dyad 6.11 comparing to the dyad 6.7 (1.45 eV) which is due to the

lowering of the acceptor LUMO (from –3.26 eV for 6.7 to –4.08 eV for 6.11) energy.

The synthetic procedure [353, 356-357], however, requires modification of the acceptor

synthon before the coupling with donor moiety step. The stronger acceptor properties of

the NDI core may increase the probability of the formation of the CTC with donor

moiety during the coupling reaction or in the monolayer.

- 183 -

Conclusions

New donor-acceptor dyads based on highly stable bisEDOT donor and NDI

acceptor moieties have been synthesized. Thiol functionality on either the donor or the

acceptor parts enables anchoring of the dyads to a gold electrode in two different

orientations. Such design allows confirming the molecular origin of the rectification

behaviour. The rigid geometry of the molecular core and a large twist between the

acceptor and the phenylene bridge allows for efficient separation of the HOMO and

LUMO orbitals, despite their close proximity. Accordingly, donor-acceptor interaction

in these dyads is taking place only in the excited state of these dyads, not in their ground

state. The HOMO-LUMO gap of ~1.4 eV provides for sufficient chemical and electronic

stability [72], while a low dipole moment of ~2.7 D is expected to lead to orientational

stability of the molecules in monolayer junctions. A chemically reversible dimerization

of the bisEDOT moieties, established through spectroelectrochemistry and EPR

spectroscopy, offers potential opportunities for the design of molecular switches based

on NDI-bisEDOT dyads. The rectification behaviour of the SAMs of the dyads 6.8 and

6.9 was studied in mercury-drop junctions. The Metal-SAM-Metal junctions for both

dyads are characterized by the significant current asymmetry (for 6.8 log(RR)=2.1±0.6

and for 6.9 log(RR)=1.5±0.7). No difference in the rectification direction for the

junctions made with dyad 6.8 and 6.9 was observed. The current preferentially flows

from the gold electrode to the mercury electrode, despite the inverse orientation of the

donor and acceptor moieties within the junction. Thus, we cannot conclude that the

observed rectification is due to the electronic asymmetry in the donor-acceptor dyads;

however, it is clear that this behaviour is based on their presence in the junctions, which

was confirmed in the control experiments in Au-SC12/C12S-Hg junctions. We propose

that the observed rectification behaviour of the junctions is due to the asymmetrical

positioning of the dyad within the junction. Further studies of charge transport behaviour

of the dyads 6.8 and 6.9, using conductive AFM and other single molecule junction

techniques are necessary.

- 184 -

Experimental part

Cyclic voltammetry measurements were done using a CHI-670 potentiostat under

nitrogen in a CH2Cl2 solution of electrolyte (0.1 M Bu4NPF6) with a Ag/AgCl reference

electrode and platinum disk (d=1.6 mm) as a working electrode for solution experiments

and a gold disk electrode (BAS, d=1.6 mm) for the SAM experiments. Fc/Fc+ (0.50 V

vs. Ag/AgCl in these conditions) were used as an internal reference.

Calculations of geometry and electronic structure of the dyads were done using density

functional theory (DFT) with hybrid B3LYP functional and 6–31G(d) basis set, as

implemented in Gaussian W03 [228]. The alkyl substituents on both donor and acceptor

moieties were modeled using methyl groups.

Absorbance/Emission spectroscopy. Absorption spectra were recorded with a Jasco V-

670 spectrophotometer in CH2Cl2 and MeCN solutions. Fluorescence was recorded on

Cary Eclipse fluorimeter in MeCN and toluene solutions. For measurements of the solid

state samples, two kinds of samples were prepared: 1) spin-coated thin film of the 2 on

the clean glass slide (similar clean glass slide was used as a reference); 2) self-assembled

monolayer of the 3 and 4 on very thin (≤ 50 nm) gold film, evaporated on the glass slide.

A part of this slide, without the SAM was used as a reference.

FTIR Spectroscopy. See Chapter III for details

Ellipsometry. See Chapter III for details.

Contact angle measurements. The static contact angles of deionized water (>18 MΩ

cm) were measured on a homemade contact angle goniometer and averaged over 3–5

spots. The plasma-cleaned Au surface produced a static contact angle of 0°.

Spectroelectrochemistry experiments were performed in thin layer

spectroelectrochemial cell CHI 760C from CH Instruments equipped with platinum grid

as a working, platinum wire as a counter and Ag/AgCl as a reference electrode. UV-Vis

spectra were recorded with a Jasco V-670 spectrophotometer and a BASi Epsilon

potentiostat and the static potential mode was used for oxidation/reduction of the

molecules. Potentials were applied in 50-100 mV steps and equilibrated by allowing the

current to drop until a negligible current change was achieved (less than 1% of initial

current per minute).

SAM preparation. Slides of the gold, evaporated on the silicon wafer or glass substrate

were immersed in the 10–3

M solution of the bisEDOT-NDI (a few drops of NH4OH

were added to facilitate the cleavage of the acetyl protecting group from the thiol) dyads

for 12–48 hours. After that period gold slides were washed with THF with sonication for

a few seconds and dried in vacuum. Mixed monolayers on the gold surface were

prepared by subsequent immersing of the substrate in the dilute solution of

dodecanethiol (10-3

M) in ethanol for 8-12 hours, and then in the solution of 6.8a in THF

(with 1-3 drops of NH4OH).

Preparation of mercury-drop junctions. Rectification measurements in mercury-drop

junctions were done as described previously for SAMs in the Chapter V.

N-2-Ethylhexyl-1,4,5,8-naphthalenetetracarboxyimide anhydride (6.2).

Monoimide 6.2 was prepared following a literature procedure [358], from

commercially available 1,4,5,8-naphthalenetetracarboxylic dianhydride 6.1 (10.0 g,

37.3 mmol) and 2-ethyl-1-hexylamine (5.35 g, 41.3 mmol). The product was purified

by column chromatography (silica; hexane/EtOAc gradient) to afford the desired

- 185 -

product 6.2 as yellowish solid (first fraction) (6.1g, 43%) along with symmetric

diimide 6.2a as a side product (second fraction) (5.9g, 32%). 6.2: m.p. 166–169 C; 1H NMR (400 MHz, CDCl3) 8.81 (4H, s), 4.15 (2H, m), 1.92 (1H, t, J 6.0 Hz), 1.33

(8H, m), 0.95 (3H, t, J 7.4 Hz), 0.89 (3H, t, J 7.2 Hz); 13C NMR (125.0 MHz,

CDCl3) 162.3, 158.8, 133.2, 131.3, 128.9, 127.9, 126.9, 122.8, 44.8, 37.9, 30.6, 28.6,

24.0, 23.0, 14.1, 10.6. 6.2a: m.p. 203-204 C, 1H NMR (500 MHz, CDCl3) 8.75 (4H,

s), 4.15 (4H, m), 1.94 (2H, t, J 6.0 Hz), 1.43-1.22 (16H, m), 0.95 (6H, t, J 7.4 Hz),

0.89 (6H, t, J 7.2 Hz).

N-(6-(tert-butylsulfanyl)hexyl-1,4,5,8-naphthalenetetracarboxyimide anhydride

(6.3). To a solution of 6.1 (1.0 g, 3.7mmol) in dry DMF (150 ml) was slowly added

(during 3 h) a solution of 4.21 (0.65 g, 3.5 mmol) in dry DMF (40 ml) under nitrogen

atmosphere and reaction mixture was refluxed overnight. After cooling to room

temperature, reaction mixture was placed in the fridge for 2-3 h (–10°C). Precipitate

(corresponding diimide side product) was filtered out and then the solvent was

removed under reduced pressure. Crude product was dissolved in acetone and

resulted solution was kept in the fridge overnight to precipitate more of diimide. The

filtrate was concentrated and resulted solid was purified by column chromatography

on silica (CH2Cl2:EtOAc eluent, gradient) resulting in desired monoimide 6.3 as a

yellow solid (0.88 g, 59%). M.p. 218–220C; 1H NMR (400 MHz, CDCl3) 8.81 (4H,

s), 4.21 (2H, t, J 7.6), 2.53 (2H, t, J 7.2), 1.77 (2H, t, br), 1.7–1.4 (6H, m), 1.31 (9H,

s); 13

C NMR (75.0 MHz, CDCl3) 162.1, 158.8, 133.1, 131.2, 128.8, 127.9, 122.7,

41.7, 41.1, 30.9, 29.7, 29.1, 28.2, 27.9, 26.9.

N-(2-Ethylhexyl)-N'-p-bromophenyl 1,4,5,8-naphthalenetetracarboxydiimide (6.4).

Compound 6.2 (6.1 g, 16.1 mmol) and p-bromoaniline (2.80 g, 16.1 mmol) were

dissolved in dry DMF (150 ml) and the reaction mixture was stirred at reflux overnight

under nitrogen atmosphere. After cooling to room temperature, the solvent was removed

under reduced pressure. After column chromatography on silica (CH2Cl2:EtOAc eluent,

gradient), afforded the desired diimide 6.4 (6.5 g, 76%). M.p.: 213–214°C; 1H NMR

(400 MHz, CDCl3) 8.82 (4H, m), 7.72 (2H, d, J 8.8 Hz), 7.21 (2H, d, J 8.8 Hz), 4.16

(2H, m), 1.96 (1H, m), 1.36 (8H, m), 0.95 (3H, t, J 7.4 Hz), 0.89 (3H, t, J 7.2 Hz); 13

C

NMR (75.0 MHz, CDCl3) 163.1, 162.8, 133.5, 132.8, 131.5, 131.1, 130.2, 127.1, 126.4,

123.3, 44.7, 37.9, 30.7, 28.6, 24.0, 23.0, 14.1, 10.6; HR-MS (ESI): calculated for

C28H25BrN2O4 533.1070, found 533.1058.

N-(6-(tert-butylsulfanyl)hexyl-N’-p-bromophenyl-1,4,5,8-

naphthalenetetracarboxydiimide (6.5). NDI 6.3 (0.88 g, 2.0 mmol) and p-

bromoaniline (1.0 g, 5.8 mmol) were dissolved in dry DMF (100 mL) and the

reaction mixture was stirred under nitrogen atmosphere at reflux overnight. After

cooling to room temperature, the solvent was removed under reduced pressure and

the crude product was dried under vacuum. After column chromatography on silica

(CH2Cl2:EtOAc eluent, gradient), desired diimide 6.5 was obtained as dark yellow

solid (0.75 g, 63%). M.p.: 237–239°C; 1H NMR (400 MHz, CDCl3) 8.81 (4H, s),

7.72 (2H, d, J 8.8 Hz), 7.23 (2H, d, J 8.8 Hz), 4.21 (2H, m,), 2.53 (2H, t, J 7.2), 1.77

(2H, m), 1.62-1.39 (6H, m), 1.32 (9H, s); 13

C NMR (75.0 MHz, CDCl3) 163.1, 162.8,

131.5, 131.0, 130.2, 127.1, 126.4, 123.3, 41.5, 40.3, 31.0, 29.2, 28.9, 28.2, 27.3,

26.8; HR-MS (ESI): calculated for C30H29BrN2O4S 592.1026, found 592.1033.

- 186 -

NDI-EDOT dyad (6.6). To a solution of NDI 6.4 (0.50 g, 0.94 mmol) and 2-

tributylstannyl-EDOT [344] 6.15 (0.50 g, 1.2 mmol) in dry toluene under nitrogen

atmosphere was added a catalyst Pd(PPh3)4 (0.054 g, 0.05 mmol), and the reaction

mixture was stirred at reflux for 12 h. After all starting compound 6.4 has reacted, the

mixture was cooled down and the solvent was evaporated under reduced pressure. The

residue was dissolved in CH2Cl2 and washed with water and brine, and the organic phase

was dried over MgSO4. Purification by column chromatography on silica using CH2Cl2

as an eluent afforded desired product 6.6 as an orange solid (0.47 g, 85%). M.p. 220–222

С; 1H NMR (400 MHz, CDCl3) 8.81 (4H, m), 7.91 (2H, d, J 8.8 Hz), 7.31 (2H, d, J 8.8

Hz), 6.36 (1H, s), 4.26 (4H, m), 4.17 (2H, m), 1.92 (1H, m), 1.36 (8H, m), 0.95 (3H, t, J

7.4 Hz), 0.89 (3H, t, J 7.2 Hz); 13

C NMR (75.0 MHz, CDCl3) 163.2, 163.0, 142.2, 138.8,

134.2, 132.4, 131.4, 131.1, 128.6, 127.06, 126.95, 126.89, 126.86, 126.68, 116.5, 98.4,

64.8, 64.4, 44.7, 37.9, 30.7, 28.6, 24.0, 23.0, 14.1, 10.6; HR-MS (ESI): calculated for

C34H31N2O6S (M+1) 595.1897, found 595.1880.

NDI–bis-EDOT dyad (6.7). To a solution of NDI 6.4 (0.82 g, 1.52 mmol) and 5-

tributylstannyl-5'-hexyl-bis-EDOT (1.00 g, 1.52 mmol) in dry toluene (20 ml) under

nitrogen atmosphere was added a catalyst Pd(PPh3)4 (0.088 g, 0.076 mmol) and the

reaction mixture was stirred at 110°C for 24 h. After all starting compounds has reacted

as monitored by TLC (silica, CH2Cl2), the reaction was cooled down and the solvent was

evaporated under reduced pressure. Residue was redissolved in CH2Cl2, washed with

water and brine, and the organic phase was dried over MgSO4. Column chromatography

on silica eluting with CH2Cl2 resulted in product 6.7 as a dark green solid (0.25 g, 20%).

M.p. 305–306 C; 1H NMR (400 MHz, CDCl3) 8.75 (4H, m), 7.87 (2H, d, J 8.8 Hz),

7.28 (2H, d, J 8.8 Hz), 4.36-4.22 (8H, m), 4.15 (2H, m), 2.64 (2H, t, J 7.6 Hz), 1.96 (1H,

m), 1.61 (2H, m), 1.35 (14H, m), 0.95 (3H, t, J 7.4 Hz), 0.89 (6H, m, m); 13

C NMR (75.0

MHz, CDCl3) 163.2, 163.0, 138.6, 137.2, 137.1, 136.5, 134.2, 131.9, 131.3, 130.9,

128.5, 127.0, 126.80, 126.77, 126.6, 126.5, 117.4, 112.9, 109.5, 105.4, 65.2, 64.7, 64.5,

44.6, 37.9, 31.6, 30.7, 30.5, 28.9, 28.6, 25.8, 24.0, 23.1, 22.6, 14.1, 10.6; HR-MS (ESI):

calculated for C46H46N2O8S2 818.2696, found 818.2680.

NDI–bis-EDOT dyad (6.8a). To a solution of NDI 6.5 (0.070 g, 0.134 mmol) and 5-

tributylstannyl-5'-(6-(tert-butylsulfanyl)hexyl)-bis-EDOT (0.1 g, 0.134 mmol) in dry

toluene under nitrogen atmosphere was added catalyst Pd(PPh3)4 (0.010 g, 0.007 mmol),

and the reaction mixture was stirred at 85°C for 24 h. After all starting material has

reacted (followed by TLC on silica), the reaction was cooled down and the solvent was

evaporated under reduced pressure. The residue was dissolved in CH2Cl2 and washed

with water and brine, and then organic phase was dried with MgSO4. Column

chromatography on silica with CH2Cl2 as an eluent resulted in desired product 6.8a as a

dark green solid (0.035 g, 28%). M.p. 272–274 C; 1H NMR (400 MHz, CDCl3) 8.81

(4H, two d), 7.92 (2H, d, J 8.4 Hz), 7.29 (2H, d, J 8.4 Hz), 4.49-4.15 (8H, m), 4.16 (2H,

m), 2.66 (2H, t, J 7.5 Hz), 2.52 (2H, t, J 7.2 Hz), 1.96 (1H, m), 1.70-1.50 (4H, m), 1.50-

1.35 (14H, m), 1.31 (9H, s), 0.95 (3H, t, J 7.4 Hz), 0.89 (3H, t, J 7.2 Hz); 13

C NMR

(125.0 MHz, CDCl3) 163.2, 163.0, 138.6, 137.3, 136.6, 134.3, 131.9, 131.4, 131.1,

128.5, 127.0, 126.9, 126, 7, 126.6, 117.0, 113.1, 109.5, 105.6, 65.2, 64.8, 64.5, 44.7,

37.9, 31.0 30.70, 30.66, 30.2, 29.4, 29.1, 28.64, 28.56, 28.50, 25.7, 24.0, 23.0, 14.1,

10.6; HR-MS (ESI): calculated for C50H54N2O8S3 906.3042, found: 906.3037.

- 187 -

NDI–bis-EDOT dyad (6.8b). To a solution of tert-Bu protected dyad 6.8a (0.035 g,

0.039 mmol) in dry CH2Cl2 (15 ml) at –78C was added acetyl bromide (0.2 ml, excess)

followed by 0.1 M solution of BBr3 in CH2Cl2 (0.8 ml, 0.08 mmol) under nitrogen

atmosphere. The reaction mixture was stirred at room temperature for 4 hours and

poured into ice. The resulted solution was extracted with CH2Cl2. The organic phase was

separated and washed with water, brine, and dried over MgSO4. Crude product was

recrystallized from CH2Cl2–hexanes mixture to afford green solid of 6.8b (0.018 g, 52

%). M.p. 290–293 C. 1H NMR (400 MHz, CDCl3) 8.80 (4H, two d), 7.93 (2H, d, J 8.4

Hz), 7.29 (2H, d, J 8.4 Hz), 4.5–4.2 (8H, m), 4.18 (2H, m), 2.86 (2H, t, J 7.2), 2.65 (2H,

t, J 7.2), 2.32 (3H, s), 1.97 (1H, m), 1.71–1.50 (4H, m), 1.50–1.20 (14H, m), 0.96 (3H, t,

J 7.4 Hz), 0.90 (3H, t, J 7.2 Hz); 13

C NMR (125.0 MHz, CDCl3) 163.2, 163.0, 138.6,

137.3, 136.6, 134.3, 131.9, 131.4, 131.1, 128.5, 127.0, 126.9, 126, 7, 126.6, 117.0,

113.1, 109.5, 105.6, 65.2, 64.8, 64.5, 44.7, 37.9, 30.70, 30.66, 30.2, 29.4, 29.1, 28.64,

28.56, 28.50, 25.7, 24.0, 23.0, 14.1, 10.6; HR-MS (ESI): calculated for C48H48O9N2S3

892.2522, found 892.2526.

