Next Article in Journal
Can Pyrolysis Oil Be Used as a Feedstock to Close the Gap in the Circular Economy of Polyolefins?
Next Article in Special Issue
Studies of the Mechanism of Nucleosome Dynamics: A Review on Multifactorial Regulation from Computational and Experimental Cases
Previous Article in Journal
Curing, Properties and EMI Absorption Shielding of Rubber Composites Based on Ferrites and Carbon Fibres
Previous Article in Special Issue
Comprehensive Analysis of Chromatin Accessibility and Transcriptional Landscape Identified BRCA1 Repression as a Potential Pathological Factor for Keloid
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Development in Biomedical Applications of Oligonucleotides with Triplex-Forming Ability

by
Incherah Bekkouche
1,
Alexander Y. Shishonin
2 and
Alexandre A. Vetcher
1,2,*
1
Nanotechnology Scientific and Educational Center, Institute of Biochemical Technology and Nanotechnology, Peoples’ Friendship University of Russia (RUDN), Miklukho-Maklaya Str. 6, Moscow 117198, Russia
2
Complementary and Integrative Health Clinic of Dr. Shishonin, 5, Yasnogorskaya Str., Moscow 117588, Russia
*
Author to whom correspondence should be addressed.
Polymers 2023, 15(4), 858; https://doi.org/10.3390/polym15040858
Submission received: 25 November 2022 / Revised: 31 January 2023 / Accepted: 2 February 2023 / Published: 9 February 2023
(This article belongs to the Special Issue Nucleic Acids as Polymers)

Abstract

:
A DNA structure, known as triple-stranded DNA, is made up of three oligonucleotide chains that wind around one another to form a triple helix (TFO). Hoogsteen base pairing describes how triple-stranded DNA may be built at certain conditions by the attachment of the third strand to an RNA, PNA, or DNA, which might all be employed as oligonucleotide chains. In each of these situations, the oligonucleotides can be employed as an anchor, in conjunction with a specific bioactive chemical, or as a messenger that enables switching between transcription and replication through the triplex-forming zone. These data are also considered since various illnesses have been linked to the expansion of triplex-prone sequences. In light of metabolic acidosis and associated symptoms, some consideration is given to the impact of several low-molecular-weight compounds, including pH on triplex production in vivo. The review is focused on the development of biomedical oligonucleotides with triplexes.

1. Introduction

Triplexes are non-canonical DNA structures that contain an extra single-stranded RNA or DNA binding region that is unique to the primary groove of double-stranded DNA. Triplex-forming oligonucleotides (TFOs) bind in the duplex major groove and produce a triple-helical structure. By creating hydrogen bonds with exposed groups on Watson-Crick base pairs, substances that directly interact with duplex DNA are of interest as gene targeting techniques. Research on these substances has centered on the use of triplex-forming oligonucleotides (TFOs), which interact with one another from the major groove side of duplex DNA to generate triplex DNA in a sequence-specific way [1,2]. The ability to inactivate and activate gene expression, recombination, and repair through the production of triplex DNA make this technique potentially useful for analyzing activities involving the genome [3,4,5]. With every sequence of duplex DNA, the ability to generate stable triplex DNA is, however, inherently constrained. With guanine at the GC base pair and adenine at the AT base pair from the main groove side of the target duplex DNA, respectively, guanine and adenine nucleobases in TFOs rich in purines create two reverse Hoogsteen-type hydrogen bonds (Figure 1) (e.g., [6]). The formation of stable triplex DNA using TFOs made of natural nucleosides is not conceivable, though, since the CG and TA base pairs only have one hydrogen bonding site at the main groove side of the cytosine and thymine nucleobases.
Triplex technology still has some flaws [7], most of which are related to (i) their low stability, (ii) sequence restrictions due to the requirement of polypurine tracks in the triplex target sequence, (iii) susceptibility to nucleases, and (iv) difficulties in delivering TFOs in the cellular nucleus. A substantial amount of chemical, biochemical, and biotechnological research is presently being directed toward resolving the practical issues associated with triplex technology [8,9,10,11]. In this study, we conduct a thorough examination of the most recent advances in biological applications of oligonucleotides with triplex-forming capacity.

2. Results and Discussion

2.1. Oligonucleotides with Triplex-Forming Ability in Targeted Delivery

Genomic DNA is frequently under torsional stress, which can cause the DNA double helix to be both over- and underwound. A decrease in the number of links (Lk) connecting the two strands of a closed-circular DNA (a negative Lk) causes negative superhelical stress. The untwisting of the helix (change in twist; Tw) and a coiling distortion of the DNA backbone (writhe; Wr) are the two components of the conformational response to this stress, known as negative supercoiling [12,13,14,15]. Genomic DNA in prokaryotes supercoils with an average density of (Lk/original Lk) of 0.065. Nuclear-associated proteins and supercoiling collaborate to control bacterial gene expression [16]. Supercoiling produced by transcription in eukaryotes is thought to be involved in the control of oncogenes such as c-Myc. It affects the positioning of RNA guide sequences by the CRISPR-Cas9 gene editing toolkit [17] and plays a crucial role in the development and stability of looped DNA structures [18] and DNA R-loops [19].

2.1.1. Substantial RNA–DNA Interactions

RNAs have been assigned many regulatory roles, some of which involve interactions with both DNA and proteins [20,21]. The addressing and recruitment of transcription factors [22,23], the structuring of the transcription factor machinery [22,24], and the mediating of histone modifications are all tasks performed at the RNA–protein interface. The functional role of the XIST transcript in the silencing of the X chromosome during dosage compensation is just one example of the epigenomic effects associated with RNA–DNA interactions [25]. G-quadruplex [26] creation is another further instance of RNA–DNA interactions. R-loop [27]. R-loops, which are associated with chromatin condensation and cancer, are interactions between single-stranded DNA and RNA that are made possible by Watson-Crick base pairing [28,29]. In addition to these, there is another type of RNA–DNA interaction in which the double helix’s primary groove is occupied by single-stranded RNA, maintaining the double helix’s structural integrity, which results in the development of a triple helix (triplex) comprising RNA and DNA. triplex production, is an aspect of epigenetics that, despite long-standing knowledge in the biophysical sense [30,31], is still not fully understood this is due to a few factors, primary among which is the difficulty of conducting experiments to investigate triplex formation on a genome-wide level in living cells. In the past, approaches for capturing the genomic interaction locations of individual transcripts have been used to experimentally investigate triplex formation in the setting of cells [32,33]. However, across a wide range of species, tissues, and cell types, regulatory transcripts have been connected to epigenetic processes [34,35,36,37]. Herein lies the significance of creating instruments that can precisely forecast RNA–DNA interaction points as possible sites of triplex formation. Hoogsteen base pairing rules [38], which are canonically accountable for triplex formation, have been used as the basis for previously reported computational tools (Triplexator/Triplex Domain Finder [39,40], as well as LongTarget [41]). These are tools for applying Hoogsteen rules. A comparison of Triplexator and LongTarget utilizing MEG3 ChOP-seq data [32] shows significant space for improvement in this area, even though the use of canonical rules to predict triplex formation offers insight into potential RNA–DNA interactions [42]. These benchmarking data imply that base pairing rules that go beyond the Hoogsteen base pairing rules now in use may be necessary to implement reliable prediction of triplex formation in a cellular context. In relation to this, research has been conducted on the prediction of triplex-forming RNA and DNA sequences using previously documented triplex interactions [43], albeit this method cannot predict entire triplex interactions. Experiments with comparable goals have been used in conjunction with computational methods to anticipate RNA–DNA interactions across the entire genome. In terms of triplex creation, the genome-wide isolation of triplexes, followed by sequencing (triplex seq), are the most important of these [44]. This technique enables the identification of triplex-forming sequences throughout the genome (triplexDNA-seq) and transcriptome (triplexRNA-seq), but it does not provide information on how the sequences couple up with one another. Numerous approaches have been described to find all to all contacts between transcripts and chromatin aside from specifically triplex-mediated RNA–DNA interactions [45,46,47,48]. However, RNA and DNA interacting complexes ligated and sequenced (RAD-ICL-seq) [49] and RedC [50] have nucleotide processing processes that are the most similar to those of triplex-seq. Through the proximity-based ligation of RNA and DNA via a linker sequence, these approaches together discover relationships between transcripts and areas of the genome. Although they are still relatively new and experimentally challenging, RADICL-seq and RedC are valuable sources of information on RNA–DNA interaction. Therefore, it is not now viable to conduct such studies under a variety of steady-state and differential situations. As a result, the most broadly applicable use for this data may be as inputs for machine learning algorithms, which could enable the prediction of RNA–DNA interactions under a condition of interest.
A type of direct RNA–DNA interaction mechanism known as a triplex is created by the binding of RNA sites to a purine-rich strand of duplex DNA in accordance with either the forward or reverse Hoogsteen base-pairing rule. It has been established that some lncRNAs create DNA to carry out certain functions. For instance, promoter-associated lncRNA interacts with TTF-I to inhibit the transcription of rRNA [51], and FENDRR increases PRC2 occupancy at the triplex formation sites [52]. MEG3 forms a DNA-lncRNA triplex with the TGF gene to regulate the activity of the gene [32], and the particle binds to the MAT2A promoter CpG island as a triplex to support gene-silencing mechanisms [53]. KHPS1 and SPHK1 interact to bind the lncRNA and associated effector proteins to the gene promoter [54]. HOTAIR and PCDH7 and HOXB2 form a triplex to control adipogenic differentiation [55]. MIR100HG controls p27 via the formation of a triplex [56], and the promoter and pre-rRNA antisense direct associated CHD4/NuRD to the rDNA promoter [57] chromatin isolation by RNA purification (ChIRP-seq) [33]. Capture hybridization analysis of RNA targets (CHART-seq) [58], RNA antisense purification (RAP-seq) [59], and chromatin oligo affinity precipitation (ChOP-seq) [52] are examples of recently developed high-throughput techniques that have assisted in creating the genome-wide map of lncRNA chromatin interactions for specific ones. As a result, they are unable to offer trustworthy sources for research into DNA: RNA triplex production. Cetin et al. [44] devised a method to map the genome-wide DNA: RNA triplexes by eliminating the chromatin crosslinking to demonstrate the existence of the DNA: RNA triplex interactions in vivo. The physiological significance of DNA: RNA triplex complexes was demonstrated using this method. Currently, base paring rules related to mathematical statistics are primarily used to predict DNA: RNA triplex. Hoogsteen and reverse-Hoogsteen base-pairing are taken into consideration in the proposed triplexator to systematically discover the probable triplex formation sites of RNA and the targeting sites on DNA [39]. Triplex-Inspector was created to choose sequence-specific ligands and targets while taking genomic architecture and gene location into account [60]. A long target was created to find motifs and binding sites in triplex formation while taking non-canonical rules into account [41], and triplex Domain Finder (TDF) was created to predict triplexes and to characterize lncRNA and the corresponding DNA targets [40].
In several biologically significant processes, including DNA replication, transcription, telomere replication, and HIV replication by reverse transcription [61,62,63,64,65], DNA: RNA hybrids are generated as intermediary structures. Additionally, they have been found to form R-loops in a variety of organisms, including bacteria and humans, and these loops are essential for controlling gene expression, DNA and histone modifications, immunoglobulin class switch recombination, DNA replication, and genome stability [66]. Small organic compounds having the capacity to specifically restrict DNA replication via Okazaki fragments, therefore also preventing transcription, offer a great deal of promise in the treatment of cancer because the replication is frequently enhanced in cancer cells furthermore, telomerase and RNase inhibitors, which are tiny compounds selective for hybrid duplexes, may be used therapeutically [67,68,69,70].
The discovery of substances that specifically attach to DNA: RNA hybrids have only been briefly discussed in the literature [71,72,73]. The existence of a few ligands that selectively bind DNA: RNA hybrids are shown by the literature sources [71,72,73,74,75,76]. A common structural motif that preferentially binds to the hybrid structures was discovered in extensive research by Arya and Chaires using quick screening assays, competition dialysis, and thermal denaturation of mixtures [71,72,73,74,75,76,77]. Ethidium bromide, coralyne, amino glycoside, propidium, thiazole orange, and ellipticine, among others, showed preferential binding to hybrid duplexes among other nucleic acid structures. These substances all possess the common motif planar aromatic ring system with a “bay” area. Triplex structures have been discovered in biologically important RNA molecules, such as telomerase RNAs, [78,79], long noncoding RNAs (lncRNAs) [80,81], RNA pseudoknots responsible for 1 programmed ribosomal frameshifting (PRF) [82], catalytic centers of spliceosomes [83], bacterial SAM-II riboswitch [1], or self-splicing group I [84].
Given the importance of triplexes in RNA biology, tiny compounds capable of recognizing, binding, and stabilizing triple helical RNA structures are gaining attention as potential molecular biology tools and therapeutic agents in contrast to many investigations on DNA triplexes [85,86]. The interaction of small molecules with RNA triplexes has received significantly less attention than organic compounds, aminoglycosides, and various dyes, which are examples of compounds that stabilize RNA triplexes, such as neomycin [87] tunable diphenylfuran based scaffold (furamidine) [88], imidazoles, benzimidazoles [89], flavonoids [90], alkaloids, and their analogs [91,92]. Proflavine and its conjugate [93] and transition metal [94,95,96] intercalation and groove binding are the key mechanisms by which these chemicals interact with RNA triplexes.

