Levitated Plasmonic Nanoantennas in an Aqueous Environment

Jul 11, 2017 - Max Planck Institute for the Science of Light, Staudt-Straße 2, 91058 Erlangen, Germany. ‡ Department of Physics, Friedrich Alexande...
1 downloads 8 Views 849KB Size
Subscriber access provided by UNIV OF NEWCASTLE

Article

Levitated Plasmonic Nanoantennas in an Aqueous Environment Yazgan Tuna, Ji Tae Kim, Hsuan-Wei Liu, and Vahid Sandoghdar ACS Nano, Just Accepted Manuscript • DOI: 10.1021/acsnano.7b03310 • Publication Date (Web): 11 Jul 2017 Downloaded from http://pubs.acs.org on July 12, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Nano is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 18

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

Levitated Plasmonic Nanoantennas in an Aqueous Environment Yazgan Tuna,†,‡,k Ji Tae Kim,¶,§,k Hsuan-Wei Liu,†,‡ and Vahid Sandoghdar∗,†,‡ †Max Planck Institute for the Science of Light, Staudt-Str 2, 91058 Erlangen, Germany ‡Department of Physics, Friedrich Alexander University of Erlangen-Nürnberg, 91058 Erlangen, Germany ¶Max Planck Institute for the Science of Light, Staudt-Str 2, 91058 wErlangen, Germany §Present address: Department of Mechanical Engineering, The University of Hong Kong, Pokfulam Road, Hong Kong, China kThese authors contributed equally to this work. E-mail: [email protected] Phone: +49 9131 7133 300. Fax: +49 9131 7133 309 Abstract We report on the manipulation of a plasmonic nanoantenna in an aqueous solution using an electrostatic trap created between a glass nanopipette and a substrate. By scanning a trapped gold nanosphere in the near field of a single colloidal quantum dot embedded under the substrate surface, we demonstrate about eight-fold fluorescence enhancement over a lateral full width at half-maximum of about 45 nm. We analyze our results with the predictions of numerical electromagnetic simulations under consideration of the electrostatic free energy in the trap. Our approach could find applications in a number of experiments, where plasmonic effects are employed at liquid-solid interfaces.

1

ACS Paragon Plus Environment

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Keywords plasmonics, plasmofluidics, electrostatic trapping, fluorescence enhancement, quantum dot Plasmonic nanoantennas have attracted much attention in the past decade both from the fundamental and technological points of view because of their potential in a number of areas such as the enhancement of the Raman scattering cross section 1 or the enhancement of the absorption cross section and the modification of the spontaneous emission rate. 2 While the great majority of the existing efforts have concerned dry conditions, plasmonic effects have also begun to be used in the liquid phase, giving rise to the emergence of plasmofluidics. 3–7 A particularly promising area of application for plasmofluidics is in biophysics and biosensing, 8–11 where biological nanoparticles and macromolecules diffuse or undergo directed transport. 12,13 A critical and central feature of plasmonic interactions is their near-field character, which demands subwavelength separations between a nanoantenna and the substance to be affected. 2 Various methods such as self-assembly, 14 lithographic nanofabrication, 15–17 random distribution of emitters, 18–20 chemical synthesis, 21,22 and use of scanning probe technology for nano-positioning 23–26 have been employed to achieve nanometer spacing between emitters and antennas. Among these, nano-positioning with scanning probes has delivered the most quantitative results because it allows one to study the very same emitter at different orientations and separations from the nanoantenna. Scanning a nanoantenna also allows the controlled realization of apertureless near-field microscopy 27–29 in a complementary fashion to conventional scanning near-field optical microscopy (SNOM) using subwavelength apertures. 30,31 Indeed, spatial resolution down to the order of 20 nm has been achieved using a simple spherical antenna geometry. 32 One can imagine different approaches for the manipulation of a plasmonic nanoantenna close to an aqueous interface. Attaching an antenna 23 or fabricating it 33–35 at the end of a fiber tip is one possibility. Using optical tweezers 36,37 can be another option although trapping of very small metallic nanoparticles 38,39 within a few tens of nanometers comes at 2

ACS Paragon Plus Environment

Page 2 of 18

Page 3 of 18

the cost of prohibitive heating. Recently, we reported on trapping of small gold nanoparticles at the junction between a glass nanopipette and a glass substrate in water 40 based on a geometry-induced electrostatic potential. 41 Here, we show that this technique can yield a considerable plasmonic enhancement of fluorescence from a single semiconductor quantum dot (qdot) in water within a lateral range well below 100 nm.

