12
University of Groningen Nucleic acid amphiphiles Kwak, Minseok; Herrmann, Andreas; Clever, Guido; Mao, Chengde; Shionoya, Mitsuhiko; Stulz, Eugen Published in: Chemical Society Reviews DOI: 10.1039/c1cs15138j IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite from it. Please check the document version below. Document Version Publisher's PDF, also known as Version of record Publication date: 2011 Link to publication in University of Groningen/UMCG research database Citation for published version (APA): Kwak, M., Herrmann, A., Clever, G. (Ed.), Mao, C. (Ed.), Shionoya, M. (Ed.), & Stulz, E. (Ed.) (2011). Nucleic acid amphiphiles: synthesis and self-assembled nanostructures. Chemical Society Reviews, 40(12), 5745-5755. DOI: 10.1039/c1cs15138j Copyright Other than for strictly personal use, 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), unless the work is under an open content license (like Creative Commons). Take-down policy If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim. Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons the number of authors shown on this cover page is limited to 10 maximum. Download date: 10-02-2018

University of Groningen Nucleic acid amphiphiles Kwak ... · Nucleic acid amphiphiles: synthesis and self-assembled nanostructures w Minseok Kwak and Andreas Herrmann* Received 24th

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

  • University of Groningen

    Nucleic acid amphiphilesKwak, Minseok; Herrmann, Andreas; Clever, Guido; Mao, Chengde; Shionoya, Mitsuhiko;Stulz, EugenPublished in:Chemical Society Reviews

    DOI:10.1039/c1cs15138j

    IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite fromit. Please check the document version below.

    Document VersionPublisher's PDF, also known as Version of record

    Publication date:2011

    Link to publication in University of Groningen/UMCG research database

    Citation for published version (APA):Kwak, M., Herrmann, A., Clever, G. (Ed.), Mao, C. (Ed.), Shionoya, M. (Ed.), & Stulz, E. (Ed.) (2011).Nucleic acid amphiphiles: synthesis and self-assembled nanostructures. Chemical Society Reviews,40(12), 5745-5755. DOI: 10.1039/c1cs15138j

    CopyrightOther than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of theauthor(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons).

    Take-down policyIf you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediatelyand investigate your claim.

    Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons thenumber of authors shown on this cover page is limited to 10 maximum.

    Download date: 10-02-2018

    http://dx.doi.org/10.1039/c1cs15138jhttps://www.rug.nl/research/portal/en/publications/nucleic-acid-amphiphiles(8e89625b-4b04-418b-988c-c5877789a0b3).html

  • This journal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev., 2011, 40, 5745–5755 5745

    Cite this: Chem. Soc. Rev., 2011, 40, 5745–5755

    Nucleic acid amphiphiles: synthesis and self-assembled nanostructuresw

    Minseok Kwak and Andreas Herrmann*

    Received 24th May 2011

    DOI: 10.1039/c1cs15138j

    This review provides an overview of a relatively new class of bio-conjugates, DNA amphiphiles,

    which consist of oligonucleotides covalently bonded to synthetic hydrophobic units. The reader

    will find the basic principles for the structural design and preparation methods of the materials.

    Moreover, the self-assembly into superstructures of higher order will be highlighted. Finally, some

    potential applications will be described.

    1 Introduction

    Five decades have passed since the discovery of the structure

    of deoxyribonucleic acids (DNAs), ‘‘the secret of life’’, by

    Watson, Crick and coworkers, key among them Franklin. This

    elegant double helix has ever since fascinated researchers

    throughout the natural and applied sciences. Thanks to the

    invention of chemical DNA synthesis in the 1960s, broader

    studies involving nucleic acids became possible, not only

    in biological and medical fields but also beyond. From a chemical

    point of view, the driving interest of DNA is its programmability

    and self-recognition properties. These properties have been

    utilized to particularly great effect in materials science to build

    programmed architectures. As such, it is now recognized that

    DNA is ‘‘not merely the secret of life’’.1

    DNA nanotechnology is now a common term to describe

    the use of DNA as a structural material rather than as a

    medium for genetic information. Since Seeman’s artistic

    Department of Polymer Chemistry, Zernike Institute for AdvancedMaterials, University of Groningen, Nijenborgh 4, 9747 AGGroningen, The Netherlands. E-mail: [email protected];Fax: +31 50 363 4400; Tel: +31 50 363 6318w Part of a themed issue on the advances in DNA-based nano-technology.

    Minseok Kwak

    Minseok Kwak studied

    chemistry at the Ajou Univer-

    sity in Suwon (South Korea)

    and received his Master’s

    degree in molecular science

    and technology in 2006. He

    started his PhD studies at the

    Max-Planck Institute for

    Polymer Research in Mainz

    (Germany) under the supervi-

    sion of Professor A. Herrmann

    and Professor K. Müllen. In

    2007, he moved to the Zernike

    Institute for Advanced Materials

    with the group of Andreas

    Herrmann and obtained his

    PhD in 2011 from the University of Groningen. He is currently

    employed as a postdoctoral researcher in the group of Professor

    W. M. Shih at Harvard Medical School in Boston with a Rubicon

    fellowship from the Netherlands Organisation for Scientific

    Research (NWO).

    Andreas Herrmann

    Andreas Herrmann studied

    chemistry at the University of

    Mainz (Germany). From 1997

    to 2000 he pursued his graduate

    studies at the Max-Planck

    Institute for Polymer Research

    in the group of Professor

    K. Müllen. Then he was

    employed as a consultant for

    Roland Berger Management

    Consultants in Munich (2001).

    In the years 2002 and 2003 he

    returned to academia working

    on protein engineering at the

    Swiss Federal Institute of Tech-

    nology, Zurich, with Professor

    D. Hilvert. In 2004 he was appointed as a head of a junior

    research group at the Max-Planck Institute for Polymer

    Research. In 2007 he moved to the Zernike Institute

    for Advanced Materials at the University of Groningen,

    The Netherlands, where he holds a chair for Polymer Chemistry

    and Bioengineering. Andreas Herrmann was awarded the

    Reimund-Stadler Prize (2008) and the Dr Hermann-Schnell

    Prize (2009) from the German Chemical Society (GDCh).

    Moreover, in 2009 and 2010, he received prestigious funds, the

    ERC Starting Grant from the European Commission and the

    Vici grant from NWO, respectively.

    Chem Soc Rev Dynamic Article Links

    www.rsc.org/csr TUTORIAL REVIEW

    Dow

    nloa

    ded

    by U

    nive

    rsit

    y of

    Gro

    ning

    en o

    n 07

    Feb

    ruar

    y 20

    12

    Pub

    lish

    ed o

    n 22

    Aug

    ust

    2011

    on

    http

    ://p

    ubs.

    rsc.

    org

    | doi

    :10.

    1039

    /C1C

    S15

    138J

  • 5746 Chem. Soc. Rev., 2011, 40, 5745–5755 This journal is c The Royal Society of Chemistry 2011

    building strategy for unusual DNA structures was reported

    in 1991,2 countless beautiful DNA architectures have been

    described—from intricate 2D patterns to highly precise nano-

    scopic tubes, convex polyhedra, strings and 3D origami. These

    new geometries, the increasingly sophisticated understanding

    of assembly mechanisms they generated and the opportunities

    for further manipulation which they revealed opened the door

    to working with this material in multidimensional systems.3–8

    The key structural property that makes these designs possible

    is the persistence length of double-stranded (ds) DNA of 50 nm.

