PROCEEDINGS OF SPIE Field-effect electro-plasmonics: a quantum leap in neurotechnologies - NSF PAR
←
→
Page content transcription
If your browser does not render page correctly, please read the page content below
PROCEEDINGS OF SPIE SPIEDigitalLibrary.org/conference-proceedings-of-spie Field-effect electro-plasmonics: a quantum leap in neurotechnologies Habib, Ahsan, Zhu, Xiangchao, Can, Uryan, McLanahan, Maverick, Zorlutuna, Pinar, et al. Ahsan Habib, Xiangchao Zhu, Uryan I. Can, Maverick McLanahan, Pinar Zorlutuna, Ahmet Ali Yanik, "Field-effect electro-plasmonics: a quantum leap in neurotechnologies," Proc. SPIE 11461, Active Photonic Platforms XII, 1146129 (4 September 2020); doi: 10.1117/12.2569154 Event: SPIE Nanoscience + Engineering, 2020, Online Only Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 26 Feb 2021 Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
Invited Paper Field-Effect Electro-Plasmonics: A Quantum Leap in Neurotechnologies Ahsan Habiba, Xiangchao Zhua, Uryan I. Canb, Maverick L. McLanahanc, Pinar Zorlutunad, Ahmet Ali Yanika* a Baskin School of Engineering, University of California Santa Cruz, Santa Cruz, CA, 95064, USA; d Aerospace and Mechanical Engineering, University of Notre Dame, Notre Dame, IN, 46556, USA; c Physics Department, University of California Santa Cruz, Santa Cruz, CA, 95064, USA. Correspondence to: Email: yanik@ ucsc.edu ABSTRACT State-of-art electro-optic translators cannot provide high signal-to-noise ratio electric-field measurement capability due to whether low photon counts (e.g., voltage sensitive dyes) or low electric-field sensitivities (e.g., quantum dots). Here, we demonstrate that electrochromic loading of plasmonic (electro-plasmonic) nanoantenna allows us to overcome these limitations and to realize extremely bright, sensitive and fast nanoscale electric-field probes. Our electro-plasmonic nanoantennas have 10-100 million times larger cross sections than fluorescence dye molecules and provide three orders of magnitude enhanced electric-field sensitivities than conventional plasmonic nanoantennas with sub-millisecond temporal response times (~ 0.2 ms). In our experiments, we demonstrated label-free optical detection of electrogenic activity of stem cell derived cardiomyocytes with high signal-to-noise ratios at low light intensity conditions. Our novel approach presents a remarkable technological leap for label-free optical imaging of electric-field dynamics with high spatiotemporal resolution. Keywords: Active plasmonics, electro-plasmonics, nanoantenna, neurophotonics, label-free detection, electrophysiology. 1. INTRODUCTION Plasmonic nanoantenna can dramatically concentrate light (down to105 times smaller volumes than the diffraction limited volume) and offer an abundance of exciting opportunities for biomedical applications [1-10]. Strongly enhanced near- fields of localized surface plasmons (LSPs) allow transduction of small fluctuations in the local refractive index to easily detectable spectral signals in the far-field spectra [2, 11-15]. Label-free biosensing technologies based on these concepts are well established; down to a single molecule detection capabilities are shown [16-18]. However, there is very limited progress towards real-time optical detection of local electrical field dynamics, which is highly desirable for the detection of the extracellular electrophysiological signals. During the past decade, there has been a concentrated effort to develop active plasmonic devices using inherent voltage sensitivity of noble metals for local field measurements. However, due to the high electron densities, the inherent voltage sensitivities of the metals are low. Electro-optic effects in metals are extremely weak [19]. Here we are presenting a new class of extremely bright non-fluorescent optical voltage sensors that