Top Banner
Identifying single bases in a DNA oligomer with electron tunnelling Shuo Huang 1,2 , Jin He 1 , Shuai Chang 1,2 , Peiming Zhang 1 , Feng Liang 1 , Shengqin Li 1 , Michael Tuchband 1 , Alexander Fuhrmann 2 , Robert Ros 2 and Stuart Lindsay 1,2,3 * It has been proposed that single molecules of DNA could be sequenced by measuring the physical properties of the bases as they pass through a nanopore 1,2 . Theoretical calculations suggest that electron tunnelling can identify bases in single- stranded DNA without enzymatic processing 3–5 , and it was recently experimentally shown that tunnelling can sense indi- vidual nucleotides 6 and nucleosides 7 . Here, we report that tun- nelling electrodes functionalized with recognition reagents can identify a single base flanked by other bases in short DNA oligomers. The residence time of a single base in a recognition junction is on the order of a second, but pulling the DNA through the junction with a force of tens of piconewtons would yield reading speeds of tens of bases per second. Changes in the ion current through a nanopore can be used to identify translocating nucleotides. This opens the way to DNA sequencing if an exonuclease can pass each cleaved nucleotide into the pore sequentially 8 . Alternatively, it has been proposed that the high spatial resolution of electron tunnelling would allow direct reading of bases in an intact DNA polymer 3–5 . Recent progress in measuring electron tunnelling through nucleotides or nucleosides shows that they can be identified by means of character- istic current signals 6,7 . Recognition tunnelling 7,9 is an approach in which electrodes are functionalized with reagents that bind the target DNA bases. Contact via molecular adsorbates has been used to produce extraordinarily high spatial resolution in atomic force microscopy 10 and, as we show here, single bases can be resolved in a DNA polymer when read by means of a selective chemical contact. To extend recognition tunnelling to reads in buffered aqueous electrolyte, we synthesized the reagent 4-mercaptobenzamide (Fig. 1a and Methods), which presents two hydrogen-bond donor sites (on the nitrogen) and one hydrogen-bond acceptor site (the carbonyl). Likely binding modes to the four bases are shown in Fig. 2a (ref. 7). A gold (111) substrate and a partially insulated gold scanning tunnelling microscope (STM) probe were functiona- lized with this reagent (Methods; see also Supplementary Information) and characterized in an electron tunnelling junction formed in the STM (PicoSPM, Agilent). Figure 1a shows a d(CCACC) oligomer trapped in a tunnel gap through hydrogen bonding to one mercaptobenzamide molecule on the probe and another on the substrate. In reality, the oligomer is probably held by many contacts, but only those that complete a short tunnelling path (highlighted) will contribute significantly to the current. In our measurements, the probe is not deliberately scanned, but moves over the substrate as the microscope drifts. Alternatively, molecules may diffuse through the gap. Characteristic bursts of current are observed (an example is shown in Fig. 1b). As we show in the following, the low-frequency, large-amplitude pulses indicate a C, and the high-frequency, small-amplitude pulses an A. Figure 1c shows a sliding average of the spike amplitudes. Values below the red line identify an A base unambiguously. Figure 1d shows a sliding average over the pulse frequencies (as defined for each adjacent pair of spikes). The low-frequency CCACC a Probability Time (s) Time (s) Time (s) Time (s) e C A 0 0 0.5 1.5 2 1 0.2 0.4 0.6 0.8 1 · f Ò (Hz) d 0 0.5 1.5 2 1 0 100 200 300 400 ·iÒ (nA) c 0 0.5 1.5 2 1 0 0.02 0.04 0.06 Current (nA) b * 0.0 0.08 0.06 0.04 0.02 0.00 0.5 1.5 2 1.0 Figure 1 | Reading a single base within a heteropolymer. a, Benzamide groups on the probe and substrate bind bases in the polymer to give a signal dominated by the shortest tunnelling path (highlighted for the connection to the single A in d(CCACC)). b, Characteristic bursts of tunnelling noise with large, infrequent spikes signalling C and smaller, more frequent spikes signalling A. (Background tunnel current, 10 pA; bias, þ0.5 V.) The spike labelled with an asterisk (off-scale at 0.11 nA) is non- specific, and is rejected from the analysis. c,d, Rolling average of the spike height (c, 0.25 s window, 0.125 s steps) and spike frequency (d). C bases generate a negligible number of spikes below 0.015 nA (c, red line). e, Probability that the signal comes from an A (shown by the red line) or a C (blue line). For signal amplitudes .0.015 nA, probabilities are calculated as described in the Supplementary Information (values are not normalized to add to 1 in these regions). This burst of signal was chosen to show a clear example of transitions between A and C bases. Longer time traces (Supplementary Fig. S9a) are dominated by signals from A bases, which are preferentially trapped in the junction. 1 Biodesign Institute, Arizona State University, Tempe, Arizona 85287, USA, 2 Department of Physics, Arizona State University, Tempe, Arizona 85287, USA, 3 Department of Chemistry and Biochemistry, Arizona State University, Tempe, Arizona 85287, USA. *e-mail: [email protected] LETTERS PUBLISHED ONLINE: 14 NOVEMBER 2010 | DOI: 10.1038/NNANO.2010.213 NATURE NANOTECHNOLOGY | ADVANCE ONLINE PUBLICATION | www.nature.com/naturenanotechnology 1 © 2010 Macmillan Publishers Limited. All rights reserved.
6

