the use of luminescent quantum

Upload: semanu-tcheyi

Post on 03-Apr-2018

226 views

Category:

Documents


0 download

TRANSCRIPT

  • 7/29/2019 The Use of Luminescent Quantum

    1/12

    The use of luminescent quantum

    dots for optical sensingJose M. Costa-Fernandez, Rosario Pereiro, Alfredo Sanz-MedelSemiconductor nanocrystals, known as quantum dots (QDs), have demon-

    strated several remarkable, attractive optoelectronic characteristics espe-

    cially suited to analytical applications in the (bio)chemical field. We review

    progress in exploiting the attractive luminescent properties of QDs in

    designing novel probes for chemical and biochemical optical sensing.

    2005 Elsevier Ltd. All rights reserved.

    Keywords: (Bio)chemical sensor; Nanostructure; Photoluminescence; Quantum dot

    1. Introduction

    Quantum dots (QDs) are nanostructured

    materials [1], also known as zero-dimen-

    sional materials, semiconductor nano-

    crystals or nanocrystallites. These colloidal

    nanocrystalline semiconductors, compris-

    ing elements from the periodic groups

    II-VI, III-V or IV-VI, are roughly spherical

    and with sizes typically in the range 112

    nanometer (nm) in diameter. At such

    reduced sizes (close to or smaller than thedimensions of the exciton Bohr radius

    within the corresponding bulk material),

    these nanoparticles behave differently

    from bulk solids due to quantum-

    confinement effects [2,3]. Quantum

    confinements are responsible for the

    remarkable attractive optoelectronic

    properties exhibited by QDs, including

    their high emission quantum yields, size-

    tunable emission profiles and narrow

    spectral bands [3,4]. Moreover, their

    strong size-dependent properties result in a

    tunability emission that leads to new

    applications in science and technology.

    The past 20 years have seen intense

    research activity in the fundamental study

    of the synthesis and the photophysical

    properties of QDs [58]. Different groups

    have studied II-VI semiconductor QDs,

    such as CdSe or CdS nanocrystals, in order

    to characterize the relationship between

    size, shape and electronic properties [2,4].

    However, most applications so far have

    focused on their use in microelectronics

    and opto-electrochemistry (e.g., light-

    emitting diodes, solar energy conversion

    or quantum cascade lasers) [3,9,10].

    The application of luminescent QDs as

    biological labels was first reported in 1998

    in two breakthrough papers [11,12]. Both

    groups simultaneously demonstrated that

    highly luminescent QDs can be madewater-soluble and biocompatible by

    surface modification and bioconjugation.

    They also showed the high potential of

    QDs as highly sensitive fluorescent bio-

    markers and (bio)chemical probes.

    Other key advances enabling the

    emerging practical applications of QDs in

    biochemistry and medicine included the

    synthesis of high-quality colloidal QDs in

    large quantities [13] or recent advances

    on surface chemistry of QDs by conjuga-

    tion with appropriate functional molecules

    [14]. The surface modification of QDs can

    increase their luminescent quantum yields

    [14], improve stability of the nanocrystals

    and prevent them from aggregating [15],

    and make QDs available for interactions

    with target analytes [16], all of crucial

    interest for chemical sensor or biosensor

    applications.

    This article deals with work on the

    analytical applications of QDs in develop-

    ing novel (bio)chemical sensors, an area of

    growing interest in the past few years. We

    include a brief discussion on the attractiveoptical properties of QDs and on the

    importance of adequate control of the

    synthesis and surface modification of

    the luminescent QDs, in order to achieve

    the desired selectivity and sensitivity for

    sensing target analytes.

    2. Optical properties

    Studies of the physical properties of QDs

    have revealed that strong confinement of

    Jose M. Costa-Fernandez,

    Rosario Pereiro,

    Alfredo Sanz-Medel*

    Department of Physical and

    Analytical Chemistry,

    University of Oviedo, c/ Julian

    Clavera, 8, E-33006 Oviedo,

    Spain

    *Corresponding author.

    Tel.: +34 985 10 34 74;

    Fax: +34 985 10 31 25;

    E-mail: [email protected]

    Trends in Analytical Chemistry, Vol. 25, No. 3, 2006 Trends

    0165-9936/$ - see front matter 2005 Elsevier Ltd. All rights reserved. doi:10.1016/j.trac.2005.07.008 2070165-9936/$ - see front matter 2005 Elsevier Ltd. All rights reserved. doi:10.1016/j.trac.2005.07.008 207

    mailto:[email protected]:[email protected]
  • 7/29/2019 The Use of Luminescent Quantum

    2/12

    excited electrons and holes in these nanocrystals exists

    at such reduced sizes and led to observations of unique

    optical and electronic properties [2,3]. These com-

    pounds, which are usually non-fluorescing, develop an

    intense, long-lasting luminescent emission when syn-

    thesized on an nm scale. Semiconductor QDs are char-

    acterized by a band-gap between their valence andconduction electron bands. When a photon having an

    excitation energy exceeding the semiconductor band-gap

    is absorbed by a QD, electrons are promoted from the

    valence band to the high-energy conduction band. The

    excited electron may then relax to its ground state by

    the emission of another photon with energy equal to the

    band-gap [4].

    The results of quantum confinement are that the

    electron and hole energy states within the nanocrystals

    are discrete, but the electron and hole energy levels (and

    therefore the band-gap) is a function of the QD diameter

    as well as composition [17]. The band-gap of semicon-

    ductor nanocrystals increases as their size decreases,resulting in shorter emission wavelengths [18,19]. This

    effect is analogous to the quantum mechanical particle

    in a box, in which the energy of the particle increases

    as the size of the box decreases.

    The size-dependent emission is probably the most

    striking and the most studied optical property of QDs. As

    the emission properties of semiconductor nanocrystals

    depend strongly upon the energy and the density of the

    electron states, they can be altered by engineering the

    size and the shape of these tiny structures. For example,

    Fig. 1 shows the fluorescence spectra of CdSe QDs with

    different nanoparticle diameter sizes. As can be seen,

    differently-sized CdSe nanocrystals can be tuned in the

    500700-nm range. Moreover, as each material has

    tunability limits, which depend on the physical limita-

    tions of the dot size, other materials have been employed

    in QD synthesis (see Table 1) (e.g., Zn-based QDs emitbelow 400 nm while Pb-based QDs have an emission in

    the near-infrared spectral region).

    As a result of their discrete, atom-like electronic

    structure, QDs have typically very narrow emission

    spectra with full width at half-maximum (FWHM) of the

    luminescent emission of around 1540 nm. (QDs with

    bandwidths as narrow as 12.716.9 nm FWHM have

    been reported [20]). Since the emission lines are com-

    paratively much narrower that those of organic dyes,

    detection of the QDs suffers much less from cross-talk

    that might result from the emission of a different fluo-

    rophore bleeding into the detection channel of the fluo-

    rophore of interest (analyte).On the other hand QDs typically exhibit higher fluo-

    rescence quantum yields than conventional organic

    fluorophores, allowing for greater analytical sensitivity.

    The quantum yield of a luminophor is a function of the

    relative influences of radiative recombination (producing

    light) and non-radiative recombination mechanisms.