Results and discussion Figure 1a shows the schematics of our experiment. A gold nanoparticle (GNP) is trapped and manipulated in the near field of individual semiconductor CdSe/CdS core/thick shell qdots, 42,43 which were embedded at the upper interface of a thin polymer film 44,45 at large lateral separations of several tens of micrometers. The GNPs used in this experiment had a diameter of 80 nm and were trapped out of the aqueous solution at the end of a glass nanopipette with surface charge density ∼1016 em−2 [Ref. 46] and an opening of about 100 nm. A three-dimensional piezo-electric stage allowed us to move and position the trapped particle at will. 40 z

a nanopipette

y

b x

x

c

water GNP

h

z (nm)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

30 15 0 -15

qdot 532 nm excitation

-30 -800

-400 0 400 Lateral position (nm)

800

Figure 1: a, Schematic illustration of the experimental setup. An 80 nm GNP electrostatically trapped between a scannable nanopipette and a substrate is scanned across a CdSe/CdS giant qdot embedded in PMMA layer in an aqueous bath. The fluorescence signal of individual qdots is collected by a total internal reflection microscope. b, Interferometric scattering (iSCAT) detection images showing a guided single GNP by lateral scanning of a nanopipette placed at height h ∼ 140 nm. c, Axial variations of the GNP in the trap during the scan trajectory of (b). To visualize the GNPs, we performed wide-field interferometric scattering microscopy 3

ACS Paragon Plus Environment

ACS Nano

(iSCAT) 47,48 at a frame rate of 1.9 KHz using a CMOS camera at wavelength λ = 532 nm. Figure 1b displays two iSCAT images of a GNP before and after it was laterally displaced at a speed of 250 nm s−1 by 1.4 µm in a trap with stiffness 10∼100 fN/nm (calculated with equipartition theorem). 40 As presented by Figure 1c , the interferometric nature of the iSCAT contrast also allows one to monitor the axial height variations of a GNP 41,49 without the need for an external shear-force control. 50 We note that the electrostatic repulsion in the trap protects the GNP from coming in contact with the PMMA substrate since it carries a

d

x = -400nm

8

a

10

surface charge density of about 6.6×1015 em−2 in deionized water. 51

12 8

c

6 4

0 nm

Intensity (a.u.)

b

FWHM = 45 nm

2

16

4 400 nm

1

0

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 18

-400

-200 0 200 400 Lateral displacement (nm)

Figure 2: a-c, A series of CCD camera images of a CdSe/CdS qdot with different lateral positions of a trapped 80 nm GNP a, -400 nm, b, 0 nm, and c, 400 nm at a constant pipettesubstrate gap, h ∼95 nm. The inset shows the unsaturated image of qdot confirming a donut shape. Scale bar is 500 nm. d, Dependency of the fluorescence signal on the lateral position of a GNP. The fluorescence enhancement about 8 and the FWHM = 45 nm are obtained. To excite the qdots, we changed the laser beam path to achieve total internal reflection through the microscope objective. The fluorescence from the qdots was collected with the same objective and imaged on a sensitive CCD camera. More details of the experimental procedure can be found in the M ethods section. Figure 2a shows the fluorescence image of an unperturbed qdot whereas in Figure 2b we present the image of the same qdot when a trapped GNP was positioned above it at a pipette-substrate gap of h ∼ 95 nm. Figure 2c displays the qdot fluorescence after the particle is scanned away. The intensity minimum in the middle of the image in Figure 2b (see inset) indicates that the emission dipole associated with the antenna-qdot composite lies dominantly along the normal to the sample plane. 52,53 In Figure 2d, we plot the integrated signal under the Gaussian fits of the fluorescence 4

