Covalent Modification of Silicon Nitride Nanopore ... - ACS Publications

Oct 25, 2017 - beam sculpting, track-etch technique, focused-ion-beam milling, ... Other characterizations such as energy-dispersive spec- trometry (E...
6 downloads 8 Views 8MB Size
This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Article Cite This: ACS Omega 2017, 2, 7127-7135

http://pubs.acs.org/journal/acsodf

Covalent Modification of Silicon Nitride Nanopore by Amphoteric Polylysine for Short DNA Detection Bohua Yin,† Wanyi Xie,† Liyuan Liang,* Yunsheng Deng, Shixuan He, Feng He, Daming Zhou, Chaker Tlili,* and Deqiang Wang* Chongqing Institute of Green and Intelligent Technology, Chinese Academy of Sciences, Chongqing 400714, P. R. China S Supporting Information *

ABSTRACT: In this work, we demonstrate a chemical modification approach, by means of covalent-bonding amphoteric poly-L-lysine (PLL) on the interior nanopore surface, which could intensively protect the pore from etching when exposed in the electrolyte under various pH conditions (from pH 4 to 12). Nanopore was generated via simple current dielectric breakdown methodology, covalent modification was performed in three steps, and the functional nanopore was fully characterized in terms of chemical structure, hydrophilicity, and surface morphology. I−V curves were recorded under a broad range of pH stimuli to evaluate the stability of the chemical bonding layer; the plotted curves demonstrated that nanopore with a covalent bonding layer has good pH tolerance and showed apparent reversibility. In addition, we have also measured the conductance of modified nanopore with varied KCl concentration (from 0.1 mM to 1 M) at different pH conditions (pHs 5, 7, 9, and 11). The results suggested that the surface charge density does not fluctuate with variation in salt concentration, which inferred that the SiNx nanopore was fully covered by PLL. Moreover, the PLL functionalized nanopore has realized the detection of single-stranded DNA homopolymer translocation under bias voltage of 500 mV, and the 20 nt homopolymers could be evidently differentiated in terms of the current amplitude and dwell time at pHs 5, 8, and 11. nanochannel.67−71 Nanopore fabrication technologies along with a wide range of supporting membrane materials have been intensively developed, including electric beam lithography, ion beam sculpting, track-etch technique, focused-ion-beam milling, and dielectric current breakdown.72−81 Our group has successfully generated nanopore with tunable diameter on silicon nitride and graphene based on dielectric breakdown methodology, and eventually realized the detection of short single-stranded DNA (ssDNA) homopolymers and tetracycline; a theoretical simulation of the ion transport properties in the pore channel has also been investigated in the light of Poisson and Nernst−Planck equation.82−85 To further exploit the application of abiotic nanopore toward tailoring single nucleotide resolution and specific target sensing, numerous modification strategies (containing covalent bonding86−93 and physical absorption94−98) have been employed to help the understanding of fundamental chemical interactions at the nanoscale/single-molecule level and modulate the surface properties as well. In the present work, we displayed the design, preparation, characterization, and application of a synthetic nanopore based on SiNx membrane by covalent attachment of hydrophilic, positively charged poly-L-lysine (PLL) on the wall

1. INTRODUCTION Nanopore sensing is an ultra-sensitive, flexible, and label-free approach for single-molecule measurement and DNA sequencing.1−11 In the past decade, the single-nanometer-scale pores demonstrated great capability for the detection, identification, and characterization of a variety of analytes, such as nanoparticles,12−17 and biomolecules like bacteria,18 protein, 19−26 antibody, 27,28 nucleic acid, 29−32 DNA, 33−35 RNA,36,37 and so on. In this approach, a membrane with a nanopore was sandwiched in a flow cell containing an electrolyte solution, and the analytes (usually some charged molecules or particles) were driven to translocate through the nanopore by applying an electric field across the monolayer membrane, during which a temporary resistive pulse was produced for the characterization of molecular information of analytes. Solid-state nanopores exhibited remarkable chemical, thermal, and mechanical stability and extraordinary versatility in terms of the size, shape, and surface properties, as well as structural robustness compared with their biological counterparts.38,39 A large number of researches have been reported on DNA sequencing with solid-state nanopores based on different substrates, such as silicon oxide, 40−42 silicon nitride (SiNx),43−46 aluminum oxide,47,48 molybdenum sulfide,49−52 boron nitride,53−56 graphene,57,58 and polymer,59,60 whereas other work was also displayed by means of nanotube61−66 and © 2017 American Chemical Society

Received: August 24, 2017 Accepted: October 6, 2017 Published: October 25, 2017 7127

DOI: 10.1021/acsomega.7b01245 ACS Omega 2017, 2, 7127−7135

ACS Omega

Article

Figure 1. AFM pictures for bare (A) and PLL-modified (B) SiNx nanochips with a mean height of the particles of 1.4 nm. TEM pictures for bare SiNx chip with nanopore 16 nm in diameter (C) and PLL-modified (D) SiNx nanopore 11 nm in diameter.

SiNx membrane on which nanopore was generated via a current-stimulus dielectric breakdown. The freshly prepared nanopore underwent three steps of chemical reactions, the buffer or water rinse was applied for each step to remove the excess reactants and any physical absorption on the interior surface. In terms of the characterization of the modified functional groups on nanopore, IR was firstly applied to check the characteristic groups of the chemical skeleton. Unfortunately, the expected absorption for −NH− and CO stretch was not observed, this might be due to the very limited amount of functional molecules on the chip surface, which exhibits an extremely low absorption, and is thus invisible on the IR spectrum. Also, the hydrophilicity evidently increased after the modification based on the contact angle, as presented in Figure S1. Other characterizations such as energy-dispersive spectrometry (EDS) and X-ray photoelectron spectroscopy (XPS)

of nanopore. The high biocompatibility and pH sensitivity of amphoteric PLL provide an insight for further investigation of the interaction between DNA and protein with the nanopore fabricated by the dielectric breakdown.99,100 The chemical modification procedure was performed in three steps according to the reported procedure,101 complete characterizations have been implemented for the modified nanopore. We have looked into the influence of conductance by varying the concentration and pH value of the electrolyte in situ and unveiled the correlation of conductance and surface properties of modified nanopores with salt concentration and pH value. The PLLcoated nanopore was applied ultimately for short ssDNA recognition under 500 mV as the input voltage.

