Abstract

Pore-based structures occur widely in living organisms. Ion channels embedded in cell membranes, for example, provide pathways, where electron and proton transfer are coupled to the exchange of vital molecules. Learning from mother nature, a recent surge in activity has focused on artificial nanopore architectures to effect electrochemical transformations not accessible in larger structures. Here, we highlight these exciting advances. Starting with a brief overview of nanopore electrodes, including the early history and development of nanopore sensing based on nanopore-confined electrochemistry, we address the core concepts and special characteristics of nanopores in electron transfer. We describe nanopore-based electrochemical sensing and processing, discuss performance limits and challenges, and conclude with an outlook for next-generation nanopore electrode sensing platforms and the opportunities they present.
Short abstract
We highlight the development of nanopore-confined electrochemistry, address the core concepts of electron transfer in nanopores, and provide an outlook for next-generation nanopore electrochemistry.
Introduction
Chemical measurement science benefited tremendously from the post-1960 emergence of sophisticated instruments coupled with automatic data analysis, which enabled researchers to separate, identify, and quantify targets at unprecedented levels.1 Nowadays, the extensive overlap between fundamental research and practical applications2,3 has brought within reach the ultimate goal of measurements which simultaneously exhibit ultrahigh sensitivity and selectivity in physical formats of reasonable cost and complexity. One powerful approach to realizing this ideal is to introduce mass-limited samples in ultrasmall confined volumes, sort the resulting mixtures, and quantify each entity one-by-one. Nanopores are biomimetic architectures that mimic the behavior of ion channels by confining targets to nanoscale volumes from which measurable signals can be generated in situ, thus recapitulating the key features of an ideal detection system.4 In this Introduction, we trace the histories of nanopores and nanoelectrodes, the merging of the two streams in the early 2000s, and finally the genesis of nanopore electrode systems.
Pore-based analysis originated in 1953 with the resistive pulse detection scheme of Wallace Coulter,5 which made possible the analysis of sub-μm particles and macromolecules in the Coulter Counter.6,7 Soon after, the quest to characterize single biomolecules stimulated researchers to use nanopores commensurate in size with target analytes.8,9 The human genome project then triggered the development of fast, cheap, and label-free DNA detection, and in 1996, Kasianowicz et al. detected single DNA strands using an α-hemolysin nanopore.10 Follow-up work led to the emergence of solid-state nanopore sensors exhibiting performance characteristics comparable to those of biological nanopores.11,12 Subsequently, Bayley and co-workers achieved single-nucleobase discrimination13,14 leading Oxford Nanopore Technologies to launch a portable real-time DNA sequencing platform in 2014.
At the same time, enhanced understanding of nanofluidics opened the way to controlling molecular transport at unprecedented levels. Rice and Whitehead first described electrokinetic transport in nanoscale capillaries,15 and the theory of potential driven flow was developed by Levine and co-workers.16 During the 1980s and 1990s, fabrication of nanochannels became easier, less expensive, and more versatile.17,18 In 1999, Ramsey and co-workers demonstrated the first in-plane nanoporous structure for sample preconcentration,19 and later surface charge in nanocapillary array membranes was exploited to effect digital nanofluidic coupling between microfluidic chanels.20 In 2004, Dekker and co-workers developed ion transport platforms that were governed by nanochannel surface charge.21 Then, Yang and Majumdar developed ionic diodes, i.e., rectified nanofluidic ion currents.22,23 Recently, Jiang and co-workers reported a biomimetic nanochannel sensing platform in which the response to ions and molecules is controlled by surface functionalization.24
In contrast, nanoelectrodes (so-called ultramicroelectrodes) were developed in the 1980s, when electrochemists started using them to detect trace analytes and perform transient electrochemical measurements.25−27 During the 1990s Martin’s group developed a robust synthesis of pore-based nanomaterials;28 Bard and Fan designed an elegant purely electrochemical detection of a single molecule;29 and Murray’s group observed the charging of gold nanoparticles (quantized capacitors).30 Inevitably, nanopores and nanoelectrodes were combined to yield single nanopore electrodes, by Zhang and White in 2004.31 Later, single nanopore electrodes and nanopore electrode arrays with well-defined, reproducible pore geometry and size were fabricated lithographically.32,33 Compared with early work, in which electrodes were embedded in track-etched or anodic aluminum oxide membranes, these well-defined nanopore electrodes enabled an additional level of control over transport and reactivity that was exploited to yield enhanced nanoscale electrochemical measurements. Subsequently, the term “nanoelectrochemistry” was coined to describe phenomena ranging from fabrication and characterization of nanoelectrodes to the applications of nanoelectrodes as ultrasensitive tools for electroanalysis.34−38
Here, we examine the overlapping regimes of nanopore sensing, nanoscale transport, and nanoelectrochemistry, shown in Figure 1. We start by introducing the core characteristics of the nanopore electrode, review progress over the past decade, and finally discuss the remaining limits and challenges and propose the outlook for next-generation nanopore electrodes and electrode arrays. Because solid-state nanopores, unlike biological nanopores, can take advantage of mature nanofabrication processes, flexible choice of materials, and easily altered functionality,39,40 they will be the focus of this outlook.
Figure 1.
Schematic illustration showing recent progress in nanopore sensing (a), nanoelectrochemistry (b), and nanoscale transport (c), including nanopore-based DNA sequencing (a1); single nucleotide identification by nanopore tunneling current (a2); stochastic protein sensing by receptor modified nanopore (a3); single molecule detection by redox cycling (b1); multiple collisions or gated transport of nanoparticles in nanopores (b2); correlated optical and electrchemical analysis of single entities (b3); molecular sieving in nanopore arrays (c1); asymmetric nanopores and bipolar nanopore as ionic diodes (c2); and surface-charge governed ion transport in nanochannels compared to microchannels (c3).
Defining Characteristics
Benefits of Nanoelectrodes
Nanoscale electrodes exhibit enhanced mass transport, enhanced faradaic currents, and negligible iR drop during electrochemical measurements.34 As shown in Figure 2a, the benefits of nanopore electrodes can be classified as follows. (1) Nanopore electrodes provide small confinement volumes, which significantly enhance collision frequencies with the electrode surface.41,42 A molecule with a diffusion coefficient of 10–5 cm2/s will collide with the wall of a 1000 nm3 nanopore millions of times more frequently than in a 1000 μm3 micropore. (2) Mass transport driven by unscreened electric fields is efficient and tunable at the nanoscale.43,44 For example, the field strength between two electrodes with a 10 nm gap is 106 V/cm at ΔE = 1 V. In addition, the direction of mass transport can be easily switched.45 Nanopores can be fabricated to be size-commensurate with the Debye length, producing strong coupling between the solution ion distribution and the nanopore surface charge, i.e., permselectivity.21,46 (3) Nanopores can serve as nanoscale reactors. Individually addressable electrodes can be inserted into the nanopore and used to control electron transfer processes, thereby achieving vectorial coupling of reactions.47 (4) Stochastic phenomena dominate when nanopore electrochemical measurements involve only one or a few molecules, making it possible to resolve the dynamics of single electron transfer events.48
Figure 2.
Schematic illustration of the defining nanopore electrode characteristics (a), fabrication methods (b), and surface modification (c) of nanopore-based sensors. Characteristic behaviors include the confinement effect (a1); strong electric fields (a2); vectorially coupled reactions (a3); and stochastic processes (a4). Pertinent fabrication methods include ion beam milling (b1); electron beam lithography (b2); nanoimprint lithography (b3); direct self-assembly of block copolymers (b4); nanosphere lithography (b5); and anodic aluminum oxide nanoporous templates (b6). Chemical modification strategies include orthogonal chemical processes for surface modification (c1) and nanopores with external stimuli, e.g., pH, temperature, ion strength, light, electric or magnetic field, and bioaffinity agents, e.g., protein, DNA, RNA, and metabolites (c2).
