Probing surface charge fluctuations with solid-state nanopores

Similar documents
Improving signal-to-noise performance for DNA translocation in solid-state nanopores at MHz bandwidths

Origin of the Electrophoretic Force on DNA in Nanopores. Biological and Soft Systems - Cavendish Laboratory

Nanopores and Nanofluidics for Single DNA Studies Derek Stein Department of Physics Brown University

Graphene: A sub-nanometer trans-electrode membrane

Super-sensitive Molecule-hugging Graphene Nanopores

Simulation of ionic current through the nanopore in a double-layered semiconductor

Numerical Modeling of the Bistability of Electrolyte Transport in Conical Nanopores

SUPPLEMENTARY INFORMATION

Determining the Elastic Modulus and Hardness of an Ultrathin Film on a Substrate Using Nanoindentation

Origin of the Electrophoretic Force on DNA in a Nanopore

Salt Dependence of Ion Transport and DNA Translocation through Solid-State Nanopores

PCCP PAPER. Investigation of field effects in a solid-state nanopore transistor. Introduction. Yong Youn and Seungwu Han*

Detecting Single Stranded DNA with a Solid State Nanopore

Nanopores: Solid-state nanopores for these experiments were produced by using the

Graphene photodetectors with ultra-broadband and high responsivity at room temperature

(a) (b) Supplementary Figure 1. (a) (b) (a) Supplementary Figure 2. (a) (b) (c) (d) (e)

Electrolyte Concentration Dependence of Ion Transport through Nanochannels

Non-Faradaic Impedance Characterization of an

SUPPLEMENTARY NOTES Supplementary Note 1: Fabrication of Scanning Thermal Microscopy Probes

Voc impact of orientation-dependent x in anisotropic PV absorbers.

3/24/11. Introduction! Electrogenic cell

2.6 The Membrane Potential

Analyzing Ion channel Simulations

Probing Access Resistance of Solid-State Nanopores with a Scanning-Probe Microscope Tip

Energy position of the active near-interface traps in metal oxide semiconductor field-effect transistors on 4H SiC

Supplementary information for

Observation of ionic Coulomb blockade in nanopores

Resistance Thermometry based Picowatt-Resolution Heat-Flow Calorimeter

Spatially resolving density-dependent screening around a single charged atom in graphene

Atomic Layer Deposition to Fine- Tune the Surface Properties and Diameters of Fabricated Nanopores

Supporting Information. Probing DNA Translocations with Inplane Current Signals in a Graphene Nanoribbon with a Nanopore

SUPPLEMENTARY INFORMATION

Exploiting pattern transformation to tune phononic band gaps in a two-dimensional granular crystal

6.5 mm. ε = 1%, r = 9.4 mm. ε = 3%, r = 3.1 mm

INTERMOLECULAR AND SURFACE FORCES

Supplementary Figure 2 Photoluminescence in 1L- (black line) and 7L-MoS 2 (red line) of the Figure 1B with illuminated wavelength of 543 nm.

A General Equation for Fitting Contact Area and Friction vs Load Measurements

SUPPLEMENTARY INFORMATION

M R S Internet Journal of Nitride Semiconductor Research

(name) Electrochemical Energy Systems, Spring 2014, M. Z. Bazant. Final Exam

Lecture 3 Charged interfaces

There's Plenty of Room at the Bottom

STRUCTURE OF IONS AND WATER AROUND A POLYELECTROLYTE IN A POLARIZABLE NANOPORE

PREPARATION OF LUMINESCENT SILICON NANOPARTICLES BY PHOTOTHERMAL AEROSOL SYNTHESIS FOLLOWED BY ACID ETCHING

Demystifying Transmission Lines: What are They? Why are They Useful?