2. RESULTS AND DISCUSSION 2.1. Fabrication, Modification, and Characterization of SiNx Nanopore. This work was performed based on the 7128

DOI: 10.1021/acsomega.7b01245 ACS Omega 2017, 2, 7127−7135

ACS Omega

Article

have been applied for the element analysis of the modified and unmodified surface (Tables S1 and S2). The comparative results demonstrated that elements C and N were increased on the modified surface than on the bare one because of the introduction of abundant aliphatic groups on both the surface and inside the nanopore; elements O and Si were decreased, as these two elements were not the main components of the modified structure and thus occupied less percentage of the total elements on the modified nanochip than on the bare one. Although change in the value is different, the variation tendency is consistent between the two techniques. Optical property was examined by Raman spectroscopy, the result displayed no obvious distinction for the modified and unmodified chips. Both atomic force microscopy (AFM) and transmission electron microscopy (TEM) technologies have been employed to analyze the surface topography and morphology of modified nanochip. The little white particles with an average height of 1.4 nm in Figure 1B revealed the agglomeration of PLL chain on a dry nanopore surface, which differs from the relatively clearer bare chip surface in Figure 1A. Also, the TEM images demonstrated a characteristic size difference before (16 nm, Figure 1C) and after the covalent modification of the nanopore (11 nm, Figure 1D); a white ring observed at the periphery of nanopore indicated that the functional layer was immobilized around the pore. 2.2. Investigation of the Stability of Modified Nanopore. I−V characterization was conducted on a patch clamp amplifier for the freshly modified nanopore mounted on the fluidic setup with a buffer solution of pH 8. The pore diameter was estimated based on an empirical formula:102 G = σ[(4L /πd 2) + (l/d)]−1

Figure 2. I−V curves of unmodified and PLL-modified SiNx nanopore after a month (black: bare SiNx nanopore; red: PLL-modified nanopore; blue and violet: PLL-modified nanopore after 10 and 30 days).

the stability of chemical bonding layer, the I−V curves were recorded with a broad range of pH stimuli (from 4 to 12) with 1 M KCl. The results in Figure S3 indicate that the slope of I− V curves of modified nanopore does not change significantly at the pH from 4 to 8, but gradually became larger under very basic condition (pH from 9 to 12), presumably owing to the polarity switch of the modified nanopore from acidic to basic condition. Furthermore, the I−V curve returns to the initial status when switching the pH value to 8. It could be elucidated that the covalent layer prevents the nanopore from etching by alteration of pH value. Table S3 summarizes the details of pore size at different pHs with 1 M KCl as the electrolyte. It is worth mentioning that no current rectification was observed on the fundamental characteristics of the current response between bare and PLL-bonded nanopore, which might depend on the surface charge and nanopore geometry. 2.4. Ionic Conductance Behavior and Surface Charge Dependence of PLL-Modified Nanopore. To disclose the correlation of amphoteric PLL and surface charge property of functional pore with the salt concentration and pH condition, we have recorded the I−V curves for the modified pore with varying KCl concentration (0.1 mM, 1 mM, 10 mM, 100 mM, and 1 M) at each pH condition (pHs 5, 7, 9, and 11). Figure 3 depicts the conductance versus KCl concentration for modified pore (pore size was calculated to be around 2.5 nm according to formula 1) at four pH conditions. The conductance of the device at all of the four pH conditions matches that of the model described as fixed charge density behavior.102−105 The experimentally extracted conductance of the modified pore does not fluctuate at low salt concentration but gradually increases with the augmentation of salt concentration at all of the four pH conditions; presumably, the surface charge is dominated by PLL at low salt concentrations. 2.5. Short ssDNA Homopolymers Translocation through PLL-Modified SiNx Nanopore. The SiNx nanopore fabricated by the dielectric breakdown approach was subsequently modified by the covalent bonding PLL and applied afterward for the detection of 20 nt short homopolymers (with chain length of around 7 nm). Figure 4 displays the baseline and current blockades determined for 10 pM of single-stranded poly(dG)20, poly(dT)20 from a PLL-modified SiNx nanopore

(1)

In the formula, σ is the electrical conductivity (normally 11.93 S/m for 1 M KCl, pH 8), L is the thickness of SiNx membrane (20 nm in this work), G stands for the pore conductance, and d represents the pore diameter. The diameter of a bare SiNx nanopore (all of the diameters presented in the text were calculated according to the aforementioned formula) reduced from 3.5 nm to approximately 1 nm after covalently bonding PLL on the interior surface; the I−V curve was collected discontinuously for this pore during more than a month to monitor the variation in pore size with time under the same condition. As correlation between the slope of I−V curve and the nanopore size is reported, the greater the slope, the larger the pore size; so, we could intuitively deduce the size change from the I−V curve. Figure 2 displays the I−V characterization acquired on the unmodified and PLL-modified SiNx nanopores. The slope of the curves illustrated that nanopore significantly dwindled after PLL modification; it turned to approximately 1 nm and sustained for more than a month. Figure S2 demonstrates the correlation between modified nanopore size and time. The diameter maintained at around 1 nm for a month, but slightly enlarged to above 1 nm afterward, which might be on account of the damage during frequent electrolyte/water exchange for detection and storage. This work was scaled up on tens of PLL-modified nanopores and proved to be a reliable modification procedure for further application on other nanopore substrates. 2.3. pH Tolerance of Modified SiNx Nanopore. Poly-Llysine (PLL) is a type of amphoteric amino acid with a pKa value of 9. Its introduction on the wall of SiNx nanopore could potentially realize the rectification of ion current and surface charge of the interior pore at different pH conditions. To testify 7129

