Tissue factor (TF), also known as coagulation factor III or thromboplastin, is a single-pass transmembrane glycoprotein that functions as the principal cellular initiator of blood coagulation in the extrinsic pathway.[1] Encoded by the F3 gene on chromosome 1p21.3, TF is normally expressed on the surfaces of perivascular cells such as fibroblasts, smooth muscle cells, and pericytes, but not on circulating blood cells or endothelial cells under steady-state conditions, thereby preventing inappropriate clotting in the vasculature.[1] Upon vascular injury, TF is exposed to circulating factor VII (FVII), which it binds with high affinity to form the TF-FVIIa complex; this complex allosterically activates FVII to FVIIa and catalyzes the activation of factors IX and X, leading to thrombin generation, fibrin clot formation, and platelet activation for hemostasis.[1] TF's discovery dates back to the early 20th century when it was identified as "thromboplastin" by researchers like Morawitz and Loeb, with its molecular cloning achieved in 1981.[1]Structurally, mature human TF consists of 263 amino acids, comprising an extracellular domain (residues 1–219) that interacts with FVIIa, a transmembrane domain (220–242) anchoring it to the lipid bilayer, and a short cytoplasmic C-terminal tail (243–263) involved in intracellular signaling.[1] The extracellular region features two fibronectin type III-like domains with key disulfide bonds, such as Cys186-Cys209, essential for FVIIa binding and procoagulant activity; the complex enhances FVIIa enzymatic efficiency by approximately 10^5- to 10^7-fold compared to FVIIa alone.[1] TF expression is tightly regulated: it is constitutively produced in extravascular tissues like the brain, lungs, and placenta, but inducible in monocytes, endothelial cells, and other cells by stimuli including cytokines (e.g., TNF-α, IL-1), lipopolysaccharide (LPS), hypoxia, and shear stress.[2] Posttranslational modifications, such as N-linked glycosylation and palmitoylation, influence its localization to lipid rafts and microparticle shedding, which can disseminate procoagulant activity.[1] An alternatively spliced soluble form of TF (asHTF) lacks the transmembrane domain and exhibits reduced coagulant potential but retains signaling functions.[1]Beyond hemostasis, TF contributes to diverse pathophysiological processes through its protease-activated receptor (PAR) signaling pathways, particularly via the TF-FVIIa-Xa complex, which cleaves PAR1 and PAR2 to trigger intracellular cascades.[2] In thrombosis, dysregulated TF expression on circulating microparticles or endothelial cells promotes hypercoagulability, arterial and venous thromboembolism, and is implicated in conditions like sepsis, atherosclerosis, and cardiovascular disease.[2] TF also drives inflammation by inducing cytokine release (e.g., IL-6, IL-8) and adhesion molecule expression, linking coagulation to immune responses in disorders such as inflammatory bowel disease and acute lung injury.[2] In cancer, TF overexpression in tumors (e.g., breast, pancreatic, colorectal) correlates with poor prognosis, enhancing angiogenesis via VEGF upregulation, tumor invasion, metastasis, and chemoresistance through non-hemostatic mechanisms.[2] Its inhibition by endogenous regulators like tissue factor pathway inhibitor (TFPI) or therapeutic agents (e.g., monoclonal antibodies) represents a target for antithrombotic and anticancer strategies.[2]
Structure and Genetics
Molecular Structure
Tissue factor (TF), also known as coagulation factor III, is a 46 kDa transmembrane glycoprotein encoded by the F3gene.[3] It consists of 263 amino acids organized into three principal domains: an extracellular domain comprising the first 219 residues, a transmembrane domain spanning residues 220–242 (23 amino acids), and a short cytoplasmic C-terminal tail of 21 amino acids (residues 243–263).[1] The extracellular domain features two fibronectin type III-like folds, which form a compact structure essential for protein-protein interactions.[4]Post-translational modifications significantly contribute to TF's stability and function. The extracellular domain undergoes N-linked glycosylation at three asparagine residues (Asn11, Asn124, and Asn137), with two sites fully occupied and the third approximately 90% utilized, resulting in heterogeneous glycoforms that influence protein folding and cellular trafficking.[5] Additionally, two disulfide bonds—Cys49–Cys57 in the N-terminal domain and Cys186–Cys209 in the C-terminal domain—stabilize the fibronectin-like domains and maintain the overall tertiary structure, the latter being an allosteric disulfide essential for procoagulant activity.