Electrostatic Cycling of Hybridization Using ... - ACS Publications

Jun 15, 2017 - Department of Chemical and Biomolecular Engineering, New York University Tandon School of Engineering, 6 MetroTech Center,. Brooklyn ...
0 downloads 0 Views 553KB Size
Subscriber access provided by CORNELL UNIVERSITY LIBRARY

Letter

Electrostatic Cycling of Hybridization Using Nonionic DNA Mimics Sade Ruffin, Isabella A Hung, Ursula M Koniges, and Rastislav Levicky ACS Sens., Just Accepted Manuscript • DOI: 10.1021/acssensors.7b00100 • Publication Date (Web): 15 Jun 2017 Downloaded from http://pubs.acs.org on June 19, 2017

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

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

Page 1 of 16

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

ACS Sensors

Electrostatic Cycling of Hybridization Using Nonionic DNA Mimics Sade Ruffin, Isabella A. Hung, Ursula M. Koniges and Rastislav Levicky* New York University Tandon School of Engineering, Dept. of Chemical and Biomolecular Engineering, 6 MetroTech Center, Brooklyn, NY 11201. *Corresponding author: [email protected]

ABSTRACT: This study demonstrates efficient electrostatic control of surface hybridization through use of morpholinos, a charge-neutral DNA mimic, as the immobilized "probes". In addition to being compatible with low ionic strengths, use of uncharged probes renders the field interaction specific to the nucleic acid analyte. In contrast to DNA probes, morpholino probes enable facile cycling between hybridized and dehybridized states within minutes. Impact of ionic strength and temperature on effectiveness of electrostatics to direct progress of hybridization is evaluated. Optimal electrostatic control is found when stability of probe-analyte duplexes is set so that electrostatics can efficiently switch between the forward (hybridization) and reverse (dehybridization) directions.

KEYWORDS: electrochemical, PNA, electrode, bioaffinity, assay, surface, monolayer

ACS Paragon Plus Environment

ACS Sensors

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

The charged character of nucleic acids has naturally stimulated interest in using electric fields to control their behavior at interfaces for sensing and related applications. Electric fields can provide orientational control, allowing shifting between adsorbed and upright conformations.1–3 Such conformational switching was subsequently pursued for biosensing.4 Electric fields have been also proposed for direct control of hybridization between surface-bound "probe" and solution "target" strands.5–18 In these applications electric fields are intended to interact with the oligonucleotide charges to facilitate or hinder hybridization. Conde and co-workers reported that attractive biases of 1 V accelerated hybridization by nearly a billion fold, resulting in hybridization times of less than a msec.5,6 These authors considered that the fields facilitated hybridization of targets after they already adsorbed to the probe-modified support, rather than influence mass transport of targets to the probe layer. Using more modest biases of 0.2 V, Su et al observed a similar instantaneous like onset of hybridization after a "lag period" of no hybridization.8 In both these studies hybridization was determined from fluorescence signals after washing and drying of the DNA layers. In contrast, when hybridization was followed in-situ under attractive 0.3 V potentials, such strong acceleration was not observed.7 A number of studies also demonstrated that repulsive biases can assist the release of targets from probe-target duplexes.7–16 One explanation for dehybridization under repulsive biases is through electrostatic mechanisms,18 but other mechanisms were also put forth. Johnson et al observed dehybridization of duplexes formed by two complementary strands of peptide nucleic acids (PNA) when sufficiently negative (-0.8 V) potentials were used. Since PNAs are uncharged, electrostatics should not be responsible; rather, the authors suggested that electron transfer to the duplex pi stack caused its destabilization.19 The authors further suggested that

