CANCER Y TROMBOSIS

Page 1

Journal of Thrombosis and Haemostasis, 5 (Suppl. 1): 246–254

INVITED REVIEW

Cancer and thrombosis: from molecular mechanisms to clinical presentations H . R . B U L L E R , F . F . V A N D O O R M A A L , G . L . V A N S L U I S and P . W . K A M P H U I S E N Department of Vascular Medicine, Academic Medical Center, Amsterdam, the Netherlands

To cite this article: Buller HR, Van Doormaal FF, Van Sluis GL, Kamphuisen PW. Cancer and thrombosis: from molecular mechanisms to clinical presentations. J Thromb Haemost 2007; 5 (Suppl. 1): 246–54.

Summary. Although the bidirectional association between cancer and venous thromboembolism (VTE) has been known for almost two centuries, recent advances in our understanding of the clinical, laboratory, and epidemiologic aspects of this association have created a renewed interest in this topic. This review consists of two parts. The first part discusses the occurrence, determinants and significance of VTE in those with cancer, as well as the risk of developing and the possible need to detect cancer in those presenting with VTE. The second part reviews the role of hemostatic constituents (coagulation and fibrinolytic proteins and platelets) in promoting growth and progression of cancer, as well as the effects and possible mechanisms of the low molecular weight heparins (LMWH) in this process. Keywords: cancer, (low molecular weight) heparins, hemostasis, venous thromboembolism.

Cancer and thrombosis Historically, Armand Trousseau, the French physician (1801– 1867), is often considered to have been the first scientist to describe the association between cancer and venous thrombosis. This is incorrect for both the relationship of cancer with subsequent complicating VTE as well as for occult cancer in thrombosis patients. A careful literature analysis revealed that already in 1823 Bouillaud described three cancer patients with deep venous thrombosis [1,2]. He speculated in his monogram that the peripheral edema in the legs of these cancer patients was the result of obstruction of the veins by Ôcaillot fibrineuxÕ (fibrin clots) induced by the cancer process. Trousseau was 22 years old at that time. In the frequently quoted book by Trousseau, published in 1865, he indeed described in detail the Correspondence: Harry R. Buller, Department of Vascular Medicine, Academic Medical Center, Room F4-211, Meibergdreef 9, 1105 AZ Amsterdam, the Netherlands. Tel.: +31 205 665 9765; Fax.: +31 206 968 833 e-mail: m.m.veendorp@amc.uva.nl Received 14 February 2007, accepted 21 February 2007

relationship between cancer and VTE, but that was 42 years after Bouillaud [3]. Trousseau is also often quoted for the link between venous thrombosis and occult cancer. Close reading of his papers reveal that all his patients described with VTE already had extensive evidence of cancer at the time of the diagnosis of thrombosis [4,5]. The honor of the first description of a patient with deep venous thrombosis and the manifestation of a gastric cancer several months later goes to Illtyd James and Matheson [6]. They published their paper in 1935. Interestingly, the first proper cohort study in patients with VTE to assess the incidence of occult cancer was only published in 1982 [7]. These authors coined the question of whether an extensive search for cancer at the time of VTE diagnosis was indicated. The focus in this part of the review will be on deep vein thrombosis of the leg (DVT) or pulmonary embolism (PE) and the two-way association with cancer, with special emphasis on the epidemiologic aspects, the magnitude and consequences. Other manifestations of venous thrombosis, such as superficial flebitis, arm vein thrombosis, and catheter-related thrombosis, as well as the prevention and treatment of DVT and PE are outside the scope of this contribution.

Cancer and venous thromboembolism

In cancer patients VTE complications are common and the second leading cause of death [8–10]. Of every seven patients with cancer who die in hospital, one dies of pulmonary embolism [11]. Compared with non-cancer patients the risk of developing symptomatic VTE is six to seven times higher in cancer patients, with similar risks for the components of VTE (i.e. DVT and PE) [12,13]. The precise incidence and time course of VTE complications among cancer patients have until recently largely been unknown [14]. Initially, small cohort studies estimated the incidence among patients with various types and stages of cancer to be approximately 4% [14]. Subsequent larger epidemiologic studies found incidences of clinically manifest VTE of 110–120 events per 10 000 patients [15,16]. Recently, a large population-based study determined the incidence of VTE among patients diagnosed with specific types and stages of cancer [14]. Among 235 149 cancer cases, confirmed symptomatic VTE events were diagnosed within 2007 International Society on Thrombosis and Haemostasis


Cancer and thrombosis 247

2 years in 1.6%. Twelve percentage of these events occurred at the time cancer was diagnosed and the remaining 88% were observed subsequently. In another recent large populationbased study the duration between the diagnosis of cancer and the occurrence of VTE was determined [13]. In the first 3 months the risk of VTE was the highest, with an odds ratio of 54, whereas in the period between 3 and 12 months following the cancer diagnosis, the odds ratio of VTE was 14, and between 1 and 3 years this figure was 4 [13]. Table 1 details the determinants for the risk of symptomatic episodes of VTE in patients with cancer. In risk-adjusted models Chew and colleagues found metastatic disease at the time of cancer diagnosis the strongest predictor of VTE [14]. The risk was four to 13 times higher as compared with cases with localized disease. In patients with metastatic-stage disease the highest incidences (expressed as VTE events per 100 patient-years) were observed for pancreatic (20.0), stomach (10.7), bladder (7.9), uterine (6.4), renal (6.0), and lung (5.0) cancer. Most of the VTE complications occurred in the first year of follow-up [14]. Blom and colleagues also determined the significance of distant metastases for the risk of VTE and found an odds ratio of 20 as compared with patients without distant metastases [13]. All hematologic and solid tumor types have been associated with VTE. However, conflicting data exist about the relative risk of VTE according to tumor histology and site of primary location [8]. It is, however, clear that the VTE risk is not equal among the various types of cancer. Adjusted for age and sex, Blom et al. observed, in a population-based study, the highest risk of VTE among patients with hematological malignancies (odds ratio 28), followed by lung cancer (odds ratio 22) and gastrointestinal cancer (odds ratio 20) [13]. In another population-based study, Heit and colleagues noted that pancreatic cancer, lymphoma, and brain cancer have relative risks of VTE >25, followed by liver cancer, leukemia, and other gastrointestinal (esophagus, gallbladder) and gynecological cancer, with a relative risk above 17 [17]. In a study of hospitalized patients the highest incidence of VTE occurred among women with ovarian cancer, followed by brain and pancreas cancer and lymphoma [15,16]. The third and probably at present the most changing determinant for the risk of VTE is anticancer therapy (Table 1). Chemotherapy is a well-recognized independent risk factor for VTE and the annual incidence has been estimated to be 10%– 20% depending on the type of drugs given [8,18]. However, one Table 1 Determinants for the risk of venous thromboembolism in cancer patients Tumor stage Tumor type Anticancer therapy Chemotherapy Hormonal therapy Angiogenesis inhibitors Supportive therapy Surgery Prothrombotic abnormalities

