10 1002@smtd 201900595luchian

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Full Paper

www.small-methods.com

Nanopore-Based Protein Sequencing Using Biopores:


Current Achievements and Open Challenges
Alina Asandei, Giovanni Di Muccio, Irina Schiopu, Loredana Mereuta, Isabela S. Dragomir,
Mauro Chinappi,* and Tudor Luchian*

degradation.[7] Known limitations affecting


The synergy of life sciences discoveries, biomolecular and protein engineering these methods, are: i) mass spectrometry
advances, and groundbreaking nanofabrication technologies, has intro- often fails to reveal the complete sequence
duced over the past years the wide use of the nanopore-based investigations information of a protein;[8] ii) Edman deg-
of matter at the molecular level. This review focuses on the fundamental radation is optimally applied to sequences
of 30–50 residues long without modified
principles of α-hemolysin (α-HL) protein-based nanopores, as sensitive
or buried N-terminal amino acid;[9] iii) in
investigative tools that combine single-molecule detection with the ability to certain instances, accurate sequencing
simultaneously manipulate single molecules, in otherwise complex samples. becomes challenging with mass spec-
Herein, there are presented some of the efforts directed to control the capture trometry, since some amino acids have
dynamics and translocation speed of tailored polypeptides through the α-HL the similar mass and charge (e.g., leucine
and isoleucine);[10] iv) the fractionation
nanopore, by harnessing the electro-osmotic flow and nanopore-tweezing
step (i.e., breaking apart a longer sequence
influence on individual molecules, which are engineered to resemble mac- into many smaller peptides), preceding
rodipoles. The reported applications of this approach suggest a solution to the analysis of the mass-to-charge ratio of
enhance the temporal resolution of nanopore detection, prove the capability of each new peptide with the mass spectro­
the system in distinguishing between groups of distinct amino acids from the meter, is slow and tedious.[8,11,12]
studied poly­peptides, and propose a strategy to translate such single-molecule Built on the success of nucleic acids
detection and characterization,[13–26]
sensors in devices suitable for polypeptide sequencing at unimolecular level.
nanopores became a viable alterna-
tive for single-molecule-based, primary
sequence recognition, and conformational
1. Introduction analysis on peptides and proteins, thus precluding the need
for proteolysis or labeling.[27–51] Through an enhanced analysis
Primary structure identification of peptides and proteins is cen- throughput, the possibility of primary structure identification
tral for the advancement of proteomics, as it determines how in peptides and proteins with nanopores, opens new vistas in
these biopolymers fold and perform their biological function, deciphering the proteome, since in vitro peptide or protein
and previous work established that even subtle changes in a pro- amplification assays lack. To this end, the underlying para-
tein’s primary sequence can cause debilitating pathologies.[1–4] digm is that the polypeptide-mediated ionic current fluctua-
Traditional methods for proteome analysis and protein tions through a voltage-biased nanopore are correlated with the
sequencing include mass spectrometry[5,6] and Edman distinct physicochemical properties of individual amino acids
(Figure 1).[26,52–55]
Dr. A. Asandei, Dr. I. Schiopu, I. S. Dragomir
Within simplifying assumptions,[29] the analyte-induced
Interdisciplinary Research Department blockade amplitude on a single nanopore modeled as an elec-
Alexandru I. Cuza University trically neutral, uniform cylinder, can be correlated with the
700506 Iasi, Romania analyte volume (Figure 1I,d and the corresponding text inset),
G. Di Muccio, Prof. M. Chinappi σ buffer ∆Vδ
Department of Industrial Engineering ∆I block = I blocked − I open ≈ (σbuffer is the conductivity of
l p2
University of Rome Tor Vergata
Via del Politecnico 1, 00133 Rome, Italy the buffer, lp is the nanopore’s length, and δ quantifies the
E-mail: mauro.chinappi@uniroma2.it volume of the analyte). Note that for other shapes other than
Dr. L. Mereuta, Prof. T. Luchian cylinders, the current blockade varies with the molecule posi-
Department of Physics tion inside the nanopore.[56] The formula above embodies
Alexandru I. Cuza University also simplifications of the formalism described originally by
700506 Iasi, Romania
E-mail: luchian@uaic.ro
DeBlois and Bean[57] regarding the relative geometry of the ana-
lyte and the nanopore, impact of the extension of the electric
The ORCID identification number(s) for the author(s) of this article
can be found under https://doi.org/10.1002/smtd.201900595. field lines beyond the nanopore inner region, or distortion of
the electric field lines inside the nanopore created by the pres-
DOI: 10.1002/smtd.201900595 ence of an analyte. Such an approach has been widely used for

Small Methods 2020, 1900595 1900595 (1 of 13) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.small-methods.com

Figure 1. Nanopore-based characterization of individual biopolymer primary structures. I,a) In the absence of nanopore–analyte interactions, the
ionic current through a single protein nanopore present in a reconstituted lipid membrane, and held at a constant transmembrane potential (ΔV), is
relatively constant I,b). Electrophoretic capture of a single biopolymer at the mouth of the nanopore, described by the association rate rateon, followed
by biopolymer trapping inside the nanopore I,c), are seen as reversible changes of the ionic current through the nanopore between the open state
(Iopen—free nanopore) and blocked state (Iblocked—nanopore transiently occupied by the biopolymer) I,d). In the simplest embodiment (I,d), the three
main descriptors used to characterize such analyte–nanopore interactions, are: the time intervals between consecutive blockade events (τon), blockade
duration (τoff ), and blockade amplitude (ΔIblock = Iblocked – Iopen) (see also text). The primary structure of nucleic acids and proteins II,a) may be obtained
from the degree to which the bases or amino acid residues reduce the ionic current flowing through the nanopore II,b).

quantitative and qualitative estimations of individual analytes interactions, and the capture rate is treated within the classical
volume,[41,58–61] and it was successfully extended to more com- Eyring’s transition state theory.[43]
plicated scenarios, in which the electric potential profile along It may be argued that for analytical purposes, τon time inter-
the nanopore axis assumed a piecewise linear dependence, and vals (Figure 1I,d) could be measured between the beginning
the relative amplitude of blockade current events ΔIblock repre- of successive blockade events. Within the data analysis theory
sented not only the volume excluded by the analyte, but mobile applied to Markov models as in Figure 1, open time intervals
cation binding to it as well.[21,62] can be analyzed in terms of exponentially distributed values,
From an analytical standpoint and when evaluated within 1
leading to τon’s arithmetic average τ on = , only if the
the frame of a diffusion-controlled process, the analyte cap- rate on
ture rate (rateon) by the nanopore is proportional to ΔV (see measurement of individual τon’s starts from the moment when
the text inset in Figure 1),[60,63] whereas the mean transit time the nanopore enters the “open” state and it lasts until the first
of the analyte across the nanopore (τtransit), as derived from 1D “open” → “blocked” transition occurs.[64] We therefore adopted
drift-diffusion models based on individual time translocating as a rule that quantification of τon time intervals should start
values (τoff),[29,59] is inversely proportional to ΔV (see the text immediately when a previously bound analyte to the nanopore
inset in Figure 1). In these expressions, the parameters have dissociates from it (Figure 1I,d), as done previously in our labo-
their usual thermodynamic and geometric meanings, namely: ratory[53] and others.[16,26,31,43]
kB—the Boltzmann’s constant, Tm—absolute temperature Despite successful implementation of this technique to stud-
of the buffer, Dbuffer and Dnanopore—analyte’s diffusion coef- ying structural features on polypeptides[39,42,47,65–71] and recent
ficient in buffer and inside of the nanopore, respectively, z efforts directed to polypeptide sequencing,[46] it still remains
and e−—effective valence of the analyte and electronic charge, an open challenge to accurately deconvolve the ion current
rp—nanopore’s radius on the cis side. For analytes comparable fluctuations to unequivocally deduce the polypeptide’s primary
in dimension to the nanopore’s inner volume, the electro-diffu- sequence. In a reserved optimistic tone on this approach,
sion-driven analyte to the nanopore mouth does not necessarily Kennedy et al.[27] have concluded that despite the use of a sub-
ensures capture and translocation. In such cases, the analyte nanometer pore which allows the detection of post-translational
requires climbing across the free energy barrier that involves modification on a single amino acid residue, the method still
the entropy penalties for confining the analyte inside the nano- lacks the sensitivity needed to discriminate between all of the
pore and enthalpic contributions from the analyte–nanopore amino acids.