can sensitively detect local electric-field dynamics. With more than 3.25´103 times enhanced field sensitivities over conventional plasmonic nanoantennas, our electrochromically loaded plasmonic (electro-plasmonic) nanoantennas enable high signal-to-shot-noise ratio (SSNR~40-250) local electric field measurements with light in a label-free fashion. Our electro-plasmonic probes compare demonstrate superior characteristics with respect to the state-of-the-art optical probes, such as genetically encoded voltage indicators (GEVIs), which have a low SSNR [20-22] due to small cross-sections (~10−2 nm2) [23] and low quantum yields (~10−3 to 10−2)[24]. Furthermore, we have demonstrated a label-free and non-invasive optical measurement of the electrogenic activity of cardiac muscle cell (CM) through our in vitro measurements using low-intensity light, which is two to three orders of magnitude lower than the typical light intensities used for GEVIs [22]. Active Photonic Platforms XII, edited by Ganapathi S. Subramania, Stavroula Foteinopoulou, Proc. of SPIE Vol. 11461, 1146129 · © 2020 SPIE · CCC code: 0277-786X/20/$21 · doi: 10.1117/12.2569154 Proc. of SPIE Vol. 11461 1146129-1 Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 26 Feb 2021 Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
2. MATERIALS AND METHODS Electron beam lithography (EBL) was used to fabricate the plasmonic nanoantenna array (Fig.1a). A positive photoresist, 495 PMMA A4 (Microchem, USA), was spin coated (4000 rpm) on the ITO glass slides (Structure probe, Inc., US A) and then baked at 180° C on a hot plate for 90 seconds. The nanometer pattern generation system (NPGS) was used for electron beam patterning (90 nm in diameter gold nanoantennas with a periodicity of 500 nm) with an FEI Quanta 3D field emission microscope. Subsequently, the sample was developed for 1 minute in isobutyl ketone (MIBK)-isopropanol (IPA) solution (MIBK: IPA=1:3) (Microchem, USA) and was placed in an IPA solution for 1 minute to stop the development. The sample was then dried under a stream of high-purity nitrogen. A 45 nm thick gold layer (Kurt J. Lesker, USA) electron beam evaporation process was performed. Deposition was done at 1.2 x 10-6 pressure and 0.5 Å/sec evaporation rate. The sample was dipped into acetone and subsequently a brief (~5 s) metal liftoff process was performed with ultrasonication. Figure 1. Electro-plasmonic nanoantenna fabrication and characterization. (a) Fabrication steps of the electric field probe using electron beam lithography (EBL) and electrochemical deposition. (b) Scanning electron microscope (SEM) image of the electro-plasmonic field probe. PEDOT: PSS was selectively deposited on nanoantenna in a monomer of 10 mM EDOT (Sigma-Aldrich, USA) and 0.1M NaPSS (Sigma-Aldrich, USA) into an aqueous solvent using a custom developed electrochemical deposition technique. The deposition was carried out in a three-electrode electrochemical cell under the potentiostatic conditions: the working electrode was the nanoantenna array, the counter-electrode was the platinum coil (Alfa Aesar, USA) and the reference electrode was Ag/AgCl (Warner instruments, USA). The electro-plasmonic nanoantenna consisted of PEDOT: PSS layer as shown in Figure 1b. In the final step, neonatal rat ventricular cardiac muscle cell (CM) was seeded on the electro- plasmonic nanoantenna according to a previously established protocol [25]. 3. RESULTS AND DISCUSSION We performed electro-optic measurements to determine the field sensitivity of our device within an external 10-2-10-3 mV/nm electric field range, which is comparable to extracellular fields [26]. Figure 2a illustrates the optical setup used. The measurement was performed with controlled electric fields generated through a transparent counter electrode. Proc. of SPIE Vol. 11461 1146129-2 Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 26 Feb 2021 Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