    Non-radiative recombination, which largely occurs at

    the nanocrystal surface, is a faster mechanism than

    radiative recombination and is greatly influenced by the

    surface chemistry. In this context, it has for example

    5

    0

    0.2

    0.4

    0.6

    0.8

    1

    500 550 600 650 700

    Normalizedfluoresce

    nce

    intensity

    Fluorescence

    h(=400 nm)

    3 nm~ 4 nm~ 7 nm~

    155= 155= mnmn 095= 095= mnmn 746= 746= mnmn

    Wavelength, nm

    Figure 1. Size-tuneable fluorescence spectra of CdSe QDs. The diameter sizes of the nanoparticles are shown over the fluorescence spectrum.

    Trends Trends in Analytical Chemistry, Vol. 25, No. 3, 2006

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

  • 7/29/2019 The Use of Luminescent Quantum

    3/12

    been demonstrated that capping the nanocrystal with a

    shell of an inorganic wide-band semiconductor (e.g.,

    ZnS) reduces such non-radiative deactivations and re-

    sults in brighter emission [21]. Chan and Nie estimated

    that single ZnS-capped CdSe QDs are about 20 times

    brighter that single rhodamine 6G molecules [12].There is also evidence that QDs, suitably surface-

    derivatized for protection, have also enhanced photolu-

    minescent stability as compared to typical fluorescent

    organic dyes. Several studies have demonstrated that the

    photoluminescence properties of CdSe nanocrystals

    (including the quantum yields, peak position and

    FWHM) did not show any detectable change upon aging

    in air for several months [22]. Moreover, QDs were ob-

    served to be 100 times more stable that conventional

    organic fluorophors against photobleaching [12].

    3. Synthesis and surface chemistry

    Progress on the synthesis of high-quality semiconductor

    nanocrystals has played and is still playing a critical role

    in the progress of QDs applications. Lithography-based

    technologies have been widely used for QDs grown onto

    adequate substrates [23], but have mainly been re-

    stricted to the preparation of optoelectronic devices.

    However, colloidal nanocrystals with single crystalline

    structure and well-controlled size and size distribution

    can be prepared by relatively simple nanocrystal-growth

    processes, starting from organometallic precursors in a

    mixed solvent [2,3], the latter approach being more

    familiar to chemists.

    Due to the availability of precursors and the simplicity

    of crystallization, CdS and CdSe have been the most well-

    studied colloidal QDs. Murray et al. [8] reported the

    synthesis of high-quality Cd-chalcogenide nanocrystals

    using dimethylcadmium as QD precursor in the presence

    of a coordinating solvent at high temperatures. The most

    common coordinating solvents used are trioctylphos-

    phine oxide, trioctylphosphine and hexadecylamine

    (frequently used together). Such solvents, which cap

    the nanocrystal and stabilize its surface, determine the

    particle solubility in organic media and prevent irre-

    versible aggregation of the nanocrystals. However, QDs

    capped with these hydrophobic coatings are incompati-

    ble with aqueous assay conditions. Consequently, in

    order to extend the field of application of the QDs,

    hydrophilic capping agents must be introduced.A landmark in the development of wet chemical

    routes for Cd-chalcogenide nanocrystals was the use of

    thiols as stabilizing agents in aqueous solution [24].

    Water-soluble nanoparticles were prepared by synthe-

    sizing thiol-capped crystalline nanoparticles in aqueous

    solution by using mercapto-alcohols (e.g., 2-mercap-

    toethanol or 1-thioglicerol) and mercapto-acids (e.g.,

    thioglycolic acid or thiolactic acid) as stabilizers [24].

    In an important paper [13], Pengs group reported the

    synthesis of high-quality CdTe, CdSe and CdS nano-

    crystals using CdO as precursor instead of Cd(CH3)2. This

    latter compound is toxic, unstable, explosive andexpensive, rendering QD-synthesis schemes based on its

    use unsuitable for large-scale synthesis (due to the need

    of critical experimental conditions). The quality of the

    QDs synthesized with this new approach [13] was found

    to be comparable or superior to the best previously re-

    ported. Moreover, the reported synthesis scheme (see

    Fig. 2) proved to be reproducible, were based on mild and

    simple conditions, and had great potential to be scaled

    up for industrial applications.

    In recent years, other alternative routes for synthesis of

    highly mono-dispersed QDs have been investigated. For

    example, the use of stable non-air-sensitive precursors

    based on selenocarbamate derivatives of Zn or Cd [25] or

    on the air-stable complex Cd imino-bis(diisopropylphos-

    phine selenide) [26] have been proposed to synthesize

    monodispersed luminescent QDs of comparable quality to

    those prepared by more conventional methods.

    However, to ensure efficient emission, any traps for

    the photogenerated electron and hole should be avoided.

    Possible traps in QDs are generally surface atoms that

    are missing at least one chemical bond. The surface

    atoms must be optimally constructed or reconstructed

    and passivated with some ligands to get rid of traps.

    Coating nanoparticles with a different semiconducting

    Table 1. Nanocrystal materials and range of tunability

    QD core materiala Fluorescence-emission range (nm) QD-diameter range (nm)

    Zinc sulfide (ZnS) 300410 Zinc selenide (ZnSe) 370430

    Cadmium sulfide (CdS) 355490 1.96.7

    Cadmium telluride (CdTe) 620710

    Lead sulfide (PbS) 700950 2.39Lead selenide (PbSe) 12002340Lead telluride (PbTe) 18002500

    aEmission fluorescence of core-shell QDs is also within the range given in the second column.

    Trends in Analytical Chemistry, Vol. 25, No. 3, 2006 Trends

    http://www.elsevier.com/locate/trac 209

  • 7/29/2019 The Use of Luminescent Quantum

    4/12

    material was shown to have a profound impact on the

    photophysics of the nanocrystalline core [2,14,21].

    Deposition of a semiconductor layer with a large band-

    gap (Eg) relative to the core typically results in the

    enhancement of the QD emission due to the suppressionof radiationless recombination mediated by surface states

    [2,21], while the degree of charge-carrier confinement

    does not change. Conversely, an outer layer from a

    semiconductor with a small Eg provides an additional

    area of delocalization for electron and hole [2,14,21]. Of

    course, relaxation of the confinement regime results in a

    red shift of the spectral features.

    The exciting size-dependent and surface-dependent

    properties of nm-sized QDs have stimulated research on

    surface modification of QDs, aiming to expand their

    practical applications. In this context, the conjugation of

    a semiconductor nanoparticle with an organic molecule

    [27], able to interact selectively with a target molecule or

    (bio)chemical species, extends the area of applications

    from the electronic or optical devices to the biological or

    chemical systems, such as the preparation of non-

    radioactive biological labels [11,12] or chemical opto-

    sensors [16].

