ArticlePDF AvailableLiterature Review

Abstract

DNA sequencing is transforming the field of medical diagnostics and personalized medicine development by providing a pool of genetic information. Recent advancements have propelled solid‐state material‐based sequencing into the forefront as a promising next‐generation sequencing (NGS) technology, offering amplification‐free, cost‐effective, and high‐throughput DNA analysis. Consequently, a comprehensive framework for diverse sequencing methodologies and a cross‐sectional understanding with meticulous documentation of the latest advancements is of timely need. This review explores a broad spectrum of progress and accomplishments in the field of DNA sequencing, focusing mainly on electrical detection methods. The review delves deep into both the theoretical and experimental demonstrations of the ionic blockade and transverse tunneling current methods across a broad range of device architectures, nanopore, nanogap, nanochannel, and hybrid/heterostructures. Additionally, various aspects of each architecture are explored along with their strengths and weaknesses, scrutinizing their potential applications for ultrafast DNA sequencing. Finally, an overview of existing challenges and future directions is provided to expedite the emergence of high‐precision and ultrafast DNA sequencing with ionic and transverse current approaches.
REVIEW
www.small-journal.com
Advancement of Next-Generation DNA Sequencing through
Ionic Blockade and Transverse Tunneling Current Methods
Rameshwar L. Kumawat, Milan Kumar Jena, Sneha Mittal, and Biswarup Pathak*
DNA sequencing is transforming the field of medical diagnostics and
personalized medicine development by providing a pool of genetic
information. Recent advancements have propelled solid-state material-based
sequencing into the forefront as a promising next-generation sequencing
(NGS) technology, offering amplification-free, cost-effective, and
high-throughput DNA analysis. Consequently, a comprehensive framework
for diverse sequencing methodologies and a cross-sectional understanding
with meticulous documentation of the latest advancements is of timely need.
This review explores a broad spectrum of progress and accomplishments in
the field of DNA sequencing, focusing mainly on electrical detection methods.
The review delves deep into both the theoretical and experimental
demonstrations of the ionic blockade and transverse tunneling current
methods across a broad range of device architectures, nanopore, nanogap,
nanochannel, and hybrid/heterostructures. Additionally, various aspects of
each architecture are explored along with their strengths and weaknesses,
scrutinizing their potential applications for ultrafast DNA sequencing. Finally,
an overview of existing challenges and future directions is provided to
expedite the emergence of high-precision and ultrafast DNA sequencing with
ionic and transverse current approaches.
1. Introduction
The genome refers to the complete set of genetic material (DNA
and RNA) present in an organism. It encompasses all the genes,
regulatory elements, and non-coding regions that determine
an individual’s traits, functions, and hereditary information.[1]
These genetic materials carry a broad range of biological infor-
mation and genetic instructions at the molecular level, which
is crucial for the growth, development, and reproduction of a
living being. A DNA molecule consists of four kinds of nu-
cleotides: deoxyadenosine monophosphate (dAMP), deoxythymi-
dine monophosphate (dTMP), deoxyguanosine monophosphate
(dGMP), and deoxycytidine monophosphate (dCMP).[2]Each nu-
cleotide is composed of a nucleoside (nucleobase +deoxyribose
R. L. Kumawat, M. K. Jena, S. Mittal, B. Pathak
Department of Chemistry
Indian Institute of Technology (IIT) Indore
Indore, Madhya Pradesh 453552, India
E-mail: biswarup@iiti.ac.in
The ORCID identification number(s) for the author(s) of this article
can be found under https://doi.org/10.1002/smll.202401112
DOI: 10.1002/smll.202401112
sugar) and a phosphate group. The
dAMP and dGMP are known as purine-
based nucleotides, whereas dTMP and
dCMP are called pyrimidine-based nu-
cleotides. The specific order of these
four nucleotides in DNA essentially
decodes all the biological phenomena.
DNA sequencing is the extraction of
precise and accurate order of nucleotides
in a DNA strand.[3]This sequencing in-
formation is vital for early disease de-
tection and facilitating efficient ways of
maintaining human health. However, the
complexity of genomes, genetic variations,
and large sizes (for example, the hu-
man genome consists of 3.2 billion
nucleotide pairs) are quite challenging
and require capable sequencing methods
to identify each nucleotide accurately.[4]
The evolution of high throughput next-
generation sequencing (NGS) technology
and increasing demand for personalized
medicine development have revolution-
ized the field of genomics and disease
diagnostics.[5–8]The NGS technology can
provide an opportunity to extract genomic
information to prevent, diagnose, and cure human diseases that
would lead medical research and medical care to a new era.
It continuously motivates the scientific and research com-
munity to develop a controlled, rapid, and cheap nanotechnol-
ogy that can sequence the whole human genome accurately.
During the COVID-19 pandemic, NGS sequencing has been a
boon for humanity as it helped in gathering crucial information
about the SARS-CoV-2 virus and its deadly variants, which are
vital for vaccine production and delivering fast and cost-effective
testing methods like real-time reverse transcription-polymerase
chain reaction (RT-PCR).[9–12]However, the NGS method is a
very time-consuming and expensive process. Hence, the devel-
opment of inexpensive and faster DNA sequencing techniques
is an immediate requirement. Currently, scientists are more fo-
cused on electrical detection methods to achieve fast and cost-
effective DNA sequencing with high precision. Figure 1rep-
resents the schematic depiction of the latest electrical DNA
sequencing methods along with their applications in various
fields.
The electrical detection methods address various challenges of
previous traditional sequencing methods, namely, the use of haz-
ardous chemicals, scalability, and technicalcomplexity.[13]Among
the two electrical detection methods, the ionic current method
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (1 of 43)
www.advancedsciencenews.com www.small-journal.com
Figure 1. Schematic illustration of electrical detection methods for DNA sequencing and related applications in the fields of biology, agriculture, and
health science. The electrical detection methods are mainly divided into two categories: ionic current and transverse tunneling current methods. The
ionic current method is based on the blockade of longitudinal ionic current as DNA passes through the nanopore. In contrast, the transverse tunneling
current method detects the variations in the transverse electrical signals resulting from interactions between the nucleotides and nanoelectrodes. The
ionic current method is broadly divided into three parts: biological nanopore, solid-state nanopore, and hybrid nanopore and heterostructure. Similarly,
the transverse current method is also divided into four parts: nanopore, nanogap, nanochannel, and heterostructure.
with biological nanopores has been extensively explored and
successfully commercialized with industrial scalability.[14–16]
However, the lack of long readability inhibits its routine ap-
plication in healthcare and medical science. To address the
challenges of the ionic current technique, a complemen-
tary approach, known as the transverse tunneling current
method, is proposed, which detects the specific molecular fin-
gerprints of each translocating nucleotide through variations in
the transverse tunneling current. It considers solid-state two-
dimensional (2D) nanomaterials with four types of potential
DNA sequencing devices: nanopore, nanogap, nanochannel, and
heterostructures.
While numerous reviews have delved into DNA sequencing,
there remains a gap in systematically documenting the breadth
of developments and breakthroughs in this field, encompass-
ing both experimental and theoretical reports.[17–26]This review
seeks to address this by providing a comprehensive overview
of the evolution and advancements in single-molecule DNA
sequencing, specifically focusing on electrical detection meth-
ods. Our review aims to encapsulate the historical progres-
sion, various developmental stages, and potential challenges en-
countered in DNA sequencing methodologies. Given the sig-
nificant strides made and the promising applications of elec-
trical detection methods utilizing solid-state materials, our fo-
cus lies in elucidating the key milestones in this area. We
underscore the potential of the transverse tunneling current
approach, particularly with solid-state nanodevices, as an ad-
vanced NGS method. This method holds promise in overcom-
ing challenges associated with existing techniques, particularly
the ionic current method. By consolidating the advancements,
challenges, and future prospects in single-molecule DNA se-
quencing with electrical detection methods, this review en-
deavors to provide valuable insights for researchers in the
field.
2. History
Since the pioneering discovery of three-dimensional double helix
DNA by Watson and Crick in 1953, there has been a rapidly grow-
ing interest in developing a potential DNA sequencing technique
capable of decoding all the biological phenomena occurring in
the human body at the molecular level.[2]The advancement in
DNA sequencing techniques has a diverse and rich history with
several paradigms. On the basis of different principles, DNA
sequencing methods are mainly divided into four categories:
first-generation (basic), second-generation, third-generation, and
next-generation sequencing.[27–29]Below, we provide a brief sum-
mary of key milestones of DNA sequencing technologies.
In 1965, for the very first time, Holley et al. isolated the com-
plete nucleotide sequence of transfer RNA from yeast. That was
the first nucleic acid for which the structure was known.[30]In
the same year, Sanger et al. determined the related sequence
of nucleotide by partial digestion.[31]Later, Wu et al. intro-
duced a primer extension method to determine twelve bases
of bacteriophage lambda.[32]Maxam and Gilbert reported a
method to sequence RNA transcription copies of 24 base pairs
of lactose repressor binding sites.[33]In 1975, Sanger and co-
workers developed a rapid and straightforward chain termina-
tion method for the determination of the sequence of single-
stranded DNA by Escherichia coli DNA polymerase I and DNA
polymerase from bacteriophage T4.[3]In an another study, they
reported a similar strategy based on chain termination using
2,3-dideoxy, and arabinose nucleoside analogs of the normal de-
oxynucleotide triphosphates, acting as specific chain-terminating
inhibitors of DNA polymerase.[34]Further, a sequencing method
based on nucleobase-specific chemical modification and subse-
quent termination has been demonstrated that partially breaks
a terminally labeled DNA molecule at each base repetition.[35]
First-generation DNA sequencing is mainly about these two
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (2 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
above-discussed methods known as Sanger’s chain-termination
method and the Maxam–Gilbert chemical cleavage method. One
common thing in both these methods is that they both used the
distances of a radioactive label along a DNA molecule to deter-
mine the nucleotide order at each base position. At that time,
these two methods were instantaneously used for sequencing.
However, first-generation sequencing methods suffer from sev-
eral key challenges from the perspective of cost, speed, and in-
dustrial scalability.[13,27]
After that, a massive improvement in advanced DNA se-
quencing techniques came into existence based on the Shotgun
method, which utilizes the libraries of cloned, randomly frag-
mented DNA.[36]This method was used in complete sequencing
of DNA strands with 4257 base pairs (bp). Here, the advantage is
that, unlike early sequencing methods, this method was capable
of analyzing longer fragments. Such generation DNA sequencing
techniques also include Illumina sequencing and pyrosequenc-
ing methods. The former method is based on reversible-dye ter-
minators, which enable the identification of single nucleobases
having different fluorescent properties,[37]and later is based on
the sequencing by synthesis principle, including the detection of
nucleotide incorporated by DNA polymerase.[38]Such sequenc-
ing methods are used for single-molecule resolution without
the necessity for amplification or labeling. Both these methods
are based on nanotechnology approaches that lead to low-cost
and procedure simplification without the necessity for ampli-
fication or labeling. In 2005, Roche 454 GS20 technology de-
veloped a second-generation sequencing method that can se-
quence the DNA fragments equivalent to 1 billion nucleobases
in a single day (that is 1/3 of the human genome) and is much
cheaper than the methods incorporated in the “Human Genome
Project”.[39–41]This method was based on in vitro amplification of
DNA strands and sequencing via synthesis of micro-beads, DNA
clusters, and DNA nanoballs. The above-discussed Shotgun, Il-
lumina, and pyrosequencing methods fall under the umbrella of
second-generation sequencing or massively parallel sequencing
(MSP) technology.
Third-generation sequencing techniques include stepwise
and real-time single-molecule sequencing by synthesis, Raman
scattering-based sequencing, molecular force microscopy, elec-
tron microscopy, single-molecule motion sequencing, and so
on.[42–44]Several brands like Illumina, Ion Torrent (Thermo Fis-
cher Scientific), Beijing Genomics Institute (BGI), pacBio, and
Oxford Nanopore Technologies are among the top DNA se-
quencing companies that build the ‘third-generation’ sequenc-
ing technologies with the promise to deliver human genomes
sequence in three minutes.[45]However, one of the third-
generation sequencing machines made by “Helicos Biosciences
of Cambridge,” Massachusetts, has been plagued by sequencing
errors.
All the recently developed sequencing methods, including
ionic current and transverse current methods, fall into the cat-
egory of next-generation sequencing (NGS) techniques. The
nanopore sequencing utilizing the ionic current method is cur-
rently a gold standard as it is trusted to provide reliable, cost-
effective, and high-throughput DNA sequencing.[20,21]There
are several companies, such as Oxford Nanopore Technolo-
gies, Electronic BioSciences, Genia, NABsys, Life Sequencing,
and Roche and IBM, actively involved in nanopore sequencing
techniques.[20]An overview of the development of DNA sequenc-
ing techniques till 2021 has been shown in Figure 2.
It is also equally important to understand how the shift in
the different paradigms of the development of DNA sequenc-
ing techniques affects the cost. In the very beginning, the cost
of whole human genome sequencing was $100 million. The
US government launched the “Human Genome Project” in 1990,
led by international scientific research with the goal of determin-
ing all 3.2 billion sequences of nucleotides building the human
genome. The project was announced to be completed on 14th
April 2003.[39,40]In Figure 3, we show the reduction of sequenc-
ing cost as a measure of sequencing technology progress with
time and compared with the curve of Moore’s Law, which signi-
fies a long-term trend of doubling transistor number every two
years.[46]
Undoubtedly, Moore’s law is specifically applied to the comput-
ing and semiconductor hardware industry, which predicts that
the computing power and number of transistors will double ev-
ery 2 years and that the cost will be halved. DNA sequencing
technology evolved at a similar pace and pattern until 2007. How-
ever, thanks to the rapid development of new NGS technologies,
a stiff reduction in the cost is observed from 2008 in both prices
per mega base of DNA sequence and total cost per genome.
There is a tremendous reduction in the cost of DNA sequenc-
ing from $100 million in 2001 to $1000 in 2021. Despite all
these achievements, the adoption of these technologies into clin-
ical practice has yet to be achieved. Scientists are hopeful that in
the near future, we can perform a whole-genome sequence in a
desktop sequencer. For a better understanding of advancement
in DNA sequencing technology, the strengths and weaknesses of
each first, second, third, and next-generation DNA sequencing
are tabulated in Table 1.
3. Electrical Detection Techniques
The electrical detection techniques involve the identification of
DNA nucleotides by analyzing the variations in their correspond-
ing electrical signals, also known as “electrical fingerprints,”
which are measured using sequencing devices.[47–49]The basic
principle is that when single-stranded DNA (ssDNA) translocates
through the sequencing device, it leads to a change in the electric
current signals. Electrical detection techniques are broadly clas-
sified into two categories: ionic current and transverse tunneling
current methods. The ionic current method detects the blockade
of the longitudinal ionic current flux. The different geometrical
shapes and sizes of each nucleotide alter the depth and width of
the ionic current read-outs differently, and on that basis, the iden-
tification of each nucleotide is made possible. Figure 4ashows the
schematic of the ionic blockade current method. The nucleotide
sequence is determined by monitoring the distinctive fluctua-
tions (both magnitude and duration) of ionic current when a
ssDNA passes through a nanopore. Nanopore devices consid-
ered for ionic current-based DNA sequencing can be divided into
three parts: biological, solid-state, and hybrid/heterostructure
nanopores, which have been discussed in section 3.1.
On the other hand, the transverse current method deter-
mines the modulation in the transverse conductance and cur-
rent signals while ssDNA is translocating through the device.
The variations in the tunneling current signals depend on the
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (3 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
Figure 2. An overview of advancement in DNA sequencing technology with the timeline. Here, we have highlighted all the notable and revolutionary
events in DNA sequencing techniques reported so far.
Figure 3. A graphical representation illustrates the decrease in the cost of whole-genome sequencing vs. time with the development of different se-
quencing technologies. Copyright 2021 National Human Genome Research Institute.
Figure 4. Schematics of electrical detection techniques: a) the ionic blockade current method-based nanopore that detects changes in the magnitude and
duration of blockade current signals as single-stranded DNA (ssDNA) molecules traverse through a nanopore. The transverse current method measures
the change in the transverse current signals of a trapped single nucleotide as a result of interactions between the nucleotide and the nanoelectrodes in
b) nanogap, c) nanopore, and d) nanochannel devices.
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (4 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
Tabl e 1 . Comparison of each first-generation, second-generation, third-generation, and next-generation DNA sequencing.
Parameters First generation Second generation Third generation Next-generation sequencing
Development Developed by Frederick Sanger in the
1970s
Introduced in the mid-2000s Introduced in the late 2000s to
early 2010
Introduced in the mid 2015
Working Principle Utilizes chain termination method Utilizes massively parallel sequencing
techniques
Sequencing is performed directly
on single DNA molecules
without amplification
Sequencing through electrical
detection techniques, including
ionic current and transverse
current methods
Read Length Relatively long reads (less than 1kb) Short reads (0.075–0.15 kb) Longer reads (10 20 kb) from short to ultra-long (longest >
4Mb)
Throughput Rate per flow cell Low throughput Low to Moderate throughput
(16–30 Gb)
Moderate to high throughput
(30 Gb)
High throughput up to 48 Gb
Data Analysis Complexity Moderate High Moderate to high Moderate to high
Cost˜ ˜13000 USD 50–63 USD 43–86 USD 21–42 USD
Speed Relatively slow Faster than first-generation Real-time or near real-time
sequencing
Real-time sequencing
Sample Preparation Complex sample preparation Less complex sample preparation
compared to first-generation
Reduced sample preparation
requirements
Simplified sample preparation
Platforms Sanger sequencing (capillary
electrophoresis)
Illumina, Ion Torrent, PacBio, etc. PacBio SMRT sequencing, etc. Oxford Nanopore sequencing
platforms (e.g., MinION)
Instrument Size and
Portability
Relatively large and less portable Varies offer compact, portable devices
suitable for fieldwork and
point-of-care applications
Compact, portable device suitable
for various applications (<450
g)
Base Modification Detection Not applicable Limited Capable of detecting DNA base
modifications during sequencing
Capable of detecting DNA base
modifications during sequencing
Accuracy (%) High accuracy, suitable for precise base
calling (>99.9%)
error rates can vary between platforms
and may require sophisticated
bioinformatics tools for error
correction
(>99.9%)
higher error rates compared to first
and second-generation
sequencing, particularly for raw
reads (>99%)
Variable accuracy, affected by
various factors such as
base-calling algorithms and
signal quality
Applications Targeted sequencing, small genomes whole-genome sequencing,
transcriptomics, epigenomics, and
metagenomics
sequencing various-omics Complex genomic regions,
structural variations, phasing,
real-time applications
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (5 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
chemical and electronic properties of nucleotides. Based on the
sequencing device architectures, this method can be broadly sub-
divided into three categories: nanogap, nanopore, and nanochan-
nel. Figure 4b,c shows a schematic illustration of nanogap and
nanopore devices, respectively. In a nanogap, there exists a gap
between the two device electrodes, whereas, in the nanopore, a
nanometer size pore is created over the nanomaterial membrane
through which a ssDNA can be translocated. The nanochannel
device is a pristine nanomaterial surface over which a ssDNA
strand can be driven by an external electric field, as shown in
Figure 4d. The variation of the interactions between the aromatic
groups of nucleotides with pristine nanochannel surface modu-
lates the electric current and conductance signals of the device,
which helps in the identification of each DNA nucleotide. Apart
from these architectures, heterostructures (with two or three
solid-state materials) have also emerged as promising devices for
DNA sequencing. Further discussion of transverse tunneling cur-
rent techniques utilizing nanopore, nanogap, nanochannel, and
heterostructure devices has been provided in Section 3.2.
3.1. Ionic Current Blockade Technique
As mentioned above, DNA sequencing can be realized by mea-
suring ionic blockade current signals. In this technique, when
a longitudinal external voltage is applied perpendicular to the
nanopore, an ionic current is generated across the nanopore. By
application of an electric field, the charged ions of the salt solu-
tion migrate in opposite directions through the nanopore, cre-
ating an ionic current. However, when a DNA strand is placed
in the cis chamber of the device, the negative charge drives it
in the head-to-tail fashion. The presence of DNA nucleotide in-
side the nanopore exhibits a sudden drop in the ionic current
signals known as ionic blockade current. The volume of each nu-
cleotide inside the nanopore determines the blockade height of
the ionic current signal. The temporal decrease time of the ionic
current depends on the interaction strength of the DNA molecule
with the nanopore, also called the translocation/dwell time of the
DNA molecule. The nucleotide sequence of the DNA strand can
be calculated by statistically analyzing the duration and height of
the ionic blockade current signals.
Among the developed NGS techniques, nanopores are the
most explored architectures for DNA sequencing, a comprehen-
sive discussion of which is provided below.
3.1.1. Biological Nanopores
In nature, the biological cell controls the movement of ions and
molecules into and outside of the cell membrane. Biological
nanopore sequencing mostly depends on the protein membrane,
which is coined as “porins.” Porins are embedded in the lipid
membranes and liposomes to create size-dependent porous sur-
faces with nanometer-scale “pores” distributed across the mem-
branes. Biological nanopore technology delivers high-resolution-
based sequencing with transmembrane protein channels. Suf-
ficiently low translocation velocity can be attained through the
incorporation of various proteins that facilitate the movement
of DNA through the pores of the lipid membranes.[50,51]Ma-
jor advantages include ease in reproducibility and modification
that can be done at pace with modern molecular biology tech-
niques. Among biological nanopores, 𝛼-Hemolysin (HL), My-
cobacterium smegmatis (MspA), and Bacteriophage Phi29 are
among the most exclusively studied nanopores, as shown in
Figure 5.
𝛼-Hemolysin (HL) Pore: In 1996, for the very first time, the
capture and translocation of ssDNA and RNA were demonstrated
through an 𝛼-HL biological nanopore (Figure 5a).[52,53]When the
oligomers of polyuridylic acid translocate through such a chan-
nel/pore, a distinct ionic current signal is observed. Such an
ionic current signal varies as a function of nucleotide length.
The polynucleotides cause short-lived ionic current blockade sig-
nals. In principle, based on such ionic current blockade signals,
DNA/RNA sequence can be obtained. The pyrimidine bases (C
and T for DNA or U for RNA) provide light signals compared to
purine bases (A and G). Later, Akeson and co-workers confirmed
the same by studying the signal differences of DNA and RNA
polymerase translocation through 𝛼-HL nanopore device.[54]For
instance, in the presence of purine polyadenylic acid molecules,
there is less blockage in ionic current compared to that of poly-
cytidylic acid and polyuridylic acid. The secondary structure of a
homopolymer is accountable for the observed ionic current sig-
nals, not its primary structure. Because polyadenylic acid forms
a bulkier secondary structure, and the structure is unstable in-
side the nanopore, and therefore, it must disassemble. However,
polycytidylic acid and polyuridylic acid have narrower secondary
structures, and thus they block more ions than polyadenylic acid.
Meller et al. demonstrated the translocation speed of ssDNA
based on the ionic blockade current and time distributions.[57]
The translocation speed varies non-linearly with the applied elec-
tric field. These studies were first theoretically supported by in-
troducing a statistical physical model of polymer translocation
through a pore in the membrane.[58]In an 𝛼-HL biological pore,
there are three recognition sites for distinguishing DNA nucle-
obases, and to improve the base identification ability, it is manda-
tory to enhance recognition at one or two points.[59,60]It has been
reported that amino acid side chain-assisted R1 site-directed mu-
tagenesis is helpful in improving base recognition, as shown in
Figure 5b.[55]The 𝛼-HL biological pore is sufficiently stable to se-
quence genomic dsDNA at pH (<11.7), a value at which dsDNA is
denatured.[61]One of the studies on 𝛼-HL biological pores reveals
the dynamics of ssDNA in a protein pore at a single molecule
level.[62]Two different time-scales have been noticed, and a longer
timescale has been found to be associated with the binding and
unbinding of DNA from the pore.
𝛼-HL combined with molecule adapter is reported to increase
the recognition ability of nanopore with accuracy averaging
99.8%.[63,64]Later, the translocation of ssDNA and ssRNA with
secondary structure was achieved through 𝛼-HL pore by adding
urea.[65]Nanopore force microscopy has been demonstrated to
identify oligonucleotides of equal length and sequence differed
by a single nucleotide.[66]In a recent report, the effect of molec-
ular crowding on the translocation behavior of DNA through 𝛼-
HL pore is studied as given in Figure 5c.[56]It has been noticed
that translocation event frequency has a nonmonotonic depen-
dence on the size of the crowding agent. However, both event
frequency and translocation time increase monotonically with in-
creasing crowder concentration. 𝛼-HL has shown great potential
in stochastic sensing of various analytes, including small organic
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (6 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
Figure 5. Schematic of 𝛼-HL biological nanopore utilizing ionic current. a) 𝛼-HL biological nanopore and ionic current blockade signals caused by
oligomers of polyuridylic acid;[52]Reproduced with permission.[52 ]Copyright 1996 Proceedings of the National Academy of Sciences, b) schematic
of R1 site-directed mutagenesis in 𝛼-HL biological nanopore;[55]Reproduced with permission.[55 ]Copyright 2010 American Chemical Society, and c)
schematic illustrating the effect of crowding agent on translocation event of DNA translocation through 𝛼-HL biological nanopore;[56]Reproduced with
permission.[56]Copyright 2020 American Chemical Society.
molecules, metal ions, RNA, DNA, and proteins, among others.
However, 𝛼-HL suffers from structural limitations such as a lim-
ited pore size of 1.4 nm and 𝛽-barrel, which takes a very long
to identify individual nucleotides.[67]Hence, to improve the sen-
sitivity and exploit the application of polynucleotide detection,
some other biological pores have also been explored.