NDI-bis-EDOT dyad (6.9a). To a solution of NDI 6.5 (0.30 g, 0.48 mmol) and 6.13

[344] (0.34 g, 0.52 mmol) in dry toluene under nitrogen atmosphere was added catalyst

Pd(PPh3)4 (0.030 g, 0.025 mmol), and the reaction mixture was stirred at refluxing for 6

h. After all starting compound has reacted (followed by TLC, silica, eluent CH2Cl2), the

reaction was cooled down and the solvent was evaporated under reduced pressure. The

residue was purified by column chromatography on silica (eluent: CH2Cl2, followed by

20:1 CH2Cl2-EtOAc) resulting in the desired compound 6.9a as a dark green solid (0.255

g, 54% yield). M.p. 330–333 С; 1H NMR (400 MHz, CDCl3) 8.79 (4H, d), 7.92 (2H, d,

J 8.4Hz), 7.29 (2H, d, J 8.4 Hz), 4.5-4.1 (10H, m), 2.65 (2H, t, J 7.2), 2.53 (2H, t, J 7.2),

1.77 (2H, br), 1.63 (m, 4H), 1.49 (m, ), 1.36 (t, br), 1.32 (s, 9H), 0.89 (t, 3H, J 7.0 Hz); 13

C NMR (75.0 MHz, CDCl3) 163.1, 162.8, 138.6, 137.3, 137.1, 136.6, 134.3, 131.9,

131.4, 131.0, 128.5, 127.0, 126.9, 126.8, 126.7, 126.6, 117.4, 113.0, 109.5, 105.5, 65.2,

64.8, 64.5, 41.8, 40.9, 31.6, 31.0, 30.5, 29.7, 29.0, 28.9, 28.2, 28.0, 26.8, 25.8, 22.6,

14.1. HR-MS (ESI): calculated for C48H50O8N2S3 878.2729, found 878.2737.

NDI-bis-EDOT dyad (6.9b). To a solution of tert-Bu protected dyad 6.9a (0.093 g, 0.1

mmol) in dry CH2Cl2 (30 mL) at –78C and acetyl bromide (0.1 ml) was added 0.1 M

solution of BBr3 in CH2Cl2 (1 mL, 0.1 mmol) dropwise and the reaction mixture was

slowly warmed to room temperature. After stirring for 4 h the reaction was quenched

with water and the crude product was extracted with CH2Cl2. The organic phase was

washed with water, brine and dried with MgSO4. Purification by column

chromatography on silica (CH2Cl2:EtOAc eluent, gradient) gave desired product as a

green solid (0.047 g, 51%). M.p. 291–294 C. 1H NMR (500 MHz, CDCl3) 8.80 (m,

4H), 7.92 (d, 2H, J 9.0 Hz), 7.29 (d, 2H, J 9.0 Hz), 4.42-4.20 (m, 10H), 2.86 (t, 2H, J

7.5), 2.33 (s, 3H), 1.76 (m, 2H), 1.57-1.45 (m, 2H), 1.45-1.20 (m, 14H) 0.92 (t, 3H, J 7.5

Hz); 13

C NMR (125.0 MHz, CDCl3) 196.0, 163.0, 162.8, 138.6, 137.3, 137.2, 137.1,

134.1, 132.1, 131.3, 130.9, 128.5, 127.0, 126.81, 126.78, 126.68, 126.58, 126.52, 118.9,

115.5, 113.7, 109.4, 106.4, 65.2, 64.7, 64.6, 40.9, 32.9, 31.5, 30.6, 29.4, 29.1, 29.0, 28.9,

28.7, 28.7, 28.0, 26.9, 22.6, 22.2, 13.9. HRMS (ESI) calculated for C46H44O9N2S3

864.2209 found 864.2219.

tert-Butylsulfanylhexyl-6-chloride (6.16). To a solution of 1-bromo-2-chlorohexane

(8.4 g, 42.1 mmol) and tert-butyl mercaptane (3.8 g, 41.2 mmol) in DMF well-grounded

- 188 -

potassium carbonate (6.0 g, 55.6 mmol) was added and the reaction mixture was stirred

at room temperature overnight. After all starting 6-bromohexyl chloride was consumed,

reaction mixture was diluted with water and resulted solution was extracted with EtOAc,

washed with water and brine. Organic phase was dried over MgSO4. Evaporation of the

solvent gave desired product as colorless oil (8.37g. 95%), which had sufficient purity

(GC-MS) to use it for further transformation. 1H NMR(400 MHz, CDCl3) 3.52 (t, 2H, J

6.8Hz), 2.52 (t, 2H, J 7.2Hz), 1.86 (2H, m), 1.57 (m, 2H), 1.50-1.38 (m, 4H), 1.37 (s,

9H). 13

C NMR (75.0 MHz, CDCl3) 44.9, 41.7, 32.4, 30.9, 29.6, 28.4, 28.0, 26.5. HRMS

(APCI) calculated for C10H22ClS 209.1125 found 209.1128.

1-tert-Butyl-sulfanyl-6-iodohexane (6.17). To a solution of tert-butylsulfanylhexyl-6-

chloride (14) (8.37g, 40.2 mmol) from previous step in acetone was added NaI (7.0 g, 47

mmol) and the reaction mixture was stirred at reflux overnight. Completion of the

reaction was followed by 1H NMR and GC-MS analysis. Then reaction mixture was

diluted with water and resulted solution was extracted with EtOAc. Organic phase was

washed with water, brine, and then was dried over MgSO4. Solvent was evaporated

resulting in desired product as a yellow oil (11.0 g, 91%); 1H NMR(400 MHz, CDCl3)

3.18 (t, 2H, 6.9 Hz), 2.52 (t, 2H, 7.2 Hz), 1.94-1.80 (m, 2H), 1.6-1.57 (m, 2H), 1.45-1.38

(m, 4H), 1.37 (s, 9H); 13

C NMR(125.0 MHz, CDCl3) 41.9, 33.3, 30.9, 30.1, 29.5, 28.2,

28.1, 7.0. HRMS (APCI) calculated for C10H22IS 301.0481 found 301.0477.

- 189 -

Conclusions

During the course of presented work we successfully synthesized four new types

of donor-acceptor dyads, studied their properties using spectroscopic and

electrochemical methods, prepared their monolayers on electrode surfaces and, using

mercury drop junctions technique, interrogated their rectification behavior.

As was already mentioned in the thesis, coupling together strong donor and

strong acceptor moieties is a challenging task. Nevertheless, we have successfully

synthesized TTF--polynitrofluorene dyads. The advantage of the polynitrofluorene as

an acceptor synthon is in the possibility to significantly increase its acceptor ability after

coupling with donor moiety by converting fluorenone moiety into dicyanomethylene

derivative. The synthesized donor-acceptor dyad 2.4 is characterized by very low

HOMO-LUMO gap (~0.3eV) which makes it the closest analog of the original Aviram-

Ratner molecular rectifier model. Another advantage of the polynitrofluorene building

block for donor-acceptor dyads was shown in Chapter IV in synthesis of D--A

molecules functionalized with ―anchor‖ group for self-assembly on the gold electrode

surface. Sequential substitution of two nitro groups of TNF in the positions 2 and 7

allows for creation of complex asymmetric structure with desired functionalities (e.g.

other electroactive moiety and ―anchor group‖). Importantly, replacing of two nitro

groups with sulfonyl substituents leads to only 0.13eV reduction of electron affinity

allowing to maintain strong electro-acceptor character. Remarkably small (0.34 eV)

HLG was attained upon conversion of the dyad 4.24 into its dicyanomethylene

derivative 4.25. Our work shows that ―multi-functionalization‖ in such strongly

amphoteric molecules is a challenging task. We met with several difficulties, related to

synthesis of fluorene-based acceptor with self-assembly capable groups, such as

incompatibility of the thiol-terminated molecules with nitrofluorene moiety,

polymerization of dithiolidene cycle (4.5). Low solubility of the acceptor synthons with

many other investigated ―anchor’ groups limited their applications. During the work,

presented in Chapter III and IV, a series of new fluorene derivatives with different

―anchor‖ functionalities was synthesized. We demonstrated unique electrochemical

characteristics of the strong electron acceptor (3.5) in the SAM, such as its multiredox

- 190 -

behavior and stability of corresponding highly charged species (up to radical trianion) on

the surface. As a perspective direction in this work it is worth considering different (not

thiol based) ―anchor‖ groups (e.g. pyridine-based tripodal anchors).

Relative flexibility of the -bridge in TTF-fluorene dyads is a serious issue that

causes an unwanted conformational freedom of the molecule and, for long bridges such

as in dyads 4.25, can even lead to formation of intramolecular CTC. Chapter VI presents

synthesis and study of series of D-A dyads that eliminates this problem. Using highly

stable nEDOT and NDI moieties as donor and acceptor, respectively, and phenylene

group as a linker, we were able to make dyad molecules that are robust,

conformationally rigid and possess linear geometry. Such linker efficiently separates the

molecular orbitals of donor and acceptor due to the large dihedral angle between NDI

core and phenyl ring (72). This type of the dyad also provides ample opportunities for

tuning the electronic properties of the molecule. Together with a remarkable stability,

this makes nEDOT-NDI a very interesting system for molecular electronics application.

In Chapters V we described a series of D-π-A dyads, nEDOT-3CNQ. -

Conjugation, enabled by a relatively small dihedral angle (~30) between the donor and

acceptor moieties, does not provide complete separation of HOMO and LUMO orbitals

on the corresponding parts of the molecule. Such design is thus similar to the push-pull

molecular rectifier C16H33Q-3CNQ extensively studied in literature. However, replacing

the amine-based n-donor with EDOT -donor decreases zwitterionic character of the

dyad, as is shown by lower dipole moment of nEDOT-3CNQ vs. C16H33Q-3CNQ and

rather moderate solvatochromism of the former. In spite of substantial orbitals

delocalization, shown by DFT calculations and manifested in strong ICT band in

electronic spectra of nEDOT-3CNQ, clear rectification behavior was observed.

Comparing our preliminary results of current transport measurements for all

studied dyads we can conclude that careful design of all aspects of the molecular

junctions is important (HLG/MO/work function alignment, bridge flexibility, linker to

the electrode, symmetric position of redox centers vs. the electrode). As we showed,

stability of the junctions, direction of the current rectification and also the value of the

RR are directly related to the nature of the junctions. The current rectification can be

observed even for molecules lacking obvious donor-acceptor structure, due to the

- 191 -

asymmetric position of the electroactive center vs. the electrodes (molecules 4.5, 6.8 and

6.9). Weak binding of the LB film to the surface can allow reorientation of the molecule

within the junction which leads to decrease of its rectification behavior. Different

orientations of the dyad in the junction, such as D--A vs. A--D allows to eliminate the

effect of the contacts on the rectification. This strategy was successfully applied for the

TTF-fluorene dyad 2.4 deposited by LB technique (X and Z deposition). Unexpectedly,

a similar approach for SAM-based dyads 6.8 and 6.9 shows no change in the

rectification direction, which is likely attributable to either asymmetric position of the

molecule (close to Au electrode) or the dominating effect of the contact (Au-S bonding

on one side vs. Van der Waals contact on the other). Further progress in the study of the

rectification of the discussed donor-acceptor dyads, especially 6.8 and 6.9, can be done

by designing the junction with symmetric position of the redox centers between

electrodes, using different anchor groups (e.g. isocyano) or by using different methods

(e.g. conductive probe AFM).

- 192 -

References

1. http://www.zyvex.com/nanotech/feynman.html.

2. Kwok, K. S.; Ellenbogen, J. C., Moletronics: future electronics. Mater Today

2002, 5 (2), 28-37.

3. Tour, J. M.; Reinerth, W. A.; Jones, L.; Burgin, T. P.; Zhou, C.-W.; Muller, C. J.;

Deshpande, M. R.; Reed, M. A., Recent Advances in Molecular Scale Electronics. Ann

Ny Acad Sci 1998, 852 (1), 197-204.

4. Ellenbogen, J. C. In Advances toward molecular-scale electronic digital logic

circuits: a review and prospectus, VLSI, 1999. Proceedings. Ninth Great Lakes

Symposium on, 4-6 Mar 1999; 1999; pp 392-393.

5. Moore, G. E., Electronics 1965, 38 (8), 114.

6. Williams, E. L.; Haavisto, K.; Li, J.; Jabbour, G. E., Excimer-based white

phosphorescent organic light emitting diodes with nearly 100 % internal quantum

efficiency. Adv Mater 2007, 19 (2), 197-202.

7. Perepichka, D. F.; Perepichka, I. F.; Meng, H.; Wudl, F., Light-Emitting

Polymers. In Organic Light-Emitting Diodes, Li, Z. R., Ed. CRC Press: Boca Raton, FL,

2006.

8. Dimitrakopoulos, C. D.; Malenfant, P. R. L., Organic thin film transistors for

large area electronics. Adv Mater 2002, 14 (2), 99-117.

9. Horowitz, G., Organic field-effect transistors. Adv Mater 1998, 10 (5), 365-377.

10. Katz, H. E.; Bao, Z., The physical chemistry of organic field-effect transistors. J

Phys Chem B 2000, 104 (4), 671-678.

11. Reese, C.; Bao, Z. N., Organic single-crystal field-effect transistors. Mater Today

2007, 10 (3), 20-27.

12. Facchetti, A., Semiconductors for organic transistors. Mater Today 2007, 10 (3),

28-37.

13. Lloyd, M. T.; Anthony, J. E.; Malliaras, G. G., Photovoltaics from soluble small

molecules. Mater Today 2007, 10 (11), 34-41.

- 193 -

14. Mayer, A. C.; Scully, S. R.; Hardin, B. E.; Rowell, M. W.; McGehee, M. D.,

Polymer-based solar cells. Mater Today 2007, 10 (11), 28-33.

15. Bird, J. O., Electrical circuit theory and technology. Rev. 2nd ed.; Newnes:

Oxford ; New York, 2003; p xviii, 984 p.

16. Wang, Q.; Ward, S.; Duda, A.; Hu, J.; Stradins, P.; Crandall, R. S.; Branz, H. M.;

Perlov, C.; Jackson, W.; Mei, P.; Taussig, C., High-current-density thin-film silicon

diodes grown at low temperature. Appl Phys Lett 2004, 85 (11), 2122-2124.

17. Seki, K.; Yamamoto, H.; Sasano, A.; Tsukada, T., Hydrogenated Amorphous-

Silicon Pin Diodes with High Rectification Ratio. J Non-Cryst Solids 1983, 59-6 (Dec),

1179-1182.

18. Taube, H., Electron-Transfer between Metal-Complexes - a Retrospective View

(Nobel Lecture). Angew Chem Int Edit 1984, 23 (5), 329-339.

19. Marcus, R. A., On the Theory of Oxidation-Reduction Reactions Involving

Electron Transfer .1. J Chem Phys 1956, 24 (5), 966-978.

20. Marcus, R. A., Electron-Transfer Reactions in Chemistry - Theory and

Experiment (Nobel Lecture). Angewandte Chemie-International Edition in English

1993, 32 (8), 1111-1121.

21. Jortner, J.; Ratner, M. A.; International Union of Pure and Applied Chemistry.,

Molecular electronics. Blackwell Science: Osney Mead, Oxford England ; Malden, MA,

USA, 1997; p viii, 485 p.

22. Closs, G. L.; Calcaterra, L. T.; Green, N. J.; Penfield, K. W.; Miller, J. R.,

Distance, Stereoelectronic Effects, and the Marcus Inverted Region in Intramolecular

Electron-Transfer in Organic Radical-Anions. J Phys Chem-Us 1986, 90 (16), 3673-

3683.

23. Miller, J. R.; Calcaterra, L. T.; Closs, G. L., Intramolecular Long-Distance

Electron-Transfer in Radical-Anions - the Effects of Free-Energy and Solvent on the

Reaction-Rates. J Am Chem Soc 1984, 106 (10), 3047-3049.

24. Closs, G. L.; Miller, J. R., Intramolecular Long-Distance Electron-Transfer in

Organic-Molecules. Science 1988, 240 (4851), 440-447.

25. Aviram, A.; Ratner, M. A., Molecular Rectifiers. Chem Phys Lett 1974, 29 (2),

277-283.

- 194 -

26. Wudl, F.; Smith, G. M.; Hufnagel, E. J., Bis-1,3-Dithiolium Chloride - an

Unusually Stable Organic Radical Cation. Journal of the Chemical Society D-Chemical

Communications 1970, (21), 1453-1454.

27. Acker, D. S.; Hertler, W. R., Substituted Quinodimethans .1. Preparation and

Chemistry of 7,7,8,8-Tetracyanoquinodimethan. J Am Chem Soc 1962, 84 (17), 3370-&.

28. Metzger, R. M., Unimolecular electrical rectifiers. Chem Rev 2003, 103 (9),

3803-3834.

29. Aviram, A.; Joachim, C.; Pomerantz, M., Evidence of Switching and

Rectification by a Single Molecule Effected with a Scanning Tunneling Microscope.

Chem Phys Lett 1988, 146 (6), 490-495.

30. Aviram, A.; Joachim, C.; Pomerantz, M., Evidence of switching and rectification

by a single molecule effected by a scanning tunneling microscope : Chem. Phys. Letters

146 (1988) 490. Chem Phys Lett 1989, 162 (4-5), 416-416.