Identifying single bases in a DNA oligomer with electron tunnelling

May 16, 2023

Download

Documents

Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: Identifying single bases in a DNA oligomer with electron tunnelling

Identifying single bases in a DNA oligomerwith electron tunnellingShuo Huang1,2, Jin He1, Shuai Chang1,2, Peiming Zhang1, Feng Liang1, Shengqin Li1, Michael Tuchband1,

Alexander Fuhrmann2, Robert Ros2 and Stuart Lindsay1,2,3*

It has been proposed that single molecules of DNA could besequenced by measuring the physical properties of the basesas they pass through a nanopore1,2. Theoretical calculationssuggest that electron tunnelling can identify bases in single-stranded DNA without enzymatic processing3–5, and it wasrecently experimentally shown that tunnelling can sense indi-vidual nucleotides6 and nucleosides7. Here, we report that tun-nelling electrodes functionalized with recognition reagents canidentify a single base flanked by other bases in short DNAoligomers. The residence time of a single base in a recognitionjunction is on the order of a second, but pulling the DNAthrough the junction with a force of tens of piconewtonswould yield reading speeds of tens of bases per second.

Changes in the ion current through a nanopore can be used toidentify translocating nucleotides. This opens the way to DNAsequencing if an exonuclease can pass each cleaved nucleotideinto the pore sequentially8. Alternatively, it has been proposedthat the high spatial resolution of electron tunnelling would allowdirect reading of bases in an intact DNA polymer3–5. Recentprogress in measuring electron tunnelling through nucleotides ornucleosides shows that they can be identified by means of character-istic current signals6,7. Recognition tunnelling7,9 is an approach inwhich electrodes are functionalized with reagents that bind thetarget DNA bases. Contact via molecular adsorbates has beenused to produce extraordinarily high spatial resolution in atomicforce microscopy10 and, as we show here, single bases can beresolved in a DNA polymer when read by means of a selectivechemical contact.

To extend recognition tunnelling to reads in buffered aqueouselectrolyte, we synthesized the reagent 4-mercaptobenzamide(Fig. 1a and Methods), which presents two hydrogen-bond donorsites (on the nitrogen) and one hydrogen-bond acceptor site (thecarbonyl). Likely binding modes to the four bases are shown inFig. 2a (ref. 7). A gold (111) substrate and a partially insulatedgold scanning tunnelling microscope (STM) probe were functiona-lized with this reagent (Methods; see also SupplementaryInformation) and characterized in an electron tunnelling junctionformed in the STM (PicoSPM, Agilent). Figure 1a shows ad(CCACC) oligomer trapped in a tunnel gap through hydrogenbonding to one mercaptobenzamide molecule on the probe andanother on the substrate. In reality, the oligomer is probably heldby many contacts, but only those that complete a short tunnellingpath (highlighted) will contribute significantly to the current. Inour measurements, the probe is not deliberately scanned, butmoves over the substrate as the microscope drifts. Alternatively,molecules may diffuse through the gap. Characteristic bursts ofcurrent are observed (an example is shown in Fig. 1b). As weshow in the following, the low-frequency, large-amplitude pulses

indicate a C, and the high-frequency, small-amplitude pulses anA. Figure 1c shows a sliding average of the spike amplitudes.Values below the red line identify an A base unambiguously.Figure 1d shows a sliding average over the pulse frequencies (asdefined for each adjacent pair of spikes). The low-frequency

CCACC

a

Prob

abili

tyTime (s)Time (s)

Time (s)Time (s)

eC

A

00 0.5 1.5 21

0.20.40.60.8

1

· fÒ (

Hz)

d

0 0.5 1.5 210

100200300400

·iÒ (n

A)

c

0 0.5 1.5 210

0.02

0.04

0.06C

urre

nt (n

A)b *

0.0

0.080.060.040.020.00

0.5 1.5 21.0

Figure 1 | Reading a single base within a heteropolymer. a, Benzamide

groups on the probe and substrate bind bases in the polymer to give a

signal dominated by the shortest tunnelling path (highlighted for the

connection to the single A in d(CCACC)). b, Characteristic bursts of

tunnelling noise with large, infrequent spikes signalling C and smaller, more

frequent spikes signalling A. (Background tunnel current, 10 pA; bias,

þ0.5 V.) The spike labelled with an asterisk (off-scale at 0.11 nA) is non-

specific, and is rejected from the analysis. c,d, Rolling average of the spike

height (c, 0.25 s window, 0.125 s steps) and spike frequency (d). C bases

generate a negligible number of spikes below 0.015 nA (c, red line).

e, Probability that the signal comes from an A (shown by the red line) or a

C (blue line). For signal amplitudes .0.015 nA, probabilities are calculated

as described in the Supplementary Information (values are not normalized

to add to 1 in these regions). This burst of signal was chosen to show

a clear example of transitions between A and C bases. Longer time traces

(Supplementary Fig. S9a) are dominated by signals from A bases,

which are preferentially trapped in the junction.

