ArticlePDF AvailableLiterature Review

The discovery and development of rivaroxaban, an oral, direct factor Xa inhibitor

Authors:

Abstract and Figures

The activated serine protease factor Xa is a promising target for new anticoagulants. After studies on naturally occurring factor Xa inhibitors indicated that such agents could be effective and safe, research focused on small-molecule direct inhibitors of factor Xa that might address the major clinical need for improved oral anticoagulants. In 2008, rivaroxaban (Xarelto; Bayer HealthCare) became the first such compound to be approved for clinical use. This article presents the history of rivaroxaban's development, from the structure-activity relationship studies that led to its discovery to the preclinical and clinical studies, and also provides a brief overview of other oral anticoagulants in advanced clinical development.
Simplified schematic for the blood coagulation cascade.The figure identifies the target points of various anticoagulants and illustrates that factor X (FX) can be activated through either the intrinsic or the extrinsic pathway. The factor VIIa–tissue factor (TF) complex (extrinsic tenase) activates both factor IX and factor X, as well as factor VII itself (dashed arrow)150, 151. Initiation of either pathway activates the inactive precursor, factor X, to factor Xa. This makes factor Xa a desirable intervention point for novel anticoagulants, because it occupies a central role in the blood-coagulation pathway20. Furthermore, in addition to initiation, both the intrinsic and extrinsic pathways lead to the propagation and amplification of coagulation through the activation of factor X. During the initiation phase of coagulation, the factor Xa produced generates some thrombin (factor IIa). This initial thrombin activates factor XI, and factors V and VIII, to factor XIa and the activated cofactors, factor Va and VIIIa, respectively. It also activates platelets (not shown), which are required for the formation of the intrinsic tenase (factor VIIIa–factor IXa) and the prothrombinase (factor Va–factor Xa) complexes. The prothrombinase complex, on the platelet surface, is substantially more efficient than free factor Xa at activating prothrombin to thrombin; the rate of thrombin formation is increased by approximately 300,000-fold over the rate with factor Xa alone152. Thrombin is the principal enzyme involved in the formation, growth and stabilization of thrombi. Thrombin mediates the conversion of fibrinogen to fibrin, the activation of factor XIII (which crosslinks and stabilizes fibrin), the activation of platelets and the above-mentioned feedback-activation of upstream coagulation factors, factor V, factor VIII and factor XI151, 153, resulting in the amplification of its own formation150, 154. Because one molecule of factor Xa catalyses the formation of approximately 1,000 thrombin molecules25, 154, this amplification can be substantial. Compared with thrombin, factor Xa is thought to have fewer effects outside coagulation and, together with factor Va (as the prothrombinase complex), acts mainly to convert prothrombin to thrombin. Specific inhibition of factor Xa does not affect pre-existing thrombin but does inhibit thrombin generation155. Conversely, although thrombin generation is delayed in the presence of a thrombin inhibitor, the amount of thrombin generated is only reduced at higher inhibitor concentrations, albeit in a concentration-dependent manner156. LMWH, low-molecular-weight heparin.
… 
Content may be subject to copyright.
Anticoagulants are used for the prevention and treatment
of venous and arterial thromboembolic disorders. Many
approaches have been explored in the development of
antithrombotic drugs that inhibit enzymes in the coagu-
lation pathways. However, most currently approved drugs
for the prevention and treatment of thromboembolic dis-
orders have been on the market for a long time, in some
cases for decades. Unfractionated heparin (UFH), which
was discovered in 1916 (REF. 1) targets multiple factors in
the coagulation cascade2, but has a number of limitations,
including a parenteral route of administration, frequent
laboratory monitoring of coagulation activity and the
risk for patients of developing potentially life-threatening
heparin-induced thrombocytopaenia2,3. Low-molecular-weight
heparins (LMWHs), which were developed in the 1980s,
promote the inactivation of both thrombin (factor IIa)
and, to a greater extent, factor Xa2.
LMWHs have largely replaced UFH owing to their
lower risk of causing bleeding, lower levels of binding to
plasma proteins and endothelium, good bioavailability,
longer half-life and superior pharmacokinetic properties
compared with UFH2. However, their use remains lim-
ited because of the need for parenteral administration2,
which can be inconvenient, especially in an outpatient
setting, with patients needing to be trained to self-inject
after discharge, and nurse visits required for those unable
to do so4. Both UFH and LMWHs are indirect inhibitors
of coagulation, and their activity is mediated by plasma
cofactors, principally antithrombin and, to a lesser extent
for UFH, heparin cofactor II (REF. 2).
Warfarin, the prototype vitamin K antagonist (VKA),
was originally discovered in 1941 (REF. 5). Until recently,
the VKAs were the only available oral anticoagulants, as
well as the most frequently prescribed. However, VKAs
have numerous well-documented drawbacks, includ-
ing unpredictable pharmacokinetics and pharmaco-
dynamics, a slow onset and offset of action, a narrow
therapeutic window, multiple food–drug and drug–drug
interactions6, and considerable inter-individual and
intra-individual variability in dose response. In addition,
regular coagulation monitoring and dose adjustment are
required to keep patients within the target international
normalized ratio (INR) range, usually 2.0–3.0, which can
be costly6. Furthermore, establishing the optimal dose
of warfarin is complicated by variations in warfarin
sensitivity due to common genetic polymorphisms,
particularly in CYP2C9 and VKORC1 (REF. 7). Attempts
have been made to use pharmacogenetics to estimate
dose, although the clinical value of such assessments
is debatable8. Nonetheless, LMWHs and VKAs are still
the basis of contemporary thromboprophylaxis and treat-
ment. However, the difficulties and shortcomings that
surround the practicalities and clinical management
of these established anticoagulants — particularly
parenteral administration, the need for monitoring
and the lack of predictable response — has spurred the
development of new agents that are less burdensome for
the patient and health-care system, and address both
patients’ and physicians’ unmet needs. TABLE 1 con-
trasts the characteristics of the VKAs with those of a
*Global Therapeutic
Research, Global Lead
Generation and Optimization,
§Global Clinical
Pharmacology,
||Global Clinical Development,
Cardiovascular
Pharmacology, Pharma R&D,
Bayer HealthCare, Aprather
Weg 18a, D‑42096
Wuppertal, Germany.
Correspondence to E.P.
e‑mail: elisabeth.perzborn@
bayer.com
doi:10.1038/nrd3185
Published online
10 December 2010
Thromboembolic disorders
A group of conditions
characterized by an increased
incidence of thrombi in the
vasculature, such as deep-vein
thrombosis, pulmonary
embolism, systemic embolism
or coronary and cerebral
ischaemia.
Unfractionated heparin
(UFH). An anticoagulant
administered intravenously or
subcutaneously. It binds to
antithrombin, greatly
increasing its activity and
resulting in the inhibition of
factors Xa, IXa, XIa, XIIa and
thrombin (factor IIa).
The discovery and development
of rivaroxaban, an oral, direct
factor Xa inhibitor
Elisabeth Perzborn*, Susanne Roehrig, Alexander Straub, Dagmar Kubitza§ and
Frank Misselwitz||
Abstract | The activated serine protease factor Xa is a promising target for new anticoagulants.
After studies on naturally occurring factor Xa inhibitors indicated that such agents could be
effective and safe, research focused on small-molecule direct inhibitors of factor Xa that
might address the major clinical need for improved oral anticoagulants. In 2008, rivaroxaban
(Xarelto; Bayer HealthCare) became the first such compound to be approved for clinical use.
This article presents the history of rivaroxaban’s development, from the structure–activity
relationship studies that led to its discovery to the preclinical and clinical studies, and also
provides a brief overview of other oral anticoagulants in advanced clinical development.
CASE HISTORY
REVIEWS
NATURE R EVIEWS
|
DRUG DISCOVERY VOLUME 10
|
JANUARY 2011
|
61
© 2011 Macmillan Publishers Limited. All rights reserved
Heparin-induced
thrombocytopaenia
The process by which
antibodies against the complex
of heparin and platelet factor 4
activate platelets, resulting in a
decrease in platelet numbers
of more than 50%.
Low-molecular-weight
heparins
(LMWHs). A class of
anticoagulants derived from
unfractionated heparin by
chemical or enzymatic
degradation. They induce a
conformational change in
antithrombin that greatly
increases its anticoagulant
activity.
Thrombin
Thrombin (also known as factor
IIa) is the terminal enzyme of
the coagulation cascade and
converts fibrinogen into fibrin,
which forms clot fibres.
Thrombin also activates several
other coagulation factors, as
well as protein C.
hypothetical ‘ideal’ anticoagulant, and FIG. 1 chronicles
the historical development of anticoagulants.
Despite the accumulated knowledge on the coag-
ulation system (FIG. 2), its complexity has presented
numerous obstacles to the discovery and development
of potent anticoagulants that are both effective and safe.
In recent years, research has focused on new classes of
anticoagulants that target a specific coagulation enzyme
or step in the coagulation cascade9,10, including inhibi-
tors of the factor VIIa–tissue factor complex10, inhibi-
tors of factor IXa11,12 and factor XIa13,14, direct thrombin
inhibitors15,16, synthetic indirect and direct inhibitors
of factor Xa (activated factor X)17–20, and recombinant
soluble thrombomodulin21,22. In addition, recombinant
activated protein C (APC) mitigates the procoagulant
state associated with sepsis23,24.
In the search for new anticoagulant drugs, the acti-
vated serine protease factor Xa is a particularly promising
target and has attracted great interest in recent years18,25–27.
This article describes the discovery and development of
the first oral, direct factor Xa inhibitor to be approved for
clinical use — rivaroxaban (Xarelto; Bayer HealthCare).
Rivaroxaban was approved by the European authorities
in 2008 for the prevention of venous thromboembolism
(VTE; comprising deep-vein thrombosis (DVT) and pul-
monary embolism) after elective hip or knee replacement.
The article then briefly considers the future clinical
potential of rivaroxaban and other factor Xa inhibitors
currently in advanced clinical development.
Factor Xa: function and biology
Factor X has long been known to have a key role in hae-
mostasis28 and factor Xa plays a central part in the blood
coagulation pathway by catalysing the production of
thrombin, which leads to clot formation and wound
closure20 (FIG. 2). Conversely, deficiency of factor Xa may
disturb haemostasis. In the very rare factor X deficiency
disorder (for which 1 in 500,000 is homozygous and
1 in 500 heterozygous), very low plasma and activity levels
of factor Xa manifest as severe bleeding tendencies29–31.
Studies of variants of factor X deficiency indicate that fac-
tor X plasma activity levels must be as low as 6–10% of
the normal range (approximately 50–150% of the popu-
lation average) to be considered a mild deficiency; cases
with factor X activity levels below 1% are considered to
be severe29,32. Thus it seems that factor X activity can be
markedly suppressed without affecting haemostasis. An
ideal anticoagulant would prevent thrombosis without
inducing systemic hypocoagulation, and would thereby
avoid unintended bleeding complications. Therefore, a
factor Xa inhibitor could potentially have the properties
of a desirable anticoagulant.
Validating factor Xa as a drug target
The first factor Xa inhibitors. Although factor Xa was
identified as a promising target for the development of
new anticoagulants in the early 1980s, the viability of
factor Xa inhibition was not tested before the end of that
decade. In 1987, the first factor Xa inhibitor, the natu-
rally occurring compound antistasin, was isolated from
the salivary glands of the Mexican leech Haementeria
officinalis33,34. Antistasin is a 119 amino-acid polypep-
tide; kinetic studies revealed that it is a slow, tight-bind-
ing, potent factor Xa inhibitor (inhibition constant (Ki)
of 0.3–0.6 nM) that also inhibits trypsin (half maximal
inhibitory concentration (IC50) is 5 nM in the presence
of 1 nM trypsin)35. Another naturally occurring factor
Xa inhibitor, the tick anticoagulant peptide (TAP), a sin-
gle-chain, 60 amino-acid peptide, was isolated in 1990
from extracts of the soft tick Ornithodoros moubata36.
Similarly to antistasin, TAP is a slow, tight-binding
inhibitor of factor Xa (Ki of ~0.6 nM).
TAP 37 and recombinant forms of antistasin38 and
TAP 38–40 were used to validate factor Xa as a viable drug
target and to improve understanding of the role of fac-
tor Xa in thrombosis. The antithrombotic effects of
these compounds were compared with those of direct
thrombin inhibitors and of indirect thrombin and fac-
tor Xa inhibitors (that is, UFH) in animal models of
thrombosis. These studies suggested that direct factor
Xa inhibitors might be a more effective approach to
anticoagulation37,39, and might also offer a wider thera-
peutic window, particularly with regard to primary
haemostasis40,41.
In vitro and in vivo studies. Clot-bound factor Xa was
shown to be enzymatically active in v itro and able to acti-
vate prothrombin to thrombin42. In addition, factor Xa was
found to be an important contributor to clot-associated
procoagulant activity in vitro43. Clot-bound factor Xa
activity was resistant to inhibition by antithrombin42, sug-
gesting that the ability to directly inhibit clot-associated
factor Xa, with no requirement for a cofactor, could
provide an effective and highly localized approach to
the prevention of thrombus growth. The clot-associated
Table 1 | Comparison of an ideal anticoagulant and a vitamin K antagonist
Property Ideal anticoagulant Vitamin K antagonist
Administration Oral Oral
Onset/offset of
action
Rapid (several hours) Slow (several days)
Therapeutic window Wide Narrow
Variability in dose
response
Little or no
inter-individual or
intra-individual
variability
Considerable inter-individual and
intra-individual variability
Interactions Little or no interaction
with food or other drugs
Multiple interactions with food
and other drugs
PK/PD Predictable Unpredictable and variable
Coagulation
monitoring
No routine monitoring
required
Regular monitoring required
Dose adjustment None required Required
Efficacy Highly effective
in reducing
thromboembolic events
Effective in reducing
thromboembolic events when
properly controlled
Safety profile Good, especially with
regard to bleeding
Difficulties in maintaining patients
within the target therapeutic
range (INR 2.0–3.0); contributes to
an increased risk of bleeding
INR, international normalized ratio; PK/PD, pharmacokinetics/pharmacodynamics.
REVIEWS
62
|
JANUARY 2011
|
VOLUME 10 www.nature.com/reviews/drugdisc
© 2011 Macmillan Publishers Limited. All rights reserved
Parenteral
Oral
Parenteral
Parenteral
Parenteral
Oral
Oral
Unfractionated heparins: antithrombin-dependent
inhibition of factor Xa and IIa in a 1:1 ratio
Vitamin K antagonists: indirectly affect
synthesis of multiple coagulation factors
LMWH: antithrombin-dependent inhibition
of factor Xa > inhibition of factor IIa (2:1 to 4:1)
Direct factor IIa inhibitors
Indirect factor Xa inhibitors
Direct factor
IIa inhibitors
Direct factor
Xa inhibitors
1930s
1940s
1980s
1990s
2000s
2008
Nature Reviews | Drug Discovery
Antithrombin
An endogenous glycoprotein
that binds covalently to
thrombin and other
coagulation factors, resulting in
their inhibition. Antithrombin
functions as a natural
anticoagulant, and its
inhibitory action is accelerated
by heparin.
Warfarin
A vitamin K antagonist that is
currently the most commonly
used oral anticoagulant.
Vitamin K antagonist
A class of compounds that
inhibit the vitamin K-dependent
carboxylation of specific
coagulation factors, resulting in
decreased levels of the
affected coagulation factors,
leading to anticoagulation.
Therapeutic window
The interval between the
lowest dose of a drug that is
sufficient for clinical
effectiveness and a higher dose
at which adverse events or
toxicity become unacceptable.
generation of fibrinopeptide A (a byproduct of fibrino-
gen cleavage by thrombin) was inhibited to the same
extent by both recombinant TAP and hirudin, a direct
thrombin inhibitor. This observation suggested that
procoagulant activity in the clot was due to de novo acti-
vation of prothrombin to thrombin, rather than to the
activity of pre-existing thrombin43. It was subsequently
shown that inhibition of factor Xa by recombinant TAP
provided sustained in vitro44 and in vivo45 inhibition
of clot-associated procoagulant activity, which may, in
turn, protect against ongoing coagulation after cessation
of anticoagulant treatment.
Overall, these data suggested that direct factor Xa
inhibitors might not be linked to the phenomenon of
‘rebound’ thrombosis that was associated with the direct
and indirect thrombin inhibitors previously under inves-
tigation46,47. Rebound thrombosis can be thought of as
a transient increase in thromboembolic events occur-
ring shortly after the withdrawal of an antithrombotic
medication46,48. Furthermore, low concentrations of a
direct thrombin inhibitor may partly suppress the nega-
tive feedback on coagulation by APC, in contrast to fac-
tor Xa inhibition, which does not seem to measurably
affect the thrombin–thrombomodulin–APC system49.
However, whether these experimental observations have
any clinical relevance remains to be determined.
The first synthetic factor Xa inhibitors
Although antistasin and TAP provided support for the
concept of factor Xa inhibition, development of these
compounds was discontinued. The reasons were never
disclosed. Nonetheless, the encouraging results from
studies using recombinant versions of the natural fac-
tor Xa inhibitors prompted several pharmaceutical
companies to initiate chemistry programmes to develop
selective, small-molecule, direct inhibitors of factor Xa,
such as DX-9065a50 and YM-60828 (REFS 51,52) (FIG. 3).
Both these compounds contain a highly basic amidine
residue, designed as mimics for the arginine of the
natural substrate prothrombin.
DX-9065a, a widely studied non-peptidic small
molecule, shows rapid, direct and reversible binding
kinetics for factor Xa (Ki of 41 nM)53. At physiologi-
cal pH it is a zwitterion with high water solubility and
low lipophilicity54. However, human oral bioavailability
was only 2–3%54. A small Phase II study was conducted
in patients with non-ST-elevation acute coronary
syndrome (ACS) who were randomized to low- or
high-dose intravenous DX-9065a or UFH55. A non-
significant trend was observed towards a reduction in
ischaemic events and bleeding for DX-9065a compared
with UFH55.
Because of the success of indirect dual factor Xa and
thrombin inhibitors, such as LMWHs, indirect inhibi-
tors of factor Xa with greater selectivity, such as fonda-
parinux (Arixtra; GlaxoSmithKline)56,57, were developed
in parallel with direct factor Xa inhibitors.