Small Methods 2020, 1900595 1900595 (2 of 13) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.small-methods.com

Major hurdles that limit the accurate polypeptide sequencing agent.[27,29,90] On the other hand, biological nanopores with
at single amino acid resolution, include: i) the dimension of known crystal structure are amenable to atomic-level engi-
the sensitivity region inside the nanopore, usually the most neering and functionalization, as their surface can be finely
constricted domain of the inner volume, does not scale with tuned through site mutation on individual single amino acid
the size of individual amino acids, implying that current fluc- groups.[85,91–94]
tuations schematically shown in Figure 1II,b, are not correlated From an historical perspective and within the framework
with the passage of individual residues;[53] ii) with the excep- described above, the homo-heptameric alpha-hemolysin (α-HL)
tion of simplified, engineered peptide constructs,[72,73] the het- protein nanopore proved successful for structural identifica-
erogenous charge distribution along the polypeptide’s sequence tions in polypeptides, due to intrinsic benefits, such as known
hinders the controlled, electrophoretic-driven translocation of crystal structure at atomic resolution,[95] structural stability, and
the polypeptide through nanopore. To add to this complexity, relative absence of gating transitions accompanying the ion
peptide-nanopore nonspecific interactions, as well as and the transport through it.[76,96] In early studies, efforts were devoted
peptide length, are relevant factors during such peptide recogni- to capturing small peptides and then correlating some of their
tion attempts;[74,75] iii) the high velocity of the polymer’s passage features (e.g., conformation, length) with the transient block-
across the nanopore poses serious challenges to unraveling the ades on the ion current through the nanopore and dwell time
identity of the moving monomers. In relation to this, two chal- statistics[43,65,97] α-HL was also shown to be a powerful tool for
lenges are critical to the subsequent statistical analysis of the analyzing the unfolding of entire proteins[98–100] and recently, it
blockade events. First, due the extremely short dwell residence was proposed to hold potential for the detection of post-trans-
of small analytes inside the nanopore (e.g., a freely diffusing lational modifications[28,101] and the analysis of ion–peptide
molecule as small as sucrose with the diffusion coefficient interactions.[102]
D = 0.5 × 10−9 m2 s−1, situated in the middle of the α-HL, Here, we will summarize some of the theoretical insights
spends ≈6.3 ns before escaping the nanopore) and depending and experimental results obtained in our laboratories, for the
upon the applied voltage and salt concentration, the number goal of unveiling new perspectives in the realm of polypeptide
of charge carriers from the electrolyte solution flowing across identification and sequence recognition.
the transiently-blocked nanopore is very small (≈hundreds).
This entails a low signal-to-noise ratio blockade signal in the
low-frequency regime, thus degrading the signal before sup- 2. Reading the Primary Structure on Macrodipole-
plementary amplifier electronics and circuitry parasitic noises
Like Polypeptides
add up.[25,76–79] Second, the extremely short durations of such
events hinder their accurate resolving upon convolution with To achieve amino acid discrimination on polypeptides from
the microsecond time-resolution of most commonly available current blockade events through the α-HL with the approach
patch-clamp amplifiers.[77] In a recent work, authors provide a outlined above, we sought to: i) control the dynamics of poly-
comprehensive account on noise-and bandwidth-related chal- peptides through the α-HL, and ii) unravel and analyze ionic
lenges during recordings on ion channels and nanopores.[80] current blockades stemming from the polypeptide passage
To alleviate these shortcomings, in an alternative approach across the nanopore’s most sensitive, constriction region.
authors described an analytical tool to characterize such short As described, a critical challenge plaguing the nanopore-
analyte–nanopore interactions. Briefly, they modeled the tran- based analysis to distinguish between and possibly identify
sitions of the nanopore-mediated current to individual states, specific groups of amino acids, is the relatively short residence
as caused by reversible analyte–nanopore interactions, with an time of analytes inside the nanopore.[25,53] The existing litera-
equivalent electrical circuit. Subsequently, such a model allowed ture presents methods to alter the mean analyte residence time
to quantify the system response, and eventually enabled the in the nanopore, such as: changing the temperature,[51,103,104] by
estimation and characterization of short-lived blockade states as increasing analyte–nanopore wall interactions,[105] or modifying
induced by polyethylene glycols with as few as 8 mono­mers, the geometry of the nanopore,[60] by modification of the trans-
approximately the size of sucrose.[81] locating molecule,[106,107] by altering viscosity and electrolyte
In terms of structural details, there are three major classes concentration of the buffer,[21,99,108,109] rapid switching or modu-
of nanopores useful for single-molecule detection and recogni- lating the transmembrane potential,[110] controlling the balance
tion: i) biological (usually protein-based) nanopores; ii) solid- between the electrostatic and electro-osmotic forces,[59,69,111]
state nanopores, and iii) hybrids of biological and solid-state using Li+ as counterions in the electrolyte solution,[112] coating
nanopores. They each display complementary benefits and nanopores with a fluid lipid bilayer where specific mobile
common drawbacks for the task outlined above.[26,53,77,82–88] ligands able to bind analytes are embedded,[58]or using optical
Notably, the recent description of synthetic membrane-span- tweezers.[113]
ning nanopores constructed with DNA nanotechnology, holds A straightforward approach to control an electrically charged
the game-changing potential in many biotechnology and bio- biopolymer trafficking across a nanopore, is to alter the applied
medicine applications.[89] voltage (ΔV). A reduced voltage across the nanopore reflects
As a brief comparison on their particular benefits, solid-state as a correspondingly diminished electric force acting upon
nanopores, although nontransferrable to lipid-based mem- a charged molecule passing through the nanopore, entailing
branes, are more stable than biological ones in extreme buffer increased dwell times of the analyte inside the nanopore (τtransit;
conditions (e.g., urea, sodium dodecyl sulfate (SDS)) and can see also Figure 1, text in the inset).[114] From the same physical
be used in sensing protocols that require strong denaturating principles however (Figure 1, text in the inset), a reduced ΔV

Small Methods 2020, 1900595 1900595 (3 of 13) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.small-methods.com

Figure 2. Schematics of the nanopore-tweezer paradigm. a1) A trans-positive voltage applied across a membrane containing a single nanopore gener-
ates an inhomogeneous electric field E (gray dashed lines) oriented toward the negative electrode. The electrical field is larger inside the nanopore and
goes to zero, far from the nanopore. A dipolar-like analyte is attracted by the maximum of the electrical field (dielectrophoresis). a2) Sketch of the free-
energy profiles Gtot as a function of the position of the dipolar analyte center. To enter the nanopore, the analyte has to overcome a capture free-energy
barrier (ΔGcp). When the analyte enters the nanopore, it gets trapped in a metastable state where the forces acting on it equilibrate. The escape from this
metastable state is characterized by an escape barrier ΔGex. a3) As the voltage increases, the capture barrier ΔGcp decreases (i.e., capture is enhanced),
while the escape barrier ΔGex increases, leading to longer dwell times of the analyte inside the nanopore. b–e) Different stages of a macrodipole-like
peptide capture and trapping inside a α-HL by dielectrophoresis. A macrodipole-like, trans-added peptide gets oriented with the positive moiety toward
the β-barrel entry of the nanopore, gets imported into the nanopore mainly under the influence of the electric force F+ b), moves across the nanopore,
and eventually protrudes to the cis side of the membrane c–e). During the course of motion, the negative moiety of the peptide will enter the β-barrel
domain of the nanopore, where it experiences an electric force F− oriented toward the trans side of the membrane, opposite to F+. At some moment
during this transit, the overall number of positive and negative residues from the peptide which experience the electric field inside the nanopore, will
become such that |F+| approximately balance |F−|. This determines an approximately net zero electric force acting on the peptide, which will assume a
metastable state inside the nanopore c). Small thermal movements of the peptide toward the cis d) or trans side e) of the membrane determine a net
electrical force which drives the peptide back to the metastable state, before it eventually escapes the nanopore to either side. A theoretical descrip-
tion of the nanopore tweezer (dielectrophoretic trapping) mechanism in term of free energy profiles can be found in previously described work.[118,119]
Panels (a1)–(a3) and (b)–(e) are adapted from[118] and,[119] respectively.

decreases the capture rate of the analyte inside the nanopore entrance (Figure 2b). On the other hand, the residence time
(rateon). More specifically, for charged biopolymers (e.g., DNA), of the polypeptide captured transiently inside the nanopore is
the dependence of the capture rate and residence time with ΔV enhanced with the increase in ΔV, due to an electrostatic tug-
is nonlinear. Experimental data[115] and a theoretical descrip- of-war between the charges on opposite sides of the polypeptide
tion[116] were described earlier. To a certain extent, the reduc- and the ΔV (Figure 2c–e). The strategy described above to con-
tion of the analyte capture rate with decreasing ΔV’s can be trol and slow-down the polypeptide passage across the nano-
mitigated by the use of salt gradients between the cis and trans pore, constitutes a productive solution to tackling the challenge
sides of the system.[42,117] From our experience though, steep of correlating ionic current fluctuations through the nanopore
salt gradients maintained across a lipid membrane containing a with the peptide’s primary structure, within a reasonable band-
single α-HL, have a destabilizing effect on the system. width of the amplifier (≈100 kHz). From a geometrical rea-
As an alternative solution to this challenge, dubbed “the soning, a metastable, captured polypeptide (Figure 2a,c) spans
nanopore-tweezer approach,” we used model polypeptides the entire nanopore and it gets positioned nearly symmetri-
consisting of 12 asparagines, whose N- and C-termini were cally relative to the nanopore’s constriction domain.[95] As the
engineered to contain 12 patches of glutamates and arginines α-HL’s vestibule and β-barrel excluded volumes remain largely
(Figure 2). We demonstrated experimentally that for such constant during small position displacements of an entrapped
macro­dipole-like peptides, an increase in ΔV leads simultane- polypeptide, they contribute to a constant residual current
ously to an increase in the capture rate by the nanopore and across the nanopore, Thus, the amino acid residues occupying
of the residence time of the polypeptides inside the nanopore, the geometrically-sensitive constriction region of the nano-
regardless of the applied potential polarity.[118,119] pore are key in affecting the ensuing ionic current blockades.
From a physical perspective,[115] it is clear that increased In other words, the random displacements of the polypeptide
polypeptides capture rates by the nanopore at increasing ΔVs, inside the nanopore and across the constriction region gen-
reflected the augmented electric interactions between the erate current fluctuations sensitive to the identity of the central
polypeptide’s end and the potential drop near the nanopore amino acid residues from the polypeptide (vide infra).