ACS Paragon Plus Environment

Page 5 of 18

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

camera images from a single qdot as a GNP was laterally scanned across it. We find that the fluorescence signal was increased by about 8 times over a profile with a full width at half-maximum (FWHM) of 45 nm, confirming a near-field interaction. We attribute the asymmetric shape of the observed peak in Figure 2d to the variation of the relative orientation of the qdot’s dipole moment with respect to the GNP’s radius at the point of closest contact during the scan as illustrated in Figure 3a. 24,54 To gain some quantitative insight into the observed effects, we performed finite-difference time-domain (FDTD) simulations, where we took into account the antenna-induced modifications of the radiative and nonradiative decay rates as well as changes in the excitation rate and collection efficiency (see Supplementary Materials). Figure 3b shows the outcome for the overall enhancement of fluorescence for emitters of different initial quantum efficiencies. We see that the behavior observed in Figure 2d is well reproduced for an emitter with an internal quantum efficiency of 40% placed at 15 nm under the surface with a tilt angle of 25◦ with respect to the z-axis. It is important to note, however, that this agreement is only semi-quantitative because qdots are known to possess two emission dipole moments 55,56 and that we do not have a precise knowledge of the depth of the qdot under the GNP although the finite size of the qdot (about 15-20 nm in diameter 42,43 ) puts a lower bound on this quantity.

5

ACS Paragon Plus Environment

ACS Nano

a

d 70 intensity (a.u.)

FWHM (nm)

75

65 60 55 5 10 15 20 25 Antenna-substrate gap, g (nm)

8

i = 40% 60% 80% 100%

6 4 2 0

Fluorescence enhancement factor

10

10 i = 40% 60% 80% 100%

8 intensity (a.u.)

c

b Fluorescence enhancement factor

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 18

6 4 2 0

5 10 15 20 25 Vertical displacement (nm)

-100 -50 0 50 100 Lateral displacement (nm)

Figure 3: a, Schematic illustration of the asymmetry in the relative orientation of the emitter dipole moment with respect to the radial normal to the antenna surface. b, Total enhancement of the fluorescence as function of the lateral displacement of a GNP with 5 nm steps. c, The calculated enhancement factor versus the axial displacement of the GNP at the peak position (−15 nm). d, FWHM of the lateral scan profile (see (b)) for different antenna-substrate gap values g (see (a)). The results of the simulations in Figure 3c show the expected drop of the fluorescence signal as a function of the axial separation of the GNP. To verify this behaviour experimentally, we adjusted the trap height h (see the M ethods section for experimental details), and scanned the antenna at different antenna-substrate gaps g which is calculated from h according to the Helmholtz free energy as shown in Figure 5. In Figure 4a-f, we present the fluorescence signal from a single qdot for h varied between 115 nm and 90 nm at 5 nm steps. We note in passing that contrary to conventional shear-force controlled SNOM tip, our method does not rely on a hard contact of the tip with the sample surface. We also remark that the qdot bleached in the latter part of the measurement presented in Figure 4f. The signal recorded after photobleaching (see red symbols) provides a convenient measure for the small fluorescence background, which was subtracted from all data for estimating the enhancement factors.

6

ACS Paragon Plus Environment

a

c

12 g = 13 nm

10

g

e 9.5 nm

6.5 nm

8 6 4 2 0

b

d

12 11 nm

10

10 8 6 4 2

h

f 8 nm

5 nm

140 FWHM (nm)

Intensity (a.u.)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

Fluorescent enhancement factor

Page 7 of 18

8 6 4 2

120 100 80 60 40

0 -200 -100

0

100 200

-200 -100 0 100 200 Lateral displacement (nm)

-200 -100

0

100 200

20

4 6 8 10 12 14 Antenna-substrate gap, g (nm)

Figure 4: Studies at different nanopipette heights. a-f, Fluorescence signal of a qdot as a function of the lateral position of a trapped GNP at different antenna-substrate gaps g. Scanning speed was 0.25 µm s−1 and the length of each scan was 1.25 µm. The fluorescence signal was monitored at a frequency of 10 Hz. The red symbols in (f) indicate the region after the qdot was bleached. g, Enhancement factor of a qdot fluorescence as a function of g. h, FWHM of the Gaussian fitted to each scan. To examine the relationship between h and g, we calculated the Helmholtz free energy (F ) in the system by summing the electrostatic field energies and entropies over all charges in three dimensions. 57 Here, we considered deionized water with 0.04 mM ionic strength, a 80 nm GNP (σgnp ∼7×1015 em−2 ), a glass nanopipette opening of 100 nm in diameter (σglass ∼ 1016 em−2 ), and a PMMA substrate (σP M M A ∼ 6.6×1015 em−2 ). Figure 5a displays the profiles of the calculated free energy difference (∆F ) as a function of g for traps with different h as indicated by the color codes in each curve. As shown in Figure 5b, g at the minimum free energy (∆Fmin ) varies nearly linearly with h. Using the outcome of Figure 5, we can now analyze the enhancement factor and the FWHM of the scan profiles obtained in Figure 4a-f in Figure 4g, h in terms of g. We find that the general trend agrees well with the simulated data in Figure 3c. However, here we observe a steeper dependence on the emitter-antenna gap. We attribute this to the weaker trap stiffness at larger h and, thus, higher fluctuations as indicated by the simulated trap potentials in Figure 5a.

7

ACS Paragon Plus Environment

ACS Nano

130

50

120

40 110

30 20

100

10

90

0 80

0

10 20 30 Antenna-substrate gap, g

40

b 18 g at Fmin (nm)

60

Pipette-substrate gap, h (nm)

a

Fmin (kBT)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 18

14 10 6 80

100 120 h (nm)

140

Figure 5: a, Dependence of the trapping free energy, ∆F , on particle-substrate gaps, g for different pipette-substrate gaps, h, ranging from 80 to 130 nm. b, The extracted value of g at the minima of the trapping potentials, ∆Fmin , as a function of the pipette-substrate gap, h.

Conclusion In this work, we demonstrated the application of an electrostatically-confined gold nanoantenna for plasmonic enhancement of fluorescence in aqueous environments. While we demonstrated this for stationary semiconductor qdots embedded at an interface, our experimental platform can also be used for diffusing fluorophores, e.g. when they act as label for proteins or lipids in biomembranes. Adaptation of the trap geometry to support asymmetric antennas such as nanorods 58 will also allow rotational diffusion studies.

Methods Sample preparation: Glass nanopipettes with about ∼100 nm aperture diameter and optimized taper shapes were fabricated by a multi-step pulling process using a laser pipette puller (P-2000, Sutter Instrument). The nanopipettes were cleaned by standard rinsing and O2 plasma process. Gold nanospheres with diameter of 80 nm (British Biocell International) were centrifuged and resuspended in deionized water (18 MΩcm−1 ) twice to remove traces of salt or other contaminants. These nanoparticles carry charge 59 of the order of ∼ 100 e. Topography-free samples of giant CdSe/CdS core/thick shell qdots 42,43 embedded in PMMA film were fabri8

ACS Paragon Plus Environment

Page 9 of 18

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

cated as follows: The CdSe/CdS quantum dots were deposited on a cleaned cover glass by spin coating at 2000 rpm for 60 s. In addition, the quantum dots were covered with a PMMA film of 800 nm thickness by spin coating at 2000 rpm for 60 s. To place the qdots at the PMMA-air interface, the resulting film was floated in water, subsequently flipped and placed on a glass substrate. 45 A customized liquid chamber on a scanning stage and an inverted microscope was used for trapping and nanoantenna experiment.

Scanning the nanopipette: The position of the glass nanopipette was controlled with 3-axis piezo scanner (P-620.2CD, P-620.ZCD, Physik Instrumente). The pipette-substrate gaps were precisely controlled via iSCAT detection. The contact point between the nanopipette and the substrate was detected by approaching the nanopipette to the surface at 10 nm steps until the interference pattern created by the light scattered of the nanopipette and the light reflected from the substrate no longer changes. Then, we pulled back the pipette and set the initial gap of 120 nm. The stepwise approach of the pipette during fluorescence imaging was 5 nm. The speed and range of lateral scanning were typically 250 nm s−1 and 1.25 µm.

Interferometric scattering detection and fluorescence imaging: The detection of gold nanoparticles (80 nm) and CdSe/CdS qdots was performed using an inverted optical microscope designed for interferometric scattering detection (iSCAT) and total internal reflection fluorescence imaging. For iSCAT, a 10 mW CW laser at λ = 532 nm illuminated the sample using an oil-immersion microscope objective (100x, NA 1.4, Olympus). The light scattered from the GNP and reflected from the substrate were collected by the objective and imaged on a CMOS camera (MV-D1024-160-CL-12, PhotonFocus). The qdots were excited under total internal reflection illumination and their fluorescence 9

ACS Paragon Plus Environment

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

was collected using the same objective, a 600 nm long pass filter, and a CCD camera (Andor Luca) with 10 frames per second.

Data analysis: CCD images were cut to create 615 nm x 615 nm box around localised qdot in the absence of the GNP. Then the fluorescence image was fitted to a 2D Gaussian for each image separately and the fit function was integrated over a range of 5µm × 5µm. The resulting total intensity values were plotted as in Figure 4a-f. Average background values of these fits were taken as a reference and the enhancement factor was calculated by dividing the peak values to the reference.

Acknowledgement This work was financed by the Max Planck Society and an Alexander von Humboldt professorship. We thank S. Götzinger, K. Matsuzaki, X.-L. Chu and T. Utikal for experimental support. Thick-shell qdots were synthesized at the Center for Integrated Nanotechnologies, an Office of Science User Facility operated for the U.S. Department of Energy (DOE) Office of Science by Los Alamos National Laboratory (Contract DE-AC52-06NA25396) and Sandia National Laboratories (Contract DE-AC04-94AL85000).

Supporting Information Available Calculations of the electrostatic potential landscape via Helmholtz free energy and Comsol Multiphysics 5.2a simulation details. Discussion about the calculation of enhancement factor. Finite-Difference Time-Domain (FDTD) simulation results with parametric scan. This material is available free of charge via the Internet at http://pubs.acs.org/.

10

ACS Paragon Plus Environment

Page 10 of 18

Page 11 of 18

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

References 1. Stiles, P. L.; Dieringer, J. A.; Shah, N. C.; Duyne, R. P. V. Surface-Enhanced Raman Spectroscopy. Annu. Rev. Anal. Chem. 2008, 1, 601–626. 2. Sandoghdar, V.; Agio, M.; Chen, X.-W.; Götzinger, S.; Lee, K.-G. In Optical Antennas; Agio, M., Alú, A., Eds.; Cambridge University Press, 2012; Chapter Antennas, Quantum Optics and Near-Field Microscopy. 3. Donner, J. S.; Baffou, G.; McCloskey, D.; Quidant, R. Plasmon-Assisted Optofluidics. ACS Nano 2011, 5, 5457–5462. 4. Punj, D.; Mivelle, M.; Moparthi, S. B.; van Zanten, T. S.; Rigneault, H.; van Hulst, N. F.; García-Parajó, M. F.; Jérôme Wenger, J. A Plasmonic ’Antenna-in-Box’ Platform for Enhanced Single-Molecule Analysis at Micromolar Concentrations. Nat. Nanotechnol. 2013, 8, 512–516. 5. Zhao, C.; Liu, Y.; Zhao, Y.; Fang, N.; Jun Huang, T. A Reconfigurable Plasmofluidic Lens. Nat. Commun. 2013, 4, 2305. 6. Ku, B.; Shin, J.-S.; Kwon, M.-S. Experimental Investigation of Plasmofluidic Waveguides. Appl. Phys. Lett. 2015, 107, 201104. 7. Wang, F.; Karan, N. S.; Nguyen, H. M.; Ghosh, Y.; Sheehan, C. J.; Hollingsworth, J. A.; Htoon, H. Correlated Structural-Optical Study of Aingle Nanocrystals in a Gap-Bar Antenna: Effects of Plasmonics on Excitonic Recombination Pathways. Nanoscale 2015, 7, 9387. 8. Anker, J. N.; Hall, W. P.; Shah, L. O.; Zhao, N. C.; J. Van Duyne, R. P. Biosensing with Plasmonic Nanosensors. Nat. Mater. 2008, 7, 442–453. 9. Garcia-Parajo, M. F. Optical Antennas Focus in on Biology. Nat. Photonics 2008, 2, 201–203. 11