Both UFH and LMWHs contain a unique pen-
tasaccharide sequence that mediates binding to anti-
thrombin. Binding induces a conformational change
in antithrombin that potentiates its ability to inhibit
coagulation factors. Inhibition of thrombin occurs
through heparin chains of sufficient length to bridge
antithrombin to thrombin in a ternary complex, after
which antithrombin binds covalently to thrombin and
the heparin chain dissociates2. However, such bridging
is not necessary for antithrombin to be able to inhibit
factor Xa. Most LMWH chains are too short to cata-
lyse thrombin inhibition, but can nonetheless promote
the inhibition of factor Xa. Thus, UFH has an anti-Xa
to anti-IIa ratio of 1:1, LMWHs have anti-Xa to anti-
IIa ratios from 2:1 to 4:1, depending on the molecular
weight distribution of the preparation2.
Fondaparinux is an analogue of the pentasaccha-
ride sequence required to promote the binding of anti-
thrombin to factor Xa2. The pentasaccharide structure is
too short to enable bridging between antithrombin and
thrombin. As a result, fondaparinux exclusively potenti-
ates the anti-factor Xa activity of antithrombin and has
no effect on thrombin2. The efficacy and safety of fon-
daparinux for the prevention of VTE after major ortho-
paedic surgery were investigated in four randomized,
Phase III trials in patients undergoing surgery for hip
fracture58, hip replacement59,60 and knee replacement61.
A meta-analysis of these trials showed the superior
efficacy of fondaparinux over the LMWH enoxaparin
(Lovenox/Clexane; Sanofi–Aventis) in reducing the
incidence of VTE. However, major bleeding occurred
more frequently in the fondaparinux-treated group
(P=0.008), although the incidence of clinically relevant
bleeding (bleeding leading to death, reoperation or in
a critical organ) did not differ between the treatment
groups62. Fondaparinux provided the proof of princi-
ple that selective inhibition of factor Xa could provide
clinically effective anticoagulation.
Figure 1 | Development of anticoagulants over the past century. The timeframe
starts with the discovery of the heparins and the vitamin K antagonists, followed by
the low-molecular-weight heparins. These were followed, in turn, by the discovery
of the parenteral direct thrombin inhibitors, the development of the indirect factor Xa
inhibitor, fondaparinux, and the work of the current decade that has resulted in oral
direct inhibitors of both thrombin (factor IIa) and factor Xa. The inverted triangle
reflects the narrowing of action and increasing specificity of the anticoagulants.
LMWH, low-molecular-weight heparin.
REVIEWS
NATURE R EVIEWS
|
DRUG DISCOVERY VOLUME 10
|
JANUARY 2011
|
63
© 2011 Macmillan Publishers Limited. All rights reserved
FXIIFXIIa
FXI FXIa
FIXFVII
FIXa
FVIII FVIIIa
FVIIa
TF
FX FXa
Intrinsic
pathway
Extrinsic
pathway
FV
FIIFIIa
Fibrinogen Fibrin
FVa
Unfractionated heparin
(+ antithrombin)
Vitamin K antagonists
Direct factor Xa inhibitors
Fondaparinux (+ antithrombin)
LMWHs (+ antithrombin)
Direct thrombin inhibitors
Nature Reviews | Drug Discovery
Numerous investigations, performed with both direct
and indirect factor Xa inhibitors, showed that inhibi-
tion of factor Xa produces antithrombotic effects by
decreasing the generation of thrombin, thus diminish-
ing thrombin-mediated activation of both coagulation
and platelets without affecting the activity of existing
thrombin. However, the residual thrombin generated
seems to be sufficient to ensure normal systemic haem-
ostasis, possibly because thrombin has a very high
affinity for platelet receptors and minimal amounts of
thrombin can provide sufficient platelet activation, thus
contributing to a favourable efficacy/safety ratio20. On
the basis of these findings, in the mid-1990s, it seemed
that small-molecule, direct factor Xa inhibitors could
potentially provide an advantage over the antithrom-
botic therapies available at that time. Several new factor
Xa inhibitors are now in clinical development. These
compounds, which include rivaroxaban and the other
direct factor Xa inhibitors discussed later, represent
new chemical entities with similar binding modes at the
active site of factor Xa.
The discovery of rivaroxaban
From the first screening hit to the oxazolidinone lead.
When we initiated the factor Xa programme at Bayer
HealthCare in 1998, no orally active factor Xa inhibi-
tors with sufficient antithrombotic activity were known.
All known potent inhibitors that had previously been
investigated contained an amidine group or other highly
basic residues, which were designed to act as mimics
for an arginine present in the natural substrate, pro-
thrombin. For a long time, these mimics were thought
to be a prerequisite for high binding affinity63. However,
we found that strongly basic mimics contribute to poor
oral absorption, an observation also later published by
others64,65.
High-throughput screening of approximately 200,000
compounds revealed several hits that selectively inhib-
ited the cleavage of a chromogenic substrate by human
factor Xa63. The most potent of these hits was on a minor
impurity in a combinatorial library — a phosphonium
salt with an IC50 of 70 nM (compound 1 in FIG. 4). We
proposed that this positively charged phosphonium
moiety might serve as an arginine mimic and could be
interchangeable with an amidine group. This resulted in
the synthesis of lead compound 2 with similar potency
(IC50 of 120 nM)63. Further optimization resulted in
the synthesis of compounds of the isoindolinone class,
among which imidazoline 3 (FIG. 4) was the most potent
(IC50 of 2 nM).
To achieve oral bioavailability, we explored less basic
or non-basic amidine replacements. Aminopyridines,
such as compound 4 (IC50 of 8 nM; FIG. 4), showed IC50
values in the low nanomolar range; however, although
they were less basic, oral absorption remained insuf-
ficient. The less potent pyridylpiperazine derivative 5
(IC50 of 48 nM), meanwhile, revealed a significantly
improved oral bioavailability of 38% in rats. Although
we had demonstrated that improved oral bioavailability
could be achieved using less basic amidine replacements,
we were not able to meet our target of identifying factor
Xa inhibitors with both high potency and sufficient oral
bioavailability in the isoindolinone class of compounds.
However, from these studies we did learn that broad
variations in the benzylamidine part were permissible,
but found that there was a very steep structure–activity
relationship (SAR) for the chlorothiophene carboxamide
moiety, which was already present in the lead structure.
Figure 2 | Simplified schematic for the blood coagulation cascade. The figure
identifies the target points of various anticoagulants and illustrates that factor X (FX) can
be activated through either the intrinsic or the extrinsic pathway. The factor VIIa–tissue
factor (TF) complex (extrinsic tenase) activates both factor IX and factor X, as well as
factor VII itself (dashed arrow)150,151. Initiation of either pathway activates the inactive
precursor, factor X, to factor Xa. This makes factor Xa a desirable intervention point for
novel anticoagulants, because it occupies a central role in the blood-coagulation
pathway20. Furthermore, in addition to initiation, both the intrinsic and extrinsic
pathways lead to the propagation and amplification of coagulation through the
activation of factor X. During the initiation phase of coagulation, the factor Xa produced
generates some thrombin (factor IIa). This initial thrombin activates factor XI, and factors
V and VIII, to factor XIa and the activated cofactors, factor Va and VIIIa, respectively. It
also activates platelets (not shown), which are required for the formation of the intrinsic
tenase (factor VIIIa–factor IXa) and the prothrombinase (factor Va–factor Xa) complexes.
The prothrombinase complex, on the platelet surface, is substantially more efficient than
free factor Xa at activating prothrombin to thrombin; the rate of thrombin formation is
increased by approximately 300,000-fold over the rate with factor Xa alone152. Thrombin
is the principal enzyme involved in the formation, growth and stabilization of thrombi.
Thrombin mediates the conversion of fibrinogen to fibrin, the activation of factor XIII
(which crosslinks and stabilizes fibrin), the activation of platelets and the
above-mentioned feedback-activation of upstream coagulation factors, factor V, factor
VIII and factor XI151,153, resulting in the amplification of its own formation150,154. Because
one molecule of factor Xa catalyses the formation of approximately 1,000 thrombin
molecules25,154, this amplification can be substantial. Compared with thrombin, factor Xa
is thought to have fewer effects outside coagulation and, together with factor Va (as the
prothrombinase complex), acts mainly to convert prothrombin to thrombin. Specific
inhibition of factor Xa does not affect pre-existing thrombin but does inhibit thrombin
generation155. Conversely, although thrombin generation is delayed in the presence of a
thrombin inhibitor, the amount of thrombin generated is only reduced at higher inhibitor
concentrations, albeit in a concentration-dependent manner156. LMWH, low-molecular-
weight heparin.
REVIEWS
64
|
JANUARY 2011
|
VOLUME 10 www.nature.com/reviews/drugdisc
© 2011 Macmillan Publishers Limited. All rights reserved
N
N
H3C
O O
OH
NH2
HN
FF
N
N
H3C
OOH
N NH2
NH
S
O
O
HO O
O
N
H3C
NH
N
O
O
N
O
O
H
N
O
S
Cl
N N
NN
NH2
O
O
O
O
CH3
N
H3C
NH
CH3
H
N
O
H
N
ON
Cl
O
CH3
Fidexaban
(ZK-807834, CI-1031)
Otamixaban
(FXV673, RPR130673)
DX-9065a YM-60828
Rivaroxaban
(BAY 59-7939)
Apixaban
(BMS 562247)
Betrixaban
(PRT054021)
Edoxaban
(DU-176b)
Fondaparinux
N
H
O
CH3
O CH3O
HN NH2
N+
O
NH3C
NH
O
HO O
NH2
NH
NN
S
H3C
N
H
O
O N
CH3
CH3
HN
O
O
HN
N
Cl
O
NaO3SO
NaO3SO
NaO3SO
O
NHSO3Na
NHSO3Na
NHSO3Na
HO
HO
O
NaO2C
NaO2C
O
OH
HO
O
O
NaO3SO
O
HO OSO3Na
O
O
O
CH3
HO
Nature Reviews | Drug Discovery
International normalized
ratio
(INR). Because prothrombin
time-test results vary according
to the activity of the
thromboplastin used, the INR
conversion is used to normalize
results for any thromboplastin
preparation. It is valid only with
vitamin K antagonists.
Thromboprophylaxis
A measure taken to prevent
the development of a
thrombus. It can be
pharmaceutical or mechanical.
Tissue factor
A cell-membrane-bound
receptor protein that is
exposed to the circulating
blood during vessel injury.
Pre-existing factor VIIa in the
blood binds to tissue factor,
initiating the coagulation
cascade.
Direct thrombin inhibitors
A class of anticoagulants that
bind directly to thrombin and
block the interaction with its
substrate, fibrinogen, thereby
inhibiting the generation of
fibrin and clot formation.
Thrombomodulin
A membrane-bound thrombin
receptor that, when bound to
thrombin, functions as a
cofactor in the
thrombin-induced activation of
protein C.
Protein C
The inactive precursor of
activated protein C (APC). APC,
with its cofactor protein S,
inactivates factor Va and factor
VIIIa, thus providing an
important anticoagulant
feedback function.
Factor Xa inhibitor
A class of anticoagulants that
inhibit factor Xa in the
coagulation cascade, either by
binding directly, or indirectly
through antithrombin.
Inhibition of factor Xa reduces
the production of thrombin.
Venous thromboembolism
(VTE). A condition in which a
blood clot (thrombus) that has
formed in the venous system
breaks free (becoming an
embolus) and migrates through
the circulation to lodge in and
block another blood vessel.
The failure to find a compound with sufficient
potency and bioavailability could have ended the
project. However, we decided instead to re-evaluate
the weaker screening hits. The oxazolidinone (com-
pound 6; FIG. 4) was a weak factor Xa inhibitor, with
an IC50 of 20,000 nM. Considering the importance of
the chlorothiophene residue in the class of compounds
studied previously, we replaced the thiophene moiety of
compound 6 with a 5-chlorothiophene group, thereby
creating lead compound 7 (FIG. 4). This had a >200-fold
higher potency (IC50 of 90 nM) and did not include any
basic group63.
Figure 3 | Structures of various factor Xa inhibitors. Fidexaban, otamixaban, DX-9065a, YM-60828, rivaroxaban
(Xarelto; Bayer HealthCare), apixaban (Pfizer/Bristol-Myers Squibb), betrixaban and edoxaban are shown. The structure
of YM150 has not been published. Not all of these compounds are still in clinical development (for example, fidexaban
and DX-9065a) but study of their structures contributed to our understanding of the structure–activity relationships
and pharmacology of factor Xa inhibitors. Key features to note are the highly basic arginine mimetic amidine groups as
P1 moieties in fidexaban, otamixaban, DX-9065a and YM-60828, which contribute to poor oral bioavailability. These
can be contrasted with the non-basic P1 moieties: the chlorothiophene moiety in rivaroxaban, the methoxyaryl group
in apixaban and the chloro-substituted pyridine rings in edoxaban and betrixaban, all of which allow improved oral
bioavailability. Other synthetic factor Xa inhibitors have also been developed26. The structure of the synthetic indirect
factor Xa inhibitor, fondaparinux (Arixtra; GlaxoSmithKline), is also shown. This is an analogue of the heparin
pentasaccharide sequence required to mediate the conformational change in antithrombin and subsequent binding
to factor Xa2.
REVIEWS
NATURE R EVIEWS
|
DRUG DISCOVERY VOLUME 10
|
JANUARY 2011
|
65
© 2011 Macmillan Publishers Limited. All rights reserved
6 R = H IC50 = 20,000 nM
7 R = Cl IC50 = 90 nM
NS N
O
HN
O
S
R
O
F
R N
O
HN
O
S
Cl
O
F
*
8 R = IC50 = 32 nM
Oral bioavailability: 94%
NO
*
9 R = IC50 = 40 nM
N
R N
O
HN
O
S
Cl
O
*
10 R = IC50 = 4 nM
Oral bioavailability: 65%
N
O
*
11 R = IC50 = 0.7 nM
Oral bioavailability: 60%
NO
O
Oxazolidinones
O
R
N
HN
O
O
O
S
Cl
NP *
+
N
N
CF3COO
1 R = IC50 = 70 nM
5
IC50 = 48 nM
Oral bioavailability: 38%
*
2 R = IC50 = 120 nM
H2N
NH
R
N
HN
O
O
S
Cl
*
3 R =
IC50 = 2 nM
Oral bioavailability: <1%
N
HN
*
4 R = IC50 = 8 nM
Oral bioavailability: <1%
N
HN
N N N
O
N
OHN
O
S
Cl
Isoindolinones
Nature Reviews | Drug Discovery
Deep-vein thrombosis
(DVT). A blood clot in a deep
vein, usually in the leg. Distal
DVT occurs in the calf, whereas
proximal DVT occurs above the
knee.
Pulmonary embolism
A blood clot or
thromboembolus in a
pulmonary blood vessel. Such
emboli generally originate from
a deep-vein thrombosis and
can cause permanent lung
damage, chronic pulmonary
hypertension and death.
Haemostasis
The complex process that leads
to the formation of a blood clot,
causing bleeding to stop.
On the basis of this promising lead, a medicinal
chemistry programme was followed that focused on fur-
ther improving the potency of the oxazolidinone class
without compromising its pharmacokinetic profile.
The optimization programme leading to rivaroxa-
ban. The starting point of these investigations was the
thiomorpholine group; morpholine and pyrrolidine
derivatives (compounds 8 and 9; FIG. 4) showed some
improvement in potency (IC50 values of 32 nM and
40 nM, respectively). Although not sufficiently potent,
compound 8 was the first compound to show a favour-
able pharmacokinetic profile, with a high oral bioavail-
ability of 94% in rats. An ortho-substitution led to the
pyrrolidinone derivative 10, with significantly improved
potency (IC50 of 4 nM) and an oral bioavailability of
65%. However, the in vivo antithrombotic potency of
this compound, evaluated in an arteriovenous shunt
model in anaesthetized rats, was estimated to be too
low63. Our knowledge of the SAR then led us to design
the morpholinone residue, resulting in compound 11
(rivaroxaban; FIG. 4), for which binding to factor Xa
depends on the (S)-configuration at the oxazolidinone
core. Rivaroxaban showed increased potency in vitro
(IC50 of 0.7 nM) and in vivo after oral administration in
rats63. This compound also showed favourable oral bio-
availability (60% in rats and 60–86% in dogs)63.
The optimization programme was continued in
order to gain a more in-depth understanding of the
SAR. However, further variations evaluated at this time,
such as substitution of the aryl ring or less lipophilic
replacements for the chlorothiophene carboxamide
moiety, did not result in further improvements63, and
the morpholinone derivative 11 (rivaroxaban) remained
the most attractive candidate. The X-ray crystal struc-
ture of rivaroxaban in complex with human factor Xa63
(BOX 1) helped us to understand its binding mode and to
explain the steep SAR observed earlier. A similar bind-
ing mode has been reported for several other factor Xa
inhibitors66–70 (BOX 1).
With its combination of high binding affinity and
good oral bioavailability, rivaroxaban was identified as
the drug candidate for further development.
Preclinical studies
Rivaroxaban is a direct, specific factor Xa inhibitor that,
unlike indirect agents such as fondaparinux, does not
require a cofactor71. In vitro kinetic studies showed that
the inhibition of human factor Xa by rivaroxaban was
competitive (Ki of 0.4 nM), with >10,000-fold greater
selectivity for factor Xa than for other serine proteases71.
Rivaroxaban inhibits prothrombinase (IC50 of 2.1 nM)71
and initial data suggest that it also inhibits clot-bound
factor Xa activity (IC50 of 75 nM)72. In human plasma,
Figure 4 | Optimization of oxazolidinone factor Xa inhibitors. Optimization of oxazolidinone factor Xa inhibitors
resulted in the discovery of rivaroxaban (compound 11)63. Half-maximal inhibitory concentration (IC50) values are for the
inhibition of factor Xa activity; oral bioavailability is shown for rats.
REVIEWS
66
|
JANUARY 2011
|
VOLUME 10 www.nature.com/reviews/drugdisc
© 2011 Macmillan Publishers Limited. All rights reserved
Tyr99
Asp102
His57
Tyr228
Ser195
S1
Asp189
Gly219
Phe174
S4
Trp215
Nature Reviews | Drug Discovery
Fibrinogen
A soluble plasma protein that,
in the final phase of the
coagulation process, is
converted to fibrin by
thrombin. Fibrin then
polymerizes and forms the
fibrous network base of a clot.
Oral bioavailability
The total proportion of
pharmacologically active drug
that enters the systemic
circulation after oral
administration. It is affected by
both absorption and local
metabolic inactivation.
Prothrombin time
A laboratory test that
measures clotting time in the
presence of tissue factor
(thromboplastin). It is used to
assess the activity of the
extrinsic coagulation pathway.
