205
UvA-DARE is a service provided by the library of the University of Amsterdam (https://dare.uva.nl) UvA-DARE (Digital Academic Repository) Intercomponent interactions and mobility in hydrogen-bonded rotaxanes Jagesar, D.C. Publication date 2010 Document Version Final published version Link to publication Citation for published version (APA): Jagesar, D. C. (2010). Intercomponent interactions and mobility in hydrogen-bonded rotaxanes. General rights It is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s), other than for strictly personal, individual use, unless the work is under an open content license (like Creative Commons). Disclaimer/Complaints regulations If you believe that digital publication of certain material infringes any of your rights or (privacy) interests, please let the Library know, stating your reasons. In case of a legitimate complaint, the Library will make the material inaccessible and/or remove it from the website. Please Ask the Library: https://uba.uva.nl/en/contact, or a letter to: Library of the University of Amsterdam, Secretariat, Singel 425, 1012 WP Amsterdam, The Netherlands. You will be contacted as soon as possible. Download date:28 Jun 2021

UvA-DARE (Digital Academic Repository) Intercomponent ...stereoisomerism and arises in molecules containing substituted double bonds (E-Z isomerism) and asymmetric carbon atoms (optical

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

  • UvA-DARE is a service provided by the library of the University of Amsterdam (https://dare.uva.nl)

    UvA-DARE (Digital Academic Repository)

    Intercomponent interactions and mobility in hydrogen-bonded rotaxanes

    Jagesar, D.C.

    Publication date2010Document VersionFinal published version

    Link to publication

    Citation for published version (APA):Jagesar, D. C. (2010). Intercomponent interactions and mobility in hydrogen-bondedrotaxanes.

    General rightsIt is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s)and/or copyright holder(s), other than for strictly personal, individual use, unless the work is under an opencontent license (like Creative Commons).

    Disclaimer/Complaints regulationsIf you believe that digital publication of certain material infringes any of your rights or (privacy) interests, pleaselet the Library know, stating your reasons. In case of a legitimate complaint, the Library will make the materialinaccessible and/or remove it from the website. Please Ask the Library: https://uba.uva.nl/en/contact, or a letterto: Library of the University of Amsterdam, Secretariat, Singel 425, 1012 WP Amsterdam, The Netherlands. Youwill be contacted as soon as possible.

    Download date:28 Jun 2021

    https://dare.uva.nl/personal/pure/en/publications/intercomponent-interactions-and-mobility-in-hydrogenbonded-rotaxanes(e07d24cb-7441-476b-befa-5e2693ee60e9).html

  • D h i r e d j C. J a g e s a r

    Intercomponent Interactions and Mobility

    in Hydrogen-Bonded Rotaxanes

    Inte

    rcom

    po

    ne

    nt In

    tera

    ction

    s an

    d M

    ob

    ility in

    Hy

    dro

    ge

    n-B

    on

    de

    d R

    ota

    xa

    ne

    s Dh

    ired

    j C. Ja

    ge

    sar

    ISBN 978-90-9025766-2

  • Intercomponent Interactions and Mobility in Hydrogen-Bonded

    Rotaxanes

    Dhiredj C. Jagesar

  • Intercomponent Interactions and Mobility in Hydrogen-Bonded

    Rotaxanes

    ACADEMISCH PROEFSCHRIFT

    Ter verkrijging van de graad van doctor

    aan de Universiteit van Amsterdam,

    op gezag van de Rector Magnificus

    Prof. Dr. D.C. van den Boom

    ten overstaan van een door het college voor promoties

    ingestelde commissie,

    in het openbaar te verdedigen in de Aula der Universiteit

    op vrijdag 26 november 2010, te 11.00 uur

    door

    Dhiredj Chandre Jagesar geboren te Nickerie, Suriname

  • Promotiecommissie

    Promotor: Prof. dr. A.M. Brouwer

    Co-promotor: Prof. dr. W.J. Buma

    Overige leden: Dr. E. Eiser

    Prof. dr. C.J. Elsevier

    Prof. dr. F. Hartl

    Dr. G. Lodder

    Prof. dr. F. Paolucci

    Prof. dr. J.N.H. Reek

    Dr. S. Woutersen

    Faculteit der Natuurwetenschappen, Wiskunde en Informatica

    Het onderzoek beschreven in dit proefschrift werd uitgevoerd binnen het Van ’t Hoff

    Institute for Molecular Sciences, Faculteit der Natuurwetenschappen, Universiteit van

    Amsterdam.

    Dit werk kwam tot stand met financiele ondersteuning van de Nederlandse Organisatie

    voor Wetenschappelijk Onderzoek (NWO).

    Printed by Wöhrman Print Service

    ISBN 978-90-9025766-2

  • v

    C o n t e n t s

    C h a p t e r 1 1 Introduction Intercomponent Interactions and Mobility in Rotaxanes

    1.1 Interlocked Molecules..................................................................................................................... 2 1.2 Intercomponent Interactions in Rotaxanes ................................................................................ 3

    1.2.1 Synthesis ................................................................................................................................. 3 1.2.2 Hydrogen-Bonded Rotaxanes ............................................................................................ 4 1.2.3 π-Stacking............................................................................................................................... 9 1.2.4 Hydrophobic Interactions .................................................................................................10 1.2.5 Transition Metal Coordination.........................................................................................11

    1.3 Intercomponent Dynamics in Rotaxanes.................................................................................. 11 1.3.1 Chemically Driven Molecular Shuttles............................................................................13 1.3.2 Electron-Driven Molecular Shuttles................................................................................14 1.3.3 Light-Driven Molecular Shuttles......................................................................................17

    1.4 Scope of this Thesis ...................................................................................................................... 21 1.5 References....................................................................................................................................... 22

    C h a p t e r 2 31

    Photoinduced Shuttling Dynamics of Rotaxanes in Viscous Polymer Solutions

    2.1 Introduction.................................................................................................................................... 32 2.2 Results and Discussion ................................................................................................................. 35

    2.2.1 Rheological Behavior .........................................................................................................35 2.2.2 Shuttling ...............................................................................................................................37 2.2.3 Hydrodynamic Scaling Model...........................................................................................41 2.2.4 Correlation with Macroscopic Viscosity .........................................................................42 2.2.5 Power-Law Relationship....................................................................................................43

    2.3 Conclusion ...................................................................................................................................... 46 2.4 Photophysics of the Naphthalimide Rotaxane......................................................................... 46 2.5 Experimental Details..................................................................................................................... 49 2.6 References....................................................................................................................................... 51

    C h a p t e r 3 57 Naphthalimide Rotaxanes Infrared Study of Intercomponent Interactions in a Switchable Hydrogen-Bonded Rotaxane

    3.1 Introduction.................................................................................................................................... 58 3.2 Results and Discussion ................................................................................................................. 61

    3.2.1 Amide I.................................................................................................................................61 3.2.2 Solvent effect.......................................................................................................................66 3.2.3 NH stretching......................................................................................................................68 3.2.4 Infrared Spectroelectrochemistry.....................................................................................69

    3.3 Conclusion ...................................................................................................................................... 75 3.4 Appendix: Calculations N-Methylacetamide ............................................................................ 76 3.5 Experimental Details..................................................................................................................... 76 3.6 References....................................................................................................................................... 79

  • vi

    C h a p t e r 4 83 Pyromellitimide Rotaxanes Selective Translocation of the Rings in a [3]Rotaxane

    4.1 Introduction ....................................................................................................................................84 4.2 Results and Discussion..................................................................................................................87

    4.2.1 Infrared Spectra .................................................................................................................. 87 4.2.2 Infrared Spectroelectrochemistry .................................................................................... 89

    4.3 Conclusion.......................................................................................................................................99 4.4 Experimental Details .................................................................................................................. 100 4.5 References .................................................................................................................................... 102

    C h a p t e r 5 105 Perylene Diimide Rotaxanes Hydrogen Bonding and Macrocycle Translocation

    5.1 Introduction ................................................................................................................................. 106 5.2 Results and Discussion............................................................................................................... 108

    5.2.1 Perylene Diimide Rotaxanes [2]8 and [3]8 ................................................................... 108 5.2.2 Perylene Diimide Shuttles [2]9 and [3]9 ....................................................................... 111

    5.3 Conclusion.................................................................................................................................... 122 5.4 Experimental Details .................................................................................................................. 123 5.5 References .................................................................................................................................... 124

    C h a p t e r 6 129 Photoinduced Dynamics in Structurally Modified Naphthalimide Rotaxanes

    6.1 Introduction ................................................................................................................................. 130 6.2 Co-conformer Distribution ....................................................................................................... 132 6.3 Photophysical Behavior ............................................................................................................. 134 6.4 Photoinduced Shuttling.............................................................................................................. 137

    6.4.1 niGly Rotaxane 11 ............................................................................................................. 137 6.4.2 C9 Rotaxane 12.................................................................................................................. 139

    6.5 Temperature Dependence of the Shuttling Rate................................................................... 140 6.6 Discussion .................................................................................................................................... 142 6.7 Conclusion.................................................................................................................................... 147 6.8 Experimental Details .................................................................................................................. 148 6.9 References .................................................................................................................................... 150

    C h a p t e r 7 153 Fullerene Rotaxanes Reverse Shuttling in Fullerene-Stoppered Rotaxane

    7.1 Introduction ................................................................................................................................. 154 7.2 Results and Discussion............................................................................................................... 155

    7.2.1 Solvent Induced Shuttling in C60R ................................................................................ 155 7.2.2 UV-Vis Absorption and Fluorescence ......................................................................... 157 7.2.3 Transient Absorption ...................................................................................................... 160 7.2.4 Driving Force for Shuttling ............................................................................................ 162

    7.3 Conclusion.................................................................................................................................... 163 7.4 Experimental Details .................................................................................................................. 164 7.5 References .................................................................................................................................... 165

  • vii

    C h a p t e r 8 169 Pressure-Induced Absorption and Fluorescence Shifts of Solvatochromic Probes

    8.1 Introduction..................................................................................................................................170 8.1.1 Pressure-Viscosity Relationship .....................................................................................170 8.1.2 Pressure-Induced Polarity Change.................................................................................171

    8.2 Results and Discussion ...............................................................................................................173 8.2.1 Absorption Probe ET30...................................................................................................173 8.2.2 Fluorescence Probe 5PI ...................................................................................................176

    8.3 Conclusion ....................................................................................................................................178 8.4 Experimental Details...................................................................................................................179

    8.4.1 High-Pressure Setup.........................................................................................................179 8.4.2 Absorption and Fluorescence.........................................................................................181

    8.5 References.....................................................................................................................................182

    Summary 185

    Samenvatting 189

    Dankwoord 193

  • C h a p t e r 1

    Introduction Intercomponent Interactions and Mobility in Rotaxanes

    Abstract

    Rotaxanes are among the most prominent members of the class of interlocked molecules. In the field of research on rotaxanes and other interlocked architectures, scientists are facing many challenges, but also many interesting opportunities are recognized. In this introductory chapter the structural aspects of rotaxanes in terms of mechanical bonding and interactions between components are discussed. The focus of this thesis is on hydrogen-bonded rotaxanes; therefore special attention is reserved for intercomponent hydrogen bonding interactions. Also the different approaches for the synthesis of rotaxanes are discussed.

