scholarly journals Fibrinolysis of Contracted Blood Clots Depends on Whether Plasminogen Activator Acts from inside or Outside

Blood ◽  
2018 ◽  
Vol 132 (Supplement 1) ◽  
pp. 3773-3773
Author(s):  
Valerie Tutwiler ◽  
Alina D. Peshkova ◽  
Giang Le Minh ◽  
Sergei Zaitsev ◽  
Rustem I. Litvinov ◽  
...  

Abstract Fibrinolysis involves the dissolution of polymeric fibrin networks that is required to restore blood flow through vessels obstructed by clots and thrombi. The efficiency of lysis depends on the susceptibility of fibrin to enzymatic digestion, which is governed by the structure and spatial organization of fibrin fibers as well as porosity and composition of the clot. Platelet-driven clot contraction results in compaction of the erythrocytes into the core of the clot, effectively reducing the permeability of the clot, and influences fibrin network structure. We have shown that clot contraction is reduced in blood from patients with thrombotic conditions such ischemic stroke and deep vein thrombosis, which points to the clinical importance of understanding the influence of clot contraction on efficacy of fibrinolysis. Here, we examined the effects of clot contraction on the rate of internal fibrinolysis emanating from within the clot to simulate (patho)physiological conditions, and external fibrinolysis initiated from the clot exterior to simulate therapeutic thrombolysis. Fibrinolysis was induced and the kinetics of lysis was measured in parallel in contracted versus uncontracted clots from the same citrated human blood samples. Clot formation and platelet activation were initiated with 1 U/ml thrombin and 2 mM CaCl2. Clot contraction was either unaffected or impaired by inhibiting platelet non-muscle myosin IIa (blebbistatin), actin polymerization (latrunculin A), and platelet-fibin(ogen) binding (abciximab). To examine internal fibrinolysis, 75 ng/ml of human recombinant tissue plasminogen activator (tPA) was added prior to initiation of clotting, allowing for tPA to be uniformly distributed through the clot volume and for fibrinolysis occur after the clot has formed. We used optical tracking to follow clot size in a time dependent manner. Contracted clots were completely lysed at a rate that was at least 2 times faster than clots with impaired contraction. Specifically, the average time to complete lysis was 33±4 minutes for contracted clots versus 59±3, 84±4, 75±3 minutes when contraction was impaired by blebbistatin, latrunculin A, and abciximab, respectively (p<0.001). To examine external fibrinolysis, blood spiked with purified human 125I-fibrinogen was allowed to clot and contract (unless contraction was inhibited) prior to the addition of 75 ng/ml tPA. Clots with impaired contraction released 2-4-fold more radiolabeled soluble degradation products during the first 30 minutes and continued to lyse at a rate 4-fold faster than contracted clots over the initial 4 hours following addition of tPA. This reduction of the fibrinolysis rate in contracted clots was not due to the expulsion of serum-soluble anti-fibrinolytic compounds during the contraction process because serum replacement with a buffer did not affect the lysis rate. This difference in the susceptibility of contracted and uncontracted clots to internal versus external lysis suggests that the lysis rate is dominated by the interplay of clot permeability to fibrinolytic enzymes and the spatial proximity of the fibrin fibers themselves. Despite limitations of in vitro experimental models, numerous studies on fibrinolysis have demonstrated the relevance of experimental findings to pathophysiological fibrinolysis and therapeutic thrombolysis. Enhancement of fibrinolysis in contracted blood clots is consistent with the need to dissolve mature clots once they have performed their hemostatic function in a vessel on in a wound. The reduced rates of dissolution of contracted clots in our model of externally applied tPA could account for the inefficacy of therapeutic thrombolysis of old thrombi that likely underwent more compaction compared to newer thrombi. Our studies point to the clinical importance of understanding how mechanical remodeling of clots and thrombi may influence their fibrinolytic resolution and could inform the development of improved thrombolytic therapies. This work, in part, was supported by the Program for Competitive Growth at KFU. Disclosures No relevant conflicts of interest to declare.

Blood ◽  
2011 ◽  
Vol 118 (21) ◽  
pp. 2236-2236
Author(s):  
Rocco Romagnuolo ◽  
Michael B Boffa ◽  
Marlys L Koschinsky

Abstract Abstract 2236 Lipoprotein(a) [Lp(a)] has been identified as an independent risk factor for cardiovascular diseases such as coronary heart disease. Lp(a) levels vary over 1000-fold within the human population and Lp(a) possesses both proatherogenic and prothrombotic properties due to the LDL-like moiety and apolipoprotein(a) [apo(a)] components, respectively. Apo(a) is highly homologous to plasminogen and thus can potentially interfere with plasminogen activation. Plasmin generated in the context of fibrin mediates the breakdown of blood clots, which are the causative factors in heart attacks and strokes. Plasmin generated on the surface of vascular cells plays a role in cell migration and proliferation, two of the fibroproliferative inflammatory events that underlie atherosclerosis. Previous studies have suggested that apo(a) may inhibit pericellular plasminogen activation on the basis of observations that apo(a) decreases plasminogen binding to cells. We have undertaken analysis of the mechanism by which apo(a) may interfere with pericellular plasminogen activation to allow for a more definitive description of the role of Lp(a) within the vasculature. Plasminogen activation was found to be markedly inhibited by the recombinant apo(a) variant 17K, in a dose dependent manner, on human umbilical vein endothelial cells (HUVECs), human monocytic leukemia cells (THP-1), THP-1 macrophages, and smooth muscle cells. The strong lysine binding site in kringle IV type 10, as well as kringle V appear to be required for this effect since apo(a) variants lacking these elements (17KΔAsp and 17KΔV, respectively) failed to inhibit activation. However, the role of lysine-dependent binding of apo(a) itself to the cells is not clear. Carboxypeptidase treatment of cells did not decrease apo(a) binding, and apo(a) does not compete directly for plasminogen binding to the cells. Rather, apo(a) and plasminogen may bind to the cells as a complex. We next attempted to identify the cell-surface receptor(s) that mediate plasminogen activation on the cell surface as well as its inhibition by apo(a). Urokinase-type plasminogen activator receptor (uPAR) has been previously shown to bind to urokinase-type plasminogen activator (uPA), vitronectin, and β3 integrins. uPAR is involved in the remodeling of the extracellular matrix (ECM) through regulation of plasminogen activation. We found evidence that uPAR is a potential receptor for both plasminogen and apo(a). Knockdown of uPAR in HUVECs results in decreased binding of plasminogen, 17K and, to a lesser extent, 17KΔAsp and 17KΔV. Similar experiments in SMCs revealed no changes in binding. A decrease in tPA-mediated plasminogen activation following uPAR knockdown occurred in HUVECs, and addition of 17K did not result in any further decrease. Overexpression of uPAR in THP-1 macrophages leads to greater than a two fold increase in 17K and plasminogen binding. Plasminogen activation increases over two-fold as a result of overexpression of uPAR, while 17K blunts the effect of uPAR overexpression. These results indicate that uPAR plays a crucial role in both plasminogen and apo(a) binding to the cell surface of specific cells and inhibition by apo(a) of plasminogen activation. Macrophage-1-antigen (Mac-1) receptor consists of CD11b (αM) and CD18 (β2) integrin and has been previously shown to recognize uPA and control migration and adhesion. Furthermore, αVβ3 has been previously shown to bind to vitronectin and the uPA-uPAR complex which promotes cell adhesion through binding of both vitronectin and αVβ3 integrins. We found that blocking the αM, β2, or αVβ3 receptors with monoclonal antibodies in THP-1 cells leads to a decrease in plasminogen activation, as well as a blunting of the inhibitory effects of apo(a) on plasminogen activation. These results indicate a role for Mac-1 and αVβ3 in apo(a) binding and inhibition of plasminogen activation. In conclusion, we have demonstrated, for the first time, the role of specific receptors in binding of apo(a) to vascular cell surfaces and in mediating the inhibitory effect of apo(a) on pericellular plasminogen activation. Disclosures: No relevant conflicts of interest to declare.


