ArticlePDF Available

Functionalized DNA Origami‐Enabled Detection of Biomarkers

Wiley
ChemBioChem
Authors:

Abstract and Figures

Biomarkers are crucial physiological and pathological indicators in the host. Over the years, numerous detection methods have been developed for biomarkers, given their significant potential in various biological and biomedical applications. Among these, the detection system based on functionalized DNA origami has emerged as a promising approach due to its precise control over sensing modules, enabling sensitive, specific, and programmable biomarker detection. We summarize the advancements in biomarker detection using functionalized DNA origami, focusing on strategies for DNA origami functionalization, mechanisms of biomarker recognition, and applications in disease diagnosis and monitoring. These applications are organized into sections based on the type of biomarkers – nucleic acids, proteins, small molecules, and ions – and concludes with a discussion on the advantages and challenges associated with using functionalized DNA origami systems for biomarker detection.
Functionalized DNA origami for nucleic acid detection. (A) Label‐free detection of RNA based on morphological changes detectable by AFM. Scale bars,150 nm. Reproduced with permission.[64] Copyright: 2008, AAAS. (B) Label‐free unambiguous detection and symbolic display of single nucleotide polymorphisms based on morphological changes detectable by AFM. Scale bars, 50 nm. Reproduced with permission.[65] Copyright: 2008, American Chemical Society. (C) Paper‐folding mechanism for miRNA detection based on morphological changes detectable by AFM. Scale bars,200 nm. Reproduced with permission.[74] Copyright: 2023, Springer Nature. (D) Tumor‐associated microRNA detection and fluorescence imaging in live Cells. Reproduced with permission.[69] Copyright: 2021, Elsevier. (E) DNA origami book biosensor for the detection of cancer‐associated nucleic acids based on Förster resonance energy transfer (FRET). Reproduced with permission.[77] Copyright: 2022, The Royal Society of Chemistry. (F) DNA origami sensor and DNA‐PAINT based miRNA detection. Scale bars, 100 nm. Reproduced with permission.[78] Copyright: 2023, Elsevier. (G) Reconfigurable 3D plasmonic metamolecules. Reproduced with permission.[79] Copyright: 2014, Springer Nature. (H) DNA origami plasmonic nanoantenna for detection of miRNA. Reproduced with permission.[80] Copyright: 2022, American Chemical Society. (I) Detection of SARS‐CoV‐2 RNA with a plasmonic chiral biosensor. Reproduced with permission.[81] Copyright: 2023, Elsevier. (J) DNA origami gatekeepers for solid‐state nanopores. Scale bars, 20 nm. Reproduced with permission.[88] Copyright: 2012, Wiley‐VCH. (K) Label‐free electrochemical biosensors for microRNA detection. Reproduced with permission.[72] Copyright: 2019, American Chemical Society. (L) Electrochemical DNA Biosensors. Reproduced with permission.[90] Copyright: 2023, American Chemical Society.
… 
Functionalized DNA origami for proteins detection. (A) Detection of human thrombin based on DNA protein binding cavities by AFM imaging. Reproduced with permission.[68] Copyright: 2020, Springer Nature. (B) Detection of human C‐reactive protein (CRP) by AFM imaging. Scale bars, 30 nm. Reproduced with permission.[95] Copyright: 2020, Springer Nature. (C) Detection of TATA‐box binding protein by TEM imaging. Reproduced with permission.[99] Copyright: 2023, The Royal Society of Chemistry. (D) Detection of Gox and HRP by single molecule fluorescence signal. Scale bars, 50 nm. Reproduced with permission.[102] Copyright: 2016, Springer Nature. (E) Detection of antibody by fluorescence signal. Reproduced with permission.[103] Copyright: 2021, Elsevier. (F) Detection of antigen on T cell.[104] Copyright: 2022, Springer Nature. (G) The detection of single peridinin‐chlorophyll a‐protein complex based on plasmonic. Reproduced with permission.[108] Copyright: 2018 American Chemical Society (H) Detection of streptavidin and thrombin based on plasmonic. Scale bars, 5 μm. Reproduced with permission.[109] Copyright: 2022, Springer Nature. (I) The simultaneous detection of four different IgG antibodies based on solid‐state nanopore. Reproduced with permission.[110] Copyright: 2016, Springer Nature. (J) Trapping of small proteins in the nanopore electro‐osmotic trap. Reproduced with permission.[73] Copyright: 2022, American Chemical Society. (K) Detection of streptavidin and PDGF‐BB based on square wave voltammetry signal. Scale bars, 100 nm. Reproduced with permission.[82]
… 
Functionalized DNA origami for small molecule or ion detection. (A) Nanomechanical DNA origami ‘single‐molecule beacons’ directly imaged by AFM. Scale bars, 300 nm. Reproduced with permission.[91] Copyright: 2011, Springer Nature. (B) DNA logic gates (OR, YES, and AND) are performed in response to the stimuli of adenosine triphosphate (ATP) and cocaine by AFM imaging. Scale bars, 200 nm. Reproduced with permission.[113] Copyright: 2016, American Chemical Society (C) Nanocaliper for detecting H⁺ by TEM imaging. Scale bars, 100 nm. Reproduced with permission.[115] Copyright: 2022, Royal Society of Chemistry. (D) The closed and opened state of the device loaded with a fluorophore (6‐FAM) and a quencher (BHQ‐1) for detecting K⁺. Reproduced with permission.[101] Copyright: 2011, Springer Nature. (E) Fluorescence energy transfer for bicolor readout in a split aptamer for detecting ATP. Reproduced with permission.[116] Copyright: 2017, American Chemical Society. (F) DNA origami‐based duplexed aptamers nanoarrays to detect ATP and E2. Reproduced with permission.[119] Copyright: 2022 Elsevier. (G) DNA origami directed assembly of gold bowtie nanoantennas for single‐molecule Surface‐Enhanced Raman Scattering. Reproduced with permission.[122] Copyright: 2018, Wiley‐VCH. (H) Detection of ATP and cocaine (COC) based on dynamic plasmonic system. Reproduced with permission.[123] Copyright:2018, American Chemical Society. (I) Determination of microcystin‐LR using Surface‐Enhanced Raman Spectroscopy. Reproduced with permission.[128] Copyright: 2022 American Chemical Society. (J) Shaped DNA origami carrier nanopore translocation influenced by aptamer‐based surface modification. Reproduced with permission.[131] Copyright:2021, Elsevier.
… 
This content is subject to copyright. Terms and conditions apply.
01/2020
Combining Chemistry and Biology
Accepted Article
Title: Functionalized DNA Origami-Enabled Detection of Biomarkers
Authors: Caiqing Yuan, Fei Zhou, Zhihao Xu, Dunkai Wu, Pengfei Hou,
Donglei Yang, Li Pan, and Pengfei Wang
This manuscript has been accepted after peer review and appears as an
Accepted Article online prior to editing, proofing, and formal publication
of the final Version of Record (VoR). The VoR will be published online
in Early View as soon as possible and may be different to this Accepted
Article as a result of editing. Readers should obtain the VoR from the
journal website shown below when it is published to ensure accuracy of
information. The authors are responsible for the content of this Accepted
Article.
To be cited as: ChemBioChem 2024, e202400227
Link to VoR: https://doi.org/10.1002/cbic.202400227
1
Functionalized DNA Origami-Enabled Detection of Biomarkers
Caiqing Yuan1,2#, Fei Zhou2,3#, Zhihao Xu2,3#, Dunkai Wu1,2, Pengfei Hou1,2, Donglei Yang2, Li
Pan2*, Pengfei Wang2*
1. College of Chemistry and Materials Science, Shanghai Normal University, Shanghai 200233,
China.
2. Institute of Molecular Medicine, Department of Laboratory Medicine, Shanghai Key Laboratory
for Nucleic Acid Chemistry and Nanomedicine, Renji Hospital, School of Medicine, Shanghai Jiao
Ton g Univ ersit y, Shan ghai 200127 , Chi na.
3. School of Life Sciences, Shanghai University, Shanghai 200444, China.
*To whom correspondence should be addressed. Email: pengfei.wang@sjtu.edu.cn,
panli2020@sjtu.edu.cn
#Equal contribution
Abstract
Biomarkers are crucial physiological and pathological indicators in the host. Over the years,
numerous detection methods have been developed for biomarkers, given their significant potential
in various biological and biomedical applications. Among these, the detection system based on
functionalized DNA origami has emerged as a promising approach due to its precise control over
sensing modules, enabling sensitive, specific, and programmable biomarker detection. We
summarize the advancements in biomarker detection using functionalized DNA origami, focusing
on strategies for DNA origami functionalization, mechanisms of biomarker recognition, and
applications in disease diagnosis and monitoring. These applications are organized into sections
based on the type of biomarkersnucleic acids, proteins, small molecules, and ionsand concludes
with a discussion on the advantages and challenges associated with using functionalized DNA
origami systems for biomarker detection.
Key words
DNA self-assembly, functionalized DNA origami, biomarker detection, diagnostics
10.1002/cbic.202400227
Accepted Manuscript
ChemBioChem
This article is protected by copyright. All rights reserved.
14397633, ja, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cbic.202400227 by Shanghai Jiaotong University, Wiley Online Library on [09/06/2024]. 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
2
1. Introduction
Biomarker detection is essential for diagnosing diseases at an early stage, monitoring disease
progression, and assessing the efficacy of treatments [1]. Biomarkers, as quantifiable indicators of
biological states or conditions, are biological molecules found in blood, body fluids, or tissues that
reflect normal, abnormal, or diseased states [2]. They encompass a variety of substances, including
nucleic acids, proteins, small molecules, and ions [3,4].
The field of DNA materials science, particularly in the realm of structural DNA nanotechnology,
has experienced significant growth. This growth is due to a focus on utilizing self-assembled DNA
scaffolds and staples. These assemblies allow for the precise localization and arrangement of
functional molecules and nanomaterials, which is crucial for the development of sophisticated
biosensors. The field is greatly indebted to the pioneering work of Ned Seeman, whose foundational
research established key DNA motifs and self-assembly strategies [5-8]. Today, DNA nanotechnology
diversifies into the construction of nanoscale architectures and functional materials with
applications spanning from molecular mechanics and computing to synthetic chemistry and biology,
all adapting to meet evolving technology needs [9-16]. A part icularly noteworthy b ranch of thi s
technology is DNA origami, an innovative system that facilitates the assembly of two-dimensional
(2D) and three-dimensional (3D) nanostructures with specific shapes and sizes. These structures can
successfully incorporate functional molecules and nanoparticles [15-25].
In diagnostic applications, nucleic acids, including DNA and RNA, serve as prominent
biomarkers, offering critical insights into a broad spectrum of medical conditions [26,27]. Assessing
their presence, quantity, mutations, and expression levels helps healthcare providers to provide
accurate diagnoses, customize treatment plans, and monitor disease progression [28,29]. Traditional
detection methods such as real-time quantitative PCR (qPCR), microarray hybridization, RNA
sequencing, and Northern blotting, are effective but can be complex, costly, and time-consuming
[30-37]. The advent of DNA origami platforms has marked a significant advancement in the detection
of nucleic acid biomarkers. This innovative approach facilitates the integration of multiple binding
sites, allowing for the concurrent capture of various target miRNAs at the nanoscale. The precision
and stability of DNA origami structures notably enhance the sensitivity of disease-related nucleic
acid biomarker detection [38,39].