ACS Paragon Plus Environment

Page 2 of 16

Page 3 of 16

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

ACS Sensors

such nonelectrostatic mechanisms could also explain denaturation of DNA-DNA duplexes under conditions of weak electrostatics, such as at high ionic strengths. More recent work considered the effects of dynamic fields on hybridization. Gebala and coworkers reported that application of potential cycles at 5 Hz under high ionic strengths (~ 1 mol L-1) approximately doubled the hybridization yield compared to applying a constant attractive potential.17 This interesting effect was not seen at lower (~ 0.02 mol L-1) ionic strengths. The physical mechanism was thought to involve structural dynamics of the DNA probe layer that facilitated hybridization when the layer organization was forced to constantly respond to switching fields. The primary motivation for using electric fields to modulate rates and equilibria of surface hybridization is convenience for diagnostic applications. For fully general capability it is desirable to accelerate as well as reverse hybridization at will. As noted above, use of DNA probes has led to inconsistent outcomes for attractive fields showing both extreme5,6,8 and little impact on forward rates,7 while achievement of fast (e.g. on a minute timescale) dehybridization has required assistance from buffer exchanges12,15 or clever use of charged nanoparticle tags to amplify electrostatic repulsions.14 Use of charged probes like DNA imposes certain limitations on electrostatic regulation of hybridization. One issue is that charged probes retain an excess of mobile counterions at the surface. This excess leads to enhanced local ionic strength, and thus to stronger screening of electric fields. In addition, charged probes can also respond to applied fields. Such responsiveness could interfere with hybridization; for example, attractive biases could cause the probes to adsorb to the solid support and block their binding to target strands. Lastly, hybridization to DNA probes requires that salt concentration be sufficiently high.20 This

ACS Paragon Plus Environment

ACS Sensors

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

Page 4 of 16

requirement limits optimization of electrostatic control by preventing evaluation of lower ionic strength conditions. These

considerations

motivated

exploration

of

uncharged

morpholino

(MO)

oligonucleotides as probes to facilitate electrostatic regulation of hybridization. In contrast to negatively charged DNA, morpholinos have a nonionic backbone consisting of alternating morpholino rings connected by neutral phosphorodiamidate linkages, Fig. 1A, and their hybridization with DNA is insensitive to salt conditions.21–23 Selection of morpholinos over other uncharged DNA analogues, such as PNA, was motivated by their good hybridization properties as well as synthetic flexibility with regard to length and base composition.22,24 The favorable properties of morpholinos have motivated a number of recent studies into their sensing applications.25–34 These applications have benefitted from the morpholinos’ ability to hybridize under low ionic strengths to, for example, destabilize interfering secondary structure in the sample25,26 or to improve sensitivity of charge-based transduction.30,35 DNA and MO probe monolayers of an identical 20mer base sequence (5' TTA AAT TCT GCA AGT GAT TT) were prepared on mercaptohexanol (MCH) passivated gold rotating disk electrodes. The first three thymine residues on the probes served as a linker to facilitate hybridization. Probe coverages were in the range of 1.5×1012 to 2×1012 cm-2, corresponding to a 7 to 8 nm average separation between probes. Thus modified electrodes were placed into 10 nmol L-1 solutions of complementary 17mer targets (5' TCA CTT GCA GAA TTT AA). Hybridization was monitored as a function of time while the potential applied to the electrode was stepped between various settings. At each potential, the extent of hybridization was determined every two minutes by calculating the surface coverage of targets from the charge Q

ACS Paragon Plus Environment

Page 5 of 16

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

ACS Sensors

needed to switch the oxidation state of their ferrocene labels. Q was calculated from fast cyclic voltammetry (CV) sweeps by integration of the corresponding peaks (Fig. S1). As shown in Fig. 1B, only slight if any effects on hybridization from applied fields were noticeable for either DNA or MO probe monolayers in 0.1 mol L-1 sodium phosphate buffer (SPB). Markedly better control was observed for MO probe layers if the SPB concentration was lowered to 0.01 mol L-1, with clearly delineated periods of acceleration under attractive (positive) potentials and halting under repulsive (negative) biases. The corresponding CV data are presented in Fig. 1C. DNA probes did not hybridize sufficiently in 0.01 mol L-1 SPB to allow quantification, as shown in Fig. S2, an observation attributed to destabilization of DNA-DNA duplexes at the low ionic strength. Based on these results subsequent experiments focused on MO probes and low salt conditions. The results for MO monolayers in Fig. 1B show that lower ionic strengths improved responsiveness of hybridization to applied potentials. This is attributed to further reach of surface electric fields at these salt concentrations. The electrostatic Debye screening length,36 rD, in 0.01 mol L-1 pH 7 SPB is about 2 nm; therefore, surface fields extend several nm into solution. As the length of a 17mer duplex is about 6 nm, this means that duplexes are substantially exposed to the field over most of their length. In comparison, longer probes would place more base pairs outside reach of the field, which would reduce the ability to regulate hybridization since a smaller fraction of duplex base pairs would be affected. A mismatch of probe size and extent of the fields can also be realized by increasing ionic strength; for example, in 0.1 mol L-1 buffer rD shrinks to 0.7 nm, resulting in field-duplex interaction over only a small portion of a duplex and a concomitant loss of control, as seen in Fig. 1B.

ACS Paragon Plus Environment

ACS Sensors

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

The data in Fig. 1 did not demonstrate reversal of hybridization, only its halting. It would be useful to actuate both the forward and reverse reactions, so that the probe layer can be brought from a given initial state to a desired final state. This requires the field-induced perturbation to switch between conditions accelerating hybridization and those that reverse it. This requirement can be best met if duplex stability, in the absence of applied fields, is marginal. Then, electrostatics can be used to tilt the balance in favor of the forward or the reverse direction, stabilizing or denaturing probe-target duplexes as needed. Fig. 2A presents results of a temperature series for a fixed buffer strength of 0.01 mol L-1. These experiments used alternating 6 min rounds of attractive +0.3 and repulsive -0.3 V potential pulses. For 35 oC and below, MO-DNA duplexes remained sufficiently stable so that repulsive fields triggered only modest dehybridization within the allotted period. Increasing the temperature to 40 oC, however, lowered duplex stability sufficiently so that repulsive rounds clearly tipped the balance toward denaturation. To confirm that this reversal reflected changes in duplex stability, these experiments were repeated for a target with a central C/T mismatch (5' TCA CTT GCC GAA TTT AA; mismatch underlined). According to the proposed mechanism a duplex destabilized by a mismatch should display a lower threshold temperature for dehybridization. Indeed, the mismatch target exhibited hybridization switching already at 35 oC as shown in Fig. S3. Use of a partly matched target with just a six-nucleotide complementary region (5' ACT TGC ATC ATC ATA ACT CAC; complementary region is underlined) did not result in significant signal even after 16 minutes at +0.3 V (Fig. S4), verifying that the cycling involved sequence-specific hybridization. Fig. 2B contrasts the full match and mismatch data at 35 oC, illustrating the pronounced effect of the mismatch. The strong contrast in responses of the full match and mismatch

ACS Paragon Plus Environment

Page 6 of 16

Page 7 of 16

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

ACS Sensors

sequences is potentially useful for controlling sequence stringency in applications. This control could be realized, for example, by using trains of potential pulses to accelerate replacement of mismatched by fully matched sequences during competitive hybridization. Development of such functionality will require optimization with respect to pulse amplitude, duration, mismatch position, and target length. Fig. S5 illustrates the expected variation of full match to mismatch signal with applied potential, as estimated from the noncompetitive hybridization results of Fig. 2B. Ultimately, electrostatic regulation of hybridization is constrained by magnitude of the available perturbation. The electrostatic energy Eel of an MO-DNA duplex interacting with an applied surface field was estimated by integrating the product of the local potential V(x), expressed relative to V = 0 in bulk solution, with the duplex charge

Eel =

−e V ( x)dx l duplex B



(1)

layer

where x is measured normal to the electrode surface. In Eqn. 1, the duplex linear charge density e/lB has been set at one elementary charge -e per Bjerrum length37 lB, in accordance with counterion condensation theory.38 The Bjerrum length is lB = NAe2/(4πεrεoRT) where NA is Avogadro's number and εr and εo are the relative and vacuum permittivities; physically, lB represents the distance at which interaction between two elementary charges in a dielectric medium is comparable to thermal energy. In water at room temperature lB is approximately 0.7 nm. V(x) was obtained numerically from a Poisson-Boltzman calculation following the model in reference 39. For the calculation the MCH layer was assigned a dielectric constant of 4.0 and a thickness of 1.2 nm, as estimated from fits to capacitance of MCH monolayers following reference 35. The MO-DNA duplex layer was approximated as a slab with a dielectric constant