2.1.2. PNA

Noncoding RNA (ncRNA) participates in crucial but poorly understood biological processes, and the emergence of illness [97,98,99] research tools for sequence-specific recognition, detection, and functional inhibition of such RNAs would be very beneficial for both theoretical applications in biology and real-world biotechnology and medical applications. A lot of non-coding RNAs acquire complicated tertiary structures with double helical sections that can be targeted via triple helical recognition. The molecular identification of such helical RNA structures has not, however, received much research [100,101].
Triplex forming oligonucleotides (TFOs), or their chemically modified counterparts, can identify double-helical DNA and RNA in a sequence-specific manner [86], but a significant obstacle to the synthesis of triple helices is that only the purine nucleobases can be identified through two Hoogsteen hydrogen bonds in the naturally occurring T*A-T (or U*A-U) and C + *G-C triplets. Because inverted U-A and C-G base pairs only have one hydrogen bond donor or acceptor in the main groove of the duplex, it is extremely difficult to recognize pyrimidine nucleobases in these base pairs. Due to the electrostatic attraction between a TFO’s negatively charged phosphates and the target nucleic acid duplex, low binding affinity and sluggish kinetics of TFOs are also issues, as is the requirement for protonation of cytosine (pKa 4.5) to generate the unfavorable C + *G-C triplet. In recent developments of cationic triplex-forming peptide nucleic acid, the latter two issues have been resolved (PNA) [102,103,104,105,106,107]. The nucleobases of PNA are joined by a pseudopeptide backbone, making it a neutral DNA counterpart. PNA was first designed as a double-stranded DNA (dsDNA) triplex-forming ligand [108]. However, it was later shown that PNA can invade dsDNA by dislodging a pyrimidine rich DNA strand and generating a PNA-DNA-PNA triplex, opening a brand new and fascinating mode of DNA recognition. Although triple-helical PNA binding to dsDNA has also been investigated, the first report of PNA binding to dsRNA was made by Rozners and colleagues in 2010 [109,110,111,112,113], along with others [104], who demonstrated that dsRNA and nucleobase-modified PNA produced triple helices with great affinity and sequence specificity [109,110,111,112,113].
PNA-dsRNA triplex production at physiological pH and salt concentration was made possible by swapping out the more basic cytosine nucleobase for the more basic 2-aminopyridine (M) nucleobase (pKa 6.7, hence partially protonated at physiological circumstances). Building on the work of Corey [114,115] and Gait [116,117], conjugation with cationic lysine residues improved the binding affinity and accelerated cellular absorption of the cationic triplex forming PNA without impairing the binding specificity [106,107]. The most striking finding was that PNA exhibited an approximately 10-fold greater affinity for complementary dsRNA than for the same dsDNA sequence. However, sequences with most purine nucleobases on one strand of the dsRNA were still required for the creation of extremely stable PNA-dsRNA triplexes (polypurine tracts) [105,106,107].
One may argue that the most important restriction on triple helical identification of nucleic acids is the requirement for polypurine tracts [118]. A few research teams have created heterocyclic nucleobases that can join the exocyclic -NH2 of cytosine to form a single hydrogen bond. Leumann demonstrated that the C–G inversion was preferentially recognized by 5-methylpyrimidin-2-one (4HT, R2 = CH3) modified TFOs over other base pairs, but the binding affinity was lower than that of conventional Hoogsteen triplets [119] when pyrimidin-2-one (P, R2 = H) was added to triplex forming PNA targeting dsRNA [120] despite having a decreased affinity, as well as P-modified PNA still bound to mRNA in a triplex with a hairpin structure, which prevented it from being translated both in vitro and in living cells [118]. The latter investigation was the initial proof of the biological impact of PNA-dsRNA triplex production in living cells [121].
A tried-and-true technique for low-cost, high throughput relative quantification is the use of DNA microarrays these arrays are used for quick screening of up/down-regulation. However, they have a limited dynamic range, and their ability to hybridize is heavily dependent on factors, including temperature, ionic strength, and probe affinity [122]. Another well-known technique, qRT-PCR, provides for the accurate quantification of numerous miRs and is frequently used to validate the results of microarray analysis on a particular collection of targets. Finally, RNA sequencing is a well-known technology used for the accurate identification and quantification of miR sequences, as well as for the discrimination of closely related sequences.
Some drawbacks to PNA use should be considered while designing probes. Reduced solubility in aqueous media is a side effect of the neutral poly-amidic backbone, which ensures the production of robust duplexes with natural nucleic acid targets. By adding charged and hydrophilic residues, such as charged amino acids, this can be easily prevented (i.e., arginine, glutamic acid, etc.) when operating at the low micromolar concentrations typically used for oligonucleotide detection, this lowered solubility limitation is typically less significant, but it could still be a problem when it comes to the manufacture of the device. For instance, it has been demonstrated that the polyamidic backbone of PNA causes issues when thiolated probes are conjugated to the surface of gold nanoparticles [123], and to address this, alternate gold decorating protocols must be developed [124,125]. Furthermore, enzymes, such as ligases or polymerases, which are currently used for amplification-based detection methods depending on DNA probes, are unable to recognize PNAs due to their artificial origin. PNAs have found many uses in nucleic acid-targeting strategies and have been proposed as biomolecule-based drugs and as probes for diagnostic tools [126,127,128,129]. Despite these drawbacks, which may deter their application, given their straightforward conjugation with peptide sequences and small functional groups and the benefits in terms of duplex stability, PNAs have found a variety of applications in these fields. The objective of this review is to provide a summary of current advancements in PNA-based miR detection methods, paying particular focus to the key advantages over conventional DNA-based approaches.

2.2. The Impact of TFOs on Gene Expression

The sequence of RNA and DNA defines its secondary and tertiary structure. Their affinity and specificity are finally defined by the structure [130]. Single stranded DNA or RNA molecules are purpose-built to attach to a certain target molecule. As a result, they might be used as components of macromolecular devices for basic research and technological applications. Because of interactions with complementary portions of the chain and stacking, each aptamer has a distinct three-dimensional structure.
The in-cell NMR approach has been proven to be a reliable tool for determining the intracellular structures of oligodeoxynucleotides. Several groups, including ours, have used in-cell NMR to directly identify the production of DNA hairpins, [131] RNA hairpins, I-motifs [132], G-quadruplexes [133], and Z-DNA [134] structures in living human cells. The NMR approach was also used to assess tRNA changes in real-time in yeast cell lysate [135,136,137].
According to research, approximately 97% of human genes have at least one triplex forming sequence, with 86% having a unique arrangement [88]. Because of this, as well as their programmability and selectivity, the creation of efficient TFOs is an essential problem because it gives a mechanism to directly control gene expression [138] with further applications in bionanotechnology and synthetic biology [1].

Triplex-Mediated Modulation of Gene Transcription

A multistep process called transcription involves the dynamic interaction of DNA with various protein partners as well as the synthesis of DNA-protein complexes and structures with various inherent stabilities. It has long been believed that it is simpler to inhibit RNA synthesis during transcription elongation than to negatively (or positively) modulate transcription initiation via triplex competition with transcription activators (or repressors) [139].

2.3. Biosensor Based on Aptamers

Aptamer based biosensor designs have received a lot of interest in the disciplines of biotechnology and chemistry [140,141,142,143]. Single stranded RNA or DNA oligonucleotide aptamers with distinct intramolecular structures have been utilized to influence chemical production, and their selectivity and affinity for target have been used in various types of biosensors [144,145]. Aptamers are usually intended to recognize and bind to the target when paired with a biosensor to create an aptasensor, with the conformational change employed to generate many detectable signals. Aptasensors have frequently been used to test mycotoxins because of their capacity to concurrently measure bind constants for most mycotoxins, as well as dissociation constants ranging within the nanomole [146,147,148,149]. Wu et al., for example, created a one-step electrochemical aptasensor for the assessment of OTA utilizing a thiol and methylene blue dual-labeled aptamer-modified gold electrode and demonstrated a detection limit of 0.095 pg/mL [150]. Lv et al. reported a fluorescence test for OTA detection based on aptamers and gold nanoparticles with a detection limit of 5 nM [151]. Lin et al. created a simple, low-cost, and sensitive liposome-based colorimetric aptasensor to detect OTA, with a detection limit of 0.023 ng/mL [152]. A triple helix aptamer probe (TAP) was introduced as a biosensing strategy, with critical reference to Watson-Crick and Hoogsteen base pairings. The probe consisted of an intended aptamer sequence with two arm-like segments flanked by the sides of the aptamer and a triplex forming oligonucleotide [153] This sensor offered various benefits over traditional duplex DNA biosensors. Aside from its duplex DNA-like stability, the tap was painstakingly developed with a longer aptamer sequence within the loop, strengthening selectivity, binding affinity, and sensitivity to the targets [154,155,156]. Furthermore, the aptamer could be changed without having to modify the probes for signal transduction, allowing it to be used to detect a variety of targets. Because of these advantages, the tap has received increased interest from scientists and has been used in a variety of innovative biosensing technologies.