The cellular function and physiological processes are significantly influenced by proteins,
highlighting the importance of detecting their presence and concentrations for cellular study and
disease diagnostics [40]. Conventional methods for protein detection, such as western blotting,
enzyme-linked immunosorbent assay (ELISA), immunoprecipitation (IP), mass spectrometry (MS)-
based techniques, and protein microarrays employ antibodies or similar agents to capture
biomarkers, with the cumulative interactions contributing to the detection signal. These traditional
methods, despite their widespread use, fall short in distinguishing single immunological interactions
and reflect an average of the combined signals [41-44]. In contrast, DNA origami-based detection
systems can discern proteins at the single-molecule level [45,46]. By assembling precise DNA-based
3D structures, this method can specifically capture and recognize individual protein molecule,
offering unprecedented sensitivity and specificity, especially at very low concentration levels. This
advantage is essential for early disease detection, monitoring treatments, and enabling personalized
medicine, thereby significantly optimizing disease management. Furthermore, DNA origami's
adaptability for portable diagnostics holds global application potential. As the field develops, DNA
origami structures are increasingly utilized in protein structure and function analysis, as well as
10.1002/cbic.202400227
Accepted Manuscript
ChemBioChem
This article is protected by copyright. All rights reserved.
14397633, ja, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cbic.202400227 by Shanghai Jiaotong University, Wiley Online Library on [09/06/2024]. 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
3
biomolecular interaction studies, expanding its utility in protein detection [47].
Beyond nucleic acids and proteins, small molecules and ions are also pivotal biomarkers, serving
as components or by-products of cellular metabolism. They reflect physiological and pathological
states, genetic traits, and exposure to environmental factors [48]. For instance, abnormal glucose
concentrations can suggest diabetes, while variations in ion levels can indicate signal cardiovascular
diseases, electrolyte imbalances, and renal dysfunctions [49,50]. The early detection of metabolic
changes is crucial for aiding in diagnosis, prognosis, and disease management [51,52]. Metabolomic
studies reveal disease mechanisms at the molecular level and can identify potential biomarkers and
therapeutic targets. Traditional detection methods for small molecules and ionic biomarkers, which
are based on spectroscopy, chromatography, and electrochemistry, although capable of quantitative
analysis, provide collective signals from multiple molecules or ions, making it challenging to
achieve detection at the single-molecule level.
Despite current reviews on DNA-based nanomaterials addressing certain aspects of sensor
detection [53-57], the field of detection technology is rapidly evolving, with new methodologies
continuously emerging. Moreover, there appears to be no comprehensive literature review reported
on the sensor detection of biomarkers using functionalized DNA origami, highlighting the ongoing
importance and necessity of such a review. This review provides the latest research advancements
in biomarker detection using functionalized DNA origami. It offers the latest research advancements
in biomarker detection utilizing functionalized DNA origami. It synthesizes strategies for
functionalization, outlines the fundamental mechanisms of biomarker recognition, and explores
significant breakthroughs in the detection techniques for nucleic acids, proteins, small molecules,
and ions.
2. Strategies for DNA origami functionalization
In biosensing applications, DNA origami platforms are engineered through various
functionalization approaches to achieve enhanced recognition capabilities. These methods
encompass the integration of biomolecules such as peptides, nucleic acids, and antibodies; the
incorporation of spectral tags like fluorescent probes; and the attachment of metal nanoparticles,
particularly gold nanoparticles (AuNPs), along with other small molecules such as cholesterol and
methylene blue [25] (Figure 1).
Peptides and nucleic acids are invaluable due to their neutral structural backbones and strong
binding affinity for specific enzymatic, DNA, or RNA substrates. This molecular affinity grants
them unmatched specificity during hybridization processes, rendering them exceptionally suited for
the high-fidelity detection of genetic markers. Antibodies are similarly indispensable, offering a
high degree of selectivity toward antigens. Their precise targeting allows for the development of
DNA origami-based biosensors that can discern proteins and pathogens with remarkable sensitivity,
serving as effective conduits for signal transduction.
Metallic entities, such as AuNPs, are renowned for their ability to enhance biosensor performance
through mechanisms like localized surface plasmon resonance (LSPR) and the amplification of
electrical signals. The conjugation of AuNPs to DNA origami structures has been shown to markedly
increase the detection sensitivity of biosensing assays.
Furthermore, the modification of DNA origami with other molecules, including cholesterol and
methylene blue, is pioneering innovative electrochemical sensing methodologies. [58,59]. These
strategies collectively contribute to the creation of highly sensitive and selective biosensing
10.1002/cbic.202400227
Accepted Manuscript
ChemBioChem
This article is protected by copyright. All rights reserved.
14397633, ja, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cbic.202400227 by Shanghai Jiaotong University, Wiley Online Library on [09/06/2024]. 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
4
platforms tailored for a wide range of analytical applications.
Figure 1. Schematic diagram of functionalized DNA origami for biomarker detection.
3. Mechanisms of biomarker detection
Figure 2. Detection mechanism based on functionalized DNA origami. (A) Watson-Crick base pairing
recognition. (B) Aptamer recognition. (C) Antibody recognition. (D) Morphology-based signal output. (E)
Fluorescence-based signal output. (F) Plasmonics-based signal output. (G) Electrochemistry based signal output.
The recognition of biomarkers using DNA origami as a sensing platform relies on precise
molecular interactions that enhance specificity and sensitivity. The mechanisms underlying this
sensing technology are predominantly categorized into three distinct approaches [60]. The primary
method involves complementary base pairing, utilized predominantly for the detection of nucleic
10.1002/cbic.202400227
Accepted Manuscript
ChemBioChem
This article is protected by copyright. All rights reserved.
14397633, ja, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cbic.202400227 by Shanghai Jiaotong University, Wiley Online Library on [09/06/2024]. 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
5
acid biomarkers (Figure 2A). This entails the strategic design of DNA origami structures with
specific sequences that are complementary to the biomarker of interest. Upon the presence of the
biomarker, specific binding occurs through Watson-Crick base pairing, which is the cornerstone of
this method, and this event is transduced into a measurable signal (Figure 2B). The second approach
capitalizes on the integration of aptamers into DNA origami frameworks. Aptamers, short strands
of DNA or RNA, have an affinity for a broad range of targets and can specifically bind to proteins,
small molecules, and ions [61-63]. This binding is attributed to the aptamers' ability to adopt three-
dimensional conformations that precisely accommodate their targets. The third mechanism employs
antibodies, also called immunoglobulins, which are used as precise recognition elements (Figure
2C). They can be conjugated to DNA origami constructs to achieve targeted detection of antigens
associated with biomarkers, such as proteins or small molecules.
To tr ansl ate the r ecog nition event s i nto o bserva ble out puts, s ever al stra tegies h ave b een
developed, classified into four main types. The initial signal transduction method is based on
morphological alterations of the DNA origami, detectable via Atomic Force Microscopy (AFM) and
Transmission Electron Microscopy (TEM) (Figure 2D). There are two sub-categories here: label-
free detection, where the DNA origami undergoes shape transformations [64] or changes in height
due to hybridization with longer strands facilitated by toehold-mediated strand displacement [65,66],
and labeled detection, manifested by biotin-streptavidin interactions that induce height variations
[67,68]. The second signal output method is fluorescence based, predominantly employing
Fluorescence Resonance Energy Transfer (FRET) (Figure 2E). This involves the incorporation of
donor and acceptor fluorophores within the DNA origami, where binding of a biomarker affects the
proximity or separation of the fluorophores, resulting in measurable alterations in the fluorescence
output [69]. The third method of signal output leverages plasmonic responses through the integration
of metallic nanoparticles such as gold (Figure 2F). These particles can be organized in defined
patterns on the DNA origami scaffolds [70] and may respond to the binding of biomarkers with
assembly or disassembly, detectable by optical property changes [71]. The fourth and final approach
is electrical signal transduction, which is divided into electrochemical detection, using mediators
like methylene blue in DNA origami-based sensors analyzed by cyclic voltammetry [72], and DNA
nanopore technology, where the translocation of molecules through DNA-origami-constructed
nanopores triggers detectable shifts in electrical current [73] (Figure 2G). Individually or in
combination, these mechanisms offer versatile and robust approaches for biomarker detection using
DNA origami. This flexibility allows for customization based on the specific characteristics of the
biomarker, the detection requirements, and the potential applications.
4. Applications of functionalized DNA origami-enabled biomarker detection
In recent years, the employment of functionalized DNA origami constructs has been a focus of
exploration for the targeted detection of an array of biomarkers, encompassing nucleic acids,
proteins, and a diversity of molecular or ionic entities as outlined in Ta b le 1. These sophisticated
DNA origami platforms are adeptly designed to incorporate specific molecular recognition elements
such as aptamers and antibodies. The incorporation of these elements is instrumental in conferring
selectivity and sensitivity, enabling these constructs to selectively bind with and detect their
corresponding biomarkers. A prin cip al ad van tag e of D NA origami lies in it s e xceptional
customizability, offering researchers the ability to meticulously tailor the geometry and the spatial
configuration of molecular recognition moieties within the nanostructures. This precise engineering
augments both the sensitivity and specificity of biomarker detection assays. Moreover, DNA
10.1002/cbic.202400227
Accepted Manuscript
ChemBioChem
This article is protected by copyright. All rights reserved.
14397633, ja, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cbic.202400227 by Shanghai Jiaotong University, Wiley Online Library on [09/06/2024]. 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
6
origami’s inherent programmability permits the construction of multiplexed detection systems.
These systems are sophisticated enough to concurrently screen for multiple biomarkers within a
single assay, thereby increasing throughput and analytical efficiency. The potential of DNA origami
in biomarker detection extends across various sectors, with significant implications for medical
diagnostics, pharmaceutical development, and the burgeoning arena of personalized medicine. The
capacity to design customized platforms that cater to specific biomarker profiles is poised to
revolutionize the methodologies for early disease diagnosis, for tracking the effectiveness of
therapeutic regimens, and for steering the creation of targeted treatment strategies. This innovative
approach represents a stride towards more personalized, precise, and preventive healthcare solutions.
4.1 Functionalized DNA origami for nucleic acid detection.
10.1002/cbic.202400227
Accepted Manuscript
ChemBioChem
This article is protected by copyright. All rights reserved.
14397633, ja, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cbic.202400227 by Shanghai Jiaotong University, Wiley Online Library on [09/06/2024]. 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
7
Figure 3. Functionalized DNA origami for nucleic acid detection. (A) Label-free detection of RNA based on
morphological changes detectable by AFM. Scale bars,150 nm. Reproduced with permission. [64] Copyright: 2008,
AAAS. (B) Label-free unambiguous detection and symbolic display of single nucleotide polymorphisms based on
morphological changes detectable by AFM. Scale bars, 50 nm. Reproduced with permission. [65] Copyright: 2008,
American Chemical Society. (C) Paper-folding mechanism for miRNA detection based on morphological changes
detectable by AFM. Scale bars,200 nm. Reproduced with permission. [74] Copyright: 2023, Springer Nature. (D)
Tumor-associated microRNA detection and fluorescence imaging in live Cells. Reproduced with permission. [69]
Copyright: 2021, Elsevier. (E) DNA origami book biosensor for the detection of cancer-associated nucleic acids
based on Förster resonance energy transfer (FRET). Reproduced with permission. [77] Copyright: 2022, The Royal
Society of Chemistry. (F) DNA origami sensor and DNA-PA IN T ba se d m iR NA d et ec t io n. Scale bars, 100 nm.
Reproduced with permission. [78] Copyright: 2023, Elsevier. (G) Reconfigurable 3D plasmonic metamolecules.
Reproduced with permission. [79] Copyright: 2014, Springer Nature. (H) DNA origami plasmonic nanoantenna for
detection of miRNA. Reproduced with permission. [80] Copyright: 2022, American Chemical Society. (I) Detection
of SARS-CoV-2 RNA with a plasmonic chiral biosensor. Reproduced with permission. [81] Copyright: 2023, Elsevier.
(J) DNA origami gatekeepers for solid-state nanopores. Scale bars, 20 nm. Reproduced with permission. [88]
Copyright: 2012, Wil ey -VCH. (K) Label-free electrochemical biosensors for microRNA detection. Reproduced with
permission. [72] Copyright: 2019, American Chemical Society. (L) Electrochemical DNA Biosensors. Reproduced
with permission. [90] Copyright: 2023, American Chemical Society.
