Ming
Li
a,
Scott K.
Cushing
ab and
Nianqiang
Wu
*a
aDepartment of Mechanical and Aerospace Engineering, West Virginia University, Morgantown, WV 26506-6106, USA. E-mail: nick.wu@mail.wvu.edu
bDepartment of Physics and Astronomy, West Virginia University, Morgantown, WV 26506-6315, USA
First published on 6th October 2014
Surface plasmon resonance (SPR) has found extensive applications in chemi-sensors and biosensors. Plasmons play different roles in different types of optical sensors. SPR transduces a signal in a colorimetric sensor through shifts in the spectral position and intensity in response to external stimuli. SPR can also concentrate the incident electromagnetic field in a nanostructure, modulating fluorescence emission and enabling plasmon-enhanced fluorescence to be used for ultrasensitive detection. Furthermore, plasmons have been extensively used for amplifying a Raman signal in a surface-enhanced Raman scattering sensor. This paper presents a review of recent research progress in plasmon-enhanced optical sensing, giving emphasis on the physical basis of plasmon-enhanced sensors and how these principles guide the design of sensors. In particular, this paper discusses the design strategies for nanomaterials and nanostructures to plasmonically enhance optical sensing signals, also highlighting the applications of plasmon-enhanced optical sensors in healthcare, homeland security, food safety and environmental monitoring.
SPR occurs in two distinct forms: localized SPR (LSPR) and propagating surface plasmon polaritons (SPPs). LSPR occurs when the dimensions of a metallic nanostructure are less than the wavelength of incident light, leading to collective but non-propagating oscillations of surface electrons in the metallic nanostructure. The LSPR strongly depends on the refractive index of the surrounding medium, providing the basis for colorimetric plasmonic sensors. LSPR also concentrates the incident electromagnetic (EM) field around the nanostructure. The local EM field can influence optical processes such as fluorescence, Raman scattering and infrared absorption, resulting in plasmon-enhanced fluorescence (PEF), surface-enhanced Raman scattering (SERS), and surface-enhanced infrared absorption spectroscopy (SEIAS). The LSPR-associated EM field extends into the surrounding medium (generally ∼30 nm) and decays roughly exponentially for a dipole. In contrast to LSPR, SPPs are the propagating charge oscillations on the surface of thin metal films. SPPs cannot be excited by free-space radiation, and instead require momentum matching, such as through periodicity in a nanostructure, for resonance excitation. SPPs are modulated by the refractive index of the surrounding medium, transducing the sensor's signal. SPP can also play a role in modulating radiation in PEF and SERS. The evanescent EM field of SPP decays with a longer length (generally ∼200 nm) than LSPR, allowing the SPP to be modulated by change at a distance farther from the nanostructure surface.
By utilizing LSPR and/or SPP, numerous plasmonic metallic nanostructures have already been developed as signal amplifiers and transducers for sensitive optical sensing. This paper will start with a summary of concepts and principles of plasmonics. Next, three principal types of optical sensors built on plasmonic nanostructures will be discussed, including plasmonic sensors, PEF sensors and SERS sensors. The goal of this review is to give a summary of the underlying physics first, and then apply these principles to guide the design of each type of plasmon-enhanced optical sensor. The design strategies which maximize signal transduction and amplification will be discussed. In addition, this paper will highlight the application of plasmon-enhanced optical sensing in chemical detection and in vivo/in vitro biological sensing.
![]() | (1) |
The plasma frequency depends on the density of electrons (n) and the effective mass (meff), and corresponds to how easily the electrons can move in response to the incident field. Additionally, e is the charge of an electron and ε0 is the permittivity of free space.11 On average, the free conduction electrons in the bulk of the metal do not oscillate against a restoring force, so unlike a mass on a spring, there is not a single resonant frequency. Instead the motion is similar to a mass being dragged in a viscous fluid, and will differ based on whether the electrons can respond quickly enough to the driving force of the incident field. If the light has a frequency above the plasma frequency (in the ultraviolet (UV) range for metals), the electrons will not oscillate and the light will simply be transmitted or absorbed in interband transitions.11 If the light has a frequency smaller than the UV range, the electrons will oscillate 180° out of phase with the incident light, causing a strong reflection.11 The combination of plasma frequency and interband transitions gives metals their characteristic color. Mathematically this behavior is described by the real part of the dielectric constant (ε′metal)12
![]() | (2) |
When the frequency of light is greater than the plasma frequency, the real part of the dielectric constant is positive and light is transmitted.12 When the frequency of light is less than the plasma frequency, the real part of the dielectric constant is negative, and the majority of the light is reflected (Fig. 1d and e).12 The dielectric constant therefore decides whether or not the metal electrons can oscillate at the given frequency of light. In Fig. 1, the convention is used that the positive imaginary part of the dielectric constant corresponds to the Ohmic losses. To avoid confusion, it should be noted that the opposite convention of a negative imaginary part of the dielectric constant referring to loss can also be found frequently in the literature.
If the bulk metal is now shrunk to a thin film, the oscillations will only exist at the surface, leading to propagating charge waves known as surface plasmon polaritons (SPPs) (Fig. 1b).13,14 The word “polariton” refers to the transformation of the bulk, volume oscillations to travelling surface charge waves. The interface between the metal (εmetal) and the surrounding medium (εdiel) places additional constraints on what frequencies the metal electrons can oscillate in the incident field. This limits the continuous spectrum of eqn (2) (all frequencies below ωp) for all incident angles to a fixed wave vector and frequency for a given interface. The resulting quantization is why the name is switched from plasma to plasmon.