    Interlocked molecules possess unique properties arising from the restriction of degrees of freedom in comparison with their separate components. In appropriately designed systems, the mechanical movements of the components with respect to each other can be made to occur between different well-defined states in a controlled manner. These molecular scale switches are referred to as molecular motors, for they are able to convert chemical, electrochemical and photochemical energy into controllable molecular motion. The molecular motor functionality of rotaxanes is discussed and illustrated with several state-of-the art prototypes.

  • C h a p t e r 1

    2

    1.1 Interlocked Molecules The structure of organic molecules is traditionally described in terms of the number and

    types of atoms they contain and the sequence and nature of the connecting bonds. Two

    molecules containing the same atoms, linked in a different sequence, are described as

    constitutional isomers, for example n-butane and 2-methylpropane. Even when two

    molecules contain the same atoms and sequence it is still possible for isomers to exist,

    originating from different spatial arrangements of atoms. This isomerism is referred to as

    stereoisomerism and arises in molecules containing substituted double bonds (E-Z

    isomerism) and asymmetric carbon atoms (optical isomerism). Over the years, as the range

    of molecules prepared became more complex, many special forms of isomerism (e.g. in

    octahedral metal complexes) have been identified. However, these are all variants upon one

    of the fundamental types of isomerism.

    A new type of isomerism was first described in the early 1960’s, namely topological

    isomerism.[1] This type of isomerism comes into play when considering the structure of

    mechanically interlocked architectures such as catenanes and rotaxanes (Figure 1-1).

    [2]rotaxane

    [2]catenane

    trefoil knot

    Solomon knot

    Borromean rings

    Figure 1-1 Schematic representation of some interlocked molecules.

    An [n]catenane, from the Latin word "catena" meaning chain, consists of n rings which

    are mechanically interlocked. The two rings in the [2]catenane in Figure 1-1 do not differ

    from the unlinked rings in terms of the atoms or bonds they contain, yet they are chemically

    different. The two structures, the separate rings and the catenane, are called topological

    isomers, because it is impossible to convert the catenane into its two separate rings without

    breaking chemical bonds. Other examples of interlocked topological isomers, of which

    synthetically created molecular embodiments are known, are trefoil knots,[2-5] Solomon

    knots (doubly interlocked catenanes)[6-9] and molecular Borromean rings.[10,11]

    [n]Rotaxanes, from the Latin words “rota”, meaning wheel and “axis”, meaning axle,

    consist of a dumbbell-shaped molecule encircled by n-1 macrocycles. The macrocycle is

    mechanically trapped on the axle due to the bulky stoppers on each end which prevent the

  • Introduction

    3

    macrocycle to slip away. Despite the fact that the macrocycle is mechanically locked onto

    the thread, rotaxanes are not considered topological isomers of their components, because

    they can be separated without breaking of chemical bonds: dissociation of a rotaxane into

    its separate parts is in principle possible by simply slipping the macrocycle over a stopper.

    In the literature, several examples of [n]rotaxanes with more than one macrocycle (n > 2)

    or thread have been reported. Some examples of these rotaxanes with more complex

    structures are (Figure 1-2): daisy chains[12-15] (A), polyrotaxanes[16-18] (B) and doubly threaded

    rotaxanes[19-22] (C).

    (A) (B) (C)

    Figure 1-2 Examples of rotaxanes with more complex structures. (A): Daisy chain, (B): polyrotaxane and (C): doubly threaded rotaxane.

    Daisy chains are referred to as “molecular muscles” because their components can

    perform contraction and extension movements, reminiscent of the shortening and

    lengthening of the functional elements in muscle fibres (see also Figure 1-10).

    1.2 Intercomponent Interactions in Rotaxanes

    1.2.1 Synthesis In contrast with the synthesis of ordinary organic molecules, in which chemical bonds

    between atoms are formed, for the construction of rotaxanes one has to deal with the

    challenge of creating mechanical bonds. Two main strategies can be distinguished in the

    synthesis of rotaxanes: threading and clipping (Figure 1-3). In the threading approach, the

    axle without the end groups is threaded through the cavity of a macrocycle to form a so-

    called pseudorotaxane, which after endcapping affords the rotaxane. Another approach is

    clipping, in which the macrocycle is assembled around the axle. A third and less often used

    strategy is slipping (not shown). In this approach the macrocycle is slided onto an already

    existing thread, the slipping can be induced by heat[23,24] or pressure.[25]

    The major event in the abovementioned approaches is reaching a specific spatial

    arrangement of the molecular components’ precursors with respect to each other. The first

    successful synthesis of a rotaxane relied on the low statistical probability of the thread

    precursor threading through a resin-immobilized macrocycle, followed by capping.[26] This

    process had to be repeated 70 times to arrive at a yield of only 6%.

  • C h a p t e r 1

    4

    Figure 1-3 Two different approaches generally used in the template-directed synthesis of rotaxanes. (A): Threading and capping and (B): clipping. The template which serves to pre-organize the precursors is represented by the rectangle.

    The recognition and development of templates in synthesis has dramatically improved

    the accessibility to a wide variety of rotaxanes. In template-directed synthesis the required

    spatial arrangement of the components is induced by noncovalent interactions between

    recognition sites incorporated into the precursors of the macrocycle and the thread. In the

    past two decades several templating mechanisms have been developed for rotaxane

    synthesis. In fact, all types of noncovalent interactions known in chemistry are suitable for

    this purpose and have indeed been utilized for the template-directed synthesis of rotaxanes.

    In the majority of the template-directed approaches, the noncovalent interactions that are

    used to correctly align the precursors, survive the chemical reactions necessary to link the

    components, and live on in the resulting rotaxanes. As will be discussed in section 1.3, the

    manipulation of these interactions is the starting point for the application of rotaxanes as

    molecular motors or switches. The pioneering work of several research groups, each

    specialized in a certain type of templating interaction, has resulted in a huge number of

    different rotaxanes. Three of the four major groups of rotaxanes rely on specific

    interactions between the thread and the macrocycle, namely hydrogen bonding, π-stacking,

    and transition metal ion coordination. Another important class is assembled through the

    hydrophobic effect. In the next sections, an overview is presented of the rotaxanes based

    on these different types of interactions.

    1.2.2 Hydrogen-Bonded Rotaxanes Hydrogen bonds are formed between a donor with an available “acidic” hydrogen atom,

    e.g. NH or OH groups, and an acceptor carrying available non-bonding lone pairs of

    electrons. The first example of a hydrogen-bonded rotaxane was reported in the literature

    by F. Vögtle and coworkers.[27] This rotaxane with hydrogen bonding between amide groups

    in the macrocycle and the thread was synthesized using templating effects following the

    threading and capping approach. This research group also developed a trapping approach,

    based on hydrogen bonding of the macrocycle with an alkoxy anion and followed by

    reaction with an alkylbromide to form an ether bond.[28,29] A typical feature of these

  • Introduction

    5

    rotaxanes with ether linkages is that the interaction that is applied to pre-align the axle

    precursors is lost once the rotaxane is formed. An example of such a rotaxane is discussed

    in Chapter 5.

    In the 1990’s, the research group of D.A. Leigh developed another powerful synthetic

    strategy for the synthesis of a new group of hydrogen-bonded rotaxanes.[30-32] Their five-

    component clipping strategy utilizes a dipeptide motif in the axle as template for the

    formation of the benzylic tetraamide macrocycle from its precursors (isophthaloyl

    dichloride and xylylene diamine, Figure 1-4A).

    (A)

    (B)

    Macrocycles Binding motifs

    amide glycylglycine (GlyGly) succinamide (succ)

    fumaramide (fum) glycyl glycyl ester succinic amide ester

    X = Y = CH, R = H X = N, Y = CH[33-35] X = CH, Y = N[36,37] X = Y = CH, R = NO2

    [37,38]

    adipamide[39] nitrone squaraine

    Figure 1-4 (A): Example of hydrogen-bond assisted rotaxane synthesis using a fumaramide motif as template. (B): Overview of the different possible binding motifs for the macrocycle and the structural variations of the macrocycle.

    Several binding motifs have been found to be efficient templates for rotaxane formation

    following this approach, including even a simple mono-amide.[40] The best results are

    obtained with templates containing at least two hydrogen bond accepting amide or ester

    groups: glycylglycine (GlyGly),[41] fumaramides (fum),[39,42-45] succinamides (succ)[37,39,42,46-49] and

  • C h a p t e r 1

    6

    succinic amide esters[50] (Figure 1-4B). The isophthaloyl dichloride is tolerant towards

    introduction of meta-substituents onto the aromatic ring, such as nitro groups.[37,38]

    Substitution of the benzene ring with a pyridine ring[33-37] allows further transformations

    after the formation of the rotaxane, including protonation[36] and alkylation.[51]

    The rotaxanes described in this thesis are Leigh-type hydrogen-bonded rotaxanes. These

    rotaxanes exhibit a great diversity in cooperative and multipoint hydrogen bonding between

    their components. Due to the tolerance of the used clipping reaction (Figure 1-4) towards

    different binding motifs and macrocycles, it is possible to tune the intercomponent

    hydrogen-bonding interactions, simply by selecting the appropriate components with the

    desired binding properties. The diverse binding can be traced back to the structural

    characteristics of the binding template and the macrocycle. The binding strengths depend

    on the hydrogen bond acceptor strengths, the spatial arrangements of the hydrogen bond

    accepting C=O and donating NH groups of the macrocycle and the binding motif, and on

    the conformational flexibility of the components.