ACS Paragon Plus Environment

ACS Sensors

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

that of water and a uniformly smeared density of negative charge as specified by the measured duplex coverage and charges per duplex L/lB ≈ 9, where L = 6 nm is the duplex length. The potential at the electrode/MCH interface was set to -0.25 V, representing the difference between -0.3 V used for electrostatic denaturation and the potential of zero charge (pzc) for morpholinomodified electrodes. The pzc was determined experimentally from the minimum in the double layer capacitance to be -0.05 V (Fig. S6). For simplicity in evaluating Eqn. 1 it was assumed that under repulsive biases duplexes orient perpendicularly to the surface. Other orientations could be included through Boltzmann partitioning based on electrostatic energy of a given tilt angle.2 The assumption of a normal orientation provides a lower limit on Eel. Fig. 3 shows the V(x) profile calculated for 0.01 mol L-1 phosphate buffer and typical duplex coverage of 5 × 1011 cm-2, with the hatched area representing the integration region over the 6 nm thickness of the duplex layer. The calculation yields Eel = 13 kJ mol-1, or about 5RT. This magnitude can be compared to an equivalent change in temperature. Using ∆Go = ∆Ho T∆So, treating ∆Ho and ∆So as independent of temperature, and taking ∆So for MO-DNA surface hybridization as about half its solution value23 (Supporting Information) leads to ∆So ≈ -0.5 kJ mol-1 K-1. A ∆Go change of 13 kJ mol-1 then corresponds to a temperature increase of 13 kJ mol1

/ 0.5 kJ mol-1 K-1 = 26 oK. Since this is approximately the width of an MO-DNA surface melting

transition,23 the prediction is that if the temperature is set just at the low side of a transition then applying a -0.3 V bias should be sufficient for achieving nearly complete denaturation. Such a prediction is in reasonable agreement with the trends in Fig. 2. Repeating the same calculation for 0.1 mol L-1 buffer leads to Eel = 1.4 kJ mol-1, or about 0.5RT. This low value, compared to thermal energy, agrees with Fig. 1B in that field effects were too weak to be significant in 0.1 mol L-1 buffer.

ACS Paragon Plus Environment

Page 8 of 16

Page 9 of 16

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

ACS Sensors

The present study demonstrated facile switching between forward and reverse directions of surface hybridization when charge neutral DNA mimics were used as the probe molecules. At the lower ionic strengths found to be effective, the control mechanism was consistent with simple electrostatics in which surface charges on the electrode interact with backbone charges on the analyte to stabilize or destabilize its presence at the surface. Other mechanisms of electrochemically influencing duplex stability, such as electron transfer to the duplex pi stack19 or local perturbation of buffer pH as in certain commercial technologies,40 are not expected to be significant at the modest potentials used. The combination of biases gentle to biomolecular monolayers and rapid, within minutes, cycling of hybridization opens up opportunity for development of direct field effect protocols for controlling sequence stringency and other aspects of surface nucleic acid bioassays.

Acknowledgment This work was supported by the National Science Foundation (DMR 12-06754, CBET 1600584), and by New York University.

Supporting Information Description of experimental methods including analysis of CV data, hybridization to DNA probes at low ionic strengths, hybridization to mismatched targets, determination of the potential of zero charge, and characterization of hybridization thermodynamics in solution. This material is available free of charge via the Internet at http://pubs.acs.org.

Notes The authors declare no competing financial interest.