1Biodesign Institute, Arizona State University, Tempe, Arizona 85287, USA, 2Department of Physics, Arizona State University, Tempe, Arizona 85287, USA,3Department of Chemistry and Biochemistry, Arizona State University, Tempe, Arizona 85287, USA. *e-mail: [email protected]

LETTERSPUBLISHED ONLINE: 14 NOVEMBER 2010 | DOI: 10.1038/NNANO.2010.213

NATURE NANOTECHNOLOGY | ADVANCE ONLINE PUBLICATION | www.nature.com/naturenanotechnology 1

© 2010 Macmillan Publishers Limited. All rights reserved.

Page 2: Identifying single bases in a DNA oligomer with electron tunnelling

regions at each end enhance the confidence with which thoseregions can be assigned to a C base. The probability of an assign-ment to A (red line) or C (blue line) is shown in Fig. 1e.Calculation of these probabilities is based on our study of nucleo-tides, homopolymers and heteropolymers, as described sub-sequently. This example clearly shows that a single A base can beidentified with high confidence when flanked by C bases in anintact DNA molecule.

We first characterized the tunnel gap using doubly distilled waterand 1 mM phosphate buffer (PB, pH¼ 7.4). Small signals were

observed from buffer alone with bare electrodes, but they weremuch rarer when both electrodes were functionalized and thetunnel gap conductance set to 20 pS or less (Fig. 2b andSupplementary Information). The tunnel decay was much morerapid (decay constant, b¼ 14.2+3.2 nm21) with both electrodesfunctionalized than for unfunctionalized electrodes11 (b≈ 6.1+0.7 nm21; see also Supplementary Information), and we estimatethat the tunnel gap at i¼ 10 pA and V¼þ0.5 V is a little overthe length of two benzamide molecules (that is, a little greaterthan 2 nm).

k ton

toff

Burstduration, TB

Burst frequency, fB

Spikefrequency, fs

Total count rate, cps

I thresh = 1.5 σ

a

Cur

rent

(nA

)

fdGMP

Time (s)0.000 0.005 0.010 0.015 0.020 0.025 0.030

0.05

0.04

0.03

0.02

0.01

0.00

b

Cur

rent

(nA

)

Time (s)

Control

120 Hz

0.0000.00

0.01

0.02

0.005 0.010 0.015 0.020 0.025 0.030

Cur

rent

(nA

)

edmeCMP

0.000 0.005 0.010 0.015 0.020 0.025 0.030

0.05

0.04

0.03

0.02

0.01

0.00

Time (s)

25 cps1

0.8

0.6

0.4

0.2

00 0.02 0.04 0.06 0.08 0.1

i

Rela

tive

coun

ts

Current (nA)

ddCMP

Cur

rent

(nA

)0.000 0.005 0.010 0.015 0.020 0.025 0.030

0.05

0.04

0.03

0.02

0.01

0.00

0.05

0.04

0.03

0.02

0.01

0.00

Time (s)

h

Rela

tive

coun

ts

0.4 cps1

0.8

0.6

0.4

0.2

00 0.02 0.04 0.06 0.08 0.1

Current (nA)

cdAMP

0.000 0.005 0.010 0.015 0.020 0.025 0.030

Cur

rent

(nA

)

Time (s)

g

Rela

tive

coun

ts

15 cps1

0.8

0.6

0.4

0.2

00 0.02 0.04 0.06 0.08 0.1

Current (nA)

jRe

lativ

e co

unts

Current (nA)

6.7 cps1

0.8

0.6

0.4

0.2

00 0.02 0.04 0.06 0.08 0.1

ONH

H

S

ONH

H

S

O

O

HO

PO

−O

O−

NN

O

O

HTO

O

HO

PO

−O

O−

NN

N

N N

H

H

A

ONH

H

S

ONH

H

S

H

OO

OH

P

O

O−

O−

N

N

N

O

H H

C

O NH

H

S

ONH

H

S

OO

OH

P

O

O−

O−

N

NN

N

O

N

H

H

H

G

ONH

H

S

ONH

H

S

Figure 2 | Tunnelling signals from nucleotides trapped in a functionalized tunnel gap. a, Proposed hydrogen-bonding modes for all four bases. In practice,

water must play a role, because the observed difference between C and meC would not be accounted for by these structures alone. b, In phosphate buffered

saline, but in the absence of analyte, a 20 pS gap (i¼ 10 pA, V¼þ0.5 V) gave a signal free of features, except for some a.c. coupled line noise (indicated by

arrows). c–f, Characteristic current spikes produced when nucleotides dAMP, dCMP, dmCMP and dGMP were introduced (longer signal runs are given in the

Supplementary Information). dTMP produced no signals. g–j, Corresponding distribution of pulse heights. Red lines are fits to two Gaussian distributions in

the logarithm of current. k, Definition of the parameters used to characterize the tunnelling signals. Spikes are counted if they exceed a threshold equal to

1.5× the standard deviation of the noise on the local background. The signals occur in bursts (duration TB; frequency fB), each containing current spikes at a

frequency fS. The spikes stay high for a period ton and low for period toff. The total count rate (indicated in g–j) is the number of spikes in all bursts divided by

the measurement time.

LETTERS NATURE NANOTECHNOLOGY DOI: 10.1038/NNANO.2010.213

NATURE NANOTECHNOLOGY | ADVANCE ONLINE PUBLICATION | www.nature.com/naturenanotechnology2

© 2010 Macmillan Publishers Limited. All rights reserved.