Recent advancements in biomolecular detection have seen the innovative application of
functionalized DNA origami structures for the specific recognition and visualization of nucleic acids.
Ke et al. highlighted the use of self-assembling rectangular DNA origami 'tiles,' which hybridize
with matching probes, allowing for their immobilization on mica substrates and subsequent
detection via AFM imaging (Figure 3A). This approach has evolved to facilitate label-free detection
of RNA molecules [64]. Building on this technology, researchers such as Hari et al. have developed
methods that integrate kinetic precision with DNA origami for the direct visualization of SNPs,
which are common genetic variations (Figure 3B). Their technique employed AFM to visually
decode target nucleotides within a probe sequence through DNA origami designs that graphically
represent nucleotide alphabets, eliminating certain visual elements upon hybridization with the
corresponding probe [65]. In a novel approach, Kim et al. introduced a method to create
reconfigurable DNA origami structures using a paper-folding mechanism, allowing for orthogonal
and reversible folding transitions (Figure 3C). This design has facilitated the detection of miRNAs,
such as miR-107 and miR-155, by altering the origami's configuration selective for the miRNA in
question. These miRNAs are implicated in diseases like Alzheimer's and breast cancer [74].
Complementing structural methods, fluorescence-based techniques have gained prominence [75,76].
Dai et al. integrated DNA origami-localized catalytic hairpin assembly reactions for intracellular
miR-21 detection via fluorescence imaging (Figure 3D). They engineered DNA origami with an
array of hairpins, restricting the reaction and enhancing the spatial resolution [69]. Similarly,
Domljanovic et al. introduced a dynamic DNA origami-based biosensor with arrays of donor and
acceptor fluorophores (Figure 3E). The biosensor produces significant fluorescence changes in
response to miRNA target binding due to alterations in FRET efficiency or fluorescence intensity
[77]. Kocabey et al. developed a DNA origami nanoarray platform for multiplex miRNA detection
using DNA-PAI NT su pe r-resolution microscopy (Figure 3F). This platform can isolate miRNAs
from the blood (plasma) of individuals with breast cancer, offering detection at femtomolar levels
devoid of amplification prerequisites. Notably, this sensor distinguishes between miRNAs and
10.1002/cbic.202400227
Accepted Manuscript
ChemBioChem
This article is protected by copyright. All rights reserved.
14397633, ja, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cbic.202400227 by Shanghai Jiaotong University, Wiley Online Library on [09/06/2024]. 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
8
analogous sequences based on single-nucleotide variances [78].
Utilizing fluorescence for signal output of detecting nucleic acids encounters inherent limitations,
including susceptibility to high background interference, which may compromise the accuracy of
experimental outcomes. Exploring alternative chiral signals, such as optical signals, presents a
promising avenue for enhancing detection performance. Kuzyk et al. engineered an adjustable
plasmonic nanostructure using two gold nanorods on a DNA origami template to actively
manipulate plasmonic signals for 3D configuration control [79] (Figure 3G). Xu et al. and Yu et al.
designed plasmonic nanosensors using gold nanorods and DNA origami for early diagnosis and
targeted therapy of acute kidney injury and precision identification of SARS-CoV-2 RNA,
respectively [80,81] (Figure 3H&I).
Electrochemical biosensors have emerged as a rapid, user-friendly, and promising modality for
the detection of nucleic acids, yielding an enhanced sensitivity [82,83]. Presently, the integration of
nanopores within third-generation sequencing technologies represents a burgeoning area of research,
where DNA origami methodologies are increasingly being leveraged to facilitate nanopore-based
sequencing efforts [84-87]. Wei et al. introduced a sensing mechanism incorporating nanopores within
a DNA origami framework, allowing for ion current signal analysis through solid-state nanopores
(SS-nanopores) [88] (Figure 3J). Concurrently, Farimani et al. explored DNA origami-graphene
heterostructures for enhanced DNA detection [89]. In the realm of conventional electrochemical
detection, Han et al. have developed an innovative, direct, label-free, and amplification-free
electrochemical biosensor, which utilizes DNA probes conjugated with methylene blue (MB) acting
as the redox indicator during hybridization, thereby facilitating the detection of miRNA (Figure
3K). Particularly, this biosensor incorporates cross-shaped DNA origami nanostructures at
strategically targeted locations on each DNA construct, which accommodate multiple single-
stranded DNA probes, thereby enhancing the accessibility and effectiveness of these probes in target
recognition [72]. Moreover, Williamson et al. have delineated a method for electrochemical signal
amplification that is associated with DNA hybridization (Figure 3L), leading to a significant
elevation in the charge transfer resistance (RCT), which is pertinent to the precise identification of
target molecules [90].
4.2 Functionalized DNA origami for proteins detection.
10.1002/cbic.202400227
Accepted Manuscript
ChemBioChem
This article is protected by copyright. All rights reserved.
14397633, ja, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cbic.202400227 by Shanghai Jiaotong University, Wiley Online Library on [09/06/2024]. 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
9
Figure 4. Functionalized DNA origami for proteins detection. (A) Detection of human thrombin based on DNA
protein binding cavities by AFM imaging. Reproduced with permission. [68] Copyright: 2020, Springer Nature. (B)
Detection of human C-reactive protein (CRP) by AFM imaging. Scale bars, 30 nm. Reproduced with permission. [95]
Copyright: 2020, Springer Nature. (C) Detection of TATA-box binding protein by TEM imaging. Reproduced with
permission. [99] Copyright: 2023, The Royal Society of Chemistry. (D) Detection of Gox and HRP by single molecule
fluorescence signal. Scale bars, 50 nm. Reproduced with permission. [102] Copyright: 2016, Springer Nature. (E)
Detection of antibody by fluorescence signal. Reproduced with permission. [103] Copyright: 2021, Elsevier. (F)
Detection of antigen on T cell.[104] Copyright: 2022, Springer Nature. (G) The detection of single peridinin-
chlorophyll a-protein complex based on plasmonic. Reproduced with permission. [108] Copyright: 2018 American
Chemical Society (H) Detection of streptavidin and thrombin based on plasmonic. Scale bars, 5μm. Reproduced
with permission. [109] Copyright: 2022, Springer Nature. (I) The simultaneous detection of four different IgG
antibodies based on solid-state nanopore. Reproduced with permission. [110] Copyright: 2016, Springer Nature. (J)
Trapping of small prot eins in the nanopore electro-osmotic trap. Reproduced with permission. [73] Copyright: 2022,
American Chemical Society. (K) Detection of streptavidin and PDGF-BB based on square wave voltammetry signal.
Scale bars, 100 nm. Reproduced with permission. [82]
10.1002/cbic.202400227
Accepted Manuscript
ChemBioChem
This article is protected by copyright. All rights reserved.
14397633, ja, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cbic.202400227 by Shanghai Jiaotong University, Wiley Online Library on [09/06/2024]. 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
10
Morphological changes in DNA origami serve as an effective means for detecting specific
proteins at the individual level. The advent of nanomechanical devices, capable of interrogating
target proteins on a single-molecule basis, is instrumental in such detection. The binding event
between a target protein and an aptamer on a functionalized DNA origami structure results in a
discernible alteration of the origami's morphology. This alteration can be precisely captured in both
two and three dimensions via advanced imaging techniques such as AFM or TEM, facilitating the
accurate identification of the protein in interest. Extensive literature documents numerous strategies
employed for this type of detection. Notably, Kuzuya et al. engineered a nanomechanical DNA
origami device that operated as a beacon for single molecules, enabling the visualization of
individual streptavidin (SA) and Immunoglobulin G (IgG) molecules through the origami structure's
transformative reshaping [91]. Likewise, Godonoga et al. integrated DNA aptamers with
programmable molecular mechanisms to detect Plasmodium falciparum lactate dehydrogenase
(PfLDH), a significant biomarker in malaria research [92]. Fujita et al. developed DNA origami-
based thick filaments that allowed for direct observation of the movement of molecular motor
protein heads, providing valuable insights into their force-generating mechanisms [93]. Moving
further, Zhang et al. assembled a triangular DNA origami scaffold with strategically anchored
artificial epitopes to capture the transient conformations of IgG molecules at ambient temperatures
using AFM [94]. Rafat et al. demonstrated the potential of DNA scaffold structures to construct
multivalent binding sites, exhibiting multiple aptamer ligands, tailored for capturing human
thrombin (Figure 4A) [68]. Raveendran et al. used a DNA-origami biosensor platform that had a
central cavity with a target-specific DNA aptamer to enable detection of human C-reactive protein
(CRP) in clinically relevant fluids (Figure 4B) [95]. Ve n ez i a n o e t a l . e x h ib i t e d eOD-GT8 clinical
vaccine immunogen on a DNA origami nanoparticle to systematically study the effects of nanoscale
parameters on B-cell activation in vitro [96]. Moreover, Hellmeier et al. created a DNA origami-based
biological interface that could experimentally ascertain distances among T-cell ligands, thus
correlating with molecular dynamics of antigen interaction and T-cell responses [97]. To all eviat e th e
costs associated with functionalized DNA origami, Oktay et al. suggested a cost-effective strategy
employing asymmetric asymmetric polymerase chain reaction PCR (aPCR) for direct synthesis of
customized scaffolds which can be precisely modified [98]. Natarajan et al. exploited the DNA
origami method to develop a nanoscale device to examine proteins that induce DNA bending, using
TATA-box binding protein (TBP) as a model system whose interaction with the TATA box induces
a measurable curvature (Figure 4C) [99].
Fluorescence-based methods have emerged as a popular approach for proteins detection, with
DNA origami offering a versatile platform for such analyses. Iwaki et al. introduced a programmable
DNA origami nanospring, which, when applied to human myosin VI, demonstrated nanometer-
precision single-molecule fluorescence imaging of individual motor domains under force [100]. Ke et
al. developed a rhombus-shaped nanoactuator integrated with split eGFP (enhanced green
fluorescent protein) to form a DNA-protein hybrid nanostructure. This structure exhibits adjustable
fluorescent properties and functions as a sensor, responding to enzymes [101]. Zhao et al. utilized a
DNA nanocage for the efficient encapsulation of enzymes such as Gox and HRP, achieving accurate
quantification of encapsulated enzymes through single-molecule fluorescence characterization
(Figure 4D) [102]. Pfeiffer et al. integrated DNA origami nanoantennas with a label-free antibody
detection system, featuring a nanoswitch that generates a fluorescent signal upon antibody binding,
thereby enabling single-antibody detection (Figure 4E) [103]. Sun et al. created a multimeric T cell
10.1002/cbic.202400227
Accepted Manuscript
ChemBioChem
This article is protected by copyright. All rights reserved.
14397633, ja, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cbic.202400227 by Shanghai Jiaotong University, Wiley Online Library on [09/06/2024]. 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
11
identification system using 2D DNA origami scaffolds to arrange pMHCs (dorimers) with nanoscale
precision, significantly enhancing the binding avidity for low-affinity antigen-specific T cell
receptors (TCRs) (Figure 4F) [104]. Selnihhin et al. employed DNA nanotechnology to engineer
barrel-shaped DNA-origami nanobeads for quantifying fluorescence and antigens in flow cytometry,
obtaining a count of 48 ± 11 envelope surface protein (Env) trimers per MLV [105]. Seitz et al. used
an optically responsive protein coating on DNA origami for stimulus-triggered antigen targeting [106].
Chen et al. developed a chemical nose sensor using multiple-aptamer-integrated DNA origami
(MADO) probes to discern and distinguish cancer cells. By adjusting the composition and quantity
of aptamers, they realized five distinct MADO probes with varying binding affinities ranging from
3.08 to 78.92nM for five specific types of cells (HeLa, MDA-MB-468, MCF-7, HepG2, and MCF-
10A) [107].
In the realm of plasmonics, Kaminska et al. assembled hybrid structures with natural light-
harvesting complexes positioned within hotspot zones facilitated by DNA origami (Figure 4G) [108].
