Seperations Nano

Embed Size (px)

Citation preview

  • 7/30/2019 Seperations Nano

    1/12

    Applications of nanomaterials

    in enantioseparation and relatedtechniquesCuilan Chang, Xin Wang, Yu Bai, Huwei Liu

    Chirality is an important, universal phenomenon in nature. For the in-depth study of pharmacology and biology, efficient

    enantioselective technologies are indispensable. Nanomaterials with large surface-to-volume ratio and specific physical and

    chemical properties have demonstrated great potential in chiral discrimination. Many publications show that utilization of

    nanomaterials could improve the selectivity, the stability and the efficiency of enantioseparation.

    This review summarizes the applications of various enantioselective nanomaterials, including mesoporous silica, organicpolymers, metal-organic frameworks, metal nanomaterials, magnetic nanomaterials, carbon nanotubes and well-oriented chiral

    nanolayers. After proper preparation and modification, these functionalized nanomaterials are effective for chiral separation

    through their specific enantioselective adsorption, especially when they are used as stationary or pseudo-stationary phases in

    chiral chromatographic separation, such as thin-layer chromatography, high-performance liquid chromatography, gas

    chromatography and capillary electrophoresis.

    2012 Elsevier Ltd. All rights reserved.

    Keywords: Carbon nanotube; Enantioseparation; Magnetic nanomaterial; Mesoporous silica; Metal nanomaterial; Metal-organic framework;

    Nanomaterial; Organic polymer; Stationary phase

    Abbreviations: AuNP, Gold nanoparticle; b-CD, b-cyclodextrin; BSA, Bovine serum albumin; CE, Capillary electrophoresis; CEC, Capillary

    electrochromatography; CNT, Carbon nanotube; EOF, Electroosmotic flow; GC, Gas chromatography; HPLC, High performance liquid

    chromatography; HP-b-CD, Hydroxypropyl-b-cyclodextrin; MNP, Magnetic nanoparticle; MOF, Metal-organic framework; MWCNT, Multi-walled

    carbon nanotube; SPR, Surface-plasmon resonance; STM, Scanning tunneling microscope; SWCNT, Single-walled carbon nanotube; TLC, Thin-layer

    chromatography; XRD, X-ray diffraction

    1. Introduction

    Chirality, which refers to the property of a

    molecule or a system not identical to its

    mirror image, has been confirmed to be

    related to the origin of life [1,2]. In

    chemistry, chirality is used to describe the

    lack of a symmetric plane in molecules.

    This property exists in various types of

    molecules, including chiral drugs, aminoacids, carbohydrates, DNAs and proteins.

    These chiral compounds, especially chiral

    drugs, often exhibit remarkably different

    effects in pharmacological activity, trans-

    port mechanism, metabolism pathway

    and toxicity. Taking thalidomide as an

    example, the R-(+)-enantiomer is an

    effective sedative, whereas the S-(-)-enan-

    tiomer was found to cause fetal abnor-

    malities. So obtaining enantiomeric pure

    compounds is significantly important, but

    it is still a big challenge because of the

    identical physical and chemical properties

    of enantiomers in an achiral environment.

    Although great success has been attained

    in asymmetric synthesis, enantioselective

    reactions are still restricted to the use of

    enantiomeric pure materials (e.g., chiral

    substrates, chiral auxiliary and chiral

    catalysts) [3]. The development of more

    efficient enantioseparation techniques istherefore greatly desired for obtaining

    pure enantiomers, monitoring asymmetric

    reactions and ensuring quality control of

    chiral drugs.

    Chiral separation has been developed for

    several decades, but the separation theory

    is still based on the three-point interac-

    tion [4], which was pointed out by

    Dalgliesh in 1952. The theory states that

    at least three configuration-dependent

    points are needed for the recognition

    Cuilan Chang, Xin Wang,

    Yu Bai, Huwei Liu*,

    Beijing National Laboratory for

    Molecular Sciences, the Key

    Laboratory of Bioorganic

    Chemistry and Molecular

    Engineering of Ministry of

    Education, Institute ofAnalytical Chemistry, College

    of Chemistry and Molecular

    Engineering, Peking University,

    Beijing 100871, China

    *Corresponding author.

    Tel.: +86 10 62754976;

    Fax: +86 10 62751708;

    E-mail: [email protected],

    Trends in Analytical Chemistry, Vol. 39, 2012 Trends

    0165-9936/$ - see front matter 2012 Elsevier Ltd. All rights reserved. doi:http://dx.doi.org/10.1016/j.trac.2012.07.002 195

    mailto:[email protected],http://dx.doi.org/10.1016/j.trac.2012.07.002http://dx.doi.org/10.1016/j.trac.2012.07.002mailto:[email protected],
  • 7/30/2019 Seperations Nano

    2/12

    between chiral selectors and enantiomeric analytes, and

    one of them must be an enantioselective interaction. The

    final enantiomeric separation comes from multiple syn-

    ergetic interactions, including electrostatic, ion ex-

    change, hydrogen bonding, inclusion complex,

    hydrophobic, p-p, dipole-dipole, and steric interactions.

    Based on the above theory, many efficient chiral selec-tors with multiple interaction sites (e.g., proteins, poly-

    saccharide, cellulose, cyclodextrins and their derivatives)

    have been used as stationary phases or pseudo-station-

    ary phases in thin-layer chromatography (TLC), high-

    performance liquid chromatography (HPLC), gas

    chromatography (GC) and capillary electrophoresis (CE)

    through chemical bonding or physical coating.

    Recently, Wards et al. [5] published a critical review

    on chiral separation, which concluded that the devel-

    opment of monolithic chiral stationary phases, ionic li-

    quid phases and derivatized cyclodextrin phases would

    be the trend in this field. However, these conventional

    chromatographic or electromigration techniques are re-stricted by their low efficiency, high cost and poor uni-

    versality. Alternatively, nanomaterials with large

    surface-to-volume ratio and specific physical and

    chemical properties attracted more and more attention

    in recent years for their potential application in this field.

    Nanomaterials, with at least one dimension in the

    range 1100 nm, have become more and more popular

    in modern science. Due to the techniques of controllable

    synthesis of nanomaterials [6], they are widely used in

    efficient separation [7,8], sensitive measurement [9] and

    selective catalysis [10,11]. As demonstrated by Zhang

    et al. [12,13], the usage of nanomaterials provided spe-cial opportunities for the development of selective, stable

    and efficient techniques in separation science. There are

    several advantages of using nanomaterials in enantio-

    separation and related techniques. First, nanomaterials

    can be easily modified with chiral selectors and used as

    enantioselective adsorbents, especially chiral stationary

    or pseudo-stationary phases. Second, nanomaterials can

    increase column capacity, separation selectivity, stability

    and efficiency for chiral chromatographic separation.