2007 International Society on Thrombosis and Haemostasis

of the most prominent reasons for the increasing importance of anticancer therapy as a determinant for the thrombosis risk is that in addition to chemotherapy, other frequently combined therapies to treat the cancer, such as hormonal therapy, angiogenesis inhibitors, and supportive therapy, carry their own risk to induce a hypercoagulable state [8]. Probably the best-studied influence of chemotherapy on the risk of thrombosis is in patients with breast cancer. In early stage breast cancer the risk of VTE is similar to that in the general population (approximately 0.2%). With chemotherapy or hormonal treatment this increases to 1%–2%. When in stages I–II chemotherapy and hormonal therapy are combined the incidence rises to 5%–7% [8,19,20]. For patients with stage IV breast cancer these rates may be as high as 18% with the combined treatments [8]. Other chemotherapeutic agents, such as cisplatin (reported VTE rates of 8%–18%), L-asparaginase (reported VTE rates in adults of 4%–14%) and fluorouracil (reported VTE rates of 15%– 17%), often in combination with other forms of chemotherapy, have been clearly associated in retro and prospective studies with an increased risk of VTE [8]. Thalidomide, particularly in combination with dexamethasone or doxorubicin, may induce VTE rates as high as 20%–40% [8,21]. Whether the newer thalidomide analogs (IMiDs) really have lower VTE complication rates remains to be demonstrated, in particular when combined with other forms of chemotherapy. Initially, angiogenesis inhibitors (such as bevacizumab), again in combination with chemotherapy, were reported to induce higher rates of VTE, but more recent comparisons (with and without bevacizumab) suggest that the contribution of angiogenesis inhibitors may be marginal [8,22]. Finally, supportive therapies associated with an increased risk of VTE include erythropoietin, hemotopoietic colony-stimulating factors, and high-dose corticosteroids. Cancer patients undergoing surgical procedures have an approximately 2-fold higher risk of developing VTE as compared with non-cancer patients [12,23]. Incidences of symptomatic VTE within 91 days after the procedure in cancer patients were as high as 3%–4% for radical cystectomy, nephrostomy, brain surgery, and total hip replacement [23]. Other determinants of the VTE risk in cancer patients include the classical risk factors such as immobility, age, and also prothrombotic abnormalities. Carriers of the factor (F) V Leiden mutation who also had cancer had a 12-fold increased risk of developing VTE as compared with individuals with the mutation but no cancer [13]. This is in agreement with earlier observation and also applies to other genetic thrombophilias, such as the prothrombin mutation [13,24,25]. Cancer patients who have developed VTE have an approximately 3-fold higher risk of experiencing a recurrent thrombosis in the first 12 months as compared with VTE patients without cancer [12,15,26]. Prandoni and colleagues observed recurrent VTE in 6.8% of patients without cancer, whereas in those with cancer the incidence was 20.7% [26]. Developing VTE predicts a worse prognosis in cancer patients [15,27]. The 1-year survival was 12% in patients diagnosed with cancer and VTE at the same time, compared with 36% in cancer


1.00 DVT/PE and Malignant disease

Probability of death

0.80

0.60

Malignant disease 0.40 DVT/PE Only

0.20

Nonmalignant disease

0.00 0

20

40

60

80

100 120 140 160 180

Number of days Fig. 1. Probability of death within 183 days of initial hospital admission [15]. Reproduced with permission of the publisher.

patients without VTE in a population-based study [27]. Figure 1 shows the mortality rates in the first 180 days in patients with cancer and VTE, which are more than 2-fold higher than those in cancer patients without VTE [27]. Venous thromboembolism and cancer

In patients with symptomatic VTE, the incidence of concomitant cancer (i.e. cancer not known before the diagnosis of VTE and discovered by routine investigation) at the time of VTE diagnosis, varies in the larger studies between 4% and 12% [28]. The cancer stage in patients with concurrently diagnosed VTE is often advanced [27,29]. Furthermore, the risk of concomitant cancer is 3- to 4-fold increased in patients with idiopathic VTE compared with secondary VTE [28]. The risk of occult cancer (i.e. cancer that becomes clinically apparent during follow-up) is also increased in patients with VTE. Summarizing 17 cohort and two population-based studies, Otten and Prins calculated an incidence of new cancer in patients with idiopathic VTE of 4%–10%, diagnosed within 3 years after the thrombotic event [28]. Again, this incidence was higher in idiopathic VTE than in patients with secondary VTE. The incidence of cancer is the highest shortly after VTE has been diagnosed. During the first year, the standardized incidence ratio for cancer in patients with VTE is 2.1–4.6, while after this period the risk gradually decreases [30]. In the large Californian registry, White et al. showed that the incidence of cancer was the highest during the first 60 days after an unprovoked episode of VTE (Fig. 2), with a gradual

Total cases with venous thromboembolism

248 H. R. Buller et al 200

Observed

Expected

180 160 140 120 100 80 60 40 20 0

365–305 304–244 243–183 182–122 121–61 Days before diagnosis of cancer

60–0

Fig. 2. Timing of diagnosis of unprovoked venous thromboembolism events relative to the cancer diagnosis date, using 2-month (61-day) increments [29]. Reproduced with permission of the publisher.

decline afterwards [29]. In the first 4 months after VTE diagnosis, the standardized incidence ratio was 2-fold higher than the expected incidence, whereas the observed and expected incidences of cancer 4–12 months after VTE were virtually the same. Also the occurrence of metastasized disease was the highest in the first 4 months after an unprovoked VTE. This indicates that the risk of cancer is only increased in the first months to a year after a VTE, while many patients already have metastases. The risk of developing overt cancer after VTE also depends on the type of cancer. In the large Danish registry, Sørensen and colleagues showed that around 15% of the patients with VTE and cancer within 1 year had lung cancer, followed by prostate (11.4%), pancreas (7.9%), colon (7%), and breast (4.3%) cancer [27]. Leukemia and non-Hodgkin lymphoma were found in 2.5% of the patients [27]. In the Californian cancer registry, White et al. revealed the highest standardized incidence ratios for acute myelogenous leukemia (4.2), followed by ovarian cancer (2.8), non-Hodgkin lymphoma (2.7), pancreatic (2.6), renal cell (2.5), stomach and lung cancer (1.8) [29]. Considering the high incidence of cancer in the first months after VTE, screening for an underlying malignancy may be clinically relevant. The literature is, however, not concordant on whether extensive screening for occult malignancy is indicated. Retrospective studies indicated that a careful medical history, physical examination, and laboratory testing detected most occult cancers in patients with VTE [31–33], suggesting that routine examination at the time of VTE diagnosis is sufficient to detect most underlying malignancies and that additional testing would not be required. Prospective studies seem to indicate that extensive screening is able to detect more cancers than routine examination alone. The six prospective studies that systematically evaluated the added value of more extensive screening compared with routine examination are summarized in Fig. 3 [34–39]. Routine examination was in most studies defined as a combination of careful medical history taking, thorough physical examination, laboratory 2007 International Society on Thrombosis and Haemostasis