    4. Optical sensing with quantum dots

    More than five years have elapsed since QDs were first

    proposed as stable luminescent probes in biological

    labeling applications. In that pioneer work, Alivisatoss

    group [11] reported a link between biomolecules and

    CdS or ZnS core-shell CdSe QDs via surface coating with

    an additional layer of silica in order to make them bio-

    compatible and water soluble, and established the utilityof the nanocrystals for biological staining. Simulta-

    neously, Chan and Nie [12] linked biomolecules to

    water-soluble and biocompatible QDs surface-modified

    with mercaptoacetic acid for ultrasensitive detection at

    the single-dot level. They demonstrated that conjuga-

    tion of the QDs with appropriate immunomolecules

    can be used for recognition of specific antibodies or

    antigens by measuring the luminescence emission of the

    nanoparticles.

    Many authors have stressed the distinct advantages of

    QD bioconjugates over conventional organic dyes (such

    as rhodamine), namely greater brightness, greater sta-

    bility with respect to photobleaching and narrower

    spectral line-widths. However, QD biological labeling has

    been slow to emerge into common practice, partly due to

    the difficulty in producing stable QD-biomolecule com-

    plexes. Developments have stressed the importance of

    adequate surface modifications in developing lumines-

    cent QDs for labeling in bioanalysis and diagnostics, as

    tags for protein and DNA immunoassays or as biocom-

    patible labels for in vivo imaging studies. Several reviews

    have summarized the use of luminescent QDs in such

    biochemical applications [2731]. Moreover, the fast

    development and improvements in the synthesis of QDs

    (a) NANOPARTICLE SYNTHESIS

    (b) SURFACE MODIFICATION

    -S-CH2-COO(-)

    QD

    Water-soluble QDs

    TOP/TOPOCdSe QDs

    (in CH4)

    Reflux 12 h.

    Thermom

    eter syrin

    ge

    H2HS-C-COOH

    Purification

    QDs separation(at~ 15,000 rpm)

    andre-dispersion in H2O

    P

    P

    P

    O

    P

    OQD

    P

    O

    P TOP

    TOPO

    TOPO+

    CdO+

    HPA

    TOP-Se

    320 C

    Thermom

    eter syringe

    Argon

    Nucleation GrowthQD

    20 C270 C

    Argon

    ~ 20 min

    Purification

    QDs separation(at~ 15,000 rpm)

    andre-dispersion in CH4

    P

    P

    P

    O

    P

    OQD

    Dilution with~ 10 mL CHCl3

    Figure 2. Schematic illustration of a typical synthesis process and surface-modification of a luminescent QD based on the use of CdO asprecursor. TOP: trioctylphosphine; TOPO: trioctylphosphine oxide; HPA: Hexylphosphonic acid.

    Trends Trends in Analytical Chemistry, Vol. 25, No. 3, 2006

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

  • 7/29/2019 The Use of Luminescent Quantum

    5/12

    have uncovered possibilities that analytical chemists

    have also started to explore in developing these

    nanomaterials for a new generation of optical sensors

    based on luminescence.

    4.1. Fluorescence-based transduction

    As the luminescence of QDs is very sensitive to the sur-face states of the QDs, it is reasonable to expect that the

    chemical or physical interactions between a given

    chemical species and the surface of the nanoparticles

    would result in changes in the efficiency of the core

    electron-hole recombination [32]. This has been the

    basis of the increase in research activity on the devel-

    opment of novel optical sensors based on QD probes.

    Following this approach, Cd-based QDs have been re-

    ported for optical sensing of small molecules and ions

    (Table 2). In a pioneering work, the addition of Cd ions

    to a basic aqueous solution containing unpassivated CdS

    nanoparticles resulted in important enhancement of the

    luminescence quantum yield of the nanoparticles,without detectable changes in particle sizes [7]. This

    effect was attributed to the formation of a Cd(OH)2 shell

    on the CdS core, which effectively eliminates the non-

    radiative recombination of charge carriers.

    A similar photoluminescence-activation effect (attrib-

    uted to passivation of surface trap sites that are either

    being filled or energetically moved closer to the band

    edges by this simple chemical process) was also induced

    after adding Zn and Mn ions to colloidal solutions of CdS

    or ZnS QDs [32,33]. This behavior provided the basis for

    optical sensing of such metallic cations with QDs.

    Besides the activation effect, QD-based optical sensingquenching strategies (based on the quenching by the

    analyte that affects the luminescence emission of the

    nanoparticle) have been proposed. Quenching mecha-

    nisms to explain how metal ions quench fluorescence

    of QDs include inner filter effects, non-radiative recom-

    bination pathways, electron-transfer processes and

    ion-binding interactions. Measurement of the lumines-

    cence-deactivation ratio of peptide-coated CdS QDs has

    been proposed for the optical sensing of Cu(II) and Ag(I)

    [34]. Similarly, the effect of three different ligands

    (L-cysteine, thioglycerol and polyphosphate) was evalu-

    ated on the luminescence deactivation of water-soluble

    CdS QDs with respect to several cations, including Zn

    and Cu ions [35]. This latter work was one of the first

    references to the use of luminescent QDs as selective ion

    probes in aqueous samples.

    Isarov and Chrysochoos [36] observed that the addi-

    tion of Cu(II) perchlorate in 2-propanol to CdS nano-

    particles led to the binding of copper ions onto the QD

    surface, accompanied by rapid reduction of Cu2+ to Cu+.

    It was proposed that copper ions bound onto the surface

    of the QDs facilitate non-radiative electron/hole (e/h+)

    annihilation, thus resulting in a quenching of the

    luminescence from the nanoparticles. It was shown thatTable2.

    QD-basedfluorescentprobes

    forchemicaldeterminationofsmallmoleculesandions

    QD

    material

    QD

    coating

    Analyte

    Matrix

    Detectionlimit

    Measuringsignal

    Ref.

    CdS

    Cly-His-Leu-Leu-Cys

    Cu(II)

    Phosphatebuffer

    0.5lM

    Fluorescencequenching

    [34]

    Ag(II)

    CdS

    Polyphosphate

    Cu(II)

    Water

    0.8mMZn(II)

    Fluorescencequenching

    [35]

    L-cysteine

    Fe(III)

    0.1mMCu(II)

    Thioglycerol

    Zn(II)

    CdSe

    2-mercaptoethanesulfonicacid

    Cu(II)

    Water

    3.2nM

    Fluorescencequenching

    [37]

    CdSe-ZnS

    Bovineserumalbumin

    Cu(II)

    Water

    10n

    M

    Fluorescencequenching

    [38]

    CdSe

    Mercaptoaceticacid+bovineserumalbumin

    Ag(I)

    Water

    70n

    M

    Fluorescencequenching

    [39]

    CdTe

    3-mercaptopropionic

    acid

    Cu(II)

    Water

    0.19

    ng/mL

    Fluorescencequenching

    [40]

    CdTe

    Thioglycolicacid

    Zn(II),

    Mn(II),

    Ni(II),Co(II)

    Water

    Fluorescencequenching-en

    hancement

    [41]

    CdS

    Polyphosphate

    I

    Methanol

    Fluorescencequenching

    [42]

    CdSe

    Tert-butyl-n-(2-mercaptoethyl)-carbamate

    CN

    Methanol

    0.1lM

    Fluorescencequenching

    [44]

    CdSe

    2-mercaptoethanesulfonicacid

    CN

    Water

    1.1lM

    Fluorescencequenching

    [45]

    CdS

    L-cysteine

    Ag

    +

    Water

    5.0nM

    Fluorescenceenhancement

    [46]

    CdSe

    Incorporatedinpolym

    erfilms

    Triethylamine

    Gasmedia

    Fluorescencequenching-en

    hancement

    [47]

    Benzylamine

    CdSe-ZnS

    Thioglycolic+organo

    phosphoroushydrolase

    Paraoxon

    Water

    10n

    M

    Fluorescencequenching

    [49]

    Trends in Analytical Chemistry, Vol. 25, No. 3, 2006 Trends

    http://www.elsevier.com/locate/trac 211

  • 7/29/2019 The Use of Luminescent Quantum

    6/12

    the quenching could be employed for chemical sensing of

    Cu ions in the organic solution.