Small Methods 2020, 1900595 1900595 (4 of 13) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.small-methods.com

Figure 3. Representative traces showing the reversible ion current alterations induced by macrodipole-like peptides through an open α-HL nanopore.
Distinct peptide constructs, termed (Pe1 Ac-(R)12 – (A)6 – (E)12 – NH2, panel a); Pe3 Ac-(R)12 – (W)6 – (E)12 – NH2, panel b); or Pe5 Ac-(R)12 – (A)3 –
(W)3 – (E)12 – NH2, panel c), were added to the trans side of the membrane (20 × 10−6 m), in buffer containing 2 m KCl, 10 × 10−3 m HEPES, pH = 7. An
increase in the applied potential from ΔV = +40 mV (panels (a)–(c)) to + 80 mV (panels (d)–(f)), leads to the peptide capture rate enhancement (the
average of time intervals separating successive blockade events τon decreases), while the average duration of the blockade events (τoff ) increases. The
expanded views shown in panels (d)–(f) illustrate ionic current fluctuations through the α-HL, elicited by a metastable peptide inside the nanopore.
Panels (g)–(i) display the voltage-dependent, average values of the interevents time intervals (τon; “free α-HL”) and the blockade-events durations (τoff;
“blocked α-HL”) for the peptide tested, from measurements on single molecule ion-current blockades. These graphs can be qualitatively compared
with the free-energy barriers reported in Figures 2 and a3. Indeed, the larger the capture (or escape) time, the larger the free-energy barrier associated
to the capture (or escape) process. Reprinted with permission.[72] Copyright 2017 American Chemical Society.

As a first application of the nanopore-tweezer approach, we studied polypeptide spans 0.38 nm,[27] we posit that an amino
demonstrated its potential for single-molecule interrogation of acid residue lodged precisely in the constriction as well as the
primary structure on model polypeptides. To this end, we engi- neighboring ones just outside it, govern the magnitude of ionic
neered polypeptides whose residues under study presented a current fluctuations through the nanopore, while fully occupied
marked difference in their physical size, namely six A (alanine), by the gliding polypeptide (Figure 5I,b,c). This is consistent
six W (tryptophan), or a combination of three A and three W with previous MspA experiments, demonstrating that current
residues, flanked at the N- and C-termini by oppositely charged fluctuations across the nanopore were critically affected by
segments at neutral pH, each containing 12 R (arginine) and ≈three to four nucleotides moving in and out of the constriction
12 E (glutamate) residues (Figure 3).[72] region, of approximately similar length.[122]
In a first series of experiments, we demonstrated that the Within this model, we proposed that data displayed in the
transmembrane potential augments such dipolar-like polypep- expanded traces in Figure 3d–f stem from the successive traf-
tides capture rate by the α-HL, and simultaneously increases ficking across the α-HL’s constriction region of at least three
their residence time inside the nanopore (Figure 3). residues located in the middle section of a captured poly­
The slowed-down passage of these polypeptides across the peptide, situated in a balanced force regime stemming from
nanopore, which is better seen at increased potentials (see the tug of war between the oppositely oriented electric forces
Figure 3, compare traces shown in panels (a)–(c) at ΔV = +40 mV exerted at its ends.
to those illustrated in panels (d)–(f) at ΔV = +80 mV), provides For a detailed view of this phenomenon, we show in Figure 4,
an opportunity to measure with increased signal-to-noise ratio representative excerpts from original recordings that illustrate
ionic currents fluctuations, as in expanded traces in Figure 3d–f. the time unfolding of the fractional blockade currents ensued
By considering the essential geometrical descriptors of the by a single Pe1 (Ac-(R)12 – (A)6 – (E)12 – NH2) a), Pe3 (Ac-(R)12 –
α-HL’s constriction region situated between M-113 and alter- (W)6 – (E)12 – NH2) b), or Pe5 polypeptide (Ac-(R)12 – (A)3 –
nating K-147 and E-111 from each of the seven monomers of (W)3 – (E)12 – NH2) c), while assuming a metastable state inside
the heptameric nanopore (length of ≈0.6 nm and diameter of the α-HL. To account for the differences in the current block-
≈1.4 ÷ 1.5 nm),[120,121] and knowing that each amino acid on the ades shown in Figure 4II,a–c, we calculated the relative ionic

Small Methods 2020, 1900595 1900595 (5 of 13) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.small-methods.com

Figure 4. Testing the ability of the α-HL to distinguish between representative amino acids in a peptide, from ion current fluctuations across the
nanopore. I) A macrodipole-like polypeptide captured in a metastable state inside the α-HL, presents the amino acid residues from its middle region
to the nanopore’s constriction domain a). From this state, thermal position fluctuations of the polypeptide, leading to the partial displacement from
the α-HL’s constriction region of a group of ≈3 amino acids, followed by the re-entry of another such group b,c), generate corresponding ionic current
flickering d,e). II) In a real-life example, the oppositely oriented electric forces acting at the termini of the transiently captured Pe1 (Ac-(R)12 – (A)6 –
(E)12 – NH2) a), Pe3 (Ac-(R)12 – (W)6 – (E)12 – NH2) b), or Pe5 polypeptide (Ac-(R)12 – (A)3 – (W)3 – (E)12 – NH2) c), diminishes the analyte’s kinetics
inside the nanopore, and reveals distinct blockade events associated to the polypeptide movement along the nanopore’s constriction region. The all-
points histograms pinpoint the amplitude distribution of the sub-blockade events (“&” and “$” levels), distinct from the fully blocked substate (“#”),
suggestive of a group of ≈3 residues from the peptide’s middle region of a captured peptide trafficking across the nanopore’s constriction domain.
Reprinted with permission.[72] Copyright 2017 American Chemical Society.

  The positioning of the polypeptide inside the nanopore asso-


current blockade  ∆I blocked  , where ΔIblocked = Iblocked − Iopen,
 I open  ciated to the “#” substate is unstable, as its movement toward
the cis or trans side caused mainly by thermal fluctuations
Iblocked corresponds to either of the “#,” “$,” or “&” substates,
tends to unbalance the electric forces acting at its termini.
and Iopen measures the ionic current through the free α-HL
We proposed that the reversible current fluctuations around
in the absence of interactions (see Figure 3a–f, “free α-HL”
level “#,” to substates of lower blockade seen as “$” (Pe3) and
substate).
“&” (Pe1), respectively, reflect polypeptide occupancy states
We recall that within a volume-exclusion model (vide supra)
inside the nanopore, whereby the constriction region har-
and simplifying assumptions,[58,59] an electrically neutral ana-
bors a lesser number of amino acids as compared to the “#”
lyte captured inside a cylinder-like nanopore determines cur-
substate (see also the schematics in Figure 4I,c). This stems
rent blockades, ΔIblocked, directly proportional to the electrolyte
from the fact that α-HL’s constriction region is the narrowest
volume excluded by the analyte (δ) (see also Figure 1, inset).[29]
domain inside the nanopore (volume of ≈924 Å3), so that the
More accurate descriptions based on Poisson–Nernst–Planck
net ionic current blockade caused by a captured polypeptide is
equation,[21] Ohm law with local conductivities[123] and quasi-
critically dependent on the occupancy extent of it. Experimental
1D models for pore resistance[124,125] have been developed for
observations support this assertion, as the relative blockade
estimating the current blockade. In accordance to the scenario
amplitude of the “$” substate measured for the W-containing
depicted in Figure 4I,b, ΔIblocked estimated for the “#” substate
∆I blocked
reflects the current blockade determined by a peptide captured peptide ( ($, Pe3) = 0.93 ± 4 × 10−3), is slightly larger
I open
inside the nanopore, precisely when a group of approximately
than that of the “&” substrate measured for the A-containing
three amino acids from the middle section of it occupies the con- ∆I blocked
striction region of the α-HL. As the relative blockade amplitude peptide ( (&, Pe1) = 0.83 ± 6 × 10−3). Another worth-
I open
of the “#” substate is slightly larger for the W-containing pep- mentioning benefit of this approach lies in that the oppositely-
∆I blocked
tide Pe3 ( (#; Pe3) = 0.98 ± 9.1 × 10−4), as compared to the oriented forces acting at the polypeptide’s ends exert a confor-
I open
∆I blocked mational control over the trapped molecule, removing folds
A-containing peptide Pe1 ( (#; Pe1) = 0.95 ± 6 × 10−3), it
I open and maintaining it in a largely linear topology.[119] This has the
reflects well the fact that individual W residues displays a larger potential to augment the accuracy of volumetric identification
volume of electrolyte (≈239 Å3) than A residues (≈100 Å3) do.[126] of amino acids.