    Third, nanomaterials, with unique surface-plasmon

    resonance (SPR) properties, can facilitate the develop-

    ment of simple, rapid and sensitive sensors for chiral

    recognition.In general, there are two main ways to prepare

    enantioselective nanomaterials. One is post modification

    of achiral nanomaterials using chiral selectors, and the

    other is synthesis of nanomaterials using chiral tem-

    plates. The enantioselective nanomaterials are expected

    Figure 1. (a) Chiral chromatographic separation; and, (b) enantioselective adsorption.

    Trends Trends in Analytical Chemistry, Vol. 39, 2012

    196 http://www.elsevier.com/locate/trac

  • 7/30/2019 Seperations Nano

    3/12

  • 7/30/2019 Seperations Nano

    4/12

    to interact with every enantiomer, forming locally-chiral

    structures according to different interaction mechanisms

    (e.g., van der Waals dispersion, short-range repulsions,

    and hydrogen-bonding interaction). The different sta-

    bility of the two diastereomeric complexes is the basis of

    enantioseparation. Fig. 1 describes the scheme of

    nanomaterials for chiral chromatographic separationand enantioselective adsorption.

    In this review, we focus on the applications of various

    functionalized nanomaterials in enantioseparation,

    which work through their specific enantioselective

    adsorption, especially used as stationary or pseudo-sta-

    tionary phases in chiral chromatographic columns (e.g.,

    TLC, HPLC, GC and CE). Table 1 sets out applications of

    both traditional and novel nanomaterials, including

    mesoporous silica, organic polymers, metal-organic

    frameworks (MOFs), metal nanoparticles (NPs), mag-

    netic nanoparticles (MNPs), carbon nanotubes (CNTs)

    and well-oriented chiral nanolayers.

    2. Mesoporous silica nanomaterials

    Mesoporous silica materials have been widely used in

    separations because of their stability, fast mass exchange

    and high thermal resistance [14]. The diversity of prep-

    aration and modification makes silica-based chiral

    nanomaterials promising in enantioseparation and

    enantioselective adsorption or release.

    Silica nanomaterials could be used as the support for

    chiral selectors to increase the phase ratio by non-

    covalent coating or chemical bonding. Physical coatingis commonly used due to its simplicity and convenience.

    Dong et al. [15] immobilized crystalline mesoporous sil-

    ica NPs onto the inner wall of a capillary, and then used

    them as the support for cellulose derivative through

    hydrogen bonding. Due to the increased surface area

    inside the capillary column, the silica-NP-modified cap-

    illary offered much higher enantioselectivity for eight

    pairs of enantiomers than a bare capillary with the same

    coating.

    In most cases, chemical bonding is more stable and

    robust than physical coating. Lee et al. [16] synthesized

    antibody-based silica-NT membranes for enantiomeric

    drug separations. These silica NTs were synthesized

    within the pores of alumina films, and then the antibody

    was attached to the inner walls of the NTs through

    chemical bonding. The enantiomer separation was

    realized through selective binding of drug enantiomers

    and antibodies with different affinities. The selective

    coefficient, a, which was defined as the ratio of two

    enantiomers content, could be as high as 2.0 0.2.

    Apart from the post-chiral modifications by chiral

    selectors, the template-based approach [17] was well

    recognized for the direct preparation of chiral silica-

    based nanomaterials. The templates could be chiral

    surfactants, chiral/achiral block copolymers and chiral

    complexes.

    Fireman-Shoresh et al. [18] prepared chiral silica

    nanomaterials by doping or imprinting the chiral sur-

    factant. These chiral nanomaterials exhibited good

    enantioselectivities with the chiral selective factors in the

    range 1.221.34 for three pairs of model enantiomers.Similarly, chiral mesoporous silica, synthesized by

    using a chiral block copolymer of polyethylene oxide and

    DD/LL-aspartic acid as the template [19], showed high

    enantioselectivity toward valine enantiomers with a

    chiral selective factor of 7.52.

    To some extent, it was not easy to obtain the chiral

    block copolymers and chiral surfactants, hence the

    synthesis of chiral NPs using achiral block copolymers

    and conventional surfactants provided new insight. Guo

    et al. [20] synthesized chiral silica-based porous mate-

    rials using a conventional achiral surfactant and a chiral

    cobalt complex as co-template. These chiral silica

    nanomaterials were successfully used for enantioselec-tive release of chiral drugs due to the chiral interactions

    between enantiomers and the chiral nanomaterials.

    Molecular imprinting [21], a technique to create

    template-shaped cavities in polymer matrices with

    memory of the template molecules to be used in molec-

    ular recognition, presents a novel approach for prepar-

    ing chiral silica nanomaterials, which possess high

    affinity and selectivity for a given target molecule. Fire-

    man-Shoresh et al. [22] synthesized silica-based chiral

    template-molecule-imprinted sol-gel thin films and

    measured the selective adsorption properties. Due to the

    high configuration match between the cavity and theadsorbed molecules, adsorption of a preferred enantio-

    mer was obtained and non-specific adsorption was

    remarkably suppressed.

    However, this traditional hydrolytic sol-gel technology

    often suffers the problems associated with cracking and

    shrinking of the beds during the drying process [23]. To

    overcome these limitations, Wang et al. [24] developed a

    novel room-temperature ionic liquid (RTIL)-mediated,

    non-hydrolytic sol-gel methodology for the preparation

    of silica-based molecular imprinting materials with the

    template of (S)-naproxen. RTILs with low vapor pressure

    could assist in reducing gel shrinkage, and their high

    ionic strength could increase the aggregation rate. With

    these chiral nanomaterials as the stationary phase in

    capillary eletrochromatography (CEC), the resolution

    between (S)-naproxen and (R)-naproxen reached 3.96

    under optimized conditions.

    3. Organic polymer nanomaterials

    Recently, the synthesis of polymer NPs with controlled

    characteristics has become an appealing topic [25]. Due

    to their easy processability, good permeability, diverse

    Trends Trends in Analytical Chemistry, Vol. 39, 2012

    198 http://www.elsevier.com/locate/trac

  • 7/30/2019 Seperations Nano

    5/12

    composition and satisfactory stability over a wide range

    of pH, polymer nanomaterials, including traditional

    polystyrene to novel core/shell NPs, have been exten-

    sively used in chiral separation techniques.