Probing into the Electrical Double Layer Using a Potential Nano-Probe

New experimental evidence for the role of long- range potential fluctuations in the mechanism of 1/f noise in a-si:h

Scanning capacitance spectroscopy of an Al x Ga 1Àx NÕGaN heterostructure field-effect transistor structure: Analysis of probe tip effects

Supporting Information. by Hexagonal Boron Nitride

Electronic transport in low dimensional systems

Nanopore sculpting with noble gas ions

Electro-Thermal Transport in Silicon and Carbon Nanotube Devices E. Pop, D. Mann, J. Rowlette, K. Goodson and H. Dai

The Sensitivity Limits of Nanowire Biosensors

PIC/MCC Simulation of Radio Frequency Hollow Cathode Discharge in Nitrogen

Basic Laboratory. Materials Science and Engineering. Atomic Force Microscopy (AFM)

Scanning Tunneling Microscopy

arxiv:cond-mat/ v1 [cond-mat.soft] 9 Aug 1997

Carbon Nanotube Synaptic Transistor Network for. Pattern Recognition. Supporting Information for

SUPPLEMENTARY INFORMATION

Nanofluidics and 2D Materials Based Nanosensors. Ivan Vlassiouk Oak Ridge National Laboratory, TN, USA

V. Electrostatics Lecture 24: Diffuse Charge in Electrolytes

Photovoltage phenomena in nanoscaled materials. Thomas Dittrich Hahn-Meitner-Institute Berlin

pickup from external sources unwanted feedback RF interference from system or elsewhere, power supply fluctuations ground currents

Time Scaling with System Size for Resistor-Only Warm Up for Photovoltaic Device Simulation

This article has been accepted and published on J-STAGE in advance of copyediting. Content is final as presented.

ECE 340 Lecture 39 : MOS Capacitor II

General Physics. Nerve Conduction. Newton s laws of Motion Work, Energy and Power. Fluids. Direct Current (DC)

Dependence of Potential and Ion Distribution on Electrokinetic Radius in Infinite and Finite-length Nano-channels

SUPPLEMENTARY INFORMATION

Nanopore sculpting with noble gas ions

single-electron electron tunneling (SET)

SUPPLEMENTARY INFORMATION

A. Optimizing the growth conditions of large-scale graphene films

Solving the Poisson Boltzmann equation to obtain interaction energies between confined, like-charged cylinders

Comment on: Depolarization corrections to the coercive field in thin-film ferroelectrics. Abstract

Characterization of a high-performance Ti/Au TES microcalorimeter with a central Cu absorber

Lognormal distribution of firing time and rate from a single neuron?

CH676 Physical Chemistry: Principles and Applications. CH676 Physical Chemistry: Principles and Applications

Module 8: "Stability of Colloids" Lecture 38: "" The Lecture Contains: Calculation for CCC (n c )

Charge spectrometry with a strongly coupled superconducting single-electron transistor

Nanoscale electrochemistry

Supporting Information

Characterization of a MEMS BioChip for planar patch-clamp recording

Optimization of the Dielectric Constant of a Blocking Dielectric in the Nonvolatile Memory Based on Silicon Nitride

Electronic Supplementary Information

CLARKSON UNIVERSITY. Poisson-Nernst-Planck Model for Simulating Ionic Current through the Nanopore in a Semiconductor Membrane

M. Maliński M. Pawlak Ł. Chrobak S. Pal A. Ludwig

Supporting Information for Conical Nanopores. for Efficient Ion Pumping and Desalination

Electron Interferometer Formed with a Scanning Probe Tip and Quantum Point Contact Supplementary Information

Title of file for HTML: Supplementary Information Description: Supplementary Figures and Supplementary References

Citation PHYSICAL REVIEW LETTERS (2000), 85( RightCopyright 2000 American Physical So

Chapter 3 Properties of Nanostructures

MICRO-SCALE SHEET RESISTANCE MEASUREMENTS ON ULTRA SHALLOW JUNCTIONS

Electrokinetic assembly and manipulation II Lecture by Chung, Jae-Hyun

Effect of Spiral Microwave Antenna Configuration on the Production of Nano-crystalline Film by Chemical Sputtering in ECR Plasma

