all-optical imaging of gold nanoparticle geometry using ...abstract: we demonstrate the all-optical...

8
All-optical imaging of gold nanoparticle geometry using super- resolution microscopy Citation for published version (APA): Taylor, A., Verhoef, R., Beuwer, M., Wang, Y., & Zijlstra, P. (2018). All-optical imaging of gold nanoparticle geometry using super-resolution microscopy. Journal of Physical Chemistry C, 122(4), 2336-2342. https://doi.org/10.1021/acs.jpcc.7b12473 Document license: CC BY-NC-ND DOI: 10.1021/acs.jpcc.7b12473 Document status and date: Published: 01/02/2018 Document Version: Publisher’s PDF, also known as Version of Record (includes final page, issue and volume numbers) Please check the document version of this publication: • A submitted manuscript is the version of the article upon submission and before peer-review. There can be important differences between the submitted version and the official published version of record. People interested in the research are advised to contact the author for the final version of the publication, or visit the DOI to the publisher's website. • The final author version and the galley proof are versions of the publication after peer review. • The final published version features the final layout of the paper including the volume, issue and page numbers. Link to publication General rights Copyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright owners and it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights. • Users may download and print one copy of any publication from the public portal for the purpose of private study or research. • You may not further distribute the material or use it for any profit-making activity or commercial gain • You may freely distribute the URL identifying the publication in the public portal. If the publication is distributed under the terms of Article 25fa of the Dutch Copyright Act, indicated by the “Taverne” license above, please follow below link for the End User Agreement: www.tue.nl/taverne Take down policy If you believe that this document breaches copyright please contact us at: [email protected] providing details and we will investigate your claim. Download date: 25. Mar. 2020

Upload: others

Post on 19-Mar-2020

7 views

Category:

Documents


0 download

TRANSCRIPT

All-optical imaging of gold nanoparticle geometry using super-resolution microscopyCitation for published version (APA):Taylor, A., Verhoef, R., Beuwer, M., Wang, Y., & Zijlstra, P. (2018). All-optical imaging of gold nanoparticlegeometry using super-resolution microscopy. Journal of Physical Chemistry C, 122(4), 2336-2342.https://doi.org/10.1021/acs.jpcc.7b12473

Document license:CC BY-NC-ND

DOI:10.1021/acs.jpcc.7b12473

Document status and date:Published: 01/02/2018

Document Version:Publisher’s PDF, also known as Version of Record (includes final page, issue and volume numbers)

Please check the document version of this publication:

• A submitted manuscript is the version of the article upon submission and before peer-review. There can beimportant differences between the submitted version and the official published version of record. Peopleinterested in the research are advised to contact the author for the final version of the publication, or visit theDOI to the publisher's website.• The final author version and the galley proof are versions of the publication after peer review.• The final published version features the final layout of the paper including the volume, issue and pagenumbers.Link to publication

General rightsCopyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright ownersand it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights.

• Users may download and print one copy of any publication from the public portal for the purpose of private study or research. • You may not further distribute the material or use it for any profit-making activity or commercial gain • You may freely distribute the URL identifying the publication in the public portal.

If the publication is distributed under the terms of Article 25fa of the Dutch Copyright Act, indicated by the “Taverne” license above, pleasefollow below link for the End User Agreement:www.tue.nl/taverne

Take down policyIf you believe that this document breaches copyright please contact us at:[email protected] details and we will investigate your claim.

Download date: 25. Mar. 2020

All-Optical Imaging of Gold Nanoparticle Geometry Using Super-Resolution MicroscopyAdam Taylor, Rene Verhoef, Michael Beuwer, Yuyang Wang, and Peter Zijlstra*

Molecular Biosensing for Medical Diagnostics, Faculty of Applied Physics, and Institute of Complex Molecular Systems, EindhovenUniversity of Technology, PO Box 513, 5600 MB Eindhoven, The Netherlands

*S Supporting Information

ABSTRACT: We demonstrate the all-optical reconstruction ofgold nanoparticle geometry using super-resolution microscopy.We employ DNA-PAINT to get exquisite control over the(un)binding kinetics by the number of complementary basesand salt concentration, leading to localization accuracies of ∼5nm. We employ a dye with an emission spectrum strongly blue-shifted from the plasmon resonance to minimize mislocalizationdue to plasmon-fluorophore coupling. We correlate the all-optical reconstructions with atomic force microscopy imagesand find that reconstructed dimensions deviate by no more than∼10%. Numerical modeling shows that this deviation isdetermined by the number of events per particle, and the signal-to-background ratio in our measurement. We further findgood agreement between the reconstructed orientation and aspect ratio of the particles and single-particle scatteringspectroscopy. This method may provide an approach to all-optically image the geometry of single particles in confined spacessuch as microfluidic circuits and biological cells, where access with electron beams or tip-based probes is prohibited.

■ INTRODUCTION

As a consequence of their plasmon resonance, metal nano-particles confine incident optical fields to subdiffraction limitedvolumes.1 Concentrating these optical fields enhances the linearand nonlinear optical response of nearby emitters,2,3 whilebinding of biomolecules in these high-field regions results inmodifications of the plasmonic response.4−6 The opticalproperties and performance of metal particles in theseapplications strongly depends on their size and shape.7,8

Nanoparticle geometry is usually studied using methods such asatomic force microscopy (AFM) and electron microscopy(EM). Correlation between optical properties and nanoparticlegeometry now requires colocalization schemes across differenttechniques to allow unambiguous comparisons.9,10 Electronmicroscopy usually requires the sample to be dried and exposedto ultrahigh vacuum, potentially perturbing surface functional-ization. AFM on the other hand requires nanoparticles to bephysically accessible to the tip, prohibiting studies in confinedspaces such as fluidic circuits and biological cells.To overcome the limitations imposed by correlative

techniques, there is a need to all-optically reconstruct thegeometry of plasmonic nanoparticles and their assemblies.Pioneering studies employed PALM,11 ground-state depletionmicroscopy,12−14 or immobilized dyes combined withstochastic optical reconstruction microscopy.15,16 However,resonant coupling between the plasmon and the fluorophoreresulted in a localization bias toward the center of thenanoparticle. Recently the mislocalization of a fluorophorethat is resonantly coupled to a gold nanosphere was

quantified17 to be up to 50 nm, depending on the particle-fluorophore distance. Lim et al.18 and Raab et al.19 performedsimilar experiments finding a fluorophore near a resonantnanoparticle can induce localization errors of up to 90 nm.It was recently shown that this plasmon-fluorophore coupling

can be minimized by choosing a fluorophore with an emissionspectrum strongly blue-shifted from the plasmon resonance andby keeping the fluorophore at a distance of >6 nm from thegold surface to prevent quenching.20−22 Although clever designof the experiments have minimized plasmon−fluorophorecoupling, these approaches employed fluorophores electrostati-cally stuck to a polyelectrolyte coated particle.20 This offerslittle control over the fluorophore density and location, andlimits the localization accuracy to several tens of nanometersdue to bleaching.Here we overcome these limitations by employing DNA-

PAINT (points accumulation for imaging nanoscale top-ography) to all-optically reconstruct gold nanoparticle geom-etry. The unique advantage of this method lies in the exquisitecontrol over the (un)binding kinetics by the number ofcomplementary bases and salt concentration, leading tolocalization accuracies of a few nanometers.23,24 We demon-strate all-optical reconstruction of the geometry of dozens ofnanorods in parallel in a wide-field microscope. We correlatethese super-resolution reconstructions to atomic force micros-copy images and to single-particle white-light spectroscopy. We

Received: December 19, 2017Published: January 3, 2018

Article

pubs.acs.org/JPCCCite This: J. Phys. Chem. C 2018, 122, 2336−2342

© 2018 American Chemical Society 2336 DOI: 10.1021/acs.jpcc.7b12473J. Phys. Chem. C 2018, 122, 2336−2342

This is an open access article published under a Creative Commons Non-Commercial NoDerivative Works (CC-BY-NC-ND) Attribution License, which permits copying andredistribution of the article, and creation of adaptations, all for non-commercial purposes.

find excellent correlation between the orientation and plasmonwavelength obtained from super-resolution microscopy andsingle-particle spectroscopy. Numerical calculations show thatparticle dimensions can be reconstructed with an accuracy of afew nanometers, determined by the number of events perparticle and the photon count per event.