    Polystyrene is one of the most widespread polymeric

    NPs appreciated for its excellent stiffness and good pro-

    cessability. Na et al. [26] dispersed polystyrene NPs intothe running buffer of CE employing hydroxypropyl-

    b-cyclodextrin (HP-b-CD) as chiral selector to study the

    chiral separation of propranolol. Polystyrene NPs pro-

    vided a large surface for HP-b-CD adsorption and en-

    hanced the interaction between HP-b-CD and solutes,

    leading to higher separation efficiency at low HP-b-CD

    concentrations. The effects of polystyrene were most

    pronounced especially when the concentration of NPs in

    the running buffer was much less than that of the sol-

    utes. Above a certain concentration, further increase of

    NPs resulted in counteraction because of conglomera-

    tion. However, the buffer containing polystyrene NPs

    was not always superior to that without polystyreneNPs. For example, in the buffer without polystyrene NPs,

    the enantiomer resolution did not change remarkably as

    pH increased from 2.14 to 3.73, whereas it was obvi-

    ously influenced and decreased dramatically above pH

    2.50 for the buffer containing polystyrene NPs.

    When used in chiral separation, in situ polymerized

    molecularly imprinted polymers (MIPs) are still limited

    by the severe peak broadening and tailing caused by the

    slow dynamic process [27,28]. Boer et al. [29] developed

    a chiral separation method using NP-sized MIPs as a

    free-moving pseudo-stationary phase in CE to minimize

    these undesirable effects. They prepared

    100-nmspherical MIP particles and used them for the enantio-

    separation of ephedrine. With a partial filling technique,

    the ephedrine enantiomers were separated with accept-

    able efficiency, and the band broadening was reduced

    because of the smaller size of the NPs. Also, by incor-

    porating two different templates during the preparation

    process, a multiple-template method was developed to

    increase the binding capability of MIPs [30]. The affinity

    for the targets was strongly affected by the relative

    amounts of the two templates. Using a partial-filling

    technique in CE, the enantiomers of propranolol and

    ropivacaine were simultaneously separated using the

    multiple-template method. However, these MIPs are bulk

    materials and have to be triturated to desired sizes. Yin

    et al. [31] synthesized S-propranolol-imprinted NTs in

    the pores of anodic aluminum oxide to avoid trituration.

    These molecularly-imprinted NTs can be used for con-

    trolled release of target chiral drug with high capacity

    and good site accessibility.

    Similar to silica-based nanomaterials, chiral organic

    nanomaterials could also be obtained through template-

    based methods. Chiral mesoporous polypyrrole NPs were

    prepared using a chiral block copolymer of polyethylene

    oxide as the template [32]. After extracting the template,

    the chiral resolution toward racemic valine and alanine

    was confirmed by circular dichroism and optical polar-

    imetry. As the chiral block copolymers were limited by

    their species and preparation, novel helical poly(phen-

    ylacetylene) was designed with bulky crown ether as

    pendant [33]. This polymer showed high sensitivity to

    chiral amino acid and could detect an extremely smallenantiomeric imbalance in a-amino acids (less than

    0.005% enantiomeric excess of alanine).

    Hybrid materials consisting of organic and inorganic

    building blocks, especially core/shell NPs, could combine

    the advantages of the individual components. In some

    cases, using silica as the shell can protect the core from

    aggregation and minimize the environmental influences.

    Chen et al. [34] reported chiral hybrid NPs consisting of

    an optically active helical polyacetylene core and a silica

    shell. The hybrid NPs possessed large optical activities,

    arising from the helical core, and at the same time, the

    silica shell provided desirable protection for the core. The

    chiral recognition ability was confirmed by enantiose-lective crystallization of amino-acid enantiomers. Re-

    cently, the same group improved the preparation

    method and obtained hollow two-layered chiral NPs [35]

    by removing the optically active helical polymer. These

    hollow NPs exhibited low density, high stability, good

    dispersity and considerable enantioselectivity toward

    racemic alanine.

    4. Metal-organic frameworks

    Metal-organic frameworks (MOFs), known as a sub-classof the coordination polymers, are crystalline materials

    generated by the association of metal ions and organic

    ligands [36]. They have emerged as a promising type of

    material for their potential applications in gas storage,

    separation and catalysis due to their fascinating struc-

    tures [37,38]. In recent years, many chiral MOFs, syn-

    thesized by incorporating enantiomerically pure

    homochiral building blocks and occasionally achiral

    building blocks, have been used in enantioselective sep-

    aration [36,38].

    Xiong et al. [39] first reported a homochiral 3D zeolite-

    like MOF, which was prepared by treating chiral bridging

    ligand quitenine with Cd(OH)2, for the enantioselective

    adsorption of racemic 2-butanol. The single-crystal

    X-ray diffraction (XRD) study performed on the complex

    of chiral MOFs(S)-2-butanol clearly revealed that (S)-2-

    butanol was wrapped in the chiral cavities of the

    network. After being heated to a suitable temperature,

    (S)-2-butanol could be completely removed, and the ee

    value of the desorbed (S)-2-butanol was estimated to be

    as high as 98.2%.

    Bradshaw et al. [40] recently reported several homo-

    chiral MOFs whose enantioselective sorption strongly

    depended on guest size. For example, the binaphthol

    Trends in Analytical Chemistry, Vol. 39, 2012 Trends

    http://www.elsevier.com/locate/trac 199

  • 7/30/2019 Seperations Nano

    6/12

    guest was enantioselectively adsorbed onto MOFs with

    an enantiomeric excess of 8.3%, while no enantioselec-

    tion toward ethyl-3-hydroxybutyrate, terpenes menth-

    one and fenchone was observed. Also, many other chiral

    MOFs with high enantioselectivity were synthesized

    using amino-acid backbones or cyclic clusters with en-

    zyme-like chiral cavities [4143]. Moreover, creatinghomochiral MOFs from the achiral building blocks for

    the resolution of racemic molecules was always a great

    challenge. In 2006, Xiong et al. [44] prepared chiral

    MOFs by unstirred crystallization [45] of achiral building

    blocks, and described the enantiomeric resolution of

    racemic 2-butanol. This work provided new insight into

    the enantioseparation methods by chiral MOFs.

    Compared with enantioselective adsorption, chiral

    chromatographic separation is more efficient. Nuzhdin

    et al. [46] developed a column chromatography with

    chiral MOFs as the chiral stationaryphase. A 33-cm chiral

    column was prepared by loading a glass tube (8 mm inner

    diameter) with the suspension of chiral MOFs. The modelracemic compounds were directly loaded onto the top of

    the column, and complete baseline separation of sulfoxide

    was achieved. They studied the separation mechanism by

    introducing electro-withdrawing substituents in the aro-

    matic ring of sulfoxides. The results revealed that both the

    sorption constants and ees of sorption were decreased due

    to the reduced coordinating ability of the sorbate induced

    by the electronic effects.