ACS Paragon Plus Environment

ACS Sensors

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

References (1)

(2)

(3) (4)

(5)

(6)

(7)

(8)

(9) (10)

(11) (12)

(13)

(14) (15)

Kelley, S. O.; Barton, J. K.; Jackson, N. M.; McPherson, L. D.; Potter, A. B.; Spain, E. M.; Allen, M. J.; Hill, M. G. Orienting DNA Helices on Gold Using Applied Electric Fields. Langmuir 1998, 14, 6781–6784. Rant, U.; Arinaga, K.; Fujita, S.; Yokoyama, N.; Abstreiter, G.; Tornow, M. Electrical Manipulation of Oligonucleotides Grafted to Charged Surfaces. Org. Biomol. Chem. 2006, 4, 3448–3455. Rant, U.; Arinaga, K.; Fujita, S.; Yokoyama, N.; Abstreiter, G.; Tornow, M. Dynamic Electrical Switching of DNA Layers on a Metal Surface. Nano Lett 2004, 4, 2441–2445. Rant, U.; Arinaga, K.; Scherer, S.; Pringsheim, E.; Fujita, S.; Yokoyama, N.; Tornow, M.; Abstreiter, G. Switchable DNA Interfaces for the Highly Sensitive Detection of LabelFree DNA Targets. Proc Natl Acad Sci USA 2007, 104, 17364–17369. Cabeca, R.; Rodrigues, M.; Prazeres, D. M. F.; Chu, V.; Conde, J. P. The Effect of the Shape of Single, Sub-Ms Voltage Pulses on the Rates of Surface Immobilization and Hybridization of DNA. Nanotechnology 2009, 20, art-015503. Fixe, F.; Branz, H. M.; Louro, N.; Chu, V.; Prazeres, D. M. F.; Conde, J. P. Electric-Field Assisted Immobilization and Hybridization of DNA Oligomers on Thin-Film Microchips. Nanotechnology 2005, 16, 2061–2071. Heaton, R. J.; Peterson, A. W.; Georgiadis, R. M. Electrostatic Surface Plasmon Resonance: Direct Electric Field- Induced Hybridization and Denaturation in Monolayer Nucleic Acid Films and Label-Free Discrimination of Base Mismatches. Proc Natl Acad Sci USA 2001, 98, 3701–3704. Su, H. J.; Surrey, S.; McKenzie, S. E.; Fortina, P.; Graves, D. J. Kinetics of Heterogeneous Hybridization on Indium Tin Oxide Surfaces with and without an Applied Potential. Electrophoresis 2002, 23, 1551–1557. Bartosik, M.; Palecek, E. Square Wave Stripping Voltammetry of Unlabeled Single- and Double-Stranded DNAs. Electroanalysis 2011, 23, 1311–1319. Johnson, R. P.; Richardson, J. A.; Brown, T.; Bartlett, P. N. A Label-Free, Electrochemical SERS-Based Assay for Detection of DNA Hybridization and Discrimination of Mutations. J. Am. Chem. Soc. 2012, 134, 14099–14107. Mahajan, S.; Richardson, J.; Brown, T.; Bartlett, P. N. SERS-Melting: A New Method for Discriminating Mutations in DNA Sequences. J Am Chem Soc 2008, 130, 15589–15601. Spehar-Deleze, A. M.; Schmidt, L.; Neier, R.; Kulmala, S.; de Rooij, N.; Koudelka-Hep, M. Electrochemiluminescent Hybridization Chip with Electric Field Aided Mismatch Discrimination. Biosens. Bioelectron. 2006, 22, 722–729. Cao, S. H.; Xie, T. T.; Cai, W. P.; Liu, Q. A.; Li, Y. Q. Electric Field Assisted Surface Plasmon-Coupled Directional Emission: An Active Strategy on Enhancing Sensitivity for DNA Sensing and Efficient Discrimination of Single Base Mutation. J. Am. Chem. Soc. 2011, 133, 1787–1789. Wang, Q.; Yang, X.; Wang, K.; Tan, W.; Gou, J. Recognition of Single-Base Mismatch DNA by Au Nanoparticle-Assisted Electroelution. Analyst 2008, 133, 1274–1279. Wei, F.; Qu, P.; Zhai, L.; Chen, C.; Wang, H.; Zhao, X. S. Electric Potential Induced Dissociation of Hybridized DNA with Hairpin Motif Immobilized on Silicon Surface. Langmuir 2006, 22, 6280–6285.