ACS Paragon Plus Environment

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

10. Höppener, C.; Novotny, L. Antenna-Based Optical Imaging of Single Ca2+ Transmembrane Proteins in Liquids. Nano Lett. 2008, 8, 642–646. 11. Limaj, O.; Etezadi, D.; Wittenberg, N. J.; Rodrigo, D.; Yoo, D.; Oh, S.-H.; Altug, H. Infrared Plasmonic Biosensor for Real-Time and Label-Free Monitoring of Lipid Membranes. Nano Lett. 2016, 16, 1502–1508. 12. Simons, K.; Gerl, M. J. Revitalizing Membrane Rafts: New Tools and Insights. Nat. Rev. Mol. Cell Biol. 2010, 11, 688–699. 13. Spindler, S.; Ehrig, J.; König, K.; Nowak, T.; Piliarik, M.; Stein, H. E.; Taylor, R. W.; Garanger, E.; Lecommandoux, S.; Alves, I. D.; Sandoghdar, V. Visualization of Lipids and Proteins at High Spatial and Temporal Resolution via Interferometric Scattering (iSCAT) Microscopy. J. Phys. D: Appl. Phys. 2016, 49, 349601. 14. Acuna, G. P.; Möller, F. M.; Holzmeister, P.; Beater, S.; Lalkens, B.; Tinnefeld, P. Fluorescence Enhancement at Docking Sites of DNA-Directed Self-Assembled Nanoantennas. Science 2012, 338, 506–510. 15. Curto, A. G.; Volpe, G.; Taminiau, T. H.; Kreuzer, M. P.; Quidant, R.; van Hulst, N. F. Unidirectional Emission of a Quantum Dot Coupled to a Nanoantenna. Science 2010, 329, 930–933. 16. Belacel, C.; Habert, B.; Bigourdan, F.; Marquier, F.; Hugonin, J.-P.; de Vasconcellos, S. M.; Lafosse, X.; Coolen, L.; Schwob, C.; Javaux, C.; Dubertret, B.; Greffet, J.J.; Senellart, P.; Maitre, A. Controlling Spontaneous Emission with Plasmonic Optical Patch Antennas. Nano Lett. 2013, 13, 1516–1521. 17. Meixner, A. J.; Jäger, R.; Jäger, S.; Bräuer, A.; Scherzinger, K.; Fulmes, J.; zur Oven Krockhaus, S.; Gollmer, D. A.; Kern, D. P.; Fleischer, M. Coupling Single Quantum Dots to Plasmonic Nanocones: Optical Properties. Faraday Discuss. 2015, 184, 321–337. 12

ACS Paragon Plus Environment

Page 12 of 18

Page 13 of 18

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

18. Song, J.-H.; Atay, T.; Shi, S.; Urabe, H.; Nurmikko, A. V. Large Enhancement of Fluorescence Efficiency from CdSe/ZnS Quantum Dots Induced by Resonant Coupling to Spatially Controlled Surface Plasmons. Nano Lett. 2005, 5, 1557–1561. 19. Kinkhabwala, A.; Yu, Z.; Fan, S.; Avlasevich, Y.; Müllen, K.; Moerner, W. E. Large Single-Molecule Fluorescence Enhancements Produced by a Bowtie Nanoantenna. Nat. Photonics 2009, 3, 654–657. 20. Zengin, G.; Johansson, G.; Johansson, P.; Antosiewicz, T. J.; Käll, M.; Shegai, T. Approaching the Strong Coupling Limit in Single Plasmonic Nanorods Interacting with J-Aggregates. Sci. Rep. 2013, 3, 3074. 21. Seelig, J.; Leslie, K.; Renn, A.; Kuhn, S.; Jacobsen, V.; van de Corput, M.; Wyman, C.; Sandoghdar, V. Nanoparticle-Induced Fluorescence Lifetime Modification as Nanoscopic Ruler: Demonstration at the Single Molecule Level. Nano Lett. 2007, 7, 685–689. 22. Ji, B.; Giovanelli, E.; Habert, B.; Spinicelli, P.; Nasilowski, M.; Xu, X.; Lequeux, N.; Hugonin, J.-P.; Marquier, F.; Greffet, J.-J.; Dubertret, B. Non-Blinking Quantum Dot with a Plasmonic Nanoshell Resonator. Nat. Nanotechnol. 2015, 10, 170–175. 23. Kalkbrenner, T.; Ramstein, M.; Mlynek, J.; Sandoghdar, V. A Single Gold Particle as a Probe for Apertureless Scanning Near-Field Optical Microscopy. J. Microscopy 2001, 202, 72–76. 24. Kühn, S.; Håkanson, U.; Rogobete, L.; Sandoghdar, V. Enhancement of Single-Molecule Fluorescence Using a Gold Nanoparticle as an Optical Nanoantenna. Phys. Rev. Lett. 2006, 97, 017402. 25. Anger, P.; Bharadwaj, P.; Novotny, L. Enhancement and Quenching of Single-Molecule Fluorescence. Phys. Rev. Lett. 2006, 96, 113002–1–4.