2.4. G-Quadruplex Forming Aptamer

The finding is related to nanobiotechnology and molecular medicine and involves DNA aptamers to the thrombin exosite I interact with prothrombin. These chemicals have the potential to be exploited to develop treatments that prevent intravascular thrombus development. Aptamer-based biosensors for sensing lateral flow are described here [157]. Aptamers are extremely valuable and include biorecognizing chemicals for the construction of biosensors. Aptamers are single-stranded strands of DNA that are used to select protein targets. SELEX is a selection approach that permits the separation of motivated nucleic acid molecules from a huge collection (more than 1015) of individual molecules, known as a combinatorial library [158,159]. Some of the aptamers in Figure 2 (modified by authors from [160]) have a good probability of not only identifying their targets, but also of destroying their biological vitality.
The SELEX technique is used to find and select aptamers. It happens because of exponential enrichment in the systematic development of ligands. The oligonucleotide library is gradually enriched with sequences that have a higher affinity for the target molecule. The process of producing target incubation, elution, and amplification of the binding sequence is continued in classical SELEX [161] until the great majority of the stored pool consists of target binding sequences.

2.5. Nucleic Acid Motifs with pH Sensitivity

Many pH-responsive motifs have been discovered in DNA structure research that may be carefully tailored to achieve conformation flipping at a chosen pH value (see Figure 3) [162]. Other than the classic Watson-Crick orientation, nucleic acids can attach in a variety of orientations, including reverse Watson-Crick, Hoogsteen, and “wobble” base pairs. Many of these interactions may be found in aptamer structures. The G-quartet motif, for example, is created by Hoogsteen interactions between guanines from four different strands, with each guanine at right angles to the next on a linear plane [163] (Figure 3). When numerous G-quartets (typically three) are layered, the resultant helical structure is known as a G-quadruplex [164]. G-quadruplexes are often found in aptamers and other synthetic DNA nanostructures, as well as mRNA and gene promoters [164]. Protonation of nucleobases can considerably enhance the stability of some base pair mismatches [163]. Adenine and cytosine, in particular, have been seen to become protonated at moderately acidic pH depending on the structure of the folded nucleic acid. The pKa of protonation can be adjusted up to neutrality, allowing mispairs to be stably absorbed into a structure [165,166]. Because protonation allows various base pairs to become stable, altering the pH of the environment can allow a DNA structure, such as an aptamer, to change conformation reversibly [163]. This conformational shift influences features, such as target binding.
The I-motif is a frequent pH-responsive DNA motif. The I-motif is made up of two intercalated duplexes connected by antiparallel C + C base pairing [167]. Isolated cytosine has a pKa of 4.58 [168] at neutral pH. I-motifs can develop due to crowding or super helicity. A new study suggests that this structure affects replication and transcription in vivo [169]. I-motifs, on the other hand, are completely folded at pH 5–6 and unstable at neutral pH under most in vitro circumstances [168,169]. Because of its pH sensitivity, the I-motif has become popular for nanodevices engineered to flip conformation under moderately acidic circumstances [168]. Bielecka and Juskowiak developed a fluorescent probe for pH monitoring based on I-motif creation in 2015, modifying the protonation equilibrium of a fluorescent cytosine analog and thereby quenching it under more acidic circumstances [170]. I-motifs have also been used in more complicated nanostructures, such as logic gates and motors [162]. Two aptamers that employ the I-motif to integrate pH-dependent binding are mentioned below.
A second common DNA pattern that reacts to pH is the triplex structure, several base combinations can produce a triplex, which consists of two strands in a Watson-Crick duplex and a third, extra strand connected to one of the other two through Hoogsteen or reverse Hoogsteen interactions [171]. The triplex motif has been utilized to generate structural change depending on a variety of triggers, the most prevalent of which is strand displacement [170]. However, changes in pH can be utilized to maintain or disrupt a triplex: the C + GC triplex motif, for example, is dependent on the protonation of the cytosine on the third strand [171,172].
The possible pH range for triplex DNA production can be varied by using different triplex DNA sequences. Idili et al. disclosed a logical design of a triplex DNA structure with varying quantities of T A T/C G triplets that “opened” or “closed” based on an intramolecular triplex structure [173]. In addition to constructing a pH-responsive sensor with a large pH window, they were also able to fine-tune the pH below which the switch was closed by altering the percentage of C + GC triplets (pKa 4.5) and TAT triplets (pKa 10.2). These are two aptamers that govern target binding via pH-controlled triplex synthesis. The A-motif, a parallel duplex comprised of A + A + base pairs that develop at pH 3–4, is another pH-responsive motif. Because the motif is more acidic than the I-motif or triplex, it has been proposed that the A-motif may be beneficial for developing pH-sensitive probes below the 5.5–7.6 range, most easily identified by I-motif probes [174]. Other non-canonical base pairs, in addition to the motifs mentioned above, can be employed to introduce pH-controlled aptamer binding when protonated, A + G, A + C, and C + C base pairs all become substantially more stable [173].

2.6. Triplex Cations

For single-molecule force spectroscopy, a rescue rope technique was created that allowed us to repeatedly alter the structures generated at free ends of DNA, such as triplex DNA, telomeric DNA, or broken dsDNA following click chemistry conjugation. DNA of interest split apart from the rope DNA handle. The rescue rope DNA assumed a shaped conformation, and DNA of interest creates structures at the crossing point, according to AFM imaging. Further proof that the DNA of interest may repeatedly form structures after being ruptured by stresses came from single-molecule magnetic tweezers. It has been also shown how to mechanically probe DNA triplexes using the rescue rope approach. DNA triplexes can be either YR*Y or YR*R (Y for pyrimidine and R for purine) arrangements, depending on whether the TFOs attach to the dsDNA’s purine strand via forward or reverse Hoogsteen pairings (*) in YR*Y or YR*R triplexes. The mutual orientation of the chemically homologous strands might be either parallel or antiparallel. It has long been suggested that a telomeric triplex model in a YR*R arrangement prevents digestion and recombination with a dsDNA of (TTGGGG)3 and a 3 overhang of (TTGGGG)2. A DNA triplex occurs in a small tetrahymena telomere, where the G rich overhang folds back and resides in the main groove of the duplex by generating the CG*G triads in an antiparallel YR*R configuration [175].
Using rescue rope strategy-assisted single molecule force spectroscopy, the production of triplexes by a short human telomeric DNA with eight repetitions of TTAGGG in the dsDNA and the ssDNA overhang has been looked at. The findings showed that the three overhangs of TTAGGG, which are most likely formed by CG*G and TA*T triads, can occasionally fold back to bind dsDNA and create mechanically stable triplexes. The interactions between the overhang and the dsDNA may change because of the TTAGGG repeating motif, resulting in a folding turn of different proportions. When folding a DNA triplex, the turn might be as tight as one that is devoid of nucleotides [176]. The telomeric duplex with repeated patterns cooperatively unfolds under the influence of forces without creating a triplex. A force ramp assay or a force jump experiment may produce a hopping characteristic or a stopping signal if there is a stabilizing triplex present. The remaining duplex area, which is around 17 bp long, may unfold before the triplex in this case. We saw hopping signals in addition to halting signals for all the mutants. A telomere sequence’s folding back triplex structures may hold together for an average of 3 s at a force of 20 pN. Telomeric triplexes may obstruct DNA cleavage and recombination, as well as force-regulated helicases [177,178] and force-sensitive RNA polymerases, due to their great mechanical stability and extended endurance [179]. Thus, two thymines significantly weaken the human telomeric triplex. This is further corroborated by the finding that AA mutation of TTAGGG to AAAGGG improved the chance of triplex formation in force-jump experiments by a factor of five. Our findings on GA TFOs revealed even greater triplex formation probabilities than those on AA mutants, indicating that DNA sequences significantly influence the likelihood that triplex structures will develop (Figure 4) (e.g., [175]).

3. Conclusions

The research on oligonucleotides with triplex-forming ability (TFO) in biological applications continues to profit tremendously from triplex technology. Even though the processes processing TFO-directed triplex formations and related lesions are not completely understood, it is known that proteins from numerous repair pathways are involved.
This repair method, as demonstrated by TFO targeted mutation generation, may be prone to error. Because of developments in the manufacture of chemically modified TFOs, we now have a better understanding of the cellular DNA repair mechanism, where proteins from distinct DNA repair pathways can work together or compete with one another in processing DNA lesions caused by TFOs. Triplex technology still has disadvantages, particularly for in vivo applications. These disadvantages include TFO absorption by cells, affinity and specificity for the binding site, and in vivo stability. As a result, current research is primarily focused on improving the efficacy of this approach of genome editing by chemical alterations to TFOs and improved cellular delivery mechanisms. Although there are still limits, triplex technology has played an important role in developing the science of targeted genome editing. While aptamers have been successfully integrated into pH-activatable systems and methods for altering aptamer binding capacities exist, the applications of designed pH-activatable aptamers remain largely unexplored. Aptamers that activate at weakly acidic pH might operate as an additional layer of control for the drug-delivery system as an extension of the drug-delivery systems mentioned above.

Author Contributions

Conceptualization, I.B.; methodology, I.B., A.A.V. and A.Y.S.; writing—original draft preparation, I.B. and A.A.V.; writing—review and editing, A.A.V. All authors have read and agreed to the published version of the manuscript.

Funding

This paper has been supported by the RUDN University Strategic Academic Leadership Program (recipient A.A.V.).