Schuknecht et al. reported utilization of a DNA origami scaffold to achieve tip-to-tip alignment of
gold nanorods, resulting in an average gap size of 8 nm. These gaps provided access to streptavidin
and thrombin, which were selectively captured at the plasmonic hotspot through specific anchoring
sites on the origami template. The achieved field enhancement for the nanorod dimers was deemed
adequate for conducting single-protein Surface-Enhanced Raman Spectroscopy (SERS) with
integration times of less than one second (Figure 4H) [109].
Solid-state nanopores serve as single-molecule protein sensors with the capability of swiftly
obtaining substantial statistical data on a sample in solution, without the need for labeling. The
fundamental detection method involves analyzing fluctuations in ionic current as molecules traverse
a nanopore under an applied potential. It yields information regarding different properties of proteins,
including its charge, molecular weight, and conformation. The single-molecule nature of the
measurement implies the potential to discern the characteristics of individual subpopulations within
a complex mixture, thereby enabling the simultaneous detection of multiple proteins. Bell et al.
engineered a series of DNA nanostructures, each featuring a distinct barcode, with individual bits
represented by the presence or absence of multiple DNA dumbbell hairpins. They demonstrated that
a 3-bit barcode can be accurately identified with 94% precision by electrophoretically transporting
the DNA structures through a solid-state nanopore. Furthermore, this enabled the simultaneous
detection of four different IgG antibodies at nanomolar concentration levels (Figure 4I) [110].
Comparatively, the NEOtrap is comprised of a DNA-origami sphere that was immobilized onto a
passivated solid-state nanopore under the application of a positive bias voltage (on the trans side).
Upon immobilization, the DNA-origami sphere, possessing a highly negative charge, induced an
electro-osmotic flow (EOF), facilitating the entrapment of a target protein. Wen et al. employed site-
specific cholesterol functionalization of the origami sphere to establish a linkage with the lipid-
coated nanopore, affording the ability to immobilize the origami in either a vertical or horizontal
orientation, thereby effectively modulating the electro-osmotic flow (EOF). The optimized EOF
substantially improves trapping efficiency, resulting in reduced noise levels, decreased
measurement heterogeneity, an enhanced capture rate, and an extension of observation times by
100-fold. They demonstrated the trapping of various single proteins, including those as small as 14
kDa. By incorporating cholesterol functionalization, the NEOtrap technology's application range
has been significantly expanded (Figure 4J) [73]. Jeon et al. developed modular electrochemical
sensors based on DNA origami, using standard electrochemical techniques such as square wave
10.1002/cbic.202400227
Accepted Manuscript
ChemBioChem
This article is protected by copyright. All rights reserved.
14397633, ja, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cbic.202400227 by Shanghai Jiaotong University, Wiley Online Library on [09/06/2024]. 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
12
voltammetry to measure signals and achieve detection of streptavidin and Platelet-Derived Growth
Factor (PDGF-BB) (Figure 4K) [82].
4.3 Functionalized DNA origami for small molecules or ions detection.
Figure 5. Functionalized DNA origami for small molecule or ion detection. (A) Nanomechanical DNA origami
‘single-molecule beacons’ directly imaged by AFM. Scale bars, 300 nm. Reproduced with permission. [91] Copyright:
2011, Springer Nature. (B) DNA logic gates (OR, YES, and AND) are performed in response to the stimuli of
adenosine triphosphate (ATP) and cocaine by AFM imaging. Scale bars, 200 nm. Reproduced with permission. [113]
Copyright: 2016, American Chemical Society (C) Nanocaliper for detecting H+ by TEM imaging. Scale bars, 100
nm. Reproduced with permission. [115] Copyright: 2022, Royal Society of Chemistry. (D) The closed and opened
state of the device loaded with a fluorophore (6-FAM ) an d a que nc he r (BH Q-1) for detecting K+. Reproduced with
permission. [101] Copyright: 2011, Springer Nature. (E) Fluorescence energy transfer for bicolor readout in a split
aptamer for detecting ATP. Reproduced with permission. [116] Copyright: 2017, American Chemical Society. (F)
10.1002/cbic.202400227
Accepted Manuscript
ChemBioChem
This article is protected by copyright. All rights reserved.
14397633, ja, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cbic.202400227 by Shanghai Jiaotong University, Wiley Online Library on [09/06/2024]. 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
13
DNA origami-based duplexed aptamers nanoarrays to detect ATP and E2. Reproduced with permission. [119]
Copyright: 2022 Elsevier. (G) DNA origami directed assembly of gold bowtie nanoantennas for single-molecule
Surface-Enhanced Raman Scattering. Reproduced with permission. [122] Copyright: 2018, Wiley-VCH. (H) Detection
of ATP and cocaine (COC) based on dynamic plasmonic system. Reproduced with permission. [123] Copyright:2018,
American Chemical Society. (I) Determination of microcystin-LR using Surface-Enhanced Raman Spectroscopy.
Reproduced with permission. [128] Copyright: 2022 American Chemical Society. (J) Shaped DNA origami carrier
nanopore translocation influenced by aptamer-based surface modification. Reproduced with permission. [131]
Copyright:2021, Elsevier.
The use of functionalized DNA origami for visualizing morphological changes to detect small
molecules or ions as biomarkers has been a subject of exploration [111-115]. It is found that literatures
on the detection of small molecules and ions biomarkers as biomarkers using DNA origami are
scarce; however, the principles of the methodology are globally applicable. Therefore, the reviewed
literature encompasses broader applications, including detection of small molecules and ions even
not biomarkers using DNA origami. Vo i g t e t a l . e m p l o y e d A F M t o s t u d y c h e m i c a l c o u p l i n g
reactions on DNA origami structures, effectively demonstrating that these nanostructures are
amenable to various post-assembly chemical modifications [111]. Kuzuya et al. innovatively designed
a functional DNA origami mechanism featuring a clamp, which changes structure in response to
environmental changes or specific molecular interactions, enabling the detection of Na+ and Ag+
through AFM imaging [112]. In 2014, they introduced a pH sensor utilizing three distinct forms of
DNA origami devices, which requires only a minimal sample volume (approximately 0.1 µL) for
effective pH detection, with sensitivity rivaling that of conventional meters (Figure 5A) [91]. Ya ng
et al. developed an adenosine triphosphate (ATP)-responsive DNA logic gate sensor with a
geometric configuration that is programmable and adjustable (Figure 5B) [113]. To a ddre ss t he heal th
risks of Aflatoxin B1 (AFB1), known for its carcinogenicity and toxicity, Lu et al. combined adapter-
labelled DNA origami with single-stranded DNA functionalized gold nanoparticles (AuNP) for the
visual detection of AFB1 using AFM [114]. This integrated approach leverages the targeting precision
of DNA origami, the recognition capabilities of aptamers, and the detectability of AuNP, opening
new possibilities for DNA origami in small molecule detection. Zhang et al. presented the DNA
origami nanocaliper, a nanomechanical device with pH-responsive triple-stranded DNA arms that
change shape in response to local pH levels, as visualized through transmission electron microscopy
(TEM) imaging (Figure 5C) [115]. These nanocalipers hold potential as universal platforms for
nanoscale pH measurement.
In the field of functionalized DNA detection for molecules or ions, fluorescence offers a more
accessible detection method compared to techniques that require complex and costly instruments
like AFM and TEM. Ke et al. developed a DNA-protein hybrid nanostructure that responds to
specific stimuli, such as K+, by inducing conformational changes, which in turn modulate
coordinated fluorescence behavior (Figure 5D) [101]. Wal te r et al. also de si gn ed a sys tem wh er e th e
presence of ATP as a target molecule triggers a shape transformation in DNA origami, resulting in
a fluorescence color change from green to red as a clear readout (Figure 5E) [116]. In another
innovative approach, Huang et al. created a polarized light sensing device based on DNA origami
[117]. This device takes advantage of the spatial recombinability of DNA origami nanostructures to
generate customized sensing probes, integrating the unique optical response of polarized light with
the high affinity and selectivity of the adapter. Cervantes et al. utilized DNA nanostructures to
address instability issues in steroid hormone detection, achieving this through the ingenious design
10.1002/cbic.202400227
Accepted Manuscript
ChemBioChem
This article is protected by copyright. All rights reserved.
14397633, ja, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cbic.202400227 by Shanghai Jiaotong University, Wiley Online Library on [09/06/2024]. 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
14
of single-molecule sensing platforms [118]. Additionally, Wang et al. reported the development of a
self-assembled DNA origami biosensor that enables the precise localization of fluorescent
oligonucleotides [119]. The binding of ATP dimer oligonucleotides to triangular DNA origami
through complementary base pairing enhances fluorescence signals on the origami array. This
method has also shown promise in detecting 17β-estradiol (E2) (Figure 5F).
Within the field of plasmonics, several methods utilizing functionalized DNA origami for the
detection of small molecules and ions have been developed. Thacker et al. developed SERS sensors
using DNA origami technology, based on the dimerization of gold nanoparticles [120]. This approach
enabled the detection of trace amounts of dye molecules with a significant local field enhancement.
Puchkova et al. introduced self-assembled DNA origami-based optical nanoantennas that, through
enhanced robustness and optimized quantum yield, achieved over 5000-fold fluorescence
enhancement, facilitating single-molecule detection even in the presence of a 25 μM background
fluorophore concentration [121]. Zhan et al. fabricated plasmonic bowtie nanostructures using a DNA
origami-based bottom-up assembly approach [122], allowing precise control over the bowtie's
geometric arrangement and creating an approximate 5 nm gap. Within this gap, a single Raman
probe was positioned, enabling the observation of single-molecule surface-enhanced Raman
scattering (SM-SERS) from individual nanostructures, including those with an alkyne group
(Figure 5G). Zhou et al. proposed a dual response system capable of non-invasive regulation over
a broad temperature range, linking thermal control with light signal changes for the detection of
small molecules like ATP (Figure 5H) [123]. Despite the development of customized metal
nanoclusters for manipulating light at the sub-wavelength scale in nanophotonics, positioning
molecules in hotspots with a fixed number and specific arrangement remains challenging. Fang et
al. showed that DNA origami supramolecular structures with Fermi resonance (DMFR) could
accurately locate individual dye molecules, generating quantifiable SERS reactions [124]. Kaur et al.
conducted experiments combining DNA origami technology with plasma-coupled bimetallic
nanostars. In 2020, they detected pyocyanin with a sensor recognition and signal amplification
approach, achieving a detection limit of 335 pM [125]. The following year, they investigated the
enhancement effect of dimers and trimers on Raman signals, enabling sensitive detection of
dopamine at the picomol level [126]. In 2023, it was observed that the fluorescence signal of Cy3 dye
in the sensor increased by 65 times [127]. Huo et al. enhanced SERS detection technology by
combining the sensing strategy of DNA origami/AuNPs [128]. A detection tec hnology ba sed on
adenosine triphosphate (ATP) responsive-strand displacement (ARSD) design was developed for
rapid microcystin-LR (MC-LR) analysis, providing a linear response within the concentration range
of 1.56-50 μg·L-1, with a limit of detection (LOD) of 0.29 μg·L-1 (Figure 5I). Similarly, Kaur et al.
designed customized arrangements of gold nanocone (Au NBP) monomers and dimer nanoantennas
using DNA origami technology for selective recognition of the amyloid protein marker thioflavin T
(ThT) [129]. Li et al. leveraged the specificity and spatial characteristics of DNA origami to form
plasmonic dimer nanoantennas for ultra-sensitive detection of trace amounts of diethylstilbestrol
(DES) [130].
DNA origami has been extensively used as a translocation conduit in the analysis of SS-nanopores,
including both soft linear origami carriers and specially designed structures. In the field of nanopore
sensing with linear origami carriers, molecular modifications that induce minor structural and
charge changes can lead to significant fluctuations in translocation signals, thus enhancing the
sensitivity of single-molecule detection. Ding et al. reported a surface modification strategy that
10.1002/cbic.202400227
Accepted Manuscript
ChemBioChem
This article is protected by copyright. All rights reserved.