ACS Paragon Plus Environment

Page 10 of 16

Page 11 of 16

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

ACS Sensors

(16) Wong, I. Y.; Melosh, N. A. Directed Hybridization and Melting of DNA Linkers Using Counterion-Screened Electric Fields. Nano Lett. 2009, 9, 3521–3526. (17) Tymoczko, J.; Schuhmann, W.; Gebala, M. Electrical Potential-Assisted DNA Hybridization. How to Mitigate Electrostatics for Surface DNA Hybridization. ACS Appl Mater Interfaces 2014, 6, 21851–21858. (18) Vainrub, A.; Pettitt, B. M. Surface Electrostatic Effects in Oligonucleotide Microarrays: Control and Optimization of Binding Thermodynamics. Biopolymers 2003, 68, 265–270. (19) Johnson, R. P.; Gale, N.; Richardson, J. A.; Brown, T.; Bartlett, P. N. Denaturation of dsDNA Immobilised at a Negatively Charged Gold Electrode Is Not Caused by Electrostatic Repulsion. Chem Sci 2013, 4, 1625–1632. (20) Gong, P.; Levicky, R. DNA Surface Hybridization Regimes. Proc Natl Acad Sci USA 2008, 105, 5301–5306. (21) Summerton, J. Uncharged Nucleic Acid Analogs for Therapeutic and Diagnostic Applications: Oligomers Assembled from Ribose-Derived Subunits. In Discoveries in Antisense Nucleic Acids (Advances in Applied Biotechnology); Brakel, C., Ed.; Portfolio Publishing Co.: The Woodlands, TX, 1989; pp. 71–80. (22) Summerton, J. E. Morpholinos and PNAs Compared. Lett Pept Sci 2004, 10, 215–236. (23) Qiao, W.; Chiang, H.-C.; Xie, H.; Levicky, R. Surface vs Solution Hybridization: Effects of Salt, Temperature, and Probe Type. Chem Commun 2015, 51, 17245–17248. (24) Micklefield, J. Backbone Modification of Nucleic Acids: Synthesis, Structure and Therapeutic Applications. Curr. Med. Chem. 2001, 8, 1157–1179. (25) Qiao, W.; Kalachikov, S.; Liu, Y.; Levicky, R. Charge-Neutral Morpholino Microarrays for Nucleic Acid Analysis. Anal Biochem 2013, 434, 207–214. (26) Zu, Y.; Ting, A. L.; Yi, G.; Gao, Z. Sequence-Selective Recognition of Nucleic Acids under Extremely Low Salt Conditions by Using Nanoparticle Probes. Anal Chem 2011, 83, 4090–4094. (27) Gao, Z.; Deng, H.; Shen, W.; Ren, Y. A Label-Free Biosensor for Electrochemical Detection of Femtomolar microRNAs. Anal Chem 2013, 85, 1624–1630. (28) Hu, W.; Hu, Q.; Li, L.; Kong, J.; Zhang, X. Detection of Sequence-Specific DNA with a Morpholino-Functionalized Silicon Chip. Anal Methods 2015, 7, 2406–2412. (29) Gao, H.-L.; Wang, M.; Wu, Z.-Q.; Wang, C.; Wang, K.; Xia, X.-H. MorpholinoFunctionalized Nanochannel Array for Label-Free Single Nucleotide Polymorphisms Detection. Anal Chem 2015, 87, 3936–3941. (30) Li, S. J.; Li, J.; Wang, K.; Wang, C.; Xu, J. J.; Chen, H. Y.; Xia, X. H.; Huo, Q. A Nanochannel Array-Based Electrochemical Device for Quantitative Label-Free DNA Analysis. ACS Nano 2010, 4, 6417–6424. (31) Wang, X.; Smirnov, S. Label-Free DNA Sensor Based on Surface Charge Modulated Ionic Conductance. ACS Nano 2009, 3, 1004–1010. (32) Zhang, G. J.; Luo, Z. H.; Huang, M. J.; Tay, G. K.; Lim, E. J. Morpholino-Functionalized Silicon Nanowire Biosensor for Sequence-Specific Label-Free Detection of DNA. Biosens Bioelectron 2010, 25, 2447–2453. (33) Lagendijk, A. K.; Moulton, J. D.; Bakkers, J. Revealing Details: Whole Mount microRNA in Situ Hybridization Protocol for Zebrafish Embryos and Adult Tissues. Biol. Open 2012, 1, 566–569.