The resonance condition to excite the SPP is given by13,14
![]() | (3) |
When the metal electrons oscillate, the real part of the dielectric constant is negative, therefore cancelling out the denominator in eqn (3), leading to a resonance condition at13,14
![]() | (4) |
The dependence of the SPP frequency on the dielectric constant at the interface transforms the bulk plasma oscillations into a useful transducer for a sensor. The local EM field resulting from the charge oscillations of the SPP extends ∼100–200 nm into the dielectric.15 If the local environment changes within this distance, the dielectric constant will differ, and the SPP frequency will shift from that in air. This can be understood conceptually as the dielectric screening the charge at the interface and reducing repulsion between adjacent electrons, effectively reducing the energy needed to drive oscillations and red-shifting the oscillation frequency. As the SPP frequency changes, so do both the dispersion curve and the angle at which the SPP can be excited, modulating the experimentally measured reflectance. The narrow absorption line shape and high angular specificity of the SPP allow excellent signal-to-noise ratio and figure of merit to be obtained for SPP-based sensors.13 This sensitivity comes at the trade-off of experimental simplicity because of the complex geometries needed for detection.13
The restrictions of SPP can be overcome by changing a two-dimensional (2D) metal film to a zero-dimensional (0D) nanoparticle. The incident electric field will be constant across the nanoparticle if it is smaller than the wavelength of light, inducing a uniform displacement of the electron density and a strong restoring force from the positive ionic core background (Fig. 1c).8 The restoring force leads to a characteristic oscillation frequency in the metal electrons similar to a simple harmonic oscillator. This phenomenon is known as LSPR.8 LSPR can be excited directly by the incident field because the geometry of the nanoparticle supplies the additional momentum.8 The local environment-induced change in the LSPR peak position can therefore be detected using a simple UV-visible spectrometer without the need for additional gratings or prisms (Fig. 1h).
The exact conditions for LSPR can be solved for a nanosphere using the Mie theory or a simple harmonic oscillator model,16–18 as the extinction (absorption + scattering) cross-section is expressed by,
![]() | (5) |
![]() | (6) |
Several key differences exist between LSPR and SPP that must be taken into account when designing a sensor. First, the factor of 2 in front of the interfacial dielectric constant depends on the geometry of the nanoparticle. The LSPR peak position will change with the shape in addition to the metal used and the local environment (Fig. 2).19–21 The larger the nanoparticles, the smaller the repulsion for electrons at opposite surfaces, and the more red-shifted the plasmon will be (Fig. 3).22–24 Second, the confined electron oscillations in LSPR lead to an intense local EM field, which can be several orders of magnitude stronger than the incident field strength. In nanoparticles with sharp edges, the field will be concentrated similar to a lightning rod, increasing the local field intensity further and improving the sensitivity to changes in the local environment.19–21,25,26 The EM field in LSPR decays in ∼10–30 nm and is therefore more sensitive to changes in distance from the surface of the metal and the local refractive index of the surrounding environment than the SPP that decays in ∼100–200 nm.6,13,18
![]() | ||
| Fig. 2 TEM images and electric field distributions of (a and d) Au nanosphere, (b and e) Au nanorod and (c and f) Au nanostar synthesized by wet-chemistry methods (reprinted with permission from ref. 19, Copyright 2012, IOP Publishing). | ||
![]() | ||
| Fig. 3 Representative plasmonic nanostructures for plasmon-enhanced sensing. (a) Extinction spectra (top), and optical images of different sized Ag nanospheres in aqueous solutions (a: 3.1 ± 0.6 nm, b: 13.4 ± 5.8 nm, c: 46.4 ± 6.1 nm and d: 91.1 ± 7.6 nm) (reprinted with permission from ref. 22, Copyright 2005, the Royal Society of Chemistry). (b) Extinction spectra (top) and optical images of Au nanorods with various aspect ratios (reprinted with permission from ref. 23, Copyright 2010, Elsevier B.V.). (c) Size- and shape-tunable localized extinction spectra of various Ag nanosphere and triangle arrays prepared by nanosphere lithography (top), and the representative AFM image of a Ag triangle array (reprinted with permission from ref. 24, Copyright 2005, Materials Research Society). | ||
Since the LSPR can be excited by incident light, the plasmon can also re-radiate its energy into the far field as scattering, with the size of the particle determining if absorption or scattering dominates (Fig. 1h and i).16 In small metal nanoparticles (less than ∼15 nm) electron–electron scattering quickly converts the energy of the LSPR into heat, which translates into a strong absorption.8,27,28 In larger particles the electron–electron surface scattering is reduced, and the energy of the plasmons will be re-radiated, leading to a strong scattering cross-section.8,27,28 The radiative dampening and electron–electron scattering make the lifetime of the LSPR much shorter than that of the SPP.29 Since the spectral width is inversely related to the lifetime, LSPR has a broader absorption peak than SPP, decreasing the figure of merit of these sensors. The absorption line width can be improved by optimizing the geometry and using planar arrays of metal nanoparticles, decreasing this disadvantage.
Finally, since both SPP and LSPR have local fields, coupling can occur between SPP/SPP, SPP/LSPR, and LSPR/LSPR when the supporting metal structures are brought within the local field decay length.13 The coupling can lead to an enhanced local field and a shift in the spectral position due to hybridization between the modes.30 For example, the local field enhancement of two spheres goes from ∼10 to ∼104 when they are aggregated.8,13 A shift in the SPP or LSPR frequency can be achieved by aggregating several plasmonic structures upon addition of an analyte, allowing lower detection levels than single particle or film-based designs.
One way to improve the refractive index sensitivity of a plasmonic sensor is to overlap the SPR peak of the metal film with the absorption band of the chromophore that binds to the plasmonic metal.38,39 Alternatively, the refractive index sensitivity can be improved by replacing the planar metal film with a large-area periodic nano-array pattern such as a nanosphere array, nano-disc array, or nano-triangle array which supports both a stronger local EM field and higher sensing area than planar films. Conjugated plasmonic nanoparticles can also be coupled to a Au film in a sandwich configuration when an analyte linker is present, further enhancing the shift of the SPR peak relative to the Au film alone.40
The rapid development of nanolithography technology has enabled the controllable fabrication of large-area (∼cm2) plasmonic nano-array patterns which combine the advantages of colloidal and planar substrates, including (i) a high density of “hot spots” on the order of a billion or more per cm2 with a tunable SPR band and field distribution, resulting in increased sensitivity compared to a bulk approach, (ii) good repeatability, (iii) facile integration with other components (e.g., microfluidics), and (iv) easier accommodation into a portable analytical instrument due to the miniaturized configuration. The nano-hole array is of particular interest to plasmonic sensing. As reported previously,41 LSPR can occur in an individual nano-hole, leading to a highly concentrated EM field near the edge. When the nano-holes are fabricated periodically over a large-area, SPP can be excited, leading to “extraordinary optical transmission” through the nano-holes at certain resonance wavelengths.42 The SPP mode is dependent on the periodicity of the nano-hole array. For a non-periodic nano-hole array, SPP mode disappears while LSPR mode still exists.41 The SPP and LSPR modes can be detected directly by incident light without the use of a prism, especially in the case of “extraordinary optical transmission” where front side illumination/back side detection is possible.41 Periodic structures significantly simplify the configuration of SPP-based sensors, eliminating the typical angle-resolved reflection arrangement used in the Kretschmann configuration, and allowing a nano-hole array-based plasmonic sensor to have a small footprint for miniaturization.