    Hydrogen Bonding Donor-Acceptor Strength

    The acceptor strength of the binding motif can be modified by incorporation of

    electron-donating or withdrawing moieties on the amide nitrogen atoms. For example,

    additional alkylation of the amide groups results in decreased hydrogen bond affinity of the

    thus obtained tertiary amides compared to the secondary amides. Replacement of an amide

    group by an ester group is another way to influence the acceptor strength; in this case the

    hydrogen bonding becomes weaker.

    Also other types of binding motifs with hydrogen bonding groups other than amides are

    capable of binding the macrocycle. An important category of such acceptors is that of

    functional groups containing negatively charged oxygen atoms. For instance, nitrones[52] and

    squaraines[24,34,35,53] are found to possess excellent affinity towards the macrocycle and have

    been used to template the assembly of rotaxanes. Apart from these stabile oxides, another

    interesting group that can firmly bind the tetraamides macrocycle is that of “transient”

    oxides. These can be created chemically (e.g. N-oxides,[54] or alkoxy anions[40,48]), electro-

    chemically or photochemically. Examples of the latter two classes are radical anions and

    dianions of aromatic imides. The corresponding neutral imides are generally poor hydrogen

    bond acceptors. This principle makes imide-based binding motifs attractive functional units

    to induce macrocycle translocation in hydrogen-bonded rotaxanes, because the hydrogen

    bond affinity of the imide can be switched on and off by the interconversion between the

    neutral and the anion states. Examples of such imide binding motifs (naphthalimides,

    pyromellitimides and perylene diimides), and their application in macrocycle translocation in

    rotaxanes are described in Chapters 3 – 6.

    The hydrogen bonding affinity of the macrocycle can be tuned by modification of the

    isophthaloyl moiety. Replacement of a CH-unit in the phenyl ring by a nitrogen atom at the

  • Introduction

    7

    2-position or the 5-position leads to the “endopyridine”[33-35] or “exopyridine”[36,37] analogs,

    respectively. In these pyridines the acidity of the NH protons is increased due to the

    electron-withdrawing nitrogen atoms; hence these macrocycles bind more strongly to the

    thread. Alkylation or protonation of the nitrogen of the pyridine leads to a further increase

    of the hydrogen bonding affinity. Nitro groups on the phenyl ring induce the same

    effect.[37,38]

    Conformational Flexibility

    Due to its flexibility, the diamide binding motif can adopt several conformations. These

    internal degrees of freedom are to a large extent lost in the rotaxane.[55] It is likely that the

    unfavorable loss of entropy in going from the flexible binding template to the much stricter

    conformational requirements of the rotaxane could be partly overcome by pre-organizing

    the hydrogen bonding sites of the template in a spatial arrangement already suited for

    binding the macrocycle. This is indeed the case. For example, the spatial arrangements of

    the amide groups in the succinamide and the fumaramide motifs are the same, but still the

    binding to the fumaramide motif is stronger. This is due to the rigidity of the latter and is

    reflected by the higher yields generally obtained with a fumaramide template (97%)

    compared to a succinamide template (53%).[39] Another illustrative example is provided by

    [2]rotaxanes containing both a fumaramide and succinamide binding site. In these systems

    the macrocycle resides almost exclusively at the fumaramide site (> 95%).[39] These

    examples demonstrate that the structural rigidity of the thread binding sites has a major

    influence on hydrogen bonding with the macrocycle.

    The spatial arrangement of the amides in the binding motif determines the geometry of

    the hydrogen bonds and thus the binding strength. The best fit is obtained if the two amide

    C=O groups are separated by two bonds, like in the succinamide and fumaramide motifs.

    X-ray crystal structures show that in this case the macrocycle adopts a chair-like

    conformation and forms two sets of bifurcated hydrogen bonds with the diamide motif, as

    is depicted in Figure 1-4.[39] If the distance is larger, the macrocycle NH groups can hardly

    reach both amides to form two sets of bifurcated hydrogen bonds. This is illustrated by the

    low yield of 8% if an adipamide template is used.[39] Another example of the importance of

    the spatial orientation of the binding motif is a rotaxane in which complexation of one of

    the amide N-atoms to a transition metal forces the diamide binding motif to adopt a spatial

    arrangement which is unfavorable for binding the macrocycle.[50] The result of complexation

    is that the macrocycle can no longer bind to the amide, due to disturbed multipoint

    hydrogen bonds interactions.

    Another important parameter that governs the hydrogen bond interactions is the

    conformational flexibility of the macrocycle. The flexibility of the macrocycle gives rise to a

    variety of binding geometries. In general, the conformation of the macrocycle is to a large

    extent determined by the binding motif, and vice versa, because the system will try to reach

  • C h a p t e r 1

    8

    a compromise between energy-minimizing hydrogen bonds and the extent of

    conformational strain of the different components. This mutual dependence is illustrated by

    several examples reported in the literature. In principle, the amides of the macrocycle can

    function as both hydrogen-bond donor (NH) and acceptor (C=O). In the fumaramide

    rotaxane depicted in Figure 1-4, the macrocycle adopts a chair-like conformation in order to

    facilitate two sets of bifurcated hydrogen bonds with the fumaramide template. In rotaxanes

    with a GlyGly binding motif, a variety of conformations of the macrocycle and different

    arrangements of the hydrogen bonds can be distinguished.[56] The X-ray crystal structure

    reveals that the macrocycle amide groups also act as hydrogen-bond acceptors. In these

    instances, the C=O groups are rotated and point inwards to form hydrogen bonds with the

    NH groups of the GlyGly dipeptide motif (Figure 1-5). Macrocycles containing endo-

    pyridine moieties have been shown to adopt a boat-conformation when attached to a

    squaraine template (Figure 1-5).[35] The bifurcated arrangement of hydrogen bonds in

    fumaramide rotaxanes is confirmed by the X-ray crystal structures.[39] This binding geometry

    is also found in rotaxanes containing a succinamide binding motif.[57]

    Figure 1-5 Different binding geometries of the tetra-amide macrocycle to different binding motifs.

    It should be noted that the crystal structure is to a large extent determined by

    intermolecular interactions. In solution on the other hand, these interactions play a minor

    role and the intercomponent and solvent-solute interactions become dominant. Therefore it

    is likely that more than one conformation and hydrogen-bonding geometry can exist in

    solution, while generally only one conformation is found in the crystalline solid state.

    Hydrogen-Bonded Crown Ether Rotaxanes

    Another important group of hydrogen-bonded rotaxanes is that of the secondary

    ammonium/crown ether type. These systems are characterized by the formation of

    hydrogen bonds between a protonated secondary amine in the thread and a crown ether

    macrocycle (Figure 1-6B). These rotaxanes are easy to synthesize with high yields using the

  • Introduction

    9

    threading and capping approach. Several examples of this type of rotaxanes are reported in

    the literature, with structural variations in the crown ether macrocycles and threads.[58-61] An

    elegant example of a system containing this binding type is the molecular cage-based

    rotaxane depicted in Figure 1-6B.[60]

    (A) (B)

    Figure 1-6 (A): Binding motif in rotaxanes based on hydrogen bonding between a secondary ammonium and a crown ether macrocycle. (B): Molecular cage-based rotaxane.[60]

    The hydrogen bond interaction between the thread and macrocycle can be manipulated

    easily by deprotonation of the ammonium ions, which leads to weaker hydrogen bonds with

    the crown ether. This principle can be used to induce translational motions of the

    macrocycle along the thread.[13,60-63]

    1.2.3 ππππ-Stacking Aromatic π-stacking interaction between electron-deficient and electron-rich com-

    ponents is another important binding type in many rotaxanes. This type of rotaxanes was

    developed in the research group of Stoddart during the 1980’s. The synthesis of these

    rotaxanes comprises the formation of a pseudorotaxane host-guest complex by π-π interactions (Figure 1-7), followed by endcapping to obtain the [2]rotaxane. This approach

    has led to the synthesis of numerous rotaxanes and catenanes. Various electron-deficient

    and electron-rich compounds have proven to be efficient combinations for creating

    interlocked structures. An overview of the diversity of electron-rich structures capable of

    binding the electron-deficient cyclobis(paraquat-p-phenylene) macrocycle (CBPQT4+) is

    presented in Figure 1-7A.

    Of special interest is the tetrathiafulvalene unit (TTF). Among the summarized electron-

    rich guests, TTF exhibits the strongest binding affinity towards the CBPQT4+ macrocycle. It

    also displays a very favorable electrochemical behavior. The TTF2+ cation obtained after

    electrochemical oxidation is an extremely weak binding station for the tetra-cationic

    macrocycle due to electrostatic repulsion. Therefore, TTF and derivates thereof have been

    used in a large number of rotaxanes with the purpose of inducing movement of the

    macrocycle along the thread. Applications of this principle are discussed in section 1.3. The

  • C h a p t e r 1

    10

    binding type with the electron-rich component in the macrocycle (Figure 1-7B) is also

    tolerant towards several structural modifications. For example, substitution of the phenyl

    ring with other aromatic ring systems, e.q. 1,5-naphthalene,[70,71,89] can be applied to tune the

    strength of the intercomponent interactions. Other examples of electron-deficient guest are

    the N,N-dialkyl-2,7-diazapyrenium dication[84] and naphthalene diimides.[89] Besides π-stacking interactions, due to the presence of the crown ether part, this macrocycle is also

    capable of forming stable host-guest complexes with secondary ammonium ions through

    hydrogen bonding (see also Figure 1-6). This dual binding mode is illustrated by the

    rotaxane depicted in Figure 1-9.

    (A) (B)

    [64,65] [66-72] [65,66,68-75] [76-79]

    [80] [81,82] [65,83] [84]

    [85,86] [87,88]

    Figure 1-7 Two different types of binding motifs in rotaxanes based on π-stacking. (A): Overview of the different electron-rich guests capable of binding the electron-deficient cyclobis(paraquat-p-phenylene) macrocycle. (B): Electron-rich dibenzo-crown ether macrocycle and examples of electron-deficient guests. The reference numbers are given between square brackets.

    1.2.4 Hydrophobic Interactions

    The most common example for hydrophobic interaction driven formation of rotaxanes

    is provided by cyclodextrins. Cyclodextrins are cyclic oligosaccharides, characterised by a

    hydrophilic exterior and a hydrophobic interior. The hydrophilic nature of the exterior is

    due to the presence of hydroxyl groups, while the aliphatic carbons are located in the cavity,

    which is the origin of the hydrophobicity. The two most familiar and frequently used

    cyclodextrins for rotaxane synthesis are α-cyclodextrins and β-cyclodextrins. The property

  • Introduction

    11

    of cyclodextrins of forming inclusion complexes with various substrates has led to the

    synthesis of numerous rotaxanes, with different axles.[12,90-95] An example of a rotaxane based

    on hydrophobic interactions with cyclodextrins is shown in Figure 1-12.

    1.2.5 Transition Metal Coordination Due to the variety in coordination number and geometry, transition metal ions are

    powerful templates in various reactions where a specific alignment of reactants or reaction

    intermediates is required to obtain selectivity. Transition metal ion templating strategies for

    the synthesis of interlocked molecules were developed during the 1980’s by the research

    group of Sauvage.[96] Sauvage used the tetrahedral coordination geometry of a CuI ion which

    provides the basis for a correct alignment of the axle within the macrocycle. An example of

    such a rotaxane synthesized with the assistance of a transition metal ion is depicted in

    Figure 1-10. In the course of the years, several other transition metals have been employed

    in this approach, including iron,[20] cobalt,[20,97] copper,[98] ruthenium,[99] palladium[16,100,101] and

    gold.[102]

    An interesting development in this field is the so-called active-metal template approach.

    In this approach the transition metal is not only used to assist the correct alignment of the

    thread precursors in the macrocycle cavity, but it also functions as an active catalyst for the

    coupling of the precursors. Examples of such couplings are CuI catalyzed azide-alkyne 1,3-

    cycloadditions,[103] PdII catalyzed coupling of terminal alkynes[104] and PdII catalyzed oxidative

    Heck cross-couplings.[105]

    1.3 Intercomponent Mobility in Rotaxanes Initially, the design and synthesis of rotaxanes were purely performed because of the

    challenge to develop synthetic approaches for the construction of these appealing

    structures. However, in the course of the past two decades, the perspective on interlocked

    molecules, and in particular rotaxanes, has undergone a transformation from novel to

    functional molecules. The interlocked nature of rotaxanes has been recognized as a tool for

    the expression of functionalities that are not achievable with “ordinary” molecules.

    This idea was prompted by the interesting property of mechanically interlocked

    molecules that they offer a unique set of additional degrees of freedom, arising from

    motions of the components with respect to each other. Three different types of

    intercomponent motions can be distinguished in rotaxanes. The macrocycle can undergo

    translocation along the axle, the so-called shuttling motion. The rotational motion of the

    macrocycle along the axle is referred to as pirouetting. Also pivoting motions, changing the

    angle between the axle and the plane of the macrocycle, are possible.

    The position of the macrocycle in rotaxanes is determined by the relative strength of the

    possible intercomponent interactions with the different parts of the thread. In rotaxanes

  • C h a p t e r 1

    12

    containing two binding sites (also referred to as binding stations), an equilibrium will exist

    between the states in which the ring is bound to station 1 or to station 2, depending on the

    Gibbs free energy difference (∆G) of both co-conformers (Figure 1-8). As illustrated in the previous section, in appropriately designed systems the equilibrium can be adjusted, simply

    by incorporating different types of binding sites with different binding strengths. The

    system will minimize its energy by adopting a co-conformation in which the macrocycle

    predominantly resides on the station with the highest binding strength, i.e. corresponding to

    the lower Gibbs energy.

    Figure 1-8 Macrocycle shuttling in a rotaxane containing two different binding stations, induced by an external stimulus.

    Binding to the other station can be made favorable after modification by an external

    stimulus of station 1, leading to a relatively lower binding affinity, or station 2, leading to a

    relatively higher binding affinity. This external stimulus can induce changes in geometric

    configuration (e.g. E-Z isomerization) or electronic arrangement (e.g. electrochemical

    reduction) of the binding stations, or changes in the environmental properties (e.g. polarity,

    temperature) which influence the noncovalent intercomponent interactions. In both cases,

    after modification the energetically favorable situation is the macrocycle residing at station

    2*. This situation is reached by translocation of the macrocycle. The rate of the

    translocation process is determined by the Gibbs energy of activation (∆G‡). The height of this free energy barrier is influenced by internal (e.g. attractive forces between the

    macrocycle and station 2) and external factors (e.g. the polarity and viscosity of the

    medium). If the stimulus has altered the relative binding strength of the station irreversibly,

  • Introduction

    13

    the endpoint is co-conformer C. If, on the other hand, the change is reversible, the original

    stations can be restored. The macrocycle then shuttles back to its initial position,

    completing the shuttling cycle. A rotaxane in which reversible ring translocation between

    two distinct binding stations is possible, is referred to as a molecular shuttle.

    A wide variety of activation methods, comprising chemical, electrochemical and

    photonic stimuli, can be employed to modulate the noncovalent interactions. The suitability

    of these stimuli depends much on the nature of the binding stations and the interactions

    involved. Chemical stimuli may be less elegant from a practical point of view, because they

    require manipulation with solvents. In the majority of molecular shuttles designed so far,

    the shuttling motion is made to occur by a change in redox states of the binding station.

    Redox states can be changed electrochemically or photochemically. Also, (reversible)

    photoisomerization is an attractive approach to induce and control the macrocycle

    shuttling. In the next paragraphs some examples of chemically, electrochemically and

    photoinduced shuttling in rotaxanes are discussed.

    1.3.1 Chemically Driven Molecular Shuttles The modification of the relative binding affinities of different stations on the thread and

    subsequent macrocycle translocation can be achieved by using a chemical reaction.

    Examples of chemical reactions suitable for reversible activation of rotaxane-based shuttles

    are formation of imine bonds[106] and Diels-Alder reactions.[107] However, the most often

    used chemical approach for the reversible activation of rotaxane-based shuttles is

    protonation-deprotonation of secondary amines in rotaxanes containing crown ether

    macrocycles.[48,60,61,76,82,108-110] A spectacular example of such a pH-switchable system is the

    [4]rotaxane depicted in Figure 1-9. [77,78]

    This system contains thrice the features of a previously reported pH-switchable

    [2]rotaxane.[76] It contains three legs, each containing two different types of recognition

    sites, a dialkylammonium centre (―RR’NH2+―) and a bipyridinium unit (BIPY2+). The three

    2,3-naphtho[24]crown-8 macrocycles fused onto a hexaoxytriphenylene core can interact

    with ― RR’NH2+― or the BIPY2+ stations via hydrogen bonding or π-stacking, respectively.

    The position switching of the platform containing the three [24]crown-8 macrocycles from

    the ―RR’NH+― station to the BIPY2+ stations was induced chemically by deprotonation of

    the dialkylammonium stations using a strong phosphazene base (t-BuNP(NMe2)3). In the

    deprotonated state (―RR’NH―) the hydrogen bonding interaction is lost and the

    macrocycles move to the BIPY2+ where donor-acceptor interactions and hydrogen bonding

    become stabilizing. The shuttling is fully reversible, upon protonation of the ―RR’NH―

    with trifluoroacetic acid, the platform moves back to its original position.

  • C h a p t e r 1

    14

    Figure 1-9 Base-acid controlled operation of a molecular elevator. Upon deprotonation of the dialkylammonium ions, the platform switches position to bind with the energetically more attractive BIPY2+ stations. The picture was redrawn from reference [77].

    The coherent movement of the three macrocycles in this rotaxane resembles the

    operation of an elevator. The shuttling in both directions was monitored with 1H NMR,

    cyclic voltammetry and UV-Vis spectroscopy.

    1.3.2 Electron-Driven Molecular Shuttles Switching between redox states, by electrochemical oxidation and reduction is a versatile

    tool to induce translational motions in rotaxanes. In rotaxanes containing more than one

    binding station, it can be employed to modify the initial binding station such that binding

    becomes energetically less favorable. Alternatively, electrochemical oxidation or reduction

    can also be used to enhance the binding affinity of an unoccupied station, such that binding

    to this station becomes more attractive. In both cases, the macrocycle will switch position

    to form predominantly the co-conformer with the lowest energy. The requirement for the

    operation of electron-driven molecular shuttles is the presence of at least one redox-active

    component in the molecule.

    Transition metals form one group of such redox-active units. Oxidation or reduction of

    transition metals can change the affinity towards different ligands. The change of affinity

    can generally be attributed to a change of the electron configuration, which determines the

    coordination geometry around the metal centre. This is illustrated by the copper ion based

    molecular shuttles designed by Sauvage and co-workers;[98,111,112] an example is shown in

  • Introduction

    15

    Figure 1-10.[98] In these systems, the binding geometry can be switched between a four-

    coordinate geometry around a CuI centre to a five-coordinate CuII complex. Upon

    electrochemical oxidation of the CuI centre, the metal ion and the bidentate macrocycle

    migrate from the bidentate phenanthroline to the tridentate terpyridine station. The rate

    constants of the shutting process were determined from cyclic voltammetry experiments,

    values of 0.4 s-1 and 50 s-1 were found for the forward and backward process, respectively.

    This switching principle has also been applied for controlling the pirouetting motion of the

    macrocycle in rotaxanes.[113]

    Figure 1-10 A reversible, electron-switchable molecular shuttle based on oxidation and reduction of a coordinated transition metal ion.[98]

    Redox-active components based on organic molecules form the second group which has

    been extensively used in electron-driven molecular shuttles. The majority of these molecules

    contain an aromatic core or conjugated double bonds. This is no surprise, because due to

    their ability to accommodate an excess charge by delocalization, the oxidation or reduction

    potentials of these molecules are well within the potential window in which commonly used

    solvents are electrochemically inert (e.g. for acetonitrile ca. -2.5 – 3.5 V vs. NHE).[114] In the

    literature, a large number of electron-driven shuttles have been reported.

    The research group of Stoddart has created several examples in which electrons are used

    to modulate π-stacking interactions between a binding station and the macrocycle, leading

    to ring movement. Among the redox-active stations in these examples are benzidines[82] and

    tetrathiafulvalenes (TTF).[68,72,74] Upon electrochemical oxidation of these neutral electron-

    rich stations, the ring is electrostatically repelled and migrates along the thread to bind with

    another station further up the axle (e.g. naphthalene, NP). An example of such a system,

  • C h a p t e r 1

    16

    which contains the typical features of the Stoddart-type shuttles, is the [3]rotaxane depicted

    in Figure 1-11.[66,115] In solution, the rings of this shuttle can be switched between the TTF

    and NP units by oxidation and reduction of the TTF unit. The movement of the

    macrocycle from the TTF2+ unit is driven by electrostatic repulsion and the π-donor ability of the NP station. The return to the recovered TTF station is a thermally activated diffusive

    process. The shuttling of both macrocycles in solution was confirmed by UV-Vis spectro-

    electrochemical methods and cyclic voltammetry. The electron-induced switching in

    rotaxanes between TTF and NP units has also been exploited as the working mechanism in

    the operation of nanovalves[116,117] and for data storage purposes.[74]

    (A)

    (B)

    Figure 1-11 (A): Reversible electron-driven shuttle that functions as a molecular muscle.[66,115] After oxidation of the TTF units, the tetra-cationic CBPQT4+ macrocycles switch position to bind with the naphthalene stations. Upon back-reduction of the TTF2+ units, the rings move back to their initial position at the TTF station. (B): Schematic representation of the operation of the molecular muscle anchored on a gold-coated flexible silicon cantilever.

    This [3]rotaxane is able to deliver work and is one of the rare examples of artificial

    molecular motors whose performance is expressed on the macroscopic level. The shuttle

    was anchored to a gold-coated (thickness 20 nm) flexible silicon cantilever via a self-

    assembly process. The disulfide tethers, which are covalently attached to the macrocycle,

    form Au-S bonds with the surface, resulting in a monolayer of rotaxanes. Oxidation of the

  • Introduction

    17

    TTF unit with four equivalents of Fe(ClO4)3 and subsequent reduction with ascorbic acid

    causes the coated cantilever to bend and relax (Figure 1-11B). The bending could be

    detected by measuring the difference in deflection angle of a light beam in the oxidized and

    reduced state. The fact that the bending can be detected on the macroscopic level is

    remarkable, because the amplitude of the motion of the macrocycles is only 2.8 nm. The

    macroscopic physical change is explained by the fact that the bending is a cumulative effect

    of individual nanoscale movements.

    1.3.3 Light-Driven Molecular Shuttles The most elegant way to operate molecular motors is probably by using photons as fuel.

    One of the advantages of using photoactivation instead of the chemical or electrochemical

    methods is that the stimulus can be applied without the need of making physical contact

    with the sample. In this sense, photons are genuine external stimuli. Other advantages are

    the high time and spatial resolutions that can be achieved with light excitation: it leads to a

    fast response and can be performed in small spaces. The mechanism of photoactivation can

    be divided into two categories which are related to the events following the absorption of a

    photon.

    The excited state can be the endpoint of the activation process if it already has the

    desired difference in binding affinity compared to the ground state. In this case shuttling

    occurs in the excited state. Examples of this type are rare, because the lifetime of the

    excited state is generally too short (due to fast radiative or non-radiative decay) to allow

    completion of the shuttling motion which generally occurs on a much longer timescale.

    Therefore, the application of this principle is restricted to systems in which small-amplitude

    motions occur. The only known example is the hydrogen-bonded rotaxane with an active

    anthracene-carboxamide chromophore. In this rotaxane, the macrocycle moves towards an

    amide binding station close to the photo-excited chromophore. The higher affinity for

    binding to this station is due to a change in charge distribution and a geometrical change of

    the anthracene-amide in the excited state.[46] The translation occurs in less than 5

    nanoseconds and the travelled distance is only ca. three C-C bonds.

    Generally, in photoactivation strategies for inducing motions in rotaxane-based shuttles,

    the excited state only serves as a high-energy intermediate in subsequent processes leading

    to the desired modified binding station. An example of the first kind is the E-Z

    isomerization of C=C double bonds in fumaramide[118] or stilbene,[119] induced by direct UV-

    excitation or by sensitization with a triplet energy donor. In this case, the formation of the

    Z-isomer leads to translation of the macrocycle, which moves to an energetically more

    attractive binding station located on the thread. Alternatively, the chromophore in the

    excited state can act as a reactant and undergo a chemical reaction. These photochemical

    reactions may sometimes require the addition of reagents (e.g. reductants). The activation of

    the majority of the photo-controlled molecular shuttles is based upon this mechanism.

  • C h a p t e r 1

    18

    E-Z Photoisomerization

    Photoisomerization of C=C or N=N double bonds are often used photochemical

    reactions for the operation of molecular shuttles and switches. The mechanism of activation

    involves change of the binding affinity due to substantial geometric changes upon

    isomerization. The two main examples of the first type, isomerization of C=C bonds, are

    fumaramides and stilbenes. E-fumaramides are excellent binding stations for the Leigh-type

    tetraamide macrocycles, while the Z-isomers bind the macrocycle only poorly.

    Photoisomerization is therefore an easy tool to reversibly change the binding affinity of

    fumaramides and has been applied in several cases to induce macrocycle

    translation.[33,39,120,121]

    Stilbenes and azobenzenes (the latter contains N=N double bonds) are often associated

    with cyclodextrin containing rotaxane shuttles based on hydrophobic interactions. A nice

    example of a shuttling induced by photoisomerization is depicted in Figure 1-12.[93] This

    light-driven shuttle contains two isomerizable units, a stilbene and an azobenzene, and two

    different fluorescent naphthalimide stoppers. The α-cyclodextrin (α-CD) macrocycles of

    this [3]rotaxane can be switched selectively by selecting the appropriate wavelength of the

    excitation light. In this way the stilbene or azobenzene can be isomerized selectively and

    thus each of the macrocycles can be moved individually to the centre part of the thread.

    The positions of the rings along the thread are signalled by the fluorescence of the two

    different naphthalimide stoppers. The fluorescence intensity of either station increases

    drastically if the adjacent α-CD ring moves away in the Z-isomer.

    Figure 1-12 A light-driven molecular shuttle.[93] The directions of shuttling motions of the α-cyclodextrin macrocycles can be controlled by selective isomerization of the stilbene or azobenzene unit.

  • Introduction

    19

    Although photoisomerization is a clean way to activate molecular shuttles, because no

    auxiliary reactants are required, a drawback can be that a selective and complete conversion

    to each of the isomers is usually difficult to achieve. This is because the isomers have

    strongly overlapping absorption bands, which makes the selective excitation of either one of

    the isomers difficult. Better selectivity can sometimes be achieved by avoiding the singlet

    excited state, and allowing the isomerization to occur via the triplet state (T1). The T1 levels

    of the E and Z isomers can be at distinctly different energies, so the enrichment of the Z-

    isomer (with the higher triplet state energy) is possible with the use of triplet sensitizers with

    an energy between those of the two isomers.[122-124] This principle has been applied in

    molecular shuttles in an elegant way by incorporating the sensitizer in the molecular

    structure of the shuttle, close to the isomerizable double bond.[118,119]

    Photoinduced Electron Transfer

    Photoinduced electron transfer is the most basic photochemical reaction because only a

    single electron is transferred from a donor to an acceptor (Figure 1-13). The excited state

    (singlet or triplet) of a chromophore, incorporated in the molecular shuttle, can function as

    electron donor or acceptor. In appropriately designed systems electron transfer can occur

    on a very short timescale, ranging from picoseconds to nanoseconds. Also, in the majority

    of the cases the process is fully reversible.

    Due to these characteristics, photoinduced electron transfer can be a powerful tool for

    fast and reversible generation of radical anions and cations, suitable for the activation of

    molecular shuttles. Yet, this process has not been exploited very often in photoactivation

    strategies for inducing large-amplitude and controllable motions in molecular shuttles. The

    main reason is the rapid decay of the charge-separated state by charge recombination, in

    intramolecular cases or in contact ion pairs. The lifetime of the activated state might be too

    short to allow completion of the shuttling process. Therefore, in the few examples that use

    photoinduced electron transfer as activation process, it is associated with fast shuttling on a

    timescale of roughly < 100 µs (see the examples below).

    Figure 1-13 The mechanism of photoinduced electron transfer from a electron donor (D) to an acceptor (A). D* represents a singlet or triplet excited state.

  • C h a p t e r 1

    20

    One of the rare working examples of systems driven by photoinduced electron transfer

    is based on single-step photoinduced electron transfer from a Ru2+ metal centre to a

    bipyridinium binding station (Figure 1-14).[125,126] Upon photoactivation with visible light,

    the dibenzo-crown ether ring undergoes displacement from the bipyridinium to the 3,3’-

    dimethylbipyridinium station with a time constant of 48 µs at 303 K, but only 14% of the

    macrocycles can complete with the fast charge recombination recovery process (7.7 µs).

    The charge recombination process could be slowed down by reduction of the Ru3+ using an

    auxiliary electron donor. This resulted in a higher shuttling efficiency of 76%.

    (A)

    (B)

    Figure 1-14 (A): Molecular structure of a rotaxane-based molecular shuttle driven by photoinduced electron transfer.[125,126] (B): Operation mechanism.

    The lifetime of the charge-separated state can be extended by retardation of the back-

    electron transfer process. This can be achieved in triad systems in which sequential electron

    transfer steps occur. The dissociation of the radical ion pair is mimicked by making the

    distance between the ions as large as possible in order to inhibit charge recombination. A

    drawback of this approach is an inherent one generally associated with multi-step electron

    transfer processes: the low quantum efficiency. That means that the number of absorbed

    photons which are actually converted into mechanical movement, will be low. This

    principle has been proven to be effective to create long-lived charge-separated states in

    rotaxanes, but its actual application for inducing controlled intercomponent motions

  • Introduction

    21

    remains a challenge. Recently, a prototype of a light-driven shuttle was presented, based on

    multi-step electron transfer and containing the required ingredients to function in a ‘stand

    alone’ manner.[127] The photoinduced shuttling process in this shuttle, containing an array of

    electron donors and acceptors, was postulated on the basis of the properties and

    arrangements of the functional components.

    Another approach for extending the lifetime of the charge-separated state, and

    successfully implemented in molecular shuttles, is by designing systems in which back-

    electron transfer is spin-forbidden. Following this approach, in our research group a

    hydrogen-bonded molecular shuttle, based on a naphthalimide chromophore as binding

    station, has been studied. This shuttle exhibits the fastest shuttling dynamics ever observed

    for controlled large-amplitude macrocycle shuttling in artificial molecular shuttles.[128] The

    key element in this system is a naphthalimide chromophore that undergoes rapid

    intersystem crossing to the triplet state after which it is reduced by an electron donor

    present in solution. The back electron transfer is a slow process because of its spin-

    forbidden nature. Due to this, the dissociation of the ion-pair is very efficient. The lifetime

    of the naphthalimide radical anion (which functions as a new binding station for the

    macrocycle) is very long: in the range of hundreds of microseconds. This is long enough to

    allow macrocycle shutting; this process occurs on a timescale of several microseconds. The

    activation mechanism, shuttling dynamics and its detection are described in Chapter 2. The

    performance of these shuttles in response to environmental variations (viscosity) and some

    structural modifications is the scope of Chapters 2 and 6, respectively.

    1.4 Scope of this Thesis This thesis describes the operation of molecular shuttles based on hydrogen-bonded

    rotaxanes. The main topic is the correlation between the intercomponent mobility in these

    shuttles and the internal noncovalent interactions between their components, and external

    interactions with the surrounding medium. So, the main objective of this thesis is to provide

    an answer to the question: How does the structure of the rotaxane components and its

    environment influence the intercomponent dynamics? In order to answer this question, the

    shuttling process in photo- and electron-driven hydrogen-bonded rotaxane shuttles was

    studied after structural modification of their components and direct surroundings.

    Chapter 2 describes the effect of viscosity on the photoinduced shuttling in a

    naphthalimide-based rotaxane. This molecular shuttle is operated in viscous polymer

    solutions and the retardation effects on the shuttling dynamics are evaluated in the context

    of several theoretical and empirical models that describe the viscosity effect on the

    transport properties of solutes.

    The topic of Chapters 3 – 5 is hydrogen-bond interaction in rotaxanes and its effect on

    the macrocycle switching behavior. The influence of these interactions on the co-conformer

  • C h a p t e r 1

    22

    distribution is evaluated in terms of hydrogen-bond affinity of the binding stations towards

    the macrocycle. For this purpose, the macrocycle shutting in [2] and [3]rotaxanes with the

    same basis structure, namely consisting of a succinamide and a redox-active aromatic imide

    station connected via an alkyl spacer, was investigated with infrared and UV-Vis

    spectroscopy, and with cyclic voltammetry. Three different imides are incorporated in the

    rotaxanes in these studies: naphthalimide (ni, Chapter 3), pyromellitimide (pmi, Chapter 4)

    and perylene diimide (pdi, Chapter 5). The basic principles of detection of the shuttling by

    means of infrared spectroscopy are established in Chapter 3.

    In Chapter 6, the mechanism of the photoinduced shuttling dynamics in naphthalimide-

    based molecular shuttles is discussed. The activation parameters of the shuttling process in

    structurally modified naphthalimide rotaxanes are used as basis to propose a model for the

    shuttling process. The structural variations comprise the distance between the binding

    stations and the binding affinity of the naphthalimide station. The consequences of these

    variables for the activation parameters of the shuttling process were surprising, but could be

    explained in the context of the proposed model.

    Chapter 7 describes the macrocycle switching in a molecular shuttle containing a

    functional fullerene stopper. Spectroscopic and cyclic voltammetry data revealed an

    unexpected shuttling behavior which can be ascribed to the presence of the fullerene

    stopper.

    The work described in Chapter 8 is to some extent a continuation of Chapter 2. The idea

    is to use hydrostatic pressure to tune the viscosity of a medium suitable for the operation of

    imide-based molecular shuttles. In contrast with the approach used in Chapter 2, this

    alternative approach guarantees a microscopically homogenous increase of the viscosity. In

    this chapter some preliminary experiments are described with a home-built high-pressure

    set-up, which unfortunately was not available in time to realize the planned measurements.

    The preliminary experiments at least allowed to obtain a picture of the polarity changes of

    alkyl nitriles associated with the increase of pressure, which are shown to counteract the

    expected effect of viscosity.

    1.5 References [1] Breault, G. A.; Hunter, C. A. and Mayers, P. C. Supramolecular Topology. Tetrahedron 1999,

    55, 5265-5293.

    [2] Carina, R. F.; Dietrich-Buchecker, C. and Sauvage, J. P. Molecular Composite Knots. J. Am. Chem. Soc. 1996, 118, 9110-9116.

    [3] Brüggemann, J.; Bitter, S.; Müller, S.; Müller, W. M.; Müller, U.; Maier, N. M.; Lindner, W. and Vögtle, F. Spontaneons Knotting - from Oligoamide Threads to Trefoil Knots. Angew. Chem. Int. Edit. 2007, 46, 254-259.

  • Introduction

    23

    [4] Guo, J.; Mayers, P. C.; Breault, G. A. and Hunter, C. A. Synthesis of a Molecular Trefoil Knot by Folding and Closing on an Octahedral Coordination Template. Nat. Chem. 2010, 2, 218-222.

    [5] Lukin, O. and Vögtle, F. Knotting and Threading of Molecules: Chemistry and Chirality of Molecular Knots and Their Assemblies. Angew. Chem. Int. Ed. 2005, 44, 1456-1477.

    [6] Ibukuro, F.; Fujita, M.; Yamaguchi, K. and Sauvage, J. P. Quantitative and Spontaneous Formation of a Doubly Interlocking [2]Catenane Using Copper(I) and Palladium(II) as Templating and Assembling Centers. J. Am. Chem. Soc. 1999, 121, 11014-11015.

    [7] McArdle, C. P.; Vittal, J. J. and Puddephatt, R. J. Molecular Topology: Easy Self-Assembly of an Organometallic Doubly Braided [2]Catenane. Angew. Chem. Int. Edit. 2000, 39, 3819-3822.

    [8] Pentecost, C. D.; Chichak, K. S.; Peters, A. J.; Cave, G. W. V.; Cantrill, S. J. and Stoddart, J. F. A Molecular Solomon Link. Angew. Chem. Int. Edit. 2007, 46, 218-222.

    [9] Peinador, C.; Blanco, V. and Quintela, J. M. A New Doubly Interlocked [2]Catenane. J. Am. Chem. Soc. 2009, 131, 920-921.

    [10] Loren, J. C.; Yoshizawa, M.; Haldimann, R. F.; Linden, A. and Siegel, J. S. Synthetic Approaches to a Molecular Borromean Link: Two-Ring Threading with Polypyridine Templates. Angew. Chem. Int. Edit. 2003, 42, 5702-5705.

    [11] Cantrill, S. J.; Chichak, K. S.; Peters, A. J. and Stoddart, J. F. Nanoscale Borromean Rings. Acc. Chem. Res. 2005, 38, 1-9.

    [12] Dawson, R. E.; Lincoln, S. F. and Easton, C. J. The Foundation of a Light Driven Molecular Muscle Based on Stilbene and α-Cyclodextrin. Chem. Commun. 2008, 3980-3982.

    [13] Fang, L.; Hmadeh, M.; Wu, J. S.; Olson, M. A.; Spruell, J. M.; Trabolsi, A.; Yang, Y. W.; Elhabiri, M.; Albrecht-Gary, A. M. and Stoddart, J. F. Acid-Base Actuation of [c2]Daisy Chains. J. Am. Chem. Soc. 2009, 131, 7126-7134.

    [14] Clark, P. G.; Day, M. W. and Grubbs, R. H. Switching and Extension of a [c2]Daisy-Chain Dimer Polymer. J. Am. Chem. Soc. 2009, 131, 13631-13633.

    [15] Jiménez, M. C.; Dietrich-Buchecker, C. and Sauvage, J. P. Towards Synthetic Molecular Muscles: Contraction and Stretching of a Linear Rotaxane Dimer. Angew. Chem. Int. Edit. 2000, 39, 3284-3287.

    [16] Fuller, A. M. L.; Leigh, D. A. and Lusby, P. J. One Template, Multiple Rings: Controlled Iterative Addition of Macrocycles onto a Single Binding Site Rotaxane Thread. Angew. Chem. Int. Edit. 2007, 46, 5015-5019.

    [17] Wu, J.; Leung, K. C. F. and Stoddart, J. F. Efficient Production of [n]Rotaxanes by Using Template-Directed Clipping Reactions. Proc. Natl. Acad. Sci. U. S. A. 2007, 104, 17266-17271.

    [18] Harada, A.; Hashidzume, A.; Yamaguchi, H. and Takashima, Y. Polymeric Rotaxanes. Chem. Rev. 2009, 109, 5974-6023.

    [19] Klotz, E. J. F.; Claridge, T. D. W. and Anderson, H. L. Homo- and Hetero-[3]Rotaxanes with Two π-Systems Clasped in a Single Macrocycle. J. Am. Chem. Soc. 2006, 128, 15374-15375.

    [20] Prikhod'ko, A. I. and Sauvage, J. P. Passing Two Strings through the Same Ring Using an Octahedral Metal Center as Template: A New Synthesis of [3]Rotaxanes. J. Am. Chem. Soc. 2009, 131, 6794-6807.

    [21] Lee, C. F.; Leigh, D. A.; Pritchard, R. G.; Schultz, D.; Teat, S. J.; Timco, G. A. and Winpenny, R. E. P. Hybrid Organic-Inorganic Rotaxanes and Molecular Shuttles. Nature 2009, 458, 314-318.

  • C h a p t e r 1

    24

    [22] Goldup, S. M.; Leigh, D. A.; McGonigal, P. R.; Ronaldson, V. E. and Slawin, A. M. Z. Two Axles Threaded Using a Single Template Site: Active Metal Template Macrobicyclic [3]Rotaxanes. J. Am. Chem. Soc. 2010, 132, 315-320.

    [23] Raymo, F. M. and Stoddart, J. F. Slippage - a Simple and Efficient Way to Self-Assemble [n]Rotaxanes. Pure Appl. Chem. 1997, 69, 1987-1997.

    [24] Lee, J. J.; White, A. G.; Baumes, J. M. and Smith, B. D. Microwave-Assisted Slipping Synthesis of Fluorescent Squaraine Rotaxane Probe for Bacterial Imaging. Chem. Commun. 2009, 46, 1068-1069.

    [25] Tokunaga, Y.; Wakamatsu, N.; Ohbayashi, A.; Akasaka, K.; Saeki, S.; Hisada, K.; Goda, T. and Shimomura, Y. The Effect of Pressure on Pseudorotaxane Formation by Using the Slipping Method. Tetrahedron Lett. 2006, 47, 2679-2682.

    [26] Harrison, I. T. and Harrison, S. Synthesis of a Stable Complex of a Macrocycle and a Threaded Chain. J. Am. Chem. Soc. 1967, 89, 5723-5724.

    [27] Vögtle, F.; Händel, M.; Meier, S.; Ottens-Hildebrandt, S.; Ott, F. and Schmidt, T. Template Synthesis of the First Amide-Based Rotaxanes. Liebigs Ann. 1995, 739-743.

    [28] Hübner, G. M.; Gläser, J.; Seel, C. and Vögtle, F. High-Yielding Rotaxane Synthesis with an Anion Template. Angew. Chem. Int. Edit. 1999, 38, 383-386.

    [29] Seel, C. and Vögtle, F. Templates, "Wheeled Reagents", and a New Route to Rotaxanes by Anion Complexation: The Trapping Method. Chem. Eur. J. 2000, 6, 21-24.

    [30] Johnston, A. G.; Leigh, D. A.; Murphy, A.; Smart, J. P. and Deegan, M. D. The Synthesis and Solubilization of Amide Macrocycles Via Rotaxane Formation. J. Am. Chem. Soc. 1996, 118, 10662-10663.

    [31] Leigh, D. A.; Murphy, A.; Smart, J. P. and Slawin, A. M. Z. Glycylglycine Rotaxanes - the Hydrogen Bond Directed Assembly of Synthetic Peptide Rotaxanes. Angew. Chem. Int. Edit. Engl. 1997, 36, 728-732.

    [32] Clegg, W.; Gimenez-Saiz, C.; Leigh, D. A.; Murphy, A.; Slawin, A. M. Z. and Teat, S. J. "Smart" Rotaxanes: Shape Memory and Control in Tertiary Amide Peptido[2]Rotaxanes. J. Am. Chem. Soc. 1999, 121, 4124-4129.

    [33] Bottari, G.; Dehez, F.; Leigh, D. A.; Nash, P. J.; Pérez, E. M.; Wong, J. K. Y. and Zerbetto, F. Entropy-Driven Translational Isomerism: A Tristable Molecular Shuttle. Angew. Chem. Int. Edit. 2003, 42, 5886-5889.

    [34] Arunkumar, E.; Forbes, C. C.; Noll, B. C. and Smith, B. D. Squaraine-Derived Rotaxanes: Sterically Protected Fluorescent near-IR Dyes. J. Am. Chem. Soc. 2005, 127, 3288-3289.

    [35] Fu, N.; Baumes, J. M.; Arunkumar, E.; Noll, B. C. and Smith, B. D. Squaraine Rotaxanes with Boat Conformation Macrocycles. J. Org. Chem. 2009, 74, 6462-6468.

    [36] Cecchet, F.; Rudolf, P.; Rapino, S.; Margotti, M.; Paolucci, F.; Baggerman, J.; Brouwer, A. M.; Kay, E. R.; Wong, J. K. Y. and Leigh, D. A. Structural, Electrochemical, and Photophysical Properties of a Molecular Shuttle Attached to an Acid-Terminated Self-Assembled Monolayer. J. Phys. Chem. B 2004, 108, 15192-15199.

    [37] Leigh, D. A.; Morales, M. A. F.; Pérez, E. M.; Wong, J. K. Y.; Saiz, C. G.; Slawin, A. M. Z.; Carmichael, A. J.; Haddleton, D. M.; Brouwer, A. M.; Buma, W. J.; Wurpel, G. W. H.; Leon, S. and Zerbetto, F. Patterning through Controlled Submolecular Motion: Rotaxane-Based Switches and Logic Gates That Function in Solution and Polymer Films. Angew. Chem. Int. Edit. 2005, 44, 3062-3067.

    [38] Bermudez, V.; Gase, T.; Kajzar, F.; Capron, N.; Zerbetto, F.; Gatti, F. G.; Leigh, D. A. and Zhang, S. Rotaxanes - Novel Photonic Molecules. Opt. Mater. 2003, 21, 39-44.

  • Introduction

    25

    [39] Altieri, A.; Bottari, G.; Dehez, F.; Leigh, D. A.; Wong, J. K. Y. and Zerbetto, F. Remarkable Positional Discrimination in Bistable Light- and Heat-Switchable Hydrogen-Bonded Molecular Shuttles. Angew. Chem. Int. Edit. 2003, 42, 2296-2300.

    [40] Berná, J.; Brouwer, A. M.; Fazio, S. M.; Haraszkiewicz, N.; Leigh, D. A. and Lennon, C. M. A Rotaxane Mimic of the Photoactive Yellow Protein Chromophore Environment: Effects of Hydrogen Bonding and Mechanical Interlocking on a Coumaric Amide Derivative. Chem. Commun. 2007, 1910-1912.

    [41] Marlin, D. S.; Cabrera, D. G.; Leigh, D. A. and Slawin, A. M. Z. Complexation-Induced Translational Isomerism: Shuttling through Stepwise Competitive Binding. Angew. Chem. Int. Edit. 2006, 45, 77-83.

    [42] Gatti, F. G.; Leigh, D. A.; Nepogodiev, S. A.; Slawin, A. M. Z.; Teat, S. J. and Wong, J. K. Y. Stiff, and Sticky in the Right Places: The Dramatic Influence of Preorganizing Guest Binding Sites on the Hydrogen Bond-Directed Assembly of Rotaxanes. J. Am. Chem. Soc. 2001, 123, 5983-5989.

    [43] Kidd, T. J.; Loontjens, T. J. A.; Leigh, D. A. and Wong, J. K. Y. Rotaxane Building Blocks Bearing Blocked Isocyanate Stoppers: Polyrotaxanes through Post-Assembly Chain Extension. Angew. Chem. Int. Edit. 2003, 42, 3379-3383.

    [44] Alvarez-Pérez, M.; Goldup, S. M.; Leigh, D. A. and Slawin, A. M. Z. A Chemically-Driven Molecular Information Ratchet. J. Am. Chem. Soc. 2008, 130, 1836-1838.

    [45] Moretto, A.; Menegazzo, I.; Crisma, M.; Shotton, E. J.; Nowell, H.; Mammi, S. and Toniolo, C. A Rigid Helical Peptide Axle for a [2]Rotaxane Molecular Machine. Angew. Chem. Int. Edit. 2009, 48, 8986-8989.

    [46] Wurpel, G. W. H.; Brouwer, A. M.; van Stokkum, I. H. M.; Farran, A. and Leigh, D. A. Enhanced Hydrogen Bonding Induced by Optical Excitation: Unexpected Subnanosecond Photoinduced Dynamics in a Peptide-Based [2]Rotaxane. J. Am. Chem. Soc. 2001, 123, 11327-11328.

    [47] Da Ros, T.; Guldi, D. M.; Morales, A. F.; Leigh, D. A.; Prato, M. and Turco, R. Hydrogen Bond-Assembled Fullerene Molecular Shuttle. Org. Lett. 2003, 5, 689-691.

    [48] Keaveney, C. M. and Leigh, D. A. Shuttling through Anion Recognition. Angew. Chem. Int. Edit. 2004, 43, 1222-1224.

    [49] Pérez, E. M.; Dryden, D. T. F.; Leigh, D. A.; Teobaldi, G. and Zerbetto, F. A Generic Basis for Some Simple Light-Operated Mechanical Molecular Machines. J. Am. Chem. Soc. 2004, 126, 12210-12211.

    [50] Marlin, D. S.; Cabrera, D. G.; Leigh, D. A. and Slawin, A. M. Z. An Allosterically Regulated Molecular Shuttle. Angew. Chem. Int. Edit. 2006, 45, 1385-1390.

    [51] Gadret, G.; Zamboni, R.; Schouwink, P.; Mahrt, R. F.; Thies, J.; Loontjens, T. and Leigh, D. A. Solid-State Optical Properties of the Methyl-Exopyridine-Anthracene Rotaxane. Chem. Phys. 2001, 269, 381-388.

    [52] Bermudez, V.; Capron, N.; Gase, T.; Gatti, F. G.; Kajzar, F.; Leigh, D. A.; Zerbetto, F. and Zhang, S. W. Influencing Intramolecular Motion with an Alternating Electric Field. Nature 2000, 406, 608-611.

    [53] Xiao, S. Z.; Fu, N.; Peckham, K. and Smith, B. D. Efficient Synthesis of Fluorescent Squaraine Rotaxane Dendrimers. Org. Lett. 2010, 12, 140-143.

    [54] Mateo-Alonso, A.; Brough, P. and Prato, M. Stabilization of Fulleropyrrolidine N-Oxides through Intrarotaxane Hydrogen Bonding. Chem. Commun. 2007, 1412-1414.

  • C h a p t e r 1

    26

    [55] Rijs, A. M.; Crews, B. O.; de Vries, M. S.; Hannam, J. S.; Leigh, D. A.; Fanti, M.; Zerbetto, F. and Buma, W. J. Shaping of a Conformationally Flexible Molecular Structure for Spectroscopy. Angew. Chem. Int. Edit. 2008, 47, 3174-3179.

    [56] Brancato, G.; Coutrot, F.; Leigh, D. A.; Murphy, A.; Wong, J. K. Y. and Zerbetto, F. From Reactants to Products Via Simple Hydrogen-Bonding Networks: Information Transmission in Chemical Reactions. Proc. Natl. Acad. Sci. U. S. A. 2002, 99, 4967-4971.