DOI: 10.1021/acsomega.7b01245 ACS Omega 2017, 2, 7127−7135

ACS Omega

Article

consistent with their physical dimension, except in the case pH 8, when the current amplitude for poly(dG)20 is 208.9 ± 2.4 pA, lower than that for poly(dT)20 (254.1 ± 4.7 pA). At pH 8, which is near the isoelectric point of PLL, the modified surface is close to neutral. The lower surface charge together with the conformation of G may be responsible for the smaller current blockade; in addition, base current was also significantly influenced by the distinction of surface charges under diverse pH. Furthermore, the current amplitude decreases with increasing pH value. This could be explained by the fact that under acidic condition, the interaction of the positively charged PLL with the negatively charged DNA is stronger. The DNA probably showed a more complicated conformation rather than a straight chain, thus exhibiting a higher current amplitude during translocation. Moreover, the corresponding dwell time has a similar tendency as the current blockade for poly(dT)20, whereas it is inverse for poly(dG)20, probably due to the conformation diversity of G-quadruplexes.106 It has been reported that ions and pH have the largest effects on the formation of the Gquadruplex.107 In addition, the translocation rate in our system is around 50 μs per base according to the values presented in Table 1, which signifies the existence of a significant interaction of DNA with the covalent bonding PLL on the wall of nanopore; here, the translocation rate has been increased more than an order of magnitude compared with our previously reported results.82,84 We anticipate that our further work will be the exploration of this modification technique on other nanopore substrates.

Figure 3. Conductance of PLL-modified SiNx nanopore with a pore size of around 2.5 nm, which measured at varied KCl concentrations under four pH conditions. The two solid lines are derived from the models called as a fixed charge density (red) and bulk behavior (green); the dash lines stand for the status that is close to the fixed charge density model.

with 500 mV as the applied voltage in the buffer solution of 1 M KCl, pH 8. The original translocation during 20 s is provided in Figure S4; the observed open-pore current is around 700 pA. The presented stochastic set of downward spikes in Figure 4B inferred the translocation of single-molecule DNA homopolymer through the nanopore (Figure 4A is the baseline ionic current without the addition of analytes), and the amplified blockade shown alongside (Figure 4C) indicated that the current amplitude for poly(dT)20 is higher than that for poly(dG)20 at pH 8. This was further exhibited by the histograms in Figure 5, which were extracted from hundreds of events of the two homopolymers at pHs 5, 8, and 11, the mean current amplitudes, and the corresponding dwell time (Figure 6, with the fitting method of exponential decay) for poly(dG)20 and poly(dT)20 at the three different pH conditions listed in Table 1. It is not difficult to discover that the current amplitude for poly(dG)20 is higher than that for poly(dT)20, which is

3. EXPERIMENTAL SECTION 3.1. Materials. The nanochips (SiNx membrane growing on both sides of the silica substrate and free standing on top of window) were purchased from Nanopore Solution, Portugal, with the thickness of 20 nm and the window size of 10 μm × 10 μm. 3-(Triethoxysilyl) propylamine, glutaraldehyde (25% solution in H2O), poly-L-lysine (PLL, Mw 1000−5000), methanol anhydrous were ordered from Aldrich, Shanghai. Phosphate buffer was prepared by using monopotassium phosphate and dipotassium phosphate obtained from Admas,

Figure 4. Baseline (A) and events (B) of 10 pM single-stranded poly(dT)20, poly(dG)20 translocation through PLL-modified SiNx nanopore with 500 mV voltage in the buffer solution of 1 M KCl, pH 8. (C) Zoomed-in event from the trace in (B). 7130

DOI: 10.1021/acsomega.7b01245 ACS Omega 2017, 2, 7127−7135

ACS Omega

Article

Figure 5. Gaussian distributions of normalized histograms of current amplitudes (a1−b3) for poly(dG)20 and poly(dT)20 at pHs 5, 8, and 11.

with PBS at 50 °C to remove the physically absorbed PLL on the interior pore surface, and finally washed with deionized water for the subsequent use. 3.3. Characterizations and Measurement. The modified nanopore was characterized in terms of chemical structure (IR, Agilent Cary 630), element analysis (EDS, JSM-7800F, and XPS, Escalab 250Xi), optical properties (Raman, inVia Reflex), surface morphology (AFM, Brucker Dimension EDGE, and TEM, FEI tecnai F20), and hydrophilicity (Contact Angle, DSA100). Electric pulse breakdown was performed on a Keithley 2450 equipment, which was controlled by a Labview program to measure the conductivity of the SiNx membrane through current−voltage (I−V) curves. I−V characteristics and ion current blockade measurements were recorded on a patch clamp amplifier (Axopatch 200B) with the supersensitive electronics housed inside a Faraday cage. 3.4. Stability of Modified Nanopore. The polylysinemodified nanopore was investigated in terms of the stability of the covalent coating layer under varied pH conditions. Nanochip with functional nanopore generated by electric pulse breakdown was assembled inside a PET or Teflon flow cell. Electrolyte containing 10 mM Tris−HCl, 1 M KCl, and 1 mM EDTA (pH 8) was filled in both reservoirs (cis and trans) before the insertion of Ag/AgCl electrode, which was connected to the resistive feedback amplifier. The I−V curves were recorded over 30 days without changing the pH value and

Shanghai. Sulfuric acid, hydrogen peroxide, and ethanol absolute were obtained from Admas, Shanghai. All of the chemicals and solvents were used as received without any further treatment. HPLC-grade water was from Molecular 1850D. 3.2. Method of Covalent Modification of PLL. The experiment was carried out on the nanochip with a freestanding SiNx membrane. First, the nanopore was fabricated on SiNx membrane via the electric pulse breakdown approach by embedding the chip in flow cell, which was filled with electrolyte of 10 mM Tris−HCl, 1 M KCl, and 1 mM ethylenediaminetetraacetic acid (EDTA, pH 8) and by using Ag/AgCl as a reference electrode. Second, the freshly prepared nanopore was treated with piranha solution for 30 min followed by rinsing with deionized water to clean and simultaneously introduce hydroxyl group on the interior nanopore surface. The schematic diagram of the chemical modification of SiNx nanopore by PLL is displayed in Scheme 1. Afterwards, the chip with nanopore was soaked in 5% of (3aminopropyl)triethoxysilane anhydrous methanol solution for 30 min, followed by washing with ethanol absolute before treating with 5% glutaraldehyde in phosphate-buffered saline (PBS, 10 mM, pH 7.4) for 1 h and then washing with PBS. The terminated aldehyde group on the wall of the nanopore was finally cross-linked with the lateral amine group on polylysine (Mw 1000−5000, 1 mg/mL) in PBS for 2 h, followed by rinsing 7131