    Water-soluble CdSe QDs with their surface modified

    with 2-mercaptoethane sulfonic acid can be used for the

    sensitive and selective determination of copper (II) ions

    in aqueous solutions, based on fluorescence-quenching

    measurements [37].In addition, based on the photoluminescence

    quenching of the nanocrystals, CdSe-ZnS QDs modified

    with bovine serum albumin (BSA) were investigated for

    the determination of copper [38], CdSe QDs modified

    with mercaptoacetic acid and BSA were assayed for the

    analysis of silver [39] and CdTe nanocrystals modified

    with mercaptopropionic acid were proposed for the

    determination of Cu(II) ions [40]. It was observed that

    the change in the absorption spectra caused by Cu(II)

    can be reversed by the addition of EDTA, a good com-

    plexing agent for Cu(II) ions. Thus, the authors proposed

    that the interaction between Cu(II) ions and the QD

    surface should be of the ion-binding type.Li et al. [41] synthesized water-soluble luminescent

    thiol-capped CdTe QDs and investigated the effect of

    divalent metal ions on their photoluminescence re-

    sponses. They found that zinc ions enhanced the lumi-

    nescence emission of the QDs. However, other metals

    (e.g., calcium, magnesium, manganese, nickel and cad-

    mium) quenched luminescence.

    Apart from research on QD-based fluorescent sensors

    for ion metals, work on other chemical species (e.g., io-

    dide [42] or cyanide [43]) has reported quenching the

    emission of CdS or CdSe QDs. A polyphosphate-stabilized

    CdS QD was evaluated for optical sensing of iodide [42]and found strong decay of luminescence intensity (decay

    times 10 ls), brought about by the analyte. Such

    quenching effects [42] were attributed to inner filter

    effects, non-radiative recombination pathways and

    electron-transfer processes.

    The strong, reversible adsorption of negatively-

    charged CN onto the QD surface, with the consequent

    increased location due to compression of the electron-

    wave function in the QDs, was used to explain the

    quenching effect of cyanide [43].

    Following this mechanism, the synthesis of red

    photoluminescent CdSe QDs, with their surface modified

    with tert-butyl-n-(2-mercaptoethyl)-carbamate, has

    been proposed for the selective, sensitive determinationof free cyanide in methanol after a photoactivation of the

    QDs [44].

    In a further work [45], the authors reported the syn-

    thesis of water-soluble luminescent CdSe QDs, surface-

    modified with 2-mercaptoethane sulfonate, for the

    selective determination of free cyanide in aqueous solu-

    tion (see Fig. 3). The addition of surfactant agents to the

    measurement aqueous solution was found to further

    greatly stabilize the QDs. In this way, the fluorescent

    signals observed allowed for high sensitivity (detection

    limit 1.1 106 M) and also for great selectivity of the

    proposed cyanide detection (over many other anionic

    species).As can be seen, most of the methods described so far

    rely on the chemical sensing of small molecules and ions

    with QDs via analyte-induced deactivation of photolu-

    minescence. However, Zhu and Chen [46] proposed a

    method for the determination of trace levels of silver ion

    based on luminescence enhancement of water-soluble

    CdS QDs modified with L-cysteine. The authors showed

    detection limits as low as 5.0 109 M. They proposed

    that the fluorescence-enhancement effect could be

    attributed to the formation of a complex between silver

    ions and the RS- groups adsorbed on the surface of the

    modified QDs, which resulted in the creation of radiativecenters at the CdS/Ag-SR complex.

    The interactions between some reactive gas molecules

    and the surface of CdSe QDs have also been exploited in

    developing gas-sensing technologies [47]. Nazzal et al.

    found that the photoluminescence of CdSe QDs incor-

    porated into polymer thin films is reversibly enhanced

    or quenched by the presence of certain gases in the

    0

    100

    200

    300

    450 550 650

    Wavelength, nm

    IF

    0

    100

    200

    300

    450 550 650

    Wavelength, nm

    IF

    + CN-

    CdSeCdSeCdSeCdSeCdSe CdSeCdSeCdSeCdSeCdSeC N-

    C N-

    C N-

    Figure 3. Effect of the addition of 0.65 mg/l of cyanide to the luminescence emission spectra of CdSe QDs surface-modified with MES.

    Trends Trends in Analytical Chemistry, Vol. 25, No. 3, 2006

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

  • 7/29/2019 The Use of Luminescent Quantum

    7/12

    environment. After QD synthesis, photostimulation was

    found to be necessary to obtain a stabilized emission

    profile and to provide reliable responses to the presence

    of the gases. This effect, shown in Fig. 4, was also re-

    ported by other groups [37,44,45]. Although the

    mechanism(s) explaining this photoactivation is (are)

    not clear, it is thought that a reconstruction of the sur-face atoms of the nanoparticle, or an optimization of

    surface-ligand passivation, could lead to the observed

    enhancement of the measured luminescence [47].

    QDs were also proposed for the design of sensing

    assemblies for selective detection of paraoxon [48].

    Water-soluble CdSe QDs, surface functionalized with

    thioglycolic acid, were synthesized and incorporated to-

    gether with organophosphorous hydrolase (OPH) in a

    thin film prepared by the layer-by-layer technique. The

    presence of paraoxon in the sample solution was de-

    tected by changes in the photoluminescence emission of

    the QDs, attributed to an interaction of the analyte with

    the OPH included in the sensing film, changing itsconformation.

    Following this approach, the synthesis of (CdSe)ZnS

    nanocrystals and their conjugation with organophos-

    phorous hydrolase (through electrostatic interaction

    between negatively-charged QD surfaces and the posi-

    tively-charged protein side chain and -NH2 ending

    groups) has been proposed in developing a biosensor to

    detect paraoxon, obtaining detection limits as low as

    108 M [49]. The photoluminescence intensity of the

    OPH/QD bioconjugate was quenched in the presence of

    paraoxon, matching very well with the Michaelis-

    Menten equation. This result indicated that the quenching

    was caused by the conformational change in the en-

    zyme, which was confirmed by gas-chromatography

    measurements. Although such a strategy of QD-surfacebioconjugation has not yet been exploited frequently for

    sensing, it holds great potential for further developments

    in optical sensing with QDs.