Page 3: Identifying single bases in a DNA oligomer with electron tunnelling

Introducing DNA nucleotides (10 mM in PB) into the tunnel gapyielded characteristic noise spikes as shown in Fig. 2c–f. The signalcount rate (defined in Fig. 2k) varied considerably from 25 countsper second (cps; 5-methyl-deoxycytidine 5′-monophosphate,dmCMP) to less than 1 cps (deoxycytidine 5′-monophosphate,dCMP). No signals were recorded at all with thymidine 5′-mono-phosphate (dTMP), and the signal looked exactly like the control(Fig. 2b). STM images suggest that this nucleotide binds to thesurface (and presumably the probe) very strongly, blocking inter-actions in which a single molecule spans the junction.

The current occurs in bursts of spikes (longer signal runs aregiven in the Supplementary Information), and the distributions ofthe spike heights were fitted quite well with two Gaussians distri-butions of the logarithm of current7, as shown in Fig. 2g–j(fitting parameters are given in the Supplementary Information).These histograms were generated by counting only pulses thatexceeded 1.5× the standard deviation of the local noise background,that is, pulses typically above 6 pA (a full description of the analysisprocedure is given in ref. 7.

dCMP generates the highest signals and the lowest count rate,and deoxyadenosine 5′-monophophate (dAMP) and dmCMPproduce the smallest signals and the highest count rate (we havefound little difference between cytidine and 5-methylcytidine inorganic solvent7; see Supplementary Information). The three baseswith narrower pulse height distributions (dAMP, dmCMP anddGMP) often show bursts of ‘telegraph noise’ characteristic ofsources that fluctuate between two levels9 (particularly marked fordAMP). Such a two-level distribution is a strong indication thatthe tunnelling signals are generated by a single molecule trappedin the tunnel junction9. The characteristics of the tunnelling noisefrom the nucleotides are summarized in Table 1.

dAMP signals are well separated from dCMP signals, anddmCMP signals are well separated from dCMP signals in spikeamplitude and in the time distribution of their signals (Table 1;see also Supplementary Information). For this reason, we chose toinvestigate DNA oligomers composed of A, C and mC bases.

Figure 3a,c,e shows representative tunnelling noise traces ford(A)5, d(C)5 and d(mC)5 with the corresponding current peak dis-tributions shown in Fig. 3b,d,f. Comparing Fig. 3b (d(A)5) withFig. 2g (dAMP), Fig. 3d (d(C)5) with Fig. 2h (dCMP) and Fig. 3f(d(mC)5) with Fig. 2i (dmCMP) leads to the following surprisingconclusion: most of the polymer binding events in the tunnel junc-tion generate signals that resemble those generated by single nucleo-tides. That this should be so is not obvious. It requires (i) that singlebases are being read and (ii) that steric constraints resulting from thepolymer backbone do not prevent base-binding events from domi-nating the signals.

There are some (small) differences between nucleotide and oligo-mer signals. First, peak positions, widths and relative intensities arealtered somewhat (see Supplementary Information for details ofthe fits, and also the nucleotide distributions, which have beenreplotted on top of the homopolymer distributions as the blacklines on Fig. 3b,d,f ). Second, almost all of the signals generated by

nucleotides are less than 0.1 nA at 0.5 V bias (Table 1). In contrast,20% of the total signals generated by d(A)5 and d(mC)5 are largerthan 0.1 nA at this bias (Table 2). This is not obvious in Fig. 2,where distributions are plotted only up to 0.1 nA (the highcurrent regions are shown in the Supplementary Information).These high current (.0.1 nA) features in d(A)5 and d(C)5 are con-tinuously distributed, so they do not represent parallel reads of morethan one base at a time (where currents would be distributed in mul-tiples of the single molecule values12). Rather, they are new featuresassociated with the presence of the polymeric structure in the tunnelgap. Such a non-specific, large-amplitude spike is labelled by anasterisk in Fig. 1b.

Features at I . 0.1 nA appear much less frequently in oligomers ofmixed sequence, suggesting that they are associated with base stackingin the homopolymers. Figure 3h shows a current distribution ford(ACACA), in which 95% of events are below 0.1 nA. Figure 3jshows a current distribution for d(CmCCmCC), where 99% of eventsare below 0.1 nA. The solid red lines are the sums of the distributionsmeasured for the homopolymers corresponding to the constituentswith, scaling aside, only one fitting parameter. This parameter is theratio rfit of the A/C (rfit¼ 0.48) or mC/C (rfit¼ 0.66) contributions.These values differ from the known composition ratios (0.6 forACACA and 0.4 for CmCCmCC), but are surprising in that the spikerate for dCMP alone is very small, yet C appears to be quite well rep-resented in the mixed sequence oligomer data. This suggests that Cbases surrounded by A bases are read more frequently, possiblybecause the C-containing oligomer is better attached to the substratethan the isolated dCMP.

Most importantly, mixed oligomers generate signals that arelargely described as the sum of the individual base signals. (Someintermediate current reads, labelled ‘1’ in Fig. 3h,j, and a smallnumber of additional high current features, labelled ‘2’, show thatsequence context plays a small role.)