Acknowledgments

The authors thank Vasilisa D. Bystrykh for her assistance with the edition of the submission’s final version. Alexandre A. Vetcher is grateful to the RUDN University Strategic Academic Leadership Program for the support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hu, Y.; Cecconello, A.; Idili, A.; Ricci, F.; Willner, I. Triplex DNA nanostructures: From basic properties to applications. Angew. Chem. Int. Ed. 2017, 56, 15210–15233. [Google Scholar] [CrossRef] [PubMed]
  2. Chandrasekaran, A.R.; Rusling, D.A. Triplex-forming oligonucleotides: A third strand for DNA nanotechnology. Nucleic Acids Res. 2018, 46, 1021–1037. [Google Scholar] [CrossRef] [PubMed]
  3. Jain, A.; Wang, G.; Vasquez, K.M. DNA triple helices: Biological consequences and therapeutic potential. Biochimie 2008, 90, 1117–1130. [Google Scholar] [CrossRef] [PubMed]
  4. Taniguchi, Y.; Sasaki, S. Enhancements in the utilization of antigene oligonucleotides in the nucleus by booster oligonucleotides. Org. Biomol. Chem. 2012, 10, 8336–8341. [Google Scholar] [CrossRef] [PubMed]
  5. Graham, M.K.; Brown, T.R.; Miller, P.S. Targeting the human androgen receptor gene with platinated triplex-forming oligonucleotides. Biochemistry 2015, 54, 2270–2282. [Google Scholar] [CrossRef]
  6. Goñi, J.R.; De La Cruz, X.; Orozco, M. Triplex-forming oligonucleotide target sequences in the human genome. Nucleic Acids Res. 2004, 32, 354–360. [Google Scholar] [CrossRef]
  7. Robles, J.; Grandas, A.; Pedroso, E.; Luque, F.; Eritja, R.; Orozco, M. Nucleic acid triple helices: Stability effects of nucleobase modifications. Curr. Org. Chem. 2002, 6, 1333–1368. [Google Scholar] [CrossRef]
  8. Soyfer, V.N.; Potaman, V.N. Triple-Helical Nucleic Acids; Springer: New York, NY, USA, 1996. [Google Scholar]
  9. Giovannangeli, C.; Helene, C. Triplex technology takes off. Nat. Biotechnol. 2000, 18, 1245–1246. [Google Scholar] [CrossRef]
  10. Ebbinghaus, S.W.; Vigneswaran, N.; Miller, C.R.; Chee-Awai, R.A.; Mayeld, C.A.; Curiel, D.T.; Miller, D.M. Efficient delivery of triplex forming oligonucleotides to tumor cells by adenovirus-polylysine complexes. Gene Ther. 1996, 3, 287–297. [Google Scholar]
  11. Zendegui, J.G.; Vasquez, K.M.; Tinsley, J.H.; Kessler, D.J.; Hogan, M.E. In vivo stability and kinetics of absorption and disposition of 3’phosphopropyl amine oligonucleotides. Nucleic Acids Res. 1992, 20, 307–314. [Google Scholar] [CrossRef]
  12. Bates, A.D.; Maxwell, A. DNA Topology; Oxford University Press: Oxford, UK, 2005. [Google Scholar]
  13. Fogg, J.M. Bullied no more: When and how DNA shoves proteins around. Q. Rev. Biophys. 2012, 45, 257–299. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Mirkin, S.M. DNA topology: Fundamentals. Encycl. Life Sci. 2001, 111. [Google Scholar] [CrossRef]
  15. Schvartzman, J.B.; Hernández, P.; Krimer, D.B.; Dorier, J.; Stasiak, A. Closing the DNA replication cycle: From simple circular molecules to supercoiled and knotted DNA catenanes. Nucleic Acids Res. 2019, 47, 7182–7198. [Google Scholar] [CrossRef] [PubMed]
  16. Sobetzko, P. Transcription-coupled DNA supercoiling dictates the chromosomal arrangement of bacterial genes. Nucleic Acids Res. 2016, 44, 1514–1524. [Google Scholar] [CrossRef] [PubMed]
  17. Newton, M.D. DNA stretching induces Cas9 off-target activity. Nat. Struct. Mol. Biol. 2019, 26, 185–192. [Google Scholar] [CrossRef]
  18. Travers, A.; Muskhelishvili, G. A common topology for bacterial and eukaryotic transcription initiation? EMBO Rep. 2007, 8, 147–151. [Google Scholar] [CrossRef]
  19. Van Aelst, K.; Martínez-Santiago, C.J.; Cross, S.J.; Szczelkun, M.D. The effect of DNA topology on observed rates of R-loop formation and DNA strand cleavage by CRISPR Cas12a. Genes 2019, 10, 169. [Google Scholar] [CrossRef]
  20. Holoch, D.; Moazed, D. RNA-mediated epigenetic regulation of gene expression. Nat. Rev. Genet. 2015, 16, 71–84. [Google Scholar] [CrossRef]
  21. Oo, J.A.; Brandes, R.P.; Leisegang, M.S. Long non-coding RNAs: Novel regulators of cellular physiology and function. Pflügers Arch.-Eur. J. Physiol. 2021, 474, 191–204. [Google Scholar] [CrossRef]
  22. Takemata, N.; Ohta, K. Role of non-coding RNA transcription around gene regulatory elements in transcription factor recruitment. RNA Biol. 2017, 14, 1–5. [Google Scholar] [CrossRef]
  23. Andric, V.; Nevers, A.; Hazra, D.; Auxilien, S.; Menant, A.; Graille, M.; Palancade, B.; Rougemaille, M. A scaffold lncRNA shapes the mitosis to meiosis switch. Nat. Commun. 2021, 12, 770. [Google Scholar] [CrossRef] [PubMed]
  24. Shimada, Y.; Mohn, F.; Bühler, M. The RNA-induced transcriptional silencing complex targets chromatin exclusively via interacting with nascent transcripts. Genes Dev. 2016, 30, 2571–2580. [Google Scholar] [CrossRef] [PubMed]
  25. Csankovszki, G.; Nagy, A.; Jaenisch, R. Synergism of Xist RNA, DNA methylation, and histone hypoacetylation in maintaining X chromosome inactivation. J. Cell Biol. 2001, 153, 773–784. [Google Scholar] [CrossRef] [PubMed]
  26. Zhang, J.Y.; Zheng, K.w.; Xiao, S. Mechanism, and manipulation of DNA: RNA hybrid G-quadruplex formation in the transcription of G-rich DNA. J. Am. Chem. Soc. 2014, 136, 1381–1390. [Google Scholar] [CrossRef]
  27. García Muse, T.; Aguilera, A. R loops: From physiological to pathological roles. Cell 2019, 179, 604–618. [Google Scholar] [CrossRef]
  28. Santos Pereira, J.M.; Aguilera, A. R loops: New modulators of genome dynamics and function. Nat. Rev. Genet. 2015, 16, 583–597. [Google Scholar] [CrossRef]
  29. Crossley, M.P.; Bocek, M.; Cimprich, K.A. R-loops as cellular regulators and genomic threats. Mol. Cell 2019, 73, 398–411. [Google Scholar] [CrossRef]
  30. Felsenfeld, G.; Davies, D.R.; Rich, A. Formation of a three-stranded polynucleotide molecule. J. Am. Chem. Soc. 1957, 79, 2023–2024. [Google Scholar] [CrossRef]
  31. Escudeé, C.; Francçois, J.C.; Sun, J.-S. Stability of triple helices containing RNA and DNA strands: Experimental and molecular modeling studies. Nucleic Acids Res. 1993, 21, 5547–5553. [Google Scholar] [CrossRef]
  32. Mondal, T.; Subhash, S.; Vaid, R.; Enroth, S.; Uday, S.; Reinius, B.; Mitra, S.; Mohammed, A.; James, A.R.; Hoberg, E.; et al. MEG3 long noncoding RNA regulates the TGF-β pathway genes through formation of RNA–DNA triplex structures. Nat. Commun. 2015, 6, 7743. [Google Scholar] [CrossRef]
  33. Chu, C.; Qu, K.; Zhong, F.L.; Artandi, S.E.; Chang, H.Y. Genomic maps of long noncoding RNA occupancy reveal principles of RNA-chromatin interactions. Mol. Cell 2011, 44, 667–678. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Vrba, L.; Futscher, B.W. Epigenetic silencing of lncRNA MORT in 16 TCGA cancer types. F1000Research 2018, 7, 211. [Google Scholar] [CrossRef]
  35. Jiang, M.C.; Ni, J.J.; Cui, W.Y. Emerging roles of lncRNA in cancer and therapeutic opportunities. Am. J. Cancer Res. 2019, 9, 1354. [Google Scholar]
  36. Zhou, Z.; Giles, K.E.; Felsenfeld, G. DNA· RNA triple helix formation can function as a cis-acting regulatory mechanism at the human β-globin locus. Proc. Natl. Acad. Sci. USA 2019, 116, 6130–6139. [Google Scholar] [CrossRef] [PubMed]
  37. Garratt, H.; Ashburn, R.; Sopic, M. Long Non-Coding RNA Regulation of Epigenetics in Vascular Cells. Non-Coding RNA 2021, 7, 62. [Google Scholar] [CrossRef]
  38. Maldonado, R.; Filarsky, M.; Grummt, I. Purine–and pyrimidine–triple-helix-forming oligonucleotides recognize qualitatively different target sites at the ribosomal DNA locus. RNA 2018, 24, 371–380. [Google Scholar] [CrossRef] [PubMed]
  39. Buske, F.A.; Bauer, D.C.; Mattick, J.S.; Bailey, T.L. Triplexator: Detecting nucleic acid triple helices in genomic and transcriptomic data. Genome Res. 2012, 22, 1372–1381. [Google Scholar] [CrossRef]
  40. Kuo, C.-C.; Hänzelmann, S.; Cetin, N.S.; Frank, S.; Zajzon, B.; Derks, J.-P.; Akhade, V.S.; Ahuja, G.; Kanduri, C.; Grummt, I.; et al. Detection of RNA–DNA binding sites in long noncoding RNAs. Nucleic Acids Res. 2019, 47, e32. [Google Scholar] [CrossRef]
  41. He, S.; Zhang, H.; Liu, H.; Zhu, H. LongTarget: A tool to predict lncRNA DNA-binding motifs and binding sites via Hoogsteen base-pairing analysis. Bioinformatics 2015, 31, 178–186. [Google Scholar] [CrossRef]
  42. Antonov, I.V.; Mazurov, E.; Borodovsky, M. Prediction of lncRNAs and their interactions with nucleic acids: Benchmarking bioinformatics tools. Brief. Bioinform. 2019, 20, 551–564. [Google Scholar] [CrossRef]
  43. Zhang, Y.; Long, Y.; Kwoh, C.K. Deep learning-based DNA: RNA triplex forming potential prediction. Brief. Bioinform. 2020, 21, 522. [Google Scholar] [CrossRef] [PubMed]
  44. Sentürk Cetin, N.; Kuo, C.C.; Ribarska, T.; Li, R.; Costa, I.G.; Grummt, I. Isolation, and genome-wide characterization of cellular DNA: RNA triplex structures. Nucleic Acids Res. 2019, 47, 2306–2321. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Zhou, B.; Li, X.; Luo, D. GRID-seq for comprehensive analysis of global RNA–chromatin interactions. Nat. Protoc. 2019, 14, 2036–2068. [Google Scholar] [CrossRef] [PubMed]