MspA Channel Nanopore: Mycobacterium smegmatis porin
A (MspA) is a potential alternative to 𝛼-HL pore, which has gen-
erated a great deal of interest from both theoretical and experi-
mental aspects. It is more robust under experimental conditions,
varying pH from 0 to 14, and under high temperatures, it keeps
the channel active.[68]MspA is a funnel-shaped octameric chan-
nel nanopore that was first prepared by the E. coil prokaryotic
expression system, which allows the transport of water-soluble
biomolecules across bacterial cell membranes.[69]This finding
has enabled all the follow-up reports with MspA, including the
site-specific mutagenesis that has enabled DNA nucleotide dif-
ferentiation using the MspA pore (Figure 6a).[70]
The authors demonstrated the protein pore MspA has great
potential in advancing next-generation DNA sequencing with im-
proved spatial resolution. They considered hairpin duplex DNA
for their study. During translocation, the sole presence of the
hairpin tail inside the pore measured the residual ionic current,
which majorly originates from the size and composition of the
ssDNA confined inside the pore. Just after 10 ms on applying
the driving voltage of 180 mV, the hairpin duplex started dissoci-
ating to facilitate the translocation event. The translocation rate
of DNA through the MspA pore is too high to clearly identify a
single nucleotide without error. For that, DNA-cutting enzymes
were added to the system, which further enhanced the complex-
ity of DNA sequencing experiments. A mutated form of MspA
with phi29 DNA polymerase has also been reported to control
the translocation speed of the pore.[74]Here, phi29 DNA poly-
merase synthesizes DNA and acts as a motor to pull the single-
stranded template. Another report on MspA nanopore shows that
it can sequence the phi X 174 genome precisely up to 4500 nu-
cleobases in length.[75]It was the first nanopore quadromer map
where ionic current signals of all possible 256 quadromers of four
DNA nucleotides were determined, enabling the sequencing of
long, complex, natural DNA strands.
Apart from this, several theoretical studies have also been re-
ported on polynucleotide detection through the MspA nanopore.
A report based on an MD simulation study reveals that sev-
eral arginine mutations can lead to 10–30-fold reduction in
the translocation speed of DNA through the MspA pore that too
without eliminating the nucleotide-induced ionic current block-
ade (Figure 6b).[71]Based on this, Bhattacharya et al. have re-
ported a study to determine the physical mechanism enabling
ionic current blockade in the MspA pore.[76]The mechanism re-
veals that the displacement of water molecules inside the MspA
nanopore determines the ionic current, while the steric and
base stacking nature of DNA determines the amount of wa-
ter displaced via the steric exclusion effect. To provide more
insight into the experimental phenomenon and other possible
sensing positions of MspA nanopore, the underlying mecha-
nism of current variation for ssDNA translocation through the
MspA nanopore has further been investigated through all-atom
MD simulations (Figure 6c).[72]It has been observed that the
region below constriction is more promising for single nu-
cleotide resolution, while ssDNA translocation is in the base-
constriction-base meshing and ratcheting across the nanopore
constriction.
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (7 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
Figure 6. Schematic of MspA biological nanopore utilizing ionic current. a) DNA translocation through MspA nanopore and resulting ionic current
blockade signals;[70]Reproduced with permission.[70 ]Copyright 2010 Proceedings of the National Academy of Sciences, b) DNA translocation through
MspA nanopore and simulated effect of arginine mutations on DNA translocation rate;[71]Reproduced with permission.[71 ]Copyright 2012 American
Chemical Society, c) schematic illustration of MspA pore with residual current data for each nucleobase;[72]Reproduced with permission.[72]Copyright
2021 American Chemical Society and d) schematic illustration of machine learning integrated MspA pore with residual current data for identification of
natural and chemically modified nucleotides;[73]Reproduced with permission.[73 ]Copyright 2022 American Chemical Society.
Tabatabaei et al. reported that the different combinations
and ordered sequences of natural and chemically modified nu-
cleotides in custom-designed oligomers can be distinguished by
using a MspA nanopore, as shown in Figure 6d.[73]It is ob-
served that MspA pore can also detect 77 combinations and or-
dering of monomers with chemical diversity. The group further
reported a deep-learning neural network to classify combinatorial
patterns of natural and modified nucleotides that operate on raw
current signals generated by GridION of Oxford Nanopore Tech-
nologies. The machine learning technique facilitates denoising,
spike recognition, extracting, and analyzing the complex multidi-
mensional information of electric signals. With recent advance-
ments, the integration of machine learning tools with biologi-
cal nanopores is an emerging area of research that is needed to
achieve reliable and high throughput DNA sequencing. Despite
the significant achievements of MspA nanopore in the field, it
suffers from certain limitations. Here, the major limitation is
its narrow channel that allows only ssDNA and ssRNA chains
to pass through them. Thus, alternative nanopores are required
for sequencing both ssDNA and dsDNA.
Bacteriophage phi29 Nanopore: Bacteriophage Phi29 biologi-
cal nanopore is another potential candidate for sequencing DNA
through ionic current blockade measurements. Here, the advan-
tage is that compared to 𝛼-HL and MspA nanopores, the phi29
pore is more flexible for biochemical modifications and has a
larger diameter, allowing the translocation of molecules larger in
size, such as dsDNA, DNA complexes, and proteins.[67]In 2009,
for the very first time, the channel of the phi29 connector mo-
tor protein was implanted into lipid bilayers inside a microflu-
idic channel array for DNA sequencing purposes, as shown in
Figure 7a.[77]The bacteriophage phi29 DNA-packaging motor
acts as a path for the translocation of double-stranded DNA. The
length of the connector is 7 nm, with a cross-sectional area of
the channel of 10 nm2(3.6 nm in diameter) at the narrow end
and 28 nm2(6 nm in diameter) at its wider end.[78]Phi29 DNA-
packaging motor has the largest diameter, facilitating the translo-
cation of dsDNA, but cannot identify ssDNA under experimental
conditions.
According to an experimental report, channel-size conversion
of the phi29 motor can facilitate the translocation of both ssDNA
and dsDNA (Figure 7b).[79]Reengineering of the phi29 motor
channel is reported to result in two classes: one with the same
size as the wild-type channel and the other with a cross-sectional
area that is 60% of the wild-type. The smaller channel can de-
tect ssDNA at the single-nucleotide level, the wild-type connector
channel allows dsDNA translocation, and the loop-deleted con-
nector is reported to translocate both ssDNA and ssRNA with
equal competencies from both terminals.
CsgG Pore: In 2016, for the very first time, Oxford Nanopore
Technologies (ONT) reported the Escherichia coli curli trans-
port channel CsgG nanopore for sequencing, which utilizes he-
licases as the motor enzyme to control the translocation rate of
DNA.[82]The CsgG nanopore has a well-defined 1 nm wide con-
striction centrally located in a 4 nm wide channel.[83,84]Mutants
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (8 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
Figure 7. Schematic of bacteriophage phi29 biological nanopore utilizing ionic current. a) Schematic of phi29 DNA packaging motor connector
channel;[77]Reproduced with permission.[77 ]Copyright 2009 Nature Publishing Group, b) illustration of phi29 DNA packaging motor connector channel
and recorded current data;[79]Reproduced with permission.[79 ]Copyright 2013 American Chemical Society, c) schematic illustration of aerolysin biolog-
ical nanopore with contour plots and total internal reflection fluorescence (TIRF) images;[80]Reproduced with permission.[80 ]Copyright 2016 American
Chemical Society, and d) Alpha-hederin nanopore and corresponding normalized histograms of conductance blockade for homopolymers and each
single DNA nucleotide;[81]Reproduced with permission.[81 ]Copyright 2019 American Chemical Society.
of CsgG mutants present several benefits, including enhanced
interactions with analytes, an extended range of electrical
currents, and improved precision in sequencing modified
nucleobase.[85]The nanopore is commercially available for nucle-
obase modification sequencing. Later, the prototype CsgG–CsgF
protein pore is reported for improved signal and base calling ac-
curacy (25–70%) of DNA molecules for homopolymer region
due to the double constriction effect.[86]
Other than that, aerolysin pore (diameter ranges from 1.0 to
1.7 nm) has been reported as a prospective candidate to de-
tect nucleic acid at a single nucleotide level, as can be seen in
Figure 7c.[80]Aerolysin nanopore has several advantages over
𝛼-HL pore, such as high spatial resolution, slow translocation
speed, high current sensitivity, and prolonged duration time for
polynucleotide detection. It is demonstrated that the detection of
unlabeled random ssDNA is feasible through wild-type aerolysin
pore with high temporal resolution and current. The transloca-
tion of ssDNA through the pore is confirmed by total internal
reflection fluorescence (TIRF). Besides, alpha-hederin nanopore
is also reported for single nucleotide identification, as shown in
Figure 7d.[81]Here, the advantage is its intrinsic pore-forming
properties in the cellular membrane. Spontaneous formation of
alpha-hederin nanopore in a lipid membrane with a determin-
istic lipid composition facilitates the identification of single nu-
cleotides with high spatial and temporal resolution. Owing to
the small diameter and adequate thickness of the alpha-hederin
nanopore, identification of ssDNA homopolymer and each DNA
nucleotide is found to be feasible.
Undoubtedly, biological nanopores offer several advantages
for single molecule-based DNA sequencing. These advantages
include biocompatibility of set-up, easy and large production
of nanopores with atomic-level precision, uniform nanopore
structure, and easy tailoring of physical and chemical proper-
ties through genetic engineering techniques. With recent ad-
vancements, biological nanopores utilizing the ionic current
method have now been commercialized. The MinION nanopore
sequencer based on biological pores has finished sequencing a
human genome.[16]However, despite high sensitivity and easy
production of biological nanopores, there are certain inarguable
intrinsic disadvantages: i) low-stress durability, ii) strict tempera-
ture, pH, electrolyte concentration, and controlled environmen-
tal conditions to keep up biological pore active, iii) easy break-
down at high voltages, iv) incompatibility with the semiconduc-
tor fabrication process, and v) sensitivity to ambient conditions
and pore functionalization. In vitro studies reveal that the speed
of DNA through the nanopores is quite high (1 bp/1–10 μs),
which means very few ions are present inside the nanopore while
the nucleotide is passing. The challenges in biological nanopores
motivated scientists to search for new types of materials for
high-throughput DNA sequencing. In the next section, we have
discussed the potential usages of solid-state-based robust and
durable materials for the sequencing of canonical nucleobases.
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (9 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
Figure 8. Solid-state Si3N4and graphene nanopore-based DNA sequencing with the ionic current approach. a) Schematic of the experimental setup
with Si3N4pore;[91]Reproduced with permission.[91 ]Copyright 2003 Nature Publishing Group, b) schematic of graphene nanopore with longitudinal
ionic current blockade method and current trace signals of DNA translocation through graphene nanopore at unfolded, partially folded, and fully folded
stage;[93]Reproduced with permission.[93 ]Copyright 2010 American Chemical Society, c) ionic current density distribution around the small and large
graphene nanopore;[94]Reproduced with permission.[94 ]Copyright 2013 Proceedings of the National Academy of Sciences, and d) open pore currents
through a series of graphene nanopores with different diameters;[94]Reproduced with permission.[94 ]Copyright 2013 Proceedings of the National
Academy of Sciences.
3.1.2. Solid-State Nanopores
Shortcomings of the biological nanopores motivated researchers
to explore the applicability of solid-state materials toward the de-
velopment of advanced nanopore devices for DNA sequencing.
With the advancement of nanofabrication technologies, the solid-
state nanopore has developed as a potential alternative to the bi-
ological nanopore. Solid-state materials offer a range of advan-
tages, including cost-effectiveness, robustness, and high durabil-
ity. They also possess superior mechanical, chemical, and ther-
mal properties.[87–89]In 2001, Li et al., for the very first time, fabri-
cated the silicon nitride (Si3N4) solid-state nanopore as an alterna-
tive to biological nanopore by drilling with the focused ion-beam
sculpting technique.[90]This fabrication of Si3N4nanopore inves-
tigated the measurement of ionic current blockage measurement
and monitored the translocation events of DNA molecules, as
shown in Figure 8a.[91]The Si3N4nanopore membrane is sep-
arated by two chambers that conduct an electrolyte solution. The
authors demonstrated that the Si3N4nanopore is capable of ob-
serving individual molecules of dsDNA and their folding behav-
ior. In 2004, Chen et al. reported that the fabrication of an atomic
layer of Al2O3on Si3N4nanopore modifies surface properties,
which leads to a reduction in the electrical noise and enhances
the DNA capture rate.[92]
Heng et al. demonstrated the relationship between DNA
translocation and blocking current and reported that the mag-
nitude and duration of ionic blockade current can lead to the
identification of DNA nucleotides.[95]In 2009, Venkatesan et al.
demonstrated a mechanically robust and highly sensitive Al2O3
nanopore for DNA analysis purposes.[88]The major advantages of
this nanopore are easy pore size control, chemical modification,
and improved signal performance. Subsequently, the solid-state
Al2O3nanopore sensor with enhanced surface properties is stud-
ied for real-time analysis and identification of DNA molecules.[96]
Owing to high surface-charge density and enhanced surface
properties, an order of magnitude reduction in the transloca-
tion speed is observed relative to other solid-state Si3N4and
SiO2architectures. Later, they studied the Al2O3coated hybrid
nanopore sensor.[97]A lipid bilayer coated with Al2O3shows ex-
cellent electrical properties and enhanced mechanical stability.
Moreover, compared to black lipid membranes, Al2O3-coated hy-
brid nanopores are found to be more stable in ionic solution with
a fivefold increased lifetime.
Besides silicon nitride nanopores, the DNA translocation
events through ‘kinked’ silica nanopores (SiO2) have also been
demonstrated with nearly perfect selectivity for ssDNA over
dsDNA.[98]By application of a gate electric field, a reduction in the
DNA translocation speed at a rate of 55 μms
1per 1 mV nm1
is observed.[99]Even though the solid-state nanopores offered
several advantages and achieved a similar nanopore dimen-
sion to the biological nanopore, their higher thickness prevents
achieving the required high-precision identification of DNA
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (10 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
nucleotides. The thickness of silicon nitride is usually higher
than 10 nm, which is equivalent to 30 DNA molecules, a
big challenge for silicon nitride nanopores to achieve single nu-
cleotide identification. The challenges, such as incapability to sin-
gle nucleobase resolution, low signal-to-noise ratio, and weaker
capacitive coupling, have pushed researchers to explore other 2D
solid-state nanomaterials for DNA sequencing.
Graphene is regarded as a highly promising substrate for
solid-state pores due to its membrane thickness of 0.34 nm,
which is equivalent to the height of a nucleotide. This fea-
ture enables the possibility of discriminating between individ-
ual nucleotides.[100,101]In 2010, three different groups, Merchant
et al., Schneider et al., and Garaj et al., independently reported
the measurement of the fluctuations of ionic current when
a ssDNA translocates through an atomically thick graphene-
based nanopore.[93,102,103]Figure 8b shows the schematic set of
graphene nanopores with a dsDNA translocation through it. In
the absence of DNA, the ionic current signal remains uniform
and continuous. However, a sharp drop in the current signal
(black color signals) can be observed as soon as the dsDNA
starts translocating through the nanopore. Interestingly, the cur-
rent signals of DNA translocation through graphene nanopores
are larger than the earlier reported Si3N4nanopores with the
same diameter due to the atomic thickness of the graphene
membrane.[93]Golovchenko and co-workers demonstrated that
the average translocation time is also slightly larger for graphene
nanopore devices compared to other solid-state nanopores due
to DNA-graphene interactions.[103]Garaj et al. studied the mono-
layer graphene nanopore with a diameter equivalent to the di-
ameter of the dsDNA molecule and observed a highly sensitive
ionic current with a small change in translocating DNA molecule
diameter.[94]Additionally, they found that when the nanopore
length (L) is less than the diameter (D), a strong ionic current
distribution peak is observed near the nanopore perimeter. How-
ever, there is a uniform current distribution around the pore for
LD, as shown in Figure 8c. A linear increase in ionic current
measurement with an increase in the graphene nanopore diam-
eter is observed, as shown in Figure 8d.
Drndi´
c and co-workers coated the graphene monolayers with
TiO2, which acts as an insulating layer and has significantly re-
duced the signal noise.[93]The TiO2layer over the graphene sur-
face provides a hydrophilic surface to the DNA strand for ex-
ploring low entropy dynamics during translocation through the
nanopore. However, coating of the layer increases the graphene
nanopore thickness and decreases the single-base resolution ca-
pability of graphene.
In an MD simulation study with a graphene nanopore device,
Wells et al. demonstrated that the hydrophobic interactions with
the graphene membrane lead to a dramatic reduction in the con-
formational fluctuations of the nucleotides in the nanopore.[104]
The ssDNA was found to adhere to the surface of graphene mem-
branes due to the strong hydrophobic interaction between ssDNA
and graphene. They also demonstrated that the adhesion of ss-
DNA over the multilayer nanopore leads to control of the translo-
cation speed with stepwise translocation in a single-nucleotide
manner. However, Schneider et al. reported that the strong in-
teractions between the graphene and DNA nucleotides are re-
sponsible for the irreversible pore closure.[105]Additionally, the
adhesion of ssDNA over the nanopore surfaces restricts the DNA
conformational fluctuations to a few typical stable configurations
only, depending on the nucleotide type. A recent MD study from
our group shows that the identification of DNA nucleotide is pos-
sible with the graphene nanogap device using an ionic current
approach.[106]The major concern with the nanogap-type setup is
controlling the translocation of ssDNA and the overlap of ionic
current signals due to high current flux through the nanogap.
However, there are certain drawbacks in graphene, such as
strong DNA–graphene hydrophobic interactions, limited per-
meation events of DNA, clogging of the pore, the require-
ment of additional functionalization of graphene surface to
facilitate controlled DNA translocation, and physisorption of
nucleobases on graphene edges (𝜋-𝜋interaction) making the
single-base resolution more difficult. Hence, graphene may not
be the most promising material for ionic current-based DNA
sequencing.
To resolve the above-mentioned issues with graphene
nanopore, other 2D solid-state materials that have promising
characteristics for cheap, reliable, and efficient DNA sequencing
are being explored. Yu and co-workers have reported h-BN
as a competitive candidate for DNA sequencing other than
graphene by the first electronic measurement of DNA transloca-
tion events through atomically thin insulating h-BN nanopores
(Figure 9a,b).[107]DNA translocation signals are observed to be
much larger compared to the similar-sized Si3N4nanopores, in-
dicating higher detection accuracy. The single-atomic thickness
(1.1 nm) of the h-BN membrane has a significant opportu-
nity to achieve high spatial resolution similar to a graphene
nanopore device. However, like graphene, the h-BN surface is
also hydrophobic in nature, which could be a reason for the easy
clogging of the pore in an experimental condition. Later, Zhou
et al. successfully enhanced the hydrophilicity of h-BN nanopores
through UV-ozone treatment.[111]They also observed sufficient
DNA translocation events with no easy pore clogging. In addition
to insulating nature and durability, another important feature is
the possibility of pore creation in the triangular shape, enabling
control over the nanopore geometry. The translocation speed
is found to be faster near the pore center than at the corner of
the triangle. In this regard, the well-designed triangular-shaped
h-BN pore exhibits two different translocation pathways, and the
translocation speed of DNA can be tuned across the membrane.
In 2014, Farimani et al. investigated molybdenum disulfide
(MoS2) nanopore, which is found to be a promising alterna-
tive to graphene due to the comparatively high signal-to-noise
ratio (Figure 9c).[112]The results show that DNA translocation
exhibits signal amplitude five times higher than the solid-state
Si3N4nanopores and improved signal-to-noise ratio (SNR >10).
Moreover, unlike graphene nanopore, no special surface treat-
ment is needed to avoid strong 𝜋𝜋interaction between DNA
and the surface, and the hydrophobic interaction is also less in
the case of MoS2nanopore.[108]However, the challenge of the
MoS2nanopore is that the membrane can suffer from breakdown
under electrical and mechanical stress. To counter the break-
down issue of the MoS2and control the translocation speed of
DNA, Zou et al. reported a study with graphene-MoS2bilayer
heterostructure.[113]Gu et al. reported the effects of KCl salt and
Bmim+ionic liquid on the translocation speed of ssDNA though
the MoS2nanopore. A strong interaction between Bmim+and
ssDNA is noted, which offers a considerable dragging force to
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (11 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
Figure 9. Experimental and theoretical studies of 2D solid-state nanopore-based DNA sequencing with the ionic current approach a) Schematic of the
device setup for DNA translocation through BN nanopore, b) the scattering diagram showing the depth versus length for DNA translocation events
through the BN nanopore;[107]Reproduced with permission.[107 ]Copyright 2013 Wiley Publishing Group, c) schematic illustration of a MoS2nanopore
membrane for DNA translocation, where the monolayer MoS2is suspended on a Si3N4supporting membrane and TEM image of nanopore drilled by
a focused electron beam;[108]Reproduced with permission.[108 ]Copyright 2014 American Chemical Society, d) schematic illustration of dsDNA translo-
cation through a single-layer Ti2CTxnanopore, e) concatenated ionic current spikes shown at different voltages of 250 mV (blue), 225 mV (black), and
200 mV (green);[109]Reproduced with permission.[109 ]Copyright 2019 American Chemical Society, and f) schematics of the MD simulation snapshots
of graphene, MoS2,andTi
3C2MXene nanopore and single-strand DNA passing through graphene nanopore;[110]Reproduced with permission.[110 ]
Copyright 2022 American Chemical Society.
decelerate the electrophoretic motion of ssDNA in the BmimCl
solution.[114]
In 2019, Wanunu and co-workers experimentally studied DNA
transport through the Ti3C2TxMXenes nanopore (Figure 9d).[109]
From Figure 9e, the concatenated negatively pointing spikes of
ionic current signals caused by the temporary position of DNA
molecules at different voltages. The MXene nanopore demon-
strated lower ionic current leakage and similar noise of the mem-
brane compared to other reported 2D solid-state nanomateri-
als. Recently, Farimani and co-workers performed an MD-based
study on the Ti3C2MXene and found that it could be a promising
candidate for a single-nucleotide resolution with a high signal-to-
noise ratio due to its hydrophilic surface, greater electrical con-
ductivity, and stability under electrical forces.[115]
Though a variety of 2D solid-state nanopores (graphene, hexag-
onal boron nitride, MoS2, MXenes) have been extensively studied
for DNA sequencing, each material comes with its own merits
and demerits. In this regard, a comparative study with graphene,
MoS2,andTi
3C2MXene nanopore is demonstrated for DNA se-
quencing with the ionic current approach (Figure 9f).[110]Fro m
the ionic current and residence time results, it is concluded that
graphene nanopore still stands out as a better nanopore device
for the identification of all four DNA nucleotides, whereas MoS2
and Ti3C2MXene nanopores were found to be selective.
Currently, the integration of machine learning techniques with
NGS technologies has emerged as an active area of research to
achieve ultrafast, cheap, and high throughput DNA sequencing.
Machine learning tools resolve several issues associated with the
DNA sequencing process. Differentiating single-stranded DNA
nucleotides is challenging due to the broad distribution of ionic
current signals, compounded by errors from thermal fluctua-
tions, counterions, electric field, and DNA dynamics. These re-
strictions impede the high-precision identification of nucleotide
sequences. The machine learning protocols translate the raw se-
quencing data into base calls by feature extraction and real-time
signal processing of long-read sequencing data.
Fyta and co-workers demonstrated unsupervised learning
studies to classify different molecular topologies from the ionic
current translocation events of MoS2nanopores without the use
of labels (Figure 10a).[116]They proposed a new efficient feature
of ionic blockage height compared to traditional dwell time and
mean ionic current feature and observed that this approach is
able to identify distinct types of the most probable molecular con-
formations of the DNA threading through the MoS2nanopore. In
a follow-up study, by utilizing deep neural networks and convo-
lutional neural networks (CNN) on the experimental ionic block-
ade current data, the authors tried to increase the classifica-
tion accuracy and read-out efficiency of MoS2pores with varying
diameters.[117]The ionic blockade height combined with dwell
time, mean ionic current, and the number of levels helped the
CNN and XGBoost models to classify all four nucleotides with
better accuracy (Figure 10b).
3.1.3. Hybrid Nanopores and Heterostructures
Afterward, hybrid nanodevices came into existence, which is a
coupling of both biological and solid-state entities. Here, the role
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (12 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
Figure 10. Machine learning studies with solid-state nanopore-based ionic current blockade method. a) concatenated events of dAMP translocation
through MoS2nanopore, where the red block highlights each event and the right box shows the extracted four features for a single nucleotide translocation
event;[116]Reproduced with permission.[116 ]Copyright 2019 IOP Science and b) confusion matrix showing the relation of the true identity (label) of a
nucleotide compared to the predicted identity (label) for the LSTM, XGBoost, DNN, and CNN models;[117]Reproduced with permission.[117 ]Copyright
2021 AIP Publishing Group.
of the biological nanopore is to offer an atomically precise struc-
ture capable of genetic engineering, and solid-state nanopore
provides durability, size, and shape control, which is well-suited
for wafer-scale devices. In addition, biological nanopores can
attain control of the DNA translocation time and its orienta-
tion, and solid-state nanopores can provide enhanced robust-
ness and durability. The hybrid nanostructures have the poten-
tial to quickly identify the nucleotides during the DNA sequenc-
ing process.[97,118–123]It has also been noticed that the hybrid
nanopores are easier to fabricate and stable under experimental
conditions, showing low peripheral leakage allowing sensing and
discrimination among DNA nucleobases.
In 2010, Hall et al. first introduced the hybrid pore forma-
tion by directed insertion of hemolysin into Si3N4solid-state
nanopore, as shown in Figure 11a.[123]Upon protein insertion,
the hybrid nanopore is noticed to be functional for 30–40% of
their attempts and, thus, offers a platform to create wafer-scale
device arrays for genomic analysis, including sequencing. In
principle, hybrid nanopores exploit the excellent addressability of
biological macromolecules to provide controllable chemical func-
tionality for solid-state nanopores. It shows both the robustness
of the device and the biological nanopore remaining unaffected
while providing better control of DNA translocation velocity.