[6]The three-dimensional structure of the soluble extracellular domain of TF was first elucidated in 1996 through X-ray crystallography at 2.0 Å resolution, revealing a cylindrical arrangement of the two fibronectin type III domains that serves as a cofactor scaffold for binding factor VIIa.[7] This structure highlights a binding interface involving acidic residues on TF that interact with basic regions on factor VIIa, facilitating the initiation of the coagulation cascade.[8]An alternatively spliced variant of TF, known as soluble tissue factor (asTF) or alternatively spliced TF, arises from the exclusion of exon 5 during mRNA processing, resulting in a 206-amino-acid protein that lacks the transmembrane domain and is secreted into the extracellular space.[9] Unlike full-length TF, asTF exhibits distinct signaling properties independent of membrane anchoring.[10]
Gene and Variants
The F3 gene, which encodes tissue factor, is located on the short arm of human chromosome 1 at the p21.3 locus. It spans approximately 12.4 kb of genomic DNA and consists of six exons.[11][3]The promoter region of the F3 gene contains binding sites for several transcription factors, including Sp1 and NF-κB, which contribute to its basal and inducible expression.[12][13]Mutations and polymorphisms in the F3 gene are rare and typically associated with mild bleeding phenotypes rather than severe deficiencies. For instance, a heterozygous two-nucleotide deletion (c.83_84del) in the F3 gene was identified in a patient with a mild bleeding phenotype (menorrhagia, epistaxis, easy bruising, and bleeding after dental extraction) but normal prothrombin time and activated partial thromboplastin time, marking the first reported human mutation linked to tissue factor deficiency.[14] Homozygous null mutations in F3 are embryonic lethal in mice and have not been documented in humans, though they underscore the gene's critical role in hemostasis.[15] Common polymorphisms, such as the -1812G/A variant in the promoter, may influence tissue factor expression levels but are not strongly associated with bleeding disorders.[16]The F3 gene exhibits strong evolutionary conservation, reflecting its essential function in coagulation across vertebrates. Orthologs are present in species ranging from mice to fish, with zebrafish studies in 2011 demonstrating that knockdown of the f3a ortholog disrupts vascularization and embryonic development, confirming conserved roles in hemostasis and angiogenesis.[17][18]
Physiological Functions
Role in Coagulation
Tissue factor (TF), a transmembrane glycoprotein, serves as the primary initiator of the extrinsic coagulation pathway, which is triggered physiologically by the exposure of subendothelial TF following vascular injury to arrest bleeding.[19] Upon vessel damage, circulating factor VII (FVII) binds to the exposed TF on cell surfaces and is activated to FVIIa, forming the TF-FVIIa complex that drives hemostasis.[20]The assembly of the TF-FVIIa complex occurs on phospholipid membranes, particularly those enriched with negatively charged phospholipids like phosphatidylserine, which provide the anionic surface essential for stabilizing the complex and enhancing its catalytic efficiency.[21] This membrane dependence facilitates the recruitment and orientation of substrates, enabling the complex to proteolytically activate factor X (FX) to FXa; the initial rate of this activation follows the kinetics Rate = k [TF][VIIa][X], where k represents the catalytic constant.[22]FXa, in turn, associates with activated factor V (FVa) on phospholipid surfaces to form the prothrombinase complex, which efficiently converts prothrombin to thrombin.[2] Thrombin then catalyzes the conversion of fibrinogen to fibrin monomers, which polymerize to form a stable clot, while also activating factors XI and VIII to amplify the coagulation cascade.[2]Additionally, the TF-FVIIa complex activates factor IX (FIX) to FIXa, which combines with activated factor VIII (FVIIIa) on membranes to generate the intrinsic tenase complex, further promoting FX activation and sustaining thrombin generation for robust fibrin clot formation.[23]
Non-Hemostatic Signaling