DOI: 10.1021/acsomega.7b01245 ACS Omega 2017, 2, 7127−7135

ACS Omega

Article

Figure 6. Normalized histograms of dwell time (a4−b6) for poly(dG)20 and poly(dT)20 at pHs 5, 8, and 11, with fitting method of exponential decay.

two single-stranded homopolymer nucleotides, poly(dG)20 and poly(dT)20, were respectively introduced into the cis reservoir for translocation detection under bias voltage of 500 mV at pHs 5, 8, and 11.

Table 1. Details of Current Amplitude and Dwell Time for poly(dG)20 and poly(dT)20 at pHs 5, 8, and 11 current amplitude (pA)

dwell time (ms)

pH variation

poly(dG)20

poly(dT)20

poly(dG)20

poly(dT)20

pH 5 pH 8 pH 11

421.7 ± 3.0 208.9 ± 2.4 229.2 ± 0.9

320.2 ± 2.6 254.1 ± 4.7 181.6 ± 1.0

0.54 ± 0.03 0.66 ± 0.03 0.83 ± 0.08

1.02 ± 0.10 0.94 ± 0.02 0.82 ± 0.10

4. CONCLUSIONS To sum up, we have presented the integration of positively charged amino acid PLL into SiNx nanopore via covalent bonding and have performed overall characterization for functional nanopore. The stability of chemical coating layer of nanopore has been evaluated under variation of pH value by measuring the I−V curves on the patch clamp. An obvious tendency was observed that the size of PLL bonding nanopore remains unchanged under acidic conditions but enlarged gradually at pH above 9, which depends on the surface charge

electrolyte concentration and increasing the voltage from −200 mV to +200 mV, at a scanning rate of 50 mV/min. Moreover, the pH gradient (from 4 to 12) has been employed during I−V characterization. Conductance was measured with electrolyte concentration varying from 0.1 mM to 1 M. Finally, 10 pM of

Scheme 1. Schematic Diagram of the Chemical Modification of SiNx Nanopore by Polylysine

7132

DOI: 10.1021/acsomega.7b01245 ACS Omega 2017, 2, 7127−7135

ACS Omega

Article

(7) Goodwin, S.; McPherson, J. D.; McCombie, W. R. Coming of age: ten years of next-generation sequencing technologies. Nat. Rev. Genet. 2016, 17, 333−351. (8) Shi, W.; Friedman, A. K.; Baker, L. A. Nanopore Sensing. Anal. Chem. 2017, 89, 157−188. (9) Garrido-Cardenas, J.; Garcia-Maroto, F.; Alvarez-Bermejo, J.; Manzano-Agugliaro, F. DNA Sequencing Sensors: An Overview. Sensors 2017, 17, 588−602. (10) Montagne, F.; Blondiaux, N.; Bojko, A.; Pugin, R. Molecular transport through nanoporous silicon nitride membranes produced from self-assembling block copolymers. Nanoscale 2012, 4, 5880− 5886. (11) Bojko, A.; Andreatta, G.; Montagne, F.; Renaud, P.; Pugin, R. Fabrication of thermo-responsive nano-valve by grafting-to in melt of poly(N-isopropylacrylamide) onto nanoporous silicon nitride membranes. J. Membr. Sci. 2014, 468, 118−125. (12) Astier, Y.; Datas, L.; Carney, R.; Stellacci, F.; Gentile, F.; DiFabrizio, E. Artificial Surface-Modified Si3N4 Nanopores for Single Surface-Modified Gold Nanoparticle Scanning. Small 2011, 7, 455− 459. (13) Goyal, G.; Freedman, K. J.; Kim, M. J. Gold Nanoparticle Translocation Dynamics and Electrical Detection of Single Particle Diffusion Using Solid-State Nanopores. Anal. Chem. 2013, 85, 8180− 8187. (14) Liu, S.; Yuzvinsky, T. D.; Schmidt, H. Effect of fabricationdependent shape and composition of solid-state nanopores on single nanoparticle detection. ACS Nano 2013, 7, 5621−5627. (15) Arjmandi, N.; Van Roy, W.; Lagae, L. Measuring mass of nanoparticles and viruses in liquids with nanometer-scale pores. Anal. Chem. 2014, 86, 4637−4641. (16) Nam, S.; Choi, I.; Fu, C. C.; Kim, K.; Hong, S.; Choi, Y.; Zettl, A.; Lee, L. P. Graphene nanopore with a self-integrated optical antenna. Nano Lett. 2014, 14, 5584−5589. (17) Cao, Y.; Lin, Y.; Qian, R.-C.; Ying, Y.-L.; Si, W.; Sha, J.; Chen, Y.; Long, Y.-T. Evidence of single-nanoparticle translocation through a solid-state nanopore by plasmon resonance energy transfer. Chem. Commun. 2016, 52, 5230−5233. (18) Loman, N. J.; Quick, J.; Simpson, J. T. A complete bacterial genome assembled de novo using only nanopore sequencing data. Nat. Methods 2015, 12, 733−735. (19) 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. (20) Rodriguez-Larrea, D.; Bayley, H. Multistep protein unfolding during nanopore translocation. Nat. Nanotechnol. 2013, 8, 288−295. (21) Rosen, C. B.; Rodriguez-Larrea, D.; Bayley, H. Single-molecule site-specific detection of protein phosphorylation with a nanopore. Nat. Biotechnol. 2014, 32, 179−181. (22) Rodriguez-Larrea, D.; Bayley, H. Protein co-translocational unfolding depends on the direction of pulling. Nat. Commun. 2014, 5, 4841−4847. (23) Plesa, C.; Ruitenberg, J. W.; Witteveen, M. J.; Dekker, C. Detection of Individual Proteins Bound along DNA Using Solid-State Nanopores. Nano Lett. 2015, 15, 3153−3158. (24) Kennedy, E.; Dong, Z.; Tennant, C.; Timp, G. Reading the primary structure of a protein with 0.07 nm3 resolution using a subnanometre-diameter pore. Nat. Nanotechnol. 2016, 11, 968−976. (25) Lin, X.; Ivanov, A. P.; Edel, J. B. Selective single molecule nanopore sensing of proteins using DNA aptamer-functionalised gold nanoparticles. Chem. Sci. 2017, 8, 3905−3912. (26) Wloka, C.; Van Meervelt, V.; van Gelder, D.; Danda, N.; Jager, N.; Williams, C. P.; Maglia, G. Label-Free and Real-Time Detection of Protein Ubiquitination with a Biological Nanopore. ACS Nano 2017, 11, 4387−4394. (27) Wang, S.; Haque, F.; Rychahou, P. G.; Evers, B. M.; Guo, P. Engineered Nanopore of Phi29 DNA-Packaging Motor for Real-Time Detection of Single Colon Cancer Specific Antibody in Serum. ACS Nano 2013, 7, 9814−9822.