According to another experimental report, the hybrid
nanopore formed by the insertion of non-modified 𝛼-hemolysin
at the entrance of biomimetic nanopore allows polynucleotide
distinguishability with increased dwell time.[124]Here, the main
finding is that the insertion of 𝛼-hemolysin can increase the
polynucleotide dwell time without perturbing the current block-
ades and losing the nucleotide’s distinguishing ability of protein
nanopore.
Recently, DNA-based technology is also become quite popu-
lar for functionalizing nanopores.[128–131]The coolest emerged
technique among other techniques and methods is DNA
origami,[132–134]which acts as a toolbox for engineering complex
nanostructures.[135–138]Inspired by the pioneering work of Hall
et al., Keyser and co-workers experimentally fabricated a sin-
gle artificial nanopore based on DNA origami and detected 𝜆-
DNA successfully.[139]Here, the advantage is that DNA origami
nanopores can be continuously inserted and ejected from the
solid-state nanopore and can be used as resistive pulsed sensors.
Later, Farimani et al. performed MD simulations to analyze the
ionic conductivity of DNA origami-graphene hybrid nanopore, as
showninFigure11b.[125 ]In the case of bare graphene, the same
dwell time is noticed for DNA nucleobases, while the use of DNA
origami hybrid nanopore significantly improved the dwell time
of all four nucleobases, which further increased distinguishabil-
ity. Moreover, the speed of DNA translocation is also noted to be
reduced in the presence of DNA origami due to the friction of
nucleobase near the mouth of the graphene pore. To control the
DNA translocation speed, the authors removed the bases from
the origami sheet near the nanopore. However, the effect of the
number and spatial distribution of the baits, also referred to as
overhangs (genomic DNA fragments of interest for sequencing),
was not addressed properly. Later, to encounter that, MD simula-
tions were performed on a hybrid nanopore comprising a single-
layer graphene and DNA origami layer with different numbers,
lengths, and spatial distribution of overhangs, as can be seen in
Figure 11c.[126]It has been noticed that different customizations
give rise to base-specific translocation times.
Cressiot et al. experimentally reported the stable and easy fab-
rication of lipid free-hybrid nanopore comprising a natural DNA
pore and portal protein as adapter inserted into a Si3N4nanopore
for sensing biomolecules (peptide, globular protein, ssDNA, and
dsDNA) as shown in Figure 11d.[127]The interesting thing is
that the reported hybrid pore allowed sensing and distinguisha-
bility among different biomolecules with low peripheral leak-
age. The work further motivates the researchers to find out what
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (13 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
Figure 11. Schematic representation of hybrid nanopores-based DNA sequencing utilizing ionic current approach. a) Schematic illustrating the translo-
cation of DNA through a hybrid 𝛼-hemolysin nanopore;[123]Reproduced with permission.[123 ]Copyright 2010 Nature Publishing Group, b) DNA origami-
graphene hybrid nanopore and the plot showing translocation rate through both bare and hybrid nanopore;[125]Reproduced with permission.[125 ]Copy-
right 2017 American Chemical Society, c) schematic of translocation of DNA through a hybrid DNA origami-graphene nanopore with ions and current-
voltage characteristics;[126]Reproduced with permission.[126 ]Copyright 2019 American Chemical Society, and d) schematic of bio-inspired lipid-free
hybrid nanopore;[127]Reproduced with permission.[127 ]Copyright 2018 Nature Publishing Group.
properties of 𝛼-hemolysin are retained by the lipid-free hy-
brid pore. The theoretical assessment of polynucleotide dif-
ferentiation based on a lipid-free hybrid nanopore functional-
ized with 𝛼-hemolysin promotes the stability of protein inside
the biomimetic nanopore and its potential for sequencing the
biomolecules.[140]
Apart from hybrid nanopore, heterostructures (with two
or three solid-state materials) have also been widely stud-
ied through the ionic current method. In a framework
of MD simulations, Luan and co-workers first studied the
stacked graphene/MoS2heterostructure nanopore, as shown in
Figure 12a.[141]The proper attraction between the negatively
charged phosphate group of DNA nucleotide and positively
charged Mo atoms provides a slow nucleotide-by-nucleotide
transport, which is crucial for precise and accurate sequencing
of DNA. Another report studied the geometry and shape effects
of stacked graphene/MoS2heterostructure nanopore in a theoret-
ical framework of MD simulations, as shown in Figure 12b.[113]
The results show that among the considered circular, square, and
triangular nanopores, the circular nanopore facilitates the fastest
translocation, and triangular nanopores show the slowest translo-
cation of ssDNA. Here, the difference in the translocation speed
arises from the difference in the electrostatic attraction. In the
different geometries of heterostructure nanopore, the number of
exposed Mo atoms is different, which means the electrostatic at-
tractions between the positively charged Mo atoms and the nega-
tively charged phosphate group of DNA nucleotides are different,
which significantly affects the translocation time.
Hexagonal boron nitride (h-BN) has a similar atomic struc-
ture and has been extensively studied for DNA sequencing from
both theoretical and experimental aspects.[144,145]These findings
motivated the researchers to use graphene/h-BN heterostruc-
tures for DNA sequencing. Luan et al. further checked the po-
tential of an in-plane graphene/h-BN heterostructure nanopore
for sensing ssDNA, as given in Figure 12c.[142]Both the DFT
and MD calculations revealed that due to the stronger van
der Waals attraction, ssDNA prefers to stay in the h-BN do-
main than graphene. This change in the conformation of ss-
DNA on the graphene/h-BN surface enables the sensing of
DNA through scanning tunneling microscopy (STM) and atomic
force microscopy (AFM) in the dimension perpendicular to
the heterostructure surface. Another theoretical report studied
the translocation of dsDNA through graphene/h-BN nanopore
and found that graphene/h-BN layer order is of crucial im-
portance for nucleobase-specific electrical signal variability.[146]
Here, the interesting observation is that h-BN allows the translo-
cation of dsDNA parallel to the electric field, and this vertical
configuration is the key to signal readability and single-base
sensing.
Another theoretically studied heterostructure is BC3/C3N
nanopore, where the authors reported spontaneous translocation
(without the need for an electric field) of DNA from BC3to the
C3N side with a strong affinity of C3N towards DNA bases in
comparison to the BC3surface, as given in Figure 12d.[143]In
the case of BC3/C3N nanopore, the translocation speed of ss-
DNA is noticed to be 2-fold slower than that of graphene/MoS2
heterostructure nanopore. For a better understanding, year-
wise documentation of ionic current studies reported on hy-
brid/heterostructure nanopores for DNA Sequencing is given in
Table 2.
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (14 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
Figure 12. Schematic representation of solid-state heterostructures based DNA sequencing utilizing ionic current approach. a) Schematics illustrating
the translocation of ssDNA through a graphene/MoS2heterostructure nanopore and the number of transported nucleotide vs. time plot;[141]Repro-
duced with permission.[141]Copyright 2018 American Chemical Society, b) different geometries of graphene/MoS2heterostructure nanopore and the plot
showing translocation rate through both different geometries of graphene/MoS2heterostructure nanopore;[113]Reproduced with permission.[113 ]Copy-
right 2020 American Chemical Society, c) schematic of MD simulations for ssDNA translocation through graphene/h-BN heterostructure nanopore;[142]
Reproduced with permission.[142]Copyright 2020 American Chemical Society, and d) Schematic of BC3/C3N heterostructure nanopore and translocation
rate;[143]Reproduced with permission.[143 ]Copyright 2021 The Royal Society of Chemistry.
3.2. Transverse Tunnelling Current Approach
Recent advancements in the field of 2D solid-state nanomateri-
als and field-effect transistor domain have paved the path for a
new potential technique for DNA sequencing named as trans-
verse tunneling current method. During tunneling current mea-
surement, the temporary molecular junction is created between
the electrodes, and the subsequent current is analyzed.[148]The
idea of transverse electrodes originated from the need to differ-
entiate between DNA nucleotides using tunneling current sig-
nals, which could be a promising alternative to the longitudinal
ionic current approach. In principle, the tunneling current DNA
sequencing promises to exhibit sufficiently distinct electrical cur-
rent readouts for each base, which in turn helps in deducing the
nucleotide sequence in a DNA strand. In comparison to the ionic
current technique, this technique is capable of sequencing faster
(by orders of magnitude).[20,48]
This method consists of four types of devices: nanopore,
nanogap, nanochannel, and heterostructure. In a nanopore de-
vice, the interaction of ssDNA with the edges of the nanopore
allows for the measurement of both tunneling current and trans-
verse conductance of the membrane concurrently, aiding in the
differentiation among the four nucleotides. A nanogap device
consists of a set of nanoelectrodes placed on the membrane sur-
face on both sides of the nanopore aperture. The transverse elec-
trodes allow for the measurement of electronic tunneling cur-
rent flowing through the gap between those electrodes. In a
nanochannel device, the ssDNA is pulled over the surface to mea-
sure the change in the tunneling current signal due to the interac-
tion of stacked nucleotides over the nanochannel surface. In the
last decade, various 2D solid-state nanomaterials have been in-
vestigated for DNA sequencing using transverse tunneling mea-
surements, the details of which are provided in the subsequent
sections.
3.2.1. Solid-State Nanopore
Regardless of major developments in the ionic current method
with the solid-state nanopore, the high signal-to-noise ratio (SNR)
and lack of single-base resolution are the major challenges
causing hindrances in the further advancement of the field.
Moreover, the technique requires consumables, which makes
it computationally and experimentally more expensive. The re-
searchers have developed a potential alternative to ionic current
nanopore sequencing, i.e., transverse tunneling nanopore se-
quencing, which is purely a physical method and has a high SNR.
The underlying principle of transverse tunneling nanopore se-
quencing is when a ssDNA molecule is electrophoretically driven
through a nanoscale pore (diameter of 2–5 nm), different nu-
cleotides (different in shape and size) cause different variations
in the transverse conductance and current readouts and based
on that DNA sequencing is made possible.[149]The modulation
of transverse conductance arises from the nucleotide-specific
interaction between DNA and solid-state nanopore, which af-
fects the local density of states and characteristic nucleotide-
specific coupling strength with the nanopore edges. In these tech-
niques, DNA molecules transverse through the nanopore embed-
ded in closely placed semiconducting/semi-metallic electrodes,
and current modulation is measured between the left and right
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (15 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
Tabl e 2 . Year-wise documentation of ionic and transverse current studies reported on hybrid/heterostructures for DNA sequencing.
Materials Nanodevice Molecular carrier Study Comments Year Refs.
𝛼HL +Si3N4Nanopore dsDNA Experimental The hybrid nanopore is noticed to be
functional for 30–40% of the attempts
2010 [123]
𝛼HL +biomimetic nanopore Nanopore ssDNA Experimental Discrimination of poly[A] and poly[C] with
increased dwell time
2013 [124]
DNA origami-graphene Nanopore ssDNA MD simulations Different residence time and ionic
current signals; decreased
translocation speed
2017 [125]
Graphene/MoS2Nanopore ssDNA MD simulations The chemical potential difference on
graphene and MoS2surfaces directly
drives ssDNA
2018 [141]
Natural DNA pore- Si3N4Nanopore dsDNA and ssDNA Experimental Lipid-free hybrid nanopore exhibiting low
peripheral leakage
2018 [127]
DNA origami-graphene Nanopore ssDNA MD simulations Specific interactions between overhangs
and ssDNA results in base-specific
residence time
2019 [126]
𝛼HL +biomimetic nanopore Nanopore ssDNA MD simulations The ionic current blockade depends on
the type of polynucleotide and its
conformation
2020 [140]
Graphene/MoS2Nanopore ssDNA MD simulations Reduced translocation speed in the
order: circular nanopore <square
nanopore <triangular nanopore
2020 [113]
Graphene/h-BN Nanopore ssDNA MD +DFT Manipulation of ssDNA conformation
through differential van der Waals
interaction
2020 [142]
eOmpG/MoS2Nanopore Single-stranded
polyadenine (dA30)
Experimental 32% lower noise levels with nearly 1.9%
improved signal-to-noise ratio and 6.5
times longer dwell time
2021 [147]
BC3/C3N Nanopore ssDNA MD simulations Spontaneous translocation from BC3to
C3N and reduced nucleotide
translocation
2021 [143]
Graphene/h-BN (vertically
stacked)
Nanopore dsDNA MD simulations Heterostructure layer order plays crucial
role in controlling the base-specific
signal variability
2021 [146]
electrodes through the scattering region. A bias voltage is applied
across the 2D nanopore device to obtain a finite current of the de-
vice, and a gated voltage is applied to determine tunneling con-
ductance.
Among the 2D materials, graphene has a single atomic thick-
ness (0.34 nm thick, equivalent to the spacing between two
bases in a DNA strand) and extraordinary mechanical and elec-
tronic properties complementary to single-base resolution. In
2010, Nelson and co-workers studied the interaction of four DNA
nucleobases with the graphene nanopore by calculating the elec-
tronic structure and density of states of a graphene nanorib-
bon with the ab initio density functional theory (DFT), as shown
in Figure 13a.[150]Further, they also studied the transverse con-
ductance and electrical current using the Landauer–Bϋttiker for-
mula, and the result analysis concluded that single nucleotide
detection can be achieved via measurement of the in-plane cur-
rent signals and conductance of the semiconducting graphene
nanoribbons with armchair edges. Electrical current and conduc-
tance spectra provided sufficient information for the identifica-
tion of DNA molecules inside the nanopore. Moreover, the oper-
ating current of the nanopore was noted to be insensitive to the
rotated configurations of a specific DNA nucleobase, which is a
desirable requirement for the feasibility of the device for experi-
mental realization.
However, the bare graphene nanopore has lower coupling
strength with DNA nucleotide due to the much smaller space ex-
tension of carbon outer orbitals. As the coupling strength drops
exponentially with decreasing overlap, the transverse tunneling
conductance using bare graphene nanopore will be reduced dra-
matically, which consequently deteriorates the SNR ratio. There-
fore, there occurs a need to introduce functionalized electrodes
that could improve the coupling strength between the graphene
nanopore and nucleotide.
He et al. studied the merits of hydrogen functionalization
of graphene nanopore edges on DNA sequencing with elec-
tronic transport studies (Figure 13b).[151]They demonstrated that
compared to the bare graphene nanopore, the hydrogen atom
functionalized nanopore accelerates the formation of temporary
H-bonds between nucleotide and translocating DNA molecule,
which leads to an increase in the average conductivity (3 orders
of magnitude) while exhibiting significantly reduced statistical
variance. The results also determined that the hydrogenation of
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (16 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
Figure 13. Electrical detection of DNA nucleobases by 2D solid-state nanopore devices with the transverse tunneling current approach. a) Illustra-
tion of a proposed graphene nanopore device for identification of single nucleobases with transverse tunneling current method;[150]Reproduced with
permission.[150]Copyright 2010 American Chemical Society, b) snapshot of MD simulation of the DNA translocation through the hydrogen functionalized
graphene nanopore;[151]Reproduced with permission.[151 ]Copyright 2011 Wiley Publishing Group, c) atomistic illustration of the graphene nanopore,
DNA, water, and counterions with the QM/MM-NEGF protocol;[152]Reproduced with permission.[152]Copyright 2018 American Chemical Society, and
d) schematic illustration of the Ti2C(OH)2MXene nanopore device for measuring the conductance of translocating DNA nucleotides;[153]Reproduced
with permission.[153]Copyright 2021 Springer Publishing Group.
graphene edges reduces the translocation speed of DNA and con-
siderably improves the accuracy of the measurement of individ-
ual nucleotides.
In a similar report, Saha et al. investigated the graphene
nanoribbon with zigzag edges (ZGNR) for DNA nucleobase iden-
tification with the transverse tunneling current method.[148]In-
terestingly, the tunneling current across the nanopore is observed
in the range of microampere (μA) at low bias voltage (0.1 V),
which is of several orders higher magnitude than that of
graphene nanopore with armchair edges. The higher current of
the ZGNR nanopore is predominated by the local current den-
sity around the edges. Therefore, drilling a nanopore in the cen-
ter of the device, away from the edges, will not reduce the con-
ductance. Additionally, the operating current is anticipated to be
sufficient to yield a rich SNR and expected to be much larger
than the thermal vibration of the graphene device, orientational
fluctuation of the translocating DNA strand, and solvent influ-
ence on the electronic properties of the nucleotides. The ZGNR
nanopore has also been successful in identifying the epigenetic
nucleotides from the natural DNA nucleotides.[154]However, the
above-discussed reports have excluded the environmental condi-
tions and solvent effects.
To get insight into the impact of environmental conditions on
tunneling current signals, Feliciano et al. studied the graphene
nanopore with the QM/MM electronic transport method, in-
corporating the effect of structural fluctuations of the nu-
cleotides (Figure 13c).[152]The results show that nucleotide-
specific charge fluctuation is responsible for the characteris-
tic conductance modulation of the device. However, they also
pointed out that the conductance profiles of all four nucle-
obases are similar due to higher conducting pathways across
the graphene nanopore membrane.[155]Despite extensive stud-
ies with graphene nanopores, the high translocation velocity of
DNA, noise, and pore-clogging issues hinder individual DNA
base identification. An extensive molecular dynamics simula-
tion followed by the non-equilibrium Green’s function (NEGF)
study demonstrated that an intrinsic stepwise translocation of ss-
DNA strand by mechanically stretching the ssDNA perpendic-
ular to the graphene sheet as it passes through the nanopore
improved the signal quality to ensure high fidelity detection of
nucleotides.[156]The stepwise translocation not only increased
the dwelling time of each nucleotide inside the nanopore but also
suppressed the signal noise by escaping the hydrophobic interac-
tion between the nucleotide and graphene.
Several theoretical strategies and architectures have also
been proposed with graphene membranes to amplify the high-
resolution identification of DNA nucleotides and address the ex-
isting challenges. For example, Ouyang et al. proposed that the
position of nanopores has a significant effect on their DNA nu-
cleobase detection ability.[157]The current signature read-outs
of each nucleobase are found to be remarkably distinguishable
upon designing a device with a nanopore at the edges of the
metallic graphene nanoribbon membrane. Girdhar et al. intro-
duced a peculiar concept of quantum point contact (QPC), where
the nanopore device is designed with irregular edge-shaped
graphene nanoribbons.[158]This QPC device is reported to ex-
hibit superior electrical sensitivity of DNA to the earlier reported
uniform armchair geometry by tuning the graphene carrier con-
centration. The proposed device can also be modeled by sand-
wiching the solid-state graphene QPC between two dielectrics,
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (17 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
which will isolate the electrically active graphene layer from the
electrolyte and restrain its mechanical instability. Additionally,
the QPC carrier tunability can be achieved by the application
of simultaneous gate voltage similar to the field-effect transistor
(FET) during DNA translocation. However, the above reports did
not include the translocation velocity of DNA, which has been
a key stumbling block to obtaining error-free sequencing using
nanopores.
Another proposed alternative strategy was to use a bilayer
graphene nanopore instead of a monolayer and compute the
probability distribution of electrical signals for DNA nucleotide
identification.[159,160]The motivation behind this is that the bi-
layer graphene nanopore is more stable and persistent than the
monolayer counterpart, which changes shape over time. Addi-
tionally, bilayer graphene has more conductance channels due
to the presence of higher electronic density of states compared
to monolayer graphene. In a similar report on graphene bi-
layer nanopore, interlayer conductance (top and bottom layer of
graphene) is present in the bilayer, while DNA nucleotides are
translocated through it.[161]The results revealed that this pro-
posed device gives rise to better identification of the individ-
ual DNA nucleotide conductance signals compared to the iso-
lated nanopore. Moreover, the OH group functionalization of
graphene nanopore edges improved the interaction with translo-
cating ssDNA and controlled the dynamical fluctuation of nu-
cleotides. Cuniberti and co-workers theoretically demonstrated
that the speed of DNA passage can be controlled by adding extra
graphene sheets on the top of the parent graphene nanopore.[162]
It should be mentioned that the dimensions of the pore con-
sidered for this study are in accordance with the earlier experi-
mentally reported graphene nanopore for DNA translocation pur-
poses. This report pointed out the reduction of the transmission
signal by the graphene pore as the DNA translocates through
the nanopore with a diameter similar to the experimental report.
Further, a hybrid nanopore with a so-called graphene nano-road
embedded in a hexagonal boron nitride (h-BN) sheet was devel-
oped, and the possibility of controlling the local current pathway
through the proposed device was demonstrated by applying a spe-
cific gate voltage.[163]
In recent years, the discovery of more solid-state 2D materials
beyond graphene has opened up a new possibility to exploit the
field of DNA sequencing by improving the high precision detec-
tion efficiency of solid-state nanopores. Some studies have also
been reported on DNA sequencing as an alternative to graphene
nanopore. Prasongkit et al. investigated the potential of a topo-
logically defective graphene nanopore for DNA sequencing with
the QM/MM approach and observed that the nucleotide conduc-
tance is associated with transport across specific molecular states
near the Fermi level and their coupling with the pore edges.[164]
Pathak and co-workers investigated the single-atom-thick solid-
state 2D graphdiyne nanopore setup for DNA sequencing. From
the conductance and current sensitivity analysis, it is clear that
all four nucleotides can be distinctly identified improved sensitiv-
ity compared to graphene nanopore.[165]Semi-metallic graphene
and metallic borophene nanopores are considered for the study
to compare the preference of semi-metallic and metallic 2D ten-
dency towards single nucleotide identification.[166]It is noted that
the borophene nanopore with robust condition channels across
the membrane doesn’t get affected by the presence of nucleotide
inside the nanopore, which is evident from their current-voltage
characteristics.
Recently, MXene, a new family of 2D transitional metal car-
bides, has been poised as a potential candidate for nanopore-
based DNA sequencing with a transverse tunneling current
method due to relatively easy fabrication, which can resolve pore
clogging issues and versatile electronic properties. Mojtabavi and
co-workers have experimentally demonstrated the applicability of
Ti2CTxand Ti3C2TxMXene nanopore membranes toward sin-
gle DNA nucleotide identification and found that, unlike other
2D materials, the MXene membranes have low noise levels and
less ionic current leakage compared to standard graphene and
MoS2nanopores.[109 ]Scheicher and co-workers have employed
combined MD simulation and DFT-NEGF analysis for single nu-
cleotide identification with OH functionalized Ti2C MXenes, i.e.,
Ti2C(OH)2nanopore, as shown in Figure 13d.[153]The Ti2C(OH)2
nanopore was reported to help in the stabilization of DNA nu-
cleotides as DNA passes through the nanopore and allowed con-
ductance measurements by sweeping the gate voltage at several
orientations. The OH functionalized pore allowed better interac-
tion and distinct transverse conductance properties at the Fermi
level.
Several experimental studies have also been reported in par-
allel to validate the feasibility of the above-discussed theoreti-
cal predictions of solid-state nanopores for DNA sequencing.
In 2012, Lieber and co-workers investigated a silicon nanowire
field-effect transistor (FET) for single molecule DNA sequencing
utilizing integrated silicon nanowire-nanopore FET on a Si3N4
membrane, where the signal is generated through the localized
change of electric potential during DNA translocation across the
nanopore.[167]In contrast to a Si3N4membrane, the graphene
membrane acts as a conductor and thus allows transversal cur-
rent flow through the membrane. Therefore, the nanopore FET
can be extended to atomically thin graphene membranes, which
have the potential to achieve single-base resolution. In 2013,
Traversi et al. fabricated the graphene nanoribbon of 100 nm
width for the nanopore drilling on the top of the Si3N4mem-
brane (Figure 14a).[168]Here, simultaneously, both the longitudi-
nal ionic current blockage across the nanopore with the translo-
cation of DNA molecules and the transverse tunneling current
across the graphene device are measured (Figure 14b).
In a similar report with a solid-state nanopore embedded with
a graphene nanoribbon (GNR) transistor, Puster et al. examined
the GNR nanopore composed of 50 nm (width) and 600 nm
(length) over the Si3N4membrane and noticed a significant re-
lationship between the ionic current and nanoribbon current
due to the capacitive coupling during the DNA translocation,
which is termed as “cross-talk.”[171]In another experimental re-
port, Heerema and co-workers experimentally fabricated 30 nm ×
30 nm square-shaped graphene nanoribbon (top layer was cov-
ered with h-BN membrane) with a nanopore of 5 nm diameter
and investigated the possibility of DNA sensing with in-plane
currents by using a custom-made differential current amplifier
that discriminates between the capacitive current signals and
the resistive response in graphene.[169]An important advantage
of this approach is the transverse current magnitude through
graphene nanoribbons is relatively large (μA), which facilitates
the measurements at much higher bandwidths, which opens up
the possibility to sequence DNA information at the translocation
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (18 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
Figure 14. Experimental studies of electrical detection of DNA nucleobases by 2D solid-state nanopore devices with the transverse tunneling current
approach. a) Schematic representation of the device setup. A DNA molecule is translocating through a nanopore fabricated in a graphene nanoribbon
supported by a Si3N4membrane (dark blue);[168]Reproduced with permission.[168 ]Copyright 2013 Nature Publishing Group, b) the recorded ionic
current and transverse current traces during DNA translocation;[168]Reproduced with permission.[168 ]Copyright 2013 Nature Publishing Group, c) both
in-plane current (through a graphene nanoribbon) and ionic current (through the nanopore) signals are measured, while a DNA molecule translocates
through a nanopore in that ribbon;[169]Reproduced with permission.[169 ]Copyright 2018 American Chemical Society, and d) schematic of the proposed
MoS2FET-nanopore device with selective insulation around the electrodes avoids cross-talk and insulates the metal from the electrolyte while keeping
the ribbon exposed to the solvent;[170]Reproduced with permission.[170 ]Copyright 2019 American Chemical Society.
speed that is typically observed with solid-state nanopores. From
the above experimental studies of graphene nanopore for DNA
sequencing with the transverse tunneling current method, it is
noted that sculpting graphene nanoribbon can open a bandgap,
which is essential to become ambipolar FET. However, the fab-
rication of nanoribbons with smaller nanopore is challenging.