31. Geddes, N. J.; Sambles, J. R.; Jarvis, D. J.; Parker, W. G.; Sandman, D. J.,

Fabrication and Investigation of Asymmetric Current-Voltage Characteristics of a Metal

Langmuir-Blodgett Monolayer Metal Structure. Appl Phys Lett 1990, 56 (19), 1916-

1918.

32. Geddes, N. J.; Sambles, J. R.; Jarvis, D. J.; Parker, W. G.; Sandman, D. J., The

Electrical-Properties of Metal-Sandwiched Langmuir-Blodgett Multilayers and

Monolayers of a Redox-Active Organic Molecular-Compound. J Appl Phys 1992, 71

(2), 756-768.

33. Ashwell, G. J.; Moczko, K.; Sujka, M.; Chwialkowska, A.; High, L. R. H.;

Sandman, D. J., Molecular diodes and ultra-thin organic rectifying junctions: Au-S-

CnH2n-Q3CNQ and TCNQ derivatives. Phys Chem Chem Phys 2007, 9 (8), 996-1002.

34. Mikayama, T.; Ara, M.; Uehara, K.; Sugimoto, A.; Mizuno, K.; Inoue, N.,

Nanometre-scale photoelectric characteristics of a molecular device monolayer. Phys

Chem Chem Phys 2001, 3 (16), 3459-3462.

35. Brady, A. C.; Hodder, B.; Martin, A. S.; Sambles, J. R.; Ewels, C. P.; Jones, R.;

Briddon, P. R.; Musa, A. M.; Panetta, C. A. F.; Mattern, D. L., Molecular rectification

with M vertical bar(D-sigma-A LB film)vertical bar M junctions. J Mater Chem 1999, 9

(9), 2271-2275.

- 195 -

36. Honciuc, A.; Jaiswal, A.; Gong, A.; Ashworth, H.; Spangler, C. W.; Peterson, I.

R.; Dalton, L. R.; Metzger, R. M., Current rectification in a Langmuir-Schaefer

monolayer of fullerene-bis-[4-diphenylamino-4 ''-(N-ethyl-N-2 '''-ethyl)amino-1,4-

diphenyl-1,3-butadiene]malonate between Au electrodes. J Phys Chem B 2005, 109 (2),

857-871.

37. Chabinyc, M. L.; Chen, X.; Holmlin, R. E.; Jacobs, H.; Skulason, H.; Frisbie, C.

D.; Mujica, V.; Ratner, M. A.; Rampi, M. A.; Whitesides, G. M., Molecular

Rectification in a Metal−Insulator−Metal Junction Based on Self-Assembled

Monolayers. J Am Chem Soc 2002, 124 (39), 11730-11736.

38. Metzger, R. M.; Xu, T.; Peterson, I. R., Electrical rectification by a monolayer of

hexadecylquinolinium tricyanoquinodimethanide measured between macroscopic gold

electrodes. J Phys Chem B 2001, 105 (30), 7280-7290.

39. Baldwin, J. W.; Amaresh, R. R.; Peterson, I. R.; Shumate, W. J.; Cava, M. P.;

Amiri, M. A.; Hamilton, R.; Ashwell, G. J.; Metzger, R. M., Rectification and nonlinear

optical properties of a Langmuir-Blodgett monolayer of a pyridinium dye. J Phys Chem

B 2002, 106 (47), 12158-12164.

40. Wescott, L. D.; Mattern, D. L., Donor−σ−Acceptor Molecules Incorporating a

Nonadecyl-Swallowtailed Perylenediimide Acceptor. The Journal of Organic Chemistry

2003, 68 (26), 10058-10066.

41. Shumate, W. J.; Mattern, D. L.; Jaiswal, A.; Dixon, D. A.; White, T. R.; Burgess,

J.; Honciuc, A.; Metzger, R. M., Spectroscopy and rectification of three donor-sigma-

acceptor compounds, consisting of a one-electron donor (pyrene or ferrocene), a one-

electron acceptor (perylenebisimide), and a C-19 swallowtail. J Phys Chem B 2006, 110

(23), 11146-11159.

42. Martin, A. S.; Sambles, J. R.; Ashwell, G. J., Molecular Rectifier. Phys Rev Lett

1993, 70 (2), 218-221.

43. Ashwell, G. J.; Sambles, J. R.; Martin, A. S.; Parker, W. G.; Szablewski, M.,

Rectifying Characteristics of Mg/(C16h33-Q3cnq Lb Film)/Pt Structures. J Chem Soc

Chem Comm 1990, (19), 1374-1376.

44. Metzger, R. M.; Chen, B.; Hopfner, U.; Lakshmikantham, M. V.; Vuillaume, D.;

Kawai, T.; Wu, X. L.; Tachibana, H.; Hughes, T. V.; Sakurai, H.; Baldwin, J. W.;

- 196 -

Hosch, C.; Cava, M. P.; Brehmer, L.; Ashwell, G. J., Unimolecular electrical

rectification in hexadecylquinolinium tricyanoquinodimethanide. J Am Chem Soc 1997,

119 (43), 10455-10466.

45. Vuillaume, D.; Chen, B.; Metzger, R. M., Electron transfer through a monolayer

of hexadecylquinolinium tricyanoquinodimethanide. Langmuir 1999, 15 (11), 4011-

4017.

46. Chen, B.; Metzger, R. M., Rectification between 370 and 105 K in

hexadecylquinolinium tricyanoquinodimethanide. J Phys Chem B 1999, 103 (21), 4447-

4451.

47. Xu, T.; Peterson, I. R.; Lakshmikantham, M. V.; Metzger, R. M., Rectification

by a monolayer of hexadecylquinolinium tricyanoquino-dimethanide between gold

electrodes. Angew Chem Int Edit 2001, 40 (9), 1749-1752.

48. Metzger, R. M.; Baldwin, J. W.; Shumate, W. J.; Peterson, I. R.; Mani, P.;

Mankey, G. J.; Morris, T.; Szulczewski, G.; Bosi, S.; Prato, M.; Comito, A.; Rubin, Y.,

Electrical rectification in a Langmuir-Blodgett monolayer of dimethyanilinoazafullerene

sandwiched between gold electrodes. J Phys Chem B 2003, 107 (4), 1021-1027.

49. Ashwell, G. J.; Dawnay, E. J. C.; Kuczynski, A. P.; Szablewski, M.; Sandy, I.

M.; Bryce, M. R.; Grainger, A. M.; Hasan, M., Langmuir-Blodgett Alignment of

Zwitterionic Optically Nonlinear D-Pi-a Materials. J Chem Soc Faraday T 1990, 86 (7),

1117-1121.

50. Metzger, R. M.; Chen, B.; Höpfner, U.; Lakshmikantham, M. V.; Vuillaume, D.;

Kawai, T.; Wu, X.; Tachibana, H.; Hughes, T. V.; Sakurai, H.; Baldwin, J. W.; Hosch,

C.; Cava, M. P.; Brehmer, L.; Ashwell, G. J., Unimolecular Electrical Rectification in

Hexadecylquinolinium Tricyanoquinodimethanide. J Am Chem Soc 1997, 119 (43),

10455-10466.

51. Metzger, R. M.; Tachibana, H.; Wu, X.; Höpfner, U.; Chen, B.;

Lakshmikantham, M. V.; Cava, M. P., Is Ashwell's zwitterion a molecular diode?

Synthetic Met 1997, 85 (1-3), 1359-1360.

52. Ashwell, G. J.; Paxton, G. A. N., Multifunctional properties of Z-beta-(N-

hexadecyl-4-quinolinium)-alpha-cyano-4-styryldicyanomethanide: a molecular rectifier,

optically non-linear dye, and ammonia sensor. Aust J Chem 2002, 55 (3), 199-204.

- 197 -

53. Ashwell, G. J.; Hamilton, R.; High, L. R. H., Molecular rectification: asymmetric

current-voltage curves from self-assembled monolayers of a donor-(pi-bridge)-acceptor

dye. J Mater Chem 2003, 13 (7), 1501-1503.

54. Ashwell, G. J.; Chwialkowska, A.; High, L. R. H., Rectifying Au-S-CnH2n-

P3CNQ derivatives. J Mater Chem 2004, 14 (19), 2848-2851.

55. Ashwell, G. J.; Chwialkowska, A.; High, L. R. H., Au-S-CnH2n-Q3CNQ: self-

assembled monolayers for molecular rectification. J Mater Chem 2004, 14 (15), 2389-

2394.

56. Ashwell, G. J.; Amiri, M. A., Alignment of chevron-shaped molecules for

photonic and electronic applicationst. J Mater Chem 2002, 12 (8), 2181-2183.

57. Ng, M. K.; Lee, D. C.; Yu, L. P., Molecular diodes based on conjugated diblock

co-oligomers. J Am Chem Soc 2002, 124 (40), 11862-11863.

58. Ng, M. K.; Yu, L. P., Synthesis of amphiphilic conjugated diblock oligomers as

molecular diodes. Angew Chem Int Edit 2002, 41 (19), 3598-3601.

59. Diez-Perez, I.; Hihath, J.; Lee, Y.; Yu, L. P.; Adamska, L.; Kozhushner, M. A.;

Oleynik, I. I.; Tao, N. J., Rectification and stability of a single molecular diode with

controlled orientation. Nat Chem 2009, 1 (8), 635-641.

60. Krzeminski, C.; Delerue, C.; Allan, G.; Vuillaume, D.; Metzger, R. M., Theory

of electrical rectification in a molecular monolayer. Phys Rev B 2001, 6408 (8), -.

61. Chabinyc, M. L.; Chen, X. X.; Holmlin, R. E.; Jacobs, H.; Skulason, H.; Frisbie,

C. D.; Mujica, V.; Ratner, M. A.; Rampi, M. A.; Whitesides, G. M., Molecular

rectification in a metal-insulator-metal junction based on self-assembled monolayers. J

Am Chem Soc 2002, 124 (39), 11730-11736.

62. Nijhuis, C. A.; Reus, W. F.; Whitesides, G. M., Mechanism of rectification in

tunneling junctions based on molecules with asymmetric potential drops. J Am Chem

Soc 2010, 132 (51), 18386-401.

63. Panetta, C. A.; Baghdadchi, J.; Metzger, R. M., Ttf-Nhco2(Ch2)2o-Tcnqbr and

Ttf-Co2(Ch2)2o-Tcnqbr, 2 Potential Molecular Rectifiers. Mol Cryst Liq Cryst 1984,

107 (1-2), 103-113.

64. Dumur, F.; Gautier, N.; Gallego-Planas, N.; Sahin, Y.; Levillain, E.; Mercier, N.;

Hudhomme, P.; Masino, M.; Girlando, A.; Lloveras, V.; Vidal-Gancedo, J.; Veciana, J.;

- 198 -

Rovira, C., Novel fused D-A dyad and A-D-A triad incorporating tetrathiafulvalene and

p-benzoquinone. J Org Chem 2004, 69 (6), 2164-2177.

65. Scheib, S.; Cava, M. P.; Baldwin, J. W.; Metzger, R. M., In search of D-sigma-A

molecular rectifiers: the D-sigma-A system derived from triptycenequinone and

tetrathiafulvalene. Thin Solid Films 1998, 329, 100-103.

66. Dumur, F.; Guegano, X.; Gautier, N.; Liu, S. X.; Neels, A.; Decurtins, S.;

Hudhomme, P., Approaches to Fused Tetrathiafulvalene/Tetracyanoquinodimethane

Systems. Eur J Org Chem 2009, (36), 6341-6354.

67. Bendikov, M.; Wudl, F.; Perepichka, D. F., Tetrathiafulvalenes, oligoacenenes,

and their buckminsterfullerene derivatives: The brick and mortar of organic electronics.

Chem Rev 2004, 104 (11), 4891-4945.

68. Perepichka, I. F.; Kuz'mina, L. G.; Perepichka, D. F.; Bryce, M. R.; Goldenberg,

L. M.; Popov, A. F.; Howard, J. A. K., Electron acceptors of the fluorene series. 7 2,7-

dicyano-4,5-dinitro-9-X-fluorenes: Synthesis, cyclic voltammetry, charge transfer

complexation with N-propylcarbazole in solution, and X-ray crystal structures of two

tetrathiafulvalene complexes. J Org Chem 1998, 63 (19), 6484-6493.

69. Perepichka, D. F.; Bryce, M. R.; McInnes, E. J. L.; Zhao, J. P., The first

tetrathiafulvalene-sigma-polynitrofluorene diads: Low HOMO-LUMO gap, amphoteric

redox behavior, and charge transfer properties. Org Lett 2001, 3 (10), 1431-1434.

70. Perepichka, D. F.; Bryce, M. R.; Batsanov, A. S.; McInnes, E. J. L.; Zhao, J. P.;

Farley, R. D., Engineering a remarkably low HOMO-LUMO gap by covalent linkage of

a strong pi-donor and a pi-acceptor-tetrathiafulvaleneu-sigma-polynitrofluorene diads:

Their amphoteric redox behavior, electron transfer and spectroscopic properties. Chem-

Eur J 2002, 8 (20), 4656-4669.

71. Martin, N.; Segura, J. L.; Seoane, C., Design and synthesis of TCNQ and DCNQI

type electron acceptor molecules as precursors for 'organic metals'. J Mater Chem 1997,

7 (9), 1661-1676.

72. Perepichka, D. F.; Bryce, M. R.; Pearson, C.; Petty, M. C.; McInnes, E. J. L.;

Zhao, J. P., A covalent tetrathiafulvalene-tetracyanoquinodimethane diad: Extremely

low HOMO-LUMO gap, thermoexcited electron transfer, and high-quality Langmuir-

Blodgett films. Angew Chem Int Edit 2003, 42 (38), 4636-4639.

- 199 -

73. Meijere, A. d.; Diederich, F., Metal-catalyzed cross-coupling reactions. 2nd,

completely rev. and enl. ed.; Wiley-VCH: Weinheim, 2004.

74. Stille, J. K., The Palladium-Catalyzed Cross-Coupling Reactions of Organotin

Reagents with Organic Electrophiles. Angewandte Chemie-International Edition in

English 1986, 25 (6), 508-523.

75. Casado, A. L.; Espinet, P., Mechanism of the Stille reaction. 1. The

transmetalation step. Coupling of (RI)-I-1 and (RSnBu3)-Sn-2 catalyzed by trans-

[(PdRIL2)-I-1] (R-1 = C6Cl2F3; R-2 = vinyl, 4-methoxyphenyl; L = AsPh3). J Am

Chem Soc 1998, 120 (35), 8978-8985.

76. Espinet, P.; Echavarren, A. M., The mechanisms of the Stille reaction. Angew

Chem Int Edit 2004, 43 (36), 4704-4734.

77. Steckler, T. T.; Zhang, X.; Hwang, J.; Honeyager, R.; Ohira, S.; Zhang, X. H.;

Grant, A.; Ellinger, S.; Odom, S. A.; Sweat, D.; Tanner, D. B.; Rinzler, A. G.; Barlow,

S.; Bredas, J. L.; Kippelen, B.; Marder, S. R.; Reynolds, J. R., A Spray-Processable,

Low Bandgap, and Ambipolar Donor-Acceptor Conjugated Polymer. J Am Chem Soc

2009, 131 (8), 2824-2826.

78. Zhang, X.; Steckler, T. T.; Dasari, R. R.; Ohira, S.; Potscavage, W. J.; Tiwari, S.

P.; Coppee, S.; Ellinger, S.; Barlow, S.; Bredas, J. L.; Kippelen, B.; Reynolds, J. R.;

Marder, S. R., Dithienopyrrole-based donor-acceptor copolymers: low band-gap

materials for charge transport, photovoltaics and electrochromism. J Mater Chem 2010,

20 (1), 123-134.

79. Rayleigh, P., Surface Tension. Nature 1891, 43 (1115), 437.

80. Langmuir, I., The constitution and fundamental properties of solids and liquids.

II. Liquids. J Am Chem Soc 1917, 39, 1848-1906.

81. Blodgett, K. B., Films built by depositing successive monomolecular layers on a

solid surface. J Am Chem Soc 1935, 57 (1), 1007-1022.

82. Roberts, G., Langmuir-Blodgett Films. Plenum Press: New York and London,

1990.

83. Tredgold, R. H., The physics of Langmuir-Blodgett films. Reports on Progress

in Physics 1987, 50, 1609-1656.

- 200 -

84. Metzger, R. M.; Panetta, C. A., Langmuir-Blodgett-Films of Potential

Unidimensional Organic Rectifiers. Advanced Organic Solid State Materials 1990, 173,

531-536.

85. Richard, J.; Vandevyver, M.; Lesieur, P.; Ruaudelteixier, A.; Barraud, A.; Bozio,

R.; Pecile, C., Structural-Properties of Langmuir-Blodgett-Films of Charge-Transfer

Salts - Pristine and Iodine Doped Conducting Films of (N-Docosyl-Pyridinium)Tcnq. J

Chem Phys 1987, 86 (4), 2428-2438.

86. Barraud, A.; Ruaudelteixier, A.; Vandevyver, M.; Lesieur, P., Transfer Ion-

Radical Salt with Conducting Charge in Langmuir-Blodgett Films. Nouv J Chim 1985, 9

(5), 365-367.

87. Pearson, C.; Dhindsa, A. S.; Bryce, M. R.; Petty, M. C., Alternate-Layer

Langmuir-Blodgett Films of Long-Chain Tcnq and Ttf Derivatives. Synthetic Met 1989,

31 (2), 275-279.

88. Petty, M. C., Langmuir-Blodgett films : an introduction. Cambridge University

Press: Cambridge ; New York, 1996; p xviii, 234 p.

89. Bigelow, W. C.; Pickett, D. L.; Zisman, W. A., Oleophobic Monolayers .1. Films

Adsorbed from Solution in Non-Polar Liquids. J Coll Sci Imp U Tok 1946, 1 (6), 513-

538.