Ions in concentrated electrolytes: from the screening length to energy storage

Specific ion effects on the interaction of. hydrophobic and hydrophilic self assembled

Written by David Niedzwiecki ( )

ANALYSIS OF DIELECTRIC, MODULUS, ELECTRO CHEMICAL STABILITY OF PVP ABSA POLYMER ELECTROLYTE SYSTEMS

SUPPLEMENTARY INFORMATION

Transcription:

Probing surface charge fluctuations with solid-state nanopores The Harvard community has made this article openly available. Please share how this access benefits you. Your story matters. Citation Published Version Accessed Citable Link Terms of Use Hoogerheide, David P., Slaven Garaj, and Jene A. Golovchenko. 2009. Probing surface charge fluctuations with solid-state nanopores. Physical Review Letters 102(25): 256804. doi:10.1103/physrevlett.102.256804 November 17, 2017 2:35:49 PM EST http://nrs.harvard.edu/urn-3:hul.instrepos:3293523 This article was downloaded from Harvard University's DASH repository, and is made available under the terms and conditions applicable to Other Posted Material, as set forth at http://nrs.harvard.edu/urn-3:hul.instrepos:dash.current.terms-ofuse#laa (Article begins on next page)

Probing Surface Charge Fluctuations with Solid-State Nanopores David P. Hoogerheide, Slaven Garaj, and Jene A. Golovchenko Physics Department, Harvard University, Cambridge Massachusetts 02138, USA (Received 8 April 2009; published 26 June 2009) We identify a contribution to the ionic current noise spectrum in solid-state nanopores that exceeds all other noise sources in the frequency band 0.1 10 khz. Experimental studies of the dependence of this excess noise on ph and electrolyte concentration indicate that the noise arises from surface charge fluctuations. A quantitative model based on surface functional group protonization predicts the observed behaviors and allows us to locally measure protonization reaction rates. This noise can be minimized by operating the nanopore at a deliberately chosen ph. DOI: 10.1103/PhysRevLett.102.256804 As nanodevices are increasingly used for singlemolecule detection and manipulation in the time domain, their noise characteristics at khz frequencies gain additional significance. In the case of protein nanopores, noise has elucidated both pore and analyte dynamics [1 3]; in solid-state nanopores to date, noise has simply interfered with detection of DNA and proteins [4 9]. Previous work has described a variety of noise sources in solid-state nanopores and typical measuring apparatus [10 12]. In this Letter, we identify a new source of current noise in solid-state nanopores that is dominant in the frequency range from 0.1 10 khz. A simple analysis of the noise dependence on ph and the electrolyte ionic strength, based on a few general assumptions, suggests that the noise originates from surface charge fluctuations. Finally, we develop a quantitative model based on protonization of surface functional groups [1] and compare it to the experimental results. Our technique is sensitive to a few tens of active surface groups in the nanopore, allowing us to observe variations in local surface properties from nanopore to nanopore and to guide optimization of single-molecule detectors. Nanopores were fabricated with a transmission electron microscope [13] in a thin film of low-stress, amorphous silicon nitride, resulting in single, roughly hourglassshaped [14] channels 55 nm long and 4 nm in diameter at the smallest point. The silicon nitride film containing the nanopore separates two reservoirs of potassium chloride electrolyte in a sealed fluidic cell; current measurements were made between Ag=AgCl electrodes in each reservoir (Fig. 1). Noise measurements on different nanopores show the same general features and ph spectral shapes, while the sensitivity and local nature of the technique reveals sample-to-sample variability [15]. The noise sources we observe appear to be intrinsic to the silicon nitride nanopore surface and are not as variable as 1=f noise [12]. The power spectral densities (PSDs) of typical current recordings (Fig. 1, inset) demonstrate various expected noise sources: with no applied voltage bias (lower spectrum), only thermal (Johnson) noise (thick dashed line) and high-frequency capacitive noise driven by the preamplifier PACS numbers: 85.35. p, 05.40.Ca, 66.90.+r, 82.45.Yz input are present. With applied voltage bias across the nanopore device (upper spectrum), conductance fluctuations such as 1=f noise (dotted line) become apparent. We also observe the appearance of an additional frequencyindependent ( white ) noise between 0.1 khz and our RC filter cutoff at 20 khz. In the inset to Fig. 2(b), this excess white noise, averaged over the 1 8 khz band, is plotted as a function of mean nanopore current. The average noise (closed circles) scales with the square of the dc current, indicating that the excess white noise derives from a conductance fluctuation. For an Ohmic system of conductance G, flowing current I, the PSDs of the current fluctuations S I and the conductance fluctuations S G are related by S I ¼ ðs G =G 2 ÞI 2. The parabolic coefficient S G =G 2 contains all the relevant physics of the intrinsic system fluctuations. A simple RC system of similar conductance and capacitance (500 M, 33 pf) shows no excess white noise (open FIG. 1 (color online). A nanopore is a narrow channel in a thin Si 3:5 N 4 membrane. A current of electrolyte ions flows through the nanopore in response to an applied voltage. Detail: surface protonization reactions. Inset: Typical current noise spectra at 0 mv (lower trace) and 150 mv (upper trace) applied voltage at 1 M KCl and neutral ph. 0031-9007=09=102(25)=256804(4) 256804-1 Ó 2009 The American Physical Society