  46. Zhong, S.; Sridhar, B.; Rivas Astroza, M. Mapping RNA-chromatin interactions. FASEB J. 2018, 32, 525.2. [Google Scholar] [CrossRef]
  47. Wu, W.; Yan, Z.; Nguyen, T.C. Mapping RNA–chromatin interactions by sequencing with iMARGI. Nat. Protoc. 2019, 14, 3243–3272. [Google Scholar] [CrossRef]
  48. Bell, J.C.; Jukam, D.; Teran, N.A. Chromatin-associated RNA sequencing (ChAR-seq) maps genome-wide RNA-to-DNA contacts. Elife 2018, 7, e27024. [Google Scholar] [CrossRef]
  49. Bonetti, A.; Agostini, F.; Suzuki, A.M. RADICL-seq identifies general and cell type–specific principles of genome-wide RNA-chromatin interactions. Nat. Commun. 2020, 11, 1018. [Google Scholar] [CrossRef]
  50. Gavrilov, A.A.; Zharikova, A.A.; Galitsyna, A.A. Studying RNA–DNA interactome by Red-C identifies noncoding RNAs associated with various chromatin types and reveals transcription dynamics. Nucleic Acids Res. 2020, 48, 6699–6714. [Google Scholar] [CrossRef]
  51. Schmitz, K.M.; Mayer, C.; Postepska, A.; Grummt, I. Interaction of noncoding RNA with the rDNA promoter mediates recruitment of DNMT3b and silencing of rRNA genes. Genes Dev. 2010, 24, 2264–2269. [Google Scholar] [CrossRef]
  52. Grote, P.; Herrmann, B.G. The long non-coding RNA Fendrr links epigenetic control mechanisms to gene regulatory networks in mammalian embryogenesis. RNA Biol. 2013, 10, 1579–1585. [Google Scholar] [CrossRef]
  53. O’Leary, V.B.; Ovsepian, S.V.; Carrascosa, L.G.; Buske, F.A.; Radulovic, V.; Niyazi, M.; Moertl, S.; Trau, M.; Atkinson, M.J.; Anastasov, N. PARTICLE, a triplex-forming long ncRNA, regulates locus-specific methylation in response to low-dose irradiation. Cell Rep. 2015, 11, 474–485. [Google Scholar] [CrossRef] [PubMed]
  54. Postepska Igielska, A.; Giwojna, A.; Gasri-Plotnitsky, L.; Schmitt, N.; Dold, A.; Ginsberg, D.; Grummt, I. LncRNA Khps1 regulates expression of the proto-oncogene SPHK1 via triplex-mediated changes in chromatin structure. Mol. Cell 2015, 60, 626–636. [Google Scholar] [CrossRef] [PubMed]
  55. Kalwa, M.; Hänzelmann, S.; Otto, S.; Kuo, C.C.; Franzen, J.; Joussen, S.; Fernandez-Rebollo, E.; Rath, B.; Koch, C.; Hofmann, A.; et al. The lncRNA HOTAIR impacts on mesenchymal stem cells via triple helix formation. Nucleic Acids Res. 2016, 44, 10631–10643. [Google Scholar] [CrossRef] [PubMed]
  56. Wang, S.; Ke, H.; Zhang, H.; Ma, Y.; Ao, L.; Zou, L.; Yang, Q.; Zhu, H.; Nie, J.; Wu, C.; et al. LncRNA MIR100HG promotes cell proliferation in triple-negative breast cancer through triplex formation with p27 loci. Cell Death Dis. 2018, 9, 805. [Google Scholar] [CrossRef]
  57. Zhao, Z.; Sentürk, N.; Song, C.; Grummt, I. ncRNA PAPAS tethered to the rDNA enhancer recruits hypophosphorylated CHD4/NuRD to repress rRNA synthesis at elevated temperatures. Genes Dev. 2018, 32, 836–848. [Google Scholar] [CrossRef]
  58. Simon, M.D.; Wang, C.I.; Kharchenko, P.V.; West, J.A.; Chapman, B.A.; Oleksiak, A.A.; Borowsky, M.L.; Kuroda, M.I.; Kingston, R.E. The genomic binding sites of a noncoding RNA. Proc. Natl. Acad. Sci. USA 2011, 108, 20497–20502. [Google Scholar] [CrossRef]
  59. Engreitz, J.M.; Pandya-Jones, A.; McDonel, P.; Shishkin, A.; Sirokman, K.; Surka, C.; Kadri, S.; Xing, J.; Goren, A.; Lander, E.S.; et al. The Xist lncRNA exploits three-dimensional genome architecture to spread across the X chromosome. Science 2013, 341, 1237973. [Google Scholar] [CrossRef]
  60. Buske, F.A.; Bauer, D.C.; Mattick, J.S.; Bailey, T.L. Triplex-Inspector: An analysis tool for triplex-mediated targeting of genomic loci. Bioinformatics 2013, 29, 1895–1897. [Google Scholar] [CrossRef]
  61. Houlard, M.; Artus, J.; Leguillier, T.; Vandormael-Pournin, S.; Cohen-Tannoudji, M. DNA-RNA hybrids contribute to the replication dependent genomic instability induced by Omcg1 deficiency. Cell Cycle 2011, 10, 108–117. [Google Scholar] [CrossRef]
  62. Balk, B.; Dees, M.; Bender, K.; Luke, B. The differential processing of telomeres in response to increased telomeric transcription and RNA–DNA hybrid accumulation. RNA Biol. 2014, 11, 95–100. [Google Scholar] [CrossRef]
  63. Stuckey, R.; García-Rodríguez, N.; Aguilera, A.; Wellinger, R.E. Role for RNA: DNA hybrids in origin-independent replication priming in a eukaryotic system. Proc. Natl. Acad. Sci. USA 2015, 112, 5779–5784. [Google Scholar] [CrossRef] [PubMed]
  64. Balk, B.; Maicher, A.; Dees, M.; Klermund, J.; Luke-Glaser, S.; Bender, K.; Luke, B. Telomeric RNA-DNA hybrids affect telomere-length dynamics and senescence. Nat. Struct. Mol. Biol. 2013, 20, 1199–1205. [Google Scholar] [CrossRef] [PubMed]
  65. Li, T.K.; Barbieri, C.M.; Lin, H.-C.; Rabson, A.B.; Yang, G.; Fan, Y.; Gaffney, B.L.; Jones, R.A.; Pilch, D.S. Drug Targeting of HIV-1 RNA DNA Hybrid Structures: Thermodynamics of Recognition and Impact on Reverse Transcriptase-Mediated Ribonuclease H Activity and Viral Replication. Biochemistry 2004, 43, 9732–9742. [Google Scholar] [CrossRef] [PubMed]
  66. Lombraña, R.; Almeida, R.; Alvarez, A.; Gómez, M. R-loops, and initiation of DNA replication in human cells: A missing link? Front. Genet. 2015, 6, 158. [Google Scholar] [CrossRef] [PubMed]
  67. Saretzki, G. Telomerase inhibition as cancer therapy. Cancer Lett. 2003, 194, 209–219. [Google Scholar] [CrossRef]
  68. Mergny, J.L.; Lacroix, L.; Teulade-Fichou, M.P.; Hounsou, C.; Guittat, L.; Hoarau, M.; Arimondo, P.B.; Vigneron, J.-P.; Lehn, J.-M.; Riou, J.-F. Telomerase inhibitors based on quadruplex ligands selected by a fluorescence assay. Proc. Natl. Acad. Sci. USA 2001, 98, 3062–3067. [Google Scholar] [CrossRef]
  69. Su, H.P.; Yan, Y.; Prasad, G.S.; Smith, R.F.; Daniels, C.L.; Abeywickrema, P.D.; Reid, J.C.; Loughran, H.M.; Kornienko, M.; Sharma, S. Structural basis for the inhibition of RNase H activity of HIV-1 reverse transcriptase by RNase H active site-directed inhibitors. J. Virol. 2010, 84, 7625–7633. [Google Scholar] [CrossRef] [PubMed]
  70. Tramontano, E.; Di Santo, R. HIV-1 RT-associated RNase H function inhibitors: Recent advances in drug development. Curr. Med. Chem. 2010, 17, 2837–2853. [Google Scholar] [CrossRef]
  71. Ren, J.; Qu, X.; Dattagupta, N.; Chaires, J. Molecular recognition of a RNA: DNA hybrid structure. J. Am. Chem. Soc. 2001, 123, 6742–6743. [Google Scholar] [CrossRef]
  72. Shaw, N.N.; Arya, D.P. Recognition of the unique structure of DNA: RNA hybrids. Biochimie 2008, 90, 1026–1039. [Google Scholar] [CrossRef]
  73. West, C.; Francis, R.; Friedman, S.H. Small molecule/nucleic acid affinity chromatography: Application for the identification of telomerase inhibitors which target its key RNA/DNA heteroduplex. Bioorganic Med. Chem. Lett. 2001, 11, 2727–2730. [Google Scholar] [CrossRef] [PubMed]
  74. Wheelhouse, R.T.; Chaires, J.B. Drug Binding to DNA⋅RNA Hybrid Structures. In Drug-DNA Interaction Protocols; Humana Press: Totowa, NJ, USA, 2010; pp. 55–70. [Google Scholar]
  75. Shaw, N.N.; Xi, H.; Arya, D.P. Molecular recognition of a DNA: RNA hybrid: Sub-nanomolar binding by a neomycin–methidium conjugate. Bioorganic Med. Chem. Lett. 2008, 18, 4142–4145. [Google Scholar] [CrossRef]
  76. Xi, H.; Davis, E.; Ranjan, N.; Xue, L.; Hyde-Volpe, D.; Arya, D.P. Thermodynamics of nucleic acid “shape readout” by an aminosugar. Biochemistry 2011, 50, 9088–9113. [Google Scholar] [CrossRef] [PubMed]
  77. Wheelhouse, R.T.; Garbett, N.C.; Buurma, N.J.; Chaires, J. Probing the Molecular Recognition of a DNA⋅ RNA Hybrid Duplex. Angew. Chem. 2010, 122, 3275–3278. [Google Scholar] [CrossRef]
  78. Zeraati, M.; Langley, D.B.; Schofield, P.; Moye, A.L.; Rouet, R.; Hughes, W.E.; Bryan, T.M.; Dinger, M.E.; Christ, D. I-motif DNA structures are formed in the nuclei of human cells. Nat. Chem. 2018, 10, 631–637. [Google Scholar] [CrossRef]
  79. Brogden, A.L.; Hopcroft, N.H.; Searcey, M.; Cardin, C.J. Ligand Bridging of the DNA Holliday Junction: Molecular Recognition of a Stacked-X Four-Way Junction by a Small Molecule. Angew. Chem. Int. Ed. 2007, 46, 3850–3854. [Google Scholar] [CrossRef]
  80. Fox, K.R.; Brown, T.; Rusling, D.A. DNA-targeting Molecules as Therapeutic Agents; The Royal Society of Chemistry: London, UK, 2018; pp. 1–32. [Google Scholar]
  81. Hoogsteen, K. The structure of crystals containing a hydrogen-bonded complex of 1-methylthymine and 9-methyladenine. Acta Crystallogr. 1959, 12, 822–823. [Google Scholar] [CrossRef]
  82. Carrascosa, L.G.; Gomez-Montes, S.; Avino, A.; Nadal, A.; Pla, M.; Eritja, R.; Lechuga, L.M. Sensitive and label-free biosensing of RNA with predicted secondary structures by a triplex affinity capture method. Nucleic Acids Res. 2012, 40, e56. [Google Scholar] [CrossRef]
  83. Wang, G.; Seidman, M.M.; Glazer, P.M. Mutagenesis in mammalian cells induced by triple helix formation and transcription-coupled repair. Science 1996, 271, 802–805. [Google Scholar] [CrossRef] [PubMed]