Fabrication and Surface Modification
While the fabrication of nanopore electrodes is based on well-developed approaches for nanofabrication, there are some issues that are characteristic of nanopore electrodes compared to nanobands, nanodisks, etc. First, electrodes are commonly embedded into substrates, typically as a multilayer film, before pore formation. In addition, nanopore electrodes are typically combined with confined cavities or fluidic channels. Both of these considerations determine the shape, size, and type of nanopore electrodes that can be fabricated. As shown in Figure 2b, nanopore electrodes are fabricated by one of two principal approaches. Highly precise single nanopores can be prepared at the sub-10 nm level by state-of-the-art fabrication approaches, such as focused ion beam (FIB) milling11,49 or e-beam lithography (EBL).12,50 In contrast, massively parallel approaches, such as nanoimprint lithography,51,52 nanosphere lithography,53,54 self-assembly of block copolymers,55,56 and nanoporous alumina templates,57,58 can produce nanopore arrays over large areas. Once formed, nanopores may be surface modified as long as (1) the surface remains inert to redox reactions in the applied potential window and (2) small charging currents are maintained. In addition, it is desirable if the structures admit versatile, yet specific, methods to modify the electrode surface. So far, noble metals, e.g., Pt or Au, and carbon-based materials, e.g., graphite, carbon nanotubes, etc., are the most widely used electrode materials, Figure 2c, and these admit a rich portfolio of surface modification approaches,4 imbuing the nanopore, for example, with stimulus-response (pH,59 temperature,60 light61) characteristics or molecular recognition capabilities.62,63
Nanopore Electrode Capabilities
The dramatic increase in research focused on nanopore electrodes and electrode arrays has resulted in a plethora of new and exciting capabilities for the chemical sciences—single entity electrochemistry, current and molecular rectification, scanning ion conductance mapping, concentration polarization, permselectivity, and correlated photonic and electrochemical measurements, to name a few. Rather than an exhaustive review, here we highlight a few of the forward looking nanopore electrode-enabled measurements that exhibit transformational new capabilities.
Single Molecule or Nanoparticle Electrochemistry
Single molecule electrochemistry is a holy grail, which is especially challenging due to the intrinsic Johnson noise at accessible gain–bandwidth conditions near 300 K.64−66 Currently, there are two widely used strategies to detect single molecules electrochemically, both of which rely on current amplification by factors >104 to produce detectable currents in the fA range. Nanopore electrodes, especially those with two closely placed and independently addressable electrodes, are especially powerful in this context, as they support enhanced collision frequencies and thus large current amplifications. In the pioneering work of Fan and Bard, redox molecules were confined in the ultrasmall volume between a scanning electrochemical microscopy (SECM) tip and conductive substrate, thus facilitating fast and efficient redox cycling.29 Subsequently, Sun and Mirkin used zeptoliter-volume recessed-disk electrodes to achieve quantized cyclic voltammograms of very few (n < 3) molecules, Figure 3a.41 Recently, Unwin and co-workers reported a novel quadruple nanostructure electrode in which the current traces reflect the fluctuations from the oxidation and reduction of single molecules confined in the nanopore,42Figure 3b. Another current amplification strategy relies on catalysis. The high turnover rate of a catalytic site can convert non-redox-active substrates into redox-active products, producing detectable currents, enabling detection down to <100 molecules.67,68 Importantly, the catalyst approach can extend the range of substrates to the detection of small numbers of nonredox active molecules, using designs in which the target molecule is sandwiched in a complex between component A bound to the electrode surface and component B, the catalyst that converts nonredox species into redox-active species.
Figure 3.

(a) Schematic illustration of a nanopore electrode immersed into Hg. The limiting current values of cyclic voltammograms correspond to zero (black), one (orange), two (blue), three (green), and four (red) molecules. (b) Schematic illustration of four electrode configuration, where the molecules are confined within the nanogap electrochemical cell. Current–time plots are from carbon working electrode (red line) and substrate working electrode (black line), where symmetric peaks indicate a single molecule event based on highly efficient redox cycling. (c) Schematic illustration of the self-assembly of gold nanoparticles (AuNPs) at the tip of a nanopore electrode, where the dark-field (20 μm scale) and TEM (20 nm scale) images present several and a cornel of microcyclic AuNP structures, respectively. (d) Schematic illustration of the manipulation of the fluorescent nanoparticle by nanopipettes, where the trajectories of each nanoparticle are captured in real time by the electron multiplied CCD detector for three-dimensional super-resolution imaging. Panel a reproduced with permission from ref (41). Copyright 2008 American Chemical Chemistry. Panel b reproduced with permission from ref (42). Copyright 2015 American Chemical Society. Panel c reproduced with permission from ref (79). Copyright 2017 Wiley-VCH. Panel d reproduced with permission from ref (80). Copyright 2017 American Chemical Society.
Nanoparticles, vesicles, and droplets have also been addressed in single entity experiments, which are primarily focused on two aspects. First, the heterogeneity of electron transfer from individual nanoparticles or nanoclusters provides useful kinetic information that can be used, for example, to develop more powerful electrocatalysts.69,70 Second, material released from vesicles or droplets provides a natural ex vivo mimic of extracellular release processes, e.g., neurotransmitters released from neurons.71,72 Early work addressing single nanoparticle collisions was conducted on ultramicroeletrodes by Bard et al.73 and others.74−76 Since these early reports, it has become possible to sequester a few nanoparticles, or even one.77,78 Recently, Long et al. reported an innovative nanopore bipolar electrode to control the dynamic self-assembly of gold nanoparticles,79Figure 3c. Similarly, White and co-workers proposed a super-resolution imaging method to map the trajectories of fluorescent nanoparticles around the tip of a nanopipette,80Figure 3d. These are just a few examples illustrating the broad interest in single entity electrochemistry; readers may refer to recent comprehensive reviews for additional details.81−84
High Density Nanopore Sensing Array
The experiments above highlight the push to single entity level detection in a single nanopore electrode. A natural extension is multiple nanopore electrodes on one device, i.e., nanopore electrode arrays (NEAs), either to enhance signal without losing the unique features of nanoscale electrode or to operate as multiplex sensors.85,86 In order to avoid the problem of overlapping diffusion profiles in high density nanopore arrays,34 Bohn and co-workers fabricated high density NEAs with two closely placed intrapore electrodes, Figure 4a, so that reversible redox couples undergo coupled reduction and oxidation reactions at oppositely biased top and bottom electrodes. The collection efficiency of redox species for both electrodes is close to 100%,87,88 which results in both greatly enhanced redox cycling and selectivity,89−91Figure 4b. Furthermore, electrochemical events can be efficiently converted to optical, e.g., fluorescence, readout by coupling the redox cycling signal to a distal reporter cell with a bipolar electrode.92 Recently, a high porosity permselective membrane was integrated with an NEA to mediate molecular transport, enhancing the selectivity to analytes of different charge.93 The permselective membrane serves as an ideal ion gate, controlling the access of charged analytes to the nanopore. Rectified redox cycling currents have also been observed raising the possibility of ionic diode functionality.
Figure 4.

(a) Photo of 8 pairs of nanopore electrode arrays (NEAs) on an electrochemical chip, where SEM images indicate the plan view (top) and side view (bottom) of NEAs, respectively. The gray, white, and black layers in cross sectional SEM image are SiO2, gold, and SiNx, respectively. (b) Schematic illustration of ion migration and accumulation in NEAs, as demonstrated by large current amplification at low ionic strength. (c) Schematic illustration of dual-ring NEAs serving as E-ZMWs, where the voltage-sensitive dye molecules are only excited inside the nanopore. The electrochemical and fluorescence signals are correlated, revealing single molecule population fluctuations across the nanopore array. Panel a reproduced with permission from ref (91). Copyright 2017 Royal Society of Chemistry. Panel b reproduced with permission from ref (90) (Copyright 2016 American Chemical Society) and ref (89) (Copyright 2014 American Chemical Society). Panel c reproduced with permission from ref (103). Copyright 2017 Royal Society of Chemistry.