Beyond its role in initiating coagulation, the tissue factor (TF)-factor VIIa (FVIIa) complex engages in non-hemostatic signaling by activating protease-activated receptor 2 (PAR2) on cell surfaces, particularly in endothelial cells and fibroblasts. This activation occurs through proteolytic cleavage of PAR2 by FVIIa, independent of downstream coagulation factors, leading to G-protein-coupled receptor signaling that triggers intracellular pathways such as the mitogen-activated protein kinase/extracellular signal-regulated kinase (MAPK/ERK) cascade.[24] The MAPK/ERK pathway subsequently promotes cell proliferation and survival by phosphorylating transcription factors like Elk-1, enhancing gene expression associated with growth and anti-apoptotic effects in various cell types, such as during wound healing and tissue remodeling.[25] This PAR2-mediated signaling is distinct from hemostatic functions, as it relies on the catalytic activity of FVIIa rather than fibrin formation.[26]TF also interacts directly with Eph receptors, notably EphA2, a receptor tyrosine kinase involved in cellular communication. The TF-FVIIa complex modulates EphA2 signaling by cleaving the receptor or altering its phosphorylation state, which influences downstream effects on angiogenesis and apoptosis.[27] Specifically, this interaction inhibits EphA2 autophosphorylation, reducing ephrin-A1 reverse signaling and thereby promoting angiogenic responses while suppressing pro-apoptotic pathways in endothelial cells.[28] These effects contribute to physiological processes like vascular development and remodeling.[27]The TF-FVIIa-PAR2 axis further induces the expression of pro-inflammatory and pro-angiogenic cytokines, including interleukin-8 (IL-8) and vascular endothelial growth factor (VEGF). In keratinocytes and fibroblasts, FVIIa binding to TF upregulates IL-8 transcription via PAR2 and NF-κB activation, contributing to inflammatory signaling during normal immune responses.[29] Similarly, TF-dependent VEGF production occurs through PAR2-mediated pathways, increasing vascular permeability by disrupting endothelial junctions and facilitating leukocyte extravasation in physiological contexts like tissue repair.[30] These cytokines amplify non-hemostatic effects, such as enhanced endothelial barrier dysfunction, which supports tissue remodeling.[26]Soluble variants of TF, particularly alternatively spliced tissue factor (asTF), mediate FVII-independent signaling by circulating in plasma or being secreted from cells such as monocytes and endothelial cells. Unlike full-length TF, asTF lacks the transmembrane domain and signals primarily through direct binding to integrins like αvβ3 or β1 on endothelial and immune cells, bypassing FVIIa and PAR2.[31] This interaction activates focal adhesion kinase (FAK) and promotes cell migration, adhesion, and angiogenesis without coagulant activity, contributing to physiological processes like vascular formation and monocyte recruitment.[25][32]
Expression and Regulation
Tissue Distribution
Tissue factor (TF), also known as coagulation factor III, is constitutively expressed in various extravascular tissues, forming a protective hemostatic barrier around the vasculature and at organ surfaces. In the vascular wall, TF is prominently found in adventitial fibroblasts, pericytes, and vascular smooth muscle cells, ensuring rapid coagulation initiation upon vessel injury.[33][34] Organ-specific expression includes high levels in the brain (particularly astrocytes and cerebral cortex), lungs (epithelial cells and bronchial epithelium), placenta (trophoblasts and decidual cells), heart (fibroblasts and myocytes), skin (epidermis and keratinocytes), and kidney (glomeruli and tubular epithelia in humans).[33][35][36]In contrast, resting endothelial cells lining blood vessels exhibit minimal to undetectable TF expression under normal physiological conditions, preventing unwanted intravascular coagulation. Similarly, circulating blood cells such as platelets, neutrophils, and most monocytes show negligible constitutive TF, with only a small subset (approximately 1.5%) of CD14+ monocytes harboring low intracellular levels.[33][33]During development, TF expression patterns are more widespread and intense in fetal tissues compared to adults, supporting embryonic morphogenesis and vascularization. In human and mouse embryos, strong TF is observed in ectodermal and endodermal layers from early stages (e.g., human stage 5, mouse 6.5-7.5 days post-coitum), with notable presence in fetal epidermis, myocardium, bronchial epithelium, hepatocytes, and maturing glomeruli; additionally, sites like skeletal muscle and pancreas show TF in fetuses but not in mature tissues.[36][37]Species differences in TF distribution are evident, with rodents displaying broader expression than humans; for instance, mouse neutrophils constitutively express functional TF, whereas human neutrophils show weak or absent levels, and renal TF is present in human glomeruli but mouse tubular epithelia.[33][36] While this baseline distribution maintains hemostasis, TF expression can be dynamically upregulated in certain cells upon stimulation.[33]
Regulation Mechanisms
Tissue factor (TF) expression is primarily regulated at the transcriptional level by pro-inflammatory cytokines such as tumor necrosis factor-α (TNF-α) and interleukin-1β (IL-1β), which activate the nuclear factor-κB (NF-κB) signaling pathway in endothelial cells and monocytes. Other stimuli, including lipopolysaccharide (LPS) and shear stress, also induce TF expression via NF-κB activation or increased Sp1 activity in endothelial cells.[38][39][40] Upon stimulation, TNF-α and IL-1β induce nuclear translocation of NF-κB, which binds to κB sites in the promoter region of the F3 gene encoding TF, leading to increased TF mRNA transcription and enhanced procoagulant activity.[41] This cytokine-mediated upregulation is a key mechanism in inflammatory responses, where TF expression can increase up to 100-fold within hours of exposure.[38]Hypoxia also transcriptionally regulates TF through hypoxia-inducible factor-1α (HIF-1α), which accumulates under low oxygen conditions and mediates upregulation of TF expression.[42]HIF-1α activation enhances TF transcription in various cell types, including tumor and vascular cells, promoting coagulation in hypoxic microenvironments such as those found in tumors or ischemic tissues.[42]At the post-transcriptional level, microRNAs (miRNAs) fine-tune TF expression by targeting the 3' untranslated region (3'UTR) of F3 mRNA. For instance, miR-19a suppresses TF by binding to the F3 3'UTR, reducing mRNA stability and translation in endothelial cells, thereby exerting antithrombotic effects that are diminished in diabetes.[43][44] This miRNA-mediated control helps maintain vascular homeostasis by preventing excessive TF activity under basal conditions.TF procoagulant activity is further modulated by encryption and decryption mechanisms on the cell surface, independent of expression levels. Encrypted TF, which is inactive despite being present, becomes decrypted upon exposure of phosphatidylserine (PS) on the outer leaflet of the plasma membrane, allowing full interaction with factor VIIa and subsequent coagulation initiation.[45][46] PS exposure, triggered by calcium influx or apoptotic signals, reorients TF and enhances its catalytic efficiency by facilitating cofactor binding, representing a critical post-translational switch in TF function.[45]Feedback inhibition of TF activity occurs via tissue factor pathway inhibitor (TFPI), a Kunitz-type serine protease inhibitor that binds to the TF-VIIa-Xa ternary complex, sterically blocking further factor X activation and downregulating the extrinsic coagulation pathway.[47] TFPI first inhibits free factor Xa, forming a TFPI-Xa complex that then associates with TF-VIIa, providing a potent negative feedback loop to prevent uncontrolled thrombus formation once initial coagulation is initiated.[48] This mechanism ensures localized hemostasis, with TFPI levels modulating the threshold for TF-dependent clotting.[47]
Pathophysiological Roles
Thrombosis and Hemostasis
Tissue factor (TF) overexpression plays a central role in the pathogenesis of disseminated intravascular coagulation (DIC), a life-threatening disorder characterized by widespread activation of the coagulation system leading to microvascular thrombosis and consumption of clotting factors. In sepsis, inflammatory cytokines such as tumor necrosis factor-alpha and interleukin-1 beta induce TF expression on monocytes, endothelial cells, and subendothelial tissues, initiating the extrinsic coagulation pathway and resulting in excessive thrombin generation, fibrin deposition, and organ dysfunction.[49] Similarly, in trauma, damaged tissues release TF-bearing microparticles and cellular debris into the circulation, triggering systemic coagulopathy and contributing to trauma-induced coagulopathy, which overlaps with DIC features like prolonged prothrombin time and elevated fibrin degradation products.[50] This overexpression disrupts the balance between procoagulant and anticoagulant mechanisms, exacerbating bleeding tendencies once clotting factors are depleted.