90. Nuzzo, R. G.; Allara, D. L., Adsorption of Bifunctional Organic Disulfides on

Gold Surfaces. J Am Chem Soc 1983, 105 (13), 4481-4483.

91. Bain, C. D.; Evall, J.; Whitesides, G. M., Formation of Monolayers by the

Coadsorption of Thiols on Gold - Variation in the Head Group, Tail Group, and Solvent.

J Am Chem Soc 1989, 111 (18), 7155-7164.

92. Holmesfarley, S. R.; Bain, C. D.; Whitesides, G. M., Wetting of Functionalized

Polyethylene Film Having Ionizable Organic-Acids and Bases at the Polymer Water

Interface - Relations between Functional-Group Polarity, Extent of Ionization, and

Contact-Angle with Water. Langmuir 1988, 4 (4), 921-937.

93. Ulman, A., Formation and structure of self-assembled monolayers. Chem Rev

1996, 96 (4), 1533-1554.

- 201 -

94. Allara, D. L.; Nuzzo, R. G., Spontaneously Organized Molecular Assemblies .1.

Formation, Dynamics, and Physical-Properties of Normal-Alkanoic Acids Adsorbed

from Solution on an Oxidized Aluminum Surface. Langmuir 1985, 1 (1), 45-52.

95. Schlotter, N. E.; Porter, M. D.; Bright, T. B.; Allara, D. L., Formation and

Structure of a Spontaneously Adsorbed Monolayer of Arachidic on Silver. Chem Phys

Lett 1986, 132 (1), 93-98.

96. Poirier, G. E.; Pylant, E. D., The self-assembly mechanism of alkanethiols on

Au(111). Science 1996, 272 (5265), 1145-1148.

97. Porter, M. D.; Bright, T. B.; Allara, D. L.; Chidsey, C. E. D., Spontaneously

Organized Molecular Assemblies .4. Structural Characterization of Normal-Alkyl Thiol

Monolayers on Gold by Optical Ellipsometry, Infrared-Spectroscopy, and

Electrochemistry. J Am Chem Soc 1987, 109 (12), 3559-3568.

98. Dubois, L. H.; Nuzzo, R. G., Synthesis, Structure, and Properties of Model

Organic-Surfaces. Annu Rev Phys Chem 1992, 43, 437-463.

99. Bain, C. D.; Whitesides, G. M., Molecular-Level Control over Surface Order in

Self-Assembled Monolayer Films of Thiols on Gold. Science 1988, 240 (4848), 62-63.

100. Biebuyck, H. A.; Bian, C. D.; Whitesides, G. M., Comparison of Organic

Monolayers on Polycrystalline Gold Spontaneously Assembled from Solutions

Containing Dialkyl Disulfides or Alkenethiols. Langmuir 1994, 10 (6), 1825-1831.

101. Laibinis, P. E.; Whitesides, G. M.; Allara, D. L.; Tao, Y. T.; Parikh, A. N.;

Nuzzo, R. G., Comparison of the Structures and Wetting Properties of Self-Assembled

Monolayers of Normal-Alkanethiols on the Coinage Metal-Surfaces, Cu, Ag, Au. J Am

Chem Soc 1991, 113 (19), 7152-7167.

102. Dubois, L. H.; Zegarski, B. R.; Nuzzo, R. G., Molecular Ordering of

Organosulfur Compounds on Au(111) and Au(100) - Adsorption from Solution and in

Ultrahigh-Vacuum. J Chem Phys 1993, 98 (1), 678-688.

103. Walczak, M. M.; Chung, C. K.; Stole, S. M.; Widrig, C. A.; Porter, M. D.,

Structure and Interfacial Properties of Spontaneously Adsorbed Normal-Alkanethiolate

Monolayers on Evaporated Silver Surfaces. J Am Chem Soc 1991, 113 (7), 2370-2378.

- 202 -

104. Fenter, P.; Eisenberger, P.; Li, J.; Camillone, N.; Bernasek, S.; Scoles, G.;

Ramanarayanan, T. A.; Liang, K. S., Structure of Ch3(Ch2)17sh Self-Assembled on the

Ag(111) Surface - an Incommensurate Monolayer. Langmuir 1991, 7 (10), 2013-2016.

105. Love, J. C.; Wolfe, D. B.; Haasch, R.; Chabinyc, M. L.; Paul, K. E.; Whitesides,

G. M.; Nuzzo, R. G., Formation and structure of self-assembled monolayers of

alkanethiolates on palladium. J Am Chem Soc 2003, 125 (9), 2597-2609.

106. Carvalho, A.; Geissler, M.; Schmid, H.; Michel, B.; Delamarche, E., Self-

assembled monolayers of eicosanethiol on palladium and their use in microcontact

printing. Langmuir 2002, 18 (6), 2406-2412.

107. Li, Z. Y.; Chang, S. C.; Williams, R. S., Self-assembly of alkanethiol molecules

onto platinum and platinum oxide surfaces. Langmuir 2003, 19 (17), 6744-6749.

108. Muskal, N.; Turyan, I.; Mandler, D., Self-assembled monolayers on mercury

surfaces. J Electroanal Chem 1996, 409 (1-2), 131-136.

109. Tidwell, C. D.; Ertel, S. I.; Ratner, B. D.; Tarasevich, B. J.; Atre, S.; Allara, D.

L., Endothelial cell growth and protein adsorption on terminally functionalized, self-

assembled monolayers of alkanethiolates on gold. Langmuir 1997, 13 (13), 3404-3413.

110. Holmlin, R. E.; Chen, X. X.; Chapman, R. G.; Takayama, S.; Whitesides, G. M.,

Zwitterionic SAMs that resist nonspecific adsorption of protein from aqueous buffer.

Langmuir 2001, 17 (9), 2841-2850.

111. Engtrakul, C.; Sita, L. R., Ferrocene-based nanoelectronics: 2,5-

diethynylpyridine as a reversible switching element. Nano Lett 2001, 1 (10), 541-549.

112. Gooding, J. J.; Mearns, F.; Yang, W. R.; Liu, J. Q., Self-assembled monolayers

into the 21(st) century: Recent advances and applications. Electroanal 2003, 15 (2), 81-

96.

113. Haag, R.; Rampi, M. A.; Holmlin, R. E.; Whitesides, G. M., Electrical

breakdown of aliphatic and aromatic self-assembled monolayers used as nanometer-

thick organic dielectrics. J Am Chem Soc 1999, 121 (34), 7895-7906.

114. Holmlin, R. E.; Haag, R.; Chabinyc, M. L.; Ismagilov, R. F.; Cohen, A. E.;

Terfort, A.; Rampi, M. A.; Whitesides, G. M., Electron transport through thin organic

films in metal-insulator-metal junctions based on self-assembled monolayers. J Am

Chem Soc 2001, 123 (21), 5075-5085.

- 203 -

115. Rampi, M. A.; Schueller, O. J. A.; Whitesides, G. M., Alkanethiol self-assembled

monolayers as the dielectric of capacitors with nanoscale thickness. Appl Phys Lett

1998, 72 (14), 1781-1783.

116. Love, J. C.; Estroff, L. A.; Kriebel, J. K.; Nuzzo, R. G.; Whitesides, G. M., Self-

Assembled Monolayers of Thiolates on Metals as a Form of Nanotechnology. Chem Rev

2005, 105 (4), 1103-1170.

117. Strong, L.; Whitesides, G. M., Structures of Self-Assembled Monolayer Films of

Organosulfur Compounds Adsorbed on Gold Single-Crystals - Electron-Diffraction

Studies. Langmuir 1988, 4 (3), 546-558.

118. Dubois, L. H. N., R. G., Annu. Rev. Phys. Chem. 1992, 43.

119. Bain, C. D.; Whitesides, G. M., Formation of Monolayers by the Coadsorption of

Thiols on Gold - Variation in the Length of the Alkyl Chain. J Am Chem Soc 1989, 111

(18), 7164-7175.

120. Bain, C. D.; Whitesides, G. M., Modeling Organic-Surfaces with Self-Assembled

Monolayers. Angewandte Chemie-International Edition in English 1989, 28 (4), 506-

512.

121. Dubois, L. H.; Nuzzo, R. G., Synthesis, Structure, and Properties of Model

Organic Surfaces. Annu Rev Phys Chem 1992, 43 (1), 437-463.

122. Kaifer, A. E.; Gómez-Kaifer, M., Supramolecular electrochemistry. Wiley-VCH:

Weinheim ; New York, 1999; p xiv, 241 p.

123. Schlenoff, J. B.; Li, M.; Ly, H., Stability and self-exchange in alkanethiol

monolayers. J Am Chem Soc 1995, 117 (50), 12528-12536.

124. Nuzzo, R. G.; Zegarski, B. R.; Dubois, L. H., Fundamental-Studies of the

Chemisorption of Organosulfur Compounds on Au(111) - Implications for Molecular

Self-Assembly on Gold Surfaces. J Am Chem Soc 1987, 109 (3), 733-740.

125. Lekner, J., Ellipsometry of a Thin-Film between Similar Media. J Opt Soc Am A

1988, 5 (7), 1041-1043.

126. Lekner, J., Ellipsometry of Surface-Films on a Uniform Layer. J Opt Soc Am A

1988, 5 (7), 1044-1047.

127. Theeten, J. B.; Aspnes, D. E., Ellipsometry in Thin-Film Analysis. Annu Rev

Mater Sci 1981, 11, 97-122.

- 204 -

128. Cresswell, J. P., Uniaxial Ellipsometry of Langmuir-Blodgett-Films. Langmuir

1994, 10 (10), 3727-3729.

129. Fujiwara, H., Spectroscopic Ellipsometry: Principles and Applications. John

Wiley & Sons: 2007.

130. Collins, R. W.; An, I.; Chen, C.; Ferlauto, A. S.; Zapien, J. A., Advances in

multichannel ellipsometric techniques for in-situ and real-time characterization of thin

films. Thin Solid Films 2004, 469-70, 38-46.

131. Auciello, O.; Krauss, A. R., In situ real time characterization of thin films /

edited by Orlando Auciello, Alan R. Krauss. Wiley: New York, 2001; p xi, 263 p.

132. Degennes, P. G., Wetting - Statics and Dynamics. Rev Mod Phys 1985, 57 (3),

827-863.

133. Förch, R.; Schönherr, H.; Jenkins, A. T. A., Surface design : applications in

bioscience and nanotechnology. Wiley-VCH: Weinheim, 2009; p xxiii, 511 p.

134. Bain, C. D.; Whitesides, G. M., A Study by Contact-Angle of the Acid-Base

Behavior of Monolayers Containing Omega-Mercaptocarboxylic Acids Adsorbed on

Gold - an Example of Reactive Spreading. Langmuir 1989, 5 (6), 1370-1378.

135. Efimenko, K.; Novick, B.; Carbonell, R. G.; DeSimone, J. M.; Genzer, J.,

Formation of self-assembled monolayers of semifluorinated and hydrocarbon

chlorosilane precursors on silica surfaces from liquid carbon dioxide. Langmuir 2002, 18

(16), 6170-6179.

136. Doughty, S. K.; Rowlen, K. L., Molecular Orientation at Dielectric Surfaces by

Angle-Resolved Photoacoustic Spectroscopy. The Journal of Physical Chemistry 1995,

99 (7), 2143-2150.

137. Olmstead, M. A.; Amer, N. M.; Kohn, S.; Fournier, D.; Boccara, A. C.,

Photothermal displacement spectroscopy: An optical probe for solids and surfaces.

Applied Physics A: Materials Science &amp; Processing 1983, 32 (3), 141-154.

138. Francis, S. A.; Ellison, A. H., Infrared Spectra of Monolayers on Metal Mirrors. J

Opt Soc Am 1959, 49 (2), 131-138.

139. Allara, D. L.; Swalen, J. D., An Infrared Reflection Spectroscopy Study of

Oriented Cadmium Arachidate Monolayer Films on Evaporated Silver. J Phys Chem-Us

1982, 86 (14), 2700-2704.

- 205 -

140. Petty, M. C., Molecular electronics: Prospects for instrumentation and

measurement science. Meas Sci Technol 1996, 7 (5), 725-735.

141. Mirabella, F. M., Internal-Reflection Spectroscopy. Appl Spectrosc Rev 1985, 21

(1-2), 45-178.

142. Hansen, W. N., Methods of Internal Reflection Spectroscopy of Interest to

Surface Chemistry. Appl Spectrosc 1968, 22 (4), 351.

143. Greenler, R. G., Infrared Study of Adsorbed Molecules on Metal Surfaces by

Reflection Techniques. J Chem Phys 1966, 44 (1), 310-315.

144. Debe, M. K.; Tushaus, L. A., Reflection Absorption Infrared Studies of the

Surface Diffused Residues on a Humidity Temperature Aged Urethane Psa. J Adhesion

1983, 15 (3-4), 287-306.

145. Chollet, P. A., Determination by Infrared-Absorption of Orientation of Molecules

in Mono-Molecular Layers. Thin Solid Films 1978, 52 (3), 343-360.

146. Tour, J. M.; Jones, L.; Pearson, D. L.; Lamba, J. J. S.; Burgin, T. P.; Whitesides,

G. M.; Allara, D. L.; Parikh, A. N.; Atre, S. V., Self-Assembled Monolayers and

Multilayers of Conjugated Thiols, Alpha,Omega-Dithiols, and Thioacetyl-Containing

Adsorbates - Understanding Attachments between Potential Molecular Wires and Gold

Surfaces. J Am Chem Soc 1995, 117 (37), 9529-9534.

147. Laibinis, P. E.; Bain, C. D.; Nuzzo, R. G.; Whitesides, G. M., Structure and

Wetting Properties of Omega-Alkoxy-N-Alkanethiolate Monolayers on Gold and Silver.

J Phys Chem-Us 1995, 99 (19), 7663-7676.

148. Folkers, J. P.; Gorman, C. B.; Laibinis, P. E.; Buchholz, S.; Whitesides, G. M.;

Nuzzo, R. G., Self-Assembled Monolayers of Long-Chain Hydroxamic Acids on the

Native Oxides of Metals. Langmuir 1995, 11 (3), 813-824.

149. Nuzzo, R. G.; Korenic, E. M.; Dubois, L. H., Studies of the Temperature-

Dependent Phase-Behavior of Long-Chain Normal-Alkyl Thiol Monolayers on Gold. J

Chem Phys 1990, 93 (1), 767-773.

150. Mendelsohn, R.; Brauner, J. W.; Gericke, A., External Infrared Reflection-

Absorption Spectrometry Monolayer Films at the Air-Water-Interface. Annu Rev Phys

Chem 1995, 46, 305-334.

- 206 -

151. Gericke, A.; Huhnerfuss, H., In-Situ Investigation of Saturated Long-Chain

Fatty-Acids at the Air-Water-Interface by External Infrared Reflection-Absorption

Spectrometry. J Phys Chem-Us 1993, 97 (49), 12899-12908.

152. Park, S. Y.; Franses, E., Hexadecanol Microstructures of Crystallites in Aqueous

Dispersions and of Langmuir-Blodgett Monolayers. Langmuir 1995, 11 (6), 2187-2194.

153. Gericke, A.; Mendelsohn, R., Partial chain deuteration as an IRRAS probe of

conformational order of different regions in hexadecanoic acid monolayers at the

air/water interface. Langmuir 1996, 12 (3), 758-762.

154. Terashita, S.; Ozaki, Y.; Iriyama, K., Infrared Study on Order-Disorder

Transitions in Langmuir-Blodgett-Films of 2-Dodecyl-7,7,8,8-

Tetracyanoquinodimethane, 2-Pentacecyl-7,7,8,8-Tetracyanoquinodimethane, and 2-

Octadecyl-7,7,8,8-Tetracyanoquinodimethane. J Phys Chem-Us 1993, 97 (40), 10445-

10452.

155. Jiang, C. Y.; He, W. J.; Huang, J. G.; Tai, Z. H.; Liang, Y. Q., FT-IR studies of

N-hexadecyl-5-iminomethyl-8-hydroxyquinoline Langmuir-Blodgett films. Mater Chem

Phys 2000, 62 (3), 236-240.

156. Dluhy, R. A.; Ping, Z.; Faucher, K.; Brockman, J. M., Infrared spectroscopy of

aqueous biophysical monolayers. Thin Solid Films 1998, 329, 308-314.

157. Nuzzo, R. G.; Dubois, L. H.; Allara, D. L., Fundamental-Studies of Microscopic

Wetting on Organic-Surfaces .1. Formation and Structural Characterization of a Self-

Consistent Series of Polyfunctional Organic Monolayers. J Am Chem Soc 1990, 112 (2),

558-569.

158. Finklea, H. O.; Snider, D. A.; Fedyk, J.; Sabatani, E.; Gafni, Y.; Rubinstein, I.,

Characterization of Octadecanethiol-Coated Gold Electrodes as Microarray Electrodes

by Cyclic Voltammetry and Ac-Impedance Spectroscopy. Langmuir 1993, 9 (12), 3660-

3667.

159. Sabatani, E.; Rubinstein, I., Organized Self-Assembling Monolayers on

Electrodes .2. Monolayer-Based Ultramicroelectrodes for the Study of Very Rapid

Electrode-Kinetics. J Phys Chem-Us 1987, 91 (27), 6663-6669.

- 207 -

160. Sabatani, E.; Rubinstein, I.; Maoz, R.; Sagiv, J., Organized Self-Assembling

Monolayers on Electrodes .1. Octadecyl Derivatives on Gold. J Electroanal Chem 1987,

219 (1-2), 365-371.

161. Finklea, H. O.; Snider, D. A.; Fedyk, J., Passivation of Pinholes in

Octadecanethiol Monolayers on Gold Electrodes by Electrochemical Polymerization of

Phenol. Langmuir 1990, 6 (2), 371-376.

