Ultra-sensitive vibrational spectroscopy of protein monolayers with plasmonic nanoantenna arrays
Edited by Erich P. Ippen, Massachusetts Institute of Technology, Cambridge, MA, and approved September 21, 2009
Abstract
Infrared absorption spectroscopy enabling direct access to vibrational fingerprints of the molecular structure is a powerful method for functional studies of bio-molecules. Although the intrinsic absorption cross-sections of IR active modes of proteins are nearly 10 orders of magnitude larger than the corresponding Raman cross-sections, they are still small compared to that of fluorescence-label based methods. Here, we developed a new tool based on collective excitation of plasmonic nanoantenna arrays and demonstrated direct detection of vibrational signatures of single protein monolayers. We first tailored the geometry of individual nanoantennas to form resonant structures that match the molecular vibrational modes. The tailored nanoantennas are then arranged in such a way that their in-phase dipolar coupling leads to a collective excitation of the ensemble with strongly enhanced near fields. The combined collective and individual plasmonic responses of the antenna array play a critical role in attaining signal enhancement factors of 104–105. We achieved measurement of the vibrational spectra of proteins at zeptomole levels for the entire array, corresponding to only 145 molecules per antenna. The near-field nature of the plasmonic enhancement of the absorption signals is demonstrated with progressive loading of the nanoantennas with varying protein film thicknesses. Finally, an advanced model based on nonequilibrium Green's function formalism is introduced, which explains the observed Fano-type absorption line-shapes and tuning of the absorption strengths with the antenna resonance.
Sign up for PNAS alerts.
Get alerts for new articles, or get an alert when an article is cited.
A new generation of antennas operating at the optical and infrared frequencies is emerging from the well developed concepts in microwave antenna theory. Plasmonic nanoantennas, with their unique ability of focusing light beyond the diffraction limit, are at the core of a myriad of new exciting opportunities in photonics (1–3). By exploiting extremely strong and localized fields in the visible wavelength range, signal enhancements of several orders of magnitude have been demonstrated in second harmonic generation (4), fluorescence (2, 5) and surface enhanced Raman scattering (SERS) (1, 2, 6). The plasmonic enhancement of optical near fields can also be extended to the infrared frequencies enabling dramatic signal enhancement in infrared (IR) spectroscopy. In analogy to SERS, this method is called surface enhanced infrared absorption (SEIRA) spectroscopy (7–14). Until recently, the bulk of SEIRA studies have revolved around enhancements achieved via chemically prepared or roughened metal surfaces (7–10, 14). In these substrates, however, signal enhancement factors have been limited to 10–100 range due to their random nature (14). Uncontrolled surface geometries also cause poor spectral overlap between plasmonic resonances and the molecular vibrational modes of interest. These limitations result in weaker absorption signals preventing reproducible vibrational measurements from protein monolayer films. In contrast, engineered nanoantenna arrays can support spectrally well-defined resonances with strong local fields. Furthermore, plasmonic resonances in these structures can be tuned to the spectral regions of interest with high spatial reproducibility. Recently, enhanced absorption signals have been observed from individual nanoantennas (8, 11–13). However, surprising new phenomena arise as a result of the collective resonant excitation of the nanoantenna ensembles. In well-engineered nano-antenna arrays, the radiative dipolar coupling and the interference of the multiple scattering from antennas in the array can be used for the spectral narrowing of the far field response. More importantly, calculations indicate that these collective resonances are linked to the strongly enhanced near-field intensities (15–19). In this manner, much stronger coupling between the incident field and the transition dipole moments of the proteins can be achieved compared to the individual antennas and chemically prepared substrates.
In this letter, we present an ultrasensitive collectively enhanced IR absorption (CEIRA) spectroscopy technique allowing direct identification of vibrational signatures of single protein monolayers of silk fibroin. Measurement of vibrational signatures with zeptomole level protein detection limits are achieved due to the 104–105 signal enhancements. Absorption signals far surpassing those of the individual nanoantennas are shown with high reliability and reproducibility. Experimentally observed asymmetric absorption spectra and relative tuning of the vibrational absorption bands are explained using a nonequilibrium Green's function (NEGF) technique (20–21). The near-field character of the enhanced signals is experimentally mapped by progressive loading of the nanoantenna with varying thicknesses of protein films. The large signal-to-noise ratios obtained here with CEIRA spectroscopy enables functional studies of monolayer protein films, a question of fundamental importance in biochemistry and biophysics.
Silk Fibroin.
Our protein of choice is silk fibroin, which has attracted considerable attention because of its exceptional mechanical and optical properties (22). Apart from its intrinsic properties, the ability to form uniform layers with excellent thickness control (23) makes silk protein an ideal model system for systematic studies of the infrared enhancement mechanisms in biomolecules.
Mechanisms of Collective Nanoplasmonics.
Our approach is first to engineer nano-rod antenna arrays supporting spectrally narrow collective plasmon excitations at midinfrared frequencies, specifically at the amide-I (1,660 cm−1) and amide-II (1,537 cm−1) vibrational normal modes of the protein backbone. The lithographically fabricated structures are then coated with protein molecules whose vibrational dipole moments act as a load to the nanoantenna. Nanoantenna arrays, driven by an infrared light source, efficiently funnel the incident light to their load through the collectively enhanced plasmonic excitations. This collective behavior is controlled by the local electromagnetic field driving each nanoantenna. For an individual nanoantenna, the local field is simply the incident field that excites the localized surface plasmons (LSP) serving as the electric dipoles. In an array, however, the local field includes (i) the incident electric field, and (ii) the field created by the other induced dipoles:where E0 and k = 2π/λ are the amplitude and the wavevector of the incident wave. P→j is the induced polarization of the jth antenna in the array while c→ij represents the long range dipolar interaction matrix without the phase term (see SI Text). The strength of the second term is controlled by the ensemble arrangement due to the phase component. For an arrangement where the induced dipolar field interactions are almost in phase, the local fields can become extremely large. Therefore, arrays of antennas generate a collective response resulting in strongly enhanced near-field excitations compared to the individual antenna.
Results and Discussion
Individual and Collective Resonances of Plasmonic Nanoantennas.
We engineer nanoplasmonic antenna arrays for enhanced collective response starting with the individual antenna behavior. To obtain individual response with a good signal to noise ratio using a conventional Fourier-transform infrared (FTIR) microscope, we used randomly positioned antennas. The random arrangement of the nanoantenna array cancels out any diffractive behavior (24). As shown in SI Text and Fig. S1, the spectral measurements are independent from the randomization process and clearly reflect the individual antenna response. Fig. 1A shows the infrared reflectance measurements from randomized arrays consisting of varying length and a fixed width (230 nm) of nanorods pointing in the same direction on an area of 100 × 100 μm2. The structures are fabricated on silicon substrates using electron beam lithography. A thin titanium adhesion layer is first deposited by electron beam evaporation, followed by ≈70 nm thick gold layer. After lift-off, well-defined rod shaped structures were formed as shown in the scanning electron microscope (SEM) image of an individual nanoantenna (Fig. 1A Inset). For incident light polarized parallel to the long axis of the rods (//-polarized), the nanoantenna resonance follows a linear scaling relation (Fig. 1B). This scaling behavior could be understood from antenna theory concepts. A traditional half-wave dipole antenna of length L has a tiny feed-gap in between two segments of length λ/4 at which the antenna is connected to its load. When the feed-gap impedance is matched to the impedance of the load, induced current is maximized in the antenna and the load circuit. Accordingly, resonance occurs at a wavelength corresponding to the total length of the two λ/4 segments (i.e., L = λRes/2). A nanorod antenna is fabricated as single particle with no feed-gap and the impedance matching condition is automatically met. As a result, the resonant excitation wavelength of a nanorod antenna with length L occurs approximately at λRes = 2Lneff. Here, neff is the effective refractive index, accounting for the inhomogeneous dielectric environment (substrate and air) surrounding the antenna. Slight deviations from the ideal half-wave dipole antenna behavior are due to the increasing penetration depth of the incident radiation at IR frequencies and decreasing aspect ratio of the long axis versus the short axis. Complete resonance, which is overlapping of the plasmonic excitations with the vibrational amide-I and amide-II bands, occurs for the nanoantenna length L = 1,100 nm. For incident light that is polarized perpendicular to the rod long axis, plasmon resonances are not excited over the spectral region of interest (Fig. 1A). The cross sectional near-field profile of the individual nanoantenna shown in Fig. 2A is obtained by three-dimensional finite difference time domain (3D FDTD) simulations. Enhanced field intensities with a factor of ≈102–103 are observed for parallel polarized (E//) incident light near the tips of the nanoantenna. Although these enhancement factors are quite impressive, single monolayer measurements require even higher near-field intensities. This cannot be simply achieved by changing the nanoantenna geometry (25).
Fig. 1.
Fig. 2.
Collective excitations in well engineered antenna ensembles, however, can be exploited to achieve dramatically enhanced near-field intensities with much narrower far field spectral responses (15–19). In a periodically ensembled antenna array with a period d, long range dipolar interactions among the nanoantennas become dominant over the localized near-field couplings for separations more than several tens of nanometers. At a fixed incident wavelength, the interaction term (given by the second component in Eq. 1) accumulates larger phase delays when propagating between the antennas with increasing lattice spacing. The narrow resonance occurs for d slightly smaller than a critical periodicity dc where the light fields corresponding to the diffractive grating order (i,j) change from evanescent to radiative:Here, nSi is the refractive index of the substrate and (i,j) are the diffraction grating orders. Collective excitations are created when the dipolar interactions among the antennas are almost in-phase just before the appearance of the grating order resulting in radiation damping (Fig. 2B). Tuning of the individual antenna LSPs to this incident wavelength strongly enhances the collective response and results in efficient funneling of the incident light into much stronger near-field excitations (15–19). This behavior is well captured by the FDTD simulations presented in Fig. 2A. An order of magnitude stronger near-field intensities than those of individual antennas are observed at slightly shorter wavelengths than λSi(1,0). Here, λSi(1,0) is the critical wavelength where the (1, 0) grating order of the silicon interface becomes radiative at a given lattice periodicity (λSi(1,0) = nSid). This collectively enhanced near-field behavior is reflected as a narrowing of the far-field spectral response due to the suppression of the radiation damping with the confinement of the electromagnetic field within the array (16).
To enhance absorption signals, arrays are optimized using far field reflectance measurements by spectrally tuning the narrow resonances to the protein vibrational bands. Fig. 3A shows an SEM image of a fabricated periodic array of rod shaped antennas. The arrays consist of 1,100 nm long rods with varying periodicities ranging from 1.5 μm to 2.0 μm, and a random set that serves as a control to probe the individual antenna response. The resulting reflectance spectra are shown in Fig. 3B. For periodicities smaller than 2 μm, collective resonance peaks are above the (1, 0) grating order cutoff of the silicon interface (dashed line in Fig. 3C) where the field is evanescent (d < dc). This results in narrower plasmonic far field resonances with respect to individual antenna as shown in Fig. 3C (red triangles). A progressive blue shift is expected for greater separations as the coupling among antennas is reduced, and the optical response converges back to that of an isolated particle. For d > dc, the grating order field changes from being evanescent to radiative form in nature, which is associated with strong damping of the collective resonance and broadening of the line-width. In fact, at 2 μm periodicity, resonance occurs at a wavelength below the grating cut-off (d > dc) and results in a broader linewidth. In comparison with the randomized arrays, the arrays with 1.5 μm and 1.6 μm periodicities have significantly narrower resonance line-widths (≈1 μm versus 2.75 μm for the individual particle behavior). As shown in Fig. 3C, the 1.6 μm periodic array offers the best combination of narrow line-width and spectral overlap with the protein amide-I band at 1,660 cm−1.
Fig. 3.
SEIRA Enhancement by Collective Resonances.
Direct identification of the vibrational signatures of the protein monolayers is achieved by using this collective excitations. Here, the protein layer is applied to the nanoantenna substrates by uniform spin coating of a thin film of silk fibroin (detailed in SI Text). As mentioned before, a unique advantage of silk is that it allows fine control over the protein film thickness by varying its concentration in solution (22). We coat homogeneous films on the nanorod substrates as thin as 2 nm, corresponding to essentially a single protein monolayer. Atomic force microscopy (AFM) is used to confirm the uniformity and the thickness of the film. Initially, the scanning of the tip is performed in contact mode to scratch away the film down to the silicon substrate. Subsequent measurements at reduced force are then used to determine the film thickness (Fig. 4A). A combination of AFM and profilometer measurements confirmed the linear dependence of the film thickness on solution concentration.
Fig. 4.
As shown in Fig. 4C, protein absorption bands are clearly noticeable within the optical spectra collected from the protein-coated antenna arrays with 1.6 μm periodicity. Dips in the plasmonic response as a result of the amide I and II absorption bands are indicated in the figure at 1,660 and 1,537 cm−1, respectively. Capacitive loading of the antenna with the protein layer results in slight red shifting of the plasmonic resonances (26, 27) (dashed curve in Fig. 4C). This shift is corrected using a polynomial fitting procedure in difference spectrum measurements (ΔR/R0 = Rbefore/R0 − Rafter/R0, where R0 is the reflection signal from reference mirror). Fig. 4D shows the difference spectra of the periodic and the randomized arrays. The enhanced absorption signals well above the noise level are observed for the antenna arrays with narrow line-width (1.6 μm periodicity). Control measurements were performed on the bare silicon substrate with protein films of the same thickness in a region far from any fabricated nanoantennas. No absorption features were observed from the control samples. The observed signal (ΔR/R0) of 6.8% in the periodic structures is nearly an order of magnitude higher than that of the randomized array, where the difference signals were 0.9%. This observation is in agreement with the enhanced near-field intensities predicted by the FDTD simulations (Fig. 2A).
To calculate the near-field enhancement factor, we compared the enhanced signal collected from the 1.6 μm periodic array to the expected reflectance signal from a 2 nm thick silk film on bare silicon substrate. Because the signal from protein films on bare silicon is below the noise level, instead we performed IR reflection absorption spectroscopy (IRRAS) measurements to obtain the expected value of the absorption signal for normally incident light. The frequency dependent dielectric function of the silk film was experimentally determined by fitting the measured IRRAS signal to a Lorentzian oscillator model (28). For a thin film (i.e., t ≪ λ) deposited on a metal substrate, the reflectance signal is approximately given by (10):where θ is the angle of incidence, t is the film thickness and ε2 = ε2,Re + iε2,Im is the dielectric function of the silk layer. IRRAS measurements are performed at a grazing angle (80°) for 2 nm thick protein film coated on a 100 nm thick gold layer deposited on a silicon substrate. The incident light was polarized in the plane of incidence. The resonant frequencies, oscillator strengths and full-width half maxima of the Lorentz oscillator model for ε2,were varied to obtain a fit with the data using Eq. 3. An estimated difference absorption signal of 4.7 × 10−2% is obtained by using three-layer Fresnel equations and ε2 fitted through the IRRAS measurements. For an accurate estimate of the enhancement factor, we also need to include the following factors (see SI Text for details). The enhanced signal mainly comes from a small quantity of molecules at the close vicinity of the N2 nanorod tips (n = 63 is the number of rows and columns). Second, in contrast to the studies conducted with self-assembled monolayers (SAMs) (7–13), here the protein molecules are physisorbed (see SI Text), which results in the following differences. (i) Unlike the SAMs, the transition dipole moments of the physisorbed proteins have no fixed orientation with respect to the metal surface normal. Accordingly, we expect that approximately one third of the transition dipoles of the molecules at the nanorod tips contribute to the absorption signal (see SI Text). (ii) Additionally, unlike in chemisorption, lack of molecular binding to the metal surface rules out any contribution of the chemical effects. (iii) Finally, given that the silk film is only 2 nm thick, it is unlikely that the physical deposition method results in the same degree of uniform coverage over the 70 nm high vertical sidewalls of our nanorods as would be possible with a SAM method. With these in consideration, we estimated the signal enhancement to be within the range of 104–105 (see SI Text). Such large enhancements allow us to detect absorption signals even in the raw spectral measurements from monolayer protein films with a commercial FTIR microscope. The detection volume of the antenna was calculated by considering the lateral area of the nanotips and the thicknesses of the monolayer protein film. Given that the density of the protein is 1.4 g/cm3 (29) and the molecular mass is ≈375 kDa (30), we estimate that the measured absorption signals are obtained from ≈300 zeptomoles for the entire array, corresponding to only 145 silk molecules per antenna (see SI Text). The large signal to noise ratios achieved in our measurements indicate that we should be able to observe vibrational signatures from even smaller quantities of protein molecules down to a few tens of zeptomoles.
In addition to the strong enhancements observed for the collective resonances, we observe variations in the absorption lineshapes as the spectral overlap between the antenna resonances and those of the protein amide bands are varied. We explain this experimental observation using a model based nonequilibrium Green's function formalism, which has been successfully applied in molecular/nanoelectronics (20–21). The strength of the NEGF formalism is that it provides a natural framework for describing the wave nature of the elementary excitations (electrons, phonons, vibrons, etc) in the presence of incoherent and dissipative processes. Conceptually, the system is partitioned into a nanoantenna and a protein load driven by the excitation of the plasmonic resonances as shown in Fig. 5A. Direct absorption of the incident light by the protein load is neglected because this absorption is expected to be very weak. The coupling of the incident light to the nanoantenna is expressed using a self-energy matrix [Σinc] whose anti-Hermitian component Γinc = [Σinc − Σinc+] is the broadening of the plasmonic antenna resonance due to the radiative coupling. Vibrational absorptions of the molecules are incorporated into the antenna/protein Hamiltonian [HAnt+Load] using a coupling matrix [κAnt,Pro]. This matrix is designed to give unit transmission in spectral regions far from the absorption bands. A combination of Lorentz oscillator dips in the matrix strength are used to account for the absorption bands at the spectrally related regions. Reflected light is calculated through the second self-energy term [Σref] with Γref = [Σref − Σref+], which takes into account the open boundary condition in a similar fashion to the perfectly matching layers (PML) used in FDTD simulations. Reflected light intensity is obtained by R = tr(ΓincGAnt+LoadΓrefGAnt+Load+). Here, [GAnt+Load] = [(ℏω)I − HAnt+Load − Σinc − Σref]−1 is the Green's function of the antenna/protein system coupled to the open boundaries and [I] is the identity matrix. Retarded Green's function [GAnt+Load] is a frequency dependent non-Hermitian matrix incorporating complex self-energy terms through which the radiative/nonradiative decaying of the plasmonic excitations are taken into account. The model absorption lines are defined as Lorentz oscillators at 1,660 cm−1 and 1,537 cm−1 with a width of 30 cm−1. As shown in Fig. 5B, for the amide-II band, when there is an exact match between the plasmonic antenna resonance and the absorption band, the spectral lineshape is Lorentzian. For slightly detuned absorption resonance, such as for the amide-I, an asymmetric Fano type line-shape is clearly noticeable. The relative strength of the absorption resonances is due to the relative coupling strength of the incident infrared radiation to the protein vibration band. As expected, with the detuning of the antenna resonance the relative absorption strength of the vibration bands can be modulated (Fig. 5C).
Fig. 5.
Silk Film as a Near-Field Probe.
The ability to control the thickness of the silk protein films from several nanometers to several micrometers provides a unique opportunity to probe the near-field behavior of the nanorod antenna. Due to the rapid decaying of the strongly enhanced near fields with distance from the nanorod surfaces (as can be seen in the FDTD simulations in Fig. 2A), saturation of the enhancement is expected to occur for films as thin as 40 nm. Although the numerical calculations offer insight into this observation, we use silk films as a near-field probe and experimentally demonstrate the surface nature of the enhancement by varying the film thicknesses (2, 4, 20, 40, and 100 nm) on identically patterned nanoantenna substrates. Fig. 6A shows the difference absorption spectra for the 1.6 μm periodic array with L = 1,100 nm. Measurements are also taken from a periodic structure with 2 μm perodicity and a randomized array for the same rod length. The strength of the absorption signal (amide-I) is plotted as a function of the film thickness in Fig. 6B for all of the structures (signals are normalized to take into account the difference of nanoantenna numbers in each array). Enhancement in the absorption signal strength is significant with increasing protein film thickness from 2 to 20 nm, but it appears to saturate as the thickness reaches to 40 nm. This behavior is in contrast to a steady linear dependence one would expect from Beer-Lambert's law for very thin films (see SI Text for details). Beyond 100 nm film thickness, bulk infrared absorption signals from thick protein films become observable. The other critical observations in Fig. 6 are as follows. Both of the periodic arrays with collective resonance display markedly higher signals than the randomized arrays, in agreement with the discussion presented above. The consistency of the data from periodic samples emphasizes the repeatability of the CEIRA measurements on the lithographically patterned substrates. Finally, experimentally measured enhancement saturation resulting from decaying near-field intensity distribution is confirmed by numerically calculating sampling volumes (for the 1.6 μm periodic array). For a given silk film thickness, t, the volumes are computed from the FDTD by summing over the simulation grids, which are within a distance t from the rod surface and have a near-field intensity >1/e2 of the average maximum. The agreement between the calculated mode volumes and the absorption signal saturation strongly emphasizes the surface nature of the CEIRA effect.
Fig. 6.
Conclusions
We demonstrated a new surface enhanced spectroscopy technique (CEIRA) based on collective plasmonic excitations created by tailoring of the dipolar interactions in engineered nano-antenna arrays. Up to 105-fold enhancement of vibrational/absorptional signatures of the monolayer thick protein films is obtained with high signal to noise ratios. We successfully demonstrated detection of 300 zeptomoles of proteins for the entire array, corresponding to only 145 molecules per antenna. With progressive loading, we resolved the near-field plasmonic behavior of the metallic nano-rod antenna in well agreement with our 3D FDTD simulations. We developed a Green's function model, which explains the experimentally observed asymmetric absorption spectra due to the detuned coupling between the plasmonic excitations and the molecular vibrations.
Materials and Methods
Antenna Substrate Fabrication.
Nanoparticle antenna arrays were fabricated on silicon wafers by a single layer liftoff process as described in SI Materials and Methods.
FTIR Spectroscopy Measurements.
All spectral data were taken on a Fourier transform infrared spectrometer with an IR microscope as described in SI Materials and Methods.
Silk Film Preparation.
Silk films were prepared by a spin coating procedure described in SI Materials and Methods.
Numerical Simulations.
CST Microwave Studio was used for the FDTD simulations as described in SI Materials and Methods.
Acknowledgments.
We thank Ozgur Sahin for the AFM measurements, Peng Wang/Bruker Optics for IRRAS measurements, and Mike Ekins for illustrations. This work is supported in part by National Science Foundation Small Grants for Exploratory Research Award ECCS-0849603 (to H.A.), a Massachusetts Life Science Center New Investigator Award (to H.A.), Boston University College of Engineering Dean's Catalyst Award (to H.A. and S.E.), and the Boston University Photonics Center and Army Research Laboratory (to H.A. and S.E.). Silk fibroin research is sponsored by in part by the U.S. Army Research Laboratory, U.S. Army Research Office Contract W911 NF-07-1-0618, and Defense Advanced Research Projects Agency-Defense Sciences Office.
Supporting Information
Supporting Information (PDF)
Supporting Information
- Download
- 174.94 KB
References
1
S Lal, S Link, NJ Halas, Nano-optics from sensing to waveguiding. Nature Photonics 1, 641–648 (2007).
2
C Genet, TW Ebbesen, Light in tiny holes. Nature 455, 39–46 (2007).
3
A Artar, AA Yanik, H Altug, Fabry-Perot nanocavities in multilayered plasmonic crystals for enhanced biosensing. Appl Phys Lett 95, 051105. (2009).
4
CK Chen, TF Heinz, D Ricard, YR Shen, Surface-enhanced second-harmonic generation and Raman scattering. Phys Rev B 27, 1965–1979 (1983).
5
HG Frey, S Witt, K Felderer, R Guckenberger, High-resolution imiaging of single fluorescent molecules with the optical near-field of a metal tip. Phys Rev Lett 93, 200801 (2004).
6
K Kneipp, et al., Single molecule detection using Surface-Enhanced Raman Scattering (SERS). Phys Rev Lett 78, 1667–1670 (1997).
7
M Osawa, M Ikeda, Surface-enhanced infrared absorption of p-nitrobenzoic acid deposited on silver island films: Contributions of electromagnetic and chemical mechanisms. J Phys Chem 95, 9914–9919 (1991).
8
TR Jensen, RP Van Duyne, SA Johnson, VA Maroni, Surface-enhanced infrared spectroscopy: A comparison of metal island films with discrete and non-discrete surface plasmons. Appl Spectrosc 54, 371–377 (2000).
9
SM Williams, AD Stafford, KR Rodriguez, TM Rogers, JV Coe, Accessing surface plasmons with Ni microarrays for enhanced IR absorption by monolayers. J Phys Chem B 107, 11871–11879 (2003).
10
D Enders, Pucci, A Surface enhanced infrared absorption of octadecanethiol on wet-chemically prepared Au nanoparticle films. Appl Phys Lett 88, 184104 (2006).
11
J Kundu, F Le, P Norlander, NJ Halas, Surface enhanced infrared absorption (SEIRA) spectroscopy on nanoshell aggregate substrates. Chem Phys Lett 452, 115–119 (2008).
12
F Neubrech, et al., Resonant plasmonic and vibrational coupling in a tailored nanoantenna for infrared detection. Phys Rev Lett 101, 157403 (2008).
13
R Bukasov, JS Shumaker-Parry, Silver nanocrescents with infrared plasmonic properties as tunable substrates for surface enhanced infrared absorption spectroscopy. Anal Chem 81, 4531–4535 (2009).
14
K Ataka, J Heberle, Biochemical spplications of surface-enhanced infrared absoprtion spectroscopy. Anal Bioanal Chem 308, 47–54 (2007).
15
M Meier, A Wokaun, PF Liao, Enhanced fields on rough surfaces: Dipolar interactions among particles of sizes exceeding the Rayleigh limit. J Opt Soc Am B 2, 931–949 (1985).
16
B Lamprecht, et al., Metal nanoparticle gratings: Influence of dipolar interaction on the plasmon resonances. Phys Rev Lett 84, 4721–4724 (2000).
17
S Zou, N Janel, GC Schatz, Silver nanoparticle array structures that produce remarkably narrow plasmon lineshapes. J Chem Phys 120, 10871–10875 (2004).
18
B Augie, WL Barnes, Collective resonaces in gold nanoparticle arrays. Phys Rev Lett 101, 143902 (2008).
19
S Zou, GC Schatz, Silver nanoparticle array structures that produce giant enhancements in electromagnetic field. Chem Phys Lett 403, 62–67 (2005).
20
AA Yanik, G Klimeck, S Datta, Quantum transport with spin dephasing: A nonequilibrium Green's function approach. Phys Rev B 76, 045213. (2007).
21
S Datta, Electrical resistance: An atomistic view. Nanotech 15, S433–S451 (2004).
22
FG Omenetto, DL Kaplan, A new route for silk. Nature Photonics 2, 641–643 (2008).
23
BD Lawrence, FG Omentetto, K Chi, DL Kaplan, Processing methods to control silk fibroin film biomaterial features. J Mater Sci 43, 6967–6985 (2008).
24
AA Yanik, X Wang, S Erramilli, MK Hong, H Altug, Extraordinary midinfrared transmission of rectangular coaxial nanoaperture arrays. Appl Phys Lett 93, 081104. (2008).
25
KB Crozier, A Sundaramurthy, GS Kino, CF Quate, Optical antennas: Resonators for local field enhancement. J Appl Phys 94, 4632–4642 (2003).
26
A Alu, N Engheta, Tuning the scattering response of optical nanoantennas with nanocircuit loads. Nature Photonics 2, 307–310 (2008).
27
M Schnell, et al., Controlling the near-field oscillations of loaded plasmonic nanoantennas. Nature Photonics 3, 287–291 (2009).
28
VP Tolstoy, IV Chernyshova, VA Skyrshevsky Handbook of Infrared Spectroscopy of Ultrathin Films (John Wiley and Sons Inc, Hoboken, NJ), pp. 243–286 (2003).
29
JO Warwicker, The crystal structure of silk fobroin. Acta Crystallogr 7, 565–571 (1954).
30
ES Sashina, AM Bocheck, NP Novoselov, DA Kirichenko Russian J Appl Chem 79, 869–876 (2006).
Information & Authors
Information
Published in
Classifications
Copyright
© 2009. Freely available online through the PNAS open access option.
Submission history
Received: July 10, 2009
Published online: November 17, 2009
Published in issue: November 17, 2009
Keywords
Acknowledgments
We thank Ozgur Sahin for the AFM measurements, Peng Wang/Bruker Optics for IRRAS measurements, and Mike Ekins for illustrations. This work is supported in part by National Science Foundation Small Grants for Exploratory Research Award ECCS-0849603 (to H.A.), a Massachusetts Life Science Center New Investigator Award (to H.A.), Boston University College of Engineering Dean's Catalyst Award (to H.A. and S.E.), and the Boston University Photonics Center and Army Research Laboratory (to H.A. and S.E.). Silk fibroin research is sponsored by in part by the U.S. Army Research Laboratory, U.S. Army Research Office Contract W911 NF-07-1-0618, and Defense Advanced Research Projects Agency-Defense Sciences Office.
Notes
This article is a PNAS Direct Submission.
This article contains supporting information online at www.pnas.org/cgi/content/full/0907459106/DCSupplemental.
Authors
Competing Interests
The authors declare no conflict of interest.
Metrics & Citations
Metrics
Citation statements
Altmetrics
Citations
Cite this article
106 (46) 19227-19232,
Export the article citation data by selecting a format from the list below and clicking Export.
Cited by
Loading...
View Options
View options
PDF format
Download this article as a PDF file
DOWNLOAD PDFLogin options
Check if you have access through your login credentials or your institution to get full access on this article.
Personal login Institutional LoginRecommend to a librarian
Recommend PNAS to a LibrarianPurchase options
Purchase this article to access the full text.