  84. Okamura, H.; Taniguchi, Y.; Sasaki, S. Aminopyridinyl–pseudodeoxycytidine derivatives selectively stabilize antiparallel triplex DNA with multiple CG inversion sites. Angew. Chem. Int. Ed. 2016, 55, 12445–12449. [Google Scholar] [CrossRef]
  85. Sun, J.S.; François, J.C.; Montenay-Garestier, T.; Saison-Behmoaras, T.; Roig, V.; Thuong, N.T.; Hélène, C. Sequence-specific intercalating agents: Intercalation at specific sequences on duplex DNA via major groove recognition by oligonucleotide-intercalator conjugates. Proc. Natl. Acad. Sci. USA 1989, 86, 9198–9202. [Google Scholar] [CrossRef] [PubMed]
  86. Paramasivam, M.; Cogoi, S.; Filichev, V.V.; Bomholt, N.; Pedersen, E.B.; Xodo, L.E. Purine twisted-intercalating nucleic acids: A new class of anti-gene molecules resistant to potassium-induced aggregation. Nucleic Acids Res. 2008, 36, 3494–3507. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Pedersen, E.B.; Osman, A.M.A.; Globisch, D.; Paramasivam, M.; Cogoi, S.; Bomholt, N.; Jorgensen, P.T.; Xodo, L.E.; Filichev, V.V. Triplex glue by synthesizing conjugated flexible intercalators. In Nucleic Acids Symposium Series; Oxford University Press: Oxford, UK, 2008; pp. 37–38. [Google Scholar]
  88. Klimkowski, P.; De Ornellas, S.; Singleton, D.; El-Sagheer, A.H.; Brown, T. Design of thiazole orange oligonucleotide probes for detection of DNA and RNA by fluorescence and duplex melting. Org. Biomol. Chem. 2019, 17, 5943–5950. [Google Scholar] [CrossRef]
  89. Sigman, D.S.; Bruice, T.W.; Mazumder, A.; Sutton, C.L. Targeted chemical nucleases. Acc. Chem. Res. 1993, 26, 98–104. [Google Scholar] [CrossRef]
  90. Pitié, M.; Sudres, B.; Meunier, B. Dramatic increase of the DNA cleavage activity of Cu(Clip-phen) by fixing the bridging linker on the C3 position of the phenanthroline units. Chem. Commun. 1998, 2597–2598. [Google Scholar] [CrossRef]
  91. Bales, B.C.; Kodama, T.; Weledji, Y.N.; Pitie, M.; Meunier, B.; Greenberg, M.M. Mechanistic studies on DNA damage by minor groove binding copper–phenanthroline conjugates. Nucleic Acids Res. 2005, 33, 5371–5379. [Google Scholar] [CrossRef]
  92. Panattoni, A.; El-Sagheer, A.H.; Brown, T.; Kellett, A.; Hocek, M. Oxidative DNA Cleavage with Clip-Phenanthroline Triplex-Forming Oligonucleotide Hybrids. ChemBioChem 2020, 21, 991–1000. [Google Scholar] [CrossRef]
  93. Gasser, G.; Pinto, A.; Neumann, S.; Sosniak, A.M.; Seitz, M.; Merz, K.; Heumann, R.; Metzler-Nolte, N. Synthesis, characterisation and bioimaging of a fluorescent rhenium-containing PNA bioconjugate. Dalton Trans. 2012, 41, 2304–2313. [Google Scholar] [CrossRef]
  94. Slator, C.; Molphy, Z.; McKee, V.; Long, C.; Brown, T.; Kellett, A. Di-copper metallodrugs promote NCI-60 chemotherapy via singlet oxygen and superoxide production with tandem TA/TA and AT/AT oligonucleotide discrimination. Nucleic Acids Res. 2018, 46, 2733–2750. [Google Scholar] [CrossRef]
  95. Kellett, A.; Molphy, Z.; Slator, C.; Mckee, V.; Farrell, N.P. Molecular methods for assessment of non-covalent metallodrug–DNA interactions. Chem. Soc. Rev. 2019, 48, 971–988. [Google Scholar] [CrossRef]
  96. Levine, M.; Wang, Y.; Padayatty, S.J.; Morrow, J. A new recommended dietary allowance of vitamin C for healthy young women. Proc. Natl. Acad. Sci. USA 2001, 98, 9842–9846. [Google Scholar] [CrossRef] [PubMed]
  97. Smith, K.N.; Miller, S.C.; Varani, G.; Calabrese, J.M.; Magnuson, T. Multimodal long noncoding RNA interaction networks: Control panels for cell fate specification. Genetics 2019, 213, 1093–1110. [Google Scholar] [CrossRef] [PubMed]
  98. Yao, R.W.; Wang, Y.; Chen, L.L. Cellular functions of long noncoding RNAs. Nat. Cell Biol. 2019, 21, 542–551. [Google Scholar] [CrossRef]
  99. Treiber, T.; Treiber, N.; Meister, G. Regulation of microRNA biogenesis and its crosstalk with other cellular pathways. Nat. Rev. Mol. Cell Biol. 2019, 20, 5–20. [Google Scholar] [CrossRef] [PubMed]
  100. Mercer, T.R.; Mattick, J.S. Structure and function of long noncoding RNAs in epigenetic regulation. Nat. Struct. Mol. Biol. 2013, 20, 300–307. [Google Scholar] [CrossRef] [PubMed]
  101. Novikova, I.V.; Hennelly, S.P.; Tung, C.S.; Sanbonmatsu, K.Y. Rise of the RNA machines: Exploring the structure of long non-coding RNAs. J. Mol. Biol. 2013, 425, 3731–3746. [Google Scholar] [CrossRef] [PubMed]
  102. Hansen, M.E.; Bentin, T.; Nielsen, P.E. High-affinity triplex targeting of double stranded DNA using chemically modified peptide nucleic acid oligomers. Nucleic Acids Res. 2009, 37, 4498–4507. [Google Scholar] [CrossRef]
  103. Wittung, P.; Nielsen, P.; Norden, B. Extended DNA-Recognition Repertoire of Peptide Nucleic Acid (PNA): PNA–dsDNA Triplex Formed with Cytosine-Rich Homopyrimidine PNA. Biochemistry 1997, 36, 7973–7979. [Google Scholar] [CrossRef]
  104. Li, M.; Zengeya, T.; Rozners, E. Short peptide nucleic acids bind strongly to homopurine tract of double helical RNA at pH 5. 5. J. Am. Chem. Soc. 2010, 132, 8676–8681. [Google Scholar] [CrossRef]
  105. Zengeya, T.; Gupta, P.; Rozners, E. Triple-helical recognition of RNA using 2-aminopyridine-modified PNA at physiologically relevant conditions. Angew. Chem. 2012, 124, 12761–12764. [Google Scholar] [CrossRef]
  106. Muse, O.; Zengeya, T.; Mwaura, J.; Hnedzko, D.; McGee, D.W.; Grewer, C.T.; Rozners, E. Sequence selective recognition of double-stranded RNA at physiologically relevant conditions using PNA-peptide conjugates. ACS Chem. Biol. 2013, 8, 1683–1686. [Google Scholar] [CrossRef] [PubMed]
  107. Hnedzko, D.; McGee, D.W.; Karamitas, Y.A.; Rozners, E. Sequence-selective recognition of double-stranded RNA and enhanced cellular uptake of cationic nucleobase and backbone-modified peptide nucleic acids. RNA 2017, 23, 58–69. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Nielsen, P.E.; Egholm, M.; Berg, R.H.; Buchardt, O. Sequence-selective recognition of DNA by strand displacement with a thymine-substituted polyamide. Science 1991, 254, 1497–1500. [Google Scholar] [CrossRef] [PubMed]
  109. Zhou, Y.; Kierzek, E.; Loo, Z.P.; Antonio, M.; Yau, Y.H.; Chuah, Y.W.; Geifman-Shochat, S.; Kierzek, R.; Chen, G. Recognition of RNA duplexes by chemically modified triplex-forming oligonucleotides. Nucleic Acids Res. 2013, 41, 6664–6673. [Google Scholar] [CrossRef]
  110. Devi, G.; Yuan, Z.; Lu, Y.; Zhao, Y.; Chen, G. Incorporation of thio-pseudoisocytosine into triplex-forming peptide nucleic acids for enhanced recognition of RNA duplexes. Nucleic Acids Res. 2014, 42, 4008–4018. [Google Scholar] [CrossRef]
  111. Sato, T.; Sakamoto, N.; Nishizawa, S. Kinetic and thermodynamic analysis of triplex formation between peptide nucleic acid and double-stranded RNA. Org. Biomol. Chem. 2018, 16, 1178–1187. [Google Scholar] [CrossRef]
  112. Taehtinen, V.; Granqvist, L.; Murtola, M.; Strömberg, R.; Virta, P. 19F NMR Spectroscopic Analysis of the Binding Modes in Triple-Helical Peptide Nucleic Acid (PNA)/MicroRNA Complexes. Chem.–A Eur. J. 2017, 23, 7113–7124. [Google Scholar] [CrossRef]
  113. Kim, K.T.; Chang, D.; Winssinger, N. Double-Stranded RNA-Specific Templated Reaction with Triplex Forming PNA. Helv. Chim. Acta 2018, 101, e1700295. [Google Scholar] [CrossRef]
  114. Hu, J.; Matsui, M.; Gagnon, K.T.; Schwartz, J.C.; Gabillet, S.; Arar, K.; Wu, J.; Bezprozvanny, I.; Corey, D.R. Allele-specific silencing of mutant huntingtin and ataxin-3 genes by targeting expanded CAG repeats in mRNAs. Nat. Biotechnol. 2009, 27, 478–484. [Google Scholar] [CrossRef]
  115. Hu, J.; Corey, D.R. Inhibiting gene expression with peptide nucleic acid (PNA)—Peptide conjugates that target chromosomal DNA. Biochemistry 2007, 46, 7581–7589. [Google Scholar] [CrossRef]
  116. Fabani, M.M.; Gait, M.J. MIR-122 targeting with LNA/2′-O-methyl oligonucleotide mixmers, peptide nucleic acids (PNA), and PNA–peptide conjugates. RNA 2008, 14, 336–346. [Google Scholar] [CrossRef] [PubMed]
  117. Turner, J.J.; Ivanova, G.D.; Verbeure, B.; Williams, D.; Arzumanov, A.A.; Abes, S.; Lebleu, B.; Gait, M.J. Cell-penetrating peptide conjugates of peptide nucleic acids (PNA) as inhibitors of HIV-1 Tat-dependent trans-activation in cells. Nucleic Acids Res. 2005, 33, 6837–6849. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Gupta, P.; Zengeya, T.; Rozners, E. Triple helical recognition of pyrimidine inversions in polypurine tracts of RNA by nucleobase modified PNA. Chem. Commun. 2011, 47, 11125–11127. [Google Scholar] [CrossRef] [PubMed]
  119. Hari, Y.; Obika, S.; Imanishi, T. Towards the sequence-selective recognition of double-stranded DNA containing pyrimidine-purine interruptions by triplex-forming oligonucleotides. Eur. J. Org. Chem. 2012, 2012, 2875–2887. [Google Scholar] [CrossRef]
  120. Prevot Halter, I.; Leumann, C.J. Selective recognition of a CG base-pair in the parallel DNA triple-helical binding motif. Bioorganic Med. Chem. Lett. 1999, 9, 2657–2660. [Google Scholar] [CrossRef] [PubMed]