1999 ◽  
Vol 82 (11) ◽  
pp. 1497-1503 ◽  
Author(s):  
Hajime Tsuji ◽  
Hiromi Nishimura ◽  
Haruchika Masuda ◽  
Yasushi Kunieda ◽  
Hidehiko Kawano ◽  
...  

SummaryIn the present study, we demonstrate that brain natriuretic peptide (BNP) and C-type natriuretic peptide (CNP) interact with angiotensin II (Ang II) in regulative blood coagulation and fibrinolysis by suppressing the expressions of both tissue factor (TF) and plasminogen activator inhibitor-1 (PAI-1) induced by Ang II. The expressions of TF and PAI-1 mRNA were analyzed by northern blotting methods, and the activities of TF on the surface of rat aortic endothelial cells (RAECs) and PAI-1 in the culture media were respectively measured by chromogenic assay.Both BNP and CNP suppressed the expressions of TF and PAI-1 mRNA induced by Ang II in a time- and concentration-dependent manner via cGMP cascade, which suppressions were accompanied by respective decrease in activities of TF and PAI-1. However, neither the expression of tissue factor pathway inhibitor (TFPI) nor tissue-type plasminogen activator (TPA) mRNA was affected by the treatment of BNP and CNP.


1997 ◽  
Vol 77 (04) ◽  
pp. 725-729 ◽  
Author(s):  
Mario Colucci ◽  
Silvia Scopece ◽  
Antonio V Gelato ◽  
Donato Dimonte ◽  
Nicola Semeraro

SummaryUsing an in vitro model of clot lysis, the individual response to a pharmacological concentration of recombinant tissue plasminogen activator (rt-PA) and the influence on this response of the physiological variations of blood parameters known to interfere with the fibrinolytic/thrombolytic process were investigated in 103 healthy donors. 125I-fibrin labelled blood clots were submersed in autologous plasma, supplemented with 500 ng/ml of rt-PA or solvent, and the degree of lysis was determined after 3 h of incubation at 37° C. Baseline plasma levels of t-PA, plasminogen activator inhibitor 1 (PAI-1), plasminogen, α2-anti-plasmin, fibrinogen, lipoprotein (a), thrombomodulin and von Willebrand factor as well as platelet and leukocyte count and clot retraction were also determined in each donor. rt-PA-induced clot lysis varied over a wide range (28-75%) and was significantly related to endogenous t-PA, PAI-1, plasminogen (p <0.001) and age (p <0.01). Multivariate analysis indicated that both PAI-1 antigen and plasminogen independently predicted low response to rt-PA. Surprisingly, however, not only PAI-1 but also plasminogen was negatively correlated with rt-PA-ginduced clot lysis. The observation that neutralization of PAI-1 by specific antibodies, both in plasma and within the clot, did not potentiate clot lysis indicates that the inhibitor, including the platelet-derived form, is insufficient to attenuate the thrombolytic activity of a pharmacological concentration of rt-PA and that its elevation, similarly to the elevation of plasminogen, is not the cause of clot resistance but rather a coincident finding. It is concluded that the in vitro response of blood clots to rt-PA is poorly influenced by the physiological variations of the examined parameters and that factors other than those evaluated in this study interfere with clot dissolution by rt-PA. In vitro clot lysis test might help to identify patients who may be resistant to thrombolytic therapy.


2021 ◽  
Vol 11 (1) ◽  
Author(s):  
James J. Miller ◽  
Richard N. Bohnsack ◽  
Linda J. Olson ◽  
Mayumi Ishihara ◽  
Kazuhiro Aoki ◽  
...  

AbstractPlasmin is the key enzyme in fibrinolysis. Upon interaction with plasminogen activators, the zymogen plasminogen is converted to active plasmin. Some studies indicate plasminogen activation is regulated by cation-independent mannose 6-phosphate receptor (CI-MPR), a protein that facilitates lysosomal enzyme trafficking and insulin-like growth factor 2 downregulation. Plasminogen regulation may be accomplished by CI-MPR binding to plasminogen or urokinase plasminogen activator receptor. We asked whether other members of the plasminogen activation system, such as tissue plasminogen activator (tPA), also interact with CI-MPR. Because tPA is a glycoprotein with three N-linked glycosylation sites, we hypothesized that tPA contains mannose 6-phosphate (M6P) and binds CI-MPR in a M6P-dependent manner. Using surface plasmon resonance, we found that two sources of tPA bound the extracellular region of human and bovine CI-MPR with low-mid nanomolar affinities. Binding was partially inhibited with phosphatase treatment or M6P. Subsequent studies revealed that the five N-terminal domains of CI-MPR were sufficient for tPA binding, and this interaction was also partially mediated by M6P. The three glycosylation sites of tPA were analyzed by mass spectrometry, and glycoforms containing M6P and M6P-N-acetylglucosamine were identified at position N448 of tPA. In summary, we found that tPA contains M6P and is a CI-MPR ligand.


2019 ◽  
Vol 12 (4) ◽  
pp. 399-407 ◽  
Author(s):  
A. Alizadeh ◽  
P. Akbari ◽  
S. Varasteh ◽  
S. Braber ◽  
H. Malekinejad ◽  
...  

Contamination of human and animal diets with different mycotoxins have gained significant attention over the past decade. The intestinal barrier is the first site of exposure and a primary target for nutritional contaminants and hazardous substances including mycotoxins. In this study, the potential impact of ochratoxin A (OTA) on intestinal barrier integrity was highlighted using a human intestinal Caco-2 cell line. Cell viability following OTA exposure was determined by lactate dehydrogenase release and the 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide (MTT) assay. Moreover, markers of barrier integrity, such as transepithelial electrical resistance (TEER) as well as the permeability of Lucifer Yellow (LY) and fluorescein isothiocyanate (FITC)-dextran, were assessed. Furthermore, the protein expression of different tight junction (TJ) proteins, as main constituents of barrier integrity, was evaluated by Western blot. Results show that OTA reduces TEER values in a concentration- and time-dependent manner and increase the permeability of LY through the intestinal epithelial layer, while the cell viability did not change significantly. However, the damage was not severe enough to change the permeability to larger molecules, such as FITC-dextran. OTA exposure down-regulated the expression of TJ proteins claudin-1, -3 and -4 and up-regulated the expression of zona occludens 1. The observation that OTA can disrupt the epithelial barrier is of clinical importance as it may lead to an increased passage of luminal antigens into the systemic circulation.


2015 ◽  
Vol 35 (suppl_1) ◽  
Author(s):  
Neha Goyal ◽  
Zhen Weng ◽  
Philip Fish ◽  
Tammy Strawn ◽  
Samantha Myears ◽  
...  

Introduction: Plasminogen activator inhibitor-1 (PAI-1) is the primary inhibitor of mammalian plasminogen activators and an important regulator of cell migration. We have shown that tiplaxtinin, a small molecule, specific inhibitor of PAI-1, inhibits intimal hyperplasia in a murine vein graft model. However, little is known about the effects of pharmacological inhibition of PAI-1 on vascular cell migration under physiologically relevant conditions. Methods: We studied the effects of tiplaxtinin on migration of smooth muscle cells (SMCs) and endothelial cells (ECs). Results: Tiplaxtinin significantly inhibited migration of murine SMCs through 3-dimensional (3-D) collagen matrix in a concentration-dependent manner. Tiplaxtinin did not inhibit SMC proliferation, and it did not inhibit migration of PAI-1-deficient SMCs, suggesting that tiplaxtinin’s effect on SMCs was non-toxic and PAI-1-dependent. The anti-migratory effect of tiplaxtinin on SMCs was preserved in collagen 3-D matrix containing vitronectin and other extracellular matrix molecules, further supporting the physiological significance of the effect. In contrast to SMCs, tiplaxtinin did not inhibit migration of human aortic ECs in vitro or murine ECs in vivo, the latter assessed in a murine carotid injury model. To study the basis for the differential effect of tiplaxtinin on SMCs vs. ECs, we compared expression of LDL receptor-related protein 1 (LRP1), a motogenic receptor for PAI-1, between cell types by RT-PCR and found that LRP1 gene expression was significantly lower in ECs than in SMCs. Furthermore, recombinant PAI-1 stimulated the migration of wild-type mouse embryonic fibroblasts (MEFs), but not LRP1-deficient MEFs. Conclusions: Tiplaxtinin, a pharmacological inhibitor of PAI-1, inhibits SMC migration under physiological conditions, while having no inhibitory effect on EC migration. The differential effect of PAI-1 inhibition on SMCs vs. ECs appears to be mediated by LRP1 and may be of clinical significance, as it is advantageous to prevent intimal hyperplasia by inhibiting SMC migration without inhibiting EC migration, which is key to preserving an intact, anti-thrombotic vascular endothelium.


2004 ◽  
Vol 91 (03) ◽  
pp. 473-479 ◽  
Author(s):  
Ana Guimarães ◽  
Dingeman Rijken

SummaryTAFIa was shown to attenuate fibrinolysis. In our in vitro study, we investigated how the inhibitory effect of TAFIa depended on the type and concentration of the plasminogen activator (PA). We measured PA-mediated lysis times of plasma clots under conditions of maximal TAFI activation by thrombin-thrombomodulin in the absence and presence of potato carboxypeptidase inhibitor. Seven different PAs were compared comprising both tPA-related (tPA, TNK-tPA, DSPA), bacterial PA-related (staphylokinase and APSAC) and urokinase-related (tcu-PA and k2tu-PA) PAs. The lysis times and the retardation factor were plotted against the PA concentration. The retardation factor plots were bell-shaped. At low PA concentrations, the retardation factor was low, probably due to the limited stability of TAFIa. At intermediate PA concentrations the retardation factor was maximal (3-6 depending on the PA), with TNK-tPA, APSAC and DSPA exhibiting the strongest effect. At high PA concentrations, the retardation factor was again low, possibly due to inactivation of TAFIa by plasmin or to a complete conversion of glu-plasminogen into lys-plasminogen. Using individual plasmas with a reduced plasmin inhibitor activity (plasmin inhibitor Enschede) the bell-shaped curve of the retardation factor shifted towards lower tPA and DSPA concentrations, but the height did not decrease. In conclusion, TAFIa delays the lysis of plasma clots mediated by all the plasminogen activators tested. This delay is dependent on the type and concentration of the plasminogen activator, but not on the fibrin specificity of the plasminogen activator. Furthermore, plasmin inhibitor does not play a significant role in the inhibition of plasma clot lysis by TAFI.


Blood ◽  
1989 ◽  
Vol 73 (8) ◽  
pp. 2188-2195 ◽  
Author(s):  
BS Schwartz ◽  
MC Monroe ◽  
JD Bradshaw

Abstract Peripheral blood mononuclear cells (PBMs) produce both tissue factor and plasminogen activator inhibitor type 2 (PAI-2) in response to gram- negative bacterial lipopolysaccharide (LPS). The cellular roles in the tissue factor response have been previously elucidated, and we now report those roles in PAI-2 production. Monocytes are the only cells among LPS-stimulated PBMs that produce PAI-2 as assessed by measurement of PAI-2 activity and antigen. Concomitant immunohistochemistry demonstrated that monocytes contain PAI-2, with a greater number staining positively and more intensely after exposure to LPS. LPS- stimulated monocytes produced increased amounts of PAI-2 with or without addition of lymphocytes. Lymphocytes prestimulated with LPS and then washed did not induce PAI-2 production in monocytes to which they were added. Lipid X, a precursor in the biosynthetic pathway of lipid A and LPS, was able to inhibit LPS induction of monocyte PAI-2 in a dose- dependent manner. This inhibition was not due to cellular toxicity, the phospholipidlike nature of lipid X, interference with the PAI-2 assay, or monocyte production of a substance interfering with PAI-2. Lipid X was an effective inhibitor of PAI-2 production even when added up to 30 minutes after LPS.


Blood ◽  
2016 ◽  
Vol 128 (22) ◽  
pp. 2641-2641
Author(s):  
Lorena Lobo Figueiredo-Pontes ◽  
Robert S. Welner ◽  
Miroslava Kardosova ◽  
Hong Zhang ◽  
Meritxell Alberich-Jorda ◽  
...  

Abstract Natural killer (NK) cells participate in innate and adaptive immune responses, and upon activation rapidly produce cytokines, chemokines, and growth factors, including IFNγ, TNFα, TGFβ, GM-CSF, MIP1α, MIP1β, IL-10, and others, which can affect the function of other hematopoietic cells. Considering the recent evidences that hematopoietic stem cells (HSCs) respond to cytokine signaling, we hypothesized that NK cell-mediated cytokine production could mediate HSC function. By the use of co-cultures of purified Ly5.1 murine NK cells and congenic Ly5.2 HSCs, we concluded that NK activity affects HSC frequency in vitro as well as hematopoietic reconstitution in vivo. Sorted NK cells (CD3- NK1.1+) and HSCs (Lin-, Sca1+, ckithi, CD48-, CD150+) were co-cultured in the presence or absence of IL2 over an OP9 stromal cells layer for 14 to 28 days. After 14 days, the addition of NK cells to HSC cultures resulted in an approximate 2-fold reduction of lineage negative cells (Lin-) recovered cells, as compared to control HSC cultures, as determined by flow cytometry analysis. Lin- counts were even lower in HSC+NK long-term cultures when compared to HSC only cultures. Ly5.1 HSCs and/or Ly5.2 NK cells were injected into sublethally irradiated Ly5.1/2 chimeric mice in a ratio of 105 NK to 103 HSCs per mouse. The addition of IL2-stimulated NK to injected HSCs reduced engraftment from 15.7% to 1.82% when the 16 weeks bone marrow (BM) chimerism was analyzed. In agreement, donor CD45.1 cells contribution to the LSK and HSC subpopulations was reduced in the HSC+NK transplanted mice. To test whether NK depletion from BM grafts would affect HSC function, we performed limiting dilution transplantation assays where whole BM from Ly5.2 mice was submitted to immunonagnetic NK1.1 or IgG depletion and injected into lethally irradiated Ly5.1 animals. Donor chimerism after 8 and 16 weeks of transplant showed that depleting NK cells improves the engraftment ability of HSC in a cell dose-dependent manner. When 25 x104 BM cells were injected, chimerism increased from 40 to more than 90% in NK depleted group. Of note, HSC frequency was 1 in 1595 in the control and 1 in 95 in the NK depleted group. In order to understand the mechanisms by which NK cells could regulate HSCs, we took advantage of a CCAAT/enhancer-binding protein gamma (C/ebpg) knockout (KO) conditional mouse model generated in our laboratory, considering that C/ebpg had been previously shown to regulate NK cytotoxicity. Using similar culture conditions, HSCs and NK cells isolated from control (CT) or Cebpg KO mice were injected into congenic sublethally irradiated recipients. Results showed that Cebpg-deficient NK cells do not harm HSC engraftment as CT NK cells do. For instance, after 8 weeks, the addition of CT non-stimulated and IL-2-stimulated NK cells to normal transplanted HSCs reduced the engraftment from 40% to 20% and 10%, respectively. In contrast, chimerism was not different when HSCs only or HSCs + stimulated KO NK cells were transplanted. Gene expression and cytokine profiles of deficient and normal NK cells revealed the potential players of this HSC-NK regulation. Of these, interferon gamma (IFNg), was lower produced by the C/ebpg deficient NK cells. Therefore, besides controlling NK cytotoxicity, we showed here that C/ebpg also plays a role in the regulation of HSCs by NK-mediated cytokine production. Next, we investigated whether depletion of NK cells from human BM samples would improve transplantation efficiency. NK cells were removed using CD56 antibody and transplanted into sublethally irradiated NSG mice. Sixteen weeks after transplantation, animals were sacrificed and the percentage of human CD45 cells in blood, BM, and spleen demonstrated that NK depletion from human BM favors engraftment. Altogether, these findings provide new insights to the knowledge of HSC regulation by NK cells, which are present in BM transplantation (BMT) grafts. Although the alloreactive effect of NK cells against non-identical tumor cells from BMT recipients is well known, its cytokine-mediated effects over identical progenitor cells from the graft were not previously explored. We show that NK-secreted cytokines harm stem cell function, thus suggesting that depletion of NK cells from BM donor cells preparations can improve stem cell engraftment, particularly in the setting of alternative transplants with limiting cell numbers or non-myeloablative conditioning regimens. Disclosures No relevant conflicts of interest to declare.


Blood ◽  
2014 ◽  
Vol 124 (21) ◽  
pp. 4217-4217
Author(s):  
Gabriela Chang ◽  
Helen M. Atkinson ◽  
Leslie R. Berry ◽  
Anthony K.C. Chan

Abstract Introduction: Unfractionated heparin (UFH) and low molecular weight heparin (LMWH) are widely used anticoagulants for thrombosis treatment. However, these anticoagulants have limitations such as increased bleeding, variable dose response, required frequent monitoring, and, in the case of LMWH, inability to inhibit thrombin. This has led to the development of a covalent complex of antithrombin and heparin (ATH), which has been shown to overcome many of these shortcomings. ATH has faster rates of inhibition of many coagulation factors, is able to inhibit clot-bound thrombin, and is a more effective inhibitor of both venous and arterial thrombosis in animal models. Moreover, in a rabbit thrombosis model, ATH has been shown to decrease clot mass and fibrin accretion, while the contrary was observed for UFH. From these observations, it was suggested that ATH may enhance fibrin breakdown and thus led to investigations into the effects of UFH and ATH on fibrinolysis. In vitro studies have shown that UFH enhances antithrombin inhibition of plasmin. In addition, ATH displays a slightly greater inhibition of plasmin generation and activity. Such studies were conducted in purified systems, in the absence of other plasmin inhibitors naturally present in plasma. Therefore, the aim of the present study was to compare the effects of UFH, LMWH, and ATH on plasmin generation in plasma. Methods: At 37°C tissue plasminogen activator (tPA) and soluble fibrin fragments (fib) were added to normal adult pooled platelet poor plasma supplemented with 0.35, 0.7, 1.4, or 2.1 U anti-Xa/ml UFH, LMWH, or ATH, to initiate plasmin generation (8.93nM tPA and 300µg/ml fib). At various time points, subsamples were mixed with excess plasminogen activator inhibitor 1 (PAI-1) (55.12nM) to stop further plasmin generation. The plasmin concentration at each time point was determined using a plasmin-specific chromogenic substrate and a standard curve produced from purified plasmin. Results: Comparisons of mean area under the curve (AUC) for plasmin generation displayed a significant decrease in plasmin generation in the presence of all three anticoagulants at all doses tested (p<0.05). Comparing the anticoagulants at similar doses, plasmin generation was significantly decreased in the presence of ATH (15384.66±1930.23nM/min) compared to LMWH (23892.28±3090.54nM/min) at 0.7 U/ml (p<0.05). At a dose of 1.4 U/ml, there was significantly less plasmin generated, over time, in the presence of UFH (20089.49±3022.1623nM/min) and ATH (19273.86±1805.7323nM/min) when compared to LMWH (24743.18±1265.1023nM/min) (p<0.05). There was no significant difference in plasmin inhibition between UFH and ATH at any of the doses tested. Conclusion: The present study supports previous findings that UFH and ATH can facilitate antithrombin inhibition of plasmin. It is also observed that LMWH catalyzes the inhibition of plasmin by antithrombin but possibly to a lesser extent. These findings suggest that ATH has a similar inhibitory effect on plasmin generation and activity in plasma compared to UFH, despite its overall superior anticoagulant properties. Therefore, previous in vivo observations displaying decrease in clot mass with administration of ATH was due to its enhanced anticoagulant abilities and not fibrinolysis enhancement. These findings add to our understanding of ATH mechanisms of action and aid in its development for clinical use. Disclosures No relevant conflicts of interest to declare.


Sign in / Sign up

Export Citation Format

Share Document