13

ACS Paragon Plus Environment

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

26. Matsuzaki, K.; Vassant, S.; Liu, H.-W.; Dutschke, A.; Hoffmann, B.; Chen, X.; Christiansen, S.; Buck, M. R.; Hollingsworth, J. A.; Götzinger, S.; Sandoghdar, V. Strong Plasmonic Enhancement of Biexciton Emission: Controlled Coupling of a Single Quantum Dot to a Gold Nanocone Antenna. Sci. Rep. 2017, 7, 42307. 27. Zenhausern, F.; Oboyle, M.; Wickramasinghe, H. Apertureless Near-Field Optical Microscope. Appl. Phys. Lett. 1994, 65, 1623–1625. 28. Gleyzes, P.; Boccara, A. C.; Bachelot, R. Near Field Optical Microscopy Using a Metallic Vibrating Tip. Ultramicroscopy 1995, 57, 318–322. 29. Kawata, S.; Inouye, Y. Scanning Probe Optical Microscopy Using a Metallic Probe Tip. Ultramicroscopy 1995, 57, 313–317. 30. Pohl, D. W.; Denk, W.; Lanz, M. Optical Stethoscopy: Image Recording with Resolution Lambda/20. Appl. Phys. Lett. 1984, 44, 651–653. 31. Lewis, A.; Isaacson, M.; Harootunian, A.; Muray, A. Development of a 500 Å Spatial Resolution Light Microscope. Ultramicroscopy 1984, 13, 227–231. 32. Eghlidi, H.; Lee, K.; Chen, X.-W.; Götzinger, S.; Sandoghdar, V. Resolution and Enhancement in Nanoantenna-Based Fluorescence Microscopy. Nano Lett. 2009, 9, 4007– 4011. 33. Frey, H. G.; Keilmann, F.; Kriele, A.; Guckenberger, R. Enhancing the Resolution of Scanning Near-Field Optical Microscopy by a Metal Tip Grown on an Aperture Probe. Appl. Phys. Lett. 2002, 81, 5030–5032. 34. Farahani, J.; Pohl, D. W.; Eisler, H.-J.; Hecht, B. Single Quantum Dot Coupled to Scanning Optical Antenna: A Tunable Superemitter. Phys. Rev. Lett. 2005, 95, 017402. 35. Taminiau, T. H.; Moerland, R. J.; Segerink, F. B.; Kuipers, L.; van Hulst, N. F. λ/4

14

ACS Paragon Plus Environment

Page 14 of 18

Page 15 of 18

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

Resonance of an Optical Monopole Antenna Probed by Single Molecule Fluorescence. Nano Lett. 2007, 7, 28–33. 36. Sugiura, T.; Kawata, S.; Okada, T. Fluorescence Imaging with a Laser Trapping Scanning Near-Field Optical Microscope. J. Micros. 1999, 194, 291–294. 37. Marago, O. M.; Jones, P. H.; Gucciardi, P. G.; Volpe, G.; Ferrari, A. C. Optical Trapping and Manipulation of Nanostructures. Nat. Nanotechnol. 2013, 8, 807–19. 38. Okamoto, K.; Kawata, S. Radiation Force Exerted on Subwavelength Particles near a Nanoaperture. Phys. Rev. Lett. 1999, 83, 4534–4537. 39. Pelton, M.; Liu, M.; Kim, H. Y.; Smith, G.; Guyot-Sionnest, P.; Scherer, N. F. Optical Trapping and Alignment of Single Gold Nanorods by Using Plasmon Resonances. Opt. Lett. 2006, 31, 2075–2077. 40. Kim, J.; Spindler, S.; Sandoghdar, V. Scanning-Aperture Trapping and Manipulation of Single Charged Nanoparticles. Nat. Commun. 2014, 5, 3380. 41. Krishnan, M.; Mojarad, N.; Kukura, P.; Sandoghdar, V. Geometry-Induced Electrostatic Trapping of Nanometric Objects in a Fluid. Nature 2010, 467, 692–695. 42. Chen, Y.; Vela, J.; Htoon, H.; Casson, J.; Werder, D.; Bussian, D.; Klimov, V.; Hollingsworth, J. Giant Multishell CdSe Nanocrystal Quantum Dots with Suppressed Blinking. J. Am. Chem. Soc. 2008, 130, 5026–5027. 43. Ghosh, Y.; Mangum, B.; Casson, J. L.; Williams, D. J.; Htoon, H.; Hollingsworth, J. A. New Insights into the Complexities of Shell Growth and the Strong Influence of Particle Volume in Non-Blinking “Giant” Core/Shell Nanocrystal Quantum Dots. J. Am. Chem. Soc. 2012, 134, 9634–9643. 44. Kalkbrenner, T.; Graf, M.; Durkan, C.; Mlynek, J.; Sandoghdar, V. High-Contrast