  121. Endoh, T.; Hnedzko, D.; Rozners, E.; Sugimoto, N. Nucleobase-Modified PNA Suppresses Translation by Forming a Triple Helix with a Hairpin Structure in mRNA In Vitro and in Cells. Angew. Chem. Int. Ed. 2016, 55, 899–903. [Google Scholar] [CrossRef]
  122. Ranasinghe, R.T.; Rusling, D.A.; Powers, V.E.C.; Fox, K.R.; Brown, T. Recognition of CG inversions in DNA triple helices by methylated 3H-pyrrolo [2,3-d] pyrimidin-2(7H)-one nucleoside analogues. Chem. Commun. 2005, 2555–2557. [Google Scholar] [CrossRef]
  123. Anstaett, P.; Zheng, Y.; Thai, T.; Funston, A.M.; Bach, U.; Gasser, G. Synthesis of Stable Peptide Nucleic Acid-Modified Gold Nanoparticles, and their Assembly onto Gold Surfaces. Angew. Chem. 2013, 125, 4311–4314. [Google Scholar] [CrossRef]
  124. Simon, L.; Lautner, G.; Gyurcsányi, R.E. Reliable microspotting methodology for peptide-nucleic acid layers with high hybridization efficiency on gold SPR imaging chips. Anal. Methods 2015, 7, 6077–6082. [Google Scholar] [CrossRef]
  125. Cadoni, E.; Rosa Gastaldo, D.; Manicardi, A.; Mancin, F.; Madder, A. Exploiting Double Exchange Diels-Alder Cycloadditions for Immobilization of Peptide Nucleic Acids on Gold Nanoparticles. Front. Chem. 2020, 8, 4. [Google Scholar] [CrossRef]
  126. Fabbri, E.; Manicardi, A.; Tedeschi, T.; Sforza, S.; Bianchi, N.; Brognara, E.; Finotti, A.; Breveglieri, G.; Borgatti, M.; Corradini, R. Modulation of the biological activity of microRNA-210 with peptide nucleic acids (PNAs). ChemMedChem 2011, 6, 2192–2202. [Google Scholar] [CrossRef] [PubMed]
  127. Brognara, E.; Fabbri, E.; Aimi, F.; Manicardi, A.; Bianchi, N.; Finotti, A.; Breveglieri, G.; Borgatti, M.; Corradini, R.; Marchelli, R. Peptide nucleic acids targeting miR-221 modulate p27Kip1 expression in breast cancer MDA-MB-231 cells. Int. J. Oncol. 2012, 41, 2119–2127. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Brandén, L.J.; Mohamed, A.J.; Smith, C.I. A peptide nucleic acid–nuclear localization signal fusion that mediates nuclear transport of DNA. Nat. Biotechnol. 1999, 17, 784–787. [Google Scholar] [CrossRef]
  129. Gambari, R. Peptide nucleic acids: A review on recent patents and technology transfer. Expert Opin. Ther. Pat. 2014, 24, 267–294. [Google Scholar] [CrossRef] [PubMed]
  130. Bayramoglu, G.; Ozalp, V.C.; Yilmaz, M.; Guler, U.; Salih, B.; Arica, M.Y. Lysozyme specific aptamer immobilized MCM-41 silicate for single-step purification and quartz crystal microbalance (QCM)-based determination of lysozyme from chicken egg white. Microporous Mesoporous Mater. 2015, 207, 95–104. [Google Scholar] [CrossRef]
  131. Yamaoki, Y.; Kiyoishi, A.; Miyake, M.; Kano, F.; Murata, M.; Nagata, T.; Katahira, M. The first successful observation of in-cell NMR signals of DNA and RNA in living human cells. Phys. Chem. Chem. Phys. 2018, 20, 2982–2985. [Google Scholar] [CrossRef]
  132. Dzatko, S.; Krafcikova, M.; Haensel-Hertsch, R.; Fessl, T.; Fiala, R.; Loja, T.; Krafcik, D.; Mergny, J.-L.; Foldynova-Trantirkova, S.; Trantirek, L. Evaluation of the stability of DNA i-Motifs in the nuclei of living mammalian cells. Angew. Chem. Int. Ed. 2018, 57, 2165–2169. [Google Scholar] [CrossRef]
  133. Bao, H.L.; Liu, H.S.; Xu, Y. Downregulated miRNA-26a-5p induces the apoptosis of endothelial cells in coronary heart disease by inhibiting PI3K/AKT pathway. Eur. Rev. Med. Pharmacol. Sci. 2019, 23, 4940–4947. [Google Scholar]
  134. Bao, H.L.; Masuzawa, T.; Oyoshi, T.; Xu, Y. Oligonucleotides DNA containing 8-trifluoromethyl-2′-deoxyguanosine for observing Z-DNA structure. Nucleic Acids Res. 2020, 48, 7041–7051. [Google Scholar] [CrossRef]
  135. Yamaoki, Y.; Nagata, T.; Sakamoto, T.; Katahira, M. Recent progress of in-cell NMR of nucleic acids in living human cells. Biophys. Rev. 2020, 12, 411–417. [Google Scholar] [CrossRef]
  136. Yamaoki, Y.; Nagata, T.; Sakamoto, T.; Katahira, M. Observation of nucleic acids inside living human cells by in-cell NMR spectroscop. Biophys. Phys. 2020, 17, 36–41. [Google Scholar] [CrossRef] [PubMed]
  137. Barraud, P.; Gato, A.; Heiss, M.; Catala, M.; Kellner, S.; Tisne, C. Time-resolved NMR monitoring of tRNA maturation. Nat. Commun. 2019, 10, 3373. [Google Scholar] [CrossRef] [PubMed]
  138. Moser, H.E.; Dervan, P.B. Sequence-specific cleavage of double helical DNA by triple helix formation. Science 1987, 238, 645–650. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Catapano, C.V.; McGuffie, E.M.; Pacheco, D.; Carbone, G.M. Inhibition of gene expression and cell proliferation by triple helix-forming oligonucleotides directed to the c-myc gene. Biochemistry 2000, 39, 5126–5138. [Google Scholar] [CrossRef]
  140. Mairal, T.; Cengiz Özalp, V.; Lozano Sánchez, P.; Mir, M.; Katakis, I.; O’Sullivan, C.K. Aptamers: Molecular tools for analytical applications. Anal. Bioanal. Chem. 2008, 390, 989–1007. [Google Scholar] [CrossRef]
  141. Tombelli, S.; Minunni, M.; Mascini, M. Analytical applications of aptamers. Biosens. Bioelectron. 2005, 20, 2424–2434. [Google Scholar] [CrossRef]
  142. Sharma, R.; Ragavan, K.V.; Thakur, M.S.; Raghavarao, K.S. Recent advances in nanoparticle based aptasensors for food contaminants. Biosens. Bioelectron. 2015, 74, 612–627. [Google Scholar] [CrossRef]
  143. Kim, Y.S.; Raston, N.H.A.; Gu, M.B. Aptamer-based nanobiosensors. Biosens. Bioelectron. 2016, 76, 2–19. [Google Scholar]
  144. Mazaafrianto, D.N.; Maeki, M.; Ishida, A.; Tani, H.; Tokeshi, M. Recent microdevice-based aptamer sensors. Micromachines 2018, 9, 202. [Google Scholar] [CrossRef]
  145. Razmi, N.; Baradaran, B.; Hejazi, M.; Hasanzadeh, M.; Mosafer, J.; Mokhtarzadeh, A.; de la Guardia, M. Recent advances on aptamer-based biosensors to detection of platelet-derived growth factor. Biosens. Bioelectron. 2018, 113, 58–71. [Google Scholar]
  146. Kaur, N.; Bharti, A.; Batra, S.; Rana, S.; Rana, S.; Bhalla, A.; Prabhakar., N. An electrochemical aptasensor based on graphene doped chitosan nanocomposites for determination of Ochratoxin A. Microchem. J. 2019, 144, 102–109. [Google Scholar] [CrossRef]
  147. Wang, Y.H.; Ning, G.; Wu, Y.H.; Wu, S.; Zeng, B.Q.; Liu, G.Q.; He, X.X.; Wang, K.M. Facile combination of beta-cyclodextrin host-guest recognition with exonuclease-assistant signal amplification for sensitive electrochemical assay of ochratoxin A. Biosens. Bioelectron. 2019, 124, 82–88. [Google Scholar] [CrossRef] [PubMed]
  148. Wang, X.Y.; Shan, Y.Q.; Gong, M.; Jin, X.; Lv, L.R.; Jiang, M.; Xu, J. A novel electrochemical sensor for ochratoxin A based on the hairpin aptamer and double report DNA via multiple signal amplification strategy. Sens. Actuators B Chem. 2019, 281, 595–601. [Google Scholar] [CrossRef]
  149. Wu, K.F.; Ma, C.B.; Zhao, H.; Chen, M.J.; Deng, Z.Y. Sensitive aptamer-based fluorescene assay for ochratoxin A based on RNase H signal amplification. Food Chem. 2019, 277, 273–278. [Google Scholar] [CrossRef]
  150. Wu, J.; Chu, H.; Mei, Z.; Deng, Y.; Xue, F.; Zheng, L.; Chen, W. Ultrasensitive one-step rapid detection of ochratoxin A by the folding-based electrochemical aptasensor. Anal. Chim. Acta 2012, 753, 27–31. [Google Scholar] [CrossRef] [PubMed]
  151. Lv, L.; Jin, Y.; Kang, X.; Zhao, Y.; Cui, C.; Guo, Z. PVP-coated gold nanoparticles for the selective determination of ochratoxin A via quenching fluorescence of the free aptamer. Food Chem. 2018, 249, 45–50. [Google Scholar] [CrossRef] [PubMed]
  152. Lin, C.; Zheng, H.; Sun, M.; Guo, Y.; Luo, F.; Guo, L.; Qiu, B.; Lin, Z.; Chen, G. Highly sensitive colorimetric aptasensor for ochratoxin A detection based on enzyme-encapsulated liposome. Anal. Chim. Acta 2018, 1002, 90–96. [Google Scholar] [CrossRef]
  153. Bagheri, E.; Abnous, K.; Alibolandi, M.; Ramezani, M.; Taghdisi, S.M. Triple-helix molecular switch-based aptasensors and DNA sensors. Biosens. Bioelectron. 2018, 111, 1–9. [Google Scholar] [CrossRef]
  154. Zheng, J.; Li, J.S.; Jiang, Y.; Jin, J.Y.; Wang, K.M.; Yang, R.H.; Tan, W.H. Design of aptamer-based sensing platform using triple-helix molecular switch. Anal. Chem. 2011, 83, 6586–6592. [Google Scholar] [CrossRef]
  155. Xiong, E.H.; Li, Z.Z.; Zhang, X.H.; Zhou, J.W.; Yan, X.X.; Liu, Y.Q.; Chen, J.H. Triple-helix molecular switch electrochemical ratiometric biosensor for ultrasensitive detection of nucleic acids. Anal. Chem. 2017, 89, 8830–8835. [Google Scholar] [CrossRef]
  156. Wang, H.; Zhang, Y.H.; Ma, H.M.; Ren, X.; Wang, Y.G.; Zhang, Y.; Wei, Q. Electrochemical DNA probe for Hg2+ detection based on a triple-helix DNA and Multistage Signal Amplification Strategy. Biosens. Bioelectron. 2016, 86, 907–912. [Google Scholar] [CrossRef] [PubMed]
  157. Kohli, R.M.; Zhang, Y. TET enzymes, TDG and the dynamics of DNA demethylation. Nature 2013, 502, 472–479. [Google Scholar] [CrossRef]