Activated partial
thromboplastin time
A laboratory test that measures
the clotting time of plasma
after contact activation. It
assesses the function of the
intrinsic coagulation pathway.
rivaroxaban inhibited thrombin generation — and there-
fore the amplification processes of coagulation — through
the inhibition of factor Xa generated by either the intrin-
sic or the extrinsic coagulation pathways73,74. Thrombin
generation was almost completely inhibited in platelet-
rich plasma at physiologically relevant concentrations
(80–100 nM) of rivaroxaban73. This and other studies dem-
onstrated that rivaroxaban prolonged the initiation phase
of thrombin generation, potently inhibited the physiologi-
cally relevant prothrombinase complex-bound factor Xa
on the surface of activated platelets71 and reduced the
thrombin burst produced in the propagation phase73.
In addition, preliminary work has shown that rivar-
oxaban inhibits thrombin generation in the presence
and absence of thrombomodulin in human plasma in a
concentration-dependent manner. This suggests that it
does not interfere with the thrombin–thrombomodulin–
APC system and, therefore, probably does not suppress
APC-mediated negative feedback75, as was shown for
DX-9065a49. Further preliminary data suggested that
rivaroxaban did not directly affect platelet aggregation
in platelet-rich plasma induced by thrombin, adeno-
sine diphosphate or collagen76, but potently inhibited
tissue-factor-induced platelet aggregation in an indirect
manner, by inhibiting thrombin generation77.
Rivaroxaban demonstrated anticoagulant effects in
human plasma, with the prothrombin time being more
sensitive than the activated partial thromboplastin time71, a
finding also observed with other direct factor Xa inhibi-
tors in clinical development78,79. This may be because
direct factor Xa inhibitors, including rivaroxaban71,
are highly effective inhibitors of the prothrombinase
complex, although differences in enzyme kinetics may
also be responsible20. A dose-dependent prolongation
of prothrombin time was demonstrated in vivo in rat
and rabbit models, with a strong correlation observed
between prothrombin time and plasma concentrations
of rivaroxaban (r=0.98)71.
In vivo, rivaroxaban given prophylactically had potent
and consistent antithrombotic effects in venous71,80 and
arterial71 thrombosis models in rats and rabbits. In a rab-
bit treatment model, rivaroxaban reduced the accretion
of radiolabelled fibrinogen into preformed clots in the
jugular vein, relative to untreated controls80. Bleeding
times in rats and rabbits were not significantly affected at
antithrombotic-effective doses71, indicating a favourable
efficacy/bleeding ratio.
Although rivaroxaban has a short half-life in
humans81,82 (see the Xarelto Summary of Product
Characteristics (SMPC) from the European Medicines
Box 1 | Binding mode of rivaroxaban
The figure shows the X‑ray crystal structure
of rivaroxaban (carbons coloured orange)
in complex with human factor Xa63.
Essential amino acids and binding pockets
(S1 and S4) are indicated; hydrogen bonds
are shown as dotted lines.
Two hydrogen bonds are formed between
rivaroxaban and the amino acid Gly219 of
factor Xa. The first, a strong interaction
(2.0 Å), is from the carbonyl oxygen of the
oxazolidinone core. The second, weaker
interaction (3.3 Å), is from the amino group
of the chlorothiophene carboxamide
moiety. These two hydrogen bonds support
the oxazolidinone core in directing its
substituents into the S1 and the S4 subsites
of factor Xa. This results in rivaroxaban
forming an L‑shape, a binding mode typical
of synthetic, direct factor Xa inhibitors63.
The aromatic rings of Tyr99, Phe174 and
Trp215 in factor Xa define a narrow
hydrophobic channel that comprises the S4
pocket. The morpholinone moiety of rivaroxaban is ‘sandwiched’ between Tyr99 and Phe174, while its aryl ring is oriented
perpendicularly, extending across the face of Trp215. There is no direct interaction between the morpholinone carbonyl
group and the factor Xa backbone; rather, this carbonyl group contributes to a planarization of the morpholinone ring,
further supporting the sandwich‑like arrangement63. Using the morpholinone moiety, as well as other six‑membered rings
such as lactames or pyridinones, we found new, non‑basic P4 residues (see REF. 146 for patent application), yielding
strongly increased binding to factor Xa. Such residues have since been used successfully for several other factor Xa
inhibitors69,147, including apixaban148 and PD0348292 (eribaxaban)149.
In the S1 pocket of factor Xa, the key interaction is between the chlorine substituent of the thiophene moiety and the
aromatic ring of Tyr228 at the bottom of the S1 pocket. This novel interaction obviates the need for strongly basic groups,
such as amidines, to achieve high factor Xa affinity, and therefore enables non‑basic rivaroxaban to achieve both high
potency and good oral bioavailability63. A similar binding mode has been found and reported for several other factor Xa
inhibitors66–70.
Figure modified, with permission, from REF. 63 (2005) American Chemical Society.
REVIEWS
NATURE R EVIEWS
|
DRUG DISCOVERY VOLUME 10
|
JANUARY 2011
|
67
© 2011 Macmillan Publishers Limited. All rights reserved
CYP450 isoforms
Many therapeutic drugs are
metabolized by cytochrome
P450 (CYP450) enzymes, a
‘superfamily’ of related but
distinct enzymes that differ in
their substrate specificity.
Creatinine clearance
The rate at which the kidney
clears the blood of creatinine (a
waste product from muscles
that is excreted at a fairly
constant rate). Creatinine
clearance is used as an
approximation of the
glomerular filtration rate.
Agency (EMA); Further information), it may be of
interest to determine whether the anticoagulant effect of
rivaroxaban can be reversed, because there might be rare
instances in which this would be of use — for example,
a need for emergency surgery. However, it is worth not-
ing that several established anticoagulants also lack full
antidotes. For example, the anticoagulant effect of the
LMWHs is only partly reversed by protamine sulphate2
and there is no available antidote for fondaparinux (see
Arixtra’s prescribing information; Further information).
Preliminary studies in primates and rats have been con-
ducted to evaluate possible antidotes for rivaroxaban.
These studies suggested that recombinant factor VIIa,
activated prothrombin complex concentrate (FEIBA
(factor VIII inhibitor bypassing activity); Baxter)83 and
prothrombin complex concentrate84 can partially reverse,
in a dose-dependent manner, the effects of high-dose
rivaroxaban on bleeding time. However, it is important to
note that there is no clinical experience with any of these
reversal strategies (see the EMA’s SMPC on rivaroxaban;
Further information).
Clinical pharmacology
In Phase I and Phase II studies, rivaroxaban exhibited
predictable pharmacokinetic and pharmacodynamic
profiles, as follows. It is rapidly absorbed, with maximum
plasma concentrations (Cmax) occurring 2–4 hours after
tablet intake82,85. Oral bioavailability of rivaroxaban is
decreased at higher doses, possibly owing to poor solu-
bility. However, at the currently approved 10 mg dose,
oral bioavailability is 80–100%85.
Food (a high-fat, high-calorie meal) did not affect the
area under the plasma concentration–time curve (AUC)
or Cmax for the 10 mg tablet (EMA’s SMPC on rivaroxa-
ban). Rivaroxaban was safe and well tolerated across a
wide dose range, with dose-proportional pharmacoki-
netics and pharmacodynamics. No relevant accumula-
tion was observed at any dose level after multiple dosing
in healthy subjects at steady state82.
Rivaroxaban is metabolized by CYP450 isoforms, par-
ticularly CYP3A4, which is strongly inhibited by ketoco-
nazole and ritonavir (see prescribing information from
Janssen Pharmaceutica (ketoconazole) and Abbott (riton-
avir), and the EMA’s Committee of Medicinal Products
for Human Use (CHMP) report for rivaroxaban; Further
information). These drugs were, therefore, predicted
to affect the metabolism of rivaroxaban and, because
both drugs are also strong inhibitors of P-glycoprotein
(P-gp) and CYP3A4 (REFS 86,87), may increase plasma
concentrations of rivaroxaban. This expectation was
confirmed in subsequent Phase I studies that showed a
strong interaction between rivaroxaban and these two
drugs. However, there was no clinically relevant inter-
action with clarithromycin, a strong CYP3A4 inhibitor
but only a moderate inhibitor of P-gp, indicating that
rivaroxaban may be used with substances that strongly
inhibit only one of the two elimination pathways (see
the EMA’s SMPC on rivaroxaban and prescribing infor-
mation from Janssen Pharmaceutica (ketoconazole) and
Abbott (ritonavir)). Conversely, these findings show that
concomitant use of rivaroxaban with strong inhibitors
of both CYP3A4 and P-gp such as ketoconazole or HIV
protease inhibitors such as ritonavir is not recommended.
In addition, rivaroxaban should be used with caution if
strong CYP3A4 inducers are administered concomitantly
(EMA’s CHMP report for rivaroxaban).
Rivaroxaban has a low propensity for drug–drug inter-
actions with frequently used concomitant medications
such as naproxen88 and asprin89, and no interaction with
the cardiac glycoside digoxin (Lanoxin; GlaxoSmithKline)
(D. Kubitza, unpublished observations).
Furthermore, dietary restrictions are not necessary
at the 10 mg once-daily dose; rivaroxaban was given
with and without food in the Phase III VTE-prevention
studies (RECORD studies 1–4)90–93 (EMA’s SMPC on
rivaroxaban).
Rivaroxaban is distributed heterogeneously to tissues
and organs, and exhibits only moderate tissue affinity in
rats; importantly, it does not substantially penetrate the
blood–brain barrier94. However, like many small mol-
ecules, rivaroxaban is expected to be able to cross the pla-
centa, although specific studies have not been published.
In vitro and Phase I studies showed that rivaroxaban has
a dual mode of elimination, with one-third eliminated
unchanged by the kidneys and two-thirds metabolized by
the liver to inactive metabolites, with no major or active
circulating metabolites detected in plasma95,96. Elimination
of rivaroxaban from plasma occurs with a mean terminal
half-life of 7–11 hours81,82 (EMA’s SMPC on rivaroxaban).
With a systemic clearance rate of approximately 10 L h–1,
rivaroxaban can be classified as a low-clearance drug97,98.
Low intra-individual and moderate inter-individual phar-
macokinetic variability have been observed97,98. In Pha se I
studies, the coefficient of variation for AUC ranged from
18% to 33%, and for Cmax from 16% to 39%, for inter-
individual variability. Median intra-individual variability
was only 14% and 19% for AUC and Cmax, respectively
(EMA’s CHMP report for rivaroxaban).
Rivaroxaban inhibited factor Xa activity in a dose-
dependent manner after single dosing in healthy sub-
jects. Rivaroxaban also inhibited thrombin generation
in healthy subjects, and some parameters of thrombin
generation remained inhibited for 24 hours after admin-
istration of a 30 mg dose74. These results offered the first
indication that once-daily dosing might be feasible.
There were close correlations between pharmacoki-
netic and pharmacodynamic parameters82. Rivaroxaban
plasma concentrations correlated closely with prolonga-
tion of prothrombin time and inhibition of factor Xa.
Plasma concentrations correlated with prothrombin
time both in healthy volunteers82 and in patients under-
going either total hip replacement (THR) or total knee
replacement (TKR) in Phase II studies97.
Various Phase I and II studies, using total daily doses
ranging from 5 mg to 80 mg, indicated that rivaroxa-
ban could be given irrespective of age, gender81, body
weight99 and mild (creatinine clearance: 50–79 ml min–1 )
to moderate (creatinine clearance: 30–49 ml min–1) renal
impairment100. A further preliminary clinical study sug-
gested that rivaroxaban can also be given irrespective
of mild hepatic impairment (classified as Child–Pugh
class A)101. Fixed doses of rivaroxaban were administered
REVIEWS
68
|
JANUARY 2011
|
VOLUME 10 www.nature.com/reviews/drugdisc
© 2011 Macmillan Publishers Limited. All rights reserved
Chromogenic assay
An enzymatic assay in which a
colour develops during the
course of the reaction, which
can then be measured spectro-
photometrically. Colour
development is reduced in the
presence of an inhibitor.
Venography
Radiography of the veins after
intravenous injection of a
radioactive isotope or contrast
dye. This can be used to
confirm the presence of
deep-vein thromboses.
in Phase II studies investigating rivaroxaban for the pre-
vention or treatment of VTE, with no restrictions on
gender, or mild or moderate renal impairment102105.
These parameters were shown to have no clinically
relevant effect on the pharmacokinetics and pharma-
codynamics of rivaroxaban97. In addition, rivaroxaban
had no effect on heart-rate-corrected QT interval pro-
longation106. Owing to its pharmacokinetic and phar-
macodynamic profiles, rivaroxaban can be given at
fixed doses to adult patients with no requirement for
routine coagulation monitoring. However, there may be
rare occasions when monitoring is of use — for exam-
ple, where poor compliance is suspected or emergency
surgery is required, or to confirm a possible overdose.
This concern has led to the evaluation of a number of
different assay types. Initial results suggest that, in light
of an apparent strong correlation between the extent of
factor Xa inhibition and rivaroxaban plasma concen-
tration, chromogenic assays for factor Xa may be par-
ticularly suited for the quantification of rivaroxaban in
plasma107,108.
Clinical development of rivaroxaban
Rivaroxaban is approved for the prevention of VTE after
elective hip or knee replacement in approximately 100
countries, including member states of the European
Union (as of 30 September 2008) and Canada (as of
16 September 2008). Clinical development is ongoing
for the prevention and treatment of thromboembolic
disorders in other conditions, including the treatment
of VTE, the prevention of VTE in hospitalized, medi-
cally ill patients (for example, patients hospitalized with
an acute medical condition, such as cancer, heart failure
or respiratory failure, or requiring intensive care), as well
as the prevention of stroke in patients with atrial fibril-
lation and secondary prevention in patients with ACS.
The TIMELINE highlights the development of rivaroxaban
during the past decade.
VTE prophylaxis after total hip or knee replacement.
A large, comprehensive Phase II programme involving
about 2,900 patients assessed both once-daily and twice-
daily dosing regimens102105. Collectively, these studies
demonstrated an optimal dose range of 5–20 mg per
day109, and indicated that rivaroxaban 10 mg once daily
had the optimum balance of efficacy and safety, relative
to enoxaparin 40 mg once daily103 (FIG. 5).
On the basis of the results of the Phase II studies,
rivaroxaban 10 mg once daily was selected for investiga-
tion in the Phase III RECORD programme, which com-
prised four large studies involving a total of more than
12,500 patients undergoing elective THR or TKR90–93
(TABLE 2). In all four studies, the primary efficacy end
point was the composite of any DVT, as detected by
mandatory, bilateral venography, non-fatal pulmonary
embolism and all-cause mortality. The primary safety
end point was major bleeding, which in the RECORD
programme did not include surgical-site bleeding90–93.
The RECORD1 and 3 studies were designed to compare
rivaroxaban 10 mg once daily with the standard of care
enoxaparin 40 mg once daily — both given for 31–39 days
(extended prophylaxis) after THR (RECORD1)90 or for
10–14 days after TKR (RECORD3)92. In both studies,
rivaroxaban was more effective than enoxaparin for the
prevention of VTE90,92 (TABLE 2). Furthermore, the inci-
dence of major bleeding was comparable and not signifi-
cantly different between treatment groups90,92. RECORD3
also showed a significant reduction in symptomatic VTE
for rivaroxaban92.
RECORD2 investigated the efficacy and safety
of extended thromboprophylaxis with rivaroxaban
(35 days; range 31–39 days) compared with short-term
enoxaparin treatment (10–14 days) followed by placebo
in patients undergoing THR91. The results demonstrated
that extended prophylaxis with rivaroxaban (10 mg
once daily) was superior to short-term prophylaxis with
enoxaparin (40 mg once daily) followed by placebo for
Timeline | Rivaroxaban: more than a decade of progress
1998 1999 2000 2002 2003 2004 2005 2007 2008 2009 2010
A new class of
compounds, the
oxazolidinones, is
identified and
optimized, leading
to the discovery of
rivaroxaban
Bayer HealthCare
initiates the factor
Xa programme
Preclinical studies
on rivaroxaban
are initiated
Phase II studies
initiated
The RECORD
clinical trial
programme begins
Rivaroxaban approved and
launched in the EU and Canada
for VTE prophylaxis after hip or
knee replacement
Results of the RECORD1, 2 and
3, and EINSTEIN DVT Phase II
studies are published90–92, 117
ROCKET AF clinical trial data
presented (http://sciencenews.
myamericanheart.org/pdfs/
ROCKET_AF_pslides.pdf)
EINSTEIN DVT and EINSTEIN EXT
Phase III results published119
Rivaroxaban shows
favourable pharmacokinetic
and pharmacodynamic
profiles in Phase I studies82,85
Phase I studies show
rivaroxaban to be
well tolerated81,82,85
Phase II studies
suggest a favourable
benefit–risk
profile103–105
RECORD1 and 3 show
rivaroxaban to be more
effective than enoxaparin in
the prevention of VTE, with
a comparable safety profile.
RECORD2 shows benefit of
extended prophylaxis with
rivaroxaban90–92
Regulatory file submitted
for approval to the EMA
FDA Advisory Committee
meeting
RECORD4 and ATLAS ACS
TIMI 46 data published93,114
DVT, deep-vein thrombosis; EMA, European Medicines Agency; EU, European Union; FDA, US Food and Drug Administration; VTE, venous thromboembolism.
REVIEWS
NATURE R EVIEWS
|
DRUG DISCOVERY VOLUME 10
|
JANUARY 2011
|
69
© 2011 Macmillan Publishers Limited. All rights reserved
Incidence–efficacy (%)
Incidence–safety (%)
40
30
20
10
0
30
20
10
0
10
520 30 40 Enoxaparin
Total daily dose of rivaroxaban (mg)
0
DVT, PE and all-cause death
Major postoperative bleeding
Nature Reviews | Drug Discover
y
Index event
The acute event that leads to a
patient’s initial presentation.
The term can also refer to the
initial event resulting in a
patient’s inclusion in a
follow-up study, such as a
survey of recurrent strokes.
the prevention of VTE, including symptomatic events91
(TABLE 2). Despite rivaroxaban being given for 3 weeks
longer than enoxaparin, the incidence of treatment-
emergent major bleeding (that is, up to 2 days after the
last dose of study medication) was <0.1% in both treat-
ment groups. Guidelines recommend a minimum dura-
tion for prophylaxis of 10 days, and also recommend that
prophylaxis be extended for up to 35 days for patients
undergoing THR110. The RECORD2 study provided
additional confirmation of the benefits of extended
prophylaxis over short-term prophylaxis after THR.
RECORD4 compared the efficacy and safety of oral
rivaroxaban 10 mg once daily with the commonly used
North American regimen of enoxaparin 30 mg twice
daily in patients undergoing TKR93. Rivaroxaban was sig-
nificantly superior to enoxaparin for the primary efficacy
end point, with no significant difference in the rates of
major bleeding between the two groups (TABLE 2). So far,
rivaroxaban is the only new oral anticoagulant to dem-
onstrate superior efficacy over the greater dose intensity
of the North American enoxaparin regimen93,111,112.