Cancer and thrombosis 249

Fig. 3. Detection of occult malignancy in patients with venous thromboembolism by routine screening alone or by routine screening followed by more extensive screening.

testing, and chest X-ray. In most studies, extensive testing included specific tumor markers, such as prostate-specific antigen, carcinoembryonic antigen, or cancer antigen-125 levels, and abdomino-pelvic ultrasonography. Upon indication, this screening was extended with CT-scanning of chest, abdomen or pelvis, gastro-, colono-, or sigmoidoscopy, mammography, or sputum cytology. In the six studies, routine testing followed by more extensive screening was associated with a 1.8-fold (95% CI: 1.28–2.51) higher probability of finding occult cancer, compared with routine testing alone (Fig. 3). The only randomized trial by Piccioli et al., however, found no additional value of extensive screening [38]. Furthermore, after routine screening the incidence of cancer showed a large variation in the different studies. This is likely to be due to differences in patient characteristics, but also to the type of routine screening itself. A predefined protocol focusing on the possibility of occult cancer that clearly describes history taking, physical examination, and laboratory testing is a prerequisite in order to reliably assess the additional value of extensive screening. In addition, another potential source of variation is the definition of extensive screening, which is not well defined. Finally, if extensive screening is advocated, early detection of cancer must lead to a benefit in survival, a matter that has not yet been adequately addressed. An ongoing study may soon provide more information. This multicenter study directly compares routine and extensive testing for occult malignancy in patients with VTE (The Trousseau study). Centers that only used routine testing, consisting of medical history, physical examination, basic laboratory examination, and chest X-ray, are compared with matched centers that additionally perform thoracic and abdominal CT, plus mammography in women. So far 444 patients have been included. The planned total number of patients is 1200. An interim analysis reveals that in the routine testing arm, 16 of 176 had a suspected cancer after the initial assessment, which was confirmed in one patient (0.6%, 95% CI: 0.2–3.1). During a median follow-up of 12 months, six new malignancies were diagnosed (incidence 3.4%, 95% CI: 1.3– 7.3). In the routine and additional screening arm, 30 of 268 patients were thought to have a malignancy after the initial assessment. In 11 patients a malignancy was diagnosed (4.1%, 95% CI: 2.1–7.2). Of these 268 patients, 214 underwent the additional screening tests. Follow-up diagnostic tests were ordered in 49 patients. In four a malignancy was found. During 2007 International Society on Thrombosis and Haemostasis

a median follow-up of 12 months, six new neoplasms were diagnosed (incidence 2.4%, 95% CI: 0.87–5.1) (abstract ISTH). Based on these preliminary data the additional value of CT screening appears to be limited in patients with idiopathic VTE. A standard initial assessment seems effective enough for diagnosing cancer in these patients. Cancer, hemostasis, and heparins This second part, first describes the way cancer cells utilize hemostatic constituents for their growth and progression. Recently, a rise in the interest in the association of cancer and thrombosis was driven by the discovery of potential anticancer effects of anticoagulants in particular heparins. The clinical effects of (LMW) heparins in cancer patients are summarized and the potential mechanisms of the anticancer effects of heparins are discussed. Other anticoagulants are outside the scope of this review. Hemostasis and cancer

Multiple pathways result in cancer growth and progression. Cancer cells gain capacity to proliferate, migrate, and induce proteolysis and permeability. They have a reduced sensitivity to apoptosis and develop functions of increased motility and adhesion [40]. It is now well established that the essential factors of hemostasis (e.g. coagulation and fibrinolytic proteins, and platelets), apart from their procoagulant effects, directly play a role in these processes. However, fibrin formation itself is essential as a supportive scaffold for angiogenesis that ensures the tumorÕs oxygen supply [41–43]. The effects and possible pathways through which the individual factors and their complexes work are summarized in Table 2. Tissue factor levels are associated with clinical progression of cancer and elevated levels are an unfavorable prognostic indicator. Tissue factor gene silencing can inhibit tumor growth and metastasis in vivo [44–46]. The importance of tissue factor for invasive growth is nicely illustrated by the fact that without tissue factor there can be no embryonic development [47]. The aberrant expression of tissue factor in cancer and vascular endothelial cells is stimulated by the tumor as a result of activation of the k-ras oncogene and loss of p53 tumor suppressor gene, as well as through inflammatory signals such as interleukine-1b and tumor necrosis factor-a [48]. These


250 H. R. Buller et al Table 2 The role of hemostatic constituents (coagulation and fibrinolytic proteins and platelets) in the growth and progression of cancer Hemostasis Coagulation proteins Specific interactions Tissue factor TF-FVIIa Factor Xa Thrombin

Pathway/mechanism

Effects on cancer growth and progression

VEGF expression› k-ras expression VEGF› TSPfl TF› Cancer procoagulant TF, VEGF, VEGFR, bFGF, MMP-2

Angiogenesis›, permeability› Metastasis, tumor growth Angiogenesis›, cell survival›, apoptosisfl Proliferation› Altered cell shape, proliferation›, permeability ›, migration›, cell survival›, proteolysis› Angiogenesis›, apoptosisfl

Platelets activation, growth factors release Mobilization of adhesion molecules Procoagulant interactions TF-FVIIa TF-FVIIa-FXa Fibrinolytic proteins u-PA, t-PA PAI-1, PAI-2 Platelets

Motility›

Thrombin/fibrin formation

Matrix for angiogenesis Feedback: TF expression›

Thrombin formation Proteolysis/fibrinolysis Fibrin matrix conservation

Invasion, proliferation Angiogenesis›

Release of growth factors (VEGF, PDGF, bFGF) P-selectin expression NK-cell protection (coating) COX-2›

Angiogenesis›, permeability›, apoptosisfl Tumor cell adhesion ›, migration› Tumor cell survival› Apoptosisfl

bFGF, basic fibroblast growth factor; COX-2, cyclo-oxygenase-2; NK-cell, natural killer cell; PAI-1, plasminogen activator inhibitor type 1; PDGF, platelet-derived growth factor; TF, tissue factor; t-PA, tissue type plasminogen activator; TSP-1, thrombospondin-1; TFPI, tissue factor pathway inhibitor; u-PA, urokinase plaminogen activator; VEGF, vascular endothelial growth factor; VEGFR, vascular endothelial growth factor receptor.