The spectral peaks of the SPP and LSPR in a periodic nano-hole array are highly sensitive to the dielectric properties at the interface, with the adsorption of molecules on the nano-hole leading to a shift in the peaks.43,44 The plasmonic nano-hole array is particularly suitable for integration into microfluidics,45 enabling real-time measurement of antibody-ligand binding kinetics46 and/or multiplexed detection.47 For example, Pang et al. have demonstrated a high-performance, microfluidic nano-hole array with a refractive index sensitivity of 1520 nm per RIU based on sequential injection of ethylene glycol in water.48
Plasmonic chip-based sensors have three distinct advantages:49–52 (i) label-free detection simplifies the configuration and operation of the sensor, and eliminates the use of multiple antibodies as employed in enzyme-linked immunosorbent assay (ELISA), the gold standard of analytical assays;49 (ii) real-time measurement of reaction kinetics when the plasmonic chip is integrated into a flow-cell, which provides a powerful tool for studying the binding events;50,51 (iii) long-range SPP (LRSPP) modes can be formed, which show orders of magnitude less damping than conventional SPP.52 The smaller damping in LRSPP allows narrower bandwidths and higher FOM, as well as a penetration depth of more than 1 μm,53 which is ideal for the analysis of living organisms.54 LRSPP modes are possible when a thin metal film is embedded between two dielectrics with similar refractive indices.55
Colorimetric detection is a simple and facile method for the detection of analytes in a solution because it provides a direct way to visualize the analyte concentration through the color change. Functionalized Au nanoparticles have been used in colorimetric detection of heavy metals, biological small molecules and biomacromolecules.61–71 When designing a sensor, surface modification is extremely important since it determines both the sensitivity and selectivity. The Mirkin group72 pioneered a colorimetric detection approach in which polynucleotides were selectively detected based on the distance-dependent LSPR coupling between Au nanoparticles.
Further, Hupp et al.73 have extended the colorimetric technique to detect heavy metal ions such as Pb2+, Cd2+, and Hg2+ using 11-mercaptoundecanoic acid (MUA) as a modifier for the Au nanoparticle surface. This modifier can recognize divalent metal ions by an ion-templated chelation process, wherein the binding of a divalent metal ion with a carboxylic group causes aggregation and a change in the extinction spectrum of the Au colloidal suspension. This technique has been further improved in sensitivity and specificity by using DNA as the recognition element,74–76 wherein the addition of metal ions results in the reversible association/dissociation of two complementary DNA-functionalized Au nanoparticles with intentionally designed base mismatches, forming DNA-linked aggregates, which change the color of the solution from red to blue.
Colorimetric detection of Hg2+ using DNA-functionalized Au nanoparticles is demonstrated in Fig. 5. When Hg2+ is present in the solution, it selectively coordinates to the mismatched T–T base pair and allows quantification of the Hg2+ concentration.75 Liu et al.77 have simplified the DNA–Au nanoparticle based sensing system using an optimal DNA sequence to realize colorimetric detection of Hg2+ at room temperature. This method has been extended to detect other metal ions by substituting the thymidine with synthetic artificial bases that selectively bind to other metal ions.78,79 The LSPR of nanoparticles is highly dependent on the shape and size of the nanoparticle, with nanorods and nanoshells being more sensitive to changes in the local refractive index than solid nanospheres.80,81 It has been reported that a nanorod's longitudinal LSPR mode exhibits six times higher sensitivity than a nanosphere counterpart.80 The dependence of refractive index sensitivity on the nanostructure's geometry can be understood conceptually by imagining how the concentrated charge and the resulting EM field are screened by the refractive index. The sharper the tip or more “pointy” the nanostructure is, the more localized the EM field will be, and the more sensitive the concentrated electron density will be to shift in the refractive index.
![]() | ||
| Fig. 5 (a) Schematic illustration of the colorimetric detection of Hg2+ using DNA–Au nanoparticles, and (b) color change in the DNA–Au nanoparticle solution in the presence of various representative metal ions (each at 1 μM) upon heating from room temperature (RT) to 47 °C (reprinted with permission from ref. 75, Copyright 2007, Wiley-VCH). | ||
Nanoparticle-based colorimetric sensors are typically operated in a vial, requiring professionals in a laboratory. Real-world samples also need to be pre-treated in a laboratory prior to testing. To solve these problems, nanoparticle-based colorimetric sensors have been incorporated into lateral-flow strips to make a point-of-care device. For example, this type of device has been used to detect cocaine in serum,82 proteins83 and nucleic acids.84
The sensitivity of LSPR to the distance between coupled nanoparticles has also been used to develop a plasmonic molecular ruler (PMR).85,86 PMRs can be used to measure the size of biomolecules in a label-free manner and to measure real-time molecular conformation changes and binding events. As compared to molecular rulers based on Förster resonance energy transfer (FRET), metal nanoparticle-based PMRs do not have photo-bleaching or blinking problems. In addition, PMRs are able to continuously monitor separations of up to 70 nm
86 as compared to the effective detection range of <10 nm for FRET.
An additional type of plasmonic nanoparticle-based sensor can be built according to the principles of plasmonic resonance energy transfer (PRET),87–90 in which a plasmonic metal nanoparticle serves as both the energy donor and the sensing reporter. When a plasmonic metal nanoparticle is in proximity to a molecule which has an absorption band overlapped with the LSPR, resonance energy transfer occurs between the metal and the molecule, giving rise to resonant quenching dips in the Rayleigh scattering spectrum of the metal particle. PRET sensors are suitable for both target detection and biomolecular imaging with nanoscale spatial resolution. Although the PRET method is superior in terms of sensitivity, selectivity, and multiplexing ability compared with methods based on dielectric property-induced shifts, it requires advanced instrumentation and algorithms to extract the molecule–plasmon interaction information.87–89
While plasmonic colloidal nanoparticle-based sensors provide a facile and rapid way to detect molecular or ionic targets using just the naked eye, this ease of use comes with the trade-off that their sensitivity (>nM in LOD) is lower compared with other optical or electrochemical techniques that can reach pM or lower LOD in fluorescence, SERS, and electrochemical sensors. In addition, colorimetric detection typically suffers from interference problems when used in complex matrices; and their stability in a physiological solution is of concern because aggregation can occur in the absence of an analyte.91,37 Efforts are therefore being made to develop robust colorimetric detection platforms for sensitive detection of analytes in real-world samples.