  158. Huang, L.; Tian, S.; Zhao, W.; Liu, K.; Ma, X.; Guo, J. Aptamer-based lateral flow assay on-site biosensors. Biosens. Bioelectron. 2021, 186, 113279. [Google Scholar] [CrossRef] [PubMed]
  159. Ayela, C.; Roquet, F.; Valera, L.; Granier, C.; Nicu, L.; Pugnière, M. Antibody–antigenic peptide interactions monitored by SPR and QCM-D: A model for SPR detection of IA-2 autoantibodies in human serum. Biosens. Bioelectron. 2007, 22, 3113–3119. [Google Scholar] [CrossRef]
  160. Sinitsyna, V.V.; Vetcher, A.A. Nucleic Acid Aptamers in Nanotechnology. Biomedicines 2022, 10, 1079. [Google Scholar] [CrossRef] [PubMed]
  161. Wu, Y.; Midinov, B.; White, R.J. Electrochemical aptamer-based sensor for real-time monitoring of insulin. ACS Sens. 2019, 4, 498–503. [Google Scholar] [CrossRef] [PubMed]
  162. Angell, C.; Kai, M.; Xie, S.; Dong, X.; Chen, Y. Bioderived DNA nanomachines for potential uses in biosensing, diagnostics, and therapeutic applications. Adv. Healthc. Mater. 2018, 7, 1701189. [Google Scholar] [CrossRef]
  163. Zhao, J.; Zhai, Q. Recent advances in the development of ligands specifically targeting telomeric multimeric G-quadruplexes. Bioorganic Chem. 2020, 103, 104229. [Google Scholar] [CrossRef]
  164. Belleperche, M.; DeRosa, M.C. pH-control in aptamer-based diagnostics, therapeutics, and analytical applications. Pharmaceuticals 2018, 11, 80. [Google Scholar] [CrossRef]
  165. Nguyen, T.Q.N.; Lim, K.W.; Phan, A.T. Folding kinetics of G-quadruplexes: Duplex stem loops drive and accelerate G-quadruplex folding. J. Phys. Chem. B 2020, 124, 5122–5130. [Google Scholar] [CrossRef]
  166. Cecerska-Heryć, E.; Surowska, O.; Heryć, R.; Serwin, N.; Napiontek-Balińska, S.; Dołęgowska, B. Are antioxidant enzymes essential markers in the diagnosis and monitoring of cancer patients–a review. Clin. Biochem. 2021, 93, 1–8. [Google Scholar] [CrossRef] [PubMed]
  167. Wu, Y.; Li, G.; Zou, L.; Lei, S.; Yu, Q.; Ye, B. Highly active DNAzyme-peptide hybrid structure coupled porous palladium for high-performance electrochemical aptasensing platform. Sens. Actuators B Chem. 2018, 259, 372–379. [Google Scholar] [CrossRef]
  168. Luo, Z.; Sun, D.; Tong, Y.; Zhong, Y.; Chen, Z. DNA nanotetrahedron linked dual-aptamer based voltammetric aptasensor for cardiac troponin I using a magnetic metal-organic framework as a label. Microchim. Acta 2019, 186, 374. [Google Scholar] [CrossRef]
  169. Sun, D.P.; Lin, X.A.; Lu, J.; Wei, P.; Luo, Z.B.; Lu, X.E.; Chen, Z.G.; Zhang, L.Y. DNA nanotetrahedron-assisted electrochemical aptasensor for cardiac troponin I detection based on the co-catalysis of hybrid nanozyme, natural enzyme and artificial DNAzyme. Biosens. Bioelectron. 2019, 142, 111578. [Google Scholar] [CrossRef] [PubMed]
  170. Tintoré, M.; Gállego, I.; Manning, B.; Eritja, R.; Fàbrega, C. DNA Origami as a DNA Repair Nanosensor at the Single-Molecule Level. Angew. Chem. 2013, 125, 7901–7904. [Google Scholar] [CrossRef]
  171. Yuminova, A.V.; Smirnova, I.G.; Arutyunyan, A.M.; Kopylov, A.M.; Golovin, A.V.; Pavlova, G.V. The structure of G-quadruplex thrombine-binding DNA aptamer RA36. Mosc. Univ. Chem. Bull. 2015, 70, 43–46. [Google Scholar] [CrossRef]
  172. Alieva, R.R.; Zavyalova, E.G.; Tashlitsky, V.N.; Kopylov, A.M. Quantitative characterization of oligomeric state of G-quadruplex antithrombin aptamers by size exclusion HPLC. Mendeleev Commun. 2019, 29, 424–425. [Google Scholar] [CrossRef]
  173. Zein, S.S.; Vetcher, A.A.; Levene, S. PCR-Based Synthesis of Repetitive Single-Stranded DNA for Applications to Nanobiotechnology. Int. J. Nanosci. 2005, 4, 287–294. [Google Scholar] [CrossRef]
  174. Kretz, C.K.; Cuddy, K.R.; Stafford, A.R. HD1, a thrombin-and prothrombin-binding DNA aptamer, inhibits thrombin generation by attenuating prothrombin activation and thrombin feedback reactions. Thromb. Haemost. 2010, 103, 83–96. [Google Scholar] [CrossRef]
  175. Li, N.; Wang, J.; Ma, K.; Liang, L.; Mi, L.; Huang, W.; Yu, Z. The dynamics of forming a triplex in an artificial telomere inferred by DNA mechanics. Nucleic Acids Res. 2019, 47, e86. [Google Scholar] [CrossRef]
  176. Liu, C.P.; Wey, M.T.; Chang, C.C.; Kan, L.S. Direct observation of single molecule conformational change of tight-turn paperclip DNA triplex in solution. Appl. Biochem. Biotechnol. 2009, 159, 261–269. [Google Scholar] [CrossRef] [PubMed]
  177. Ma, J.B.; Jia, Q.; Xu, C.H. Asynchrony of base-pair breaking and nucleotide releasing of helicases in DNA unwinding. J. Phys. Chem. B 2018, 122, 5790–5796. [Google Scholar] [CrossRef] [PubMed]
  178. Lin, C.T.; Ha, T. Direct visualization of helicase dynamics using fluorescence localization and optical trapping. Methods Enzymol. 2017, 582, 121–136. [Google Scholar] [PubMed]
  179. Lee, J.; Crickard., J.B.; Reese, J.C.; Lee, T.H. Single-molecule Frest method to investigate the dynamics of transcription elongation through the nucleosome by RNA polymerase II. Methods 2019, 159, 51–58. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Hoogsteen-based triads are shown schematically in parallel and anti-parallel triplexes, as well as in triads based on the reverse Hoogsteen algorithm.
Figure 1. Hoogsteen-based triads are shown schematically in parallel and anti-parallel triplexes, as well as in triads based on the reverse Hoogsteen algorithm.
Polymers 15 00858 g001
Figure 2. Aptamer secondary structure (a) and aptamer with a marker (b).
Figure 2. Aptamer secondary structure (a) and aptamer with a marker (b).
Polymers 15 00858 g002
Figure 3. Nucleic acid base-pair binding and motifs that respond to pH. Illustration of the triplex binding motif and C + GC triplex non-canonical base pair binding in (a). Illustration of the i-motif, which is made up of C + C bonds, and non-canonical C + C base pair binding in (b). (c) Left: A non-canonical base pair binding is shown. Right: Illustration of A + G non-canonical base pair binding.
Figure 3. Nucleic acid base-pair binding and motifs that respond to pH. Illustration of the triplex binding motif and C + GC triplex non-canonical base pair binding in (a). Illustration of the i-motif, which is made up of C + C bonds, and non-canonical C + C base pair binding in (b). (c) Left: A non-canonical base pair binding is shown. Right: Illustration of A + G non-canonical base pair binding.
Polymers 15 00858 g003
Figure 4. A rescue rope approach for investigating an artificial telomere. (a) The DNA in telomeres adopts a triplex shape. The G-rich and C-rich strands are denoted by red and black, respectively. (b) Using a rescue rope to probe a telomere with a free end. Structures having a closed end can be repeatedly manipulated by conventional mechanical pulling experiments, such as a DNA hairpin (left). A construction with an open end cannot rotate mechanically in circles before relaxing (middle). A rescue-rope technique supported by dsDNA enables recurrent operations on a structure with a free end, such as a telomere (right). (c). Designing a synthetic telomere. The identical configuration of the artificial telomere in melting/reannealing rings is guaranteed by a random sequence at the blunt upstream end. In a duplex and a three overhang, the TTAGGG motif count might change. With n = 5 and m = 2.
Figure 4. A rescue rope approach for investigating an artificial telomere. (a) The DNA in telomeres adopts a triplex shape. The G-rich and C-rich strands are denoted by red and black, respectively. (b) Using a rescue rope to probe a telomere with a free end. Structures having a closed end can be repeatedly manipulated by conventional mechanical pulling experiments, such as a DNA hairpin (left). A construction with an open end cannot rotate mechanically in circles before relaxing (middle). A rescue-rope technique supported by dsDNA enables recurrent operations on a structure with a free end, such as a telomere (right). (c). Designing a synthetic telomere. The identical configuration of the artificial telomere in melting/reannealing rings is guaranteed by a random sequence at the blunt upstream end. In a duplex and a three overhang, the TTAGGG motif count might change. With n = 5 and m = 2.
Polymers 15 00858 g004
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Bekkouche, I.; Shishonin, A.Y.; Vetcher, A.A. Recent Development in Biomedical Applications of Oligonucleotides with Triplex-Forming Ability. Polymers 2023, 15, 858. https://doi.org/10.3390/polym15040858

AMA Style

Bekkouche I, Shishonin AY, Vetcher AA. Recent Development in Biomedical Applications of Oligonucleotides with Triplex-Forming Ability. Polymers. 2023; 15(4):858. https://doi.org/10.3390/polym15040858

Chicago/Turabian Style

Bekkouche, Incherah, Alexander Y. Shishonin, and Alexandre A. Vetcher. 2023. "Recent Development in Biomedical Applications of Oligonucleotides with Triplex-Forming Ability" Polymers 15, no. 4: 858. https://doi.org/10.3390/polym15040858

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop