[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

Journals

Article Types

Countries / Regions

Search Results (76)

Search Parameters:
Keywords = factor Xa inhibitors

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
11 pages, 994 KiB  
Article
Laboratory Evaluation of Interferences Associated with Factor XIa Inhibitors Asundexian and Milvexian in Routine Coagulation Assays
by Gavin T. Buckley, Maeve P. Crowley and James V. Harte
Diagnostics 2024, 14(17), 1991; https://doi.org/10.3390/diagnostics14171991 - 9 Sep 2024
Viewed by 479
Abstract
Direct oral anticoagulants (DOACs) are increasingly used for the treatment of thrombosis. While inhibitors of factor IIa and factor Xa have shown effectiveness, the risk of bleeding remains a significant concern. Recently, direct factor XIa inhibitors—including asundexian and milvexian—have emerged as potential anticoagulation [...] Read more.
Direct oral anticoagulants (DOACs) are increasingly used for the treatment of thrombosis. While inhibitors of factor IIa and factor Xa have shown effectiveness, the risk of bleeding remains a significant concern. Recently, direct factor XIa inhibitors—including asundexian and milvexian—have emerged as potential anticoagulation therapies, based on clinical observations that patients with factor XIa deficiencies seldom present with spontaneous bleeding tendencies. The interferences associated with DOACs in routine and specialised coagulation assays are well-described; however, the interferences associated with emerging FXIa inhibitors are largely uncharacterised. Here, we briefly report the impact of asundexian and milvexian in routine coagulation assays using in vitro plasma-based systems. Asundexian and milvexian induce concentration-dependent prolongations in APTT-based assays with curvilinear regressions, which may be suitable for the measurement of pharmacodynamic effects at peak levels ex vivo. We also report differential sensitivities of APTT-based assays—particularly at higher FXIa inhibitor concentrations—highlighting the clinical need for an extensive evaluation of interferences associated with FXIa inhibitors in coagulation assays. Full article
(This article belongs to the Special Issue Recent Advances in Hematology and Oncology)
Show Figures

Figure 1

Figure 1
<p>Chemical structures of investigational FXIa inhibitors asundexian (<b>a</b>) and milvexian (<b>b</b>). Structure visualized with MolView (Available at: <a href="https://molview.org/" target="_blank">https://molview.org/</a> [Accessed: 22 August 2024]).</p>
Full article ">Figure 2
<p>Interference associated with FXIa inhibitors in routine coagulation assays. The impact of select concentrations of asundexian and milvexian (0–1000 ng/mL) on Dade<sup>®</sup> Innovin<sup>®</sup> PT (<b>a</b>), Thromborel S<sup>®</sup> PT (<b>b</b>), Dade<sup>®</sup> Actin<sup>®</sup> FS APTT (<b>c</b>), Dade<sup>®</sup> Actin<sup>®</sup> FSL APTT (<b>d</b>), Dade<sup>®</sup> Thrombin (<b>e</b>), and INNOVANCE<sup>®</sup> D-dimer (<b>f</b>) assays were determined in human-derived plasma. Data are presented as the mean (bars) with standard deviation, using a scatter dot plot. For each assay, five independent replicates of plasma spiked with asundexian or milvexian, respectively, were assayed (<span class="html-italic">n</span> = 5). Horizontal dashed lines represent normal ranges (<a href="#app1-diagnostics-14-01991" class="html-app">Tables S1 and S2</a>). Abbreviations: APTT, activated partial thromboplastin time; APTT-R, APTT ratio; DDi, D-dimer; FIB, fibrinogen; INR, international normalised ratio; PT, prothrombin time. Probability value (one-way ANOVA with post hoc Dunnett test): ***, <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 3
<p>Interference associated with FXIa inhibitors in APTT-based assays. The impact of increasing concentrations of asundexian and milvexian (0–4000 ng/mL) on APTT clotting time (<b>a</b>,<b>b</b>) and FXI activity (<b>c</b>,<b>d</b>) assayed with Dade<sup>®</sup> Actin<sup>®</sup> FS were determined in human-derived plasma. Data are presented as the mean (closed circles) with standard deviation, using an XY plot. For each assay, five independent replicates of plasma spiked with asundexian or milvexian, respectively, were assayed (N = 5; <span class="html-italic">n</span> = 5). Horizontal dashed lines represent normal ranges (<a href="#app1-diagnostics-14-01991" class="html-app">Tables S5 and S6</a>). Abbreviations: APTT, activated partial thromboplastin time.</p>
Full article ">
13 pages, 1412 KiB  
Review
Cerebral Vein Thrombosis and Direct Oral Anticoagulants: A Review
by Johanna Umurungi, Francesca Ferrando, Daniela Cilloni and Piera Sivera
J. Clin. Med. 2024, 13(16), 4730; https://doi.org/10.3390/jcm13164730 - 12 Aug 2024
Viewed by 781
Abstract
Cerebral venous thrombosis (CVT) is a rare type of cerebrovascular event in which the thrombosis occurs in a vein of the cerebral venous system. The diagnosis could be challenging due to the great clinical variability, but the outcome is favourable in most cases, [...] Read more.
Cerebral venous thrombosis (CVT) is a rare type of cerebrovascular event in which the thrombosis occurs in a vein of the cerebral venous system. The diagnosis could be challenging due to the great clinical variability, but the outcome is favourable in most cases, especially in the case of early diagnosis. Anticoagulant therapy is the core of CVT management and currently consists of heparin in the acute phase followed by vitamin K antagonists (VKAs) in the long term. The ideal duration of anticoagulant therapy is still unclear, and the same criteria for the treatment of extracerebral venous thromboembolism currently apply. In this paper, we reviewed the literature regarding the use of direct oral anticoagulants (DOACs) in CVT since in recent years a considerable number of studies have been published on the use of these drugs in this specific setting. DOACs have already been shown to be equally effective with VKAs in the treatment of venous thromboembolism. In addition to efficacy, DOACs appear to have the same safety profile, being, on the other hand, more manageable, as they do not require close monitoring with continuous personalised dose adjustments. In addition, a further advantage of DOACs over VKAs is the possibility of anticoagulant prophylaxis using a reduced dosage of the drug. In conclusion, although the use of DOACs appears from preliminary studies to be effective and safe in the treatment of CVT, additional studies are needed to include these drugs in the treatment of CVT. Full article
(This article belongs to the Section Hematology)
Show Figures

Figure 1

Figure 1
<p>Anatomy of the cerebral venous system and distribution of CVT [<a href="#B4-jcm-13-04730" class="html-bibr">4</a>].</p>
Full article ">Figure 2
<p>Percentage of complete, partial and non-recanalisations in our study patients.</p>
Full article ">
20 pages, 3270 KiB  
Article
Chemical Synthesis and Structure–Activity Relationship Studies of the Coagulation Factor Xa Inhibitor Tick Anticoagulant Peptide from the Hematophagous Parasite Ornithodoros moubata
by Vincenzo De Filippis, Laura Acquasaliente, Andrea Pierangelini and Oriano Marin
Biomimetics 2024, 9(8), 485; https://doi.org/10.3390/biomimetics9080485 - 12 Aug 2024
Viewed by 917
Abstract
Tick Anticoagulant Peptide (TAP), a 60-amino acid protein from the soft tick Ornithodoros moubata, inhibits activated coagulation factor X (fXa) with almost absolute specificity. Despite TAP and Bovine Pancreatic Trypsin Inhibitor (BPTI) (i.e., the prototype of the Kunitz-type protease inhibitors) sharing a [...] Read more.
Tick Anticoagulant Peptide (TAP), a 60-amino acid protein from the soft tick Ornithodoros moubata, inhibits activated coagulation factor X (fXa) with almost absolute specificity. Despite TAP and Bovine Pancreatic Trypsin Inhibitor (BPTI) (i.e., the prototype of the Kunitz-type protease inhibitors) sharing a similar 3D fold and disulphide bond topology, they have remarkably different amino acid sequence (only ~24% sequence identity), thermal stability, folding pathways, protease specificity, and even mechanism of protease inhibition. Here, fully active and correctly folded TAP was produced in reasonably high yields (~20%) by solid-phase peptide chemical synthesis and thoroughly characterised with respect to its chemical identity, disulphide pairing, folding kinetics, conformational dynamics, and fXa inhibition. The versatility of the chemical synthesis was exploited to perform structure–activity relationship studies on TAP by incorporating non-coded amino acids at positions 1 and 3 of the inhibitor. Using Hydrogen–Deuterium Exchange Mass Spectrometry, we found that TAP has a remarkably higher conformational flexibility compared to BPTI, and propose that these different dynamics could impact the different folding pathway and inhibition mechanisms of TAP and BPTI. Hence, the TAP/BPTI pair represents a nice example of divergent evolution, while the relative facility of TAP synthesis could represent a good starting point to design novel synthetic analogues with improved pharmacological profiles. Full article
(This article belongs to the Special Issue Biomimetic Approaches in Healthcare—Innovations Inspired by Nature)
Show Figures

Figure 1

Figure 1
<p>The amino acid sequence and structural similarity of TAP and BPTI. (<b>A</b>) The amino acid sequence alignment of TAP and BPTI: Conserved residues are indicated in bold. The disulphide bond topology is conserved in the two inhibitors and indicated by plain lines (orange). The secondary structure alignment of TAP and BPTI is also reported. (<b>B</b>) The three-dimensional structure of TAP and BPTI: Ribbon drawing representations are based on the best representative NMR conformers of TAP (1tcp.pdb) and BPTI (1pit.pdb). Helical regions are coloured in red, β-strands are in cyan, while segments of irregular structure are in light grey. The regions involved in trypsin or fXa binding are shown in magenta. N- and C-termini are also indicated. (<b>C</b>) The surface electrostatic potentials of TAP and BPTI: The orientation of the two inhibitors are as in panel B. The surface is coloured according to the electrostatic potential (blue, positive; red, negative) and expressed as kJ/(mol·q), as indicated. Calculations were performed using the APBS programme, run on the coordinates of the best representative conformers in the NMR structure of TAP and BPTI. Protein structure images were generated using the PyMOL ver. 1.3 Molecular Graphics System.</p>
Full article ">Figure 2
<p>Purification and disulphide oxidative renaturation of TAP. (<b>A</b>) RP-HPLC analysis of crude, synthetic TAP after resin cleavage and side chain protecting group removal. (<b>B</b>) RP-HPLC analysis of reduced (R) and oxidised (N) TAP after purification by semi-preparative RP-HPLC. (<b>C</b>) Kinetics of disulphide oxidative renaturation of TAP. Purified TAP (1 mg/mL) with Cys residues in reduced state (R) was allowed to fold under air oxidation conditions, at pH 8.3 in the presence of β-mercaptoethanol (250 μM) (see text). At fixed time points, aliquots (20 μL) were taken, acid quenched, and fractionated by analytical RP-HPLC to estimate folding yields of native TAP (N).</p>
Full article ">Figure 3
<p>Spectroscopic characterisation of N-TAP. (<b>A</b>) Fluorescence spectra of N-TAP. Protein samples (50 µg/mL) were excited at 280 and 295 nm. (<b>B</b>,<b>C</b>) Far-UV (<b>B</b>) and near-UV (<b>C</b>) circular dichroism spectra of N-TAP. Spectra were recorded at protein concentration of 0.1 mg/mL and 1 mg/mL in far- and near-UV region, respectively. All measurements were carried out at 25 ± 0.1 °C in PBS, pH 7.4, and resulting spectra were subtracted for corresponding baselines. (<b>D</b>) Temperature dependence of the relative ellipticity (θ/θ<sub>0</sub>) of TAP (●) and BPTI (<span style="color:#A4A4A4">●</span>). Ellipticity values (θ) of TAP and BPTI (1 mg/mL) were recorded at 289 nm as a function of temperature and normalised by the initial value (θ<sub>0</sub>) measured at 2 °C.</p>
Full article ">Figure 4
<p>Global HDX-MS analysis of N-TAP and BPTI. (<b>A</b>,<b>B</b>) Representative traces of global deuterium uptake of N-TAP (<b>A</b>) and BPTI (<b>B</b>) at increasing H/D exchange times. Proteins (25 μg/mL) were incubated at 25 °C with 95% D<sub>2</sub>O in PBS buffer, pD 7.43, and <span class="html-italic">m</span>/<span class="html-italic">z</span> spectra were taken at increasing labelling times, as indicated. For both N-TAP (<span class="html-italic">m</span>/<span class="html-italic">z</span> = 997.875, average value) and BPTI (<span class="html-italic">m</span>/<span class="html-italic">z</span> = 931.166, average value), multiple charged species at z = +7 were selected for monitoring deuterium uptake. (<b>C</b>) Time-course analysis of %D increase in TAP and BPTI, as indicated. Experimental conditions are those reported in panels (<b>A</b>,<b>B</b>). Data points are average of three different experiments, with error bars as standard deviations (see Methods).</p>
Full article ">Figure 5
<p>Inhibition of FXa amidolytic activity by wild-type synthetic TAP and mutated analogues. (<b>A</b>) Progress curves of pNA release by fXa. FXa solutions were pre-incubated (30 min) with increasing concentration of TAP and reaction was started by addition of chromogenic substrate RGR-pNA. Measurements were carried out at 25 °C in TBS, pH 7.4, containing 0.2 M NaCl and steady-state velocities of pNA release were determined from increase in absorbance at 405 nm. (<b>B</b>) Plot of relative velocities (v<sub>i</sub>/v<sub>0</sub>) of pNA release as function of increasing concentrations of wild-type and synthetic TAP analogues, as indicated. Notably, v<sub>i</sub> and v<sub>0</sub> are steady-state velocities of RGR-pNA hydrolysis in presence and absence of inhibitor, respectively. Data points are average of three independent measurements, with error bars as ±SD. As relevant examples, only fXa inhibition properties of Tyr1β-naphthyl-Ala and Arg3pyridyl-Ala analogues are reported. Data points were analysed according to the competitive tight-binding inhibition model, to yield K<sub>I</sub> values reported in <a href="#biomimetics-09-00485-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 6
<p>The details of the TAP-fXa interaction and the molecular structure of substituting amino acid side chains. (<b>A</b>) A close-up view of the interaction of the N-terminal amino acids 1–3 with the fXa active site. The amino acids of the TAP involved in the interaction with fXa are highlighted in stick and colour coded (carbon, grey; oxygen, red; nitrogen, blue); the catalytic amino acids of fXa are in blue, while those of the substrate specificity sites are coloured magenta. The picture was generated on the crystallographic structure of the TAP-fXa complex (1kig.pdb). (<b>B</b>) The surface electrostatic potential of fXa. The stick representation of the first three residues of TAP is also shown. The surface is coloured according to the electrostatic potential (blue, positive; red, negative) and expressed as kJ/(mol·q), as indicated. Calculations were performed using the APBS programme, run on the coordinates of the TAP-fXa complex, after removing the coordinates of TAP 4–60. Images were generated using the PyMOL ver. 1.3 Molecular Graphics System. (<b>C</b>) The structure of the substituting non-coded amino acids at position 1 and 3 of TAP.</p>
Full article ">Figure 7
<p>Comparison of the three-dimensional structure of TAP and BPTI in the free and proteases-bound form. (<b>A</b>) Ribbon drawing representation of the superposition of the NMR solution structure of free TAP (1tcp, red) with that of TAP in the crystallographic structure of the fXa-TAP complex (1kig, light gold). Ribbon drawing representation of the superposition of the NMR solution structure of free BPTI (1pit, blue) with that of BPTI in the crystallographic structure of the trypsin-BPTI complex (4y0y, light grey). (<b>B</b>) Close-up view of the superposition of the conformation of TAP sequence Tyr1-Asn2-Arg3- in the free (1tcp, red) and fXa-bound state (1kig, light gold).</p>
Full article ">
9 pages, 214 KiB  
Review
Treatment of Superficial Vein Thrombosis: Recent Advances, Unmet Needs and Future Directions
by Marcello Di Nisio, Giuseppe Camporese, Pierpaolo Di Micco, Romeo Martini, Walter Ageno and Paolo Prandoni
Healthcare 2024, 12(15), 1517; https://doi.org/10.3390/healthcare12151517 - 31 Jul 2024
Viewed by 1306
Abstract
Once considered relatively benign, superficial vein thrombosis (SVT) of the lower limbs is linked to deep vein thrombosis (DVT) or pulmonary embolism (PE) in up to one fourth of cases. Treatment goals include alleviating local symptoms and preventing SVT from recurring or extending [...] Read more.
Once considered relatively benign, superficial vein thrombosis (SVT) of the lower limbs is linked to deep vein thrombosis (DVT) or pulmonary embolism (PE) in up to one fourth of cases. Treatment goals include alleviating local symptoms and preventing SVT from recurring or extending into DVT or PE. Fondaparinux 2.5 mg once daily for 45 days is the treatment of choice for most patients with SVT. Potential alternatives include intermediate-dose low-molecular-weight heparin or the direct oral factor Xa inhibitor rivaroxaban, however, these require further evidence. Despite these treatment options, significant gaps remain, including the role of systemic or topical anti-inflammatory agents alone or combined with anticoagulants, and the optimal duration of anticoagulation for patients at varying risk levels. Additionally, the efficacy and safety of factor Xa inhibitors other than rivaroxaban, management of upper extremity SVT, and optimal treatment for SVT near the sapheno-femoral or sapheno-popliteal junctions are not well understood. This narrative review aims to summarize current evidence on anticoagulant treatment for SVT, highlight key unmet needs in current approaches, and discuss how ongoing studies may address these gaps. Full article
16 pages, 940 KiB  
Article
Clinical and Economic Consequences of a First Major Bleeding Event in Patients Treated with Direct Factor Xa Inhibitors in Spain: A Long-Term Observational Study
by Carlos Escobar, Beatriz Palacios, Miriam Villarreal, Martín Gutiérrez, Margarita Capel, Unai Aranda, Ignacio Hernández, María García, Laura Lledó and Juan F. Arenillas
J. Clin. Med. 2024, 13(14), 4253; https://doi.org/10.3390/jcm13144253 - 21 Jul 2024
Viewed by 1220
Abstract
Aims: Our aims were to describe the clinical characteristics, adverse clinical events, healthcare resource utilization (HCRU) and costs of patients with major bleeding during direct Factor Xa inhibitor (FXai) use. Methods: This is a retrospective cohort study that included secondary data from computerized [...] Read more.
Aims: Our aims were to describe the clinical characteristics, adverse clinical events, healthcare resource utilization (HCRU) and costs of patients with major bleeding during direct Factor Xa inhibitor (FXai) use. Methods: This is a retrospective cohort study that included secondary data from computerized health records of seven Spanish Autonomous Communities. Patients with a first major bleeding during treatment with a direct FXai were analyzed during a 3-year period. Results: Of 8972 patients taking a direct FXai, 470 (5.24%) had major bleeding (mean age (SD) 77.93 (9.71) years, 61.06% women). The most frequent indications for using FXais were atrial fibrillation (78.09%) and venous thromboembolism (17.66%). Among those with major bleeding, 88.94% presented with gastrointestinal bleeding, 6.81% intracranial bleeding, 2.13% trauma-related bleeding and 4.26% other major bleeding. Prothrombin complex concentrates were used in 63.19%, followed by transfusion of blood products (20.21%) and Factor VIIa (7.66%). In total, 4.26% of patients died in the hospital due to the first major bleeding. At the study end (after 3-year follow-up), 28.94% of the patients had died, 12.34% had a myocardial infarction and 9.15% an ischemic stroke. At year 3, overall bleeding cost was EUR 5,816,930.5, of which 79.74% accounted for in-hospital costs to treat the bleeding episode. Conclusions: Despite the use of replacement agents being high, major events were common, with a 29% mortality at the end of the follow up, and HCRU and costs were high, evidencing the need for new reversal treatment strategies. Full article
(This article belongs to the Section Hematology)
Show Figures

Figure 1

Figure 1
<p>Flow chart of the study.</p>
Full article ">Figure 2
<p>Cumulative incidence (per 100 patient-years) curves for events and corresponding AUCs. AMI: acute myocardial infarction; CV: cardiovascular.</p>
Full article ">
9 pages, 541 KiB  
Systematic Review
Andexanet Alfa versus Four-Factor Prothrombin Complex Concentrate for the Reversal of Factor Xa (FXa) Inhibitor-Associated Intracranial Hemorrhage: A Systematic Review of Retrospective Studies
by Luan Oliveira Ferreira, Ricardo Andres León Oldemburg, João Monteiro Leitão Filho, Rodrigo Arcoverde Cerveira, Victoria Winkler Vasconcelos, Giovana Escribano da Costa, Roseny dos Reis Rodrigues and Dielly Catrina Favacho Lopes
J. Clin. Med. 2024, 13(11), 3077; https://doi.org/10.3390/jcm13113077 - 24 May 2024
Cited by 1 | Viewed by 1160
Abstract
Background/Objectives: There are limited data on the risks and benefits of using Andexanet alfa (AA) compared with four-factor prothrombin complex concentrate (4F-PCC) for the reversal of factor Xa inhibitor-associated intracranial hemorrhage (ICH). Our aim was to describe a compilation of the information [...] Read more.
Background/Objectives: There are limited data on the risks and benefits of using Andexanet alfa (AA) compared with four-factor prothrombin complex concentrate (4F-PCC) for the reversal of factor Xa inhibitor-associated intracranial hemorrhage (ICH). Our aim was to describe a compilation of the information available in the literature to date. Methods: PubMed, Embase, Web of Science (Clarivate Analytics) and the Cochrane Central Register of Controlled Trials were searched until December 2023. Following the “Preferred Reporting Items for Systematic Reviews and Meta-Analyses (PRISMA)” guidelines, our systematic literature review included studies that were retrospective in design and evaluated both drugs to control bleeding and complications (death and thromboembolic events). Two researchers re-examined the studies for relevance, extracted the data and assessed the risk of bias. No meta-analyses were performed for the results. Results: In this limited patient sample, we found no differences between published articles in terms of neuroimaging stability or thrombotic events. However, some studies show significant differences in mortality, suggesting that one of the AAs may be superior to 4F-PCC. Conclusions: Our qualitative analysis shows that AA has a better efficacy profile compared with 4F-PCC. However, further studies monitoring these patients and a multicenter collaborative network dedicated to this topic are needed. Full article
(This article belongs to the Special Issue Clinical Treatment for Intracerebral Hemorrhage)
Show Figures

Figure 1

Figure 1
<p>PRISMA flow diagram of study screening and selection.</p>
Full article ">
14 pages, 1090 KiB  
Article
Ex Vivo Antiplatelet Effects of Oral Anticoagulants
by Giulia Renda, Valentina Bucciarelli, Giulia Barbieri, Paola Lanuti, Martina Berteotti, Gelsomina Malatesta, Francesca Cesari, Tanya Salvatore, Betti Giusti, Anna Maria Gori, Rossella Marcucci and Raffaele De Caterina
J. Cardiovasc. Dev. Dis. 2024, 11(4), 111; https://doi.org/10.3390/jcdd11040111 - 31 Mar 2024
Viewed by 1340
Abstract
Background: The impact of non-vitamin K antagonist oral anticoagulants (NOACs) on platelet function is still unclear. We conducted a comprehensive ex vivo study aimed at assessing the effect of the four currently marketed NOACs on platelet function. Methods: We incubated blood samples from [...] Read more.
Background: The impact of non-vitamin K antagonist oral anticoagulants (NOACs) on platelet function is still unclear. We conducted a comprehensive ex vivo study aimed at assessing the effect of the four currently marketed NOACs on platelet function. Methods: We incubated blood samples from healthy donors with concentrations of NOACs (50, 150 and 250 ng/mL), in the range of those achieved in the plasma of patients during therapy. We evaluated generation of thrombin; light transmittance platelet aggregation (LTA) in response to adenosine diphosphate (ADP), thrombin receptor-activating peptide (TRAP), human γ-thrombin (THR) and tissue factor (TF); generation of thromboxane (TX)B2; and expression of protease-activated receptor (PAR)-1 and P-selectin on the platelet surface. Results: All NOACs concentration-dependently reduced thrombin generation compared with control. THR-induced LTA was suppressed by the addition of dabigatran at any concentration, while TF-induced LTA was reduced by factor-Xa inhibitors. ADP- and TRAP-induced LTA was not modified by NOACs. TXB2 generation was reduced by all NOACs, particularly at the highest concentrations. We found a concentration-dependent increase in PAR-1 expression after incubation with dabigatran, mainly at the highest concentrations, but not with FXa inhibitors; P-selectin expression was not changed by any drugs. Conclusions: Treatment with the NOACs is associated with measurable ex vivo changes in platelet function, arguing for antiplatelet effects beyond the well-known anticoagulant activities of these drugs. There are differences, however, among the NOACs, especially between dabigatran and the FXa inhibitors. Full article
Show Figures

Figure 1

Figure 1
<p>Thrombin generation parameters after the addition of increasing concentrations of dabigatran, rivaroxaban, apixaban and edoxaban. Dot plots report all single values of endogenous thrombin potential (ETP, panel (<b>A</b>)) and peak of thrombin generation (panel (<b>B</b>)). The horizontal lines represent the mean values. * <span class="html-italic">p</span> &lt; 0.0001, ° <span class="html-italic">p</span> &lt; 0.001, <sup>#</sup> <span class="html-italic">p</span> &lt; 0.01, <sup>§</sup> <span class="html-italic">p</span> &lt; 0.05 vs. the corresponding controls. C = control; 50 = 50 ng/mL; 150 = 150 ng/mL; 250 = 250 ng/mL.</p>
Full article ">Figure 2
<p>Platelet aggregation induced by different agents after the addition of increasing concentrations of dabigatran, rivaroxaban, apixaban and edoxaban. Box and whiskers Tukey plot reporting the median, 25th and 75th percentile and range for values of platelet aggregation in platelet-rich plasma (optical transmittance, in percent of maximum aggregation) induced by ADP (<b>A</b>), thrombin receptor-activating peptide (TRAP) (<b>B</b>), gamma-thrombin (THR) (<b>C</b>) and tissue factor (TF) (<b>D</b>). The asterisk within the bars reports the mean values. * <span class="html-italic">p</span> &lt; 0.0001, <sup>#</sup> <span class="html-italic">p</span> &lt; 0.01, vs. corresponding controls.</p>
Full article ">Figure 3
<p>Serum TXB<sub>2</sub> generation after the addition of increasing concentrations of dabigatran, rivaroxaban, apixaban and edoxaban. Box and whiskers Tukey plot reporting the median, 25th and 75th percentile and range for values of serum TXB<sub>2</sub>. The asterisk within the bars reports the mean values. <sup>#</sup> <span class="html-italic">p</span> &lt; 0.01, <sup>§</sup> <span class="html-italic">p</span> &lt; 0.05 vs. corresponding control.</p>
Full article ">Figure 4
<p>Expression of PAR-1 on platelet surface using flow cytometry without and with the addition of increasing concentrations of dabigatran. Box and whiskers Tukey plot reporting the median, 25th and 75th percentile and range for values of mean fluorescence intensity (MFI) ratio. The «+» within the bars indicate the mean values. ° <span class="html-italic">p</span> &lt; 0.001 vs. control; <sup>§</sup> <span class="html-italic">p</span> &lt; 0.05 vs. control.</p>
Full article ">
11 pages, 7308 KiB  
Article
Changes in Internal Structure and Dynamics upon Binding Stabilise the Nematode Anticoagulant NAPc2
by Elaine Woodward and Brendan M. Duggan
Biomolecules 2024, 14(4), 421; https://doi.org/10.3390/biom14040421 - 30 Mar 2024
Viewed by 894
Abstract
Abnormal blood coagulation is a major health problem and natural anticoagulants from blood-feeding organisms have been investigated as novel therapeutics. NAPc2, a potent nematode-derived inhibitor of coagulation, has an unusual mode of action that requires coagulation factor Xa but does not inhibit it. [...] Read more.
Abnormal blood coagulation is a major health problem and natural anticoagulants from blood-feeding organisms have been investigated as novel therapeutics. NAPc2, a potent nematode-derived inhibitor of coagulation, has an unusual mode of action that requires coagulation factor Xa but does not inhibit it. Molecular dynamics simulations of NAPc2 and factor Xa were generated to better understand NAPc2. The simulations suggest that parts of NAPc2 become more rigid upon binding factor Xa and reveal that two highly conserved residues form an internal salt bridge that stabilises the bound conformation. Clotting time assays with mutants confirmed the utility of the salt bridge and suggested that it is a conserved mechanism for stabilising the bound conformation of secondary structure-poor protease inhibitors. Full article
(This article belongs to the Section Molecular Structure and Dynamics)
Show Figures

Figure 1

Figure 1
<p>NAPc2 inhibition of fVIIa. NAPc2 is shown in magenta, fXa in green, fVIIa in red and TF in blue. In fXa and fVIIa circles represent serine protease domains, rectangles EGF domains and trapezoids γ-carboxyglutamate (Gla) domains. Tissue factor is shown embedded in a lipid membrane via a transmembrane domain (striped rectangle). The rectangular TF domains are fibronectin type III modules and the oval is the cytoplasmic signalling domain. NAPc2 binds to the fXa serine protease domain without blocking the active site. The NAPc2–fXa complex binds to the fVIIa-TF complex on the cell surface using interactions between the fXa Gla domain and the phosphatidylserine head groups of the membrane, thereby delivering NAPc2 to fVIIa where it can block the active site of the protease.</p>
Full article ">Figure 2
<p>RMSDs and total energy during the three simulations. RMSDs of NAPc2 are shown in magenta and of fXa in green. RMSDs were calculated by fitting the backbone heavy atoms to the initial frame of each trajectory. (<b>A</b>) RMSDs of the free proteins. The mean RMSD of free NAPc2 was 3.9 ± 0.5 Å and of free fXa 1.4 ± 0.1 Å (<b>B</b>) RMSDs of the bound proteins. The mean RMSD of bound NAPc2 was 3.8 ± 0.8 Å, and of bound fXa 1.1 ± 0.1 Å (<b>C</b>) Total energy of NAPc2 in magenta, fXa in green, and the NAPc2-fXa complex in dashed magenta and green. The mean total energy of NAPc2 was − 38,500 ± 100 kcal/mol, of fXa − 89,500 ± 200 kcal/mol, and of the NAPc2-fXa complex − 138,400 ± 300 kcal/mol.</p>
Full article ">Figure 3
<p>RMSDs of NAPc2 interface residues in free and bound simulations. (<b>A</b>) Selected frames from the NAPc2 simulation are superimposed on the backbone atoms of the residues forming the NAPc2 β-strand (M75-I78) that binds to fXa in the NAPc2-fXa complex. Backbone atoms in the crystal structure of the NAPc2-fXa complex (2H9E) are shown in CPK colour scheme. The closest frame from the NAPc2 simulation is shown in purple (78.5 ps, RMSD = 2.99 Å), the furthest frame in pink (3.3 ps, RMSD = 3.44 Å), and a frame close to the mean in magenta (78.5 ps, RMSD = 3.19 Å). (<b>B</b>) RMSD of M75-I78 backbone heavy atoms in the NAPc2 simulation to the same atoms in the first frame of the NAPc2-fXa simulation plotted against time. Excluding the first frame with an RMSD of 1.08 Å, the mean RMSD was 3.19 ± 0.07 Å.</p>
Full article ">Figure 4
<p>Correlation analysis of NAPc2 in free and bound simulations. Dynamic cross-correlation matrices were calculated using Cα atoms. A correlation of 1.0 (red) indicates correlated motions, i.e., the atoms move in the same direction. A correlation of −1.0 (blue) indicates anti-correlated motions, i.e., the atoms move in opposite directions. A correlation of 0.0 (green) indicates no correlated motions. Open circles indicate disulfides. (<b>A</b>) Correlation matrix of NAPc2 from the NAPc2 simulation. (<b>B</b>) Correlation matrix of NAPc2 from the NAPc2-fXa simulation.</p>
Full article ">Figure 5
<p>Internal salt bridge found in the NAPc2-fXa simulation. (<b>A</b>) Residues E10 and R58 at 20.0 ps in the NAPc2-fXa simulation are shown in CPK colour scheme. Numbers indicate the distance between the E10 oxygen and R58 hydrogen atoms. (<b>B</b>) The distance between the E10 Oε and R58 Nη atoms plotted against simulation time. The upper, light magenta line is from the NAPc2 simulation while the lower, heavy line is from the NAPc2-fXa simulation. The horizontal black line at 4.0 A indicates the cutoff below which a salt bridge is considered to be present. In the NAPc2 simulation, the salt bridge did not form, whereas in the NAPc2-fXa simulation, the salt bridge was present 70.2% of the time. When the salt bridge was formed, the mean distance between the E10 Oε and R58 Nη atoms was 3.2 ± 0.4 Å. In the free NAPc2 simulation, the mean distance between the E10 Oε and R58 Nη atoms was 16 ± 3 Å.</p>
Full article ">Figure 6
<p>Inhibition of coagulation by wild-type NAPc2 and NAP5 and their E10Q mutants. Clotting times for wild-type NAPc2 are shown in filled magenta squares, for E10Q NAPc2 in open magenta squares, for wild-type NAP5 in filled purple triangles, and for E10Q NAP5 in open purple triangles. Error bars are the standard deviation of at least three measurements.</p>
Full article ">Figure 7
<p>Representative structures from the simulations. Structures were taken every 10 ns from (<b>A</b>) the NAPc2 and (<b>B</b>) NAPc2-fXa simulations. NAPc2 is shown in magenta with the E10 and R58 sidechains in stick form. fXa is shown in green with the serine protease domain in front and the EGF domain at the rear. Structures from the NAPc2-fXa simulation were superimposed on the fXa backbone atoms. Structures from the NAPc2 simulation were superimposed on the NAPc2-fXa structures using the backbone atoms of rigid NAPc2 residues (E8-C24, C46-M75), then translated to the left. The location of the NAPc2 cleavage loop that occupies and inhibits the active site of fVIIa is indicated. Notice how the disordered NAPc2 C-terminus forms a β-strand to bind to fXa and much of the NAPc2 structure becomes more ordered.</p>
Full article ">
19 pages, 2620 KiB  
Article
Clinical Characteristics and Incidence of Hemorrhagic Complications in Patients Taking Factor Xa Inhibitors in Spain: A Long-Term Observational Study
by Carlos Escobar, Beatriz Palacios, Miriam Villarreal, Martín Gutiérrez, Margarita Capel, Ignacio Hernández, María García, Laura Lledó and Juan F. Arenillas
J. Clin. Med. 2024, 13(6), 1677; https://doi.org/10.3390/jcm13061677 - 14 Mar 2024
Cited by 2 | Viewed by 1230
Abstract
Objective. To analyze the clinical characteristics of patients taking Factor Xa inhibitors (FXai), either direct FXai or enoxaparin (only in active cancer patients), and to estimate the incidence of and risk factors for major bleeding during FXai use. Methods. A retrospective cohort study, [...] Read more.
Objective. To analyze the clinical characteristics of patients taking Factor Xa inhibitors (FXai), either direct FXai or enoxaparin (only in active cancer patients), and to estimate the incidence of and risk factors for major bleeding during FXai use. Methods. A retrospective cohort study, which included secondary data from computerized health records of primary care centers and hospitals in seven Spanish Autonomous Communities. Results. 9374 patients were analyzed, with 8972 taking direct FXai and 402 enoxaparin. At baseline, the mean age (SD) was 71.8 (9.4) years, 56.0% were women, 76.3% had hypertension, 33.6% had type 2 diabetes, and 25.5% had heart failure. The most common indication for FXai use was atrial fibrillation (72.3%), followed by venous thromboembolism (22.2%) and non-mechanical cardiac–valve replacement (5.6%). At the end of the follow-up period, the incidence rates of major bleeding overall, gastrointestinal, and intracranial were 10.2, 9.0, and 0.8 per 100 person-years, respectively. The total incidence of fatal major bleeding was 0.5 per 100 person-years. Incidence rates of all bleedings progressively decreased over time, with 62.5% of the first events occurring in the initial three months and reaching 76.8% within six months following initiation of treatment. Only 4.8% of the 1st major bleedings led to death, 2.3% in the case of major gastrointestinal bleeding, and 30.8% after an intracranial bleeding. 65.9% of patients discontinued anticoagulation after experiencing major bleeding. Conclusions. In Spain, patients taking FXai were old and had many comorbidities. Despite incidence rates of major bleeding were high, incidence rates of intracranial and fatal bleedings were low, but more efforts are required due to their relevant clinical impact. Full article
(This article belongs to the Section Cardiology)
Show Figures

Figure 1

Figure 1
<p>Flow chart of the study. DOACs: direct oral anticoagulants; FXai: factor Xa inhibitors.</p>
Full article ">Figure 2
<p>Kaplan–Meier curves for major and fatal bleedings for the overall population, DOACs, and enoxaparin groups. DOACs: direct oral anticoagulants.</p>
Full article ">Figure 3
<p>Kaplan–Meier curves for major and fatal bleedings for DOACs, according to label prescription. DOACs: direct oral anticoagulants.</p>
Full article ">Figure 4
<p>Kaplan–Meier curves for major and fatal gastrointestinal bleedings for the overall population, DOACs, and enoxaparin groups. DOACs: direct oral anticoagulants.</p>
Full article ">Figure 5
<p>Kaplan–Meier curve for major and fatal intracranial bleedings for the overall population, DOACs, and enoxaparin groups. DOACs: direct oral anticoagulants.</p>
Full article ">Figure 6
<p>Risk factors for major bleeds for DOAC patients. * within 180 days prior to the index date. AF: atrial fibrillation; BMI: body mass index; DOACs: direct oral anticoagulants; eGFR: estimated glomerular filtration rate; FXai; Factor Xa inhibitors; MPR: medical possesion ratio; NSAIDs: non-steroidal anti-inflammatory drugs; OR: Odds Ratio; VTE: venous thromboembolism.</p>
Full article ">
17 pages, 7284 KiB  
Article
Novel Thiourea and Oxime Ether Isosteviol-Based Anticoagulants: MD Simulation and ADMET Prediction
by Marcin Gackowski, Mateusz Jędrzejewski, Sri Satya Medicharla, Rajesh Kondabala, Burhanuddin Madriwala, Katarzyna Mądra-Gackowska and Renata Studzińska
Pharmaceuticals 2024, 17(2), 163; https://doi.org/10.3390/ph17020163 - 28 Jan 2024
Viewed by 1478
Abstract
Activated blood coagulation factor X (FXa) plays a critical initiation step of the blood-coagulation pathway and is considered a desirable target for anticoagulant drug development. It is reversibly inhibited by nonvitamin K antagonist oral anticoagulants (NOACs) such as apixaban, betrixaban, edoxaban, and rivaroxaban. [...] Read more.
Activated blood coagulation factor X (FXa) plays a critical initiation step of the blood-coagulation pathway and is considered a desirable target for anticoagulant drug development. It is reversibly inhibited by nonvitamin K antagonist oral anticoagulants (NOACs) such as apixaban, betrixaban, edoxaban, and rivaroxaban. Thrombosis is extremely common and is one of the leading causes of death in developed countries. In previous studies, novel thiourea and oxime ether isosteviol derivatives as FXa inhibitors were designed through a combination of QSAR studies and molecular docking. In the present contribution, molecular dynamics (MD) simulations were performed for 100 ns to assess binding structures previously predicted by docking and furnish additional information. Moreover, three thiourea- and six oxime ether-designed isosteviol analogs were then examined for their drug-like and ADMET properties. MD simulations demonstrated that four out of the nine investigated isosteviol derivatives, i.e., one thiourea and three oxime ether ISV analogs, form stable complexes with FXa. These derivatives interact with FXa in a manner similar to Food and Drug Administration (FDA)-approved drugs like edoxaban and betrixaban, indicating their potential to inhibit factor Xa activity. One of these derivatives, E24, displays favorable pharmacokinetic properties, positioning it as the most promising drug candidate. This, along with the other three derivatives, can undergo further chemical synthesis and bioassessment. Full article
Show Figures

Figure 1

Figure 1
<p>The crystal structure of <span class="html-italic">H. sapiens</span> FXa (red) in complex with a calcium cation (yellow) and an inhibitor; PDB code: 2P16. The inset shows the binding pocket, with key residues highlighted. The binding pocket comprises four subpockets: S1 (constructed by C191, Q192, D194, and I227), S2 (C219), S3 (E147 and G218), and S4 (Y99, F174, W215, G216, and E217).</p>
Full article ">Figure 2
<p>Structures of complexes between ISV derivatives and FXa (red) obtained through docking (<b>A</b>–<b>I</b>). These structures served as the initial configurations for the subsequent MD simulations.</p>
Full article ">Figure 3
<p>RMSD values of protein backbone (red) and ligand (blue) for the simulation of an FXa complex with different ligands (<b>A</b>–<b>I</b>). While the RMSD for the protein backbone remains stable with consistently small values across all simulations, the RMSD of the ligand varies between simulations, indicating distinct stability levels of the complexes.</p>
Full article ">Figure 4
<p>The distance between the CoM of the ligand and the CoM of FXa throughout the simulations (<b>A</b>–<b>I</b>). Typically, the distance remains around 15 to 20 Å, except for the FXa complex with E04. In this case, after 15 ns, the distance increases rapidly, stabilizing at 25 Å. This observation implies the dissociation of the ligand and its potential rebinding at a different site.</p>
Full article ">Figure 5
<p>The RMSD changes of amino acids forming the binding pocket (Y99, E147, F174, C191, Q192, D194, W215, G216, E217, G218, C219, and I227) (<b>A</b>–<b>I</b>). The RMSD for FXa complexes with E10, E20, and E21 remains relatively stable, with low values ranging from 2 to 3 Å. In contrast, other complexes show larger RMSD changes, indicating diverse mobility of the ligand-binding site, dependent on the specific type of ligand bound.</p>
Full article ">Figure 6
<p>Representative ligand conformations bound to FXa. Conformations were chosen through ligand clustering using the GROMOS algorithm with a cutoff of 1.0 Å, based on frames sampled every 100 ps. Conformations found in less than 5% of frames are not shown for clarity. (<b>A</b>): E10, (<b>B</b>): E15, (<b>C</b>): E20, and (<b>D</b>): E24.</p>
Full article ">Figure 7
<p>Snapshots from MD simulations for the most representative conformations of the ligands shown in <a href="#pharmaceuticals-17-00163-f006" class="html-fig">Figure 6</a>. The stick representation depicts the side chains of key residues and ligands. Hydrogen bonds are shown with dashed lines. (<b>A</b>): E10, (<b>B</b>): E15, (<b>C</b>): E20, and (<b>D</b>): E24.</p>
Full article ">Figure 8
<p>Chemical structures of FDA-approved FXa inhibitors.</p>
Full article ">
18 pages, 2476 KiB  
Article
Design, Synthesis, and Evaluation of New Hybrid Derivatives of 5,6-Dihydro-4H-pyrrolo[3,2,1-ij]quinolin-2(1H)-one as Potential Dual Inhibitors of Blood Coagulation Factors Xa and XIa
by Anna A. Skoptsova, Athina Geronikaki, Nadezhda P. Novichikhina, Alexey V. Sulimov, Ivan S. Ilin, Vladimir B. Sulimov, Georgii A. Bykov, Nadezhda A. Podoplelova, Oleg V. Pyankov and Khidmet S. Shikhaliev
Molecules 2024, 29(2), 373; https://doi.org/10.3390/molecules29020373 - 11 Jan 2024
Cited by 1 | Viewed by 1252
Abstract
Cardiovascular diseases caused by blood coagulation system disorders are one of the leading causes of morbidity and mortality in the world. Research shows that blood clotting factors are involved in these thrombotic processes. Among them, factor Xa occupies a key position in the [...] Read more.
Cardiovascular diseases caused by blood coagulation system disorders are one of the leading causes of morbidity and mortality in the world. Research shows that blood clotting factors are involved in these thrombotic processes. Among them, factor Xa occupies a key position in the blood coagulation cascade. Another coagulation factor, XIa, is also a promising target because its inhibition can suppress thrombosis with a limited contribution to normal hemostasis. In this regard, the development of dual inhibitors as new generation anticoagulants is an urgent problem. Here we report the synthesis and evaluation of novel potential dual inhibitors of coagulation factors Xa and XIa. Based on the principles of molecular design, we selected a series of compounds that combine in their structure fragments of pyrrolo[3,2,1-ij]quinolin-2-one and thiazole, connected through a hydrazine linker. The production of new hybrid molecules was carried out using a two-stage method. The reaction of 5,6-dihydropyrrolo[3,2,1-ij]quinoline-1,2-diones with thiosemicarbazide gave the corresponding hydrazinocarbothioamides. The reaction of the latter with DMAD led to the target methyl 2-(4-oxo-2-(2-(2-oxo-5,6-dihydro-4H-pyrrolo[3,2,1-ij]quinolin-1(2H)-ylidene)hydrazineyl)thiazol-5(4H)-ylidene)acetates in high yields. In vitro testing of the synthesized molecules revealed that ten of them showed high inhibition values for both the coagulation factors Xa and XIa, and the IC50 value for some compounds was also assessed. The resulting structures were also tested for their ability to inhibit thrombin. Full article
(This article belongs to the Special Issue Anticoagulant and Antithrombotic Therapy)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Docking pose of <b>3b</b> and <b>3f</b> in an active site of factor Xa.</p>
Full article ">Figure 2
<p>Docking poses of <b>3k</b> and <b>3n</b> in the active site of factor Xa. The blue dotted lines indicate π-stacking. Yellow dotted lines indicate hydrogen bonds.</p>
Full article ">Figure 3
<p><b>3n</b> docked into an active site of FXIa. Green lines and blue lines indicate pi-cation and pi-pi-stacking interaction, respectively.</p>
Full article ">Figure 4
<p><b>3c</b> docked into an active site of FXIa. Blue line indicates pi stacking interaction.</p>
Full article ">Scheme 1
<p>Preparation of 2-oxo-5,6-dihydropyrrolo[3,2,1-<span class="html-italic">ij</span>]quinolin-1-ylidene)hydrazineyl)-4-oxothiazol-5-ylidene)acetates <b>3a-q</b>. Reagents and conditions: (i) thiosemicarbazide, MeOH/HCl, reflux, 1 h; (ii) DMAD, MeOH/AcOH (<span class="html-italic">v</span>/<span class="html-italic">v</span> 4:1), reflux, 30–60 min.</p>
Full article ">
12 pages, 1495 KiB  
Article
Effectiveness of Early Direct Oral Anticoagulant Monotherapy within One Year of Coronary Stent Implantation in Patients with Atrial Fibrillation: A Nationwide Population-Based Study
by Youmi Hwang, Soyoon Park, Soohyun Kim, Sung-Hwan Kim, Yong-Seog Oh, Kiyuk Chang and Young Choi
J. Clin. Med. 2023, 12(23), 7487; https://doi.org/10.3390/jcm12237487 - 4 Dec 2023
Viewed by 1085
Abstract
We evaluated the effectiveness of early direct oral anticoagulant (DOAC) monotherapy within one year after percutaneous coronary intervention (PCI) in patients with atrial fibrillation (AF) using Korean National Health Insurance Service data. AF patients who underwent PCI were included and divided into the [...] Read more.
We evaluated the effectiveness of early direct oral anticoagulant (DOAC) monotherapy within one year after percutaneous coronary intervention (PCI) in patients with atrial fibrillation (AF) using Korean National Health Insurance Service data. AF patients who underwent PCI were included and divided into the DOAC monotherapy group and the combination therapy group (DOAC with an antiplatelet agent) based on the medications used at 6 months after PCI. A major adverse cardiovascular event (MACE) was defined as a composite of cardiovascular death, acute myocardial infarction (AMI), stroke, or systemic thromboembolic event between 6 and 12 months after PCI. In the overall study population, the DOAC dose reduction rate was high in both the monotherapy group (70.8%) and the combination therapy group (79.1%). After propensity score matching, the MACE incidence was not significantly different between the two groups (hazard ratio [HR] 1.42 [0.90–2.24]). The numerical trend for higher MACE in the monotherapy group was mainly driven by the difference in stroke incidence (HR 1.84 [0.97–3.46]). All-cause death (HR 1.29 [0.61–2.74] or the incidence of major bleeding (HR 1.07 [0.49–2.35]) results were similar in the two groups. In conclusion, early DOAC monotherapy was not significantly associated with MACE risk between 6 and 12 months after PCI. Full article
(This article belongs to the Section Cardiovascular Medicine)
Show Figures

Figure 1

Figure 1
<p>Flow chart for patient selection. AF = atrial fibrillation; ESRD = end-stage renal disease; DOAC = direct acting oral anticoagulant; PCI = percutaneous coronary intervention.</p>
Full article ">Figure 2
<p>Clinical outcomes between 6 and 12 months after PCI in the propensity score-matched cohort. (<b>A</b>) Freedom from MACE, (<b>B</b>) freedom from all-cause mortality, (<b>C</b>) freedom from major bleeding (bleeding event requiring hospitalization, and (<b>D</b>) freedom from any bleeding into CAS. MACE = major adverse cardiovascular event; PCI = percutaneous coronary intervention; CAS = critical anatomical site.</p>
Full article ">Figure 3
<p>Subgroup analysis for risk of major adverse cardiac events. PCI = percutaneous coronary intervention; DOAC = direct oral anticoagulant.</p>
Full article ">Figure 4
<p>Per-protocol treatment comparison between 6 and 12 months after PCI in the propensity-score-matched cohort. (<b>A</b>) MACE-free survival, and (<b>B</b>) freedom from major bleeding event. MACE = major adverse cardiovascular event; PCI = percutaneous coronary intervention.</p>
Full article ">
10 pages, 1678 KiB  
Article
Usefullness of Heparin Calibrated Anti-Xa Activity to Assess Anticoagulant Activity of Apixaban and Rivaroxaban in Emergency Patients Scheduled for Acute Interventions
by Nada Riahi, Laurence Rozen and Anne Demulder
J. Clin. Med. 2023, 12(21), 6785; https://doi.org/10.3390/jcm12216785 - 26 Oct 2023
Cited by 2 | Viewed by 1069
Abstract
(1) Background: Direct oral anticoagulants (DOACs) require monitoring in some critical clinical situations. The specific tests for DOAC monitoring are not yet available in all labs. The aim of this study was to evaluate if a unique, more widespread heparin-calibrated anti-Xa assay could [...] Read more.
(1) Background: Direct oral anticoagulants (DOACs) require monitoring in some critical clinical situations. The specific tests for DOAC monitoring are not yet available in all labs. The aim of this study was to evaluate if a unique, more widespread heparin-calibrated anti-Xa assay could be suitable to estimate the concentrations of apixaban and rivaroxaban in order to establish an algorithm helping our clinicians in their therapeutic decision for patients treated with DOACs in emergencies. (2) Methods: A first retrospective part allowed us to determine of a conversion factor between the measured DOAC concentration and the deducted anti-Xa heparin activity based on optic density. During the second prospective part, both DOAC concentration (ng/mL) and anti-Xa activity heparin (UI/mL) were measured on the same sample, and the previously determined conversion factor was applied to each UI/mL value. We then compared the calculated and measured DOAC concentration values. (3) Results: The analysis of the derivation cohort confirmed a good correlation, especially between the anti-Xa heparin activity and the apixaban concentrations (r = 0.97). Additionally, we determined heparin-calibrated anti-Xa assay cut-offs for invasive procedures at 0.3 UI/mL and for intravenous thrombolysis at 0.51 UI/mL using ROC curves with a sensitivity at 98% and specificity at 95% for 0.3 UI/mL and a sensitivity at 97.7% and specificity at 88.2% for the cut-off of 0.51 UI/mL. In the validation cohort, we confirmed the agreement between measured and calculated DOAC concentrations for the low values, especially around cut-offs with an excellent negative predictive value for 0.51 UI/mL (94% for apixaban and 100% for rivaroxaban) and a good negative predictive value for 0.3 UI/mL (83.3% for apixaban and 85.7% for rivaroxaban). (4) Conclusions: Our results confirm that it is possible to correctly predict or exclude the presence of apixaban/rivaroxaban in emergency situations when specific tests are not readily available. Full article
(This article belongs to the Section Clinical Laboratory Medicine)
Show Figures

Figure 1

Figure 1
<p>Passing Bablok regression comparing heparin anti-Xa activity (UI/mL) and concentrations of (<b>a</b>) apixaban low (ng/mL) and (<b>b</b>). rivaroxaban low (ng/mL).</p>
Full article ">Figure 2
<p>ROC curves for the determination of cut-offs (“hemostatic” security) expressed in heparin anti-Xa activity units (UI/mL) (<b>a</b>) for 30 ng/mL and (<b>b</b>) for 50 ng/mL.</p>
Full article ">Figure 3
<p>Passing Bablok regression curve and Bland–Altmann plot for the comparison of measured and calculated apixaban.</p>
Full article ">Figure 4
<p>Passing Bablok regression curve and Bland–Altmann plot for the comparison of measured and calculated rivaroxaban.</p>
Full article ">Figure 5
<p>Algorithm of decision in pre-invasive clinical situations for patients treated with apixaban or rivaroxaban.</p>
Full article ">
14 pages, 20389 KiB  
Article
Molecular Dynamics Simulation Study of the Selective Inhibition of Coagulation Factor IXa over Factor Xa
by Hyun Jung Yoon, Sibsankar Kundu and Sangwook Wu
Molecules 2023, 28(19), 6909; https://doi.org/10.3390/molecules28196909 - 2 Oct 2023
Viewed by 1607
Abstract
Thromboembolic disorders, arising from abnormal coagulation, pose a significant risk to human life in the modern world. The FDA has recently approved several anticoagulant drugs targeting factor Xa (FXa) to manage these disorders. However, these drugs have potential side effects, leading to bleeding [...] Read more.
Thromboembolic disorders, arising from abnormal coagulation, pose a significant risk to human life in the modern world. The FDA has recently approved several anticoagulant drugs targeting factor Xa (FXa) to manage these disorders. However, these drugs have potential side effects, leading to bleeding complications in patients. To mitigate these risks, coagulation factor IXa (FIXa) has emerged as a promising target due to its selective regulation of the intrinsic pathway. Due to the high structural and functional similarities of these coagulation factors and their inhibitor binding modes, designing a selective inhibitor specifically targeting FIXa remains a challenging task. The dynamic behavior of protein–ligand interactions and their impact on selectivity were analyzed using molecular dynamics simulation, considering the availability of potent and selective compounds for both coagulation factors and the co-crystal structures of protein–ligand complexes. Throughout the simulations, we examined ligand movements in the binding site, as well as the contact frequencies and interaction fingerprints, to gain insights into selectivity. Interaction fingerprint (IFP) analysis clearly highlights the crucial role of strong H-bond formation between the ligand and D189 and A190 in the S1 subsite for FIXa selectivity, consistent with our previous study. This dynamic analysis also reveals additional FIXa-specific interactions. Additionally, the absence of polar interactions contributes to the selectivity for FXa, as observed from the dynamic profile of interactions. A contact frequency analysis of the protein–ligand complexes provides further confirmation of the selectivity criteria for FIXa and FXa, as well as criteria for binding and activity. Moreover, a ligand movement analysis reveals key interaction dynamics that highlight the tighter binding of selective ligands to the proteins compared to non-selective and inactive ligands. Full article
(This article belongs to the Section Computational and Theoretical Chemistry)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The binding mode of the co-crystallized ligand (PDB 2P16) in the binding site of FXa. The surface colors are based on hydrophobicity [<a href="#B34-molecules-28-06909" class="html-bibr">34</a>], from blue, which is the most hydrophilic, to white and red, which are the most hydrophobic. S1, S2, S3, and S4 are the binding site subsites. (<b>b</b>) All the ligands in the protein–ligand complex are aligned to FIXa. P1, P2, P3, and P4 are the corresponding regions of the ligands that occupy the binding site subsites.</p>
Full article ">Figure 2
<p>Ligand movement patterns upon binding with FIXa and FXa were evaluated. To assess ligand dynamics, three specific ligand atoms were selected from the P1, P2, and P4 regions of the ligand (blue: nitrogen, red: oxygen, green: chlorine, others: carbon). The atom coordinates were extracted from the entire trajectory of the three replicated 100 ns simulations and are presented in the figure. Each figure shows atomic movement patterns for ligands that are (<b>a</b>) FIXa selective, (<b>b</b>) FXa selective, (<b>c</b>) active for both factors, and (<b>d</b>) inactive for both factors.</p>
Full article ">Figure 3
<p>The average and standard deviation (in brackets) of protein–ligand contacts for ligand binding with FIXa. The average contact was calculated by determining interactions with more than 40% contact in each run for protein–ligand complex systems. Hydrogen bonds are marked in blue and hydrophobic contacts are marked in red.</p>
Full article ">Figure 4
<p>The average and standard deviation (in brackets) of protein–ligand contacts for ligand binding with FXa. The average contact was calculated by determining interactions with more than 40% contact in each run for protein–ligand complex systems. Hydrogen bonds are marked in blue and hydrophobic contacts are marked in red.</p>
Full article ">Figure 5
<p>The 2D chemical structure of the 11 ligands.</p>
Full article ">
20 pages, 3562 KiB  
Article
Intersection of Coagulation and Fibrinolysis by the Glycosylphosphatidylinositol (GPI)-Anchored Serine Protease Testisin
by Marguerite S. Buzza, Nisha R. Pawar, Amando A. Strong and Toni M. Antalis
Int. J. Mol. Sci. 2023, 24(11), 9306; https://doi.org/10.3390/ijms24119306 - 26 May 2023
Cited by 2 | Viewed by 1587
Abstract
Hemostasis is a delicate balance between coagulation and fibrinolysis that regulates the formation and removal of fibrin, respectively. Positive and negative feedback loops and crosstalk between coagulation and fibrinolytic serine proteases maintain the hemostatic balance to prevent both excessive bleeding and thrombosis. Here, [...] Read more.
Hemostasis is a delicate balance between coagulation and fibrinolysis that regulates the formation and removal of fibrin, respectively. Positive and negative feedback loops and crosstalk between coagulation and fibrinolytic serine proteases maintain the hemostatic balance to prevent both excessive bleeding and thrombosis. Here, we identify a novel role for the glycosylphosphatidylinositol (GPI)-anchored serine protease testisin in the regulation of pericellular hemostasis. Using in vitro cell-based fibrin generation assays, we found that the expression of catalytically active testisin on the cell surface accelerates thrombin-dependent fibrin polymerization, and intriguingly, that it subsequently promotes accelerated fibrinolysis. We find that the testisin-dependent fibrin formation is inhibited by rivaroxaban, a specific inhibitor of the central prothrombin-activating serine protease factor Xa (FXa), demonstrating that cell-surface testisin acts upstream of factor X (FX) to promote fibrin formation at the cell surface. Unexpectedly, testisin was also found to accelerate fibrinolysis by stimulating the plasmin-dependent degradation of fibrin and enhancing plasmin-dependent cell invasion through polymerized fibrin. Testisin was not a direct activator of plasminogen, but it is able to induce zymogen cleavage and the activation of pro-urokinase plasminogen activator (pro-uPA), which converts plasminogen to plasmin. These data identify a new proteolytic component that can regulate pericellular hemostatic cascades at the cell surface, which has implications for angiogenesis, cancer biology, and male fertility. Full article
Show Figures

Figure 1

Figure 1
<p>GPI-anchored testisin promotes fibrin generation and degradation. (<b>a</b>) Time course analysis of cell-based fibrin generation. TsWT and Ctl cells were treated with 1 mg/mL fibrinogen and 10 nM prothrombin zymogen at 0 h and turbidity at OD<sub>350</sub> monitored over 24 h. Graph shows a representative experiment of average turbidity ± SEM from quadruplicate wells at the indicated times. (<b>b</b>) TsWT cells significantly accelerate both fibrin generation and its subsequent degradation compared to Ctl cells. Graphs show average turbidity measurements ± SEM at 2 h (<b>left</b> panel) and 16 h (<b>right</b> panel) from 5 independent experiments. ** <span class="html-italic">p</span> &lt; 0.01; **** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 2
<p>GPI-anchored testisin accelerates fibrin polymerization and degradation. Urea-solubilized lysates prepared from cell-based fibrin generation experiments from Ctl cells (<b>a</b>) and TsWT cells (<b>b</b>) were analyzed by reducing SDS-PAGE and Coomassie blue staining for detection of fibrinogen and fibrin species over 24 h. For reference, cell-free control wells that contain fibrinogen alone (Fg, lane 1) or polymerized fibrin (Fg + Thr, lane 2) and polymerized fibrin treated with plasmin (Fg + Thr + Pln, lane 3, generated by treatment of polymerized fibrin with 20 nM plasminogen and 2 nM uPA) are included. As reported [<a href="#B30-ijms-24-09306" class="html-bibr">30</a>], commercial preparations of fibrinogen contain trace amounts of contaminating FXIII, which covalently crosslinks fibrin monomers into insoluble fibrin polymers [<a href="#B19-ijms-24-09306" class="html-bibr">19</a>], so that the conversion of fibrinogen to fibrin by thrombin results in higher-molecular weight cross-linked species and the loss of the fibrinogen monomer α- and γ-chains (<b>a</b>,<b>b</b>, lane 1 vs. lane 2, arrow heads). A small decrease in molecular mass of the Bβ-chain also occurs due to the removal of fibrinopeptide B (<b>a</b>,<b>b</b>, lane 2, asterisk). The molecular masses of unpolymerized fibrinogen subunits (Aα, ~62 kDa; Bβ, ~48 kDa; γ, ~52 kDa), those of polymerized fibrin (γ-γ dimers, α-polymers, and the thrombin-cleaved β-chain), and those of fibrin that has been degraded by plasmin (γ-γ deg and β-deg) are shown. Fibrinogen/fibrin species and fibrin degradation products were identified as reported in [<a href="#B31-ijms-24-09306" class="html-bibr">31</a>,<a href="#B32-ijms-24-09306" class="html-bibr">32</a>]. Quantification of unpolymerized fibrinogen chain monomers and fibrin/fibrin degradation species by densitometry are shown below each gel. Levels of individual species are expressed as the % of the total fibrin(ogen) bands quantified in each sample.</p>
Full article ">Figure 3
<p>Catalytically active testisin accelerates prothrombin-dependent fibrin generation through a FXa-dependent mechanism. (<b>a</b>) Testisin-induced fibrin generation is dependent upon its catalytic activity. TsWT, TsMut, and Ctl cells were treated with 1 mg/mL fibrinogen with 10 nM prothrombin zymogen and turbidity at OD<sub>350</sub> monitored over 5 h. Graph shows average OD<sub>350</sub> from quadruplicate wells ± SEM and is representative of at least 5 independent experiments. (<b>b</b>) SDS-PAGE of urea-solubilized lysates confirms faster generation of fibrin in TsWT cells compared to Ctl and TsMut cells. Lysates taken at 0.5, 1, and 3 h from the experiment (<b>a</b>) stained with Coomassie blue are shown. Shown also are γ-γ dimers (black arrows), monomer fibrinogen γ-chain, α-polymers (white arrows), and the thrombin-cleaved β-chain (*). Control lanes show cell-free fibrinogen alone (Fg, lane 1), fibrinogen and prothrombin (Fg + PT, lane 2), and polymerized fibrin (Fg + Thr, lane 3). (<b>c</b>) TsWT cells stimulate prothrombin activation. Thrombin activity was measured using the Bz-FVR-AMC fluorogenic substrate on cells treated with fibrinogen alone (−PT) or with 10 nM prothrombin (+PT) and fluorescence measured at Ex370 nm/Em450 nm. Graph shows average thrombin activity normalized to Ctl + Fg alone ± SEM at 3 h from 4–6 independent experiments. (<b>d</b>) Compared to thrombin, testisin catalyzes only a low level of proteolytic cleavage of the fluorogenic thrombin substrate Bz-FVR-AMC. Activity (relative fluorescence units, RFU) of α-thrombin (1 nM), rTs (10 nM) or rTs control (Cont., (10 nM) see Materials and Methods) was measured from duplicate wells over 20 min. Graph is representative of 4 independent experiments. (<b>e</b>) Average turbidity measurements at 3 h from 5–6 independent experiments in Ctl, TsWT, and TsMut cells in the presence of fibrinogen, with (+) or without (−) 10 nM prothrombin (PT). (<b>f</b>) Rivaroxaban does not inhibit cell-expressed testisin. Testisin activity was assayed in the presence of 1 mg/mL fibrinogen with (+) or without (−) 50 nM rivaroxaban using the Boc-QAR-AMC substrate. Graph shows average testisin activity at 3 h ± SEM from 3 independent experiments. Data are normalized to Ctl + Fg alone. (<b>g</b>) Rivaroxaban dose-dependently inhibits prothrombin-dependent induction of turbidity mediated by TsWT cells. Turbidity assay in the presence of fibrinogen and prothrombin in which TsWT cells were treated with 20 or 50 nM rivaroxaban or no inhibitor. Turbidity was monitored over 3 h. Graph represents mean ± SEM from triplicate wells. (<b>h</b>) Average turbidity induced by Ctl, TsWT and TsMut cells in the absence (−) or presence (+) of 50 nM rivaroxaban (RIV) at 3 h ± SEM from 3–4 independent experiments. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.005; **** <span class="html-italic">p</span> &lt; 0.001; ns, non-significant.</p>
Full article ">Figure 4
<p>Acceleration of fibrinolysis by GPI-anchored testisin is dependent upon plasmin (ogen). (<b>a</b>) Time course analysis of cell-based fibrinolysis assays. TsWT, TsMut, and Ctl cells were treated with fibrinogen, without (closed symbols) or with 250 µM TXA (open symbols) and fibrin polymerization initiated with thrombin. Turbidity was monitored over 24 h. Data are representative of 3 independent experiments and show average turbidity ± SEM of quadruplicate wells. (<b>b</b>) TXA does not inhibit the activity of cell-expressed testisin, assessed in the presence of 1 mg/mL fibrinogen with (+) or without (−) 250 µM TXA using the peptide substrate Boc-QAR-AMC. Graph shows average activity ± SEM from 5 independent experiments at 3 h after adding substrate and is normalized to Ctl -TXA. (<b>c</b>) Average turbidity at the 24 h endpoint in the absence (−) or presence (+) of 250 µM TXA from 3 independent experiments. Data are expressed as % maximal turbidity reached after addition of thrombin in each experimental treatment. (<b>d</b>) SDS-PAGE confirms faster fibrin degradation by TsWT cells compared to Ctl and TsMut. Urea-solubilized lysates prepared at the 24 h endpoint analyzed by reducing SDS-PAGE stained with Coomassie blue. Control lanes include fibrin alone (Fn, lane 1), fibrin generated in the presence of 20 nM plasminogen (Fn + Plg, lane 2), and fibrin generated in the presence of plasminogen to which 5 nM uPA was added after polymerization to generate plasmin (Fn + Plg + uPA, lane 3). Fibrin degradation is shown by loss of insoluble fibrin (gray triangle) and α-polymers (white triangles). γ-γ dimers (black triangle) and the β-chain are also cleaved, producing γ-γ deg and β-deg fibrin degradation products, respectively (lane 3). Fibrin degradation species were identified as reported in [<a href="#B32-ijms-24-09306" class="html-bibr">32</a>]. Graphs show quantitation of fibrin degradation products by densitometry, expressed as the percentage of the total protein quantified in each sample. (<b>e</b>) Cellular testisin facilitates plasmin(ogen)-dependent cell invasion through fibrin. TsWT, TsMut, and Ctl cells were plated onto fibrin or Matrigel-coated Transwells in serum-free medium with (+) or without (−) 250 µM TXA, and invasion towards 10% serum-containing media assessed. Representative images shown are stitched, whole-well scans of the underside of Transwells showing invaded cells stained with KwikDiff after 10 h invasion (4× original magnification). The two left panels (−TXA) show wells from 2 independent experiments; right panel (+TXA) shows wells from a representative experiment showing inhibition of testisin-mediated fibrin invasion. Images are representative of 3 independent experiments for both conditions. (<b>f</b>) Quantitation of cell invasion analyzed by manual cell counting of all invaded cells/membrane using Image J. TsWT invaded through fibrin ~13-fold faster than Ctl and TsMut cells, but invasion through Matrigel was not enhanced by expression of wildtype testisin. Data show average fold invasion relative to Ctl from 2–3 independent experiments, * <span class="html-italic">p</span> &lt; 0.05; *** <span class="html-italic">p</span> &lt; 0.005; **** <span class="html-italic">p</span> &lt; 0.001; ns, non-significant.</p>
Full article ">Figure 5
<p>Testisin accelerates fibrin degradation but does not activate plasminogen directly. (<b>a</b>) Plasminogen accelerates TsWT-mediated fibrin degradation. Fibrin was generated by the addition of thrombin as in <a href="#ijms-24-09306-f004" class="html-fig">Figure 4</a>a, in the absence (−) or presence (+) of 20 nM plasminogen (Plg). Graph shows average turbidity measured over 6 h from quadruplicate wells ± SEM and is representative of 3 independent experiments. (<b>b</b>) Average turbidity ±SEM at the 6 h endpoint from 3 independent experiments. ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.005; ns, non-significant. (<b>c</b>) Unlike uPA, testisin does not activate the plasminogen zymogen in solution. 1.5 µM plasminogen (Plg) was incubated alone or with 30 nM rTs or uPA (50:1 zymogen-to-activating protease ratio) for 2 h at 37 °C, and reactions analyzed by reducing SDS-PAGE and silver staining. Gel shown is representative of 3 independent experiments and shows activation cleavage of the plasminogen zymogen to release the 55 kDa heavy chain, the 26 kDa serine protease domain (SPD), and the 8.2 kDa pre-activation peptide (PAP). (<b>d</b>) Plasmin activity assays after treatment of plasminogen with rTs or uPA as in (<b>c</b>). Reactions were diluted ~40-fold and incubated with 100 µM Boc-EKK-AMC plasmin substrate, and fluorescence monitored at 1 min for 20 min. Graph shows average relative fluorescence units (RFU) from duplicate wells and is representative of 4 independent experiments.</p>
Full article ">Figure 6
<p>Cell-surface testisin induces pro-uPA activation. (<b>a</b>) Testisin directly activates the pro-uPA zymogen in solution. An amount of 1.5 µM pro-uPA was incubated alone or with 30 nM rTs or plasmin (50:1 zymogen-to-activating protease ratio) for 2 h at 37 °C, and reactions analyzed by reducing SDS-PAGE and Coomassie blue staining. Since the commercial preparation of rTs is in zymogen form and is activated by trace amounts of the metalloprotease thermolysin, reactions were incubated under identical conditions with equivalent final concentrations of thermolysin. Under these conditions, thermolysin does not cleave pro-uPA (<span class="html-italic">Cont</span>), demonstrating the pro-uPA cleavage is mediated specifically by testisin. Gel shown is representative of 4 independent experiments. The ~55 kDa band in the pro-uPA is likely a contaminating <span class="html-italic">Drosophila</span> protein that sometimes co-purifies with the recombinant pro-uPA produced in S2 cells as reported in [<a href="#B45-ijms-24-09306" class="html-bibr">45</a>]. (<b>b</b>) uPA activity assay of reactions in (<b>a</b>) using the Glt-GR-AMC uPA substrate. Graph shows average relative fluorescence units (RFU) from duplicate wells over 20 min and is representative of 3 independent experiments. (<b>c</b>) TsWT cells show increased activation of endogenous pro-uPA compared to Ctl and TsMut cells. Cell lysates and 10x-concentrated 24 h-conditioned media (Cond. Media) were immunoblotted for uPA, testisin, and β-tubulin loading control by stripping and reprobing. Purified recombinant pro-uPA (5 ng, 50 kDa) and active uPA proteins (10 ng, 33 kDa) were resolved in the first two lanes as molecular mass controls. Panels separating the blots indicate lanes in between samples that have been spliced out. For the uPA blot, Cond. Media is a shorter exposure than other two panels to enable comparison between Ctl, TsWT, and TsMut. Data are representative of 3 independent experiments. (<b>d</b>) TsWT-conditioned media has high uPA activity compared to that of Ctl and TsMut cells, which is not altered by the presence of the plasmin(ogen) inhibitor TXA. Conditioned media was collected from cells incubated for 24 h with (+) or without (−) 250 µM TXA and assayed for uPA activity using the Glt-GR-AMC uPA substrate. Graph shows average uPA activity ± SEM at 20 min from 3 independent experiments expressed relative to Ctl media, **** <span class="html-italic">p</span> &lt; 0.001, ns, non-significant.</p>
Full article ">Figure 7
<p>Zymogen activation pathways activated by membrane-anchored testisin that promote coagulation and fibrinolysis. GPI-anchored testisin promotes thrombin activation via stimulating the activation of the prothrombin activator FXa, which may occur via activation of an upstream protease of the intrinsic or extrinsic coagulation pathways or via a novel mechanism. Generation of active thrombin cleaves fibrinogen to induce its polymerization into fibrin. GPI-anchored testisin promotes plasmin generation via the activation of cell-surface pro-uPA, which then activates plasminogen to degrade polymerized fibrin. Shown also are inhibitors of FXa rivaroxaban (RIV) and plasminogen (TXA). Schematic is simplified for clarity.</p>
Full article ">
Back to TopTop