and polymer extension. This process was reproducible and it has been demonstrated that the nanopore size is reversible when the pH value decreases down to the acidic range. The functional nanopore was extremely stable and could remain unchanged evidently for a month. We have investigated the salt dependence of modified nanopore conductance and discovered that our device followed a fixed charge density model in a lowsalt regime. The modified pore was ultimately employed for the detection of ssDNA homopolymer translocation under bias voltage of 500 mV, thereby realizing the distinction between the 20 nt homopolymers by ionic current amplitude and dwell time. Furthermore, the amphoteric PLL on the wall of nanopore showed virtually no current rectification in this system, whereas it could provide an insight into further investigation of the interaction between DNA and protein and suggest future applications in biosensor technology.



ASSOCIATED CONTENT

* Supporting Information S

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsomega.7b01245. Details of characterization of PLL-modified nanopore and its correlation with time and distinct pH value, the translocation events with the modified nanopore (PDF)



AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected] (L.L.). *E-mail: [email protected] (C.T.). *E-mail: [email protected] (D.W.). ORCID

Liyuan Liang: 0000-0003-3990-064X Author Contributions †

B.Y. and W.X. contributed equally to the paper.

Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We would address our gratitude to Mr. Jianqiang Fu for TEM characterization. The authors acknowledge the support from the National Natural Science Foundation of China (Grant No. 51503207, 51503206, 41603089, 61471336), the Key Basic Sciences and Frontier Project of Chongqing (Grant No. cstc2017jcyjB0105), and the Joint-Scholar 2015 of West Light Foundation of the Chinese Academy of Sciences (CAS) and Instrument development program of CAS (YZ201568).



REFERENCES

(1) Shendure, J. A.; Porreca, G. J.; Church, G. M.; Gardner, A. F.; Hendrickson, C. L.; Kieleczawa, J.; Slatko, B. E. Overview of DNA sequencing strategies. Curr. Protoc. Mol. Biol. 2011. Chapter 7, Unit 7.1., 1−23. DOI: 10.1002/0471142727.mb0701s96. (2) Gracheva, M. E. Nanopore-Based Technology; Humana Press, 2012; pp 1−265. (3) Wanunu, M. Nanopores: A journey towards DNA sequencing. Phys. Life Rev. 2012, 9, 125−158. (4) Marx, V. Nanopores: a sequencer in your backpack. Nat. Methods 2015, 12, 1015−1018. (5) Liu, L.; Wu, H. C. DNA-Based Nanopore Sensing. Angew. Chem., Int. Ed. 2016, 55, 15216. (6) Deamer, D.; Akeson, M.; Branton, D. Three decades of nanopore sequencing. Nat. Biotechnol. 2016, 34, 518−524. 7133