14397633, ja, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cbic.202400227 by Shanghai Jiaotong University, Wiley Online Library on [09/06/2024]. 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
15
employs aptamer/target binding to affect the translocation of shaped origami ribbon carriers through
SS-nanopores [131]. (Figure 5J). Their study showed that translocation signal variations were due to
AT P /a p t a m er i nt e r a ct i o n s o n t he c ar r i e r s u r f a ce a nd m od i f i c at i o n s f r om s p at i a l d i s t ri b u t io n a nd
enzyme catalysis. These findings indicate the potential for detecting small spatial and electronic
changes on DNA origami through SS-nanopore analysis. The surface aptamer-binding strategy to
influence origami translocation could be applied to a broader range of SS-nanopore sensing and
detection applications.
4.4 Methods Summary
Given the above discussion on the use of functionalized DNA origami for nucleic acid, proteins,
and small molecule detection, we have summarized and categorized the information based on the
type of functionalized origami, the biological markers detected, the signal output mode, and the
detection performance. The results are shown in Ta bl e 1 .
Biomarkers
Signal Output
Detection Performance
Ref
Nucleic Acid
Detection
C-myc, Rag-1, β-actin
AFM
LOD: ~ 1000 molecules
64
Single nucleotide
polymorphisms
AFM
A direct visual readout of the target
nucleotide contained in the probe
sequence
65
miR-107, miR-155
AFM
Separate or orthogonal miRNA detection
74
miR-21
Fluorescence
LOD: 20.9 pM
69
miR-21, let-7a
Fluorescence
LOD: 1-10 pM
77
miR-153, let-7a, miR-
155, miR-142
Fluorescence
LOD:11 fM - 388 fM
78
DNA fuel strands
Circular
dichroism
changes
CD signal is as large as 250 mdeg at a
sample concentration of only 0.5 nM
79
miR-21
Photoacoustic
(PA) imaging
LOD:2.8 nM
80
SARS-CoV-2 RNA
Plasmonic chiral
signal
LOD:0.133 aM
81
dsDNA
Electrical signal
Single-molecule sensing
88
miRNA-21
Electrochemical
signal
LOD:79.8 fM
72
DNA, RNA
Electrochemical
signals
LOD: 8.86 pM
Linear range:10 pM-10nM
90
Proteins
Detection
Human thrombin, SA
AFM
A strong increase in binding yield: ~95%
for streptavidin
68
CRP
AFM
LOD: 3 nM
95
TBP, TF(II)A, TF(II)B
TEM
Direct observation of the changes caused
by DNA bending proteins
99
HRP, MDH, G6PDH,
LDH, GOx, β-Gal
Single-molecule
fluorescence
Five of them (GOx, HRP, G6PDH, MDH
and LDH) exhibited higher activity in
nanocages than the free enzyme, with
enhancements ranging from 3- to 10-fold.
102
Anti-digoxigenin
Fluorescence
Fluorescence enhancement values up to
63-fold
103
Antigen-specific CD8+
T cell receptors(TCRs)
Fluorescence
Enhance the binding avidity (Kd:1.8-86.9
nM; Koff: 2.55 × 10−4 -0.56 × 10−4 s−1)
104
Peridinin-Chlorophyll
a-Protein
Fluorescence
Fluorescence enhancement of 500-fold
108
SA, thrombin
SERS
Linear range: 330 nM-4.2 µM (SA)
6.9-61 µM (thrombin)
109
Anti-Biotin, anti-
BrdU, anti-puromycin
and anti-digoxigenin
Electrical signal
Multiple antibodies can be simultaneously
detected at nanomolar concentration
levels.
110
Ribonuclease A,
carbonic anhydrase,
ovalbumin, avidin,
dCas9, and ClpP
Electrical signal
Single proteins, small ones down
to14kDa
73
SA, PDGF-BB
Electrochemical
signal
Linear range: 5 pM-1nM (SA);
LOD<1pM (SA).
The resulting sensor achieved detection of
PDGF-BB at concentrations as low as 500
pM with a signal change of 20%.
82
Small Molecules
or Ions
Detection
Na+K+
AFM
Single-molecule level
91
ATP a nd Cocaine
AFM
DNA logic gates (OR, YES, and AND)
are performed in response to the stimuli
113
10.1002/cbic.202400227
Accepted Manuscript
ChemBioChem
This article is protected by copyright. All rights reserved.
14397633, ja, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cbic.202400227 by Shanghai Jiaotong University, Wiley Online Library on [09/06/2024]. 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
16
of ATP ((1 mM)) and cocaine (0.4 mM).
H+
TEM
pH sensing at the nanoscale
115
K+
Fluorescence
100 mM K+ shows a characteristic angle
distribution ranging from 10° to 25°
101
ATP
Fluorescence
It is sensitive in a concentration range
between 0.10 mM and 1.00 mM ATP
116
ATP
Fluorescence
Linear range: 0.1-100 ng·mL-1
LOD: 0.29 ng·mL-1
119
Cy5
SM-SERS
Single-molecule level
122
ATP
CD spectra
Upon addition of 0.05 mM ATP molecules,
the CD intensity exhibits an obvious
increase.
123
MC-LR
SERS
Linear range: 1.56-50 μg·L-1
LOD: 0.29 μg·L-1
128
ATP
Electrical signal
The binding of ATP with aptamer can
induce structural and electronic charge
changes on the origami carrier surface.
131
Ta bl e 1 . Applications of functionalized DNA origami-enabled biomarker detection.
5. Summary and Outlook
The utilization of DNA origami for biomarker detection represents a promising and rapidly
evolving area of research with substantial implications for the future of diagnostics. As the field
continues to advance, further exploration of DNA origami-based detection platforms is expected to
contribute to the development of highly sensitive, specific, and multiplexed biomarker detection
technologies with broad applications in biomedical research and clinical practice. However, despite
its compelling potential, the application of DNA origami in biomarker detection is not without its
limitations and challenges. Challenges in the field include optimizing the stability and
reproducibility of DNA origami-based detection platforms, further improving the detection
sensitivity and multiplex capability, as well as integrating these platforms with readout technologies
for signal transduction and quantification, and expanding scalability for commercial application.
Despite these challenges, the potential of DNA origami for biomarker detection has sparked
significant research efforts, driving the exploration of innovative detection strategies and
bioanalytical applications. Based on the existing research, we propose the following areas for
improvement and clinical potential:
Improving stability and robustness of DNA origami sensors. DNA origami structures, while
robust under specific conditions, can be susceptible to degradation by nucleases present in biological
samples or the environment. Therefore, improving the stability and robustness of DNA origami
sensors is crucial for their practical applications. Several strategies can be employed to achieve this
goal. Firstly, optimization of DNA origami design. Careful design of the DNA origami structure
can enhance its stability and robustness. This includes optimizing staple strand lengths, sequences,
and crossovers to minimize structural defects and undesirable interactions. Secondly, incorporation
of stabilizing modifications, such as phosphorothioate linkages or nucleotide analogs, can be
introduced into the DNA strands to enhance stability against nuclease degradation and improve
overall robustness. Thirdly, coating the DNA origami structures with protective layers or
encapsulating them within inert materials can provide additional stability and protection against
environmental factors, such as changes in pH, ionic strength, or temperature. Fourthly, introducing
chemical crosslinking or utilizing stabilizing agents can help reinforce the structural integrity of
DNA origami, making it more resistant to denaturation and degradation. Fifthly, developing
improved purification techniques can remove impurities and incomplete structures, leading to a
higher percentage of intact and stable DNA origami sensors. Lastly, engineering DNA origami
structures with self-healing capabilities, such as dynamic reconfiguration or responsiveness to
external stimuli, can contribute to their stability and robustness by repairing structural damages.
10.1002/cbic.202400227
Accepted Manuscript
ChemBioChem
This article is protected by copyright. All rights reserved.
14397633, ja, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cbic.202400227 by Shanghai Jiaotong University, Wiley Online Library on [09/06/2024]. 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
17
Improving the detection sensitivity. Further developing signal amplification techniques based
on functional DNA origami. This development could involve using enzymatic reactions or
incorporating amplification strands, which can significantly increase the signal in the presence of
the target, making it easier to detect multiple targets at low concentrations. Developing more
sensitive signal probes by exploring the integration of DNA origami structures with novel
nanomaterials such as 2D materials (e.g., graphene, MXenes), quantum dots with unique optical
properties, or upconversion nanoparticles. These materials could provide new pathways for signal
amplification and detection. Enhancing signal processing algorithms by improving those used to
process and analyze the signals obtained from DNA origami sensors. Utilizing sophisticated data
analysis techniques to filter noise and enhance signal detection could significantly boost sensitivity.
Improving multiplexed detection capabilities. Given their sizable and customizable surface
area, DNA origami structures can host multiple distinct recognition sites on a single construct. This
multiplexing capability enables the simultaneous detection of several targets, facilitating complex
analyses that can be particularly beneficial in applications like metabolomics or environmental
monitoring, where a broad spectrum of analytes needs to be assessed. There are several strategies
to improve multiplex detection capabilities. One method is to optimize DNA origami design.
Enhance the design of DNA origami structures to include unique and specific binding sites for
different target molecules. This involves careful planning of the spatial arrangement to minimize
cross-reactivity and ensure efficient target binding. Another potential method is to develop systems
capable of simultaneous detection and differentiation of multiple targets. This can be achieved by
assigning specific fluorescent markers or electrochemical signals to each target, allowing for parallel
processing and analysis. Moreover, emerging bioinformatics analysis technologies offer powerful
tools for signal analysis, with great potential to significantly improve multiplex detection
capabilities
Integration with microfluidics and other analytical platforms. Functionalized DNA origami
holds the potential to be seamlessly integrated into microfluidic devices or combined with other
nanomaterials to develop advanced analytical tools. This integration can lead to enhanced
performance, including faster analysis times and reduced sample requirements, thus opening new
avenues for point-of-care diagnostics and real-time environmental monitoring.
Potential impact on clinical diagnostics. DNA origami sensing technology may facilitate the
development of point-of-care diagnostic devices that can deliver rapid and accurate results at the
bedside or in remote settings. This could improve patient access to timely diagnostic information
and enable faster decision-making for healthcare providers. Moreover, DNA origami sensors can be
utilized for the continuous monitoring of disease progression, therapeutic efficacy, and the
recurrence of specific biomarkers over time. This could aid in disease management and the ongoing
assessment of treatment outcomes. And it also could lead to personalized health monitoring devices
that detect and potentially respond to changes in an individual's health status.
Acknowledgements
The authors thank the National Key Research and Development Program of China
(2021YFA0910100), the National Natural Science Foundation of China (82372349, 22304112), and
the Innovative Research Team of High-Level Local Universities in Shanghai (SHSMU-
ZLCX20212602) for their support.
10.1002/cbic.202400227
Accepted Manuscript
ChemBioChem
This article is protected by copyright. All rights reserved.
14397633, ja, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cbic.202400227 by Shanghai Jiaotong University, Wiley Online Library on [09/06/2024]. 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
18
Conflict of interest
The authors declare no conflict of interest
References
[1] R. L. Siegel, K. D. Miller, N. S. Wagle, A. Jemal, CA A Cancer J Clinicians 2023, 73, 17-48.
[2] Biomarkers Definitions Working Group, Clinical Pharmacology & Therapeutics 2001, 69, 89-
95.
[3] S. Li, H. Zhang, M. Zhu, Z. Kuang, X. Li, F. Xu, S. Miao, Z. Zhang, X. Lou, H. Li, F. Xia, Chem.
Rev. 2023, 123, 7953-8039.
[4] Y. Y. Br o z a , X . Z h o u , M. Yu a n , D. Q u , Y. Z h e n g , R. V i s h i n k i n , M . K h a t i b , W. W u , H . Ha i c k ,
Chem. Rev. 2019, 119, 11761-11817.