Correlated Electrochemical and Optical Detection
Direct electronic detection of quantized events characterized by the passage of a few electrons is limited by the Johnson noise floor. In contrast, shifting to the more tractable problem of photon detection would allow the sensitivity issue at low analyte numbers to be addressed.94−97 To achieve this, the bottom ring of dual-ring NEAs can be used both as a working electrode and as the optical cladding layer of a zero mode waveguide (ZMW).98,99 The resulting electrochemical ZMWs (E-ZMWs) are ideal systems to investigate the singe molecule spectroelectrochemistry100 and have been used to probe single molecule dynamics of immobilized101 and freely diffused enzymes102,103 of the fluorigenic flavoenzyme monomeric sarcosine oxidase, by modulating the fluorescence ON and OFF with applied potential, Figure 4c. The electrochemical and fluorescence signals are correlated, revealing single molecule fluctuations across the nanopore array. This method holds great promise for the study of vectorially coupled enzyme reactions at single molecule sensitivity.
Challenges and Limits
As successful as nanopore electrochemical structures have been, there exist both practical and fundamental limits to performance. The resulting challenges and limits constitute a list of attractive candidates for new breakthroughs.
High Bandwidth Limits and Current Fluctuation
One of the stiffest technical challenges is to measure ultrasmall currents (<1 pA) at bandwidths (>1 MHz) corresponding to nanopore residence times at reasonable cost.104 Even though 100 kHz is sufficient for stochastic experiments of most small target molecules, biomacromolecular targets such as DNA or proteins require higher frequencies to distinguish internal composition.105 Achieving these performance goals will require careful device design that simultaneously maximizes sensitivity while minimizing parasitic capacitance.106
Nanoelectrode Design and Fabrication
Despite substantial improvements in nanofabrication, there is still an urgent need to develop methods to fabricate nanopore electrode structures with sub-10 nm feature sizes. Direct-write techniques, such as FIB and EBL, are limited to lab-scale structures. In addition, FIB milling implants conductive impurities, leading to current leakage problems, especially at high frequency.107,108 EBL is limited by e-beam scattering during exposure, the development of e-beam resists after exposure, and subsequent pattern transfer to the underlayer.109,110 Nanotemplate-based parallel processing approaches, e.g., nanoimprint lithography, nanosphere lithography, and block copolymer nanotemplates, are promising, but they need further development to provide high precision nanopatterning at production size scales.111−114
Specificity and Multiplex Sensing
Another significant factor affecting performance of nanopore electrode systems is the specificity between target and interferences. One straightforward approach is to exploit biomolecular recognition at the surface of a nanoelectrode.115,116 However, surface modification inside the ultrasmall confined volume of a nanopore is still tedious and inefficient, and nanopore electrode sensing constructs must ensure efficient electron transfer, even after surface modification—a particularly challenging problem for biorecognition motifs, such as enzymes.117,118 If these problems can be solved, then multiple sensing units may be realized within a single nanopore to effect vectorially coupled reactions, or alternatively to differentially modify different regions of nanopore arrays for high throughput multiplex sensing.
Next-Generation Nanopore Electrodes
Nanopore-based and nanopore electrode based sensors have benefited from the growth and maturation of nanotechnology. In linear succession, Wallace Coulter’s 1950s idea of counting particles in a fluid was followed by Richard Feynman’s oft-quoted 1959 essay “There is Plenty of Room at the Bottom”,119 which had a tremendous catalyzing impact on the scientific community. The technological developments flowing from these two seminal events were the intellectual ancestors to the human genome project, and after two decades the $1,000 genome has ushered in the era of personalized genomics and precision medicine.120 What will come next? Nanopore electrode systems are certainly poised to be integrated into contemporary point-of-care devices—not only reading DNA but identifying a range of proteomic and metabolomic biomarkers related to human health and wellness.121−123
Incorporating new passive and active electrode materials is one area for potential elaboration of nanopore electrode characteristics. Over the past decade, solid-state nanopores and two-dimensional nanopore arrays have advanced to exhibit excellent performance, in some cases competing with biological nanopores. However, the insertion of electronic components into nanopore systems has the potential to extend the contemporary capabilities to efficiently control molecular transport, directly monitor electron transfer processes, and rapidly record electrical signals.124−126 Gold and carbon are dominant electrode materials in nanopore electrode systems, and there are a number of interesting examples using carbon nanotubes or, more recently, graphene as electrode materials in nanopores.127−129 However, there are now a myriad of newly characterized two-dimensional materials, e.g., molybdenum disulfide (MoS2)130 and hexagonal boron nitride,131 promising candidates that exemplify new opportunities for nanopore sensing.
Lastly, the emergence of novel transport-reaction models suitable for application at the nanoscale and the development of powerful simulations together provide experimentalists with a direct way to predict the performance of new nanopore electrode sensors before testing as well as a way to assess performance afterward.132,133 The modeling of graphene-based nanopore sensors is just one example.134−136 Calculations can guide optimization of the number of graphene layers, pore diameter, and graft density of surface functional groups before fabricating graphene nanopores in the lab.
The topics highlighted here necessarily represent only a small fraction of the innovative work at the nexus where nanopore electrodes connect electron transfer and molecular control. Overall, there is a great deal of synergy in the opportunities before the nanopore community, and the exciting new directions that nanopore electrochemistry is poised to take should lead to a bright future and even more transformative surprises.
Acknowledgments
This work is supported by the National Science Foundation, Grant 1404744. We gratefully acknowledge Notre Dame Nanofabrication Facility and Integrated Imaging Facility for providing fabrication and structural characterization of the devices used in our studies in recent years. K.F. would like to thank Xin Zhang for her continued support of his research work.
The authors declare no competing financial interest.
References
- Larive C. K.; Sweedler J. V. Celebrating the 75th Anniversary of the ACS Division of Analytical Chemistry: A Special Collection of the Most Highly Cited Analytical Chemistry Papers Published between 1938 and 2012. Anal. Chem. 2013, 85 (9), 4201–4202. 10.1021/ac401048d. [DOI] [PubMed] [Google Scholar]
- Murray R. W. An Editor’s View of Analytical Chemistry (the Discipline). Annu. Rev. Anal. Chem. 2010, 3 (1), 1–18. 10.1146/annurev.anchem.111808.073555. [DOI] [PubMed] [Google Scholar]
- Sweedler J. V.; Armstrong D. W.; Baba Y.; Desmet G.; Dovichi N.; Ewing A.; Fenselau C. C.; Kennedy R. T.; Larive C. K.; Ligler F. S.; McCreery R. L.; Niessner R.; Pemberton J. E.; Tan W. H.; Walt D. R.; Yates J. R.; Zenobi R.; Zhang X. R. The Scope of Analytical Chemistry. Anal. Chem. 2015, 87 (13), 6425–6425. 10.1021/acs.analchem.5b02231. [DOI] [PubMed] [Google Scholar]
- Hou X.; Guo W.; Jiang L. Biomimetic smart nanopores and nanochannels. Chem. Soc. Rev. 2011, 40 (5), 2385–2401. 10.1039/c0cs00053a. [DOI] [PubMed] [Google Scholar]
- Coulter W. H.Means for counting particles suspended in a fluid. U.S. Patent 2,656,508 A, Oct 20, 1953.