DOI: 10.1021/acsomega.7b01245 ACS Omega 2017, 2, 7127−7135

ACS Omega

Article

(28) Farimani, A. B.; Heiranian, M.; Min, K.; Aluru, N. R. Antibody Subclass Detection Using Graphene Nanopores. J. Phys. Chem. Lett. 2017, 8, 1670−1676. (29) Nelson, T.; Zhang, B.; Prezhdo, O. V. Detection of nucleic acids with graphene nanopores: ab initio characterization of a novel sequencing device. Nano Lett. 2010, 10, 3237−3242. (30) Venkatesan, B. M.; Bashir, R. Nanopore sensors for nucleic acid analysis. Nat. Nanotechnol. 2011, 6, 615−624. (31) Ying, Y.-L.; Zhang, J.; Gao, R.; Long, Y.-T. Nanopore-based sequencing and detection of nucleic acids. Angew. Chem., Int. Ed. 2013, 52, 13154−13161. (32) Shi, J.; Hou, J.; Fang, Y. Recent advances in nanopore-based nucleic acid analysis and sequencing. Microchim. Acta 2016, 183, 925− 939. (33) Howorka, S.; Cheley, S.; Bayley, H. Sequence-specific detection of individual DNA strands using engineered nanopores. Nat. Biotechnol. 2001, 19, 636−639. (34) Stoddart, D.; Heron, A. J.; Mikhailova, E.; Maglia, G.; Bayley, H. Single-nucleotide discrimination in immobilized DNA oligonucleotides with a biological nanopore. Proc. Natl. Acad. Sci. 2009, 106, 7702−7707. (35) Rincon-Restrepo, M.; Mikhailova, E.; Bayley, H.; Maglia, G. Controlled Translocation of Individual DNA Molecules through Protein Nanopores with Engineered Molecular Brakes. Nano Lett. 2011, 11, 746−750. (36) Clamer, M.; Höfler, L.; Mikhailova, E.; Viero, G.; Bayley, H. Detection of 3′-End RNA Uridylation with a Protein Nanopore. ACS Nano 2014, 8, 1364−1374. (37) Zahid, O. K.; Wang, F.; Ruzicka, J. A.; Taylor, E. W.; Hall, A. R. Sequence-Specific Recognition of MicroRNAs and Other Short Nucleic Acids with Solid-State Nanopores. Nano Lett. 2016, 16, 2033−2039. (38) Dekker, C. Solid-state nanopores. Nat. Nanotechnol. 2007, 2, 209−215. (39) Haque, F.; Li, J.; Wu, H. C.; Liang, X. J.; Guo, P. Solid-State and Biological Nanopore for Real-Time Sensing of Single Chemical and Sequencing of DNA. Nano Today 2013, 8, 56−74. (40) Steinbock, L. J.; Bulushev, R. D.; Krishnan, S.; Raillon, C.; Radenovic, A. DNA Translocation through Low-Noise Glass Nanopores. ACS Nano 2013, 7, 11255−11262. (41) Bulushev, R. D.; Marion, S.; Radenovic, A. Relevance of the Drag Force during Controlled Translocation of a DNA−Protein Complex through a Glass Nanocapillary. Nano Lett. 2015, 15, 7118− 7125. (42) He, H.; Xu, X.; Wang, P.; Chen, L.; Jin, Y. The facile surface chemical modification of a single glass nanopore and its use in the nonenzymatic detection of uric acid. Chem. Commun. 2015, 51, 1914− 1917. (43) Squires, A. H.; Hersey, J. S.; Grinstaff, M. W.; Meller, A. A nanopore-nanofiber mesh biosensor to control DNA translocation. J. Am. Chem. Soc. 2013, 135, 16304−16307. (44) Venta, K.; Shemer, G.; Puster, M.; Rodríguez-Manzo, J. A.; Balan, A.; Rosenstein, J. K.; Shepard, K.; Drndić, M. Differentiation of short, single-stranded DNA homopolymers in solid-state nanopores. ACS Nano 2013, 7, 4629−4636. (45) Liang, Z.; Tang, Z.; Li, J.; Hu, R.; Yu, D.; Zhao, Q. Interaction prolonged DNA translocation through solid-state nanopores. Nanoscale 2015, 7, 10752−10759. (46) Shekar, S.; Niedzwiecki, D. J.; Chien, C. C.; Ong, P.; Fleischer, D. A.; Lin, J.; Rosenstein, J. K.; Drndic, M.; Shepard, K. L. Measurement of DNA Translocation Dynamics in a Solid-State Nanopore at 100 ns Temporal Resolution. Nano Lett. 2016, 16, 4483− 4489. (47) Venkatesan, B. M.; Dorvel, B.; Yemenicioglu, S.; Watkins, N.; Petrov, I.; Bashir, R. Highly Sensitive, Mechanically Stable Nanopore Sensors for DNA Analysis. Adv. Mater. 2009, 21, 2771−2776. (48) Venkatesan, B. M.; Shah, A. B.; Zuo, J.-M.; Bashir, R. DNA Sensing Using Nanocrystalline Surface-Enhanced Al2O3Nanopore Sensors. Adv. Funct. Mater. 2010, 20, 1266−1275.

(49) Farimani, A. B.; Min, K.; Aluru, N. R. DNA Base Detection Using a Single-Layer MoS2. ACS Nano 2014, 8, 7914−7922. (50) Liu, K.; Feng, J.; Kis, A.; Radenovic, A. Atomically thin molybdenum disulfide nanopores with high sensitivity for DNA translocation. ACS Nano 2014, 8, 2504−2511. (51) 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, 1070−1076. (52) Smolyanitsky, A.; Yakobson, B. I.; Wassenaar, T. A.; Paulechka, E.; Kroenlein, K. A MoS2-Based Capacitive Displacement Sensor for DNA Sequencing. ACS Nano 2016, 10, 9009−9016. (53) Zhou, Z.; Hu, Y.; Wang, H.; Xu, Z.; Wang, W.; Bai, X.; Shan, X.; Lu, X. DNA Translocation through Hydrophilic Nanopore in Hexagonal Boron Nitride. Sci. Rep. 2013, 3, No. 3287. (54) Gu, Z.; Zhang, Y.; Luan, B.; Zhou, R. DNA translocation through single-layer boron nitride nanopores. Soft Matter 2016, 12, 817−823. (55) Zhang, L.; Wang, X. DNA Sequencing by Hexagonal Boron Nitride Nanopore: A Computational Study. Nanomaterials 2016, 6, 111−121. (56) de Souza, F. A. L.; Amorim, R. G.; Scopel, W. L.; Scheicher, R. H. Electrical detection of nucleotides via nanopores in a hybrid graphene/h-BN sheet. Nanoscale 2017, 9, 2207−2212. (57) 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, 2915−2921. (58) Schneider, G. F.; Kowalczyk, S. W.; Calado, V. E.; Pandraud, Gg; Zandbergen, H. W.; Vandersypen, L. M. K.; Dekker, C. DNA Translocation through Graphene Nanopores. Nano Lett. 2010, 10, 3163−3167. (59) Siwy, Z. S. Ion-Current Rectification in Nanopores and Nanotubes with Broken Symmetry. Adv. Funct. Mater. 2006, 16, 735−746. (60) Buchsbaum, S. F.; Nguyen, G.; Howorka, S.; Siwy, Z. S. DNAModified Polymer Pores Allow pH- and Voltage-Gated Control of Channel Flux. J. Am. Chem. Soc. 2014, 136, 9902−9905. (61) Fan, R.; Karnik, R.; Yue, M.; Li, D.; Majumdar, A.; Yang, P. DNA Translocation in Inorganic Nanotubes. Nano Lett. 2005, 5, 1633−1637. (62) Liu, L.; Yang, C.; Zhao, K.; Li, J.; Wu, H.-C. Ultrashort singlewalled carbon nanotubes in a lipid bilayer as a new nanopore sensor. Nat. Commun. 2013, 4, No. 2989. (63) Kim, H. S.; Lee, S. J.; Kim, Y.-H. Distinct Mechanisms of DNA Sensing Based on N-Doped Carbon Nanotubes with Enhanced Conductance and Chemical Selectivity. Small 2014, 10, 774−781. (64) Liu, L.; Xie, J.; Li, T.; Wu, H.-C. Fabrication of nanopores with ultrashort single-walled carbon nanotubes inserted in a lipid bilayer. Nat. Protoc. 2015, 10, 1670−1678. (65) Zhang, S.; Wang, X.; Li, T.; Liu, L.; Wu, H. C.; Luo, M.; Li, J. Sensitive Detection of a Modified Base in Single-Stranded DNA by a Single-Walled Carbon Nanotube. Langmuir 2015, 31, 10094−10099. (66) Yang, N.; Jiang, X. Nanocarbons for DNA sequencing: A review. Carbon 2017, 115, 293−311. (67) Chang, S.; Huang, S.; He, J.; Liang, F.; Zhang, P.; Li, S.; Chen, X.; Sankey, O.; Lindsay, S. Electronic Signatures of all Four DNA Nucleosides in a Tunneling Gap. Nano Lett. 2010, 10, 1070−1075. (68) Wang, D.; Harrer, S.; Luan, B.; Stolovitzky, G.; Peng, H.; AfzaliArdakani, A. Regulating the Transport of DNA through Biofriendly Nanochannels in a Thin Solid Membrane. Sci. Rep. 2014, 4, No. 3985. (69) Gao, H. L.; Wang, M.; Wu, Z. Q.; Wang, C.; Wang, K.; Xia, X. H. Morpholino-functionalized nanochannel array for label-free single nucleotide polymorphisms detection. Anal. Chem. 2015, 87, 3936− 3941. (70) Göpfrich, K.; Li, C.-Y.; Mames, I.; Bhamidimarri, S. P.; Ricci, M.; Yoo, J.; Mames, A.; Ohmann, A.; Winterhalter, M.; Stulz, E.; Aksimentiev, A.; Keyser, U. F. Ion Channels Made from a Single Membrane-Spanning DNA Duplex. Nano Lett. 2016, 16, 4665−4669. 7134