cytokines also influence leukocyte, platelet, and endothelial cell function, resulting in decreased thrombomodulin expression on the endothelial cell surface, impairment of the protein C anticoagulant pathway and enhanced expression of adhesion molecules [46]. Furthermore, tissue factor (isolated as well as in complex with FVIIa) induces vascular-derived growth factor (VEGF) expression, down-regulation of thrombospondin and further tissue factor expression [49–52]. Together, this stimulates angiogenesis, adhesion, and migration [53,54]. The TF–FVIIa complex induces also survival of cells that have been directed toward apoptosis [55,56]. Cancer procoagulant, a cystein protease that is expressed on the surface of many cancer cells, can directly induce activation of FX independent of FVIIa. Hereby, vascular cell proliferation is promoted [46]. Thrombin enhances the expression of tissue factor, VEGF, VEGF-receptor, basic fibroblast growth factor, and matrix metalloproteinase through signaling via the protease activated receptors on the cell surface of cancer cells and/or vascular endothelial cells [57–61]. This leads to changes in endothelial cell shape and proliferation, increased vascular permeability, migration, and enhanced survival of cancer cells [54,62,63]. The factors mentioned above are all well known for their classic role in the hemostatic system, which naturally occurs at the same time. It appears that activating the coagulation system

has advantages for cancer growth and progression, as fibrin is an excellent matrix for angiogenesis. The fibrin matrix also protects against proteolysis. Furthermore, fibrin formation leads to proliferation and migration and in reverse to tissue factor expression, which is probably signaled through the proangiogenetic cytokine IL-8 [46,64,65]. The complex of tissue factor-FVIIa-FXa stays active because of impaired TFPI binding, which results in prothrombin conversion and thrombin formation [43]. Cancer cells can express many fibrinolytic proteins (uPA, tPA, PAI-1, PAI-2) and both promotion as well as inhibition occurs [66]. Cancer cells overall impair the fibrinolytic system, which results in fibrin matrix conservation, tumor invasion, cell proliferation, and metastasis. Activation of platelets by thrombin leads to release of growth factors such as VEGF, platelet-derived growth factor, and basic fibroblast growth factor. These growth factors contribute to angiogenesis and inhibit apoptosis [67,68]. Cancer cells enhance P-selectin expression on monocytes, macrophages, endothelial cells, and platelets [69]. In addition, cancer cells express P-selectin ligands to bind those cells. The latter provides protection against natural killer cells and therefore leads to tumor cell survival and enhanced adhesion to the endothelium [70]. Platelets promote COX-2 synthesis in monocytes and it has been shown that COX-2 overexpression leads to inhibited apoptosis, abnormal cell-to-cell interactions and a more 2007 International Society on Thrombosis and Haemostasis


Cancer and thrombosis 251

invasive phenotype. Furthermore, prostaglandins (COX-2 products and mediators of classic inflammation) might suppress host immunity against tumors [71,72]. More recently, coagulation system activation has been related to hypoxia driven overexpression or mutation of the MET-oncogene (physiologic invasive growth). Boccaccio and colleagues induced disseminated intravascular coagulation and the Trousseau syndrome by targeting the human METoncogene in the mouse liver, resulting in enhanced PAI-1 and COX-2 expression [73–75]. In glioblastomas it was found that a tumor suppressor gene (PTEN) and hypoxia-regulated tissue factor expression can induce intravascular thrombotic occlusion [76]. Cancer and heparins

For many decades, experimental and clinical studies have evaluated the effects of anticoagulants on tumor growth with different outcomes. In 1992, Prandoni et al. observed an important reduction in total mortality in the subgroup of VTE patients with cancer at enrolment that received LMWH. At 3 months, 44% (eight of 18) of the cancer patients died in the standard heparin group vs. 7% (one of 15) in the LMWH group (P = 0.021) [77]. These findings of a survival benefit of LMWH in cancer patients were confirmed in a meta-analysis of nine studies that compared LMWH with standard heparin in the treatment of VTE [78]. These results have led to clinical trials that evaluated the effect of (LMW) heparin on survival in cancer patients without thrombosis. Figure 4A presents the overall survival results at 6 months after study entry of the five randomized controlled trials that primarily evaluated the effects of (LMW) heparin on cancer progression [79–83]. Overall, there is no difference in survival (odds ratio 0.91; 95% CI: 0.71–1.16). These studies, however,

differ considerably in study design. In one study UFH was administered, in the others LMWH [79]. In two of the studies only patients with small cell lung cancer were included [79,80], whereas patients with different cancer types were enrolled in the other studies. The duration of heparin therapy differed, from 5 to 52 weeks. In one study a therapeutic dose of LMWH was given [82]. Three studies defined a priori a subgroup consisting of patients with a prognosis of at least 6 months at study entry (also referred to as limited disease) [79,80,82]. Figure 4B shows the survival in this subgroup. Patients treated with (LMW) heparin have a significant survival benefit (odds ratio 0.47; 0.30–0.75). Subgroup analysis indicated no benefit for patients with advanced disease (odds ratio 0.96; 0.77–1.21). Post hoc analyses of other studies are in agreement with these positive findings in those with limited disease [81,83,84]. These clinical results are supported by proposed mechanisms of heparin on cancer progression, mainly based on in vitro and in vivo studies. The possible mechanisms by which heparins impair cancer growth and progression include decrease of thrombin and fibrin formation, binding to adhesion molecules, antiheparanase effects and increase of apoptosis. Heparins have well-known anticoagulant effects resulting in inhibition of thrombin and fibrin formation. As discussed above, both thrombin and fibrin are important for cancer growth and progression. Besides its coagulant effects, heparin is thought to have other properties. This is illustrated by the antimetastatic effects of non-anticoagulant heparins [85]. The first non-anticoagulant mechanism of heparin is interfering with the interaction between cancer cells and platelets. Specific oligosaccharide structures in heparin can bind to P-selectins exposed on activated platelets [86]; these structures show minimal anticoagulant activity [87]. Ludwig and colleagues tested the impact of UFH, LMWH (nadroparine and enoxaparin), and fondaparinux on P-selectin-dependent tumor

A

B

Fig. 4. (A) Heparin vs. placebo or no anticoagulants in patients with cancer. Overall survival after 6 months. (B) Heparin vs. placebo or no anticoagulants in patients with cancer. Survival after 6 months in patients with limited disease, with a similar definition at study entry. UFH, unfractioned heparin; LMWH, low-molecular weight heparin. 2007 International Society on Thrombosis and Haemostasis


252 H. R. Buller et al

interactions in vitro and metastasis formation in vivo. These agents differ widely in their potential to interfere with Pselectin-mediated cell binding. This inhibitory function strongly correlates with the potency in inhibiting experimental lung metastasis in vivo [88,89]. Selectins are also expressed on leukocytes (L-selectin) and the vascular endothelium (E- and Pselectin). Heparin can inhibit L-selectin binding to its ligands, resulting in inhibition of the inflammatory response. Extravasation of cancer cells is also hindered by binding of heparin to selectins. [86,90,91]. The second non-anticoagulant property of heparin is inhibition of cancer cell heparanase [92]. Heparanase is expressed in many human cancers. This enzyme breaks down the polysaccharide barrier, facilitating cancer-cell invasion through the vascular basement membrane; also angiogenesis and metastases are promoted [93]. In humans elevated heparanase expression by the tumor has been correlated with a more aggressive behavior in breast, colon, ovary, pancreas, non-small lung cancer, acute myeloid leukemia, and myeloma tumors [94–101]. Heparanase also cleaves the growth factor bearing heparan sulfate groups from heparan sulfate proteoglycans in the extracellular matrix and cellular membranes [102– 104]. Heparins can directly inhibit the effect of growth factors as well. LMWH hinders binding of growth factors to highaffinity receptors to a greater degree than UFH [105]. Heparin, especially LMWH, is thought to increase the apoptosis index in cancer cells. Nasir et al. observed in heparin knockout mice a 4-fold increase in the apoptosis index with LMWH as compared with controls [106]. Heparin probably enters the cell by endocytosis and induces apoptosis by interfering with transcription factor [107]. Disclosure of conflict of interest HRB has received grants from Sanofi-Aventis, GSK, and Bayer. The other authors have not declared any conflicts of interest. References 1 Bouillaud S. De lÕObliteration des veines et de son influence sur la formation des hydropisies partielles: consideration sur la hydropisies passive et general. Arch Gen Med 1823; 1: 188–204. 2 Otten HM. Thrombosis and cancer, Thesis, Academic Medical Center, University of Amsterdam, 2002. 3 Trousseau A. Phlegmasia alba dolens. In: Trousseau A, ed. Clinique medicinale de lÕHotel-Dieu de Paris. Paris: Bailliere J.-B. et fils, 1865: 645–712. 4 Trousseau A. Ulcere chronique simple de lÕestomac. In: Peter M, ed. Clinique Medicinale de lÕHotel-Dieu de Paris.Paris: Bailliere J.-B. et fils: 1877: p. 80–107. 5 Trousseau A. Phlegmasia alba dolens. In: Peter M, ed. Clinique Medicinale de lÕHotel-Dieu de Paris.Paris: Bailliere J.-B. et fils: 1877: p. 695–739. 6 Illtyd James T, Matheson N. Thromboplebitis in cancer. Practitioner 1935; 134: 683–4. 7 Gore JM, Appelbaum JS, Greene HL, Dexter L, Dalen JE. Occult cancer in patients with acute pulmonary embolism. Ann Intern Med 1982; 96: 556–60.

8 Haddad TC, Greeno EW. Chemotherapy-induced thrombosis. Thromb Res 2006; 118: 555–68. 9 Prandoni P, Falanga A, Piccioli A. Cancer and venous thromboembolism. Lancet Oncol 2005; 6: 401–10. 10 Rickles FR, Edwards RL. Activation of blood coagulation in cancer: TrousseauÕs syndrome revisited. Blood 1983; 62: 14–31. 11 Shen VS, Pollak EW. Fatal pulmonary embolism in cancer patients: is heparin prophylaxis justified? South Med J 1980; 73: 841–3. 12 Heit JA, Mohr DN, Silverstein MD, Petterson TM, OÕFallon WM, Melton LJ III. Predictors of recurrence after deep vein thrombosis and pulmonary embolism: a population-based cohort study. Arch Intern Med 2000; 160: 761–8. 13 Blom JW, Doggen CJ, Osanto S, Rosendaal FR. Malignancies, prothrombotic mutations, and the risk of venous thrombosis. JAMA 2005; 293: 715–22. 14 Chew HK, Wun T, Harvey D, Zhou H, White RH. Incidence of venous thromboembolism and its effect on survival among patients with common cancers. Arch Intern Med 2006; 166: 458–64. 15 Levitan N, Dowlati A, Remick SC, Tahsildar HI, Sivinski LD, Beyth R, Rimm AA. Rates of initial and recurrent thromboembolic disease among patients with malignancy vs. those without malignancy. Risk analysis using Medicare claims data. Medicine (Baltimore) 1999; 78: 285–91. 16 Thodiyil PA, Kakkar AK. Variation in relative risk of venous thromboembolism in different cancers. Thromb Haemost 2002; 87: 1076–7. 17 Heit JA. Cancer and venous thromboembolism: scope of the problem. Cancer Control 2005; 12 (Suppl. 1): 5–10. 18 Otten HM, Mathijssen J, Ten Cate H, Soesan M, Inghels M, Richel DJ, Prins MH. Symptomatic venous thromboembolism in cancer patients treated with chemotherapy: an underestimated phenomenon. Arch Intern Med 2004; 164: 190–4. 19 Otten HM, Prins MH, Smorenburg SM, Hutten BA. Risk assessment and prophylaxis of venous thromboembolism in non-surgical patients: cancer as a risk factor. Haemostasis 2000; 30 (Suppl. 2): 72–6. 20 Levine MN. Prevention of thrombotic disorders in cancer patients undergoing chemotherapy. Thromb Haemost 1997; 78: 133–6. 21 Zangari M, Barlogie B, Anaissie E, Saghafifar F, Eddlemon P, Jacobson J, Lee CK, Thertulien R, Talamo G, Thomas T, Van RF, Fassas A, Fink L, Tricot G. Deep vein thrombosis in patients with multiple myeloma treated with thalidomide and chemotherapy: effects of prophylactic and therapeutic anticoagulation. Br J Haematol 2004; 126: 715–21. 22 Hurwitz H, Fehrenbacher L, Novotny W, Cartwright T, Hainsworth J, Heim W, Berlin J, Baron A, Griffing S, Holmgren E, Ferrara N, Fyfe G, Rogers B, Ross R, Kabbinavar F. Bevacizumab plus irinotecan, fluorouracil, and leucovorin for metastatic colorectal cancer. N Engl J Med 2004; 350: 2335–42. 23 White RH, Zhou H, Romano PS. Incidence of symptomatic venous thromboembolism after different elective or urgent surgical procedures. Thromb Haemost 2003; 90: 446–55. 24 Otterson GA, Monahan BP, Harold N, Steinberg SM, Frame JN, Kaye FJ. Clinical significance of the FV:Q506 mutation in unselected oncology patients. Am J Med 1996; 101: 406–12. 25 Pihusch R, Danzl G, Scholz M, Harich D, Pihusch M, Lohse P, Hiller E. Impact of thrombophilic gene mutations on thrombosis risk in patients with gastrointestinal carcinoma. Cancer 2002; 94: 3120–6. 26 Prandoni P, Lensing AW, Piccioli A, Bernardi E, Simioni P, Girolami B, Marchiori A, Sabbion P, Prins MH, Noventa F, Girolami A. Recurrent venous thromboembolism and bleeding complications during anticoagulant treatment in patients with cancer and venous thrombosis. Blood 2002; 100: 3484–8. 27 Sorensen HT, Mellemkjaer L, Olsen JH, Baron JA. Prognosis of cancers associated with venous thromboembolism. N Engl J Med 2000; 343: 1846–50. 28 Otten HM, Prins MH. Venous thromboembolism and occult malignancy. Thromb Res 2001; 102: V187–94. 2007 International Society on Thrombosis and Haemostasis


Cancer and thrombosis 253 29 White RH, Chew HK, Zhou H, Parikh-Patel A, Harris D, Harvey D, Wun T. Incidence of venous thromboembolism in the year before the diagnosis of cancer in 528,693 adults. Arch Intern Med 2005; 165: 1782–7. 30 Lee AY. Thrombosis and cancer: the role of screening for occult cancer and recognizing the underlying biological mechanisms. Hematol Am Soc Hematol Educ Program 2006; 438–43. 31 Cornuz J, Pearson SD, Creager MA, Cook EF, Goldman L. Importance of findings on the initial evaluation for cancer in patients with symptomatic idiopathic deep venous thrombosis. Ann Intern Med 1996; 125: 785–93. 32 Nordstrom M, Lindblad B, Anderson H, Bergqvist D, Kjellstrom T. Deep venous thrombosis and occult malignancy: an epidemiological study. BMJ 1994; 308: 891–4. 33 Hettiarachchi RJ, Lok J, Prins MH, Buller HR, Prandoni P. Undiagnosed malignancy in patients with deep vein thrombosis: incidence, risk indicators, and diagnosis. Cancer 1998; 83: 180–5. 34 Bastounis EA, Karayiannakis AJ, Makri GG, Alexiou D, Papalambros EL. The incidence of occult cancer in patients with deep venous thrombosis: a prospective study. J Intern Med 1996; 239: 153–6. 35 Monreal M, Lafoz E, Casals A, Inaraja L, Montserrat E, Callejas JM, Martorell A. Occult cancer in patients with deep venous thrombosis. A systematic approach. Cancer 1991; 67: 541–5. 36 Monreal M, Fernandez-Llamazares J, Perandreu J, Urrutia A, Sahuquillo JC, Contel E. Occult cancer in patients with venous thromboembolism: which patients, which cancers. Thromb Haemost 1997; 78: 1316–8. 37 Monreal M, Lensing AW, Prins MH, Bonet M, Fernandez-Llamazares J, Muchart J, Prandoni P, Jimenez JA. Screening for occult cancer in patients with acute deep vein thrombosis or pulmonary embolism. J Thromb Haemost 2004; 2: 876–81. 38 Piccioli A, Lensing AW, Prins MH, Falanga A, Scannapieco GL, Ieran M, Cigolini M, Ambrosio GB, Monreal M, Girolami A, Prandoni P. Extensive screening for occult malignant disease in idiopathic venous thromboembolism: a prospective randomized clinical trial. J Thromb Haemost 2004; 2: 884–9. 39 Sannella NA, OÕConnor DJ Jr. ‘‘Idiopathic’’ deep venous thrombosis: the value of routine abdominal and pelvic computed tomographic scanning. Ann Vasc Surg 1991; 5: 218–22. 40 Bogenrieder T, Herlyn M. Axis of evil: molecular mechanisms of cancer metastasis. Oncogene 2003; 22: 6524–36. 41 Rak J, Yu JL, Luyendyk J, Mackman N. Oncogenes, trousseau syndrome, and cancer-related changes in the coagulome of mice and humans. Cancer Res 2006; 66: 10643–6. 42 Bromberg ME, Konigsberg WH, Madison JF, Pawashe A, Garen A. Tissue factor promotes melanoma metastasis by a pathway independent of blood coagulation. PNAS 1995; 92: 8205–9. 43 Rickles FR, Patierno S, Fernandez PM. Tissue factor, thrombin, and cancer. Chest 2003; 124 (Suppl. 3): 58S–68S. 44 Yu J, May L, Klement P, Weitz J, Janusz MD. Oncogenes as regulators of tissue factor expression in cancer: implications for tumor angiogenesis and anti-cancer therapy. Semin Thromb Hemost 2004; 1: 21–30. 45 Kakkar AK, Lemoine NR, Scully MF, Tebbutt S, Williamson RC. Tissue factor expression correlates with histological grade in human pancreatic cancer. Br J Surg 1995; 82: 1101–4. 46 Rickles FR, Falanga A. Molecular basis for the relationship between thrombosis and cancer. Thromb Res 2001; 102: V215–24. 47 Bugge TH, Xiao Q, Kombrinck KW, Flick MJ, Holmback K, Danton MJ, Colbert MC, Witte DP, Fujikawa K, Davie EW, Degen JL. Fatal embryonic bleeding events in mice lacking tissue factor, the cell-associated initiator of blood coagulation. PNAS 1996; 93: 6258–63. 48 Grignani G, Maiolo A. Cytokines and hemostasis. Haematologica 2000; 85: 967–72. 49 Rao LVM, Pendurthi UR. Tissue factor-factor VIIa signaling. Arterioscler Thromb Vasc Biol 2005; 25: 47–56. 2007 International Society on Thrombosis and Haemostasis

50 Belting M, Ahamed J, Ruf W. Signaling of the tissue factor coagulation pathway in angiogenesis and cancer. Arterioscler Thromb Vasc Biol 2005; 25: 1545–50. 51 Mechtcheriakova D, Wlachos A, Holzmuller H, Binder BR, Hofer E. Vascular endothelial cell growth factor-induced tissue factor expression in endothelial cells is mediated by EGR-1. Blood 1999; 93: 3811– 23. 52 Abe K, Shoji M, Chen J, Bierhaus A, Danave I, Micko C, Casper K, Dillehay DL, Nawroth PP, Rickles FR. Regulation of vascular endothelial growth factor production and angiogenesis by the cytoplasmic tail of tissue factor. PNAS 1999; 96: 8663–8. 53 Ott I, Fischer EG, Miyagi Y, Mueller BM, Ruf W. A role for tissue factor in cell adhesion and migration mediated by interaction with actin-binding protein 280. J Cell Biol 1998; 140: 1241–53. 54 Mueller BM, Ruf W. Requirement for binding of catalytically active factor VIIa in tissue factor-dependent experimental metastasis. J Clin Invest 1998; 101: 1372–8. 55 Flynn AN, Buret AG. Proteinase-activated receptor 1 (PAR-1) and cell apoptosis. Apoptosis 2004; V9: 729–37. 56 Sorensen BB, Rao LVM, Tornehave D, Gammeltoft S, Petersen LC. Antiapoptotic effect of coagulation factor VIIa. Blood 2003; 102: 1708–15. 57 Coughlin SR. Thrombin signalling and protease-activated receptors. Nature 2000; 407: 258–64. 58 Dvorak HF, Detmar M, Claffey KP, Nagy JA, van de WL, Senger DR. Vascular permeability factor/vascular endothelial growth factor: an important mediator of angiogenesis in malignancy and inflammation. Int Arch Allergy Immunol 1995; 107: 233–5. 59 Ruf W, Dorfleutner A, Riewald M. Specificity of coagulation factor signaling. J Thromb Haemost 2003; 1: 1–495. 60 Clauss M, Gerlach M, Gerlach H, Brett J, Wang F, Familletti PC, Pan YC, Olander JV, Connolly DT, Stern D. Vascular permeability factor: a tumor-derived polypeptide that induces endothelial cell and monocyte procoagulant activity, and promotes monocyte migration. J Exp Med 1990; 172: 1535–45. 61 Brooks PC, Stromblad S, Sanders LC, von Schalscha TL, Aimes RT, Stetler-Stevenson WG, Quigley JP, Cheresh DA. Localization of matrix metalloproteinase MMP-2 to the surface of invasive cells by interaction with integrin alpha v beta 3. Cell 1996; 85: 683– 93. 62 Fischer EG, Riewald M, Huang HY, Miyagi Y, Kubota Y, Mueller BM, Ruf W. Tumor cell adhesion and migration supported by interaction of a receptor-protease complex with its inhibitor. J Clin Invest 1999; 104: 1213–21. 63 Hu L, Lee M, Campbell W, Perez-Soler R, Karpatkin S. Role of endogenous thrombin in tumor implantation, seeding, and spontaneous metastasis. Blood 2004; 104: 2746–51. 64 Koolwijk P, Sidenius N, Peters E, Sier CFM, Hanemaaijer R, Blasi F, van Hinsbergh VWM. Proteolysis of the urokinase-type plasminogen activator receptor by metalloproteinase-12: implication for angiogenesis in fibrin matrices. Blood 2001; 97: 3123–31. 65 Qi J, Kreutzer DL. Fibrin activation of vascular endothelial cells. Induction of IL-8 expression. J Immunol 1995; 155: 867–76. 66 Kwaan HC. The plasminogen-plasmin system in malignancy. Cancer Metastasis Rev 1992; V11: 291–311. 67 Kennedy SG, Kandel ES, Cross TK, Hay N. Akt/protein kinase B inhibits cell death by preventing the release of cytochrome c from mitochondria. Mol Cell Biol 1999; 19: 5800–10. 68 Detmar M. Tumor angiogenesis. J Investig Dermatol Symp Proc 2000; 5: 20–3. 69 Gupta GP, Massague J. Platelets and metastasis revisited: a novel fatty link. J Clin Invest 2004; 114: 1691–3. 70 Nieswandt B, Hafner M, Echtenacher B, Mannel DN. Lysis of tumor cells by natural killer cells in mice is impeded by platelets. Cancer Res 1999; 59: 1295–300. 71 Dixon DA, Tolley ND, Bemis-Standoli K, Martinez ML, Weyrich AS, Morrow JD, Prescott SM, Zimmerman GA. Expression of


254 H. R. Buller et al

72

73

74 75

76

77

78

79

80

81

82

83

84

85

86

87 88

COX-2 in platelet-monocyte interactions occurs via combinatorial regulation involving adhesion and cytokine signaling. J Clin Invest 2006; 116: 2727–38. Tsuji S, Tsujii M, Kawano S, Hori M. Cyclooxygenase-2 upregulation as a perigenetic change in carcinogenesis. J Exp Clin Cancer Res 2001; 20: 117–29. Boccaccio C, Sabatino G, Medico E, Girolami F, Follenzi A, Reato G, Sottile A, Naldini L, Comoglio PM. The MET oncogene drives a genetic programme linking cancer to haemostasis. Nature 2005; 434: 396–400. Boccaccio C, Comoglio PM. A functional role for hemostasis in early cancer development. Cancer Res 2005; 65: 8579–82. Pennacchietti S, Michieli P, Galluzzo M, Mazzone M, Giordano S, Comoglio PM. Hypoxia promotes invasive growth by transcriptional activation of the met protooncogene. Cancer Cell 2003; 3: 347–61. Rong Y, Post DE, Pieper RO, Durden DL, Van Meir EG, Brat DJ. PTEN and hypoxia regulate tissue factor expression and plasma coagulation by glioblastoma. Cancer Res 2005; 65: 1406–13. Prandoni P, Lensing AW, Buller HR, Carta M, Cogo A, Vigo M, Casara D, Ruol A, ten Cate JW. Comparison of subcutaneous lowmolecular-weight heparin with intravenous standard heparin in proximal deep-vein thrombosis. Lancet 1992; 339: 441–5. Hettiarachchi RJ, Smorenburg SM, Ginsberg J, Levine M, Prins MH, Buller HR. Do heparins do more than just treat thrombosis? The influence of heparins on cancer spread Thromb Haemost 1999; 82: 947–52. Lebeau B, Chastang C, Brechot JM, Capron F, Dautzenberg B, Delaisements C, Mornet M, Brun J, Hurdebourcq JP, Lemarie E. Subcutaneous heparin treatment increases survival in small cell lung cancer. ‘‘Petites Cellules’’ Group. Cancer 1994; 74: 38–45. Altinbas M, Coskun HS, Er O, Ozkan M, Eser B, Unal A, Cetin M, Soyuer S. A randomized clinical trial of combination chemotherapy with and without low-molecular-weight heparin in small cell lung cancer. J Thromb Haemost 2004; 2: 1266–71. Kakkar AK, Levine MN, Kadziola Z, Lemoine NR, Low V, Patel HK, Rustin G, Thomas M, Quigley M, Williamson RC. Low molecular weight heparin, therapy with dalteparin, and survival in advanced cancer: the fragmin advanced malignancy outcome study (FAMOUS). J Clin Oncol 2004; 22: 1944–8. Klerk CP, Smorenburg SM, Otten HM, Lensing AW, Prins MH, Piovella F, Prandoni P, Bos MM, Richel DJ, van TG, Buller HR. The effect of low molecular weight heparin on survival in patients with advanced malignancy. J Clin Oncol 2005; 23: 2130–5. Sideras K, Schaefer PL, Okuno SH, Sloan JA, Kutteh L, Fitch TR, Dakhil SR, Levitt R, Alberts SR, Morton RF, Rowland KM, Novotny PJ, Loprinzi CL. Low-molecular-weight heparin in patients with advanced cancer: a phase 3 clinical trial. Mayo Clin Proc 2006; 81: 758–67. Lee AY, Rickles FR, Julian JA, Gent M, Baker RI, Bowden C, Kakkar AK, Prins M, Levine MN. Randomized comparison of low molecular weight heparin and coumarin derivatives on the survival of patients with cancer and venous thromboembolism. J Clin Oncol 2005; 23: 2123–9. Kragh M, Loechel F. Non-anti-coagulant heparins: a promising approach for prevention of tumor metastasis (review). Int J Oncol 2005; 27: 1159–67. Wang L, Brown JR, Varki A, Esko JD. HeparinÕs anti-inflammatory effects require glucosamine 6-O-sulfation and are mediated by blockade of L- and P-selectins. J Clin Invest 2002; 110: 127–36. Esko JD, Lindahl U. Molecular diversity of heparan sulfate. J Clin Invest 2001; 108: 169–73. Ludwig RJ, Alban S, Bistrian R, Boehncke WH, Kaufmann R, Henschler R, Gille J. The ability of different forms of heparins to suppress P-selectin function in vitro correlates to their inhibitory capacity on bloodborne metastasis in vivo. Thromb Haemost 2006; 95: 535–40.

89 Stevenson JL, Choi SH, Varki A. Differential metastasis inhibition by clinically relevant levels of heparins–correlation with selectin inhibition, not antithrombotic activity. Clin Cancer Res 2005; 11: 7003–11. 90 Varki NM, Varki A. Heparin inhibition of selectin-mediated interactions during the hematogenous phase of carcinoma metastasis: rationale for clinical studies in humans. Semin Thromb Hemost 2002; 28: 53–66. 91 Nelson RM, Cecconi O, Roberts WG, Aruffo A, Linhardt RJ, Bevilacqua MP. Heparin oligosaccharides bind L- and P-selectin and inhibit acute inflammation. Blood 1993; 82: 3253–8. 92 Parish CR, Freeman C, Brown KJ, Francis DJ, Cowden WB. Identification of sulfated oligosaccharide-based inhibitors of tumor growth and metastasis using novel in vitro assays for angiogenesis and heparanase activity. Cancer Res 1999; 59: 3433–41. 93 Vlodavsky I, Friedmann Y, Elkin M, Aingorn H, Atzmon R, IshaiMichaeli R, Bitan M, Pappo O, Peretz T, Michal I, Spector L, Pecker I. Mammalian heparanase: gene cloning, expression and function in tumor progression and metastasis. Nat Med 1999; 5: 793–802. 94 Maxhimer JB, Quiros RM, Stewart R, Dowlatshahi K, Gattuso P, Fan M, Prinz RA, Xu X. Heparanase-1 expression is associated with the metastatic potential of breast cancer. Surgery 2002; 132: 326–33. 95 Friedmann Y, Vlodavsky I, Aingorn H, Aviv A, Peretz T, Pecker I, Pappo O. Expression of heparanase in normal, dysplastic, and neoplastic human colonic mucosa and stroma. Evidence for its role in colonic tumorigenesis. Am J Pathol 2000; 157: 1167–75. 96 Ginath S, Menczer J, Friedmann Y, Aingorn H, Aviv A, Tajima K, Dantes A, Glezerman M, Vlodavsky I, Amsterdam A. Expression of heparanase, Mdm2, and erbB2 in ovarian cancer. Int J Oncol 2001; 18: 1133–44. 97 Gohji K, Okamoto M, Kitazawa S, Toyoshima M, Dong J, Katsuoka Y, Nakajima M. Heparanase protein and gene expression in bladder cancer. J Urol 2001; 166: 1286–90. 98 Koliopanos A, Friess H, Kleeff J, Shi X, Liao Q, Pecker I, Vlodavsky I, Zimmermann A, Buchler MW. Heparanase expression in primary and metastatic pancreatic cancer. Cancer Res 2001; 61: 4655–9. 99 Bitan M, Polliack A, Zecchina G, Nagler A, Friedmann Y, Nadav L, Deutsch V, Pecker I, Eldor A, Vlodavsky I, Katz BZ. Heparanase expression in human leukemias is restricted to acute myeloid leukemias. Exp Hematol 2002; 30: 34–41. 100 Takahashi H, Ebihara S, Okazaki T, Suzuki S, Asada M, Kubo H, Sasaki H. Clinical significance of heparanase activity in primary resected non-small cell lung cancer. Lung Cancer 2004; 45: 207–14. 101 Yang Y, Macleod V, Bendre M, Huang Y, Theus AM, Miao HQ, Kussie P, Yaccoby S, Epstein J, Suva LJ, Kelly T, Sanderson RD. Heparanase promotes the spontaneous metastasis of myeloma cells to bone. Blood 2005; 105: 1303–9. 102 Ishai-Michaeli R, Eldor A, Vlodavsky I. Heparanase activity expressed by platelets, neutrophils, and lymphoma cells releases active fibroblast growth factor from extracellular matrix. Cell Regul 1990; 1: 833–42. 103 Whitelock JM, Murdoch AD, Iozzo RV, Underwood PA. The degradation of human endothelial cell-derived perlecan and release of bound basic fibroblast growth factor by stromelysin, collagenase, plasmin, and heparanases. J Biol Chem 1996; 271: 10079–86. 104 Sanderson RD, Yang Y, Suva LJ, Kelly T. Heparan sulfate proteoglycans and heparanase – partners in osteolytic tumor growth and metastasis. Matrix Biol 2004; 23: 341–52. 105 Jayson GC, Gallagher JT. Heparin oligosaccharides: inhibitors of the biological activity of bFGF on Caco-2 cells. Br J Cancer 1997; 75: 9– 16. 106 Nasir FA, Patel HK, Scully MF, Fareed J, Lemoine NR, Kakkar AK. The low molecular weight heparins dalteparin sodium inhibits angiogenesis and induces apoptosis in an experimental tumour model (abstract 2993). Blood 2003; 102. 107 Linhardt RJ. Heparin-induced cancer cell death. Chem Biol 2004; 11: 420–2.

2007 International Society on Thrombosis and Haemostasis


Issuu converts static files into: digital portfolios, online yearbooks, online catalogs, digital photo albums and more. Sign up and create your flipbook.