- Bull B. S.; Schneiderman M. A.; Brecher G. Platelet counts with the Coulter counter. Am. J. Clin. Pathol. 1965, 44 (6), 678–688. 10.1093/ajcp/44.6.678. [DOI] [PubMed] [Google Scholar]
- DeBlois R. W.; Bean C. P. Counting and Sizing of Submicron Particles by the Resistive Pulse Technique. Rev. Sci. Instrum. 1970, 41, 909–916. 10.1063/1.1684724. [DOI] [Google Scholar]
- Simon S. M.; Blobel G. A protein-conducting channel in the endoplasmic reticulum. Cell 1991, 65 (3), 371–380. 10.1016/0092-8674(91)90455-8. [DOI] [PubMed] [Google Scholar]
- Bezrukov S. M.; Vodyanoy I.; Parsegian V. A. Counting Polymers Moving through a Single-Ion Channel. Nature 1994, 370, 279–281. 10.1038/370279a0. [DOI] [PubMed] [Google Scholar]
- Kasianowicz J. J.; Brandin E.; Branton D.; Deamer D. W. Characterization of Individual Polynucleotide Molecules Using a Membrane Channel. Proc. Natl. Acad. Sci. U. S. A. 1996, 93, 13770–13773. 10.1073/pnas.93.24.13770. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li J.; Stein D.; McMullan C.; Branton D.; Aziz M. J.; Golovchenko J. A. Ion-Beam Sculpting at Nanometre Length Scales. Nature 2001, 412, 166–169. 10.1038/35084037. [DOI] [PubMed] [Google Scholar]
- Storm A. J.; Chen J. H.; Ling X. S.; Zandbergen H. W.; Dekker C. Fabrication of solid-state nanopores with single-nanometre precision. Nat. Mater. 2003, 2 (8), 537–540. 10.1038/nmat941. [DOI] [PubMed] [Google Scholar]
- Ashkenasy N.; Sánchez-Quesada J.; Bayley H.; Ghadiri M. R. Recognizing a Single Base in an Individual DNA Strand: A Step Toward DNA Sequencing in Nanopores. Angew. Chem., Int. Ed. 2005, 44 (9), 1401–1404. 10.1002/anie.200462114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Clarke J.; Wu H. C.; Jayasinghe L.; Patel A.; Reid S.; Bayley H. Continuous base identification for single-molecule nanopore DNA sequencing. Nat. Nanotechnol. 2009, 4 (4), 265–270. 10.1038/nnano.2009.12. [DOI] [PubMed] [Google Scholar]
- Rice C. L.; Whitehead R. Electrokinetic Flow in a Narrow Cylindrical Capillary. J. Phys. Chem. 1965, 69 (11), 4017–4024. 10.1021/j100895a062. [DOI] [Google Scholar]
- Levine S.; Marriott J. R.; Robinson K. Theory of electrokinetic flow in a narrow parallel-plate channel. J. Chem. Soc., Faraday Trans. 2 1975, 71 (0), 1–11. 10.1039/f29757100001. [DOI] [Google Scholar]
- Tonucci R. J.; Justus B. L.; Campillo A. J.; Ford C. E. Nanochannel Array Glass. Science 1992, 258 (5083), 783–785. 10.1126/science.258.5083.783. [DOI] [PubMed] [Google Scholar]
- Masuda H.; Fukuda K. Ordered Metal Nanohole Arrays Made by a Two-Step Replication of Honeycomb Structures of Anodic Alumina. Science 1995, 268 (5216), 1466–1468. 10.1126/science.268.5216.1466. [DOI] [PubMed] [Google Scholar]
- Khandurina J.; Jacobson S. C.; Waters L. C.; Foote R. S.; Ramsey J. M. Microfabricated Porous Membrane Structure for Sample Concentration and Electrophoretic Analysis. Anal. Chem. 1999, 71 (9), 1815–1819. 10.1021/ac981161c. [DOI] [PubMed] [Google Scholar]
- Kuo T.-C.; Sloan L. A.; Sweedler J. V.; Bohn P. W. Manipulating Molecular Transport through Nanoporous Membranes by Control of Electrokinetic Flow: Effect of Surface Charge Density and Debye Length. Langmuir 2001, 17 (20), 6298–6303. 10.1021/la010429j. [DOI] [Google Scholar]
- Stein D.; Kruithof M.; Dekker C. Surface-charge-governed ion transport in nanofluidic channels. Phys. Rev. Lett. 2004, 93 (3), 035901. 10.1103/PhysRevLett.93.035901. [DOI] [PubMed] [Google Scholar]
- Karnik R.; Fan R.; Yue M.; Li D.; Yang P.; Majumdar A. Electrostatic Control of Ions and Molecules in Nanofluidic Transistors. Nano Lett. 2005, 5 (5), 943–948. 10.1021/nl050493b. [DOI] [PubMed] [Google Scholar]
- Karnik R.; Duan C.; Castelino K.; Daiguji H.; Majumdar A. Rectification of Ionic Current in a Nanofluidic Diode. Nano Lett. 2007, 7 (3), 547–551. 10.1021/nl062806o. [DOI] [PubMed] [Google Scholar]
- Hou X.; Guo W.; Xia F.; Nie F.-Q.; Dong H.; Tian Y.; Wen L.; Wang L.; Cao L.; Yang Y.; Xue J.; Song Y.; Wang Y.; Liu D.; Jiang L. A Biomimetic Potassium Responsive Nanochannel: G-Quadruplex DNA Conformational Switching in a Synthetic Nanopore. J. Am. Chem. Soc. 2009, 131 (22), 7800–7805. 10.1021/ja901574c. [DOI] [PubMed] [Google Scholar]
- Wehmeyer K. R.; Wightman R. M. Cyclic Voltammetry and Anodic-Stripping Voltammetry with Mercury Ultramicroelectrodes. Anal. Chem. 1985, 57 (9), 1989–1993. 10.1021/ac00286a046. [DOI] [Google Scholar]
- Bard A. J.; Crayston J. A.; Kittlesen G. P.; Varco Shea T.; Wrighton M. S. Digital-Simulation of the Measured Electrochemical Response of Reversible Redox Couples at Microelectrode Arrays - Consequences Arising from Closely Spaced Ultramicroelectrodes. Anal. Chem. 1986, 58 (11), 2321–2331. 10.1021/ac00124a045. [DOI] [Google Scholar]
- Morris R. B.; Franta D. J.; White H. S. Electrochemistry at platinum bane electrodes of width approaching molecular dimensions: breakdown of transport equations at very small electrodes. J. Phys. Chem. 1987, 91 (13), 3559–3564. 10.1021/j100297a019. [DOI] [Google Scholar]
- Martin C. R. Nanomaterials: a membrane-based synthetic approach. Science 1994, 266 (5193), 1961–1966. 10.1126/science.266.5193.1961. [DOI] [PubMed] [Google Scholar]
- Fan F. R.; Bard A. J. Electrochemical detection of single molecules. Science 1995, 267 (5199), 871–874. 10.1126/science.267.5199.871. [DOI] [PubMed] [Google Scholar]
- Chen S.; Ingram R. S.; Hostetler M. J.; Pietron J. J.; Murray R. W.; Schaaff T. G.; Khoury J. T.; Alvarez M. M.; Whetten R. L. Gold nanoelectrodes of varied size: transition to molecule-like charging. Science 1998, 280 (5372), 2098–2101. 10.1126/science.280.5372.2098. [DOI] [PubMed] [Google Scholar]
- Zhang B.; Zhang Y.; White H. S. The Nanopore Electrode. Anal. Chem. 2004, 76 (21), 6229–6238. 10.1021/ac049288r. [DOI] [PubMed] [Google Scholar]
- Lemay S. G.; van den Broek D. M.; Storm A. J.; Krapf D.; Smeets R. M. M.; Heering H. A.; Dekker C. Lithographically Fabricated Nanopore-Based Electrodes for Electrochemistry. Anal. Chem. 2005, 77 (6), 1911–1915. 10.1021/ac0489972. [DOI] [PubMed] [Google Scholar]
- Lanyon Y. H.; De Marzi G.; Watson Y. E.; Quinn A. J.; Gleeson J. P.; Redmond G.; Arrigan D. W. M. Fabrication of Nanopore Array Electrodes by Focused Ion Beam Milling. Anal. Chem. 2007, 79 (8), 3048–3055. 10.1021/ac061878x. [DOI] [PubMed] [Google Scholar]
- Arrigan D. W. M. Nanoelectrodes, nanoelectrode arrays and their applications. Analyst 2004, 129 (12), 1157–1165. 10.1039/b415395m. [DOI] [PubMed] [Google Scholar]
- Murray R. W. Nanoelectrochemistry: Metal Nanoparticles, Nanoelectrodes, and Nanopores. Chem. Rev. 2008, 108 (7), 2688–2720. 10.1021/cr068077e. [DOI] [PubMed] [Google Scholar]
- Bae J. H.; Han J.-H.; Chung T. D. Electrochemistry at nanoporous interfaces: new opportunity for electrocatalysis. Phys. Chem. Chem. Phys. 2012, 14 (2), 448–463. 10.1039/C1CP22927C. [DOI] [PubMed] [Google Scholar]
- Oja S. M.; Wood M.; Zhang B. Nanoscale Electrochemistry. Anal. Chem. 2013, 85 (2), 473–486. 10.1021/ac3031702. [DOI] [PubMed] [Google Scholar]
- Ying Y.-L.; Ding Z.; Zhan D.; Long Y.-T. Advanced electroanalytical chemistry at nanoelectrodes. Chem. Sci. 2017, 8 (5), 3338–3348. 10.1039/C7SC00433H. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dekker C. Solid-state nanopores. Nat. Nanotechnol. 2007, 2 (4), 209–215. 10.1038/nnano.2007.27. [DOI] [PubMed] [Google Scholar]
- Miles B. N.; Ivanov A. P.; Wilson K. A.; Dogan F.; Japrung D.; Edel J. B. Single molecule sensing with solid-state nanopores: novel materials, methods, and applications. Chem. Soc. Rev. 2013, 42 (1), 15–28. 10.1039/C2CS35286A. [DOI] [PubMed] [Google Scholar]
- Sun P.; Mirkin M. V. Electrochemistry of Individual Molecules in Zeptoliter Volumes. J. Am. Chem. Soc. 2008, 130 (26), 8241–8250. 10.1021/ja711088j. [DOI] [PubMed] [Google Scholar]
- Byers J. C.; Paulose Nadappuram B.; Perry D.; McKelvey K.; Colburn A. W.; Unwin P. R. Single Molecule Electrochemical Detection in Aqueous Solutions and Ionic Liquids. Anal. Chem. 2015, 87 (20), 10450–10456. 10.1021/acs.analchem.5b02569. [DOI] [PubMed] [Google Scholar]
- Siwy Z. S.; Howorka S. Engineered voltage-responsive nanopores. Chem. Soc. Rev. 2010, 39 (3), 1115–1132. 10.1039/B909105J. [DOI] [PubMed] [Google Scholar]
- Lan W. J.; Edwards M. A.; Luo L.; Perera R. T.; Wu X. J.; Martin C. R.; White H. S. Voltage-Rectified Current and Fluid Flow in Conical Nanopores. Acc. Chem. Res. 2016, 49 (11), 2605–2613. 10.1021/acs.accounts.6b00395. [DOI] [PubMed] [Google Scholar]
- van der Heyden F. H. J.; Bonthuis D. J.; Stein D.; Meyer C.; Dekker C. Power generation by pressure-driven transport of ions in nanofluidic channels. Nano Lett. 2007, 7 (4), 1022–1025. 10.1021/nl070194h. [DOI] [PubMed] [Google Scholar]
- van der Heyden F. H.; Stein D.; Dekker C. Streaming currents in a single nanofluidic channel. Phys. Rev. Lett. 2005, 95 (11), 116104. 10.1103/PhysRevLett.95.116104. [DOI] [PubMed] [Google Scholar]
- Kuchler A.; Yoshimoto M.; Luginbuhl S.; Mavelli F.; Walde P. Enzymatic reactions in confined environments. Nat. Nanotechnol. 2016, 11 (5), 409–420. 10.1038/nnano.2016.54. [DOI] [PubMed] [Google Scholar]
- Singh P. S.; Lemay S. G. Stochastic Processes in Electrochemistry. Anal. Chem. 2016, 88 (10), 5017–5027. 10.1021/acs.analchem.6b00683. [DOI] [PubMed] [Google Scholar]
- Garaj S.; Hubbard W.; Reina A.; Kong J.; Branton D.; Golovchenko J. A. Graphene as a subnanometre trans-electrode membrane. Nature 2010, 467, 190–193. 10.1038/nature09379. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wanunu M.; Dadosh T.; Ray V.; Jin J.; McReynolds L.; Drndić M. Rapid electronic detection of probe-specific microRNAs using thin nanopore sensors. Nat. Nanotechnol. 2010, 5, 807–814. 10.1038/nnano.2010.202. [DOI] [PubMed] [Google Scholar]
- Chou S. Y.; Krauss P. R.; Zhang W.; Guo L. J.; Zhuang L. Sub-10 nm imprint lithography and applications. J. Vac. Sci. Technol., B: Microelectron. Process. Phenom. 1997, 15 (6), 2897–2904. 10.1116/1.589752. [DOI] [Google Scholar]
- Im H.; Lee S. H.; Wittenberg N. J.; Johnson T. W.; Lindquist N. C.; Nagpal P.; Norris D. J.; Oh S.-H. Template-Stripped Smooth Ag Nanohole Arrays with Silica Shells for Surface Plasmon Resonance Biosensing. ACS Nano 2011, 5 (8), 6244–6253. 10.1021/nn202013v. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Whitney A. V.; Myers B. D.; Van Duyne R. P. Sub-100 nm Triangular Nanopores Fabricated with the Reactive Ion Etching Variant of Nanosphere Lithography and Angle-Resolved Nanosphere Lithography. Nano Lett. 2004, 4 (8), 1507–1511. 10.1021/nl049345w. [DOI] [Google Scholar]
- Jiang P.; McFarland M. J. Wafer-Scale Periodic Nanohole Arrays Templated from Two-Dimensional Nonclose-Packed Colloidal Crystals. J. Am. Chem. Soc. 2005, 127 (11), 3710–3711. 10.1021/ja042789+. [DOI] [PubMed] [Google Scholar]
- Thurn-Albrecht T.; Schotter J.; Kastle G. A.; Emley N.; Shibauchi T.; Krusin-Elbaum L.; Guarini K.; Black C. T.; Tuominen M. T.; Russell T. P. Ultrahigh-density nanowire arrays grown in self-assembled diblock copolymer templates. Science 2000, 290 (5499), 2126–2129. 10.1126/science.290.5499.2126. [DOI] [PubMed] [Google Scholar]
- Jung Y. S.; Ross C. A. Well-Ordered Thin-Film Nanopore Arrays Formed Using a Block-Copolymer Template. Small 2009, 5 (14), 1654–1659. 10.1002/smll.200900053. [DOI] [PubMed] [Google Scholar]
- Kustandi T. S.; Loh W. W.; Gao H.; Low H. Y. Wafer-Scale Near-Perfect Ordered Porous Alumina on Substrates by Step and Flash Imprint Lithography. ACS Nano 2010, 4 (5), 2561–2568. 10.1021/nn1001744. [DOI] [PubMed] [Google Scholar]
- Lee W.; Park S.-J. Porous Anodic Aluminum Oxide: Anodization and Templated Synthesis of Functional Nanostructures. Chem. Rev. 2014, 114 (15), 7487–7556. 10.1021/cr500002z. [DOI] [PubMed] [Google Scholar]
- Ali M.; Ramirez P.; Mafe S.; Neumann R.; Ensinger W. A pH-Tunable Nanofluidic Diode with a Broad Range of Rectifying Properties. ACS Nano 2009, 3 (3), 603–608. 10.1021/nn900039f. [DOI] [PubMed] [Google Scholar]
- Zhang Z.; Xie G.; Xiao K.; Kong X.-Y.; Li P.; Tian Y.; Wen L.; Jiang L. Asymmetric Multifunctional Heterogeneous Membranes for pH- and Temperature-Cooperative Smart Ion Transport Modulation. Adv. Mater. 2016, 28 (43), 9613–9619. 10.1002/adma.201602758. [DOI] [PubMed] [Google Scholar]
- Zhang M.; Hou X.; Wang J.; Tian Y.; Fan X.; Zhai J.; Jiang L. Light and pH Cooperative Nanofluidic Diode Using a Spiropyran-Functionalized Single Nanochannel. Adv. Mater. 2012, 24 (18), 2424–2428. 10.1002/adma.201104536. [DOI] [PubMed] [Google Scholar]
- Siwy Z.; Trofin L.; Kohli P.; Baker L. A.; Trautmann C.; Martin C. R. Protein Biosensors Based on Biofunctionalized Conical Gold Nanotubes. J. Am. Chem. Soc. 2005, 127 (14), 5000–5001. 10.1021/ja043910f. [DOI] [PubMed] [Google Scholar]
- Abelow A. E.; Schepelina O.; White R. J.; Vallee-Belisle A.; Plaxco K. W.; Zharov I. Biomimetic glass nanopores employing aptamer gates responsive to a small molecule. Chem. Commun. 2010, 46 (42), 7984–7986. 10.1039/c0cc02649b. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Watkins J. J.; Chen J.; White H. S.; Abruña H. D.; Maisonhaute E.; Amatore C. Zeptomole Voltammetric Detection and Electron-Transfer Rate Measurements Using Platinum Electrodes of Nanometer Dimensions. Anal. Chem. 2003, 75 (16), 3962–3971. 10.1021/ac0342931. [DOI] [PubMed] [Google Scholar]
- Zhang J.; Kuznetsov A. M.; Medvedev I. G.; Chi Q.; Albrecht T.; Jensen P. S.; Ulstrup J. Single-Molecule Electron Transfer in Electrochemical Environments. Chem. Rev. 2008, 108 (7), 2737–2791. 10.1021/cr068073+. [DOI] [PubMed] [Google Scholar]
- Lemay S. G.; Kang S.; Mathwig K.; Singh P. S. Single-Molecule Electrochemistry: Present Status and Outlook. Acc. Chem. Res. 2013, 46 (2), 369–377. 10.1021/ar300169d. [DOI] [PubMed] [Google Scholar]
- Hoeben F. J. M.; Meijer F. S.; Dekker C.; Albracht S. P. J.; Heering H. A.; Lemay S. G. Toward Single-Enzyme Molecule Electrochemistry: [NiFe]-Hydrogenase Protein Film Voltammetry at Nanoelectrodes. ACS Nano 2008, 2 (12), 2497–2504. 10.1021/nn800518d. [DOI] [PubMed] [Google Scholar]
- Rassaei L.; Mathwig K.; Kang S.; Heering H. A.; Lemay S. G. Integrated Biodetection in a Nanofluidic Device. ACS Nano 2014, 8 (8), 8278–8284. 10.1021/nn502678t. [DOI] [PubMed] [Google Scholar]
- Li Y.; Cox J. T.; Zhang B. Electrochemical Responses and Electrocatalysis at Single Au Nanoparticles. J. Am. Chem. Soc. 2010, 132 (9), 3047–3054. 10.1021/ja909408q. [DOI] [PubMed] [Google Scholar]
- Shan X.; Díez-Pérez I.; Wang L.; Wiktor P.; Gu Y.; Zhang L.; Wang W.; Lu J.; Wang S.; Gong Q.; Li J.; Tao N. Imaging the electrocatalytic activity of single nanoparticles. Nat. Nanotechnol. 2012, 7, 668. 10.1038/nnano.2012.134. [DOI] [PubMed] [Google Scholar]
- Phan N. T. N.; Li X.; Ewing A. G. Measuring synaptic vesicles using cellular electrochemistry and nanoscale molecular imaging. Nat. Rev. Chem. 2017, 1, 0048. 10.1038/s41570-017-0048. [DOI] [Google Scholar]
- Wang K.; Xiao T.; Yue Q.; Wu F.; Yu P.; Mao L. Selective Amperometric Recording of Endogenous Ascorbate Secretion from a Single Rat Adrenal Chromaffin Cell with Pretreated Carbon Fiber Microelectrodes. Anal. Chem. 2017, 89 (17), 9502–9507. 10.1021/acs.analchem.7b02508. [DOI] [PubMed] [Google Scholar]
- Xiao X.; Bard A. J. Observing Single Nanoparticle Collisions at an Ultramicroelectrode by Electrocatalytic Amplification. J. Am. Chem. Soc. 2007, 129 (31), 9610–9612. 10.1021/ja072344w. [DOI] [PubMed] [Google Scholar]
- Zhou Y.-G.; Rees N. V.; Compton R. G. The Electrochemical Detection and Characterization of Silver Nanoparticles in Aqueous Solution. Angew. Chem., Int. Ed. 2011, 50 (18), 4219–4221. 10.1002/anie.201100885. [DOI] [PubMed] [Google Scholar]
- Dasari R.; Robinson D. A.; Stevenson K. J. Ultrasensitive Electroanalytical Tool for Detecting, Sizing, and Evaluating the Catalytic Activity of Platinum Nanoparticles. J. Am. Chem. Soc. 2013, 135 (2), 570–573. 10.1021/ja310614x. [DOI] [PubMed] [Google Scholar]
- Fosdick S. E.; Anderson M. J.; Nettleton E. G.; Crooks R. M. Correlated Electrochemical and Optical Tracking of Discrete Collision Events. J. Am. Chem. Soc. 2013, 135 (16), 5994–5997. 10.1021/ja401864k. [DOI] [PubMed] [Google Scholar]
- Zhou M.; Yu Y.; Hu K.; Xin H. L.; Mirkin M. V. Collisions of Ir Oxide Nanoparticles with Carbon Nanopipettes: Experiments with One Nanoparticle. Anal. Chem. 2017, 89 (5), 2880–2885. 10.1021/acs.analchem.6b04140. [DOI] [PubMed] [Google Scholar]
- Gao R.; Ying Y.-L.; Li Y.-J.; Hu Y.-X.; Yu R.-J.; Lin Y.; Long Y.-T. A 30 nm Nanopore Electrode: Facile Fabrication and Direct Insights into the Intrinsic Feature of Single Nanoparticle Collisions. Angew. Chem., Int. Ed. 2017, 10.1002/anie.201710201. [DOI] [PubMed] [Google Scholar]
- Gao R.; Lin Y.; Ying Y.-L.; Liu X.-Y.; Shi X.; Hu Y.-X.; Long Y.-T.; Tian H. Dynamic Self-Assembly of Homogenous Microcyclic Structures Controlled by a Silver-Coated Nanopore. Small 2017, 13 (25), 1700234. 10.1002/smll.201700234. [DOI] [PubMed] [Google Scholar]
- Yu Y.; Sundaresan V.; Bandyopadhyay S.; Zhang Y.; Edwards M. A.; McKelvey K.; White H. S.; Willets K. A. Three-Dimensional Super-resolution Imaging of Single Nanoparticles Delivered by Pipettes. ACS Nano 2017, 11 (10), 10529–10538. 10.1021/acsnano.7b05902. [DOI] [PubMed] [Google Scholar]
- Mathwig K.; Aartsma T. J.; Canters G. W.; Lemay S. G. Nanoscale Methods for Single-Molecule Electrochemistry. Annu. Rev. Anal. Chem. 2014, 7 (1), 383–404. 10.1146/annurev-anchem-062012-092557. [DOI] [PubMed] [Google Scholar]
- Brasiliense V.; Berto P.; Combellas C.; Tessier G.; Kanoufi F. Electrochemistry of Single Nanodomains Revealed by Three-Dimensional Holographic Microscopy. Acc. Chem. Res. 2016, 49 (9), 2049–2057. 10.1021/acs.accounts.6b00335. [DOI] [PubMed] [Google Scholar]
- Clausmeyer J.; Schuhmann W. Nanoelectrodes: Applications in electrocatalysis, single-cell analysis and high-resolution electrochemical imaging. TrAC, Trends Anal. Chem. 2016, 79, 46–59. 10.1016/j.trac.2016.01.018. [DOI] [Google Scholar]
- Wang Y.; Shan X.; Tao N. Emerging tools for studying single entity electrochemistry. Faraday Discuss. 2016, 193 (0), 9–39. 10.1039/C6FD00180G. [DOI] [PubMed] [Google Scholar]
- LaFratta C. N.; Walt D. R. Very high density sensing arrays. Chem. Rev. 2008, 108 (2), 614–637. 10.1021/cr0681142. [DOI] [PubMed] [Google Scholar]
- Wolfrum B.; Kätelhön E.; Yakushenko A.; Krause K. J.; Adly N.; Hüske M.; Rinklin P. Nanoscale Electrochemical Sensor Arrays: Redox Cycling Amplification in Dual-Electrode Systems. Acc. Chem. Res. 2016, 49 (9), 2031–2040. 10.1021/acs.accounts.6b00333. [DOI] [PubMed] [Google Scholar]
- Ma C.; Contento N. M.; Gibson L. R.; Bohn P. W. Redox Cycling in Nanoscale-Recessed Ring-Disk Electrode Arrays for Enhanced Electrochemical Sensitivity. ACS Nano 2013, 7 (6), 5483–5490. 10.1021/nn401542x. [DOI] [PubMed] [Google Scholar]
- Fu K.; Han D.; Ma C.; Bohn P. W. Electrochemistry at single molecule occupancy in nanopore-confined recessed ring-disk electrode arrays. Faraday Discuss. 2016, 193 (0), 51–64. 10.1039/C6FD00062B. [DOI] [PubMed] [Google Scholar]
- Ma C.; Contento N. M.; Bohn P. W. Redox Cycling on Recessed Ring-Disk Nanoelectrode Arrays in the Absence of Supporting Electrolyte. J. Am. Chem. Soc. 2014, 136 (20), 7225–7228. 10.1021/ja502052s. [DOI] [PubMed] [Google Scholar]
- Ma C.; Xu W.; Wichert W. R. A.; Bohn P. W. Ion Accumulation and Migration Effects on Redox Cycling in Nanopore Electrode Arrays at Low Ionic Strength. ACS Nano 2016, 10 (3), 3658–3664. 10.1021/acsnano.6b00049. [DOI] [PubMed] [Google Scholar]
- Fu K.; Han D.; Ma C.; Bohn P. W. Ion selective redox cycling in zero-dimensional nanopore electrode arrays at low ionic strength. Nanoscale 2017, 9 (16), 5164–5171. 10.1039/C7NR00206H. [DOI] [PubMed] [Google Scholar]
- Ma C.; Zaino Iii L. P.; Bohn P. W. Self-induced redox cycling coupled luminescence on nanopore recessed disk-multiscale bipolar electrodes. Chem. Sci. 2015, 6 (5), 3173–3179. 10.1039/C5SC00433K. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fu K.; Bohn P. W. Nanochannel Arrays for Molecular Sieving and Electrochemical Analysis by Nanosphere Lithography Templated Graphoepitaxy of Block Copolymers. ACS Appl. Mater. Interfaces 2017, 9 (29), 24908–24916. 10.1021/acsami.7b06794. [DOI] [PubMed] [Google Scholar]
- Cortés E.; Etchegoin P. G.; Le Ru E. C.; Fainstein A.; Vela M. E.; Salvarezza R. C. Monitoring the Electrochemistry of Single Molecules by Surface-Enhanced Raman Spectroscopy. J. Am. Chem. Soc. 2010, 132 (51), 18034–18037. 10.1021/ja108989b. [DOI] [PubMed] [Google Scholar]
- Shan X.; Patel U.; Wang S.; Iglesias R.; Tao N. Imaging local electrochemical current via surface plasmon resonance. Science 2010, 327 (5971), 1363–1366. 10.1126/science.1186476. [DOI] [PubMed] [Google Scholar]
- Liu Z.; Ding S.-Y.; Chen Z.-B.; Wang X.; Tian J.-H.; Anema J. R.; Zhou X.-S.; Wu D.-Y.; Mao B.-W.; Xu X.; Ren B.; Tian Z.-Q. Revealing the molecular structure of single-molecule junctions in different conductance states by fishing-mode tip-enhanced Raman spectroscopy. Nat. Commun. 2011, 2, 305. 10.1038/ncomms1310. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang W.; Caldarola M.; Pradhan B.; Orrit M. Gold Nanorod Enhanced Fluorescence Enables Single-Molecule Electrochemistry of Methylene Blue. Angew. Chem., Int. Ed. 2017, 56 (13), 3566–3569. 10.1002/anie.201612389. [DOI] [PubMed] [Google Scholar]
- Levene M. J.; Korlach J.; Turner S. W.; Foquet M.; Craighead H. G.; Webb W. W. Zero-mode waveguides for single-molecule analysis at high concentrations. Science 2003, 299 (5607), 682–686. 10.1126/science.1079700. [DOI] [PubMed] [Google Scholar]
- Goldschen-Ohm M. P.; White D. S.; Klenchin V. A.; Chanda B.; Goldsmith R. H. Observing Single-Molecule Dynamics at Millimolar Concentrations. Angew. Chem., Int. Ed. 2017, 56 (9), 2399–2402. 10.1002/anie.201612050. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Han D.; Zaino L. P.; Fu K.; Bohn P. W. Redox Cycling in Nanopore-Confined Recessed Dual-Ring Electrode Arrays. J. Phys. Chem. C 2016, 120 (37), 20634–20641. 10.1021/acs.jpcc.6b01287. [DOI] [Google Scholar]
- Zhao J.; Zaino L. P. III; Bohn P. W. Potential-dependent single molecule blinking dynamics for flavin adenine dinucleotide covalently immobilized in zero-mode waveguide array of working electrodes. Faraday Discuss. 2013, 164, 57–69. 10.1039/c3fd00013c. [DOI] [PubMed] [Google Scholar]
- Zaino L. P.; Grismer D. A.; Han D.; Crouch G. M.; Bohn P. W. Single occupancy spectroelectrochemistry of freely diffusing flavin mononucleotide in zero-dimensional nanophotonic structures. Faraday Discuss. 2015, 184 (0), 101–115. 10.1039/C5FD00072F. [DOI] [PubMed] [Google Scholar]
- Han D.; Crouch G.; Fu K.; Zaino Iii L. P.; Bohn P. W. Single-molecule spectroelectrochemical cross-correlation during redox cycling in recessed dual ring electrode zero-mode waveguides. Chem. Sci. 2017, 8 (8), 5345–5355. 10.1039/C7SC02250F. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shi W.; Friedman A. K.; Baker L. A. Nanopore Sensing. Anal. Chem. 2017, 89 (1), 157–188. 10.1021/acs.analchem.6b04260. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lindsay S. The promises and challenges of solid-state sequencing. Nat. Nanotechnol. 2016, 11 (2), 109–111. 10.1038/nnano.2016.9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dimitrov V.; Mirsaidov U.; Wang D.; Sorsch T.; Mansfield W.; Miner J.; Klemens F.; Cirelli R.; Yemenicioglu S.; Timp G. Nanopores in solid-state membranes engineered for single molecule detection. Nanotechnology 2010, 21 (6), 065502. 10.1088/0957-4484/21/6/065502. [DOI] [PubMed] [Google Scholar]
- Tseng A. A. Recent developments in micromilling using focused ion beam technology. J. Micromech. Microeng. 2004, 14 (4), R15–R34. 10.1088/0960-1317/14/4/R01. [DOI] [Google Scholar]
- MoberlyChan W. J.; Adams D. P.; Aziz M. J.; Hobler G.; Schenkel T. Fundamentals of Focused Ion Beam Nanostructural Processing: Below, At, and Above the Surface. MRS Bull. 2007, 32 (5), 424–432. 10.1557/mrs2007.66. [DOI] [Google Scholar]
- Grigorescu A. E.; Hagen C. W. Resists for sub-20-nm electron beam lithography with a focus on HSQ: state of the art. Nanotechnology 2009, 20 (29), 292001. 10.1088/0957-4484/20/29/292001. [DOI] [PubMed] [Google Scholar]
- Chen Y. Nanofabrication by electron beam lithography and its applications: A review. Microelectron. Eng. 2015, 135, 57–72. 10.1016/j.mee.2015.02.042. [DOI] [Google Scholar]
- Henzie J.; Barton J. E.; Stender C. L.; Odom T. W. Large-Area Nanoscale Patterning: Chemistry Meets Fabrication. Acc. Chem. Res. 2006, 39 (4), 249–257. 10.1021/ar050013n. [DOI] [PubMed] [Google Scholar]
- Stewart M. E.; Anderton C. R.; Thompson L. B.; Maria J.; Gray S. K.; Rogers J. A.; Nuzzo R. G. Nanostructured Plasmonic Sensors. Chem. Rev. 2008, 108 (2), 494–521. 10.1021/cr068126n. [DOI] [PubMed] [Google Scholar]
- Jones M. R.; Osberg K. D.; Macfarlane R. J.; Langille M. R.; Mirkin C. A. Templated Techniques for the Synthesis and Assembly of Plasmonic Nanostructures. Chem. Rev. 2011, 111 (6), 3736–3827. 10.1021/cr1004452. [DOI] [PubMed] [Google Scholar]
- Claridge S. A.; Liao W.-S.; Thomas J. C.; Zhao Y.; Cao H. H.; Cheunkar S.; Serino A. C.; Andrews A. M.; Weiss P. S. From the bottom up: dimensional control and characterization in molecular monolayers. Chem. Soc. Rev. 2013, 42 (7), 2725–2745. 10.1039/C2CS35365B. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zen J. M.; Kumar A. S.; Tsai D. M. Recent updates of chemically modified electrodes in analytical chemistry. Electroanalysis 2003, 15 (13), 1073–1087. 10.1002/elan.200390130. [DOI] [Google Scholar]
- Labib M.; Sargent E. H.; Kelley S. O. Electrochemical Methods for the Analysis of Clinically Relevant Biomolecules. Chem. Rev. 2016, 116 (16), 9001–9090. 10.1021/acs.chemrev.6b00220. [DOI] [PubMed] [Google Scholar]
- Léger C.; Bertrand P. Direct Electrochemistry of Redox Enzymes as a Tool for Mechanistic Studies. Chem. Rev. 2008, 108 (7), 2379–2438. 10.1021/cr0680742. [DOI] [PubMed] [Google Scholar]
- Noll T.; Noll G. Strategies for ″wiring″ redox-active proteins to electrodes and applications in biosensors, biofuel cells, and nanotechnology. Chem. Soc. Rev. 2011, 40 (7), 3564–3576. 10.1039/c1cs15030h. [DOI] [PubMed] [Google Scholar]
- Feynman R. P. There's Plenty of Room at the Bottom. Eng. Sci. 1960, 23, 22–36. [Google Scholar]
- Hayden E. C. The $1,000 genome. Nature 2014, 507 (7492), 294–295. 10.1038/507294a. [DOI] [PubMed] [Google Scholar]
- Wei R.; Gatterdam V.; Wieneke R.; Tampé R.; Rant U. Stochastic sensing of proteins with receptor-modified solid-state nanopores. Nat. Nanotechnol. 2012, 7, 257–263. 10.1038/nnano.2012.24. [DOI] [PubMed] [Google Scholar]
- Ohshiro T.; Tsutsui M.; Yokota K.; Furuhashi M.; Taniguchi M.; Kawai T. Detection of post-translational modifications in single peptides using electron tunnelling currents. Nat. Nanotechnol. 2014, 9, 835–840. 10.1038/nnano.2014.193. [DOI] [PubMed] [Google Scholar]
- Zhao Y.; Ashcroft B.; Zhang P.; Liu H.; Sen S.; Song W.; Im J.; Gyarfas B.; Manna S.; Biswas S.; Borges C.; Lindsay S. Single-molecule spectroscopy of amino acids and peptides by recognition tunnelling. Nat. Nanotechnol. 2014, 9, 466–473. 10.1038/nnano.2014.54. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nam S. W.; Rooks M. J.; Kim K. B.; Rossnagel S. M. Ionic field effect transistors with sub-10 nm multiple nanopores. Nano Lett. 2009, 9 (5), 2044–2048. 10.1021/nl900309s. [DOI] [PubMed] [Google Scholar]
- Rutkowska A.; Edel J. B.; Albrecht T. Mapping the ion current distribution in nanopore/electrode devices. ACS Nano 2013, 7 (1), 547–555. 10.1021/nn304695y. [DOI] [PubMed] [Google Scholar]
- Ren R.; Zhang Y.; Nadappuram B. P.; Akpinar B.; Klenerman D.; Ivanov A. P.; Edel J. B.; Korchev Y. Nanopore extended field-effect transistor for selective single-molecule biosensing. Nat. Commun. 2017, 8 (1), 586. 10.1038/s41467-017-00549-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Merchant C. A.; Healy K.; Wanunu M.; Ray V.; Peterman N.; Bartel J.; Fischbein M. D.; Venta K.; Luo Z.; Johnson A. T. C.; Drndić M. DNA Translocation through Graphene Nanopores. Nano Lett. 2010, 10 (8), 2915–2921. 10.1021/nl101046t. [DOI] [PubMed] [Google Scholar]
- Traversi F.; Raillon C.; Benameur S. M.; Liu K.; Khlybov S.; Tosun M.; Krasnozhon D.; Kis A.; Radenovic A. Detecting the translocation of DNA through a nanopore using graphene nanoribbons. Nat. Nanotechnol. 2013, 8, 939–945. 10.1038/nnano.2013.240. [DOI] [PubMed] [Google Scholar]
- Heerema S. J.; Dekker C. Graphene nanodevices for DNA sequencing. Nat. Nanotechnol. 2016, 11, 127–136. 10.1038/nnano.2015.307. [DOI] [PubMed] [Google Scholar]
- Feng J.; Liu K.; Bulushev R. D.; Khlybov S.; Dumcenco D.; Kis A.; Radenovic A. Identification of single nucleotides in MoS2 nanopores. Nat. Nanotechnol. 2015, 10 (12), 1070–1076. 10.1038/nnano.2015.219. [DOI] [PubMed] [Google Scholar]
- Liu S.; Lu B.; Zhao Q.; Li J.; Gao T.; Chen Y.; Zhang Y.; Liu Z.; Fan Z.; Yang F.; You L.; Yu D. Boron Nitride Nanopores: Highly Sensitive DNA Single-Molecule Detectors. Adv. Mater. 2013, 25 (33), 4549–4554. 10.1002/adma.201301336. [DOI] [PubMed] [Google Scholar]
- Momotenko D.; Cortes-Salazar F.; Josserand J.; Liu S.; Shao Y.; Girault H. H. Ion current rectification and rectification inversion in conical nanopores: a perm-selective view. Phys. Chem. Chem. Phys. 2011, 13 (12), 5430–5440. 10.1039/c0cp02595j. [DOI] [PubMed] [Google Scholar]
- Yusko E. C.; Bruhn B. R.; Eggenberger O. M.; Houghtaling J.; Rollings R. C.; Walsh N. C.; Nandivada S.; Pindrus M.; Hall A. R.; Sept D.; Li J.; Kalonia D. S.; Mayer M. Real-time shape approximation and fingerprinting of single proteins using a nanopore. Nat. Nanotechnol. 2017, 12, 360–367. 10.1038/nnano.2016.267. [DOI] [PubMed] [Google Scholar]
- Wells D. B.; Belkin M.; Comer J.; Aksimentiev A. Assessing Graphene Nanopores for Sequencing DNA. Nano Lett. 2012, 12 (8), 4117–4123. 10.1021/nl301655d. [DOI] [PMC free article] [PubMed] [Google Scholar]
- McFarland H. L.; Ahmed T.; Zhu J.-X.; Balatsky A. V.; Haraldsen J. T. First-Principles Investigation of Nanopore Sequencing Using Variable Voltage Bias on Graphene-Based Nanoribbons. J. Phys. Chem. Lett. 2015, 6 (13), 2616–2621. 10.1021/acs.jpclett.5b01014. [DOI] [PubMed] [Google Scholar]
- Qiu H.; Sarathy A.; Leburton J.-P.; Schulten K. Intrinsic Stepwise Translocation of Stretched ssDNA in Graphene Nanopores. Nano Lett. 2015, 15 (12), 8322–8330. 10.1021/acs.nanolett.5b03963. [DOI] [PMC free article] [PubMed] [Google Scholar]