    Padmanaban et al. [47] also developed a chiral HPLC

    column employing highly ordered chiral MOFs as the

    stationary phase for enantioseparation. Usually, these

    MOF-packed columns could not provide high resolutionbecause of the considerable diffusion resistance of bulky

    packing. As a result, the MOF-coated capillary chiral

    column was recently developed. Xie et al. [48] prepared

    chiral open-tubular columns for high-resolution GC

    separation of chiral compounds. The chiral MOF with a

    single-handed helical structure was coated onto the

    capillary by the dynamic coating method. After being

    heated, the single-stranded helices lost water molecules

    and cross-linked to a chiral open framework, which was

    critical for subsequent chiral separation. The columns

    showed excellent selectivity and possessed good recog-

    nition ability toward a wide range of organic compounds

    (e.g., racemates, isomers, alkanes and alcohols).

    5. Metal nanomaterials

    In recent years, chiral metal NPs have shown various

    unexpected advantages in asymmetry catalysis [4952],

    chiral separation [5355] and enantioselective detection

    [56,57]. Here, we mainly discuss the reports about metal

    nanomaterials applied in chiral separation or enantiose-

    lective detection. For the applications on asymmetry

    synthesis, readers can refer to the recent reviews [5860].

    Gold NPs (AuNPs) are very attractive in nanotech-

    nology [6163] due to their ease of preparation, con-

    trollable particle size, narrow size distribution, good

    solubility in buffer and convenient modification via the

    Au-S bond. Dispersing chiral selector-modified AuNPs in

    running buffer to act as the chiral pseudo-stationary

    phase in CE is easy. Yang et al. synthesized thiolatedbCD modified AuNPs and dispersed them in running

    buffer for chiral separation of enantiomers [53]. Efficient

    baseline separation with theoretical plate numbers up to

    2.4 105 and chiral separation resolution up to 4.7 was

    achieved with very low concentration of thiolated bCD

    modified AuNPs. The corresponding concentration of

    thiolated bCD modified AuNPs in the running buffer

    was only 0.300.53 mM, which was much lower than

    the optimum concentration of 15 mM if pure bCD was

    used. These results indicated that thiolated bCD-modi-

    fied AuNPs can provide sufficient interaction with the

    analytes through the increased surface area. This

    method was restricted from repeated use as the runningbuffer had to be updated during experiments. Hence, the

    same group developed another enantioselective CE

    method by immobilizing the thiolated b-CD-modified

    AuNPs onto the inner wall of a capillary [54]. The fused-

    silica capillary was first coated with positively-charged

    poly(diallydimethylammonium chloride), and then the

    negatively-charged thiolated b-CD-modified AuNPs were

    assembled onto the capillary by electrostatic interaction.

    The capillary column constructed had high stability,

    steady electroosmotic flow (EOF) mobility over a wide pH

    range 3.09.2 and good column-to-column reproduc-

    ibility.Chiral-modified AuNPs could also be chemically

    bonded to the pre-derivatized capillary by the Au-S bond,

    to get a more stable separation performance. Generally,

    the capillary or the microchannel was pre-derived with

    (3-mercaptopropyl)-trimethoxysilane to provide thiol

    groups, and then chiral-modified AuNPs could be linked

    onto the surface by the Au-S bond. Based on this

    chemistry, a chip-based enantioselective CE employing

    bovine serum albumin (BSA)-modified AuNPs as chiral

    stationary phase was developed [55]. Compared with the

    untreated capillary, the incorporation of AuNPs greatly

    increased the phase ratio and favored the generation of

    sufficient EOF via the negatively-charged citrate. This

    micro-device exhibited excellent enantiomeric selectivity

    and good run-to-run repeatability in the enantiosepa-

    ration of ephedrine and norephedrine. In addition, this

    capillary could keep an enantioselective lifetime for more

    than 1 month. Similarly, the BSA-modified AuNPs were

    also incorporated into a silica monolith and used as

    chiral stationary phase for chiral separation [64].

    AuNPs were also used for enantioselective adsorption

    [56] and recognition [65]. Shukla et al. [56] employed

    either DD- or LL-cysteine-modified AuNPs for enantioselec-

    tive adsorption of propylene oxide. The enantioselective

    Trends Trends in Analytical Chemistry, Vol. 39, 2012

    200 http://www.elsevier.com/locate/trac

  • 7/30/2019 Seperations Nano

    7/12

    adsorption process was monitored by optical rotation

    measurements, which were based on the rotation of

    polarized light by (R)- and (S)-propylene oxide being en-

    hanced through interaction with AuNPs. The results

    demonstrated that LL-cysteine (DD-cysteine)-modified

    AuNPs could selectively adsorb the (R)-propylene oxide

    ((S)-propylene oxide). When exposed to racemic propyl-

    ene oxide, the chiral AuNPs selectively adsorbed one

    enantiomer and gave an enantiomeric excess in the

    solution phase, thereby inducing enantioselective sepa-

    ration. Similarly, an electrochemical sensor for enantio-

    selective recognition of 3,4-dihydroxyphenylalanine was

    developed employing penicillamine-modified AuNPs [65].

    With regard to the method discussed above, working

    with specific instruments is necessary, so it is still highly

    desirable to develop a simple, rapid, sensitive and

    instrument-independent method for chiral recognition.

    One of the most urgent challenges is to achieve visual

    discrimination of enantiomers by an appropriate color

    change. Metal nanomaterial-based colorimetric sensors,

    due to their simplicity, high sensitivity and lower cost,

    were promising in this field. An ultra-efficient enantio-separation and detection platform for DD- and LL-cysteine

    was therefore developed by employing uridine 5-tri-

    phosphate-capped AgNPs (Fig. 2) [57]. The aggregation

    of 5-triphosphate-capped AgNPs was selectively induced

    by DD-cysteine, which allowed the rapid enantiomeric

    discrimination of racemic cysteine without any prior

    derivatization or specific instruments. After centrifuga-

    tion, an excess of the other enantiomer was left in the

    solution and resulted in an enantioseparation. Simplicity

    and efficiency made this method promising for the

    future.

    6. Magnetic nanomaterials

    Derivatized magnetic NPs (MNPs) with large surface

    areas could form complexes with target molecules via

    multiple interactions. At the same time, the complexes

    could be conveniently separated from the solution by

    applying magnetic field [14]. Compared with conven-

    tional separation methods, this process is noted for its

    speed, simplicity and cost effectiveness. MNPs, immobi-

    lized with an appropriate chiral catalyst or chiral selec-

    tor, have been successfully used for asymmetric

    synthesis [66] and enantioseparation [6770]. Here, we

    focus on the work employing chiral selector-modified

    MNPs for enantioseparation. MNPs tagged with an

    appropriate chiral selector were expected to interact

    selectively with a single enantiomer and form complexes,

    which were easily removed by an external magnet. The

    enantioseparation process is just like enantioselective

    fishing.

    Ghosh et al. [67] synthesized carboxymethyl-b-cyclo-

    dextrin-bonded Fe3O4/SiO2 core-shell NPs, and applied

    these MNPs in the enantioselective adsorption of chiralamino acids. Adsorption-equilibrium experiments dem-

    onstrated that the adsorption capacities were higher for

    LL-enantiomers than the corresponding DD-enantiomers.