15

ACS Paragon Plus Environment

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Topography-Free Sample for Near-Field Optical Microscopy. Appl. Phys. Lett. 2000, 76, 1206. 45. Lee, K.; Eghlidi, H.; Chen, X.-W.; Renn, A.; Götzinger, S.; Sandoghdar, V. Spontaneous Emission Enhancement of a Single Molecule by a Double-Sphere Nanoantenna across an Interface. Opt. Express 2012, 20, 23331–23338. 46. Behrens, S.; Grier, D. The Charge of Glass and Silica Surfaces. J. Chem. Phys. 2001, 115, 6716–6721. 47. Lindfors, K.; Kalkbrenner, T.; Stoller, P.; Sandoghdar, V. Detection and Spectroscopy of Gold Nanoparticles Using Supercontinuum White Light Confocal Microscopy. Phys. Rev. Lett. 2004, 93, 037401. 48. Ortega-Arroyo, J.; Kukura, P. Interferometric Scattering Microscopy (iSCAT): New Frontiers in Ultrafast and Ultrasensitive Optical Microscopy. Phys. Chem. Chem. Phys. 2012, 14, 15625–15636. 49. Mojarad, N.; Sandoghdar, V.; Krishnan, M. Measuring Three-Dimensional Interaction Potentials Using Optical Interference. Opt. Express 2013, 21, 9377–89. 50. Karrai, K.; Tiemann, I. Interfacial Shear Force Microscopy. Phys. Rev. B 2000, 62, 13174–13181. 51. Kirby, B.; Hasselbrink, E. Zeta Potential of Microfluidic Substrates: 2. Data for Polymers. Electrophoresis 2004, 25, 203–213. 52. Novotny, L.; Beversluis, M.; Youngworth, K.; Brown, T. Fluorescence Imaging with a Laser Trapping Scanning Near-Field Optical Microscope. Phys. Rev. Lett. 2001, 86, 5251–5255. 53. Pfab, R. J.; Zimmermann, J.; Hettich, C.; Gerhardt, I.; Renn, A.; Sandoghdar, V.

16

ACS Paragon Plus Environment

Page 16 of 18

Page 17 of 18

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

Aligned Terrylene Molecules in a Spin-Coated Ultrathin Crystalline Film of p-terphenyl. Chem. Phys. Lett. 2004, 387, 490–495. 54. Kühn, S.; Mori, G.; Agio, M.; Sandoghdar, V. Modification of Single Molecule Fluorescence Close to a Nanostructure: Radiation Pattern, Spontaneous Emission and Quenching. Mol. Phys. 2008, 106, 893–908. 55. Chung, I.; Shimizu, K. T.; Bawendi, M. Room Temperature Measurements of the 3D Orientation of Single CdSe Quantum Dots Using Polarization Microscopy. Proc. Natl. Acad. Sci. U. S. A. 2003, 100, 405–408. 56. Brokmann, X.; Ehrensperger, M.-V.; Hermier, J.-P.; Triller, A.; Dahan, M. Orientational Imaging and Tracking of Single CdSe Nanocrystals by Defocused Microscopy. Chem. Phys. Lett. 2005, 406, 210–214. 57. Krishnan, M. Electrostatic Free Energy for a Confined Nanoscale Object in a Fluid. J. Chem. Phys. 2013, 138, 114906. 58. Celebrano, M.; Rosman, C.; Sönnichsen, C.; Krishnan, M. Angular Trapping of Anisometric Nano-Objects in a Fluid. Nano Lett. 2012, 12, 5791–5796. 59. Mojarad, N.; Krishnan, M. Measuring the Size and Charge of Single Nanoscale Objects in Solution Using an Electrostatic Fluidic Trap. Nat. Nanotechnol. 2012, 7, 448–452.

17

ACS Paragon Plus Environment

ACS Nano

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

TOC Graphics 76x41mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 18 of 18