Small Methods 2020, 1900595 1900595 (6 of 13) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.small-methods.com

Based on these findings, a next objective was to assess the 3. Insights on Protein Sequencing from Atomistic
system’s ability to discriminate between groups of three distinct Simulation
amino acids, from fluctuations in a single blockade event. In
such experiments, we used another polypeptide construct (Pe5), Molecular dynamics (MD) simulations have proved a useful
presenting in its middle section three A followed by three W computational tool to investigate transport phenomena in bio-
residues (Ac-(R)12 – (A)3 – (W)3 – (E)12 – NH2). The analysis of logical nanopore sensing systems and to interpret experimental
the ionic current fluctuations accompanying such a construct data. Examples are the estimation of α-HL ionic conductance
being transiently captured inside the nanopore, revealed three and electro-osmotic flow,[120,131,132] investigation of single-
distinct peaks, which were denoted in Figure 4II,c by “#,” “$,” stranded DNA molecules inside α-HL,[133] the cotranslocational
and “&.” Based on their relative amplitudes, which for the deep unfolding of thioredoxin,[134] and DNA base distinguishability
(“#”) and, respectively, shallower events (“$” and “&”) equal in MspA.[135]
∆I block ∆I block A core question in nanopore protein sequencing is the capa-
(#, Pe5) = 0.94 ± 10−3, ($, Pe5) = 0.89 ± 3 × 10−3, bility to distinguish among the 20 amino acids. In the sensing
I open I open
∆I block approach based on ion current blockade, this amounts to check
and (&, Pe5) = 0.84 ± 2 × 10−3, which are close to those
I open if each amino acid can be unambiguously associated with a
measured in the case of Pe3 and Pe1 polypeptides (vide supra), specific ion current blockade level. As a preliminary step in
these data were interpreted as evidence of the system being assessing the capability of α-HL to distinguish between all the
able to distinguish between groups of distinct amino acids on 20 standard amino acids, we estimated the reduction of avail-
polypeptides containing a mixture of three A and W residues in able space for ionic transport through the nanopore induced by
their middle section. 20 different small homopeptides, one for each standard amino
By using a similar paradigm, we proved the ability of the acid, using all-atom MD simulations.[124] The system, sketched
α-HL nanopore to identify three-amino acids long patches in Figure 5a, is constituted by the α-HL nanopore (blue)
of polar S (serine) and aliphatic I (isoleucine) residues in the embedded into a lipid membrane (gray). A 35-residues homo-
primary structure of polypeptides.[73] Recently, the robustness of peptide (orange chain) is imported into the nanopore with the
our approach was also confirmed by independent experimental central residue close to the nanopore constriction. The simula-
results by Zhao et al.[127] and Pérez et al.[128] that employed a tion box is filled up by 2 m KCl electrolyte solution.
FraC nanopore, and by coarse-grained computational analysis In principle, ionic fluxes can be directly sampled through
by Ghosh and Chaudhury.[129] nonequilibrium MD by applying an external electric field to
If combined with click addition of positive and negative the system.[136] However, due to high thermal noise, this pro-
tails to the termini of a short peptide, this approach for resi- cedure requires very long simulations to reduce the statistical
dues identification can be applied to any peptide construct, error on the estimation of the average current. Moreover, dif-
regardless of its charge. In this scenario, the click addition ferent metastable states of the peptide inside the channel may
of a single strand DNA tail to a peptide terminal introduced give rise to significant current blockade differences. Hence,
by Biswas et al.[130] can potentially foster the use of the nano- multiple replicas of each system are needed, making the com-
pore tweezer method for other nanopore protein sensing putational cost even larger and impractical for the estimation of
applications. all the 20 standard amino acids. To overcome this difficulty, we

Figure 5. Atomistic simulation of nanopore steric exclusion by homopeptides. a) System setup. The system is constituted by an α-HL blue nanopore
(blue) embedded into a lipid membrane (gray). A 35-residues homopeptide (orange chain) is imported into the nanopore with the central residue close
to the nanopore constriction. The simulation box is filled up by 2 m KCl electrolyte solution, that, for the sake of clarity, is not shown. b) Nanopore steric
exclusion estimator for all residues (Va is the amino acid volume). Gray circles correspond to hydrophobic residues, yellow squares to polar, blue up-
triangles to positively charged residues, and red down-triangles to negatively charged ones. The dashed line is the least square regression line. Error
bars represent the standard error of the mean over five independent replicas, and they are reported only when larger than symbols. Figure adapted
from Di Muccio et al.[124]

Small Methods 2020, 1900595 1900595 (7 of 13) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.small-methods.com

formulated a simple theoretical model for the estimation of the that need to be fulfilled by a working nanopore protein sensing
electrical resistance of the nanopore. This approach requires device. The other crucial issue is the control of protein cap-
relatively short equilibrium runs that allowed us to reduce by ture and its sequential translocation.[36] In order to address the
a factor of four the computational requirements. We tested this difficulties associated with the control of protein capture and
formulation against direct ionic current measurements from translocation, it is instrumental to briefly recall how these prob-
MD simulations, for four selected amino acids (A, Q, F, W), and lems were solved for the nucleic acids sequencing. For DNA,
found that our nanopore steric exclusion estimator correlates capture is achieved by electrophoresis induced by the poten-
quite well with the measured current blockade,[124] allowing us tial difference applied across the nanopore, as the molecule is
to extend this procedure to all the 20 canonical amino acids. highly charged (e.g., in 1 m salt, the effective charge of ssDNA
Figure 5b reports the nanopore steric exclusion estimator as is ≈30% of the charge of bare ssDNA.[138] Precise translocation
a function of the amino acid volume. As expected, our results control, however, required more efforts.[139,140] In brief, the
show that the amino acid volume is the main feature that rules most effective solution was to employ a processive enzyme.[141]
the nanopore steric exclusion. In addition, we also find that the Indeed, several enzymes are naturally able to bind the DNA
amino acid charge significantly influences the nanopore steric molecule and to move along its chain to perform their bio-
exclusion estimator. Indeed, charged residues show a significant logical function, such as DNA repairing and replication (e.g.,
minor steric hindrance with respect to uncharged amino acids polymerase). These enzymes can be adapted to push the DNA
of similar volume as apparent, for instance, by the comparison chain through the nanopore ensuring a unidirectional single-
between D (aspartate) and N (asparagine) in Figure 5b. We also monomer translocation on a millisecond time scale, that, in the
find that ion concentration inside the nanopore is higher for DNA case, allows to recognize single bases.[140] Notably though,
charged residues (i.e., the charged homopeptides bring a coun- electrophoretic capture and enzyme-controlled translocation
terion shell into the nanopore), and that charged homopeptides as mentioned above cannot be straightforwardly extended to
lie in a slightly stretched conformation. Both these observa- peptide chains. If we exclude extreme acidic and alkaline pHs,
tions suggest that charged homopeptides leave more room for proteins can be positively or negatively charged, with the charge
the electrolyte passage and may tentatively explain the smaller not being uniformly distributed along the chain. This ren-
value of the nanopore steric exclusion estimator with respect to ders electrophoresis not a robust method to capture a generic
residues of the same size. A similar result was recently reported (positive, neutral, or negative) protein on nanopores. Several
in a computational work by Si et al.,[137] where ionic currents strategies were proposed to overcome this issue and achieve
of different short homo- and hetero-peptides translocating an efficient electrophoresis-induced capture. One possibility
through a graphene nanopore are studied via nonequilibrium is to add a charged tail to one protein terminal.[130,142–144] In
MD. In particular, it is reported that two different protonation this approach, the charged tail enters the nanopore by electro-
states of the histidine residue show two significantly different phoresis and it induces the translocation of the whole peptide
current signals, the larger current being related to the charged chain, supplementary assisted in certain cases by a molecular
state. Hence, for similar amino acids dimensions, the charged motor.[143] A second strategy is to employ SDS, an anionic com-
residue give rise to a smaller blockade. Another recent compu- pound that in combination with heat and reducing agents is
tational work by Wilson et al.,[123] compares the blockades of all able to denature the protein and create a negative charged shell
20 individual amino acids in MspA nanopore, reporting again around it.[27,51,90,104,145,146]
volume-dependent blockades (i.e., blockade increases with Our groups focused on a third strategy, namely adjusting the
volume), but not a significant charge effect. This work employs electro-osmotic flow (EOF).[69,132] EOF refers to the net motion
a reduced steric exclusion model to estimate the current block- of the solvent (usually water) induced by an applied voltage. If
ades computing a 3D local conductivity map of the electrolyte the nanopore is selective for cations (or anions), the positive and
inside the channel, based on the distance of each point from the negative ion distributions inside the nanopore differ and, con-
nearest nonsolvent atom. This approach provides ionic current sequently, in some regions of the nanopore, the electrolyte solu-
blockade estimation in good agreement with nonequilibrium tion presents an average net charge. Consequently, the external
runs. However, the local conductivity map does not take into field induced by the applied voltage results in a net force that
account the effects of surface charge of the biomolecules, this acts on the solvent, thus generating the EOF, see Figure 6a,b.
may be one of the possible reasons for the different behavior For biological nanopores, it is also feasible to tune their surface
of charged residues when compared with our study,[124] and charge by specific residues mutations and/or changing the pH
Si et al. work.[137] These recent studies, together with the con- of the solution. For the specific case of α-HL, we showed that at
stant increase in the computational performance, suggest that low pHs, the EOF is so intense that it is able to capture charged
in the near future MD simulations will play an important role peptides against electrophoresis,[69] see Figure 6c,d. Atomistic
to elucidate the main mechanism responsible for the current simulations provided a qualitative explanation of the process
blockades. and, more importantly, a quantitative estimation of the EOF.[132]
The reduction of pH from 7 to 2.8 strongly alters the charge
distribution at the β-barrel entrance, where the surface charge
4. Capture and Translocation Control switches from negative (pH = 7) to positive (pH = 2.8) see
Figure 6e,f. Hence, α-HL is more anion selective at low pH
In Sections 1–3, we discussed mainly about the capability of and, consequently, the EOF flux increases (Figure 6g). The pos-
nanopore systems to distinguish the different amino acids. sibility to employ EOF to capture molecules irrespective of their
However, distinguishability is only one of the requirements charge was also recently explored by other research groups for