A comparison of rivaroxaban with enoxaparin in
these four studies showed the efficacy and safety of rivar-
oxaban in the prevention of VTE in patients undergoing
elective THR or TKR. The superiority of rivaroxaban
for the primary efficacy end point was demonstrated in
all four studies. Rivaroxaban also showed a good safety
profile, with a low incidence of major bleeding, com-
parable to that observed with enoxaparin90–93, and no
evidence of compromised liver function attributable to
rivaroxaban113. Furthermore, the incidence of haemor-
rhagic wound complications (composite of excessive
wound haematoma and reported surgical-site bleeding)
was similar in both treatment groups113. Liver safety is
also an important consideration in regulatory review. In
a Phase II trial of rivaroxaban in patients with ACS, no
patient who had received rivaroxaban for 6 months had
an alanine-amino-transferase level greater than three
times the upper limit of normal or total bilirubin greater
than twice the upper limit of normal114.
These studies were also conducted with no routine
coagulation monitoring and no dose adjustment for
demographic variables, consistent with preliminary
results from pooled subgroup analyses115.
Treatment of VTE. The efficacy and safety of rivaroxaban
for the treatment of VTE were assessed in two Phase IIb
dose-ranging studies116,117. These studies suggested
that rivaroxaban had good efficacy and a similar safety
profile, compared with standard therapy, for the treat-
ment of acute symptomatic DVT. An initial intensified
twice-daily regimen (rivaroxaban 15 mg twice daily for
3 weeks) followed by convenient 20 mg once-daily dos-
ing for 3, 6 or 12 months was selected for investigation
in Phase III studies. The efficacy and safety of rivaroxa-
ban for the treatment of VTE are being assessed in three
Phase III studies involving approximately 9,000 patients
in total— EINSTEIN DVT (ClinicalTrials.gov identifier:
NCT00440193), EINSTEIN PE (NCT00439777) and
EINSTEIN EXT (NCT00439725).
EINSTEIN PE (pulmonary embolism) is still ongo-
ing. However, data for EINSTEIN DVT (presented at
the 2010 American Society of Hematology meeting118
and recently published119) showed that rivaroxaban
(15 mg twice daily for 21 days followed by 20 mg once
daily) was non-inferior for the prevention of recurrent
symptomatic VTE in comparison to the current standard
of care (enoxaparin 1.0 mg kg–1 twice daily for ≥5 days
followed by VKA titrated to INR 2.0–3.0). First symp-
tomatic recurrent VTEs occurred in 2.1% of patients
receiving rivaroxaban compared with 3.0% of those in
the enoxaparin/VKA treatment arm. Major or non-major
clinically relevant bleeding occurred in 8.1% of patients
in each treatment arm119.
In the EINSTEIN EXT trial, 1,196 patients (intention-
to-treat population) who had completed 6–12 months
of anticoagulant therapy for the acute index event (VTE)
were randomized to an additional 6–12 months of therapy
with either rivaroxaban, 20 mg once daily, or placebo119.
Study medication was administered for a mean period
of 190 days in each treatment group. Recurrent sympto-
matic VTE events were observed in 7.1% and 1.3% of the
placebo and rivaroxaban treatment groups, respectively
(hazard ratio 0.18, P<0.0001), a relative risk reduction of
82% with rivaroxaban119 (hazard ratio, 0.68; 95% confi-
dence interval, 0.44–11.04; P<0.001 for non-inferiority).
Major bleeding did not occur in any placebo-treated
patients and occurred in four (0.7%) rivaroxaban-treated
patients, although none of these occurred in a critical
location or proved fatal119.
Thromboprophylaxis in medically ill patients. A
Phase III study (NCT00571649) has been initiated to
investigate the efficacy and safety of VTE prophylaxis
Figure 5 | Efficacy and safety dose–response relationships. The figure shows the
dose–response relationships between rivaroxaban total daily dose and the primary
efficacy end point (any deep-vein thrombosis (DVT); non-fatal pulmonary embolism (PE)
and all-cause death) and primary safety end point (major bleeding) in the once-daily
study investigating rivaroxaban for the prevention of venous thromboembolism after
total hip replacement103. Solid lines show the dose–response curves (logistic regression),
blue hatched lines represent 95% confidence intervals for efficacy and red hatched
lines represent 95% confidence intervals for safety. Details for 40 mg enoxaparin once
daily, the current European standard of care, are shown for comparison. Figure
reproduced, with permission, from REF.103 (2006) Lippincott Williams & Wilkins.
REVIEWS
70
|
JANUARY 2011
|
VOLUME 10 www.nature.com/reviews/drugdisc
© 2011 Macmillan Publishers Limited. All rights reserved
with rivaroxaban 10 mg once daily (for 35 ± 4 days),
compared with short-term 40 mg once-daily enoxaparin
(administered for 10 ± 4 days followed by placebo), in
hospitalized, medically ill patients.
Stroke prevention in non-valvular atrial fibrillation.
Rivaroxaban 20 mg once daily has been compared with
warfarin for the prevention of stroke and non-central-
nervous-system systemic embolism in approximately
14,000 patients with non-valvular atrial fibrillation in a
Phase III study (NCT00403767)120. Patients with moderate
renal impairment (creatinine clearance: 30–49 ml min–1)
received a fixed dose of 15 mg once daily. It was recently
reported at the 2010 American Heart Association
meeting that, in the ROCKET AF study, rivaroxaban
significantly reduced the risk of stroke and non-central-
nervous-system thromboembolism in patients with atrial
fibrillation compared to warfarin, with comparable rates
of bleeding (http://sciencenews.myamericanheart.org/
pdfs/ROCKET_AF_pslides.pdf).
Secondary prevention in ACS. A Phase III study inves-
tigating secondary prevention of ischaemic events in
patients with ACS (NCT00809965) was started in late
2008, and is expected to enrol up to 16,000 patients. Two
doses of rivaroxaban, 2.5 mg and 5 mg twice daily, are
being investigated on the basis of the results of a Phase IIb
study that assessed safety and efficacy in approximately
3,500 patients with recent, non-ST-elevation myocar-
dial infarction, ST-elevation myocardial infarction or
unstable angina114. As noted above, this trial also showed
no evidence of compromised liver function in patients
receiving rivaroxaban for up to 6 months.
Other direct factor Xa inhibitors in development
Below, we discuss other direct factor Xa inhibitors that
are in advanced clinical development (Phase III).
Apixaban. Apixaban (developed by Pfizer/Bristol-Myers
Squibb) is a small-molecule, oral, direct factor Xa inhibi-
tor (FIG. 3) that selectively and reversibly inhibits both free
factor Xa (Ki of 0.08 nM) and prothrombinase activity79,121.
Another study reported that apixaban reacts rapidly with
factor Xa (kon 2 × 107 M–1 s–1) with high-affinity binding
(Ki of 0.3 nM at 37 °C)122. Preclinical studies have shown
that apixaban was well absorbed in chimpanzees, dogs
and rats; mean oral bioavailability was 51% (chimpan-
zees), 88% (dogs) and 34% (rats)79. The drug is currently
Table 2 | Incidence of venous thromboembolism and bleeding events across the four RECORD* studies
End pointsEfficacy end point (% (n/N)) Safety end points (patients with bleeding events) (% (n/N))
Total VTE Major VTE Symptomatic
VTE
Major bleeding Major and non-major
clinically relevant bleeding
RECORD1 (THR: prophylaxis administered for 35 days)
Enoxaparin (40 mg once
daily)
3.7 (58/1,558) 2.0 (33/1,678) 0.5 (11/2,206) 0.1 (2/2,224) 2.5 (56/2,224)
Rivaroxaban (10 mg
once daily)
1.1 (18/1,595) 0.2 (4/1,686) 0.3 (6/2,193) 0.3 (6/2,209) 3.2 (70/2,209)
P value <0.001 <0.001 0.22 0.18 NS§
RECORD2 (THR: extended rivaroxaban prophylaxis, 35 days, versus short-term enoxaparin, 14 days, followed by placebo)
Enoxaparin (40 mg once
daily/placebo)
9.3 (81/869) 5.1 (49/962) 1.2 (15/1,207) <0.1 (1/1,229) 2.8 (34/1,229)
Rivaroxaban (10 mg
once daily)
2.0 (17/864) 0.6 (6/961) 0.2 (3/1,212) <0.1 (1/1,228) 3.3 (41/1,228)
P value <0.0001 <0.0001 0.004 0.98 (REF. 157) NR
RECORD3 (TKR: prophylaxis administered for 14 days)
Enoxaparin (40 mg once
daily)
18.9 (166/878) 2.6 (24/925) 2.0 (24/1,217) 0.5 (6/1,239) 2.7 (34/1,239)
Rivaroxaban (10 mg
once daily)
9.6 (79/824) 1.0 (9/908) 0.7 (8/1,201) 0.6 (7/1,220) 3.3 (40/1,220)
P value <0.001 0.01 0.005 0.77 0.44
RECORD4 (TKR: prophylaxis administered for 14 days)
Enoxaparin (30 mg
twice daily)
10.1 (97/959) 2.0 (22/1,112) 1.2 (18/1,508) 0.3 (4/1,508) 2.3 (34/1,508)
Rivaroxaban (10 mg
once daily)
6.9 (67/965) 1.2 (13/1,112) 0.7 (11/1,526) 0.7 (10/1,526) 3.0 (46/1,526)
P value 0.012 0.124 0.187 0.11 0.18
N, total number of patients evaluable in each treatment group; NR, not reported; NS, not significant; THR, total hip replacement; TKR, total knee replacement;
VTE, venous thromboembolism. *The four RECORD studies90–93 compared rivaroxaban regimens with enoxaparin regimens (the current standard of care); summary
results are shown in the table. Results shown as number (n) of patients experiencing an event; patients could have more than one event. §P value not reported, but
95% confidence interval for risk difference includes zero.
REVIEWS
NATURE R EVIEWS
|
DRUG DISCOVERY VOLUME 10
|
JANUARY 2011
|
71
© 2011 Macmillan Publishers Limited. All rights reserved
being evaluated in a number of thromboembolic indica-
tions, including the prevention and treatment of VTE, the
prevention of stroke in patients with atrial fibrillation and
the prevention of cardiovascular events in ACS.
In the Phase III programme for the prevention of VTE
after major orthopaedic surgery, apixaban did not meet
the prespecified criteria for non-inferiority compared
with enoxaparin 30 mg twice daily (NCT00371683) for
VTE prevention after TKR111. However, a second study
(NCT00452530) conducted in patients undergoing TKR
demonstrated superior efficacy for apixaban against
enoxaparin 40 mg once daily. Clinically relevant bleeding
(major or non-major) occurred in fewer patients given
apixaban, although the differences were not signifi-
cant123. A third study (NCT00423319), presented as an
abstract, assessed extended prophylaxis with apixaban
versus enoxaparin (40 mg once daily) in patients under-
going THR and also demonstrated superior efficacy for
apixaban with similar rates of major and non-major
clinically relevant bleeding124.
A Phase II trial in patients with ACS (NCT00313300)
assessed efficacy and safety in patients taking apixa-
ban compared with placebo125. The developing compa-
nies recently reported that a subsequent Phase III trial
(NCT00831441) in patients with ACS investigating
whether apixaban (5 mg twice daily) is superior to pla-
cebo has been halted (press release, Bristol-Myers Squibb;
Further information). In addition, other apixaban trials are
ongoing for VTE prevention in patients with acute medical
illness (NCT00457002), or recently completed for malig-
nant disease (NCT00320255), as well as Phase III trials
for the treatment of VTE (NCT00633893, NCT00643201)
and trials for the prevention of stroke in patients with
atrial fibrillation (NCT00412984, NCT00496769)126,127.
Data for the AVERROES trial (NCT00496769) were
recently presented at the European Society of Cardiology
Congress. Patients with atrial fibrillation who have either
been demonstrated to be or were expected to be unsuita-
ble for treatment with VKAs received apixaban 5 mg twice
daily or aspirin 81–324 mg per day. Preliminary results
showed that the primary end point of stroke or systemic
embolism occurred at a significantly lower rate in patients
receiving apixaban, with an absolute risk reduction of
approximately 2% versus aspirin128.
Edoxaban. Edoxaban (developed by Daiichi Sankyo) is
an oral, direct and specific factor Xa inhibitor with a Ki
of 0.56 nM (FIG. 3); Ki values for other coagulation factors
are >10,000-fold higher78. In healthy volunteers, peak
plasma levels of edoxaban were observed at 1.5 hours
after a single oral dose, corresponding to the maximum
inhibition of factor Xa activity; and ex vivo thrombus
formation was reduced at 1.5 hours and 5 hours after
administration129.
Phase II trials in patients undergoing TKR130 or
THR131 have been completed, as well as a Phase II trial
for stroke prevention in atrial fibrillation132, and two
Phase III trials are now under way. One is designed to
compare two different doses of edoxaban with warfarin
for stroke prevention in patients with atrial fibrillation
(NCT00781391) and another is evaluating edoxaban for
the secondary prevention of recurrent VTE in patients
with acute symptomatic proximal DVT or pulmonary
embolism (NCT00986154).
YM150. YM150 (developed by Astellas) is also an oral,
direct factor Xa inhibitor (Ki of 31 nM). Preliminary
data show that both YM150 and its major metabolite
YM-222714 (Ki of 20 nM) have antithrombotic effects
at doses that do not prolong template bleeding time in
animal models of venous and arterial thrombosis133.
Phase II trials for VTE prevention after THR134 or TKR
(NCT00408239) have been completed, as has a warfarin-
controlled Phase II trial for stroke prevention in patients
with atrial fibrillation (NCT00448214). Three Phase III
trials are now in progress. One will assess once-daily and
twice-daily doses of YM150 against enoxaparin for VTE
prevention in patients undergoing elective hip replace-
ment (NCT00902928). In another Phase III trial, YM150
is being compared with mechanical prophylaxis for
the prevention of VTE after major abdominal surgery
(NCT00942435). A third study is being conducted in
Japan to evaluate YM150 for the prevention of VTE in
the 28 days following hospitalization for an acute medical
illness (NCT01028950).
Outlook
For more than 65 years, VKAs have been the only avail-
able oral anticoagulants, and although effective, the need
for dose adjustment and periodic coagulation monitoring
considerably complicates their clinical management. New
and improved anticoagulants could potentially address
the shortcomings associated with the current standard of
care and it should be noted that other approaches, besides
the development of oral, direct factor Xa inhibitors, are
also being evaluated. For example, the parenteral direct
factor Xa inhibitor, otamixaban (developed by Sanofi–
Aventis), has completed Phase II trials for short-term use
in non-ST-elevation ACS135 and in percutaneous coronary
intervention136. The encouraging results obtained in these
studies have been followed by an ongoing Phase III trial in
patients with unstable angina or non-ST elevation myo-
cardial infarction undergoing an invasive intervention
(NCT01076764). Conversely, idraparinux, and its succes-
sor idrabiotaparinux137 (Sanofi–Aventis; both now discon-
tinued) are parenterally administered pentasaccharides
(indirect factor Xa inhibitors) with very long half-lives138
that were developed to permit once-weekly dosing in
indications requiring long-term or chronic therapy139,140.
In addition, direct thrombin inhibitors are also in
development, most notably dabigatran (Pradaxa/Pradax;
Boehringer Ingelheim), which, like rivaroxaban, has
completed several Phase III thrombosis prevention and
treatment studies141–144 and has been approved in mem-
ber states of the European Union and other countries for
the prevention of VTE after elective hip or knee replace-
ment. Dabigatran is also in trials for a number of other
indications, as is AZD0837 (developed by AstraZeneca),
which is in Phase II145. Although indirect comparisons,
based on meta-analyses, can be conducted, direct com-
parative trials are required for a comprehensive evalua-
tion of one particular drug regimen versus another.
REVIEWS
72
|
JANUARY 2011
|
VOLUME 10 www.nature.com/reviews/drugdisc
© 2011 Macmillan Publishers Limited. All rights reserved
In the future, it is anticipated that long-term antico-
agulant therapy will favour oral agents with a wide thera-
peutic window and a predictable anticoagulant response
that do not require routine coagulation monitoring or
dose adjustment. Emerging data suggest that direct fac-
tor Xa inhibitors are both effective antithrombotic agents
for short-term use and promising agents for long-term
usage. As they have shown a predictable pharmacological
profile, are given orally and do not require routine
coagulation monitoring, these new agents may also
facilitate patient compliance and adherence to clinical
guidelines. Thus, they are likely to improve anticoagula-
tion treatment in thromboembolic disorders and reduce
the burden associated with long-term therapy, thereby
providing important alternative options for the manage-
ment of these conditions.
1. McLean, J. The thromboplastic action of cephalin.
Am. J. Physiol. 41, 250–257 (1916).
2. Hirsh, J. et al. Parenteral anticoagulants: American
College of Chest Physicians evidence-based clinical
practice guidelines (8th Edition). Chest 133,
141S–159S (2008).
3. Warkentin, T. E., Greinacher, A., Koster, A., Lincoff,
A. M. & American College of Chest Physicians.
Treatment and prevention of heparin-induced
thrombocytopenia: American College of Chest
Physicians evidence-based clinical practice guidelines
(8th Edition). Chest 133, 340S–380S (2008).
4. Warwick, D., Dahl, O. E. & Fisher, W. D. Orthopaedic
thromboprophylaxis: limitations of current guidelines.
J. Bone Joint Surg. Br. 90, 127–132 (2008).
5. Campbell, H. A. & Link, K. P. Studies on the
hemorrhagic sweet clover disease. IV. The isolation
and crystallization of the hemorrhagic agent. J. Biol.
Chem. 138, 21–33 (1941).
6. Ansell, J. et al. Pharmacology and management of the
vitamin K antagonists: American College of Chest
Physicians evidence-based clinical practice guidelines
(8th Edition). Chest 133, 160S–198S (2008).
7. Schwarz, U. I. et al. Genetic determinants of response
to warfarin during initial anticoagulation. N. Engl.
J. Med. 358, 999–1008 (2008).
8. Klein, T. E. et al. Estimation of the warfarin dose with
clinical and pharmacogenetic data. N. Engl. J. Med.
360, 753–764 (2009).
9. Hirsh, J., O’Donnell, M. & Weitz, J. I. New
anticoagulants. Blood 105, 453–463 (2005).
10. Weitz, J. I. Emerging anticoagulants for the treatment
of venous thromboembolism. Thromb. Haemost. 96,
274–284 (2006).
11. Dyke, C. K. et al. First-in-human experience of an
antidote-controlled anticoagulant using RNA aptamer
technology: a phase 1a pharmacodynamic evaluation
of a drug-antidote pair for the controlled regulation of
factor IXa activity. Circulation 114 , 2490–2497
(2006).
12. Eriksson, B. I. et al. Partial factor IXa inhibition with
TTP889 for prevention of venous thromboembolism:
an exploratory study. J. Thromb. Haemost. 6,
457–463 (2008).
13. Goto, S. Factor XIa as a possible new target of
antithrombotic therapy. J. Thromb. Haemost. 4,
1494–1495 (2006).
14. Schumacher, W. A. et al. Antithrombotic and hemostatic
effects of a small molecule factor XIa inhibitor in rats.
Eur. J. Pharmacol. 570, 167–174 (2007).
15. Ho, S. J. & Brighton, T. A. Ximelagatran: direct
thrombin inhibitor. Vasc. Health Risk Manag. 2,
49–58 (2006).
16. Nishio, H., Ieko, M. & Nakabayashi, T. New
therapeutic option for thromboembolism —
dabigatran etexilate. Expert Opin. Pharmacother. 9,
2509–2517 (2008).
17. Turpie, A. G. G., Eriksson, B. I., Lassen, M. R. & Bauer,
K. A. Fondaparinux, the first selective factor Xa
inhibitor. Curr. Opin. Hematol. 10, 327–332 (2003).
18. Carreiro, J. & Ansell, J. Apixaban, an oral direct factor
Xa inhibitor: awaiting the verdict. Expert Opin.
Investig. Drugs 17, 1937–1945 (2008).
19. Piccini, J. P., Patel, M. R., Mahaffey, K. W., Fox, K. A.
& Califf, R. M. Rivaroxaban, an oral direct factor Xa
inhibitor. Expert Opin. Investig. Drugs 17, 925–937
(2008).
20. Leadley, R. J. Coagulation factor Xa inhibition:
biological background and rationale. Curr. Top. Med.
Chem. 1, 151–159 (2001).
This comprehensive review describes some of the
key studies that generated interest in factor Xa as
a therapeutic target and highlights the significant
findings that provided the rationale for the
development of direct factor Xa inhibitors.
21. Kearon, C. et al. Dose-response study of recombinant
human soluble thrombomodulin (ART-123) in the
prevention of venous thromboembolism after total hip
replacement. J. Thromb. Haemost. 3, 962–968
(2005).
22. Saito, H. et al. Efficacy and safety of recombinant
human soluble thrombomodulin (ART-123) in
disseminated intravascular coagulation: results of a
phase III, randomized, double-blind clinical trial.
J. Thromb. Haemost. 5, 31–41 (2007).
23. Bernard, G. R. et al. Efficacy and safety of
recombinant human activated protein C for severe
sepsis. N. Engl. J. Med. 344, 699–709 (2001).
24. Levi, M. & van der Poll, T. Recombinant human
activated protein C: current insights into its
mechanism of action. Crit. Care 11 (Suppl. 5), S3
(2007).
25. Ansell, J. Factor Xa or thrombin: is factor Xa a better
target? J. Thromb. Haemost. 5 (Suppl. 1), 60–64
(2007).
26. Gerotziafas, G. T. & Samama, M. M. Heterogeneity of
synthetic factor Xa inhibitors. Curr. Pharm. Des. 11,
3855–3876 (2005).
27. Turpie, A. G. G. Oral, direct factor Xa inhibitors in
development for the prevention and treatment of
thromboembolic diseases. Arterioscler. Thromb. Vasc.
Biol. 27, 1238–1247 (2007).
28. Koller, F. History of factor X. Thromb. Diath.
Haemorrh. 4, 58–65 (1960).
29. Brown, D. L. & Kouides, P. A. Diagnosis and treatment
of inherited factor X deficiency. Haemophilia 14,
1176–1182 (2008).
30. Hougie, C., Barrow, H. M. & Graham, J. B.
Stuart clotting defect. I. Segregation of a hereditary
hemorrhagic state from the heterozygous
heretofore called ‘stable factor’ (SPCA, proconvertin
factor VII deficiency). J. Clin. Invest. 36, 485–493
(1957).
31. Telfer, T. P., Denson, K. W. & Wright, D. R. A ‘new’
coagulation defect. Br. J. Haematol. 2, 308–316
(1956).
32. Butenas, S., van’t Veer, C. & Mann, K. G. “Normal”
thrombin generation. Blood 94, 2169–2178 (1999).
33. Nutt, E. et al. The amino acid sequence of antistasin.
A potent inhibitor of factor Xa reveals a repeated
internal structure. J. Biol. Chem. 263, 10162–10167
(1988).
34. Tuszynski, G. P., Gasic, T. B. & Gasic, G. J. Isolation
and characterization of antistasin. An inhibitor of
metastasis and coagulation. J. Biol. Chem. 262,
9718–9723 (1987).
35. Dunwiddie, C. et al. Antistasin, a leech-derived
inhibitor of factor Xa. Kinetic analysis of enzyme
inhibition and identification of the reactive site.
J. Biol. Chem. 264, 16694–16699 (1989).
36. Waxman, L., Smith, D. E., Arcuri, K. E. & Vlasuk, G. P.
Tick anticoagulant peptide (TAP) is a novel inhibitor of
blood coagulation factor Xa. Science 248, 593–596
(1990).
37. Nicolini, F. A. et al. Selective inhibition of factor Xa
during thrombolytic therapy markedly improves
coronary artery patency in a canine model of coronary
thrombosis. Blood Coagul. Fibrinolysis 7, 39–48
(1996).
38. Vlasuk, G. P. et al. Comparison of the in vivo
anticoagulant properties of standard heparin and the
highly selective factor Xa inhibitors antistasin and tick
anticoagulant peptide (TAP) in a rabbit model of
venous thrombosis. Thromb. Haemost. 65, 257–262
(1991).
39. Lynch, J. J. et al. Maintenance of canine coronary
artery patency following thrombolysis with front
loaded plus low dose maintenance conjunctive
therapy. A comparison of factor Xa versus thrombin
inhibition. Cardiovasc. Res. 28, 78–85 (1994).
40. Sitko, G. R. et al. Conjunctive enhancement of
enzymatic thrombolysis and prevention of thrombotic
reocclusion with the selective factor Xa inhibitor,
tick anticoagulant peptide. Comparison to hirudin
and heparin in a canine model of acute coronary
artery thrombosis. Circulation 85, 805–815
(1992).
41. Lefkovits, J. et al. Selective inhibition of factor Xa is
more efficient than factor VIIa-tissue factor complex
blockade at facilitating coronary thrombolysis in the
canine model. J. Am. Coll. Cardiol. 28, 1858–1865
(1996).
42. Eisenberg, P. R., Siegel, J. E., Abendschein, D. R. &
Miletich, J. P. Importance of factor Xa in determining
the procoagulant activity of whole-blood clots. J. Clin.
Invest. 91, 1877–1883 (1993).
43. Prager, N. A., Abendschein, D. R., McKenzie, C. R. &
Eisenberg, P. R. Role of thrombin compared with
factor Xa in the procoagulant activity of whole blood
clots. Circulation 92, 962–967 (1995).
44. McKenzie, C. R., Abendschein, D. R. & Eisenberg, P. R.
Sustained inhibition of whole-blood clot procoagulant
activity by inhibition of thrombus-associated factor Xa.
Arterioscler. Thromb. Vasc. Biol. 16, 1285–1291
(1996).
45. Kotze, H. F., Lamprecht, S., Badenhorst, P. N.,
Roodt, J. P. & van Wyk, V. Transient interruption of
arterial thrombosis by inhibition of factor Xa results in
long-term antithrombotic effects in baboons. Thromb.
Haemost. 77, 1137–1142 (1997).
46. Gold, H. K. et al. Evidence for a rebound coagulation
phenomenon after cessation of a 4-hour infusion of a
specific thrombin inhibitor in patients with unstable
angina pectoris. J. Am. Coll. Cardiol. 21, 1039–1047
(1993).
47. Theroux, P., Waters, D., Lam, J., Juneau, M. &
McCans, J. Reactivation of unstable angina after the
discontinuation of heparin. N. Engl. J. Med. 327,
141–145 (1992).
48. Hermans, C. & Claeys, D. Review of the rebound
phenomenon in new anticoagulant treatments. Curr.
Med. Res. Opin. 22, 471–481 (2006).
49. Furugohri, T. et al. Different antithrombotic properties
of factor Xa inhibitor and thrombin inhibitor in rat
thrombosis models. Eur. J. Pharmacol. 514, 35–42
(2005).
50. Becker, R. C., Alexander, J., Dyke, C. K. & Harrington,
R. A. Development of DX-9065a, a novel direct factor
Xa antagonist, in cardiovascular disease. Thromb.
Haemost. 92, 1182–1193 (2004).
51. Sato, K. et al. YM-60828, a novel factor Xa inhibitor:
separation of its antithrombotic effects from its
prolongation of bleeding time. Eur. J. Pharmacol.
339, 141–146 (1997).
52. Taniuchi, Y. et al. Biochemical and pharmacological
characterization of YM-60828, a newly synthesized
and orally active inhibitor of human factor Xa. Thromb.
Haemost. 79, 543–548 (1998).
53. Hara, T. et al. DX-9065a, a new synthetic, potent
anticoagulant and selective inhibitor for factor Xa.
Thromb. Haemost. 71, 314–319 (1994).
54. Fujii, Y. et al. Characteristics of gastrointestinal
absorption of DX-9065a, a new synthetic
anticoagulant. Drug Metab. Pharmacokinet. 22,
26–32 (2007).
55. Alexander, J. H. et al. First experience with direct,
selective factor Xa inhibition in patients with
non-ST-elevation acute coronary syndromes: results
of the XaNADU-ACS Trial. J. Thromb. Haemost. 3,
439–447 (2005).
56. Bauer, K. A. et al. Fondaparinux, a synthetic
pentasaccharide: the first in a new class of
antithrombotic agents — the selective factor Xa
inhibitors. Cardiovasc. Drug Rev. 20, 37–52
(2002).
REVIEWS
NATURE R EVIEWS
|
DRUG DISCOVERY VOLUME 10
|
JANUARY 2011
|
73
© 2011 Macmillan Publishers Limited. All rights reserved
57. Bergqvist, D. Review of fondaparinux sodium injection
for the prevention of venous thromboembolism in
patients undergoing surgery. Vasc. Health Risk
Manag. 2, 365–370 (2006).
58. Eriksson, B. I., Bauer, K. A., Lassen, M. R., Turpie,
A. G. G. & Steering Committee of the Pentasaccharide
in Hip-Fracture Surgery Study. Fondaparinux
compared with enoxaparin for the prevention of
venous thromboembolism after hip-fracture surgery.
N. Engl. J. Med. 345, 1298–1304 (2001).
59. Lassen, M. R., Bauer, K. A., Eriksson, B. I. & Turpie,
A. G. G. Postoperative fondaparinux versus
preoperative enoxaparin for prevention of venous
thromboembolism in elective hip-replacement
surgery: a randomised double-blind comparison.
Lancet 359, 1715–1720 (2002).
60. Turpie, A. G. G., Bauer, K. A., Eriksson, B. I., Lassen,
M. R. & PENTATHLON 2000 Study Steering
Committee. Postoperative fondaparinux versus
postoperative enoxaparin for prevention of venous
thromboembolism after elective hip-replacement
surgery: a randomised double-blind trial. Lancet 359,
1721–1726 (2002).
61. Bauer, K. A., Eriksson, B. I., Lassen, M. R. & Turpie,
A. G. G. Fondaparinux compared with enoxaparin for
the prevention of venous thromboembolism after
elective major knee surgery. N. Engl. J. Med. 345,
1305–1310 (2001).
62. Turpie, A. G. G., Bauer, K. A., Eriksson, B. I., Lassen,
M. R. & Steering Committees of the Pentasaccharide
Orthopedic Prophylaxis Studies. Fondaparinux vs
enoxaparin for the prevention of venous
thromboembolism in major orthopedic surgery: a
meta-analysis of 4 randomized double-blind studies.
Arch. Intern. Med. 162, 1833–1840 (2002).
63. Roehrig, S. et al. Discovery of the novel
antithrombotic agent 5-chloro-N-([(5S)-2-oxo-3-
[4-(3-oxomorpholin-4-yl)phenyl]-1,3-oxazolidin-5-yl]
methyl)thiophene-2-carboxamide (BAY 59-7939): an
oral, direct factor Xa inhibitor. J. Med. Chem. 48,
5900–5908 (2005).
Describes the chemical development of
rivaroxaban, including the important
structure–activity relationship that led to the
combination of high binding affinity with high oral
bioavailability.
64. Lam, P. Y. et al. Structure-based design of novel
guanidine/benzamidine mimics: potent and orally
bioavailable factor Xa inhibitors as novel
anticoagulants. J. Med. Chem. 46, 4405–4418
(2003).
65. Pinto, D. J. et al. Discovery of 1-[3-(aminomethyl)
phenyl]-N-3-fluoro-2-(methylsulfonyl)-[1,1-biphenyl]-4-
yl]-3-(trifluoromethyl)-1H-pyrazole-5-carboxamide
(DPC423), a highly potent, selective, and orally
bioavailable inhibitor of blood coagulation factor Xa.
J. Med. Chem. 44, 566–578 (2001).
66. Adler, M. et al. Crystal structures of two potent
nonamidine inhibitors bound to factor Xa. Biochem.
Mosc. 41, 15514–15523 (2002).
67. Choi-Sledeski, Y. M. et al. Discovery of an orally
efficacious inhibitor of coagulation factor Xa which
incorporates a neutral P1 ligand. J. Med. Chem. 46,
681–684 (2003).
68. Maignan, S. et al. Molecular structures of human
factor Xa complexed with ketopiperazine inhibitors:
preference for a neutral group in the S1 pocket.
J. Med. Chem. 46, 685–690 (2003).
69. Mederski, W. W. et al. Chlorothiophenecarboxamides
as P1 surrogates of inhibitors of blood coagulation
factor Xa. Bioorg. Med. Chem. Lett. 14, 5817–5822
(2004).
70. Nazare, M. et al. Probing the subpockets of factor Xa
reveals two binding modes for inhibitors based on a
2-carboxyindole scaffold: a study combining structure–
activity relationship and X-ray crystallography. J. Med.
Chem. 48, 4511–4525 (2005).
71. Perzborn, E. et al. In vitro and in vivo studies of the
novel antithrombotic agent BAY 59-7939 — an oral,
direct factor Xa inhibitor. J. Thromb. Haemost. 3,
514–521 (2005).
This paper reports the in vitro properties of
rivaroxaban, its antithrombotic efficacy in animal
models of arterial and venous thrombosis and its
effect on haemostasis — the key findings that
established the rationale for its clinical
development.
72. Depasse, F. et al. Effect of BAY 59-7939 — a novel,
oral, direct factor Xa inhibitor — on clot-bound factor
Xa activity in vitro. J. Thromb. Haemost. Abstr. 3
(Suppl. 1), P1104 (2005).
73. Gerotziafas, G. T., Elalamy, I., Depasse, F., Perzborn, E.
& Samama, M. M. In vitro inhibition of thrombin
generation, after tissue factor pathway activation, by
the oral, direct factor Xa inhibitor rivaroxaban.
J. Thromb. Haemost. 5, 886–888 (2007).
74. Graff, J. et al. Effects of the oral, direct factor Xa
inhibitor rivaroxaban on platelet-induced thrombin
generation and prothrombinase activity. J. Clin.
Pharmacol. 47, 1398–1407 (2007).
75. Perzborn, E. & Harwardt, M. Direct thrombin inhibitors,
but not factor Xa inhibitors, enhance thrombin formation
in human plasma by interfering with the thrombin-
thrombomodulin-protein C system. Blood (ASH Annual
Meeting Abstracts) 112 , 198–199 (2008).
76. Perzborn, E. et al. Biochemical and pharmacologic
properties of BAY 59–7939, an oral, direct factor Xa
inhibitor. Pathophysiol. Haemost. Thromb. 33 (Suppl.
2), PO079 (2004).
77. Perzborn, E. & Lange, U. Rivaroxaban — an oral,
direct factor Xa inhibitor — inhibits tissue factor-
mediated platelet aggregation. J. Thromb. Haemost. 5
(Suppl. 2), P-W-642 (2007).
78. Furugohri, T. et al. DU-176b, a potent and orally
active factor Xa inhibitor: in vitro and in vivo
pharmacological profiles. J. Thromb. Haemost. 6,
1542–1549 (2008).
79. Shantsila, E. & Lip, G. Y. Apixaban, an oral, direct
inhibitor of activated factor Xa. Curr. Opin. Investig.
Drugs 9, 1020–1033 (2008).
80. Biemond, B. J. et al. Prevention and treatment of
experimental thrombosis in rabbits with rivaroxaban
(BAY 59-7939) — an oral, direct factor Xa inhibitor.
Thromb. Haemost. 97, 471–477 (2007).
81. Kubitza, D., Becka, M., Roth, A. & Mueck, W.
Dose-escalation study of the pharmacokinetics and
pharmacodynamics of rivaroxaban in healthy elderly
subjects. Curr. Med. Res. Opin. 24, 2757–2765
(2008).
82. Kubitza, D., Becka, M., Wensing, G., Voith, B. &
Zuehlsdorf, M. Safety, pharmacodynamics, and
pharmacokinetics of BAY 59-7939 — an oral, direct
factor Xa inhibitor — after multiple dosing in healthy
male subjects. Eur. J. Clin. Pharmacol. 61, 873–880
(2005).
This Phase I study reported the safety, tolerability,
pharmacodynamics and pharmacokinetics of
rivaroxaban when administered as multiple doses
to healthy male subjects. The favourable results of
this study, and others, supported the progression
to Phase II clinical trials.
83. Gruber, A., Marzec, U. M., Buetehorn, U., Hanson, S.
& Perzborn, E. Potential of activated prothrombin
complex concentrate and activated factor VII to
reverse the anticoagulant effects of rivaroxaban in
primates. Blood (ASH Annual Meeting Abstracts) 112 ,
1307 (2008).
84. Perzborn, E., Trabandt, A., Selbach, K. & Tinel, H.
Prothrombin complex concentrate reverses the effects
of high-dose rivaroxaban in rats. J. Thromb. Haemost.
7 (Suppl. 2), 379 (2009).
85. Kubitza, D., Becka, M., Voith, B., Zuehlsdorf, M. &
Wensing, G. Safety, pharmacodynamics, and
pharmacokinetics of single doses of BAY 59-7939, an
oral, direct factor Xa inhibitor. Clin. Pharmacol. Ther.
78, 412–421 (2005).
86. Zhang, L., Reynolds, K. S., Zhao P. & Huang. S. M.
Drug interactions evaluation: an integrated part of
risk assessment of therapeutics. Toxicol. Appl.
Pharmacol. 243, 134–145 (2010).
87. Zhang, L., Zhang, Y. D., Strong, J. M., Reynolds, K. S.
& Huang, S. M. A regulatory viewpoint on transporter-
based drug interactions. Xenobiotica 38, 709–724
(2008).
88. Kubitza, D., Becka, M., Mueck, W. & Zuehlsdorf, M.
Rivaroxaban (BAY 59-7939) — an oral, direct factor
Xa inhibitor — has no clinically relevant interaction
with naproxen. Br. J. Clin. Pharmacol. 63, 469–476
(2007).
89. Kubitza, D., Becka, M., Mueck, W. & Zuehlsdorf, M.
Safety, tolerability, pharmacodynamics, and
pharmacokinetics of rivaroxaban — an oral, direct
factor Xa inhibitor — are not affected by aspirin.
J. Clin. Pharmacol. 46, 981–990 (2006).
90. Eriksson, B. I. et al. Rivaroxaban versus enoxaparin
for thromboprophylaxis after hip arthroplasty. N. Engl.
J. Med. 358, 2765–2775 (2008).
91. Kakkar, A. K. et al. Extended duration rivaroxaban
versus short-term enoxaparin for the prevention of
venous thromboembolism after total hip arthroplasty:
a double-blind, randomised controlled trial. Lancet
372, 31–39 (2008).
92. Lassen, M. R. et al. Rivaroxaban versus enoxaparin for
thromboprophylaxis after total knee arthroplasty.
N. Engl. J. Med. 358, 2776–2786 (2008).
93. Turpie, A. G. G. et al. Rivaroxaban versus enoxaparin
for thromboprophylaxis after total knee arthroplasty
(RECORD4): a randomised trial. Lancet 373,
1673–1680 (2009).
94. Weinz, C. et al. Pharmacokinetics of BAY 59-7939 —
an oral, direct factor Xa inhibitor — in rats and dogs.
Xenobiotica 35, 891–910 (2005).
This preclinical study established some of the key
pharmacokinetic characteristics of rivaroxaban,
including a fully reversible, high level of plasma
protein binding, a dual mode of excretion and
moderate tissue affinity.
95. Lang, D., Freudenberger, C. & Weinz, C. In vitro
metabolism of rivaroxaban — an oral, direct factor Xa
inhibitor — in liver microsomes and hepatocytes
of rat, dog and man. Drug Metab. Dispos. 37,
1046–1055 (2009).
96. Weinz, C., Schwarz, T., Kubitza, D., Mueck, W. &
Lang, D. Metabolism and excretion of rivaroxaban,
an oral, direct factor Xa inhibitor, in rats, dogs and
humans. Drug Metab. Dispos. 37, 1056–1064
(2009).
97. Mueck, W. et al. Population pharmacokinetics and
pharmacodynamics of rivaroxaban — an oral, direct
factor Xa inhibitor — in patients undergoing major
orthopaedic surgery. Clin. Pharmacokinet. 47,
203–216 (2008).
The results showed that rivaroxaban has
predictable, dose-dependent pharmacokinetics and
pharmacodynamics and indicated that rivaroxaban
could be administered in fixed doses without
routine coagulation monitoring.
98. Mueck, W., Becka, M., Kubitza, D., Voith, B. &
Zuehlsdorf, M. Population model of the
pharmacokinetics and pharmacodynamics of
rivaroxaban — an oral, direct factor Xa inhibitor — in
healthy subjects. Int. J. Clin. Pharmacol. Ther. 45,
335–344 (2007).
99. Kubitza, D., Becka, M., Zuehlsdorf, M. & Mueck, W.
Body weight has limited influence on the safety,
tolerability, pharmacokinetics, or pharmacodynamics
of rivaroxaban (BAY 59-7939) in healthy subjects.
J. Clin. Pharmacol. 47, 218–226 (2007).
100. Kubitza, D. et al. Effects of renal impairment on the
pharmacokinetics, pharmacodynamics and safety of
rivaroxaban — an oral, direct factor Xa inhibitor. Br.
J. Clin. Pharmacol. 70, 703–712 (2010).
101. Halabi, A. et al. Effect of hepatic impairment
on the pharmacokinetics, pharmacodynamics and
tolerability of rivaroxaban — an oral, direct factor Xa
inhibitor. J. Thromb. Haemost. 5 (Suppl. 2), P-M-635
(2007).
102. Eriksson, B. I. et al. Dose-escalation study of
rivaroxaban (BAY 59-7939) — an oral, direct factor Xa
inhibitor — for the prevention of venous
thromboembolism in patients undergoing total hip
replacement. Thromb. Res. 120, 685–693 (2007).
103. Eriksson, B. I. et al. A once-daily, oral, direct factor Xa
inhibitor, rivaroxaban (BAY 59-7939), for
thromboprophylaxis after total hip replacement.
Circulation 114 , 2374–2381 (2006).
The results from this study, and others, led to the
selection of 10 mg rivaroxaban once daily for use in
Phase III studies for the prevention of venous
thromboembolism after hip or knee replacement.
104. Eriksson, B. I. et al. Oral, direct factor Xa inhibition
with BAY 59-7939 for the prevention of venous
thromboembolism after total hip replacement.
J. Thromb. Haemost. 4, 121–128 (2006).
105. Turpie, A. G. G. et al. BAY 59-7939: an oral, direct
factor Xa inhibitor for the prevention of venous
thromboembolism in patients after total knee
replacement. A phase II dose-ranging study.
J. Thromb. Haemost. 3, 2479–2486 (2005).
106. Kubitza, D., Mueck, W. & Becka, M. Randomized,
double-blind, crossover study to investigate the effect
of rivaroxaban on QT-interval prolongation. Drug Saf.
31, 67–77 (2008).
107. Samama, M. M. et al. Suitability of chromogenic anti-
FXa methods to measure rivaroxaban in human
plasma. J. Thromb. Haemost. 7 (Suppl. 2), 693
(2009).
108. Karst, A., Bakowski-Enzian, B. & Perzborn, E.
Monitoring of rivaroxaban: suitability of a well-
established chromogenic anti-factor Xa assay.
J. Thromb. Haemost. 7 (Suppl. 2), 372 (2009).
109. Fisher, W. D. et al. Rivaroxaban for
thromboprophylaxis after orthopaedic surgery:
REVIEWS
74
|
JANUARY 2011
|
VOLUME 10 www.nature.com/reviews/drugdisc
© 2011 Macmillan Publishers Limited. All rights reserved
pooled analysis of two studies. Thromb. Haemost. 97,
931–937 (2007).
110. Geerts, W. H. et al. Prevention of venous
thromboembolism: American College of Chest
Physicians evidence-based clinical practice guidelines
(8th Edition). Chest 133, 381S–453S (2008).
111. Lassen, M. R. et al. Apixaban or enoxaparin for
thromboprophylaxis after knee replacement. N. Engl.
J. Med. 361, 594–604 (2009).
112. The RE-MOBILIZE Writing Committee. Oral thrombin
inhibitor dabigatran etexilate vs North American
enoxaparin regimen for prevention of venous
thromboembolism after knee arthroplasty surgery.
J. Arthroplasty 24, 1–9 (2009).
113. Eriksson, B. I. et al. Oral rivaroxaban for the
prevention of symptomatic venous thromboembolism
after elective hip and knee replacement. J. Bone Joint
Surg. Br. 91, 636–644 (2009).
A pooled analysis of data from the pivotal Phase III
clinical trials that, together with other publications,
demonstrated the clinical efficacy and safety of
rivaroxaban for thromboprophylaxis after total
knee- or hip-replacement surgery.
114. Mega, J. L. et al. Rivaroxaban versus placebo in
patients with acute coronary syndromes (ATLAS ACS-
TIMI 46): a randomised, double-blind, phase II trial.
Lancet 374, 29–38 (2009).
A Phase II dose-ranging study investigating the use
of rivaroxaban for the secondary prevention of
ischaemic events in patients with ACS receiving
aspirin with or without a thienopyridine.
115. Bauer, K. A., Homering, M. & Berkowitz, S. D. Effects
of age, weight, gender and renal function in a pooled
analysis of four phase III studies of rivaroxaban for
prevention of venous thromboembolism after major
orthopedic surgery. Blood (ASH Annual Meeting
Abstracts) 112 , 166–167 (2008).
116. Agnelli, G. et al. Treatment of proximal deep-vein
thrombosis with the oral direct factor Xa inhibitor
rivaroxaban (BAY 59-7939): the ODIXa-DVT (oral
direct factor Xa inhibitor BAY 59–7939 in patients
with acute symptomatic deep-vein thrombosis) study.
Circulation 116 , 180–187 (2007).
117. Buller, H. R. et al. A dose-ranging study evaluating
once-daily oral administration of the factor Xa
inhibitor rivaroxaban in the treatment of patients with
acute symptomatic deep vein thrombosis. The
EINSTEIN-DVT Dose-Ranging Study. Blood 11 2,
2242–2247 (2008).
This Phase II study, with one other, provided the
basis for the dosing strategy in the Phase III VTE
treatment trials and the dose selection for the
Phase III ROCKET AF trial.
118. Buller, H.R. Oral rivaroxaban for the acute and
continued treatment of symptomatic venous
thromboembolism. The Einstein-DVT and Einstein-
Extension Study (Abstract 187). ASH Annual Meeting
Abstracts [online], http://ash.confex.com/ash/2010/
webprogram/Paper32683.html (2010).
119. The EINSTEIN Investigators. Oral rivaroxaban
for symptomatic venous thromboembolism. New
Engl. J. Med. 4 Dec 2010 (doi:10.1056/
NEJMoa1007903).
120. ROCKET AF Study Investigators. Rivaroxaban — once
daily, oral, direct factor Xa inhibition compared with
vitamin K antagonism for prevention of stroke and
embolism trial in atrial fibrillation: rationale and
design of the ROCKET AF study. Am. Heart J. 159,
340–347 (2010).
121. Wong, P. C. et al. Apixaban, an oral, direct and highly
selective factor Xa inhibitor: in vitro, antithrombotic
and antihemostatic studies. J. Thromb. Haemost. 6,
820–829 (2008).
122. Luettgen, J. M. et al. Inhibition of measured thrombin
generation in human plasma by apixaban: a predictive
mathematical model based on experimentally
determined rate constants. J. Thromb. Haemost. 5
(Suppl. 1), P-T-633 (2007).
123. Lassen, M. R. et al. Apixaban versus enoxaparin for
thromboprophylaxis after knee replacement
(ADVANCE-2): a randomised double-blind trial. Lancet
375, 807–815 (2010).
124. Lassen, M. R. et al. Randomized double-blind
comparison of apixaban and enoxaparin for
thromboprophylaxis after hip replacement: the
ADVANCE-3 trial. Pathophysiol. Haemost. Thromb.
37, A20 (2010).
125. Alexander, J. H. et al. Apixaban, an oral, direct,
selective factor Xa inhibitor, in combination with
antiplatelet therapy after acute coronary syndrome:
results of the Apixaban for Prevention of Acute
Ischemic and Safety Events (APPRAISE) trial.
Circulation 119 , 2877–2885 (2009).
126. Eikelboom, J. W. et al. Rationale and design of
AVERROES: apixaban versus acetylsalicylic acid to
prevent stroke in atrial fibrillation patients who have
failed or are unsuitable for vitamin K antagonist
treatment. Am. Heart J. 159, 348–353 (2010).
127. Lopes, R. D. et al. Apixaban for Reduction In Stroke
and Other ThromboemboLic Events in Atrial
Fibrillation (ARISTOTLE) trial: design and rationale.
Am. Heart J. 159, 331–339 (2010).
128. Connolly, S. J. & Arnesen, H. AVERROES: apixaban
versus acetylsalicylic acid to prevent strokes.
European Society of Cardiology website [online]
http://www.escardio.org/congresses/esc-2010/
congress-reports/Documents/Tuesday/AVERROES-
discussant-slides-Connolly.pdf (2010).
129. Zafar, M. U. et al. Antithrombotic effects of factor Xa
inhibition with DU-176b: phase-I study of an oral,
direct factor Xa inhibitor using an ex‑vivo flow
chamber. Thromb. Haemost. 98, 883–888 (2007).
130. Fuji, T., Fujita, S., Tachibana, S. & Kawai, Y. A dose-
ranging study evaluating the oral factor Xa inhibitor
edoxaban for the prevention of venous
thromboembolism in patients undergoing total knee
arthroplasty. J. Thromb. Haemost. 8, 2458–2468
(2010).
131. Raskob, G. et al. Oral direct factor Xa inhibition with
edoxaban for thromboprophylaxis after elective total
hip replacement. A randomised double-blind dose-
response study. Thromb. Haemost. 104, 642–649
(2010).
132. Weitz, J. I. et al. Randomised, parallel-group,
multicentre, multinational phase 2 study comparing
edoxaban, an oral factor Xa inhibitor, with warfarin for
stroke prevention in patients with atrial fibrillation.
Thromb. Haemost. 104, 633–641 (2010).
133. Iwatsuki, Y. et al. Biochemical and pharmacological
profiles of YM150, an oral direct factor Xa inhibitor.
Blood (ASH Annual Meeting Abstracts) 108, 273a
(2006).
134. Eriksson, B. I. et al. Prevention of venous
thromboembolism with an oral factor Xa inhibitor,
YM150, after total hip arthroplasty. A dose finding
study (ONYX-2). J. Thromb. Haemost. 8, 714–721
(2010).
135. Sabatine, M. S. et al. Otamixaban for the treatment of
patients with non-ST-elevation acute coronary
syndromes (SEPIA-ACS1 TIMI 42): a randomised,
double-blind, active-controlled, phase 2 trial. Lancet
374, 787–795 (2009).
136. Cohen, M. et al. Randomized, double-blind, dose-
ranging study of otamixaban, a novel, parenteral,
short-acting direct factor Xa inhibitor, in percutaneous
coronary intervention: the SEPIA-PCI trial. Circulation
115 , 2642–2651 (2007).
137. Harenberg, J. Development of idraparinux and
idrabiotaparinux for anticoagulant therapy. Thromb.
Haemost. 102, 811–815 (2009).
138. Veyrat-Follet, C., Vivier, N., Trellu, M., Dubruc, C. &
Sanderink, G. J. The pharmacokinetics of idraparinux,
a long-acting indirect factor Xa inhibitor: population
pharmacokinetic analysis from Phase III clinical trials.
J. Thromb. Haemost. 7, 559–565 (2009).
139. Amadeus Investigators et al. Comparison of
idraparinux with vitamin K antagonists for prevention
of thromboembolism in patients with atrial fibrillation:
a randomised, open-label, non-inferiority trial. Lancet
371, 315–321 (2008).
140. van Gogh Investigators et al. Idraparinux versus
standard therapy for venous thromboembolic disease.
N. Engl. J. Med. 357, 1094–1104 (2007).
141. Connolly, S. J. et al. Dabigatran versus warfarin in
patients with atrial fibrillation. N. Engl. J. Med. 361,
1139–1151 (2009).
142. Eriksson, B. I. et al. Dabigatran etexilate versus
enoxaparin for prevention of venous
thromboembolism after total hip replacement: a
randomised, double-blind, non-inferiority trial. Lancet
370, 949–956 (2007).
143. Eriksson, B. I. et al. Oral dabigatran etexilate vs.
subcutaneous enoxaparin for the prevention of venous
thromboembolism after total knee replacement: the
RE-MODEL randomized trial. J. Thromb. Haemost. 5,
2178–2185 (2007).
144. Schulman, S. et al. Dabigatran versus warfarin in the
treatment of acute venous thromboembolism. N. Engl.
J. Med. 361, 2342–2352 (2009).
145. Squizzato, A., Dentali, F., Steidl, L. & Ageno, W. New
direct thrombin inhibitors. Intern. Emerg. Med. 4,
479–484 (2009).
146. Straub, A. et al. Substituted oxazolidinones and their
use in the field of blood coagulation. WO0147919 (A1)
(2001).
147. Mederski, W. W. et al. Halothiophene benzimidazoles
as P1 surrogates of inhibitors of blood coagulation
factor Xa. Bioorg. Med. Chem. Lett. 14, 3763–3769
(2004).
148. Pinto, D. J. et al. Discovery of
1-(4-methoxyphenyl)-7-oxo-6-(4-(2-oxopiperidin-1-yl)
phenyl)-4,5,6,7-tetrahydro-1H-pyrazolo[3,4-c]
pyridine-3-carboxamide (Apixaban, BMS-562247), a
highly potent, selective, efficacious, and orally
bioavailable inhibitor of blood coagulation factor Xa.
J. Med. Chem. 50, 5339–5356 (2007).
149. Kohrt, J. T. et al. The discovery of (2R,4R)-N-(4-chloro
phenyl)-N-(2-fluoro-4-(2-oxopyridin-1(2H)-yl)
phenyl)-4-methoxypyrrolidine-1,2-dicarboxamide (PD
0348292), an orally efficacious factor Xa inhibitor.
Chem. Biol. Drug Des. 70, 100–112 (2007).
150. Mann, K. G., Butenas, S. & Brummel, K. The dynamics
of thrombin formation. Arterioscler. Thromb. Vasc.
Biol. 23, 17–25 (2003).
151. Furie, B. & Furie, B. C. Mechanisms of thrombus
formation. N. Engl. J. Med. 359, 938–949 (2008).
152. Mann, K. G., Jenny, R. J. & Krishnaswamy, S.
Cofactor proteins in the assembly and expression
of blood clotting enzyme complexes. Annu. Rev.
Biochem. 57, 915–956 (1988).
153. Licari, L. G. & Kovacic, J. P. Thrombin physiology
and pathophysiology. J. Vet. Emerg. Crit. Care
(San Antonio) 19, 11–22 (2009).
154. Mann, K. G., Brummel, K. & Butenas, S. What is all that
thrombin for? J. Thromb. Haemost. 1, 1504–1514
(2003).
155. Kaiser, B. Factor Xa-a promising target for drug
development. Cell. Mol. Life Sci. 59, 189–192
(2002).
156. Wienen, W., Stassen, J. M., Priepke, H.,
Ries, U. J. & Hauel, N. In‑vitro profile and ex‑vivo
anticoagulant activity of the direct thrombin
inhibitor dabigatran and its orally active prodrug,
dabigatran etexilate. Thromb. Haemost. 98,
155–162 (2007).
