throbber
Factor Xa Inhibitors: Next-Generation Antithrombotic Agents
`
`Donald J. P. Pinto, Joanne M. Smallheer, Daniel L. Cheney, Robert M. Knabb, and Ruth R. Wexler*
`
`Research and Development, Bristol-Myers Squibb Company, P.O. Box 5400, Princeton, New Jersey, 08543
`
`Received February 3, 2010
`
`1. Introduction
`
`Thrombosis is the underlying cause of a host of common,
`debilitating, and often fatal cardiovascular disorders. Forma-
`tion of thrombi in the arterial circulation can lead to acute
`myocardial infarction (MIa) or ischemic stroke. In the venous
`circulation, deep vein thrombosis (DVT) may result in chronic
`leg pain, swelling, and ulceration and can, if partially or fully
`dislodged, be followed by life-threatening pulmonary embolism
`(PE). The public health consequences of thromboembolic
`disease are vast. For example, approximately 2.5 million people
`in the United States are affected by atrial fibrillation (AF), a
`cardiac arrhythmia associated with a 4- to 5-fold increase in
`the risk of stroke of primarily cardioembolic origin.1 Another
`group at elevated risk of ischemic stroke, as well as recurrent
`acute MI, is the large and growing population of patients with
`acute coronary syndrome (ACS).2 Furthermore, it has been
`estimated that DVT and PE, which together comprise venous
`thromboembolism (VTE), afflict up to 600 000 individuals in
`the United States each year and are implicated in at least
`100 000 deaths.3
`Despite the continued morbidity and mortality caused
`by thromboembolic disease, recent advances in drug develop-
`ment provide cause for optimism that we are about to enter a
`new era in antithrombotic therapy. One of the most important
`advances has been the development, and recent introduction
`into clinical practice, of a new class of anticoagulants, the
`direct factor Xa (FXa) inhibitors.4,5 In this Perspective, we
`provide a detailed insight into the development of these
`important new agents, describe how structure-based design
`played a pivotal role in this process, and review the wealth of
`preclinical and clinical data that have emerged to date.
`Finally, we consider the issues that will determine the future
`impact of the direct FXa inhibitors on clinical practice. We
`
`*To whom correspondence should be addressed. Phone: (609) 818-
`5450. Fax: (609) 818-3331. E-mail: ruth.wexler@bms.com.
`aAbbreviations: ACS, acute coronary syndrome; AF, atrial fibrilla-
`tion; aPTT, activated partial thromboplastin time; ATIII, antithrombin
`III; AV, arteriovenous; DTI, direct thrombin inhibitor; DVT, deep vein
`thrombosis; ECAT, electric-current-induced arterial thrombosis; FVa,
`factor Va; FVIIa, factor VIIa; FVIII, factor FVIII; FVIIIa, factor
`FVIIIa; FIX, factor IX; FX, factor X; FXa, factor Xa ; FXIa, factor
`XIa; GPIIb/IIIa, glycoprotein IIb/IIIa; HLM, human liver micro-
`somes; hERG, human ether-a-go-go related gene; HTS, high-through-
`put screen; LMWH, low-molecular-weight heparin; MI, myocardial
`infarction; PD, pharmacodynamic; PDB, Protein Data Bank; PE,
`pulmonary embolism; PK, pharmacokinetic; PT, prothrombin time;
`SAR, structure-activity relationship; TAFI, thrombin-activatable fi-
`brinolysis inhibitor; TAP, tick anticoagulant peptide; TF, tissue factor;
`TG, thrombin generation; THR, total hip replacement; TKR, total knee
`replacement; UFH, unfractionated heparin; VKA, vitamin K antago-
`nist; VTE, venous thromboembolism; vWF, von Willebrand factor.
`
`begin by briefly highlighting the limitations of the current
`standards of care in antithrombotic therapy, reviewing some
`key concepts in hemostasis and thrombosis, and explaining
`the rationale for targeting FXa.