Small Methods 2020, 1900595 1900595 (8 of 13) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.small-methods.com

Figure 6. Electro-osmotic flow (EOF) and peptide capture. Nanopore surface charges alter the ion distribution in the electrolyte solution. Ions of
opposite charge with respect to surface accumulate in a thin layer close to the wall, e.g., a positively charged surface attracts a negative cloud of ions
a). The external electric field induced by the applied voltage ΔV results in a net force on the charged fluid layer that puts in motion the fluid. For positively
charged surface a), the fluid velocity ueo is opposed to the direction of the electrical field E, while, for negative charged surfaces b), ueo and E have the
same direction. c,d) EOF can be used to capture a peptide against electrophoresis. In the reported example,[69] a positively charged peptide is initially
present on the α-HL’s cis side, and the electrical field as well as electrophoretic forces Fepl are directed from the trans to cis side. The experiment is
performed at pH 2.8, where α-HL is highly positively charged. Current traces adapted from[69] show a multilevel signal that is interpreted as the move-
ment of the peptide along different regions of the nanopore, hence the capture is induced by EOF against electrophoresis. e,f) Ion concentration maps
from equilibrium molecular dynamics, MD, simulations. The regions not accessible to ions (membrane and channel) are in dark blue. The red peaks
at both pH in the Cl− density correspond to the anions accumulation caused by the presence of positive residue rings at the constriction. A marked
increase of Cl− ions in the trans region for pH = 2.8 is responsible of the enhancement of the nanopore selectivity and, consequently, of the EOF.
g) EOF from MD simulations as a function of ΔV at pH 7 and pH 2.8 for 2 m KCl, adapted from Bonome et al.[132]

FraC[147,148] and Aerolysin[149] nanopores. Even more, achieve- of the three main requirements, peptide capture, transloca-
ments were accomplished for the goal of the efficient use of tion control, and amino acid distinguishability can be con-
EOF in molecule capturing and investigation.[150,151] sidered, to date, fully solved. Biological nanopores can be
Concerning translocation control, there is no direct equiva- engineered to increase their capability to distinguish among
lent of the DNA processive enzyme (roughly speaking, there is the different amino acids and/or to improve the tuning pos-
no protein-polymerase) that may be easily engineered to control sibility of the electro-osmotic flow. These two requirements
translocation. An interesting approach was to use an unfoldase can be fulfilled independently since distinguishability takes
and in particular, the ClpX unfoldase was placed in solution place is mainly at the nanopore constriction, while EOF is
on the trans side of α-HL, while the protein to be translocated sensitive to surface charges alterations in wider regions of the
was present on the cis side.[143] A peptide tag was added to nanopore, such as the α-HL vestibule, the large open of the
the protein terminal allowing ClpX to specifically bind the C FraC funnel,[147] and the two wide vestibules of CsgG.[152,153]
terminus of the protein, when translocating to the trans com- Concerning distinguishability at single amino acid level with
partment. At this point, ClpX induces the translocation of the the α-HL, the ionic current blockade signal corresponding to
protein. The approach enabled discrimination among distinct the amino acid in the constriction will be also affected by the
protein domains and among variants of these protein domains amino acids that occupy the β-barrel. In other biopores such
due to sequence differences.[32] With particular appeal for such as Mpsa,[93] and the already mentioned FraC and CsgG, this
future endeavors, gold nanoparticle plasmon heating may be effect should be less relevant since the section of the vestibule
employed for the purposes of small proteins unfolding and is much larger than the section of the constriction. Another
analysis with nanopores.[51] possibility is to modify the α-HL by cutting a portion of the
β-barrel region close to the trans side, as reported by Stod-
dart et al.[154] and numerically analyzed by Di Muccio et al.[124]
5. Open Challenges and Opportunities Beside the progresses in the capability to modify biological
pores, the possibility to distinguish all the 20 amino acids
Despite the recent progresses, the route toward nanopore pro- (or, at least, a large subset of them) still remains a formidable
tein sensing is still full of interesting open challenges. None open challenge.

Small Methods 2020, 1900595 1900595 (9 of 13) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.small-methods.com

Theoretical insights on the feasibility of exopeptidase


sequencing are complex since the transport of single monomers
into the nanopore is ruled by the interplay among EOF, dielec-
trophoresis, Brownian diffusion, specific chemical interactions
and, for charged residues, electrophoresis. A theoretical work
by Reiner at al.,[159] analyzed the exonuclease/nanopore-based
DNA sequencing. They studied the capture of a monomer
under the effect of electrophoresis and electroosmosis showing
limitations of the exonuclease approach for DNA sequencing.
In particular, for applied voltage typically used in biological
nanopore systems, they found a capture probability of below
0.5 for a molecule released at the pore entrance under typical
voltage employed for biological nanopores. We expect that these
limitations will be relevant also for the protein sequencing
case. A main obstacle is the Brownian motion which for small
molecules may overwhelm all the other effects, introducing a
mismatch between the protein sequence and the order (and
the number) of the residues traversing the nanopore. For this
Figure 7. Exopepdidase-based nanopore protein sequencing. a) A pos- purpose, any method that can reduce the Brownian diffusion
sible approach to nanopore protein sequencing is based on sequentially (e.g., increasing the viscosity of the solvent) would be highly
cutting single amino acids from one protein terminal and then importing beneficial. In addition, pressure gradients can be exploited to
them one by one into the nanopore. Three main conditions are needed: further increase the incoming solvent flow and, hence, the cap-
i) an exopeptidase able to cut all the 20 amino acids, ii) a way to bind the ture probability. Other possible strategies to limit the effect of
exopeptidase close to the nanopore entrance, and iii) a method to import
Brownian diffusion may involve modifications to the pore set-
all the residues into the nanopore preserving the chain order of the trans-
locating monomers. This last requirement may be obtained via dielec- up, such as an additional confinement around the exopeptidase
trophoresis b). Indeed, there are no conceptual limitations that impede or different pore design. A preliminary analysis neglecting
the dielectrophoretic capture and trapping, described for macrodipole in EOF and dielectrophoresis was presented a few years ago for a
Sections 2–3, for small dipole, such as a single amino acid. Another pos- tandem nanopore set-up.[160] Further analysis in this direction
sibility is EOF (electro-osmotic flow) (c) that proved efficient to capture could allow to get first insights on the feasibility of exopepti-
molecules irrespective of their charge.
dase nanopore sequencing.
Although not described in this review, solid-state nano-
Molecular motors, as protein unfoldase ClpX employed pores may also play an important role in the development of
in Nivala et al.,[143] can be employed for translocation control. protein sequencers. Indeed, solid-state nanopore allows more
To be effective, these motors probably need to be stuck at the options in the selection of the electrolyte solution features,
nanopore when translocating the protein, in order to reduce such as the above cited SDS approach.[27,90,145] Moreover,
fluctuations and ensure a strictly unidirectional translocation. solid-state nanopore, constructed in particular from 2D mate-
Another interesting approach is to employ a molecular hopper rials, may be employed in tunneling sensing where the single
moved in an electric field and controlled by a chemical ratchet monomer is detected measuring the transverse electronic cur-
as proposed by Qing et al.[94] rent.[37] Concerning protein sequencing, recent computational
An alternative direction, inspired from the exonuclease studies have shown the possibility to detect specific fingerprints
DNA sequencing approach,[155] is to preserve the chain order of the atoms involved in the peptide bonds,[38,161] although
of the translocating monomers sequentially cutting single experimental evidences still need to be provided. Progress on
amino acids from one protein terminal and then importing solid-state nanopore fabrication may also open the way to com-
them one by one into the nanopore, Figure 7. This exopepti- bined approaches that merge the longitudinal ion current and
dase method requires three main ingredients: i) an exopepti- transversal tunneling current tunneling to improve the signal
dase able to cut all the 20 amino acids, ii) a way to bind the distinguishability.
exopeptidase close to the nanopore entrance, and iii) a robust As a final comment, the efforts spent on protein sequencing
method to import all the residues into the nanopore. Con- may lead to innovative approaches potentially useful for other
cerning the first requirement, computational approaches may applications on protein sensing. Examples are single molecule
help in the design of such an enzyme as recently did for the protein identification,[145,162] and detection of post-translational
development of an Edmanase, an engineered enzyme capable modifications.[28,101] Notably, during the writing of this work,
to perform the stepwise removal of N-terminal amino acids a landmark paper on the topic reviewed herein appeared in
in an Edman degradation.[156] The second requirement, trans- which authors reported the single-molecule detection of all 20
location control, appears in principle slightly simpler since it proteinogenic amino acids in an aerolysin nanopore, with the
may be achieved by covalently attaching proper linkers to both help of a short polycationic carrier.[163]
nanopore and exopeptidase. A possible method it to apply a Despite all existing challenges reviewed herein, we firmly
complementary DNA strand; indeed, the binding of DNA believe that proof-of-concept experiments involving the α-HL
strands to nanopore is quite a robust technique in the nano- nanopore to demonstrate amino acids recognition of individual
pore field.[157,158] polypeptides, could pave the way toward the nanopore-based