[5] N. C. Seeman, Journal of Theoretical Biology 1982, 99, 237-247.
[6] J. Chen, N. C. Seeman, Nature 1991, 350, 631-633.
[7] T. J. Fu, N. C. See ma n, Biochemistry 1993, 32, 3211-3220.
[8] S. Dey, C. Fan, K. V. Gothelf, J. Li, C. Lin, L. Liu, N. Liu, M. A. D. Nijenhuis, B. Saccà, F. C.
Simmel, H. Yan, P. Zhan, Nature Reviews Methods Primers 2021, 1, 13.
[9] S. Liu, Q. Jiang, X. Zhao, R. Zhao, Y. Wang, Y. Wang, J. Liu, Y. Shang, S. Zhao, T. Wu, Y. Zhang,
G. Nie, B. Ding, Nat. Mater. 2021, 20, 421-430.
[10] A.-K. Pumm, W. Engelen, E. Kopperger, J. Isensee, M. Vogt, V. Kozina, M. Kube, M. N.
Honemann, E. Bertosin, M. Langecker, R. Golestanian, F. C. Simmel, H. Dietz, Nature 2022, 607,
492-498.
[11] S. Jiang, Z. Ge, S. Mou, H. Yan, C. Fan, Chem 2021, 7, 1156-1179.
[12] L. Shen, P. Wang, Y. Ke, Adv Healthcare Materials 2021, 10, 2002205.
[13] Y. Wa n g , E . B e n s o n , F. F ö r d ő s , M . L o l a i c o , I. Baars, T. Fang, A. I. Teixeira, B. Högberg,
Advanced Materials 2021, 33, 2008457.
[14] E. Silvester, B. Vollmer, V. Pražák, D. Vasishtan, E. A. Machala, C. Whittle, S. Black, J. Bath,
A. J. Turberfield, K. Grünewald, L. A. Baker, Cell 2021, 184, 1110-1121.e16.
[15] F. Xu , Q. X ia , J. Ye, L . Do ng , D. Ya ng , W. Xu e, P. Wa ng , Nano Lett. 2022, 22, 9935-9942.
[16] Z. Zhou, Y. S. Sohn, R. Nechushtai, I. Willner, ACS Nano 2020, 14, 9021-9031.
[17] P. W . K . Ro t h e mu n d , Nature 2006, 440, 297-302.
[18] P. W an g , T. A . M e y e r, V. P a n , P. K . D u t t a, Y. K e , Chem 2017, 2, 359-382.
[19] E. S. Andersen, M. Dong, M. M. Nielsen, K. Jahn, R. Subramani, W. Mamdouh, M. M. Golas,
B. Sander, H. Stark, C. L. P. Oliveira, J. S. Pedersen, V. Birkedal, F. Besenbacher, K. V. Gothelf, J.
Kjems, Nature 2009, 459, 73-76.
[20] Y. K e , J . S h a r m a , M . L i u , K . J a h n , Y. L i u , H . Ya n , Nano Lett 2009, 9, 2445-2447.
[21] S. M. Douglas, H. Dietz, T. Liedl, B. Högberg, F. Graf, W. M. Shih, Nature 2009, 459, 414-
418.
[22] D. Han, S. Pal, J. Nangreave, Z. Deng, Y. Liu, H. Yan, Science 2011, 332, 342-346.
[23] B. Kou, Z. Wang, S. Mousavi, P. Wang, Y. Ke, Small 2023, 2308862.
[24] Y. Y a n g , Q . L u , C . H u a n g , H . Q i a n , Y. Z h a n g , S . D e s h p a n d e , G. A r y a , Y. K e , S . Z a u s c h e r ,
Angewandte Chemie 2021, 133, 23429-23435.
[25] M. Dass, F. N. Gür, K. Kołątaj, M. J. Urban, T. Liedl, J. Phys. Chem. C 2021, 125, 5969-5981.
[26] K. Zheng, Z.-H. You, J.-Q. Li, L. Wang, Z.-H. Guo, Y.-A. Huang, PLoS Comput Biol 2020, 16,
e1007872.
10.1002/cbic.202400227
Accepted Manuscript
ChemBioChem
This article is protected by copyright. All rights reserved.
14397633, ja, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cbic.202400227 by Shanghai Jiaotong University, Wiley Online Library on [09/06/2024]. 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
19
[27] Y. N a k a b e p p u , D . T s u c h i m o t o , H . Y a m a g u c h i , K . S a k u m i , J of Neuroscience Research 2007,
85, 919934.
[28] Q. Ren, L. Jiang, S. Ma, T. Li, Y. Zhu, R. Qiu, Y. Xing, F. Yin, Z. Li, X. Ye, Y. Zhang, M. Zhang,
Advanced Materials 2023, 35, 2304119.
[29] L.-Y. L i u , Y. Z h a o , N . Z h a n g , K . -N. Wang, M. Tian, Q. Pan, W. Lin, Anal. Chem. 2021, 93,
16121619.
[30] S.-S. Lee, J.-H. Park, Y.-K. Bae, Clinica Chimica Acta 2021, 521, 9-18.
[31] J. Aschenbrenner, A. Marx, Nucleic Acids Res 2016, 44, 34953502.
[32] A. A. Kojabad, M. Farzanehpour, H. E. G. Galeh, R. Dorostkar, A. Jafarpour, M. Bolandian,
M. M. Nodooshan, Journal of Medical Virology 2021, 93, 4182-4197.
[33] J. Andreani, J. Lupo, B. Nemoz, A. Truffot, S. Larrat, P. Morand, R. Germi, Infectious Diseases
Now 2022, 52, 453-455.
[34] M. Awashra, P. Elomaa, T. Ojalehto, P. Saavalainen, V. Jokinen, Adv Materials Inter 2024, 11,
2300596.
[35] G. Ventimiglia, M. Pesaturo, A. Malcolm, S. Petralia, Biosensors 2022, 12, 563.
[36] G. Oudeng, M. Benz, A. A. Popova, Y. Zhang, C. Yi, P. A. Levkin, M. Yang, ACS Appl. Mater.
Interfaces 2020, 12, 55614-55623.
[37] A. Rumlow, E. Keunen, J. Klein, P. Pallmann, A. Riemenschneider, A. Cuypers, J. Papenbrock,
PLoS ONE 2016, 11, e0163679.
[38] W. Ji an g , J . L i, Z. L in , J. Gu o, J . M a, Z. Wan g , M . Z h an g, Y. Wu , Small Structures 2023, 4,
2200376.
[39] M. Shahhosseini, P. E. Beshay, E. Akbari, N. Roki, C. R. Lucas, A. Avendano, J. W. Song, C.
E. Castro, ACS Appl. Mater. Interfaces 2022, 14, 55307-55319.
[40] L. Westergard, H. M. Christensen, D. A. Harris, Biochimica et Biophysica Acta (BBA) -
Molecular Basis of Disease 2007, 1772, 629-644.
[41] E. Sinkala, E. Sollier-Christen, C. Renier, E. Rosàs-Canyelles, J. Che, K. Heirich, T. A.
Duncombe, J. Vlassakis, K. A. Yamauchi, H. Huang, S. S. Jeffrey, A. E. Herr, Nat Commun 2017,
8, 14622.
[42] S. B. Breitkopf, M. Yuan, G. A. Pihan, J. M. Asara, Proc. Natl. Acad. Sci. U.S.A. 2012, 109,
16190-16195.
[43] Z. Chen, J. Huang, L. Li, TrAC Tren ds in Analy tical Ch emist ry 2019, 11 8, 880-892.
[44] G.-D. Syu, J. Dunn, H. Zhu, Molecular & Cellular Proteomics 2020, 19, 916-927.
[45] A. Rajendran, M. Endo, H. Sugiyama, Angew Chem Int Ed 2012, 51, 874-890.
[46] D. Daems, I. Rutten, J. Bath, D. Decrop, H. Van Gorp, E. P. Ruiz, S. De Feyter, A. J. Turberfield,
J. Lammertyn, ACS Sens. 2019, 4, 2327-2335.
[47] Y. Z h o u , J . D o n g , Q . W a n g , NPG Asia Mater 2023, 15, 25.
[48] M. Petracca, R. O. Vancea, L. Fleysher, L. E. Jonkman, N. Oesingmann, M. Inglese, Brain
2016, 139, 795-806.
[49] Y. H u a n g , D . H u , C . H u a n g , C . G . N i c h o l s , Circ: Arrhythmia and Electrophysiology 2019, 12,
e007322.
[50] A. Hrvat, M. Schmidt, B. Wagner, D. Zwanziger, R. Kimmig, L. Volbracht, S. Brandau, N.
Mallmann-Gottschalk, J Exp Clin Cancer Res 2023, 42, 235.
[51] Z. Sun, M. Jia, Z. Fu, M. Zhang, H. Wang, Y. Xu, Chemical Engineering Journal 2021, 406,
126755.
10.1002/cbic.202400227
Accepted Manuscript
ChemBioChem
This article is protected by copyright. All rights reserved.
14397633, ja, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cbic.202400227 by Shanghai Jiaotong University, Wiley Online Library on [09/06/2024]. 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
20
[52] Y. T a n g , S . L i , L . H u , X. S u n , B . Z h a n g , W . J i , L . M a , W. Q i a n , A . K a n g , D . Z h u , Advanced
Healthcare Materials 2021, 10, 2100764.
[53] M. Xiao, W. Lai, T. Man, B. Chang, L. Li, A. R. Chandrasekaran, H. Pei, Chem. Rev. 2019,
119, 11631-11717.
[54] A. R. Chandrasekaran, H. Wady, H. K. K. Subramanian, Small 2016, 12, 2689-2700.
[55] S. Ranallo, A. Porchetta, F. Ricci, Anal. Chem. 2019, 91, 44-59.
[56] W. Wang, S. Yu, S. Huang, S. Bi, H. Han, J.-R. Zhang, Y. Lu, J.-J. Zhu, Chem. Soc. Rev. 2019,
48, 4892-4920.
[57] P. Zhan, A. Peil, Q. Jiang, D. Wang, S. Mousavi, Q. Xiong, Q. Shen, Y. Shang, B. Ding, C.
Lin, Y. Ke, N. Liu, Chem. Rev. 2023, 123, 3976-4050.
[58] T. Pa tino, Nature Nanotechnology 2024, 19, 1-2.
[59] P. I v a s ko v i c , A. Ya ma d a , J . E l e z g ar a y, D . Ta l a ga , S. B on h o m me a u , M . Bl a n c ha r d -Desce, R. A.
L. Vallée, S. Ravaine, Nanoscale 2018, 10, 16568-16573.
[60] J. D. WATSON, F. H. C. CRICK, Nature 1953, 171, 964-967.
[61] R. Stoltenburg, C. Reinemann, B. Strehlitz, Biomolecular Engineering 2007, 24, 381-403.
[62] F. Ma, S. Wei, C. Zhang, Anal. Chem. 2019, 91, 7505-7509.
[63] Q. Zhang, X. Gao, Y.-P. Ho, M. Liu, Y. Han, D. Li, H. Yuan, C. Zhang, Nano Lett. 2024, 24,
2360-2368.
[64] Y. K e , S . L i n d s a y , Y. C h a n g , Y. L i u , H . Ya n , Science 2008, 319, 180-183.
[65] H. K. K. Subramanian, B. Chakraborty, R. Sha, N. C. Seeman, Nano Lett. 2011, 11, 910-913.
[66] M. Scherf, F. Scheffler, C. Maffeo, U. Kemper, J. Ye, A. Aksimentiev, R. Seidel, U. Reibetanz,
Nanoscale 2022, 14, 1804118050.
[67] E. Weinhold, B. Chakraborty, Nanoscale 2021, 13, 2465-2471.
[68] A. Aghebat Rafat, S. Sagredo, M. Thalhammer, F. C. Simmel, Nat. Chem. 2020, 12, 852-859.