157. Kakkar, A. K. et al. Extended thromboprophylaxis
with rivaroxaban compared with short-term
thromboprophylaxis with enoxaparin after total hip
arthroplasty: the RECORD2 trial. Blood (ASH Annual
Meeting Abstracts) 110, 307 (2007).
Acknowledgements
The authors would like to acknowledge S. Salaria and S.
McMillan, who provided medical writing services, with
funding from Bayer HealthCare and Johnson & Johnson
Pharmaceutical Research & Development. We would also like
to thank Proteros Biostructures, Planegg-Martinsried,
Germany, for performing the X-ray crystal structure work on
rivaroxaban.
Competing interests statement
The authors declare competing financial interests: see web
version for details.
FURTHER INFORMATION
ClinicalTrials.gov: http://clinicaltrials.gov
Bayer HealthCare Xarelto (rivaroxaban) Summary of
Product Characteristics: http://www.ema.europa.eu/docs/
en_GB/document_library/EPAR_-_Product_Information/
human/000944/WC500057108.pdf
GlaxoSmithKline, Arixtra (fondaparinux) prescribing
information: http://us.gsk.com/products/assets/us_arixtra.pdf
Janssen Pharmaceutica Products, Nizoral (ketoconazole)
tablets prescribing information: https://www.ortho-mcneil.
com/ortho-mcneil/shared/pi/nizoral_tablets.pdf
Abbott Laboratories, Norvir (ritonavir) prescribing
information: http://www.rxabbott.com/pdf/norvirtab_pi.pdf
EMA CHMP Assessment Report for Xarelto (rivaroxaban):
http://www.ema.europa.eu/docs/en_GB/document_library/
EPAR_-_Public_assessment_report/human/000944/
WC500057122.pdf
Bristol-Myers Squibb press release 18 Nov 2010
(APPRAISE-2 study with investigational compound
apixaban in acute coronary syndrome discontinued):
http://www.bms.com/news/press_releases/pages/default.
aspx?RSSLink=http://www.businesswire.com/news/
bms/20101118007161/en&t=634257431381752059
Rocket AF study: http://sciencenews.myamericanheart.org/
pdfs/ROCKET_AF_pslides.pdf
ALL LINKS ARE ACTIVE IN THE ONLINE PDF
REVIEWS
NATURE R EVIEWS
|
DRUG DISCOVERY VOLUME 10
|
JANUARY 2011
|
75
© 2011 Macmillan Publishers Limited. All rights reserved
... [1][2][3][4] Rivaroxaban is one of the most used DOACs approved by the United States Food and Drugs Administration in 2011; for the past few years, it has also been frequently used in Lithuania. 5 The drug is the first orally administered inhibitor of factor Xa. 5 Rivaroxaban interacts with factor Xa, disrupting the conversion of prothrombin to thrombin. 5 Even though adverse reactions caused by rivaroxaban are relatively rare, there are now case reports of various adverse reactions caused by the drug (Table 1). ...
... [1][2][3][4] Rivaroxaban is one of the most used DOACs approved by the United States Food and Drugs Administration in 2011; for the past few years, it has also been frequently used in Lithuania. 5 The drug is the first orally administered inhibitor of factor Xa. 5 Rivaroxaban interacts with factor Xa, disrupting the conversion of prothrombin to thrombin. 5 Even though adverse reactions caused by rivaroxaban are relatively rare, there are now case reports of various adverse reactions caused by the drug (Table 1). ...
... 5 The drug is the first orally administered inhibitor of factor Xa. 5 Rivaroxaban interacts with factor Xa, disrupting the conversion of prothrombin to thrombin. 5 Even though adverse reactions caused by rivaroxaban are relatively rare, there are now case reports of various adverse reactions caused by the drug (Table 1). These include eosinophilia, urticaria, angioedema, rashes, anaphylaxis, etc. (Table 1). ...
Article
Full-text available
Here, we describe a case of anaphylaxis secondary to rivaroxaban in a 61-year-old woman 24 hours after orthopedic surgery. 10–15 minutes after ingestion of rivaroxaban and nimesulide, the patient’s palms started itching, her face and lips swelled, her face flushed, she developed shortness of breath and subsequently lost consciousness. Serum tryptase levels at the time of the anaphylactic reaction were elevated, with subsequent measurement one month later returning a value within the normal range. Dabigatran and meloxicam were identified as suitable alternative drugs by oral provocation at an allergy clinic. Even though rivaroxaban rarely causes serious allergic reactions, when prescribing it, it is important to analyze patients’ medical history for possible previously experienced drug-induced allergic reactions and to be aware of the risks of possible undesired drug interactions.
... In 2008, Rivaroxaban was approved to prevent postoperative venous thromboembolism following elective hip or knee arthroplasty in approximately 100 countries, including Canada and European Union member states. [34] Based on our observations, the most frequent and recent studies are focused on aspirin, LMWHs, and new oral anticoagulants like rivaroxaban, apixaban, edoxaban and dabigatran. Their research mainly comprises clinical randomized controlled trials, focusing on the incidence of VTE and bleeding events in patients undergoing lower limb joint arthroplasty. ...
Article
Full-text available
This study aims to visualize publications related to venous thromboembolism (VTE) and lower limb joint arthroplasty to identify research frontiers and hotspots, providing references and guidance for further research. We retrieved original articles published from 1985 to 2022 and their recorded information from the Web of Science Core Collection. The search strategy used terms related to knee or hip arthroplasty and thromboembolic events. Microsoft Excel was used to analyze the annual publications and citations of the included literature. The rest of the data were analyzed using the VOSviewer, citespace and R and produced visualizations of these collaborative networks. We retrieved 3543 original articles and the results showed an overall upward trend in annual publications. The United States of America had the most significant number of publications (Np) and collaborative links with other countries. McMaster University had the greatest Np. Papers published by Geerts WH in 2008 had the highest total link strength. Journal of Arthroplasty published the most articles on the research of VTE associated with lower limb joint arthroplasty. The latest research trend mainly involved “general anesthesia” “revision” and “tranexamic acid.” This bibliometric study revealed that the research on VTE after lower limb joint arthroplasty is developing rapidly. The United States of America leads in terms of both quantity and quality of publications, while European and Canadian institutions and authors also make significant contributions. Recent research focused on the use of tranexamic acid, anesthesia selection, and the VTE risk in revision surgeries.
... However, hepatotoxicity, on the other hand, led to ximelagatran withdrawal [178]. It was reported that DOACs have more significant predictable pharmacokinetic and pharmacodynamic properties than VKAs [179,180]. Randomized clinical studies reported that other DOACs (rivaroxaban, edoxaban, etexilate, apixaban, and dabigatran) are the best alternatives to VKA in patients without malignancy [156,181]. These alternatives have comparable or superior antithrombotic efficacy. ...
Article
Full-text available
Venous thromboembolism (VTE) is one of the life-threatening complications in cancer patients, the incidence of which is affected by the patient and malignancy-related variables. Location, type, therapeutic route, stage, grade, and non-supportive treatment of the cancer are the most important VTE risk factors. Patient age, ethnicity, and concomitant genetic or acquired comorbidities or thrombophilias are known risk factors for VTE in cancer. All high-risk cancer patients admitted to hospitals or treated as outpatients should receive VTE prophylaxis. Low molecular weight heparin is the main treatment for active malignant VTE. Vitamin K antagonists and non-Vitamin K-dependent oral anticoagulants are used in stable, non-bleeding cancer patients. Anticoagulation should be continued until the cancer is treated or at least controlled. Over the past two decades, randomized clinical and observational trials have improved the pathogenesis and therapeutic knowledge of VTE, but many challenges remain. The lack of an optimal primary prophylaxis method for inpatients and outpatients in oncology and the care of cancer-associated VTE in standard and high-bleeding risk groups are examples for which more clinical research on cancer-associated thrombosis is necessary to address these issues. In this review, we describe the pathogenesis, factors that increase the risk of VTE, and the comparison between the effectiveness of available anticoagulants in the treatment and prevention of VTE in cancer patients.
... The first anticoagulants (vitamin K antagonists and heparins) non-specifically targeted the procoagulant serine proteases of coagulation [2,3]. Then, at the dawn of the 21st century, the development of oral medications specifically targeting thrombin (FIIa) and activated factor X (FXa), formerly referred to as DOACs for direct oral anticoagulants, improved the management of patients with indication for anticoagulant therapy [4,5]. Over time, these specific antithrombotics have become the first-line treatment for most thrombosis patients [6,7]. ...
Article
Full-text available
Direct oral anticoagulants against activated factor X and thrombin were the last milestone in thrombosis treatment. Step by step, they replaced antivitamin K and heparins in most of their therapeutic indications. As effective as the previous anticoagulant, the decreased but persistent risk of bleeding while using direct oral anticoagulants has created space for new therapeutics aiming to provide the same efficacy with better safety. On this basis, drug targeting factor XI emerged as an option. In particular, cancer patients might be one of the populations that will most benefit from this technical advance. In this review, after a brief presentation of the different factor IX inhibitors, we explore the potential benefit of this new treatment for cancer patients.
Article
Full-text available
El síndrome de Stevens-Johnson (SSJ) y la necrólisis epidérmica tóxica (NET) son enfermedades raras que se caracterizan por necrosis epidérmica generalizada y desprendimiento de la piel. Se asocian con una morbilidad y mortalidad significativas, y el diagnóstico y tratamiento tempranos son fundamentales para lograr resultados favorables para los pacientes. En esta revisión se abordaran los avances recientes en el diagnóstico y tratamiento de la enfermedad. Se identificaron múltiples proteínas (galectina 7 y RIP3) que son biomarcadores poNETciales prometedores para SSJ/NET, aunque ambos aún se encuentran en fases iniciales de investigación. En cuanto al tratamiento, la ciclosporina es la terapia más eficaz para el tratamiento del SSJ, y una combinación de inmunoglobulina intravenosa (IVIg) y corticosteroides es más eficaz para la superposición SSJ/NET y NET. Debido a la naturaleza rara de la enfermedad, hay una falta de ensayos controlados aleatorios prospectivos y realizarlos en el futuro proporcionaría información valiosa sobre el tratamiento de esta enfermedad.
Article
Full-text available
This mini-review theoretically illustrates the in silico methods used in the pharmacy field to enhance drug discovery and development and reduce preclinical studies. It is shown that in silico methods are computational-based approaches that study the structure, properties, and activities of molecules using computer simulations and mathematical algorithms. These results highlight the importance of obtaining data that can affect the prediction of in vivo results. Artificial intelligence and machine learning development enhance in silico methods such as quantitative structure-activity relationship, molecular placement, and physiological-based pharmacokinetics, which are usually used. This approach not only saves time and costs but also offers ease of application. Studies conducted to evaluate the use of in silico methods in areas such as pharmacology, toxicology, and pharmaceuticals are provided as examples. It was concluded that over time, in silico methods usage and development increased due to their ability to predict the in vivo performance of the drug.
Article
The effects of food on the pharmacokinetics (PKs) and safety of 10‐mg rivaroxaban tablets in healthy Chinese subjects were investigated from 1 bioequivalence trial. The bioequivalence trial was designed as randomized, open‐label, 2‐sequence, 4‐period crossover under both fasted and fed conditions. A total of 56 healthy subjects were enrolled, 62.5% were male. These subjects received a single oral 10‐mg dose of rivaroxaban with a 7‐day washout between 4 periods. Serial PK samples were collected and plasma concentrations were analyzed using validated high‐performance liquid chromatography–mass spectrometry. Pharmacokinetic parameters were calculated by noncompartmental methods. The BE module of WinNonLin was used for statistical analysis of the maximum concentration (C max ), the area under the concentration–time curve from zero to the final measurable concentration (AUC 0‐t ), and the area under the concentration–time curve from time zero to infinity (AUC 0‐∞ ) of rivaroxaban in plasma. Compared with the fasted state, the C max , AUC 0‐t , and AUC 0‐∞ of rivaroxaban significantly increased by 47%, 28%, and 26%, respectively, with oral administration of rivaroxaban 10 mg in the fed state. The incidence of adverse events (AEs) was similar between the fasted and fed states, and no serious AEs were observed. Food significantly increased the exposure to rivaroxaban 10 mg in Chinese subjects.
Article
BCS class II drugs exhibit low aqueous solubility and high permeability. Such drugs often have an incomplete or erratic absorption profile. This study aimed to predict the effects of β-cyclodextrin (βCD) and different hydrophilic polymers (poloxamer 188 (PXM-188), polyvinyl pyrrolidone (PVP) and soluplus (SOLO)) on the saturated solubility and dissolution profile of hydrophobic model drug rivaroxaban (RIV). Binary inclusion complex with βCD were prepared by kneading and solvent evaporation method, at drug to cyclodextrin weight molar ratios of 1:1, 1:2, and 1:4. Saturated solubility of the hydrophobic model moiety was evaluated with βCD to explore the increment in saturated solubility. Dissolution test was carried out to assess the drug release from the produced binary inclusion complex in the aqueous medium. Solid state analysis was performed using Fourier transform infrared spectroscopy (FTIR), Differential scanning calorimetry (DSC), X-ray diffraction (XRD), and Scanning electron microscopy (SEM) techniques. When compared to pure drug, the binary complex (Drug: βCD at molar ratio of 1:2 w/w) demonstrated the best performance in terms of enhanced solubility and drug release. Furthermore, ternary inclusion complex was prepared with hydrophilic polymers SOLO, PVP K-30 and PXM-188 at 0.5%,1%,2.5%,5% and 10% w/w to optimized binary formulation RIV:βCD (1:2) prepared by kneading (KN) and solvent evaporation (S.E) method. The findings demonstrated that among ternary formulations (1:2 Drug: βCD: SOLO 10% S.E) manifested greatest improvement in saturated solubility and dissolution rate. Results of solubility enhancement and improvement in dissolution profile of model drug by ternary inclusion complexation were also supported by FTIR, DSC, XRD, and SEM analysis. So, it can be concluded that the ternary inclusion systems were more effective compared to the binary combinations in improving solubility as well as dissolution of hydrophobic model drug rivaroxaban.
Article
Background: Establishing the appropriate rivaroxaban dose in older patients with non-valvular atrial fibrillation (NVAF) is important because of the high risk of adverse events. In this EXPAND study subanalysis, we examined the safety and efficacy of standard-dose (15 mg/day) and non-recommended reduced-dose (10 mg/day) rivaroxaban in patients aged ≥65 years with NVAF and preserved renal function. Methods: The entire analysis population (ALL cohort [n = 3982]; ≥65 years) was divided into early elderly (ELD) (65-74 years [n = 1444]) and late ELD (≥75 years [n = 2386]) sub-cohorts. Each sub-cohort was divided into reduced-dose and standard-dose groups. Kaplan-Meier survival curves with adjusted hazard ratios (aHRs) and 95% confidence intervals (CIs) were used to assess efficacy (thromboembolic events) and safety (hemorrhagic events) outcomes. Results: The aHR for major bleeding did not differ between the dosages in any of the cohorts (aHRs: 0.86-0.93). There were no significant differences in the occurrence of stroke + systemic embolism (SE) or stroke + SE + myocardial infarction (MI) + cardiovascular (CV) death among the cohorts. The aHR for MI/unstable angina + interventional/CV surgery + CV death was higher with 10-mg/day rivaroxaban than 15-mg/day rivaroxaban in the ALL cohort (aHR: 1.56 [95% CI 1.02-2.37], p = 0.039) and the late ELD sub-cohort (aHR: 1.86 [95% CI 1.01-3.42], p = 0.045). Conclusions: Reduced-dose rivaroxaban may increase the risk of coronary artery events. The use of rivaroxaban 15 mg/day in patients with NVAF aged ≥75 years with preserved renal function was supported.
Article
The RECORD program comprised four pivotal phase III trials comparing the oral direct Factor Xa inhibitor, rivaroxaban, with subcutaneous enoxaparin for the prevention of venous thromboembolism (VTE) after total hip and knee replacement (THR and TKR). A total of 12,729 patients were randomized to receive rivaroxaban 10 mg once daily (od) starting 6–8 hours after surgery, or enoxaparin 40 mg od (RECORD1–3) starting the evening before surgery, or 30 mg every 12 hours (RECORD4) starting 12–24 hours after surgery. In RECORD1, patients undergoing THR received extended prophylaxis for 31–39 days. In RECORD2, patients undergoing THR received rivaroxaban for 31–39 days, or enoxaparin for 12±2 days followed by placebo up to 35 days. Patients undergoing TKR in RECORD3 and 4 received prophylaxis for 12±2 days. In all four trials the rivaroxaban regimens showed superior efficacy to the enoxaparin regimens with no significant increase in bleeding. The RECORD1–4 pooled analysis showed that rivaroxaban was significantly superior to enoxaparin for the primary endpoint, symptomatic VTE and all-cause mortality (0.8% vs 1.6%, respectively; p<0.001), with similar bleeding rates in both groups. Pooled subgroup analyses were performed to determine the influence of age, weight, gender, or renal function (calculated creatinine clearance [CrCl] at baseline) on the studies primary and main secondary efficacy outcomes, and any adjudicated bleeding events in RECORD1–4. Treatment effect (rivaroxaban vs enoxaparin regimens) was expressed on the odds-ratio (OR) scale for total VTE (the composite of any deep vein thrombosis, non-fatal pulmonary embolism and all-cause mortality, modified-intention-to-treat [mITT] population) and major VTE (the composite of proximal deep vein thrombosis, non-fatal pulmonary embolism and VTE-related death, mITT population for major VTE) at the end of the planned medication period. For any bleeding (major and non-major bleeding after the start of study medication up to 2 days after the last dose, safety population) the hazard ratio (HR) was used. Separate analyses were done for each subgroup. These pooled analyses show consistently superior efficacy to enoxaparin, irrespective of patients’ age (<65, 65–75, >75 yrs), weight (≤70, >70–90, >90kg), gender, or renal function (CrCl >80, 50–80, <50 ml/min). Interaction testing in logistic regression models provided no indication of treatment effect differences in these subgroups, except for total VTE and the CrCl subgroups, where treatment effect with rivaroxaban appeared to be larger for the >80 and <50 ml/min group compared with the 50–80 ml/min group. With respect to any bleeding, the risk on rivaroxaban relative to enoxaparin regimens was comparable across gender and the different age, body weight and renal function (>30 ml/min CrCl) groups studied. Similar trends were seen in the smaller subgroups of body weight extremes (>110 kg subgroup - total VTE: 5/100 [5.0%] vs 8/119 [6.7%], respectively; ≤50 kg subgroup - any bleeding: 12/147 [8.2%] vs 8/162 [4.9%], respectively). These results suggest that age, body weight, gender and renal function (>30 ml/min CrCl) have no clinically relevant effect on the efficacy or safety of rivaroxaban as thrombroprophylaxis following major orthopedic surgery. Subgroup Total VTE Odds ratio (rivaroxaban vs enoxaparin regimens) (95% CI Major VTE Odds ratio (rivaroxaban vs enoxaparin regimens) (95% CI) Any major or non-major bleeding Hazard ratio (rivaroxaban vs enoxaparin regimens) (95% CI) *Because of study exclusion criteria, few patients had CrCl <30 ml/min, CI = confidence interval Age <65 years 0.40 (0.28–0.55) 0.22 (0.10–0.44) 1.05 (0.86–1.29) 65–75 years 0.39 (0.29–0.52) 0.18 (0.08–0.34) 1.11 (0.90–1.38) >75 years 0.54 (0.36–0.81) 0.51 (0.21–1.15) 1.10 (0.79–1.52) Weight ≤70kg 0.43 (0.31–0.60) 0.25 (0.11–0.52) 1.17 (0.91–1.51) >70–90kg 0.45 (0.34–0.59) 0.21 (0.11–0.39) 0.95 (0.77–1.17) >90kg 0.38 (0.25–0.57) 0.32 (0.13–0.69) 1.22 (0.95–1.58) Gender Male 0.35 (0.25–0.48) 0.21 (0.10–0.40) 1.12 (0.93–1.36) Female 0.47 (0.37–0.59) 0.27 (0.16–0.45) 1.06 (0.87–1.28) CrCl (ml/min) >80 0.35 (0.27–0.46) 0.22 (0.12–0.38) 1.12 (0.94–1.35) 50–80 0.60 (0.45–0.81) 0.30 (0.15–0.56) 0.99 (0.79–1.24) <50* 0.32 (0.15–0.64) 0.23 (0.02–1.15) 1.07 (0.64 – 1.79)
Article
Activation of protein C is dependent on thrombin complexed with thrombomodulin (TM). Activated protein C (APC), together with its cofactor protein S, degrades coagulation Factors Va and VIIIa, thereby limiting further thrombin formation. Thus, in addition to suppressing the procoagulant effects of thrombin, direct thrombin inhibitors (DTIs) may also downregulate anticoagulant effects of thrombin-mediated feedback mechanisms. By contrast, direct Factor Xa (FXa) inhibitors block the formation of thrombin, but not its actions. The objective of this study was to investigate whether the direct FXa inhibitor, rivaroxaban, and the DTIs, dabigatran and melagatran, inhibit the negative-feedback reaction of the thrombin–TM complex/APC (thrombin–TM/APC) system and thereby increase thrombin formation. Experiments were conducted in plateletpoor plasma from healthy donors (normal plasma) and in pooled protein C-deficient plasma, both substituted with 1.33 μM phospholipids, in the presence or absence of 10 nM TM with increasing concentrations of rivaroxaban, dabigatran, melagatran, or the appropriate vehicles. Thrombin formation was initiated by adding 1.67 pM tissue factor (TF) and assessed by measuring the cleavage of the fluorogenic substrate Z-Gly-Gly-Arg-AMC (Bachem) using the Calibrated Automated Thrombogram (CAT, Thrombinoscope® BV) method. The parameters assessed were lag time, time to peak thrombin generation (tmax), peak thrombin generation (Cmax), and endogenous thrombin potential (ETP). In addition, formation of prothrombin fragments 1+2 (F1+2) was determined by ELISA (Enzygnost® F1+2 monoclonal [Dade Behring]). Rivaroxaban potently inhibited thrombin formation in the absence and presence of TM across all parameters in a concentration-dependent manner in both normal plasma and protein C-deficient plasma (see Table). In the absence of TM, melagatran and dabigatran also inhibited thrombin formation in a concentration-dependent manner, both in normal plasma and protein C-deficient plasma. In the presence of TM, DTIs prolonged lag time and tmax in a concentration-dependent manner. However, only high concentrations of the DTIs reduced ETP, Cmax, and F1+2, In normal plasma,lower concentrations even increased ETP, Cmax, and F1+2. Increased thrombin formation was observed with melagatran 119–474 nM or dabigatran 68–545 nM. DTIs did not increase thrombin formation in protein C-deficient plasma, suggesting that both protein C and TM are needed for the DTI-mediated increase in thrombin formation. The results suggest that low concentrations of DTIs suppress the anticoagulant effects of the thrombin–TM/APC system by inhibiting activation of protein C by the thrombin–TM complex, and thereby enhance thrombin formation. Conversely, rivaroxaban does not increase thrombin formation, suggesting that it does not suppress the negative-feedback reaction by inhibition of protein C activation. This hypothesis is supported by the absence of enhanced thrombin formation in protein C-deficient plasma. Enhanced thrombin formation might explain the hypercoagulation observed with DTIs in a rat model of TF-induced intravascular coagulation (Furugohri, et al. 2005; Morishima, et al. 2005; Perzborn, et al. 2008) and suggests that DTIs could cause activation of coagulation at lower plasma concentrations. Table. Effect of rivaroxaban, dabigatran, and melagatran on peak thrombin formation (Cmax [nM thrombin]) in the absence or presence of thrombomodulin (TM) in normal plasma (NP) from healthy volunteers (n=8–12), and in pooled protein C-deficient plasma (PPC) in the presence of TM (n=3). Results were obtained by the CAT method and are a mean of n plasma samples. Prothrombin fragments F1+2 (nM F1+2) results were obtained in the presence of TM (mean results; n=3 [NP]). Thrombin (nM) in normal and in protein C-deficient plasma n.d:, no data. Rivaroxaban (nM) 0 18 91 182 363 1,090 Cmax: − TM in NP 273 222 113 71 43 15 Cmax: + TM in NP 100 67 32 18 10 3 Cmax: + TM in PPC 290 249 168 117 80 38 F1+2: + TM (in NP) 184 89 43 18 n.d. 2 Dabigatran (nM) 0 68 136 273 545 1,090 Cmax: − TM in NP 261 287 289 282 239 102 Cmax: + TM in NP 81 156 203 240 218 96 Cmax: + TM in PPC 298 301 300 292 253 116 F1+2: + TM in NP 252 n.d. 362 422 351 98 Melagatran(nM) 0 24 119 237 474 948 Cmax: − TM in NP 276 290 297 289 251 127 Cmax: + TM in NP 101 133 215 251 237 113 Cmax: + TM in PPC 294 296 299 292 256 132 F1+2: + TM in NP 213 n.d. 389 427 393 123
Article
Venous thromboembolism (VTE) is a common, potentially fatal complication of major orthopaedic surgery. Pharmacologic thromboprophylaxis is recommended for patients undergoing total hip arthroplasty (THA) for a minimum of 10 days, and up to 35 days. However, extended thromboprophylaxis is not universally used. Therefore, this trial was conducted to evaluate the potential benefits of extended thromboprophylaxis after THA. RECORD2 is the largest, prospective, randomized clinical trial conducted to date, in this indication. This global, phase III, double-blind trial, was designed to compare short-term thromboprophylaxis with a low molecular weight heparin - enoxaparin - with extended thromboprophylaxis for up to 5 weeks with a novel, oral, direct Factor Xa inhibitor - rivaroxaban after THA. Patients received subcutaneous enoxaparin 40 mg once daily (od), beginning the evening before surgery, continuing for 10–14 days (short-term prophylaxis), and followed by placebo until day 35±4, or oral rivaroxaban 10 mg od beginning 6–8 hours after surgery and continuing for 35±4 days (extended prophylaxis). Mandatory, bilateral venography was conducted at the end of the extended treatment period. The primary efficacy endpoint was the composite of any deep vein thrombosis (DVT), non-fatal pulmonary embolism (PE), and all-cause mortality. The main secondary efficacy endpoint was major VTE; the composite of proximal DVT, non-fatal PE, and VTE-related death. Major and non-major bleeding during double-blind treatment were the primary and secondary safety endpoints, respectively. A total of 2509 patients were randomized; 2457 were included in the safety population and 1733 in the modified intention-to-treat (mITT) population. Extended thromboprophylaxis with rivaroxaban was associated with a significant reduction in the incidence of the primary efficacy endpoint and major VTE, compared with short-term thromboprophylaxis with enoxaparin (Table). The incidences of major and non-major bleeding were similar in both groups (Table). In conclusion, extended duration rivaroxaban was significantly more effective than short term enoxaparin for the prevention of VTE, including major VTE, in patients undergoing THA. Furthermore, this large trial demonstrated that extended thromboprophylaxis provides substantial benefits for patients undergoing THA, and that the oral, direct Factor Xa inhibitor rivaroxaban provides a safe and effective option for such a strategy. Short-term s.c. enoxaparin 40 mg od % (n/N) Extended oral rivaroxaban 10 mg od % (n/N) Relative risk reduction (%) P-value for difference DVT, non-fatal PE, and all-cause mortalitya 9.3% (81/869) 2.0% (17/864) 79% P<0.001 Major VTEb 5.1% (49/962) 0.6% (6/961) 88% P<0.001 Major bleedingc 0.1% (1/1229) 0.1% (1/1228) - P=0.980 Non-major bleedingc 5.5% (67/1229) 6.5% (80/1228) - P=0.246
Article
187 Background New oral anticoagulants hold the promise of simple fixed-dose regimens without the need for monitoring and could make extended use more attractive. Current guidelines advise indefinite therapy in a substantial proportion of DVT patients. The Einstein-DVT study was designed to compare rivaroxaban, a direct oral factor Xa inhibitor, to enoxaparin followed by oral vitamin K antagonist (VKA) treatment in patients with acute DVT for either 3, 6, or 12 months. In Einstein-Extension, patients who had completed 6 to 12 months of anticoagulant treatment for either DVT or PE were randomized to receive rivaroxaban or placebo for an additional 6 or 12 months. Study Design Einstein-DVT was an open label, event-driven (target 88 confirmed recurrent VTEs) non-inferiority study. Subjects with a confirmed acute symptomatic DVT without symptomatic PE were randomized to receive either oral rivaroxaban 15 mg twice-daily for 3 weeks followed by 20 mg once-daily or initial treatment with enoxaparin (1 mg/kg twice daily) followed by oral VKA treatment (warfarin or acenocoumarol, target INR 2.5, range 2.0 to 3.0). Einstein-Extension was a randomized, double-blind, event-driven (target 30 confirmed recurrent VTEs), placebo-controlled, superiority study that evaluated rivaroxaban 20 mg once-daily for an additional 6 or 12 months. The primary efficacy outcome for both studies was recurrent non-fatal or fatal symptomatic venous thromboembolism (VTE). The principal safety outcome was clinically relevant bleeding (major or clinically relevant non-major bleeding) in Einstein-DVT and major bleeding only in Einstein-Extension. All study outcomes were adjudicated by a central and blinded committee. Results Einstein-DVT: the ITT-population consisted of 1,731 rivaroxaban and 1,718 enoxaparin/VKA recipients and rivaroxaban demonstrated non-inferior efficacy to enoxaparin/VKA for the primary outcome (rivaroxaban 36 events (2.1%), enoxaparin/VKA 51 events (3.0%), hazard ratio (HR), 0.68; 95% CI 0.44 –1.04, p <0.0001 for non-inferiority, 0.076 for superiority). Major and non-major clinically relevant bleeding occurred in 8.1% of subjects in both treatment groups (HR 0.97; 95% CI 0.76 –1.22, p =0.77) and major bleedings occurred in 14 (0.8%, 1 fatal) and 20 (1.2%, 5 fatal) of the rivaroxaban and enoxaparin/VKA recipients, respectively (HR 0.65; 95% CI 0.33 –1.28, p =0.21). The net clinical benefit defined as the primary efficacy outcome plus major bleeding showed a HR of 0.67 (95% CI 0.47 – 0.95; p=0.027). In the rivaroxaban group, 38 (2.2%) subjects died versus 49 (2.9%) in the enoxaparin/VKA group (HR 0.67; 95% CI 0.44 – 1.02). The time spent in the therapeutic INR range during VKA treatment was 58%. Einstein-Extension: the ITT population consisted of 602 rivaroxaban and 594 placebo subjects and rivaroxaban demonstrated superiority to placebo for the primary outcome (rivaroxaban 8 events (1.3%), placebo 42 events (7.1%), HR 0.18; 95% CI, 0.09 – 0.39; p<0.0001; number needed to treat: 15) over a mean study treatment period of approximately 6.5 months. Major bleeding did not occur in placebo subjects and was observed in 4 (0.7%, none were fatal) rivaroxaban subjects (p=0.11). Clinically relevant non-major bleeding was noted in 7 (1.2%) and 32 (5.4%) of the placebo and rivaroxaban recipients, respectively (p<0.0001). Two (0.3%) patients in the placebo group died versus 1 (0.2%) in the rivaroxaban group. Efficacy and safety results were consistent across all pre-specified subgroups in both studies. Conclusions Against a background of prolonging anticoagulant treatment for many months to years, this study indicates that oral rivaroxaban, 15 mg twice-daily for 3 weeks followed by 20 mg once-daily, could provide clinicians and patients with a simple, single-drug approach for the acute and continued treatment of DVT that potentially improves the benefit–risk profile of anticoagulation. Disclosures Buller: Bayer Schering Pharma: Consultancy, Research Funding.
Article
YM150 is an oral direct FXa inhibitor used as prophylaxis for venous thromboembolism in patients undergoing elective primary hip replacement surgery (Blood106: 530a (abstract#1865), 2005). No preclinical data has been reported for this compound so far. The biochemical and pharmacological properties of YM150 were evaluated in this study. In addition, the anticoagulation activity of orally administered YM150 was compared with that of YM466, a 1st generation FXa inhibitor, in fed cynomolgus monkeys and in bile duct-cannulated rats. The Ki values for YM150 and YM-222714, its major metabolite, against human FXa were 0.031 and 0.020 μM, respectively (n=4). Those for other serine proteases, such as trypsin, plasmin, and thrombin, were greater than 10 μM. YM150 and YM-222714 doubled the FXa clotting time and PT at 2.0 and 1.8 μM, and 1.2 and 0.95 μM, respectively (n=4). They also strongly inhibited prothrombin activation induced by free Xa, prothrombinase, and whole-blood clots with similar IC50 values (0.025–0.082 μM, n=5). In contrast, enoxaparin was much less effective at inhibiting prothrombin activation induced by prothrombinase or clots than prothrombin activation induced by free Xa (IC50 values: 330, 120, and 3.5 mU/mL, respectively, n=5). In the thromboplastin-induced venous thrombosis model in rats, YM150 (0.3–10 mg/kg i.d.) exerted its antithrombotic effects dose-dependently, with significance at 1 mg/kg (ED50: 0.97 mg/kg, n=6). YM150 prolonged the PT slightly at 10 and 30 mg/kg (1.2 and 1.4 times that of the control group), but the template bleeding time was not affected at 30 mg/kg. Although warfarin also exerted antithrombotic effects dose-dependently and with significance at 0.2 mg/kg (ED50: 0.12 mg/kg, n=6), this dose level markedly prolonged PT and bleeding time (4.4 and 2.2 times that of the control group). In an arterio-venous shunt thrombosis model in rabbits, YM150 (1–10 mg/kg p.o.) exerted antithrombotic effects dose-dependently and with significance at 10 mg/kg (ED50: 4.8 mg/kg, n=6), but did not prolong bleeding time at any dose level. Warfarin also exerted antithrombotic effects dose-dependently and with significance at 0.1 mg/kg/day (ED50: 0.29 mg/kg, n=6). Bleeding time was prolonged significantly at this dose level (control: 3.5 min warfarin: 5.8 min). The plasma concentrations of YM-222714 were 129+/−73.7, 396+/−224, and 3,641+/−902 ng/mL after dosing 1, 3, and 10 mg/kg, respectively, while those of YM150 was substantially lower (less than 125+/−265 ng/mL at 10 mg/kg). In fasted cynomolgus monkeys, oral administration of either YM150 (3–30 mg/kg) or YM466 (1–10 mg/kg) dose-dependently prolonged PT. The anticoagulation activity of YM466 was 3 times that of YM150, but this activity decreased significantly in the presence of food, while that of YM150 did not. The peak plasma anti-FXa activity after oral administration of 3 mg/kg YM150 to bile duct-cannulated rats and sham-operated rats were 67.7% and 68.5%, respectively. In contrast, those of 3 mg/kg YM466 were 57.4% and 26.2%, respectively. These data suggest that food or bile interferes with YM150 less than it does with YM466. In conclusion, YM150 is a promising oral FXa inhibitor that carries a bleeding risk that is less than that of warfarin. It also seems be well-absorbed without interference by food or bile. The in vivo antithrombotic activity of YM150 after oral administration was also determined to be produced by its active metabolite, YM-222714.
Article
The risk of venous thromboembolism is high after total hip arthroplasty and could persist after hospital discharge. Our aim was to compare the use of rivaroxaban for extended thromboprophylaxis with short-term thromboprophylaxis with enoxaparin. METHODS: 2509 patients scheduled to undergo elective total hip arthroplasty were randomly assigned, stratified according to centre, with a computer-generated randomisation code, to receive oral rivaroxaban 10 mg once daily for 31-39 days (with placebo injection for 10-14 days; n=1252), or enoxaparin 40 mg once daily subcutaneously for 10-14 days (with placebo tablet for 31-39 days; n=1257). The primary efficacy outcome was the composite of deep-vein thrombosis (symptomatic or asymptomatic detected by mandatory, bilateral venography), non-fatal pulmonary embolism, and all-cause mortality up to day 30-42. Analyses were done in the modified intention-to-treat population, which consisted of all patients who had received at least one dose of study medication, had undergone planned surgery, and had adequate assessment of thromboembolism. This study is registered at ClinicalTrials.gov, number NCT00332020. FINDINGS: The modified intention-to-treat population for the analysis of the primary efficacy outcome consisted of 864 patients in the rivaroxaban group and 869 in the enoxaparin group. The primary outcome occurred in 17 (2.0%) patients in the rivaroxaban group, compared with 81 (9.3%) in the enoxaparin group (absolute risk reduction 7.3%, 95% CI 5.2-9.4; p<0.0001). The incidence of any on-treatment bleeding was much the same in both groups (81 [6.6%] events in 1228 patients in the rivaroxaban safety population vs 68 [5.5%] of 1229 patients in the enoxaparin safety population; p=0.25). INTERPRETATION: Extended thromboprophylaxis with rivaroxaban was significantly more effective than short-term enoxaparin plus placebo for the prevention of venous thromboembolism, including symptomatic events, in patients undergoing total hip arthroplasty. Comment in Selective factor Xa inhibition for thromboprophylaxis.