-9 2 S I (10 na /Hz) σ 1.2 0.8 0.4 0.0-0.5 0.0 0.5 Current (na) FIG. 2 (color online). (a) Dependence of the ionic conductance through the nanopore on bulk electrolyte concentration (open circles), with a fit to the Poisson-Boltzmann model (solid line). Insets: (upper left) A schematic of the geometry [14] used for the Poisson-Boltzmann model, (lower right) The effect of surface fluctuations over the Debye length D, predicting the c 3=2 dependence of the noise level. (b) Conductance fluctuations vs bulk ionic concentration (open circles). The solid line is the fit, using the same parameters as in (a), to the charge fluctuation model outlined in the text. Inset: Parabolic dependence of the frequency-averaged PSD of S I (closed circles) on current, with fit (solid line). circles), demonstrating that the noise source is not the electronics or analysis used to obtain our results; likewise, a long capillary with Ag=AgCl electrodes of similar conductance and capacitance yields no excess white noise, indicating that the electrodes are also not a source of noise. To determine the physical source of the excess white noise, we measured the PSD as a function both of electrolyte concentration [Fig. 2(b)] near neutral ph, and of ph [Fig. 3(b)] at constant ionic strength (110 mm unbuffered KCl), using a fluidic cell capable of preparing gradient concentration flows and an in situ ph-sensitive half-cell [15]. For each point, full I-V and S I -V curves were collected; the conductance is determined from the slope of the I-V curve at zero bias, while the conductance fluctuation S G =G 2 is determined from the parabolic fit of the excess white noise vs the current [Fig. 2(b), inset]. There are two distinct regimes of the noise as a function of electrolyte concentration c [Fig. 2(b)]. At high concentrations, the noise varies as c 3=2 (dotted line), while the conductance is proportional to c [Fig. 2(a), dashed line]. Below about 100 mm KCl, both the noise and the conductance deviate strongly from their high-c behavior. This effect can be understood by considering Debye screening in the context of fluctuating surface charges on the walls of 256804-2 FIG. 3 (color online). (a) Measured conductance (open circles) vs ph at constant 110 mm ionic strength. The minimum at ph 4:1 corresponds to the point of zero charge. The solid line represents the predicted conductance from a fluctuating 3-level system. Inset: Surface charge as induced from the data (open circles) and as predicted by the model (solid line). (b) Experimental noise spectral density (open circles) vs ph. The solid line is a fit to the charge fluctuation model presented in the text. Both curves are modeled by the same set of parameters that was used to fit the variable electrolyte concentration data (see Fig. 2). the nanopore, which for simplicity we assume to be cylindrical. For high electrolyte concentrations, the Debye screening length ( D c 1=2 ) is smaller than the radius of the nanopore R p. When D R p, the number of charge carriers affected by the fluctuating charges [Fig. 2(a), inset] is given by c A ring ¼ cð2 D R p 2 D Þc D c 1=2 : (1) Conductance is proportional to the concentration in this regime, so S G =G 2 c 3=2. At low concentrations, the Debye length becomes comparable to or larger than the radius of the nanopore, and ions in the entire cross section of the nanopore are affected by the fluctuations of the electric field originating from the surface charge. In the simplest approximation, the local ion concentration and hence the conductance should be independent of the bulk concentration and proportional to the surface charge required to maintain charge neutrality inside the nanopore. Because the surface charge itself slightly decreases with electrolyte dilution, the concentration dependence of the noise is bounded between the limit of constant surface charge (S G ¼ const, G ¼ const) and the limit of no surface charge effects (S G c, G c). To summarize,