In these experiments, the probe drifts randomly over the samples,so the sequence is not ‘read’ deterministically. Nonetheless, we canreadily find traces in which the signals alternate between ‘A-like’ and‘C-like’ (Fig. 3g) and ‘mC-like’ and ‘C-like’ (Fig. 3i). The duration ofthese ‘bursts’ (Fig. 2k) of signals is long (0.14+0.02 s in ACACAand 0.15+0.02 s in CmCCmCC). Similar bursts are seen in the homo-polymers (Table 2) and the nucleotides (Table 1). This leads us to oursecond unexpected conclusion: the lifetime of the bound complex inthe tunnel gap is very long (fraction of a second) compared to eitherthe interval between noise spikes (ms) or the lifetime of the boundstate in solution (very short; see Supplementary Information).

Dynamic force spectroscopy was used as an independent test ofthe unexpectedly long lifetime of the benzamide–base–benzamidecomplex confined to a nanoscale gap. In these measurements(Fig. 4a), one of the recognition molecules was bound to an AFMprobe via a 34-nm-long polyethyleneglycol (PEG) linker (seeMethods), while the other formed a monolayer on an Au(111) sub-strate. dAMP was used as the target analyte to bridge the gap. In theabsence of dAMP, adhesion between the probe and substrate wasextremely small, presumably because the hydrogen-bonding sites

Table 1 | Nucleotide tunnelling noise characteristics.

Nucleotide dAMP dGMP dCMP dmCMP

Burst duration TB (s) 0.19+0.05* 0.13+0.02* 0.12+0.02* 0.06+0.01*Burst frequency fB (Hz) 732+82† 574+67† 306+23† 1,305+100†

Fraction of reads .0.1 nA 0.02 0.001 0.02 0.01ton (ms) 0.38+0.01* 0.48+0.02* 0.42+0.02* 0.31+0.09*toff (ms) 0.35+0.01* 0.56+0.04* 0.71+0.06* 0.41+0.11*ton/toff �1 0.9 0.6 0.8DG (kT units) 0 0.1 0.51 0.22

Parameters are defined in Fig. 2k. *Error in fit to exponential distribution. †Standard error.

NATURE NANOTECHNOLOGY DOI: 10.1038/NNANO.2010.213 LETTERS

NATURE NANOTECHNOLOGY | ADVANCE ONLINE PUBLICATION | www.nature.com/naturenanotechnology 3

© 2010 Macmillan Publishers Limited. All rights reserved.

Page 4: Identifying single bases in a DNA oligomer with electron tunnelling

on the benzamide recognition molecules were stably bound by watermolecules. Adhesion features were observed in the presence of asmall amount of dAMP, falling as the concentration of dAMPincreased (resulting in binding of both probe and substrate bydAMP; see Supplementary Information). Stretching of the PEGtether generated a characteristic signal that permitted multiplebinding events (Fig. 4b, top trace) to be separated from single mol-ecule events (Fig. 4b, bottom trace), so that only single-moleculebond-breaking events were analysed13. Single-molecule bond-break-ing forces as a function of pulling speed are summarized in Fig. 4c(solid lines are maximum likelihood fits to a heterogeneous bondmodel13,14), and the bond survival probability as a function ofbond-breaking force is shown in Fig. 4d. The solid lines are fits tothe same heterogeneous bond model14. They yield an off-rate atzero force, K off

0 ¼ 0.28 s21. Thus the intrinsic (zero-force) survivaltime of this complex is on the order of seconds, not milliseconds.The analysis also yields the distance to the transition state for dis-sociation, a¼ 0.78 nm (as well as its variance, s¼ 0.19 nm). Weconclude that each base resides in the tunnel junction for a signifi-cant fraction of a second, while generating tunnelling signals at kilo-hertz rates. The entire cluster of signals that occur in one burst(burst durations are listed in Tables 1 and 2) can be used to charac-terize a base.

Long-bound-state lifetimes accompanied by rapid fluctuations inelectronic signatures have been reported previously in STMimages15, and also in the effect of single-molecule reactions on trans-port in carbon nanotubes16. The origin of this noise is unclear, exceptthat it appears to be very sensitive to temperature, indicative of smallenergy barriers to the motion that causes the noise15. Following thework of Goldsmith and colleagues16, we analysed (seeSupplementary Information) the distribution of ‘on’ and ‘off’ times(Fig. 2k). In a limited range of times, determined by the amplifierresponse at one end and the servo response time at the other7, thesedistributions are exponential (as expected for a Poisson source); the1/e times (ton and toff ) are listed in Tables 1 and 2. They do notdiffer much, and calculating an energy difference DG between theon and off states from DG¼ kBT ln(toff/ton ) yields the values listedin the tables (in units of thermal energy, kBT, at 300 K). These valuesare all a fraction of kBT. The ‘switching’ cannot therefore representthermal activation over a significant barrier (the normal source oftwo-level noise). One possible explanation is Brownian motion in abound state sampled by an exponentially sensitive matrix element (seeSupplementary Information).