    But the enantioselective adsorption ability for racemic

    solution was not investigated.

    Choi et al. [68] developed a method employing race-

    mic amino-acid solutions as model compounds for fur-

    ther study (Fig. 3). First, MNPs with (S)-chiral selector

    (MNPs/(S)-CS) were added into the racemic solution for

    enantioselective adsorption. The enantiomeric excess

    was evaluated by chiral HPLC and confirmed the

    Figure 2. Colorimetric discrimination of LL-cysteine and DD-cysteine using 5-triphosphate-capped Ag nanoparticles [57] (Reprinted with permis-sion from [57], 2012, American Chemical Society).

    Trends in Analytical Chemistry, Vol. 39, 2012 Trends

    http://www.elsevier.com/locate/trac 201

  • 7/30/2019 Seperations Nano

    8/12

    enantioselectivity of MNPs/(S)-CS. But the enantioselec-

    tivity was not stratified. Consequently, another method

    was developed, employing equal non-magnetic silica

    particles tagged to the antipode (R)-chiral selector

    (MNPs/(R)-CS) and MNPs/(S)-CS at the same time. After

    extraction, the MNPs/(S)-CS-enantiomer complexes were

    collected by a magnet, while the MNPs/(R)-CS-enantio-

    mer complexes were collected by decantation after cen-

    trifugation. These results of chiral HPLC separation

    demonstrated that the enantioselectivity was improved

    when MNPs/(R)-CS was simultaneously used. This

    method indicated that the two contrary chiral selectors

    Figure 3. The separation of enantiomers using (a) MNPs/(S)-CS only; and, (b) MNPs/(S)-CS and NMPs/(R)-CS at the same time [68] (Reprintedwith permission from [68], 2012, Royal Society of Chemistry).

    Figure 4. Preparation procedures and separation mechanism ofb-CD modified Fe3O4 nanoparticles for amino acids enantiomers [69] (Reprintedwith permission from [69], 2012, Royal Society of Chemistry).

    Trends Trends in Analytical Chemistry, Vol. 39, 2012

    202 http://www.elsevier.com/locate/trac

  • 7/30/2019 Seperations Nano

    9/12

    could have some synergistic effects for enantioseparation

    and provided a new insight to the development of a high-

    efficiency enantioselective method.

    In most conditions, washing to retrieve the target

    molecules was necessary for further experiments. But

    elution made the method complicated, and the organic

    solvent was not economic or environmental friendly. Asan alternative, Chen et al. [69] developed a photo-con-

    trolled inclusion-and-exclusion method to induce the

    release of the adsorbed enantiomers in reverse (Fig. 4).

    The chiral selective system also employed b-CD-func-

    tionalized Fe3O4 NPs as chiral adsorbents. After extrac-

    tion, pure optical enantiomer adsorbed on the NPs was

    released via the photo-controlled inclusion-and-

    exclusion reactions of azobenzene derivatives with b-CD.

    The reversible photo-controlled inclusion-and-exclusion

    process enabled recycled use of b-CD-functionalized

    Fe3O4 NPs in enantioseparation. It is worth mentioning

    that, after repeated use for more than 5 times, b-CD-

    modified Fe3O4 NPs still showed good enantioselectivityfor LL-tryptophan, indicating that b-CD-modified Fe3O4NPs were highly applicable in repeated recycling.

    Another solution, packing chiral MNPs into microchip

    channels, was developed to achieve complete enantio-

    separation [70]. Molecularly-imprinted MNPs, with S-

    ofloxacin as the template molecule, were located as

    stationary phase in the microchannel of the microfluidic

    device. Under optimized conditions, efficient molecular

    recognition of ofloxacin enantiomers was achieved in

    195 s with good reproducibility and precision. Compared

    with a conventional device, the MNPs could be conve-

    niently localized to the predetermined position byapplying an external magnetic field, and the length of

    the packing zone could be easily tuned by changing the

    length of the magnet.

    7. Carbon nanotubes

    CNTs with good absorbability and electrical properties

    have been used in many fields {e.g., electrochemical

    biosensors [71] and chromatographic separation [72]}.

    Physical studies revealed the existence of chirality in the

    structure of CNTs [73], but CNTs alone have not so far

    been successfully used in enantioseparation. Theoretical

    studies and experimental results showed that CNTs

    cannot supply enough energy differences between

    enantiomers [74,75], so adding chiral selectors (e.g.,

    b-CD) or an achiral assistant was necessary when CNTs

    were used for chiral separation.

    In general, CNTs are insoluble in most common sol-

    vents, due to the aggregation induced by van der Waals

    attractions. In order to produce pseudo-stationary pha-

    ses in CE, one effective solution was using surfactant-

    coated CNTs {e.g., the CE separation of ()-clenbuterol

    was performed using b-CD-modified multi-walled CNTs

    (MWCNTs) [76]}. Compared with the other three

    nanomaterials, b-CD-MWCNTs showed better results,

    because MWCNTs formed a network in the running

    buffer and presented a larger surface area to allow better

    contact with the analytes.

    However, the addition of long-chain surfactants was

    not always suitable for specific systems. Another solutionwas to cut the single-walled CNTs (SWCNTs) into short

    pipes by chemical oxidation in a mixture of concentrated

    sulfuric and nitric acids [77]. On the one hand, the

    shortened SWCNTs could be well dispersed in distilled

    water after ultrasonic agitation, benefiting from the

    activated hydrophilic groups on the surface. On the

    other hand, the shortened SWCNTs could offer more

    carboxylic groups for binding chiral selectors. When the

    short SWCNTs were functionalized with BSA to form

    BSA-SWCNT conjugates, this showed that, without the

    BSA-SWCNT conjugates, no separation of tryptophan

    enantiomers was achieved. After the addition of 0.075

    0.1 mg/mL of BSA-SWCNT conjugates, baseline sepa-ration was achieved, and the resolution increased along

    with the increase of BSA-SWCNT conjugates. Under

    optimal conditions, successful separation of tryptophan

    enantiomers was achieved in less than 70 s with a res-

    olution factor of 1.35, illustrating the acceptable

    enantioselectivity of the BSA-SWCNT conjugates.

    Similarly, HP-b-CD was successfully bonded to

    MWCNTs, and the HP-b-CD-modified MWCNTs were

    used as additives in the TLC stationary phase for chiral

    separation [78]. Clenbuterol enantiomers were separated

    with better resolution. This method could benefit from

    the advantages of TLC, including high sample through-put, accessibility and cost effectiveness, to construct a

    rapid, robust chiral separation method.

    Apart from the above methods, CNTs could be bonded

    to the inner wall of a capillary and applied as the support

    for chiral selectors. To investigate the effect of SWCNTs

    on enantioseparation, Zhao et al. [79] prepared two

    capillary columns for GC, one containing the chiral ionic

    liquid, and the other containing the complex of chiral

    ionic liquid and SWCNTs. Taking 12 racemates as model

    compounds, the column with SWCNTs was able to sep-

    arate eight chiral compounds, while the column without

    SWCNTs was only able to separate four racemates. The

    SEM images showed that most of the bonded SWCNTs

    were linked end to end and formed a layer with a skeletal

    network structure. Therefore, the SWCNT-modified

    capillary column possessed a larger surface for the chiral

    ionic liquid stationary phase, which enhanced the

    interaction between the chiral stationary phases and the

    analytes. Compared with dispersal in running buffer, the

    chemical-bonding approach provided more stable, reus-

    able capillaries.

    In addition to the above chiral-modified CNTs, recent

    research demonstrated that achiral-modified CNTs also

    exhibited ability for chiral recognition. Martnez et al.

    Trends in Analytical Chemistry, Vol. 39, 2012 Trends

    http://www.elsevier.com/locate/trac 203

  • 7/30/2019 Seperations Nano

    10/12

    [80] prepared achiral surfactant-coated MWCNTs via

    sonication of MWCNTs and SDS until they formed a final

    homogeneous phase. The modified MWCNTs were stable

    under high electric field and compatible with the detec-

    tor. What is more, sonication allowed modifications of

    the chiral properties of CNTs. The resulting material was

    then used for chiral separation using the partial-fillingtechnique. The enantiomers (+)- and (-)-norephedrine,

    (-)-ephedrine and (+)-N-methylephedrine were com-

    pletely resolved, demonstrating that the inherent chi-

    rality of MWCNTs was amplified through the

    modification and sonication. This work demonstrated

    the possibility of directly using carbon nanostructures as

    chiral selectors for the separation of racemates.

    8. Well-oriented nanolayers

    Many kinds of well-oriented metal solid surfaces, with

    modification or not, have the ability of chiral recognitionwhen they interact with chiral molecules [81]. Chiral

    surfaces can be classified into two main categories: sur-

    faces derived from achiral bulk structures; and, those

    produced by chiral molecules self-assembling onto an

    achiral surface.

    In the first method, a chiral surface was obtained by

    cutting single crystals of bulk achiral metals along a

    specific crystal face. The periodic arrayed surfaces ob-

    tained can be used for chiral recognition. For example,

    Au (110) surfaces can adsorb enantioselectively. Con-

    sidering the affinity between the thiol group and Au,

    cysteine was selected as a model compound [82]. Thescanning tunneling microscope (STM) image showed

    that the deposition of LL-cysteine molecules on the Au

    (110) surface resulted in the formation of bright pro-

    trusions at the sides of the close-packed rows of gold

    atoms. The protrusions always existed as pairs, and the

    main axis running through the two bright protrusions

    was always rotated 20 clockwise with respect to the

    close-packed direction. When the DD-cysteine was depos-

    ited, similar molecule pairs rotated 20 anticlockwise.

    Moreover, for racemic mixtures, the STM images also

    showed molecular dimers identical to the previous

    measurements. These exclusively homochiral molecular

    pairs indicated the chiral recognition ability of the Au

    (110) surface. For further studies, Au (111) and Pt (111)

    surfaces were also confirmed to be able to adsorb

    enantioselectively [83,84].

    Another method of obtaining a chiral surface is to

    immobilize chiral molecules onto an achiral surface by

    self-assembling. The chirality of the surface is derived

    from the adsorbed chiral ligands that retain their chi-

    rality. Through this method, five molecules, including

    LL-cysteine, DD-cysteine, N-acetyl-LL-cysteine, LL-glutathione

    and DD-penicillamine, were assembled on a gold surface.

    The self-assembled monolayers were used in crystalliza-

    tion experiments for supersaturated racemic solutions

    and supersaturated solutions with enantiomeric excess.

    [85]. Crystallization experiments indicated that, when

    starting with racemic solutions, one enantiomer in ex-

    cess was obtained, while crystals of the pure enantiomer

    were obtained when starting with a solution having an

    eevalue of 50%. The results revealed that these chiralself-assembling monolayers had the ability to adsorb

    enantioselectively, but this ability was still restricted to

    the enantiomeric excess of the solution.

    Similarly, many other chiral molecules (e.g., phenyl-

    alanine) were also distinguished by this method [86].

    Self-assembling of chiral molecules with higher selec-

    tivity may improve the enantioselectivity of well-oriented

    nanolayers in the future.

    9. Conclusions

    As an important and universal phenomenon in nature,chirality is related to many significant biological activi-

    ties. The development of efficient enantioseparation

    technologies is extraordinarily significant for the phar-

    maceutical industry and asymmetric synthesis. A wide

    range of nanomaterials with large surface area and

    specific chemical and physical properties have been used

    in enantioseparation, including mesoporous silica, or-

    ganic polymers, MOFs, metal NPs, MNPs, CNTs and well-

    oriented chiral nanolayers, to improve the selectivity,

    stability and efficiency. The chiral recognition of

    nanomaterials is realized through enantioselective

    adsorption, especially when they are used as stationary/pseudo-stationary phases in chiral chromatographic

    (HPLC, TLC and GC) and CE separation.

    Although many novel nanomaterials have been ap-

    plied in chiral separation and remarkable separations

    have been achieved, the mechanism of nanomaterials in

    chiral separation remains unclear and untouched in the

    literature. The most likely explanation is still based on

    the three-point interaction theory, which lacks theo-

    retical and experimental illustration and need to be

    further investigated. Under the guidance of current

    theory, the design and the synthesis of nanomaterials

    with high selectivity, stability, reusability and efficiency

    are the future orientation of nanomaterials applied in

    enantioseparation. Magnetic and metal nanomaterials

    will have high potential in this field. Magnetic nanom-

    aterials allow fast, convenient separation from the

    solution. Metal-based nanomaterials, with unique SPR

    properties, are suitable for the development of simple,

    rapid and sensitive sensors for chiral recognition.

    Acknowledgements

    This work was financially supported by the National

    Natural Science Foundation of China (Grant No.

    Trends Trends in Analytical Chemistry, Vol. 39, 2012

    204 http://www.elsevier.com/locate/trac

  • 7/30/2019 Seperations Nano

    11/12

    21027012) and the fundamental research funds for the

    central universities.

    Appendix A. Supplementary data

    Supplementary data associated with this article can befound, in the online version, at http://dx.doi.org/

    10.1016/j.trac.2012.07.002.

    References[1] T.L.V. Ulbricht, Nature (London) 258 (1975) 383.