[69] J. Dai, C. Xing, Y. Lin, Y. Huang, Y. Yang, Z. Chen, C. Lu, H. Yang, Sensors and Actuators B:
Chemical 2021, 344, 130195.
[70] R. Schreiber, J. Do, E.-M. Roller, T. Zhang, V. J. Schüller, P. C. Nickels, J. Feldmann, T. Liedl,
Nature Nanotechnology 2014, 9, 74-78.
[71] M. J. Urban, P. K. Dutta, P. Wang, X. Duan, X. Shen, B. Ding, Y. Ke, N. Liu, J. Am. Chem. Soc.
2016, 138, 54955498.
[72] S. Han, W. Liu, S. Yang, R. Wang, ACS Omega 2019, 4, 11025-11031.
[73] C. Wen, E. Bertosin, X. Shi, C. Dekker, S. Schmid, Nano Lett. 2023, 23, 788-794.
[74] M. Kim, C. Lee, K. Jeon, J. Y. Lee, Y.-J. Kim, J. G. Lee, H. Kim, M. Cho, D.-N. Kim, Nature
2023, 619, 78-86.
[75] D. Selnihhin, S. M. Sparvath, S. Preus, V. Birkedal, E. S. Andersen, ACS Nano 2018, 12, 5699-
5708.
[76] C. Xing, S. Chen, Q. Lin, Y. Lin, M. Wang, J. Wang, C. Lu, Nanoscale 2022, 14, 1327-1332.
[77] I. Domljanovic, M. Loretan, S. Kempter, G. P. Acuna, S. Kocabey, C. Ruegg, Nanoscale 2022,
14, 15432-15441.
[78] S. Kocabey, G. Chiarelli, G. P. Acuna, C. Ruegg, Biosensors and Bioelectronics 2023, 224,
115053.
[79] A. Kuzyk, R. Schreiber, H. Zhang, A. O. Govorov, T. Liedl, N. Liu, Nature Mater 2014, 13,
862-866.
[80] Y. X u , Q . Z h a n g , R . C h e n , H . C a o , J . T a n g , Y. Wu , X . L u , B . C h u , B . S o n g , H . W a n g , Y. H e , J.
10.1002/cbic.202400227
Accepted Manuscript
ChemBioChem
This article is protected by copyright. All rights reserved.
14397633, ja, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cbic.202400227 by Shanghai Jiaotong University, Wiley Online Library on [09/06/2024]. 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
21
Am. Chem. Soc. 2022, 144, 23522-23533.
[81] Z. Yu, L. Pan, X. Ma, T. Li, F. Wang, D. Yang, M. Li, P. Wang, Biosensors and Bioelectronics
2023, 237, 115526.
[82] B. Jeon, M. M. Guareschi, J. M. Stewart, E. Wu, A. Gopinath, N. Arroyo-Currás, P. Dauphin-
Ducharme, P. S. Lukeman, K. W. Plaxco, P. W. K. Rothemund, 2023.
[83] E. Sameiyan, E. Bagheri, M. Ramezani, M. Alibolandi, K. Abnous, S. M. Taghdisi, Biosensors
and Bioelectronics 2019, 143, 111662.
[84] E. Pibiri, P. Holzmeister, B. Lalkens, G. P. Acuna, P. Tinnefeld, Nano Lett. 2014, 14, 3499-
3503.
[85] E. Beamish, V. Tabard-Cossa, M. Godin, ACS Sens. 2017, 2, 1814-1820.
[86] L. Liang, F. Qin, S. Wang, J. Wu, R. Li, Z. Wang, M. Ren, D. Liu, D. Wang, D. Astruc,
Coordination Chemistry Reviews 2023, 478, 214998.
[87] U. F. Keyser, Nature Nanotech 2016, 11, 106-108.
[88] R. Wei, T. G. Martin, U. Rant, H. Dietz, Angew Chem Int Ed 2012, 51, 4864-4867.
[89] A. Barati Farimani, P. Dibaeinia, N. R. Aluru, ACS Appl. Mater. Interfaces 2017, 9, 92-100.
[90] P. W i ll i a m so n , P. P i sk u n e n, H . I j ä s, A. B ut t e r wo r t h , V. L i n k o, D . K . C o r r ig a n , ACS Sens. 2023,
8, 1471-1480.
[91] A. Kuzuya, Y. Sakai, T. Yamazaki, Y. Xu, M. Komiyama, Nat Commun 2011, 2, 449.
[92] M. Godonoga, T.-Y. L i n , A . O s h i m a , K . S u m i t o m o , M . S . L . Ta n g , Y. -W. C h eu ng , A . B .
Kinghorn, R. M. Dirkzwager, C. Zhou, A. Kuzuya, J. A. Tanner, J. G. Heddle, Sci Rep 2016, 6,
21266.
[93] K. Fujita, M. Ohmachi, K. Ikezaki, T. Yanagida, M. Iwaki, Commun Biol 2019, 2, 437.
[94] P. Zh a n g , X . L iu , P. L iu , F. Wa n g, H . A r i ya m a , T . A n do , J . L i n , L. Wa n g , J . H u, B . Li , C . F a n ,
Nat Commun 2020, 11, 3114.
[95] M. Raveendran, A. J. Lee, R. Sharma, C. Wälti, P. Actis, Nat Commun 2020, 11, 4384.
[96] R. Veneziano, T. J. Moyer, M. B. Stone, E.-C. Wamhoff, B. J. Read, S. Mukherjee, T. R.
Shepherd, J. Das, W. R. Schief, D. J. Irvine, M. Bathe, Nat. Nanotechnol. 2020, 15, 716-723.
[97] J. Hellmeier, R. Platzer, J. B. Huppa, E. Sevcsik, in The Immune Synapse (Eds.: C.T. Baldari,
M.L. Dustin), Springer US, New York, NY, 2023, pp. 277-302.
[98] E. Oktay, J. Bush, M. Vargas, D. V. Scarton, B. O’Shea, A. Hartman, C. M. Green, K. Neyra,
C. M. Gomes, I. L. Medintz, D. Mathur, R. Veneziano, ACS Appl. Mater. Interfaces 2023, 15, 27759-
27773.
[99] A. K. Natarajan, J. Ryssy, A. Kuzyk, Nanoscale 2023, 15, 3212-3218.
[100] M. Iwaki, S. F. Wickham, K. Ikezaki, T. Yanagida, W. M. Shih, Nat Commun 2016, 7,
13715.
[101] Y. K e , T. M e y e r , W. M . S h i h , G . B e l l o t , Nat Commun 2016, 7, 10935.
[102] Z. Zhao, J. Fu, S. Dhakal, A. Johnson-Buck, M. Liu, T. Zhang, N. W. Woodbury, Y. Liu, N. G.
Walter, H . Yan , Nat Commun 2016, 7, 10619.
[103] M. Pfeiffer, K. Trofymchuk, S. Ranallo, F. Ricci, F. Steiner, F. Cole, V. Glembockyte, P.
Tinnefeld, iScience 2021, 24, 103072.
[104] Y. S u n , L . Y a n , J . S u n , M . X i a o , W . L a i , G . S o n g , L . L i , C . F a n , H . P e i , Nat Commun 2022,
13, 3916.
[105] D. Selnihhin, K. I. Mortensen, J. B. Larsen, J. B. Simonsen, F. S. Pedersen, Small Methods
2022, 6, 2101364.
10.1002/cbic.202400227
Accepted Manuscript
ChemBioChem
This article is protected by copyright. All rights reserved.
14397633, ja, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cbic.202400227 by Shanghai Jiaotong University, Wiley Online Library on [09/06/2024]. 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
22
[106] I. Seitz, H. Ijäs, V. Linko, M. A. Kostiainen, ACS Appl. Mater. Interfaces 2022, 14, 38515-
38524.
[107] Q. Chen, X. Wang, J. Chen, Y. Xiang, M. Xiao, H. Pei, L. Li, Anal. Chem. 2022, 94, 10192-
10197.
[108] I. Kaminska, J. Bohlen, S. Mackowski, P. Tinnefeld, G. P. Acuna, ACS Nano 2018, 12, 1650-
1655.
[109] F. Schuknech t, K. Kołą ta j, M. St ei nb er ge r, T. Li ed l, T. Lo hm ue ll er, Nat Commun 2023, 14,
7192.
[110] N. A. W. Bell, U. F. Keyser, Nature Nanotech 2016, 11, 645651.
[111] N. V. Voigt, T. Tørring, A. Rotaru, M. F. Jacobsen, J. B. Ravnsbæk, R. Subramani, W.
Mamdouh, J. Kjems, A. Mokhir, F. Besenbacher, K. V. Gothelf, Nature Nanotech 2010, 5, 200-203.
[112] A. Kuzuya, R. Watanabe, Y. Yamanaka, T. Tamaki, M. Kaino, Y. Ohya, Sensors 2014, 14,
1932919335.
[113] J. Yang, S. Jiang, X. Liu, L. Pan, C. Zhang, ACS Appl. Mater. Interfaces 2016, 8, 34054-34060.
[114] Z. Lu, Y. Wang, D. Xu, L. Pang, Chem. Commun. 2017, 53, 941-944.
[115] X. Zhang, L. Pan, R. Guo, Y. Zhang, F. Li, M. Li, J. Li, J. Shi, F. Qu, X. Zuo, X. Mao, Chem.
Commun. 2022, 58, 3673-3676.
[116] H.-K. Walter, J. Bauer, J. Steinmeyer, A. Kuzuya, C. M. Niemeyer, H.-A. Wagenknecht, Nano
Lett. 2017, 17, 2467-2472.
[117] Y. H u a n g , M . -K. Nguyen, A. K. Natarajan, V. H. Nguyen, A. Kuzyk, ACS Appl. Mater.
Interfaces 2018, 10, 44221-44225.
[118] K. Cervantes-Salguero, M. Freeley, J. L. Chávez, M. Palma, J. Mater. Chem. B 2020, 8, 6352-
6356.
[119] X. Wang, Z. Mao, R. Chen, S. Li, S. Ren, J. Liang, Z. Gao, Biosensors and Bioelectronics
2022, 211, 114336.
[120] V. V. T h a c k e r , L . O . H e r r m a n n , D . O . S i g l e , T . Z h a n g , T . L i e d l , J . J . B a u m b e r g , U . F . K e y s e r ,
Nat Commun 2014, 5, 3448.
[121] A. Puchkova, C. Vietz, E. Pibiri, B. Wünsch, M. Sanz Paz, G. P. Acuna, P. Tinnefeld, Nano
Lett. 2015, 15, 8354-8359.
[122] P. Z h a n, T. W en , Z . Wa n g , Y. H e , J . S h i , T . Wa n g , X . L i u , G . L u , B. D in g , Angew Chem Int Ed
2018, 57, 28462850.
[123] C. Zhou, L. Xin, X. Duan, M. J. Urban, N. Liu, Nano Lett. 2018, 18, 7395-7399.
[124] W. Fan g, S . J ia , J . Ch a o, L . Wa ng , X . Du a n, H . L iu , Q . Li , X. Z u o, L . Wa ng , L . Wan g , N. Li u,
C. Fan, Sci. Adv. 2019, 5, eaau4506.
[125] V. K a u r , S . T a n w a r , G . K a u r , T . S e n , ChemPhysChem 2021, 22, 160-167.
[126] V. K a u r , M . S h a r m a , T . S e n , Front. Chem. 2021, 9, 772267.
[127] V. K a u r , C . K a u r , T . S e n , J. Phys. Chem. C 2023, 127, 7308-7318.
[128] B. Huo, L. Xia, Z. Gao, G. Li, Y. Hu, Anal. Chem. 2022, 94, 11889-11897.
[129] C. Kaur, V. Kaur, S. Rai, M. Sharma, T. Sen, Nanoscale 2023, 15, 6170-6178.
[130] S. Li, B. Shi, D. He, H. Zhou, Z. Gao, Journal of Hazardous Materials 2023, 458, 131874.