The on and off times are so broadly distributed that they are notvery useful for identifying base signals. However, the frequencywithin a burst ( fS, Tables 1 and 2) is a much more simple par-ameter. Supplementary Fig. S19a,b shows the current distributionsand frequency distributions for the three homopolymers, normal-ized so that the area under each curve is unity. The frequency distri-bution for d(mC)5 is bimodal, with many reads in the ‘C’ frequencyrange and a number at the very fast rate (�1,300 Hz) observed fordCMP alone (labelled f(mCMP) on the figure). This suggests thatthe binding modes of mC are altered significantly in a polymercontext (consistent with the larger shift of the polymer signal

a b

c d

e f

g h

i j

Cur

rent

(nA

)C

urre

nt (n

A)

Cur

rent

(nA

)C

urre

nt (n

A)

Cur

rent

(nA

)1.2

1

0.8

0.6

0.4

0.2

00 0.02 0.04 0.06 0.08 0.1

2

1.2

1

0.8

0.6

0.4

0.2

00 0.02 0.04 0.06 0.08 0.1

1

1.2

0.8

0.6

0.4

0.2

00 0.02 0.04 0.06 0.08 0.1

1

1

1.2

0.8

0.6

0.4

0.2

00 0.02 0.04 0.06 0.08 0.1

1

2

Current (nA)

0.00

0.02

0.04

0.06

0.0 0.5 1.0 2.52.01.5

d(A)5

0.00

0.02

0.04

0.06

0.0 0.5 1.0 2.52.01.5

d(C)5

0.00

0.02

0.04

0.06

0.0 0.5 1.0 2.52.01.5

d(ACACA)

Time (s)

Current (nA)Time (s)

Rela

tive

coun

tsRe

lativ

e co

unts

Rela

tive

coun

tsRe

lativ

e co

unts

Rela

tive

coun

ts

1

1.2

0.8

0.6

0.4

0.2

00 0.02 0.04 0.06 0.08 0.1

2

Current (nA)

0.00

0.02

0.04

0.06

0.0 0.5 1.0 2.52.01.5

d(5meC)5

Time (s)

Current (nA)Time (s)

Current (nA)Time (s)

0.00

0.02

0.04

0.06

0.0 0.5 1.0 2.52.01.5

d(CmCCmCC)

Figure 3 | Tunnelling signal distributions from oligomers resemble those of

the constituent nucleotides. a–f, Representative current traces from d(A)5

(a), d(C)5 (c) and d(mC)5 (e), with the corresponding distributions shown in

b, d and f, respectively. Red lines are fits with parameters similar to those

used for nucleotides (see Supplementary Information). The black lines are

the fits to the corresponding nucleotide distributions shown in Fig. 2g–i.

Label ‘2’ indicates some of the high current features seen in homopolymers

but not nucleotides (b,f). g–j, Current traces from mixed oligomers

d(ACACA) (g) and d(CmCCmCC) (i), with corresponding current

distributions (h,j). Red lines are scaled homopolymer fits, with the green-

dashed line showing the ‘C’ contribution, the orange-dashed line showing

the ‘A’ contribution and the purple-dashed line the mC contribution. The data

are well described by the homopolymer parameters, although some

intermediate signals (‘1’) and new high current features (2) show that the

sequence context affects the reads a little. The coloured bars on the current

traces mark bursts of A-like signals (orange), C-like signals (green) andmC-like signals (purple).

Table 2 | Oligomer tunnelling noise characteristics.

Oligomer d(A)5 d(C)5 d(mC)5

Burst duration TB (s) 0.14+0.02* 0.15+0.03* 0.41+0.03*Burst frequency fB (Hz) 738+100† 320+85† 662+116†

Fraction of reads .0.1 nA 0.20 0.0 0.23ton (ms) 0.33+0.01* 0.34+0.02* 0.26+0.01*toff (ms) 0.52+0.02* 0.42+0.01* 0.47+0.01*ton/toff 0.6 0.8 0.6DG (kT units) 0.51 0.22 0.51

Parameters are defined in Fig. 2k. *Error in fit to exponential distribution. †Standard error.

LETTERS NATURE NANOTECHNOLOGY DOI: 10.1038/NNANO.2010.213

NATURE NANOTECHNOLOGY | ADVANCE ONLINE PUBLICATION | www.nature.com/naturenanotechnology4

© 2010 Macmillan Publishers Limited. All rights reserved.

Page 5: Identifying single bases in a DNA oligomer with electron tunnelling

compared to the nucleotide signal, Fig. 3f), so we chose to analyseoligomers containing A and C, in particular the d(CCACC)sequence shown in Fig. 1a.

Given an average current in a burst, kil, and frequency k f l, thedistributions shown in Supplementary Fig. S20, IA,C(kil)(Supplementary Fig. S20a) and FA,C(k f l) (SupplementaryFig. S20b) determine independent probabilities that a base isan A or a C: Pi

A,C = IA,C kil( )

/ IA kil( )

+ IC kil( )[ ]

and P fA,C =

FA,C k f l( )

/ FA k f l( )

+ FC k f l( )[ ]

. The current distribution fromd(CCACC) (inset in Supplementary Fig. S9a) is almost completelydominated by A spikes (the component of the C distribution inthis fit is 7% or less). This is a surprising result, that more C basesin the sequence give a smaller number of C spikes. However, it is con-sistent with our hypothesis that the frequency of C reads is increasedwhen the base is flanked by A bases (c.f. the increase in C reads ind(ACACA) compared to the dCMP versus dAMP count rate).