[51]TF-bearing microparticles also contribute to hypercoagulability and venous thromboembolism (VTE) in non-malignant conditions such as sepsis and trauma. These microparticles, derived from activated leukocytes, endothelial cells, and platelets, circulate in plasma and express functional TF that complexes with factor VIIa to promote thrombin formation and fibrin clot development on venous endothelium. In septic patients, elevated levels of TF-positive microparticles correlate with increased VTE risk, as they enhance local procoagulant activity in low-flow venous sites, independent of malignancy.[52] Post-trauma, microparticle-associated TF sustains a prothrombotic state, facilitating deep veinthrombosis through endothelial activation and platelet aggregation, with studies showing higher microparticle counts in patients developing VTE compared to those who do not.[53] This mechanism underscores TF's role in shifting hemostasis toward pathological thrombosis in acute inflammatory states.Congenital TF deficiencies are exceedingly rare. The first reported human case involves a heterozygous mutation in the F3 gene, leading to mild bleedingdiathesis with prolonged prothrombin time and reduced procoagulant activity. Affected individuals exhibit symptoms such as menorrhagia, epistaxis, and easy bruising, highlighting TF's role in hemostasis.[14][54] A 2025 study identified rare heterozygous missense variants in F3 (e.g., p.Gly196Arg) that impair TF-factor VIIa interaction and basal coagulation, resulting in reduced factor VIIa-antithrombin complexes and D-dimer levels without severe bleeding.[55]Recent advances from 2020 to 2025 have elucidated TF's involvement in COVID-19-associated coagulopathy, where viral infection upregulates TF on monocytes and endothelial cells, driving hypercoagulability and microvascular thrombosis. Elevated TF activity correlates with high D-dimer levels, indicating fibrinolysis of widespread clots, and predicts severe outcomes like acute respiratory distress syndrome in hospitalized patients.[56] Studies during the pandemic demonstrated that TF pathway inhibition in preclinical models reduced D-dimer elevation and thrombosis incidence, informing targeted anticoagulant strategies for COVID-19-related hemostatic imbalances.[57]
Cancer and Angiogenesis
Tissue factor (TF) is overexpressed in a high percentage of solid tumors, ranging from 50% to over 90% in cases such as breast cancer (including 50-85% in triple-negative subtypes), pancreatic cancer, and glioblastoma (up to 95%), where elevated levels strongly correlate with tumor aggressiveness, increased metastasis risk, and poor patient prognosis.[58] In pancreatic cancer, TF overexpression promotes tumor progression and venous thromboembolism, while in breast cancer, it is linked to stromal neovascularization and reduced survival; similarly, in glioblastoma, it associates with higher malignancy grades, microvessel density, and a 50% decrease in survival for wild-type IDH1 cases.[58] This overexpression often exceeds that in normal tissues by several fold, serving as a marker of advanced disease stages across these malignancies.[58]The TF-VIIa complex activates protease-activated receptor 2 (PAR2) on tumor cells, driving pro-tumorigenic signaling that enhances cell migration, invasion, and metastasis through pathways involving ERK activation, cofilin regulation, and integrin engagement (such as α3β1 and α6β1).[59] This signaling promotes tumor growth by inducing proangiogenic cytokines like IL-8 and VEGFC, as well as immune modulators such as CSF1, while also facilitating metastatic niche formation via thrombin-mediated platelet recruitment and β-arrestin-dependent motility.[59] In breast cancer models, inhibition of this TF-VIIa-PAR2 axis reduces invasion and metastasis by disrupting chemokine-induced pathways, highlighting its non-coagulant role in oncogenesis.[60]TF further contributes to angiogenesis by upregulating vascular endothelial growth factor (VEGF) expression—particularly isoforms like VEGF165 and VEGF189—via PAR2-mediated pathways, which stimulates endothelial cell proliferation and new vessel formation in the tumor microenvironment.[58] Additionally, TF-initiated coagulation leads to fibrin deposition in the tumor stroma, providing a provisional matrix that supports cancer cell survival, invasion, and vascular stabilization, independent of full thrombus formation.[60] The alternatively spliced form of TF (asTF), lacking coagulant activity, directly binds endothelial integrins (αvβ3 and α6β1) to enhance microvascular density and tumor vascularization.[60]Recent studies from 2020-2025 have identified TF as a potential biomarker in lymphomas, with elevated tissue factor-carrying monocytes (particularly intermediate subtypes) observed in diffuse large B-cell lymphoma (DLBCL), correlating with increased thromboinflammatory risk and disease progression.[61] In breast cancer, thrombin generation assays—triggered by low-dose TF—have shown predictive value for early recurrence, with higher endogenous thrombin potential (e.g., 1,843 nM·min) in patients relapsing within two years post-surgery, integrating TF-driven coagulation as a prognostic indicator alongside subtypes like triple-negative and Luminal B.[62] These findings underscore TF's role in stratifying high-risk patients for targeted interventions.[62]