162. Murray, R. W., Chemically Modified Electrodes. Accounts Chem Res 1980, 13

(5), 135-141.

163. Kondo, T.; Takechi, M.; Sato, Y.; Uosaki, K., Adsorption Behavior of

Functionalized Ferrocenylalkane Thiols and Disulfide onto Au and Ito and

Electrochemical Properties of Modified Electrodes - Effects of Acyl and Alkyl-Groups

Attached to the Ferrocene Ring. J Electroanal Chem 1995, 381 (1-2), 203-209.

164. Chidsey, C. E. D.; Bertozzi, C. R.; Putvinski, T. M.; Mujsce, A. M.,

Coadsorption of Ferrocene-Terminated and Unsubstituted Alkanethiols on Gold -

Electroactive Self-Assembled Monolayers. J Am Chem Soc 1990, 112 (11), 4301-4306.

165. Collard, D. M.; Fox, M. A., Use of Electroactive Thiols to Study the Formation

and Exchange of Alkanethiol Monolayers on Gold. Langmuir 1991, 7 (6), 1192-1197.

166. Creager, S. E.; Rowe, G. K., Redox Properties of Ferrocenylalkane Thiols

Coadsorbed with Linear Normal-Alkanethiols on Polycrystalline Bulk Gold Electrodes.

Anal Chim Acta 1991, 246 (1), 233-239.

167. Rowe, G. K.; Creager, S. E., Chain-Length and Solvent Effects on Competitive

Self-Assembly of Ferrocenylhexanethiol and 1-Alkanethiols onto Gold. Langmuir 1994,

10 (4), 1186-1192.

168. Rowe, G. K.; Creager, S. E., Interfacial Solvation and Double-Layer Effects on

Redox Reactions in Organized Assemblies. J Phys Chem-Us 1994, 98 (21), 5500-5507.

169. Smalley, J. F.; Feldberg, S. W.; Chidsey, C. E. D.; Linford, M. R.; Newton, M.

D.; Liu, Y. P., The Kinetics of Electron-Transfer through Ferrocene-Terminated

Alkanethiol Monolayers on Gold. J Phys Chem-Us 1995, 99 (35), 13141-13149.

170. Everett, W. R.; Fritschfaules, I., Factors That Influence the Stability of Self-

Assembled Organothiols on Gold under Electrochemical Conditions. Anal Chim Acta

1995, 307 (2-3), 253-268.

- 208 -

171. Everett, W. R.; Welch, T. L.; Reed, L.; Fritschfaules, I., Potential-Dependent

Stability of Self-Assembled Organothiols on Gold Electrodes in Methylene-Chloride.

Anal Chem 1995, 67 (2), 292-298.

172. Lee, L. Y. S.; Lennox, R. B., Electrochemical desorption of n-alkylthiol SAMs

on polycrystalline gold: Studies using a ferrocenylalkylthiol probe. Langmuir 2007, 23

(1), 292-296.

173. Cai, L. T.; Skulason, H.; Kushmerick, J. G.; Pollack, S. K.; Naciri, J.;

Shashidhar, R.; Allara, D. L.; Mallouk, T. E.; Mayer, T. S., Nanowire-based molecular

monolayer junctions: Synthesis, assembly, and electrical characterization. J Phys Chem

B 2004, 108 (9), 2827-2832.

174. Vilan, A.; Cahen, D., Soft contact deposition onto molecularly modified GaAs.

Thin metal film flotation: Principles and electrical effects. Adv Funct Mater 2002, 12

(11-12), 795-807.

175. Dunbar, T. D.; Cygan, M. T.; Bumm, L. A.; McCarty, G. S.; Burgin, T. P.;

Reinerth, W. A.; Jones, L.; Jackiw, J. J.; Tour, J. M.; Weiss, P. S.; Allara, D. L.,

Combined scanning tunneling microscopy and infrared spectroscopic characterization of

mixed surface assemblies of linear conjugated guest molecules in host alkanethiolate

monolayers on gold. J Phys Chem B 2000, 104 (20), 4880-4893.

176. Wold, D. J.; Frisbie, C. D., Fabrication and characterization of metal-molecule-

metal junctions by conducting probe atomic force microscopy. J Am Chem Soc 2001,

123 (23), 5549-5556.

177. Tran, E.; Rampi, M. A.; Whitesides, G. M., Electron transfer in a Hg-

SAM//SAM-Hg junction mediated by redox centers. Angew Chem Int Edit 2004, 43

(29), 3835-3839.

178. Akkerman, H. B.; Blom, P. W. M.; de Leeuw, D. M.; de Boer, B., Towards

molecular electronics with large-area molecular junctions. Nature 2006, 441 (7089), 69-

72.

179. Reed, M. A.; Zhou, C.; Muller, C. J.; Burgin, T. P.; Tour, J. M., Conductance of

a molecular junction. Science 1997, 278 (5336), 252-254.

- 209 -

180. Jung, D. R.; Czanderna, A. W.; Herdt, G. C., Interactions and penetration at

metal/self-assemble organic monolayer interfaces. J Vac Sci Technol A 1996, 14 (3),

1779-1787.

181. Collier, C. P.; Wong, E. W.; Belohradsky, M.; Raymo, F. M.; Stoddart, J. F.;

Kuekes, P. J.; Williams, R. S.; Heath, J. R., Electronically configurable molecular-based

logic gates. Science 1999, 285 (5426), 391-394.

182. Chang, S. C.; Li, Z. Y.; Lau, C. N.; Larade, B.; Williams, R. S., Investigation of a

model molecular-electronic rectifier with an evaporated Ti-metal top contact. Appl Phys

Lett 2003, 83 (15), 3198-3200.

183. Zhou, C.; Deshpande, M. R.; Reed, M. A.; Jonesll, L.; Tour, J. M., Nanoscale

metal self-assembled monolayer metal heterostructures (vol 71, pg 611, 1997). Appl

Phys Lett 1997, 71 (19), 2857-2857.

184. Chen, J.; Reed, M. A.; Rawlett, A. M.; Tour, J. M., Large on-off ratios and

negative differential resistance in a molecular electronic device. Science 1999, 286

(5444), 1550-1552.

185. Mann, B.; Kuhn, H., Tunneling through Fatty Acid Salt Monolayers. J Appl Phys

1971, 42 (11), 4398-4405.

186. Slowinski, K.; Fong, H. K. Y.; Majda, M., Mercury-mercury tunneling junctions.

1. Electron tunneling across symmetric and asymmetric alkanethiolate bilayers. J Am

Chem Soc 1999, 121 (31), 7257-7261.

187. Slowinski, K.; Majda, M., Mercury-mercury tunneling junctions - Part II.

Structure and stability of symmetric alkanethiolate bilayers and their effect on the rate of

electron tunneling. J Electroanal Chem 2000, 491 (1-2), 139-147.

188. Holmlin, R. E.; Ismagilov, R. F.; Haag, R.; Mujica, V.; Ratner, M. A.; Rampi, M.

A.; Whitesides, G. M., Correlating electron transport and molecular structure in organic

thin films. Angew Chem Int Edit 2001, 40 (12), 2316-2320.

189. Selzer, Y.; Salomon, A.; Cahen, D., Effect of molecule-metal electronic coupling

on through-bond hole tunneling across metal-organic monolayer-semiconductor

junctions. J Am Chem Soc 2002, 124 (12), 2886-2887.

190. Weiss, E. A.; Chiechi, R. C.; Kaufman, G. K.; Kriebel, J. K.; Li, Z.; Duati, M.;

Rampi, M. A.; Whitesides, G. M., Influence of Defects on the Electrical Characteristics

- 210 -

of Mercury-Drop Junctions:  Self-Assembled Monolayers of n-Alkanethiolates on

Rough and Smooth Silver. J Am Chem Soc 2007, 129 (14), 4336-4349.

191. Nijhuis, C. A.; Reus, W. F.; Whitesides, G. M., Molecular Rectification in Metal-

SAM-Metal Oxide-Metal Junctions. J Am Chem Soc 2009, 131 (49), 17814-17827.

192. Chiechi, R. C.; Weiss, E. A.; Dickey, M. D.; Whitesides, G. M., Eutectic

gallium-indium (EGaIn): A moldable liquid metal for electrical characterization of self-

assembled monolayers. Angew Chem Int Edit 2008, 47 (1), 142-144.

193. Xue, Y. Q.; Datta, S.; Ratner, M. A., Charge transfer and "band lineup" in

molecular electronic devices: A chemical and numerical interpretation. J Chem Phys

2001, 115 (9), 4292-4299.

194. Kushmerick, J. G.; Whitaker, C. M.; Pollack, S. K.; Schull, T. L.; Shashidhar, R.,

Tuning current rectification across molecular junctions. Nanotechnology 2004, 15 (7),

S489-S493.

195. Mujica, V.; Kemp, M.; Ratner, M. A., Electron conduction in molecular wires. I.

A scattering formalism. The Journal of Chemical Physics 1994, 101 (8), 6849-6855.

196. Kushmerick, J. G.; Holt, D. B.; Yang, J. C.; Naciri, J.; Moore, M. H.; Shashidhar,

R., Metal-molecule contacts and charge transport across monomolecular layers:

Measurement and theory. Phys Rev Lett 2002, 89 (8), -.

197. Datta, S.; Tian, W. D.; Hong, S. H.; Reifenberger, R.; Henderson, J. I.; Kubiak,

C. P., Current-voltage characteristics of self-assembled monolayers by scanning

tunneling microscopy. Phys Rev Lett 1997, 79 (13), 2530-2533.

198. Tian, W.; Datta, S.; Hong, S.; Reifenberger, R.; Henderson, J. I.; Kubiak, C. P.,

Conductance spectra of molecular wires. The Journal of Chemical Physics 1998, 109

(7), 2874-2882.

199. Petty, M. C., Molecular electronics : from principles to practice. John Wiley &

Sons: Chichester, England ; Hoboken, NJ, 2007; p xxiii, 517 p.

200. Ferraris, J.; Walatka, V.; Perlstei.Jh; Cowan, D. O., Electron-Transfer in a New

Highly Conducting Donor-Acceptor Complex. J Am Chem Soc 1973, 95 (3), 948-949.

201. Carroll, R. L.; Gorman, C. B., The Genesis of Molecular Electronics.

Angewandte Chemie International Edition 2002, 41 (23), 4378-4400.

- 211 -

202. Maruccio, G.; Cingolani, R.; Rinaldi, R., Projecting the nanoworld: Concepts,

results and perspectives of molecular electronics. J Mater Chem 2004, 14 (4), 542-554.

203. Mukherjee, T. K., Electron affinities of polynuclear acceptors : Nitro derivatives

of fluoren-[Delta]9,[alpha]-malononitrile. Tetrahedron 1968, 24 (2), 721-728.

204. Perepichka, D. F.; Bryce, M. R.; Batsanov, A. S.; McInnes, E. J. L.; Zhao, J. P.;

Farley, R. D., Engineering a Remarkably Low HOMO–LUMO Gap by Covalent

Linkage of a Strong π-Donor and a π-Acceptor—Tetrathiafulvalene-σ-Polynitrofluorene

Diads: Their Amphoteric Redox Behavior, Electron Transfer and Spectroscopic

Properties. Chemistry – A European Journal 2002, 8 (20), 4656-4669.

205. Perepichka, D. F.; Bryce, M. R.; Batsanov, A. S.; Howard, J. A. K.; Cuello, A.

O.; Gray, M.; Rotello, V. M., Trialkyltetrathiafulvalene-sigma-

tetracyanoanthraquinodimethane (R3TTF-sigma-TCNAQ) diads: Synthesis,

intramolecular charge-transfer properties, and X-ray crystal structure. J Org Chem 2001,

66 (13), 4517-4524.

206. Baerends, E. J.; Gritsenko, O. V., A quantum chemical view of density functional

theory. J Phys Chem A 1997, 101 (30), 5383-5403.

207. Ishii, H.; Sugiyama, K.; Ito, E.; Seki, K., Energy level alignment and interfacial

electronic structures at organic metal and organic organic interfaces. Adv Mater 1999, 11

(8), 605-625.

208. Perepichka, I. F.; Kuz'mina, L. G.; Perepichka, D. F.; Bryce, M. R.; Goldenberg,

L. M.; Popov, A. F.; Howard, J. A. K., Electron Acceptors of the Fluorene Series. 7.1

2,7-Dicyano-4,5-dinitro-9-X-fluorenes: Synthesis, Cyclic Voltammetry, Charge Transfer

Complexation with N-Propylcarbazole in Solution, and X-ray Crystal Structures of Two

Tetrathiafulvalene Complexes. The Journal of Organic Chemistry 1998, 63 (19), 6484-

6493.

209. Perepichka, D. F.; Bryce, M. R., Molecules with Exceptionally Small HOMO–

LUMO Gaps. Angewandte Chemie International Edition 2005, 44 (34), 5370-5373.

210. Cognard, J., Adhesion to Gold: A Review. Gold Bulletin 1984, 17 (4), 131-139.

211. Chappell, J. S.; Bloch, A. N.; Bryden, W. A.; Maxfield, M.; Poehler, T. O.;

Cowan, D. O., Degree of Charge-Transfer in Organic Conductors by Infrared-

Absorption Spectroscopy. J Am Chem Soc 1981, 103 (9), 2442-2443.

- 212 -

212. Lakshmikantham, M. V.; Garito, A. F.; Cava, M. P., S-Oxides of

Tetrathiafulvalenes. J Org Chem 1978, 43 (22), 4394-4395.

213. Lafrance, C. P.; Nabet, A.; Prudhomme, R. E.; Pezolet, M., On the Relationship

between the Order-Parameter [P-2(Cos Theta)] and the Shape of Orientation

Distributions. Can J Chem 1995, 73 (9), 1497-1505.

214. Picard, F.; Buffeteau, T.; Desbat, B.; Auger, M.; Pezolet, M., Quantitative

orientation measurements in thin lipid films by attenuated total reflection infrared

spectroscopy. Biophys J 1999, 76 (1), 539-551.

215. Harrick, N. J.; Carlson, A. I., Internal Reflection Spectroscopy - Validity of

Effective Thickness Equations. Appl Optics 1971, 10 (1), 19-23.

216. Konstadinidis, K.; Zhang, P.; Opila, R. L.; Allara, D. L., An in-Situ X-Ray

Photoelectron Study of the Interaction between Vapor-Deposited Ti Atoms and

Functional-Groups at the Surfaces of Self-Assembled Monolayers. Surf Sci 1995, 338

(1-3), 300-312.

217. Walker, A. V.; Tighe, T. B.; Stapleton, J.; Haynie, B. C.; Upilli, S.; Allara, D. L.;

Winograd, N., Interaction of vapor-deposited Ti and Au with molecular wires. Appl

Phys Lett 2004, 84 (20), 4008-4010.

218. de Boer, B.; Frank, M. M.; Chabal, Y. J.; Jiang, W. R.; Garfunkel, E.; Bao, Z.,

Metallic contact formation for molecular electronics: interactions between vapor-

deposited metals and self-assembled monolayers of conjugated mono- and dithiols.

Langmuir 2004, 20 (5), 1539-1542.

219. Philipp, G.; Muller-Schwanneke, C.; Burghard, M.; Roth, S.; von Klitzing, K.,

Gold cluster formation at the interface of a gold Langmuir-Blodgett film gold

microsandwich resulting in Coulomb charging phenomena. J Appl Phys 1999, 85 (6),

3374-3376.

220. Walker, A. V.; Tighe, T. B.; Cabarcos, O. M.; Reinard, M. D.; Haynie, B. C.;

Uppili, S.; Winograd, N.; Allara, D. L., The dynamics of noble metal atom penetration

through methoxy-terminated alkanethiolate monolayers. J Am Chem Soc 2004, 126 (12),

3954-3963.

221. Ashwell, G. J.; Bonham, J. S.; Lyons, L. E., Photoconduction in Magnesium

Phthalocyanine Thin-Films. Aust J Chem 1980, 33 (7), 1619-1623.

- 213 -

222. McCreery, R.; Dieringer, J.; Solak, A. O.; Snyder, B.; Nowak, A. M.; McGovern,

W. R.; DuVall, S., Molecular rectification and conductance switching in carbon-based

molecular junctions by structural rearrangement accompanying electron injection. J Am

Chem Soc 2003, 125 (35), 10748-10758.

223. McCreery, R.; Dieringer, J.; Solak, A. O.; Snyder, B.; Nowak, A. M.; McGovern,

W. R.; DuVall, S., Molecular rectification and conductance switching in carbon-based

molecular junctions by structural rearrangement accompanying electron injection (vol

125, pg 10748, 2003). J Am Chem Soc 2004, 126 (19), 6200-6200.

224. Ho, G.; Heath, J. R.; Kondratenko, M.; Perepichka, D. F.; Arseneault, K.;

Pezolet, M.; Bryce, M. R., The first studies of a tetrathiafulvalene-sigma-acceptor

molecular rectifier. Chem-Eur J 2005, 11 (10), 2914-2922.

225. Collier, C. P.; Mattersteig, G.; Wong, E. W.; Luo, Y.; Beverly, K.; Sampaio, J.;

Raymo, F. M.; Stoddart, J. F.; Heath, J. R., A [2]catenane-based solid state electronically

reconfigurable switch. Science 2000, 289 (5482), 1172-1175.

226. Collier, C. P.; Jeppesen, J. O.; Luo, Y.; Perkins, J.; Wong, E. W.; Heath, J. R.;

Stoddart, J. F., Molecular-based electronically switchable tunnel junction devices. J Am

Chem Soc 2001, 123 (50), 12632-41.

227. Ron, H.; Matlis, S.; Rubinstein, I., Self-assembled monolayers on oxidized

metals. 2. Gold surface oxidative pretreatment, monolayer properties, and depression

formation. Langmuir 1998, 14 (5), 1116-1121.