Small Methods 2020, 1900595 1900595 (10 of 13) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.small-methods.com

polypeptide sequencing in the near future, and help pinpointing [19] J. A. Huang, M. Z. Mousavi, Y. Zhao, A. Hubarevich, F. Omeis,
post-translational modifications or quickly and cheaply identify G. Giovannini, M. Schütte, D. Garoli, F. De Angelis, Nat. Commun.
the sequence of proteins which are key for the onset and devel- 2019, 10, 5321.
opment of many crippling diseases. [20] J. W. F. Robertson, C. G. Rodrigues, V. M. Stanford, K. A. Rubinson,
O. V. Krasilnikov, J. J. Kasianowicz, Proc. Natl. Acad. Sci. USA 2007,
104, 8207.
[21] J. E. Reiner, J. J. Kasianowicz, B. J. Nablo, J. W. F. Robertson, Proc.
Acknowledgements Natl. Acad. Sci. USA 2010, 107, 12080.
[22] S. Kumar, C. Tao, M. Chien, B. Hellner, A. Balijepalli,
A.A. and G.D.M. contributed equally to this work. The work was supported J. W. F. Robertson, Z. Li, J. J. Russo, J. E. Reiner, J. J. Kasianowicz,
by UEFISCDI Grant Nos. PN-III-P4-ID-PCE-2016-0026, PN-IIIP1-
J. Ju, Sci. Rep. 2012, 2, 684.
1.1-TE-2016-0508, PN-III-P1-1.1-PD-2016-0737, 34PFE/19.10.2018, and
[23] C. W. Fuller, S. Kumar, M. Porel, M. Chien, A. Bibillo,
PN-III-P1-1.2-PCCDI-2017-0010/74PCCDI⁄2018 (PNCDI III).
P. B. Stranges, M. Dorwart, C. Tao, Z. Li, W. Guo, S. Shi,
D. Korenblum, A. Trans, A. Aguirre, E. Liu, E. T. Harada, J. Pollard,
A. Bhat, C. Cech, A. Yang, C. Arnold, M. Palla, J. Hovis, R. Chen,
Conflict of Interest I. Morozova, S. Kalachikov, J. J. Russo, J. J. Kasianowicz, R. Davis,
S. Roever, G. M. Church, J. Ju, Proc. Natl. Acad. Sci. USA 2016, 113,
The authors declare no conflict of interest. 5233.
[24] E. A. Manrao, I. M. Derrington, A. H. Laszlo, K. W. Langford,
M. K. Hopper, N. Gillgren, M. Pavlenok, M. Niederweis,
J. H. Gundlach, Nat. Biotechnol. 2012, 30, 349.
Keywords [25] J. J. Kasianowicz, J. W. F. Robertson, E. R. Chan, J. E. Reiner,
electrophysiology, molecular dynamics, polypeptide sequencing, protein V. M. Stanford, Annu. Rev. Anal. Chem. 2008, 1, 737.
nanopores, single-molecule sensing [26] S. Howorka, Z. Siwy, Chem. Soc. Rev. 2009, 38, 2360.
[27] E. Kennedy, Z. Dong, C. Tennant, G. Timp, Nat. Nanotechnol.
Received: August 28, 2019 2016, 11, 968.
Revised: January 14, 2020 [28] C. B. Rosen, D. Rodriguez-Larrea, H. Bayley, Nat. Biotechnol. 2014,
Published online: 32, 179.
[29] D. S. Talaga, J. Li, J. Am. Chem. Soc. 2009, 131, 9287.
[30] A. K. Thakur, L. Movileanu, Nat. Biotechnol. 2019, 37, 96.
[31] R. Wei, V. Gatterdam, R. Wieneke, R. Tampe, U. Rant, Nat. Nano-
[1] R. McDaniell, D. M. Warthen, P. A. Sanchez-Lara, A. Pai, technol. 2012, 7, 257.
I. D. Krantz, D. A. Piccoli, N. B. Spinner, Am. J. Hum. Genet. 2006, [32] J. Nivala, L. Mulroney, G. Li, J. Schreiber, M. Akeson, ACS Nano
79, 169. 2014, 8, 12365.
[2] M. Katoh, M. Katoh, Clin. Cancer Res. 2007, 13, 4042. [33] P. Waduge, R. Hu, P. Bandarkar, H. Yamazaki, B. Cressiot,
[3] J. S. Valastyan, S. Lindquist, Dis. Models Mech. 2014, 7, 9. Q. Zhao, P. C. Whitford, M. Wanunu, ACS Nano 2017, 11, 5706.
[4] I. Dalle-Donne, G. Aldini, M. Carini, R. Colombo, R. Rossi, [34] L. R. Pérez, C. Joo, C. Dekker, Nat. Nanotechnol. 2018, 13, 786.
A. Milzani, J. Cell. Mol. Med. 2006, 10, 389. [35] J. W. Robertson, J. E. Reiner, Proteomics 2018, 18, 1800026.
[5] B. Domon, R. Aebersold, Science 2006, 312, 212. [36] M. Chinappi, F. Cecconi, J. Phys.: Condens. Matter 2018, 30, 204002.
[6] T. E. Angel, U. K. Aryal, S. M. Hengel, E. S. Baker, R. T. Kelly, [37] M. Di Ventra, M. Taniguchi, Nat. Nanotechnol. 2016, 11, 117.
E. W. Robinson, R. D. Smith, Chem. Soc. Rev. 2012, 41, 3912. [38] A. E. Rossini, F. Gala, M. Chinappi, G. Zollo, Nanoscale 2018, 10,
[7] P. Edman, Acta Chem. Scand. 1950, 4, 283. 5928.
[8] K. G. Standing, Curr. Opin. Struct. Biol. 2003, 13, 595. [39] L. Mereuta, I. Schiopu, A. Asandei, Y. Park, K. S. Hahm, T. Luchian,
[9] G. Allen, in Sequencing of Proteins and Peptides, Vol. 9 (Eds: Langmuir 2012, 28, 17079.
T. S. Work, R. H. Burdon), Elsevier, Amsterdam, Netherlands 2011, [40] A. Asandei, L. Mereuta, T. Luchian, Biophys. Chem. 2008, 135, 32.
pp. 161–234. [41] E. C. Yusko, P. Prangkio, D. Sept, R. C. Rollings, J. Li, M. Mayer,
[10] X. Yongsheng, M. M. Vecchi, D. Wen, Anal. Chem. 2016, 88, 10757. ACS Nano 2012, 6, 5909.
[11] R. A. Laursen, Eur. J. Biochem. 1971, 20, 89. [42] L. Mereuta, A. Asandei, C. H. Seo, Y. Park, T. Luchian, ACS Appl.
[12] P. Schmitt-Kopplin, M. Frommberger, Electrophoresis 2003, 24, 3837. Mater. Interfaces 2014, 6, 13242.
[13] J. J. Kasianowicz, E. Brandin, D. Branton, D. W. Deamer, Proc. [43] L. Movileanu, J. P. Schmittschmitt, J. M. Scholtz, H. Bayley,
Natl. Acad. Sci. USA 1996, 93, 13770. Biophys. J. 2005, 89, 1030.
[14] C. Cao, Y. L. Ying, Z. L. Hu, D. F. Liao, H. Tian, Y. T. Long, Nat. [44] Y. Zhao, B. Ashcroft, P. Zhang, H. Liu, S. Sen, W. Song, J. Im,
Nanotechnol. 2016, 11, 713. B. Gyarfas, S. Manna, S. Biswas, C. Borges, S. Lindsay, Nat.
[15] D. Branton, D. W. Deamer, A. Marziali, H. Bayley, Nanotechnol. 2014, 9, 466.
S. A. Benner, T. Butler, M. Di Ventra, H. Garaj, A. Hibbs, [45] L. J. Steinbock, S. Krishnan, R. D. Bulushev, S. Borgeaud,
X. Huang, S. B. Jovanovich, P. S. Krstic, S. Lindsay, X. S. Ling, M. Blokesch, L. Feletti, A. Radenovic, Nanoscale 2014, 6, 14380.
C. H. Mastrangelo, A. Meller, J. S. Oliver, Y. V. Pershin, [46] F. Piguet, H. Ouldali, M. Pastoriza-Gallego, P. Manivet, J. Pelta,
J. M. Ramsey, R. Riehn, G. V. Soni, V. Tabard-Cossa, M. Wanunu, A. Oukhaled, Nat. Commun. 2018, 9, 966.
M. Wiggin, J. A. Schloss, Nat. Biotechnol. 2008, 26, 1146. [47] A. Asandei, S. Iftemi, L. Mereuta, I. Schiopu, T. Luchian, Am. J.
[16] C. Cao, J. Yu, M. Y. Li, Y. Q. Wang, H. Tian, Y. T. Long, Small 2017, Hum. Genet. 2014, 247, 523.
13, 1702011. [48] V. Van Meervelt, M. Soskine, S. Singh, G. K. Schuurman-Wolters,
[17] N. An, A. M. Fleming, H. S. White, C. J. Burrows, Proc. Natl. Acad. H. J. Wijma, B. Poolman, G. Maglia, J. Am. Chem. Soc. 2017, 139,
Sci. USA 2012, 109, 11504. 18640.
[18] D. Stoddart, A. Heron, E. Mikhailova, G. Maglia, H. Bayley, Proc. [49] N. Varongchayakul, J. Song, A. Meller, M. W. Grinstaff, Chem. Soc.
Natl. Acad. Sci. USA 2009, 106, 7702. Rev. 2018, 47, 8512.