The energy transfer between the plasmon and the fluorophore is dominated by dipole–dipole interactions, with the exact mechanism being determined by the separation distance. First, if the plasmon and the fluorophore are within ∼1–10 nm of each other, the non-radiative local field of one dipole can excite the second one (Fig. 6). This is known as Förster resonance energy transfer (FRET). The efficiency of energy transfer in FRET depends on two factors.98
![]() | (7) |
![]() | ||
| Fig. 6 Dependence on distance and sphere radius of plasmon-enhanced fluorescence. (a) If the plasmon overlaps with the absorption of the fluorophore, an excitation enhancement is possible through the near field and FRET or scattering. (b) If the plasmon overlaps with the emission of the fluorophore, an emission enhancement is possible through the Purcell effect or FRET. (c) The excitation enhancement (red line) falls off quickly with distance, while the emission enhancement (blue line) is quenched at short distances but increases rapidly. The combined photoluminescence enhancement, equal to the emission enhancement times the excitation enhancement, peaks at around 10–30 nm (adapted from ref. 111). (d) The optimal sphere radius for excitation (red line), emission (green line), and total photoluminescence enhancement (black line) varies with the balance between absorption and scattering (reprinted with permission from ref. 101, Copyright 2009, Optical Society of America). | ||
The most obvious being the separation distance, R, which decays as 1/R6 because each dipole has a 1/R3 near field. The distance behavior is scaled by the factor R0, which depends on the spectral overlap between the emission of the donor's excited state and the acceptor's ground state absorption. The value of R0 is usually in the range of 3–8 nm in plasmon/fluorophore FRET. The FRET process is very efficient, approaching 100% for R < R0 in the presence of plasmon, because of the amplified local field and large absorption cross-section inherent to LSPR. FRET can occur from plasmon to fluorophore or from fluorophore to plasmon.
Secondly, the plasmon can enhance the radiative rate of the fluorophore through the Purcell effect (Fig. 6).99 The Purcell effect can be understood as follows. If a radiative dipole is placed in a resonant cavity, the emission intensity will be amplified on-resonance and quenched off-resonance when compared to free space. This is because the cavity modifies the local density of optical states (LDOS) as follows99
| ρLDOS(ω) ∼ |Eloc(ω)|2 | (8) |
The plasmon has a local field |Eloc|2 stronger than the free space light incident on the nanoparticle, so it increases the LDOS compared to vacuum and acts like a resonant cavity for the fluorophore. If the plasmon's absorption or scattering spectrally overlaps with the fluorophore's emission, the fluorophore's emission rate will be enhanced by the plasmon due to the change in the LDOS. The plasmon can either absorb the transferred energy or re-radiate it as scatter, resulting in a quenching or enhancement of the fluorophore's emission intensity. For a plasmonic dipole, the LDOS goes roughly as 1/R3, so the Purcell effect is usually seen outside the ∼10 nm range of FRET where coupling goes as 1/R6.100–102 It is possible that the plasmon's radiation or scattering rate can be enhanced by the fluorophore's LDOS. However, since the |Eloc|2 of the few electron oscillator dipole is usually on the order of free space, the resulting Purcell effect is negligible.101
The spectral overlap between the plasmon and the fluorophore determines whether FRET or the Purcell effect is present and whether these mechanisms lead to an enhancement or quenching of the fluorophore emission (Fig. 6). If the plasmon overlaps with the fluorophore's absorption, the excitation rate of the fluorophore will be enhanced from its free space value (Fig. 6a).28,30,100–105 For metal nanoparticles which primarily absorb, usually smaller than ∼15 nm, the fluorophore will be excited through FRET by the intense local field of the plasmon. For larger metal nanoparticles which primarily scatter, usually greater than ∼15 nm, an enhancement will be possible through both FRET at close distances (∼10 nm) and the Purcell effect at longer distances (∼10–50 nm) (Fig. 6c).28,30,100–105 As mentioned, the Purcell enhancement for the plasmon's radiation is usually negligible because the LDOS for the fluorophore is negligible, meaning that the excitation enhancement is primarily dominated by FRET at distances of a few nanometers.101 The excitation enhancement is maximized by using nanoparticles which absorb and do not scatter light, plus concentrate the local field in gaps or sharp points (Fig. 6d). Given the trade off between FRET and the Purcell effect, changes in the distance of a few nanometers will greatly change the excitation enhancement, allowing a sensor to be built which is highly responsive to minimal changes in the local environment.
If the plasmon is overlapped with the fluorophore's emission, an enhancement or quenching of the emission intensity is possible (Fig. 6b).28,30,95,100–105 If the fluorophore is within a few nanometers of the plasmon, its emission will be quenched by FRET into the plasmon. Although the plasmon could re-radiate this energy enhancing the emission intensity, the near field of the dipole also excites higher order modes in the plasmon which cannot re-radiate into the far field, leading to an overall quenching at separation distances of a few nanometers (Fig. 6c). At distances beyond FRET, a strong Purcell enhancement will be induced because of the enhanced LDOS of the plasmonic field relative to free space. This will lead to an increase of the radiative rate of the fluorophore compared to free space, and if the plasmon can scatter more efficiently than it absorbs, will result in a fluorescence emission enhancement. Metal nanoparticles with higher scattering than absorption efficiencies and separation distances beyond where non-radiative transfer is efficient will maximize the emission enhancement (Fig. 6d). It is usually very difficult, however, to obtain a pure excitation or emission enhancement because of the limited Stoke's shift of the dye and large absorption and scattering line widths of the plasmon relative to the fluorophore. Instead a balance is present between each effect. Generally this leads to a strong quenching being present at a few nanometer distance, which transitions into a large enhancement of 10–100 times in the range of ∼10–30 nm, and then slowly returns to normal emission strengths as the separation distance approaches hundreds of nanometers (Fig. 6c).106–115
When designing a sensor, the PEF theory can be simplified to manageably guide fabrication choices by considering three factors including (i) the enhancement of EM field induced by the plasmon, (ii) the spectral overlap of the plasmon's absorption and scattering with the absorption and emission bands of the fluorophore, and (iii) the gap (or space) between the plasmonic metal and the fluorophore. First, the total enhancement in excitation or emission can be thought of as |
dipole|4/|
higher order|2 ∼ |
|2 for most plasmonic structures.116 The quantity |
|2 can be easily estimated by FDTD simulations, helping to predict the enhancement in more complex structures.19,117 Second, the optimal positioning of the spectral overlaps will be defined by the separation distances possible and the size of the metal nanostructure (whether it absorbs or scatters). If metal nanoparticles which primarily absorb (less than ∼15 nm) can be used over distances less than ∼10 nm, the plasmon should be overlapped with the excitation of the fluorophore to avoid emission quenching. If nanoparticles which primarily scatter can be used, and precise control over distance is possible, the plasmon should overlap both peaks with a separation of around ∼10–30 nm to achieve the maximum fluorescence enhancement (Fig. 6d).28,30,95,100–105,110–112 The balance of quenching and fluorescence, while difficult from an optimization standpoint, allows sensors to be constructed that heavily modulates the fluorescence intensity through the interaction with a single analyte.
Finally, it should be noted that the plasmon can also absorb or scatter the fluorophore's emission in the far field beyond the range of PEF, distorting the measured spectrum. The far field scattering of the plasmon can also increase the path length of light in the solution or matrix holding the fluorophore, increasing the measured absorption efficiency. The far field interactions are not useful for sensors since they do not depend on the inter-nanoparticle distance allowing modulation, however, they can be useful as a passive component of a pre-existing fluorescence sensor to increase the measured signal. It should also be noted that the quantum efficiency of the fluorophore is also important, as PEF cannot occur if the efficiency of the dye is already 100%.101
Hu et al. have developed a plasmon-enhanced fluorescence probe based on these principles (Fig. 7a).131 First, an anti-aptamer linked with a Cyanine 3 dye is coupled to a prion protein. Second, the Ag@SiO2 core–shell nanoparticles are conjugated with an aptamer, allowing capture of the prion protein. This means that the fluorescence emission is enhanced by the LSPR of the Ag nanoparticles only when captured, thus producing a sensor. This PEF probe was used to successfully image the prion protein in live SK-N-SH cells (Fig. 7b). In addition, Lu et al. have developed a Ag@SiO2@aptamer–Cyanine 5 nanoparticle for detection of adenosine-5′-triphosphate (ATP).132 The aptamer linked with Cyanine 5 dye was immobilized onto the nanoparticle surface via hybridization with the complementary DNA in the absence of ATP, showing a 32-fold enhancement compared to free-standing Cyanine 5 in solution. When ATP was present in the solution, the binding of the aptamer with the DNA led to detachment of the aptamer–Cyanine 5 from the nanoparticle surface, modulating the fluorescence. This PEF sensor exhibited a linear response from 0 to 0.5 mM and a limit of detection (LOD) of 8 μM towards ATP.
![]() | ||
| Fig. 7 (a) Schematic illustration of plasmon-enhanced fluorescence detection of prion proteins; (b) distribution of the Ag@SiO2–Cyanine-labeled prion protein after endocytosis in live SK-N-SH cells (reprinted with permission from ref. 131, Copyright 2013, Elsevier). | ||
Fluorescence can also be enhanced via excitation of the LSPR of randomly distributed metal nanoparticles (such as Ag island film) or a periodic nano-array pattern on a solid substrate.137–139 Recently, Xie et al. have utilized a Ag nano-triangle array to enhance the fluorescence of near-infrared (NIR) dyes.140 NIR dyes in the “water window” (700–900 nm) are of particular interest to bio-sensing and bio-imaging since this spectral window allows deep penetration into biological fluids, cells, and tissues. In addition, at these wavelengths, the auto-fluorescence of biomolecules is minimal, reducing interference. Xie's results have shown that the fluorescence of a low quantum-yield (4%) NIR dye (Alexa Fluor 790) can be enhanced by two orders of magnitude using the Ag triangle array pattern. The SPP in a nano-hole array pattern allows for extraordinary optical transmission, which has been used to enhance the fluorescence of a dye by two orders of magnitude compared to a reference counterpart on a glass support.141
As mentioned before, SPP-based sensors were initially engineered using a metal film on a prism. The SPP field decays exponentially with a length of hundreds of nanometers,13 leading to a fluorescence excitation enhancement over a longer distance than LSPR.28,30,95 The SPP cannot re-radiate the absorbed energy into free space due to the momentum mismatch, so when a fluorophore is coupled into the SPP, the SPP re-radiates the fluorescence into the coupling prism or grating, leading to an angle-resolved emission (Fig. 1g).13,28,30,95 This well-known phenomenon is referred to as surface plasmon-coupled emission and occurs when the excited fluorophores are positioned up to 200 nm from a continuous thin metallic film (∼20–50 nm thick).96,142 If the fluorophore is at distances of a few nanometers to the metal film, the fluorophore can induce charge oscillations in the metal film, leading to quenching through FRET or coupling to lossy surface waves. Energy cannot be coupled into the SPP in this manner therefore no re-radiation through the coupling prism or grating exists, only quenching of the fluorescence. SPP-enhanced fluorescence offers the ability to selectively detect analytes with many advantages such as high spatial resolution, suppressed background interference, p-polarized emission, and a large effective detection distance. SPP-enhanced fluorescence has been used for detection of nucleic acids, proteins and other chemicals.143–145 LRSPP-based PEF sensors with a penetration depth on the micron scale are of particular interest in the analysis of large analytes such as bacteria and tissues.146–148 For example, Huang et al. have developed an LRSPP-based PEF sandwich immunoassay that was able to detect E. coli O157:H7 with an LOD of 6 colony forming units (CFU mL−1).147 In addition, an LRSPP-based PEF sensor exhibited an LOD of 34 fM and 330 fM towards prostate specific antigen (PSA) in a buffer and in human serum, respectively.149 It is worth noting that the LOD of the LRSPP-based PEF sensor was four orders of magnitude better than the detection using the plasmon peak shift with refractive index.
When the plasmon acts only as the energy acceptor, it is ideal to use nanoparticles with high absorption cross-sections and intense local fields but little scattering. This ensures that the plasmon only quenches the fluorescence and does not enhance the fluorescence intensity.28,30,100–105 Plasmonics offers two key advantages over organic dyes as an acceptor. First is the larger absorption cross-section which allows higher quenching efficiencies. The higher quenching efficiency leads to sensing distances up to twice that found in a fluorophore–fluorophore FRET system.170,171 Second, when designed properly, the plasmon does not re-radiate energy, reducing the background signal levels. These advantages are demonstrated in a recent study on the CdSe/ZnS quantum dot (QD)–Au system which showed that the energy transfer efficiency and mechanism depend on the size of the Au nanoparticles (Fig. 8).163 For 3 nm Au nanoparticles where LSPR is absent due to quantum size effects, energy transfer proceeds through nano-surface energy transfer (NSET) rather than FRET. In NSET, the energy is transferred into individual electron–hole pairs instead of collective LSPR mode, leading to a quenching rate of 1/R4.163,172 According to the NSET mechanism, a QD–DNA–Au ensemble was designed to detect Hg(II) ions.173 This nanosensor exhibited an LOD of 1.2 ppb towards Hg(II) in river water, which was lower than the maximum allowable levels of Hg(II) in drinking water (2 ppb) regulated by the U.S. Environmental Protection Agency (EPA).
![]() | ||
| Fig. 8 (a) Schematic illustration of CdSe/ZnS quantum dot–Au nanoparticle energy transfer. The CdSe/ZnS quantum dots with fluorescence emission at 572 nm were used as the energy donor while different sized Au nanoparticles (3, 15 and 80 nm) were used as the energy acceptor. Two complementary single stranded DNA strands with deliberately designed T–T base mismatches are employed to control the separation distance. (b) Normalized fluorescence emission intensity at 572 nm as a function of Hg2+ ion concentration. (c) Stern-Volmer plots showing the quenching efficiencies by the Au nanoparticles (reprinted with permission from ref. 163, Copyright 2011, American Chemical Society). | ||
Biological molecules can modulate and switch on/off the energy transfer process in FRET. This strategy has been extensively used to study intermolecular interactions, quantify analytes, and generate in vivo biological images.174,175 Plasmonic nanoparticles (i.e. Au and Ag) have many advantages as the energy acceptor in FRET sensors,163,176,177 including (i) high fluorescence quenching efficiency compared to organic dyes because of the stronger LSPR absorption (molar extinction coefficients up to 1010 cm−1 M−1), (ii) tunable quenching, (iii) limited photo-stability problems, and (iv) ease of labeling for biocompatible applications. Therefore, incorporation of plasmonic nanoparticles into energy transfer-based fluorescence sensors has led to great success in improving the sensitivity, stability, and reproducibility of FRET based sensors.
Another route is plasmon-mediated FRET in which a plasmonic nanostructure is used as a mediator rather than a donor or an acceptor. The plasmon can lead to both excitation and emission enhancements for both the donor and the acceptor. For PEF, these two effects are balanced vs. the distance for each fluorophore. The distance must be balanced to either individually increase the excitation and emission of the donor and acceptor, or collectively mediate the emission of the donor into the acceptor.168,178–181 For plasmon-mediated FRET, the donor's emission spectrum and the acceptor's absorption need to overlap with the plasmon absorption. The energy is first absorbed by the donor and then transferred to the plasmon. If at a close distance of a few nanometers, this transfer will occur by FRET and to higher order modes, which mainly lead to quenching but can also lead to re-radiation of the plasmon into the acceptor.28,30,100–105 If outside the FRET range, the donor can radiate into the plasmon by the Purcell effect, similar to the emission enhancement.168,178,181 The plasmon can then re-radiate this energy into the acceptor by FRET or the Purcell effect, just like PEF with a plane wave incident field, leading to an excitation enhancement for the donor.28,30,100–105 Unfortunately, the small Stoke's shift of most quantum dots and dyes means that the donor's emission and absorption strongly overlap, as does the acceptor's emission and absorption, preventing the ideal chain of energy transfer. Instead, the combined excitation and emission enhancements of the donor and acceptor must be taken into account as well as the plasmon mediated field.
A plasmon-mediated FRET sensor has been developed for the detection of human platelet-derived growth factor-BB (PDGF-BB).182 A fluorophore-linked single-stranded DNA and a quencher-terminated complementary DNA formed a DNA duplex with both the fluorophore and the quencher at the same end, enabling the FRET process. This FRET sensor was immobilized onto the surface of a Ag nanoparticle so that when PDGF-BB bound to the DNA strand was functionalized with the quencher, the DNA duplex was disrupted, leading to the detachment of the quencher from the fluorophore. FRET is then disabled and the fluorophore re-emits, modulating the signal. The sensor had a linear range of 6.2–50 ng mL−1 with an LOD of 0.8 ng mL−1. When the Ag nanoparticle was not used, replaced with a Au nanoparticle, or substituted with a shifted, non-matching LSPR peak, the sensitivity of the sensor was reduced, thereby proving that the spectral position of the LSPR played a key role in determining if the energy transfer was possible. Zhou et al. have recently reported a fascinating plasmon-mediated energy transfer sensor for Hg(II) detection.183 A 6-carboxyfluorescein (FAM) dye and a Ag nanoparticle were linked to two opposite ends of an aptamer to form a hairpin-shaped molecular beacon. Initially, the fluorescence emission of the FAM dye was quenched due to its proximity to the plasmonic Ag nanoparticle. When Hg(II) ions were added, the aptamer stretched due to DNA hybridization, leading to the formation of a gap larger than the FRET distance between the FAM dye and the Ag nanoparticle. The large gap relative to the FRET distance means that the LSPR of the Ag nanoparticles only enhanced but not quenched the fluorescence of the FAM dye when Hg(II) ions were present, signalling the presence of the ions. This sensor has achieved an LOD of 1 nM towards Hg(II) in phosphate saline buffer.
higher order|2 negligible and leading to |Eloc|2 enhancement during both excitation and emission. The difference in excitation and scattered Raman frequency is usually taken as negligible, leading to an overall enhancement of |
dipole|4/|
higher order|2 ≅ |
loc|4.116,200 Since the local field of the plasmon can be several orders of magnitude higher than that of the incident field, SERS signals can be detected even though the Raman scattering cross-section is itself very small (∼10−30 cm2 per molecule), leading to the EM enhancement mechanism being preferred in most SERS-based sensors. The design and fabrication of plasmonic nanostructures is the key for high-performance SERS sensors since the maximum field enhancement determines the sensitivity, reproducibility and applicability of the sensor.
For example, DNA bases can be identified by their intrinsic SERS spectral signature after directly binding onto a Au nanoshell nanoparticle (Fig. 9a).202–205 The extrinsic method is popular as it allows detection of analytes with weak or non-existent SERS signals. One method for achieving a strong SERS signal in the absence of an intrinsic signal is to conjugate Raman reporter molecules onto a plasmonic colloidal metal nanoparticle, forming an SERS label (tag) similar to a fluorescent dye tag. Nitrogen- and sulfur-containing molecules are usually used as Raman reporters because of their large affinity to typical plasmonic metals such as Au and Ag. The plasmonic field of the nanoparticle increases the SERS signal of abundant Raman reporter molecules, increasing the minimum detectable concentration compared to the intrinsic signal of a single molecule. Additionally, once an analyte is present, the distance between nanoparticles can be reduced and the plasmonic field increased (Fig. 9b), multiplying the SERS signal of abundant Raman reporter molecules covering the surface of the nanoparticle. This results in an amplified signal. Even though only one analyte molecule may cause the aggregation of nanoparticles, the signals of abundant Raman reporter molecules are enhanced. The amplification effect is maximal when the excitation wavelength of the laser source is overlapped with the LSPR of the nanoparticle,19 a condition referred to as resonance enhancement.
![]() | ||
| Fig. 9 (a) Assay of DNA hybridization using SERS via direct readout of spectral signatures of DNA bases (reprinted with permission from ref. 204, Copyright 2013, American Chemical Society). (b) Analyte-induced SERS enhancement through aggregation of plasmonic nanoparticles where Raman reporters are directly adsorbed on the particle surface. (c) Shape-dependent SERS enhancement (reprinted with permission from ref. 19, Copyright 2005, IOP Publishing). | ||
Maximizing the plasmonic EM field is critical to the development of a sensitive SERS probe. Many efforts have therefore been devoted to tailoring the SERS substrate.206–217 A recent study has investigated the effect of the Au nanoparticle shape on the SERS enhancement.19 The results have shown that nanorods have a larger enhancement than nanospheres (Fig. 9c), and Au nanostars exhibited the strongest SERS enhancement, which is attributed to the highly concentrated EM field in the sharp tips, similar to a lightning rod. Recently, a colloidal gold nanostar-based SERS sensor was developed for detection of glucose in saliva using optimization of the SERS substrate.218 In this sensor design, Au nanostars were conjugated with the glucose oxidase enzyme. When glucose was present, the enzyme molecules produced H2O2, which was detectable by SERS because of amplification by the Au nanostar's local EM field. In another approach, an interesting SERS substrate has been developed by an “On-Wire” lithography method. Multiple nano-gaps were fabricated in individual metal nanowires to form “hot spots”, resulting in a stronger SERS enhancement.219,220
Poor reproducibility often exists when Raman reporter molecules are directly adsorbed onto a bare nanoparticle's surface as a SERS tag. One problem is that the SERS tags can aggregate in a high strength ionic solution to the point of becoming sediment. Also, the Raman reporter molecules can easily desorb from the nanoparticle surface, leading to a loss of signal during detection. To overcome these issues, the Raman reporter molecules can be encapsulated between a metal core and a thin protective layer (e.g., polyethylene glycol (PEG) or SiO2) to form a sandwich-structured SERS tag.19,221–229 The sandwiched SERS tag offers several appealing features,19,226–228 such as (i) encapsulating a large number of Raman reporter molecules in a single particle, leading to a strong SERS signal; (ii) preventing leaching out of the Raman reporter molecules; (iii) enabling excellent water-solubility due to high hydrophility of the shell surface, and (iv) providing further flexibility for bio-conjugation. SiO2 is of particular interest as the shell material because current synthetic chemistry techniques are able to obtain a very thin SiO2 layer (1–2 nm thick) on the surface of Au nanoparticles using sodium silicate as a precursor; and the SiO2 layer is dense, biocompatible, water-soluble, and easy to surface-functionalize.19,224,225,229 For example, sandwich nanoparticles with a thin SiO2 outer-layer have been used for selective imaging of a target protein in prostate tissue.229
Biocompatible and nontoxic sandwich SERS probes have been demonstrated in drug delivery and in vivo tumor detection. Delivery and release of drugs can be tracked by SERS signals, which is advantageous since a near-infrared laser source can penetrate into biological fluids and thick tissues.230–233 Nie and coworkers234 have reported in vivo tumor targeting and spectroscopic detection using a sandwich SERS tag with a PEG shell (Fig. 10). The pegylated sandwich SERS tag exhibited excellent in vivo bio-distribution and pharmacokinetic properties as well as good colloidal stability over a wide range of pH and ionic strength, and thus superior performance to near-infrared semiconductor quantum dots. When conjugated to tumor-targeting ligands, the conjugated SERS tags were able to target tumor biomarkers such as the epidermal growth factor receptors on human cancer cells and in xenograft tumor models.