c ðs G =G 2 3=2 for Þ surface D R p c : (2) ; 0 < 1 D * R p Despite the simplicity of our model, it agrees very well with the experimental results in Fig. 2(b). For the highconcentration regime, we also note that for geometric fluctuations at the surface (such as those postulated for polymer nanopores [16]), and for volume fluctuations (as suggested by the Hooge relation for 1=f noise [17]), S G would scale with c, so S G =G 2 c 1. Only for surfacebound electrical potential fluctuations does the normalized PSD scale as c 3=2. Excess white noise varies with ph [Fig. 3(b)]. The noise peaks at ph 7.5 with a long tail extending to ph 2. The conductance has minimum at ph 4.1, but there is no obvious correspondence between the noise and conductance curves. A noise level peak similar to that depicted in Fig. 3(b) has been observed in single OmpF channels [3]. Bezrukov and Kasianowicz, working with -hemolysin protein ion channels [1], attribute the peak to protonization reactions of individual protein residues in the ion channel and calculate the number of ionizable residues and the reaction kinetics of the protein channel. The silicon nitride surface consists of a large fraction of amphoteric silicon oxide groups [18,19], which are known to be active in the experimentally accessible ph range. The minimum of the conductance corresponds to the ph at which the densities of positive and negative surface groups are equal, the so-called point of zero charge. At low or high phs, respectively, either positive or negative surface groups dominate, increasing the local concentration of charge carriers and hence the total conductance. The relevant reactions are SiO þ H þ! k R SiOH þ H þ! l R k D SiOH (3) l D SiOH þ 2 : (4) The equilibrium constants of these reactions are defined by K k D =k R ¼ N SiO ½Hþ Š 0 N SiOH 10 pk L l D =l R ¼ N SiOH½H þ Š 0 10 pl N SiOH þ 2 where ½H þ Š 0 is the hydrogen ion activity at the surface and N i is the density of surface sites in state i. The hydrogen ion activity at the surface is related to the bulk ph by ½H þ Š 0 ¼½H þ Š bulk expð ec 0 Þ, where ½H þ Š bulk ¼ 10 ph, c 0 is the electrical potential at the nanopore surface, ¼ ðktþ 1 is the thermodynamic factor, and e is the unit of elementary charge, and ec 0 is the energy cost for a positively charged ion to approach a charged surface. For example, if the surface has a net negative surface charge, hydrogen ions are attracted to the nanopore walls, resulting in a lower local ph. Combining the above expressions with Eq. (5), and taking ¼ P N i to be the total surface density (5) 256804-3 of active sites and to be the surface charge density, we obtain a generalized form of the Behrens-Grier equation [20]: 10 pl ph ð e Þ expð 2ec 0 Þ þ expð ec 0 Þþ10 ph pk ð e Þ ¼0: (6) The basic Stern model relates the potential at the nanopore surface to the potential at the double layer, c D, by the Stern capacitance C s, which captures the dielectric properties and effective thickness of the water-solid interface: c 0 c D ¼ =C s. Finally, the surface charge as a function of double layer potential is approximated by the Grahame equation [21]: ðc D Þ¼ 2 0 e D sinh ec D 2 : (7) Solving Eqs. (6) and (7) self-consistently, we obtain and the surface density of each species N i as a function of ph, electrolyte concentration, and the four chemical parameters, C s, pk, and pl. A plot of vs ph is given in the inset to Fig. 3(a). To model the conductance data in Figs. 2(a) and 3(a), we calculate GðÞ from the numerical solution of the Poisson- Boltzmann equation in a cylindrically symmetric hourglass geometry. Access resistance is small enough to be subsumed into the geometric parameters of the nanopore. A thorough investigation of the conductance of the nanopore is beyond the scope of this Letter and will be published elsewhere. The set of coupled chemical reactions 3 and 4 define a 3-level protonization system. To calculate the spectral density, we note that L K, allowing us to decouple these reactions into two independent 2-level systems defined by the individual reactions. The contribution to the spectral density from each chemical reaction is calculated by combining the solution to the Kolmogorov equations for a discrete 2-level Markov system [1,22,23] with the chemical equilibrium and surface proton activity Eqs. (3) (7), yielding the expression G ðph;cþ¼s 10 pk ph expð ec 0 Þ 0 ½1 þ 10 pk ph expð ec 0 ÞŠ 3 : (8) S ðþ This is a generalization of the Bezrukov-Kasianowicz equation for protonization noise in protein channels [1]. The additional surface potential factor c 0 ðph;cþ combines contributions from all species and is responsible for the dependence of the noise on ionic strength. The solid curves in Figs. 2 and 3 are calculated using geometric parameters (determined from conductance measurements and transmission electron microscope imaging) R p ¼ 3:3 nm, L ¼ 54 nm, h ¼ 7 nm, and ¼ 30, and the chemical parameters pl ¼ 2:7, pk ¼ 6:6, C s ¼ 0:7 F=m 2, and ¼ 0:4 nm 2, corresponding to 150 protonizable surface sites in the cylindrical region of the nanopore. The chemical parameters are consistent with