■ EXPERIMENTAL METHODSThe experimental layout is shown in Figure 1a−c, withimmobilized gold nanorods excited in an objective-TIR

configuration. After being spin-coated onto glass coverslips,gold nanorods are functionalized with thiolated dsDNA with asingle-stranded toe-hold, so-called docking strands (Figure 1b).The toe-hold provides binding-sites for fluorophore-labeledimager strands. With the first 20 bases of the docking strandhybridized, a 10 nucleotide toe-hold is rigidly placed away fromthe gold surface which reduces quenching of the emission.20,25

Docking sites are deposited as a mixed monolayer with thiolfunctionalized PEG4 to control docking site density. From theassociation rate of imager stands we deduce that each particleon average contains ∼145 docking strands (see SupportingInformation). Considering the surface area of the particles, eachdocking strand occupies ∼200 nm2, leaving them free to pivotaround the flexible linker attaching the thiol to the dsDNA.This results in a time-averaged particle-fluorophore spacing of∼7.5 nm.The geometry of the nanorods used here is depicted in the

TEM image in Figure 2a, with measured average dimensions of120 nm x 38 nm. A dark-field scattering image of spin-coatedparticles is shown in Figure 2b, illustrating multiple singleparticles can be monitored in parallel. We perform scatteringspectroscopy using hyperspectral imaging (see Supporting

Information), two representative scattering spectra are plottedin Figure 2c. The particles exhibit a longitudinal surfaceplasmon resonance (LSPR) in the near-infrared with single-particle plasmon wavelengths of 790 ± 40 nm. We use the lineshape and line width of the single-particle scattering spectra toidentify clusters of particles,26 evidenced by multiple peaks inthe near-infrared or line widths exceeding 190 meV. Clusters ofparticles were omitted from further analysis.We employ imager strands functionalized with ATTO 532,

its emission spectrum is plotted in dark green in Figure 2c. Theemission is detuned by >150 nm to the blue from thelongitudinal plasmon peak to reduce plasmon-fluorophorecoupling. Fluorescence is excited at 532 nm (bright green line),and collected between 545 and 605 nm (gray shaded region).Although this filter bandwidth suppresses the blue and red tailsof the fluorophore emission and reduces the photon count perevent, it also minimizes spectral overlap between the detectedemission and the transverse (∼520 nm) or longitudinal (>750nm) plasmon resonance.

Figure 1. (a) Immobilized gold nanorods are coated with DNAdocking sites, and excited in an objective-TIR configuration in thepresence of a continuous flow of imager strands. (b) Zoom illustratesdocking site coating across the nanorod surface. (c) Docking-strandsurface chemistry. Docking strands consist of two prehybridized andthiol-terminated strands, 20 and 30 nt in length. This leaves a 10 nttoehold for imager strands to bind. We employed a mixed monolayerof docking strands and thiol-PEG4 to control the docking site density.

Figure 2. (a) TEM image of gold nanorods employed here, with meandimensions 120 nm × 38 nm, mean AR = 3.1. (b) Dark-field scatteringimage of immobilized gold nanorods, recorded at 800 nm. (c, top) Redand blue points show measured scattering spectra of the nanoparticlesindicated in part b, together with Lorentzian fits. The dark green curveshows the emission spectrum of ATTO 532, the light green verticalline indicates the excitation wavelength of 532 nm. Fluorescence iscollected in the spectral range 545−605 nm. (c, bottom) Photonenergy dependent radiative rate enhancement experienced by anemitter 7.5 nm away from the tip (orange curve) or side (blue curve)of a 120 nm × 38 nm nanorod (inset). The enhancement is calculatedfor a single-wavelength emitter and averaged over all dipoleorientations.

The Journal of Physical Chemistry C Article

DOI: 10.1021/acs.jpcc.7b12473J. Phys. Chem. C 2018, 122, 2336−2342

2337

This spectral overlap between a plasmon resonance andfluorophore emission may induce mislocalization due to theplasmonic antenna effect, whereby the emission of thefluorophore is enhanced by coupling to plasmonic modes inthe particle.12−22 This coupling results in modification of theradiative and nonradiative rates of the fluorophore. Mislocaliza-tion originates from enhancements in the radiative rate, ξrad, ofthe complex.27 We have quantified ξrad using the boundaryelement method (BEM).28 For a range of dipole emissionwavelengths (here approximated as a single-wavelengthemitter) the orientation-averaged ξrad values are plotted inFigure 2d. A detailed description of the calculation can befound in the Supporting Information.For an emitter resonant with the LSPR of the nanorod, high

ξrad values of ∼90 and 13 are found for respectively tip and sidebinding positions. This high ξrad may explain previousunderestimation of reconstructed nanorod dimensions using aresonant emitter,14 with largest mislocalization occurring fortip-bound emitters. To minimize fluorophore-plasmon couplingwe choose a fluorophore emitting in the wavelength windowthat minimizes ξrad, occurring on the blue side of the LSPR forboth for tip- and side-bound emitters. We therefore chooseATTO532 as fluorophore, for which we expect a reduction ofξrad of one to 2 orders of magnitude compared to resonantlycoupled emitters.