    Qualifying it as a semi-rigid polymer, the double helix serves as a

    rigid scaffold for short nucleic acid stretches.

    At the same time, attempts have been made to develop

    applications for DNA by incorporating functionalities and

    utilizing its information-carrying capability. For instance,

    fluorescently labeled DNAs can be used to detect particular

    sequences or monitor the polymerase chain reaction (PCR).9

    Today a range of modifications are commercially available for

    daily laboratory use: chemical functionalities for covalent-

    bond formation, optical/fluorescent labels for detection, and

    others for further molecular recognition. While the above

    modifications simply introduce an additional function, a

    particular focus in materials science has been to chemically

    couple nucleic acid strands with new moieties that alter the

    molecule’s structural, morphological or self-assembly properties.

    Such combinations of DNA with other classes of materials,

    including chemical functional groups, (bio)polymers and

    inorganic nanoobjects, are known as DNA hybrid materials.

    This class of materials finds application in a wide range of

    research fields, for instance diagnostics and electronics, as the

    properties and advantages of each component can be tailored

    to meet particular needs. This has been well demonstrated

    and reviewed for hybrids incorporating gold nanoparticles10

    and proteins.11

    In this review, the class of amphiphilic DNA hybrids will be

    covered in depth. Their amphiphilic nature arises from the

    hydrophilicity of the DNA backbone containing charged

    phosphodiester bonds, and chemical attachment of hydro-

    phobic moieties. Here, amphiphilic structures with only a single

    or very short nucleotides will be excluded. Such materials most

    often lack the property of predetermined DNA hybridization

    so that rarely more complex superstructures are formed. This

    class of modified nucleic acids is covered in an excellent

    review.12 First, the classification and techniques for the syn-

    thesis of amphiphilic oligonucleotides (AONs) will be intro-

    duced. Then their self-assembled structures and potential

    applications will be described.

    2 Structural classification and synthesis

    Solid-phase synthesis (SPS) and solution coupling are two

    general ways for the preparation of DNA (or RNA) modified

    with functional groups.13–15 Short up to long fragments of

    DNA (up to 200 mers), called oligonucleotides (ONs), can be

    built up by iterative chemical synthesis on solid supports using

    activated phosphoramidite building blocks. The four-step

    cycle includes: (1) deblocking, (2) coupling, (3) capping and

    (4) oxidation.16 The SPS technique makes it possible not only

    to prepare pristine ONs but also to introduce functional

    groups while the ONs are still present on the support. Solution

    coupling is a more widely used method to make covalent

    bonds between reaction-ready DNAs (e.g. amine- or thiol-

    modified, prepared by SPS) that are available from commer-

    cial suppliers and the complementary functionalities (e.g.

    carboxylic acid, thiol or maleimide) of the target molecules.

    However, solution coupling with hydrophobic molecules often

    results in low yields due to the solvent incompatibility of the

    two components. For this reason, solid-phase oligonucleotide

    synthesis, which affords a wide choice of organic solvents and

    restricts reactions to the solid supports, is more suitable for

    preparing AONs. SPS methods can be further divided into two

    classes: (1) fully automated synthesis within an ON synthesizer

    and (2) modified techniques including syringe synthesis and

    in-flask reactions utilizing DNA as synthesized on solid

    supports. Fully automated synthesis offers precise control

    and monitoring of the entire procedure, relatively large scales

    and higher reproducibility despite of a few disadvantages such

    as a high level of expertise and high investment costs for equip-

    ment. Often the modified techniques can be used to mitigate some

    of these difficulties, as well as to meet more specialized needs (i.e.

    longer reaction times or solvents incompatible with a synthesizer).

    This review will focus on the use of SPS methods to prepare

    two structural classes of AONs. The first architecture contains

    terminal hydrophobic moieties while the second type of structures

    consists of a brush-configuration.

    Terminally-functionalized AONs

    Solution coupling has been used to prepare amphiphilic

    DNAs with hydrophobic molecules at either terminus of the

    DNA. While yields can often be low for the reasons discussed

    above, the introduction of an oligoethyleneglycol linker

    between the hydrophobic moiety and the DNA enhances

    solvent compatibility, giving moderate coupling yields.15 SPS

    represents the method of choice for the preparation of these

    materials. Unlike solution coupling, though, the incorporation

    of small hydrophobic moieties does not notably lower the

    coupling yield. One classic and essential method to prepare

    AONs using SPS with the hydrophobic moiety covalently

    connected to the 50-end of the ONs is the use of 2-cyanoethyl-

    N,N-diisopropylphosphoramidite (CEPA) groups.17 Here, a

    hydrophobic molecule is functionalized with CEPA groups

    and subsequently applied onto the detritylated 50-end of ONs

    grown on solid supports (Fig. 1A). Introducing a hydrophobe

    at the 30-end of the DNA requires different strategies. One

    technique necessitates starting with custom solid supports which

    already contain the desired hydrophobic moieties, connected via

    a cleavable linker (e.g. carboxylic ester) (Fig. 1B).15 The other

    method is reverse synthesis, in which the DNA sequence is

    elongated in the opposite (i.e. from the 50- to 30-end) direction

    with a final addition of the desired hydrophobe at the 30-end.

    In this case, the CEPA group is located at the 50-end and the

    4,40-dimethoxytrityl (DMT) group at the 30-end of each nucleo-

    side, as opposed to conventional DNA synthesis (compare the

    structures shown in Fig. 1B). Fig. 1C portrays a range of CEPA

    compounds to introduce hydrophobicity for specific purposes.

    In simple hydrophobic functionalities, linear lipophilic (or

    aliphatic) chains are directly connected to CEPA (1 and 2)18,19

    Dow

    nloa

    ded

    by U

    nive

    rsit

    y of

    Gro

    ning

    en o

    n 07

    Feb

    ruar

    y 20

    12

    Pub

    lish

    ed o

    n 22

    Aug

    ust

    2011

    on

    http

    ://p

    ubs.

    rsc.

    org

    | doi

    :10.

    1039

    /C1C

    S15

    138J

  • This journal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev., 2011, 40, 5745–5755 5747

    Fig. 1 Synthetic schemes and precursors of DNA amphiphiles with terminal hydrophobic moieties. General solid-phase synthesis for amphiphilic

    DNAs with hydrophobes at (A) 50- and (B) 30-end of the DNA; (a) deblocking of DMT; (b) coupling of activated CEPA to 50-end; (c) standard

    synthesis with nucleoside phosphoramidite. After the synthesis, cleavage from the solid supports and removal of the 2-cyanoethyl group

    on phosphorus yield amphiphilic DNAs. (C) Chemical structures of small hydrophobic compounds modified as CEPA. (D) Hydrophobic

    CEPA-functionalized polymers. (E) Graphical representations of AONs with a hydrophobic functionality and polymer (DNA block copolymer),

    left and right, respectively, at a terminus of DNA. Note that primary amines (–NH2) at nucleobases (denoted as N) are accordingly protected and

    CEPA will be the linker to a DNA in AON products.

    Dow

    nloa

    ded

    by U

    nive

    rsit

    y of

    Gro

    ning

    en o

    n 07

    Feb

    ruar

    y 20

    12

    Pub

    lish

    ed o

    n 22

    Aug

    ust

    2011

    on

    http

    ://p

    ubs.

    rsc.

    org

    | doi

    :10.

    1039

    /C1C

    S15

    138J

  • 5748 Chem. Soc. Rev., 2011, 40, 5745–5755 This journal is c The Royal Society of Chemistry 2011

    or attached to the nucleobases of the phosphoramidite building

    blocks (3 and 4).20,21 Hydrophobic moieties such as the fluorous

    DMT group (FDMT) of 3 offer the additional advantage of fast,

    highly efficient purification using a fluorous affinity column or

    cartridge. In a similar synthetic manner, cholesterol derivatives

    (5, 6 and 7)22 are introduced into DNA as terminal chain motifs.

    For the synthesis of 30-end conjugated AONs, compounds 7a

    and 7b are attached to the solid support via an ester bond at the

    hydroxy groups. A useful means to tune the degree of hydro-

    phobicity in these materials is by incorporating bis-adducts of

    the hydrophobe connected to CEPA or altering the length of the

    linear hydrophobic chain.23

    Due to the high charge density of nucleic acids, it is often

    desirable in AONs to have relatively large hydrophobic seg-

    ments rather than small moieties (compare Fig. 1E, especially

    when self-assembled nanostructures should be achieved.

    Linear DNA block copolymers (DBCs) are good candidates

    for attaining balanced amphiphilicity by changing the mole-

    cular weight (Mw) of the polymer segments. In general, DBCs

    consist of an ON covalently connected at either end to a

    hydrophobic polymer or a block copolymer, yielding amphi-

    philic diblock (Fig. 1E, right image) or multiblock structures.

    Since Jeong and Park generated a polymeric AON with

    poly(D,L-lactic-co-glycolic acid) (PLGA) as the hydrophobic

    block using solution coupling,24 analogous routes were com-

    monly used. Note that in these solution based methods, there

    are always two or more reacting groups for the coupling of the

    hydrophobes to DNA. Moreover, the coupling of highly

    hydrophobic polymers to ONs entails very low yields due to

    the incompatibility of solvents for hydrophobes and DNA.

    To overcome solvent compatibility limitations on the choice of

    polymers, solid-phase synthesis was also adopted to prepare

    DBCs, yielding AONs with more hydrophobic polymers like

    poly(styrene) (PS) (8).25 Here, the hydroxy-terminal group of

    PS was converted to 2-cyanoethyl-N,N-diisopropylphosphor-

    amidite (CEPA) and was subsequently applied onto the

    detritylated 50-end of elongated ONs on solid supports

    (Fig. 1A and D). Nowadays, fully automated SPS is optimized

    and thus commonly used for large-scale preparation of DBCs

    in DNA synthesizers with several different hydrophobic

    polymers including poly(propyleneoxide) (PPO) (9)26 and

    poly(9,9-di-n-octylfluorene) (PFO) (10) (Fig. 1D and E, right

    image).27 The reader is directed to two recent reviews for more

    comprehensive information about other polymer hybrids of

    biomolecules, including DNA.28,29

    Brush-type AONs

    Brush-type is a structural term for macromolecules consisting

    of a linear backbone onto which multiple oligomer or polymer

    segments are grafted. Such macromolecular architectures

    containing DNA can be generated through SPS. As shown

    in the general scheme of DNA synthesis (Fig. 1B), the presence

    of CEPA and DMT groups is critical for performing the

    reaction cycle. Thus both groups need to be present in a

    molecule if it is to be incorporated within the DNA sequence.

    To build up brush-type AONs, monomers must contain either

    protruding hydrophobic moieties or functional groups which

    can be converted into hydrophobic units through additional

    transformations. One such building block (11) fulfilling the

    former requirement was recently reported,30 resulting in lipid-

    grafted ONs (Fig. 2A, left). Here, a C10-alkyl chain was

    introduced at C5 of the uracil nucleobase via an ethynyl group

    Fig. 2 Brush-type AONs. (A) Chemical structures of nucleoside- and polyaza crown ether monomers (11 and 12, respectively) bearing CEPA and

    DMT to yield hydrophobe(s)-grafted AONs. (B) Three-dimensional representation of modified uracil nucleotide (precursor: 11) base-pairing. The

    hydrophobic chain (white) does sterically not interfere with the hydrogen bond formation between modified-U and A nucleobases. Side (left) and

    top view (right). (C) Brush-type 12 mer AONs containing the hydrophobic (Uracil) building block (11) at varying positions (Middle or Termini)

    and in different numbers (2 or 4). (D) Representation of DNA-grafted polymers in homopolymer (left, precursor: 14a) and block copolymer motifs

    (right, precursors: 14b and 15; see Fig. 3C for the chemical structures). Each sphere represents a monomer unit and red strings are grafted-ONs.

    Dow

    nloa

    ded

    by U

    nive

    rsit

    y of

    Gro

    ning

    en o

    n 07

    Feb

    ruar

    y 20

    12

    Pub

    lish

    ed o

    n 22

    Aug

    ust

    2011

    on

    http

    ://p

    ubs.

    rsc.

    org

    | doi

    :10.

    1039

    /C1C

    S15

    138J

  • This journal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev., 2011, 40, 5745–5755 5749

    Fig. 3 Post-DNA-synthesis preparation of AONs. (A) After SPS, hydrophobes in organic solvents are attached to the 50-end of the ribose (*) or

    at the nucleobase (**) of nucleotides on the solid support. After the coupling, liberation from the resin and deprotection of primary amines of

    nucleobases are performed in concentrated ammonia to yield AON products. (B) List of reactions that allow functionalization of the DNA with

    hydrophobic units. (a) Chemical functionalities (R1 and R2) can be exchanged with each other in a particular case. (b) The CEPA group is only

    introduced in the hydrophobes (R2). (c) Reaction conditions may vary. (C) Hydrophobic moieties for the attachment on the solid support after

    DNA synthesis. (D) Synthetic route for the on-resin-modification of ONs with hydrophobes by Huisgen cycloaddition.

    Dow

    nloa

    ded

    by U

    nive

    rsit

    y of

    Gro

    ning

    en o

    n 07

    Feb

    ruar

    y 20

    12

    Pub

    lish

    ed o

    n 22

    Aug

    ust

    2011

    on

    http

    ://p

    ubs.

    rsc.

    org

    | doi

    :10.

    1039

    /C1C

    S15

    138J

  • 5750 Chem. Soc. Rev., 2011, 40, 5745–5755 This journal is c The Royal Society of Chemistry 2011

    using Sonogashira–Hagihara coupling. This 5-(dodec-1-ynyl)-

    uracil-containing nucleoside phosphoramidite has both DMT

    and CEPA at the 50- and 30-ends, respectively. This structure

    allows multiple addition of the modified nucleobases at precise

    positions in the target ON.31 Moreover, the relatively small

    hydrophobic functional group does sterically not hinder

    duplex formation with its complementary nucleobase, in this

    case adenine (Fig. 2B). In spite of the short alkyl chain, the

    morphological behavior of the hybrids confirms their classifi-

    cation as brush-type AONs. Due to the fact that the alkyl-

    modification does not lower the coupling yield compared to

    commercial reagents during SPS, virtually no additional

    synthetic effort is needed to include the desired number of this

    modified nucleotide in the desired positions (Fig. 2C). Using

    this strategy, 12 mer DNA amphiphiles with two or four lipid

    tails positioned in the middle or terminus (U2M, U2T and

    U4T) could be isolated with reasonable yields of 60 to 91%,

    based on analytical HPLC. As such, this method represented

    a facile means to tune the hydrophobicity of AONs in a fully

    controlled, automated manner. In an analogous strategy, a non-

    nucleoside building block (12) is also capable of linking two

    nucleotides in SPS. These azacrown derivatives possess varied

    hydrophobic segments in two aza-N positions.32 The duplexes

    formed from such compounds exhibited tolerable stability.

    Just as hydrophobic functional groups can be introduced

    along a DNA chain, one can imagine a structure in which

    DNAs are grafted onto a hydrophobic polymer backbone

    (Fig. 2D). The first such reported hybrid was prepared via

    SPS exhibiting a norbornene polymer (PNB) backbone.33

    Therein, two types of brush macromolecules were introduced:

    DNA and DNA/ferrocene (14a and b respectively, see

    Fig. 2D). Recently, Gianneschi et al. reported an analogous

    brush block copolymer (15), in which one block contained

    carboxylic acid in each repeating unit, which could be con-

    jugated to amino-modified DNA, while in the other block

    methylbenzene was incorporated as a hydrophobic function-

    ality.34 Here, amino-modified DNA on solid-supports was

    reacted with the diblock copolymers to realize an amide bond

    between the DNA and the polymer backbone.

    Preparation of AONs without a DNA synthesizer

    This ex situ conjugation method, entailing post-DNA-synthesis

    reactions while still on solid supports, particularly helps to

    achieve couplings in cases of incompatible solvents or catalysts,

    longer reaction times or more extreme reaction conditions than

    are accessible with a DNA synthesizer. This strategy can be

    used to prepare both brush-type (14 and 15) and linear (13 and

    16) AONs (see Fig. 3A–C). An emerging synthetic method

    along these lines is the Huisgen cycloaddition, the so-called

    ‘‘Click chemistry’’,35 employing either azide- or acetylene-

    modified DNA that is reacted with the correspondingly

    functionalized hydrophobic moieties (Fig. 3D).36 However,

    the conditions for this ex situ method vary depending on the

    chemical structure, catalyst and chemostability and thus need

    to be optimized separately for each reactant.37–39

    Regardless of the type of SPS method used for synthesis, the

    resulting AONs offer additional advantages in the purification

    stage compared to pristine ODNs. When hydrophobic units

    are introduced, the AON materials interact strongly with

    chromatographic stationary phases. Thus, high performance

    liquid chromatography (HPLC) with affinity columns, such

    as reverse-phase or anion exchange resins, can be used to

    efficiently isolate target AONs.40

    3 Self-assembled structures of AONs

    DNA amphiphiles on their own generally self-assemble into

    structures with simple geometries: spherical or cylindrical micelle

    aggregates or vesicles. In addition, these structures can be altered

    by mixing in other materials of interest, by hybridization or by

    enzymatic reactions, as will be discussed below.

    In aqueous media, the hydrophobic units of AONs tend to

    undergo microphase separation, to minimize the exposure to

    water. The linear DBC with poly(butadiene), DNA-b-PB,

    formed vesicles of about 80 nm in diameter (Fig. 4A).41

    Thereby, the isolation of the vesicular interior from the

    surrounding medium by a DBC double-layer was demon-

    strated using fluorescence probes. Although the material

    stability and duplex formation capability have not yet been

    studied, the novel behavior of this DNA hybrid suggests

    further investigations of such AON structures.42 Another

    vesicular structure involving AONs used terminal bilipid

    moieties to form the hydrophobic core of the bilayer.43

    Through the hybridization of fluorescently labelled comple-

    mentary DNAs (cDNAs), the reversible decoration of the

    surface of the self-assembled aggregates by hybridization was

    demonstrated. Instead of constructing vesicle structures from

    Fig. 4 Microscopic images and illustrations of self-assembled

    amphiphilic DNAs. (A) A TEM image (top) of PB-b-DNA (from

    precursor: 16, italicized b such as X-b-Y denotes a block copolymer

    composed of polymer segments X and Y that are covalently

    connected) vesicles. Reproduced from ref. 41 with permission of the

    Royal Society of Chemistry. (B) An AFM image (top) of PS-b-DNA

    spherical micelles (unpublished). Analogous aggregates were also

    formed with single-stranded DBCs (precursors: 8–10) and lipid-DNAs

    (see Fig. 2C, precursor: 11) with 4–10 nm height. Scale bars: 100 nm;

    height color scale: 10 nm.

    Dow

    nloa

    ded

    by U

    nive

    rsit

    y of

    Gro

    ning

    en o

    n 07

    Feb

    ruar

    y 20

    12

    Pub

    lish

    ed o

    n 22

    Aug

    ust

    2011

    on

    http

    ://p

    ubs.

    rsc.

    org

    | doi

    :10.

    1039

    /C1C

    S15

    138J

  • This journal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev., 2011, 40, 5745–5755 5751

    AONs exclusively, it is much easier to incorporate AONs into

    the membrane of artificial liposomes fabricated from lipids or

    polymers.42

    Simpler and more extensively studied systems are DNA

    micelles (Fig. 4B). Many AONs have been observed to form

    spherical micelles with a diameter range of 5 to 12 nm, as

    measured by dynamic light scattering (DLS) and atomic force

    microscopy (AFM).18,26,31 These AFM measurements further

    revealed that deformation of such micelles on the surface is

    depending on the hydrophobic segments. AONs containing

    hydrophobic segments with higher glass transition tempera-

    tures (Tg), in particular PS and PFO, endure the mechanical

    forces, during the AFM measurement, more than those with

    lower-Tg segments like PPO, PLGA and lipids.

    One big advantage of constructing nanostructures with

    DNA is that once assembled, the architectures can be further

    manipulated by enzymatic reactions. Indeed, enzymatic

    manipulations were also applied to linear AONs and DNA

    brushes. For instance, micelles consisting of a DNA diblock

    copolymer, DNA-b-PPO, were treated with the enzyme,

    terminal deoxynucleotidyl transferase (TdT), which catalyzes

    nucleotidic elongation of the terminal 30-end of single-stranded

    (ss) DNA in the presence of deoxyribonucleoside triphosphate

    (dNTP) under isothermal conditions at 37 1C. In this experi-

    ment, over the course of 16 hours, the sequence-independent

    DNA polymerase added approximately 60 T nucleotides to the

    termini of DNA chains in the corona of the micelles (Fig. 5A),

    which resulted in a height increase from 5 to 11 nm on a mica

    surface.44,45 This method offers a convenient post-synthesis size

    control of nucleic acid nanoobjects. Precise post-synthesis

    shape control has also been demonstrated with such materials

    through the addition of ssDNA: AON micelles were trans-

    formed into rod-shaped aggregates through hybridization with

    a long cDNA template (Fig. 5B).46 The sequence of the long

    DNA template was designed to contain five repeats of the

    complementary sequence of the DNA corona of spherical

    micelles. Alignment of five AONs along the template, after

    disaggregation of spherical micelles, resulted in rod-like aggre-

    gates, where the hydrophobic polymer units were uniformly

    placed between two long dsDNAs glueing the two duplexes

    together. Another example of changing supramolecular AON

    structures by very mild external stimuli was realized with the

    enzyme, phosphodiesterase (Fig. 5C). Starting from a spherical

    micelle consisting of brush-type polymeric AONs, digestion

    of the DNA segments reduced the volume occupied by the

    grafted nucleotides. The resulting change in the hydrophilic/

    hydrophobic volume ratio induced a transformation of the

    spherical nanoobjects into long cylindrical aggregates. The

    second stimulus applied on the AON tube was hybridization

    of longer ONs with the remaining segments of the DNA brush.

    This lengthening restored the initial structural parameters,

    producing spherical aggregates again.34 Such reversible tech-

    niques for structural manipulation should eventually produce

    potential applications of AONs in the biomedical field with

    possibly the same impact as peptide amphiphiles.47

    4 Applications of AONs and their nanocomposites

    Utilization of pristine AONs

    Many advances of AON-micellar systems in the last five years

    were presented in the fields of DNA detection,25 templated

    reactions,26 and drug delivery due to their self-recognition

    properties and the presence of a hydrophobic core to serve as a

    carrier unit.24 In particular, an AON micelle has shown great

    potential in cancer therapy.48–50 PPO-b-DNA micelles were

    equipped both with a hydrophobic anticancer drug and a

    targeting unit. Therby, doxorubicin accumulated in the core

    of the AON aggregates by simple mixing while the folic acid

    targeting units were incorporated into the carrier system by

    hybridization (Fig. 6A, top). These aggregates were taken up

    by CaCo-2 cancerous cells through receptor mediated endo-

    cytosis (Fig. 6A, bottom) and, as anticipated, resulted in

    efficient cytotoxicity and high mortality.48 In a different

    approach, Tan et al. demonstrated that RNA aptamers,

    ONs that can specifically bind to a target molecule, can be

    conjugated to lipids, resulting in micelle formation and

    efficient targeting ability to a specific cancerous cell line

    Fig. 5 Engineering self-assembled structures of AONs. (A) Isothermal enzymatic growth of DNA-b-PPO (precursor: 9) micelles incubated

    with TdT and dTTP at 37 1C (top). AFM images of growing nanoparticles with increasing reaction times (bottom). (B) DNA-templated shape

    transition from spherical micelles to rod-like aggregates (top) and an AFM image of rod-like structures (bottom). (C) Enzyme induced reversible

    transformation of a DNA-brush block copolymer (top, precursor: 15) and its TEM images (bottom). Note that the outer-segments of

    DNA-brushes in both spherical aggregates (top left and right of the representation) have ethylene glycol chains to lower electrostatic repulsion

    between the DNA brushes. Reproduced from ref. 44, 46 and 34, respectively, with permission of Wiley-VCH Verlag GmbH & Co KGaA.

    Dow

    nloa

    ded

    by U

    nive

    rsit

    y of

    Gro

    ning

    en o

    n 07

    Feb

    ruar

    y 20

    12

    Pub

    lish

    ed o

    n 22

    Aug

    ust

    2011

    on

    http

    ://p

    ubs.

    rsc.

    org

    | doi

    :10.

    1039

    /C1C

    S15

    138J

  • 5752 Chem. Soc. Rev., 2011, 40, 5745–5755 This journal is c The Royal Society of Chemistry 2011

    (Ramos) (Fig. 6B).50 The multimerization of the aptamer unit

    at the surface of the AON aggregates significantly increased the

    binding of the nanoobjects compared to the pristine nucleic

    acid sequences. Although these are powerful new tools for

    cancer therapy, there is an inherent limitation to the use of

    micellar carriers in vivo: upon dilution below the critical micelle

    concentration (CMC), they tend to disassemble. One potential

    solution to this issue involves an additional nanocontainer, a

    virus capsid.30 In addition to the advantages of micellar

    delivery described above, it was demonstrated that micelles

    of DNA amphiphiles can act as a template for the self-

    assembly of Cowpea Chlorotic Mottle Virus capsids with

    T = 1 and 2 geometry at neutral pH (Fig. 7A). The resulting

    structures were then stable against dilution. At the same time,

    this encapsulation around micelles of DNA amphiphiles

    represented a general and facile supramolecular loading strategy

    for hydrophobic or hydrophilic small molecules within the protein

    nanocontainer, an essential process for the application of these

    nanoobjects in biomedicine. The virus encapsulation of AONs is a

    good example of how the properties, in this case the micellar

    stability, can be greatly improved by combining AONs with other

    materials.

    AONs employed together with other materials

    Composites are made from two or more materials to achieve

    desired synergic properties out of the individual constituents.

    Although the synthesis and designed self-assembly of a single

    material is one of the beauties and goals of fundamental

    research, more focused applications can be realized through

    combination with other materials of interest. Since the essence

    of DNA nanotechnology is the precise programmed assembly

    of DNAs through Watson–Crick basepairing, the other com-

    ponent of AONs, the hydrophobic moieties, can play an

    important role as well by interacting with other hydrophobic

    materials to produce versatile composites for particular

    applications.

    Fig. 6 Medical applications of micellar AONs in cancer therapy.

    (A) A multifunctional micellar drug-carrier consisting of a DBC

    (precursor: 9). The targeting unit, folic acid (FA), to cancerous cells

    is attached to the outer surface of the DNA corona through hybridiza-

    tion while a drug, doxorubicin (Dox), is loaded into a hydrophobic

    core of the micelle (top). Fluorescence microscopy image of CaCo-2

    cells that have taken up the micellar vehicle (bottom). Reproduced

    from ref. 49 with permission of Wiley-VCH Verlag GmbH & Co

    KGaA. (B) Graphical representation of a targeting micelle (precursor:

    11) with an RNA aptamer as the corona (top). The aptamer micelle

    exhibited a high cell-specific affinity to lymphoma cells (Ramos) as

    shown in the fluorescence micrograph (bottom). Reproduced from

    ref. 50 with permission of the National Academy of Sciences.

    Fig. 7 Composite micelles consisting of AONs and viral capsids or synthetic polymers, respectively. (A) Micelles of DNA amphiphiles

    (precursors: 9 and 11) loaded with either small hydrophobic compounds (top left) or with hydrophilic compounds by hybridization (top right)

    templated virus capsid formation at neutral pH. TEM images of micelles incorporated in virus capsids with T = 1 or 2 geometry (bottom) and an

    empty capsid formed at pH 5.0 as control (inset). Scale bars: 40 nm (adapted from ref. 30). (B) Graphical representation of a blend micelle.

    A diblock DNA copolymer, PPO-b-DNA (precursor: 9), was mixed with a triblock copolymer Pluronic (PEO-b-PPO-b-PEO) that exhibits the same

    hydrophobic block, PPO. Both block copolymers form mixed micelles. The hydrophobic core was further stabilized through UV-induced semi-

    interpenetrating network formation, which inhibits disaggregation or precipitation under dilution or low temperature, respectively. In a similar

    manner as shown in (A), the core and corona could be equipped with desired functionalities also with distance-control (adapted from ref. 53).

    Dow

    nloa

    ded

    by U

    nive

    rsit

    y of

    Gro

    ning

    en o

    n 07

    Feb

    ruar

    y 20

    12

    Pub

    lish

    ed o

    n 22

    Aug

    ust

    2011

    on

    http

    ://p

    ubs.

    rsc.

    org

    | doi

    :10.

    1039

    /C1C

    S15

    138J

  • This journal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev., 2011, 40, 5745–5755 5753

    With regard to micellar aggregates, this principle has

    resulted in doping electrostatically or magnetically active

    materials in the hydrophobic core.51,52 In a more elegant

    example, DNA diblock copolymers (DNA-b-PPO) formed

    mixed micelles with Pluronict triblock architectures (PEO-b-

    PPO-b-PEO) (Fig. 7B).53 This polymer–AON composite, with

    a DNA/PEO corona and a PPO core, could be loaded with

    hydrophobic molecules and stabilized against dissociation by

    the formation of a semi-interpenetrating network, resulting in

    highly stable soft particles. Furthermore, the potential of the

    DNA for further self-assembly remained unchanged—the

    corona could be conveniently modified by hybridization either

    with dye-modified complementary DNA, demonstrating dis-

    tance controlled functionalization, or with cDNA-labelled

    gold nanoparticles. This composite is an excellent candidate

    for a smart drug delivery vehicle, with possibly reduced

    immunotoxicity due to the presence of the PEO chains on

    the surface.

    Vesicles or liposomes have been intensively investigated to

    understand and manipulate biological events and processes.42,54

    Due to the fact that their bilayer is composed of amphiphiles,

    it is no surprise that AONs can also play a role in decorating

    these nanocontainers with DNA. More precisely, the hydro-

    phobic components of AONs can be tailored to anchor into

    liposomes, leaving the DNA strands available for subsequent

    supramolecular events, whether for functionalizing the surface

    (Fig. 8A), anchoring the liposome or even tethering vesicles

    together through hybridization (Fig. 8B. The AONs described

    above, in which the hydrophobes are conjugated to the DNA

    and peptide nucleic acid (PNA)55 termini, are most commonly

    combined with phospholipid bilayers. Thus AONs have been

    used to achieve liposome tagging,21,56,57 intervesicular

    fusion,19,58 the assembly of larger containers59 and attachment

    to lipid bilayers, mimicking cellular systems.22,60 With regard

    to applications of AONs in animal experiments, lipophilic

    units and cholesterol-conjugated small interfering (si) RNA

    have been utilized to study siRNA-AON uptake into cells and

    gene silencing in vivo.61

    A more integrated application of AON composites was

    reported with single-walled carbon nanotubes (SWNTs). In

    the study, a single polymeric AON was demonstrated to be an

    all-in-one solution for SWNT technology.27 The hydrophobic

    segment of the AON, PFO (11), was covalently connected to a

    short 22 mer via a phosphodiester bond and interacted

    strongly with SWNT sidewalls by p–p interactions. This

    allowed highly diameter-selective dispersion (0.8–1.1 nm) of

    SWNT species out of as-produced HiPCO SWNTs (Fig. 9A).

    The dispersion selected for a relatively small subset of chiral

    indices, which determine the electronic properties of the

    materials. In addition, the ssDNA block provided solubility

    in aqueous media and, most importantly, remained available

    for straightforward duplex formation with its cDNA (Fig. 9B).

    Through base-pairing between the free DNA block and cDNA

    conjugated to a gold nanoparticle, alignment of the metallic

    particles along the SWNT was demonstrated (Fig. 9C).

    A similar procedure allowed attachment of the dispersion to

    a cDNA-functionalized surface. The same surface immobiliza-

    tion onto lithographically prepared electrodes (Fig. 9D)

    resulted in 98% fabrication efficiency for SWNT field effect

    transistors. The entire process proved highly selective for

    semiconducting SWNT species and represented a solution-

    based, scalable method achieved purely by self-assembly. This

    composite powerfully combined all the advantageous proper-

    ties of the individual polymer, SWNT and DNA constituents.

    5 Conclusions and outlook

    Synthetic methods, chemical variety, structural parameters

    and outlined functions of AONs are summarized in Table 1

    for the reader’s convenience. These AONs are indeed a new

    Fig. 8 AONs incorporated into lipid bilayers of liposomes.

    (A) Hydrophobic parts (orange, precursors: 1–2, 4–7, 13) of AONs

    could be used as anchors (left) to bilayers of liposomes. Simple tagging

    of the liposomes was realized by hybridizing fluorescently labeled

    cDNA (blue string, right) to the DNA segment (red string) of the

    AON. (B) AONs could induce moderate vesicular fusion via duplex

    formation of AONs anchored in liposomes.

    Fig. 9 Utilization of AONs in a SWNT composite. (A) A polymeric

    AON (precursor: 10), PFO-b-DNA, was used to disperse HiPCO

    SWNTs as produced. (B) Resulting dispersions were enriched in

    semiconducting SWNT species with narrow diameter distribution

    due to the strong and specific interaction between the PFO block

    and the SWNT surface. The presence of the protruding DNA strand

    was exploited to introduce functionalities by hybridization with

    cDNA-modified (blue) Au-nanoparticles (yellow spheres) (C) and to

    sequence-specifically immobilize SWNTs on defined cDNA surfaces,

    such as electrodes of field-effect transistors (FETs). (D) The device

    yield (98%) achieved exclusively by aqueous solution processing and

    hybridization was extremely high (adapted from ref. 27).Dow

    nloa

    ded

    by U

    nive

    rsit

    y of

    Gro

    ning

    en o

    n 07

    Feb

    ruar

    y 20

    12

    Pub

    lish

    ed o

    n 22

    Aug

    ust

    2011

    on

    http

    ://p

    ubs.

    rsc.

    org

    | doi

    :10.

    1039

    /C1C

    S15

    138J

  • 5754 Chem. Soc. Rev., 2011, 40, 5745–5755 This journal is c The Royal Society of Chemistry 2011

    class of materials that has seen rapid growth with great future

    potential. Regarding the choice of synthetic methods, in

    general coupling hydrophobic units to DNA in solution is a

    difficult task mostly resulting in low overall yields. In contrast,

    synthetic strategies based on SPS approaches are the preferred

    route to fabricate AON materials. While preparing AONs

    in-line guarantees good reproducibility and allows large scale

    synthesis, the modification of ONs on solid supports with

    hydrophobic moieties offers more flexibility regarding experi-

    mental conditions and does not necessitate large equipment

    costs. Within the context of polymeric AON synthesis, as of

    yet no 30-amphiphilic DBCs have been reported using a fully

    automated SPS approach showing that there are still synthetic

    challenges to be overcome in the field. The same holds true for

    the DNA part of AONs. ONs of up to hundred nucleotides,

    as shown in Table 1, could be chemically synthesized. There

    are still ongoing efforts to overcome this current limit, for

    instance, by enzymatic reactions employing ss- and ds

    ON-hybrids in combination with polymerases, ligases or

    endonucleases to extend the size of nucleic acid segments.62–64

    Another future task in AON synthesis is to incorporate

    hydrophobic compounds that have interesting optoelectronic

    properties like conjugated polymers, thereby widening the field

    for potential applications in diagnostics and (nano)electronics.

    The amphiphilicity of AONs allows them to aggregate into

    superstructures, mostly micelle aggregates or vesicles. Thereby,

    well established strategies for structure formation of small

    molecule amphiphiles and block copolymers can be adopted

    for AONs and their composites with other amphiphiles. And

    indeed those methods are much simpler than the sophisticated

    design and annealing techniques known for pristine DNA

    nanoconstruction. Moreover, self-assembled nanostructures

    of AONs consist of a very limited set of building blocks

    compared to complex DNA nanoobjects.65,66 The aggregation

    behavior of AONs can be conveniently controlled by adjusting

    the molecular parameters like the length of the DNA segment

    and the molecular weight or the nature of the hydrophobic

    unit. Additional opportunities for manipulating their self-

    assembly properties are hybridization and enzymatic modi-

    fications. From another point of view, introducing amphiphilic

    properties into the already well-developed field of pristine

    DNA nanostructures should greatly enhance their appli-

    cability. This can range from simply depositing them on

    hydrophobic surfaces or generating well defined compartments

    of different polarities.

    Regarding applications, great potential of AONs can be

    foreseen for the biomedical field. Promising examples have

    demonstrated basic functions such as targeted-drug delivery or

    transfection with AON nanoobjects in vitro. However, the

    cytotoxicity, immunogenicity and uptake of AON nano-

    structures are far from being well understood. In the future,

    elucidating these issues will bring advances with AONs in the

    pharmaceutical sector especially in regard to stability and

    delivery of biologicals. Those challenges can only be mastered

    in a multidisciplinary effort, especially with input from medical

    and clinical researchers. In this respect, one can be excited

    about the future demonstrations of AON nanostructures

    in vivo.

    Acknowledgements

    We thank the EU (ERC starting grant, ECCell), the Netherlands

    Organization for Scientific Research (NWO-Vici), the German

    Research Foundation (DFG), and the Zernike Institute for

    Advanced Materials for financial support.

    References and notes

    1 This quote is Nadrian C. Seeman’s, the founder of DNA nano-technology, title of his lecture since a decade.

    2 J. H. Chen and N. C. Seeman, Nature, 1991, 350, 631–633.3 B. Yurke, A. J. Turberfield, A. P. Mills, F. C. Simmel andJ. L. Neumann, Nature, 2000, 406, 605–608.

    Table 1 Summary of AONs covered in this review

    No. Hydrophobea Func. ONb ON-Fuct. ON lengthc Couplingd Ref.

    AONs terminally functionalized with a hydrophobic unit1 Lipid CEPA D, R 50-OH 20–81 SPS 18,502 Bislipid CEPA D 30-/50-OH 24–48 SPS 193 FDMT CEPA D 50-OH 30–100 SPS 204 Bislipid CEPA D 50-OH 15 SPS 215–7 Mono-/bischol CEPA D, R, P 30-/50-OH 12–42 SPS 22,56,588 PS (10K) CEPA D 50-OH 20, 25 postSPS 25,51,529 PPO (7K) CEPA D 50-OH 11, 22 SPS 26,30,4810 PFO (6K) CEPA D 50-OH 22 SPS 2713 Bislipid COOH D 50-OH 9 postSPS 4316 PB (2K) NH2 D 5

    0-COOH 12 postSPS 41N/A PLGA (10K) COOH D 50-NH2 15 Solution 24AONs with DNA chains attached to a hydrophobic polymer backbone14a–b PNB (-b-Fc) CEPA D 50-OH 18, 22 postSPS 3315 PNB COOH D 50-NH2 19 postSPS 34AONs with DNA hydrophobic units attached to a DNA backbone11 Lipid CEPA/DMT D N/A 11, 12 (2, 4) SPS 30,31AONs containing a hydrophobic unit within the nucleic acid structure12 Azacrown ether CEPA/DMT D N/A 9–17 (1–3) SPS 32,60

    a Numbers in parentheses represent number-averaged molecular weight (g molÿ1) of the polymers. b D: DNA, R: RNA, P: PNA. c Unit: number

    of nucleotides. Numbers in parentheses represent the number of hydrophobes included in the AON. d SPS: solid-phase synthesis, postSPS: post-

    synthesis after SPS of ONs.

    Dow

    nloa

    ded

    by U

    nive

    rsit

    y of

    Gro

    ning

    en o

    n 07

    Feb

    ruar

    y 20

    12

    Pub

    lish

    ed o

    n 22

    Aug

    ust

    2011

    on

    http

    ://p

    ubs.

    rsc.

    org

    | doi

    :10.

    1039

    /C1C

    S15

    138J

  • This journal is c The Royal Society of Chemistry 2011 Chem. Soc. Rev., 2011, 40, 5745–5755 5755

    4 H. Liu, Y. Chen, Y. He, A. Ribbe and C. Mao, Angew. Chem., Int.Ed., 2006, 45, 1942–1945.

    5 P. Rothemund, Nature, 2006, 440, 297–302.6 P. Yin, R. F. Hariadi, S. Sahu, H. M. T. Choi, S. H. Park,T. H. Labean and J. H. Reif, Science, 2008, 321, 824–826.

    7 S. Douglas, H. Dietz, T. Liedl, B. Högberg, F. Graf and W. Shih,Nature, 2009, 459, 414–418.

    8 D. Han, S. Pal, J. Nangreave, Z. Deng, Y. Liu and H. Yan,Science, 2011, 332, 342–346.

    9 A. N. Kapanidis and S. Weiss, J. Chem. Phys., 2002, 117,10953–10964.

    10 D. A. Giljohann, D. S. Seferos, W. L. Daniel, M. D. Massich,P. C. Patel and C. A. Mirkin, Angew. Chem., Int. Ed., 2010, 49,3280–3294.

    11 C. M. Niemeyer, Angew. Chem., Int. Ed., 2010, 49, 1200–1216.12 H. Rosemeyer, Chem. Biodiversity, 2005, 2, 977–1063.13 J. Goodchild, Bioconjugate Chem., 1990, 1, 165–187.14 S. Verma and F. Eckstein, Annu. Rev. Biochem., 1998, 67, 99–134.15 E. Defrancq, Y. Singh and N. Spinelli, Curr. Org. Chem., 2008, 12,

    263–290.16 M. H. Caruthers, Acc. Chem. Res., 1991, 24, 278–284.17 S. L. Beaucage and M. H. Caruthers, Tetrahedron Lett., 1981, 22,

    1859–1862.18 H. Liu, Z. Zhu, H. Kang, Y. Wu, K. Sefan and W. Tan,

    Chem.–Eur. J., 2010, 16, 3791–3797.19 Y.-H. M. Chan, B. van Lengerich and S. G. Boxer, Proc. Natl.

    Acad. Sci. U. S. A., 2009, 106, 979–984.20 W. Pearson, D. Berry, P. Stoy, K. Jung and A. D. Sercel, J. Org.

    Chem., 2005, 70, 7114–7122.21 A. Gissot, C. Di Primo, I. Bestel, G. Giannone, H. Chapuis and

    P. Barthelemy, Chem. Commun., 2008, 5550–5552.22 I. Pfeiffer and F. Hook, J. Am. Chem. Soc., 2004, 126,

    10224–10225.23 G. Stengel, L. Simonsson, R. A. Campbell and F. Hook, J. Phys.

    Chem. B, 2008, 112, 8264–8274.24 J. Jeong and T. Park, Bioconjugate Chem., 2001, 12, 917–923.25 Z. Li, Y. Zhang, P. Fullhart and C. Mirkin, Nano Lett., 2004, 4,

    1055–1058.26 F. Alemdaroglu, K. Ding, R. Berger and A. Herrmann, Angew.

    Chem., Int. Ed., 2006, 45, 4206–4210.27 M. Kwak, J. Gao, D. K. Prusty, A. J. Musser, V. A. Markov,

    N. Tombros, M. C. A. Stuart, W. R. Browne, E. J. Boekema,G. ten Brinke, H. T. Jonkman, B. J. van Wees, M. A. Loi andA. Herrmann, Angew. Chem., Int. Ed., 2011, 50, 3206–3210.