    In recent years, QDs have been used as inorganic,

    non-specific, DNA-binding proteins that act as lumines-

    cent labels for different applications (e.g., multicolor gene

    mapping on the nm scale) [50]. Moreover, fluorescence

    quenching of water-soluble CdSe QDs, surface modified

    with mercaptoacetic acid, has also been used to develop

    a fluorescence probe for rapid, sensitive determination of

    DNA in a neutral medium [51]. The mechanism for the

    binding of the nucleic acids to the QDs was investigated

    and it was concluded that nanoparticles bind to the helixstructure of the DNA in a non-intercalative way,

    resulting in the observed deactivation of the lumines-

    cence emission of the QDs.

    4.2. Fluorescence (or Forster) resonance

    energy-transfer-based sensors

    Energy-transfer mechanisms have been widely used in

    different fields and are the basis of a new generation of

    Figure 4. Effect of photoactivation (a) fluorescence spectra from CdSe quantum dots measured after different sunlight time exposures (from ref-erence [44]), (b) Pictures of a methanolic QDs solution (1) freshly prepared and (2) after 3 days exposed to the sunlight.

    Trends in Analytical Chemistry, Vol. 25, No. 3, 2006 Trends

    http://www.elsevier.com/locate/trac 213

  • 7/29/2019 The Use of Luminescent Quantum

    8/12

    luminescent sensors [52]. In this context, the capability

    of tailoring (via size) QD-photoemission properties should

    allow efficient energy transfer with a number of con-

    ventional organic dyes, thus suggesting the use of the

    nanoparticles in sensor or chemical assay applications

    based on photochemically-induced fluorescence (or

    Forster) resonance-energy-transfer (FRET) mechanisms.However, the QD emission spectrum is narrower and

    more symmetric than the emission from conventional

    organic fluorophores, making it much easier to distin-

    guish the emission of the donor from that of the accep-

    tor. Moreover, the high quantum yields of QDs make

    energy transfer very efficient.

    Several studies have already confirmed that QDs are

    excellent candidates for use in the design of novel FRET-

    based strategies. As an example, specific binding of dif-

    ferent proteins was observed via measurements of FRET

    between a CdSe-ZnS QD donor, attached to one of the

    proteins, and some organic acceptor dyes attached to the

    other protein under study. In the presence of specificinteractions between both proteins, strong enhancement

    of the acceptor-dye fluorescence was observed [53].

    In a more fundamental study, conjugation of BSA

    with luminescent CdTe nanoparticles (capped with L-

    cysteine) resulted in a significant increase in the CdTe

    fluorescent emission, attributed to an efficient reso-

    nance-energy transfer from the tryptophan moieties of

    the protein units to the CdTe nanoparticles acting as

    acceptors [54]. However, despite the demonstrated

    favourable properties of luminescent QDs for FRET

    experiments, only very few studies on the synthesis of

    QD bioconjugates and their applications for QD-FRET-based optical sensors have been published so far.

    Luminescent CdSe QDs have been used as energy do-

    nors in developing a competitive FRET assay for maltose

    [55] (see Fig. 5). Semiconductor nanoparticles biocon-

    jugated to different maltose-binding proteins, formed

    using a non-covalent self-assembly scheme, act as the

    resonance-energy-transfer donors, while non-fluorescent

    dyes bound to cyclodextrin serve as the energy-transfer

    acceptors. In the absence of maltose, cyclodextrin-dye

    complexes occupy the protein binding sites. Energy

    transfer from the QDs to the dyes quenches the QD

    fluorescence. When maltose is present, it replaces the

    cyclodextrin complexes, and the QD fluorescence recov-

    ers [55]. This approach has been successfully employed

    in developing a prototype QD-based sensor for sensing

    maltose in solution [56].

    CdSe-ZnS core-shell biocompatible QDs, stabilized by

    mercaptopropionic acid modified with a thiolated oligo-

    nucleotide, have been proposed as energy donors for

    lighting up the dynamics of telomerization or of DNA

    replication occurring on the nanoparticles, using FRET

    to dye units incorporated into the new synthesized

    telomere or DNA replica [57]. After addition of telome-

    rase, during the progression of the telomerization, the

    fluorescence emission from the QDs at 560 nm decreaseswith the concomitant increase of the 610 nm emission of

    the dye (using 400-nm excitation). Emission observed

    upon telomerization is attributed by the authors to FRET

    from the QDs to the dye molecules incorporated into the

    telomeric units by telomerase. The CdSe-ZnS QDs func-

    tionalized with M 13~ DNA also enabled the detection of

    a viral DNA by following the DNA-replication process by

    FRET. Results can be applied to the fast, sensitive

    detection of cancer cells [57]. It could be also applied to

    the development of chip-based DNA sensors as it func-

    tions like logic gates, where FRET readout occurs when

    hybridization and replication proceed.

    In one application, luminescent ZnS-capped CdSe QDs,

    covered with mercaptoacetic acid, have been conjugated

    to amine-terminated molecular beacons (MBs) at the 5 0

    end for probing DNA sequences [58]. Connected to the 3 0

    end of the molecular beacon, there is a quencher mole-

    cule [4-(40-dimethylaminophenylazo) benzoic acid,

    DABCYL]. In the absence of the target DNA sequence,

    MBs form a hairpin structure in which the QDs and

    DABCYL are in such close proximity that energy from

    the QDs is transferred to the quencher and no fluorescent

    signal is observed. After adding the target DNA se-

    quence, the MB structure opens. Since QD and quencher

    DQ PBM

    excitation(350nm)

    FRET Quenching

    Dye

    DC

    DQ PBM

    excitation(350nm)

    maltose+

    ecnecseroulF

    Dye

    DC

    a

    b

    Figure 5. Schematic diagram of the quantum-dot based FRET malt-ose sensor (adapted fromreference [56]). QDs conjugated to around10 maltose-binding proteins function as the FRET donors. Non-fluorescent dyes bound to a cyclodextrin serve as the acceptorsand in the absence of maltose are filling the protein binding sitesresulting in a quenching of the luminescence. Whenmaltose is pres-ent, it removes the cyclodextrin-dye complex and the fluorescenceis recovered. MBP: maltose binding protein, CD: cyclodextrin.

    Trends Trends in Analytical Chemistry, Vol. 25, No. 3, 2006

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

  • 7/29/2019 The Use of Luminescent Quantum

    9/12

    are then separated from each other, no FRET occurs and

    QD emission can be detected. The authors demonstrated

    that using QDs in this probe resulted in an improved

    lifetime during imaging, as compared to using organic

    fluorophores.

    The application of water-soluble ZnS nanoparticles,

    surface-modified with sodium thioglycolate, as fluores-cence probes has been described for specific determination

    of protein content in a serum sample (e.g., human serum

    albumin, BSA and gamma globulin) with detection limits

    of the order of 10 pg/mL [59]. Energy transfer from sur-

    face-adsorbed proteins to the nanoparticles has been

    proposed as the mechanism responsible of enhancing the

    QD luminescence used for sensing. The methodology was

    applied to the analysis of human serum samples, and re-

    sults obtained were in good agreement with those given by

    alternative, conventional techniques.

    The use of luminescent QDs conjugated to appropriate

    antibody fragments has been employed to develop solu-

    tion-phase, nanoscale, sensing assemblies for detectingthe explosive 2,4,6-trinitrotoluene (TNT) in aqueous

    environments based on FRET measurements [60]. The

    presence of TNT was detected by displacing the dye-

    labeled analogue bonded to the QD surface, resulting in

    the elimination of FRET and in the concentration-

    dependent recovery of QD photoluminescence.

    It should be mentioned that QD-FRET assays can be

    designed so that FRET is the dominant energy-transfer

    process, but FRET efficiency is still inherently low com-

    pared to that of conventional dyes (due to the compar-

    atively large size of QDs, it is too difficult to secure close

    enough proximity for FRET to occur efficiently). How-ever, several studies have been carried out in order to

    gain a better understanding of the process, showing that

    enhanced efficiency can be obtained by careful design of

    the QD-bioconjugation scheme. Using the FRET scheme

    for maltose determination [55], mentioned previously,

    the authors found that, by attaching several active dye-

    labeled proteins to the QD surface, the overall FRET

    signal was improved substantially over a simple one

    donor-to-one acceptor FRET pair [61]. Efficiency was

    further enhanced by increasing the number of dye

    acceptors in the QD bioconjugate, where QDs functioned

    as efficient energy donors [61]. Furthermore, the large

    size of QD fluorophores, compared to organic dyes,

    allowed design of configurations where, for example,

    multiple acceptors could coordinate around and interact

    with a single QD donor. This strategy, already demon-

    strated for multiple detection in immunoassays [62],

    suggests the possibility of achieving mutianalyte opto-

    sensing using a single QD donor and multiple acceptors

    in a FRET-assay format.

    4.3. Surface-plasmon-resonance applications

    QDs have been also investigated by measuring surface-

    plasmon resonance (SPR). Redox transformations

    occurring on chemically-modified surfaces may signifi-

    cantly alter the refractive index of the interface and thus

    induce changes in the plasmon angle of the SPR spectra

    [63]. This approach was followed in the design of an SPR

    sensor for acetylcholine-esterase inhibitors based on the

    photoelectrochemical-charging effect of Au nanoparti-

    cles in an Au-nanoparticle/CdS-QD array (coating anAu/glass surface), which was followed by means of SPR

    changes upon continuous irradiation of the sample [63].

    The fact that other enzymes may be coupled to the Au-

    semiconductor-nanoparticle array and so activate

    photoelectrochemical functions suggests that using SPR

    spectroscopy combined with surface-modified QDs could

    provide an alternative tool for new SPR (bio)sensor

    probes.

    4.4. Phosphorescence transduction

    Only fluorescence transduction has so far been employed

    for photoluminescence sensing in combination with QDs.

    However, investigation of the luminescence properties ofQDs is slowly expanding into phosphorescence, a detec-

    tion principle that may provide several advantages for

    the design of reliable optical sensors [64,65].

    The dopage of sol-gel porous matrices with Tb2S3 QDs

    has been found to produce photoluminescent materials

    with an emission comprising two well-defined bands,

    one at 440 nm (that corresponds to the undoped sol-gel)

    and the other at 600 nm that the authors attributed to

    the Tb2S3 nanoparticles in the silica xerogel [66]. This

    last emission presents characteristics typical of a room-

    temperature phosphorescence (RTP) emission, although

    the origin of the luminescence and the emission mech-anism is not yet understood [66].

    Moore et al. from Mercer University have also reported

    phosphorescence emission from aqueous mixed sulfide

    QD matrices, QD-CdxZn1-xS, doped with manganese(II)

    [67]. The authors evaluated the impact of matrix com-

    position on the QD phosphorescence ($590 nm). They

    found that the observed RTP intensity for the CdxZn1-

    xS:Mn QDs was very sensitive to matrix composition

    (e.g., the 590 nm emission increased with the Zn con-

    centration of the matrix).

    Although very preliminary, those studies seem to open

    the door to novel transduction schemes and applications

    of QDs for optical sensing.

    4.5. Immobilization techniques

    Most of the work on QD applications so far has been

    restricted to solution-sensing assays. A step further to-

    wards developing useful optosensing approaches [68]

    consists of immobilizing those QDs in appropriate solid

    supports to fabricate active solid phases for working in

    flowing solutions [65].

    In this context, sol-gel materials have been demon-

    strated to be especially suited to the development of

    luminescent optical sensors by trapping the indicator

    Trends in Analytical Chemistry, Vol. 25, No. 3, 2006 Trends

    http://www.elsevier.com/locate/trac 215

  • 7/29/2019 The Use of Luminescent Quantum

    10/12

    molecules inside the inorganic structure during the

    polymerization process [69]. Several approaches have

    been also proposed for synthesis of sol-gel materials

    doped with QDs [7072]. Most of the synthetic routes

    involve preparing and surface modifying the QDs in

    solution followed by sol-gel processing in order to obtain

    an inorganic material doped with the luminescentnanocrystals. Considering that QDs are very sensitive to

    changes in the environment, the transfer of these

    materials into glasses, through sol-gel processes, is not a

    simple task. Changes in solvent polarity or QD surface

    reactions during sol-gel polymerization would result in

    an undesirable quenching of the QD photoluminescence.

    In order to overcome such limitations, several ap-

    proaches have been investigated, including the use of

    alkyl amines as a bifunctional aid in QD-glass synthesis

    (amines act as gelation catalysts and stabilizers) [70].

    Such QD glasses have demonstrated high stabilities

    and resistance to degradation, so they have been mainly

    used for optoelectronic applications (e.g., solar concen-trators or as active media in tunable lasers [73]). How-

    ever, these sol-gel materials, doped with luminescent

    QDs, are also expected to be also for optosensing appli-

    cations in the near future.

    A related approach is to incorporate QDs into

    molecularly imprinted polymers (MIPs) [74], which act

    as artificial receptors/antibodies exhibiting tailor-made

    selectivity for a given template molecule. Following

    this approach, Lin et al. [75] synthesized different

    MIPs with several templates incorporating CdSe/ZnS

    core-shell QDs, derivatized with 4-vinylpiridine. Adding

    the functionalized QDs to the monomers, cross-linkersand template molecules in the precursor mixture

    incorporated the nanocrystals into the MIP during

    polymerization. Optosensing of the analytes is achieved

    by measuring the quenching of the photoluminescent

    emission from the QDs included in the polymeric

    structure. Such quenching is attributed to fluores-

    cence-energy-transfer processes between the QDs and

    the template molecules. The approach has been suc-

    cessfully tested for caffeine detection, although addi-

    tional work needs to be carried out to characterize this

    optosensor analytically. It is clear that this approach

    also opens up a new avenue for the development of

    new QD-based optical sensors.

    The advantages and the disadvantages of using the

    different optical transduction strategies already at-

    tempted in developing chemical sensors based on QDs

    can be summarized as follows:

    i Methods based on chemical or physical interactions

    between target chemical species and the surface of

    the nanoparticles are very simple, easy to develop

    and have demonstrated very high sensitivity and

    selectivity features. However, those methods appear

    to be restricted to sensing just a few reactive small

    molecules or ions.

    ii Quantum dots have been demonstrated to be espe-

    cially suited to the development of new chemical

    sensors based on energy-transfer phenomena. This

    approach will probably be widely used as a general

    strategy to develop new QD-based sensor systems

    for analytes unsuitable for direct analysis via interac-

    tion with QD particles. Of course, those methods arenot as simple as those in i) because many different

    parameters need to be carefully controlled (e.g.,

    distance of the acceptor/donor dyes to the QD sur-

    face, and orientation and number of groups) in order

    to achieve an analytically useful energy-transfer

    process.

    iii RTP methods offer general, exceptional characteristics

    in terms of sensitivity and selectivity and some other

    advantages over fluorescence methods. However,

    work on the development of chemical sensors based

    on QDs using phosphorescence transduction is still

    at its very early preliminary development stages.

    Thus, the practical usefulness of RTP methods hasnot yet been demonstrated.

    5. Conclusions and future prospects

    The popularity of QDs as photoluminescent probes

    for optical sensing is steadily increasing, as research-

    ers move to exploit the unique properties of this new

    class of luminophores. Optosensing technologies will

    probably combine the important advantages of QDs

    with flow-analysis techniques and perhaps fibre-opticinstrumentation.

    Chemical-sensing developments will benefit from the

    continuous advances taking place in the science of

    QDs. Thus, future improvements in the nature, range

    and quality of prepared nanomaterials can be ex-

    pected. Chemical-surface modifications of the QDs

    have still to be perfected in order to enhance the

    selectivity of the systems and to profit from their

    favorable emission features. In this context, the con-

    jugation of selective reagents to the surface of lumi-

    nescent QDs (a strategy well established for imaging,

    immunoassay and labeling applications in biological

    science) appears to be a most promising strategy in

    further developing bioactive fluorescent probes for

    sensing applications.

    Moreover, approaches such as the combination of the

    nanoparticles with energy-transfer processes, phospho-

    rescence detection or inclusion on MIPs are promising

    possibilities that are now being investigated.

    Last, but not least, QDs should now be integrated into

    appropriate solid supports, a process that has only just

    begun, in order to develop reliable active phases and

    optosensors able to provide useful flow-through optical

    sensing or fiber-optic-based sensing applications.

    Trends Trends in Analytical Chemistry, Vol. 25, No. 3, 2006

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

  • 7/29/2019 The Use of Luminescent Quantum

    11/12

    In brief, the future of QDs for optical sensing looks

    bright, as their analytical potential in the field now starts

    to be realized.

    Acknowledgement

    Financial support from the EU Project SWIFT-WFD

    (Contract SSPI-CT-2003-502492) and MAT2003-

    09074-C02 (Feder Programme and Ministerio de Ciencia

    y Tecnologa, Spain) is gratefully acknowledged.

    References

    [1] C.M. Niemeyer, Angew. Chem. Int. Ed. Engl. 40 (2001) 4128.

    [2] A.P. Alivisatos, Science (Washington, DC) 271 (1996) 933.

    [3] C.J. Murphy, J.L. Coffer, Appl. Spectrosc. 56 (2002) 16A.

    [4] H. Weller, Angew. Chem. Int. Ed. Engl. 32 (1993) 41.

    [5] L.E. Brus, J. Chem. Phys. 79 (1983) 5566.

    [6] L.E. Brus, J. Chem. Phys. 80 (1984) 4403.

    [7] L. Spanhel, M. Haase, H. Weller, A. Henglein, J. Am. Chem. Soc.

    109 (1987) 5649.

    [8] C.B. Murray, D.J. Norris, M.G. Bawendi, J. Am. Chem. Soc. 115

    (1993) 8706.

    [9] S. Chaudhary, M. Ozkan, W.C.W. Chan, Appl. Phys. Lett. 84

    (2004) 2925.

    [10] V. Colin, M.C. Schlamp, A.P. Alivisatos, Nature (London) 370

    (1994) 374.

    [11] M. Bruchez, M. Moronne, P. Gin, S. Weiss, A.P. Alivisatos, Science

    (Washington, DC) 281 (1998) 2013.

    [12] W.C.W. Chan, S.M. Nie, Science (Washington, DC) 281 (1998)

    2016.

    [13] Z.A. Peng, X.G. Peng, J. Am. Chem. Soc. 123 (2001) 183.

    [14] X. Peng, M.C. Schlamp, A.V. Kadavanich, A.P. Alivisatos, J. Am.

    Chem. Soc. 119 (1997) 7019.

    [15] A.R. Kortan, R. Hull, R.L. Opila, M.G. Bawendi, M.L. Steigerwald,

    P.J. Carroll, L.E. Brus, J. Am. Chem. Soc. 112 (1990) 1327.

    [16] C.J. Murphy, Anal. Chem. 74 (2002) 520A.

    [17] L.E. Brus, J. Chem. Phys. 90 (1986) 2555.

    [18] C.B. Murray, D.J. Norris, M.G. Bawendi, J. Am. Chem. Soc. 115

    (1993) 8706.

    [19] J.E. Bowen-Katari, V.L. Colvin, A.P. Alivisatos, J. Phys. Chem. 98

    (1994) 411.

    [20] P. Reiss, G. Quemard, S. Carayon, J. Bleuse, F. Chandezon,

    A. Pron, Mater. Chem. Phys. 84 (2004) 10.

    [21] B.O. Dabbousi, J. Rodriguez-Viejo, F.V. Mikulec, J.R. Heine,

    H. Mattoussi, R. Ober, K.F. Jensen, M.G. Bawendi, J. Phys. Chem.

    B 101 (1997) 9463.

    [22] L. Qu, X. Peng, J. Am. Chem. Soc. 124 (2002) 2049.

    [23] M. Henini, S. Sanguinetti, L. Brusaferri, E. Grilli, M. Guzzi,

    M.D. Upward, P. Moriarty, P.H. Beton, Microelectron. J. 28 (1997)

    933.

    [24] A.L. Rogach, A. Kornowski, M. Gao, A. Eychmuller, H. Weller,

    J. Phys. Chem. B 103 (1999) 3065.

    [25] B. Ludolph, M.A. Malik, P. OBrien, N. Revaprasadu, Chem.

    Commun. (1998) 1849.

    [26] D.J. Crouch, P. OBrien, M.A. Malik, P.J. Skabara, S.P. Wright,

    Chem. Commun. (2003) 1454.

    [27] A.J. Sutherland, Curr. Opin. Solid State Mater. Sci. 6 (2002) 365.

    [28] W.C.W. Chan, D.J. Maxwell, X. Gao, R.E. Bailey, M. Han, S. Nie,

    Curr. Opin. Biotechnol. 13 (2002) 40.

    [29] A.M. Smith, S. Nie, Analyst (Cambridge, UK) 129 (2004) 672.

    [30] J. Riegler, T. Nann, Anal. Bioanal. Chem. 379 (2004) 913.

    [31] P. Alivisatos, Nat. Biotechnol. 22 (2004) 47.

    [32] D.E. Moore, K. Patel, Langmuir 17 (2001) 2541.

    [33] K. Sooklal, B.S. Cullum, S.M. Angel, C.J. Murphy, J. Phys. Chem.

    100 (1996) 4551.

    [34] K.M. Gatas-Asfura, R.M. Leblanc, Chem. Commun. (2003) 2684.

    [35] Y. Chen, Z. Rosenzweig, Anal. Chem. 74 (2002) 5132.

    [36] A.V. Isarov, J. Chrysochoos, Langmuir 13 (1997) 3142.

    [37] M.T. Fernandez-Arguelles, W.J. Jin, J.M. Costa-Fernandez,

    R. Pereiro, A. Sanz-Medel, Anal. Chim. Acta (2005) 20.

    [38] H.Y. Xie, J.G. Liang, Z.L. Zhang, Y. Liu, Z.K. He, D.W. Pan,

    Spectrochim. Acta Part A 60 (2004) 2527.

    [39] J.G. Liang, X.P. Ai, Z.K. He, D.W. Pang, Analyst (Cambridge, UK)

    129 (2004) 610.

    [40] C. Bo, Z. Ping, Anal. Bioanal. Chem. 381 (2005) 986.

    [41] J. Li, D. Bao, X. Hong, D. Li, J. Li, Y. Bai, T. Li, Colloids Surf A 257

    258 (2005) 267.

    [42] J.R. Lakowicz, I. Gryczynski, Z. Gryczynski, C.J. Murphy, J. Phys.

    Chem. B 103 (1999) 7613.

    [43] S.K. Sarkar, N. Chandrasekharan, S. Gorer, G. Hodes, Appl. Phys.

    Lett. 81 (2002) 5045.

    [44] W.J. Jin, J.M. Costa-Fernandez, R. Pereiro, A. Sanz-Medel, Anal.

    Chim. Acta 522 (2004) 1.

    [45] W.J. Jin, M.T. Fernandez-Arguelles, J.M. Costa-Fernandez,

    R. Pereiro, A. Sanz-Medel, Chem. Commun. (2005) 883.

    [46] J.-L. Chen, C.-Q. Zhu, Anal. Chim. Acta 546 (2005) 147.

    [47] A.Y. Nazzal, L. Qu, X. Peng, M. Xiao, Nano Lett. 3 (2003) 819.

    [48] C.A. Constantine, K.M. Gattas-Asfura, S.V. Mello, G. Crespo,

    V. Rastogi, T.C. Cheng, J.J. DeFrank, R.M. Leblanc, J. Phys. Chem.

    B 107 (2003) 13762.

    [49] X. Ji, J. Zheng, J. Xu, V.K. Rastogi, T.-C. Cheng, J.J. DeFrank, R.M.

    Leblanc, J. Phys. Chem. B 109 (2005) 3793.

    [50] D. Gerion, W.J. Parak, S.C. Williams, D. Zanchet, C.M. Michael,

    A.P. Alivisatos, J. Am. Chem. Soc. 124 (2002) 7070.

    [51] L.-Y. Wang, L. Wang, F. Gao, Z.-Y. Yu, Z.-M. Wu, Analyst

    (Cambridge, UK) 127 (2002) 977.

    [52] J.M. Traviesa, J.M. Costa, R. Pereiro, A. Sanz-Medel, Talanta 62

    (2004) 827.

    [53] D.M. Willard, L.L. Carillo, J. Jung, A. Van Orden, Nano Lett. 1

    (2001) 469.

    [54] N.N. Mamedova, N.A. Kotov, A.L. Rogach, J. Studer, Nano Lett. 1

    (2001) 281.

    [55] H. Mattoussi, J.M. Mauro, E.R. Goldman, G.P. Anderson, V.C.

    Sundar, F.V. Mikulec, M.G. Bawendi, J. Am. Chem. Soc. 122

    (2000) 12142.

    [56] I.L. Medintz, A.R. Clapp, H. Mattoussi, E.R. Goldman, B. Fisher,

    J.M. Mauro, Nat. Mater. 2 (2003) 630.

    [57] F. Patolski, R. Gill, Y. Weizmann, T. Mokari, U. Banin, I. Willner,

    J. Am. Chem. Soc. 125 (2003) 13918.

    [58] J.H. Kim, D. Morikis, M. Ozkan, Sens. Actuators B 102 (2004)

    315.

    [59] L.-Y. Wang, X.-W. Kan, M.-C. Zhang, C.-Q. Zhu, L. Wang, Analyst

    (Cambridge, UK) 127 (2002) 1531.

    [60] E.R. Goldman, I.L. Medintz, J.L. Whitley, A. Hayhurst, A.R. Clapp,

    H.T. Uyeda, J.R. Deschamps, M.E. Lassman, H. Mattoussi, J. Am.Chem. Soc. 127 (2005) 6744.

    [61] A.R. Clapp, I.L. Medintz, J.M. Mauro, B.R. Fisher, M.G. Bawendi,

    H. Mattoussi, J. Am. Chem. Soc. 126 (2004) 301.

    [62] E.R. Goldman, A.R. Clapp, G.P. Anderson, H.T. Uyeda,

    J.M. Mauro, I.L. Medintz, H. Mattoussi, Anal. Chem. 76 (2004) 684.

    [63] M. Zayats, A.B. Kharitonov, S.P. Pogorelova, O. Lioubashevski

    E. Katz, I. Willner, J. Am. Chem. Soc. 125 (2003) 16006.

    [64] J. Kuijt, F. Ariese, U.A.Th. Brinkman, C. Gooijer, Anal. Chim. Acta

    488 (2003) 135.

    [65] A. Sanz-Medel, Anal. Chim. Acta 283 (1993) 367.

    [66] P. Yang, M.K. Lu, C.F. Song, G.J. Zhou, D. Xu, D.R. Yan, Inorg.

    Chem. Commun. 5 (2002) 187.

    [67] http://chemistry.mercer.edu/dem/demres.html .

    [68] J. Ruzicka, E.M. Hansen, Anal. Chim. Acta 173 (1985) 3.

    Trends in Analytical Chemistry, Vol. 25, No. 3, 2006 Trends

    http://www.elsevier.com/locate/trac 217

    http://chemistry.mercer.edu/dem/demres.htmlhttp://chemistry.mercer.edu/dem/demres.html
  • 7/29/2019 The Use of Luminescent Quantum

    12/12

    [69] J.M. Costa Fernandez, A. Sanz Medel, Anal. Chim. Acta 407

    (2000) 61.

    [70] S.T. Selvan, C. Bullen, M. Ashokkumar, P. Mulvaney, Adv. Mater.

    13 (2001) 985.

    [71] C. Bullen, P. Mulvaney, C. Sada, M. Ferrari, A. Chiasera,

    A. Martucci, J. Mater. Chem. 14 (2004) 1112.

    [72] H. Jiang, X.Y.J. Che, M. Wang, F. Kong, Ceramics Int. 30 (2004)

    1685.

    [73] R. Reisfeld, Opt. Mater. (Amsterdam) 16 (2001) 1.

    [74] K. Yano, I. Karube, Trends Anal. Chem. 18 (1999) 199.

    [75] C.I. Lin, A.K. Joseph, C.K. Chang, Y.D. Lee, Biosens. Bioelectron.

    20 (2004) 127.

    Trends Trends in Analytical Chemistry, Vol. 25, No. 3, 2006

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