`1.1. Current Antithrombotic Therapy. Numerous clinical
`trials have confirmed the efficacy of traditional anticoagu-
`lants, including vitamin K antagonists (VKAs), unfraction-
`ated heparin (UFH), and low-molecular-weight heparins
`(LMWHs, fractionated heparin with reduced activity toward
`thrombin compared to UFH), in the prevention and treat-
`ment of a range of arterial and venous thromboembolic
`diseases.1,6-8 Despite the fact that these drugs are the current
`standard of care and despite their proven efficacy, these
`anticoagulants all possess significant limitations that restrict
`their usefulness in the clinic and have created the need for
`new therapies. Use of warfarin and other VKAs is especially
`problematic, even though these anticoagulants offer the
`convenience of oral administration.9 For example, warfarin
`is associated with numerous drug and food interactions, an
`unpredictable pharmacokinetic (PK) and pharmacodynamic
`(PD) profile, and considerable intra- and interpatient varia-
`bility in drug response.9 As a result, the appropriate ther-
`apeutic dose varies, necessitating monitoring and frequent
`dose adjustment. Monitoring of warfarin therapy is critical
`because of this variability and relatively narrow therapeutic
`index, which often leads to subtherapeutic anticoagulation
`and a higher risk of thromboembolism or to excessive anti-
`coagulation and an increased risk of bleeding.9 Furthermore,
`the delayed onset of action of warfarin means that in critical
`situations therapy must be initiated with a rapid-acting,
`parenteral anticoagulant. Urgent surgical or invasive proce-
`dures may also be complicated by the fact that the anti-
`coagulant effects of warfarin are retained for several days
`after discontinuation of treatment.
`Agents for short-term anticoagulation include UFH,
`LMWHs, the indirect FXa inhibitor fondaparinux, and
`the direct thrombin inhibitors (DTIs) argatroban, bivali-
`rudin, and hirudin. These anticoagulants all require paren-
`teral administration, which makes their use outside the
`hospital problematic and which can also be associated with
`injection-site hematomas. UFH and, to a lesser extent,
`LMWHs carry the risk of thrombocytopenia, and since they
`are produced from animal tissue, they are sometimes asso-
`ciated with serious side effects.10 In addition, UFH has an
`unpredictable PK profile and anticoagulant response that
`necessitate monitoring.5 The limitations of parenteral anti-
`coagulants, particularly the need for injection, mean that
`warfarin and other VKAs, which in many countries are still
`the only orally administered anticoagulants approved for
`
`r 2010 American Chemical Society
`
`Published on Web 05/26/2010
`
`pubs.acs.org/jmc
`
`J. Med. Chem. 2010, 53, 6243–6274 6243
`DOI: 10.1021/jm100146h
`
`BMS 2001
`CFAD v. BMS
`IPR2015-01723
`
`

`
`use, represent the sole viable option for long-term anti-
`coagulation therapy.
`As discussed below, hemostasis and thrombosis are inter-
`related processes, and it is not surprising that anticoagulant
`and antiplatelet agents may have hemorrhagic complica-
`tions. Dose selection, therefore, is primarily influenced by
`the maximum dose that can be safely administered rather
`than the dose providing greatest efficacy. Although this will
`remain true for new antithrombotic agents on the horizon,
`benefit-to-risk profiles do vary.11 These limitations result in
`the underutilization of existing therapies in the very patients
`who are at highest risk of serious thrombotic events. For
`example, use of antithrombotic drugs is often avoided in the
`very elderly, despite the fact that advanced age is a major risk
`factor for both arterial and venous thromboembolism.12,13
`Concerns over bleeding have also restricted long-term use of
`warfarin and other anticoagulants in combination with
`antiplatelet drugs in patients with ACS and AF.4
`All of these factors have been the driving force for the
`development of new anticoagulants, including the direct
`FXa inhibitors. For any new anticoagulant to be considered
`ideal, it would need to possess a predictable PK profile that
`allows fixed oral dosing without routine monitoring and to
`possess a relatively wide therapeutic index with low peak-
`to-trough plasma concentrations to provide high levels of