    28 M. Kwak and A. Herrmann, Angew. Chem., Int. Ed., 2010, 49,8574–8587.

    29 J.-F. Lutz and H. G. Boerner, Prog. Polym. Sci., 2008, 33,1–39.

    30 M. Kwak, I. Minten, D.-M. Anaya, A. Musser, M. Brasch,R. Nolte, K. Müllen, J. Cornelissen and A. Herrmann, J. Am.Chem. Soc., 2010, 132, 7834–7835.

    31 M. Anaya, M. Kwak, A. J. Musser, K. Müllen and A. Herrmann,Chem.–Eur. J., 2010, 16, 12852–12859.

    32 K. Rohr and S. Vogel, ChemBioChem, 2006, 7, 463–470.33 K. Watson, S. Park, J. Im, S. Nguyen and C. Mirkin, J. Am. Chem.

    Soc., 2001, 123, 5592–5593.34 M.-P. Chien, A. M. Rush, M. P. Thompson and N. C. Gianneschi,

    Angew. Chem., Int. Ed., 2010, 49, 5076–5080.35 H. Kolb, M. Finn and K. Sharpless, Angew. Chem., Int. Ed., 2001,

    40, 2004–2021.36 G. Godeau, H. Arnion, C. Brun, C. Staedel and P. Barthelemy,

    Med. Chem. Commun., 2010, 1, 76–78.37 F. Amblard, J. H. Cho and R. F. Schinazi, Chem. Rev., 2009, 109,

    4207–4220.38 A. H. El-Sagheer and T. Brown, Chem. Soc. Rev., 2010, 39, 1388.39 P. M. E. Gramlich, C. T. Wirges, A. Manetto and T. Carell,

    Angew. Chem., Int. Ed., 2008, 47, 8350–8358.

    40 For rapid and efficient purification, anion-exchange (AIEX)chromatography is a commonly employed technique using a low-cost Tris-based buffer system. ‘‘Tris’’ is an abbreviated name oftris(hydroxymethyl)aminomethane. During AIEX chromato-graphy, Tris-HCl is used as a low saline buffer in combinationwith other chemicals: ethylenediaminetetraacetic acid, acetic acid,boric acid. In a high saline buffer, NaCl is added in high concen-tration (i.e. 1.0 M).

    41 F. Teixeira Jr, P. Rigler and C. Vebert-Nardin, Chem. Commun.,2007, 1130–1132.

    42 S. F. M. van Dongen, H.-P. M. de Hoog, R. J. R. W. Peters,M. Nallani, R. J. M. Nolte and J. C. M. van Hest, Chem. Rev.,2009, 109, 6212–6274.

    43 M. P. Thompson, M.-P. Chien, T.-H. Ku, A. M. Rush andN. C. Gianneschi, Nano Lett., 2010, 10, 2690–2693.

    44 F. E. Alemdaroglu, J. Wang, M. Börsch, R. Berger andA. Herrmann, Angew. Chem., Int. Ed., 2008, 47, 974–976.

    45 J. Wang, F. E. Alemdaroglu, D. K. Prusty, A. Herrmann andR. Berger, Macromolecules, 2008, 41, 2914–2919.

    46 K. Ding, F. E. Alemdaroglu, M. Börsch, R. Berger andA. Herrmann, Angew. Chem., Int. Ed., 2007, 46, 1172–1175.

    47 Y. Lim, K. Moon and M. Lee, Chem. Soc. Rev., 2009, 38,925–934.

    48 F. E. Alemdaroglu, C. N. Alemdaroglu, P. Langguth andA. Herrmann, Adv. Mater., 2008, 20, 899–902.

    49 F. E. Alemdaroglu, C. N. Alemdaroglu, P. Langguth andA. Herrmann, Macromol. Rapid Commun., 2008, 29, 326–329.

    50 Y. Wu, K. Sefah, H. Liu, R. Wang and W. Tan, Proc. Natl. Acad.Sci. U. S. A., 2010, 107, 5–10.

    51 M. Sowwan, M. Faroun, E. Mentovich, I. Ibrahim, S. Haboush,F. E. Alemdaroglu, M. Kwak, S. Richter and A. Herrmann,Macromol. Rapid Commun., 2010, 31, 1242–1246.

    52 X.-J. Chen, B. L. Sanchez-Gaytan, S. E. N. Hayik, M. Fryd,B. B. Wayland and S.-J. Park, Small, 2010, 6, 2256–2260.

    53 M. Kwak, A. J. Musser, J. Lee and A. Herrmann,Chem. Commun.,2010, 46, 4935–4937.

    54 J. Voskuhl and B. J. Ravoo, Chem. Soc. Rev., 2009, 38,495–505.

    55 M. Loew, R. Springer, S. Scolari, F. Altenbrunn, O. Seitz,J. Liebscher, D. Huster, A. Herrmann and A. Arbuzova, J. Am.Chem. Soc., 2010, 132, 16066–16072.

    56 F. Bombelli, F. Betti, F. Gambinossi, G. Caminati, T. Brown,P. Baglioni and D. Berti, Soft Matter, 2009, 5, 1639–1645.

    57 P. Beales and T. Vanderlick, J. Phys. Chem. B, 2009, 113,13678–13686.

    58 G. Stengel, R. Zahn and F. Hook, J. Am. Chem. Soc., 2007, 129,9584–9585.

    59 M. Loew, L. Kang, L. Daehne, R. Hendus-Altenburger,O. Kaczmarek, J. Liebscher, D. Huster, K. Ludwig, C. Boettcher,A. Herrmann and A. Arbuzova, Small, 2009, 5, 320–323.

    60 U. Jakobsen, A. C. Simonsen and S. Vogel, J. Am. Chem. Soc.,2008, 130, 10462–10463.

    61 C. Wolfrum, S. Shi, K. N. Jayaprakash, M. Jayaraman, G. Wang,R. K. Pandey, K. G. Rajeev, T. Nakayama, K. Charrise,E. M. Ndungo, T. Zimmermann, V. Koteliansky, M. Manoharanand M. Stoffel, Nat. Biotechnol., 2007, 25, 1149–1157.

    62 M. Safak, F. E. Alemdaroglu, Y. Li, E. Ergen and A. Herrmann,Adv. Mater., 2007, 19, 1499–1505.

    63 S. Keller, J. Wang, M. Chandra, R. Berger and A. Marx, J. Am.Chem. Soc., 2008, 130, 13188–13189.

    64 M. S. Ayaz, M. Kwak, F. E. Alemdaroglu, J. Wang, R. Berger andA. Herrmann, Chem. Commun., 2011, 47, 2243.

    65 C. X. Lin, Y. Liu, S. Rinker and H. Yan, ChemPhysChem, 2006, 7,1641–1647.

    66 S. M. Douglas, A. H. Marblestone, S. Teerapittayanon,A. Vazquez, G. M. Church and W. M. Shih, Nucleic Acids Res.,2009, 37, 5001–5006.

    Dow

    nloa

    ded

    by U

    nive

    rsit

    y of

    Gro

    ning

    en o

    n 07

    Feb

    ruar

    y 20

    12

    Pub

    lish

    ed o

    n 22

    Aug

    ust

    2011

    on

    http

    ://p

    ubs.

    rsc.

    org

    | doi

    :10.

    1039

    /C1C

    S15

    138J