228. Gaussian 03, R. C., M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria,

M. A. Robb, J. R. Cheeseman, J.; A. Montgomery, J., T. Vreven, K. N. Kudin, J. C.

Burant, J. M. Millam, S. S. Iyengar, J. Tomasi, V. Barone, B. Mennucci,; M. Cossi, G.

S., N. Rega, G. A. Petersson, H. Nakatsuji, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J.

Hasegawa, M.; Ishida, T. N., Y. Honda, O. Kitao, H. Nakai, M. Klene, X. Li, J. E. Knox,

H. P. Hratchian, J. B. Cross, V. Bakken, C.; Adamo, J. J., R. Gomperts, R. E. Stratmann,

O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, P. Y.; Ayala, K. M., G.

A. Voth, P. Salvador, J. J. Dannenberg, V. G. Zakrzewski, S. Dapprich, A. D. Daniels,

M. C.; Strain, O. F., D. K. Malick, A. D. Rabuck, K. Raghavachari, J. B. Foresman, J. V.

Ortiz, Q. Cui, A. G. Baboul, S.; Clifford, J. C., B. B. Stefanov, G. Liu, A. Liashenko, P.

Piskorz, I. Komaromi, R. L. Martin, D. J. Fox, T. Keith, M.; A. Al-Laham, C. Y. P., A.

- 214 -

Nanayakkara, M. Challacombe, P. M. W. Gill, B. Johnson, W. Chen, M. W. Wong, C.;

Gonzalez, a. J. A. P., Gaussian, Inc., Wallingford CT, 2004.

229. Perepichka, I. F.; Popov, A. F.; Orekhova, T. V.; Bryce, M. R.; Vdovichenko, A.

N.; Batsanov, A. S.; Goldenberg, L. M.; Howard, J. A. K.; Sokolov, N. I.; Megson, J. L.,

Electron acceptors of the fluorene series .5. Intramolecular charge transfer in nitro-

substituted 9-(aminomethylene)fluorenes. J Chem Soc Perk T 2 1996, (11), 2453-2469.

230. Gittins, D. I.; Bethell, D.; Schiffrin, D. J.; Nichols, R. J., A nanometre-scale

electronic switch consisting of a metal cluster and redox-addressable groups. Nature

2000, 408 (6808), 67-69.

231. Ashwell, G. J.; Tyrrell, W. D.; Whittam, A. J., Molecular rectification: Self-

assembled monolayers in which donor-(pi-bridge)-acceptor moieties are centrally

located and symmetrically coupled to both gold electrodes. J Am Chem Soc 2004, 126

(22), 7102-7110.

232. Flink, S.; van Veggel, F. C. J. M.; Reinhoudt, D. N., Sensor Functionalities in

Self-Assembled Monolayers. Adv Mater 2000, 12 (18), 1315-1328.

233. Moore, A. J.; Goldenberg, L. M.; Bryce, M. R.; Petty, M. C.; Monkman, A. P.;

Marenco, C.; Yarwood, J.; Joyce, M. J.; Port, S. N., Cation Recognition by Self-

Assembled Layers of Novel Crown-Annelated Tetrathiafulvalenes. Adv Mater 1998, 10

(5), 395-398.

234. Liu, H.; Liu, S.; Echegoyen, L., Remarkably stable self-assembled monolayers of

new crown-ether annelated tetrathiafulvalene derivatives and their cation recognition

properties[dagger]. Chem Commun 1999, (16), 1493-1494.

235. Liu, S.-G.; Liu, H.; Bandyopadhyay, K.; Gao, Z.; Echegoyen, L., Dithia-Crown-

Annelated Tetrathiafulvalene Disulfides:  Synthesis, Electrochemistry, Self-Assembled

Films, and Metal Ion Recognition. The Journal of Organic Chemistry 2000, 65 (11),

3292-3298.

236. Cooke, G., Electrochemical and Photochemical Control of Host–Guest

Complexation at Surfaces. Angewandte Chemie International Edition 2003, 42 (40),

4860-4870.

- 215 -

237. Hatton, R. A.; Willis, M. R.; Chesters, M. A.; Briggs, D., A robust ultrathin,

transparent gold electrode tailored for hole injection into organic light-emitting diodes. J

Mater Chem 2003, 13 (4), 722-726.

238. Campbell, I. H.; Kress, J. D.; Martin, R. L.; Smith, D. L.; Barashkov, N. N.;

Ferraris, J. P., Controlling charge injection in organic electronic devices using self-

assembled monolayers. Appl Phys Lett 1997, 71 (24), 3528-3530.

239. Yamada, H.; Imahori, H.; Nishimura, Y.; Yamazaki, I.; Fukuzumi, S.,

Enhancement of Photocurrent Generation by ITO Electrodes Modified Chemically with

Self-Assembled Monolayers of Porphyrin–Fullerene Dyads. Adv Mater 2002, 14 (12),

892-895.

240. Imahori, H.; Norieda, H.; Yamada, H.; Nishimura, Y.; Yamazaki, I.; Sakata, Y.;

Fukuzumi, S., Light-Harvesting and Photocurrent Generation by Gold Electrodes

Modified with Mixed Self-Assembled Monolayers of Boron−Dipyrrin and

Ferrocene−Porphyrin−Fullerene Triad. J Am Chem Soc 2000, 123 (1), 100-110.

241. Chidsey, C. E. D.; Bertozzi, C. R.; Putvinski, T. M.; Mujsce, A. M.,

Coadsorption of ferrocene-terminated and unsubstituted alkanethiols on gold:

electroactive self-assembled monolayers. J Am Chem Soc 1990, 112 (11), 4301-4306.

242. CHIDSEY, C. E. D., Free Energy and Temperature Dependence of Electron

Transfer at the Metal-Electrolyte Interface. Science 1991, 251 (4996), 919-922.

243. Liu, H. Y.; Liu, S. G.; Echegoyen, L., Remarkably stable self-assembled

monolayers of new crown-ether annelated tetrathiafulvalene derivatives and their cation

recognition properties. Chem Commun 1999, (16), 1493-1494.

244. Yip, C. M.; Ward, M. D., Self-Assembled Monolayers with Charge-Transfer

Groups: n-Mercaptoalkyl Tetrathiafulvalenecarboxylate on Gold. Langmuir 1994, 10

(2), 549-556.

245. Nakai, H.; Yoshihara, M.; Fujihara, H., New Electroactive Tetrathiafulvalene-

Derivatized Gold Nanoparticles and Their Remarkably Stable Nanoparticle Films on

Electrodes. Langmuir 1999, 15 (25), 8574-8576.

246. Yuge, R.; Miyazaki, A.; Enoki, T.; Tamada, K.; Nakamura, F.; Hara, M.,

Fabrication of TTF−TCNQ Charge-Transfer Complex Self-Assembled Monolayers: 

- 216 -

Comparison between the Coadsorption Method and the Layer-by-Layer Adsorption

Method. The Journal of Physical Chemistry B 2002, 106 (27), 6894-6901.

247. Trippe, G.; Ocafrain, M.; Besbes, M.; Monroche, V.; Lyskawa, J.; Le Derf, F.;

Salle, M.; Becher, J.; Colonna, B.; Echegoyen, L., Self-assembled monolayers of a

tetrathiafulvalene-based redox-switchable ligand. New J Chem 2002, 26 (10), 1320-

1323.

248. Bryce, M. R.; Cooke, G.; Duclairoir, F. M. A.; John, P.; Perepichka, D. F.;

Polwart, N.; Rotello, V. M.; Fraser Stoddart, J.; Tseng, H.-R., Surface confined

pseudorotaxanes with electrochemically controllable complexation properties. J Mater

Chem 2003, 13 (9), 2111-2117.

249. Herranz, M. Á.; Yu, L.; Martín, N.; Echegoyen, L., Synthesis, Electrochemistry

and Self-Assembled Monolayers of Novel Tetrathiafulvalene (TTF) and π-Extended

TTF (exTTF) Disulfides. The Journal of Organic Chemistry 2003, 68 (22), 8379-8385.

250. Liedberg, B.; Yang, Z.; Engquist, I.; Wirde, M.; Gelius, U.; Gotz, G.; Bauerle, P.;

Rummel, R. M.; Ziegler, C.; Gopel, W., Self-Assembly of α-Functionalized

Terthiophenes on Gold. The Journal of Physical Chemistry B 1997, 101 (31), 5951-

5962.

251. de Boer, B.; Meng, H.; Perepichka, D. F.; Zheng, J.; Frank, M. M.; Chabal, Y. J.;

Bao, Z., Synthesis and Characterization of Conjugated Mono- and Dithiol Oligomers

and Characterization of Their Self-Assembled Monolayers. Langmuir 2003, 19 (10),

4272-4284.

252. Otsubo, T.; Aso, Y.; Takimiya, K., Functional oligothiophenes as advanced

molecular electronic materials. J Mater Chem 2002, 12 (9), 2565-2575.

253. Bong, D.; Tam, I.; Breslow, R., Oligothiophene Isocyanides for Platinum-Based

Molecular Electronic Applications. J Am Chem Soc 2004, 126 (38), 11796-11797.

254. Morita, T.; Kimura, S.; Kobayashi, S.; Imanishi, Y., Photocurrent Generation

under a Large Dipole Moment Formed by Self-Assembled Monolayers of Helical

Peptides Having an N-Ethylcarbazolyl Group. J Am Chem Soc 2000, 122 (12), 2850-

2859.

- 217 -

255. Imahori, H.; Nishimura, Y.; Norieda, H.; Karita, H.; Yamazaki, I.; Sakata, Y.;

Fukuzumi, S., Photoinduced energy transfer in mixed self-assembled monolayers of

pyrene and porphyrin. Chem Commun 2000, (8), 661-662.

256. Skulason, H.; Frisbie, C. D., Self-assembled monolayers with charge-transfer

functional groups: Immobilization of the electron donor TMPD and the electron acceptor

TCNQ. Langmuir 1998, 14 (20), 5834-5840.

257. Zak, J.; Yuan, H.; Ho, M.; Woo, L. K.; Porter, M. D., Thiol-derivatized

metalloporphyrins: monomolecular films for the electrocatalytic reduction of dioxygen

at gold electrodes. Langmuir 1993, 9 (11), 2772-2774.

258. Hutchison, J. E.; Postlethwaite, T. A.; Murray, R. W., Molecular films of thiol-

derivatized tetraphenylporphyrins on gold: film formation and electrocatalytic dioxygen

reduction. Langmuir 1993, 9 (11), 3277-3283.

259. Boeckl, M. S.; Bramblett, A. L.; Hauch, K. D.; Sasaki, T.; Ratner, B. D.; Rogers,

J. W., Self-Assembly of Tetraphenylporphyrin Monolayers on Gold Substrates.

Langmuir 2000, 16 (13), 5644-5653.

260. Yasseri, A. A.; Syomin, D.; Malinovskii, V. L.; Loewe, R. S.; Lindsey, J. S.;

Zaera, F.; Bocian, D. F., Characterization of Self-Assembled Monolayers of Porphyrins

Bearing Multiple Thiol-Derivatized Rigid-Rod Tethers. J Am Chem Soc 2004, 126 (38),

11944-11953.

261. Shi, X.; Caldwell, W. B.; Chen, K.; Mirkin, C. A., A well-defined surface-

confinable fullerene: monolayer self-assembly on Au(111). J Am Chem Soc 1994, 116

(25), 11598-11599.

262. Fibbioli, M.; Bandyopadhyay, K.; Liu, S.-G.; Echegoyen, L.; Enger, O.;

Diederich, F.; Buhlmann, P.; Pretsch, E., Redox-active self-assembled monolayers as

novel solid contacts for ion-selective electrodes. Chem Commun 2000, (5), 339-340.

263. Hoang, V. T.; Rogers, L. M.; D'Souza, F., Synthesis and formation of monolayer

self-assembly of thiol appended fullerenes and fullerene-ferrocene dyads on gold

electrode. Electrochem Commun 2002, 4 (1), 50-53.

264. Gu, T.; Whitesell, J. K.; Fox, M. A., Electrochemical Charging of a Fullerene-

Functionalized Self-Assembled Monolayer on Au(111). The Journal of Organic

Chemistry 2004, 69 (12), 4075-4080.

- 218 -

265. Kim, K.-S.; Kang, M.-S.; Ma, H.; Jen, A. K. Y., Highly Efficient Photocurrent

Generation from a Self-Assembled Monolayer Film of a Novel C60-Tethered 2,5-

Dithienylpyrrole Triad. Chem Mater 2004, 16 (24), 5058-5062.

266. Hayes, W. A.; Shannon, C., Electrochemistry of Surface-Confined Mixed

Monolayers of 4-Aminothiophenol and Thiophenol on Au. Langmuir 1996, 12 (15),

3688-3694.

267. Yousaf, M. N.; Chan, E. W. L.; Mrksich, M., The Kinetic Order of an Interfacial

Diels–Alder Reaction Depends on the Environment of the Immobilized Dienophile.

Angewandte Chemie International Edition 2000, 39 (11), 1943-1946.

268. Kwan, W. S. V.; Atanasoska, L.; Miller, L. L., Oligoimide monolayers

covalently attached to gold. Langmuir 1991, 7 (7), 1419-1425.

269. Revell, D. J.; Chambrier, I.; Cook, M. J.; Russell, D. A., Formation and

spectroscopic characterisation of self-assembled phthalocyanine monolayers. J Mater

Chem 2000, 10 (1), 31-37.

270. De Long, H. C.; Buttry, D. A., Ionic interactions play a major role in determining

the electrochemical behavior of self-assembling viologen monolayers. Langmuir 1990, 6

(7), 1319-1322.

271. Creager, S. E.; Collard, D. M.; Fox, M. A., Mediated electron transfer by a

surfactant viologen bound to octadecanethiol on gold. Langmuir 1990, 6 (10), 1617-

1620.

272. Gittins, D. I.; Bethell, D.; Nichols, R. J.; Schiffrin, D. J., Redox-Connected

Multilayers of Discrete Gold Particles: A Novel Electroactive Nanomaterial. Adv Mater

1999, 11 (9), 737-740.

273. Skulason, H.; Frisbie, C. D., Direct Detection by Atomic Force Microscopy of

Single Bond Forces Associated with the Rupture of Discrete Charge-Transfer

Complexes. J Am Chem Soc 2002, 124 (50), 15125-15133.

274. Ho, G.; Heath, J. R.; Kondratenko, M.; Perepichka, D. F.; Arseneault, K.;

Pézolet, M.; Bryce, M. R., The First Studies of a Tetrathiafulvalene-σ-Acceptor

Molecular Rectifier. Chemistry – A European Journal 2005, 11 (10), 2914-2922.

275. Perepichka, I. F.; Popov, A. F.; Orekhova, T. V.; Bryce, M. R.; Andrievskii, A.

M.; Batsanov, A. S.; Howard, J. A. K.; Sokolov, N. I., Electron acceptors of the fluorene

- 219 -

series. 10. Novel acceptors containing butylsulfanyl, butylsulfinyl, and butylsulfonyl

substituents: Synthesis, cyclic voltammetry, charge-transfer complexation with

anthracene in solution, and X-ray crystal structures of two tetrathiafulvalene complexes.

J Org Chem 2000, 65 (10), 3053-3063.

276. Bryce, M. R.; Cooke, G.; Duclairoir, F. M. A.; John, P.; Perepichka, D. F.;

Polwart, N.; Rotello, V. M.; Stoddart, J. F.; Tseng, H. R., Surface confined

pseudorotaxanes with electrochemically controllable complexation properties. J Mater

Chem 2003, 13 (9), 2111-2117.

277. Herranz, M. A.; Yu, L.; Martin, N.; Echegoyen, L., Synthesis, electrochemistry

and self-assembled monolayers of novel tetrathiafulvalene (TTF) and pi-extended TTF

(exTTF) disulfides. J Org Chem 2003, 68 (22), 8379-8385.

278. Perepichka, D. F.; Kondratenko, M.; Bryce, M. R., Self-assembly and multistage

redox chemistry of strong electron acceptors on metal surfaces: Polynitrofluorenes on

gold and platinum. Langmuir 2005, 21 (19), 8824-8831.

279. Beisswenger, T. L., G.; Landgraf, K. F. Oestreich, E.; Rischer, M. November 30,

1999.

280. Shepherd, J. L.; Kell, A.; Chung, E.; Sinclar, C. W.; Workentin, M. S.; Bizzotto,

D., Selective reductive desorption of a SAM-coated gold electrode revealed using

fluorescence microscopy. J Am Chem Soc 2004, 126 (26), 8329-8335.

281. Goldenberg, L. M.; Williams, G.; Bryce, M. R.; Monkman, A. P.; Petty, M. C.;

Hirsch, A.; Soi, A., Electrochemical Studies on Langmuir-Blodgett-Films of 1-Tert-

Butyl-1,9-Dihydrofullerene-60. J Chem Soc Chem Comm 1993, (17), 1310-1312.

282. Jakubowicz, A.; Jia, H.; Wallace, R. M.; Gnade, B. E., Adsorption kinetics of p-

nitrobenzenethiol self-assembled monolayers on a gold surface. Langmuir 2005, 21 (3),

950-955.

283. Moore, A. J.; Bryce, M. R.; Batsanov, A. S.; Cole, J. C.; Howard, J. A. K.,

Functionalised Trimethyltetrathiafulvalene (TriMe-TTF) Derivatives via Reactions of

Trimethyltetrathiafulvalenyllithium with Electrophiles: X-ray Crystal Structures of

Benzoyl-TriMe-TTF and Benzoylthio-TriMe-TTF. Synthesis 1995, 1995 (06), 675,682.

284. Perepichka, D. F.; Bryce, M. R.; Perepichka, I. F.; Lyubchik, S. B.; Christensen,

C. A.; Godbert, N.; Batsanov, A. S.; Levillain, E.; McInnes, E. J. L.; Zhao, J. P., A (pi-

- 220 -

extended tetrathiafulvalene)-fluorene conjugate. Unusual electrochemistry and charge

transfer properties: The first observation of a covalent D2+-sigma-A(center dot-) redox

state. J Am Chem Soc 2002, 124 (47), 14227-14238.