Armed with our analysis of the burst signals, we can now makequantitative assignments of mixed signals (this was done ‘by eye’ inFig. 3g,i). d(C)5 produces no signals below 0.015 nA, so bursts ofcurrent below this level (but above the noise) can be unambiguously

assigned to A. For larger-amplitude signals we use both the fre-quency and amplitude data as described in the SupplementaryInformation. The result is the pair of curves shown in Fig. 1e.

Using this approach to sequence DNA requires several furtherdevelopments. First, the polymer must be pulled through a tunneljunction at a controlled speed, particularly if homopolymer runsare to be read. Because DNA passes through unfunctionalized nano-pores too rapidly to be read2, the long residence time of bases in afunctionalized tunnel junction is an asset. At present, movementfrom one site to another is driven by uncontrolled mechanicaldrift, which generates unknown forces on the reading complex.Our force spectroscopy data can be used to give a crude estimateof the ‘pulling’ force that would be needed to achieve a given readrate (assuming the measured off rate for dAMP to be representativefor all bases). The Bell equation gives the off rate at a force Fas Koff = K0

off exp(Fa/kBT) so, with Koff0 ¼ 0.28 s21 and a¼

0.78 nm, 19 pN would result in the passage of 10 bases persecond. A rate of 10 bases per second gives about 30 data spikes(on average) for a ‘C’ read, enough to generate an assignmentwith a reasonable level of confidence. A force of 19 pN can be

20 nm50

pN

Forc

e

z-extension

Fmin

a

b

121086420

100806040200

Freq

uenc

y

Force (pN)

200 nm s−1

[19]

100806040200

30

20

10

0

Freq

uenc

y

Force (pN)

500 nm s−1

[62]

Freq

uenc

y

6050403020100

100806040200Force (pN)

2,000 nm s−1

[114]

Freq

uenc

y

80

60

40

20

0100806040200

Force (pN)

5,000 nm s−1

[184]

c

101

102

103

104

100806040200

200 nm s−1

500 nm s−1

2,000 nm s−1

5,000 nm s−1

Force (pN)

−v ln

(nv(f)

) (nm

s−1

)

d

i

ii

O

O

HO

P

O

O–

O–

N

N

N

N N

H

H

A

ON

H

H

S

ON

H

H

S

H

Figure 4 | The lifetime of the reading complex is on the order of a second at zero force. a, AFM gap functionalization, where the blue line represents a

34-nm PEG linker. b, Representative force curves: pulling on more than one molecule at a time (top trace; the force baseline is not restored after each break,

and the z-extension, corrected for tip displacement, is .34 nm) and a single molecule curve (bottom trace) of the type accepted by the software. The force

returns to the baseline after the bond breaks and the corrected extension is �34 nm. c, Histograms of bond breaking forces at the pulling speeds marked.

The solid lines are maximum likelihood fits to the heterogeneous bond model. Numbers in square brackets are sample size. d, Bond survival probability

plotted versus bond-breaking force for the four pulling speeds, fitted by the same heterogeneous bond model parameters (solid lines). These fits yield a

zero-force off rate of 0.28 s21, implying that the assembly lives for times on the order of seconds in a nanogap, much longer than the lifetime in solution.

For details see ref. 13.

NATURE NANOTECHNOLOGY DOI: 10.1038/NNANO.2010.213 LETTERS

NATURE NANOTECHNOLOGY | ADVANCE ONLINE PUBLICATION | www.nature.com/naturenanotechnology 5

© 2010 Macmillan Publishers Limited. All rights reserved.

Page 6: Identifying single bases in a DNA oligomer with electron tunnelling

generated by a bias of just 80 mV across a nanopore17, so read ratesof 10 bases per second per tunnel junction seem feasible.

The second requirement for a practical sequencing system is abetter recognition chemistry in which there is a much larger separ-ation of the current distributions from all five bases. New com-pounds are presently under study in our laboratory.

MethodsNucleoside 5′-monophosphates (Sigma-Aldrich) were used as supplied. DNAoligomers were synthesized and characterized by Integrated DNA Technologies andused without further purification. Synthesis and characterization of other materialsare described in the Supplementary Information. Gold probes were etched asdescribed previously7 and coated with high-density polyethylene18 to leave a fractionof a micrometre of exposed gold (optical and transmission electron microscopy(TEM) characterization is described in the Supplementary Information). Theseprobes gave no measureable d.c. leakage, which is important, as this can be a sourceof distortion of the tunnelling signal7. Capacitative coupling of 120 Hz switchingsignals was a problem (Fig. 2b) that was minimized by careful control of the coatingprofile. The gaps were characterized by recording current decay curves as a functionof distance, starting at 20 pS, a distance that gave no signals in buffer alone withfunctionalized electrodes. Current signals were recorded using an Agilent PicoSPMtogether with a digital oscilloscope controlled by a custom Labview program. Theservo response time was set to �30 ms, as described previously7. This places anupper limit on undistorted measurements of pulse widths of a few milliseconds.Analysis of current distributions was automated using software described elsewhere7.Force spectroscopy was carried out with an MFP3D AFM (Asylum Research).Heterobifunctional PEG linkers (MAL-SVA 3400, Lysan Bio) with an extendedlength of 34 nm were attached at one end to a silicon nitride AFM probe (VeecoMSNL; spring constant, 0.02 N m21) and mercaptobenzamide molecules attachedto the remaining maleimide as described elsewhere19. Force curves were taken in1 mM PB buffer with an initial 10 mM concentration of dAMP in the gap adjustedby rinsing (see Supplementary Information). Force curves were analysed usingcustom software20.