Small Methods 2020, 1900595 1900595 (11 of 13) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.small-methods.com

[50] J. S. Yu, S. C. Hong, S. Wu, H. M. Kim, C. Lee, J. S. Lee, J. E. Lee, [85] C. Cao, Y. T. Long, Acc. Chem. Res. 2018, 51, 331.
K. B. Kim, Nanoscale 2019, 11, 2510. [86] L. Q. Gu, J. W. Shim, Analyst 2010, 135, 441.
[51] J. E. Reiner, J. W. Robertson, D. L. Burden, L. K. Burden, [87] R. Wei, T. G. Martin, U. Rant, H. Dietz, Angew. Chem., Int. Ed.
A. Balijepalli, J. J. Kasianowicz, J. Am. Chem. Soc. 2013, 135, 3087. 2012, 51, 4864.
[52] J. E. Reiner, A. Balijepalli, J. W. F. Robertson, J. Campbell, J. Suehle, [88] A. R. Hall, A. Scott, D. Rotem, K. K. Mehta, H. Bayley, C. Dekker,
J. J. Kasianowicz, Chem. Rev. 2012, 112, 6431. Nat. Nanotechnol. 2010, 5, 874.
[53] T. Luchian, Y. Park, A. Asandei, I. Schiopu, L. Mereuta, A. Apetrei, [89] T. Diederichs, G. Pugh, A. Dorey, Y. Xing, J. R. Burns,
Acc. Chem. Res. 2019, 52, 267. Q. H. Nguyen, M. Tornow, R. Tampé, S. Howorka, Nat. Commun.
[54] D. K. Lubensky, D. R. Nelson, Biophys. J. 1999, 77, 1824. 2019, 10, 5018.
[55] J. Ettedgui, J. J. Kasianowicz, A. Balijepalli, J. Am. Chem. Soc. 2016, [90] L. R. Pérez, S. John, A. Aksimentiev, C. Joo, C. Dekker, Nanoscale
138, 7228. 2017, 9, 11685.
[56] S. C. Kim, S. K. Kannam, S. Harrer, Phys. Rev. E 2014, 89, 042702. [91] Y. L. Ying, C. Cao, Y. X. Hu, Y. T. Long, Natl. Sci. Rev. 2018, 5, 450.
[57] R. W. DeBlois, C. P. Bean, Rev. Sci. Instrum. 1970, 41, 909. [92] Y. Q. Wang, C. Cao, Y. L. Ying, S. Li, M. B. Wang, J. Huang,
[58] E. C. Yusko, J. M. Johnson, S. Majd, P. Prangkio, R. C. Rollings, Y. T. Long, ACS Sens. 2018, 3, 779.
J. Li, J. Yang, M. Mayer, Nat. Nanotechnol. 2011, 6, 253. [93] T. Z. Butler, M. Pavlenok, I. M. Derrington, M. Niederweis,
[59] L. Mereuta, M. Roy, A. Asandei, J. K. Lee, Y. Park, I. Andricioaei, J. H. Gundlach, Proc. Natl. Acad. Sci. USA 2008, 105, 20647.
T. Luchian, Sci. Rep. 2015, 4, 3885. [94] Y. Qing, S. A. Ionescu, G. S. Pulcu, H. Bayley, Science 2018, 361,
[60] M. Davenport, K. Healy, M. Pevarnik, N. Teslich, S. Cabrini, 908.
A. P. Morrison, Z. S. Siwy, S. E. Letant, ACS Nano 2012, 6, 8366. [95] L. Song, M. R. Hobaugh, C. Shustak, S. Cheley, H. Bayley,
[61] H. Wang, J. Ettedgui, J. Forstater, J. W. F. Robertson, J. E. Reiner, J. E. Gouaux, Science 1996, 274, 1859.
H. Zhang, S. Chen, J. J. Kasianowicz, ACS Sens. 2018, 3, 251. [96] X. F. Kang, L. Q. Gu, S. Cheley, H. Bayley, Angew. Chem., Int. Ed.
[62] A. Balijepalli, J. W. F. Robertson, J. E. Reiner, J. J. Kasianowicz, 2005, 44, 1495.
R. W. Pastor, J. Am. Chem. Soc. 2013, 135, 7064. [97] R. Stefureac, Y. T. Long, H. B. Kraatz, P. Howard, J. S. Lee, Bio-
[63] A. Y. Grosberg, Y. Rabin, J. Chem. Phys. 2010, 133, 165102. chemistry 2006, 45, 9172.
[64] D. Colquhoun, A. G. Hawkes, in Single-Channel Recording, 2nd [98] R. Stefureac, L. Waldner, P. Howard, J. S. Lee, Small 2008, 4, 59.
ed. (Eds: B. Sakmann, E. Neher), Plenum Press, New York, 1995, [99] G. Oukhaled, J. Mathe, A. L. Biance, L. Bacri, J. M. Betton,
pp. 397–482. D. Lairez, J. Pelta, L. Auvray, Phys. Rev. Lett. 2007, 98, 158101.
[65] T. C. Sutherland, Y. T. Long, R. I. Stefureac, I. Bediako-Amoa, [100] B. Cressiot, A. Oukhaled, L. Bacri, J. Pelta, BioNanoScience 2014, 4,
H. B. Kraatz, J. S. Lee, Nano Lett. 2004, 4, 1273. 111.
[66] H. Y. Wang, Y. L. Ying, Y. Li, H. B. Kraatz, Y. T. Long, Anal. Chem. [101] E. L. Bonome, F. Cecconi, M. Chinappi, Nanoscale 2019, 11, 9920.
2011, 83, 1746. [102] G. M. Roozbahani, X. Chen, Y. Zhang, R. Xie, R. Ma, D. Li, H. Li,
[67] H. Y. Wang, Z. Gu, C. Cao, J. Wang, Y. T. Long, Anal. Chem. 2013, X. Guan, ACS Sens. 2017, 2, 703.
85, 8254. [103] D. V. Verschueren, M. P. Jonsson, C. Dekker, Nanotechnology 2015,
[68] Y. X. Hu, Y. L. Ying, Z. Gu, C. Cao, B. Y. Yan, H. F. Wang, Y. T. Long, 26, 234004.
Chem. Commun. 2016, 52, 5542. [104] L. Payet, M. Martinho, M. Pastoriza-Gallego, J. M. Betton,
[69] A. Asandei, I. Schiopu, M. Chinappi, C. H. Seo, Y. Park, T. Luchian, L. Auvray, J. Pelta, J. Mathé, Anal. Chem. 2012, 84, 4071.
ACS Appl. Mater. Interfaces 2016, 8, 13166. [105] D. Fologea, J. Uplinger, B. Thomas, D. S. McNabb, J. Li, Nano Lett.
[70] I. Schiopu, S. Iftemi, T. Luchian, Langmuir 2015, 31, 387. 2005, 5, 1734.
[71] A. Asandei, I. Schiopu, S. Iftemi, L. Mereuta, T. Luchian, Langmuir [106] R. F. Purnell, K. K. Mehta, J. J. Schmidt, Nano Lett. 2008, 8, 3029.
2013, 29, 15634. [107] J. Nakane, M. Wiggin, A. Marziali, Biophys. J. 2004, 87, 615.
[72] A. Asandei, A. E. Rossini, M. Chinappi, Y. Park, T. Luchian, Lang- [108] C. G. Rodrigues, D. C. Machado, S. F. Chevtchenko,
muir 2017, 33, 14451. O. V. Krasilnikov, Biophys. J. 2008, 95, 5186.
[73] A. Asandei, I. S. Dragomir, G. Di Muccio, M. Chinappi, Y. Park, [109] R. Kawano, A. E. P. Schibel, C. Cauley, H. S. White, Langmuir 2009,
T. Luchian, Polymers 2018, 10, 885. 25, 1233.
[74] S. Li, C. Cao, J. Yang, Y. T. Long, ChemElectroChem. 2019, 6, 126. [110] M. Schiel, Z. S. Siwy, J. Phys. Chem. C 2014, 118, 19214.
[75] Y. Lu, X. Y. Wu, Y. L. Ying, Y. T. Long, Chem. Commun. 2019, 55, [111] M. Firnkes, D. Pedone, J. Knezevic, M. Doblinger, U. Rant, Nano
9311. Lett. 2010, 10, 2162.
[76] H. Bayley, T. Luchian, S. H. Shin, M. B. Steffensen, in Single [112] S. W. Kowalczyk, D. B. Wells, A. Aksimentiev, C. Dekker, Nano Lett.
Molecules and Nanotechnology, Vol. 12 (Eds: R. Rigler, H. Vogel), 2012, 12, 1038.
Springer, Heidelberg, Germany 2008, pp. 251–227. [113] U. F. Keyser, J. van der Does, C. Dekker, N. H. Dekker, Rev. Sci.
[77] J. J. Kasianowicz, A. K. Balijepalli, J. Ettedgui, J. H. Forstater, Instrum. 2006, 77, 105105.
H. Wang, H. Zhang, J. W. Robertson, Biochim. Biophys. Acta, [114] M. Zwolak, M. Di Ventra, Rev. Mod. Phys. 2008, 80, 141.
Biomembr. 2016, 1858, 593. [115] S. E. Henrickson, M. Misakian, B. Robertson, J. J. Kasianowicz,
[78] M. Wanunu, C. A. Merchant, M. Drndic, K. L. Shepard, Nat. Phys. Rev. Lett. 2000, 85, 3057.
Methods 2012, 9, 487. [116] T. Ambjörnsson, S. P. Apell, Z. Konkoli, E. A. Di Marzio,
[79] F. Wohnsland, R. Benz, J. Membr. Biol. 1997, 158, 77. J. J. Kasianowicz, J. Chem. Phys. 2002, 117, 4063.
[80] A. J. W. Hartel, S. Shekar, P. Ong, I. Schroeder, G. Thiel, [117] M. Wanunu, W. Morrison, Y. Rabin, A. Y. Grosberg, A. Meller, Nat.
K. L. Shepard, Anal. Chim. Acta 2019, 1061, 13. Nanotechnol. 2010, 5, 160.
[81] A. Balijepalli, J. Ettedgui, A. T. Cornio, J. W. F. Robertson, [118] M. Chinappi, T. Luchian, F. Cecconi, Phys. Rev. E 2015, 92,
K. P. Cheung, J. J. Kasianowicz, C. Vaz, ACS Nano 2014, 8, 1547. 032714.
[82] B. M. Venkatesan, R. Bashir, Nat. Nanotechnol. 2011, 6, 615. [119] A. Asandei, M. Chinappi, J. K. Lee, C. H. Seo, L. Mereuta, Y. Park,
[83] K. Lee, K. B. Park, H. J. Kim, J. S. Yu, H. Chae, H. M. Kim, T. Luchian, Sci. Rep. 2015, 5, 10419.
K. B. Kim, Adv. Mater. 2018, 30, 1704680. [120] A. Aksimentiev, K. Schulten, Biophys. J. 2005, 88, 3745.
[84] F. Haque, J. Li, H. C. Wu, X. J. Liang, P. Guo, Nano Today 2013, 8, [121] S. E. Henrickson, E. A. Di Marzio, Q. Wang, V. M. Stanford,
56. J. J. Kasianowicz, J. Chem. Phys. 2010, 132, 135101.