285. Kricheldorf, H. R.; Kaschig, J., Depsipeptides and Polyesteramides .1. Syntheses

of Omega-Hydroxy-Carboxylic and Mercaptocarboxylic Acids Tert-Butyl Esters. Justus

Liebigs Annalen Der Chemie 1976, (5), 882-890.

286. Nuzzo, R. G.; Fusco, F. A.; Allara, D. L., Spontaneously organized molecular

assemblies. 3. Preparation and properties of solution adsorbed monolayers of organic

disulfides on gold surfaces. J Am Chem Soc 1987, 109 (8), 2358-2368.

287. Pfammatter, M. J.; Siljegovic, V.; Darbre, T.; Keese, R., Synthesis of omega-

substituted alkanethiols and (bromomethyl)methylthiomalonates. Helv Chim Acta 2001,

84 (3), 678-689.

288. Stuhr-Hansen, N., The tert-butyl moiety - A base resistent thiol protecting group

smoothly replaced by the labile acetyl moiety. Synthetic Commun 2003, 33 (4), 641-646.

289. Blaszczyk, A.; Elbing, M.; Mayor, M., Bromine catalyzed conversion of S-tert-

butyl groups into versatile and, for self-assembly processes accessible, acetyl-protected

thiols. Org Biomol Chem 2004, 2 (19), 2722-2724.

290. Gazal, S.; Gellerman, G.; Glukhov, E.; Gilon, C., Synthesis of novel protected N-

alpha(omega-thioalkyl) amino acid building units and their incorporation in backbone

cyclic disulfide and thioetheric bridged peptides. J Pept Res 2001, 58 (6), 527-539.

291. Balfe, M. P.; Kenyon, J.; Searle, C. E., 645. Fission of the alkyl-oxygen bond in

carbinols and of the alkyl-sulphur bond in sulphides and sulphones. Journal of the

Chemical Society (Resumed) 1950, 3309-3312.

292. Kubatkin, S.; Danilov, A.; Hjort, M.; Cornil, J.; Bredas, J. L.; Stuhr-Hansen, N.;

Hedegard, P.; Bjornholm, T., Single-electron transistor of a single organic molecule with

access to several redox states. Nature 2003, 425 (6959), 698-701.

293. Ribera, E.; Rovira, C.; Veciana, J.; Tarres, J.; Canadell, E.; Rousseau, R.; Molins,

E.; Mas, M.; Schoeffel, J. P.; Pouget, J. P.; Morgado, J.; Henriques, R. T.; Almeida, M.,

The [(DT-TTF)(2)M(mnt)(2)] family of radical ion salts: From a spin ladder to

delocalised conduction electrons that interact with localised magnetic moments. Chem-

Eur J 1999, 5 (7), 2025-2039.

- 221 -

294. Ie, Y.; Hirose, T.; Nakamura, H.; Kiguchi, M.; Takagi, N.; Kawai, M.; Aso, Y.,

Nature of Electron Transport by Pyridine-Based Tripodal Anchors: Potential for Robust

and Conductive Single-Molecule Junctions with Gold Electrodes. J Am Chem Soc 2011,

133 (9), 3014-3022.

295. Jiang, P.; Morales, G. M.; You, W.; Yu, L. P., Synthesis of diode molecules and

their sequential assembly to control electron transport. Angew Chem Int Edit 2004, 43

(34), 4471-4475.

296. Morales, G. M.; Jiang, P.; Yuan, S. W.; Lee, Y. G.; Sanchez, A.; You, W.; Yu, L.

P., Inversion of the rectifying effect in diblock molecular diodes by protonation. J Am

Chem Soc 2005, 127 (30), 10456-10457.

297. Ashwell, G. J.; Urasinska-Wojcik, B.; Phillips, L. J., In Situ Stepwise Synthesis

of Functional Multijunction Molecular Wires on Gold Electrodes and Gold

Nanoparticles. Angew Chem Int Edit 2010, 49 (20), 3508-3512.

298. Okazaki, N.; Sambles, J. R.; Jory, M. J.; Ashwell, G. J., Molecular rectification at

8 K in an Au/C(16)H(33)Q-3CNQ LB film/Au structure. Appl Phys Lett 2002, 81 (12),

2300-2302.

299. Ashwell, G. J.; Gandolfo, D. S., Molecular rectification: dipole reversal in a

cationic donor-(pi-bridge)-acceptor dye. J Mater Chem 2002, 12 (3), 411-415.

300. Nalwa, H. S., Handbook of organic conductive molecules and polymers. Wiley:

Chichester ; New York, 1997.

301. Groenendaal, B. L.; Jonas, F.; Freitag, D.; Pielartzik, H.; Reynolds, J. R.,

Poly(3,4-ethylenedioxythiophene) and its derivatives: Past, present, and future. Adv

Mater 2000, 12 (7), 481-494.

302. Roncali, J.; Blanchard, P.; Frere, P., 3,4-Ethylenedioxythiophene (EDOT) as a

versatile building block for advanced functional p-conjugated systems. J Mater Chem

2005, 15 (16), 1589-1610.

303. Kirchmeyer, S.; Reuter, K., Scientific importance, properties and growing

applications of poly(3,4-ethylenedioxythiophene) (vol 15, pg 2077, 2005). J Mater

Chem 2005, 15 (23), 2338-2338.

- 222 -

304. Wheland, R. C.; Martin, E. L., Synthesis of substituted 7,7,8,8-

tetracyanoquinodimethanes. The Journal of Organic Chemistry 1975, 40 (21), 3101-

3109.

305. Uno, M.; Seto, K.; Masuda, M.; Ueda, W.; Takahashi, S., A new route to

phenylenedimalononitrile and the analogues using palladium-catalyzed carbon-carbon

bond formation. Tetrahedron Lett 1985, 26 (12), 1553-1556.

306. Aumüller, A.; Hünig, S., Mehrstufige, reversible Redoxsysteme, XXXVI.

11,11,12,12-Tetracyan-9,10-anthrachinodimethan (TCNAQ) und seine Derivate:

Synthese und Redoxeigenschaften. Liebigs Ann Chem 1984, 1984 (3), 618-621.

307. Hertler, W. R.; Hartzler, H. D.; Acker, D. S.; Benson, R. E., Substituted

Quinodimethans. III. Displacement Reactions of 7,7,8,8-Tetracyanoquinodimethan. J

Am Chem Soc 1962, 84 (17), 3387-3393.

308. Bespalov, B. P.; Getmanova, E. V.; Titov, V. V., Reaction du tetracyano-7,7,8,8

quinodimethane avec les phenois, pyrroles et indoles. Tetrahedron Lett 1976, 17 (22),

1867-1870.

309. Zaidi, N. A.; Bryce, M. R.; Cross, G. H., Synthesis and optical properties of new

tricyano-p-quinodimethane dyes: molecular and polymeric systems. Tetrahedron Lett

2000, 41 (23), 4645-4649.

310. Boldt, P.; Bourhill, G.; Brauchle, C.; Jim, Y.; Kammler, R.; Muller, C.; Rase, J.;

Wichern, J., Tricyanoquinodimethane derivatives with extremely large second-order

optical nonlinearities. Chem Commun 1996, (6), 793-795.

311. Turbiez, M.; Frère, P.; Roncali, J., Stable and Soluble Oligo(3,4-

ethylenedioxythiophene)s End-Capped with Alkyl Chains. The Journal of Organic

Chemistry 2003, 68 (13), 5357-5360.

312. Akoudad, S.; Roncali, J., Electrochemical synthesis of poly(3,4-

ethylenedioxythiophene) from a dimer precursor. Synthetic Met 1998, 93 (2), 111-114.

313. Mohanakrishnan, A. K.; Hucke, A.; Lyon, M. A.; Lakshmikantham, M. V.; Cava,

M. P., Functionalization of 3,4-ethylenedioxythiophene. Tetrahedron 1999, 55 (40),

11745-11754.

314. Sotzing, G. A.; Reynolds, J. R.; Steel, P. J., Poly(3,4-ethylenedioxythiophene)

(PEDOT) prepared via electrochemical polymerization of EDOT, 2,2′-Bis(3,4-

- 223 -

ethylenedioxythiophene) (BiEDOT), and their TMS derivatives. Adv Mater 1997, 9 (10),

795-798.

315. Kagan, J.; Arora, S. K., The Synthesis of Alpha-Thiophene Oligomers by

Oxidative Coupling of 2-Lithiothiophenes. Heterocycles 1983, 20 (10), 1937-1940.

316. Metzger, R. M.; Heimer, N. E.; Ashwell, G. J., Crystal and Molecular-Structure

and Properties of Picolyltricyanoquinodimethan, the Zwitterionic Donor-Pi-Acceptor

Adduct between Li+Tcnq- and 1,2-Dimethylpyridinium Iodide. Mol Cryst Liq Cryst

1984, 107 (1-2), 133-149.

317. M. Turbiez, P. F., J. Roncali., J. Org. Chem. 2003, 68, 5357-5360.

318. Kondratenko, M.; Moiseev, A. G.; Perepichka, D. F., New stable donor-acceptor

dyads for molecular electronics. J Mater Chem 2011, 21 (5), 1470-1478.

319. Halls, M. D.; Velkovski, J.; Schlegel, H. B., Harmonic frequency scaling factors

for Hartree-Fock, S-VWN, B-LYP, B3-LYP, B3-PW91 and MP2 with the Sadlej pVTZ

electric property basis set. Theor Chem Acc 2001, 105 (6), 413-421.

320. Vilan, A.; Yaffe, O.; Biller, A.; Salomon, A.; Kahn, A.; Cahen, D., Molecules on

Si: Electronics with Chemistry. Adv Mater 2010, 22 (2), 140-159.

321. Salomon, A.; Boecking, T.; Seitz, O.; Markus, T.; Amy, F.; Chan, C.; Zhao, W.;

Cahen, D.; Kahn, A., What is the Barrier for Tunneling Through Alkyl Monolayers?

Results from n- and p-Si–Alkyl/Hg Junctions. Adv Mater 2007, 19 (3), 445-450.

322. Aswal, D. K.; Lenfant, S.; Guerin, D.; Yakhmi, J. V.; Vuillaume, D., Self

assembled monolayers on silicon for molecular electronics. Anal Chim Acta 2006, 568

(1-2), 84-108.

323. Mathieu Turbiez, D., Pierre Frère, Magali Allain, Christine Videlot, Jörg

Ackermann, Jean Roncali., Chem. Eur. J. 2005, 11 (12), 3742-3752.

324. Bhosale, S. V.; Jani, C. H.; Langford, S. J., Chemistry of naphthalene diimides.

Chem Soc Rev 2008, 37 (2), 331-342.

325. Levanon, H.; Galili, T.; Regev, A.; Wiederrecht, G. P.; Svec, W. A.;

Wasielewski, M. R., Determination of the energy levels of radical pair states in

photosynthetic models oriented in liquid crystals with time-resolved electron

paramagnetic resonance. J Am Chem Soc 1998, 120 (25), 6366-6373.

- 224 -

326. Hasharoni, K.; Levanon, H.; Greenfield, S. R.; Gosztola, D. J.; Svec, W. A.;

Wasielewski, M. R., Radical pair and triplet state dynamics of a photosynthetic reaction-

center model embedded in isotropic media and liquid crystals. J Am Chem Soc 1996,

118 (42), 10228-10235.

327. Debreczeny, M. P.; Svec, W. A.; Marsh, E. M.; Wasielewski, M. R.,

Femtosecond optical control of charge shift within electron donor-acceptor arrays: An

approach to molecular switches. J Am Chem Soc 1996, 118 (34), 8174-8175.

328. Kelley, R. F.; Tauber, M. J.; Wasielewski, M. R., Intramolecular electron transfer

through the 20-position of a chlorophyll a derivative: An unexpectedly efficient conduit

for charge transport. J Am Chem Soc 2006, 128 (14), 4779-4791.

329. Mi, Q. X.; Chernick, E. T.; McCamant, D. W.; Weiss, E. A.; Ratner, M. A.;

Wasielewski, M. R., Spin dynamics of photogenerated triradicals in fixed distance

electron donor-chromophore-acceptor-TEMPO molecules. J Phys Chem A 2006, 110

(23), 7323-7333.

330. Colvin, M. T.; Giacobbe, E. M.; Cohen, B.; Miura, T.; Scott, A. M.;

Wasielewski, M. R., Competitive Electron Transfer and Enhanced Intersystem Crossing

in Photoexcited Covalent TEMPO-Perylene-3,4:9,10-bis(dicarboximide) Dyads:

Unusual Spin Polarization Resulting from the Radical-Triplet Interaction. J Phys Chem

A 2010, 114 (4), 1741-1748.

331. Katz, H. E.; Lovinger, A. J.; Johnson, J.; Kloc, C.; Siegrist, T.; Li, W.; Lin, Y.

Y.; Dodabalapur, A., A soluble and air-stable organic semiconductor with high electron

mobility. Nature 2000, 404 (6777), 478-481.

332. Jones, B. A.; Facchetti, A.; Marks, T. J.; Wasielewski, M. R., Cyanonaphthalene

diimide semiconductors for air-stable, flexible, and optically transparent n-channel field-

effect transistors. Chem Mater 2007, 19 (11), 2703-2705.

333. Shukla, D.; Nelson, S. F.; Freeman, D. C.; Rajeswaran, M.; Ahearn, W. G.;

Meyer, D. M.; Carey, J. T., Thin-Film Morphology Control in Naphthalene-Diimide-

Based Semiconductors: High Mobility n-Type Semiconductor for Organic Thin-Film

Transistors. Chem Mater 2008, 20 (24), 7486-7491.

- 225 -

334. Jones, B. A.; Facchetti, A.; Wasielewski, M. R.; Marks, T. J., Tuning orbital

energetics in arylene diimide semiconductors. Materials design for ambient stability of

n-type charge transport. J Am Chem Soc 2007, 129 (49), 15259-15278.

335. Wang, Y. F.; Chen, Y. L.; Li, R. J.; Wang, S. Q.; Su, W.; Ma, P.; Wasielewski,

M. R.; Li, X. Y.; Jiang, J. Z., Amphiphilic perylenetretracarboxyl diimide dimer and its

application in field effect transistor. Langmuir 2007, 23 (10), 5836-5842.

336. Grzegorczyk, W. J.; Ganesan, P.; Savenije, T. J.; van Bavell, S.; Loos, J.;

Sudholter, E. J. R.; Siebbeles, L. D. A.; Zuilhof, H., Photoconductance of Bulk

Heterojunctions with Tunable Nanomorphology Consisting of P3HT and Naphthalene

Diimide Siloxane Oligomers. J Phys Chem C 2009, 113 (18), 7863-7869.

337. Angadi, M. A.; Gosztola, D.; Wasielewski, M. R., Characterization of

photovoltaic cells using poly(phenylenevinylene) doped with perylenediimide electron

acceptors. J Appl Phys 1998, 83 (11), 6187-6189.

338. Ortiz, R. P.; Herrera, H.; Blanco, R.; Huang, H.; Facchetti, A.; Marks, T. J.;

Zheng, Y.; Segura, J. L., Organic n-Channel Field-Effect Transistors Based on

Arylenediimide-Thiophene Derivatives. J Am Chem Soc 2010, 132 (24), 8440-8452.

339. Yan, H.; Chen, Z. H.; Zheng, Y.; Newman, C.; Quinn, J. R.; Dotz, F.; Kastler,

M.; Facchetti, A., A high-mobility electron-transporting polymer for printed transistors.

Nature 2009, 457 (7230), 679-686.

340. Shumate, W. J.; Mattern, D. L.; Jaiswal, A.; Dixon, D. A.; White, T. R.; Burgess,

J.; Honciuc, A.; Metzger, R. M., Spectroscopy and rectification of three donor-sigma-

acceptor compounds, consisting of a one-electron donor (pyrene or ferrocene), a one-

electron acceptor (perylenebisimide), and a C19 swallowtail. J Phys Chem B 2006, 110

(23), 11146-11159.

341. Jones, B. A.; Ahrens, M. J.; Yoon, M. H.; Facchetti, A.; Marks, T. J.;

Wasielewski, M. R., High-mobility air-stable n-type semiconductors with processing

versatility: Dicyanoperylene-3,4 : 9,10-bis(dicarboximides). Angew Chem Int Edit 2004,

43 (46), 6363-6366.

342. Weitz, R. T.; Amsharov, K.; Zschieschang, U.; Villas, E. B.; Goswami, D. K.;

Burghard, M.; Dosch, H.; Jansen, M.; Kern, K.; Klauk, H., Organic n-channel transistors

- 226 -

based on core-cyanated perylene carboxylic diimide derivatives. J Am Chem Soc 2008,

130 (14), 4637-4645.

343. Metzger, R. M., All about (N-hexadecylquinolin-4-ium-1-

yl)methylidenetricyanoquinodimethanide, a unimolecular rectifier of electrical current. J

Mater Chem 2000, 10 (1), 55-62.

344. Turbiez, M.; Frere, P.; Roncali, J., Stable and soluble oligo(3,4-

ethylenedioxythiophene)s end-capped with alkyl chains. J Org Chem 2003, 68 (13),

5357-5360.

345. Apperloo, J. J.; Groenendaal, L.; Verheyen, H.; Jayakannan, M.; Janssen, R. A.

J.; Dkhissi, A.; Beljonne, D.; Lazzaroni, R.; Bredas, J. L., Optical and redox properties

of a series of 3,4-ethylenedioxythiophene oligomers. Chem-Eur J 2002, 8 (10), 2384-

2396.

