I-motif DNA

Last updated

i-motif DNA, short for intercalated-motif DNA, are cytosine-rich four-stranded quadruplex DNA structures, similar to the G-quadruplex structures that are formed in guanine-rich regions of DNA.

Contents

History

This structure was first discovered in 1993 by Maurice Guéron at École Polytechnique in Palaiseau, France. It was found when two antiparallel doubled stranded DNA complexes with cytosine-protonated cytosine (C·C*) base pairs became associated with one another. This formed a complex 4- stranded DNA complex. [1] The structure was originally found only in vitro , usually at a slightly acidic pH, but was recently discovered in the nuclei of human cells. [2] A new antibody fragment was created, and was found to have highly specific binding affinity for I-motif complexes, but did not bind to other DNA structures, making it optimal for identifying i-motif structures in cells. [3]

During their media release in April 2018, Dr. Mahdi Zeraati and colleagues mentioned that these complexes are constantly forming and dissociating due to their constantly changing temperatures, which could play a role in its function in regulation of gene expression and cell reproduction. Although the exact function of these structures is unknown, the transient nature of these molecules gives insight regarding the biological function of these molecules. Found primarily in the G1 phase of the cell cycle and in promoter regions, i-motif complexes could potentially affect which gene sequences are read and could play a role in determining which genes are switched on or off. [4] Other experimentation is in progress to determine the role of i-motif DNA in nanotechnology using i-motifs as biosensors and nanomachines, [5] and it has even been seen to play a role in the advancement of cancer therapy.

Structural overview

RNA i-motif structure showing cytosine bases in the quadruplex. PDB: 1I9K RNA i-motif.gif
RNA i-motif structure showing cytosine bases in the quadruplex. PDB: 1I9K
C*C base pairing found in i-motif structures. Base-pairing energy=169.7 kJ/mol. C-C pairing.svg
C·C base pairing found in i-motif structures. Base-pairing energy=169.7 kJ/mol.

Similar to G-quadruplex DNA structures with intercalated guanine residues, i-motifs consists of antiparallel tracts of oligodeoxynucleotides strands that contain mostly cytosine residues. The interactions between these molecules occur by the hemi protonation of cytosine residues and non-Watson Crick base pairing, more specifically Hoogsteen base pairing. There are two main intercalated topologies that i-motifs can be classified in: 3'-E, when the outmost C:C+ base pair is at the 3'-end, and 5'-E, where the outermost C:C+ base pair is at the 5'-end. When comparing the two topologies, the 3'-E topology is more stable due to increased sugar-sugar contacts. [9] This occurs due to the difference in Van der Waals energy contribution between the two topologies. The interactions of the sugar-sugar contacts along the narrow grooves allows for optimal backbone twisting, which ultimately contributes to formation of stacking bases and the stability of the molecule. [10] However, the overall stability of i-motif structures is dependent on the number of cytosine residues that are interacting with each other. This means that as more cytosine residues interact through hydrogen bonding, the more stable the molecule will be. [11] Other factors that affect the stability of the molecules include temperature, salt concentration and pH of the environment. [12]

While many i-motif complexes are most stable at a slightly acidic pH (between 4.2 and 5.2), [11] some i-motifs have been found to form at neutral pH, when a free proton is used by the nucleic acids during the folding process. These particular i-motif complexes are found under particular conditions, including low temperature (4 °C), molecular crowding, negative super helicity, and the introduction of silver(I) cations. Maintaining a negative super helicity is crucial for the stabilization of i-motifs at a neutral pH. [13]

i-motif structures have also been found to form under biological conditions. These structures have been discovered in many different locations of the cell including the nuclei, [2] the cytoplasm, and in telomeres and promoter sights. [14] It can also be found in cell processes such as the G1 phase of the cell cycle.

Stability of i-motif DNA

As a nucleic acid structure, i-motif DNA stability is dependent on the nature of the sequence, temperature, and ionic strength. The structural stability of i-motif DNA is mainly reliant on the fact that there is minimal overlap between the six-membered aromatic pyrimidine bases due to the consecutive base pairs' intercalative geometry. Exocyclic carbonyls and amino groups stacked in an antiparallel formation are essential to C:C+ base pairs' stability due to the lack of compensation for the electrostatic repulsion between their charged amino groups. [15] Other factors, including sugar and phosphate backbone interactions, C-tract length, capping and connecting loops interaction, ionic interactions, molecular crowding, and super helicity, all affect the stability of i-motif DNA. [16]

C:C+ base pairs

The C:C+ base pairs contribute most to i-motif stability due to three hydrogen bonds. This stability is exhibited by the base-pairing energy (BPE) of i-motif being 169.7 kJ/mol, which is relatively high compared to neutral C·C and canonical Watson-Crick G·C, which have BPEs of 68.0 kJ/mol and 96.6 kJ/mol, respectively. [17] The most stable central hydrogen bond in the C:C+ base pair (N3··H··N3) has been denoted as having double-well potential due to the proton's capability of oscillating between the two nitrogen base wells [18] with a proton transfer rate found to be 8 × 104 s-1. [19]

The results of two studies by Waller's group and Mir et al. emphasized the importance of electrostatic interactions contributing to the stability of the C:C+ base pair. [20] Waller's group wished to determine the effect of 2 - deoxyriboguanylurea (GuaUre-dR), a chemotherapeutic agent, on i-motif DNA formation in human telomeres. Waller's group found that the addition of GuaUre-dR led to a decrease in pH when compared to i-motifs without it. [21] Mir et al. showed that the addition of pseudoiso-deoxycytidine (psC) increased the stability of head-to-head and head-to-tail dimeric i-motif structures when a neutrally charged psC:C was found at the end of the C:C stack. [22] Both studies ultimately found that the existence of positive charges in the core of these structures contributed most to the stability of the C:C+ base pair. [20]

Alterations to the environment conditions of C:C+ were studied by Watkins et al. to observe changes in overall stability. [23] Chemical modifications to the C:C+ base pair in which halogenated analogs (5-fluoro, 5-bromo, and 5-iodo) took the place of cytosine increased i-motif DNA stability in acidic environments. [24] This study initiated an investigation into the methylation of cytosine and its effect on pH. Methylation of cytosine at position 5 increased pH of mid-transition and Tm of i-motifs. On the other hand, hydroxymethylation leads to a decrease in the pH of mid-transition and Tm. [25]

Phosphate backbone and sugar interactions

The minor groove of i-motif DNA consists of a phosphate backbone in which two negatively charged sides repel each other, requiring balance to stabilize the overall structure. [20] Hydrogen bonding and Van der Waals interactions between sugars of the minor groove from the sequence d(CCCC) of tetrameric i-motif DNA stabilizes the narrow grooves of the i-motif structure. The stability of 3'E and 5'E topologies from the sequence d(CCCC) was observed through molecular dynamic simulations to determine the effect of repulsion between the phosphate backbones. [15] The stability observed in the simulations is derived from the supportive sugar interactions, so much so that the stability of any i-motif is dependent on the balance between sugar interactions and connective loop activity. This is due to the low free energy of the hydrogen bond (CHO) in the i-motif structure with a value of 2.6 kJ/mol. [19]

Modifications to the phosphate backbone have been seen in research in which an alternative to the phosphate backbone was studied. Oligodeoxycytidine phosphorothioates can form intramolecular and intermolecular i-motifs. [26] Mergny and Lacroix determined that the addition of a bulky methyl group had a destabilizing effect on the i-motif formation when they compared phosphorothioate, the natural phosphodiester, methylphosphonate, and peptide linkages and determined that only phosphodiester and phosphorothioate oligodeoxynucleotides were capable of forming stable i-motifs. [27]