those cited in the literature for SiO 2 surfaces (see [24] and references within). We note an excellent general correspondence between the experimental data and the theoretical dependence of measurables on electrolyte concentration. For the ph-dependence, the model correctly predicts the noise peak close to pk, with two discrepancies: the relative positions of peaks of experimental and modeled data are shifted by ph 0:5, and we cannot account for the increased experimental noise level between ph 4 and ph 6. These deviations could have various origins. The basic Stern model may not be sophisticated enough to explain the details of the S G ðphþ curve: additional effects, such as incorporation of electrolyte ions into the Stern layer, or the high surface curvature in the nanopore, might be significant [25]. Additional contributions to the conductance fluctuations may interfere with the low-ph peak, particularly charge carrier number fluctuations [26]. It is also likely that the protonizable chemical sites have a distribution of equilibrium constants (pks), due to variations in the local chemistry and the existence of species involving both N and O atoms, multiple Si atoms, and multiple O atoms. The conductance measurements with ph support this idea, as the two-peak model underestimates the conductance (and hence the surface charge) between phs 4 and 6 and overestimates the conductance above ph 7. Nanopore noise measurement is a sensitive local probe, not only for surface composition, but also for surface charge fluctuations at time scales much shorter than the electronic measurement bandwidth. The chemical dynamics information is contained in the normalization factor [1] of Eq. (8), S ðkþ 0 ¼ 4 P iðg i Þ 2 k 1 D 4ðGÞ2 N 1 k 1 D, where g i is the fractional change of conductance due to protonization of an individual chemical site i, N is the total number of protonizable sites, and G is the maximum enhancement of the ionic conductance due to the surface reactions 3 and 4. Using the geometric and chemical parameters of the nanopore, we obtain reaction parameters k D ¼ 1:8 10 5 s 1 and k R ¼ 7 10 11 M 1 s 1 for reaction 3. The mean lifetime of the proton bonding in SiO-H, SiOH ¼ k 1 D ¼ 5:6 s, is comparable to, though smaller than, the proton binding lifetime in a protein channel [1], perhaps due to additional conformational changes induced in the protein upon protonization. Equivalently, we deduce reaction parameters for reaction 4: l D > 3:9 10 5 s 1 and l R > 2 10 8 M 1 s 1. Because we cannot actually resolve this peak from the background, l D and l R must be interpreted as lower bounds on the low-ph reaction kinetics. Practically, surface charge fluctuations will decrease the signal to noise