`efficacy and low rates of bleeding. Since the target is in the
`central compartment, the ideal PK profile also features a low
`volume of distribution (to reduce the potential for off-target
`liabilities) and low systemic clearance (to minimize drug-
`drug interactions). Other key features of the “ideal” new
`anticoagulant include a rapid onset of action and an ability
`to bind clot-bound coagulation factors.14
`1.2. Hemostasis and Thrombosis. Hemostasis is the phy-
`siologic process during which bleeding is antagonized, and
`possibly stopped, in order to minimize blood loss.15 Unlike
`hemostasis, which is a necessary physiologic response to
`bleeding, thrombosis is a pathological process involving an
`exaggerated hemostatic response and is often ascribed to the
`combined influence of a triad of causative factors, namely,
`prothrombotic vascular endothelial changes or injury, stasis
`of blood flow, and/or other causes of hypercoagulability.16
`Primary hemostasis begins immediately after endothelial
`damage and involves localized vasoconstriction and the
`activation and adhesion of platelets to form a soft aggre-
`gate plug. In contrast, secondary hemostasis comprises a
`complex series of reactions during which several trypsin-like
`serine proteases,
`including FXa, are formed from their
`respective proenzymes. Figure 1A depicts how the “coagula-
`tion cascade” has classically been viewed, with two parallel
`and largely independent pathways, the extrinsic and in-
`trinsic pathways, converging at the point of factor X (FX)
`activation.4 Upon the formation of FXa, the common path-
`way is initiated, leading to the activation of prothrombin to
`thrombin and the subsequent conversion of fibrinogen to
`fibrin (described in more detail in section 1.3). Polymerized
`fibrin strands, together with activated platelets, then form
`stable clots that seal the breach at the site of injury. While this
`classic view of the coagulation cascade aids understanding of
`some of the key processes involved in secondary hemostasis,
`recent advances have led to the development of a cell-based
`model that may describe the time course and physiology of
`thrombogenesis more accurately,
`including the roles of
`tissue-factor (TF)-bearing cells and platelets. In this model,
`shown schematically in Figure 1B, coagulation occurs in
`
`Figure 1. (A) Classic view of the coagulation cascade. Sites of
`action of traditional and new anticoagulants are also depicted
`(coagulation factors affected by VKAs are indicated by an asterisk).
`Adapted by permission from Macmillan Publishers Ltd.: Clinical
`Pharmacology & Therapeutics (http://www.nature.com/clpt/index.
`html) (Gross, P. L.; Weitz, J. I. New antithrombotic drugs. Clin.
`Pharmacol.Ther. 2009, 86, 139-146),4 Copyright 2009. (B) Cell-
`based model of coagulation. Adapted by permission from Wolters
`Kluwer Health/Lippincott, Williams & Wilkins: Journal of Cardi-
`ovascular Pharmacology (Hammw€ohner, M.; Goette, A. Will war-
`farin soon be passe? New approaches to stroke prevention in atrial
`fibrillation. Journal of Cardiovascular Pharmacology 2008, 52,
`18-27),17 Copyright 2008, and Arteriosclerosis, Thrombosis and
`Vascular Biology (Monroe, D. M.; Hoffman, M.; Roberts, H. R.
`Platelets and thrombin generation. Arterioscler., Thromb., Vasc.
`Biol. 2002, 22, 1381-1389),18 Copyright 2002.
`
`three overlapping phases (initiation, amplification, and
`propagation) and is critically dependent on the contribution
`of cell surface proteins such as TF.17,18 In the initiation
`
`6244 Journal of Medicinal Chemistry, 2010, Vol. 53, No. 17
`
`Pinto et al.
`
`

`
`phase, TF-factor VIIa (FVIIa) complexes activate factor IX
`(FIX) and FX; subsequently formed FXa-factor Va (FVa)
`complexes then generate small amounts of thrombin. During
`the amplification phase, thrombin generated in the initiation
`phase activates platelets, leading to release of factor VIIIa
`(FVIIIa) from factor VIII (FVIII)-von Willebrand factor
`(vWF) complexes on the platelet surfaces and also genera-
`tion of FVa and factor XIa (FXIa). In the final propagation
`phase, factor XIa (FXIa) generates more FIXa; this FIXa,
`together with that generated in the amplification phase,
`activates FX, leading to the formation of numerous FXa-
`FVa complexes and a burst of thrombin generation.