    [2] J.B. Youatt, R.D. Brown, Science (Washington, DC) 212 (1981)

    1145.

    [3] G. Procter, Stereoselectivity in Organic Synthesis, Oxford Univer-

    sity Press, New York, USA, 1996.

    [4] C.E. Dalgliesh, J. Chem. Soc. (1952) 3940.

    [5] T.J. Ward, K.D. Ward, Anal. Chem. 84 (2012) 626.

    [6] T.K. Sau, A.L. Rogach, Adv. Mater. 22 (2010) 1781.

    [7] A.H. Duan, S.M. Xie, L.M. Yuan, Trends Anal. Chem. 30 (2011)

    484.

    [8] R. Sancho, C. Minguillon, Chem. Soc. Rev. 38 (2009) 797.

    [9] F.P. Zamborini, L. Bao, R. Dasari, Anal. Chem. 84 (2012) 541.

    [10] J.M. Campelo, D. Luna, R. Luque, J.M. Marinas, A.A. Romero,

    ChemSusChem 2 (2009) 18.

    [11] M. Zahmakiran, S. Ozkar, Nanoscale 3 (2011) 3462.

    [12] Z.X. Zhang, B. Yan, Y.P. Liao, H.W. Liu, Anal. Bioanal. Chem.

    391 (2008) 925.

    [13] Z.X. Zhang, B. Yan, K.L. Liu, Y. Liao, H.W. Liu, Electrophoresis 30

    (2009) 379.

    [14] L.G. Chen, T. Wang, J. Tong, Trends Anal. Chem. 30 (2011)

    1095.

    [15] X.L. Dong, R.A. Wu, J. Dong, M.H. Wu, Y. Zhu, H.F. Zou,

    Electrophoresis 29 (2008) 3933.

    [16] S.B. Lee, D.T. Mitchell, L. Trofin, T.K. Nevanen, H. Soderlund, C.R.

    Martin, Science (Washington, DC) 296 (2002) 2198.

    [17] N.K. Raman, M.T. Anderson, C.J. Brinker, Chem. Mater. 8 (1996)

    1682.

    [18] S. Fireman-Shoresh, S. Marx, D. Avnir, Adv. Mater. 19 (2007)

    2145.

    [19] P. Paik, A. Gedanken, Y. Mastai, Microporous Mesoporous Mater.

    129 (2010) 82.

    [20] Z. Guo, Y. Du, X.B. Liu, S.C. Ng, Y. Chen, Y.H. Yang, Nanotech-

    nology 21 (2010) 285706.

    [21] C. Alexander, H.S. Andersson, L.I. Andersson, R.J. Ansell, N.

    Kirsch, I.A. Nicholls, J. OMahony, M.J. Whitcombe, J. Mol.

    Recognit. 19 (2006) 106.

    [22] S. Fireman-Shoresh, D. Avnir, S. Marx, Chem. Mater. 15 (2003)

    3607.

    [23] A.M. Siouffi, J. Chromatogr., A 1000 (2003) 801.[24] H.F. Wang, Y.Z. Zhu, X.P. Yan, R.Y. Gao, J.Y. Zheng, Adv. Mater.

    18 (2006) 3266.

    [25] M.K. Aiertza, I. Odriozola, G. Cabanero, H.J. Grande, I. Loinaz,

    Cell. Mol. Life Sci. 69 (2012) 337.

    [26] N. Na, Y.P. Hu, J. Ouyang, W.R.G. Baeyens, J.R. Delanghe, T. De

    Beer, Anal. Chim. Acta 527 (2004) 139.

    [27] R.J. Ansell, Adv. Drug Delivery Rev. 57 (2005) 1809.

    [28] N.M. Maier, W. Lindner, Anal. Bioanal. Chem. 389 (2007) 377.

    [29] T. de Boer, R. Mol, R.A. de Zeeuw, G.J. de Jong, D.C.

    Sherrington, P.A.G. Cormack, K. Ensing, Electrophoresis 23

    (2002) 1296.

    [30] P. Spegel, L. Schweitz, S. Nilsson, Anal. Chem. 75 (2003) 6608.

    [31] J.F. Yin, Y. Cui, G.L. Yang, H.L. Wang, Chem. Commun. (2010)

    7688.

    [32] P. Paik, A. Gedanken, Y. Mastai, J. Mater. Chem. 20 (2010) 4085.

    [33] R. Nonokawa, E. Yashima, J. Am. Chem. Soc. 125 (2003) 1278.

    [34] B. Chen, J.P. Deng, L.Y. Tong, W.T. Yang, Macromolecules 43

    (2010) 9613.

    [35] B. Chen, J.P. Deng, W.T. Yang, Adv. Funct. Mater. 21 (2011)

    2345.

    [36] Y. Liu, W.M. Xuan, Y. Cui, Adv. Mater. 22 (2010) 4112.

    [37] A.U. Czaja, N. Trukhan, U. Muller, Chem. Soc. Rev. 38 (2009)

    1284.

    [38] G. Nickerl, A. Henschel, R. Grunker, K. Gedrich, S. Kaskel, Chem.

    Ing. Tech. 83 (2011) 90.

    [39] R.G. Xiong, X.Z. You, B.F. Abrahams, Z.L. Xue, C.M. Che, Angew.

    Chem., Int. Edit. Engl. 40 (2001) 4422.

    [40] D. Bradshaw, T.J. Prior, E.J. Cussen, J.B. Claridge, M.J. Rosseinsky,

    J. Am. Chem. Soc. 126 (2004) 6106.

    [41] R. Vaidhyanathan, D. Bradshaw, J.N. Rebilly, J.P. Barrio, J.A.

    Gould, N.G. Berry, M.J. Rosseinsky, Angew. Chem., Int. Edit. Engl.

    45 (2006) 6495.

    [42] G. Li, W.B. Yu, J. Ni, T.F. Liu, Y. Liu, E.H. Sheng, Y. Cui, Angew.

    Chem., Int. Edit. Engl. 47 (2008) 1245.

    [43] G. Li, W.B. Yu, Y. Cui, J. Am. Chem. Soc. 130 (2008) 4582.

    [44] Y.M. Song, T. Zhou, X.S. Wang, X.N. Li, R.G. Xiong, Cryst.

    Growth Des. 6 (2006) 14.

    [45] D.K. Kondepudi, R.J. Kaufman, N. Singh, Science (Washington,

    DC) 250 (1990) 975.

    [46] A.L. Nuzhdin, D.N. Dybtsev, K.P. Bryliakov, E.P. Talsi, V.P. Fedin,

    J. Am. Chem. Soc. 129 (2007) 12958.

    [47] M. Padmanaban, P. Muller, C. Lieder, K. Gedrich, R. Grunker, V.

    Bon, I. Senkovska, S. Baumgartner, S. Opelt, S. Paasch, E.