    [57] Altieri, A.; Gatti, F. G.; Kay, E. R.; Leigh, D. A.; Martel, D.; Paolucci, F.; Slawin, A. M. Z. and Wong, J. K. Y. Electrochemically Switchable Hydrogen-Bonded Molecular Shuttles. J. Am. Chem. Soc. 2003, 125, 8644-8654.

    [58] Hirose, K.; Shiba, Y.; Ishibashi, K.; Doi, Y. and Tobe, Y. A Shuttling Molecular Machine with Reversible Brake Function. Chem. Eur. J. 2008, 14, 3427-3433.

    [59] Zhou, W. D.; Li, J. B.; He, X. R.; Li, C. H.; Lv, J.; Li, Y. L.; Wang, S.; Liu, H. B. and Zhu, D. B. A Molecular Shuttle for Driving a Multilevel Fluorescence Switch. Chem. Eur. J. 2008, 14, 754-763.

    [60] Chuang, C. J.; Li, W. S.; Lai, C. C.; Liu, Y. H.; Peng, S. M.; Chao, I. and Chiu, S. H. A Molecular Cage-Based [2]Rotaxane That Behaves as a Molecular Muscle. Org. Lett. 2009, 11, 385-388.

    [61] Zheng, H. Y.; Zhou, W. D.; Lv, J.; Yin, X. D.; Li, Y. J.; Liu, H. B. and Li, Y. L. A Dual-Response [2]Rotaxane Based on a 1,2,3-Triazole Ring as a Novel Recognition Station. Chem. Eur. J. 2009, 15, 13253-13262.

    [62] Vella, S. J.; Tiburcio, J. and Loeb, S. J. Optically Sensed, Molecular Shuttles Driven by Acid-Base Chemistry. Chem. Commun. 2007, 4752-4754.

    [63] Leigh, D. A. and Thomson, A. R. An Ammonium/Bis-Ammonium Switchable Molecular Shuttle. Tetrahedron 2008, 64, 8411-8416.

    [64] Ashton, P. R.; Bissell, R. A.; Spencer, N.; Stoddart, J. F. and Tolley, M. S. Towards Controllable Molecular Shuttles. 1. Syn. Lett. 1992, 914-918.

    [65] Anelli, P. L.; Asakawa, M.; Ashton, P. R.; Bissell, R. A.; Clavier, G.; Górski, R.; Kaifer, A. E.; Langford, S. J.; Mattersteig, G.; Menzer, S.; Philp, D.; Slawin, A. M. Z.; Spencer, N.; Stoddart, J. F.; Tolley, M. S. and Williams, D. J. Toward Controllable Molecular Shuttles. Chem. Eur. J. 1997, 3, 1113-1135.

    [66] Liu, Y.; Flood, A. H.; Bonvallett, P. A.; Vignon, S. A.; Northrop, B. H.; Tseng, H. R.; Jeppesen, J. O.; Huang, T. J.; Brough, B.; Baller, M.; Magonov, S.; Solares, S. D.; Goddard, W. A.; Ho, C. M. and Stoddart, J. F. Linear Artificial Molecular Muscles. J. Am. Chem. Soc. 2005, 127, 9745-9759.

    [67] Spruell, J. M.; Dichtel, W. R.; Heath, J. R. and Stoddart, J. F. A One-Pot Synthesis of Constitutionally Unsymmetrical Rotaxanes Using Sequential CuI-Catalyzed Azide-Alkyne Cycloadditions. Chem. Eur. J. 2008, 14, 4168-4177.

    [68] Zheng, Y. B.; Yang, Y. W.; Jensen, L.; Fang, L.; Juluri, B. K.; Flood, A. H.; Weiss, P. S.; Stoddart, J. F. and Huang, T. J. Active Molecular Plasmonics: Controlling Plasmon Resonances with Molecular Switches. Nano Lett. 2009, 9, 819-825.

    [69] Yoon, I.; Benítez, D.; Zhao, Y. L.; Miljanić, O. S.; Kim, S. Y.; Tkatchouk, E.; Leung, K. C. F.; Khan, S. I.; Goddard, W. A. and Stoddart, J. F. Functionally Rigid and Degenerate Molecular Shuttles. Chem. Eur. J. 2009, 15, 1115-1122.

    [70] Trabolsi, A.; Khashab, N.; Fahrenbach, A. C.; Friedman, D. C.; Colvin, M. T.; Cotí, K. K.; Benítez, D.; Tkatchouk, E.; Olsen, J. C.; Belowich, M. E.; Carmielli, R.; Khatib, H. A.; Goddard, W. A.; Wasielewski, M. R. and Stoddart, J. F. Radically Enhanced Molecular Recognition. Nat. Chem. 2010, 2, 42-49.

  • Introduction

    27

    [71] Trabolsi, A.; Fahrenbach, A. C.; Dey, S. K.; Share, A. I.; Friedman, D. C.; Basu, S.; Gasa, T. B.; Khashab, N. M.; Saha, S.; Aprahamian, I.; Khatib, H. A.; Flood, A. H. and Stoddart, J. F. A Tristable [2]Pseudo[2]Rotaxane. Chem. Commun. 2010, 46, 871-873.

    [72] Yasuda, T.; Tanabe, K.; Tsuji, T.; Coti, K. K.; Aprahamian, I.; Stoddart, J. F. and Kato, T. A Redox-Switchable [2]Rotaxane in a Liquid-Crystalline State. Chem. Commun. 2010, 46, 1224-1226.

    [73] Ashton, P. R.; Bissell, R. A.; Spencer, N.; Stoddart, J. F. and Tolley, M. S. Towards Controllable Molecular Shuttles. 3. Syn. Lett. 1992, 923-926.

    [74] Luo, Y.; Collier, C. P.; Jeppesen, J. O.; Nielsen, K. A.; DeIonno, E.; Ho, G.; Perkins, J.; Tseng, H. R.; Yamamoto, T.; Stoddart, J. F. and Heath, J. R. Two-Dimensional Molecular Electronics Circuits. ChemPhysChem 2002, 3, 519-525.

    [75] Choi, J. W.; Flood, A. H.; Steuerman, D. W.; Nygaard, S.; Braunschweig, A. B.; Moonen, N. N. P.; Laursen, B. W.; Luo, Y.; DeIonno, E.; Peters, A. J.; Jeppesen, J. O.; Xu, K.; Stoddart, J. F. and Heath, J. R. Ground-State Equilibrium Thermodynamics and Switching Kinetics of Bistable 2 Rotaxanes Switched in Solution, Polymer Gels, and Molecular Electronic Devices. 2006, 12, 261-279.

    [76] Ashton, P. R.; Ballardini, R.; Balzani, V.; Baxter, I.; Credi, A.; Fyfe, M. C. T.; Gandolfi, M. T.; Gómez-López, M.; Martínez-Díaz, M. V.; Piersanti, A.; Spencer, N.; Stoddart, J. F.; Venturi, M.; White, A. J. P. and Williams, D. J. Acid-Base Controllable Molecular Shuttles. J. Am. Chem. Soc. 1998, 120, 11932-11942.

    [77] Badjić, J. D.; Balzani, V.; Credi, A.; Silvi, S. and Stoddart, J. F. A Molecular Elevator. Science 2004, 303, 1845-1849.

    [78] Badjić, J. D.; Ronconi, C. M.; Stoddart, J. F.; Balzani, V.; Silvi, S. and Credi, A. Operating Molecular Elevators. J. Am. Chem. Soc. 2006, 128, 1489-1499.

    [79] Nikitin, K.; Lestini, E.; Lazzari, M.; Altobello, S. and Fitzmaurice, D. A Tripodal [2]Rotaxane on the Surface of Gold. Langmuir 2007, 23, 12147-12153.

    [80] Ikeda, T.; Higuchi, M. and Kurth, D. G. From Thiophene [2]Rotaxane to Polythiophene Polyrotaxane. J. Am. Chem. Soc. 2009, 131, 9158-9159.

    [81] Córdova, E.; Bissell, R. A.; Spencer, N.; Ashton, P. R.; Stoddart, J. F. and Kaifer, A. E. Novel Rotaxanes Based on the Inclusion Complexation of Biphenyl Guests by Cyclobis(Paraquat-p-Phenylene). J. Org. Chem. 1993, 58, 6550-6552.

    [82] Bissell, R. A.; Cordova, E.; Kaifer, A. E. and Stoddart, J. F. A Chemically and Electrochemically Switchable Molecular Shuttle. Nature 1994, 369, 133-137.

    [83] Ashton, P. R.; Bissell, R. A.; Górski, R.; Philp, D.; Spencer, N.; Stoddart, J. F. and Tolley, M. S. Towards Controllable Molecular Shuttles. 2. Syn. Lett. 1992, 919-922.

    [84] Balzani, V.; Credi, A.; Marchioni, F. and Stoddart, J. F. Artificial Molecular-Level Machines. Dethreading-Rethreading of a Pseudorotaxane Powered Exclusively by Light Energy. Chem. Commun. 2001, 1860-1861.

    [85] Katz, E.; Sheeney-Haj-Ichia, L. and Willner, I. Electrical Contacting of Glucose Oxidase in a Redox-Active Rotaxane Configuration. Angew. Chem. Int. Edit. 2004, 43, 3292-3300.

    [86] Katz, E.; Baron, R.; Willner, I.; Richke, N. and Levine, R. D. Temperature-Dependent and Friction-Controlled Electrochemically Induced Shuttling Along Molecular Strings Associated with Electrodes. ChemPhysChem 2005, 6, 2179-2189.

    [87] Abraham, W.; Grubert, L.; Schmidt-Schäffer, S. and Buck, K. Pseudorotaxanes and Rotaxanes Incorporating Diarylcycloheptatriene Stations. Eur. J. Org. Chem. 2005, 374-386.

  • C h a p t e r 1

    28

    [88] Schmidt-Schäffer, S.; Grubert, L.; Grummt, U. W.; Buck, K. and Abraham, W. A Photoswitchable Rotaxane with an Unfolded Molecular Thread. Eur. J. Org. Chem. 2006, 378-398.