DOI: 10.1021/acsomega.7b01245 ACS Omega 2017, 2, 7127−7135

ACS Omega

Article

amine-terminated polymer brushes. Nanotechnology 2012, 23, 225502−225511. (91) Anderson, B. N.; Muthukumar, M.; Meller, A. pH Tuning of DNA Translocation Time through Organically Functionalized Nanopores. ACS Nano 2013, 7, 1408−1414. (92) Zharov, I.; Khabibullin, A. Surface-modified silica colloidal crystals: nanoporous films and membranes with controlled ionic and molecular transport. Acc. Chem. Res. 2014, 47, 440−449. (93) Tagliazucchi, M.; Szleifer, I. Chemically Modified Nanopores and Nanochannels; Elsevier, 2017. (94) Xie, Y.; Xue, J.; Wang, L.; Wang, X.; Jin, K.; Chen, L.; Wang, Y. Surface Modification of Single Track-Etched Nanopores with Surfactant CTAB. Langmuir 2009, 25, 8870−8874. (95) Wang, L.; Yan, Y.; Xie, Y.; Chen, L.; Xue, J.; Yan, S.; Wang, Y. A method to tune the ionic current rectification of track-etched nanopores by using surfactant. Phys. Chem. Chem. Phys. 2011, 13, 576−581. (96) Tang, Z.; Lu, B.; Zhao, Q.; Wang, J.; Luo, K.; Yu, D. Surface Modification of Solid-State Nanopores for Sticky-Free Translocation of Single-Stranded DNA. Small 2014, 10, 4332−4339. (97) Pla-Roca, M.; Isa, L.; Kumar, K.; Reimhult, E. Selective (Bio)Functionalization of Solid-State Nanopores. ACS Appl. Mater. Interfaces 2015, 7, 6030−6035. (98) Hu, R.; Diao, J.; Li, J.; Tang, Z.; Li, X.; Leitz, J.; Long, J.; Liu, J.; Yu, D.; Zhao, Q. Intrinsic and membrane-facilitated α-synuclein oligomerization revealed by label-free detection through solid-state nanopores. Sci. Rep. 2016, 6, No. 20776. (99) Umehara, S.; Pourmand, N.; Webb, C. D.; Davis, R. W.; Yasuda, K.; Karhanek, M. Current Rectification with Poly-l-Lysine-Coated Quartz Nanopipettes. Nano Lett. 2006, 6, 2486−2492. (100) 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, 603−608. (101) Zhang, L.-X.; Cao, X.-H.; Zheng, Y.-B.; Li, Y.-Q. Covalent modification of single glass conical nanopore channel with 6carboxymethyl-chitosan for pH modulated ion current rectification. Electrochem. Commun. 2010, 12, 1249−1252. (102) Kowalczyk, S. W.; Grosberg, A. Y.; Rabin, Y.; Dekker, C. Modeling the conductance and DNA blockade of solid-state nanopores. Nanotechnology 2011, 22, 315101−315104. (103) Smeets, R. M.; Keyser, U. F.; Krapf, D.; Wu, M. Y.; Dekker, N. H.; Dekker, C. Salt dependence of ion transport and DNA translocation through solid-state nanopores. Nano Lett. 2006, 6, 89− 95. (104) Kumar, A.; Park, K. B.; Kim, H. M.; Kim, K. B. Noise and its reduction in graphene based nanopore devices. Nanotechnology 2013, 24, 495503−495509. (105) Kuan, A. T.; Lu, B.; Xie, P.; Szalay, T.; Golovchenko, J. A. Electrical pulse fabrication of graphene nanopores in electrolyte solution. Appl. Phys. Lett. 2015, 106, 203109−203113. (106) Sengar, A.; Heddi, B.; Phan, A. T. Formation of Gquadruplexes in poly-G sequences: structure of a propeller-type parallel-stranded G-quadruplex formed by a G(1)(5) stretch. Biochemistry 2014, 53, 7718−7723. (107) Selvam, S.; Mandal, S.; Mao, H. Quantification of Chemical and Mechanical Effects on the Formation of the G-Quadruplex and iMotif in Duplex DNA. Biochemistry 2017, 56, 4616−4625.