Environmental conditions

Investigations into the formation of i-motifs at physiological pH, rather than acidic pH, include simulations of molecular crowding, superhelicitification, and cationic conditions. Through the utilization of polyethylene glycols with a high molecular weight, conditions of congested molecular and nuclear environments were induced. [28] An increase in pKa of cytosine N3 [23] showed that these conditions favored quadruplex and i-motif over double [29] and single-stranded DNA when i-motif formation and protonation was Bacollainduced at neutral pH. [30] Negative super helicity assists in the formation of i-motifs under physiological conditions. [31] The formation of both G-quadruplex and i-motifs occurred at neutral pH when G-quadruplex and i-motif forming sequences of the c-MYC oncogene promoter were placed into a supercoiled plasmid, inducing super helicity of both structures. The conditions of super helicity mitigation were inspired by the fact that i-motif destabilizes double-stranded structures. [28] [31] This result reflects the transcription process in which supercoiled DNA is unwound into single-stranded structures, which causes negative super helicity. [32] The stability of i-motif DNA can be influenced by increasing ionic concentration. [33] The addition of Na has shown to increase the destabilization of the i-motif structure from the c-jun proto-oncogene at pH 4.8. A decrease in stability of i-motif corresponded with an increase in ionic concentration in a study of i-motif DNA from n-MYC. [34] However, no significant differences in stability occurred with the addition of 5 mM Mg+, Ca+, Zn+, Li+ or K+ cations in the presence of 100 mM NaCl at pH 6.4. [27]

Base modifications

Further investigation is required to determine the absolute effect on i-motif stability when bases are modified, but studies have indicated that there is potential for the modification of bases corresponding to the stability of i-motifs. [35] Two examples include replacing cytosine with 5-methylcytosine [35] and replacing thymine with 5-propynyl uracil, [23] both increasing the stability of the i-motif structure. The modification of bases may be helpful in determining the pH/temperature-dependent folding patterns of i-motifs. [28]

Formation

Intercalated motif (i-motif) DNA is formed in the nuclei of cells via a stack of intercalating hemi-protonated C-neutral C base pairs, which are optimized at a slightly negative pH. In vitro, i-motifs have been characterized with indications that the DNA is derived from telomeres. Using a variety of biophysical techniques, i-motif DNA has been characterized to be derived from centromeres and promoter regions of proto-oncogenes. An analysis of the biophysical results shows the structures' overall stability depends on the number of cytosines in the i-motif core and the length and composition of loops in the formation of both intramolecular and intermolecular structures. [36]

Although it has been largely established that C-rich sequences can form i-motif structures in vitro, there is still significant debate regarding the in vivo existence of four-stranded i-motif DNA structure in the human genome. It has been confirmed that motif DNA in vivo can be formed at physiological pH under certain molecular crowding conditions and negative super helicity induced during transcription. [36] Recent studies have shown that the formation of i-motif DNA by specific genomic sequences can occur at neutral pH. Numerous studies have demonstrated that i-motif DNA affects replication and transcription in DNA processing after its formation. [37]

G-quadruplex formation

I-motif DNA forms from any complementary strand of G-quadruplex forming sequence. G-quadruplexes are helically shaped and found in nucleic acids that are rich in guanine. These secondary structures possess guanine tetrads formed into one of three types of strands: one, two, or four. With prior knowledge of G-quadruplex forming sequences being susceptible to i-motif DNA formation, Waller's group used the algorithm Quadparser to determine the amount of i-motif forming sequences in the human genome. [21] The query consisted of four C-tracts of five cytosines distinguished by the number of nucleotides that could range from 1–19. Across the human genome, 5,125 sequences have potential i-motif formation capabilities with 12.4% (637) of the total resulting sequences found in the promoter regions of genes. Based upon the ontology codes corresponding to the promoter regions, i-motif formation is concentrated sequence-specific DNA binding, DNA templated transcription, skeletal system development and RNA polymerase II positive regulation of transcription. [38]

Interacting agents and ligands

Interacting agents derived from G4 ligands

Tetra-(N-methyl-4-pyridyl)porphyrin (TMPyP4)

The first study to determine a ligand binding to i-motif DNA was by Hurley and colleagues in 2000. They researched the interaction between Tetra-(N-methyl-4-pyridyl)porphyrin (TMPyP4) and tetramolecular i-motif DNA isolated from a human telomeric sequence. The study utilized an electrophoretic mobility shift assay (EMSA) by notably not changing the DNA melting temperature. This ligand interacts with G4 on the i-motif sequence to deregulate c-myc expression and inhibit telomerase. [39] [40] Two molecules of TMPyP4 coordinate with i-motif DNA on both the top and bottom of its structure as determined by NMR experiments. [41]

Phenanthroline and acridine derivatives

These cores characterize phenanthroline derivatives due to their G4 binding and telomerase inhibiting activity. [42] This activity leads to an overall increase in the Tm of the i-motif. Phenanthroline derivatives bind to the C:C base pair, leading to a decrease in the binding constant lower than that of a normal G-quadruplex. [43] Acridine derivatives are also G4 ligands and through fluorescence resonance energy transfer (FRET) melting assays, diethylenetriamine (BisA) was determined to increase the melting temperature of both i-motif and G4, while monomeric acridine (MonoA) had no such effect. [44]

Macrocyclic tetraconazoles, L2H2-4OTD

Inspired by telomestatin, a natural potent telomerase inhibitor, macrocyclic poly-oxazoles were synthesized. Macrocyclic poly-oxazole compounds possess the same binding mode as telomestatin when interacting with G4 in a pi-pi stack formation. [43] Smaller macrocycles, penta- (L2H2-5OTD) and tetra-oxazoles (L2H2-4OTD) were developed with amine R-groups to observe stability and binding site locations on i-motif. Reducing the size of the ligands reduced its stabilizing effect on G4-forming sequences. L2H2-4OTD molecules bind cooperatively to Loop 1 and 2 on the telomeres of the i-motif DNA sequence which induces deformities on C3-C15, C2-C14 and C8-C20 base pairs while maintaining the structure of i-motif. [45]

Mitoxantrone, tilorone and tobramycin

Mitoxantrone stabilizes the i-motif and G4 and aids in their formation under neutral conditions with a preference in binding to i-motif over double-stranded DNA. Tilorone and Tobramycin are i-motif binding ligands discovered via thiazole orange fluorescence intensity displacement (FID) assay. [46]

Carboxylic acid-modified single-walled carbon nanotubes (SWCNTs) and graphene quantum dots (GQDs)

SWCNTs stabilize i-motif DNA by attracting water molecules from the structure. GQDs intercalate with DNA to aid in the formation of i-motif DNA by end-stacking loop regions. This process allows GQDs to stabilize i-motifs by minimizing solvent-access. [47]

Ligands used for biological functions

There are several ligands for i-motif that are used for biological functions. These include IMC-48, IMC-76, Nitidine, NSC309874, acridone derivative, and PBP1. IMC-48 stabilizes the bcl-2 structure of i-motif by upregulating bcl-2 gene expression. IMC-76 stabilizes the bcl-2 hairpin structure by downregulating the bcl-2 gene expression. Nitidine destabilizes the hairpin on hybride i-motif/hairpin structure and has no significant interactions with complementary G4. Nitidine downregulates the k-ras gene expression by showing selectivity toward the k-ras structure. [43] NSC309874 stabilizes the PDGFR-b i-motif structure with no significant interaction with complementary G4 to downregulate the PDGFR-b gene expression. Acridone derivative stabilizes the c-myc i-motif structure with no significant interaction with G4 in order to downregulate the c-myc gene expression. PBP1 stabilizes the bcl-2 i-motif structure and promotes its formation in neutral pHs to upregulate the bcl-2 gene expression. [43]

Ligands used as fluorescent probes

The ligands for i-motif used as fluorescent probes include Thiazole orange, 2,2'-diethyl-9-methylselenacarbocyanine bromide (DMSB), crystal violet, berberine neutral red, thioflavin T, and perylene tetracarboxylic acid diimide derivative (PTCDI), originally seen as G4 probes. [43]

Biological function

Large tracts of G/C-rich DNA exist in regulatory regions of genes and in terminal regions of chromosomes and telomeres. These expansions of C-rich regions are present in a wide variety of organisms, and suggests that i-motifs could exist in vivo. It is postulated that i-motifs play roles in gene regulation and expression, telomerase inhibition, and DNA replication and repair. Although there are limited examples of i-motif formation in living cells, there are conditions that can be induced to create i-motifs. Coupling the examples of i-motif structures in cells with these experiments give avenues for further investigation.

Gene regulation and expression

Promoter regions of certain genes are C-rich. It is found in more than 40% of all human genes, especially in oncogenes, skeletal system development regions and areas of DNA processes, which strengthens the suggestion that i-motifs function as gene transcription regulators. [48] [49] [50] The promoter region of a transcription factor gene in silkworms, called BmPOUM2, was seen to form i-motif structures. The BmPOUM2 gene regulates another gene that affects wing disc cuticle formation during metamorphosis, and was seen to be positively regulated by i-motif formation. [12] This is an example of an important biological function in an organism being influenced by i-motif structure. Human telomeric DNA (hTelo) was also observed to form i-motif structures, also in vivo. This was confirmed by fluorescent marking with iMab. [51] These i-motif hTelos were found in regulatory regions of the human genome during the late G1 phase which indicates that i-motifs are involved in regulating genes important to development in the human genome. Even though more studies need to be conducted to validate these findings and provide specific insight to what genes are regulated, this study was important to opening the conversation of i-motif roles, and possible applications, in humans.

A similar role i-motifs can play is aiding the binding of transcription factors during gene transcription. One way this can occur is through temporary DNA unwinding into i-motif and g-quadruplex structures at promoter regions (like BCL2 ), which allows transcription of single strands of DNA.

Telomerase inhibition

The formation of g-quadruplexes and i-motifs at ends of chromosomes can lead to telomerase inhibition. The formation of i-motif structures at the ends of chromosomes inhibits telomerase from binding, which interferes with telomere lengthening. These formations result in the uncapping of telomeres, which exposes telomeres and triggers DNA damage response, ceasing rapid tumor growth. [52] Because i-motif structures are not specifically stable, the discovery of a ligand that selectively binds to i-motifs and stabilizes them was important to telomerase inhibition. Once bound with CSWNT, i-motifs were found to interfere with telomerase functions in vitro and in vivo in cancer cells, which was assessed by a TRAP assay. [53]

Ligand interaction

The binding of ligands can increase and modify i-motif functions. The first known selective ligand to bind with i-motif DNA is Carboxyl-modified single-walled carbon nanotubes (CSWNTs). These ligands binds to the 5' end major groove of DNA to induce i-motifs. The binding of CSWNT to i-motifs increases thermal stability at both acidic and biological pH by a significant amount. In this way, the CSWNT supports the formation of i-motif DNA over the Watson-Crick base pairing at pH 8.0. [54] Furthermore, many proteins and ligands fundamental to gene expression recognize C-rich oligonucleotides, such as Poly-C-binding protein (PCBP) and heterogeneous nuclear ribonucleoprotein K (HNRPK).

In the presence of C-rich single stranded oligonucleotides, PCBPs have the ability to play a variety of roles such as stabilizing mRNA and translational repression or enhancement depending on the C-rich single strand oligonucleotide that is being targeted. [55] Like PCBPs, the transcription factor heterogeneous nuclear ribonucleoprotein K (HNPRK) has the ability to selectively modulate the promoter regions of proteins such as KRAS and VGEF, in the presence of C-rich sequences such as i-motifs. [56] [57] C-rich sequences such as i-motifs exist throughout the human genome, acting as targets for a variety of proteins that can regulate gene expression in multiple ways and locations.

DNA replication and repair

There was also evidence that i-motifs could interfere with DNA repair and replication. An experiment was performed where sequences that encouraged i-motif formation in a DNA strand that was being replicated by DNA polymerase. The focus of this experiment was the visualization of i-motifs in silkworms, and it was noted DNA polymerase was stalled, which implied i-motifs can impede DNA replication and repair. [58] The stalling effect of i-motif sequences was higher than hairpin DNA, although it is thermodynamically similar. This is due to the topology of i-motif DNA. The i-motif is unique when compared to other DNA because it is intercalated, which resists unwinding. This is what stalls DNA polymerase. It may also be attributed to steric hindrance, which would not allow DNA polymerase to bind. [16]

Other considerations

The formation of G-quadruplexes can lead its complementary DNA strand to be C-rich, which can form an i-motif, but this is not always the case. This is evident due to the majority of i-motif formation occurring in the G1 phase, while G-quadruplex formation is primarily noted in the S phase.

Applications

Applications of i-motifs are centered around biomedical topics, including bio-sensing, drug delivery systems, and molecular switches. Many of the current applications for i-motif DNA are due to its sensitivity to pH. The development of pH sensitive systems, which includes ligand binding, is a field of great interest to medicine, especially in the treatment and detection of cancer.

Bio-sensors

The conformational change from B DNA to i-motif under acidic conditions makes it useful as a colormetric sensor for glucose levels. A glucose detection system, Poly(24C)-MB, was created to detect a drop in pH levels in organisms, which occurs when glucose is oxidized. The dye of the Poly(24C)-MB system, methylene blue (MB), cannot bind when i-motifs are induced, giving rise to a color change that is easily visible. This system is simple, cost-effective, and precise due to i-motif conformation. [59]

Drug delivery systems

Gold nanoparticles/i-motif conjugated systems have been developed as a pH-induced drug delivery system. A study done using DNA conjugated gold nanoparticles (DNA-GNP) created a delivery molecule with stretches of C-rich single-stranded DNA that form i-motifs in cancer cell due to their acidic endosomes. When the DNA-GNP molecule enters a normal cell, no change takes place, but when the DNA-GNP enters a cancer cell, it induces i-motif conformation, which triggers doxorubicin (DOX), an effective cancer drug against leukemia and Hodgkin's lymphoma, to be released into the cell. [60] This method not only acts as an efficient drug delivery system, but can also be modified to detect cancer cells by including a dye or fluorescent, much like the colormetric sensor.

Theranostics

Due to i-motif formation in acidic conditions and cancer cells having acidic endosomes, cancer therapy and theranostic applications have been investigated. In a study by Takahashi et al., it was found that by using carboxyl-modified single-walled carbon nanotubes (C-SWNTs), telomerase activity could be inhibited, which could potentially lead to apoptosis of cancer cells. This is due to the use of fisetin, a plant flavanol, changing the conformation of i-motif structures into hairpin structures, which is a promising result in the investigation of various cancer drug therapies. [50] The binding of fisetin to an i-motif in the promoter region of vascular endothelial growth factor (VEGS), which is a signal protein for angiogenesis, induced a conformational change to a hairpin structure that inhibited it from functioning. The fisetin was suggested to bind to the loop of the i-motif, and when bound, it fluoresced. The fluorescent nature of this bond can be used as a diagnostic for this i-motif formation, and the formation of i-motifs that contain guanine residues. Overall, the study provided new information on how i-motifs can be used as a method for cancer treatment and detection. [61]

Molecular switches

A study at the University of Bonn explained how i-motifs can be utilized as molecular switches. The study synthesized a ring of DNA with certain regions of C-rich DNA. At a pH of 5, these regions contracted to form i-motifs, tightening the ring in a fashion similar to closing a trash bag. At a pH of 8 the i-motif regions collapsed back into their linear forms, relaxing the ring. DNA rings that can tighten and loosen based on pH can be used to build more complex structures of interlocking DNA like catenanes and rotaxanes. [62] This study emphasized that the manipulation of i-motif structure can unlock new possibilities in nanomechanics. Another study showed CSWNTs could induce i-motif formation in human telomeric DNA and modify it by attaching a redox active methylene blue group to the 3' end and an electrode to the 5' end. In the i-motif conformation this modified DNA strand produces a large increase in Faradaic current, which only reacts to CSWNTs, allowing researchers to detect a specific type of carbon nanotube with a direct detection limit of 0.2 ppm.

See also

Related Research Articles

<span class="mw-page-title-main">Transcription factor</span> Protein that regulates the rate of DNA transcription

In molecular biology, a transcription factor (TF) is a protein that controls the rate of transcription of genetic information from DNA to messenger RNA, by binding to a specific DNA sequence. The function of TFs is to regulate—turn on and off—genes in order to make sure that they are expressed in the desired cells at the right time and in the right amount throughout the life of the cell and the organism. Groups of TFs function in a coordinated fashion to direct cell division, cell growth, and cell death throughout life; cell migration and organization during embryonic development; and intermittently in response to signals from outside the cell, such as a hormone. There are approximately 1600 TFs in the human genome. Transcription factors are members of the proteome as well as regulome.

<span class="mw-page-title-main">Telomere</span> Region of repetitive nucleotide sequences on chromosomes

A telomere is a region of repetitive nucleotide sequences associated with specialized proteins at the ends of linear chromosomes. Telomeres are a widespread genetic feature most commonly found in eukaryotes. In most, if not all species possessing them, they protect the terminal regions of chromosomal DNA from progressive degradation and ensure the integrity of linear chromosomes by preventing DNA repair systems from mistaking the very ends of the DNA strand for a double-strand break.

<span class="mw-page-title-main">Transcription (biology)</span> Process of copying a segment of DNA into RNA

Transcription is the process of copying a segment of DNA into RNA. The segments of DNA transcribed into RNA molecules that can encode proteins produce messenger RNA (mRNA). Other segments of DNA are transcribed into RNA molecules called non-coding RNAs (ncRNAs).

A regulatory sequence is a segment of a nucleic acid molecule which is capable of increasing or decreasing the expression of specific genes within an organism. Regulation of gene expression is an essential feature of all living organisms and viruses.

<span class="mw-page-title-main">DNA methyltransferase</span> Class of enzymes

In biochemistry, the DNA methyltransferase family of enzymes catalyze the transfer of a methyl group to DNA. DNA methylation serves a wide variety of biological functions. All the known DNA methyltransferases use S-adenosyl methionine (SAM) as the methyl donor.

<span class="mw-page-title-main">Telomerase</span> Telomere-restoring protein active in the most rapidly dividing cells

Telomerase, also called terminal transferase, is a ribonucleoprotein that adds a species-dependent telomere repeat sequence to the 3' end of telomeres. A telomere is a region of repetitive sequences at each end of the chromosomes of most eukaryotes. Telomeres protect the end of the chromosome from DNA damage or from fusion with neighbouring chromosomes. The fruit fly Drosophila melanogaster lacks telomerase, but instead uses retrotransposons to maintain telomeres.

<span class="mw-page-title-main">Triple-stranded DNA</span> DNA structure

Triple-stranded DNA is a DNA structure in which three oligonucleotides wind around each other and form a triple helix. In triple-stranded DNA, the third strand binds to a B-form DNA double helix by forming Hoogsteen base pairs or reversed Hoogsteen hydrogen bonds.

<span class="mw-page-title-main">Leucine zipper</span> DNA-binding structural motif

A leucine zipper is a common three-dimensional structural motif in proteins. They were first described by Landschulz and collaborators in 1988 when they found that an enhancer binding protein had a very characteristic 30-amino acid segment and the display of these amino acid sequences on an idealized alpha helix revealed a periodic repetition of leucine residues at every seventh position over a distance covering eight helical turns. The polypeptide segments containing these periodic arrays of leucine residues were proposed to exist in an alpha-helical conformation and the leucine side chains from one alpha helix interdigitate with those from the alpha helix of a second polypeptide, facilitating dimerization.

<span class="mw-page-title-main">Dyskeratosis congenita</span> Medical condition

Dyskeratosis congenita (DKC), also known as Zinsser-Engman-Cole syndrome, is a rare progressive congenital disorder with a highly variable phenotype. The entity was classically defined by the triad of abnormal skin pigmentation, nail dystrophy, and leukoplakia of the oral mucosa, and MDS/AML, but these components do not always occur. DKC is characterized by short telomeres. The disease initially can affect the skin, but a major consequence is progressive bone marrow failure which occurs in over 80%, causing early mortality.

<span class="mw-page-title-main">G-quadruplex</span> Structure in molecular biology

In molecular biology, G-quadruplex secondary structures (G4) are formed in nucleic acids by sequences that are rich in guanine. They are helical in shape and contain guanine tetrads that can form from one, two or four strands. The unimolecular forms often occur naturally near the ends of the chromosomes, better known as the telomeric regions, and in transcriptional regulatory regions of multiple genes, both in microbes and across vertebrates including oncogenes in humans. Four guanine bases can associate through Hoogsteen hydrogen bonding to form a square planar structure called a guanine tetrad, and two or more guanine tetrads can stack on top of each other to form a G-quadruplex.

<span class="mw-page-title-main">Telomerase reverse transcriptase</span> Catalytic subunit of the enzyme telomerase

Telomerase reverse transcriptase is a catalytic subunit of the enzyme telomerase, which, together with the telomerase RNA component (TERC), comprises the most important unit of the telomerase complex.

<span class="mw-page-title-main">Telomeric repeat-binding factor 2</span> Protein

Telomeric repeat-binding factor 2 is a protein that is present at telomeres throughout the cell cycle. It is also known as TERF2, TRF2, and TRBF2, and is encoded in humans by the TERF2 gene. It is a component of the shelterin nucleoprotein complex and a second negative regulator of telomere length, playing a key role in the protective activity of telomeres. It was first reported in 1997 in the lab of Titia de Lange, where a DNA sequence similar, but not identical, to TERF1 was discovered, with respect to the Myb-domain. De Lange isolated the new Myb-containing protein sequence and called it TERF2.

<span class="mw-page-title-main">Telomeric repeat-binding factor 1</span> Protein-coding gene in humans

Telomeric repeat-binding factor 1 is a protein that in humans is encoded by the TERF1 gene.

<span class="mw-page-title-main">Telomerase RNA component</span> NcRNA found in eukaryotes

Telomerase RNA component, also known as TR, TER or TERC, is an ncRNA found in eukaryotes that is a component of telomerase, the enzyme used to extend telomeres. TERC serves as a template for telomere replication by telomerase. Telomerase RNAs differ greatly in sequence and structure between vertebrates, ciliates and yeasts, but they share a 5' pseudoknot structure close to the template sequence. The vertebrate telomerase RNAs have a 3' H/ACA snoRNA-like domain.

<span class="mw-page-title-main">PINX1</span> Protein-coding gene in the species Homo sapiens

PIN2/TERF1-interacting telomerase inhibitor 1, also known as PINX1, is a human gene. PINX1 is also known as PIN2 interacting protein 1. PINX1 is a telomerase inhibitor and a possible tumor suppressor.

<span class="mw-page-title-main">Nucleic acid tertiary structure</span> Three-dimensional shape of a nucleic acid polymer

Nucleic acid tertiary structure is the three-dimensional shape of a nucleic acid polymer. RNA and DNA molecules are capable of diverse functions ranging from molecular recognition to catalysis. Such functions require a precise three-dimensional structure. While such structures are diverse and seemingly complex, they are composed of recurring, easily recognizable tertiary structural motifs that serve as molecular building blocks. Some of the most common motifs for RNA and DNA tertiary structure are described below, but this information is based on a limited number of solved structures. Many more tertiary structural motifs will be revealed as new RNA and DNA molecules are structurally characterized.

Telomere-binding proteins function to bind telomeric DNA in various species. In particular, telomere-binding protein refers to TTAGGG repeat binding factor-1 (TERF1) and TTAGGG repeat binding factor-2 (TERF2). Telomere sequences in humans are composed of TTAGGG sequences which provide protection and replication of chromosome ends to prevent degradation. Telomere-binding proteins can generate a T-loop to protect chromosome ends. TRFs are double-stranded proteins which are known to induce bending, looping, and pairing of DNA which aids in the formation of T-loops. They directly bind to TTAGGG repeat sequence in the DNA. There are also subtelomeric regions present for regulation. However, in humans, there are six subunits forming a complex known as shelterin.

<span class="mw-page-title-main">Telomeric repeat–containing RNA</span> Long non-coding RNA transcribed from telomeres

Telomeric repeat–containing RNA (TERRA) is a long non-coding RNA transcribed from telomeres - repetitive nucleotide regions found on the ends of chromosomes that function to protect DNA from deterioration or fusion with neighboring chromosomes. TERRA has been shown to be ubiquitously expressed in almost all cell types containing linear chromosomes - including humans, mice, and yeasts. While the exact function of TERRA is still an active area of research, it is generally believed to play a role in regulating telomerase activity as well as maintaining the heterochromatic state at the ends of chromosomes. TERRA interaction with other associated telomeric proteins has also been shown to help regulate telomere integrity in a length-dependent manner.

<span class="mw-page-title-main">Polypurine reverse-Hoogsteen hairpin</span>

Polypurine reverse-Hoogsteen hairpins (PPRHs) are non-modified oligonucleotides containing two polypurine domains, in a mirror repeat fashion, linked by a pentathymidine stretch forming double-stranded DNA stem-loop molecules. The two polypurine domains interact by intramolecular reverse-Hoogsteen bonds allowing the formation of this specific hairpin structure.

<span class="mw-page-title-main">Guanine tetrad</span> Structure in molecular biology

In molecular biology, a guanine tetrad is a structure composed of four guanine bases in a square planar array. They most prominently contribute to the structure of G-quadruplexes, where their hydrogen bonding stabilizes the structure. Usually, there are at least two guanine tetrads in a G-quadruplex, and they often feature Hoogsteen-style hydrogen bonding.

References

  1. Gehring, Kalle; Leroy, Jean-Louis; Guéron, Maurice (June 1993). "A tetrameric DNA structure with protonated cytosine-cytosine base pairs". Nature. 363 (6429): 561–565. Bibcode:1993Natur.363..561G. doi:10.1038/363561a0. ISSN   0028-0836. PMID   8389423. S2CID   388606.
  2. 1 2 Zeraati, Mahdi; Langley, David B.; Schofield, Peter; Moye, Aaron L.; Rouet, Romain; Hughes, William E.; Bryan, Tracy M.; Dinger, Marcel E.; Christ, Daniel (June 2018). "I-motif DNA structures are formed in the nuclei of human cells". Nature Chemistry. 10 (6): 631–637. Bibcode:2018NatCh..10..631Z. doi:10.1038/s41557-018-0046-3. ISSN   1755-4330. PMID   29686376. S2CID   13816298.
  3. Chaudhary, Swati; Kaushik, Mahima; Ahmed, Saami; Kukreti, Shrikant (7 December 2020). "Exploring potential of i-motif DNA formed in the promoter region of GRIN1 gene for nanotechnological applications". Results in Chemistry. 2: 100086. doi: 10.1016/j.rechem.2020.100086 . ISSN   2211-7156.
  4. SCIMEX (23 April 2018). "Found: a new form of DNA in our cells". Scimex. Archived from the original on 1 July 2018. Retrieved 10 December 2020.
  5. Benabou, S.; Aviñó, A.; Eritja, R.; González, C.; Gargallo, R. (2014). "Fundamental aspects of the nucleic acid i-motif structures". RSC Advances. 4 (51): 26956–26980. Bibcode:2014RSCAd...426956B. doi:10.1039/C4RA02129K. hdl: 10261/126489 . ISSN   2046-2069.
  6. Snoussi, K.; Nonin-Lecomte, S.; Lerou, J.L. (30 May 2001). "THE RNA I-MOTIF". Journal of Molecular Biology. 309 (1): 139–53. doi:10.2210/pdb1i9k/pdb. PMID   11491284.
  7. Gurung, Sarah P.; Schwarz, Christine; Hall, James P.; Cardin, Christine J.; Brazier, John A. (2015). "The importance of loop length on the stability of i-motif structures". Chemical Communications. 51 (26): 5630–5632. doi:10.1039/c4cc07279k. ISSN   1359-7345. PMC   4384421 . PMID   25686374.
  8. Benabou, S.; Aviñó, A.; Eritja, R.; González, C.; Gargallo, R. (2014). "Fundamental aspects of the nucleic acid i-motif structures" (PDF). RSC Advances. 4 (51): 26956–26980. Bibcode:2014RSCAd...426956B. doi:10.1039/c4ra02129k. hdl: 10261/126489 . ISSN   2046-2069.
  9. Abou Assi, Hala; Garavís, Miguel; González, Carlos; Damha, Masad J (19 September 2018). "i-Motif DNA: structural features and significance to cell biology". Nucleic Acids Research. 46 (16): 8038–8056. doi:10.1093/nar/gky735. ISSN   0305-1048. PMC   6144788 . PMID   30124962.
  10. Malliavin, Thérèse E.; Gau, Jocelyne; Snoussi, Karim; Leroy, Jean-Louis (June 2003). "Stability of the I-motif Structure Is Related to the Interactions between Phosphodiester Backbones". Biophysical Journal. 84 (6): 3838–3847. Bibcode:2003BpJ....84.3838M. doi:10.1016/S0006-3495(03)75111-X. ISSN   0006-3495. PMC   1302965 . PMID   12770889.
  11. 1 2 Dettler, Jamie M.; Buscaglia, Robert; Cui, JingJing; Cashman, Derek; Blynn, Meredith; Lewis, Edwin A. (July 2010). "Biophysical Characterization of an Ensemble of Intramolecular i-Motifs Formed by the Human c-MYC NHE III1 P1 Promoter Mutant Sequence". Biophysical Journal. 99 (2): 561–567. Bibcode:2010BpJ....99..561D. doi:10.1016/j.bpj.2010.04.042. ISSN   0006-3495. PMC   2905117 . PMID   20643075.
  12. 1 2 Wright, Elisé P.; Huppert, Julian L.; Waller, Zoë A. E. (9 February 2017). "Identification of multiple genomic DNA sequences which form i-motif structures at neutral pH". Nucleic Acids Research. 45 (6): 2951–2959. doi:10.1093/nar/gkx090. ISSN   0305-1048. PMC   5605235 . PMID   28180276.
  13. Sun, Daekyu; Hurley, Laurence H. (14 May 2009). "The importance of negative superhelicity in inducing the formation of G-quadruplex and i-motif structures in the c-Myc promoter: implications for drug targeting and control of gene expression". Journal of Medicinal Chemistry. 52 (9): 2863–2874. doi:10.1021/jm900055s. ISSN   0022-2623. PMC   2757002 . PMID   19385599.
  14. Pandit, Sarwar Ahmad; Rather, Mudasir Ahmad; Bhat, Sajad Ahmad; Jan, Roohi; Rather, Ghulam Mohd; Bhat, Mohsin Ahmad (21 June 2017). "An Insight into a Fascinating DMF-Water Mixed Solvent System: Physicochemical and Electrochemical Studies". ChemistrySelect. 2 (18): 5115–5127. doi:10.1002/slct.201700553. ISSN   2365-6549.
  15. 1 2 Malliavin, Thérèse E.; Gau, Jocelyne; Snoussi, Karim; Leroy, Jean-Louis (June 2003). "Stability of the I-motif Structure Is Related to the Interactions between Phosphodiester Backbones". Biophysical Journal. 84 (6): 3838–3847. Bibcode:2003BpJ....84.3838M. doi:10.1016/s0006-3495(03)75111-x. ISSN   0006-3495. PMC   1302965 . PMID   12770889.
  16. 1 2 Abou Assi, Hala; Garavís, Miguel; González, Carlos; Damha, Masad J (16 August 2018). "i-Motif DNA: structural features and significance to cell biology". Nucleic Acids Research. 46 (16): 8038–8056. doi:10.1093/nar/gky735. ISSN   0305-1048. PMC   6144788 . PMID   30124962.
  17. Yang, Bo; Rodgers, M. T. (19 December 2013). "Base-Pairing Energies of Proton-Bound Heterodimers of Cytosine and Modified Cytosines: Implications for the Stability of DNAi-Motif Conformations". Journal of the American Chemical Society. 136 (1): 282–290. doi:10.1021/ja409515v. ISSN   0002-7863. PMID   24320604.
  18. Lieblein, Anna Lena; Krämer, Maximilian; Dreuw, Andreas; Fürtig, Boris; Schwalbe, Harald (12 March 2012). "The Nature of Hydrogen Bonds in Cytidine⋅⋅⋅H+⋅⋅⋅Cytidine DNA Base Pairs". Angewandte Chemie International Edition. 51 (17): 4067–4070. doi:10.1002/anie.201200549. ISSN   1433-7851. PMID   22411471.
  19. 1 2 Leroy, Jean Louis; Gehring, Kalle; Kettani, Abdelali; Gueron, Maurice (15 June 1993). "Acid multimers of oligodeoxycytidine strands: Stoichiometry, base-pair characterization, and proton exchange properties". Biochemistry. 32 (23): 6019–6031. doi:10.1021/bi00074a013. ISSN   0006-2960. PMID   8389586.
  20. 1 2 3 Peng, Yinghua; Wang, Xiaohui; Xiao, Yi; Feng, Lingyan; Zhao, Chao; Ren, Jinsong; Qu, Xiaogang (30 September 2009). "i-Motif Quadruplex DNA-Based Biosensor for Distinguishing Single- and Multiwalled Carbon Nanotubes". Journal of the American Chemical Society. 131 (38): 13813–13818. doi:10.1021/ja9051763. ISSN   0002-7863. PMID   19736925.
  21. 1 2 Wright, Elisé P.; Lamparska, Katarzyna; Smith, Steven S.; Waller, Zoë A. E. (30 August 2017). "Substitution of Cytosine with Guanylurea Decreases the Stability of i-Motif DNA". Biochemistry. 56 (36): 4879–4883. doi: 10.1021/acs.biochem.7b00628 . ISSN   0006-2960. PMID   28853563.
  22. Mir, B.; Solés, X.; González, C.; Escaja, N. (5 June 2017). "The effect of the neutral cytidine protonated analogue pseudoisocytidine on the stability of i-motif structures". Scientific Reports. 7 (1): 2772. Bibcode:2017NatSR...7.2772M. doi:10.1038/s41598-017-02723-y. ISSN   2045-2322. PMC   5459817 . PMID   28584239.
  23. 1 2 3 Bhavsar-Jog, Yogini P.; Van Dornshuld, Eric; Brooks, Tracy A.; Tschumper, Gregory S.; Wadkins, Randy M. (6 March 2014). "Epigenetic Modification, Dehydration, and Molecular Crowding Effects on the Thermodynamics of i-Motif Structure Formation from C-Rich DNA". Biochemistry. 53 (10): 1586–1594. doi:10.1021/bi401523b. ISSN   0006-2960. PMC   3985701 . PMID   24564458.
  24. Lannes, Laurie; Halder, Saheli; Krishnan, Yamuna; Schwalbe, Harald (30 June 2015). "Tuning the pH Response of i-Motif DNA Oligonucleotides". ChemBioChem. 16 (11): 1647–1656. doi:10.1002/cbic.201500182. ISSN   1439-4227. PMID   26032298. S2CID   28930880.
  25. Xu, Baochang; Devi, Gitali; Shao, Fangwei (2015). "Regulation of telomeric i-motif stability by 5-methylcytosine and 5-hydroxymethylcytosine modification". Organic & Biomolecular Chemistry. 13 (20): 5646–5651. doi:10.1039/c4ob02646b. ISSN   1477-0520. PMID   25886653.
  26. Kanehara, Hideyuki; Mizuguchi, Masatsugu; Tajima, Kunihiko; Kanaori, Kenji; Makino, Keisuke (February 1997). "Spectroscopic Evidence for the Formation of Four-Stranded Solution Structure of Oligodeoxycytidine Phosphorothioate†". Biochemistry. 36 (7): 1790–1797. doi:10.1021/bi961528c. ISSN   0006-2960. PMID   9048563.
  27. 1 2 Mergny, J. (1 November 1998). "Kinetics and thermodynamics of i-DNA formation: phosphodiester versus modified oligodeoxynucleotides". Nucleic Acids Research. 26 (21): 4797–4803. doi:10.1093/nar/26.21.4797. ISSN   1362-4962. PMC   147917 . PMID   9776737.
  28. 1 2 3 Day, Henry A.; Pavlou, Pavlos; Waller, Zoë A.E. (August 2014). "i-Motif DNA: Structure, stability and targeting with ligands". Bioorganic & Medicinal Chemistry. 22 (16): 4407–4418. doi:10.1016/j.bmc.2014.05.047. ISSN   0968-0896. PMID   24957878.
  29. Miyoshi, Daisuke; Matsumura, Shizuka; Nakano, Shu-ichi; Sugimoto, Naoki (January 2004). "Duplex Dissociation of Telomere DNAs Induced by Molecular Crowding". Journal of the American Chemical Society. 126 (1): 165–169. doi:10.1021/ja036721q. ISSN   0002-7863. PMID   14709080.
  30. Cui, Jingjing; Waltman, Phillip; Le, Vu; Lewis, Edwin (15 October 2013). "The Effect of Molecular Crowding on the Stability of Human c-MYC Promoter Sequence I-Motif at Neutral pH". Molecules. 18 (10): 12751–12767. doi: 10.3390/molecules181012751 . ISSN   1420-3049. PMC   6270392 . PMID   24132198.
  31. 1 2 Sun, Daekyu; Hurley, Laurence H. (14 May 2009). "The Importance of Negative Superhelicity in Inducing the Formation of G-Quadruplex and i-Motif Structures in the c-Myc Promoter: Implications for Drug Targeting and Control of Gene Expression". Journal of Medicinal Chemistry. 52 (9): 2863–2874. doi:10.1021/jm900055s. ISSN   0022-2623. PMC   2757002 . PMID   19385599.
  32. Krishnan-Ghosh, Yamuna; Stephens, Elaine; Balasubramanian, Shankar (2005). "PNA forms an i-motif". Chemical Communications (42): 5278–80. doi:10.1039/b510405j. ISSN   1359-7345. PMID   16244727.
  33. Saxena, Sarika; Bansal, Aparna; Kukreti, Shrikant (March 2008). "Structural polymorphism exhibited by a homopurine·homopyrimidine sequence found at the right end of human c-jun protooncogene". Archives of Biochemistry and Biophysics. 471 (2): 95–108. doi:10.1016/j.abb.2008.01.015. ISSN   0003-9861. PMID   18262488.
  34. Benabou, Sanae; Ferreira, Rubén; Aviñó, Anna; González, Carlos; Lyonnais, Sébastien; Solà, Maria; Eritja, Ramon; Jaumot, Joaquim; Gargallo, Raimundo (January 2014). "Solution equilibria of cytosine- and guanine-rich sequences near the promoter region of the n-myc gene that contain stable hairpins within lateral loops". Biochimica et Biophysica Acta (BBA) - General Subjects. 1840 (1): 41–52. doi:10.1016/j.bbagen.2013.08.028. hdl: 2445/46204 . ISSN   0304-4165. PMID   24012973.
  35. 1 2 Lacroix, Laurent; Mergny, Jean-Louis (September 2000). "Chemical Modification of Pyrimidine TFOs: Effect on i-Motif and Triple Helix Formation". Archives of Biochemistry and Biophysics. 381 (1): 153–163. doi:10.1006/abbi.2000.1934. ISSN   0003-9861. PMID   11019831.
  36. 1 2 Zeraati, Mahdi; Langley, David B.; Schofield, Peter; Moye, Aaron L.; Rouet, Romain; Hughes, William E.; Bryan, Tracy M.; Dinger, Marcel E.; Christ, Daniel (23 April 2018). "I-motif DNA structures are formed in the nuclei of human cells". Nature Chemistry. 10 (6): 631–637. Bibcode:2018NatCh..10..631Z. doi:10.1038/s41557-018-0046-3. ISSN   1755-4330. PMID   29686376. S2CID   13816298.
  37. Kang, Hyun-Jin; Kendrick, Samantha; Hecht, Sidney M.; Hurley, Laurence H. (7 March 2014). "The Transcriptional Complex Between theBCL2i-Motif and hnRNP LL Is a Molecular Switch for Control of Gene Expression That Can Be Modulated by Small Molecules". Journal of the American Chemical Society. 136 (11): 4172–4185. doi:10.1021/ja4109352. ISSN   0002-7863. PMC   3985447 . PMID   24559432.
  38. Zhang, Xi Yuan; Luo, Hong Qun; Li, Nian Bing (15 June 2014). "Crystal violet as an i-motif structure probe for reversible and label-free pH-driven electrochemical switch". Analytical Biochemistry. 455: 55–59. doi:10.1016/j.ab.2014.03.015. ISSN   0003-2697. PMID   24699211.
  39. Warner, S.L.; Vankayalapati, H.; Bashyam, S.; Grand, C.L.; Han, H.; Von Hoff, D.D.; Hurley, L.H.; Bearss, D.J. (September 2004). "126 Development of a new series of tricyclic pyrimido-indole inhibitors targeting Aurora kinases". European Journal of Cancer Supplements. 2 (8): 41. doi:10.1016/s1359-6349(04)80134-4. ISSN   1359-6349.
  40. Siddiqui-Jain, A.; Grand, C. L.; Bearss, D. J.; Hurley, L. H. (23 August 2002). "Direct evidence for a G-quadruplex in a promoter region and its targeting with a small molecule to repress c-MYC transcription". Proceedings of the National Academy of Sciences. 99 (18): 11593–11598. Bibcode:2002PNAS...9911593S. doi: 10.1073/pnas.182256799 . ISSN   0027-8424. PMC   129314 . PMID   12195017.
  41. Fernández, Sergio; Eritja, Ramon; Aviñó, Anna; Jaumot, Joaquim; Gargallo, Raimundo (November 2011). "Influence of pH, temperature and the cationic porphyrin TMPyP4 on the stability of the i-motif formed by the 5'-(C3TA2)4-3' sequence of the human telomere". International Journal of Biological Macromolecules. 49 (4): 729–736. doi:10.1016/j.ijbiomac.2011.07.004. hdl: 10261/47689 . ISSN   0141-8130. PMID   21777611.
  42. Neidle, Stephen (2012), "RNA Quadruplexes", Therapeutic Applications of Quadruplex Nucleic Acids, Elsevier, pp. 139–149, doi:10.1016/b978-0-12-375138-6.00008-x, ISBN   978-0-12-375138-6 , retrieved 16 December 2020
  43. 1 2 3 4 5 Sedghi Masoud, Shadi; Nagasawa, Kazuo (1 December 2018). "i-Motif-Binding Ligands and Their Effects on the Structure and Biological Functions of i-Motif". Chemical and Pharmaceutical Bulletin. 66 (12): 1091–1103. doi: 10.1248/cpb.c18-00720 . ISSN   0009-2363. PMID   30504626.
  44. Ni, Nan; Qu, Bo; Tian, Di; Cong, Zhiyuan; Wang, Weiping; Gao, Chao; Xiao, Lixin; Chen, Zhijian; Gong, Qihuang; Wei, Wei (2 May 2013). "Macromol. Chem. Phys. 9/2013". Macromolecular Chemistry and Physics. 214 (9): 961. doi: 10.1002/macp.201370030 . ISSN   1022-1352.
  45. Nagasawa, Kazuo; Sedghi Masoud, Shadi; Tsushima, Yamato; Iida, Keisuke (2015). "Synthesis of Macrocyclic Penta- and Tetraoxazoles as G-Quadruplex Ligands". Heterocycles. 90 (2): 866. doi: 10.3987/com-14-s(k)90 (inactive 17 February 2024). ISSN   0385-5414.{{cite journal}}: CS1 maint: DOI inactive as of February 2024 (link)
  46. Sheng, Qiran; Neaverson, Joseph C.; Mahmoud, Tasnim; Stevenson, Clare E. M.; Matthews, Susan E.; Waller, Zoë A. E. (2017). "Identification of new DNA i-motif binding ligands through a fluorescent intercalator displacement assay". Organic & Biomolecular Chemistry. 15 (27): 5669–5673. doi:10.1039/c7ob00710h. ISSN   1477-0520. PMC   5708337 . PMID   28567459.
  47. Li, X.; Peng, Y.; Ren, J.; Qu, X. (13 December 2006). "Carboxyl-modified single-walled carbon nanotubes selectively induce human telomeric i-motif formation". Proceedings of the National Academy of Sciences. 103 (52): 19658–19663. Bibcode:2006PNAS..10319658L. doi: 10.1073/pnas.0607245103 . ISSN   0027-8424. PMC   1750900 . PMID   17167055.
  48. Fleming, Aaron M.; Ding, Yun; Rogers, R. Aaron; Zhu, Judy; Zhu, Julia; Burton, Ashlee D.; Carlisle, Connor B.; Burrows, Cynthia J. (5 April 2017). "4n–1 Is a "Sweet Spot" in DNA i-Motif Folding of 2'-Deoxycytidine Homopolymers". Journal of the American Chemical Society. 139 (13): 4682–4689. doi:10.1021/jacs.6b10117. ISSN   0002-7863. PMID   28290680.
  49. Wright, Elisé P.; Huppert, Julian L.; Waller, Zöe A.E. (13 November 2017). "Identification of multiple genomic DNA sequences which form i-motif structures at neutral pH". Nucleic Acids Research. 45 (22): 13095–13096. doi:10.1093/nar/gkx1178. ISSN   0305-1048. PMC   5728391 . PMID   29140476.
  50. 1 2 Takahashi, Shuntaro; Bhattacharjee, Snehasish; Ghosh, Saptarshi; Sugimoto, Naoki; Bhowmik, Sudipta (December 2020). "Preferential targeting cancer-related i-motif DNAs by the plant flavonol fisetin for theranostics applications". Scientific Reports. 10 (1): 2504. Bibcode:2020NatSR..10.2504T. doi:10.1038/s41598-020-59343-2. ISSN   2045-2322. PMC   7018961 . PMID   32054927.
  51. Zeraati, Mahdi; Langley, David B.; Schofield, Peter; Moye, Aaron L.; Rouet, Romain; Hughes, William E.; Bryan, Tracy M.; Dinger, Marcel E.; Christ, Daniel (2018). "I-motif DNA structures are formed in the nuclei of human cells". Nature Chemistry. 10 (6): 631–637. Bibcode:2018NatCh..10..631Z. doi:10.1038/s41557-018-0046-3. PMID   29686376. S2CID   13816298.
  52. Amato, Jussara; Iaccarino, Nunzia; Randazzo, Antonio; Novellino, Ettore; Pagano, Bruno (24 June 2014). "Noncanonical DNA Secondary Structures as Drug Targets: the Prospect of the i-Motif". ChemMedChem. 9 (9): 2026–2030. doi:10.1002/cmdc.201402153. ISSN   1860-7179. PMID   24962454. S2CID   39651503.
  53. Chen, Yong; Qu, Konggang; Zhao, Chuanqi; Wu, Li; Ren, Jinsong; Wang, Jiasi; Qu, Xiaogang (January 2012). "Insights into the biomedical effects of carboxylated single-wall carbon nanotubes on telomerase and telomeres". Nature Communications. 3 (1): 1074. Bibcode:2012NatCo...3.1074C. doi: 10.1038/ncomms2091 . ISSN   2041-1723. PMID   23011128.
  54. Sushmita, Nautiyal (June 2020). "I-MOTIF DNA: SIGNIFICANCE AND FUTURE PROSPECTIVE".
  55. Yoga, Y. M. K.; Traore, D. A. K.; Sidiqi, M.; Szeto, C.; Pendini, N. R.; Barker, A.; Leedman, P. J.; Wilce, J. A.; Wilce, M. C. J. (1 June 2012). "Contribution of the first K-homology domain of poly(C)-binding protein 1 to its affinity and specificity for C-rich oligonucleotides". Nucleic Acids Research. 40 (11): 5101–5114. doi:10.1093/nar/gks058. ISSN   0305-1048. PMC   3367169 . PMID   22344691.
  56. Kaiser, Christine E.; Van Ert, Natalie A.; Agrawal, Prashansa; Chawla, Reena; Yang, Danzhou; Hurley, Laurence H. (28 June 2017). "Insight into the Complexity of the i-Motif and G-Quadruplex DNA Structures Formed in the KRAS Promoter and Subsequent Drug-Induced Gene Repression". Journal of the American Chemical Society. 139 (25): 8522–8536. doi:10.1021/jacs.7b02046. ISSN   0002-7863. PMC   5978000 . PMID   28570076.
  57. Uribe, Diana J.; Guo, Kexiao; Shin, Yoon-Joo; Sun, Daekyu (10 May 2011). "Heterogeneous Nuclear Ribonucleoprotein K and Nucleolin as Transcriptional Activators of the Vascular Endothelial Growth Factor Promoter through Interaction with Secondary DNA Structures". Biochemistry. 50 (18): 3796–3806. doi:10.1021/bi101633b. ISSN   0006-2960. PMC   3119528 . PMID   21466159.
  58. Tang, Wenhuan; Niu, Kangkang; Yu, Guoxing; Jin, Ying; Zhang, Xian; Peng, Yuling; Chen, Shuna; Deng, Huimin; Li, Sheng; Wang, Jian; Song, Qisheng (5 March 2020). "In vivo visualization of the i-motif DNA secondary structure in the Bombyx mori testis". Epigenetics & Chromatin. 13 (1): 12. doi: 10.1186/s13072-020-00334-y . ISSN   1756-8935. PMC   7059380 . PMID   32138783.
  59. Wang, Qin; Dai, Tianyue; Sun, Pengfei; Wang, Xiayan; Wang, Guangfeng (1 August 2020). "A colorimetric and ratiometric glucose sensor based on conformational switch of i-motif DNA". Talanta Open. 1: 100001. doi: 10.1016/j.talo.2020.100001 . ISSN   2666-8319.
  60. Song, Lei; Ho, Vincent H. B.; Chen, Chun; Yang, Zhongqiang; Liu, Dongsheng; Chen, Rongjun; Zhou, Dejian (2013). "Drug Delivery: Efficient, pH-Triggered Drug Delivery Using a pH-Responsive DNA-Conjugated Gold Nanoparticle (Adv. Healthcare Mater. 2/2013)". Advanced Healthcare Materials. 2 (2): 380. doi:10.1002/adhm.201370008. ISSN   2192-2659.
  61. Takahashi, Shuntaro; Bhattacharjee, Snehasish; Ghosh, Saptarshi; Sugimoto, Naoki; Bhowmik, Sudipta (13 February 2020). "Preferential targeting cancer-related i-motif DNAs by the plant flavonol fisetin for theranostics applications". Scientific Reports. 10 (1): 2504. Bibcode:2020NatSR..10.2504T. doi:10.1038/s41598-020-59343-2. ISSN   2045-2322. PMC   7018961 . PMID   32054927.
  62. Li, Tao; Famulok, Michael (22 January 2013). "I-Motif-Programmed Functionalization of DNA Nanocircles". Journal of the American Chemical Society. 135 (4): 1593–1599. doi: 10.1021/ja3118224 . ISSN   0002-7863. PMID   23312021.