■ RESULTS AND DISCUSSION

An experimental time trace of fluorescent imager strandsstochastically binding and unbinding to a single gold nanorod isshown in Figure 3a. Each spike corresponds to the photoncounts of a single binding event, integrated over a region ofinterest of 3 × 3 pixels centered on the nanoparticle. Thebackground signal is plotted in red, measured similarly in a 3 ×3 region displaced 4 pixels away from the nanorod emissioncenter. This illustrates that imager strands bind predominantlyto the functionalized nanorods, with minimal nonspecificbinding to the coverslip. We find a distribution of residencetimes that follows a single exponential distribution with a meanbinding time of 3.6 s (see Figure S3). Events longer than 15 sare discarded as multiple binding events (see Figure S3). Thesmall offset between the baseline of the intensity measured onthe nanorod and off the nanorod originates from the one-photon luminescence (1PL) of the gold nanorod.29

An exemplary fluorescent binding event is marked by thegreen star, persisting above threshold for 5 frames. Mergingthese 5 frames together, and subtracting the contribution fromthe weak 1PL (see Supporting Information), returns the totalemission from only the bound imager strand, which is depictedin the inset image. The binding location is then super-resolvedby numerically fitting this intensity distribution with a Gaussianfunction using the maximum likelihood method.30 The bindinglocation is extracted from the Gaussian centroid (red dot in the

Figure 3. (a) Fluorescence time trace showing single imager strands binding to a single gold nanorod. The integrated intensity distribution of theevent marked by the green star is shown in the inset, along with a red dot indicating the fitted emitter position. Scale bar is 200 nm. (b) AFM imageof a single gold nanorod, overlaid with localized events (red dots) and with fitted nanorod outlines as blue solid lines. (c) Correlation betweenmeasured nanorod height from AFM and reconstructed width (blue dots).

The Journal of Physical Chemistry C Article

DOI: 10.1021/acs.jpcc.7b12473J. Phys. Chem. C 2018, 122, 2336−2342

2338

inset), resolved here with a precision a 1/34 of the pixel size or∼ λ/100.Fitting all binding events on each single nanorod results in a

set of spatially distributed points for each particle, which arecorrected for drift.16 For a typical nanoparticle, theselocalizations are plotted as the red dots in Figure 3b. Themean integrated photon count per event is ∼3 × 104 counts,resulting in a mean localization precision of 6 nm (seeSupporting Information). These localizations and reconstructedgeometries are correlated with AFM measurements of the sameparticles. An exemplary AFM image of a single nanorod isshown in Figure 3b, along with localized binding events (reddots). Overlaid as the blue solid line is the numericallydetermined nanorod geometry, obtained by fitting an errorellipse to the localizations to extract out the nanorod length andwidth.31,32 A nanorod shape with this fitted length and width isthen overlaid onto the localizations (see Supporting Informa-tion for full procedure to reconstruct geometry). Goodagreement is observed between the AFM resolved andreconstructed nanorod geometry. Convolution with the AFMtip potentially causes the particle to appear slightly larger in theAFM images. The height in the AFM image is therefore a betterestimate for the nanorod width, because it avoids tipconvolution effects. Here a nanorod height of 33 ± 2 nm

was measured while a width of 37 ± 4 nm was reconstructed.AFM measured heights are correlated with reconstructedwidths for 6 single nanorods in Figure 3c, where we observedeviations less than 5 nm for all particles. We will discuss theaccuracy of the reconstructions in more detail below.In addition to AFM, we correlated the reconstructed

geometry with the measured single-particle scattering spectra.Correlations are presented in Figure 4 for three nanoparticles,with reconstructions shown in Figure 4a, along with thepolarization (Figure 4b) and spectral response (Figure 4c). Inthe top two rows (i and ii), single nanorods are reconstructed(Figure 4a, blue outline) with aspect ratios of 1.7 and 3.6. Thepolarization of the scattered light indicates the angle of thenanorod, which corresponds to within 20° with the angleobtained from the super-resolution reconstruction. As expectedwe find a clearly red-shifted plasmon wavelength for the longeraspect ratio nanorod. In the third row (iii), a different pictureemerges, with localizations arranged in a T-shape, suggesting acluster. The spectral response confirms the presence of a dimerbecause two peaks are resolved. These results indicate theability of DNA-PAINT to resolve the underlying geometry andorientation all-optically, without the requirement of AFM orEM.

Figure 4. (a) Red dots show localized emitter positions on nanorod surface. Blue outlines in parts i and ii are computationally fitted using an errorellipse method, while blue outlines in part iii are fit by hand as a guide to the eye. (b) Polarization dependent scattering response and (c) single-particle scattering spectra of the respective particles, along with fit Lorentzian fits in parts i and ii (blue curves).

The Journal of Physical Chemistry C Article

DOI: 10.1021/acs.jpcc.7b12473J. Phys. Chem. C 2018, 122, 2336−2342

2339

The wide-field detection strategy demonstrated here enablessimultaneous super-resolution microscopy and spectroscopy ofmany particles and particle-assemblies. In Figure 5a, we depictthe correlation between the measured LSPR energy and thecalculated one. The LSPR energy for each nanorod is calculatedusing numerical simulations (BEM) with the reconstructeddimensions as input. Here a positive correlation is observed asexpected, with the 76% of the nanorods having an LSPR energythat deviates by less than 0.1 eV from the calculated LSPR.Apart from errors in reconstructed dimensions, we attribute theresidual spread in LSPR energy to effects of end-cap shape,which can significantly affect plasmon peak position33 but arenot captured in our calculations that assume all nanorods havehemispherical end-caps. We observe a similar picture for theorientation of the particles (Figure 5c), with 88% of themeasured nanorods having a reconstructed angle that closelyresembles the orientation measured using scattering spectros-copy. We further plot the reconstructed lengths, widths, andaspect ratios of 25 nanorods in Figure 5c−e (blue bars)together with the size distribution obtained from TEM (yellowbars). We find that both the mean and standard deviation of thedistribution matches the TEM dimensions to within 10%.Interesting to note is that this agreement between mean

reconstructed and TEM dimensions occurs despite the imagerstrand being bound an average of 7.5 nm from the gold surface.The deviations between the dimensions from TEM and super-resolution microscopy that we observe in Figure 5c−e couldarise from two phenomena: (1) Reconstructions are made frompoints which each have a finite localization precision, and (2)reconstructions are made from a finite number of binding

events. We now estimate the effect of both mechanisms toestablish an achievable “resolution” considering the exper-imental conditions. Here we achieve this by simulating thestochastic reconstruction process of a single nanorod, over anexperimentally relevant range of signal and background levels,for differing numbers of events per particle.For each event, a binding location is randomly sampled from

the 2d projected surface of a 120 nm × 38 nm nanorod. AGaussian point-spread-function plus constant background isthen generated, with the desired signal-to-background ratio(SBR). To this image shotnoise is added. The apparentemission center of this noise-affected signal is then estimatedusing the same procedure we used in the experiments. This isrepeated for the desired number of events on the particle, afterwhich the dimensions are extracted as we do in theexperiments.Three examples of the obtained spatial distributions are

shown in Figure 6a, where the original (simulated) nanorodgeometry is depicted by the black dotted line and thereconstructed geometry by the blue solid line. Cases i-iii depicthow improving both the number of events and SBR improvesthe accuracy of reconstruction. With a low number of eventsand a low SBR, particle dimensions are overestimated and theorientation is poorly reconstructed (case i). Both effects are duepoor sampling the nanorod surface due to the low number ofevents, and the low SBR leading to high localizationuncertainty. Increasing the number of events and the SBRleads to improved estimation of all parameters (case ii and iii).The absolute errors for both length and width are plotted as

heat maps in Figure 6b, reflecting both the over and under-

Figure 5. (a) Correlation between measured LSPR peak energy, and the value calculated using BEM calculations that use the reconstruceddimensions as input.(b) Correlation of measured orientation angle of nanorods with reconstructed angle. (c)-(e) Histogram comparison betweenreconstructed (blue bars) and TEM measured (yellow bars) dimensions of nanorods, in terms of length (c), width (d), and aspect ratio (e).

The Journal of Physical Chemistry C Article

DOI: 10.1021/acs.jpcc.7b12473J. Phys. Chem. C 2018, 122, 2336−2342

2340

estimation observed for case i and ii, and the near perfectreconstruction achieved for case iii. Reconstructing 120 nm ×38 nm nanorods with errors below 10% thus requires >100events and a SBR > 5, while smaller nanorods will have morestringent requirements. Case ii reflects the average number ofevents and SBR in our experiments. The simulations indicatethat the length is underestimated by ∼10 nm, while the width isunderestimated by ∼5 nm due to the limited sampling of theedges of the particle. This explains why we reconstruct particlesizes that match the dimensions from AFM and EM to 5 nm(see Figure 5a), even though the spacing of 7.5 nm betweenparticle and fluorophore should result in overestimation of thedimensions by 15 nm. The reconstruction errors observed hereare thus dominated by the limited number of events perparticle.

■ CONCLUSIONSWe have demonstrated the first method capable of all-opticallyreconstructing the dimensions of single metallic nanoparticlesusing DNA-PAINT, while correlating the particle geometry tosingle-particle spectroscopy. This correlation provides amethod to validate the reconstructions, and a tool to studyoptical properties of single particles without requiring anelectron microscope or atomic force microscopy. Numericalsimulations of our experimental conditions reveal an under-estimation of particle dimensions by 5−10 nm, caused by thelimited sampling of the particle’s edges. This error can bereduced to below 1% with for SBR > 20 and more than 500events per particle. This may be achieved through employing

e.g. a FRET-PAINT approach34 that minimize background andallows one to work at higher imager strand concentration. Thephoton budget can then be maintained while reducing thebinding duration, allowing for more binding events per second,yielding greater statistics and improved reconstructions. Thepresented method may provide an approach to all-opticallyimage the geometry of single particles in confined spaces suchas microfluidics and biological cells, where access with electronbeams or tip-based probes is prohibited.

■ ASSOCIATED CONTENT

*S Supporting InformationThe Supporting Information is available free of charge on theACS Publications website at DOI: 10.1021/acs.jpcc.7b12473.

Detailed description of sample preparation methods,including DNA sequences employed here, and howoptical and AFM experiments were carried out,furthermore, the fitting procedure for events andnanoparticle reconstruction, and electromagnetic simu-lations discussed in more detail, and finally, fluorescencetime traces and reconstructions for additionally nanorods(PDF)

■ AUTHOR INFORMATION

Corresponding Author*[email protected].

ORCIDPeter Zijlstra: 0000-0001-9804-2265NotesThe authors declare no competing financial interest.

■ ACKNOWLEDGMENTS

P.Z. and A.T. acknowledge support from the Foundation forFundamental Research on Matter (FOM), which is financiallysupported by The Netherlands Organization for ScientificResearch (NWO). P.Z. and Y.W. acknowledge financial supportfrom The Netherlands Organization for Scientific Research(NWO VIDI). Solliance and the Dutch province of NoordBrabant are acknowledged for funding the TEM facility. Wethank Ralf Jungmann for useful discussions.

■ REFERENCES(1) Maier, S. A. Plasmonics: Fundamentals and Applications; SpringerScience & Business Media: 2007.(2) Khatua, S.; Paulo, P. M. R.; Yuan, H.; Gupta, A.; Zijlstra, P.; Orrit,M. Resonant Plasmonic Enhancement of Single-Molecule Fluores-cence by Individual Gold Nanorods. ACS Nano 2014, 8, 4440−4449.(3) Simoncelli, S.; Roller, E. M.; Urban, P.; Schreiber, R.; Turberfield,A. J.; Liedl, T.; Lohmuller, T. Quantitative Single-Molecule Surface-Enhanced Raman Scattering by Optothermal Tuning of DNAOrigami-Assembled Plasmonic Nanoantennas. ACS Nano 2016, 10,9809−9815.(4) Beuwer, M. A.; Prins, M. W. J.; Zijlstra, P. Stochastic ProteinInteractions Monitored by Hundreds of Single-Molecule PlasmonicBiosensors. Nano Lett. 2015, 15, 3507−3511.(5) Baaske, M. D.; Vollmer, F. Optical Observation of Single AtomicIons Interacting with Plasmonic Nanorods in Aqueous Solution. Nat.Photonics 2016, 10, 733−739.(6) Ament, I.; Prasad, J.; Henkel, A.; Schmachtel, S.; Sonnichsen, C.Single Unlabeled Protein Detection on Individual Plasmonic Nano-particles. Nano Lett. 2012, 12, 1092−1095.

Figure 6. (a) Numerical simulation of the reconstruction processtaking into account the limited signal-to-background ratio (SBR) andlimited number of events per particle. Red dots are the positions atwhich simulated events were localized, overlaid with a fitted nanorodgeometry (blue curve). The back dashed curve illustrates the simulated(input) nanorod geometry. The simulation parameters for the threecases are indicated in the contour plots below. (b) Heat map showingthe discrepancy between the true nanorod dimensions and thereconstructed dimensions for different SBR and events per particle.

The Journal of Physical Chemistry C Article

DOI: 10.1021/acs.jpcc.7b12473J. Phys. Chem. C 2018, 122, 2336−2342

2341

(7) Chen, H.; Kou, X.; Yang, Z.; Ni, W.; Wang, J. Shape- and Size-Dependent Refractive Index Sensitivity of Gold Nanoparticles.Langmuir 2008, 24, 5233−5237.(8) Nusz, G. J.; Curry, A. C.; Marinakos, S. M.; Wax, A.; Chilkoti, A.Rational Selection of Gold Nanorod Geometry for Label-FreePlasmonic Biosensors. ACS Nano 2009, 3, 795−806.(9) Jin, R.; Jureller, J. E.; Kim, H. Y.; Scherer, N. F. CorrelatingSecond Harmonic Optical Responses of Single Ag Nanoparticles withMorphology. J. Am. Chem. Soc. 2005, 127, 12482−12483.(10) Siddiquee, A. M.; Taylor, A. B.; Syed, S.; Lim, G. H.; Lim, B.;Chon, J. W. M. Measurement of Plasmon-Mediated Two-PhotonLuminescence Action Cross Sections of Single Gold Bipyramids,Dumbbells, and Hemispherically Capped Cylindrical Nanorods. J.Phys. Chem. C 2015, 119, 28536−28543.(11) Lin, H.; Centeno, S. P.; Su, L.; Kenens, B.; Rocha, S.; Sliwa, M.;Hofkens, J.; Uji-i, H. Mapping of Surface-Enhanced Fluorescence onMetal Nanoparticles Using Super-Resolution Photoactivation Local-ization Microscopy. ChemPhysChem 2012, 13, 973−981.(12) Blythe, K. L.; Mayer, K. M.; Weber, M. L.; Willets, K. A. GroundState Depletion Microscopy for Imaging Interactions Between GoldNanowires and Fluorophore-Labeled Ligands. Phys. Chem. Chem. Phys.2013, 15, 4136−4145.(13) Blythe, K. L.; Titus, E. J.; Willets, K. A. Triplet-State-MediatedSuper-Resolution Imaging of Fluorophore-Labeled Gold Nanorods.ChemPhysChem 2014, 15, 784−793.(14) Blythe, K. L.; Willets, K. A. Super-Resolution Imaging ofFluorophore-Labeled DNA Bound to Gold Nanoparticles: A Single-Molecule, Single-Particle Approach. J. Phys. Chem. C 2016, 120, 803−815.(15) Wertz, E.; Isaacoff, B. P.; Flynn, J. D.; Biteen, J. S. Single-Molecule Super-Resolution Microscopy Reveals How Light Couplesto a Plasmonic Nanoantenna on the Nanometer Scale. Nano Lett.2015, 15, 2662−2670.(16) Su, L.; Yuan, H.; Lu, G.; Rocha, S.; Orrit, M.; Hofkens, J.; Uji-I,H. Super-Resolution Localization and Defocused FluorescenceMicroscopy on Resonantly Coupled Single-Molecule, Single-NanorodHybrids. ACS Nano 2016, 10, 2455−2466.(17) Fu, B.; Isaacoff, B. P.; Biteen, J. S. Super-Resolving the ActualPosition of Single Fluorescent Molecules Coupled to a PlasmonicNanoantenna. ACS Nano 2017, 11, 8978−8987.(18) Lim, K.; Ropp, C.; Barik, S.; Fourkas, J.; Shapiro, B.; Waks, E.Nanostructure-Induced Distortion in Single-Emitter Microscopy.Nano Lett. 2016, 16, 5415−5419.(19) Raab, M.; Vietz, C.; Stefani, F. D.; Acuna, G. P.; Tinnefeld, P.Shifting Molecular Localization by Plasmonic Coupling in a Single-Molecule Mirage. Nat. Commun. 2017, 8, 13966.(20) De Silva Indrasekara, A. S.; Shuang, B.; Hollenhorst, F.; Hoener,B. S.; Hoggard, A.; Chen, S.; Villarreal, E.; Cai, Y.-Y.; Kisley, L.; Derry,P. J.; et al. Optimization of Spectral and Spatial Conditions to ImproveSuper- Resolution Imaging of Plasmonic Nanoparticles. J. Phys. Chem.Lett. 2017, 8, 299−306.(21) Heaps, C. W.; Schatz, G. C. Modeling Super-Resolution SERSUsing a T-Matrix Method to Elucidate Molecule-NanoparticleCoupling and the Origins of Localization Errors. J. Chem. Phys.2017, 146, 224201.(22) Mack, D. L.; Cortes, E.; Giannini, V.; Torok, P.; Roschuk, T.;Maier, S. A. Decoupling Absorption and Emission Processes in Super-Resolution Localization of Emitters in a Plasmonic Hotspot. Nat.Commun. 2017, 8, 14513.(23) Jungmann, R.; Steinhauer, C.; Scheible, M.; Kuzyk, A.;Tinnefeld, P.; Simmel, F. C. Single-Molecule Kinetics and Super-Resolution Microscopy by Fluorescence Imaging of Transient Bindingon DNA Origami. Nano Lett. 2010, 10, 4756−4761.(24) Schnitzbauer, J.; Strauss, M. T.; Schlichthaerle, T.; Schueder, F.;Jungmann, R. Super-Resolution Microscopy with DNA-PAINT. Nat.Protoc. 2017, 12, 1198−1228.(25) Holzmeister, P.; Pibiri, E.; Schmied, J. J.; Sen, T.; Acuna, G. P.;Tinnefeld, P. Quantum Yield and Excitation Rate of Single MoleculesClose to Metallic Nanostructures. Nat. Commun. 2014, 5, 5356.

(26) Zijlstra, P.; Orrit, M. Single Metal Nanoparticles: OpticalDetection, Spectroscopy and Applications. Rep. Prog. Phys. 2011, 74,106401.(27) Mertens, H.; Polman, A. Strong Luminescence Quantum-Efficiency Enhancement Near Prolate Metal Nanoparticles: Dipolarversus Higher-Order Modes. J. Appl. Phys. 2009, 105, 044302.(28) Hohenester, U.; Trugler, A. MNPBEM - A Matlab Toolbox forthe Simulation of Plasmonic Nanoparticles. Comput. Phys. Commun.2012, 183, 370−381.(29) Yorulmaz, M.; Khatua, S.; Zijlstra, P.; Gaiduk, A.; Orrit, M.Luminescence Quantum Yield of Single Gold Nanorods. Nano Lett.2012, 12, 4385−4391.(30) Mortensen, K. I.; Churchman, S. L.; Spudich, J. A.; Flyvbjerg, H.Optimized Localization Analysis for Single-Molecule Tracking andSuper-Resolution Microscopy. Nat. Methods 2010, 7, 377−381.(31) Bock, R. K.; Krischer, W. The Data Analysis Briefbook. In TheData Analysis BriefBook; Springer: 1998; pp 1−181.(32) Johnson, A. J. Error Ellipse MATLAB Program. 2015, http://nl.mathworks.com/matlabcentral/fileexchange. (accessed April 2016).(33) Prescott, S. W.; Mulvaney, P. Gold Nanorod Extinction Spectra.J. Appl. Phys. 2006, 99, 123504.(34) Auer, A.; Strauss, M. T.; Schlichthaerle, T.; Jungmann, R. Fast,Background-Free DNA-PAINT Imaging Using FRET-Based Probes.Nano Lett. 2017, 17, 6428−6434.

The Journal of Physical Chemistry C Article

DOI: 10.1021/acs.jpcc.7b12473J. Phys. Chem. C 2018, 122, 2336−2342

2342