![]() | ||
| Fig. 10 (a) Preparation and schematic structures of Au nanoparticles encoded with a Raman reporter and coated with a layer of thiol–PEG. (b) Preparation of targeted SERS tags by using a mixture of thiol–PEG and a heterofunctional PEG. Covalent conjugation of an EGFR–antibody fragment occurs at the exposed terminal of the hetero-functional PEG. (c) In vivo cancer targeting and SERS detection by using ScFv–antibody conjugated Au nanoparticles that can recognize the tumor biomarker EGFR. SERS spectra were obtained from targeted and non-targeted SERS tags. Photographs show a laser beam focusing on the tumor site or on the anatomical location of the liver (reprinted with permission from ref. 234, Copyright 2007, Nature Publishing Group). | ||
An alternative way to make a SERS substrate is to generate nanostructures with the assistance of templates. For example, a Ag film can be deposited over a self-assembled polystyrene nanosphere monolayer on a solid-state support. This Ag film-on-nanosphere (AgFON) structure shows strong LSPR.237,238 The AgFON has been used to identify Bacillus subtilis spores, which are the simulants for Bacillus anthracis spores,239 achieving an LOD of 2.6 × 103 spores. A large-area periodic nano-array pattern provides better reproducibility and controllability for SERS sensing than non-periodic metal nanoparticles. For example, van Duyne and co-workers have fabricated plasmonic nano-triangle arrays using nanosphere lithography.240 The LSPR peak of a Ag nano-dot array can be systematically tuned in a wide range (400 to 6000 nm) by tailoring the gap, size, and height of the nano-dots; and the sharp tips of the nano-triangle can achieve an EM enhancement factor of ∼108.241,242 Additionally, strong “hot spots” are generated in the small gap between triangles when the nano-triangles are brought in proximity to form a bowtie structure. A gold bowtie with a gap of 8 ± 1 nm exhibited a SERS enhancement factor of 1011.243 The SPP-induced “extraordinary optical transmission” effect in a nano-hole array has also been utilized to enhance the SERS signal,244,245 with the enhancement maximized when the excitation wavelength of the laser matched the largest transmission peak of the nano-hole array. It is worth noting that LRSPP modes have also been utilized to generate angle-dependent SERS with a SERS enhancement factor of 9.2 × 108.246
It should be pointed out that the size, shape, and pitch of the nano-array define the effectiveness for SERS enhancement. For example, if a large SERS enhancement is to be formed using the strong “hot spots” that exist between the individual dots in a nano-dot array, the edge-to-edge gap must be less than 40 nm. Unfortunately, it is difficult to obtain patterns of this scale using conventional photolithography. Electron-beam lithography and focused ion beam lithography are able to produce nano-array patterns with these features, however both methods are low-throughput and cannot massively fabricate large-area nano-array patterns. More recent fabrication techniques promise to overcome this barrier, using nanosphere lithography and nanoimprinting lithography, offering a promising route to massively produce large-area periodic nano-array patterns at a relatively low cost.
As shown in Fig. 11, coupling SERS labels (tags) onto a chip is an excellent method to construct SERS sensors, showing higher sensitivity and selectivity as well as better anti-interference capability.186,187,189,206,247–255 As shown in Fig. 12a, Au nanostar@Raman reporter@SiO2 sandwich SERS tags were used for detection of adenosine triphosphate (ATP) by bringing the SERS tags in proximity to a gold chip.187 In this SERS sensor, the aptamer labeled with the SERS tag was hybridized with complementary single-stranded DNA on the Au chip to form rigid duplex DNA. When ATP was present, the binding of the aptamer with ATP led to the removal of the aptamer-linked SERS tag from the Au chip, modulating the SERS signal. This SERS sensor has achieved an LOD of 12.4 pM for ATP detection. The availability of substrates with high enhancement factors is the key to the success of SERS-based sensors. In the above-mentioned ATP sensor, the SERS enhancement mainly came from the plasmonic field of the individual Au nanostars with weak coupling between the Au nanostars and the Au chip. A different nano-architecture was therefore designed to couple the plasmonic fields and further enhance the EM field both in intensity and in space/volume.189 Three different configurations of SERS immunoassays were constructed including (i) Au sphere@MGITC@SiO2 SERS tags coupled to a planar Au film (where MGITC stands for malachite green isothiocyanate, a Raman reporter), (ii) Au sphere@MGITC@SiO2 SERS tags coupled to a Au triangle nanoarray pattern, and (iii) Au star@MGITC@SiO2 SERS tags coupled to a Au triangle nanoarray. The Au star@MGITC@SiO2/Au triangle nanoarray showed the highest sensitivity due to the generation of a larger density of “hot spots” in the 3D plasmonic nano-architecture. The SERS signal exhibited a dynamic linear range from 0.1 pg mL−1 up to 10 ng mL−1 and an LOD of 7 fg mL−1 towards the immunoglobulin G (IgG) protein in a buffer solution. The biosensor also demonstrated the ability to measure the level of vascular endothelial growth factor (VEGF) in clinical blood plasma samples taken from breast cancer patients and exhibited excellent resistance to interference from biological species in complex matrices.189 This sensor configuration provides a general detection scheme that can be adapted for measurement of a wide range of biomolecules. For example, a SERS sensor has been developed using these guidelines to detect hepatitis B virus DNA with an LOD of about 50 aM, as shown in Fig. 12c.186
![]() | ||
| Fig. 11 Scheme of SERS detection of adenosine through the plasmonic coupling of Raman reporter-labeled Au nanoparticles and a Au film (reprinted with permission from ref. 206, Copyright 2013, American Chemical Society). | ||
![]() | ||
| Fig. 12 Schemes of sandwich SERS tag-based assays of (a) ATP, (b) cancer biomarkers in blood plasma, and (c) hepatitis B virus DNA (reprinted with permission from ref. 186, 187 and 189, Copyrights 2012, 2013, American Chemical Society). | ||
Recently, paper- and fiber-based SERS substrates have become the focus of SERS research because of their flexibility, conformability, efficient uptake of analytes, and efficient transport due to hierarchical vasculature and high specific surface area.254–257 These novel SERS substrates are also easily integrated into conventional chromatography, microfluidics, and other biological assays, and therefore hold great promise in future SERS applications. In addition, one of the most promising features of future SERS is the capability of multiplexed detection in live animals.258–261 SERS therefore holds great promise for chemical sensing and in vivo biological imaging as the plasmonic-based design continues to progress.
In the future, plasmonics will continue to be developed for not only enhancing the sensing signal but also for designing new detection schemes based on fluorescence and SERS devices. To date, many plasmon-involved sensors have been constructed on an empirical basis. A “Device-by-Design” fashion is desirable to ensure the continued development of high-performance plasmon-mediated sensors, but this is impossible without first establishing an in-depth understanding of the underlying theory. The mechanisms of plasmonic transduction and plasmon-enhanced SERS have already been well studied. Although significant progress has been made in understanding the underlying mechanism of plasmon-mediated fluorescence, more systematic studies need to be performed to clarify the confusion on the issue of enhancement versus quenching. The limits of plasmonic sensors can only be pushed once the underlying theory is well understood.
Both LSPR and SPP can be tailored by using nanostructured materials and architecture, providing great opportunity and flexibility for designing plasmon-enhanced sensors. The success of future plasmon-enhanced sensors directly depends on the effective design of plasmonic materials/architectures and the development of new methods for their fabrication.
| This journal is © The Royal Society of Chemistry 2015 |