Figure 2. Electro-optic characterization of electro-plasmonic field probe. (a) Scanning electron microscope (SEM) image of the electro- plasmonic field probe. (b) Electric field sensitivity of the electro-plasmonic nanoantenna. We have developed a quasi-static model that treats Au particles and conducting polymer as Drude metals to provide a physical insight into linear dependence of the signal to the external electric field. The spectral modulation due to dielectric constant variations in the electrochromic PEDOT: PSS load can be expressed as in [7]: ⎡ ⎤ ⎢ ⎥ Δλ LSP ω λ ⎢ 2 3 (1− L ) L ⎥ = 2 Δε PEDOT EP p LSP (1) 8π c ⎢ ⎛ 2 2 ⎛ 1− L ⎞ ⎞ ⎥ ⎢ ⎜ ε ∞ + ε PEDOT ⎜ ⎥ ⎢⎣ ⎝ ⎝ L ⎟⎠ ⎟⎠ ⎥⎦ where L is the geometrical factor for the nanoantenna, ω p is the metal plasma frequency, ε ∞ is the high frequency contribution to metal dielectric function, and ePEDOT is the dielectric constant of the PEDOT film. Here, the resonance wavelength shift Δλ LSP is linearly proportional to the DePEDOT, the change in the PEDOT: PSS load permittivity with EP electric field. Following Drude model [27, 28], DePEDOT can be expressed as [7]: ω 3p,PEDOT ⎛ ε0 ⎞ Δε PEDOT =− 2 2 ⎜ PEDOT ⎟ E (2) γ PEDOT + ω ⎝ N PEDOT edTF ⎠ where dTFPEDOT = 1/ 2(ao3 / N PEDOT )1/6 is the Thomas-Fermi screening length, ao = 5.29 × 10−11 m is the Bohr radius, NPEDOT is the free electron density in PEDOT conducting polymer. wp,PEDOT is the bulk plasma frequency, e¥,PEDOT is the relative permittivity at the high-frequency limit and gPEDOT is the damping coefficient, e is the electron charge, and E is the external electric field strength. For the conventional plasmonic nanoantenna, the differential scattering signal displays a linear dependence to the applied electric field with 0.28´10-3 nm/mV (Fig. 2b, blue line). The linear relationship in between the differential scattering signal and the local electric field strength is associated to the altering plasma frequency of the metal [7]. In the case of PEDOT: PSS, following Eqs.1-2, dielectric permittivity modulations can be simply expressed as Δε PEDOT ∝ E . The LSP resonance of the electro-plasmonic nanoantenna is extremely sensitive to the external field through the PEDOT:PSS load. In our experiments, we showed that demonstrated more than three orders of magnitude enhanced sensitivity of 9.1 × 10−5 cm/V for the electro-plasmonic nanoantenna (Fig.2b, red line) [7]. Proc. of SPIE Vol. 11461 1146129-3 Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 26 Feb 2021 Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
Figure 3. Electric field sensitivity and scaling of electro-plasmonic field probe. (a) Electric field sensitivity of a 90 nm diameter electro- plasmonic nanoantenna. Differential scattering signal (blue-curve) and SSNR (red-curve) are calculated at different distances from the cell. We assumed an incident light intensity of 100 W/cm2, an integration time of 1 ms, NA of 0.45, and an overall detection efficiency of 50%. (b) Resonance wavelength shift for varying diameters of electro-plasmonic voltage probes are calculated using Eq. 1. Here, we use E=150 V/cm (~ 100 µm from the cell), ! =1.79 pHz, Qsca =5.6 and "#$%& =1.43 pHz [28]. L is calculated analytically [29]. Next, we sought to establish detection limits of our electro-plasmonic nanoantenna. The optical shot noise provides a fundamental bottleneck for optical measurements with low photon counts [e.g., the genetically encoded voltage indicators (GEVI) with tiny cross-sections and low quantum yield]. The photon counts need to be boosted to achieve high signal-to- shot-noise ratio measurements. Electro-plasmonic nanoantennas have physically much larger dimensions (90 nm in diameter), with approximately 10-million times large larger cross-sections compared to those of GEVI [23, 30]. In addition, the LSP controlled light scattering of the loaded nanoantenna results in a high incident to scattered light conversion efficiencies. We calculate the SSNR with the following relationship using experimentally obtained (DS/So)EP and 3-D FDTD simulations [7]: EP ⎛ λ ⎞ ⎛ ΔS ⎞ ( ) SSNR = QE ⎡⎣ηcollectionT ⎤⎦ I inc Qscatπ r 2 tint ⎜ ⎟ ⎜ ⎝ hc ⎠ ⎝ S0 E ⎟⎠ E (2) where hcollection is the solid angle fraction of the total scattered light collected by the microscope objective, T is the transmittance of the objective lens at the scattering wavelength, QE is the quantum efficiency of the photodetector at the scattering wavelength, Iinc is the incident light intensity, Qsca is the scattering efficiency (calculated using FDTD simulations), r is the radius of the electro-plasmonic nanoantenna, and tint is the integration time. hcollection , the solid angle fraction of the total scattered light collected by the microscope objective, is given by [31] ⎛ 2⎞ 1⎜ ⎛ NA ⎞ ηcollection = 1− 1− ⎜ ⎟ (3) ⎟ 2⎜ ⎝ ncoup ⎠ ⎟ ⎝ ⎠ where NA is the numerical aperture of the objective, and ncoup is the refractive index of the objective lens coupling medium. Proc. of SPIE Vol. 11461 1146129-4 Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 26 Feb 2021 Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
Our analysis shows that label-free optical detection of cell action potentials is possible with a remarkably high SSNR of >250 under typical illumination conditions (100 W/cm2). This measurement capability, outperforming millions of genetically tagged voltage sensitive dyes [24], could open the door to extracellular voltage imaging from diffraction limited volumes without averaging [7]. Moreover, electro-plasmonic nanoantenna detects electric fields in a similar fashion to extracellular microelectrodes without physical contact to the origin of the electrophysiological activity of the electrogenic cell. Our analysis indicates that electro-plasmonic nanoantenna can remotely measure the activity of excitable cells from distances up to 100 µm in a similar manner as extracellular microelectrodes (Fig. 3a, blue curve). Further improvements in photon counts and SSNRs could be achieved by packing electro-plasmonic nanoantenna closely. Scaling of the electro- plasmonic nanoantenna dimensions to achieve higher filling factors moderately affect the electric-field sensitivities (Fig. 3b). However, below 50 nm diameter, rapid scaling of nanoantenna scattering cross-section ( Cscat ∝ a6 ) with nanoantenna diameter ( ) poses a fundamental limitation as the photon counts drop significantly. Figure 4. Spectroelectrochemical characterization of the electro-plasmonic voltage probe. (a). Experimental scheme used for the dark- field spectroelectrochemical measurements. (b) Dynamic response of the electro-plasmonic field probe (red) due to the alternating (1 kHz) external field (blue). We conduct spectroelectrochemical measurements using a transmission dark-field microscope and probe high bandwidth measurement capabilities of our electro-plasmonic voltage probe (Figure 4a). Figure 4b shows that our electro-plasmonic field probe can closely follow alternating external field with 1kHz frequency. For electrophysiological testing, cardiac muscle cells (CMs) are seeded on the electro-plasmonic nanoantenna array (Fig. 5a). Transmission dark-field measurements are carried out with a compact spectrometer at a fixed integration time of 50 ms (Fig. 4a). We performed our experiments using light intensities (11 mW/mm2) that are ~102-103 times lower than the typical light intensities used in electrophysiological experiments with GEVIs [24]. Our results show that synchronized spiking of the CMs leads to increased light scattering (Fig. 5b). To confirm the synchronized behavior, we characterize the phenotype of the cells through immunostaining Cardiac Troponin I (cardiomyocyte marker, green), and Connexin 43 (gap junction marker, red) (Fig. 5c). In figure 5c, cardiomyocyte marker confirms that the all cells seeded on electro-plasmonic nanoantenna are CMs (no cardiac fibroblast cell). Also, gap junction marker displays that the CMs are electrically coupled and coupling is mediated by gap junction protein (Connexin units) [32]. Proc. of SPIE Vol. 11461 1146129-5 Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 26 Feb 2021 Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
Figure 5. Optoelectrochemical characterization of the electro-plasmonic field probe. (a). SEM micrograph of the cardiac muscle cells cultured on electro-plasmonic nanoantenna array (b Differential scattering signal DS/So in response to the electrophysiological activity of a network of cardiomyocyte cells. (c) Fluorescence image of Troponin-I (green), and Connexin 43 (red) immunostaining of CMs seeded on electro-plasmonic nanoantennas. 4. CONCLUSIONS We have developed a new all-optical, label free, high-throughput, and non-invasive voltage probe to measure the electrophysiological activity of the electrogenic cell based on electrochromic loading of the plasmonic nanoantenna. Differential scattering signal of the electro-plasmonic nanoantenna is very sensitive to the local electric field due to the strong light-localization beyond the diffraction limit and low electron densities of the electrochromic loading material. In our experiments, we observed three orders of magnitude enhanced sensitivities with respect to previous all-optical voltage probes and high-SSNRs without averaging. 5. ACKNOWLEDGMENTS This work is supported by National Science Foundation grants ECCS-1611290, ECCS-1611083, ECCS-1847733 (CAREER Award) and CBET-1651385 (CAREER Award). The authors acknowledge Dr. Tom Yuzvinsky (UCSC) for his assistance with device fabrication, the W.M. Keck Center for Nanoscale Optofluidics for the use of the FEI Quanta 3D, and Imran Hossain (UCSC) for help in setting up the optical measurement systems and technical assistance. REFERENCES [1] J. A. Schuller, E. S. Barnard, W. Cai et al., “Plasmonics for extreme light concentration and manipulation,” Nat Mater, 9(3), 193-204 (2010). [2] J. N. Anker, W. P. Hall, O. Lyandres et al., “Biosensing with plasmonic nanosensors,” Nature Materials, 7, 442 (2008). [3] A. Kinkhabwala, Z. Yu, S. Fan et al., “Large single-molecule fluorescence enhancements produced by a bowtie nanoantenna,” Nature Photonics, 3, 654 (2009). [4] R. Adato, A. A. Yanik, J. J. Amsden et al., “Ultra-sensitive vibrational spectroscopy of protein monolayers with plasmonic nanoantenna arrays,” Proceedings of the National Academy of Sciences, 106(46), 19227 (2009). [5] C. Wu, A. B. Khanikaev, R. Adato et al., “Fano-resonant asymmetric metamaterials for ultrasensitive spectroscopy and identification of molecular monolayers,” Nat Mater, 11(1), 69-75 (2012). [6] A. V. Kabashin, P. Evans, S. Pastkovsky et al., “Plasmonic nanorod metamaterials for biosensing,” Nature Materials, 8, 867 (2009). Proc. of SPIE Vol. 11461 1146129-6 Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 26 Feb 2021 Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
[7] A. Habib, X. Zhu, U. I. Can et al., “Electro-plasmonic nanoantenna: A nonfluorescent optical probe for ultrasensitive label-free detection of electrophysiological signals,” Science Advances, 5(10), eaav9786 (2019). [8] X. Zhu, A. Cicek, Y. Li et al., “Plasmofluidic Microlenses for Label-Free Optical Sorting of Exosomes,” Scientific Reports, 9(1), 8593 (2019). [9] C. A. R. Chapman, X. Zhu, H. Chen et al., “Nanostructure Introduces Artifacts in Quantitative Immunofluorescence by Influencing Fluorophore Intensity,” Scientific Reports, 7(1), 427 (2017). [10] X. Zhu, G. M. Imran Hossain, M. George et al., “Beyond Noble Metals: High Q-Factor Aluminum Nanoplasmonics,” ACS Photonics, 7(2), 416-424 (2020). [11] A. A. Yanik, A. E. Cetin, M. Huang et al., “Seeing protein monolayers with naked eye through plasmonic Fano resonances,” Proceedings of the National Academy of Sciences, (2011). [12] A. A. Yanik, M. Huang, O. Kamohara et al., “An Optofluidic Nanoplasmonic Biosensor for Direct Detection of Live Viruses from Biological Media,” Nano Letters, 10(12), 4962-4969 (2010). [13] F. Eftekhari, C. Escobedo, J. Ferreira et al., “Nanoholes As Nanochannels: Flow-through Plasmonic Sensing,” Analytical Chemistry, 81(11), 4308-4311 (2009). [14] A. G. Brolo, “Plasmonics for future biosensors,” Nat Photon, 6(11), 709-713 (2012). [15] X. Zhu, N. Cao, B. J. Thibeault et al., “Mechanisms of Fano-resonant biosensing: Mechanical loading of plasmonic oscillators,” Optics Communications, 469, 125780 (2020). [16] P. Zijlstra, P. M. Paulo, and M. Orrit, “Optical detection of single non-absorbing molecules using the surface plasmon resonance of a gold nanorod,” Nat Nanotechnol, 7(6), 379-82 (2012). [17] I. Ament, J. Prasad, A. Henkel et al., “Single Unlabeled Protein Detection on Individual Plasmonic Nanoparticles,” Nano Letters, 12(2), 1092-1095 (2012). [18] T. J. Antosiewicz, and M. Käll, “A Multiscale Approach to Modeling Plasmonic Nanorod Biosensors,” The Journal of Physical Chemistry C, 120(37), 20692-20701 (2016). [19] X. Liu, J.-H. Kang, H. Yuan et al., “Electrical tuning of a quantum plasmonic resonance,” Nature Nanotechnology, 12, 866 (2017). [20] M. Scanziani, and M. Hausser, “Electrophysiology in the age of light,” Nature, 461(7266), 930-939 (2009). [21] T. Knöpfel, J. Díez-García, and W. Akemann, “Optical probing of neuronal circuit dynamics: genetically encoded versus classical fluorescent sensors,” Trends in Neurosciences, 29(3), 160-166 (2006). [22] V. Emiliani, A. E. Cohen, K. Deisseroth et al., “All-Optical Interrogation of Neural Circuits,” The Journal of Neuroscience, 35(41), 13917-13926 (2015). [23] L. Kastrup, and S. W. Hell, “Absolute optical cross section of individual fluorescent molecules,” Angew Chem Int Ed Engl, 43(48), 6646-9 (2004). [24] D. R. Hochbaum, Y. Zhao, S. L. Farhi et al., “All-optical electrophysiology in mammalian neurons using engineered microbial rhodopsins,” Nat Meth, 11(8), 825-833 (2014). [25] J. Fu, J. Gao, R. Pi et al., “An Optimized Protocol for Culture of Cardiomyocyte from Neonatal Rat,” Cytotechnology, 49(2), 109-116 (2005). [26] S. Sylantyev, L. P. Savtchenko, Y.-P. Niu et al., “Electric Fields Due to Synaptic Currents Sharpen Excitatory Transmission,” Science, 319(5871), 1845-1849 (2008). [27] S. H. Chang, C. Chiang, F. Kao et al., “Unraveling the Enhanced Electrical Conductivity of PEDOT:PSS Thin Films for ITO-Free Organic Photovoltaics,” IEEE Photonics Journal, 6(4), 1-7 (2014). [28] K.-Y. Jung, W.-J. Yoon, Y. B. Park et al., “Broadband Finite-Difference Time-Domain Modeling of Plasmonic Organic Photovoltaics,” ETRI Journal, 36(4), 654-661 (2014). [29] S. A. Maier, Plasmonics: Fundamentals and applications, (2007). [30] P. K. Jain, K. S. Lee, I. H. El-Sayed et al., “Calculated Absorption and Scattering Properties of Gold Nanoparticles of Different Size, Shape, and Composition: Applications in Biological Imaging and Biomedicine,” The Journal of Physical Chemistry B, 110(14), 7238-7248 (2006). [31] C. A. DiMarzio, [Optics for Engineers] Taylor & Francis, (2011). [32] J. M. T. de Bakker, and H. V. M. van Rijen, [Cardiac Action Potentials, Ion Channels, and Gap Junctions] Springer US, Boston, MA(2010). Proc. of SPIE Vol. 11461 1146129-7 Downloaded From: https://www.spiedigitallibrary.org/conference-proceedings-of-spie on 26 Feb 2021 Terms of Use: https://www.spiedigitallibrary.org/terms-of-use
You can also read