Moreover, the measurement of simultaneous ionic current and
the transverse sheet current through the graphene membrane
led to capacitive cross-talk, which has severely restricted the in-
dustrial scale realization (Figure 14c).[171]
Several other 2D materials have been computationally ex-
plored for the transverse detection of DNA nucleotides. How-
ever, there are limited experimental studies. In 2014, Feng and
co-workers experimentally established that MoS2nanopore can
complement graphene nanopore membranes and offer poten-
tially better performance in transverse detection. The major ad-
vantages are that it doesn’t need any additional surface treatment
to avoid strong surface interaction, unlike graphene, and offers
an improved sensitivity compared to Si3N4nanopores (SNR >
10).[108]Recently, the translocating DNA nucleotides have also
been detected from the readouts of both the ionic and transverse
current modulation through the MoS2nanoribbon, as shown in
Figure 14d.[170]Here, the authors implemented single-layer MoS2
nanopore-based FET (500 nm ×2μm) on top of an aperture in a
Si3N4membrane. The area selective insulation around the elec-
trodes is incorporated to avoid cross-talk and insulate the metal
from the electrolyte while keeping the ribbon exposed to the sol-
vent. In comparison to graphene FET, the MoS2is associated with
a reduction of low-frequency 1/fnoise with a decrease in salt con-
centration.
Recently, Pathak and co-workers explored the potential of ma-
chine learning in the field of nanopore technology with the trans-
verse tunneling current method. By employing the supervised
ML regression tools with the quantum transport method, sin-
gle DNA nucleotide prediction is made possible using a model
monolayer gold nanopore (Figure 15a).[172]Training the XGBR
model with single nucleotide transmission data sets, including
electronic and chemical features, transmission spectra of other
nucleotides are predicted. Interestingly, dCMP nucleotide data
sets were successful in predicting all three other nucleotides
(dTMP, dGMP, and dAMP). On comparing both the DFT-NEGF
and ML-predicted transmission sensitivity values, it is observed
that though the sensitivity order of the nucleotides decreases,
their qualitative trend remains the same (Figure 15b).[173]More-
over, the local and global interpretable analysis by the SHapley
Additive exPlanations (SHAP) method shed light on the complex
relationship between the molecular properties of nucleotides and
their transmission function. Except for the energy feature, the
electronegative, ionization energies, and electron affinity were
the dominant features in the prediction of the output. Notably, the
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (19 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
Figure 15. Combined machine learning and transverse tunneling approach with nanopore for DNA sequencing. a) An interpretable ML prediction of the
transmission functions of single DNA nucleotides. Trained with the dCMP nucleotide data set, the XGBoost regression model predicted dAMP, dGMP,
and dTMP nucleotides;[172]Reproduced with permission.[172 ]Copyright 2022 American Chemical Society, b) DFT calculated versus machine learning
predicted conductance sensitvity;[172]Reproduced with permission.[172 ]Copyright 2022 American Chemical Society, and c) strategy for the development
of an artificially intelligent nanopore with the machine learning and quantum tunneling method;[173]Reproduced with permission.[173 ]Copyright 2023
American Chemical Society.
successful prediction of nucleotide configuration was restricted
to only the most stable configurations.
In a follow-up study, the authors developed an artificial intel-
ligent nanopore with combined machine learning and quantum
transport method, where four different XGBR models were built
that are trained on dAMP, dTMP, dGMP, and dCMP transmis-
sion function data of graphene nanopore (Figure 15c).[173]The
trained models were able to predict not only the most stable
configuration but also all the considered rotated orientations in-
side the nanogap. Further, the overlapped signals of all four nu-
cleotides are classified with the support vector machine (SVM)
classifier with an accuracy of up to 96.01%. Both combinations
of regression and classification methods boosted the efficiency of
the graphene nanopore toward the high-precision identification
of DNA nucleotides. Table 3below summarizes year-wise impor-
tant contributions from the experimental and computational re-
ports on nanopore-based DNA sequencing.
3.2.2. Solid-State Nanogaps for DNA Sequencing
Solid-state nanogap device has emerged as another wonderful al-
ternative technique for DNA sequencing with several advantages,
such as enabling single-molecule resolution owing to inherent
superiorities of electrical transduction methods, good compati-
bility with advanced field-effect transistor technology, and pro-
viding a sufficient level of sensitivity with long readability and
fidelity.[49,174–176]Solid-state nanogap devices are made up of two
electrodes separated by a distance in the magnitude of a nanome-
ter or sub-nanometer. The transverse tunneling current is moni-
tored from the left electrode to the right electrode across the gap
for single-molecule identifications.
When a DNA molecule strand traverses through the nanogap,
the electronic structure, chemical composition, and induced
dipole-moment of each nucleotide located inside the nanogap
modulate the electron tunneling current of the device with dis-
tinct magnitude and time duration. That modulation in the
tunneling current allows the identification of DNA nucleotides
with good sensitivity. Also, it is important to note that, unlike
nanopore devices, the interaction can be controlled for molecules
of different sizes by relatively easy and flexible adjustment of the
gap size to detect molecules more effectively, as some molecules
have varying interactions with the nanoelectrodes depending on
the size of the gaps.[18]In comparison to other nano-devices, a
big advantage of nanogap detection systems is that they detect
the biomolecule at a single molecular level, making it highly
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (20 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
Tabl e 3 . Year-wise documentation of DFT-NEGF and experimental studies reported on solid-state nanopore-based DNA sequencing.
Materials Nanopore Molecular carrier Study Comments Year Refs.
Graphene Armchair graphene nanoribbon nanopore
with diameter 1.45 nm
Single nucleobase DFT +NEGF Modulation of conductance and
current-voltage (I-V) spectra can identify
DNA bases
2010 [150]
Graphene Hydrogen functionalized nanopore
electrodes with diameter 1.1 and 1.3 nm
Single-stranded DNA MD +NEGF Hydrogen termination increases the
conductance by three orders than the bare
nanopore
2011 [151]
Graphene Graphene nanoribbon-based nanopore
with a diameter of 1.5 nm
Single nucleobase DFT +NEGF Distinguishability of the nanopore current
depends on the nanopore position on the
device
2011 [157]
Graphene graphene nanoribbon with zigzag edges
(ZGNR) and nanopore (1.2 nm)
Single nucleotide DFT +NEGF Transverse current in the microampere range
is achieved with an edge termination
2012 [148]
Silicon Silicon nanowire-nanopore (2 nm) Double-stranded DNA Experimental Simultaneous ionic current and FET
conductance measured
2012 [167]
Graphene Graphene pore with four sheets over the
membrane (1.6 nm)
Double-stranded DNA MD+DFT-NEGF The addition of extra graphene sheets
around the graphene nanopore reduced
the DNA translocation speed
2013 [162]
Graphene Bilayer nanopore with a diameter (of 1.2
and 3 nm)
Single nucleotides MD+DFT-NEGF A distinct change in the current signal for
bilayer nanopore noted than monolayer
during nucleobase detection
2014 [159]
Graphene Nanopore
(0.5 nm)
Natural and epigenetic
nucleobase
DFT +NEGF Graphene nanopore can detect both natural
and methylated nucleotides
2014 [154]
MoS2MoS2is suspended on a Si3N4(2-20 nm) Double-stranded DNA Experimental Unlike graphene, there is no strong surface
interaction with DNA SNR higher than 10.
2014 [108]
Graphene Nanopore diameter (1.3 nm) Single nucleobase QM/MM Charge fluctuations in nucleotides are
responsible for the conductance
modulation
2015 [155]
Graphene Quantum point contact-edged nanopore
with a diameter (1.6 nm)
Single-stranded DNA MD+NEGF Vertically stretched stepwise translocation of
ss-DNA improved detection ability with
reduced noise
2015 [156]
Graphene Graphene nanopore over Si3N4Double-stranded DNA Experimental The cross-talk originating from capacitive
coupling between the ionic current and
device membrane current
2015 [171]
Graphene The OH group terminated graphene
bilayer nanopore with a diameter (2 nm)
Single-stranded DNA MD+DFT-NEGF The OH group improved interaction and
reduced translocation speed with a better
conductance signal
2015 [161]
Graphene and h-BN Graphene with h-BN at top layer
Nanoribbon nanopore diameter (5 nm)
ssDNA Experimental Exceedingly challenging due to low yields in
device fabrication
2018 [169]
Topologically defected
graphene
Hydrogen functionalized nanopore with
diameter (1.2 nm)
Single nucleotide MD and DFT +NEGF Conductance is associated with transport
across specific molecular states near the
Fermi level and their coupling to the pore
2018 [164]
MoS2Nanopore diameter (1.3 nm) Single and
double-stranded
DNA
Experimental Better FET device with a high signal-to-noise
ratio
2019 [170]
Graphdiyne Hydrogen functionalized nanopore with
diameter (1.4 nm)
Single nucleotide DFT +NEGF The distinction between purine- and
pyrimidine-type nucleobases
2021 [165]
Graphene and Borophene Hydrogen functionalized graphene and
borophene nanopores with diameters
1.3 and 1.2 nm, respectively
Single nucleotide DFT +NEGF Metallic borophene nanopore is showing
poor distinguishability
2021 [166]
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (21 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
Figure 16. Theoretical studies of solid-state nanogaps-based DNA sequencing with transverse tunneling current approach. a) Schematic representation
of the translocation of single DNA nucleotides through a graphene nanogap (left) and illustration of transverse conductance calculation with numerical
method;[178]Reproduced with permission.[178 ]Copyright 2010 American Chemical Society, b) a single DNA nucleotide (dGMP) is placed between two
hydrogen atom functionalized graphene electrodes with a gap of 12.6 Å;[179]Reproduced with permission.[179 ]Copyright 2011 American Chemical Society,
c) current variation due to nucleotide rotation about the y-axis and translation along the z-axis at a bias of 1 V;[179]Reproduced with permission.[179 ]
Copyright 2011 American Chemical Society, d) illustration of the double-functionalized graphene nanoelectrodes;[180]Reproduced with permission.[180 ]
Copyright 2013 American Chemical Society, e) the current-voltage curves plotted on a semilogarithmic scale in the left panel for the four targeted
nucleotides;[180]Reproduced with permission.[180 ]Copyright 2013 American Chemical Society, f) QM/MM study of nitrogen functionalized graphene
nanogap showing electron transport pathway in the presence of solvent;[181]Reproduced with permission.[181 ]Copyright 2019 Elsevier, and g) schematic
of DNA sequencing through phosphorene nanogap device via transverse current approach and current sensitivity of four nucleotides calculated at the
bias of 1.6 V;[176]Reproduced with permission.[176 ]Copyright 2017 American Chemical Society.
sensitive. Furthermore, the size of the gaps can be adjusted rela-
tively easily, allowing the system to detect molecules of different
sizes.
Lagerkvist et al. reported the modality of nanopore sequenc-
ing by exploiting the quantum mechanical phenomenon of
electron tunneling, allowing the identification of individual
nucleobases.[177]Postma proposed the concept of nanogap gap-
based DNA sequencing by using the numerical simulation con-
cept with graphene electrodes (Figure 16a).[178]It is demonstrated
that the single base resolution of DNA nucleotide can be ob-
tained by measuring their electrical conductance signals while
translocating between two graphene electrodes. Scheicher and
co-workers have calculated the transmission function of each nu-
cleotide and transverse tunneling current variation at finite bias
with the DFT-NEGF approach (Figure 16b).[179]The results show
that the electrode-nucleotide coupling effect is vital for change in
the magnitude of conductance and tunneling current signals, as
shown in Figure 16c. This study with a graphene nanogap device
for DNA sequencing was crucial in addressing the technical prob-
lem of achieving single-base resolution. Additionally, the report
highlighted that the nucleotide with a HOMO peak close to the
Fermi energy of the graphene electrodes can increase the current
magnitude.
Though graphene nanogap results were promising, the low
sensitivity and high translocation speed remained an important
concern. To address these issues of graphene nanogap, the idea of
functionalizing the electrodes with different molecules or atoms
other than hydrogen was proposed. Prasongkit et al. functional-
ized the phosphate-group-grabbing guanidinium ion on the right
electrode and a reader-nucleobase (cytosine) on the left electrode
of the graphene nanogap and studied the electronic transport
properties of DNA nucleotides (Figure 16d).[180]Double function-
alization of nanogap edges created a temporary hydrogen bond-
ing that decreased their orientation fluctuations, which further
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (22 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
stabilized the nucleotide inside the nanogap and improved the
electronic coupling, leading to better distinguishability in the
current–voltage (IV) characteristics (Figure 16e). Moreover, the
improved interactions between nucleotides and functionalized
electrodes can slow down the high translocation speed of DNA
and provide more time for each nucleotide measurement. Com-
pared to non-functionalized graphene electrodes, the function-
alized electrode shifts the molecular orbitals closer to the Fermi
level, delivering better and unambiguous electrical recognition of
nucleobase entities.
However, the functionalization of nanoelectrodes with molec-
ular groups will undoubtedly add another layer of complexity to
the fabrication process. Further, efforts have also been made to
enhance the DNA recognition sensitivity of graphene nanogap
devices through nitrogen edge functionalization.[182]On compar-
ison of both hydrogen and nitrogen functionalization graphene
electrodes, it is noticed that nitrogen-functionalized electrodes
lead to increased sensitivity by up to 3 orders of magnitude with-
out affecting the qualitative trend. The higher sensitivity could
be due to strong coupling and the contribution of nitrogen-
functionalized edges to resonance states close to the Fermi level,
resulting in increased conductance.
Additionally, De et al. performed a QM/MM coupled with the
NEGF study with a nitrogen-functionalized graphene nanogap
setup and demonstrated that water plays a major role in the elec-
tronic transport of the nanogap tunneling device (Figure 16f).[181]
The local current analysis showed the solvent molecule pro-
vides localized electronic states that significantly increase the
conductance in nitrogen-functionalized nanogap devices. On
the other hand, Jung et al. also studied the potential of het-
eroatom substitution to control the dynamics of DNA translo-
cations with MD-DFT-based study on the pristine carbon nan-
otube (CNT) and nitrogen-doped carbon nanotube (capCNT)
electrodes.[183]It is observed the nitrogen doping of capCNTs
stabilized nucleotide configuration by the ‘edge on’ interaction
rather than the original ‘face-on’ type interaction in the pristine
CNT electrodes. Moreover, the ssDNA translocation and confor-
mations were found to be dominated by the 𝜋𝜋interactions of
nucleobases-CNT caps and nucleotide dipole moments-external
electric field interaction. The capCNT facilitated the entrance
of ssDNA by dominant ‘edge on’ conformation and control of
DNA translocation speed by establishing hydrogen bonds be-
tween the N dopant atoms and nucleotides by increasing the
translocation time by 300%. Instead of doping, when the left
electrode of the CNT gap is functionalized with the guanine
nucleobase and attempted to detect with the quantum trans-
port method, each nucleotide orientation is stabilized by forming
temporal hydrogen bonding with the functionalized molecule.
The continuous molecular orbitals at specific transmission peaks
represent an improved coupling effect with the CNT, which
is reflected in the distinctive electrical current signal for each
nucleotide.[184]
Fyta and co-workers investigated the diamondoid-
functionalized Au (111) electrodes for the detection of different
nucleotides through quantum tunneling measurements and
noted that the sensitivity of the device at different gate volt-
ages depends upon the coupling or decoupling effect of the
functionalized electrodes with the targeted nucleotide, which is
confirmed from the presence or absence of underlying local cur-
rent and scatter transmission channel.[185]In another report with
the first principle study, Zou et al. studied sulfur-functionalized
gold electrodes and the effect of nucleotide orientation inside
the nanogap towards single nucleotide identification.[186]The
transport properties of nucleotides increased due to better inter-
action with the terminated sulfur atoms. The report highlighted
that sulfur atom functionalization substantially enhanced the
distinguishability of nucleotides compared to the bare gold
electrodes.
Beyond graphene, the applicability of the other 2D solid-
state materials towards nanogap-based DNA sequencing with
the transverse tunneling method has also been demonstrated.
Pathak and co-workers reported a study on phosphorene
nanogap for single-molecule DNA sequencing, as shown in
Figure 16g.[176]Phosphorene offers remarkable unique proper-
ties such as biocompatibility and experimentally proven stabil-
ity in both air and water. From the DFT-NEGF study, it is ob-
served that phosphorene can effectively identify all nucleotides
at both low (0.2 V) and high (1.6 V) voltages.[187]The authors
also tried to explore the potential of BC3nanoelectrodes in the
identification of all four DNA nucleotides.[188]The results suggest
that unique identification of all four nucleotides can be achieved
in the bias region 0.3–0.4 V. A comparative study of the interac-
tion of poly-guanine (dGMP dimer) with BC3and graphene edges
shows lower interaction energy of BC3electrodes in comparison
to graphene electrodes. Additionally, BC3is polar compared to
graphene, thus providing better electronic coupling with the po-
larized nanogap edges that resulted in distinct current readouts
by one order of difference, even at low bias.
Apart from exploring alternative materials to graphene, Mit-
tal et al. theoretically introduced the labeling effect to amplify
the DNA recognition sensitivity.[189]The simplest amino acid,
glycine, is tagged to the 5’ position of all four nucleotides,
and their quantum transport properties are investigated with
graphene nanogap electrodes. The results show improved con-
ductance sensitivity, which enables better resolution. In addition,
labeling of nucleotides has the potential to control the transloca-
tion speed and dynamic fluctuations while translocating through
the nanogap. However, except for the dTMP nucleotide, the elec-
tric current of all other nucleotides is found to be similar. Apart
from glycine, three other amino acids: (basic) arginine, (neutral)
tyrosine, and aspartic acid (acidic), have also been studied for la-
beling of DNA nucleotides, and results show that arginine and
glycine labels are promising to achieve distinct current signa-
tures for nucleotide identification.[190]
The topologically defected graphene (octagons and a pair
of pentagon sp2-hybridized carbon rings embedded in a pure
graphene nanosheet) electrodes are also reported for DNA se-
quencing application with DFT-NEGF approach and found to
exhibit high order distinguishability in their current character-
istics at a wide bias range.[191]Rocha and co-workers demon-
strated the potential of different phases of MoS2electrodes for
nanogap-based DNA sequencing. The results indicate that the
1T’-phase of MoS2is selective and sensitive toward nucleobase
identification.[192]The effect of metallic 2D (single atomic thick-
ness) borophene nanogap is also explored to check the via-
bility for nucleotide identification.[193]The nanogap results of
borophene are observed to be better than their nanopore coun-
terpart.
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (23 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
Figure 17. Experimental identification of DNA nucleotides through nanogap device with transverse tunneling method. a) Illustration of Pt electrode-
based tunneling junction located at the bottom of a nanopore, where DNA electrophoretically translocated through the nanopore and the tunneling
junction;[196]Reproduced with permission.[196 ]Copyright 2011 American Chemical Society, b) Nonlinear relation of transverse current with increasing
bias for Pt electrode junction;[196]Reproduced with permission.[196 ]Copyright 2011 American Chemical Society, c) schematics of Pt/Ti electrode-based
nanogap device;[197]Reproduced with permission.[197 ]Copyright 2014 American Chemical Society, d) ionic current and transverse current recorded
during DNA translocation;[197]Reproduced with permission.[197 ]Copyright 2014 American Chemical Society, and e) schematic illustration of the experi-
mental setup of nanosized bare gold gap electrodes for detection of a nucleotide with quantum tunneling approach and chemical modification strategy
to amplify the signal difference between four nucleotides;[198]Reproduced with permission.[198 ]Copyright 2019 American Chemical Society.
Other than traditional architectures, a new architecture called
“semi/hybrid” nanogap with graphene membrane is also pro-
posed to distinguish DNA nucleobases.[194]The semi/hybrid
nanogap has resulted in better conductance and current sensi-
tivity for single nucleotide identification. Moreover, this architec-
ture could be expected to reserve the advantage of both nanopore
and nanogap systems that will make this type of device more ro-
bust.
However, all these proof-of-concept studies with MD and DFT-
NEGF methods look promising. Analyzing experimental results
is extremely important to understand the difficulties and effec-
tiveness of the demonstrated strategies for real sequencing ap-
plications. In 2010, Tsutsui et al. experimentally demonstrated
the first evidence of trapping and identification of single nu-
cleotides between tunneling electrodes.[195]However, this proof
of principle only restricted single nucleotides in the presence
of an STM environment. To develop a potentially much faster
tunneling current-based device, a nanopore combined tunneling
junction can help in both unfolding the DNA strand and precise
measuring of variations in the tunneling currents. In this regard,
Ivanov et al. experimentally reported nanopore-integrated Pt elec-
trodes and calculated concurrent tunneling and ionic current sig-
nals, as shown in Figure 17a.[196]The DNA with a translocation
speed greater than 1 ms was able to detect both ionic and tunnel-
ing current signals. Additionally, they also plotted the IVcurves
for the Pt nanogap devices with different gap sizes and observed
a nonlinear relationship between current and voltage, as shown
in Figure 17b.
Further, Tsutsui et al. experimentally demonstrated single-
molecule detection using a nucleotide-sized nanoelectrode
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (24 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
Figure 18. Machine learning studies of solid-state nanogap for DNA sequencing with transverse tunneling method. a) Detection of modified RNA
nucleotides with recognition tunneling (RT) current with combined scanning tunneling microscopy (STM) and ML techniques. The SVM, a supervised
ML method, was able to identify the individual RNA nucleotides and distinguish them from their DNA counterparts with a reasonably high accuracy of
85%;[204]Reproduced with permission.[204 ]Copyright 2018 American Chemical Society, b) principle diagram of a bare nanogap electrode and electric
current signal;[205]Reproduced with permission.[205 ]Copyright 2018 American Chemical Society,and c) single molecule classification of DNA nucleotides
with a high degree of precision via a one-electric current pulse method using Random Forest and positive unlabeled classification (PUC) method;[205]
Reproduced with permission.[205]Copyright 2018 American Chemical Society.
integrated with a solid-state nanopore device to identify DNA
oligomers.[199]Fanget et al. experimentally reported a high-
throughput fabrication of sub-10 nm nanogap electrodes com-
bined with solid-state nanopores that allowed concomitant
tunneling and ionic current detection of translocating DNA
molecules (Figure 17c,d).[197]Pang et al. fabricated a layered fixed
tunneling junction with functionalized molecules to detect DNA
nucleotides by ‘recognition tunneling’ measurements.[200]Patel
and co-workers experimentally demonstrated the translocation
of double-stranded DNA through a graphene nanogap device.[201]
The conductance and translocation signatures are found to be dif-
ferent than other solid-state nanopores owing to the unique and
flexible geometry of nanogap. However, the accurate identifica-
tion of nucleotides was seriously affected by the similar tunneling
molecular conductance among the nucleotides.
Furuhata et al. experimentally proposed a chemical strategy
that amplified the difference in molecular conductance among
four nucleotides by replacing the 2-deoxyadenosine (dA) with an
analog 7-deaza dA (dzdA) (Figure 17e).[198]This study suggested
that the nucleotide with a HOMO level closer to the electrodes
has higher conductivity compared to other nucleotides. A quan-
tum tunneling probe based on double-barrelled capillary nano-
electrodes has also been reported to identify single molecules at
low concentrations.[202]
Besides computational and experimental studies, there have
been reports on the application of the ML method, which can
address the characteristic challenges of nanogap-based DNA se-
quencing. The overlap of tunneling conductance signals due to
similarity in the shape/size of nucleotides and frontier molecular
orbital is a major problem in nanogap-based DNA sequencing.
Lindsay and co-workers first reported the use of ML methods with
the recognition tunneling (RT) current method to identify sin-
gle protein molecules.[203]The single amino acid molecule can
bind to the recognition molecules at different orientations when
trapped between two electrodes. In this case, the support vector
machine (SVM) algorithm helped to distinguish between the sets
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (25 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
Tabl e 4 . Year-wise documentation of DFT-NEGF and experimental studies reported on solid-state nanogaps for DNA sequencing.
Materials Nanodevice (gap width) Molecular carrier Study Comments Year Refs.
Gold Nanogap
(1.0 nm)
Nucleotides dissolved in
Milli-Q solution
Experimental Tunneling current depends on the
electrode distance and the
nucleotide-electrode coupling
2010 [195]
Graphene Nanogap
(1.6 nm)
ssDNA DFT +NEGF Significant differences in the
conductance spectra with
single-base resolution
2010 [178]
Gold (Au) electrodes embedded in
nanopore(SiO2) consisting a
multilayer junction(SiO2/Au/SiO2)
Electrode-embedded
nanopore
(sub-nanometer
electrode gap with a
15 nm sized pore)
DNA Oligomer Experimental Electrode-embedded in-plane
nanopore offers label-free
electrical identification of
nucleobases in a DNA oligomer
2011 [199]
Graphene Nanogap
(1.47 nm)
Single nucleotide DFT +NEGF Hydrogen-terminated edges
improved tunneling current
variation by several orders of all
four nucleotides
2011 [179]
Doule-functionalized graphene
electrodes
Nanogap
(2.38 nm)
Single nucleotide DFT +NEGF Temporal hydrogen bonds and
electronic coupling enhanced
the difference in conductance
and current signal of
nucleotides
2013 [180]
Pd nanowire with Si3N4window Nanogap
(1.8 and 2.0 nm)
Solution of single
nucleotides
Experimental When functionalized with
recognition, molecules can
capture individual DNA
nucleotides
2014 [200]
Graphene Nanogap
(1.38 nm)
Nucleobase DFT +NEGF Nitrogen functionalization
boosted the sensing ability of
the device by increasing the
transverse conductance of the
device
2016 [182]
Diamondoid-functionalized gold (111) Nanogap
(1.9 nm)
Natural, mutated, and
epigenetic nucleotide
DFT +NEGF Different nucleobases and their
modifications can be identified
by quantum tunneling
measurements across the
electrodes
2016 [185]
Graphene electrodes placed above the
SiO2substrate
Nanogap
()
Double-stranded DNA Experimental Graphene-DNA interaction
important for DNA
translocation time
2017 [201]
Phosphorene Nanogap
(1.45 nm)
Single nucleotide DFT +NEGF From the I–V signals, all four
nucleotides can be identified.
2018 [187]
(Continued)
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (26 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
Tabl e 4 . (Continued)
Materials Nanodevice (gap width) Molecular carrier Study Comments Year Refs.
Nitrogen-capped carbon nanotube
(CNT)
Nanogap
(1.20 nm)
ssDNA MD and DFT +NEGF Controlled translocation of ssDNA
strand due to formation of
hydrogen bond between the
N-dopant atom of CNT and
nucleobases
2017 [183]
Gold (111) Nanogap
()
Natural and epigenetic
nucleotides
MD and DFT +NEGF The presence of water does not
hinder the detection of DNA
nucleotides
2019 [206]
Graphene Nanogap
(1.90 nm)
Single nucleotide MD and DFT +NEGF Water plays a major role in
electronic transport and
increases the conductance of
the nanogap device
2019 [181]
BC3Nanogap
(1.26 nm)
Single nucleotide DFT +NEGF The lower interaction energy of
BC3electrodes than graphene
electrodes with DNA
nucleotides
2020 [188]
Functionalized carbon nanotube (CNT) Nanogap
(1.90 nm)
Single nucleotide DFT +NEGF Current traces differ by at least 1
order of current magnitude for
all the four target nucleotides
2020 [184]
MoS2Nanogap
(1.35 nm)
Single nucleotide DFT +NEGF 1T-phase MoS2sensitive and
towards DNA sequencing
2020 [192]
Defected Graphene Nanogap
(1.47 nm)
Single nucleotide DFT +NEGF Tunneling electric current signals
that vary by more than one
order of magnitude electric
current
2021 [191]
Gold (Au) tip of two coplanar carbon
electrode
Nanogap
(<5 nm)
Mononucleotides, DNA
oligomers, and proteins
Experimental 5-orders of magnitude increase in
event detection rates and
sub-femtomolar sensitivity
2021 [202]
Graphene Nanogap
(1.56 nm)
Labeled nucleotides DFT +NEGF Labeling of nucleotides amplifies
the conductance sensitivity and
selectivity
2022 [189]
Gold Nanogap
(1.89 nm)
Nucleobase DFT +NEGF Sulfur-functionalized gold
electrodes increase
distinguishability than bare gold
electrodes
2022 [186]
Borophene Nanogap
(1.26 nm)
Single nucleotide DFT +NEGF The current was in the
picoampere (pA) range, which
was fairly higher than the
electrical background noise
2022 [193]
(Continued)
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (27 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
Tabl e 4 . (Continued)
Materials Nanodevice (gap width) Molecular carrier Study Comments Year Refs.
Graphene Semi/hybrid nanogap
(1.30 nm)
Single nucleobase DFT +NEGF Good conductance and current
sensitivity
2022 [194]
Graphene Nanogap
(1.56 nm)
Aspartic acid, arginine, and
tyrosine labeled single
nucleotides
DFT +NEGF Arginine and glycine amino acids
are better for labeling
2023 [190]
of electronic ‘fingerprints’ associated with each binding of that
amino acid molecule. In a similar report, the modified RNA nu-
cleotides are classified by the SVM classifier model to the recogni-
tion tunneling current signal of molecules (Figure 18a).[204]The
classification of recognition tunneling current signal of DNA nu-
cleotides with the support vector classifier (SVC) is noted to have
significant (>90%) accuracy.
In another experimental report, Taniguchi et al. classified the
DNA nucleotides from the collected current signals with the ran-
dom forest classifier and the positive unlabeled classification,
as shown in Figure 18b,c.[205]The authors also studied the ef-
fect of inter-electrode distance on the electronic signal of the nu-
cleotides. With the increase of the gap between the electrodes,
the electrode-nucleotide coupling decreases, which is reflected in
their signature electric signals. The ML model classification ac-
curacy (F1 score) also dropped significantly. Hence, models are
able to extrapolate the coupling information between the elec-
trode and nucleotide. Moreover, the experimental fabrication of
solid-state nanoelectrodes is easier and more economical than
nanopores. Thus, from computational and experimental studies,
we assume that the solid-state nanogap could be realized for DNA
in the future. In Table 4, we have highlighted important reports
on DNA sequencing with the solid-state nanogap device.
3.2.3. Solid-State Nanochannels
Among the NGS methods, nanopore sequencing is the most
promising technique for achieving atomic-scale resolution with
high fidelity and long readability. However, the technique does
not address several issues that need to be resolved for the ex-
perimental purpose, such as controlling DNA translocation rate,
suppression in stochastic nucleobase motions, and controlling
orientational fluctuations of individual nucleobases.[25]Thus, re-
searchers are more focused on finding potential alternatives for
nanopore techniques capable of resolving the existing challenges.
Nanochannel is found to be a promising alternative to
nanopore sequencing. Here, the advantage is that the technique
can be used to detect long chains of DNA with controlled ori-
entational fluctuations, unlike solid-state nanopore or nanogap
architecture.[207–209]The identification process of nanochannel
devices relies on the phenomenon of adsorption at the inter-
face between DNA nucleotides and nano-scale solid-state mate-
rials. In this technique, the DNA strand is dragged through a
nanochannel membrane, and the detector reads the correspond-
ing change in the electric current. On account of different chem-
ical and electronic properties, the strength of the interactions
varies for different nucleobases.[144,207]The unique adsorption
properties of each molecule cause distinct changes in the elec-
tric current signal, enabling sequencing to occur.
Similarly, nanoribbons (1D nanochannel) are also potential
candidates for ultrafast DNA sequencing. In nanoribbon-based
DNA sequencing, the identification of individual nucleobases
is achieved by the distinction in the transmission dip. The lo-
cation of the dips is determined by the alteration in conduc-
tance when a nucleobase interacts with a nanoribbon. According
to Landauer’s theory, the flow of electrons through a molecule
occurs through elastic scattering, which involves a variety of
paths. This results in either constructive interference, where
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (28 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
Figure 19. Graphene nanochannel-based DNA sequencing with the transverse tunneling current approach. a) DNA base stacking on graphene
nanoribbon;[212]Reproduced with permission.[212 ]Copyright 2011 Nature Publishing Group and b) two-dimensional differential conduction maps for
DNA sequencing and methylated DNA detection;[213]Reproduced with permission.[213 ]Copyright 2014 American Chemical Society.
transmission is amplified through resonant enhancement (Breit-
Weigner resonance), or destructive interference that reduces
transmission (Fano resonance).[210,211]The Fano resonance is
caused by the interference between paths formed between the
ground state and a discrete excited state, as well as the ground
state and the continuum of excited states. This Fano resonance
allows the identification of targeted molecules through a nanorib-
bon device.
In 2011, Min et al. first utilized the graphene nanochan-
nel technique to electrically distinguish the DNA nucleobases
(Figure 19a).[212]Based on molecular Fano resonance, which
acts as a signature of each individual, the identification of a
single molecule is found to be feasible. A structural examina-
tion revealed that each base formed a stable stacked config-
uration throughout the entire transit. This stacking is much
stronger than the H–𝜋interactions between water molecules and
graphene. Thus, the stacked structures remain unaltered by sol-
vent effects. In this regard, the proposed nanochannel technique
is found to be promising for experimental purposes.
In another report, the authors introduced two-dimensional
molecular electronics spectroscopy for DNA and cancerous
methylated DNA nucleobase detection with armchair graphene
nanoribbons at atomic resolution (Figure 19b).[213]As the DNA
molecule passes through the nanoribbon, each nucleobase is
stacked through 𝜋𝜋interactions and exhibits different trans-
port characteristics depending on the spatial orientations and
molecular orbital energy levels. The dip in transmission origi-
nates due to Fano resonance, and the resultant 2D conductance
map at bias voltages (Vb) and electron channel energy (E-EF) allow
each nucleobase to be differentiated at spatiotemporal resolution.
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (29 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
Figure 20. Solid-state 2D materials-based nanoribbons for DNA sequencing utilizing transverse tunneling current approach. a) Schematic of armchair
h-BN nanoribbon for DNA sequencing and related zero-bias transmission function plot;[221]Reproduced with permission.[221 ]Copyright 2014 Amer-
ican Chemical Society and b) schematic of atomically thin black phosphorene nanoribbon for DNA sequencing and related zero-bias transmission
function;[224]Reproduced with permission.[224 ]Copyright 2019 American Chemical Society.
Graphene nanoribbons are of two types, armchair and zigzag,
and results suggest that zigzag nanoribbons may be more bene-
ficial for DNA sequencing.[148,214]Based on complex graphene-
DNA binding, various mechanisms have been reported, such
as electrostatic, van der Waals, 𝜋𝜋stacking, and hydropho-
bic interactions.[215]The 𝜋𝜋stacking interaction between DNA
nucleobases and graphene nanoribbons allows controlled ori-
entational fluctuations, which is important for single molecule
recognition.[215,207]
Furthermore, Ahmed et al. demonstrated that the local
density of states of DNA nucleotides deposited on graphene
can also serve as electronic fingerprints for the recognition
of individual bases in scanning tunneling spectroscopy.[216]
The origin of dips from graphene nanoribbon devices is also
investigated, and it is found that dips originate from the
interference between the molecular orbitals of nucleobases
and graphene nanoribbons.[217]Water-immersed nucleobase-
functionalized graphene nanoribbon is reported for selective
and high-speed detection of DNA nucleotides.[218]Apart from
this, boron (B), nitrogen (N), and sulfur (S) doped graphene
nanoribbons have also been explored for DNA nucleobase
identification.[219]B and S-doped graphene nanoribbons act as p-
type semiconductors, and N-doped graphene nanoribbons act as
n-type semiconductors. It is noticed that doping of heteroatoms
in armchair graphene nanoribbon at different sites causes dif-
ferent dips in the transmission curve, which in turn leads to
enhanced sensitivity. Comparative analysis of aforementioned
doped nanoribbons reveals doping of B at different sites can lead
to more distinct conductance. Rocha and co-workers theoreti-
cally studied the environmental effects of DNA detection on the
graphene surface with a hybrid QM/MM and NEGF study.[220 ]
In the presence of solvent, a significant decrease in the trans-
mission of pristine graphene is observed below the Fermi level.
Moreover, the horizontal orientation of ssDNA and dsDNA of-
fers better identification by changing the electronic environment
of the graphene surface.
Apart from graphene, other solid-state 2D materials have
also been explored, such as silicene, hexagonal boron nitride
(h-BN), phosphorene, and molybdenum disulfide (MoS2)for
nanochannel-based DNA sequencing. The potential of armchair
nanoribbons of different materials, namely, silicene, hexago-
nal boron nitride (h-BN), and molybdenum disulfide (MoS2),
has been explored in ultrafast DNA sequencing, and it is no-
ticed that edge-modified h-BN nanoribbon is a promising alter-
native for graphene to achieve single molecule DNA sequenc-
ing (Figure 20a).[221,144]In h-BN nanoribbons, the shift of dip
in transmission as a function of the bias voltage is noticed
to be relatively high, which may be a reason behind the in-
creased sensitivity. Along with nucleobase identification, the
silicene nanoribbon has also been found to be promising for
methylation detection based on electric current signals.[222,223]
The epigenetic modification of DNA nucleobases is associ-
ated with various forms of cancer. Thereby, methylation de-
tection is crucial for developing a universal cancer screen-
ing test. The current–voltage (IV) characteristics of silicene
nanoribbon with adsorbed nucleobases reveal that silicene can
identify all eight nucleobases (DNA and methylated DNA) at
two distinct 0.5 and 1.0 V bias voltages, making it a promis-
ing material for DNA sequencing and cancerous methylation
detection.
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (30 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
Figure 21. Solid-state 2D materials-based nanochannels for DNA sequencing utilizing transverse tunneling current approach. a) Schematic for DNA
sequencing through extended topological line defected graphene nanoribbons and related zero bias transmission function;[225]Reproduced with
permission.[225]Copyright 2020 The Royal Society of Chemistry and b) schematic of armchair germanene nanoribbon for DNA sequencing and related
zero-bias transmission function;[226]Reproduced with permission.[226 ]Copyright 2022 The Royal Society of Chemistry.
Pathak and co-workers further explored the potential of phos-
phorene armchair nanoribbons for individual identification of
nucleobases (Figure 20b).[224]Phosphorene has various advanta-
geous properties over graphene, such as hydrophilic surface, tun-
able band gap, large surface-to-volume ratio, and biocompatibil-
ity. On account of molecular orbital coupling between phospho-
rene nanoribbon and nucleobases, a dip in the transmission is
observed, allowing the identification of DNA nucleobases at sin-
gle molecule resolution. The authors also studied the extended
line defects-based graphene nanodevice for the identification of
DNA nucleobases through conductance measurements, as given
in Figure 21a.[225]Here, the advantage is that extended line de-
fects embedded in a graphene sheet help in precisely controlled
translocation of DNA nucleobase over the nanochannel surface,
promising spatial resolution. The results show strong Fano res-
onance in the transmission function, enabling recognition of all
four DNA nucleobases with good sensitivity and selectivity. In a
very recent study, armchair germanene nanoribbons have been
theoretically explored for sequencing both DNA and RNA nucle-
obases via two-dimensional molecular electronic spectroscopy, as
shown in Figure 21b.[226]Additionally, the germanene nanorib-
bons are also reported to be capable of recognizing cancerous
methylated DNA nucleobases. Here, the advantage is that, com-
pared to gold standard graphene, germanene nanoribbon ex-
hibits more distinct adsorption energy values for different nu-
cleobases.
Despite significant progress in the field, very few studies have
been reported for nanoribbon/nanochannel sequencing. In the
absence of experimental demonstration, the discussion so far in-
cludes only theoretical results. Nanoribbons are potential alterna-
tives to the nanopore/nanogap sequencing technique. However,
to provide more insight into real DNA sequencing, more detailed
theoretical (DFT and MD) and experimental studies are needed.
We believe that in the near future, more studies will be reported
on nanochannel/nanoribbon-based DNA sequencing to fill the
gap between nanopore/nanogap and nanochannel sequencing.
To provide a brief overview of nanochannels/nanoribbons se-
quencing, studies utilizing transverse tunneling current for DNA
sequencing are summarized in Table 5.
3.2.4. Heterostructures
So far, we have discussed nanopore, nanogap, and nanochan-
nel/nanoribbons for DNA sequencing based on transverse cur-
rent measurements. Apart from this, solid-state heterostructures
have also emerged as promising platforms for sequencing DNA.
As previously discussed, graphene has some disadvantages. To
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (31 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
Tabl e 5 . Year-wise documentation of transverse current studies reported on nanochannels/nanoribbons for DNA sequencing.
Materials Nanodevice Molecular carrier Study Comments Year Refs.
Graphene Armchair ssDNA DFT +NEGF Ultrasensitive characteristic Fano
resonance-driven conductance
2011 [212]
Graphene Armchair DNA nucleobases and
methylated DNA
DFT +NEGF Two-dimensional molecular electronics
spectroscopy is demonstrated
2014 [213]
h-BN, MoS2, Silicene,
Graphene
Armchair DNA nucleobases DFT +NEGF Distinct transmission dips for graphene
and h-BN
2014 [221]
Silicene Zigzag DNA nucleobases DFT +NEGF Guanine is distinguishable from the other
three nucleobases
2018 [223]
Doped graphene nanoribbon Armchair DNA nucleobases DFT +NEGF Doping leads to enhanced features in the
transmission and modification of
transport properties
2019 [219]
Graphene Armchair nanoribbon ssDNA and dsDNA QM/MM +NEGF The presence of water decreased
transmission function intensity.
2019 [220]
Phosphorene Armchair DNA nucleobases DFT +NEGF G >A>CT 2019 [224]
Extended line defected
graphene nanoribbon
Zigzag DNA nucleobases DFT +NEGF Precisely controlled translocation of DNA
nucleobases
2020 [225]
Silicene Zigzag Methylated nucleobases DFT +NEGF Differentiation possible of DNA and
methylated DNA nucleobases at the bias
of 0.5 and 1.0 V
2020 [222]
Germanene Armchair DNA/RNA and methylated
nucleobases
DFT+NEGF 2D and 3D conductance maps exhibit
explicitly distinct features
2022 [226]
overcome these drawbacks, several efforts are being made by re-
searchers. It has been shown that the graphene heterostructure
with other solid-state 2D materials can suppress the associated
challenges while preserving the advantages of graphene-based
heterostructures.
Souza et al. proposed a novel single modulation device com-
prising a nanopore formed within a heterostructure composed
of graphene and h-BN for DNA sequencing, as shown in
Figure 22a.[163]The results show that local current is concentrated
in the carbon chain binding nanopore and h-BN part, facilitating
unique current modulation due to the unique dipole moment
of each nucleotide. The study emphasizes that dipole-induced
conduction modulation is a promising approach for sequencing
biomolecules. Later, the device is also found to be capable of de-
tecting hachimoji or unnatural DNA nucleotides.[227]The hachi-
moji nucleobases have low binding affinity to the nanopore, re-
sulting in short resident time during translocation. Moreover,
these different artificial nucleobases in the nanopore cause sig-
nificant variations in the electron transmission properties, lead-
ingtosensitivityupto80%.
Besides this nanopore technology, heterostructures based
on nanogap have also been reported. Shukla et al. studied
the graphene/h-BN heterostructure nanogap architecture for
DNA sequencing based on transverse current measurements
(Figure 22b).[228]The transverse transmission and IVcharacter-
istics of the proposed setup are prominent for single nucleotide
sensing. Moreover, graphene edge passivation with h-BN is re-
ported to solve the major bottlenecks of graphene, such as dif-
ficulty in controlled engineering, hydrophobic interactions, and
higher noise levels. Ralph and co-workers proposed a 2D hybrid
graphene/h-BN double modulation device with a nanopore em-
bedded in a graphene part and found characteristic fingerprints
for each DNA nucleotide (Figure 22c).[229]In the proposed archi-
tecture, by applying a specific gate voltage, local current pathways
can be controlled.
Recently, a nanopore structure has been explored embedded in
graphene/h-BN/graphene heterostructure with 3, 5, and 7 h-BN
layers in the central part. The DFT calculations reveal that nu-
cleotides show better interactions with nanopores in h-BN rather
than nanopores in pristine graphene (Figure 22d).[230]In the case
of graphene/h-BN/graphene heterostructure, higher sensitivity
is observed compared to a similar structure consisting of pris-
tine graphene. Though these theoretical reports are promising
for sequencing, no experimental studies have been reported so
far on solid-state heterostructures utilizing the transverse tunnel-
ing current approach.
Solid-state hybrid nanopores resolved the resolution and
single-molecule sensing issues to a great extent. However, there
still exist some daunting challenges. These consist of high leak-
iness of the nanopore structures, uniform fabrication of het-
erostructure nanopore/nanogap, and the controlled orientation
of nucleotides. Besides, the smooth aggregation of functionalized
pores is also very difficult. These issues require further advance-
ment in the field to improve the performance and reliability of
solid-state heterostructures. To provide a brief overview of DNA
sequencing with solid-state heterostructures, transverse cur-
rent studies reported with heterostructures are summarized in
Table 6.
So far, we have provided a detailed discussion of ionic and
transverse current methods for DNA sequencing across a broad
range of next-generation architectures, including nanopore,
nanogap, nanochannel/nanoribbon, hybrid nanopores, and het-
erostructures. The merits and demerits of each electrical tech-
nique are thoroughly investigated with both experimental and
theoretical studies. For a better understanding, a summary com-
paring the advantages and disadvantages of the ionic current
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (32 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
Figure 22. Schematic representation of solid-state heterostructure DNA sequencing devices utilizing transverse tunneling current approach. a) Sin-
gle modulation graphene/h-BN nanopore;[163]Reproduced with permission.[163 ]Copyright 2017 The Royal Society of Chemistry, b) graphene/h-BN
heterostructure nanogap;[228]Reproduced with permission.[228 ]Copyright 2017 American Chemical Society, c) double modulation graphene/h-BN
nanogap;[229]Reproduced with permission.[229 ]Copyright 2019 The Royal Society of Chemistry, and d) graphene/h-BN/graphene heterostructures with
3, 5, and 7 layers of h-BN;[230]Reproduced with permission.[230 ]Copyright 2021 The Royal Society of Chemistry.
blockade and the transverse tunneling current method is shown
in Table 7.
4. Challenges of DNA Sequencing
Though biological nanopores have certain advantages, such as
reproducibility at the atomic level crucial for consistent and re-
liable sequencing, wide pH stability range (pH 2–12) desired
for various experimental conditions, suitable pore sizes facilitat-
ing efficient molecule translocation and amenability to surface
functionalization. However, despite significant progress in se-
quencing techniques and the emergence of biological nanopores
in commercial applications, achieving high-throughput and ac-
curate DNA sequencing remains a challenge that needs to
be addressed.[231,232]This includes mechanical fragility of the
lipid bilayer, complex fabrication, breakdown under thermal and
Tabl e 6 . Year-wise documentation of ionic and transverse current studies reported on hybrid/heterostructures for DNA sequencing.
Materials Nanodevice Molecular carrier Study Comments Year Refs.
Graphene/h-BN Nanopore (single
modulation
device)
DNA nucleotides DFT +NEGF Detection of DNA nucleotides based on
current-modulation signature
2017 [163]
Graphene/h-BN Nanogap DNA nucleotides DFT +NEGF Unique identification of DNA nucleotides
from current traces within two bias
windows
2017 [228]
Graphene/h-BN Nanopore (double
modulation
device)
DNA nucleotides DFT +NEGF Current modulation control and electrical
readout allow single nucleotides to be
identified
2019 [229]
Graphene/h-BN Nanopore Hachimoji DNA
nucleobases
DFT +NEGF Significant changes in the transmission
function leading to sensitivity in
distinction of up to 80%
2020 [227]
Graphene/h-BN/Graphene Nanopore DNA nucleotides DFT+NEGF Improved sensitivity compared to pure
graphene nanopore
2021 [230]
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (33 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
Tabl e 7 . Comparison of ionic and transverse current methods for next-generation DNA sequencing.
Aspect Ionic Current Blockade Method (Nanopore Sequencing) Transverse Tunneling Current Method
Principle Detection of changes in ionic current as DNA passes through a
nanopore
Detection of changes in tunneling current as DNA interacts with
electrodes
Detection Method Electrolyte solution with nanopore; electrical signal changes due to
DNA translocation
Solid-state device with electrodes; electrical signal changes due to
DNA-electrode interaction
Read Length Potentially long reads, depending on nanopore size and technology Not typically specified, but potential for high resolution down to single
nucleotide level
Resolution limited resolution, may struggle with homopolymer regions and base
calling accuracy
High resolution, potentially down to single nucleotide level
Throughput Moderate to high throughput, depending on the number of
nanopores and sequencing speed
Not typically specified, may be limited by the number of electrodes and
measurement speed
Sample Preparation Relatively simple sample preparation typically involves DNA library
preparation
Sample preparation may involve surface functionalization and
immobilization of DNA
Labeling Requirement Label-free detection; DNA molecules are detected as they pass
through the nanopore
Label-free detection; DNA molecules interact directly with electrodes
Error Rates Prone to errors, including insertions, deletions, and base
misassignments
Error rates may vary but potentially lower due to high-resolution
Commercial Platforms Multiple commercial platforms are available, including Oxford
Nanopore and others
primarily experimental at this stage
Scalability Scalable with multiple nanopores, enabling parallelization for
high-throughput sequencing
Scalable but may require optimization for parallelization
Complexity Relatively straightforward instrumentation and experimental setup More complex instrumentation and experimental setup may be required
Environmental
Sensitivity
Sensitive to environmental factors such as temperature and ionic
strength.
Sensitive to environmental factors such as temperature and humidity.
Potential Applications Whole-genome sequencing, Epigenetics studies, RNA sequencing,
real-time monitoring
High-resolution DNA sequencing, detection of base modifications,
single-molecule analysis
electrical stress, poor fabrication yield, high translocation rate,
and low sensitivity that need to be solved.
In this direction, enzyme-integrated bionanopores have been
successfully raised as a promising solution.[26,233–236]The key ad-
vantages of enzyme-integrated bionanopore lie in improving the
signal-to-noise ratio, reduction in the translocation kinetics varia-
tion, controlled translocation, and helicase activity helpful in un-
winding dsDNA or DNA-RNA duplexes. Apart from the devel-
opment of enzyme-integrated bionanopores, the above-discussed
NGS architectures: solid-state nanopore, nanogap, nanochannel,
hybrid nanopores, and heterostructures have also been realized
in experimental and theoretical arenas. However, the existing
DNA sequencing methods are still gaining their potential from
the perspective of cost and base-by-base read efficiency. Herein,
we have summarized the major challenges of DNA sequencing,
followed by the relevant discussion on how to resolve them.
4.1. Limitations of the Ionic Current Blockade Method
Although several experimental and theoretical studies have been
done on ionic current-based DNA sequencing, none of the stud-
ies demonstrate a device that can detect only one nucleotide
at a particular time. Traditional biological nanopore sequenc-
ing still needs advancement as it cannot detect the single bases
separated by 0.4 nm.[25]Biological nanopores having channels
shorter than 5 nm yet need to be realized to identify the sin-
gle nucleotide.[25]Currently, the major challenge associated with
nanopore sequencing is controlling the high translocation speed
of DNA.[99,237,238]The high velocity of DNA issues affects DNA
translocation dynamics and widens the statistical data measured
in the sequencing results.[49]As DNA translocates through a
nanopore in an ionic solution under an applied electric field,
its dwell time within the nanopore is typically less than one nu-
cleotide per microsecond (1 base μs1).[118,239]This short dwell
time presents difficulties in accurately detecting the ionic-current
signal associated with each single nucleotide using commercially
available amplifiers. To ensure satisfactory electric signal record-
ing, the optimal DNA translocation speed within a nanopore
should fall within the range of 1–100 base pairs per millisecond
(1–100 bp ms1).[240]
Various methods have been utilized in recent years to decrease
the translocation rate of DNA, including the immobilization
of DNA polynucleotides using streptavidin, the creation of a
rotaxane, and the use of enzymes as binding agents.[234–236,241]
The translocation speed of ssDNA in biological nanopores
is slower (1ntμS1) compared to solid-state nanopores
(10 nt μS1). Specifically, phi29 DNA polymerase is reported
to have decreased translocation velocity by four orders of mag-
nitude with ratcheting DNA during translocation through the
biological nanopore.[242,243]Atomic force microscopy has also
been developed to control the translocation rate.[244,245]Along
with that, various strategies have also been reported to slow
down DNA translocation and improve the temporal resolution of
ionic current-based solid-state nanopores. The optimization of
experimental physical conditions can be achieved by modulating
influencing factors: electrolyte viscosity,[246–248]temperature,[239]
electrolyte concentration,[249,250]lowering applied bias
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (34 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
voltage,[246,251]small size/double nanopore,[77,252 ]and surface
charge density of nanopore.[99]
Kowalczyk et al. reported that different ions exhibit different
binding strengths with the nucleotides.[249]When the counte-
rion size is decreased, the translocation time of DNA molecules
through a solid-state nanopore can increase dramatically. With
LiCl ion solution, the translocation speed of both ssDNA and
dsDNA has been found to be significantly reduced. This is be-
cause the smaller size of Li+ions results in stronger interactions
with the negatively charged DNA molecules, slowing down their
movement through the nanopore. This effect can be useful in
controlling the translocation rate of DNA, as it provides a way to
regulate the speed and orientation of the DNA molecule. How-
ever, it can also pose challenges in DNA sequencing applications,
where it is important to maintain a consistent and predictable
translocation rate. Therefore, careful selection of the electrolyte
solution and optimization of the experimental conditions are re-
quired to ensure reliable and accurate DNA sequencing results.
Moreover, Feng et al. reported that a viscosity gradient system
based on room-temperature ionic liquids (RTILs) can control the
dynamics of DNA and slow down DNA translocation through
MoS2nanopore.[247 ]With the introduction of high viscosity, the
nucleotide translocation velocity falls within an optimal speed
(1–50 nt ms1) while maintaining a high signal-to-noise ratio
(SNR >10).
Another major challenge is the low level of signals in the or-
der of pico-ampere. Low ionic blockage current decreases the ef-
ficiency of accurate nucleobase identification.[253,254,52,255–257]The
high-quality resolution with a high signal-to-noise ratio can be
achieved by considering the size of solid-state nanopores compa-
rable with DNA nucleotide diameter, and the nanopore thickness
should be within the range of 0.4–0.6 nm. In biological pores, the
major issue is the pore thickness, which is around 5 nm, consid-
erably longer, and can easily accommodate more than 10–15 nu-
cleotides, resulting in a low signal-to-noise ratio and a high error
rate of base identification.[258]In this regard, it is nearly impos-
sible to achieve single-nucleobase resolution with the biological
nanopore. A comparative study of current noise between biologi-
cal and solid-state nanopores suggests that solid-state nanopores
are more trustworthy for a large signal-to-noise ratio compared
to their biological counterparts.[259]
4.2. Limitations of Transverse Current-Based DNA Sequencing
The challenges of the transverse current-based approach can be
broadly classified into four categories[260]: i) optimization of a
voltage bias and solution conditions, as it is really difficult to pre-
dict exactly the electronic response of the detector with different
nucleotides in the solvent conditions, ii) fabrication of the device
enabling each base to assume a reproducible orientation and dis-
tances on the collector probe as minor variations in the orienta-
tions and positions of atoms can have a significant impact on tun-
neling current, iii) translocation of DNA must be controlled for
minimum read-out error, and iv) it still remains unclear whether
the tunneling current calculations can provide sufficient contrast
to base discrimination and signal characteristics of the gap be-
tween the nucleobases.
In addition to the aforementioned challenges, there are
other significant obstacles associated with the transverse current
method. Two of these major challenges are controlling the ve-
locity and orientation of DNA during translocation and achiev-
ing a high signal-to-noise ratio in the current signals. The sen-
sitivity of solid-state materials needs to be improved to discrim-
inate biomolecules of similar sizes. Electronic transport proper-
ties of probe-DNA interactions need to be optimized. Efforts are
still under consideration to fabricate more efficient nanopore and
nanogap-based transverse tunneling current detectors.
Single-walled carbon nanotubes (SWCNTs) have been pro-
posed as a promising method to address the challenges asso-
ciated with DNA translocation-based detection.[25,261]One ad-
vantage of SWCNTs is their ability to interact with DNA in
a nucleobase-specific manner. This means that different DNA
bases can be identified based on their distinct interaction with
the SWCNT surface, allowing for more accurate and reliable se-
quencing. In addition, the translocation rate of DNA through a
SWCNT nanopore can be controlled by modulating the tempera-
ture, ionic strength, or voltage bias. This provides a way to regu-
late the speed and orientation of the DNA molecule, enabling bet-
ter control over the sequencing process. Base-specific hydrogen
bonding interactions between chemically modified metal elec-
trodes and nucleobases may also be helpful in resolving the above
issues.
4.3. Low Signal-To-Noise Ratio
Noise is an unwanted signal that shadows the actual signals,
decreasing the efficiency of the device and signal-to-noise ratio
(SNR) during detection. Hence, there is a substantial demand
for strategies to reduce noise. It is reported that even though
biological pores generally exhibit lower noise, solid-state pores
have substantially higher signal-to-noise (SNR) than biological
nanopores.[259]It can be attributed to the higher currents that
solid-state systems offer, along with relatively low high-frequency
noise.
In solid-state nanopores, noise originates from the charge fluc-
tuations at the nanopore surface, fluctuations of the total num-
ber of charge carriers, and nanometer-sized gaseous bubbles in-
side the pore during the translocation of DNA molecules.[259,262]
Based on its frequency range, noise can be divided into four cate-
gories: a) low-frequency noise, b) thermal current noise, c) high-
frequency dielectric noise, and d) capacitive noise.[263–265]Con-
siderable research has been devoted to unraveling the origins
of noise in both biological and solid-state nanopores.[263,266,267]
To reduce the low-frequency noise, researchers are using more
mechanically stable nanopores and thicker membranes for the
studies. Graphene membranes coated with a TiO2layer or a
multi-layered graphene-Al2O3stack structure with a thickness
of approximately 20 nm are reported to be more effective in re-
ducing electrical noise than pure graphene nanopores in single
molecule-level DNA sequencing.[268]High-frequency noise is re-
ported to be reduced by integrating complementary metal-oxide
semiconductor (CMOS) preamplifiers with solid-state nanopores
in thin silicon nitride membranes.[269]Further, signal-to-noise
ratios (SNRs) in nanopore and ion channel recordings can be
improved with wavelet-denoising techniques.[270]With the ad-
vancement of ML techniques, deep learning methods are also
being used for denoising electric signals with convolutional
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (35 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
auto-encoding neural networks.[271]This network is specifically
engineered to iteratively analyze and minimize disparities be-
tween two waveforms by employing gradient descent optimiza-
tion. Utilizing a permittivity gradient approach, Tsutsui et al.
have reported an enhancement in the single-molecule sensitiv-
ity of nanopores by amplifying ionic current signals while mini-
mizing interference with other features.[272]Drawing inspiration
from concepts in both electronics and biology, an abiotic ionic
circuit demonstrates signal amplification by transforming small
ionic signals into amplified ionic outputs.[273]In the case of trans-
verse tunneling methods, the insulator-protected mechanically-
controllable break junctions (MCBJs) demonstrated fast tunnel-
ing current measurements and improved the signal-to-noise ratio
in the electrical single-molecule detection.[274]
The sensing performance of 2D materials can also be hindered
by noise resulting from tunneling current across the transmem-
brane sheet. However, till now, it is a less explored area. Accord-
ing to a recent experimental study by Graf et al., the transverse
current demonstrated a 40% improvement in signal-to-noise ra-
tio compared to the ionic current in simultaneous measurements
using a MoS2nanopore.[170]This suggests that the transverse cur-
rent approach has potential advantages over the ionic current ap-
proach for DNA sequencing.
4.4. To Achieve Long Readability
One of the potential challenges of DNA sequencing is achiev-
ing long reads for significant nucleotide identification. There
are some fluctuations in translocation kinetics owing to device-
surface interaction, which is the key challenge to achieve long
reads. In addition, the translocation of DNA in an unfolded state
is also a key requirement for more reliable DNA sequencing.
The nanopore sequencing turned out to be the most promising
in achieving long readability, as the nanopore sensor reads the
DNA molecule base by base, and the accuracy of nucleobase at
one instance of time is independent of the prior history of the
system. The MinION device of Oxford Nanopore indeed repre-
sented a significant advancement in portable DNA sequencing
technology. The MinION device, resembling a USB drive in ap-
pearance, offered real-time DNA sequencing capabilities by uti-
lizing nanopore technology.[24,275]
Nevertheless, some limitations still need to be explored in the
experimental studies, such as avoiding shearing during sample
preparation and threading exceptionally long molecules through
the pore. The experimental DNA sequencing analysis of nanogap
and nanochannel devices with the transverse tunneling current
method is still in the infant stage. A more detailed exploration
of these specific devices can shed light on their long-strand DNA
sequencing capabilities.
4.5. Challenges in the Solid-State Device Fabrication
Several strategies have been employed for solid-state nanopore
fabrication, such as ion-beam sculpting, electron-beam drilling,
and atomic layer deposition, but a potential method capable of
generating a large number of uniform solid-state nanopores with
suitable diameter ranges from 1.5 to 2.0 nm remains a daunting
challenge. The problem with focused ion beam drilling is that the
resolution is insufficient to fabricate nanopores with a diameter
smaller than 10 nm.[276]Similarly, achieving precise control over
the geometry and size of nanopores is a significant challenge, as
it is essential for ensuring the reproducibility of nanopore sens-
ing experiments. Solid-state nanopore fabrication with thin-film
growth is unscalable, costly, and time-consuming. There are ma-
jor difficulties as well in the fabrication of hybrid nanopores, such
as the condition of insertion with controlled precision and prepa-
ration of an inner volume of the nanopore.
Moreover, the reproducible and controlled fabrication of
nanogap electrodes with a sub-5nm gap remains a formidable
challenge. The potential techniques for the fabrication of
nanogap are reported to be focused ion beam (FIB) milling, pho-
tolithography, and electron-beam lithography (EBL).[277]In the
case of nanoribbons, one major challenge is the fabrication of
a device with a width of 1 nm, which is important for single-
base resolution.[278]An additional challenge is the controlled
alignment of nanoribbons within a nanofluidic channel. This
issue may be resolved by nanowire lithography[279]and the un-
zipping of carbon nanotubes.[280]Nanopore fabrication utilizing
dielectric-controlled breakdown (CBD) of thin membranes di-
rectly in an aqueous solution is currently undergoing rapid ad-
vancement and active investigation to enhance the scalability of
solid-state nanopores.[281]The CBD method presents three signif-
icant advantages: a) cost-effective in situ fabrication, b) the abil-
ity to fabricate planar nanopores within existing nanostructures,
and c) high accessibility and ease of handling.[282,283]Nonetheless,
significant hurdles persist, particularly in controlling the precise
location of pore formation and establishing scalable manufactur-
ing processes. Addressing these challenges is pivotal in the seam-
less integration of solid-state pore-based platforms into DNA se-
quencing applications.
Lastly, a strong collaboration work in both experimental and
computational research can reduce lots of effort, time, and cost
by providing simulated information for a more precise selec-
tion of devices and their applications for DNA sequencing. For
a more effective interpretation of experimental results with com-
putational reports, transmembrane-applied bias voltages should
not be higher than the experimental voltage.
4.6. Orientational Fluctuations of the DNA Molecule
Environmental factors mostly lead to conformational, structural,
and orientational fluctuations in DNA nucleotides inside the
nanopore. DNA translocation is stochastic, and the fluctuation
occurs due to translocation kinetics, which gives rise to a broad
distribution of dwell times. These variations due to translocation
kinetics cause uncertainties in the number of DNA nucleotides
translocating through nanopores. As a result of these fluctua-
tions, there can be significant variations in the measured electric
current signals, leading to a higher potential for read-out errors.
In biological nanopores, the problem of stochastic translocation
kinetics can be reduced using a processive enzyme,[74]while in
solid-state nanopores, a possible way to control the nucleobase
motions is the use of a multilayered membrane transistor with
a motion-control electrode layer that can facilitate decrease in
stochastic fluctuations of trapped DNA molecule.[158]
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (36 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
In nanopores, any fluctuation in the DNA molecule is critical
to the transverse tunneling current approach as it can consider-
ably influence the coupling strength by affecting the eigenstates
of DNA nucleotides with the nanopore/nanogap edges and af-
fecting the distribution of the recorded current values, resulting
in overlap in the signature signal of DNA nucleotides. Most of the
previous computational studies either excluded the environmen-
tal effects or simulated homogenous DNA strands by considering
only arbitrary structural and conformational orientations.[148,284]
In a realistic simulation, the effects of embedded material, tem-
perature, solvent, and salt concentration come into the picture,
which triggers thermal fluctuation. The most promising architec-
ture for reducing the orientational fluctuation is the nanochan-
nel. There are several tactical methods that have been pro-
posed to reduce conformational instabilities, and stochastic nu-
cleobase motions and resolve the signal overlap between differ-
ent DNA nucleobases. Some important methods are the func-
tionalization of nanogap electrodes,[180,285,204]labeling of DNA
nucleotides,[189,286]and double nanopore,[287,288 ]among others.
5. Conclusions
Solid-state material-based DNA sequencing has emerged as a
powerful paradigm with the potential to transform comprehen-
sive genomic and medical research. Significant efforts are be-
ing made to achieve an ultrafast, cost-effective, and reproducible
DNA sequencing technique that recognizes a single base within
a DNA strand. Though biological nanopores with ionic cur-
rent methods have already been commercialized, solid-state 2D
materials-based electrical detection techniques will be the key to
achieve single nucleotide resolution. In the present review, we
have provided a comprehensive discussion of all the prominent
aspects of various DNA sequencing techniques proposed so far,
including their challenges, opportunities, and future directions.
The review highlights the ionic and transverse tunneling current
methods involving nanopores, nanogaps, nanochannels, and hy-
brid/heterostructures for DNA sequencing.
Despite currently being the advanced method, the ionic cur-
rent approach suffers from major bottlenecks that need to be
resolved, such as scalability, slowing down DNA translocation
speed, resolving nucleotides signal overlap, reducing nucle-
obases stochastic motion, and low signal-to-noise ratio. Recently,
transverse tunneling current detection technique is also moving
fast in the direction of sequencing biomolecules. The technique
has several advantages, such as the potential to sequence many
times faster than the ionic current method, single-base resolu-
tion, high readability, and fidelity, along with lower noise levels.
Moreover, in tunneling current measurements, the recognition
can be achieved without complex enzymatic processing. How-
ever, the technique is extremely sensitive to molecule sizes, com-
position, and thermal fluctuations in an aqueous solution. Cur-
rently, as the promising frontier, more theoretical and experimen-
tal reports are needed to comment more on the potential of the
transverse current method for ultrafast DNA sequencing.
We believe that solid-state 2D materials are more promis-
ing for single-molecule DNA detection. Solid-state 2D materi-
als have the advantages of long-term stability, high surface-to-
volume ratio, atomically thin membranes, higher durability, and
superior mechanical and electrical properties. Till now, graphene
is reported to be the most promising material to achieve single
nucleotide recognition. Nevertheless, graphene nanostructures
have certain demerits, such as hydrophobic interactions, pore-
clogging, and high noise levels, which need to be resolved. We
anticipate that the commercialization of solid-state 2D materials
based on nanopore/nanogap sequencing techniques may appear
in the near future for decoding DNA with long readability in a
very short time, reasonable cost, and high accuracy.
A remaining question is which nanostructure (nanopore,
nanogap, and nanochannel) will provide sufficient distinguisha-
bility among the DNA nucleotides. At present, the technique that
is commercialized is the nanopore technique utilizing an ionic
current approach. However, the technique has certain shortcom-
ings, including irreversible pore clogging, low signal-to-noise ra-
tio, high translocation speed, and conformational fluctuations,
which need to be overcome so that more accurate sequencing
can be achieved. The nanogap sequencing technique is based
on the quantum tunneling current method. Interestingly, in this
method, each targeted nucleotide exhibits a current signal inde-
pendent of the neighbor nucleotide. Nevertheless, the drawback
is the possibility of significant fluctuations in the tunneling cur-
rent. In the nanochannel technique, the merit is controlled orien-
tational fluctuation, which may provide better signal readability.
The demerit of the nanochannel is its ultra-thin fabrication, lim-
iting its capacity to adsorb and detect only a single molecule at a
time. After thoroughly analyzing all the prominent aspects of cur-
rently available architectures, we believe that both nanopore and
nanogap architectures are promising for single nucleotide iden-
tification. Moreover, machine learning integrated nanopore de-
vices with transverse current method has shown potential for pre-
diction of nucleotide fingerprint conductance signals from train-
ing with single nucleotide datasets and helps in the classification
of overlapped complex, noisy nucleotide signals with high accu-
racy.
Despite all the achievements, there is an urgent need to dis-
cover new sensing methods, such as alternative read-out schemes
and other 2D solid-state materials, which can have the potential
to suppress stochastic nucleotide motions, control DNA translo-
cation rate, conformational fluctuations, and resolve the signal
overlap. Considering the incredible progress in nanopore DNA
sequencing, there is no doubt that it is currently the most promis-
ing sequencing technique, even capable of sequencing the hu-
man genome in a very fast and cost-effective manner. There
are still lots of things that need to be explored further in this
field. Furthermore, fabrication techniques need to be improved
to achieve uniform nanostructures and enable high-throughput
DNA sequencing.
6. Future Perspectives
The future of DNA sequencing research hinges on the pursuit
of their affordability, rapidity, and precision. Remarkable strides
have been made in the realm of solid-state nanopores. However,
it is worth noting that commercialization of DNA sequencing
is only achieved by biological nanopores with the ionic block-
ade current approach. For a successful transition from biologi-
cal nanopores to solid-state material-based devices into commer-
cial viability for DNA sequencing, an emphasis on extensive ex-
perimental investigations under realistic conditions is crucial.
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (37 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
Moreover, the transverse tunneling current method using solid-
state 2D materials has been emerging as a potential strategy for
single molecule DNA sequencing. Several theoretical investiga-
tions emphasize the promise of this approach. However, more ex-
perimental studies can provide better insights into the complex-
ities and accuracy of identification with the transverse current
approach, which is still in the nascent stage of technological evo-
lution. The solid-state 2D material-based sequencing paradigm
warrants a patient outlook, drawing parallels to the evolution-
ary trajectory of biological nanopores, which required nearly two
decades of research to achieve commercialization. From the ex-
tensive experimental and computational analysis, it is evident
that amongst the solid-state materials, the commercialization of
graphene nanopores can be achieved in the future. Recently, the
integration of machine learning and artificial intelligence tech-
nologies into sequencing devices has helped signal processing,
denoising large complex data, and facilitating accurate nucleotide
identification and classification through the extraction of elec-
tronic signal information. From our understanding, we feel the
future of DNA sequencing will be taken over by the application
of artificial intelligence and data-driven prediction of nucleotide
sequencing. Additionally, by leveraging the lessons learned from
biological pore commercialization and harnessing the capabili-
ties of machine learning, ionic current-based DNA sequencing
stands poised for substantial advancement.
Acknowledgements
M.K.J. and S.M. contributed equally to this work. This work was sup-
ported by grants DST-SERB (project number: CRG/2022/000836), CSIR
(project number: 01(3046)/21/EMR-II), and BRNS (project number: 2023-
BRNS/12356). R.L.K. and M.K.J. acknowledge MHRD, and S.M. acknowl-
edge UGC for the research fellowship, respectively.
Conflict of Interest
The authors declare no conflict of interest.
Keywords
DNA sequencing, ionic current, machine learning, nanopore, transverse
current
Received: February 13, 2024
Revised: April 5, 2024
Published online:
[1] P. S. G. Chain, D. V. Grafham, R. S. Fulton, M. G. FitzGerald, J.
Hostetler, D. Muzny, J. Ali, B. Birren, D. C. Bruce, C. Buhay, J. R. Cole,
Y. Ding, S. Dugan, D. Field, G. M. Garrity, R. Gibbs, T. Graves, C. S.
Han, S. H. Harrison, S. Highlander, P. Hugenholtz, H. M. Khouri,
C. D. Kodira, E. Kolker, N. C. Kyrpides, D. Lang, A. Lapidus, S. A.
Malfatti, V. Markowitz, T. Metha, et al., Science 2009,326, 236.
[2] S. White, J. W. Szewczyk, J. M. Turner, E. E. Baird, P. B. Dervan, Na-
ture 1998,391, 468.
[3] F. Sanger, A. R. Coulson, J. Mol. Biol. 1975,94, 441.
[4] S. Nurk, S. Koren, A. Rhie, M. Rautiainen, A. V. Bzikadze, A.
Mikheenko, M. R. Vollger, N. Altemose, L. Uralsky, A. Gershman,
S. Aganezov, S. J. Hoyt, M. Diekhans, G. A. Logsdon, M. Alonge,
S. E. Antonarakis, M. Borchers, G. G. Bouffard, S. Y. Brooks, G. V.
Caldas, N.-C. Chen, H. Cheng, C.-S. Chin, W. Chow, L. G. de Lima, P.
C. Dishuck, R. Durbin, T. Dvorkina, I. T. Fiddes, G. Formenti, et al.,
Science 2022,376, 44.
[5] J. Shendure, H. Ji, Nat. Biotechnol. 2008,26, 1135.
[6] S. Behjati, P. S. Tarpey, Arch Dis Child Educ Pract Ed 2013,98, 236.
[7] S. Goodwin, J. D. McPherson, W. R. McCombie, Nat. Rev. Genet.
2016,17, 333.
[8] D.M.Church,Science 2022,376, 34.
[9] W. Yang, F. Yan, Radiology 2020,295,E3.
[10] G. Seo, G. Lee, M. J. Kim, S.-H. Baek, M. Choi, K. B. Ku, C.-S. Lee, S.
Jun, D. Park, H. G. Kim, S.-J. Kim, J.-O. Lee, B. T. Kim, E. C. Park, S.
I. Kim, ACS Nano 2020,14, 5135.
[11] X. Chen, Y. Kang, J. Luo, K. Pang, X. Xu, J. Wu, X. Li, S. Jin, Front Cel l
Infect Microbiol 2021,11.
[12] M.-M. Aynaud, J. J. Hernandez, S. Barutcu, U. Braunschweig, K.
Chan, J. D. Pearson, D. Trcka, S. L. Prosser, J. Kim, M. Barrios-
Rodiles, M. Jen, S. Song, J. Shen, C. Bruce, B. Hazlett, S. Poutanen,
L. Attisano, R. Bremner, B. J. Blencowe, T. Mazzulli, H. Han, L.
Pelletier, J. L. Wrana, Nat. Commun. 2021,12, 1405.
[13] J. Shendure, S. Balasubramanian, G. M. Church, W. Gilbert, J.
Rogers, J. A. Schloss, R. H. Waterston, Nature 2017,550, 345.
[14] M. Jain, S. Koren, K. H. Miga, J. Quick, A. C. Rand, T. A. Sasani,
J. R. Tyson, A. D. Beggs, A. T. Dilthey, I. T. Fiddes, S. Malla, H.
Marriott, T. Nieto, J. O’Grady, H. E. Olsen, B. S. Pedersen, A. Rhie,
H. Richardson, A. R. Quinlan, T. P. Snutch, L. Tee, B. Paten, A. M.
Phillippy, J. T. Simpson, N. J. Loman, M. Loose, Nat. Biotechnol.
2018,36, 338.
[15] C. G. Brown, J. Clarke, Nat. Biotechnol. 2016,34, 810.
[16] A. D. Tyler, L. Mataseje, C. J. Urfano, L. Schmidt, K. S. Antonation,
M. R. Mulvey, C. R. Corbett, Sci. Rep. 2018,8, 10931.
[17] M. L. Metzker, Nat. Rev. Genet. 2010,11, 31.
[18] S. J. Heerema, C. Dekker, Nat. Nanotechnol. 2016,11, 127.
[19] D. Deamer, M. Akeson, D. Branton, Nat. Biotechnol. 2016,34,518.
[20] N. Yang, X. Jiang, Carbon 2017,115, 293.
[21] H. Qiu, W. Zhou, W. Guo, ACS Nano 2021,15, 18848.
[22] L. Xue, H. Yamazaki, R. Ren, M. Wanunu, A. P. Ivanov, J. B. Edel,
Nat. Rev. Mater. 2020,5, 931.
[23] M. Rahman, M. J. N. Sampad, A. Hawkins, H. Schmidt, Lab Chip
2021,21, 3030.
[24] Y. He, M. Tsutsui, Y. Zhou, X.-S. Miao, NPG Asia Mater 2021,13,1.
[25] D. Branton, D. W. Deamer, A. Marziali, H. Bayley, S. A. Benner, T.
Butler, M. Di Ventra, S. Garaj, A. Hibbs, X. Huang, S. B. Jovanovich,
P. S. Krstic, S. Lindsay, X. S. Ling, C. H. Mastrangelo, A. Meller, J. S.
Oliver, Y. V. Pershin, J. M. Ramsey, R. Riehn, G. V. Soni, V. Tabard-
Cossa, M. Wanunu, M. Wiggin, J. A. Schloss, Nat. Biotechnol. 2008,
26, 1146.
[26] Y. Wang, Y. Zhao, A. Bollas, Y. Wang, K. F. Au, Nat. Biotechnol. 2021,
39, 1348.
[27] J. M. Heather, B. Chain, Genomics 2016,107,1.
[28] A. M. Giani, G. R. Gallo, L. Gianfranceschi, G. Formenti, Comput.
Struct. Biotechnol. J 2020,18,9.
[29] Milestones in Genomic Sequencing.
[30] R. W. Holley, J. Apgar, G. A. Everett, J. T. Madison, M. Marquisee, S.
H. Merrill, J. R. Penswick, A. Zamir, Science 1965,147, 1462.
[31] F. Sanger, G. G. Brownlee, B. G. Barrell, J. Mol. Biol. 1965,13, 373.
[32] R. Wu, A. D. Kaiser, J. Mol. Biol. 1968,35, 523.
[33] W. Gilbert, A. Maxam, Proc. Natl. Acad. Sci. USA 1973,70, 3581.
[34] F. Sanger, S. Nicklen, A. R. Coulson, Proc. Natl. Acad. Sci. USA 1977,
74, 5463.
[35] A. M. Maxam, W. Gilbert, Proc. Natl. Acad. Sci. USA 1977,74, 560.
[36] S. Anderson, Nucleic Acids Res 1981,9, 3015.
[37] B.Canard,R.S.Sarfati,Gene 1994,148,1.
[38] P. Nyren, B. Pettersson, M. Uhlen, Anal. Biochem. 1993,208, 171.
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (38 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
[39] F. S. Collins, A. Patrinos, E. Jordan, A. Chakravarti, R. Gesteland, L.
Walters, Science 1998,282, 682.
[40] R. F. Service, Science 2006,311, 1544.
[41] R. A. Gibbs, Nat. Rev. Genet. 2020,21, 575.
[42] L.T.C.França,E.Carrilho,T.B.L.Kist,QRevBiophys2002,35,
169.
[43] E. E. Schadt, S. Turner, A. Kasarskis, Hum. Mol. Genet. 2010,19,
R227.
[44] J. F. Thompson, P. M. Milos, Genome Biol. 2011,12, 217.
[45] E. Check Hayden, Nature 2009,https://doi.org/10.1038/news.2009.
86.
[46] C. Lippert, D. Heckerman, XRDS 2015,21, 24.
[47] M. Zwolak, M. Di Ventra, Rev. Mod. Phys. 2008,80, 141.
[48] L. Gasparyan, I. Mazo, V. Simonyan, F. Gasparyan, Open J. Biophys.
2019,09, 169.
[49] H. S. Kim, Y.-H. Kim, Biosens. Bioelectron. 2015,69, 186.
[50] M. Wanunu, Phys Life Rev 2012,9, 125.
[51] Y.-L. Ying, J. Zhang, R. Gao, Y.-T. Long, Angew Chem Int Ed Engl 2013,
52, 13154.
[52] J. J. Kasianowicz, E. Brandin, D. Branton, D. W. Deamer, Proc. Natl.
Acad. Sci. USA 1996,93, 13770.
[53] L. Song, M. R. Hobaugh, C. Shustak, S. Cheley, H. Bayley, J. E.
Gouaux, Science 1996,274, 1859.
[54] M. Akeson, D. Branton, J. J. Kasianowicz, E. Brandin, D. W. Deamer,
Biophys. J. 1999,77, 3227.
[55] D. Stoddart, A. J. Heron, J. Klingelhoefer, E. Mikhailova, G. Maglia,
H. Bayley, Nano Lett. 2010,10, 3633.
[56] F. Yao, X. Peng, Z. Su, L. Tian, Y. Guo, X. Kang, Anal. Chem. 2020,
92, 3827.
[57] A. Meller, L. Nivon, E. Brandin, J. Golovchenko, D. Branton, Proc.
Natl. Acad. Sci. USA 2000,97, 1079.
[58] W.Sung,P.J.Park,Phys.Rev.Lett.1996,77, 783.
[59] D. Stoddart, A. J. Heron, E. Mikhailova, G. Maglia, H. Bayley, Proc.
Natl. Acad. Sci. USA 2009,106, 7702.
[60] D. Stoddart, G. Maglia, E. Mikhailova, A. J. Heron, H. Bayley, Angew.
Chem. 2010,122, 566.
[61] G. Maglia, M. Henricus, R. Wyss, Q. Li, S. Cheley, H. Bayley, Nano
Lett. 2009,9, 3831.
[62] M. Bates, M. Burns, A. Meller, Biophys. J. 2003,84, 2366.
[63] Y. Astier, O. Braha, H. Bayley, J. Am. Chem. Soc. 2006,128,
1705.
[64] J. Clarke, H.-C. Wu, L. Jayasinghe, A. Patel, S. Reid, H. Bayley, Nat.
Nanotechnol. 2009,4, 265.
[65] D. Japrung, M. Henricus, Q. Li, G. Maglia, H. Bayley, Biophys. J.
2010,98, 1856.
[66] C. Tropini, A. Marziali, Biophys. J. 2007,92, 1632.
[67] Y. Feng, Y. Zhang, C. Ying, D. Wang, C. Du, Genomics, Proteomics
Bioinf. 2015,13,4.
[68] C. Heinz, H. Engelhardt, M. Niederweis, J. Biol. Chem. 2003,278,
8678.
[69] M. Faller, M. Niederweis, G. E. Schulz, Science 2004,303, 1189.
[70] I. M. Derrington, T. Z. Butler, M. D. Collins, E. Manrao, M. Pavlenok,
M. Niederweis, J. H. Gundlach, Proc. Natl. Acad. Sci. USA 2010,107,
16060.
[71] S. Bhattacharya, I. M. Derrington, M. Pavlenok, M. Niederweis, J. H.
Gundlach, A. Aksimentiev, ACS Nano 2012,6, 6960.
[72] M. Yu, W. Si, T. Zeng, C. Chen, X. Lin, Z. Ji, F. Guo, Y. Li, J. Sha, Y.
Dong, J. Phys. Chem. Lett. 2021,12, 9132.
[73] S. K. Tabatabaei, B. Pham, C. Pan, J. Liu, S. Chandak, S. A. Shorkey,
A. G. Hernandez, A. Aksimentiev, M. Chen, C. M. Schroeder, O.
Milenkovic, Nano Lett. 2022,22, 1905.
[74] 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.
[75] A. H. Laszlo, I. M. Derrington, B. C. Ross, H. Brinkerhoff, A. Adey,
I. C. Nova, J. M. Craig, K. W. Langford, J. M. Samson, R. Daza, K.
Doering, J. Shendure, J. H. Gundlach, Nat. Biotechnol. 2014,32, 829.
[76] S. Bhattacharya, J. Yoo, A. Aksimentiev, ACS Nano 2016,10, 4644.
[77] D. Wendell, P. Jing, J. Geng, V. Subramaniam, T. J. Lee, C.
Montemagno, P. Guo, Nat. Nanotechnol. 2009,4, 765.
[78] Y. Xiang, M. C. Morais, A. J. Battisti, S. Grimes, P. J. Jardine, D. L.
Anderson, M. G. Rossmann, EMBO J. 2006,25, 5229.
[79] J. Geng, S. Wang, H. Fang, P. Guo, ACS Nano 2013,7, 3315.
[80] C. Cao, J. Yu, Y.-Q. Wang, Y.-L. Ying, Y.-T. Long, Anal. Chem. 2016,
88, 5046.
[81] K.-B. Jeong, K. Luo, H. Lee, M.-C. Lim, J. Yu, S.-J. Choi, K.-B. Kim,
T.-J. Jeon, Y.-R. Kim, ACS Nano 2019,13, 1719.
[82] P. Goyal, P. V. Krasteva, N. V. Gerven, F. Gubellini, I. V. D. Broeck, A.
Troupiotis-Tsaïlaki, W. Jonckheere, G. Péhau-Arnaudet, J. S. Pinkner,
M. R. Chapman, S. J. Hultgren, S. Howorka, R. Fronzes, H. Remaut,
Nature 2014,516, 250.
[83] L. S. Robinson, E. M. Ashman, S. J. Hultgren, M. R. Chapman, Mol.
Microbiol. 2006,59, 870.
[84] B. Cao, Y. Zhao, Y. Kou, D. Ni, X. C. Zhang, Y. Huang, Proc. Natl.
Acad. Sci. USA 2014,111, E5439.
[85] P.Lv,Y.Yang,S.Li,C.S.Tan,D.Ming,Electrochem. Sci. Adv. 2022,
2, e2100119.
[86] A. Singh, Nat. Methods 2020,17, 874.
[87] A. J. Storm, J. H. Chen, X. S. Ling, H. W. Zandbergen, C. Dekker, Nat.
Mater. 2003,2, 537.
[88] B. M. Venkatesan, B. Dorvel, S. Yemenicioglu, N. Watkins, I. Petrov,
R. Bashir, Adv. Mater. 2009,21, 2771.
[89] M. J. Kim, M. Wanunu, D. C. Bell, A. Meller, Adv. Mater. 2006,18,
3149.
[90] J. Li, Nature 2001,412.
[91] J. Li, M. Gershow, D. Stein, E. Brandin, J. A. Golovchenko, Nat.
Mater. 2003,2, 611.
[92] P. Chen, J. Gu, E. Brandin, Y.-R. Kim, Q. Wang, D. Branton, Nano
Lett. 2004,4, 2293.
[93] C. A. Merchant, K. Healy, M. Wanunu, V. Ray, N. Peterman, J. Bartel,
M. D. Fischbein, K. Venta, Z. Luo, A. T. C. Johnson, M. Drndi´
c, Nano
Lett. 2010,10, 2915.
[94] S. Garaj, S. Liu, J. A. Golovchenko, D. Branton, Proc. Natl. Acad. Sci.
USA 2013,110, 12192.
[95] J. B. Heng, C. Ho, T. Kim, R. Timp, A. Aksimentiev, Y. V. Grinkova,
S. Sligar, K. Schulten, G. Timp, Biophys. J. 2004,87, 2905.
[96] B. M. Venkatesan, A. B. Shah, J.-M. Zuo, R. Bashir, Adv. Funct. Mater.
2010,20, 1266.
[97] B. M. Venkatesan, J. Polans, J. Comer, S. Sridhar, D. Wendell, A.
Aksimentiev, R. Bashir, Biomed. Microdevices 2011,13, 671.
[98] Z. Chen, Y. Jiang, D. R. Dunphy, D. P. Adams, C. Hodges, N. Liu, N.
Zhang, G. Xomeritakis, X. Jin, N. R. Aluru, S. J. Gaik, H. W. Hillhouse,
C. Jeffrey Brinker, Nature Mater 2010,9, 667.
[99] Y. He, M. Tsutsui, C. Fan, M. Taniguchi, T. Kawai, ACS Nano 2011,
5, 5509.
[100] J. C. Meyer, A. K. Geim, M. I. Katsnelson, K. S. Novoselov, T. J. Booth,
S. Roth, Nature 2007,446, 60.
[101] G. Tsoukleri, J. Parthenios, K. Papagelis, R. Jalil, A. C. Ferrari, A. K.
Geim, K. S. Novoselov, C. Galiotis, Small 2009,5, 2397.
[102] G. F. Schneider, S. W. Kowalczyk, V. E. Calado, G. Pandraud, H. W.
Zandbergen, L. M. K. Vandersypen, C. Dekker, Nano Lett. 2010,10,
3163.
[103] S. Garaj, W. Hubbard, A. Reina, J. Kong, D. Branton, J. A.
Golovchenko, Nature 2010,467, 190.
[104] D. B. Wells, M. Belkin, A. Aksimentiev, 2012.
[105] G. F. Schneider, Q. Xu, S. Hage, S. Luik, J. N. H. Spoor, S. Malladi,
H. Zandbergen, C. Dekker, Nat. Commun. 2013,4, 2619.
[106] R. L. Kumawat, B. Pathak, J. Phys. Chem. B 2022,126, 1178.
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (39 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
[107] S. Liu, B. Lu, Q. Zhao, J. Li, T. Gao, Y. Chen, Y. Zhang, Z. Liu, Z. Fan,
F. Yang , L . Yo u , D . Yu , Adv. Mater. 2013,25, 4549.
[108] K. Liu, J. Feng, A. Kis, A. Radenovic, ACS Nano 2014,8, 2504.
[109] M. Mojtabavi, A. VahidMohammadi, W. Liang, M. Beidaghi, M.
Wanunu, ACS Nano 2019,13, 3042.
[110] Z. Cao, P. Yadav, A. Barati Farimani, Nano Lett. 2022,22, 7874.
[111] Z. Zhou, Y. Hu, H. Wang, Z. Xu, W. Wang, X. Bai, X. Shan, X. Lu, Sci.
Rep. 2013,3, 3287.
[112] A. B. Farimani, K. Min, N. R. Aluru, ACS Nano 2014,8, 7914.
[113] A. Zou, P. Xiu, X. Ou, R. Zhou, J. Phys. Chem. B 2020,124, 9490.
[114] Z. Gu, Z. He, F. Chen, L. Meng, J. Feng, R. Zhou, ACS Appl. Mater.
Interfaces 2022,14, 32618.
[115] P. Yadav, Z. Cao, A. Barati Farimani, ACS Nano 2021,15, 4861.
[116] A. D. Carral, C. S. Sarap, K. Liu, A. Radenovic, M. Fyta, 2D Mater.
2019,6, 045011.
[117] Á. Díaz Carral, M. Ostertag, M. Fyta, J. Chem. Phys. 2021,154,
044111.
[118] B. M. Venkatesan, R. Bashir, Nat. Nanotechnol. 2011,6, 615.
[119] S. M. Iqbal, D. Akin, R. Bashir, Nat. Nanotechnol. 2007,2, 243.
[120] E. C. Yusko, J. M. Johnson, S. Majd, P. Prangkio, R. C. Rollings, J. Li,
J. Yang, M. Mayer, Nat. Nanotechnol. 2011,6, 253.
[121] M. Wanunu, A. Meller, Nano Lett. 2007,7, 1580.
[122] S. Majd, E. C. Yusko, Y. N. Billeh, M. X. Macrae, J. Yang, M. Mayer,
Curr. Opin. Biotechnol. 2010,21, 439.
[123] A. R. Hall, A. Scott, D. Rotem, K. K. Mehta, H. Bayley, C. Dekker,
Nat. Nanotechnol. 2010,5, 874.
[124] S. Cabello-Aguilar, S. Balme, A. Abou Chaaya, M. Bechelany, E.
Balanzat, J.-M. Janot, C. Pochat-Bohatier, P. Miele, P. Dejardin,
Nanoscale 2013,5, 9582.
[125] A. Barati Farimani, P. Dibaeinia, N. R. Aluru, ACS Appl. Mater. Inter-
faces 2017,9, 92.
[126] R. Balasubramanian, S. Pal, H. Joshi, A. Rao, A. Naik, M.
Varma, B. Chakraborty, P. K. Maiti, J. Phys. Chem. C 2019,123,
11908.
[127] B. Cressiot, S. J. Greive, M. Mojtabavi, A. A. Antson, M. Wanunu,
Nat. Commun. 2018,9, 4652.
[128] D. H. Stoloff, M. Wanunu, Curr. Opin. Biotechnol. 2013,24, 699.
[129] N. A. W. Bell, U. F. Keyser, FEBS Lett. 2014,588, 3564.
[130] T. Ding, J. Yang, V. Pan, N. Zhao, Z. Lu, Y. Ke, C. Zhang, Nucleic Acids
Res. 2020,48, 2791.
[131] S. Hernández-Ainsa, U. F. Keyser, Nanoscale 2014,6, 14121.
[132] P. W. K. Rothemund, Nature 2006,440, 297.
[133] S. M. Douglas, H. Dietz, T. Liedl, B. Högberg, F. Graf, W. M. Shih,
Nature 2009,459, 414.
[134] P. Wang, T. A. Meyer, V. Pan, P. K. Dutta, Y. Ke, Chem 2017,2, 359.
[135] F. Hong, F. Zhang, Y. Liu, H. Yan, Chem. Rev. 2017,117, 12584.
[136] S. Julin, S. Nummelin, M. A. Kostiainen, V. Linko, J Nanopart Res
2018,20, 119.
[137] P. Piskunen, S. Nummelin, B. Shen, M. A. Kostiainen, V. Linko,
Molecules 2020,25, 1823.
[138] R. Wei, T. G. Martin, U. Rant, H. Dietz, Angew. Chem., Int. Ed. 2012,
51, 4864.
[139] C. R. Engst, M. Ablay, G. Divitini, C. Ducati, T. Liedl, U. F. Keyser,
Nano Lett. 2012,12, 512.
[140] J. Bentin, S. Balme, F. Picaud, Soft Matter 2020,16, 1002.
[141] B. Luan, R. Zhou, ACS Nano 2018,12, 3886.
[142] B. Luan, M. A. Kuroda, ACS Nano 2020,14, 13137.
[143] Y. Liu, Y. Deng, Y. Yang, Y. Qu, C. Zhang, Y.-Q. Li, M. Zhao, W. Li,
Nanoscale Adv 2021.
[144] J.-H. Lee, Y.-K. Choi, H.-J. Kim, R. H. Scheicher, J.-H. Cho, J. Phys.
Chem. C 2013,117, 13435.
[145] J. S. Lee, J. P. Oviedo, Y. M. N. D. Y. Bandara, X. Peng, L. Xia, Q.
Wang,K.Garcia,J.Wang,M.J.Kim,M.J.Kim,Electrophoresis 2021,
42, 991.
[146] R. Balasubramanian, S. Pal, A. Rao, A. Naik, B. Chakraborty, P. K.
Maiti, M. M. Varma, ACS Appl Bio Mater 2021,4, 451.
[147] P. Sen, H. Hoi, M. Gupta, ACS Appl Bio Mater 2021,4, 5416.
[148] K. K. Saha, M. Drndi´
c, B. K. Nikoli´
c, Nano Lett. 2012,12,50.
[149] S. Raza, A. Ameen, IntJofAdvinSciRes2017,3, 90.
[150] T. Nelson, B. Zhang, O. V. Prezhdo, Nano Lett. 2010,10, 3237.
[151] Y. He, R. H. Scheicher, A. Grigoriev, R. Ahuja, S. Long, Z. Huo, M.
Liu, Adv. Funct. Mater. 2011,21, 2674.
[152] G. T. Feliciano, C. Sanz-Navarro, M. D. Coutinho-Neto, P. Ordejón,
R. H. Scheicher, A. R. Rocha, J. Phys. Chem. B 2018,122, 485.
[153] J. Prasongkit, S. Jungthawan, R. G. Amorim, R. H. Scheicher, Nano
Res. 2022,15, 9843.
[154] T. Ahmed, J. T. Haraldsen, J.-X. Zhu, A. V. Balatsky, J. Phys. Chem.
Lett. 2014,5, 2601.
[155] G. T. Feliciano, C. Sanz-Navarro, M. D. Coutinho-Neto, P. Ordejón,
R. H. Scheicher, A. R. Rocha, Phys. Rev. Applied 2015,3, 034003.
[156] H. Qiu, A. Sarathy, J.-P. Leburton, K. Schulten, Nano Lett. 2015,15,
8322.
[157] P. S.-L. Ouyang Fang-Ping, P. S.-L. Ouyang Fang-Ping, Chin. Phys. B
2011,20, 58504.
[158] A. Girdhar, C. Sathe, K. Schulten, J.-P. Leburton, Proc. Natl. Acad.
Sci. USA 2013,110, 16748.
[159] H. Sadeghi, L. Algaragholy, T. Pope, S. Bailey, D. Visontai, D.
Manrique, J. Ferrer, V. Garcia-Suarez, S. Sangtarash, C. J. Lambert,
J. Phys. Chem. B 2014,118, 6908.
[160] E. McCann, M. Koshino, Rep Prog Phys 2013,76, 056503.
[161] J. Prasongkit, G. T. Feliciano, A. R. Rocha, Y. He, T. Osotchan, R.
Ahuja, R. H. Scheicher, Sci. Rep. 2015,5, 17560.
[162] S. M. Avdoshenko, D. Nozaki, C. Gomes da Rocha, J. W. González,
M. H. Lee, R. Gutierrez, G. Cuniberti, Nano Lett. 2013,13, 1969.
[163] F. A. L. de Souza, R. G. Amorim, W. L. Scopel, R. H. Scheicher,
Nanoscale 2017,9, 2207.
[164] J. Prasongkit, E. de Freitas Martins, F. A. L. de Souza, W. L. Scopel,
R. G. Amorim, V. Amornkitbamrung, A. R. Rocha, R. H. Scheicher,
J. Phys. Chem. C 2018,122, 7094.
[165] R. L. Kumawat, B. Pathak, ACS Appl. Electron. Mater. 2021,3, 3835.
[166] M. K. Jena, R. L. Kumawat, B. Pathak, ACS Appl. Nano Mater. 2021.
[167] P. Xie, Q. Xiong, Y. Fang, Q. Qing, C. M. Lieber, Nat. Nanotechnol.
2012,7, 119.
[168] F. Traversi, C. Raillon, S. M. Benameur, K. Liu, S. Khlybov, M. Tosun,
D. Krasnozhon, A. Kis, A. Radenovic, Nat. Nanotechnol. 2013,8, 939.
[169] S. J. Heerema, L. Vicarelli, S. Pud, R. N. Schouten, H. W.
Zandbergen, C. Dekker, ACS Nano 2018,12, 2623.
[170] M. Graf, M. Lihter, D. Altus, S. Marion, A. Radenovic, Nano Lett.
2019,19, 9075.
[171] M. Puster, A. Balan, J. A. Rodríguez-Manzo, G. Danda, J.-H. Ahn, W.
Parkin, M. Drndi´
c, Small 2015,11, 6309.
[172] M. K. Jena, D. Roy, B. Pathak, J. Phys. Chem. Lett. 2022,13, 11818.
[173] M. K. Jena, B. Pathak, Nano Lett. 2023.
[174] F. C. Maier, M. Fyta, ChemPhysChem 2020,21, 2068.
[175] J. Shendure, H. Ji, Nat. Biotechnol. 2008,26, 1135.
[176] R. L. Kumawat, P. Garg, S. Kumar, B. Pathak, ACS Appl. Mater. Inter-
faces 2019,11, 219.
[177] J. Lagerqvist, M. Zwolak, M. Di Ventra, Nano Lett. 2006,6, 779.
[178] H. W. Ch. Postma, Nano Lett. 2010,10, 420.
[179] J. Prasongkit, A. Grigoriev, B. Pathak, R. Ahuja, R. H. Scheicher,
Nano Lett. 2011,11, 1941.
[180] J. Prasongkit, A. Grigoriev, B. Pathak, R. Ahuja, R. H. Scheicher, J.
Phys. Chem. C 2013,117, 15421.
[181] E. de Freitas Martins, R. G. Amorim, G. T. Feliciano, R. H. Scheicher,
A. R. Rocha, Carbon 2019.
[182] R. G. Amorim, A. R. Rocha, R. H. Scheicher, J. Phys. Chem. C 2016,
120, 19384.
[183] Henk W. Ch. Postma, Nano Lett. 2010,10, 420425.
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (40 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
[184] R. L. Kumawat, B. Pathak, Nanoscale Adv 2020,2, 4041.
[185] G. Sivaraman, R. G. Amorim, R. H. Scheicher, M. Fyta, Nanoscale
2016,8, 10105.
[186] H. Zou, S. Wen, X. Wu, K.-W. Wong, C. Yam, Phys.Chem.Chem.Phys.
2022,24, 5748.
[187] R. L. Kumawat, P. Garg, S. Kumar, B. Pathak, ACS Appl. Mater. Inter-
faces 2019,11, 219.
[188] R. L. Kumawat, P. Garg, G. Bhattacharyya, B. Pathak, ACS Appl. Elec-
tron. Mater. 2020,2, 1218.
[189] S. Mittal, M. K. Jena, B. Pathak, ACS Appl. Nano Mater. 2022,5, 9356.
[190] S. Mittal, B. Pathak, ACS Appl Bio Mater 2023,6, 218.
[191] R. L. Kumawat, B. Pathak, ACS Appl Bio Mater 2021,4, 1403.
[192] A. Perez, R. G. Amorim, C. E. P. Villegas, A. R. Rocha, Phys. Chem.
Chem. Phys. 2020,22, 27053.
[193] M. Kumar Jena, B. Pathak, Phys.Chem.Chem.Phys.2022,24, 21427.
[194] S. Mittal, B. Pathak, Nanoscale 2023,15, 757.
[195] M. Tsutsui, M. Taniguchi, K. Yokota, T. Kawai, Nat. Nanotechnol.
2010,5, 286.
[196] A. P. Ivanov, E. Instuli, C. M. McGilvery, G. Baldwin, D. W. McComb,
T. Albrecht, J. B. Edel, Nano Lett. 2011,11, 279.
[197] A. Fanget, T. Floriano, S. Khlybov, P. Granjon, A. Magrez, L. Forró,
A. Radenovic, Nano Lett. 2013,14.
[198] T. Furuhata, T. Ohshiro, G. Akimoto, R. Ueki, M. Taniguchi, S. Sando,
ACS Nano 2019,13, 5028.
[199] M. Tsutsui, S. Rahong, Y. Iizumi, T. Okazaki, M. Taniguchi, T. Kawai,
Sci. Rep. 2011,1, 46.
[200] P. Pang, B. A. Ashcroft, W. Song, P. Zhang, S. Biswas, Q. Qing,
J. Yang, R. J. Nemanich, J. Bai, J. T. Smith, K. Reuter, V. S. K.
Balagurusamy, Y. Astier, G. Stolovitzky, S. Lindsay, ACS Nano 2014,
8, 11994.
[201] H. N. Patel, I. Carroll, R. L. Jr, S. Sankararaman, C. Etienne, S.
R. Kodigala, M. R. Paul, H. W. C. Postma, PLoS One 2017,12,
e0171505.
[202] L. Tang, B. P. Nadappuram, P. Cadinu, Z. Zhao, L. Xue, L. Yi, R. Ren,
J. Wang, A. P. Ivanov, J. B. Edel, Nat. Commun. 2021,12, 913.
[203] Y. Zhao, B. Ashcroft, P. Zhang, H. Liu, S. Sen, W. Song, J. Im, B.
Gyarfas, S. Manna, S. Biswas, C. Borges, S. Lindsay, Nat. Nanotech-
nol. 2014,9, 466.
[204] J. Im, S. Sen, S. Lindsay, P. Zhang, ACS Nano 2018,12, 7067.
[205] M. Taniguchi, T. Ohshiro, Y. Komoto, T. Takaai, T. Yoshida, T. Washio,
J. Phys. Chem. C 2019,123, 15867.
[206] M. Dou, F. C. Maier, M. Fyta, Nanoscale 2019,11, 14216.
[207] S. Gowtham, R. H. Scheicher, R. Ahuja, R. Pandey, S. P. Karna, Phys.
Rev. B 2007,76, 033401.
[208] D. Le, A. Kara, E. Schröder, P. Hyldgaard, T. S. Rahman, JPhysCon-
dens Matter 2012,24, 424210.
[209] J. Antony, S. Grimme, Phys.Chem.Chem.Phys.2008,10, 2722.
[210] G. Breit, E. Wigner, Phys. Rev. 1936,49, 519.
[211] U. Fano, Phys. Rev. 1961,124, 1866.
[212] S. K. Min, W. Y. Kim, Y. Cho, K. S. Kim, Nat. Nanotechnol. 2011,6,
162.
[213] A. C. Rajan, M. R. Rezapour, J. Yun, Y. Cho, W. J. Cho, S. K. Min, G.
Lee, K. S. Kim, ACS Nano 2014,8, 1827.
[214] P.-H. Chang, H. Liu, B. K. Nikoli´
c, JComputElectron2014,13, 847.
[215] A. Oliveira-Brett, A. Paquim, Langmuir 2003,19, 3830.
[216] T. Ahmed, S. Kilina, T. Das, J. T. Haraldsen, J. J. Rehr, A. V. Balatsky,
Nano Lett. 2012,12, 927.
[217] Y. Cho, S. K. Min, W. Y. Kim, K. S. Kim, Phys.Chem.Chem.Phys.2011,
13, 14293.
[218] E. Paulechka, T. A. Wassenaar, K. Kroenlein, A. Kazakov, A.
Smolyanitsky, Nanoscale 2016,8, 1861.
[219] M. Rastgoo, M. Fathipour, Appl. Surf. Sci. 2019,492, 634.
[220] E. de Freitas Martins, G. Troiano Feliciano, R. Hendrik Scheicher, A.
Reily Rocha, Molecules 2019,24, 951.
[221] S. Thomas, A. C. Rajan, M. R. Rezapour, K. S. Kim, J. Phys. Chem. C
2014,118, 10855.
[222] Q. Li, H. Liu, Y. Tian, J. Guo, G. Chen, J. Y. Lee, J. Phys. Chem. C 2020,
124, 10823.
[223] S. Alesheikh, N. Shahtahmassebi, M. R. Roknabadi, R. Pilevar
Shahri, Phys. Lett. A 2018,382, 595.
[224] R. L. Kumawat, B. Pathak, J. Phys. Chem. C 2019,123, 22377.
[225] R. L. Kumawat, B. Pathak, Mater. Adv. 2020,1, 2908.
[226] M. R. Rezapour, B. Biel, Nanoscale 2022,14, 5147.
[227] F. A. L. De Souza, G. Sivaraman, M. Fyta, R. H. Scheicher, W. L.
Scopel,R.G.Amorim,Nanoscale 2020,12, 18289.
[228] V. Shukla, N. K. Jena, A. Grigoriev, R. Ahuja, ACS Appl. Mater. Inter-
faces 2017,9, 39945.
[229] F. A. L. de Souza, R. G. Amorim, W. L. Scopel, R. H. Scheicher, Phys.
Chem. Chem. Phys. 2019,21, 24884.
[230] A. Kiakojouri, I. Frank, E. Nadimi, Phys.Chem.Chem.Phys.2021,23,
25126.
[231] F. Haque, J. Li, H.-C. Wu, X.-J. Liang, P. Guo, Nano Today 2013,8, 56.
[232] Q. Chen, Z. Liu, Sensors 2019,19, 1886.
[233] A. S. Panwar, M. Muthukumar, J. Am. Chem. Soc. 2009,131, 18563.
[234] S. Benner, R. J. A. Chen, N. A. Wilson, R. Abu-Shumays, N. Hurt, K.
R. Lieberman, D. W. Deamer, W. B. Dunbar, M. Akeson, Nat. Nan-
otechnol. 2007,2, 718.
[235] B. Hornblower, A. Coombs, R. D. Whitaker, A. Kolomeisky, S. J.
Picone, A. Meller, M. Akeson, Nat. Methods 2007,4, 315.
[236] S. L. Cockroft, J. Chu, M. Amorin, M. R. Ghadiri, J. Am. Chem. Soc.
2008,130, 818.
[237] R. S. S. De Zoysa, D. A. Jayawardhana, Q. Zhao, D. Wang, D. W.
Armstrong, X. Guan, J. Phys. Chem. B 2009,113, 13332.
[238] H. Peng, X. S. Ling, Nanotechnology 2009,20, 185101.
[239] M. Wanunu, J. Sutin, B. McNally, A. Chow, A. Meller, Biophys. J.
2008,95, 4716.
[240] R. Akahori, I. Yanagi, Y. Goto, K. Harada, T. Yokoi, K. Takeda, Sci. Rep.
2017,7, 9073.
[241] F. Olasagasti, K. R. Lieberman, S. Benner, G. M. Cherf, J. M. Dahl,
D. W. Deamer, M. Akeson, Nat. Nanotechnol. 2010,5, 798.
[242] K. R. Lieberman, G. M. Cherf, M. J. Doody, F. Olasagasti, Y. Kolodji,
M. Akeson, J. Am. Chem. Soc. 2010,132, 17961.
[243] G. M. Cherf, K. R. Lieberman, H. Rashid, C. E. Lam, K. Karplus, M.
Akeson, Nat. Biotechnol. 2012,30, 344.
[244] E. M. Nelson, H. Li, G. Timp, ACS Nano 2014,8, 5484.
[245] Z. Yuan, Y. Liu, M. Dai, X. Yi, C. Wang, Nanoscale Res. Lett. 2020,15,
80.
[246] D. Fologea, J. Uplinger, B. Thomas, D. S. McNabb, J. Li, Nano Lett.
2005,5, 1734.
[247] J. Feng, K. Liu, R. D. Bulushev, S. Khlybov, D. Dumcenco, A. Kis, A.
Radenovic, Nat. Nanotechnol. 2015,10, 1070.
[248] S. Marion, N. Vuˇ
cemilovi´
c-Alagi´
c, M. Špadina, A. Radenovi´
c, A.-S.
Smith, Small 2021,17, 2100777.
[249] S. W. Kowalczyk, D. B. Wells, A. Aksimentiev, C. Dekker, Nano Lett.
2012,12, 1038.
[250] M. Wanunu, W. Morrison, Y. Rabin, A. Y. Grosberg, A. Meller, Nat.
Nanotechnol. 2010,5, 160.
[251] J. Li, D. S. Talaga, J. Phys.: Condens. Matter 2010,22, 454129.
[252] R. Akahori, T. Haga, T. Hatano, I. Yanagi, T. Ohura, H. Hamamura,
T. Iwasaki, T. Yokoi, T. Anazawa, Nanotechnology 2014,25, 275501.
[253] T. Z. Butler, J. H. Gundlach, M. Troll, Biophys. J. 2007,93, 3229.
[254] D. W. Deamer, D. Branton, Acc. Chem. Res. 2002,35, 817.
[255] H. Bayley, Curr. Opin. Chem. Biol. 2006,10, 628.
[256] N. Ashkenasy, J. Sánchez-Quesada, H. Bayley, M. R. Ghadiri, Angew.
Chem., Int. Ed. 2005,44, 1401.
[257] N. Mitchell, S. Howorka, Angew. Chem., Int. Ed. 2008,47, 5565.
[258] T. P. Niedringhaus, D. Milanova, M. B. Kerby, M. P. Snyder, A. E.
Barron, Anal. Chem. 2011,83, 4327.
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (41 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
[259] A. Fragasso, S. Schmid, C. Dekker, ACS Nano 2020,14, 1338.
[260] Rev. Mod. Phys. 2008,80, 141.
[261] V. Meunier, P. S. Krsti´
c, J. Chem. Phys. 2008,128, 041103.
[262] C. Wen, S. Zeng, Z. Zhang, K. Hjort, R. Scheicher, S.-L. Zhang, Nan-
otechnology 2016,27, 215502.
[263] S. J. Heerema, G. F. Schneider, M. Rozemuller, L. Vicarelli, H. W.
Zandbergen, C. Dekker, Nanotechnology 2015,26, 074001.
[264] C. Wen, S. Zeng, K. Arstila, T. Sajavaara, Y. Zhu, Z. Zhang, S.-L.
Zhang, ACS Sens. 2017,2, 300.
[265] R. M. M. Smeets, N. H. Dekker, C. Dekker, Nanotechnology 2009,
20, 095501.
[266] K. Lee, K.-B. Park, H.-J. Kim, J.-S. Yu, H. Chae, H.-M. Kim, K.-B. Kim,
Adv. Mater. 2018,30, 1704680.
[267] R. M. M. Smeets, U. F. Keyser, N. H. Dekker, C. Dekker, Proc. Natl.
Acad. Sci. USA 2008,105, 417.
[268] B. M. Venkatesan, D. Estrada, S. Banerjee, X. Jin, V. E. Dorgan, M.-H.
Bae, N. R. Aluru, E. Pop, R. Bashir, ACS Nano 2012,6, 441.
[269] J. K. Rosenstein, M. Wanunu, C. A. Merchant, M. Drndic, K. L.
Shepard, Nat. Methods 2012,9, 487.
[270] S. Shekar, C.-C. Chien, A. Hartel, P. Ong, O. B. Clarke, A. Marks, M.
Drndic, K. L. Shepard, Nano Lett. 2019,19, 1090.
[271] M. Tsutsui, T. Takaai, K. Yokota, T. Kawai, T. Washio, Small Methods
2021,5, 2100191.
[272] M. Tsutsui, K. Yokota, Y. He, T. Kawai, Small Methods 2022,6,
2200761.
[273] R. A. Lucas, C.-Y. Lin, L. A. Baker, Z. S. Siwy, Nat. Commun. 2020,
11, 1568.
[274] T. Morikawa, K. Yokota, M. Tsutsui, M. Taniguchi, Nanoscale 2017,
9, 4076.
[275] R. M. Leidenfrost, D.-C. Pöther, U. Jäckel, R. Wünschiers, Sci. Rep.
2020,10, 5125.
[276] A.-L. Biance, J. Gierak, É. Bourhis, A. Madouri, X. Lafosse, G.
Patriarche, G. Oukhaled, C. Ulysse, J.-C. Galas, Y. Chen, L. Auvray,
Microelectron. Eng. 2006,83, 1474.
[277] S. Luo, B. H. Hoff, S. A. Maier, J. C. de Mello, Adv. Sci. 2021,8,
2102756.
[278] J. P. Fried, J. L. Swett, X. Bian, J. A. Mol, MRS Commun. 2018,8, 703.
[279] J. Bai, X. Duan, Y. Huang, Nano Lett. 2009,9, 2083.
[280] L. Jiao, L. Zhang, X. Wang, G. Diankov, H. Dai, Nature 2009,458,
877.
[281] M. Waugh, K. Briggs, D. Gunn, M. Gibeault, S. King, Q. Ingram, A.
M. Jimenez, S. Berryman, D. Lomovtsev, L. Andrzejewski, V. Tabard-
Cossa, Nat. Protoc. 2020,15, 122.
[282] M. Belkin, S.-H. Chao, M. P. Jonsson, C. Dekker, A. Aksimentiev,
ACS Nano 2015,9, 10598.
[283] F. Nicoli, D. Verschueren, M. Klein, C. Dekker, M. P. Jonsson, Nano
Lett. 2014,14, 6917.
[284] S. Peng, Z. Yang, X. Ni, H. Zhang, J. Ouyang, O. Fangping, Mater.
Res. Express 2014,1, 015044.
[285] S. Biswas, S. Sen, J. Im, S. Biswas, P. Krstic, B. Ashcroft, C.
Borges, Y. Zhao, S. Lindsay, P. Zhang, ACS Nano 2016,10,
11304.
[286] S. Mittal, B. Pathak, ACS Appl Bio Mater 2022.
[287] S. Pud, S.-H. Chao, M. Belkin, D. Verschueren, T. Huijben, C.
van Engelenburg, C. Dekker, A. Aksimentiev, Nano Lett. 2016,16,
8021.
[288] A. Choudhary, H. Joshi, H.-Y. Chou, K. Sarthak, J. Wilson, C. Maffeo,
A. Aksimentiev, ACS Nano 2020,14, 15566.
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (42 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.small-journal.com
Rameshwar L. Kumawat was born in Jaipur,Rajasthan, India. He obtained his Bachelor and Master
of Technology in Nanotechnology (major) and Information and Communication Technology (minor)
in 2015 and 2016 from the Centre for Converging Technologies-University of Rajasthan, Jaipur, Ra-
jasthan, India. In July 2017, he joined Prof. Biswarup Pathak’s lab as a doctoral student in the Depart-
ment of Metallurgy Engineering and Materials Science, Indian Institute of Technology Indore (IITI),
India. His Ph.D. thesis was titled “Nano-electrodes for ultrafast DNA/protein sequencing: Ab-initio
quantum transport studies.” He extensively used atomistic simulations to shed light on the DNA and
protein sensing and detection physics of solid-state nanopores, nanogaps, and nanochannels. He
also worked on tuning the physical and chemical properties of low-dimensional materials. Soon af-
ter that, he joined Regents’ Prof. C. David Sherrill’s lab as a Postdoctoral Fellow at Georgia Institute
of Technology, Georgia, Atlanta, USA. His current research focuses on developing and applying new
models for intermolecular interactions and their implementation as software and machine learn-
ing for high-throughput screening of materials/molecules and their chemical and physical property
prediction at Georgia Tech.
Milan Kumar Jena obtained his UG degree from Ravenshaw University,Cuttack, Odisha. He com-
pleted his master’s degree from the Indian Institute of Technology Guwahati, Assam. Then, he joined
Prof. Biswarup Pathak’s group as a graduate student in January 2020 at the Department of Chemistry,
Indian Institute of Technology Indore, where his primary research interest is the computational inves-
tigation of solid-state nanopores/nanogaps for next-generation DNA sequencing.
Sneha Mittal received her M.Sc. degree in chemistry from Ramjas College, University of Delhi (DU),
in 2018. Currently,she is carrying out her doctorate studies under the supervision of Prof. Biswarup
Pathak at the Department of Chemistry,Indian Institute of Technology Indore (IITI), India. Her current
research work focuses on the application of molecular electronics and machine learning in single-
molecule DNA and protein sequencing.
Biswarup Pathak was born in Bankura, West Bengal, India. He obtained his B.Sc. (Chemistry, 2000)
from Bankura Christian College and M.Sc. (Physical Chemistry, 2002) from Banaras Hindu University
(BHU). After postgraduation from BHU, he obtained his Ph.D. from Hyderabad Central University
under the supervision of Professor E.D. Jemmis. During his Ph.D., he was involved in computational
studies on the structure, bonding, and reactivity of boranes, carboranes, metallaboranes, and transi-
tion metal complexes. Soon after his Ph.D., he completed his four-year postdoctoral study with Prof.
Jerzy Leszczynski’s group at Jackson State University,USA (January 2008–July 2009) and Prof. Rajeev
Ahuja’s group at Uppsala University,Sweden (September 2009–May 2012). He joined IIT Indore as
an Assistant Professor in 2012 and became a Professor in 2022. Dr.Pathak is currently working on de-
veloping solid-state materials for clean energy (hydrogen storage, photocatalysis, fuel cells, batteries,
and solar cells) and biological (DNA sequencing) applications.
Small 2024, 2401112 © 2024 Wiley-VCH GmbH
2401112 (43 of 43)
16136829, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/smll.202401112 by The Librarian, Wiley Online Library on [08/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
The tremendous upsurge in the research of next-generation sequencing (NGS) methods has broadly been driven by the rise of the wonder material graphene and continues to dominate the futuristic approaches for fast and accurate DNA sequencing. The success of graphene has also triggered the search for many new potential NGS methods capable of ultrafast, reliable, and controlled DNA sequencing. The present study delves into the potential of a new NGS architecture utilizing graphene, namely, a 'semi/hybrid-nanogap' for the identification of DNA nucleobases with single-base resolution. In the framework of first-principles density functional theory methods, we have calculated the transmission function and current-voltage (I-V) characteristics which are of particular significance for DNA sequencing applications. It is noted that the interaction energy values are significantly reduced compared to the previously reported graphene nanodevices, which can lead to a controlled translocation during experimental measurements. Based on the transmission function, each nucleobase can be identified with pertinent sensitivity. It is noticed that the use of highly conductive nucleobase analogs can facilitate improved single nucleobase sensing by increasing the transmission sensitivity. Therefore, we believe that the present study opens up promising frontiers for sequencing applications.
Article
Full-text available
Ionic signal amplification is a key challenge for single‐molecule analyses by solid‐state nanopore sensing. Here, a permittivity gradient approach for amplifying ionic blockade characteristics of DNA in a nanofluidic channel is reported. The transmembrane ionic current response is found to change substantially through modifying the liquid permittivity at one side of a pore with an organic solvent. Imposing positive liquid permittivity gradients with respect to the direction of DNA electrophoresis, this study observes the resistive ionic signals to become larger due to the varying contributions of molecular counterions. On the contrary, negative gradients render adverse effects causing conductive ionic current pulses upon polynucleotide translocations. Most importantly, both the positive and negative gradients are demonstrated to be capable of amplifying the ionic signals by an order of magnitude with a 1.3‐fold difference in the transmembrane liquid dielectric constants. This phenomenon allows a novel way to enhance the single‐molecule sensitivity of nanopore sensing that may be useful in analyzing secondary structures and genome sequence of DNA by ionic current measurements.
Article
Full-text available
Rapid and inexpensive DNA sequencing is critical to biomedical research and healthcare for the accomplishment of personalized medicine. Solid-state nanopores and nanogaps have marshalled themselves in the fascinating paradigm of nano-research since the advent of its application in DNA sequencing by analyzing the quantum conductance and electric current signals. In this study, the feasibility of the considered borophene nanogaps for DNA sequencing purposes via the electronic tunnelling current approach was investigated by utilizing combined density functional theory with non-equilibrium Green's function (DFT-NEGF) techniques. The interaction energy (Ei) and the charge density difference (CDD) plots exploit the charge modulation around the nanogap edges due to the presence of each nucleotide. Our results revealed a distinct variation in the tunnelling conductance, as a characteristic fingerprint of each nucleotide at the Fermi level. The calculated tunnelling current variation across the nanogap under an applied bias voltage was also significant due to the effective coupling of nucleotides with the electrode edges. The current was in the picoampere (pA) range, which was fairly higher than the electrical background noise and also experimentally detectable by the canning tunnelling microscopy (STM) technique. Our findings demonstrated that in the borophene nanopore vs. nanogap scenario, the nanogap has several advantages and is a more promising nanobiosensor. Moreover, we also compared our results with various previous experimental and theoretical reports on nanogaps as well as nanopores for gaining better insights.
Article
Full-text available
Nanopore-based devices have provided exciting opportunities to develop affordable label-free DNA sequencing platforms. Over a decade ago, graphene has been proposed as a two-dimensional (2D) nanopore membrane in order to achieve single-base resolution. However, it was experimentally revealed that clogging of the graphene nanopore can occur due to the hydrophobic nature of graphene, thus hindering the translocation of DNA. To overcome this problem, the exploration of alternative 2D materials has gained considerable interest over the last decade. Here we show that a Ti 2 C-based MXene nanopore functionalized by hydroxyl groups (−OH) exhibits transverse conductance properties that allow for the distinction between all four naturally occurring DNA bases. We have used a combination of density functional theory and non-equilibrium Green’s function method to sample over multiple orientations of the nucleotides in the nanopore, as generated from molecular dynamics simulations. The conductance variation resulting from sweeping an applied gate voltage demonstrates that the Ti 2 C-based MXene nanopore possesses high potential to rapidly and reliably sequence DNA. Our findings open the door to further theoretical and experimental explorations of MXene nanopores as a promising 2D material for nanopore-based DNA sensing.
Article
Full-text available
Since its initial release in 2000, the human reference genome has covered only the euchromatic fraction of the genome, leaving important heterochromatic regions unfinished. Addressing the remaining 8% of the genome, the Telomere-to-Telomere (T2T) Consortium presents a complete 3.055 billion-base pair sequence of a human genome, T2T-CHM13, that includes gapless assemblies for all chromosomes except Y, corrects errors in the prior references, and introduces nearly 200 million base pairs of sequence containing 1956 gene predictions, 99 of which are predicted to be protein coding. The completed regions include all centromeric satellite arrays, recent segmental duplications, and the short arms of all five acrocentric chromosomes, unlocking these complex regions of the genome to variational and functional studies.
Article
Existing obstacles in next-generation DNA sequencing techniques, for instance, high noise, high translocation speed, and configurational fluctuations, call for approaches capable of reaching the goal and accelerating the process of personalized medicine development. The labeling nucleotide approach has the potential to overcome these barriers and boost the recognition sensitivity of a solid-state nanodevice. In this theoretical report, the first-principles density functional theory calculations have been employed to study the role of three different labels, tyrosine (Tyr), aspartic acid (Asp), and arginine (Arg), for labeling DNA nucleotides and study their effect in rapid and controlled DNA sequencing at atomic resolution. Remarkable differences in interaction energy values are noticed in all three cases of differently labeled nucleotides. The zero-bias transmission spectra confirm that proposed labels have the ability to detect the individual nucleotide, amplifying the tunneling current sensitivity by several orders of magnitude. The current-voltage characteristics of Arg-labeled nucleotides are found to be promising for single nucleotide recognition even at a very low bias voltage of 0.1 V.
Article
Solid-state nanopore-based electrical detection of DNA nucleotides with the quantum tunneling technique has emerged as a powerful strategy to be the next-generation sequencing technology. However, experimental complexity has been a foremost obstacle in achieving a more accurate high-throughput analysis with industrial scalability. Here, with one of the nucleotide training data sets of a model monolayer gold nanopore, we have predicted the transmission function for all other nucleotides with root-mean-square error scores as low as 0.12 using the optimized eXtreme Gradient Boosting Regression (XGBR) model. Further, the SHapley Additive exPlanations (SHAP) analysis helped in exploring the interpretability of the XGBR model prediction and revealed the complex relationship between the molecular properties of nucleotides and their transmission functions by both global and local interpretable explanations. Hence, experimental integration of our proposed machine-learning-assisted transmission function prediction method can offer a new direction for the realization of cheap, accurate, and ultrafast DNA sequencing.
Article
Despite much research on characterizing 2D materials for DNA detection with nanopore technology, a thorough comparison between the performance of different 2D materials is currently lacking. In this work, using extensive molecular dynamics simulations, we compare nanoporous graphene, MoS2 and titanium carbide MXene (Ti3C2) for their DNA detection performance and sensitivity. The ionic current and residence time of DNA are characterized in each nanoporous materials by performing hundreds of simulations. We devised two statistical measures including the Kolmogorov-Smirnov test and the absolute pairwise difference to compare the performance of nanopores. We found that graphene nanopore is the most sensitive membrane for distinguishing DNA bases. The MoS2 is capable of distinguishing the A and T bases from the C and G bases better than graphene and MXene. Physisorption and the orientation of DNA in nanopores are further investigated to provide molecular insight into the performance characteristics of different nanopores.
Article
Recent advances in cost-effective, ultra-rapid, and efficient DNA sequencing are in the saddle of the advancement of the personalization of medicines for understanding and early-stage detection of several killer diseases. Paying attention to a timely need for the development of solid-state nanodevices for rapid and controlled identification of DNA nucleotides, in this report, we theoretically explored the potential of labeling techniques in the sequencing of DNA nucleotides through solid-state graphene nanogap electrodes using the quantum tunneling current approach. Our study boasts the idea that labeling of DNA nucleotides can solve major hurdles of DNA sequencing, such as improving the signal-to-noise ratio, slowing down translocation velocity, and controlling orientational variations. Employing the first-principle density functional theory study, we identify unique interaction energy values for each labeled nucleotide having remarkable differences in the range of 0.10-0.74 eV. The zero-bias transmission spectra of the proposed setup suggest that the detection of the nucleotides is possible by applying very low gate voltages. Moreover, the labeling of nucleotides amplifies the conductance sensitivity considerably. I-V characteristics suggest that electrical recognition of each labeled nucleotide can be carried out at both lower (0.3 V) and higher (0.8 V) bias voltages with single-molecule resolution, although the maximum current sensitivity is observed at a higher bias voltage. The proposed sequencing device possesses high sensitivity and selectivity characteristics that are crucial for experimental purposes. We find that our results are rich compared to unlabeled nucleotides-based graphene nanopore/nanogap devices. Hence, the study will certainly motivate the experimentalists toward the application of a labeled DNA nucleotide system for ultra-rapid DNA sequencing by using the tunneling current approach.
Article
Nanopores in two-dimensional (2D) materials have emerged to offer in principle necessary spatial resolution for high-throughput DNA sequencing. However, their fidelity is severely limited by the fast DNA translocation. A recent experiment indicates that introducing ionic liquids could slow down DNA translocation in a MoS2 nanopore. However, the corresponding in-depth molecular mechanism underlying the experimental findings is not fully understood, which is crucial for the future improvement of rational DNA translocation control. Here, we computationally investigate and then experimentally identify the effect of BmimCl ionic liquid on the retardation of ssDNA translocation through a single-layer MoS2 nanopore. Our all-atom molecular dynamics simulations demonstrate that the strong interaction between Bmim+ and ssDNA offers a considerable dragging force to decelerate the electrophoretic motion of ssDNA in the BmimCl solution. Moreover, we show that Bmim+ ions exhibit preferential binding on the sulfur edges of the nanopore. These Bmim+ in the pore region can not only act as a steric blockage but also form π-π stackings with nucleobases, which provide a further restriction on the ssDNA motion. Therefore, our molecular dynamics simulation investigations deepen the understanding of the critical role of ionic liquid in DNA translocation through a nanopore from a molecular landscape, which may benefit practical implementations of ionic liquids in nanopore sequencing.