Small Methods 2020, 1900595 1900595 (12 of 13) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
www.advancedsciencenews.com www.small-methods.com

[122] E. A. Manrao, I. M. Derrington, M. Pavlenok, M. Niederweis, [144] D. P. Hoogerheide, P. A. Gurnev, T. K. Rostovtseva, S. M. Bezrukov,
J. H. Gundlach, PLoS One 2011, 6, e25723. Biophys. J. 2018, 114, 772.
[123] J. Wilson, K. Sarthak, W. Si, L. Gao, A. Aksimentiev, ACS Sens. [145] S. Ohayon, A. Girsault, M. Nasser, S. Shen-Orr, A. Meller, PLoS
2019, 4, 634. Comput. Biol. 2019, 15, e1007067.
[124] G. Di Muccio, A. E. Rossini, D. Di Marino, G. Zollo, M. Chinappi, [146] M. P. Jonsson, C. Dekker, Nano Lett. 2013, 13, 1029.
Sci. Rep. 2019, 9, 6440. [147] G. Huang, K. Willems, M. Soskine, C. Wloka, G.Maglia , Nat.
[125] W. Si, A. Aksimentiev, ACS Nano 2017, 11, 7091. Commun. 2017, 8, 935.
[126] A. E. Counterman, D. E. Clemmer, J. Am. Chem. Soc. 1999, 121, [148] G. Huang, A. Voet, G. Maglia, Nat. Commun. 2019, 10, 835.
4031. [149] M. Boukhet, F. Piguet, H. Ouldali, M. Pastoriza-Gallego, J. Pelta,
[127] S. Zhao, L. R. Pérez, M. Soskine, G. Maglia, C. Joo, C. Dekker, A. Oukhaled, Nanoscale 2016, 8, 18352.
A. Aksimentiev, ACS Nano 2019, 13, 2398. [150] L. Q. Gu, S. Cheley, H. Bayley, Proc. Natl. Acad. Sci. USA 2003, 100,
[128] L. R. Pérez, C. H. Wong, G. Maglia, C. Dekker, Nano Lett. 2019, 19, 15498.
7957. [151] B. Luan, A. Aksimentiev, Phys. Rev. E 2008, 78, 021912.
[129] B. Ghosh, S. Chaudhury, J. Phys. Chem. B 2019, 123, 4318. [152] P. Goyal, P. V. Krasteva, N. Van Gerven, F. Gubellini, I. Van den
[130] S. Biswas, W. Song, C. Borges, S. Lindsay, P. Zhang, ACS Nano Broeck, A. Troupiotis-Tsaïlaki, W. Jonckheere, G. Péhau-Arnaudet,
2015, 9, 9652. J. S. Pinkner, M. R. Chapman, S. J. Hultgren, S. Howorka,
[131] S. Bhattacharya, J. Muzard, L. Payet, J. Mathé, U. Bockelmann, R. Fronzes, H. Remaut, Nature 2014, 516, 250.
A. Aksimentiev, V. Viasnoff, J. Phys. Chem. C 2011, 115, 4255. [153] B. Cao, Y. Zhao, Y. Kou, D. Ni, X. C. Zhang, Y. Huang, Proc. Natl.
[132] E. L. Bonome, F. Cecconi, M. Chinappi, Microfluid. Nanofluid. Acad. Sci. USA 2014, 111, E5439.
2017, 21, 96. [154] D. Stoddart, M. Ayub, L. Höfler, P. Raychaudhuri,
[133] J. Mathé, A. Aksimentiev, D. R. Nelson, K. Schulten, A. Meller, J. W. Klingelhoefer, G. Maglia, A. Heron, H. Bayley, Proc. Natl.
Proc. Natl. Acad. Sci. USA 2005, 102, 12377. Acad. Sci. USA 2014, 111, 2425.
[134] D. Di Marino, E. L. Bonome, A. Tramontano, M. Chinappi, J. Phys. [155] J. Clarke, H. C. Wu, L. Jayasinghe, A. Patel, S. Reid, H. Bayley, Nat.
Chem. Lett. 2015, 6, 2963. Nanotechnol. 2009, 4, 265.
[135] S. Bhattacharya, Y. Jejoong, A. Aksimentiev, ACS Nano 2016, 10, [156] B. Borgo, J. J. Havranek, Protein Sci. 2015, 24, 571.
4644. [157] D. Rotem, L. Jayasinghe, M. Salichou, H. Bayley, J. Am. Chem. Soc.
[136] P. S. Crozier, D. Henderson, R. L. Rowley, D. D. Busath, Biophys. J. 2012, 134, 2781.
2001, 81, 3077. [158] R. Kawano, T. Osaki, H. Sasaki, M. Takinoue, S. Yoshizawa,
[137] W. Si, Y. Zhang, G. Wu, Y. Kan, Y. Zhang, J. Sha, Y. Chen, Small S. Takeuchi, J. Am. Chem. Soc. 2011, 133, 8474.
2019, 15, 1900036. [159] J. E. Reiner, A. Balijepalli, J. W. F. Robertson, B. S. Drown,
[138] B. Luan, G. Martyna, G. Stolovitzky, Biophys. J. 2011, 101, 2214. D. L. Burden, J. J. Kasianowicz, J. Chem. Phys. 2012, 137, 214903.
[139] M. Wanunu, Phys. Life Rev. 2012, 9, 125. [160] G. Sampath, RSC Adv. 2015, 5, 30694.
[140] M. Jain, H. E. Olsen, B. Paten, M. Akeson, Genome Biol. 2016, 17, [161] G. Zollo, A. E. Rossini, Nanoscale Adv. 2019, 1, 3547.
239. [162] Y. Yao, M. Docter, J. van Ginkel, D. de Ridder, C. Joo, Phys. Biol.
[141] G. M. Cherf, K. R. Lieberman, H. Rashid, C. E. Lam, K. Karplus, 2015, 12, 055003.
M. Akeson, Nat. Biotechnol. 2012, 30, 344. [163] H. Ouldali, K. Sarthak, T. Ensslen, F. Piguet, P. Manivet, J. Pelta,
[142] R. L. David, H. Bayley, Nat. Nanotechnol. 2013, 8, 288. J. C. Behrends, A. Aksimentiev, A. Oukhaled, Nat. Biotechnol.
[143] J. Nivala, D. B. Marks, M. Akeson, Nat. Biotechnol. 2013, 31, 247. 2019, https://doi.org/10.1038/s41587-019-0345-2.

Small Methods 2020, 1900595 1900595 (13 of 13) © 2020 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

You might also like