(71) Di Ventra, M.; Taniguchi, M. Decoding DNA, RNA and peptides with quantum tunnelling. Nat. Nanotechnol. 2016, 11, 117− 126. (72) Kant, K.; Losic, D. Focused Ion Beam (FIB) Technology for Micro- and Nanoscale Fabrications. FIB Nanostructures; Lecture Notes in Nanoscale Science and Technology; Springer: Cham, 2013; Vol. 20, pp 1−22. (73) Choi, S. S.; Park, M.-J.; Yamaguchi, T.; Kim, S.-I.; Park, K.-J.; Park, N. K. Fabrication of nanopore on pyramid. Appl. Surf. Sci. 2014, 310, 196−203. (74) Bai, J.; Wang, D.; Nam, S.-w.; Peng, H.; Bruce, R.; Gignac, L.; Brink, M.; Kratschmer, E.; Rossnagel, S.; Waggoner, P.; Reuter, K.; Wang, C.; Astier, Y.; Balagurusamy, V.; Luan, B.; Kwark, Y.; Joseph, E.; Guillorn, M.; Polonsky, S.; Royyuru, A.; Papa Rao, S.; Stolovitzky, G. Fabrication of sub-20 nm nanopore arrays in membranes with embedded metal electrodes at wafer scales. Nanoscale 2014, 6, 8900− 8906. (75) Kwok, H.; Briggs, K.; Tabardcossa, V. Nanopore Fabrication by Controlled Dielectric Breakdown. PLoS One 2014, 9, No. e92880. (76) Ahuir-Torres, J. I.; Hernández-López, J. M.; Arenas, M. A.; Conde, A.; de Damborenea, J. Synthesis of TiO2 nanopore arrays by pulsed laser treatment and anodic oxidation. Surf. Coat. Technol. 2014, 259, 408−414. (77) Briggs, K.; Charron, M.; Kwok, H.; Le, T.; Chahal, S.; Bustamante, J.; Waugh, M.; Tabard-Cossa, V. Kinetics of nanopore fabrication during controlled breakdown of dielectric membranes in solution. Nanotechnology 2015, 26, No. 084004. (78) Pud, S.; Verschueren, D.; Vukovic, N.; Plesa, C.; Jonsson, M. P.; Dekker, C. Self-Aligned Plasmonic Nanopores by Optically Controlled Dielectric Breakdown. Nano Lett. 2015, 15, 7112−7117. (79) Kudr, J.; Skalickova, S.; Nejdl, L.; Moulick, A.; Ruttkay-Nedecky, B.; Adam, V.; Kizek, R. Fabrication of solid-state nanopores and its perspectives. Electrophoresis 2015, 36, 2367−2379. (80) Arcadia, C. E.; Reyes, C. C.; Rosenstein, J. K. In Situ Nanopore Fabrication and Single-Molecule Sensing with Microscale Liquid Contacts. ACS Nano 2017, 11, 4907−4915. (81) Yanagi, I.; Fujisaki, K.; Hamamura, H.; Takeda, K. Thicknessdependent dielectric breakdown and nanopore creation on sub-10-nmthick SiN membranes in KCl aqueous solution. J. Appl. Phys. 2017, 121, 045301−045310. (82) Ying, C.; Zhang, Y.; Feng, Y.; Zhou, D.; Wang, D.; Xiang, Y.; Zhou, W.; Chen, Y.; Du, C.; Tian, J. 3D nanopore shape control by current-stimulus dielectric breakdown. Appl. Phys. Lett. 2016, 109, 063105−063109. (83) Zhang, Y.; Chen, Y.; Fu, Y.; Ying, C.; Feng, Y.; Huang, Q.; Wang, C.; Pei, D.-S.; Wang, D. Monitoring tetracycline through a solid-state nanopore sensor. Sci. Rep. 2016, 6, No. 27959. (84) Zhou, D.-M.; Deng, Y.-S.; Ying, C.-F.; Zhang, Y.-C.; Feng, Y.-X.; Huang, Q.-M.; Liang, L.-Y.; Wang, D.-Q. Rectifcation of Ion Current Determined by the Nanopore Geometry: Experiments and Modelling. Chin. Phys. Lett. 2016, 33, 108501−108504. (85) Deng, Y.; Huang, Q.; Zhao, Y.; Zhou, D.; Ying, C.; Wang, D. Precise fabrication of a 5 nm graphene nanopore with a helium ion microscope for biomolecule detection. Nanotechnology 2017, 28, 045302−045308. (86) Wanunu, M.; Meller, A. Chemically modified solid-state nanopores. Nano Lett. 2007, 7, 1580−1585. (87) Gyurcsányi, R. E. Chemically-modified nanopores for sensing. TrAC, Trends Anal. Chem. 2008, 27, 627−639. (88) Ali, M.; Schiedt, B.; Healy, K.; Neumann, R.; Ensinger, W. Modifying the surface charge of single track-etched conical nanopores in polyimide. Nanotechnology 2008, 19, No. 085713. (89) Kumar, S.; Tao, C.; Chien, M.; Hellner, B.; Balijepalli, A.; Robertson, J. W.; Li, Z.; Russo, J. J.; Reiner, J. E.; Kasianowicz, J. J.; Ju, J. PEG-labeled nucleotides and nanopore detection for single molecule DNA sequencing by synthesis. Sci. Rep. 2012, 2, No. 684. (90) Nasir, S.; Ali, M.; Ensinger, W. Thermally controlled permeation of ionic molecules through synthetic nanopores functionalized with 7135

DOI: 10.1021/acsomega.7b01245 ACS Omega 2017, 2, 7127−7135