Received 28 June 2010; accepted 4 October 2010;published online 14 November 2010

References1. Zwolak, M. & Di Ventra, M. Physical approaches to DNA sequencing and

detection. Rev. Mod. Phys. 80, 141–165 (2008).2. Branton, D. et al. Nanopore sequencing. Nature Biotechnol. 26,

1146–1153 (2008).3. Lagerqvist, J., Zwolak, M. & Di Ventra, M. Influence of the environment and

probes on rapid DNA sequencing via transverse electronic transport. Biophys. J.93, 2384–2390 (2007).

4. Zwolak, M. & Di Ventra, M. Electronic signature of DNA nucleotides viatransverse transport. Nano Lett. 5, 421–424 (2005).

5. Krstic, P. S., Wells, J. C., Fuentes-Cabrera, M., Xu, D. & Lee, J. W. Towardelectronic conductance characterization of DNA nucleotide bases. Solid StatePhenom. 121–123, 1387–1390 (2007).

6. Tsutsui, M., Taniguchi, M., Yokota, K. & Kawai, T. Identification of singlenucleotide via tunnelling current. Nature Nanotech. 5, 286–290 (2010).

7. Chang, S. et al. Electronic signature of all four DNA nucleosides in a tunnelinggap. Nano Lett. 10, 1070–1075 (2010).

8. Clarke, J. et al. Continuous base identification for single-molecule nanoporeDNA sequencing. Nature Nanotech. 4, 265–270 (2009).

9. Lindsay, S. et al. Recognition tunneling. Nanotechnology 21, 262001 (2010).10. Gross, L., Mohn, F., Moll, N., Liljeroth, P. & Meyer, G. The chemical structure of

a molecule resolved by atomic force microscopy. Science 325, 1110–1114 (2009).11. Vaught, A., Jing, T. W. & Lindsay, S. M. Non-exponential tunneling in water

near an electrode. Chem. Phys. Lett. 236, 306–310 (1995).12. Lindsay, S. M. & Ratner, M. A. Molecular transport junctions: clearing mists.

Adv. Mater. 19, 23–31 (2007).13. Fuhrmann, A., Anselmetti, D., Ros, R., Getfert, S. & Reimann, P. Refined

procedure of evaluating experimantal single-molecule force spectroscopy data.Phys. Rev. E 77, 031912 (2008).

14. Getfert, S. & Reimann, P. Optimal evaluation of single-molecule forcespectroscopy experiments. Phys. Rev. E 76, 052901 (2007).

15. Ramachandran, G. K. et al. A bond-fluctuation mechanism for stochasticswitching in wired molecules. Science 300, 1413–1415 (2003).

16. Goldsmith, B. R., Coroneus, J. G., Kane, A. A., Weiss, G. A. & Collins, P. G.Monitoring single-molecule reactivity on a carbon nanotube. Nano Lett. 8,189–194 (2008).

17. Keyser, U. F. et al. Direct force measurements on DNA in a solid-state nanopore.Nature Phys. 2, 473–477 (2006).

18. Visoly-Fisher, I. et al. Conductance of a biomolecular wire. Proc. Natl Acad. Sci.USA 103, 8686–8690 (2006).

19. Ashcroft, B. et al. An AFM/rotaxane molecular reading head for sequence-dependent DNA structure. Small 4, 1468–1475 (2008).

20. Fuhrmann, A. Force Spectroscopy from Single Molecules to Whole Cells: RefinedProcedures of Data Analysis. PhD thesis, Arizona State Univ. (2010).

AcknowledgementsThe authors acknowledge useful discussions with O. Sankey, P. Krstic and B. Gyarfus.P. Collins made helpful comments on an earlier version of this manuscript. H. Liucomposed the graphic for Fig. 1a. This work was supported by a grant from the SequencingTechnology Program of the National Human Genome Research Institute (HG004378). R.R.and A.F. were supported by a grant from the National Cancer Institute (U54CA143682).

Author contributionsS.H., S.C. and J.H. carried out tunnelling measurements and characterized the samples.P.Z., F.L. and Sq. L. designed, synthesized and characterized reagents. M.T. preparedtunnelling probes. A.F. and R.R. carried out force spectroscopy. S.L. designed experiments,analysed data and wrote the paper.

Additional informationThe authors declare competing financial interests: details accompany the full-text HTMLversion of the paper at www.nature.com/naturenanotechnology. Supplementaryinformation accompanies this paper at www.nature.com/naturenanotechnology.Reprints and permission information is available online at http://npg.nature.com/reprintsandpermissions/. Correspondence and requests for materials should beaddressed to S.L.

LETTERS NATURE NANOTECHNOLOGY DOI: 10.1038/NNANO.2010.213

NATURE NANOTECHNOLOGY | ADVANCE ONLINE PUBLICATION | www.nature.com/naturenanotechnology6

© 2010 Macmillan Publishers Limited. All rights reserved.