    Brunner, F. Glorius, E. Klemm, S. Kaskel, Chem. Commun. (2011)

    12089.

    [48] S.M. Xie, Z.J. Zhang, Z.Y. Wang, L.M. Yuan, J. Am. Chem. Soc.

    133 (2011) 11892.

    [49] D.F. Han, X.H. Li, H.D. Zhang, Z.M. Liu, G.S. Hu, C. Li, J. Mol.

    Catal. A: Chem. 283 (2008) 15.

    [50] K. Sawai, R. Tatumi, T. Nakahodo, H. Fujihara, Angew. Chem.,

    Int. Edit. Engl. 47 (2008) 6917.

    [51] S. Reimann, T. Mallat, A. Baiker, J. Catal. 254 (2008) 79.

    [52] G. Szollosi, A. Mastalir, Z. Kiraly, I. Dekany, J. Mater. Chem. 15

    (2005) 2464.

    [53] L. Yang, C.J. Chen, X. Liu, J. Shi, G.A. Wang, L.D. Zhu, L.P. Guo,

    J.D. Glennon, N.M. Scully, B.E. Doherty, Electrophoresis 31

    (2010) 1697.

    [54] M. Li, X. Liu, F.Y. Jiang, L.P. Guo, L. Yang, J. Chromatogr., A

    1218 (2011) 3725.

    [55] H.F. Li, H.L. Zeng, Z.F. Chen, J.M. Lin, Electrophoresis 30 (2009)

    1022.

    [56] N. Shukla, M.A. Bartel, A.J. Gellman, J. Am. Chem. Soc. 132

    (2010) 8575.

    [57] M. Zhang, B.C. Ye, Anal. Chem. 83 (2011) 1504.

    [58] Y.L. Huang, S. Xu, V.S.Y. Lin, Chemcatchem 3 (2011) 690.

    [59] T. Mallat, E. Orglmeister, A. Baiker, Chem. Rev. 107 (2007)

    4863.

    [60] S. Roy, M.A. Pericas, Org. Biomol. Chem. 7 (2009) 2669.

    [61] Y.L. Liu, K.B. Male, P. Bouvrette, J.H.T. Luong, Chem. Mater. 15

    (2003) 4172.

    [62] M.C. Daniel, D. Astruc, Chem. Rev. 104 (2004) 293.

    [63] C. Gautier, T. Burgi, ChemPhysChem 10 (2009) 483.

    [64] J.Y. Lu, F.G. Ye, A.Z. Zhang, Z. Wei, Y. Peng, S.L. Zhao, J. Sep. Sci.

    34 (2011) 2329.

    [65] Y.J. Kang, J.W. Oh, Y.R. Kim, J.S. Kim, H. Kim, Chem. Commun.

    (2010) 5665.

    [66] A.G. Hu, G.T. Yee, W.B. Lin, J. Am. Chem. Soc. 127 (2005)

    12486.

    [67] S. Ghosh, A.Z.M. Badruddoza, M.S. Uddin, K. Hidajat, J. Colloid

    Interface Sci. 354 (2011) 483.

    [68] H.J. Choi, M.H. Hyun, Chem. Commun. (2009) 6454.

    Trends in Analytical Chemistry, Vol. 39, 2012 Trends

    http://www.elsevier.com/locate/trac 205

    http://dx.doi.org/10.1016/j.trac.2012.07.002http://dx.doi.org/10.1016/j.trac.2012.07.002http://dx.doi.org/10.1016/j.trac.2012.07.002http://dx.doi.org/10.1016/j.trac.2012.07.002
  • 7/30/2019 Seperations Nano

    12/12

    [69] X. Chen, J.A. Rao, J. Wang, J.J. Gooding, G. Zou, Q.J. Zhang,

    Chem. Commun. (2011) 10317.

    [70] P. Qu, J.P. Lei, L. Zhang, R.Z. Ouyang, H.X. Ju, J. Chromatogr., A

    1217 (2010) 6115.

    [71] R.T. Kachoosangi, M.M. Musameh, I. Abu-Yousef, J.M. Yousef,

    S.M. Kanan, L. Xiao, S.G. Davies, A. Russell, R.G. Compton, Anal.

    Chem. 81 (2009) 435.

    [72] Y. Moliner-Martinez, S. Cardenas, B.M. Simonet, M. Valcarcel,

    Electrophoresis 30 (2009) 169.

    [73] E.B. Barros, A. Jorio, G.G. Samsonidze, R.B. Capaz, A.G. Souza, J.

    Mendes, G. Dresselhaus, M.S. Dresselhaus, Phys. Rep. 431 (2006)

    261.

    [74] T.D. Power, A.I. Skoulidas, D.S. Sholl, J. Am. Chem. Soc. 124

    (2002) 1858.

    [75] B. Suarez, B.M. Simonet, S. Cardenas, M. Valcarcel, Electropho-

    resis 28 (2007) 1714.

    [76] N. Na, Y.P. Hu, O.Y. Jin, W.R.G. Baeyens, J.R. Delanghe, Y.E.C.

    Taes, M.X. Xie, H.Y. Chen, Y.P. Yang, Talanta 69 (2006) 866.

    [77] X.X. Weng, H.Y. Bi, B.H. Liu, J.L. Kong, Electrophoresis 27 (2006)

    3129.

    [78] J.G. Yu, D.S. Huang, K.L. Huang, Y. Hong, Chin. J. Chem. 29

    (2011) 893.

    [79] L. Zhao, P. Ai, A.H. Duan, L.M. Yuan, Anal. Bioanal. Chem. 399

    (2011) 143.

    [80] Y. Moliner-Martinez, S. Cardenas, M. Valcarcel, Electrophoresis 28

    (2007) 2573.

    [81] Y. Mastai, Chem. Soc. Rev. 38 (2009) 772.

    [82] A. Kuhnle, T.R. Linderoth, B. Hammer, F. Besenbacher, Nature

    (London) 415 (2002) 891.

    [83] Q.H. Yuan, C.J. Yan, H.J. Yan, L.J. Wan, B.H. Northrop, H. Jude,

    P.J. Stang, J. Am. Chem. Soc. 130 (2008) 8878.

    [84] G. Otero, G. Biddau, T. Ozaki, B. Gomez-Lor, J. Mendez, R. Perez,

    J.A. Martin-Gago, Chem. Eur. J. 16 (2010) 13920.

    [85] A. Singh, A.S. Myerson, J. Pharm. Sci. 99 (2010) 3931.

    [86] T. Nakanishi, N. Yamakawa, T. Asahi, T. Osaka, B. Ohtani, K.

    Uosaki, J. Am. Chem. Soc. 124 (2002) 740.

    Trends Trends in Analytical Chemistry, Vol. 39, 2012

    206 http://www.elsevier.com/locate/trac