[131] T. Di ng, J. Yan g, J. Wang, V. P an , Z . L u, Y. K e, C. Zh an g, Biosensors and Bioelectronics 2022,
195, 113658.
10.1002/cbic.202400227
Accepted Manuscript
ChemBioChem
This article is protected by copyright. All rights reserved.
14397633, ja, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cbic.202400227 by Shanghai Jiaotong University, Wiley Online Library on [09/06/2024]. 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
23
TOC:
An emerging field of study involves integrating sensing modules onto DNA origami structures,
which allows for the robust detection of a wide range of biomarkers including nucleic acids, proteins,
small molecules, and ions.
10.1002/cbic.202400227
Accepted Manuscript
ChemBioChem
This article is protected by copyright. All rights reserved.
14397633, ja, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cbic.202400227 by Shanghai Jiaotong University, Wiley Online Library on [09/06/2024]. 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
24
Author Biographies
Pengfei Wang is a professor at the Institute of Molecular Medicine,
School of Medicine, Shanghai Jiao Tong University, China. He
earned his Ph.D. degree in biomedical engineering from Purdue
University, USA, in 2014. His current research interests include
molecular self-assembly, soft materials, and disease diagnosis and
therapy.
Li Pan received her PhD degree in 2019 from Jackson State
University, USA. She currently works at Shanghai Jiao Tong
University as an assistant professor. She currently focuses on the
development of biosensing methods.
Caiqing Yuan is currently a graduate student jointly trained by
Shanghai Normal University and Shanghai Jiao Tong University
School of Medicine, under the guidance of Prof. Pengfei Wang,
with a research focus on the detection of tumor and cancer-related
biomarkers.
10.1002/cbic.202400227
Accepted Manuscript
ChemBioChem
This article is protected by copyright. All rights reserved.
14397633, ja, Downloaded from https://chemistry-europe.onlinelibrary.wiley.com/doi/10.1002/cbic.202400227 by Shanghai Jiaotong University, Wiley Online Library on [09/06/2024]. 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 combination of DNA nanotechnology and Nano Gold (NG) plasmon has opened exciting possibilities for a new generation of functional plasmonic systems that exhibit tailored optical properties and find utility in various applications. In this review, the booming development of dynamic gold nanostructures are summarized, which are formed by DNA self‐assembly using DNA‐modified NG, DNA frameworks, and various driving forces. The utilization of bottom‐up strategies enables precise control over the assembly of reversible and dynamic aggregations, nano‐switcher structures, and robotic nanomachines capable of undergoing on‐demand, reversible structural changes that profoundly impact their properties. Benefiting from the vast design possibilities, complete addressability, and sub‐10 nm resolution, DNA duplexes, tiles, single‐stranded tiles and origami structures serve as excellent platforms for constructing diverse 3D reconfigurable plasmonic nanostructures with tailored optical properties. Leveraging the responsive nature of DNA interactions, the fabrication of dynamic assemblies of NG becomes readily achievable, and environmental stimulation can be harnessed as a driving force for the nanomotors. It is envisioned that intelligent DNA‐assembled NG nanodevices will assume increasingly important roles in the realms of biological, biomedical, and nanomechanical studies, opening a new avenue toward exploration and innovation.
Article
Full-text available
The label-free identification of individual proteins from liquid samples by surface-enhanced Raman scattering (SERS) spectroscopy is a highly desirable goal in biomedical diagnostics. However, the small Raman scattering cross-section of most (bio-)molecules requires a means to strongly amplify their Raman signal for successful measurement, especially for single molecules. This amplification can be achieved in a plasmonic hotspot that forms between two adjacent gold nanospheres. However, the small (≈1−2 nm) gaps typically required for single-molecule measurements are not accessible for most proteins. A useful strategy would thus involve dimer structures with gaps large enough to accommodate single proteins, whilst providing sufficient field enhancement for single-molecule SERS. Here, we report on using a DNA origami scaffold for tip-to-tip alignment of gold nanorods with an average gap size of 8 nm. The gaps are accessible to streptavidin and thrombin, which are captured at the plasmonic hotspot by specific anchoring sites on the origami template. The field enhancement achieved for the nanorod dimers is sufficient for single-protein SERS spectroscopy with sub-second integration times. This design for SERS probes composed of DNA origami with accessible hotspots promotes future use for single-molecule biodiagnostics in the near-infrared range.
Article
Full-text available
Superhydrophilic/superhydrophobic patterned surfaces can be used to create droplet microarrays. A specific challenge with the liquids needed for various biomedical applications, as compared to pure water, is their lower surface tension and potential for contaminating the surfaces through adsorption. Here, a method is shown to create biofluid droplet microarrays using discontinuous dewetting of pure water, an oil protective layer, and finally biofluid exchange with the water droplet array. With this method, a droplet array of a viscous nucleic acid amplification solution can be formed with a low surface tension of 34 mN m ⁻¹ and a contact angle of only 76° with the used hydrophobic coating. This droplet array is applied for nucleic acid detection of SARS‐CoV‐2 virus using strand invasion‐based amplification (SIBA) technology. It is shown that by using an array of 10 000 droplets of 50 µm diameter the limit of detection is 1 RNA copy µL ⁻¹ . The results demonstrate that SIBA on droplet microarrays may be a quantitative technology.
Article
Full-text available
Background Malignant ascites commonly occurs in advanced or recurrent stages of epithelial ovarian cancer during peritoneal carcinomatosis and is correlated with poor prognosis. Due to its complex composition of cellular and acellular components malignant ascites creates a unique tumor microenvironment, which mediates immunosuppression and promotes progression of disease. However, the immunosuppressive mechanisms remain poorly understood. Methods In the present study, we explored the antitumor activity of healthy donor NK and T cells directed against ovarian cancer cells in presence of malignant ascites derived from patients with advanced or recurrent peritoneal carcinomatosis. A wide range of methods was used to study the effect of ascites on NK and T cells (FACS, ELISA, EliSpot, qPCR, Live-cell and confocal microscopy, Western blot and electrolyte flux assays). The ascites components were assessed using quantitative analysis (nephelometry, potentiometry and clinical chemistry) and separation methods (dialysis, ultracentrifugal filtration and lipid depletion). Results Ascites rapidly inhibited NK cell degranulation, tumor lysis, cytokine secretion and calcium signaling. Similarly, target independent NK and T cell activation was impaired in ascites environment. We identified imbalanced electrolytes in ascites as crucial factors causing extensive immunosuppression of NK and T cells. Specifically, high sodium, low chloride and low potassium content significantly suppressed NK-mediated cytotoxicity. Electrolyte imbalance led to changes in transcription and protein expression of electrolyte channels and impaired NK and T cell activation. Selected inhibitors of sodium electrolyte channels restored intracellular calcium flux, conjugation, degranulation and transcript expression of signaling molecules. The levels of ascites-mediated immunosuppression and sodium/chloride/potassium imbalance correlated with poor patient outcome and selected molecular alterations were confirmed in immune cells from ovarian cancer patients. Conclusion Our data suggest a novel electrolyte-based mechanism of immunosuppression in malignant ascites of patients with peritoneal carcinomatosis. We show for the first time that the immunosuppression of NK cytotoxicity in coculture assays is correlated to patient poor survival. Therapeutic application of sodium channel inhibitors may provide new means for restoring immune cell activity in ascites or similar electrolyte imbalanced environments.
Article
Full-text available
Ultrasensitive identification of biomarkers in biofluids is essential for precise diagnosis of diseases. For the gold standard approaches, polymerase chain reaction and enzyme‐linked immunosorbent assay, cumbersome operational steps hinder their point‐of‐care applications. Here, w e implement a bionic biomarker entrapment system (BioES), which employs a multi‐body Y‐shaped tetrahedral DNA probe immobilized on carbon nanotube transistors. W e successfully realize clinical identification of endometriosis by detecting an estrogen receptor, ERβ, from the lesion tissue of endometriosis patients and establish a standard diagnosis procedure. The multi‐body Y‐shaped BioES achieves a theoretical limit of detection (LoD) of 6.74 aM and a LoQ of 141 aM in a complex protein milieu. Furthermore, the BioES is optimized into multi‐site recognition module for enhanced binding efficiency, realizing the first identification of monkeypox virus antigen A35R and unamplified detection of circulating tumor DNA of breast cancer in serum. The rigid and compact probe framework with synergy effect enable the BioES to target A35R and DNA with a LoD down to 991 aM and 0.21 aM, respectively. Owing to its versatility for proteins and nucleic acids as well as ease‐of‐manipulation and ultra‐sensitivity, the BioES can be leveraged as an all‐encompassing tool for population‐wide screening of epidemics and clinical disease diagnosis. This article is protected by copyright. All rights reserved
Article
Full-text available
The paper-folding mechanism has been widely adopted in building of reconfigurable macroscale systems because of its unique capabilities and advantages in programming variable shapes and stiffness into a structure1–5. However, it has barely been exploited in the construction of molecular-level systems owing to the lack of a suitable design principle, even though various dynamic structures based on DNA self-assembly6–9 have been developed10–23. Here we propose a method to harness the paper-folding mechanism to create reconfigurable DNA origami structures. The main idea is to build a reference, planar wireframe structure²⁴ whose edges follow a crease pattern in paper folding so that it can be folded into various target shapes. We realized several paper-like folding and unfolding patterns using DNA strand displacement²⁵ with high yield. Orthogonal folding, repeatable folding and unfolding, folding-based microRNA detection and fluorescence signal control were demonstrated. Stimuli-responsive folding and unfolding triggered by pH or light-source change were also possible. Moreover, by employing hierarchical assembly²⁶ we could expand the design space and complexity of the paper-folding mechanism in a highly programmable manner. Because of its high programmability and scalability, we expect that the proposed paper-folding-based reconfiguration method will advance the development of complex molecular systems.
Article
Nanopore platform holds a great fascination that attracts increasing interests in research and leads the development of various domains. Herein we first briefly introduce the development, generation and challenges in nanopore sensing. Afterward, we overview the nanopore substrates from the material perspective that is different from the general classification by categorizing the pore carrier materials into five types. They are biological pores (including protein, peptide, DNA origami and G-quadruplexes), organic nanopores (including polymer, chemically synthesized pores, COFs, and helical self-assembled pores), inorganic pores (involving silicon-based, carbon-based, and 2D pores), metallic nanopores (metal, metallic oxide, MOFs and MXenes) and hybrid nanopores. We then review the common sensing principle in nanopore devices (resistive sensing, current rectification, nano-gap tunneling, FET, electroosmosis, opto-electronic readout and multiple nanoarrays), followed by the implementations of nanopore techniques in the matter and molecular identification and interplay, catalysis, single-molecule reaction mechanism, molecular conformation dynamics, data storage, and theoretical simulation among others. We finally conclude and propose potential substrate materials, design strategies and new technique combinations in nanopore devices and sensing areas.
Article
The detection of SARS-CoV-2 infection is crucial for effective prevention and surveillance of COVID-19. In this study, we report the development of a novel detection assay named CENSOR that enables sensitive and specific detection of SARS-CoV-2 RNA using a plasmonic chiral biosensor in combination with CRISPR-Cas13a. The chiral biosensor was designed by assembling gold nanorods (AuNR) into three-dimensional plasmonic architectures of controllable chirality on a DNA origami template. This modular assembly mode enhances the flexibility and adaptability of the sensor, thereby improving its universality as a sensing platform. In the presence of SARS-CoV-2 RNA, the CRISPR-Cas13a enzyme triggers collateral cleavage of RNA molecules, resulting in a differential chiral signal readout by the biosensor compared to when there are no RNA targets present. Notably, even subtle variations in the concentration of SARS-CoV-2 RNA can provoke significant changes in chiral signals after preamplification of RNA targets (calculated LOD: 0.133 aM), which establishes the foundation for quantitative detection. Furthermore, CENSOR demonstrated high sensitivity and accuracy in detecting SARS-CoV-2 RNA from clinical samples, suggesting its potential application in clinical settings for viral detection beyond SARS-CoV-2.