ACS Paragon Plus Environment

ACS Sensors

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

(34) Martins, D.; Levicky, R.; Song, Y.-A. Enhancing the Speed of Morpholino-DNA Biosensor by Electrokinetic Concentration of DNA in a Microfluidic Chip. Biosens Bioelectron 2015, 72, 87–94. (35) Tercero, N.; Wang, K.; Levicky, R. Capacitive Monitoring of Morpholino-DNA Surface Hybridization: Experimental and Theoretical Analysis. Langmuir 2010, 26, 14351–14358. (36) Debye, V. P.; Huckel, E. Zur Theorie Der Elektrolyte. Phys. Z. 1923, 24, 185–206. (37) Stillinger Jr., F. H.; Lovett, R. Ion-Pair Theory of Concentrated Electrolytes. I. Basic Concepts. J Chem Phys 1968, 48, 3858–3868. (38) Record, M. T.; Anderson, C. F.; Lohman, T. M. Thermodynamic Analysis of Ion Effects on the Binding and Conformational Equilibria of Proteins and Nucleic Acids: The Roles of Ion Association or Release, Screening, and Ion Effects on Water Activity. Q Rev Biophys 1978, 11, 103–178. (39) Tercero, N.; Wang, K.; Gong, P.; Levicky, R. Morpholino Monolayers: Preparation and Label-Free DNA Analysis by Surface Hybridization. J Am Chem Soc 2009, 131, 4953– 4961. (40) Gurtner, C.; Tu, E.; Jamshidi, N.; Haigis, R. W.; Onofrey, T. J.; Edman, C. F.; Sosnowski, R.; Wallace, B.; Heller, M. J. Microelectronic Array Devices and Techniques for Electric Field Enhanced DNA Hybridization in Low-Conductance Buffers. Electrophoresis 2002, 23, 1543–1550.

ACS Paragon Plus Environment

Page 12 of 16

Page 13 of 16

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

ACS Sensors

Figure 1. (A) Molecular structure of the morpholino (MO) probe with detail of the end modifications. (B) Effect of applied potentials on hybridization of DNA targets to complementary DNA and MO probe layers. Background colors delineate 16 min periods of different applied potentials as indicated. Potentials are expressed relative to a Ag/AgCl/3 mol L-1 NaCl reference electrode. For clarity, the curve for DNA probes in 0.1 mol L-1 buffer was offset by multiplying it by 1.5. Experiments were carried out at room temperature, using 10 nmol L-1 target concentrations. (C) CV data for hybridization to MO probes in 0.01 mol L-1 buffer in part (B). Inset: A full CV trace. Main plot: Enlarged region corresponding to the red rectangle in the inset, showing progress of hybridization as a function of applied potential.

ACS Paragon Plus Environment

ACS Sensors

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

Page 14 of 16

Figure 2. (A) Hybridization of DNA targets to complementary MO probes under alternating attractive (0.3 V) and repulsive (-0.3 V) pulses, as a function of temperature. (B) Comparison of hybridization responses for full match targets to that for targets with a single central mismatch, at 35 oC. In all cases targets were present as a 10 nmol L-1 solution in 0.01 mol L-1 PBS, pH 7.

ACS Paragon Plus Environment

Page 15 of 16

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

ACS Sensors

Figure 3. Potential profile V(x) during a denaturation round, calculated for a 0.01 mol L-1 phosphate buffer and a MO-DNA duplex coverage of 5 × 1011 cm-2. The electrode/MCH interface is at x = 0. The potential is referenced relative to the potential of zero charge for morpholino-modified, MCH-passivated electrodes. Most of the potential drop occurs across the MCH passivation layer.

ACS Paragon Plus Environment

ACS Sensors

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

For TOC Only

ACS Paragon Plus Environment

Page 16 of 16