346. Barros, T. C.; Brochsztain, S.; Toscano, V. G.; Berci, P.; Politi, M. J.,

Photophysical characterization of a 1,4,5,8-naphthalenediimide derivative. J Photoch

Photobio A 1997, 111 (1-3), 97-104.

347. Trivedi, D. R.; Fujiki, Y.; Goto, Y.; Fujita, N.; Shinkai, S.; Sada, K., A naked-

eye colorimetric indicator to discriminate aromatic compounds by solid-state charge-

transfer complexation. Chem Lett 2008, 37 (5), 550-551.

348. Posokhov, Y.; Alp, S.; Koz, B.; Dilgin, Y.; Icli, S., Photophysical properties and

electrochemistry of the N,N '-bis-n-butyl derivative of naphthalene diimide. Turk J

Chem 2004, 28 (4), 415-424.

349. Wasserberg, D.; Marsal, P.; Meskers, S. C. J.; Janssen, R. A. J.; Beljonne, D.,

Phosphorescence and triplet state energies of oligothiophenes. J Phys Chem B 2005, 109

(10), 4410-4415.

350. Andric, G.; Boas, J. F.; Bond, A. M.; Fallon, G. D.; Ghiggino, K. P.; Hogan, C.

F.; Hutchison, J. A.; Lee, M. A. P.; Langford, S. J.; Pilbrow, J. R.; Troup, G. J.;

Woodward, C. P., Spectroscopy of naphthalene diimides and their anion radicals. Aust J

Chem 2004, 57 (10), 1011-1019.

351. Hostetler, M. J.; Stokes, J. J.; Murray, R. W., Infrared spectroscopy of three-

dimensional self-assembled monolayers: N-alkanethiolate monolayers on gold cluster

compounds. Langmuir 1996, 12 (15), 3604-3612.

- 227 -

352. Karakouz, T.; Tesler, A. B.; Bendikov, T. A.; Vaskevich, A.; Rubinstein, I.,

Highly Stable Localized Plasmon Transducers Obtained by Thermal Embedding of Gold

Island Films on Glass. Adv Mater 2008, 20 (20), 3893-3899.

353. Misek, J.; Jentzsch, A. V.; Sakurai, S. I.; Emery, D.; Mareda, J.; Matile, S., A

Chiral and Colorful Redox Switch: Enhanced pi Acidity in Action. Angew Chem Int Edit

2010, 49 (42), 7680-7683.

354. Sakai, N.; Mareda, J.; Vauthey, E.; Matile, S., Core-substituted

naphthalenediimides. Chem Commun 2010, 46 (24), 4225-4237.

355. Dawson, R. E.; Hennig, A.; Weimann, D. P.; Emery, D.; Ravikumar, V.;

Montenegro, J.; Takeuchi, T.; Gabutti, S.; Mayor, M.; Mareda, J.; Schalley, C. A.;

Matile, S., Experimental evidence for the functional relevance of anion-pi interactions.

Nat Chem 2010, 2 (7), 533-538.

356. Błaszczyk, A.; Fischer, M.; von Hänisch, C.; Mayor, M., Synthesis, Structure,

and Optical Properties of Terminally Sulfur-Functionalized Core-Substituted

Naphthalene-Bisimide Dyes. Helv Chim Acta 2006, 89 (9), 1986-2005.

357. Röger, C.; Würthner, F., Core-Tetrasubstituted Naphthalene Diimides: 

Synthesis, Optical Properties, and Redox Characteristics. The Journal of Organic

Chemistry 2007, 72 (21), 8070-8075.

358. Hayes, R. T.; Wasielewski, M. R.; Gosztola, D., Ultrafast Photoswitched Charge

Transmission through the Bridge Molecule in a Donor−Bridge−Acceptor System. J Am

Chem Soc 2000, 122 (23), 5563-5567.

359. Langhals, H., Cyclic Carboxylic Imide Structures as Structure Elements of High-

Stability - Novel Developments in Perylene Dye Chemistry. Heterocycles 1995, 40 (1),

477-500.

360. Troster, H., Studies on the Protonation of Perlene-3,4,9,10-Tetracarboxylic Acid

Alkali Salts. Dyes Pigments 1983, 4 (3), 171-177.

361. Nagao, Y.; Ishikawa, N.; Tanabe, Y.; Misono, T., Synthesis of Unsymmetrical

Perylenebis(Dicarboximide) Derivatives. Chem Lett 1979, (2), 151-154.

362. Nagao, Y.; Tanabe, Y.; Misono, T., Synthesis and Reactions of

Perylenecarboxylic Acid-Derivatives .3. Reactions of 3,4 - 9,10-Perylenetetracarboxylic

Dianhydride with Alkylamines. Nippon Kagaku Kaishi 1979, (4), 528-534.

- 228 -

363. Xue, C. M.; Sun, R. K.; Annab, R.; Abadi, D.; Jin, S., Perylene monoanhydride

diester: a versatile intermediate for the synthesis of unsymmetrically substituted

perylene tetracarboxylic derivatives. Tetrahedron Lett 2009, 50 (8), 853-856.

- 229 -

Appendix

Figure A1: TGA analysis of the dyad 6.7.

Figure A2. Structures of the proposed series of Donor-Acceptor dyads with sulfone

substituents.

- 230 -

Figure A3: Calculated molecular orbitals for the proposed series of dyads with sulfone

substituents.

We have started the work presented in the Chapter VI by exploring the

possibility of using PDI as an acceptor moiety in the dyads (dyad 6.12, Fig. A3).

However, asymmetrically disubstituted PDI is not accessible via a step-wise

condensation (as was done for NDI, Chapter VI) of PDA with two different primary

amines by controlling the stoichiometric ratio [359].

General procedure of the synthesis of asymmetrical PDI, suggests conversion of

the PTCDA into the mono-potassium salt 6.16, which later can be reacted with an amine

followed by acidification of the last carboxylate anion and reaction with the second

amine. (Scheme A1) [360]. The main disadvantage of this route is that during the

acidification of the solution the compound starts precipitating immediately after the pH

of the solution gets lower than 8-9, resulting in formation of mono-, di- or tri-acids and it

is very difficult to control the formation of desired product.

- 231 -

Scheme A1. ―Protection‖ of the one side of the PTCDA.

Another way to obtain the asymmetrically substituted PDI is by a partial

saponification of the symmetric PDI under strong basic conditions was first described by

Nagao [361-362] (Scheme A2). However, this procedure did not work for our specific

PDI derivative 6.17.

Scheme A2: Synthesis of perylene monoimide via ―half hydrolysis‖.

The solubility if the perylene intermediates could be increased by introducing

temporary or permanent functional groups that can be easily removed or will not limit

the performance of the desired dyad.

The intermediate with temporary functional groups was described in literature

not very long time ago [363]. A soluble perylene monoanhydride diester 6.19 (Scheme

A3) which has only one anhydride group that can be selectively reacted with amine; two

alkyl groups on the other side of the molecule provide necessary solubility in organic

solvents, which facilitates purification and characterization [363].

- 232 -

Scheme A3: Esterification of the PTCDA.

The condensation of the monoanhydride 6.19 with 2-ethylhexylamine or with p-

bromoaniline results in formation of monoimide diesters 6.20a (68% yield) and 6.20b

(48% yield) respectively (Scheme A4).

Scheme A4: Proposed synthetic procedure for of asymmetrically substituted PDI via

esterification.

Similarly to the NDI, the electron acceptor properties of the PDI can be tuned by

introducing electron-withdrawing substituents in the aromatic core. For example, alkyl

sulfone substituents in PTCDA can not only lower the LUMO but also increase the

solubility of the molecule which might improve the step-wise synthesis of asymmetric

PDI. Our DFT calculation of the di- and tetrasulfonesubstituted PDI (Fig. A3) showed

that already two sulfone substituents (6.13) dramatically lower the LUMO of the

perylene moiety and, thus, reduce the HLG to almost the same value as for 6.11 (0.9 and

0.84 eV respectively). Introduction of four sulfone substituents (6.14) leads to further

decrease of the HLG up to 0.7eV. However, significant distortion of the aromatic core of

the PDI may have influence on the overall acceptor properties of the moiety and packing

of the molecules in the monolayers

- 233 -

N,N-bis(2-ethylhexyl)-perylene-3,4,9,10-tetracarboxydiimide (6.17). 2-

ethylhexylamine (0.83 mL, 5.07 mmol) was added to a refluxing solution of 3,4,9,10-

perylenetetracarboxyldianhydride (0.493g, 1.26 mmol) in DMF (100mL), followed by

addition of zinc acetate (51.4 mg,0.234 mmol). The solution was refluxed at 120oC for a

few hours, followed by stirring at room temperature over the weekend. A concentrated

potassium hydroxide solution was added to the warm reaction mixture and the red

precipitate was filtered, washed with water, collected, and dried affording 6.17 (0.749g,

97%): m.p. 370-372C. 1H-NMR (400 MHz, CDCl3) 8.67 (m, 8H), 4.17-4.13(m, 4H),

1.97 (m, 4H, t, 2H), 1.42-1.32 (m, 16H), 0.97-0.88 (m, 12H).

Perylene-3,4,9,10-tetracaboxilic acid 3,4,9,10-trakishexyl ester (6.18) [363]. PTCDA

(1.03g, 2.63 mmol), KOH (0.80g, 14.3 mmol), and distilled water (13 mL) were added

to a 250-mL Erlenmeyer flask. The solution was stirred at 70oC for one hour until the

reagent dissolved. The pH was adjusted to 8-9 using 1M HCl solution, followed by

addition of Aliquat 336 (0.38 mL, 0.83 mmol) and KI (0.0665g, 0.39 mmol). The dark

red solution was stirred for 10 minutes and hexyl bromide (3 mL, 21.4 mmol) was

added. The solution was refluxed at 70oC for two hours. The mixture was extracted with

CH2Cl2 and washed with brine three times. The bright orange organic layer was dried

with magnesium sulfate, which was filtered out. The solution was concentrated and

methanol was added. The bright orange precipitate was filtered out, collected, and dried

in vacuum yielding 6.18 (1.36g, 68%): m.p. 185-186C 1H NMR (400 MHz, CDCl3)

8.31 (d, J =8.0 Hz, 4H), 8.05 (d, J = 8.0 Hz, 4H), 4.31 (t, J=6.8Hz, 8H), 1.80-1.76 (m,

8H), 1.44-1.33 (m, 24H), 0.90 (t, J=6.9 Hz); HRMS(ESI) calculated for C48H60O8Na

787.4180 found 787.4167; UV-vis (in CH2Cl2): 442.5 and 471 nm (λmax).

Perylene-3,4,9,10-tetracaboxilic acid 3,4-anhydride 9,10-dihexyl ester (6.19) [363].

The p-toluenesulfonic acid monohydrate (0.213g, 1.12 mmol) was added to a solution of

6.18 (0.830g, 1.08 mmol) in toluene (0.5 mL) and hexadecane (15 mL). The orange

reaction mixture was heated at 100oC for 6 hours. Hexanes was added to the solution

and the bright red solids were filtered out, collected, and dried under vacuum yielding

6.19 (0.597 g, 95%): Tdec>400C. 1H NMR (400 MHz, CDCl3) 8.62 (d, J = 7.6 Hz, 2H),

8.48 (s, 4H), 8.12 (d, J = 8.0 Hz, 2H), 4.34 (t, J= 6.4Hz, 4H), 1.82-1.36 (m, 16H), 0.91

(t, J=6.4 Hz, 6H); HRMS(ESI) calculated for C36H34O7Na 601.2202 found 601.2216;

UV-vis (in CH2Cl2): 476 and 506.5 nm (λmax).

N-(2-ethylhexyl)-perylene-3,4,9,10-tetracarboxylic-3,4-imide-9,10-dihexyl ester

(6.20a). 2-Ethylhexylamine (0.30mL, 1.83 mmol) was added to a flask with 6.19

(0.895g, 1.55 mmol) and DMF (20mL). The solution was refluxed at 120oC overnight.

The solvent was evaporated and redissolved in CH2Cl2. The mixture was purified on

silica gel by column chromatography using 40/1 (v/v) CH2Cl2/ethyl acetate as the eluent

affording 6.26a as a dark red powder (0.807g, 68%): m.p. 268-270C 1H NMR (400

MHz, CDCl3) 8.53 (d, J=8.0 Hz, 2H), 8.35 (dd, J=2.8, 8.0 Hz, 4H), 8.05 (d, J = 8.0 Hz,

2H), 4.35 (t, J= 6.8 Hz, 4H), 4.15-4.11(m, 2H),1.96 (m, 1H), 1.83-1.77 (m, 4H), 1.48-

1.32 (m, 20H), 0.97-0.87 (m, 12H); MS(EI) m/z 712 (100%).

- 234 -

N-(4-bromoaniline)-perylene-3,4,9,10-tetracarboxylic-3,4-imide-9,10-dihexyl ester

(6.20b). 4-bromoaniline (101.4 mg, 0.589 mmol) and 4-dimethylaminopyridine (20 mg,

0.164 mmol) was added to a solution of 6.19 (98.7 mg, 0.171 mmol) in DMF (10 mL),

followed by addition of zinc acetate (1 mg) and 4-dimethylaminopyridine (DMAP) (0.02

g, 0.16 mmol). The solution was refluxed at 120oC for three nights. Distilled water (200

mL) was added to the reaction mixture and the bright red solids were filtered off. The

solids were dissolved in CH2Cl2 and filtered through silica gel with CH2Cl2 as the eluent

yielding two fractions. The bright orange fraction was dried affording 6.26b as a red-

orange powder (0.058g, 48%): Tdec>400C. 1H NMR (400 MHz, CD2Cl2) 8.57 (d, J=8.0

Hz, 2H), 8.42 (t, J=8.4 Hz, 4H), 8.07 (d, J = 8.0 Hz, 2H), 7.73 (d, J=8.4 Hz, 2H), 7.29 (d,

J=7.6 Hz, 2H), 4.33 (t, J=7.6 Hz, 4H), 1.82-1.28 (m, 16H), 0.94 (t, 6H); MS(EI) m/z 733

(30%), 609 (100%); UV-vis (in CH2Cl2): 476 and 505.5 nm (λmax).

- 235 -

Author’s contribution

Chapter II

(Part of this Chapter was adapted with permission from: G. Ho, J. Heath, M.

Kondratenko, D. F. Perepichka, K. Arseneault, M. Pezolet, M. R. Bryce, The first

studies of a tetrathiafulvalene–σ–acceptor molecular rectifier, Chem. Eur. J. 2005, 11,

2914–2922. Copyrights 2005 Wiley-VCH Verlag GmbH&Co. KGaA, Weinheim)

M. Kondratenko synthesized the TTF--nitrofluorene dyad. In collaboration with K.

Arseneault and M. Pezolet from Laval University he deposited LB films on the gold and

Ge ATR crystal surfaces. He also built and validated mercury drop junction setup and

performed rectification study of the dyad.

K. Arseneault (group of M. Pezolet, Laval University) performed ATR IR spectroscopy

and orientation analysis of the molecules transferred on Ge crystal.

G. Ho and J. Heath (University of California, USA) performed the initial preparation

and ellipsometry study of the LB films and studied their rectification in Si/dyad/Ti

junctions.

Chapter III

(Part of this Chapter was adapted with permission from: D.F. Perepichka, M.

Kondratenko, M.R. Bryce, Self-Assembled Monolayers of Strong Electron Acceptors:

Polynitrofluorenes on Gold and Platinum, Langmuir 2005, 21, 8824–8831. Copyrights

2005 American Chemical Society)

M. Kondratenko performed synthesis, characterization and self-assembly study of all

compounds described in this Chapter III. He also performed rectification study of the

nitrofluorene based SAMs by mercury drop technique.

Chapter IV and V

Design, synthesis and all reported studies of the compounds presented in Chapter IV and

Chapter V was done by M. Kondratenko, except for EPR analysis which was performed

by A. Moiseev.

- 236 -

Chapter VI

(Part of this Chapter was adapted with permission from: M. Kondratenko, A. Moiseev,

D.F. Perepichka, New stable donor-acceptor Dyads for molecular electronics, J. Mater.

Chem. 2011, 21, 1470–1478. Copyrights 2011 The Royal Society of Chemistry)

M. Kondratenko synthesized and studied the n-EDOT-NDI dyads and together with

Jenny Liu carried out preliminary synthesis of nEDOT-PDI dyads.

- 237 -

List of publications

1. G. Ho, J. Heath, M. Kondratenko, D. F. Perepichka, K. Arseneault, M. Pezolet,

M. R. Bryce, The first studies of a tetrathiafulvalene–σ–acceptor molecular

rectifier, Chem. Eur. J. 2005, 11, 2914–2922.

2. D.F. Perepichka, M. Kondratenko, M.R. Bryce, Self-Assembled Monolayers of

Strong Electron Acceptors: Polynitrofluorenes on Gold and Platinum, Langmuir

2005, 21, 8824–8831.

3. Z.Wei, M.Kondratenko, D.F.Perepichka, L.H.Dao, Rectifying diodes from

symmetrically functionalized single wall carbon nanotubes, J. Am. Chem. Soc.

2006, 128, 3134-3135.

4. S. Clair, F. Variola, M. Kondratenko, P. Jedrzejowski, A. Nanci, F. Rosei, D.F.

Perepichka, Self-assembled monolayer of alkanephosphoric acid on nanotextured

Ti, J. Chem. Phys. 2008, 128, 144705.

5. J.A. Lipton-Duffin, J.A. Miwa, M. Kondratenko, F. Cicoira, B.G. Sumpter, V.

Meunier, D.F. Perepichka, F. Rosei, Step-by-step growth of aligned

polythiophene wires by surface-confined oligomerization, Proc. Nat. Acad. Sci.

USA. 2010, 107, 11200-11204.

6. M. Kondratenko, A. Moiseev, D.F. Perepichka, New stable donor-acceptor

Dyads for molecular electronics, J. Mater. Chem. 2011, 21, 1470–1478.


Recommended