[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (579)

Search Parameters:
Keywords = G-quadruplex

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
16 pages, 3303 KiB  
Article
Cross-Effects in Folding and Phase Transitions of hnRNP A1 and C9Orf72 RNA G4 In Vitro
by Tatiana Vedekhina, Julia Svetlova, Iuliia Pavlova, Nikolay Barinov, Sabina Alieva, Elizaveta Malakhova, Pavel Rubtsov, Alina Shtork, Dmitry Klinov and Anna Varizhuk
Molecules 2024, 29(18), 4369; https://doi.org/10.3390/molecules29184369 (registering DOI) - 14 Sep 2024
Viewed by 200
Abstract
Abnormal intracellular phase transitions in mutant hnRNP A1 may underlie the development of several neurodegenerative diseases. The risk of these diseases increases upon C9Orf72 repeat expansion and the accumulation of the corresponding G-quadruplex (G4)-forming RNA, but the link between this RNA and the [...] Read more.
Abnormal intracellular phase transitions in mutant hnRNP A1 may underlie the development of several neurodegenerative diseases. The risk of these diseases increases upon C9Orf72 repeat expansion and the accumulation of the corresponding G-quadruplex (G4)-forming RNA, but the link between this RNA and the disruption of hnRNP A1 homeostasis has not been fully explored so far. Our aim was to clarify the mutual effects of hnRNP A1 and C9Orf72 G4 in vitro. Using various optical methods and atomic force microscopy, we investigated the influence of the G4 on the formation of cross-beta fibrils by the mutant prion-like domain (PLD) of hnRNP A1 and on the co-separation of the non-mutant protein with a typical SR-rich fragment of a splicing factor (SRSF), which normally drives the assembly of nuclear speckles. The G4 was shown to act in a holdase-like manner, i.e., to restrict the fibrillation of the hnRNP A1 PLD, presumably through interactions with the PLD-flanking RGG motif. These interactions resulted in partial unwinding of the G4, suggesting a helicase-like activity of hnRNP A1 RGG. At the same time, the G4 was shown to disrupt hnRNP A1 co-separation with SRSF, suggesting its possible contribution to pathology through interference with splicing regulation. Full article
Show Figures

Figure 1

Figure 1
<p>The domains and key motifs of hnRNP A1 are involved in G4 recognition and fibrillation. (<b>a</b>) The domain structure of hnRNP A1. (<b>b</b>) The schematic representation of hnRNP A1 fibrillation and binding to the C9Orf72 RNA G4. Blue, RNA recognition motifs of hnRNP A1 (RRM1 and RRM2); gray, Gly-rich disordered region of hnRNP A1; black, RGG motif within the Gly-rich region; orange, prion-like domain (PLD); yellow, nuclear localization signal (NLS); red, C9Orf72 RNA G4 (PDB ID: 8X0S). The orange box indicates the cross-beta core of hnRNP A1 (PDB ID: 7BX7). The blue arrows indicate potential interactions, and the black arrows indicate rotation. (<b>c</b>) ThT assays illustrating fibrillation of hnRNP A1 fragments in the presence/absence of random/G4 RNA. Conditions: 40 µM peptide, 13 µM ThT, 0.5 mM DTT, and 0.1 mg/mL random RNA or 20 µM G4 in 40 mM HEPES-KOH buffer (pH 7.4), supplemented with 150 mM KCl. Error bars indicate the SD of six measurements (two biological and three technical repeats). * <span class="html-italic">p</span> &lt; 0.05 (two-tailed <span class="html-italic">t</span>-test).</p>
Full article ">Figure 2
<p>The impact of the RNA G4 on the fibrillation of hnRNP A1 (RGG)-PLD<sup>mut</sup>-NLS. AFM assays. Top: representative AFM images of the samples from the ThT assays on a modified graphite support or mica. Bottom: summary of the statistical analysis of particle heights in the AFM images obtained on a modified graphite support (500 particles for each sample type). All ThT assays and AFM scans were performed after a 3 h incubation of the peptide/RNA solutions or their mixtures at room temperature.</p>
Full article ">Figure 3
<p>The impact of RGG-PLD<sup>mut</sup>-NLS on the RNA secondary structure and vice versa. (<b>a</b>) CD spectra of RGG-PLD<sup>mut</sup>-NLS (40 µM), random RNA (0.1 mg/mL), G4wt (10 µM), G4mut (10 µM), and their mixtures in 4 mM HEPES-KOH buffer (pH 7.4) supplemented with 15 mM KCl (0.1 × HEPES buffer). (<b>b</b>) Ellipticity changes upon RGG-PLD<sup>mut</sup>-NLS titration with random RNA. Conditions: 0.2 mg/mL peptide and 0–0.2 mg/mL RNA in 0.1 × HEPES buffer. Top: UV absorption spectra; middle: CD spectra; bottom: comparison of the experimental (mixture) and theoretical (superposition) spectra in the far-UV region. (<b>c</b>) Ellipticity changes upon G4mut (<b>left</b>) or G4wt (<b>right</b>) titration with the peptides. Conditions: 1 µM G4 and 0–8 µM peptide in 0.1 × HEPES buffer. All spectra were registered at room temperature in 0.05 cm (<b>a</b>,<b>b</b>) or 1 cm (<b>c</b>) cuvettes (the average of three measurements is shown).</p>
Full article ">Figure 4
<p>The impact of the G4s on the co-separation of hnRNP A1 and SRSF<sup>fr</sup>. (<b>a</b>) A generic phase diagram of an RNA-protein mixture in non-crowded media. The arrows between points 1 and 4 summarize the experimental setup and the observed RNA-induced transitions. (<b>b</b>) Fluorescence microscopy images of the hnRNP A1/SRSF<sup>fr</sup> mixture with/without RNA. Conditions: 6 µM hnRNP A1 (5% RED-labeled) and 1 mg/mL SRSF<sup>fr</sup> (5% FITC-labeled) in the absence of RNA (1), in the presence of 3 mg/mL random RNA (2), in the presence of 3 mg/mL random RNA and 6 µM G4mut (3), or in the presence of 3 mg/mL random RNA and 6 µM G4wt (4). All images were obtained after 30 min of incubation of the mixtures at room temperature in 40 mM HEPES-KOH buffer (pH 7) supplemented with 150 mM KCl. (<b>c</b>) Enlarged images of samples 2 (random RNA only) and 4 (random RNA with G4wt) in the far-red (hnRNP A1) and green (SRSF<sup>fr</sup>) channels and their superposition. (<b>d</b>) A generic phase diagram of an RNA-protein mixture in a crowded media. (<b>e</b>) Fluorescence microscopy images of hnRNP A1/SRSF<sup>fr</sup> mixture with/without RNA in the presence of a crowding agent (PEG-400). The conditions are similar to those in (<b>b</b>), except that 10% PEG-400 was added to all samples. (<b>f</b>) Enlarged images of samples (2) and (4) in the presence of PEG obtained from the red and green channels and their superposition.</p>
Full article ">Figure 5
<p>Hypothetical intracellular transitions of hnRNP A1 in the presence of C9Orf72 RNA G4. Normally, hnRNP A1 co-separates with SRSFs and pre-mRNA/ncRNA in nuclear speckles (1), which ensures its homeostasis in the nucleoplasm and its controlled contribution to the splicing of cryptic exons (2). Under stress conditions, the relocation of hnRNP A1 to the cytoplasm (3) opens ways for its fibrillation (4). The maintenance of hnRNA 1 in the cytoplasm is possible upon the accumulation of C9Orf72 RNA, which is prone to non-AUG translation into NIR-inhibiting polydipeptides (5). At the same time, C9Orf72 RNA is likely to disrupt the co-separation of hnRNP A1 with SRSFs in the nucleoplasm (6), suggesting an increased risk of splicing deregulation (loss-of-function toxicity). In the cytoplasm, C9Orf72 RNA is likely to form irregular aggregates/condensates, but not lengthy fibrils, with hnRNP A1 (7), suggesting a decreased risk of fibrillation (gain-of-function toxicity).</p>
Full article ">
23 pages, 1593 KiB  
Article
Design, Synthesis, and In Vitro Antimalarial Evaluation of New 1,3,5-Tris[(4-(Substituted-Aminomethyl)Phenoxy)Methyl]Benzenes
by Sandra Albenque-Rubio, Jean Guillon, Patrice Agnamey, Céline Damiani, Solène Savrimoutou, Romain Mustière, Noël Pinaud, Stéphane Moreau, Jean-Louis Mergny, Luisa Ronga, Ioannis Kanavos, Mathieu Marchivie, Serge Moukha, Pascale Dozolme, Pascal Sonnet and Anita Cohen
Drugs Drug Candidates 2024, 3(3), 615-637; https://doi.org/10.3390/ddc3030035 - 13 Sep 2024
Viewed by 167
Abstract
By taking into account our previously described series of 1,3,5-tris[(4-(substituted-aminomethyl)phenyl)methyl]benzene compounds, we have now designed, prepared, and evaluated in vitro against Plasmodium falciparum a novel series of structural analogues of these molecules, i.e., the 1,3,5-tris[(4-(substituted-aminomethyl)phenoxy)methyl]benzene derivatives. The pharmacological data [...] Read more.
By taking into account our previously described series of 1,3,5-tris[(4-(substituted-aminomethyl)phenyl)methyl]benzene compounds, we have now designed, prepared, and evaluated in vitro against Plasmodium falciparum a novel series of structural analogues of these molecules, i.e., the 1,3,5-tris[(4-(substituted-aminomethyl)phenoxy)methyl]benzene derivatives. The pharmacological data showed antimalarial activity with IC50 values in the sub and μM range. The in vitro cytotoxicity of these new nitrogen polyphenoxymethylbenzene compounds was also evaluated on human HepG2 cells. The 1,3,5-tris[(4-(substituted-aminomethyl)phenoxy)methyl]benzene derivative 1m was found as one of the most potent and promising antimalarial candidates with favorable cytotoxic to antiprotozoal properties in the P. falciparum strains W2 and 3D7. In conclusion, this 1,3,5-tris[(4-(pyridin-3-ylmethylaminomethyl)phenoxyl)methyl]benzene 1m (IC50 = 0.07 μM on W2, 0.06 μM on 3D7, and 62.11 μM on HepG2) was identified as the most promising antimalarial derivative with selectivity indexes (SI) of 887.29 on the W2 P. falciparum chloroquine-resistant strain, and of 1035.17 on the chloroquine-sensitive and mefloquine decreased sensitivity strain 3D7. It has been previously described that the telomeres of P. falciparum could represent potential targets for these types of polyaromatic compounds; therefore, the capacity of our novel derivatives to stabilize the parasitic telomeric G-quadruplexes was assessed using a FRET melting assay. However, with regard to the stabilization of the protozoal G-quadruplex, we observed that the best substituted derivatives 1, which exhibited some interesting stabilization profiles, were not the most active antimalarial compounds against the two Plasmodium strains. Thus, there were no correlations between their antimalarial activities and selectivities of their respective binding to G-quadruplexes. Full article
(This article belongs to the Collection Anti-Parasite Drug Discovery)
Show Figures

Figure 1

Figure 1
<p>The structures of chloroquine (CQ), amodiaquine (AQ), piperaquine, mefloquine (MQ), bisquinoline (<b>A</b>), bisacridine (<b>B</b>), tafenoquine, series (<b>A</b>–<b>D</b>), and newly synthesized 1,3,5-<span class="html-italic">tris</span>[(4-(substituted-aminomethyl)phenoxy)methyl]benzene derivatives <b>1a–r</b> (Series (<b>E</b>)).</p>
Full article ">Figure 2
<p>The ORTEP (Oak Ridge Thermal Ellipsoid Plot) drawing of the 1,3,5-<span class="html-italic">tris</span>[(4-formylphenoxy)methyl]benzene <b>2</b> with atom labeling and thermal ellipsoids at 30% probability level. Hydrogen atoms are represented as small spheres of arbitrary radii.</p>
Full article ">Figure 3
<p>The structures of the most promising synthesized 1,3,5-<span class="html-italic">tris</span>[(4-(substituted-aminomethyl)phenoxy)methyl]benzene derivatives <b>1m</b> and <b>1p</b>. Selectivity indexes (S.I.) are provided below.</p>
Full article ">Figure 4
<p>Thermal stabilization (ΔTm) induced by selected compounds <b>1a–r</b> (at 2 μM) of <span class="html-italic">Plasmodium</span> telomeric G-quadruplexes (<b>A</b>) FPf1T and (<b>B</b>) FPf8T vs. the human telomeric quadruplex F21T.</p>
Full article ">Figure 5
<p>Stabilization specificity profile of <b>1a–r</b> (2 µM) toward various G4 and control duplex oligonucleotides. The difference in Tm in presence and absence of <b>1a–r</b>, ∆Tm, in °C is plotted for each sequence. Three quadruplexes and one duplex (FdxT) were tested.</p>
Full article ">Scheme 1
<p>General procedure for the preparation of novel compounds <b>1a–r</b>.</p>
Full article ">
13 pages, 2918 KiB  
Article
Fluorescence Turn-Off Ligand for Parallel G-Quadruplexes
by Joanna Nowak-Karnowska, Agata Głuszyńska, Joanna Kosman and Anna Dembska
Molecules 2024, 29(16), 3907; https://doi.org/10.3390/molecules29163907 - 18 Aug 2024
Viewed by 529
Abstract
Parallel-stranded G-quadruplex structures are found to be common in the human promoter sequences. We tested highly fluorescent 9-methoxyluminarine ligand (9-MeLM) binding interactions with different parallel G-quadruplexes DNA by spectroscopic methods such as fluorescence and circular dichroism (CD) titration as well as UV melting [...] Read more.
Parallel-stranded G-quadruplex structures are found to be common in the human promoter sequences. We tested highly fluorescent 9-methoxyluminarine ligand (9-MeLM) binding interactions with different parallel G-quadruplexes DNA by spectroscopic methods such as fluorescence and circular dichroism (CD) titration as well as UV melting profiles. The results showed that the studied 9-MeLM ligand interacted with the intramolecular parallel G-quadruplexes (G4s) with similar affinity. The binding constants of 9-methoxyluminarine with different parallel G4s were determined. The studies upon oligonucleotides with different flanking sequences on c-MYC G-quadruplex suggest that 9-methoxyluminarine may preferentially interact with 3′end of the c-MYC promoter. The high decrease in 9-MeLM ligand fluorescence upon binding to all tested G4s indicates that 9-methoxyluminarine molecule can be used as a selective fluorescence turn-off probe for parallel G-quadruplexes. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>A schematic illustration of binding mode between 9-MeLM ligand and c-<span class="html-italic">MYC</span> G-quadruplex.</p>
Full article ">Figure 2
<p>Circular dichroism (CD) spectra of G-quadruplexes c-<span class="html-italic">KIT1</span> (<b>A</b>), <span class="html-italic">RET</span> (<b>B</b>) and catG4 (<b>C</b>) with increasing amounts of 9-methoxyluminarine ligand; (<b>D</b>) CD signal changes at 265 nm against G-quadruplexes DNA/ligand molar ratio. Conditions: 10 mM Tris-HCl buffer (pH 7.2), 100 mM KCl, [G4] = 5 µM.</p>
Full article ">Figure 3
<p>CD spectra of c-<span class="html-italic">KIT1</span> (<b>A</b>), <span class="html-italic">RET</span> (<b>B</b>), catG4 (<b>C</b>) and oligonucleotides (5 µM) with 5 equiv. of 9-methoxyluminarine ligand in Tris–HCl buffer (10 mM, pH 7.2) and increasing amounts (0–100 mM) of KCl.</p>
Full article ">Figure 4
<p>Normalized UV melting profiles (melting curves) of 2 μM G-quadruplexes (black), with and without 3 equiv. of 9-MeLM ligand (red) in 10 mM Tris–HCl buffer (pH 7.2) containing 10 mM KCl/90 mM LiCl (<b>A</b>) <span class="html-italic">RET</span>, (<b>B</b>) c-<span class="html-italic">MYC</span>, (<b>C</b>) c-<span class="html-italic">KIT1</span> and (<b>D</b>) catG4.</p>
Full article ">Figure 5
<p>The fluorescence intensity (at 495 nm, λ<sub>ex</sub> 390 nm) of 9-methoxyluminarine (9-MeLM) vs. the increasing concentration of G-quadruplexes (G4s) (<b>A</b>,<b>B</b>) and fluorescence spectra recorded for 9-methoxyluminarine in Tris–HCl buffer (10 mM, pH 7.2), KCl (100 mM) w/o (black) and with 10 equiv. of different G4s (<b>C</b>,<b>D</b>).</p>
Full article ">
20 pages, 2434 KiB  
Review
Advancements in Telomerase-Targeted Therapies for Glioblastoma: A Systematic Review
by Giovanni Pennisi, Placido Bruzzaniti, Benedetta Burattini, Giacomo Piaser Guerrato, Giuseppe Maria Della Pepa, Carmelo Lucio Sturiale, Pierfrancesco Lapolla, Pietro Familiari, Biagia La Pira, Giancarlo D’Andrea, Alessandro Olivi, Quintino Giorgio D’Alessandris and Nicola Montano
Int. J. Mol. Sci. 2024, 25(16), 8700; https://doi.org/10.3390/ijms25168700 - 9 Aug 2024
Viewed by 740
Abstract
Glioblastoma (GBM) is a primary CNS tumor that is highly lethal in adults and has limited treatment options. Despite advancements in understanding the GBM biology, the standard treatment for GBM has remained unchanged for more than a decade. Only 6.8% of patients survive [...] Read more.
Glioblastoma (GBM) is a primary CNS tumor that is highly lethal in adults and has limited treatment options. Despite advancements in understanding the GBM biology, the standard treatment for GBM has remained unchanged for more than a decade. Only 6.8% of patients survive beyond five years. Telomerase, particularly the hTERT promoter mutations present in up to 80% of GBM cases, represents a promising therapeutic target due to its role in sustaining telomere length and cancer cell proliferation. This review examines the biology of telomerase in GBM and explores potential telomerase-targeted therapies. We conducted a systematic review following the PRISMA-P guidelines in the MEDLINE/PubMed and Scopus databases, from January 1995 to April 2024. We searched for suitable articles by utilizing the terms “GBM”, “high-grade gliomas”, “hTERT” and “telomerase”. We incorporated studies addressing telomerase-targeted therapies into GBM studies, excluding non-English articles, reviews, and meta-analyses. We evaluated a total of 777 records and 46 full texts, including 36 studies in the final review. Several compounds aimed at inhibiting hTERT transcription demonstrated promising preclinical outcomes; however, they were unsuccessful in clinical trials owing to intricate regulatory pathways and inadequate pharmacokinetics. Direct hTERT inhibitors encountered numerous obstacles, including a prolonged latency for telomere shortening and the activation of the alternative lengthening of telomeres (ALT). The G-quadruplex DNA stabilizers appeared to be potential indirect inhibitors, but further clinical studies are required. Imetelstat, the only telomerase inhibitor that has undergone clinical trials, has demonstrated efficacy in various cancers, but its efficacy in GBM has been limited. Telomerase-targeted therapies in GBM is challenging due to complex hTERT regulation and inadequate inhibitor pharmacokinetics. Our study demonstrates that, despite promising preclinical results, no Telomerase inhibitors have been approved for GBM, and clinical trials have been largely unsuccessful. Future strategies may include Telomerase-based vaccines and multi-target inhibitors, which may provide more effective treatments when combined with a better understanding of telomere dynamics and tumor biology. These treatments have the potential to be integrated with existing ones and to improve the outcomes for patients with GBM. Full article
(This article belongs to the Special Issue Novel Biomarkers and Molecular Targets in Gliomas)
Show Figures

Figure 1

Figure 1
<p>PRISMA flow diagram [<a href="#B11-ijms-25-08700" class="html-bibr">11</a>].</p>
Full article ">Figure 2
<p>Molecules that inhibit hTERT at the transcriptional level, resulting in a reduction in hTERT mRNA levels. EGCG, epigallocatechin gallate; hTERT, human telomerase reverse transcriptase; hTER, RSV, resveratrol; TQ, thymoquinone; TSA, trichostatin A; MZ-5-156: growth hormone–releasing hormone antagonist inhibitor.</p>
Full article ">Figure 3
<p>The telomerase pathway and its inhibitors are the topic of interest. BIBR1532 has direct effects on the hTERT protein, while other molecules have an indirect effect through the modulations of ROS, p53, and p21. Imetelstat and GNR163 affect hTR, while the TRF1 Inhibitor and miR-90 affect the sheltering proteins complex.</p>
Full article ">Figure 4
<p>G4 ligands bind to DNA secondary structures called the G-quadruplexes complex. These proteins interact with telomeres, induce telomere uncapping, and indirectly affect telomerase function.</p>
Full article ">Figure 5
<p>Illustration of the mechanism of action of telomerase reverse transcriptase (h<span class="html-italic">TERT</span>) promoter mutations. Two of the most common mutations in cancer are mutations in the cytidine-to-thymidine transition, which occur in two hotspots before the transcription start site. A novel binding site for the GABP transcription factor, which activates TERT expression, is created by these mutations.</p>
Full article ">
10 pages, 1963 KiB  
Article
G-Quadruplex DNA as a Macromolecular Target for Semi-Synthetic Isoflavones Bearing B-Ring Tosylation
by Giovanni Ribaudo, Margrate Anyanwu, Matteo Giannangeli, Erika Oselladore, Alberto Ongaro, Maurizio Memo and Alessandra Gianoncelli
Macromol 2024, 4(3), 556-565; https://doi.org/10.3390/macromol4030033 - 7 Aug 2024
Viewed by 398
Abstract
Guanine-rich sequences of nucleic acids, including DNA and RNA, are known to fold into non-canonical structures named G-quadruplexes (G4s). Such arrangements of these macromolecular polymers are mainly located in telomeres and in promoter regions of oncogenes and, for this reason, they represent a [...] Read more.
Guanine-rich sequences of nucleic acids, including DNA and RNA, are known to fold into non-canonical structures named G-quadruplexes (G4s). Such arrangements of these macromolecular polymers are mainly located in telomeres and in promoter regions of oncogenes and, for this reason, they represent a potential target for compounds with therapeutic applications. In fact, the ligand-mediated stabilization of G4s inhibits telomerase and the activity of transcriptional machinery and counteracts cancer cell immortalization. Flavonoids, along with other classes of small molecules, have been previously tested for their ability to stabilize G4s, but the mechanism of their interaction has not been fully elucidated. In the current work, we report a multi-technique investigation on the binding of tosylated isoflavones obtained by the B-ring modification of compounds from Maclura pomifera to a telomeric DNA sequence. Our study demonstrates that such derivatization leads to compounds showing lower binding affinity but with an increased selectivity toward G4 with respect to double-stranded DNA. The binding mode to the macromolecular target G4 was studied by combining results from electrospray mass spectrometry binding studies, nuclear magnetic resonance experiments and computational simulations. Overall, our findings show that tosylation influences the selectivity toward the macromolecular target by affecting the interaction mode with the nucleic acid. Full article
Show Figures

Figure 1

Figure 1
<p>Representative mass spectrum showing the interaction of a ligand (isoosajin <b>5</b>) with G4 DNA. In the inset, CID curves of G4–ligand complexes for compounds <b>5</b> (yellow), <b>6</b> (pink), <b>9</b> (blue), <b>10</b> (red), <b>11</b> (black) and <b>12</b> (cyan) are reported.</p>
Full article ">Figure 2
<p>Poses overlap of the multiparametric molecular docking: osajin <b>1</b> (<b>a</b>), <span class="html-italic">p</span>-toluensulfonyl osajin <b>7</b> (<b>b</b>), scandenone <b>4</b> (<b>c</b>), <span class="html-italic">p</span>-toluensulfonyl scandenone <b>10</b> (<b>d</b>). The chemical structures of the compounds are reported in the upper part of every panel.</p>
Full article ">Figure 3
<p>MD trajectories and snapshots showing the pose transition from groove to stacking binding mode for osajin <b>1</b> (blue) and scandenone <b>4</b> (green).</p>
Full article ">
19 pages, 1921 KiB  
Review
Novel Therapeutic Horizons: SNCA Targeting in Parkinson’s Disease
by Alessio Maria Caramiello and Valentina Pirota
Biomolecules 2024, 14(8), 949; https://doi.org/10.3390/biom14080949 - 6 Aug 2024
Viewed by 825
Abstract
Alpha-synuclein (αSyn) aggregates are the primary component of Lewy bodies, which are pathological hallmarks of Parkinson’s disease (PD). The toxicity of αSyn seems to increase with its elevated expression during injury, suggesting that therapeutic approaches focused on reducing αSyn burden in neurons could [...] Read more.
Alpha-synuclein (αSyn) aggregates are the primary component of Lewy bodies, which are pathological hallmarks of Parkinson’s disease (PD). The toxicity of αSyn seems to increase with its elevated expression during injury, suggesting that therapeutic approaches focused on reducing αSyn burden in neurons could be beneficial. Additionally, studies have shown higher levels of SNCA mRNA in the midbrain tissues and substantia nigra dopaminergic neurons of sporadic PD post-mortem brains compared to controls. Therefore, the regulation of SNCA expression and inhibition of αSyn synthesis could play an important role in the pathogenesis of injury, resulting in an effective treatment approach for PD. In this context, we summarized the most recent and innovative strategies proposed that exploit the targeting of SNCA to regulate translation and efficiently knock down cytoplasmatic levels of αSyn. Significant progress has been made in developing antisense technologies for treating PD in recent years, with a focus on antisense oligonucleotides and short-interfering RNAs, which achieve high specificity towards the desired target. To provide a more exhaustive picture of this research field, we also reported less common but highly innovative strategies, including small molecules, designed to specifically bind 5′-untranslated regions and, targeting secondary nucleic acid structures present in the SNCA gene, whose formation can be modulated, acting as a transcription and translation control. To fully describe the efficiency of the reported strategies, the effect of αSyn reduction on cellular viability and dopamine homeostasis was also considered. Full article
(This article belongs to the Section Biomacromolecules: Nucleic Acids)
Show Figures

Figure 1

Figure 1
<p>Common chemical modifications to improve classical ASO features.</p>
Full article ">Figure 2
<p>Chemical structure of AmNA and LNA phosphorothioate moieties.</p>
Full article ">Figure 3
<p>siRNA pathway: ribonuclease protein Dicer recognizes and cleaves DNA double-strand into small fragments (21–23 bp), known as siRNAs, which form the protein RISC complex. Then, siRNA binds the target sequence on mRNA, inducing its cleavage into small fragments (10–11 bp). This results in the suppression of mRNA translation.</p>
Full article ">Figure 4
<p>Chemical structures of ligands targeting 5′-UTR of <span class="html-italic">SNCA</span> mRNA.</p>
Full article ">Figure 5
<p>Schematic representation of (<b>A</b>) square-planar G-tetrad; (<b>B</b>) backbone of the intramolecular G-quadruplex structure; (<b>C</b>) G4 topologies: parallel-, antiparallel-, and hybrid-type G4 structures.</p>
Full article ">Figure 6
<p>Cartoon representing the three G4 motifs identified in the 5′-UTR mRNA region of the <span class="html-italic">SNCA</span> gene by Koukouraki et al. [<a href="#B78-biomolecules-14-00949" class="html-bibr">78</a>], together with all the G to A mutations tested (highlighted in yellow).</p>
Full article ">
22 pages, 1115 KiB  
Review
A Phenotypic Approach to the Discovery of Potent G-Quadruplex Targeted Drugs
by Stephen Neidle
Molecules 2024, 29(15), 3653; https://doi.org/10.3390/molecules29153653 - 1 Aug 2024
Viewed by 1485
Abstract
G-quadruplex (G4) sequences, which can fold into higher-order G4 structures, are abundant in the human genome and are over-represented in the promoter regions of many genes involved in human cancer initiation, progression, and metastasis. They are plausible targets for G4-binding small molecules, which [...] Read more.
G-quadruplex (G4) sequences, which can fold into higher-order G4 structures, are abundant in the human genome and are over-represented in the promoter regions of many genes involved in human cancer initiation, progression, and metastasis. They are plausible targets for G4-binding small molecules, which would, in the case of promoter G4s, result in the transcriptional downregulation of these genes. However, structural information is currently available on only a very small number of G4s and their ligand complexes. This limitation, coupled with the currently restricted information on the G4-containing genes involved in most complex human cancers, has led to the development of a phenotypic-led approach to G4 ligand drug discovery. This approach was illustrated by the discovery of several generations of tri- and tetra-substituted naphthalene diimide (ND) ligands that were found to show potent growth inhibition in pancreatic cancer cell lines and are active in in vivo models for this hard-to-treat disease. The cycles of discovery have culminated in a highly potent tetra-substituted ND derivative, QN-302, which is currently being evaluated in a Phase 1 clinical trial. The major genes whose expression has been down-regulated by QN-302 are presented here: all contain G4 propensity and have been found to be up-regulated in human pancreatic cancer. Some of these genes are also upregulated in other human cancers, supporting the hypothesis that QN-302 is a pan-G4 drug of potential utility beyond pancreatic cancer. Full article
(This article belongs to the Special Issue Bioorganic Chemistry in Europe)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Structures of G4 ligands and experimental drugs discussed in this text. The substituent group in QN-302 responsible for enhanced potency is circled in red, as is the methoxy group in the short-listed ND lead compound SOP1247.</p>
Full article ">Figure 2
<p>Surface representations of three well-established G4 binders, drawn to the same scale. The four side chains of QN-302 can be seen—three are oriented out of the plane of the page and the fourth, with the benzyl-pyrrolidine group, is seen on the left-hand side of the molecule. Two of the three side chains of BRACO-19 are in the plane of the page; these have similar dimensions to two of the QN-302 ones and would be expected to bind in G4 grooves in a similar mode.</p>
Full article ">Figure 3
<p>Flowchart of the history and phenotypic/structure-guided development of the experimental drug QN-302.</p>
Full article ">
13 pages, 4047 KiB  
Article
Chaperone Copolymer-Assisted Catalytic Hairpin Assembly for Highly Sensitive Detection of Adenosine
by Yazhen Liao, Xiaoxue Yin, Wenqian Liu, Zhenrui Du and Jie Du
Polymers 2024, 16(15), 2179; https://doi.org/10.3390/polym16152179 - 31 Jul 2024
Viewed by 552
Abstract
Adenosine is an endogenous molecule that plays a vital role in biological processes. Research indicates that abnormal adenosine levels are associated with a range of diseases. The development of sensors capable of detecting adenosine is pivotal for early diagnosis of disease. For example, [...] Read more.
Adenosine is an endogenous molecule that plays a vital role in biological processes. Research indicates that abnormal adenosine levels are associated with a range of diseases. The development of sensors capable of detecting adenosine is pivotal for early diagnosis of disease. For example, elevated adenosine levels are closely associated with the onset and progression of cancer. In this study, we designed a novel DNA biosensor utilizing chaperone copolymer-assisted catalytic hairpin assembly for highly sensitive detection of adenosine. The functional probe comprises streptavidin magnetic beads, an aptamer, and a catalytic chain. In the presence of adenosine, it selectively binds to the aptamer, displacing the catalytic chain into the solution. The cyclic portion of H1 hybridizes with the catalytic strand, while H2 hybridizes with the exposed H1 fragment to form an H1/H2 complex containing a G-quadruplex. Thioflavin T binds specifically to the G-quadruplex, generating a fluorescent signal. As a nucleic acid chaperone, PLL-g-Dex expedites the strand exchange reaction, enhancing the efficiency of catalytic hairpin assembly, thus amplifying the signal and reducing detection time. The optimal detection conditions were determined to be a temperature of 25 °C and a reaction time of 10 min. Demonstrating remarkable sensitivity and selectivity, the sensor achieved a lowest limit of detection of 9.82 nM. Furthermore, it exhibited resilience to interference in complex environments such as serum, presenting an effective approach for rapid and sensitive adenosine detection. Full article
(This article belongs to the Special Issue Biopolymer-Based Materials in Medical Applications)
Show Figures

Figure 1

Figure 1
<p>Chemical structure of chaperone copolymer PLL-<span class="html-italic">g</span>-Dex.</p>
Full article ">Figure 2
<p>Schematic diagram of a biosensor based on (<b>A</b>) streptavidin magnetic beads functional probe, and (<b>B</b>) PLL-<span class="html-italic">g</span>-Dex assisted catalytic hairpin assembly for adenosine detection.</p>
Full article ">Figure 3
<p>Fluorescence emission spectrum under different conditions: (a) MBs-Aptamer/tDNA + H1 + H2 + K<sup>+</sup>; (b) MBs-Aptamer/tDNA + H1 + H2 + K<sup>+</sup> + Target; (c) MBs-Aptamer/tDNA + H1 + H2 + K<sup>+</sup> + PLL-<span class="html-italic">g</span>-Dex; and (d) MBs-Aptamer/tDNA + H1 + H2 + K<sup>+</sup>+ PLL-<span class="html-italic">g</span>-Dex + Target.</p>
Full article ">Figure 4
<p>Optimization of experimental conditions. (<b>A</b>) Sequence of H1 and H2. (<b>B</b>) Concentration of H1 and H2. (<b>C</b>) Concentration of PLL-<span class="html-italic">g</span>-Dex. (<b>D</b>) ThT concentration. (<b>E</b>) K<sup>+</sup> concentration. (<b>F</b>) Reaction temperature. (<b>G</b>) Reaction time.</p>
Full article ">Figure 5
<p>(<b>A</b>) Fluorescence spectra of different adenosine concentrations (a–l: 0, 25, 50, 100, 150, 200, 300, 400, 500, 600, 1000, and 2000 nM). (<b>B</b>) The correlation between adenosine concentrations and fluorescence signal intensity (the inset shows the linear calibration curve of adenosine vs. fluorescence intensity in the range of 0–600 nM). The error bars represent standard deviations obtained from triplicate experiments.</p>
Full article ">Figure 6
<p>(<b>A</b>) Fluorescence spectrum diagram of the relationship between different types of nucleosides and fluorescence intensity. (<b>B</b>) Diagram of the relationship between the fluorescence intensity difference and different nucleoside types at the excitation wavelength of 420 nm. Statistical significance was calculated by the one-way ANOVA. (**** <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">
30 pages, 3701 KiB  
Review
Structural Unfolding of G-Quadruplexes: From Small Molecules to Antisense Strategies
by Giorgia Fracchioni, Sabrina Vailati, Marta Grazioli and Valentina Pirota
Molecules 2024, 29(15), 3488; https://doi.org/10.3390/molecules29153488 - 25 Jul 2024
Cited by 1 | Viewed by 945
Abstract
G-quadruplexes (G4s) are non-canonical nucleic acid secondary structures that have gathered significant interest in medicinal chemistry over the past two decades due to their unique structural features and potential roles in a variety of biological processes and disorders. Traditionally, research efforts have focused [...] Read more.
G-quadruplexes (G4s) are non-canonical nucleic acid secondary structures that have gathered significant interest in medicinal chemistry over the past two decades due to their unique structural features and potential roles in a variety of biological processes and disorders. Traditionally, research efforts have focused on stabilizing G4s, while in recent years, the attention has progressively shifted to G4 destabilization, unveiling new therapeutic perspectives. This review provides an in-depth overview of recent advances in the development of small molecules, starting with the controversial role of TMPyP4. Moreover, we described effective metal complexes in addition to G4-disrupting small molecules as well as good G4 stabilizing ligands that can destabilize G4s in response to external stimuli. Finally, we presented antisense strategies as a promising approach for destabilizing G4s, with a particular focus on 2′-OMe antisense oligonucleotide, peptide nucleic acid, and locked nucleic acid. Overall, this review emphasizes the importance of understanding G4 dynamics as well as ongoing efforts to develop selective G4-unfolding strategies that can modulate their biological function and therapeutic potential. Full article
(This article belongs to the Section Chemical Biology)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of G-quadruplex structures. (<b>A</b>) Square-planar G-quartet; (<b>B</b>) backbone of the intramolecular G4 structures, common to every G4s, with loops and flanking regions differentiating structural elements; and (<b>C</b>) parallel-, antiparallel-, and hybrid-type topologies of G4 structures.</p>
Full article ">Figure 2
<p>Chemical structures of TMPyP4; its positional isomers TMPyP2 and TMPyP3; its metal-complexes ZnTMPyP4, CuTMPyP4, or PtTMPyP4; and H<sub>2</sub>TCPPSpm4 derivative.</p>
Full article ">Figure 3
<p>Chemical structures of the Ru(II) complexes RuPDC3, RuS, and RuSe, and of NMM-Cisplatin combination, Eu 15-MC-5, and Tb 15-MC-5.</p>
Full article ">Figure 4
<p>Chemical structures of the main small molecules capable of unfolding G-quadruplex structures. For the two rotamers ThT and TO the rotation is evidenced by blue arrows.</p>
Full article ">Figure 5
<p>Chemical structures of the ligands that alter their affinity and unfolding capacity in response to G4’s changing experimental circumstances.</p>
Full article ">Figure 6
<p>Chemical modifications of various ASOs, 2′-OMe, MOE, PMO, LNA, and PNA.</p>
Full article ">Figure 7
<p>Schematic representation of the oligonucleotide-based strategy to disrupt G-quadruplex structures. The best-performing strategy lies in hybridizing the central loop and the two adjacent G-tracts on both sides. Blue lines represent G-tracts; red lines represent the central loop in the G4 structure. The complementary bases in the Anti-G4 ASO were represented with the same colors.</p>
Full article ">Figure 8
<p>Schematic representation of invasion of TBA quadruplex by complementary 7-mer PNA (P1), and the effect of ionic strength on thermal denaturation pathway of TBA-P1 hetero duplex.</p>
Full article ">Figure 9
<p>Schematic representation of PNA<sub>2</sub>-DNA heteroduplex and heteroquadruplex formed by Myc19 DNA quadruplex and complementary or homologous 7-mer PNA.</p>
Full article ">Figure 10
<p>Schematic representation of the mechanism htelo-fl unfolding with complementary 13-mer PNA.</p>
Full article ">Figure 11
<p>Schematic representation of the comparison of the cations’ effects (NH<sub>4</sub><sup>+</sup> versus K<sup>+</sup>) on the unfolding of cKit87up quadruplex in the presence of P2–P4 PNAs.</p>
Full article ">Figure 12
<p>Schematic representation of the effect of the number of LNA modifications (LNA5 10 modifications, LNA2 5 modification, and control DNA) on the unfolding of a G-quadruplex structure formed by a 22-mer purine-rich sequence of NHE III of the c-MYC. The percentage values represent the amount of duplex formation, consistently with G4-unfolding.</p>
Full article ">Figure 13
<p>Schematic representation of the unfolding of cKIT1 G4 by KIT_LNA3 by targeting the central guanine of each G-tract. The red tract is related to loop bases.</p>
Full article ">
16 pages, 4214 KiB  
Article
Carbazole Derivatives Binding to Bcl-2 Promoter Sequence G-quadruplex
by Agata Głuszyńska, Joanna Kosman, Shang Shiuan Chuah, Marcin Hoffmann and Shozeb Haider
Pharmaceuticals 2024, 17(7), 912; https://doi.org/10.3390/ph17070912 - 9 Jul 2024
Viewed by 575
Abstract
In this study, we used ultraviolet-visible (UV-Vis), fluorescence, and circular dichroism (CD) techniques, as well as molecular modeling, to probe the interactions between carbazole derivatives and the G-quadruplex structure formed in the promoter region of gene Bcl-2. This gene is a rational [...] Read more.
In this study, we used ultraviolet-visible (UV-Vis), fluorescence, and circular dichroism (CD) techniques, as well as molecular modeling, to probe the interactions between carbazole derivatives and the G-quadruplex structure formed in the promoter region of gene Bcl-2. This gene is a rational target for anticancer therapy due to its high expression in a variety of tumors as well as resistance to chemotherapy-induced apoptosis. We employed a sequence with a specific dual G-to-T mutation that may form a mixed-type hybrid G-quadruplex structure in the Bcl-2 P1 promoter region. The three tested carbazole compounds differing in substitution on the nitrogen atom of carbazole interact with the Bcl-2 G-quadruplex by the same binding mode with the very comparable binding affinities in the order of 105 M−1. During absorption and fluorescence measurements, large changes in the ligand spectra were observed at higher G4 concentrations. The spectrophotometric titration results showed a two-step complex formation between the ligands and the G-quadruplex in the form of initial hypochromicity followed by hyperchromicity with a bathochromic shift. The strong fluorescence enhancement of ligands was observed after binding to the DNA. All of the used analytical techniques, as well as molecular modeling, suggested the π–π interaction between carbazole ligands and a guanine tetrad of the Bcl-2 G-quadruplex. Molecular modeling has shown differences in the interaction between each of the ligands and the tested G-quadruplex, which potentially had an impact on the binding strength. Full article
(This article belongs to the Section Medicinal Chemistry)
Show Figures

Figure 1

Figure 1
<p>Chemical structures of carbazole ligands <b>1</b>–<b>3</b>.</p>
Full article ">Figure 2
<p>Structure of G-quadruplex formed by Bcl-2 (5′-GGGCGCGGGAGGAATTGGGCGGG-3′) [PDB id: 2F8U]. Guanine, adenine, cytosine, and thymine are represented as green, blue, red or yellow circles.</p>
Full article ">Figure 3
<p>Spectrophotometric titration of ligand <b>3</b> (6 µM) with G4 Bcl-2 (0–26.5 µM) in Tris–HCl buffer (10 mM, pH 7.2) containing 100 mM KCl (<b>A</b>). Panel (<b>B</b>) shows normalized absorbance variations corresponding to the increasing concentrations of Bcl-2 G-quadruplex at 517, 503, 509 nm for ligands <b>1</b>, <b>2</b> and <b>3</b>, respectively.</p>
Full article ">Figure 4
<p>Fluorescence titration spectra of ligand <b>3</b> (2 µM) with G4 Bcl-2 (0–15.2 µM) in Tris–HCl buffer (10 mM, pH 7.2) containing 100 mM KCl (<b>A</b>); fluorescence binding curves (<b>B</b>) and relative enhancement in the fluorescence intensities (<b>C</b>) of ligands vs. the increasing concentration of G-quadruplex Bcl-2; λ<sub>ex</sub>: <b>1</b>—501 nm, <b>2</b> and <b>3</b>—494 nm.</p>
Full article ">Figure 5
<p>CD spectra of G-quadruplex BCL-2 (5 µM) with increasing amounts of ligands <b>1</b> (<b>A</b>), <b>2</b> (<b>B</b>) and <b>3</b> (<b>C</b>) in Tris–HCl buffer (10 mM, pH 7.2) containing 100 mM KCl.</p>
Full article ">Figure 6
<p>Normalized CD melting profiles of Bcl-2 G-quadruplex at 265 nm (<b>A</b>) and at 293 nm (<b>B</b>), observed both in the absence and presence of 3 equiv. of ligands <b>1</b>–<b>3</b> in 10 mM Tris–HCl buffer (pH 7.2) containing 100 mM KCl.</p>
Full article ">Figure 7
<p>Docked ligands <b>1</b> (green), <b>2</b> (yellow) and <b>3</b> (cyan) on the 5′ external G-tetrad of the hybrid type G-quadruplex. Cartoons represents the side view of the complex. The loop is illustrated as white sticks, and the quartets are drawn in a surface representation.</p>
Full article ">Figure 8
<p>The most stable conformations of ligand <b>1</b> (green), <b>2</b> (yellow) and <b>3</b> (cyan) complexes. The ligands stack on the 5′ terminal tetrad and are sandwiched between T15 from the loop. Additional interactions are observed in the 2F8U/<b>3</b> complex. The illustrations in the bottom row highlight the respective binding sites of the ligands (represented as sticks) on the G-quadruplex (surface representation).</p>
Full article ">Figure 9
<p>Interaction energy computed from the simulations at every frame. The average interaction energy for 2F8U/<b>1</b> complex is −92.95 kcal/mol, for 2F8U/<b>2</b> complex is −104.36 kcal/mol and for 2F8U/<b>3</b> complex is −124.39 kcal/mol.</p>
Full article ">
6 pages, 1312 KiB  
Short Note
5,10,15,20-Tetrakis-(4-(3-carbamoyl-pyridyl)-methylphenyl)porphyrin Bromide
by Giuseppe Satta, Silvia Gaspa, Luisa Pisano, Lidia De Luca and Massimo Carraro
Molbank 2024, 2024(2), M1836; https://doi.org/10.3390/M1836 - 13 Jun 2024
Viewed by 509
Abstract
The synthesis of a new tetracationic porphyrin derivative is described. Contrary to the best known derivatives in the literature, which are derived from 5,10,15,20-tetrakis-4-pyridylporphyrin (TPyP), in this procedure we start from 5,10,15,20-tetrakis-(4-carboxymethoxyphenyl)porphyrin (TPPCOOMe), obtained by the condensation reaction between pyrrole and 4-formylbenzoate. The [...] Read more.
The synthesis of a new tetracationic porphyrin derivative is described. Contrary to the best known derivatives in the literature, which are derived from 5,10,15,20-tetrakis-4-pyridylporphyrin (TPyP), in this procedure we start from 5,10,15,20-tetrakis-(4-carboxymethoxyphenyl)porphyrin (TPPCOOMe), obtained by the condensation reaction between pyrrole and 4-formylbenzoate. The reaction is carried out in refluxed xylene, avoiding the use of halogenated solvents. The final product, 5,10,15,20-tetrakis-(4-(3-carbamoyl-pyridyl)-methylphenyl)porphyrin bromide (P15p), exhibits four cationic portions that make it soluble in water and suitable for G4 stabilization. The choice to synthesize a derivative of TPPCOOMe is based on the idea of having a possible stabilizer that, unlike those obtained from TPyP, shows the cationic moieties farther from the porphyrin core and thus closer to the phosphate groups present on the G4 loops. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Porphyrin core structure. (<b>b</b>) TMPyP4 (<b>c</b>) 5,10,15,20-tetrakis-(4-(3-carbamoyl-pyridyl)-methylphenyl)porphyrin bromide (P15p).</p>
Full article ">Figure 2
<p>Partition test: (<b>a</b>) t = t<sub>0</sub>; (<b>b</b>) t = 3 h.</p>
Full article ">Scheme 1
<p>Synthesis of 5,10,15,20-tetrakis-(4-carbomethoxyphenyl)porphyrin. (a) Salicylic acid, xylene, reflux, 2.5 h.</p>
Full article ">Scheme 2
<p>Synthetic route of 5,10,15,20-tetrakis-(4-bromomethylphenyl)porphyrin TPPCOOMe. (a) LiAlH<sub>4</sub>, THF, reflux, 1.5 h. (b) PBr<sub>3</sub>, THF, room temperature, 18 h.</p>
Full article ">Scheme 3
<p>Synthesis of 5,10,15,20-tetrakis-(4-(3-carbamoyl-pyridyl)-methylphenyl)porphyrin bromide. (a) DMF, 80 °C, 3 h.</p>
Full article ">
11 pages, 1463 KiB  
Perspective
Therapeutic Use of G4-Ligands in Cancer: State-of-the-Art and Future Perspectives
by Sara Iachettini, Annamaria Biroccio and Pasquale Zizza
Pharmaceuticals 2024, 17(6), 771; https://doi.org/10.3390/ph17060771 - 13 Jun 2024
Cited by 2 | Viewed by 956
Abstract
G-quadruplexes (G4s) are guanine-rich non-canonical secondary structures of nucleic acids that were identified in vitro almost half a century ago. Starting from the early 1980s, these structures were also observed in eukaryotic cells, first at the telomeric level and later in regulatory regions [...] Read more.
G-quadruplexes (G4s) are guanine-rich non-canonical secondary structures of nucleic acids that were identified in vitro almost half a century ago. Starting from the early 1980s, these structures were also observed in eukaryotic cells, first at the telomeric level and later in regulatory regions of cancer-related genes, in regulatory RNAs and within specific cell compartments such as lysosomes, mitochondria, and ribosomes. Because of the involvement of these structures in a large number of biological processes and in the pathogenesis of several diseases, including cancer, the interest in G4 targeting has exponentially increased in the last few years, and a great number of novel G4 ligands have been developed. Notably, G4 ligands represent a large family of heterogeneous molecules that can exert their functions by recognizing, binding, and stabilizing G4 structures in multiple ways. Regarding anti-cancer activity, the efficacy of G4 ligands was originally attributed to the capability of these molecules to inhibit the activity of telomerase, an enzyme that elongates telomeres and promotes endless replication in cancer cells. Thereafter, novel mechanisms through which G4 ligands exert their antitumoral activities have been defined, including the induction of DNA damage, control of gene expression, and regulation of metabolic pathways, among others. Here, we provided a perspective on the structure and function of G4 ligands with particular emphasis on their potential role as antitumoral agents. In particular, we critically examined the problems associated with the clinical translation of these molecules, trying to highlight the main aspects that should be taken into account during the phases of drug design and development. Indeed, taking advantage of the successes and failures, and the more recent technological progresses in the field, it would be possible to hypothesize the development of these molecules in the future that would represent a valid option for those cancers still missing effective therapies. Full article
(This article belongs to the Section Medicinal Chemistry)
Show Figures

Figure 1

Figure 1
<p>G-quadruplex structure. Schematic representation of G-quartet (<b>A</b>) and G-quadruplex structures (<b>B</b>).</p>
Full article ">Figure 2
<p>Timeline of G-quadruplex ligands publications. Timeline of publications about G-quadruplex ligands. The grey bars indicate the total amount of publications about G-quadruplex ligands, the yellow bars indicate the number of publications related to cancer. The results, derived from PubMed (<a href="https://pubmed.ncbi.nlm.nih.gov" target="_blank">https://pubmed.ncbi.nlm.nih.gov</a>), are expressed as the number of published papers per year.</p>
Full article ">Figure 3
<p>G-quadruplex ligands. List of representative G4-interacting ligands mentioned in the text. Each G4 ligand was drawn according to chemical structures available in the PubChem database (<a href="https://pubchem.ncbi.nlm.nih.gov" target="_blank">https://pubchem.ncbi.nlm.nih.gov</a>).</p>
Full article ">
14 pages, 5245 KiB  
Article
Synergistic Effects of Metal–Organic Nanoplatform and Guanine Quadruplex-Based CpG Oligodeoxynucleotides in Therapeutic Cancer Vaccines with Different Tumor Antigens
by Xia Li, Mitsuhiro Ebara, Naoto Shirahata, Tomohiko Yamazaki and Nobutaka Hanagata
Vaccines 2024, 12(6), 649; https://doi.org/10.3390/vaccines12060649 - 11 Jun 2024
Viewed by 824
Abstract
Oligodeoxynucleotides (ODNs) containing unmethylated cytosine–phosphate–guanosine (CpG) motifs are readily recognized by Toll-like receptor 9 on immune cells, trigger an immunomodulatory cascade, induce a Th1 -biased immune milieu, and have great potential as an adjuvant in cancer vaccines. In this study, a green one-step [...] Read more.
Oligodeoxynucleotides (ODNs) containing unmethylated cytosine–phosphate–guanosine (CpG) motifs are readily recognized by Toll-like receptor 9 on immune cells, trigger an immunomodulatory cascade, induce a Th1 -biased immune milieu, and have great potential as an adjuvant in cancer vaccines. In this study, a green one-step synthesis process was adopted to prepare an amino-rich metal–organic nanoplatform (FN). The synthesized FN nanoplatform can simultaneously and effectively load model tumor antigens (OVA)/autologous tumor antigens (dLLC) and immunostimulatory CpG ODNs with an unmodified PD backbone and a guanine quadruplex structure to obtain various cancer vaccines. The FN nanoplatform and immunostimulatory CpG ODNs generate synergistic effects to enhance the immunogenicity of different antigens and inhibit the growth of established and distant tumors in both the murine E.G7-OVA lymphoma model and the murine Lewis lung carcinoma model. In the E.G7-OVA lymphoma model, vaccination efficiently increases the CD4+, CD8+, and tetramer+CD8+ T cell populations in the spleens. In the Lewis lung carcinoma model, vaccination efficiently increases the CD3+CD4+ and CD3+CD8+ T cell populations in the spleens and CD3+CD8+, CD3CD8+, and CD11b+CD80+ cell populations in the tumors, suggesting the alteration of tumor microenvironments from cold to hot tumors. Full article
(This article belongs to the Special Issue Cutting-Edge Cancer Vaccines Enhanced by Nanotechnology)
Show Figures

Figure 1

Figure 1
<p>Physicochemical characterization of FN nanoplatform and cancer vaccines. (<b>a</b>) Schematic illustration of the synthesis of FN nanoplatform using 2-aminoterephthalic acid and Fe<sup>3+</sup>. (<b>b</b>) SEM images of FN nanoplatform. (<b>c</b>) FTIR spectra of reactant (2-aminoterephthalic acid) and reaction product (FN). (<b>d</b>) SEM image and EDX mapping analysis of cancer vaccines for E.G7-OVA lymphoma (OVA + ODNs + FN). Uniform distribution of Fe, P, and S elements suggests that model antigen OVA and CpG ODNs are homogeneously adsorbed into FN nanoplatform. (<b>e</b>) Zeta potentials of FN nanoplatform, model antigen OVA and cancer vaccines (OVA + ODNs + FN). (<b>f</b>) DLS analysis of FN nanoplatform and cancer vaccines (OVA + ODNs + FN). (<b>g</b>,<b>h</b>) Absorbance of proteins and concentration of CpG ODNs before and after loading into cancer vaccines for E.G7-OVA lymphoma (OVA + ODNs + FN) and cancer vaccines for Lewis lung carcinoma (dLLC + ODNs + FN).</p>
Full article ">Figure 2
<p>Quantitative analysis of cytokine contents after culture for 1 day (<b>a</b>) or 3 days (<b>b</b>). Data are presented as mean ± SD (n = 3). * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 3
<p>Representative flow cytometry plots (<b>a</b>) and quantitative analysis (<b>b</b>) of MHC I<sup>+</sup> in CD11c<sup>+</sup> cells population in lymph nodes of mice. Data are presented as mean ± SD (n = 3). * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 4
<p>Therapeutic effects in E.G7-OVA lymphoma-bearing mouse model. (<b>a</b>) Schematic diagram of antitumor experimental process; (<b>b</b>) tumor growth curves of mice administered with different formulations. Data are presented as mean ± SD (n = 6). * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 5
<p>Antitumor mechanism analysis in therapeutic E.G7-OVA lymphoma mouse model. (<b>a</b>–<b>c</b>) Representative flow cytometry plots (<b>a</b>,<b>b</b>) and populations (<b>c</b>) of CD4<sup>+</sup>, CD8<sup>+</sup>, and tetramer<sup>+</sup>CD8<sup>+</sup> T cells in spleen at the endpoint. (<b>d</b>) Quantitative analysis of cytokine content in spleen at the endpoint. Data are presented as mean ± SD (n = 3). * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 6
<p>Therapeutic effects in Lewis lung carcinoma-bearing mouse model. (<b>a</b>) Schematic diagram of antitumor experimental process; (<b>b</b>) tumor growth curves of mice administered with different formulations. Data are presented as mean ± SD (n = 4). * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 7
<p>Antitumor mechanism analysis in therapeutic Lewis lung carcinoma mouse model. (<b>a</b>–<b>c</b>) Representative flow cytometry plots (<b>a</b>,<b>b</b>) and populations (<b>c</b>) of CD3<sup>+</sup>CD4<sup>+</sup> and CD3<sup>+</sup>CD8<sup>+</sup> in spleen at the endpoint. (<b>d</b>,<b>e</b>) Quantitative analysis of cytokine content in spleen at the endpoint. Data are presented as mean ± SD (n = 4). * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 8
<p>Antitumor mechanism analysis in therapeutic Lewis lung carcinoma mouse model. Representative flow cytometry plots (<b>a</b>,<b>b</b>) and populations (<b>c</b>) of CD3<sup>+</sup>CD8<sup>+</sup>, CD3<sup>−</sup>CD8<sup>+</sup>, and CD11b<sup>+</sup>CD80<sup>+</sup> cells in tumor sites at the endpoint. Data are presented as mean ± SD (n = 4). * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">
18 pages, 1843 KiB  
Article
G-Quadruplex Forming DNA Sequence Context Is Enriched around Points of Somatic Mutations in a Subset of Multiple Myeloma Patients
by Anna S. Zhuk, Elena I. Stepchenkova, Irina V. Zotova, Olesya B. Belopolskaya, Youri I. Pavlov, Ivan I. Kostroma, Sergey V. Gritsaev and Anna Y. Aksenova
Int. J. Mol. Sci. 2024, 25(10), 5269; https://doi.org/10.3390/ijms25105269 - 12 May 2024
Viewed by 1115
Abstract
Multiple myeloma (MM) is the second most common hematological malignancy, which remains incurable despite recent advances in treatment strategies. Like other forms of cancer, MM is characterized by genomic instability, caused by defects in DNA repair. Along with mutations in DNA repair genes [...] Read more.
Multiple myeloma (MM) is the second most common hematological malignancy, which remains incurable despite recent advances in treatment strategies. Like other forms of cancer, MM is characterized by genomic instability, caused by defects in DNA repair. Along with mutations in DNA repair genes and genotoxic drugs used to treat MM, non-canonical secondary DNA structures (four-stranded G-quadruplex structures) can affect accumulation of somatic mutations and chromosomal abnormalities in the tumor cells of MM patients. Here, we tested the hypothesis that G-quadruplex structures may influence the distribution of somatic mutations in the tumor cells of MM patients. We sequenced exomes of normal and tumor cells of 11 MM patients and analyzed the data for the presence of G4 context around points of somatic mutations. To identify molecular mechanisms that could affect mutational profile of tumors, we also analyzed mutational signatures in tumor cells as well as germline mutations for the presence of specific SNPs in DNA repair genes or in genes regulating G-quadruplex unwinding. In several patients, we found that sites of somatic mutations are frequently located in regions with G4 context. This pattern correlated with specific germline variants found in these patients. We discuss the possible implications of these variants for mutation accumulation and specificity in MM and propose that the extent of G4 context enrichment around somatic mutation sites may be a novel metric characterizing mutational processes in tumors. Full article
(This article belongs to the Special Issue Genetic Variations in Human Diseases)
Show Figures

Figure 1

Figure 1
<p>Percentage of G4 strong context occurrence near mutation sites in different patients and in randomly sampled sequences. The random1 and random2 sets include 2000 randomly selected sequences from genomic intervals corresponding to the All Exon V6+UTR V6 enrichment panel (random2) or Truseq Exome panel (random1). The graph displays the percentage proportion along with the confidence interval for the proportion. The asterisk denotes a statistically significant difference between the proportions of G4 context occurrence around point of somatic mutations in patients and in randomly sampled sequences as determined by a z-test.</p>
Full article ">Figure 2
<p>Mutation signatures observed in the analyzed tumors. (<b>a</b>) Visualization of SBS proportions in each of the analyzed tumors based on SigProfilerAssignment. (<b>b</b>) Visualization of small insertions and deletions (ID) among somatic mutations determined in different patients by SigProfilerAssignment. (<b>c</b>) t-SNE analysis based on SigProfilerAssignment SBS classification, percentage of SBS in each sample used, samples with G4 strong context enrichment are salmon, samples without G4 strong context enrichment are cyan. (<b>d</b>) k-means cluster analysis based on SigProfilerAssignment SBS classification was performed for illustration of similarity between samples; percentage of SBS in each sample used. Mutational signature associations: SBS1—aging, clock-like signature, spontaneous or enzymatic deamination of 5-methylcytosine to thymine; SBS5—aging, clock-like signature, may implicate NER [<a href="#B55-ijms-25-05269" class="html-bibr">55</a>]; SBS6—defective DNA mismatch repair, is very specific to MM with high genomic risk [<a href="#B56-ijms-25-05269" class="html-bibr">56</a>]; SBS7a—DNA damage due to exposure to ultraviolet light; SBS9—activity of activation-induced deaminase (AID) in non-coding regions, mutation pattern found in B-cell cancers that develop after the germinal center stage. This signature results from the off-target activity of AID (normally working during the germinal center phase of the hypermutation of immunoglobulin genes [<a href="#B57-ijms-25-05269" class="html-bibr">57</a>], MMR, and gap repair with participation of DNA polymerase eta); SBS10b—polymerase epsilon exonuclease (POLE-Exo) domain mutations [<a href="#B58-ijms-25-05269" class="html-bibr">58</a>]; SBS11—a mutation pattern similar to that of alkylating agents; SBS12—defective mismatch repair [<a href="#B59-ijms-25-05269" class="html-bibr">59</a>]; SBS15—defective DNA mismatch repair [<a href="#B60-ijms-25-05269" class="html-bibr">60</a>]; SBS17a and b—unidentified etiology, were found in MM [<a href="#B61-ijms-25-05269" class="html-bibr">61</a>]; SBS32—treatment with azathioprine prior to induce immunosuppression, the presence of transcription-coupled nucleotide excision repair activity on damaged DNA [<a href="#B62-ijms-25-05269" class="html-bibr">62</a>]; SBS38—indicating possible secondary harm caused by UV exposure; SBS40b—related to indicators of decreased kidney function; SBS84—activity of AID [<a href="#B62-ijms-25-05269" class="html-bibr">62</a>,<a href="#B63-ijms-25-05269" class="html-bibr">63</a>]; SBS87—thiopurine chemotherapy treatment; SBS88—explore to the colibactin from <span class="html-italic">E. coli</span>-carrying pks pathogenicity island, displays heightened activity during early childhood; SBS19, SBS37, SBS93, SBS94—unknown; SBS45, SBS47, SBS58—possible sequencing artefact. ID1, ID2—indicate DNA mismatch repair deficiency; ID5—possible clock-like signature; ID6—defective homologous recombination repair; ID13—UV exposure; ID23—aristolochic acid exposure; ID4, ID9, ID11, ID12, ID20—unknown.</p>
Full article ">Figure 3
<p>Types of mutations in samples with G4 context enrichment around points of somatic mutations and without G4 context enrichment, classified by the type of context. Standard deviation of a proportion is shown as error bars.</p>
Full article ">Figure 4
<p>Consequence of somatic mutations found in different groups of samples in respect to the G4 context.</p>
Full article ">Figure 5
<p>Germline SNPs associated with MM according to publications, GWAS catalog, and Clinvar. SNPs affect such genes as <span class="html-italic">XRCC5</span>, <span class="html-italic">ULK4</span>, <span class="html-italic">ADH1B</span>, <span class="html-italic">ELL2</span>, <span class="html-italic">NDUFA8</span>, <span class="html-italic">CCND1</span>, <span class="html-italic">SLC28A2</span>, <span class="html-italic">RFWD3</span>, <span class="html-italic">CTC1</span>, <span class="html-italic">TNFRSF13B</span>, <span class="html-italic">KLF2</span>, <span class="html-italic">ZBTB46</span>, <span class="html-italic">MYNN</span>, <span class="html-italic">LRRC34</span>, <span class="html-italic">SMARCD3</span>, <span class="html-italic">ICAM1</span>, <span class="html-italic">SAA4</span>, <span class="html-italic">DCLRE1B</span>, <span class="html-italic">CASP3,</span> and <span class="html-italic">MRTFA</span>. CoMut were used for SNP visualization [<a href="#B64-ijms-25-05269" class="html-bibr">64</a>].</p>
Full article ">
13 pages, 13343 KiB  
Article
Nanoscale Interaction of Endonuclease APE1 with DNA
by Sridhar Vemulapalli, Mohtadin Hashemi, Yingling Chen, Suravi Pramanik, Kishor K. Bhakat and Yuri L. Lyubchenko
Int. J. Mol. Sci. 2024, 25(10), 5145; https://doi.org/10.3390/ijms25105145 - 9 May 2024
Viewed by 748
Abstract
Apurinic/apyrimidinic endonuclease 1 (APE1) is involved in DNA repair and transcriptional regulation mechanisms. This multifunctional activity of APE1 should be supported by specific structural properties of APE1 that have not yet been elucidated. Herein, we applied atomic force microscopy (AFM) to characterize the [...] Read more.
Apurinic/apyrimidinic endonuclease 1 (APE1) is involved in DNA repair and transcriptional regulation mechanisms. This multifunctional activity of APE1 should be supported by specific structural properties of APE1 that have not yet been elucidated. Herein, we applied atomic force microscopy (AFM) to characterize the interactions of APE1 with DNA containing two well-separated G-rich segments. Complexes of APE1 with DNA containing G-rich segments were visualized, and analysis of the complexes revealed the affinity of APE1 to G-rich DNA sequences, and their yield was as high as 53%. Furthermore, APE1 is capable of binding two DNA segments leading to the formation of loops in the DNA–APE1 complexes. The analysis of looped APE1-DNA complexes revealed that APE1 can bridge G-rich segments of DNA. The yield of loops bridging two G-rich DNA segments was 41%. Analysis of protein size in various complexes was performed, and these data showed that loops are formed by APE1 monomer, suggesting that APE1 has two DNA binding sites. The data led us to a model for the interaction of APE1 with DNA and the search for the specific sites. The implication of these new APE1 properties in organizing DNA, by bringing two distant sites together, for facilitating the scanning for damage and coordinating repair and transcription is discussed. Full article
(This article belongs to the Collection Feature Papers in Molecular Nanoscience)
Show Figures

Figure 1

Figure 1
<p>DNA substrates, AFM image, and contour length. (<b>A</b>) The schematic for the G rich-substrates (upper scheme) and the control (bottom). 22 bp G-rich motifs are located at 123–144 bp and at 561–583 bp and are shown in blue. The non-G-rich DNA substrate with 612 bp in length was used as a control. (<b>B</b>). A typical 1 × 1 μm AFM scan of G-rich DNA substrate (<b>C</b>,<b>D</b>) are histograms for the contour length measurements for G-rich DNA substrate and the control, respectively. Each distribution is approximated with single Gaussians built with a bin size of 20 bp. The contour length values in base pairs and standard deviations are indicated for each histogram.</p>
Full article ">Figure 2
<p>AFM image of complexes of APE1 with G-rich DNA complexes (1:1). (<b>A</b>) The AFM image with looped complexes of APE1-G-rich-DNA. Zoomed images of complexes circled in (<b>A</b>) are indicated in (<b>B</b>,<b>C</b>). (<b>B</b>) A set of images with no APE1 bound (frame 1) and non-looped complexes with one APE1 bound (frame 2) and two APE1 bound (frame 3). (<b>C</b>) A set of three looped complexes with different sizes of loops.</p>
Full article ">Figure 3
<p>AFM image of complexes of APE1 with non-G-rich DNA complexes (control substrate). (<b>A</b>) A typical AFM scan with 3 × 3 in size. shows the AFM image with looped complexes of APE1–non-G rich-DNA. (<b>B</b>) and (<b>C</b>) show a few examples of complexes with linear morphology and looped DNA complexes, respectively.</p>
Full article ">Figure 4
<p>Mapping of the APE1 positions on the G-rich-DNA substrate. (<b>A</b>) AFM image of APE1–G rich-DNA complex. The dotted line illustrates the contour length of the short arm measured from the DNA end to the center of the protein. (<b>B</b>) The histogram of APE1 mapping performed over 300 molecules. Vertical green lines correspond to the range of distances from both DNA ends to G-rich motifs, which includes the 22 bp size of the motifs. Locations of APE1 within the 92–156 bp range correspond to the specific binding of the protein.</p>
Full article ">Figure 5
<p>Looped complexes formed by APE1 on the G-rich DNA substrate. (<b>A</b>) The histogram for the loop sizes obtained for 200 looped complexes. Vertical green lines indicate the sizes of loops formed by bridging of two G-rich motifs, which includes their sizes. (<b>B</b>) AFM image showing the looped complex. The loop is indicated using a dotted line. (<b>C</b>) The histogram of the lengths of the long arms. (<b>D</b>) The histogram of the lengths of short arms.</p>
Full article ">Figure 6
<p>The height and volume analysis of the APE1 on G-rich DNA with the non-looped and looped complexes. (<b>A</b>) AFM image of the APE1 protein positioned on linear DNA. Circle highlights the APE1 protein, while dotted line illustrates the short DNA flank. (<b>B</b>) Histograms for height values of the APE1 protein approximated with a Gaussian distribution (1.2 ± 0.20 nm). (<b>C</b>) The histogram of the volume measurements data approximated with a Gaussian distribution (125 ± 45 nm<sup>3</sup>). (<b>D</b>) AFM image of looped complexes of APE1 protein (circled). (<b>E</b>) The histogram for the protein height approximated with a Gaussian distribution (1.1 ± 0.13 nm). (<b>F</b>) The histogram for the protein volume approximated with a Gaussian distribution (130 ± 51 nm<sup>3</sup>).</p>
Full article ">Figure 7
<p>Height and volume measurements for complexes of the APE1 on control DNA substrate with non-looped and looped complexes. (<b>A</b>) and (<b>B</b>) are the histograms for the protein heights and volume, respectively, for non-looped complexes. (<b>C</b>) and (<b>D</b>) are the histograms for the height and volume of APE1, respectively. Each histogram is approximated by single Gaussians with parameters indicated in the plots.</p>
Full article ">Figure 8
<p>Height and volume measurements of the free APE1 protein. (<b>A</b>) AFM images of the free protein with added DNA as a reference. (<b>B</b>,<b>C</b>) are the histograms for the height and volume values built for 100 measurements. The histograms are approximated with Gaussians with parameters indicated in the plots.</p>
Full article ">
Back to TopTop