Inflammation and Immunity
Tissue factor (TF) plays a pivotal role in linking inflammation to coagulation through its upregulation in monocytes and macrophages by inflammatory cytokines such as tumor necrosis factor-α (TNF-α) and interleukin-1β (IL-1β). This induction enhances procoagulant activity on these immune cells, fostering thromboinflammation, a bidirectional process where inflammatory signals promote thrombus formation and vice versa.[63] In the context of innate immunity, this upregulation facilitates localized hemostasis at sites of infection or injury but can exacerbate pathological inflammation when dysregulated.[64]A key mechanism involves TF's contribution to neutrophil extracellular trap (NET) formation during sepsis, where TF-bearing NETs serve as procoagulant scaffolds that trap pathogens while initiating the extrinsic coagulation pathway. These structures amplify the inflammation-thrombosis axis by recruiting platelets and promoting fibrin deposition, leading to microvascular thrombosis and organ damage in severe cases.[65] Functional TF on NETs, often in complex with complement factors, heightens immunothrombotic responses, underscoring TF's role in bridging antimicrobial defense with coagulopathy.[66]In atherosclerosis, TF expression in vascular cells—including endothelial cells and vascular smooth muscle cells—drives plaque instability by sustaining a prothrombotic microenvironment within the arterial wall. This expression, triggered by local inflammatory cues, leads to chronic coagulation activation that weakens plaque integrity and predisposes to rupture, resulting in acute thrombotic events.[67] Notably, TF-laden microparticles from apoptotic cells within plaques further propagate this instability.[68]Recent advances between 2020 and 2025 highlight TF as a thromboinflammatory biomarker in contexts such as lymphomas and autoimmune diseases. In lymphomas such as diffuse large B-cell lymphoma, elevated levels of TF-carrying classical monocytes correlate with increased thrombosis risk, reflecting heightened monocyte activation.[69] Similarly, in autoimmune disorders like systemic lupus erythematosus, augmented TF expression on monocytes contributes to a procoagulant state, associating with disease flares and vascular complications.[70] These findings emphasize TF's utility in monitoring immune-driven thrombotic tendencies, with cytokine regulation—as detailed in expression mechanisms—central to its pathological induction.[63]
Interactions and Pathways
Protein Interactions
Tissue factor (TF), a transmembrane glycoprotein, primarily interacts with coagulation factor VII (FVII) and its activated form, FVIIa, through its extracellular domain, forming a high-affinity complex essential for initiating the extrinsic coagulation pathway. This binding occurs with a dissociation constant (Kd) of approximately 10 nM in the presence of phospholipids, enabling the allosteric activation of FVIIa and subsequent proteolytic activity.[71]TF also associates with tissue factor pathway inhibitor (TFPI), a Kunitz-type serine protease inhibitor, to form an inhibitory quaternary complex comprising TF, FVIIa, activated factor X (FXa), and TFPI. This complex sequesters the TF-FVIIa catalytic site, providing feedback inhibition to limit excessive coagulation initiation after FXa generation.[72]Beyond coagulation factors, TF binds to cell surface integrins, such as αvβ3, facilitating non-hemostatic signaling events like angiogenesis and cell migration. This interaction occurs via specific motifs in TF's extracellular domain, promoting integrin clustering and downstream intracellular pathways independent of proteolytic activity.[10][73]TF's functionality is modulated by its interactions with membrane phospholipids, particularly phosphatidylserine (PS), which anchors the TF-FVIIa complex to the cell surface for optimal coagulant activity. Annexin V, an anticoagulant protein, competes with these coagulation complexes for PS binding sites, thereby inhibiting TF-mediated procoagulant responses by blocking access to the lipid surface.[74][75]
Downstream Pathways
Tissue factor (TF), in complex with factor VIIa (FVIIa), initiates downstream signaling primarily through protease-activated receptor 2 (PAR2), a G-protein-coupled receptor. The TF-FVIIa complex cleaves PAR2 at its N-terminus, triggering conformational changes that activate heterotrimeric G proteins, including Gαq, Gα12/13, and Gαi, which mobilize intracellular calcium and activate phospholipase C.[31] This leads to protein kinase C (PKC) activation, which phosphorylates the TF cytoplasmic domain at serine 253, enhancing receptor trafficking and sustained signaling.[76] Additionally, PAR2 engages β-arrestin pathways upon activation, recruiting β-arrestin-1 and -2 to facilitate receptor internalization and scaffold-mediated activation of extracellular signal-regulated kinase 1/2 (ERK1/2), ultimately promoting gene transcription for cellular responses such as migration and survival.[77] In pro-inflammatory contexts, PAR2 signaling activates nuclear factor kappa B (NF-κB) via IκB kinase (IKK) and stress-activated protein kinases, resulting in RelA/p65 nuclear translocation and transcription of pro-inflammatory genes like interleukin-8 (IL-8).[78]The EphA2-TF axis represents another key downstream pathway, where TF-FVIIa directly interacts with the receptor tyrosine kinase EphA2. FVIIa cleaves the EphA2 ectodomain at a conserved arginine residue in the ligand-binding domain, independent of PAR2, which disrupts ephrin-A1 binding and promotes EphA2 autophosphorylation.[27] This cleavage initiates phosphorylation cascades involving Src kinase, leading to downstream activation of RhoA/ROCK for cytoskeletal remodeling and enhanced cell migration.[79] In endothelial cells, thrombin-induced EphA2 phosphorylation via PAR1 further amplifies these effects, contributing to migratory responses without requiring ephrin ligands.[79]TF signaling exhibits significant cross-talk with the phosphoinositide 3-kinase (PI3K)/Akt pathway, which supports cell survival and proliferation. The TF-FVIIa complex induces phosphorylation of EphA2 at serine 897 through PI3K/Akt activation, providing docking sites for PI3K's SH2 domain and stabilizing β-catenin for transcriptional regulation.[27] This cross-talk is evident in cancer cells, where inhibition of PI3K with LY294002 blocks EphA2 phosphorylation and reduces motility.[27] Concurrently, TF integrates with NF-κB signaling; for instance, EphA2 knockdown suppresses thrombin-induced serine 536 phosphorylation of NF-κB p65, impairing pro-inflammatory gene expression like ICAM-1.[79] In macrophages, TF-PAR2 signaling upregulates NF-κB via c-Jun N-terminal kinase (JNK), linking coagulation to inflammatory amplification.[80]Integration with coagulation occurs through thrombin feedback, which amplifies TF signaling in a positive loop. Thrombin, generated downstream of TF-FVIIa, cleaves PAR1, leading to non-proteolytic transactivation of PAR2 via heterodimer formation, thereby boosting TF expression and procoagulant activity up to eightfold in stimulated cells.[81] This feedback sustains endothelial thromboinflammatory responses, enhancing leukocyte adhesion molecules like E-selectin and cytokines such as IL-6, while thiol-disulfide exchanges at TF cysteines (e.g., Cys186-Cys209) modulate the balance between coagulant and signaling functions.[31]
Clinical and Historical Aspects
Thromboplastin and Assays
Thromboplastin, historically known as tissue thromboplastin or coagulation factor III, was first recognized in the early 1900s as a tissue-derived substance that initiates blood clotting when added to plasma. In 1905, Paul Morawitz proposed a coagulation theory incorporating "thrombokinase," the clot-promoting principle from tissues, which required interaction with calcium, prothrombin, and fibrinogen to generate thrombin and form clots.[82] This discovery linked tissue extracts to the extrinsic pathway of coagulation, establishing the foundation for laboratory assays.[83]As a reagent in coagulation testing, thromboplastin consists of tissue factor (TF) combined with phospholipids, traditionally extracted from rabbit or human brain tissue to provide the necessary procoagulant activity.[84] In the prothrombin time (PT) assay, developed by Armand Quick in the 1930s, citrated plasma is recalcified with thromboplastin, measuring the time to fibrin clot formation as an indicator of extrinsic pathway function and factor deficiencies.[85] The phospholipids in thromboplastin, such as phosphatidylserine and phosphatidylethanolamine, facilitate TF's assembly with factor VII, enhancing assay sensitivity.[84]To address variability among thromboplastin preparations, the International Sensitivity Index (ISI) was introduced for PT standardization, calibrating reagents against World Health Organization reference standards to compute the International Normalized Ratio (INR).[86] ISI values are species-specific: rabbit brain-derived thromboplastins typically yield an ISI around 1.21, while human recombinant variants are calibrated to approximately 1.11, ensuring consistent INR reporting across laboratories.[86] For instance, the Fifth International Standard for rabbit plain thromboplastin (RBT/16) and recombinant human thromboplastin (rTF/16) were established through multicenter studies with low inter-laboratory variability (4.6–5.7% coefficient of variation).[86]Modern PT assays have evolved to use recombinant human TF incorporated into synthetic phospholipid vesicles, replacing animal-derived extracts to improve batch-to-batch consistency and reduce contamination risks.[84] These recombinant reagents maintain high sensitivity to factor VII deficiencies while achieving ISI values close to 1.0, facilitating more reliable monitoring of anticoagulant therapy.[85] The phospholipid composition, particularly 5–30 mol% phosphatidylserine, critically influences ISI without altering the molar ratio of phospholipids to TF.[84]
Therapeutics and Deficiencies
Tissue factor (TF) has emerged as a promising therapeutic target in cancer due to its overexpression on tumor cells and vascular endothelium, contributing to angiogenesis and tumor progression. Monoclonal antibodies and antibody-drug conjugates (ADCs) directed against TF, such as tisotumab vedotin, inhibit TF-mediated signaling and deliver cytotoxic payloads to TF-expressing tumors, thereby blocking tumor angiogenesis and promoting tumor regression. Tisotumab vedotin received accelerated FDA approval in 2021 and full approval in April 2024 for recurrent or metastatic cervical cancer, based on superior overall survival in phase III trials compared to chemotherapy. As of 2025, tisotumab vedotin has received approvals in the European Union, Japan, and Hong Kong for similar indications, with ongoing evaluations in other solid tumors like ovarian and endometrial cancers. Additionally, TF-VIIa complex inhibitors, such as the small-molecule PCI-27483, have shown preclinical efficacy in suppressing tumor growth and angiogenesis by disrupting non-hemostatic TF signaling pathways without significantly affecting coagulation.[87][88][89][90][91][92][93]Modulation of tissue factor pathway inhibitor (TFPI), a natural antagonist of TF-VIIa, represents a novel strategy for treating bleeding disorders like hemophilia by enhancing thrombin generation and hemostasis. Marstacimab, a monoclonal antibody that inhibits TFPI's Kunitz domain 2, promotes procoagulant activity and has been evaluated in clinical trials for prophylaxis in hemophilia A and B patients without inhibitors. Phase III trials (BASIS study) demonstrated significant reductions in annualized bleeding rates, leading to FDA approval in October 2024. This approach avoids direct factor replacement, offering a non-factor therapy that rebalances coagulation without increasing thrombotic risk.[94][95][96][97]Congenital deficiencies in TF are exceedingly rare, with no well-documented human cases reported due to the protein's essential role in hemostasis, potentially rendering homozygous null mutations embryonic lethal. In hypothetical or acquired TF-deficient states presenting with severe bleeding, management involves supportive measures such as fresh frozen plasma transfusions to provide exogenous TF and restore coagulation initiation. Heterozygotes carrying TF mutations exhibit reduced TF expression and, in preclinical mouse models, demonstrate protection against thrombosis, including decreased thrombus formation in atherosclerosis and injury-induced models, suggesting a potential antithrombotic benefit without overt bleedingdiathesis.[15][98]Recent advances from 2020 to 2025 have focused on TF-targeted nanotherapies to address hypercoagulability in COVID-19-associated thrombosis, where elevated TF expression on activated monocytes and extracellular vesicles drives microvascular clotting. Preclinical studies have explored nanoparticle platforms, such as liposomes and biomimetic carriers, to deliver TF inhibitors or anticoagulants selectively to TF-enriched thrombi, improving thrombolytic efficacy while minimizing systemic bleeding risks in SARS-CoV-2 infection models.[99][100]