ratio of nanopore biosensing devices. Our experiments demonstrate that there are large, predictable variations in excess white noise over broad ph ranges. The results suggest that the nanopores can be used as a sensitive local probe of detailed surface dynamics and chemistry: to establish such a technique, experimental evaluations of multiple films and conditions will be performed. This study establishes the importance of mitigating interference from protonization noise by the selection of appropriate surface materials and operating conditions for dynamic nanoscale devices. We thank M. Gershow, F. Albertorio, and L. Bo for technical assistance and helpful discussions. D. P. H. acknowledges support from NDSEG and the NSF Graduate program. This work was supported by NIH Grant No. 2R01HG0037. [1] S. M. Bezrukov and J. J. Kasianowicz, Phys. Rev. Lett. 70, 2352 (1993). [2] T. K. Rostovtseva and S. M. Bezrukov, Biophys. J. 74, 2365 (1998). [3] E. M. Nestorovich, T. K. Rostovtseva, and S. M. Bezrukov, Biophys. J. 85, 3718 (2003). [4] J. Li et al., Nature (London) 412, 166 (2001). [5] J. Li et al., Nature Mater. 2, 611 (2003). [6] A. J. Storm et al., Phys. Rev. E 71, 051903 (2005). [7] D. Fologea et al., Nano Lett. 5, 1905 (2005). [8] D. Fologea et al., Appl. Phys. Lett. 91, 053901 (2007). [9] M. Gershow and J. A. Golovchenko, Nature Nanotech. 2, 775 (2007). [10] V. Tabard-Cossa et al., Nanotechnology 18, 305505 (2007). [11] J. D. Uram, K. Ke, and M. Mayer, ACS Nano 2, 857 (2008). [12] R. M. M. Smeets et al., Proc. Natl. Acad. Sci. U.S.A. 105, 417 (2008). [13] A. J. Storm et al., Nature Mater. 2, 537 (2003). [14] M. J. Kim et al., Nanotechnology 18, 205302 (2007). [15] See EPAPS Document No. E-PRLTAO-103-026928 for more details on experimental methods. For more information on EPAPS, see http://www.aip.org/pubservs/ epaps.html. [16] Z. Siwy and A. Fulinski, Phys. Rev. Lett. 89, 158101 (2002). [17] F. N. Hooge, Phys. Lett. A 29, 139 (1969). [18] X. G. Zhang, Electrochemistry of Silicon and Its Oxide (Kluwer Academic/Plenum Publishers, New York, 2001). [19] M.-Y. Wu et al., Nano Lett. 9, 479 (2009). [20] S. H. Behrens and D. G. Grier, J. Chem. Phys. 115, 6716 (2001). [21] J. N. Israelachvili, Intermolecular and Surface Forces: With Applications to Colloidal and Biological Systems (Academic Press, London; Orlando, Florida, 1985). [22] S. Kogan, Electronic Noise and Fluctuations in Solids (University Press, Cambridge, 1996). [23] S. Machlup, J. Appl. Phys. 25, 341 (1954). [24] F. H. J. van der Heyden, D. Stein, and C. Dekker, Phys. Rev. Lett. 95, 116104 (2005). [25] J. Westall and H. Hohl, Adv. Colloid Interface Sci. 12, 265 (1980). [26] S. M. Bezrukov et al., Physica B (Amsterdam) 159, 388 (1989). 256804-4