`The inhibitory activity of anticoagulants can be assessed
`by a number of methods. Assays that employ purified co-
`agulation proteases and synthetic substrates are most com-
`monly used for the initial characterization of the affinity
`of synthetic inhibitors toward their target enzyme. Modifi-
`cations of these methods can also be used to measure the
`activity of inhibitors in blood samples as, for example, the
`anti-Xa assay that can be used with the LMWHs, fondapar-
`inux, or synthetic inhibitors. Anticoagulants can be further
`characterized using clotting assays (including prothrombin
`time (PT) and activated partial thromboplastin time (aPTT))
`that measure the time for formation of a fibrin clot after
`addition of an activator of coagulation to whole blood
`or plasma.19 Clotting assays can be used for either in
`vitro characterization of the potency of compounds or for
`ex vivo assessments in laboratory animals or in humans.
`In vitro potency is conventionally reported as the concentra-
`tion of inhibitor required to produce a doubling of the
`uninhibited clotting time (PT2 or aPTT2). Other useful
`laboratory methods include the thrombin generation (TG)
`assay, which measures time-dependent changes in thrombin
`concentration.20
`1.3. Targeting FXa. FXa plays a critical role in coagula-
`tion. Together with FVa and calcium ions on a phospholipid
`surface, FXa forms the prothrombinase complex, which is
`responsible for the conversion of prothrombin to thrombin,
`the final effector of coagulation. Regulation of thrombin
`generation is the primary physiologic function of FXa, and
`few other roles have been identified.21 FXa is therefore an
`attractive and potentially specific target for new anticoagu-
`lant agents. As will be discussed, experience with indirect
`inhibitors of FXa has helped to validate FXa inhibition as
`an effective and safe anticoagulant strategy. However, in-
`direct FXa inhibitors possess two significant limitations.
`First, these agents require parenteral administration; second,
`they rely on the activity of antithrombin and are therefore
`not able to inhibit FXa bound within the prothrombinase
`complex.22 Development of orally administered direct inhi-
`bitors of FXa that can effectively inhibit prothrombinase-
`associated and clot-bound FXa, and thereby offer poten-
`tially greater anticoagulant activity, is therefore a highly
`significant advance.
`Once formed via the actions of FXa, thrombin plays key
`roles in both coagulation and platelet activation. Within the
`coagulation cascade, thrombin directly cleaves fibrinopep-
`tides from fibrinogen, participates in positive feedback reac-
`tions via activation of FV and FVIII, promotes the cross-
`linking of fibrin through activation of factor XIII (FXIII),
`and renders fibrin resistant to fibrinolysis through activation
`of thrombin-activatable fibrinolysis inhibitor (TAFI).23 Oral
`anticoagulant drug discovery efforts initially focused on the
`development of small-molecule anticoagulants that target
`
`thrombin directly, the oral DTIs. Although the rationale for
`targeting thrombin is clear, there is some evidence to suggest
`that inhibition earlier in the coagulation cascade at the level
`of FXa may have greater antithrombotic potential.21
`Furthermore, preclinical studies suggest that FXa inhibitors
`may possess a wider therapeutic index than DTIs.24 There-
`fore, it is not surprising that the oral anticoagulant drug
`discovery efforts of many pharmaceutical companies ulti-
`mately focused aggressively on small-molecule, direct FXa
`inhibitors. Differences between the direct FXa inhibitors and
`DTIs, and the potential consequences of these differences for
`clinical practice, are discussed in more detail in section 5.
`Before we begin our review of the development of the
`direct FXa inhibitors, it is worth briefly considering some
`important molecular features of the target protein. FXa
`belongs to the family of trypsin-like serine proteases, the
`catalytic domain of which consists of two similar antiparallel
`β-barrel folds that together form the catalytic triad and
`substrate binding site.25 Schechter and Berger have devel-
`oped a useful nomenclature to describe the prototypical
`binding site of a serine protease that has been widely adopted
`and that we will use herein.26 Accordingly, each protein
`subsite, labeled Si, binds the corresponding substrate amino
`acid labeled Pi, with “i” increasing toward the substrate
`N-terminus. Similarly, the corresponding subsites and sub-
`strate amino acids to the left of the scissile amide bond
`0
`0
`in Figure 2 are designated as Si
`, respectively, in-
`and Pi
`creasing toward the substrate C-terminus. Substrate clea-
`0-P1 amide bond. As the discovery
`vage occurs at the P1
`of small-molecule protease inhibitors has advanced, this
`convention has been extended to denote drug substructures
`that bind in a manner similar to substrate amino acids.27
`Figure 2A depicts the serine protease subsites primarily
`responsible for the recognition and binding of substrate
`and druglike molecules. It is noteworthy that all reported
`small-molecule serine protease inhibitors for which structur-
`al data exist bind in the S1 and one or more of the remaining
`subsites.
`The FXa binding site is defined by the S1 and S4 subsites
`and surrounding residues (Figure 2B and Figure 2C). S1 is a
`deep, largely hydrophobic cleft at the bottom of which lies
`the Asp189 and Tyr228 side chains. S4 is a strongly hydro-
`phobic pocket defined principally by the side chains of
`Tyr99, Phe174, and Trp215. The most potent ligands re-
`ported in the literature invariably engage both sites. Other
`features include the catalytic triad consisting of His57,
`Asp102, and Ser195 and the β-strand region defined by the
`214-217 backbone.
`Selectivity is a significant issue in the development of
`factor Xa inhibitors, since, as discussed above, several
`trypsin-like serine proteases play key regulatory roles in the
`coagulation cascade, among them FVIIa, FIXa, FXa, FXIa,
`and thrombin. Trypsin itself is an important enzyme for
`digestion of proteins in the gastrointestinal tract. Although
`the consequences of trypsin inhibition in humans have not
`been well-studied, results in laboratory animals could be
`problematic in preclinical toxicity testing that is required of
`any new drug.30 Orally administered drugs can reach con-
`centrations many-fold higher within the GI tract compared
`to concentrations within blood; therefore, a high degree of
`selectivity for the target coagulation enzyme over trypsin
`is important. In addition, selectivity over trypsin may be
`considered as a surrogate determination for achieving selec-
`tivity over other members of the trypsin-like protease family.
`
`Perspective
`
`Journal of Medicinal Chemistry, 2010, Vol. 53, No. 17
`
`6245
`
`

`
`Figure 2. Serine protease and factor Xa structure and nomenclature. (A) Depiction of substrate/protease nomenclature based on the
`convention of Schechter and Berger.26 This figure is based on that of Leung et al.28 (B, C) Overview of the factor Xa binding site,29 with subsites
`and several residues important for ligand binding labeled.
`
`As will be discussed in section 3 in more detail, a differentiat-
`ing feature that can be exploited for selectivity among serine
`proteases is the nature of the S1 pocket, being smaller and
`lipophilic in some such as trypsin, or larger and more
`hydrophobic such as in FXa.
`
`2. Development of Direct FXa Inhibitors
`
`2.1. Precedence for FXa Inhibition as an Effective and Safe
`Anticoagulant Therapy. Proof of principle for the effectiveness
`of direct FXa inhibition was established in preclinical animal
`models of thrombosis with naturally occurring FXa inhibitors
`of the prothrombinase complex such as tick anticoagulant
`peptide (TAP)31 and antistasin.32,33 Both are highly potent
`inhibitors of FXa (Ki = 0.59 and 0.3-0.6 nM, respectively),
`with >50 000-fold selectivity for FXa over other related serine
`proteases.34,35 Compound 1 (Figure 3), a synthetic pentasac-
`charide, is selective for FXa but acts indirectly via binding
`to antithrombin and has demonstrated improved or similar
`clinical benefit over LMWHs in venous thrombotic indica-
`tions.36 Superiority in ACS patients with unstable angina/non-
`ST-segment elevation MI for reducing risk of death or recur-
`rent heart attack was also demonstrated.37,38 The safety and
`efficacy of 1 provided the first clinical proof of principle that
`targeting FXa would be an important advancement in the area
`of anticoagulation therapy.36
`More recently, Sanofi-Aventis advanced a hypermethy-
`lated derivative of 1 with a high affinity for antithrombin III
`(ATIII), in which the amino functional groups were replaced
`with hydroxyl or methoxy groups.39 This compound, idra-
`parinux (2a, Kd = 1 nM, Figure 3), interacts more strongly
`with ATIII than 1 (Kd = 50 nM) and in patients has a longer
`half-life (t1/2 ≈ 80 h), allowing for once-weekly dosing. The
`results of phase II/III trials with 2a were mixed, however, and
`
`Figure 3. Structures of the pentasaccharide indirect FXa inhibitors.
`
`did not demonstrate a clear advantage over 1.40 In a phase III
`trial, long-term treatment with 2a (once weekly) for the
`prevention of stroke and systemic embolism in patients with
`AF was noninferior to warfarin but caused more bleeding.41
`Development of 2a has since been discontinued. A second-
`generation synthetic pentasaccharide, idrabiotaparinux (2b,
`SSR126517E), is in late-stage clinical trials for treatment of
`VTE and for stroke prevention in patients with AF,4,42
`although the company recently announced discontinuation
`of development of the drug for the latter indication.43 Com-
`pound 2b incorporates a biotin moiety that enables the
`selective reversal of anticoagulant activity by intravenous
`(iv) administration of avidin, which binds the biotin group.
`
`6246 Journal of Medicinal Chemistry, 2010, Vol. 53, No. 17
`
`Pinto et al.
`
`

`
`was rapidly distributed in the plasma and rapidly eliminated
`with a half-life of 1.5-2 h.55 In a phase II trial setting, 4
`significantly reduced prothrombin fragment 1 þ 2, a marker
`of thrombin generation, when compared with UFH and a
`glycoprotein (GPIIb/IIIa) antagonist, eptifibatide.56 Two
`phase II clinical trials for the management of ACS and
`patients with ACS undergoing percutaneous coronary inter-
`vention have also been completed and showed the potential
`for reduced ischemic events with bleeding rates similar to
`those of UFH plus eptifibatide.56,57 A third parenteral agent
`that was advanced to human clinical trials was fidexaban (5,
`ZK-807834, Berlex-Pfizer; Figure 4),58 a compound that
`contains two amidine groups and a polar carboxylic acid
`moiety.59 The dihydrochloride salt of 5 (ZK-807191) is a
`potent inhibitor of FXa (Ki = 0.10 nM) and exhibits nearly
`20 000-fold selectivity over thrombin and 2500-fold selectiv-
`ity over trypsin; it has been shown to be efficacious in several
`in vivo animal models of thrombosis.60-62 In human clinical
`trials, infusion of 5 was found to be well tolerated.63 Phase II
`trials were underway in unstable angina in 2001; however,
`no results from these studies were published.64 In addition to
`the above direct FXa parenteral compounds, a dual throm-
`bin/Xa inhibitor, tanogitran (6, BIBT 986, Boehringer In-
`gelheim, FXa Ki = 26 nM, thrombin Ki = 2.7 nM, Figure 4),
`was evaluated more recently in a phase II clinical trial
`involving a human model of endotoxin-induced coagula-
`tion.65 In this study, 6 prolonged plasma aPTT, reduced in
`vivo thrombin generation in a dose-dependent manner, and
`was safe and well tolerated. No further development has
`been reported.
`Clinical success of indirect FXa inhibitors such as 1 and
`the improved therapeutic index with the early direct FXa
`inhibitors such as the parenteral inhibitors described above
`fueled an intense effort to discover and develop safer and
`more effective oral FXa inhibitors. The discovery and ad-
`vancement of orally bioavailable, direct-acting FXa inhibi-
`tors that progressed to clinical studies, including a brief
`survey of some of the early inhibitors that led up to these
`agents, are now discussed.
`2.3. Approach to Oral FXa Inhibitors. 2.3.1. Transition
`State and Peptidomimetic Approach: Covalent Inhibitors.
`Early efforts to identify inhibitors of FXa stemmed from
`the prior discoveries of thrombin inhibitors that contained
`“serine traps”, such as aldehyde or ketothiazole moieties,
`which are capable of interacting covalently with the cata-
`lytic Ser195 hydroxyl group to mimic a tetrahedral transi-
`tion state in a reversible manner. Compounds 7 (FXa IC50 =
`15 nM),66 8 (FXa Ki = 0.13 nM),67 and 9 (FXa IC50 =
`0.83 nM)68 are examples of early FXa inhibitors that contain
`a “serine trap” (Figure 5). As with the thrombin inhibitors in
`this class, several of the first transition-state FXa inhibitors
`also contained arginine or constrained arginine P1 residues
`that would interact strongly with the acidic Asp189 S1 resi-
`due, flanked by aromatic residues designed to fit into the
`hydrophobic S1 pocket of FXa.44 In the design of these
`inhibitors, it also became apparent that basic substituents
`could be tolerated in both the S1 and S4 regions, with a
`basic P4 moiety interacting in a π-cation manner with the
`hydrophobic residues in the S4 subsite. These peptide-like
`inhibitors eventually evolved to incorporate heterocyclic
`amide-bond replacements,69 such as pyridone (10, FXa
`IC50 = 3 nM), ketopiperazine (11, FXa IC50 = 2 nM), and
`caprolactam (12, FXa IC50 = 3 nM), while maintaining
`good FXa binding affinity; however, oral bioavailability was
`
`Figure 4. Parenteral direct FXa inhibitors 3, 4, and 5 and the dual
`FXa/thrombin inhibitor 6.
`
`2.2. Early Prototype Direct FXa Inhibitors: Parenteral
`Agents. The shortcomings of the LMWHs (e.g., indirect
`activity, limited ability to inhibit fibrin-bound FXa, paren-
`teral use only) provided great impetus for the discovery of
`synthetic, small-molecule direct FXa inhibitors. The first
`generation of these were iv agents, and several small-mole-
`cule, nonpeptidic, direct inhibitors of FXa were advanced
`to phase II clinical trials as parenteral agents.44,45 DX-9065a
`(3, Daiichi Sankyo, Figure 4, FXa Ki = 41 nM, thrombin
`Ki > 2000 μM, trypsin Ki = 620 nM) was clearly one of the
`first potent and selective FXa inhibitors identified, with
`good clotting activity (activated partial thromboplastin time
`twice the control [aPTT2] = 0.97 μM, prothrombin time
`twice the control [PT2] = 0.52 μM).46,47 The compound was
`studied extensively in preclinical models; found to be effica-
`cious in animal models of thrombosis after iv, subcutaneous,
`and oral administration; and did not show prolonga-
`tion of bleeding time.48 Because of its very low human oral
`bioavailability (F = 2-3%), 3 was advanced clinically
`as a parenteral agent.49 The compound was well tolerated,
`with no increase in bleeding over the dose range. Human PK
`for 3 showed a long half-life (t1/2 > 20 h), low clearance
`(120 mL/min), and low plasma protein binding (60% bound).
`Plasma concentrations correlated well with PD markers.
`In a phase II study in patients with non-ST-elevation ACS,
`dose-related trends toward reductions in ischemic events
`with high-dose 3 compared with heparin were observed.50
`Safety parameters such as bleeding increased dose propor-
`tionally.
`Otamixaban (4, FXV-673, Sanofi-Aventis, Figure 4), a
`2,3-disubstituted β-aminoester derivative, is a potent, rever-
`sible FXa inhibitor (Ki = 0.5 nM) with good in vitro antico-
`agulant activity (aPTT2 = 0.41 μM, PT2 = 1.1 μM).51,52
`Like 3, compound 4 belongs to the benzamidine class
`of molecules, and its polarity precludes significant oral
`absorption. In vivo, 4 was efficacious in canine models of
`thrombosis and demonstrated minimal effect on bleeding at
`effective doses.53,54 In phase I/II studies, 4 was administered
`intravenously and was found to be well tolerated in healthy
`volunteers and patients with coronary artery disease and
`
`Perspective
`
`Journal of Medicinal Chemistry, 2010, Vol. 53, No. 17
`
`6247
`
`

`
`benzoxazinone ring resulted in analogues, such as 21, that
`retained potent binding affinity against FXa (Ki = 6 nM).
`Kissei Laboratories reported the discovery of 22 (KFA-
`1411, FXa Ki = 1.7 nM), which is highly selective over a
`broad range of serine proteases,
`including trypsin, and
`has good in vitro anticoagulant potency (PT2 = 0.28 μM,
`aPTT2 = 0.87 μM).78 The compound was efficacious when
`compared with dalteparin and LMWH in a hemodialysis
`model in monkeys.79 Additional early bisamidine-containing
`analogues have been covered extensively in previous review
`articles.80-82
`2.4. Transition From Benzamidine to Oral Agents. 2.4.1.
`Isoxazoline, Isoxazole, and Pyrazole-Based Inhibitors. In the
`quest for potent inhibitors of FXa, Bristol-Myers Squibb
`researchers recognized the similarity between the platelet
`GPIIb/IIIa peptide sequence Arg-Gly-Asp and the pro-
`thrombin substrate FXa sequence Glu-Gly-Arg.83 Isoxazo-
`line derivative 23 (FXa Ki ≈ 39 000 nM, Figure 7) was
`
`identified from a high-throughput screen (HTS) of a library
`of proprietary GPIIb/IIIa antagonists. Modifications to
`incorporate the bisbenzamidine motif of the known FXa
`inhibitors, along with further optimization, provided 24
`with greatly enhanced affinity (FXa Ki = 94 nM).83 Guided
`by structure-based design, these researchers replaced the
`4-amidino moiety of 24 with a neutral o-phenylsulfonamide
`group to reduce the basicity of this dibasic lead and thus
`improve permeability. This afforded the first known mono-
`basic FXa inhibitor 25, which also exhibited enhanced
`potency (FXa Ki = 6.3 nM).84 The biaryl moiety of 25 was
`designed to interact with the hydrophobic S4 aryl binding
`domain of the FXa active site (see Figure 2B) and, from
`modeling experiments, was shown to be neatly stacked
`between the residues Tyr99, Phe174, and Trp215, with the
`terminal o-phenylsulfonamide ring designed to make an edge-
`to-face interaction with Trp215. Further scaffold optimiza-
`tion led to isoxazoline compound 26 (FXa Ki = 0.55 nM)
`with the benzamidine P1 and the biarylsulfonamide vicinally
`substituted on the five-member ring.85 Replacement of
`the isoxazoline ring with a planar aromatic isoxazole ring
`in 27 provided further improvement in FXa potency (Ki =
`0.15 nM).85 The lack of chirality of this template, and the
`potency, made it an attractive starting point for further
`optimization. Rapid evaluation of a variety of vicinally
`substituted five- and six-member ring scaffolds resulted in
`the determination that the five-member nitrogen-based,
`N-linked templates were most potent.86 This effort led to
`the discovery of a very potent pyrazole analogue 28 (SN429,
`FXa Ki = 0.013 nM).87 An X-ray crystal structure of 28 in
`bovine trypsin showed the following interactions: S1-Asp189
`with the benzamidine P1, the pyrazole N2 group with the
`backbone of Gln192, a lipophilic interaction of the C3 methyl
`at the outer ridge of the enzyme, the linker carboxamide
`carbonyl oxygen with the NH of Gly216, and the P4 bi-
`arylsulfonamide neatly stacked in the S4 hydrophobic box
`formed by Tyr99, Trp215, and Phe174. In vivo, in the rabbit
`arteriovenous (AV) shunt model, administration of 28 by iv
`infusion resulted in reduction of thrombus weight by 50% at
`a dose of 0.02 μmol/kg/h (ID50).87 Poor oral bioavailability
`(F < 4% in dogs), combined with a short half-life (t1/2 =
`0.82 h) and lack of selectivity over related trypsin-like serine
`proteases, precluded further development of 28 as an oral
`anticoagulant.
`The high affinity of 28 for FXa afforded a unique oppor-
`tunity to give up potency in an effort to improve the oral
`
`Figure 5. Examples of transition-state inhibitors of FXa.
`
`not achieved. There are no published reports of any transi-
`tion-state FXa inhibitors advancing into clinical trials.
`2.3.2. Early Dibasic Benzamidine Approach. Reports of
`nonpeptidic, small-molecule inhibitors of FXa such as bis-
`amidine compounds 13 (DABE, bovine FXa Ki = 570 nM)70
`and 14 (BABCH, bovine FXa Ki = 610 nM)71 set the stage
`for the evolution of additional dibasic inhibitors (Figure 6),
`of which 3 is an early example. The success of 3 prompted
`multiple research groups to further optimize compounds in
`this class. Initial attempts at Daiichi Sankyo to improve FXa
`potency and oral bioavailability in direct analogues of 3
`resulted in constrained indoline compounds 15 (FXa IC50 =
`7.6 nM)72 and 16 (FXa IC50 = 3.9 nM).73 Although these
`analogues demonstrated potent binding affinity against
`FXa, selectivity over trypsin remained an issue (trypsin Ki =
`24 and 39 nM, respectively). This was addressed by mod-
`ification of the naphthyl P1 moiety to a substituted benz-
`amidine and introduction of a hydroxyl mo

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket