Caunt 2015

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

REVIEWS

MEK1 and MEK2 inhibitors


and cancer therapy:
the long and winding road
Christopher J. Caunt1, Matthew J. Sale2, Paul D. Smith3 and Simon J. Cook2
Abstract | The role of the ERK signalling pathway in cancer is thought to be most prominent
in tumours in which mutations in the receptor tyrosine kinases RAS, BRAF, CRAF, MEK1 or
MEK2 drive growth factor-independent ERK1 and ERK2 activation and thence inappropriate
cell proliferation and survival. New drugs that inhibit RAF or MEK1 and MEK2 have recently
been approved or are currently undergoing late-stage clinical evaluation. In this Review,
we consider the ERK pathway, focusing particularly on the role of MEK1 and MEK2, the
‘gatekeepers’ of ERK1/2 activity. We discuss their validation as drug targets, the merits of
targeting MEK1 and MEK2 versus BRAF and the mechanisms of action of different inhibitors
of MEK1 and MEK2. We also consider how some of the systems-level properties
(intrapathway regulatory loops and wider signalling network connections) of the ERK
pathway present a challenge for the success of MEK1 and MEK2 inhibitors, discuss
mechanisms of resistance to these inhibitors, and review their clinical progress.

2i
The authors dedicate this article to Prof. Chris Marshall, the expression of immediate- and delayed-early genes
A cocktail of two protein kinase FRS (1949–2015), an inspirational scientist whose work such as the D‑type cyclins to promote G1/S progression
inhibitors, one inhibiting MEK1 contributed enormously to our understanding of the in the cell cycle4. ERK1 and ERK2 can also regulate cell
and MEK2, and the other RAS-regulated RAF–MEK–ERK pathway. survival by phosphorylating members of the apoptosis
inhibiting glycogen synthase
regulating BCL‑2 protein family at the mitochondria5.
kinase 3 (GSK3).
The ERK signalling pathway is activated by an array ERK1/2 signalling regulates processes that are crucial for
of receptor types, including receptor tyrosine kinases normal development, including cell proliferation, dif-
(RTKs), G protein-coupled receptors and cytokine recep- ferentiation, survival and cell motility; indeed, germline
tors, and the core components of this pathway are now deletion of some components of the ERK pathway causes
well known1,2. Activated RTKs recruit adaptor proteins embryonic lethality 6, and a MEK1/2 inhibitor (MEKi)
1
Department of Biology and
Biochemistry, University of and guanine nucleotide exchange factors (GEFs; such as forms part of the 2i protocol that maintains embryonic
Bath, Claverton Down, SOS) to activate the HRAS, KRAS or NRAS GTPases at stem cell pluripotency 7. The same cellular processes are
Bath BA2 7AY, UK. the inner leaflet of the plasma membrane (FIG. 1). Once deregulated in cancer and represent some of the key hall-
2
Signalling Laboratory, activated, GTP-bound RAS (RAS–GTP) drives the for- marks and driving characteristics of the cancer cell8,9.
The Babraham Institute,
Babraham Research Campus,
mation of high-activity homodimers or heterodimers Many human cancers contain activating mutations in
Cambridge CB22 3AT, UK. of the RAF protein kinases (ARAF, BRAF or CRAF), genes encoding RTKs, RAS, BRAF, CRAF, MEK1 or
3
AstraZeneca, Oncology which directly phosphorylate and activate MEK1 and MEK2, which act as driving oncogenes; consequently,
iMed, Cancer Biosciences, MEK2 (also known as MAPKK1 and MAPKK2). MEK1 many cancers exhibit deregulated activation of, and an
Cancer Research UK, Li Ka
and MEK2 are dual-specificity kinases that activate enhanced dependency on, ERK1/2 signalling.
Shing Centre, Cambridge
Institute, Robinson Way, ERK1 and ERK2 by phosphorylating them at conserved The discovery of the core components of the RAS–
Cambridge CB2 0RE, UK. threonine and tyrosine residues in the T‑E‑Y motif ERK pathway 2 kick-started a protein kinase drug dis-
Correspondence to found in their activation loop. Hundreds of proteins covery effort that continues today 10–12. The first ERK
P.D.S. and S.J.C. have been defined as ERK1 and ERK2 substrates and pathway inhibitor to be discovered, PD98059, was
e-mails: paul.d.smith@
astrazeneca.com;
ERK-interacting partners1,3; these include other pro- reported 20 years ago13 and was shown to act inde-
simon.cook@babraham.ac.uk tein kinases and transcription factors (such as ETS and pendently of ATP as an apparent allosteric inhibitor of
doi:10.1038/nrc4000 the activator protein 1 complex (AP1)), which regulate MEK1 and MEK2. Since then, MEKis have proved to be

NATURE REVIEWS | CANCER VOLUME 15 | O CTOBER 2015 | 577

© 2015 Macmillan Publishers Limited. All rights reserved


REVIEWS

Allosteric inhibitor invaluable research tools, underpinning our knowledge that proved to be BRAF‑V600E selective, culminating
A small molecule that inhibits of ERK1 and ERK2 biology and validating MEK1 and in 2008 in the description and subsequent approval of
the activity of an enzyme by MEK2 as cancer drug targets14,15. As they are not ATP- vemurafenib18,19 and dabrafenib20,21.
binding to a regulatory site that competitive, it was correctly anticipated that MEKis In this Review, we consider the fundamental aspects
is distinct from the active or
catalytic site.
would be more selective than conventional kinase inhibi- of the ERK pathway from a MEK perspective, and the
tors and widely hoped that this would translate into rapid evidence that supports MEK1 and MEK2 as drug targets
clinical success. In fact, it took a further 18 years until in cancer. We describe the MEKis, their unique modes
trametinib16 became the first MEKi to receive regulatory of action and innate resistance mechanisms, and how
approval. In the interim, the identification of activating tumour cells that are sensitive to MEKis can adapt and
BRAF mutations (most notably BRAFT1799A encoding acquire resistance. We also discuss how knowledge of
BRAF‑V600E) in melanoma, thyroid and colorectal MEKi resistance has informed combination strategies
cancer (CRC) in 2002 (REF. 17) galvanized the field and and review the clinical experiences of MEKis, including
focused efforts on developing BRAF inhibitors (BRAFis) successes and lessons learned.

a b c
EGFR or FGFR

SOS RAS GTP SOS RAS GTP SOS RAS GTP –


P P P – SPRY

RAF RAF RAF RAF P RAF RAF


P

P MEK1/2

KSR1

KSR1
MEK1/2 MEK1/2 – –
P

ERK1/2 ERK1/2 ERK1/2 – DUSP6


Cytoplasm

Nucleus
ERK1/2 ERK1/2 ERK1/2 – DUSP5

P P P P P P
ETS ETS ETS

Figure 1 | Scaffolds and feedback controls underpin the normal functioning of the ERK1/2 pathway. a | A simplified
linear representation of the RAS-regulated RAF–MEK–ERK signalling cascade. Activated growth factor Naturereceptors
Reviews (for
| Cancer
example, epidermal growth factor receptor (EGFR) and fibroblast growth factor receptor (FGFR)) recruit the guanine
nucleotide exchange factor (GEF) SOS to promote the release of GDP from RAS, allowing GTP to bind. Active RAS
(RAS–GTP) drives the formation of active homodimers or heterodimers of the RAF protein kinases (ARAF, BRAF or CRAF),
which directly phosphorylate and activate MEK1 and MEK2 (MEK1/2). MEK1/2 are dual-specificity protein kinases that
phosphorylate ERK1 and ERK2 (ERK1/2) on a conserved TEY motif to activate them. ERK1/2 substrates include transcription
factors of the ETS family; in this way the pathway links activated receptors at the plasma membrane to changes in gene
expression in the nucleus. The concentration of the core components typically increases down the pathway such that [RAF]
< [MEK] < [ERK]. This signal amplification allows low-level receptor occupancy to elicit meaningful ERK signals throughout
the cell166–168. b | Activation of the ERK pathway is critically dependent on scaffold proteins. Scaffolds serve several roles,
including increasing the efficiency of interactions between the enzyme and substrate at each step and insulating pathway
components against inputs from other parallel pathways to ensure signal fidelity22,169,170. For example, the kinase suppressor
of RAS 1 (KSR1) scaffold assembles RAF, MEK1/2 and ERK1/2 to increase signalling efficiency; acts allosterically to activate
the RAF kinase domain171; controls the subcellular location of the pathway; and insulates it from other pathways. Scaffolds
tend to make signal transmission more efficient but limit amplification. c | The ERK pathway is extensively regulated by
homeostatic negative feedback controls that fine-tune pathway output166. Rapid and direct feedback mechanisms involve
the phosphorylation of MEK1, CRAF, BRAF, KSR1, SOS and some receptor tyrosine kinases (RTKs) by ERK and downstream
kinases (such as RSK) to inhibit signal propagation23,101. Notably, ERK can phosphorylate CRAF and BRAF to inhibit MEK
phosphorylation172–174, and MEK1 to inhibit ERK phosphorylation41,42. Loss of ERK activity (for example, by treatment with a
MEK inhibitor (MEKi)) collapses these feedback loops and reactivates MEK and ERK; this confers ‘robustness’, allowing the
pathway to adapt to perturbations166,175 and explains the ERK reactivation that is observed in tumour cells with wild-type
BRAF75. The slower de novo expression of Sprouty (SPRY) proteins and the dual-specificity phosphatases (DUSPs) also
regulates pathway output. SPRY proteins inhibit ERK signalling at the level of RTKs, SOS and by interfering with the RAF
catalytic domain176. The DUSPs inactivate ERK by dephosphorylating the pT‑E‑pY motif. ERK-driven expression of DUSPs
provides homologous intrapathway feedback to dampen pathway activation. Additionally, different DUSPs function in
different locations, with DUSP5 residing in the nucleus and DUSP6 in the cytoplasm, allowing differential regulation of ERK
output in these different locations24.

578 | O CTOBER 2015 | VOLUME 15 www.nature.com/reviews/cancer

© 2015 Macmillan Publishers Limited. All rights reserved


REVIEWS

The role of MEK in the ERK pathway DUSP feedback in restraining the oncogenic potential
The ERK pathway is frequently represented as a lin- of MEK–ERK signalling is exemplified by the fre-
ear RAS–RAF–MEK–ERK signalling cascade (FIG. 1a) quent loss of the ERK-specific cytoplasmic phos-
but this ignores the non-enzymatic components of phatase DUSP6 in epidermal growth factor receptor
the pathway and the layers of feed-forward and feed- (EGFR)- and KRAS-driven non-small-cell lung cancers
back regulation that are vital for the role of the ERK (NSCLCs)25 and the demonstration that loss of DUSP5
pathway in information processing. Understanding (the nuclear counterpart of DUSP6) in mouse models
how the pathway is activated by different input stim- accelerates HRAS-driven skin cancer 26. These negative
uli to elicit different responses (such as proliferation feedback loops have additionally emerged as key deter-
or differentiation) and how the pathway responds minants of rapid pathway adaptation and long-term
and adapts to selective inhibition (for example, with acquired resistance to MEKis27,28. When taking into
MEKis) requires some appreciation of these systems- account scaffolding proteins and feedback loops, the
level features. First, activation of the ERK pathway canonical ERK pathway diagram looks more complex
depends on scaffold proteins such as kinase suppres- (FIG. 1c) but even this complexity fails to consider the
sor of RAS 1 (KSR1), which increase the efficiency of position of the ERK pathway within wider signalling net-
the inter­actions between the enzyme and substrate at works (FIG. 2). MEK1 and MEK2 are the only activators
each step22 (FIG. 1b). Second, ERK1/2 signalling is criti- of ERK1 and ERK2 and serve an entirely unique role as
cally regulated by homeostatic feedback controls23 that critical ‘ERK1 and ERK2 gatekeeper’ kinases, processing
include the direct phosphorylation of upstream com- inputs from multiple upstream kinases. Indeed, a RAS-
ponents and increased expression of Sprouty (SPRY) or RAF-centred view ignores the fact that RAF proteins
proteins and dual-specificity phosphatases (DUSPs); are only a subset of the ‘MEK1 and MEK2 activators’
the DUSPs inactivate ERK1 and ERK2 by dephosphory­ in cells (FIG. 2). Multiple MAP kinase kinase kinases
lating the pT‑E‑pY motif 24 (FIG. 1c). The importance of (MAP3Ks) can activate MEK1 and MEK2, and some of

SOS RAS GTP

MEKK1 MEKK3 MAP3K8 MLK2 MLK4


Integrins MEKK2
ARAF BRAF CRAF MLK1 MLK3
RAC

P P
PAK P MEK1/2

MKK3 or MKK4 or
MEK5 IKK
MKK6 MKK7
ERK1/2

P P P P P P P P ERK5 p38 JNK1/2 IκB


BIM MCL1 RSK MNK Cytoplasm

P P P P P P P P P P P P P P Nucleus
FOS ETS MYC MSK MEF2 MEF2 JUN NF-κB

P P
CCND1 CHOP

Figure 2 | MEK1 and MEK2 are the key ‘gatekeepers’ for ERK1 and ERK2 in a wider signalling Nature Reviews |The
network. Cancer
canonical pathway for ERK activation (RAS–RAF–MEK–ERK) is shown on the left. ERK1 and ERK2 (ERK1/2) substrates
include transcription factors (FOS, ETS and MYC) and other protein kinases (RSK, MNK and MSK), thereby controlling the
transcription and translation of genes that promote cell cycle progression such as CCND1 (which encodes cyclin D1);
other ERK substrates include regulators of apoptosis (BIM and MCL1). One key network feature is the convergence of
signalling at the level of MEK1 and MEK2 (MEK1/2). Although ARAF, BRAF and CRAF are the best-studied MEK activators,
a number of other MAP kinase kinase kinases (MAP3Ks; show on the right) can also fulfil this role, including MEKK1 (also
known as MAP3K1), MEKK3 (also known as MAP3K3), MAP3K8 (also known as COT) and the mixed-lineage kinases
(MLK1–4; also known as MAP3K9, MAP3K10, MAP3K11 and KIAA1804)31–35. These ‘alternative’ MEK1/2 activators can also
promote the activation of ERK5, p38, JUN N-terminal kinase (JNK) and nuclear factor-κB (NF‑κB) (through IκB kinase or IKK)
via their relevant upstream activating kinases (MEK5, MKK3 or MKK6, MKK4 or MKK7 or IκB kinase, respectively) to
regulate transcription factors such as MEF2, CHOP, JUN or NF-κB31–35. This reflects a key feature of ERK pathway
architecture: receptor tyrosine kinases (RTKs)177, RAS GTPases2 and MAP3Ks31–35 are relatively promiscuous, activating two
or more parallel signalling cascades; even RAF proteins may regulate non-MEK targets through scaffold functions178,179.
By contrast, MEK1/2 are exquisitely specific activators of ERK1/2. Because ERK1/2 have hundreds of binding partners and
substrates throughout the cell1,3, this represents an astonishing achievement in signal processing through MEK1/2 and
underscores their role as key ‘gatekeepers’ of ERK signalling.

NATURE REVIEWS | CANCER VOLUME 15 | O CTOBER 2015 | 579

© 2015 Macmillan Publishers Limited. All rights reserved


REVIEWS

Receptors Integrins region that stabilizes an inactive kinase conformation.


and RAS The extreme C terminus contains the docking site for

CFC and cancer

CFC and cancer


upstream activating MAP3Ks. Like many other kinases,

mutations

mutations
RAF and ERK PAK the MEK kinase domain consists of a small N‑terminal
MAP3Ks
lobe and a larger C‑terminal lobe; conserved regions that
Ser218 Ser222 Thr292 Ser298 are involved in ATP binding and hydrolysis, substrate
Lys97 P P P P binding, and phosphate transfer are found at the inter-
face between these lobes. Activation of MEK1 and MEK2
MEK1 1 DD NES NRR Kinase catalytic domain AL PRD DVD 393
requires conformational rearrangement of a C‑helix
Ser222 Ser226 in the N‑lobe and the activation loop in the C‑lobe to
Lys101 P P allow the correct alignment of ATP and substrate. This
rearrangement is caused by RAF- or MAP3K‑mediated
MEK2 1 DD NES NRR Kinase catalytic domain AL PRD DVD 400
phosphorylation of Ser218 and Ser222 within the MEK1
activation loop (Ser222 and Ser226 in MEK2)38,39.
Figure 3 | Linear representation of the key functional domains of the human MEK1
NatureofReviews | Cancer MEK1 is also regulated by the phosphorylation
and MEK2 proteins. Human MEK1 and MEK2 encode protein kinases 393 amino
acids and 400 amino acids, respectively. The docking domain (DD) for ERK1 and ERK2, the of additional sites that are clustered within the kinase
nuclear export sequence (NES) and the MAP kinase kinase kinase (MAP3K) docking domain. In addition to RTKs and RAS, MEK1 integrates
domain (domain of versatile docking (DVD)) are shown in red and the amino‑terminal signals from integrins via the RAC-dependent phos-
negative regulatory region (NRR) domain is shown in green. The kinase catalytic domain is phorylation of MEK1 in Ser298, which is catalysed by
shown in blue and includes the highly conserved catalytic lysine residue (Lys97 or Lys101), p21‑activated kinase (PAK1)40,41. MEK1 and MEK2 form
the activation loop (AL), with sites of activating phosphorylation by RAF and other stable heterodimers in vivo that are unable to assemble
MAP3Ks, and the proline-rich domain (PRD) that in MEK1 includes Thr292, a site of when MEK1 is phosphorylated by ERK1 and ERK2 on
negative feedback phosphorylation by ERK1 and ERK2, and Ser298, a site of Thr292; this inability to form dimers decreases MEK1
phosphorylation by p21-activated kinase (PAK). Most gain-of-function mutations in
and MEK2 kinase activity as part of a negative feedback
MEK1 and MEK2 that are found in cancer or cardio-facio-cutaneous (CFC) syndrome
loop41,42. Notably, Thr292 is absent from MEK2, adding
cluster in the NRR or the N‑terminal lobe of the kinase domain (shaded in purple). Figure
adapted: from REF. 36, Elsevier; from REF. 37, Springer; and from Bromberg-White, J. L., to a body of evidence that MEK1 and MEK2 have non-
Andersen, N. J. & Duesbery, N. S. MEK genomics in development and disease. Briefings in redundant roles, despite their high degree of homology
Functional Genomics, 2012, 11, 4, 300–310, by permission of Oxford University Press. and identical substrate specificity. For example, Mek2−/−
mice apparently develop normally 43, whereas Mek1
knockout causes embryonic death at embryonic day 10.5
these MAP3Ks are mutated in cancer 29 and can confer (E10.5) owing to placental defects44.
resistance to BRAFis30,31. Some of these MAP3Ks also
activate the JUN N-terminal kinase (JNK), p38, ERK5 MEK1 and MEK2 in cancer
and nuclear factor-κB (NF‑κB) signalling pathways31–35, The importance of MEK1 and MEK2 in cancer first
so that activation of the ERK pathway proceeds in con- emerged with the recognition of their strategic position
cert with these other pathways (FIG. 2). Thus, specific in the RAS–RAF–MEK–ERK pathway and the demon-
inhibition of MEK1 and MEK2 by allosteric inhibitors stration that activating mutations in the cDNAs encoding
may cause substantial qualitative changes to signalling MEK1 and MEK2 that mimicked activation loop phos-
networks. The extent to which these effects of MEKi phorylation could transform cells45,46. Subsequently, one
are therapeutically desirable or contribute to toxicity of the earliest allosteric MEKis, PD184352 (also known
in normal tissue is unclear. However, such qualitative as CI‑1040), was shown to inhibit tumour cell prolif-
changes in signalling are less likely to be observed in cells eration in vitro and to inhibit tumour growth in vivo14.
that harbour BRAF‑V600E and that are treated with a Since then, an array of studies have demonstrated that
BRAFi, which leaves MEK–ERK activation by other various MEKis block tumour cell growth both in vitro
MAP3Ks intact (see Supplementary information S1 and in vivo, underscoring the broad level of depend-
(figure)). Clearly, understanding the biochemistry and ency of cancer cells on MEK1 and MEK2 in preclinical
cell biology of MEK1 and MEK2 in the context of such models10–12,15. Such ‘MEK addiction’ seems to be strong-
wider signalling networks is essential to understand their est in tumours that harbour BRAFV600E (REF. 47), which
role in oncogenic signalling and to interpret the effects is consistent with the transforming effects of this onco-
of MEKis. gene being mediated via the activation of MEK–ERK.
However, a considerable number of tumour cells that
Functional domains of MEK1 and MEK2. The second- are driven by RAS mutations are also sensitive to MEK1
ary structure of MEK1 and MEK2 (FIG. 3) comprises and MEK2 inhibition in vitro and in vivo47,48. This high-
an amino‑terminal sequence, a central protein kinase lights a key difference between MEKis and BRAFis. The
domain (residues 68–361 in MEK1 and 72–367 in anti-proliferative effects of BRAFis are confined to cells
MEK addiction MEK2) that contains the kinase activation loop and a that express BRAF‑V600E or similar activating muta-
How dependent on MEK1/2 proline-rich segment, as well as a short carboxy‑terminal tions in BRAF49. However, in the presence of RAS–GTP,
activity a tumour cell is for sequence6,36,37. The N‑terminal region contains an ERK1 BRAFis, such as vemurafenib, promote paradoxical acti-
survival and proliferation; it can
broadly be measured by how
and ERK2 docking site and a strong nuclear export vation of ERK1 and ERK2. This is a consequence of the
sensitive a tumour cell is to a sequence that controls cytoplasmic MEK1 and MEK2 drug-induced allosteric transactivation of one BRAF
MEK inhibitor. localization and overlaps with a negative regulatory molecule within RAS-induced RAF dimers and/or the

580 | O CTOBER 2015 | VOLUME 15 www.nature.com/reviews/cancer

© 2015 Macmillan Publishers Limited. All rights reserved


REVIEWS

prevention of RAF-inhibitory autophosphorylation50–53. smoking-associated lung adenocarcinoma, which may


In the clinic, this paradoxical ERK activation results in account for up to 600 patients with lung cancer per
a range of secondary cutaneous lesions, including pap- year in the United States60. Although MEK1 and MEK2
illomas, squamous cell carcinomas, keratoacanthomas mutations are rare in cancer as a whole, their existence
and basal cell carcinomas that profoundly limits the use provides an important level of validation for MEK1 and
of vemurafenib in the treatment of tumours that express MEK2 as drug targets, and their incidence in lung can-
BRAF‑V600E21. cer defines a specific patient population that may ben-
The first demonstration of activating mutations in efit from MEKi therapy. In addition, MEK1 and MEK2
MEK1 or MEK2 came from cardio-facio-cutaneous mutations have emerged as drivers of acquired drug
(CFC) syndrome, a genetic disorder that is caused by the resistance (see below). Finally, the deletion of both Mek1
aberrant activation of ERK1 and ERK2 during develop- and Mek2 in mice prevents the induction of NSCLC by
ment 54. The first report of an amino acid-altering MEK endogenous KrasG12V (REF. 61). Interestingly, although
mutation (encoding MEK2‑P298L) was in a lung can- CRAF is rarely mutated in human cancer its activity is
cer cell line in 1997 but the functional consequences strongly induced by mutant RAS proteins and Craf, but
were not defined55. Activating mutations in MEK1 or not Braf, is required for KrasG12V-driven lung cancers in
MEK2 were first reported in ovarian cancer cell lines mice61,62. Therefore, both CRAF and MEK1 and MEK2
in 2007 (REF. 56); since then, gain‑of‑function mutations in are strongly validated drug targets in mouse models of
MEK1 or MEK2 have been reported in melanoma, CRC lung cancer driven by mutant Kras.
and lung cancer 57–60. Most of these mutations cluster
together with mutations that are found in CFC syndrome Mechanism of action of MEK inhibitors
in either the N‑terminal negative regulatory region or The first synthetic small-molecule inhibitor of MEK1 and
the ATP-binding region of the N‑terminal lobe (FIG. 3). MEK2 kinase activity to be discovered, PD98059, was
Notably, activating MEK1 mutations define a subset of reported in 1995 (REFS 13,63) (TABLE 1). Kinetic experiments

Table 1 | Properties and clinical progression of some widely used allosteric MEK1/2 inhibitors
MEK1/2 inhibitor Year Developer or In vitro IC50 for Ability to disrupt Clinical T0.5 (hours Refs
reported owner MEK1 (nM)* MEK phosphorylation progression or days)
PD098059 1995 Pfizer 2000 ‡ Weak Pre-clinical Not relevant 13,63,73
U0126 1998 DuPont 72‡ Weak Pre-clinical Not relevant 64,73
PD184352 (CI‑1040) 1999 Pfizer 17 ‡
Weak Phase II 20.9+/−4.8 h 14,72,180||
PD0325901 2004 Pfizer 1‡
Weak Phase II 7.8 h 10,181||
Binimetinib (MEK162, 2006 Novartis/Array 12 Weak Phase III 3.63–7.4 h 182–185
ARRY‑438162) Biopharma
Selumetinib 2007 AstraZeneca/ 14‡ Weak Phase III 5.33 h 68,73,
(AZD6244, Array Biopharma 186
ARRY‑142886)
Refametinib 2009 Bayer AG 19§ Weak Phase II 12 h 114,187,
(RDEA119, BAY 188
869766)
CH4987655 2009 Chugai 5.2 Moderate Phase I 4h 75,79,
(RO4987655) Pharmaceutical 189,190
Co
Pimasertib 2010 Merck KGaA 52‡ Not available Phase II 5h 191
(AS703026,
MSC1936369)
TAK‑733 2011 Takeda 3.2‡ Weak Phase I 48–56 h 192–194
Trametinib 2011 GlaxoSmithKline 0.7 Moderate Approved for ~4 days 16,75,195
(GSK1120212) BRAFV600E/K‑mutant
melanoma
CH5126766 2012 Chugai 160‡ Strong Phase I 60 h 75,78,196
(RO5126766) Pharmaceutical
Co
Cobimetinib 2012 Genentech 4.2 Weak Phase III 40 h 74,110,
(GDC‑0973, XL518) (Roche) 197
GDC‑0623 2013 Genentech 5 Strong Phase I 4–10 h 74,198
(Roche)
*No unified protocol was used to generate the in vitro MEK1 IC50 values stated and so these should be treated as a rough guide to potency only. IC50 values shown
for CH4987655, cobimetinib, GDC‑0623 and trametinib were generated using methods in which MEK1 was activated after incubation with inhibitor. Apparent
potency can differ greatly depending on whether constitutively active MEK1 (S218D/E S222D/E) is used (‡) or whether MEK1 is activated by RAF before inhibitor
addition (§) as well as other experimental details. ||J. Sebolt-Leopold, personal communication (2015).

NATURE REVIEWS | CANCER VOLUME 15 | O CTOBER 2015 | 581

© 2015 Macmillan Publishers Limited. All rights reserved


REVIEWS

Feedback relief with PD98059 and U0126 (REF. 64), another MEKi, dem- of PD184352 was curtailed by poor biological availability
ERK1/2‑catalysed feedback onstrated that both molecules shared a common binding and potency 70, the past 10 years have seen the discov-
phosphorylation and inhibition site and inhibited MEK1 and MEK2 non-competitively ery of highly selective allosteric MEKis with far superior
of RAF normally operates in with respect to their substrates Mg‑ATP and ERK1 and pharmacological and pharmaceutical properties (TABLE 1).
the pathway but is relieved
when MEK1/2 are inhibited
ERK2. These early MEKis served as valuable research Until recently, allosteric MEKis were thought to act
and ERK1/2 activity declines, tools but their low potency, physical properties and cer- through broadly equivalent mechanisms71; however,
resulting in the activation of tain off-target effects65,66 stimulated the development recent studies have changed this view. In cells with
RAF and further MEK1/2 of further generations of MEKi (TABLE 1). wild-type BRAF, including those with RAS mutations,
phosphorylation. Pathway
PD184352 was the first MEKi to be evaluated in vivo ERK-dependent feedback phosphorylation of BRAF and
activity also stimulates the
expression of negative
and was shown to inhibit the growth of CRC tumour CRAF inhibits the binding of both BRAF and CRAF to
regulators of pathway activity xenografts14. Crystal structures of MEK1 bound to ana- RAS–GTP and disrupts BRAF–CRAF heterodimers, thus
such as the ERK phosphatase logues of PD184352 demonstrated the presence of a reducing RAF signalling to MEK1 and MEK2 (FIG. 1c).
DUSP6; expression of these unique inhibitor-binding pocket that is separate from, As a result, the loss of ERK1 and ERK2 activity following
negative regulators is reduced
when the pathway is inhibited.
but adjacent to, the Mg‑ATP-binding site in both MEK1 MEK1 and MEK2 inhibition results in the dephosphory­
and MEK2 (REF. 67); this region of MEK1 and MEK2 lation and activation of CRAF (so‑called feedback relief)
exhibits little homology to other protein kinases, thus and a CRAF-dependent increase in phosphorylation of
providing an explanation for why this class of drugs is so the MEK1 and MEK2 activation loop and reactivation
specific68. Allosteric MEKis, such as PD184352, stabilize of ERK1 and ERK2 (FIG. 4). This has been observed for
an inactive conformation of MEK1 and MEK2 in which many MEKis, including PD098059, U0126, PD184352,
helix C and the activation loop are displaced, resulting PD0325901, selumetinib (also known as AZD6244
in the misalignment of catalytic residues and the pos- and ARRY-142886) and cobimetinib 72–75. However,
sible partial occlusion of the ERK1 and ERK2 activation enhanced MEK1 and MEK2 phosphorylation in
loop binding site67–69. Although the clinical progression response to MEKis is not observed in BRAF-mutant cells,

a b c
EGFR or FGFR

RAS GTP RAS GTP RAS GTP RAS GTP RAS GTP

RAF RAF RAF RAF


RAF RAF
RAF RAF RAF RAF
• CH5126766
• GDC-0623
P P P P P P • Trametinib
MEK1/2 MEK1/2 MEK1/2 MEK1/2

P P • PD0325901 P P
ERK1/2 • Selumetinib ERK1/2 ERK1/2
• Cobimetinib

DUSP
• Proliferation Adaptive
• Survival resistance

Figure 4 | MEKis that inhibit MEK1/2 phosphorylation suppress the rebound in ERK1/2 activation that results
Nature Reviews | Cancer
from relief of negative feedback. a | With the exception of tumour cells that harbour activating BRAF mutations, ERK1
and ERK2 (ERK1/2) signalling is subject to extensive ERK1/2-dependent negative feedback at multiple levels of the
pathway, including RAF. RAS–GTP drives the formation of high activity homodimers or heterodimers of the RAF protein
kinases (ARAF, BRAF and CRAF), whereas ERK-dependent phosphorylation of RAF proteins inhibits the binding of BRAF
and CRAF to RAS–GTP and disrupts BRAF:CRAF heterodimers, thereby inhibiting phosphorylation of MEK. b | MEK1 and
MEK2 (MEK1/2) inhibition with a MEK inhibitor (MEKi) blocks ERK1/2 activation and so relieves this negative feedback and
consequently allows stronger activation of upstream pathway components, including RAS and RAF (represented by
increased numbers of active RAS molecules and active RAF dimers). The majority of MEKis, such as PD0325901,
selumetinib and cobimetinib, do not disrupt the phosphorylation of the MEK1/2 activation loop sites and so treatment
with these MEKis and concomitant relief of negative feedback typically results in the accumulation of phosphorylated
MEK1/2. This is thought to explain the rebound in phosphorylated ERK1/2 and pathway output that is observed with
these MEKis in various contexts, notably in tumour cells with mutant RAS74,75. c | A subset of newer MEKis, so-called
‘feedback busters’, including trametinib, CH5126766 and GDC‑0623, disrupt the conformation of the MEK1/2 activation
loop sites so that they can no longer be efficiently phosphorylated by RAF. Thus, although MEK inhibition with these MEKis
is also expected to result in stronger activation of upstream pathway components, little or no increase in MEK1/2
phosphorylation or rebound in ERK1/2 activity is observed (right), translating into more durable pathway inhibition and
superior efficacy in preclinical models75.

582 | O CTOBER 2015 | VOLUME 15 www.nature.com/reviews/cancer

© 2015 Macmillan Publishers Limited. All rights reserved


REVIEWS

Feedback buster especially BRAF‑V600E melanoma, principally because but inhibit their phosphorylation74,75,78. Both trametinib
A term adopted by the field, BRAF‑V600E is fully active as a monomer and insensi- and cobimetinib promote the dissociation of RAF–
although something of a tive to ERK1- and ERK2‑dependent inhibitory phospho- MEK1/2 complexes, but whereas trametinib antago-
misnomer. All MEK inhibitors rylation and the inhibitory action of SPRY proteins, but nizes MEK1 and MEK2 phosphorylation, cobimetinib
(MEKis) relieve ERK-dependent
negative feedback to RAF,
also because the activity of RAS and CRAF is typically does not 74,75. Thus, although a firm causal relationship
resulting in RAF activation. low in BRAF‑V600E melanoma72,76,77. Intriguingly, some between MEKi activity and the association or disso-
Feedback buster MEKis newer MEKis — including trametinib, CH5126766 and ciation of the RAF–MEK1/2 complex has not yet been
mitigate some, but not all, GDC‑0623 — while inhibiting negative feedback mecha- established, the modulation of these complexes may
consequences of feedback
nisms, mitigate the consequences of this by reducing the be integral to their mode of action and may influence
relief that arise when ERK is
inhibited by interfering with the
phosphorylationn of MEK by RAF74,78 (FIG. 4). This may the depth and duration of pathway inhibition.
phosphorylation of MEK by explain why these inhibitors cause more durable sup-
RAF, thereby reducing rebound pression of ERK1 and ERK2 phosphorylation, cell pro- Feedback buster MEKis in the context of BRAFV600E
activation of the pathway. By liferation and xenograft growth in the context of RAS mutations. Intriguingly, MEKis that disrupt MEK1 and
contrast, conventional MEKis
do not prevent MEK
mutations74,75. It is important to understand the under­ MEK2 phosphorylation seem to have a relatively weaker
phosphorylation by RAF. lying mechanism of these feedback buster MEKis because affinity for dual-phosphorylated MEK1/2 than do the
KRAS-mutant tumours represent an important unmet MEKis that permit the accumulation of phosphorylated
Phosphomimetic mutant clinical need and there is a growing interest in using MEK1/2. The binding of GDC‑0623 to a constitutively
A phosphomimetic mutant of
MEKis as part of drug combinations in KRAS-mutant active phosphomimetic mutant of MEK1 and the binding
MEK1 exhibits constitutive
(MAP3K‑independent)
disease, including NSCLC. of trametinib to MEK1 that had been pre-phosphory­
activation owing to acidic lated in vitro were both markedly weaker than binding to
substitutions at Ser218 and Inhibition of activation loop phosphorylation. Differences wild-type MEK1 and unphosphorylated MEK1, respec-
Ser222 in the activation loop in the mechanism of action may reflect distinct inter­actions tively 16,74. By contrast, cobimetinib bound with similar
that mimic the negative charge
of phosphorylation.
between the MEKi and the MEK1 and MEK2 activation affinity to wild-type and phosphomimetic MEK1, and
loop residues. For example, GDC‑0623 is proposed to form suppressed ERK phosphorylation more effectively than
a stronger hydrogen bond inter­action with Ser212 than GDC‑0623 in cells expressing phosphomimetic MEK1
other inhibitors, thereby constraining the activation loop (REF. 74). Thus, for certain feedback buster MEKis, an
and disrupting phosphorylation by RAF74. However, crys- equilibrium may exist between the inhibition of MEK1
tal structures of MEK1 bound to various allosteric MEKis and MEK2 activation loop phosphorylation and the
— such as CH4987655, CH5126766, and PD184352- and attenuation of MEKi binding by the phosphorylated
PD0325901‑like inhibitors — suggest that interaction with activation loop, with the balance being determined by
Ser212 is invariably crucial for the binding of allosteric pretreatment phospho‑MEK1/2 levels. These observa-
MEKis to MEK1 and MEK2 (REFS 67,69,75,79). Instead, tions may be of particular relevance when constitutive
the ability to disrupt MEK1 and MEK2 phosphory­lation MEK1 and MEK2 phosphorylation levels are high,
correlates with the extent to which a MEKi coordinates such as in BRAF‑V600E‑positive cells; indeed, there is
with Asn221 and Ser222 to displace the activation loop75,79. some support for this proposal from BRAFV600E-mutant
Whereas CH5126766 binds both Asn221 and Ser222 and xenograft models72. Clearly, careful consideration of the
causes activation loop displacement, a near-identical signalling context and particular properties of a MEKi
enantiomer of PD0325901 does not coordinate Asn221 will be required to achieve optimal clinical responses,
and Ser222 or displace the activation loop, which is con- especially in RAS-mutant tumours.
sistent with the differing abilities of these inhibitors to
disrupt MEK1 and MEK2 phosphorylation69,75. Whether Resistance to MEKi and mitigation
these MEKis antagonize phosphorylation of MEK1 at both Preclinical and clinical studies have led to the identifica-
Ser218 and Ser222 (and of MEK2 at Ser222 and Ser226) tion of various modes of innate, adaptive and acquired
is unknown; mass spectrometry suggests that trametinib resistance to MEKis, many of which are druggable,
prevents the phosphorylation of MEK1 at Ser218, but not allowing relevant combination strategies to be tested.
at Ser222, and this monophosphorylated form of MEK1
has severely limited kinase activity compared with dual- Intrinsic resistance through parallel oncogenic path-
phosphorylated MEK1 (REF. 16). Further structural analyses ways. Primary sensitivity to MEKi correlates with the
will determine whether, and how, distinct MEKis disrupt decreased expression of cyclin D1 (CCND1), expres-
the phosphorylation of MEK1 and MEK2 and may guide sion of p27 (also known as KIP1) and cell cycle arrest.
future MEKi development. However, the deregulation of the cyclin-dependent
kinases 4 and 6 (CDK4/6)–RB axis (such as through
Modulation of RAF–MEK1/2 complexes. MEKis also the amplification of CCND1 or CDK4 or the loss of CDK
modulate the interaction between MEK1 or MEK2 and inhibitor 2A (CDKN2A)) is common in many cancers80
RAF, and although this may influence their ability to sup- and can confer resistance to ERK pathway inhibitors81.
press MEK1 and MEK2 activity, no simple correlation Indeed, activating mutations in RAS, BRAF or MEK1 and
is apparent. For example, selumetinib and PD0325901 MEK2 can co­operate with CDKN2A loss in a variety of
induce the binding of all three RAF kinases to MEK1 and tumours82–84. Such results have led to the effective combi-
MEK2, and this seems to attenuate the activity of these nation of BRAFis or MEKis with inhibitors of CDK4/6 in
MEKis75. By contrast, GDC‑0623 and CH5126766 also preclinical models85,86. Other signalling pathways — such
promote the association of RAF with MEK1 and MEK2, as PI3K, adenomatous polyposis coli (APC)–β‑catenin,

NATURE REVIEWS | CANCER VOLUME 15 | O CTOBER 2015 | 583

© 2015 Macmillan Publishers Limited. All rights reserved


REVIEWS

PDGFRβ,
EGFR or HER2 HER3 VEGFR or AXL FGFR2 or FGFR3

RAS P PI3K

RAF RAF

STAT3
MEKi MEK1/2 ERK1/2

Cytoplasm • Proliferation
• Survival
Nucleus
ERK1/2

CtBP MYC ?
Inhibit gene
BETi BETi BETi
expression
ERBB3 PDGFRB FGFR2
NRG1 VEGFR2 FGFR3
AXL

Figure 5 | Adaptive kinome reprograming arising from MEK1/2 inhibition. Active ERK1 and ERK2 (ERK1/2) chronically
Nature Reviews
inhibit signalling from an array of receptor tyrosine kinases (RTKs) by direct phosphorylation of receptors | Cancer
at inhibitory sites
(for example, ERK-mediated phosphorylation of epidermal growth factor receptor (EGFR) and HER2) or by repressing the
transcription of the genes that encode RTKs and their cognate ligands, mediated by ERK1/2‑dependent phosphorylation
of transcriptional regulators such as CtBP and MYC (solid arrows). Inhibition of MEK1 and MEK2 (MEK1/2) collapses these
feedback loops, resulting in rapid and sustained reactivation of multiple RTKs that can then sustain cell survival and
proliferation by reactivation of RAF–MEK–ERK signalling or by activation of PI3K or signal transducer and activator of
transcription 3 (STAT3)-dependent signalling28,96–99 (dashed arrows). Such kinome reprogramming validates clinical trials in
which MEK inhibitors (MEKis) are being combined with various RTK inhibitors. However, MEKis frequently cause activation
of multiple RTKs so an alternative, broader approach is to combine a MEKi with BET domain inhibitors (BETis), which act on
key chromatin reader proteins to inhibit transcription100. FGFR, fibroblast growth factor receptor; NRG1, neuregulin 1;
PDGFRβ, platelet-derived growth factor receptor-β; VEGFR, vascular endothelial growth factor receptor.

signal transducer and activator of transcription 3 For example, tumour cell lines with BRAFV600E exhibited
(STAT3) and NF-κB — converge on the same cell cycle durable inhibition of ERK1 and ERK2 and were very
regulators and can drive primary MEKi resistance. For sensitive to PD0325901, whereas tumour cell lines with
example, even in tumour cells that harbour BRAFV600E, mutations in KRAS were notably resistant to PD0325901
strong PI3K‑dependent signalling owing to mutations and exhibited strong ERK1 and ERK2 reactivation within
in PIK3CA or the loss of PTEN can maintain CCND1 24–48 hours of drug administration75. Thus, hardwired
levels in the presence of MEKis and can confer resist- homeostatic mechanisms dictate the efficacy of MEKi in
ance to MEKis87,88. This can be overcome by combining tumour cells with wild-type BRAF, including those with
MEKis with PI3K, mTOR or AKT1 (also known as PKB) mutations in RAS, and provide a clear rationale for dual
inhibitors. In addition, the frequent increase in PKB/AKT pathway inhibition: RAFi plus MEKi to prevent MEK1
activity following treatment with MEKis has prompted and MEK2 reactivation, or MEKi plus ERKi to prevent
numerous studies with these drug combinations89,90 and ERK1 and ERK2 reactivation. An alternative approach
ongoing clinical trials91. STAT3 activation has been shown may be to use feedback buster MEKis with a dual mecha-
to promote MEKi resistance; combining a STAT3 inhibi- nism — such as trametinib, CH5126766, GDC‑0623 and
tor with selumetinib overcame resistance and promoted G-573 — that inhibit MEK1 and MEK2 kinase activity
tumour cell death92,93. More recently, the Hippo pathway and prevent MEK1/2 phosphorylation by RAF74,75 (FIG. 4).
effector YAP1 has been shown to promote resistance to Interestingly, despite the existence of multiple non-RAF
RAFi or MEKi therapy, and combined inhibition of YAP1 MAP3Ks that can activate MEK1 and MEK2 (includ-
and MEKi was synthetic lethal in tumour cells94. ing MAP3K8 (also known as COT), some MEKKs and
MLKs31–35) few studies have assessed whether these alter-
Loss of feedback inhibition and MEK–ERK reactiva- native MEK activators are inhibited by ERK-dependent
tion. Although MEKis have shown great promise in feedback phosphorylation and thus might contribute to
BRAF‑V600E preclinical models16,47, as well as some MEK reactivation following MEK inhibition.
clinical activity in BRAF‑V600E melanoma95, their
more limited success in RAS-mutant tumours may well Adaptive kinome reprogramming. ERK1/2 signal-
be due in large part to the loss of ERK1/2‑dependent ling inhibits an array of RTKs such that MEK1 and
feedback and reactivation of the MEK–ERK1 pathway. MEK2 inhibition elicits rapid RTK activation (FIG. 5).

584 | O CTOBER 2015 | VOLUME 15 www.nature.com/reviews/cancer

© 2015 Macmillan Publishers Limited. All rights reserved


REVIEWS

For example, ERK1- and ERK2‑mediated phosphory­ MEKi resistance. Thus, amplification of the upstream
lation of EGFR and HER2 (also known as ERBB2) sup- driving oncogene (BRAFV600E or KRASG13D) drives resist-
presses signalling to HER3 (also known as ERBB3) so that ance to MEKi by activating a greater pool of MEK1 and
MEKis promote the rapid reactivation of EGFR–HER3 MEK2 to maintain ERK1 and ERK2 activity (FIG. 6).
and PI3K‑dependent signalling 96,97. Unbiased, non- Strikingly, most, if not all, mechanisms of MEKi
targeted chemical proteomic analysis demonstrated the resistance reinstate active ERK1 and ERK2 in the pres-
activation of multiple RTKs following MEK1 and MEK2 ence of MEKis (FIG. 6), thus highlighting a strong hard-
inhibition, which were otherwise repressed by the ERK1 wired addiction to ERK1 and ERK2 signalling. This
and ERK2 target MYC28. In BRAF-mutant thyroid cancer provides another opportunity for rational therapeutic
cells, BRAFis or MEKis promoted ERK1 and ERK2 reac- intervention through dual pathway inhibition; MEKi
tivation involving HER3 transcription and the autocrine plus ERKi in the case of MEK1 and MEK2 mutations103
secretion of neuregulin 1 (NRG1)98, and in KRAS-mutant and BRAFi plus MEKi in the case of BRAFV600E ampli-
NSCLC cells selumetinib promoted autocrine fibroblast fication106,107. Indeed, the combination of dabrafenib
growth factor (FGF) production and increased expression (a BRAFi) and trametinib (a MEKi) has now received reg-
of FGF receptors (FGFRs), which conferred MEKi resist- ulatory approval in advanced-stage melanoma, in which
ance via STAT3 activation99. These studies underscore it increased progression-free survival (compared with
the extent of such ‘kinome re‑programming’ in response either agent alone) and reduced the incidence of cutane-
to MEKis and support ongoing trials that are combining ous lesions arising from paradoxical RAF activation in
MEKis with relevant RTK inhibitors. However, as multiple tissues with wild-type BRAF21,108. However, recent studies
RTKs are often activated in cancer and because transcrip- have revealed that acquired resistance may still arise in
tional programmes underpin kinome reprogramming, an patients with BRAFV600E melanoma who are treated with
alternative strategy is to combine MEKis with inhibitors of dual BRAFi plus MEKi blockade through the acquisition
key chromatin reader proteins such as the bromodomain- of BRAFV600E ‘ultra’ amplification, MEK1 mutations and/or
containing protein 4 (BRD4). Certainly, JQ1, a BET loss of the CDKN2A locus, as well as loss of the nuclear
domain inhibitor, synergizes well with tyrosine kinase ERK-regulatory phosphatase DUSP4 (REF. 27), echo-
inhibitors in acute myelogenous leukaemia100 (FIG. 5). ing the recurring theme of ERK pathway re‑activation.
In the case of KRASG13D amplification, resistance is more
Acquired resistance to MEKis. Various studies have now complex; KRAS activates multiple effector pathways,
investigated acquired resistance to long-term MEKi suggesting that ERK pathway inhibition may need to be
exposure, both in tumour cell lines and in clinical sam- combined with AKT, PI3K or mTOR kinase inhibitors.
ples101. The first report of acquired resistance to MEKis Indeed, such combinations are currently being tested in
was the identification of a gain‑of‑function mutation preclinical disease models and clinical trials89–91.
(MEK1P124L) in a metastatic focus of a patient with More recently, the first potent, selective ERKis have
melanoma with BRAFV600E, who exhibited prolonged been described, including VTX11e, an ATP-competitive
stable disease under treatment with selumetinib102. inhibitor of ERK1 and ERK2, and SCH772984, an ATP-
Subsequently, mutations in MEK1 and MEK2 have competitive inhibitor of ERK1 and ERK2 that also pre-
been found in CRC, melanoma and breast cancer cell vents the activating phosphorylation of ERK1 and ERK2
lines (with driving mutations in BRAF or KRAS) that by MEK1 and MEK2 (REFS 103,104). Of these selective
have been treated with various MEKis (selumetinib, ERKis, SCH772984 can resensitize tumour cells with
trametinib and RO4927350), as well as in samples from acquired resistance to either BRAFis or MEKis, provid-
patients undergoing MEKi therapy 103–105. These ‘on ing proof‑of‑principle for combining ERK inhibitors
target’ mutations (MEK1F129L and MEK2Q60P) activate with BRAFis or MEKis. SCH900353 (MK08353), an
MEK1/2 or cluster at the allosteric inhibitor-binding analogue of SCH772984, has now entered clinical trials,
pocket and abrogate MEKi binding (for example, as have BVD‑523, RG7842 and CC‑90003. These new
MEK1L115P, MEK1G128D/L215P, MEK1F129L, MEK1V211D and potent and selective ERK inhibitors add new weapons to
MEK2 V215E), thereby maintaining ERK1 and ERK2 the arsenal of ERK1/2 pathway-targeted drugs for use as
activity in the presence of MEKis (FIG. 6). first-line therapies, probably in combination, and for the
Studies have also reported acquired resistance to treatment of ‘on‑pathway’ resistance.
MEKis via the amplification of BRAFV600E. Two stud-
ies in four CRC cell lines harbouring BRAFV600E dem- MEKis in the clinical setting
onstrated the amplification of the mutant BRAF allele; Pharmacokinetics and dose schedule. In addition to dif-
MEKi resistance was overcome by BRAF RNA interfer- ferences in the mechanism of inhibition, the pharmaco­
ence (RNAi) or co‑treatment with a RAF inhibitor 106,107. kinetic properties of the MEKis that are currently in
BRAFV600E amplification has also been observed in mela- clinical development vary considerably with half-lives
noma cell lines selected in trametinib105, indicating that (T0.5) ranging from 5 hours to 4–5 days (TABLE 1). Despite
BRAFV600E amplification is a common mecha­nism of this, there has been little sustained effort to optimize the
resistance irrespective of the MEKi used. Amplification dose and schedule of administration on the basis of bio-
of KRASG13D has also been described as a mechanism of logical rationale rather than pragmatic choices based
MEKi resistance in CRC cells107, either in isolation or upon pharmacokinetics and tolerance109, and most clini-
concurrent with an activating MEK1F129L mutation103, cal trials are predicated upon continuous daily dosing.
suggesting that these oncogenes may cooperate to confer The exception is cobimetinib, which is administered on

NATURE REVIEWS | CANCER VOLUME 15 | O CTOBER 2015 | 585

© 2015 Macmillan Publishers Limited. All rights reserved


REVIEWS

a MEK1 or MEK2 mutation prevents MEKi b BRAF amplification maintains or c KRAS amplification increases pool of active
binding or activates MEK1 or MEK2 increases pool of active MEK1 or MEK2 MEK1 or MEK2 but also activates other RAS effectors

KRAS KRAS KRAS KRAS KRAS PLCε


Ca2+

PKC
BRAF BRAF
BRAF BRAF ARAF BRAF CRAF PI3K RAL-GEF

MEK1/2 MEK1/2 MEK1/2 MEK1/2 MEK1/2 MEK1/2 MEK1/2 MEK1/2 PDK1 RAL

P P P P P P
ERK1/2 ERK1/2 ERK1/2 PKB mTOR

Mitigation Mitigation Mitigation


Combine MEKi Combine MEKi Combine MEKi with RAFi
with ERKi with RAFi and PI3Ki, PKBi or mTORi

Figure 6 | Mechanisms of acquired resistance to MEK1/2 inhibitors resistance. Selective amplification of the mutant BRAF V600E
allele greatly
Nature Reviews | Cancer
(MEKis). Studies in preclinical models and analysis of clinical samples increases the proportion of active MEK1/2, exceeding the drug-inhibited
have revealed two basic mechanisms of acquired resistance to MEK1 and pool, to reactivate ERK1/2; in this case, resistance can be overcome by
MEK2 (MEK1/2) inhibitors (MEKis; represented by a red X), and, in both combining a MEKi with a RAF inhibitor (RAFi)106,107. c | Amplification of
cases, tumour cells adapt to maintain or re‑activate ERK1 and ERK2 KRASG13D can also drive MEKi resistance by the same basic mechanism107
(ERK1/2) in the presence of the drug101, providing a rational basis for but provides a greater therapeutic challenge. KRASG13D can activate
treating resistance. a | The emergence of mutations in MEK1 or MEK2 multiple signalling pathways (such as PI3K, RAL, phospholipase Cε (PLCε),
(indicated by a hexagon), provides an example of ‘on‑target’ resistance to and so on) and even the combined inhibition of MEK1/2 and PI3K signalling
allosteric MEKis. Emergent MEK1/2 mutations confer resistance by fails to reverse the acquired resistance to selumetinib that is driven by
reducing MEKi binding or enhancing intrinsic MEK1/2 activity102–105 and amplification of KRASG13D (REF. 107). Hexagons indicate oncogenic
resistance can be overcome by combining MEKi with ERK1/2 inhibitors. mutations, either those that are present as the primary driving oncogene
b | Amplification of the upstream driving oncogene, BRAFV600E (REFS 106,107) and are amplified by MEKi selection (BRAFV600E, KRASG13D) or those that
or KRASG13D (REF. 107), has also emerged as a mechanism of acquired MEKi emerge upon selumetinib selection (for example, MEK1P124L and MEK1F129L).

a 2 weeks on, 1 week off schedule; a pragmatic choice mono­therapy. Trametinib has gained regulatory approval
that is based upon its T0.5 of ~40 hours110, its adverse for this indication following rapid clinical develop-
event profile and evidence of melanoma re‑growth using ment from a single-arm Phase I trial to a randomized
schedules with longer drug holidays111. Phase III trial in which treatment with this MEKi yielded
4.8 months progression-free survival compared with
Pharmacodynamics. The main biomarker for monitor- 1.5 months for dacarbazine95. In addition, the licensing
ing pathway inhibition by MEKis, phosphorylated ERK1 and supervisory authority of Switzerland has recently
and ERK2 (p‑ERK), has been assessed in normal tissues approved cobimetinib for use in combination with
(such as peripheral blood mononuclear cells) and tumour vemurafenib as a treatment for patients with advanced
biopsy samples. In multi-tumour-type Phase I studies, melanoma. Such is the pace of clinical development
selumetinib, trametinib, BAY 86–9766 and pimasertib all in advanced-stage melanoma that MEKi monotherapy in
reduced levels of p-ERK1 and p-ERK2 in tumour tissue BRAF‑V600E melanoma — which is relatively ineffective
using immunohistochemistry assays, and in some cases in patients who relapse following treatment with a first-
have shown complete inhibition112–114. On this basis and generation BRAFi115 — is being superseded by BRAFi
on the basis of data from BRAF inhibitors in BRAFV600E plus MEKi combinations, of which three have completed
melanoma18 the prevailing view is that 80% inhibition or are in clinical trials111,116,117. MEKis have been tested in
of ERK1 and ERK2 phosphorylation is the benchmark multiple tumour types with a high incidence of BRAF or
for clinically effective MEK1 and MEK2 inhibition. RAS mutations with mixed results. Some diseases seem
However, these measures of target inhibition are based notably refractory; for example, KRAS-mutant CRC118,119.
almost entirely on single time-point paired biopsies and However, others have shown sufficient activity to prompt
non-standardized assays; no study has formally evaluated Phase III trials; for example, MEK162 in NRAS-mutant
a link between target inhibition and clinical response. In advanced-stage melanoma120 and selumetinib in serous
summary, clinical dose and schedule selection for MEKis low-grade ovarian cancer 121, a disease in which MEK162
is mostly based on establishing the maximum tolerated and trametinib are currently undergoing assessment in
doses using continuous or chronic dosing regimens. randomized Phase III trials. Between these extremes,
Dacarbazine MEKis have shown clear but low response rates in most
A DNA-alkylating agent that
has commonly been used as a
Clinical activity as monotherapy. As with preclinical indications, including pancreatic cancer 109,122, biliary
single agent in the treatment of models, BRAF‑V600E‑mutant melanoma has proved tract cancer 123, NSCLC124, uveal melanoma125 and acute
metastatic melanoma. to be the most responsive adult solid tumour to MEKi myeloid leukaemia (AML)126,127, and were generally not

586 | O CTOBER 2015 | VOLUME 15 www.nature.com/reviews/cancer

© 2015 Macmillan Publishers Limited. All rights reserved


REVIEWS

Plexiform neurofibromatosis considered to warrant further monotherapy trials. For response rate and progression-free survival136. A simi-
A non-cancerous example, trametinib failed to show superiority over larly encouraging response rate for trametinib in com-
hyperproliferative disease of docetaxel in a trial in KRAS-mutant-positive NSCLC128. bination with docetaxel in NSCLC137 indicates that this
the nerve sheath that is driven One conclusion from monotherapy trials is that most combination is robustly active in this disease; it remains
by the loss of the NF1 tumour
suppressor.
KRAS-mutant tumours are not highly addicted to ERK to be seen to what extent the antitumour efficacy is
signalling, at least to the degree that it can be inhibited compromised by the tolerability burden. Other combi-
Neoadjuvant by MEKi monotherapy at tolerated doses using chronic nations of MEKi and standard-of-care cytotoxic drugs
A neoadjuvant therapy is dosing schedules. Possible solutions include better patient have fared less well: trametinib and pimasertib failed
treatment given before primary
selection, optimizing dosing schedules and combination to progress beyond Phase II trials in combination with
therapy; in the context of this
Review, selumetinib is
therapy. For example, there is a growing appreciation that gemcitabine in pancreatic cancer 138,139 and selumetinib
administered first as adjuvant relief of feedback RAF phosphorylation and reactivation in combination with irinotecan in KRAS-mutant CRC
therapy and continued until of MEK and ERK may limit the monotherapy activity of showed only modest activity 140. Combinations with
the radioactive iodine therapy the MEKis that do not prevent the phosphorylation novel targeted agents, principally targeting PI3K, AKT
has been administered.
of MEK by RAF (discussed above). Indeed, it is entirely or mTORC1, have required dose modifications, and
possible that such feedback relief and reactivation of MEK although responses have been reported (for example,
is masking the true extent of ‘MEK addiction’ in KRAS- in NSCLC, pancreatic cancer and low-grade ovarian
mutant tumours. Thus, one obvious rational combination cancer) they have not yet shown definitive, confirmed
therapy that should be tested for KRAS-mutant tumours is combination activity in the clinic91,141,142.
intrapathway dual inhibition (MEKi plus RAFi or MEKi Finally, one MEKi combination concept with intrigu-
plus ERKi). ing activity that may come close to fulfilling the criteria
Although responses to MEKi monotherapy have been of synergy, selected patient population and tolerance
modest in adult solid tumours, emerging data from the is the use of a MEKi to re‑sensitize iodine-refractory
use of selumetinib in paediatric diseases — such as low- differentiated thyroid cancer (DTC) to radioactive
grade paediatric glioma and plexiform neurofibromatosis iodine (RAI) therapy. The predominant oncogenic
(neurofibromatosis type 1 (NF1) with inoperable plexi- pathway in DTC, RAS–ERK, drives the loss of thyroid
form neurofibromas) — show an encouraging and dura- differentiation-specific gene expression, including the
ble response rate129,130. Both diseases are characterized by sodium–iodine transporter that takes up RAI. The result
ubiquitous mutational activation of RAS–ERK signalling — inadequate uptake of and response to RAI — can be
and it is tempting to speculate that as low-grade diseases reversed by MEK inhibition, and this has prompted a
these tumours are less able to adapt to MEKis either Phase III neoadjuvant trial in patients with DTC who are
through feedback ‘relief ’ or compensatory pathway acti- at high risk of early relapse143,144.
vation, perhaps owing to low somatic mutation rates, as
reported recently for pilocytic astrocytoma131. Adverse clinical events associated with MEKis. Because
MEK1 and MEK2 are rarely mutated in human cancer
Clinical activity in drug combinations. Randomized there is little prospect of the target-related, genetic thera-
Phase III trials to assess MEKi in combination with first- peutic index for a MEKi that is evident with the first gen-
generation BRAFis in BRAF‑V600E‑mutant melanoma eration of BRAFis in BRAF‑V600E/K melanoma; thus,
have recently reported a clear benefit of the combination normal tissue toxicity will constrain clinical activity, as
not only in efficacy but also in tolerability compared with evidenced by the difference in response rate between
a BRAFi or MEKi alone111,116,117. This successful combi- MEKis and BRAFis in BRAF-mutant melanoma. For
nation paradigm is, however, an outlier due to the two example, trametinib (a MEKi) and dabrafenib (a BRAFi)
drugs synergizing in the tumour while antagonizing each achieved overall response rates of 22% and 53%, respec-
other in normal tissue. Although it is highly unlikely that tively, as single agents145–147. With at least eight MEKis in
this type of combination effect can be reproduced with clinical development, the adverse event profile is becom-
other classes of drugs the result highlights the impor- ing clear and characteristic: acneform rash, diarrhoea
tant characteristics of a successful drug combination: and fatigue are prevalent and generally well managed,
mechanistic synergy, appropriate patient population whereas a range of ocular toxic effects, oedema, creatine
selection and good tolerance. It is worth noting that the phosphokinase elevations and cardiac complications are
combination of a MEKi and a first-generation BRAFi is rarer but more troublesome148. The range of ocular toxic
markedly less active in BRAF-mutant CRC132 in which effects and their management have been reviewed else-
relief of feedback and consequent activation of EGFR where in detail149, and although many of these may be
signalling is thought to limit response, prompting triple self-limiting, some effects require dose discontinuation
combination trials including anti-EGFR mono­clonal or dose modification. Therefore, in conclusion, the dose
antibodies (mAbs) with early, encouraging signs of and target engagement for all MEKis is currently limited
activity 133. Beyond dual pathway inhibition in BRAF- by normal tissue toxic effects.
mutant melanoma the furthest advanced MEKi com- Combinations of MEKis and first-generation BRAFis
bination is selumetinib and docetaxel, which showed are better tolerated than the respective monotherapies
strong preclinical activity 134,135 and is now in Phase III because the paradoxical activation of ERK signalling
trials in KRAS-mutant NSCLC following a small ran- by the BRAFi in BRAF wild-type tissue antagonizes
domized Phase II study in which the combination the pathway inhibition by the MEKi, and the MEKi
showed a benefit over docetaxel alone in terms of overall limits the paradoxical activation of ERK signalling.

NATURE REVIEWS | CANCER VOLUME 15 | O CTOBER 2015 | 587

© 2015 Macmillan Publishers Limited. All rights reserved


REVIEWS

RASness In this way, MEKi-induced skin rash and BRAFi- Future perspectives
Refers to a transcriptomic induced squamous lesions are diminished in combi- The past decade of MEKis clinical development has thrown
signature that reports nation 21. Combining MEKis with standard-of-care into sharp focus the issues that will need to be addressed
activation of the RAS pathway cytotoxic chemotherapy or various targeted agents has if their potential is to be realized. The use of MEKis with
and may predict responses to
RAS pathway drugs.
invariably increased the toxicity burden for patients. a shorter T0.5 at higher intermittent doses should test the
In the case of selumetinib in combination with cyto- hypothesis that a higher degree of pathway inhibition (with
toxic chemotherapy it has been possible to maintain an inhibitor alone or in combination) can improve anti­
the recommended monotherapy dose, albeit with tumour efficacy, delay time to resistance and also main-
enhanced toxicity 136,140,150. By contrast, combinations tain a therapeutic margin by sparing normal tissue157. The
with targeted therapies have generally required dose advent of next-generation sequencing platforms and more
reduction or dosing schedule alteration of one or both sensitive and quantifiable means for testing transcriptome
agents141,151; for example, selumetinib in combination signatures (such as those that measure ERK pathway out-
with MK‑2206, an allosteric AKT inhibitor, or the com- put and RASness)48,158 will provide a step change in the
bination of trametinib and pan-Class I PI3K inhibitor breadth and depth of molecular information around which
buparlisib (BKM120) showed some signs of clinical to develop patient selection biomarkers. Alternative com-
activity but was compromised by dose and schedule bination partners are currently progressing rapidly in clini-
modifications to accommodate the enhanced toxic cal trials. The success of intrapathway combinations to treat
effects that are associated with dual-pathway inhibi- BRAF-V600E/K melanoma in the clinic is mirrored by pre-
tion91,152. In the absence of a strong and specific clini- clinical data on KRAS-mutant tumours combining MEKis
cal efficacy signal it is possible that combination toxic with RAF dimer inhibiting BRAFis159 (WO2008120004)
effects will further limit the clinical development of or with ERKis. Indeed, the entry of ERK1 and ERK2
MEK inhibitors and agents that target PI3K–AKT inhibitors into clinical trials means that ‘on-target’ muta-
signalling. Tolerance to trametinib in combination tions in ERK1 and ERK2 should be anti­cipated as possible
with cytotoxic chemotherapy is more variable: it can mechanisms of acquired resistance160. Intrapathway com-
result in non-tolerance when trametinib is combined binations will be more challenging to develop clinically
with platinum doublet chemotherapy, and it requires owing to the probable added toxicity burden, although
growth factor support when combined with docetaxel, intermittent dosing may mitigate this added burden.
and dose reduction when combined with peme- MEK1 and MEK2 inhibition ultimately regulates the G1/S
trexed153, although it does not require dose reduction in transition, and therefore efficacy may be compromised
combination with gemcitabine138. in tumours with lesions on the cyclin D–CDK4/6–RB
pathway (for example, loss of CDKN2A)86,88,161. Indeed,
Patient populations. Although preclinical studies indi- early clinical data indicate that the combination of MEKis
cate a preferential activity for MEKis in tumours carry- and CDK4/6is holds some promise162. As MEKis predomi-
ing BRAF and RAS mutations47,48,154,155, clinical validation nantly exert cytostatic effects, identifying drug combina-
of this concept is lacking despite the demonstration of tions that harness ‘MEK addiction’ to promote tumour cell
clinical activity in selected patient populations. Even synthetic lethality will prove increasingly important and
the approval of trametinib in BRAFV600E/K-mutant mela- hopefully fruitful. Indeed, this has already been achieved
noma did not demonstrate that trametinib was pref- as a preclinical proof‑of‑principle with BH3 mimetics.
erentially active in BRAF-mutant disease, as patients MEKis modulate the apoptotic threshold by stabiliz-
who did not have mutations in BRAF were not included ing the pro-apoptotic protein BIM; in combination with
in randomized trials95,145. Across many disease types BH3 mimetics that antagonize either BCL-2 or BCL-XL,
MEKis have induced clinical responses in tumours BIM is liberated to neutralize anti-apoptotic BCL‑2 fam-
with activating mutations in KRAS, NRAS, MEK1, ily members such as MCL1, or to directly activate BAX,
GNAQ and GNA11 (guanine nucleotide-binding thus driving tumour cell death163,164 and thereby delaying
protein sub­unit αq and subunit α11) and inactivation the onset of MEKi resistance163. Most importantly, given the
of the RAS GTPase-activating protein NF1 as well as rapid recent advancement and success of tumour immu-
in tumours in which no pathway-related mutation has notherapies, it is vital that the combination of MEKis with
been detected109,118,120–122,125,127,144,145. Progress towards the immune checkpoint inhibitory antibodies (for example,
validation of patient selection biomarkers will require programmed cell death protein 1 ligand 1 (PDL1)-specific
more rigorous trials. Although small non­randomized antibodies that antagonize the immune-suppressive effects
trials of trametinib in combination with either docet­ of PDL1, thereby boosting tumour immunity) is evaluated.
axel or pemetrexed in advanced NSCLC127,128,156 indi- The complex effects of MEKis on the relationship between
cate that KRAS mutation status does not condition the tumour and the immune environment are formida-
the response rate or progression-free survival, the ble, comprising both positive (for example, tumour cell
outcome of ongoing randomized studies in NSCLC death, release of antigens, increased MHC expression and
with prospective KRAS mutation testing should help increased immune system priming) and negative effects
to demonstrate with greater certainty the degree to (for example, inhibitory effects on T cell activation and pro-
which overall clinical benefit, compared with stand- liferation)165. Optimization of dose and schedule and the
ard of care, is dependent upon KRAS mutational sta- identification of patient populations in which RAS–ERK
tus (ClinicalTrials.gov identifiers NCT01750281 and signalling may exert an immune-suppressive effect will be
NCT01933932). the key to success.

588 | O CTOBER 2015 | VOLUME 15 www.nature.com/reviews/cancer

© 2015 Macmillan Publishers Limited. All rights reserved


REVIEWS

1. Yoon, S. & Seger, R. The extracellular signal-regulated 25. Zhang, Z. et al. Dual specificity phosphatase 6 47. Solit, D. B. et al. BRAF mutation predicts
kinase: multiple substrates regulate diverse cellular (DUSP6) is an ETS-regulated negative feedback sensitivity to MEK inhibition. Nature 439,
functions. Growth Factors 24, 21–44 (2006). mediator of oncogenic ERK signaling in lung cancer 358–362 (2006).
2. Karnoub, A. E. & Weinberg, R. A. Ras oncogenes: split cells. Carcinogenesis 31, 577–586 (2010). This paper studied a panel of human cancer cell
personalities. Nat. Rev. Mol. Cell Biol. 9, 517–531 26. Rushworth, L. K. et al. Dual-specificity phosphatase 5 lines and showed that BRAF-mutated cells tended
(2008). regulates nuclear ERK activity and suppresses skin to be far more sensitive to MEK inhibition
3. von Kriegsheim, A. et al. Cell fate decisions are cancer by inhibiting mutant Harvey-Ras (HRasQ61L)- (addicted to MEK1/2) than those harbouring RAS
specified by the dynamic ERK interactome. Nat. Cell driven SerpinB2 expression. Proc. Natl Acad. Sci. USA mutations.
Biol. 11, 1458–1464 (2009). 111, 18267–18272 (2014). 48. Dry, J. R. Transcriptional pathway signatures predict
4. Meloche, S. & Pouysségur, J. The ERK1/2 mitogen- 27. Moriceau, G. et al. Tunable-combinatorial mechanisms MEK addiction and response to selumetinib
activated protein kinase pathway as a master of acquired resistance limit the efficacy of BRAF/MEK (AZD6244). Cancer Res. 70, 2264–2273 (2010).
regulator of the G1‑ to S‑phase transition. Oncogene cotargeting but result in melanoma drug addiction. 49. Joseph, E. W. et al. The RAF inhibitor PLX4032
26, 3227–3239 (2007). Cancer Cell 27, 240–256 (2015). inhibits ERK signaling and tumor cell proliferation in a
5. Balmanno, K. & Cook, S. J. Tumour cell survival 28. Duncan, J. S. et al. Dynamic reprogramming of the V600E BRAF-selective manner. Proc. Natl Acad. Sci.
signalling by the ERK1/2 pathway. Cell Death Differ. kinome in response to targeted MEK inhibition in triple- USA 107, 14903–14908 (2010).
16, 368–377 (2009). negative breast cancer. Cell 149, 307–321 (2012). 50. Heidorn, S. J. et al. Kinase-dead BRAF and oncogenic
6. Bromberg-White, J. L., Andersen, N. J. & The elegant quantitative proteomics approach used RAS cooperate to drive tumor progression through
Duesbery, N. S. MEK genomics in development and in this study reveals the extent of adaptive CRAF. Cell 140, 209–221 (2010).
disease. Brief. Funct. Genom. 11, 300–310 (2012). reprogramming induced by MEK inhibition. 51. Poulikakos, P. I., Zhang, C., Bollag, G., Shokat, K. M. &
7. Ying, Q. L. et al. The ground state of embryonic stem 29. Fawdar, S. et al. Targeted genetic dependency screen Rosen, N. RAF inhibitors transactivate RAF dimers
cell self-renewal. Nature 453, 519–523 (2008). facilitates identification of actionable mutations in and ERK signalling in cells with wild-type BRAF. Nature
8. Hanahan, D. & Weinberg, R. A. The hallmarks of FGFR4, MAP3K9, and PAK5 in lung cancer. Proc. Natl 464, 427–430 (2010).
cancer. Cell 100, 57–70 (2000). Acad. Sci. USA 110, 12426–12431 (2013). 52. Hatzivassiliou, G. et al. RAF inhibitors prime wild-type
9. Hanahan, D. & Weinberg, R. A. Hallmarks of cancer: 30. Johannessen, C. M. et al. COT drives resistance to RAF to activate the MAPK pathway and enhance
the next generation Cell 144, 646–674 (2011). RAF inhibition through MAP kinase pathway growth. Nature 464, 431–435 (2010).
10. Sebolt-Leopold, J. S. & Herrera, R. Targeting the reactivation. Nature 468, 968–972 (2010). 53. Holderfield, M. et al. RAF inhibitors activate the
mitogen-activated protein kinase cascade to treat This is the first demonstration of an alternative, MAPK pathway by relieving inhibitory
cancer. Nat. Rev. Cancer. 4, 937–947 (2004). non-RAF ‘MEK1/2 activator’ driving BRAFi autophosphorylation. Cancer Cell 23, 594–602
11. Montagut, C. & Settleman, J. Targeting the RAF– resistance. (2013).
MEK–ERK pathway in cancer therapy. Cancer Lett. 31. Marusiak, A. A. et al. Mixed lineage kinases activate The studies in references 50–53 identified the
283, 125–134 (2009). MEK independently of RAF to mediate resistance to biochemical bases for paradoxical RAF activation
12. Samatar, A. A. & Poulikakos, P. I. Targeting RAS–ERK RAF inhibitors. Nat. Commun. 5, 3901 (2014). by BRAFi, highlighting the need for MEK inhibition
signalling in cancer: promises and challenges. Nat. 32. Chiariello, M., Marinissen, M. J. & Gutkind, J. S. for more stable pathway inhibition.
Rev. Drug Discov. 13, 928–942 (2014). Multiple mitogen-activated protein kinase signaling 54. Rodriguez-Viciana, P. et al. Germline mutations in
13. Dudley, D. T., Pang, L., Decker, S. J., Bridges, A. J. & pathways connect the Cot oncoprotein to the c‑jun genes within the MAPK pathway cause cardio-facio-
Saltiel, A. R. A synthetic inhibitor of the mitogen- promoter and to cellular transformation. Mol. Cell. cutaneous syndrome. Science 311, 1287–1290
activated protein kinase cascade. Proc. Natl Acad. Sci. Biol. 20, 1747–1758 (2000). (2006).
USA 92, 7686–7689 (1995). 33. Todd, D. E. et al. ERK1/2 and p38 cooperate to induce 55. Bansal, A., Ramirez, R. D. & Minna, J. D. Mutation
This is the first report of a small-molecule inhibitor a p21CIP1-dependent G1 cell cycle arrest. Oncogene 23, analysis of the coding sequences of MEK‑1 and MEK‑2
of MEK1/2. 3284–3295 (2004). genes in human lung cancer cell lines. Oncogene 14,
14. Sebolt-Leopold, J. S. et al. Blockade of the MAP 34. Johnson, G. L., Dohlman, H. G. & Graves, L. M. MAPK 1231–1234 (1997).
kinase pathway suppresses growth of colon tumors kinase kinases (MKKKs) as a target class for small- This is the first report of mutations in MEK1 or
in vivo. Nature Med. 5, 810–816 (1999). molecule inhibition to modulate signaling networks and MEK2 in human cancer.
This is the first paper to demonstrate that a potent gene expression. Curr. Opin. Chem. Biol. 9, 325–331 56. Estep, A. L., Palmer, C. McCormick, F. & Rauen, K. A.
and specific allosteric MEK1/2 inhibitor could (2005). Mutation analysis of BRAF, MEK1 and MEK2 in 15
inhibit the growth of tumour xenografts in vivo. 35. Cuevas, B. D., Abell, A. N. & Johnson, G. L. Role of ovarian cancer cell lines: implications for therapy. PLoS
15. Frémin, C. & Meloche, S. From basic research to mitogen-activated protein kinase kinase kinases in ONE 2, e1279 (2007).
clinical development of MEK1/2 inhibitors for cancer signal integration. Oncogene 26, 3159–3171 (2007). This is the first report of activating MEK1 or MEK2
therapy. J. Hematol. Oncol. 3, 8 (2010). 36. Roskoski, R. MEK1/2 dual-specificity protein kinases: mutations in human cancer.
16. Gilmartin, A. G. et al. GSK1120212 (JTP‑74057) is an structure and regulation. Biochem. Biophys. Res. 57. Marks, J. L. et al. Novel MEK1 mutation identified by
inhibitor of MEK activity and activation with favorable Commun. 417, 5–10 (2012). mutational analysis of epidermal growth factor
pharmacokinetic properties for sustained in vivo 37. Procaccia, S. & Seger, R. in Encyclopedia of Signaling receptor signaling pathway genes in lung
pathway inhibition. Clin. Cancer Res. 17, 989–1000 Molecules (ed. Choi, S.) 1051–1058 (Springer, 2012). adenocarcinoma. Cancer Res. 68, 5524–5528
(2011). 38. Zheng, C. F. & Guan, K. L. Activation of MEK family (2008).
This study provides a detailed characterization of kinases requires phosphorylation of two conserved 58. Murugan, A. K., Dong, J., Xie, J. & Xing, M. MEK1
trametinib, the first MEK1/2 inhibitor to receive US Ser/Thr residues. EMBO J. 13, 1123–1131 (1994). mutations, but not ERK2 mutations, occur in
Food and Drug Administration (FDA) approval. 39. Alessi, D. R. Identification of the sites in MAP kinase melanomas and colon carcinomas, but none in
17. Davies, H. et al. Mutations of the BRAF gene in human kinase‑1 phosphorylated by p74raf‑1. EMBO J. 13, thyroid carcinomas. Cell Cycle 8, 2122–2124
cancer. Nature 417, 949–954 (2002). 1610–1619 (1994). (2009).
This paper provides the first description of 40. Eblen, S. T., Slack, J. K., Weber, M. J. & Catling, A. D. 59. Nikolaev, S. I. et al. Exome sequencing identifies
activating mutations in BRAF in human cancer, Rac-PAK signaling stimulates extracellular signal- recurrent somatic MAP2K1 and MAP2K2
including the particularly high incidence in regulated kinase (ERK) activation by regulating mutations in melanoma. Nat. Genet. 44, 133–139
melanoma. formation of MEK1–ERK complexes. Mol. Cell. Biol. (2011).
18. Bollag, G. et al. Clinical efficacy of a RAF inhibitor 22, 6023–6033 (2002). 60. Arcila, M. E. et al. MAP2K1 (MEK1) mutations define
needs broad target blockade in BRAF-mutant 41. Eblen, S. T. et al. Mitogen-activated protein kinase a distinct subset of lung adenocarcinoma associated
melanoma. Nature 467, 596–599 (2010). feedback phosphorylation regulates MEK1 complex with smoking. Clin. Cancer Res. 21, 1935–1943
This is the first study that attempted to formation and activation during cellular adhesion. (2015).
systematically quantify the extent of RAF–MEK– Mol. Cell. Biol. 24, 2308–2317 (2004). 61. Blasco, R. B. et al. c‑Raf, but not B‑Raf, is essential
ERK suppression required for effective tumour 42. Catalanotti, F. et al. A Mek1–Mek2 heterodimer for development of K‑Ras oncogene-driven non-small
suppression. determines the strength and duration of the Erk cell lung carcinoma. Cancer Cell. 19, 652–663
19. Bollag, G. Vemurafenib: the first drug approved for signal. Nat. Struct. Mol. Biol. 16, 294–303 (2009). (2011).
BRAF-mutant cancer. Nat. Rev. Drug Discov. 11, 43. Bélanger, L. F. et al. Mek2 is dispensable for mouse 62. Karreth, F. A., Frese, K. K., DeNicola, G. M.,
873–886 (2012). growth and development. Mol. Cell. Biol. 23, Baccarini, M. & Tuveson, D. A. C‑Raf is required for the
20. Ascierto, P. A. et al. Phase II trial (BREAK‑2) of the 4778–4787 (2003). initiation of lung cancer by K‑RasG12D. Cancer Discov.
BRAF inhibitor dabrafenib (GSK2118436) in patients 44. Giroux, S. et al. Embryonic death of Mek1‑deficient 1, 128–136 (2011).
with metastatic melanoma. J. Clin. Oncol. 31, mice reveals a role for this kinase in angiogenesis in 63. Alessi, D. R., Cuenda, A., Cohen, P., Dudley, D. T. &
3205–3211 (2013). the labyrinthine region of the placenta. Curr. Biol. 9, Saltiel, A. R. PD 098059 is a specific inhibitor of the
21. Holderfield, M., Deuker, M. M., McCormick, F. & 369–372 (1999). activation of mitogen-activated protein kinase kinase
McMahon, M. Targeting RAF kinases for cancer 45. Cowley, S., Paterson, H., Kemp, P. & Marshall, C. J. in vitro and in vivo. J. Biol. Chem. 270,
therapy: BRAF-mutated melanoma and beyond. Nat. Activation of MAP kinase kinase is necessary and 27489–27494 (1995).
Rev. Cancer. 14, 455–467 (2014). sufficient for PC12 differentiation and for 64. Favata, M. F. et al. Identification of a novel inhibitor of
22. Good, M. C., Zalatan, J. G. & Lim, W. A. Scaffold transformation of NIH 3T3 cells. Cell 77, 841–852 mitogen-activated protein kinase kinase. J. Biol. Chem.
proteins: hubs for controlling the flow of cellular (1994). 273, 18623–18632 (1998).
information. Science 332, 680–686 (2011). 46. Mansour, S. J. et al. Transformation of mammalian 65. Mody, N., Leitch, J., Armstrong, C., Dixon, J. &
23. Ramos, J. W. The regulation of extracellular signal- cells by constitutively active MAP kinase kinase. Cohen, P. Effects of MAP kinase cascade inhibitors on
regulated kinase (ERK) in mammalian cells. Int. Science 265, 966–970 (1994). the MKK5/ERK5 pathway. FEBS Lett. 502, 21–24
J. Biochem. Cell Biol. 40, 2707–2719 (2008). This paper, along with reference 45, demonstrated (2001).
24. Caunt, C. J. & Keyse, S. M. Dual-specificity MAP that phosphomimetic mutations in the activation 66. Squires, M. S., Nixon, P. M. & Cook, S. J. Cell-cycle
kinase phosphatases (MKPs): shaping the outcome of loop caused constitutive activation of MEK1; the arrest by PD184352 requires inhibition of
MAP kinase signalling. FEBS J. 280, 489–504 ‘activated’ MEK1 served as an oncogene when extracellular signal-regulated kinases (ERK) 1/2 but
(2013). expressed in cells. not ERK5/BMK1. Biochem. J. 366, 673–680 (2002).

NATURE REVIEWS | CANCER VOLUME 15 | O CTOBER 2015 | 589

© 2015 Macmillan Publishers Limited. All rights reserved


REVIEWS

67. Ohren, J. F. et al. Structures of human MAP kinase 87. Balmanno, K., Chell, S. D., Gillings, A. S., Hayat, S. & 108. Flaherty, K. T. et al. Combined BRAF and MEK
kinase 1 (MEK1) and MEK2 describe novel Cook, S. J. Intrinsic resistance to the MEK1/2 inhibitor inhibition in melanoma with BRAF V600 mutations.
noncompetitive kinase inhibition. Nat. Struct. Mol. AZD6244 (ARRY‑142886) is associated with weak N. Engl. J. Med. 367, 1694–1703 (2012).
Biol. 11, 1192–1197 (2004). ERK1/2 signalling and/or strong PI3K signalling in This is a key clinical proof-of-principle study
The structural studies of MEK1 and MEK2 in this colorectal cancer cell lines. Int. J. Cancer. 125, demonstrating the potential of dual RAF and MEK
paper reveal the presence of a unique allosteric 2332–2341 (2009). inhibition to cause more durable tumour
inhibitor binding pocket adjacent to, but distinct 88. Halilovic, E. et al. PIK3CA mutation uncouples tumor suppression.
from, the ATP-binding site; this pocket is the growth and cyclin D1 regulation from MEK/ERK and 109. Infante, J. R. et al. Safety, pharmacokinetic,
binding site for the allosteric MEKis. mutant KRAS signaling. Cancer Res. 70, 6804–6814 pharmacodynamic, and efficacy data for the oral MEK
68. Yeh, T. C. et al. Biological characterization of (2010). inhibitor trametinib: a phase 1 dose-escalation trial.
ARRY‑142886 (AZD6244), a potent, highly selective 89. Holt, S. V. et al. Enhanced apoptosis and tumor Lancet Oncol. 13, 773–781 (2012).
mitogen-activated protein kinase kinase 1/2 inhibitor. growth suppression elicited by combination of MEK 110. Musib, L. et al. Clinical pharmacokinetics of
Clin. Cancer Res. 13, 1576–1583 (2007). (selumetinib) and mTOR kinase inhibitors (AZD8055). GDC‑0973, an oral MEK inhibitor, in cancer patients:
69. Fischmann, T. O. et al. Crystal structures of MEK1 Cancer Res. 72, 1804–1813 (2012). data from a Phase 1 study. Cancer Res. 71, 1304
binary and ternary complexes with nucleotides and 90. Roberts, P. J. et al. Combined PI3K/mTOR and MEK (2011).
inhibitors. Biochemistry 48, 2661–2674 (2009). inhibition provides broad antitumor activity in faithful 111. Ribas, A. et al. Combination of vemurafenib and
70. Rinehart, J. et al. Multicenter phase II study of the murine cancer models. Clin. Cancer Res. 18, cobimetinib in patients with advanced BRAFV600-
oral MEK inhibitor, CI‑1040, in patients with advanced 5290–5303 (2012). mutated melanoma: a Phase 1b study. Lancet Oncol.
non-small-cell lung, breast, colon, and pancreatic 91. Tolcher, A. W. et al. Antitumor activity in RAS-driven 15, 954–965 (2014).
cancer. J. Clin. Oncol. 22, 4456–4462 (2004). tumors by blocking AKT and MEK. Clin. Cancer Res. 112. Adjei, A. A. et al. Phase I pharmacokinetic and
71. Smith, C. K. & Windsor, W. T. Thermodynamics of 21, 739–748 (2015). pharmacodynamic study of the oral, small-molecule
nucleotide and non-ATP-competitive inhibitor binding 92. Dai, B. et al. STAT3 mediates resistance to MEK mitogen-activated protein kinase kinase 1/2 inhibitor
to MEK1 by circular dichroism and isothermal titration inhibitor through microRNA miR‑17. Cancer Res. 71, AZD6244 (ARRY‑142886) in patients with advanced
calorimetry. Biochemistry 46, 1358–1367 (2007). 3658–3668 (2011). cancers. J. Clin. Oncol. 26, 2139–2146 (2008).
72. Delaney, A. M., Printen, J. A., Chen, H., Fauman, E. B. 93. Bid, H. K. et al. Development, characterization, and 113. Houédé, N. et al. 600 pharmacokinetics and
& Dudley, D. T. Identification of a novel mitogen- reversal of acquired resistance to the MEK1 inhibitor pharmacodynamics of a selective oral MEK1/2
activated protein kinase kinase activation domain selumetinib (AZD6244) in an in vivo model of inhibitor, pimasertib (MSC1936369B/AS703026), in
recognized by the inhibitor PD 184352. Mol. Cell. childhood astrocytoma. Clin. Cancer Res. 19, patients with advanced solid tumors. Eur. J. Cancer
Biol. 22, 7593–7602 (2002). 6716–6729 (2013). 48, S184 (2012).
73. Friday, B. B. et al. BRAF V600E disrupts 94. Lin, L. et al. The Hippo effector YAP promotes 114. Weekes, C. D. et al. Multicenter Phase I trial of the
AZD6244‑induced abrogation of negative feedback resistance to RAF- and MEK-targeted cancer mitogen-activated protein kinase 1/2 inhibitor BAY
pathways between extracellular signal-regulated therapies. Nat. Genet. 47, 250–256 (2015). 86–9766 in patients with advanced cancer. Clin.
kinase and Raf proteins. Cancer Res. 68, 6145–6153 95. Flaherty, K. T. et al. Improved survival with MEK Cancer Res. 19, 1232–1243 (2013).
(2008). inhibition in BRAF-mutated melanoma. N. Engl. 115. Kim, K. B. et al. Phase II study of the MEK1/MEK2
This is the first publication showing that MEKis J. Med. 367, 107–114 (2012). inhibitor Trametinib in patients with metastatic BRAF-
cause potent MEK1/2 phosphorylation owing to 96. Turke, A. B. et al. MEK inhibition leads to PI3K/AKT mutant cutaneous melanoma previously treated with
‘feedback relief’. activation by relieving a negative feedback on or without a BRAF inhibitor. J. Clin. Oncol. 31,
74. Hatzivassiliou, G. et al. Mechanism of MEK inhibition ERBB receptors. Cancer Res. 72, 3228–3237 482–489 (2013).
determines efficacy in mutant KRAS- versus BRAF- (2012). 116. Long, G. V. et al. Combined BRAF and MEK inhibition
driven cancers. Nature 501, 232–236 (2013). 97. Mirzoeva, O. K. et al. Subtype-specific MEK‑PI3kinase versus BRAF inhibition alone in melanoma. N. Engl.
75. Lito, P. et al. Disruption of CRAF-mediated MEK feedback as a therapeutic target in pancreatic J. Med. 371, 1877–1888 (2014).
activation is required for effective MEK inhibition in adenocarcinoma. Mol. Cancer Ther. 12, 2213–2225 117. Robert, C. et al. Improved overall survival in
KRAS mutant tumors. Cancer Cell 25, 697–710 (2013). melanoma with combined dabrafenib and trametinib.
(2014). 98. Montero-Conde, C. et al. Relief of feedback inhibition N. Engl. J. Med. 372, 30–39 (2015).
This paper and reference 74 demonstrate that of HER3 transcription by RAF and MEK inhibitors 118. Zimmer, L. et al. Phase I expansion and
newer ‘feedback buster’ MEKis prevent the attenuates their antitumor effects in BRAF-mutant pharmacodynamic study of the oral MEK inhibitor
MEK1/2 phosphorylation and associated ERK1/2 thyroid carcinomas. Cancer Discov. 3, 520–533 RO4987655 (CH4987655) in selected patients with
rebound arising from ‘feedback relief’, providing (2013). advanced cancer with RAS–RAF mutations. Clin.
more durable suppression of KRAS-mutant tumour 99. Lee, H. J. et al. Drug resistance via feedback activation Cancer Res. 20, 4251–4261 (2014).
growth. of Stat3 in oncogene-addicted cancer cells. Cancer Cell 119. Bennouna, J. et al. A Phase II, open-label, randomised
76. Pratilas, C. A. et al. V600EBRAF is associated with 26, 207–221 (2014). study to assess the efficacy and safety of the MEK1/2
disabled feedback inhibition of RAF-MEK signaling 100. Fiskus, W. et al. BET protein antagonist JQ1 is inhibitor AZD6244 (ARRY‑142886) versus capecitabine
and elevated transcriptional output of the pathway. synergistically lethal with FLT3 tyrosine kinase monotherapy in patients with colorectal cancer who
Proc. Natl Acad. Sci. USA 106, 4519–4524 (2009). inhibitor (TKI) and overcomes resistance to FLT3‑TKI in have failed one or two prior chemotherapeutic regimens.
77. Lito, P. et al. Relief of profound feedback inhibition of AML cells expressing FLT-ITD. Mol. Cancer Ther. 13, Invest. New Drugs 29, 1021–1028 (2011).
mitogenic signaling by RAF inhibitors attenuates their 2315–2327 (2014). 120. Ascierto, P. A. et al. MEK162 for patients with
activity in BRAFV600E melanomas. Cancer Cell 22, 101. Little, A. S., Smith, P. D. & Cook, S. J. Mechanisms of advanced melanoma harbouring NRAS or Val600
668–682 (2012). acquired resistance to ERK1/2 pathway inhibitors. BRAF mutations: a non-randomised, open-label
78. Ishii, N. et al. Enhanced inhibition of ERK signaling by Oncogene 32, 1207–1215 (2013). Phase 2 study. Lancet Oncol. 14, 249–256 (2013).
a novel allosteric MEK inhibitor, CH5126766, that 102. Emery, C. M. et al. MEK1 mutations confer resistance 121. Farley, J. et al. Selumetinib in women with recurrent
suppresses feedback reactivation of RAF activity. to MEK and B‑RAF inhibition. Proc. Natl Acad. Sci. low-grade serous carcinoma of the ovary or
Cancer Res. 73, 4050–4060 (2013). USA 106, 20411–20416 (2009). peritoneum: an open-label, single-arm, Phase 2 study.
79. Isshiki, Y. et al. Design and synthesis of novel allosteric This is the first report of acquired resistance to Lancet Oncol. 14, 134–140 (2013).
MEK inhibitor CH4987655 as an orally available allosteric MEKis; in this case an emergent 122. Bodocky, G. et al. A Phase II open-label randomized
anticancer agent. Bioorg. Med. Chem. Lett. 21, MEK1 mutation in a patient treated with study to assess the efficacy and safety of selumetinib
1795–1801 (2011). selumetinib. (AZD6244 [ARRY‑142886]) versus capecitabine in
80. Dickson, M. A. Molecular pathways: CDK4 inhibitors 103. Hatzivassiliou, G. et al. ERK inhibition overcomes patients with advanced or metastatic pancreatic
for cancer therapy. Clin. Cancer Res. 20, 3379–3383 acquired resistance to MEK inhibitors. Mol. Cancer cancer who have failed first-line gemcitabine therapy.
(2014). Ther. 11, 1143–1154 (2012). Invest. New Drugs 30, 1216–1223 (2012).
81. Smalley, K. S. et al. Increased cyclin D1 expression can 104. Morris, E. J. et al. Discovery of a novel ERK inhibitor 123. Bekaii-Saab, T. et al. Multi-institutional Phase II study
mediate BRAF inhibitor resistance in BRAF with activity in models of acquired resistance to BRAF of selumetinib in patients with metastatic biliary
V600E‑mutated melanomas. Mol. Cancer Ther. 7, and MEK inhibitors. Cancer Discov. 3, 742–750 cancers. J. Clin. Oncol. 29, 2357–2363 (2011).
2876–2883 (2008). (2013). 124. Hainsworth, J. D. et al. A phase II, open-label,
82. Aguirre, A. J. et al. Activated Kras and Ink4a/Arf 105. Villanueva, J. et al. Concurrent MEK2 mutation and randomized study to assess the efficacy and safety of
deficiency cooperate to produce metastatic pancreatic BRAF amplification confer resistance to BRAF and AZD6244 (ARRY‑142886) versus pemetrexed in
ductal adenocarcinoma. Genes Dev. 17, 3112–3126 MEK inhibitors in melanoma. Cell Rep. 4, 1090–1099 patients with non-small cell lung cancer who have
(2003). (2013). failed one or two prior chemotherapeutic regimens.
83. Carragher, L. A. et al. V600EBraf induces gastrointestinal 106. Corcoran, R. B. et al. BRAF gene amplification can J. Thorac Oncol. 5, 1630–1636 (2010).
crypt senescence and promotes tumour progression promote acquired resistance to MEK inhibitors in 125. Carvajal, R. D. et al. Effect of selumetinib versus
through enhanced CpG methylation of p16INK4a. EMBO cancer cells harboring the BRAF V600E mutation. Sci. chemotherapy on progression-free survival in uveal
Mol. Med. 2, 458–471 (2010). Signal 3, ra84 (2010). melanoma: a randomized clinical trial. JAMA 311,
84. Robinson, J. P. et al. Activated MEK cooperates with 107. Little, A. S. et al. Amplification of the driving 2397–2405 (2014).
Ink4a/Arf loss or Akt activation to induce gliomas oncogene, KRAS or BRAF, underpins acquired 126. Borthakur, G. et al. Phase I/II trial of the MEK1/2
in vivo. Oncogene 30, 1341–1350 (2011). resistance to MEK1/2 inhibitors in colorectal cancer inhibitor GSK1120212 (GSK212) in patients (pts) with
85. Franco, J., Witkiewicz, A. K. & Knudsen, E. S. CDK4/6 cells. Sci. Signal 4, ra17 (2011). relapsed/refractory myeloid malignancies: evidence of
inhibitors have potent activity in combination with This report and reference 106 demonstrate that activity in pts with RAS mutation. J. Clin. Oncol. 29,
pathway selective therapeutic agents in models of amplification of BRAFV600E can drive acquired 4251–4261 (2011).
pancreatic cancer. Oncotarget 5, 6512–6525 (2014). resistance to MEKis, which can be overcome by 127. Jain, N. et al. Phase II study of the oral MEK inhibitor
86. Kwong, L. N. et al. Oncogenic NRAS signaling dual RAF and MEK inhibition. This paper selumetinib in advanced acute myelogenous leukemia:
differentially regulates survival and proliferation in additionally demonstrates that KRAS amplification a University of Chicago Phase II consortium trial. Clin.
melanoma. Nat. Med. 18, 1503–1510 (2012). can also drive MEKi resistance. Cancer Res. 20, 490–498 (2014).

590 | O CTOBER 2015 | VOLUME 15 www.nature.com/reviews/cancer

© 2015 Macmillan Publishers Limited. All rights reserved


REVIEWS

128. Blumenschein, G. Jr et al. A randomized Phase 2 148. Zhao, Y. & Adjei, A. A. The clinical development of MEK 168. Voliotis, M., Perrett, R. M., McWilliams, C.
study of the MEK1/MEK2 inhibitor trametinib inhibitors. Nat. Rev. Clin. Oncol. 11, 385–400 (2014). McArdle, C. A. & Bowsher, C. G. Information transfer
(GSK1120212) compared with docetaxel in KRAS- 149. Duncan, K. E., Chang, L. Y. & Patronas, M. MEK by leaky, heterogeneous, protein kinase signaling
mutant advanced non-small cell lung cancer (NSCLC). inhibitors: a new class of chemotherapeutic agents systems. Proc. Natl Acad. Sci. USA 111, E326–E333
Ann Oncol. 26, 894–901 (2015). with ocular toxicity. Eye 29, 1003–1012 (2015). (2014).
129. Widemann, B. C. et al. Phase I study of the MEK1/2 150. Robert, C. et al. Selumetinib plus dacarbazine versus 169. Ebisuya, M., Kondoh, K. & Nishida, E. The duration,
inhibitor selumetinib (AZD6244) hydrogen sulfate in placebo plus dacarbazine as first-line treatment for magnitude and compartmentalization of ERK MAP
children and young adults with neurofibromatosis BRAF-mutant metastatic melanoma: a phase 2 kinase activity: mechanisms for providing signaling
type 1 (NF1) and inoperable plexiform neurofibromas double-blind randomised study. Lancet Oncol. 14, specificity. J. Cell Sci. 118, 2997–3002 (2005).
(PNs). J. Clin. Oncol. 32, S10018 (2014). 733–740 (2013). 170. Kholodenko, B. N., Hancock, J. F. & Kolch, W.
130. Banerjee, A. et al. A phase 1 study of AZD6244 in 151. LoRusso, P. et al. A first‑in‑human Phase Ib study to Signalling ballet in space and time. Nat. Rev. Mol. Cell
children with recurrent or refractory low-grade evaluate the MEK inhibitor GDC‑0973, combined with Biol. 11, 414–426 (2010).
gliomas: A Pediatric Brain Tumor Consortium report. the pan‑PI3K inhibitor GDC‑0941, in patients with 171. Rajakulendran, T., Sahmi, M., Lefrançois, M.,
J. Clin. Oncol. 32, 10065 (2014). advanced solid tumors. J. Clin. Oncol. 30, S2566 Sicheri, F. & Therrien, M. A dimerization-dependent
131. Jones, D. T. W. et al. Recurrent somatic alterations of (2012). mechanism drives RAF catalytic activation. Nature
FGFR1 and NTRK2 in pilocytic astrocytoma Nat. 152. Kurzrock, R. et al. Phase I dose-escalation of the 461, 542–545 (2009).
Genet. 45, 927–932 (2013). oral MEK1/2 inhibitor GSK1120212 (GSK212) 172. Rushworth, L. K., Hindley, A. D., O’Neill, E. & Kolch, W.
132. Corcoran, R. B. et al. Phase 1–2 trial of the BRAF dosed in combination with the oral AKT inhibitor Regulation and role of Raf‑1/B‑Raf heterodimerization
inhibitor dabrafenib (D) plus MEK inhibitor trametinib GSK2141795 (GSK795). J. Clin. Oncol. 29, S3085 Mol. Cell. Biol. 26, 2262–2272 (2006).
(T) in BRAF V600 mutant colorectal cancer (CRC): (2011). 173. Dougherty, M. K. et al. Regulation of Raf‑1 by direct
updated efficacy and biomarker analysis. J. Clin. 153. Becerra, C. et al. A five-arm, open-label, Phase I/lb feedback phosphorylation. Mol. Cell. 17, 215–224
Oncol. 32, S3517 (2014). study to assess safety and tolerability of the oral (2005).
133. Bendell, J. C. et al. Efficacy and tolerability in an open- MEK1/MEK2 inhibitor trametinib (GSK1120212) in 174. Ritt, D. A., Monson, D. M., Specht, S. I. &
label phase I/II study of MEK inhibitor trametinib (T), combination with chemotherapy or erlotinib in Morrison, D. K. Impact of feedback phosphorylation
BRAF inhibitor dabrafenib (D), and anti-EGFR patients with advanced solid tumors. J. Clin. Oncol. and Raf heterodimerization on normal and mutant
antibody panitumumab (P) in combination in patients 30, S3023 (2012). B‑Raf signaling. Mol. Cell. Biol. 30, 806–819
(pts) with BRAF V600E mutated colorectal cancer 154. Pratilas, C. A. et al. Genetic predictors of MEK (2010).
(CRC). J. Clin. Oncol. 32, S3515 (2014). dependence in non-small cell lung cancer. Cancer Res. 175. Fritsche-Guenther, R. et al. Strong negative feedback
134. Chen, Z. et al. A murine lung cancer co‑clinical trial 68, 9375–9383 (2008). from Erk to Raf confers robustness to MAPK
identifies genetic modifiers of therapeutic response. 155. Jing, J. et al. Comprehensive predictive biomarker signalling. Mol. Syst. Biol. 7, 489 (2011).
Nature 483, 613–617 (2012). analysis for MEK inhibitor GSK1120212. Mol. Cancer 176. McKay, M. M. & Morrison, D. K. Integrating signals
135. Holt, S. V. et al. The MEK1/2 inhibitor, selumetinib Ther. 11, 720–729 (2012). from RTKs to ERK/MAPK. Oncogene 26, 3113–3121
(AZD6244; ARRY‑142886), enhances anti-tumour 156. Kelly, K. et al. Oral MEK1/MEK2 inhibitor trametinib (2007).
efficacy when combined with conventional (GSK1120212) in combination with pemetrexed for 177. Yarden, Y. & Slikowski, M. X. Untangling the ErbB
chemotherapeutic agents in human tumour KRAS-mutant and wild-type (WT) advanced non-small signalling network. Nat. Revs Mol. Cell. Biol. 2,
xenograft models. Br. J. Cancer 106, 858–866 cell lung cancer (NSCLC): a Phase I/Ib trial. J. Clin. 127–137 (2001).
(2012). Oncol. 31, S8027 (2013). 178. Chen, J., Fujii, K., Zhang, L., Roberts, T. & Fu, H. Raf‑1
136. Jänne, P. A. et al. Selumetinib plus docetaxel for 157. Abdel-Wahab, O. et al. Efficacy of intermittent promotes cell survival by antagonizing apoptosis
KRAS-mutant advanced non-small-cell lung cancer: a combined RAF and MEK inhibition in a patient with signal-regulating kinase 1 through a MEK-ERK
randomised, multicentre, placebo-controlled, Phase 2 concurrent BRAF- and NRAS-mutant malignancies. independent mechanism. Proc. Natl Acad. Sci. USA
study. Lancet Oncol. 14, 38–47 (2013). Cancer Discov. 4, 538–545 (2014). 98, 7783–7788 (2001).
137. Gandara, D. R. et al. Oral MEK1/MEK2 inhibitor This is a case study highlighting the use of 179. Mielgo, A. et al. A MEK-independent role for CRAF in
trametinib (GSK1120212) in combination with combined BRAFis and MEKis to control both target mitosis and tumor progression. Nat. Med. 17,
docetaxel in KRAS-mutant and wild-type (WT) BRAF-mutant melanoma and prevent paradoxical 1641–1645 (2011).
advanced non-small cell lung cancer (NSCLC): a acceleration of incipient RAS-mutant tumours. 180. Lorusso, P. M. et al. Phase I and pharmacodynamic
Phase I/Ib trial. J. Clin, Oncol. 31, S8028 (2013). 158. Loboda, A. et al. A gene expression signature of RAS study of the oral MEK inhibitor CI‑1040 in patients
138. Infante, J. R. et al. A randomised, double-blind, pathway dependence predicts response to PI3K and with advanced malignancies. J. Clin. Oncol. 23,
placebo-controlled trial of trametinib, an oral MEK RAS pathway inhibitors and expands the population of 5281–5293 (2005).
inhibitor, in combination with gemcitabine for RAS pathway activated tumors. BMC Med. Genom. 3, 181. LoRusso, P. M. et al. Phase I pharmacokinetic and
patients with untreated metastatic adenocarcinoma 26 (2010). pharmacodynamic study of the oral MAPK/ERK kinase
of the pancreas. Eur. J. Cancer. 50, 2072–2081 159. Lamba, S. et al. RAF suppression synergizes with MEK inhibitor PD‑0325901 in patients with advanced
(2014). inhibition in KRAS mutant cancer cells. Cell Rep. 8, cancers. Clin. Cancer Res. 16, 1924–1937 (2010).
139. Van Cutsem, E. et al. Phase II randomized trial of 1475–1483 (2014). 182. Lee, P. A. et al. Preclinical development of ARRY‑162,
MEK inhibitor pimasertib or placebo combined 160. Goetz, E. M. Ghandi, M., Treacy, D. J., Wagle, N. & a potent and selective MEK 1/2 inhibitor. Cancer Res.
with gemcitabine in the first-line treatment of Garraway, L. A. ERK mutations confer resistance to 70, S2515 (2010).
metastatic pancreatic cancer. J. Clin. Oncol. 33, mitogen-activated protein kinase pathway inhibitors. 183. Chen, X. et al. Combined PKC and MEK inhibition in
S344 (2015). Cancer Res. 74, 7079–7089 (2014). uveal melanoma with GNAQ and GNA11 mutations.
140. Hochster, H. S. et al. Phase II study of selumetinib 161. Xing, F. et al. Concurrent loss of the PTEN and RB1 Oncogene 33, 4724–4734 (2014).
(AZD6244, ARRY‑142886) plus irinotecan as second- tumour suppressors attenuates RAF dependence in 184. Finn, R. S. et al. A Phase I study of MEK inhibitor
line therapy in patients with K‑RAS mutated colorectal melanomas harboring V600EBRAF. Oncogene 31, MEK162 (ARRY‑438162) in patients with biliary tract
cancer. Cancer Chemother. Pharmacol. 75, 17–23 446–457 (2012). cancer. J. Clin. Oncol. 30, S220 (2012).
(2015). 162. Sosman, J. A. et al. A Phase 1b/2 study of LEE011 in 185. Carter, L. et al. ARRY‑162, a novel MEK inhibitor:
141. Bedard, P. L. et al. A Phase Ib dose-escalation study of combination with binimetinib in patients with NRAS- results of a 14‑day Phase 1a study in healthy subjects
the oral pan‑PI3K inhibitor buparlisib (BKM120) in mutant melanoma: early encouraging clinical activity. and a 28‑day Phase 1b study in rheumatoid arthritis
combination with the oral MEK1/2 inhibitor J. Clin. Oncol. 32, S9009 (2014). patients. American College of Rheumatology [online],
trametinib (GSK1120212) in patients with selected 163. Sale, M. J. & Cook, S. J. The BH3 mimetic ABT‑263 https://acr.confex.com/acr/2008/webprogram/
advanced solid tumors. Clin. Cancer Res. 21, synergizes with the MEK1/2 inhibitor selumetinib/ Paper3095.html (2008).
730–738 (2015). AZD6244 to promote BIM-dependent tumour cell 186. Banerji, U. et al. The first‑in‑human study of the
142. Juric, D. et al. A phase 1b dose-escalation study of death and inhibit acquired resistance. Biochem. J. hydrogen sulfate (Hyd-sulfate) capsule of the MEK1/2
BYL719 plus binimetinib (MEK162) in patients with 450, 285–294 (2013). inhibitor AZD6244 (ARRY‑142886): a Phase I open-
selected advanced solid tumors. J. Clin. Oncol. 32, 164. Corcoran, R. B. et al. Synthetic lethal interaction of label multicenter trial in patients with advanced
S9051 (2014). combined BCL‑XL and MEK inhibition promotes tumor cancer. Clin. Cancer Res. 16, 1613–1623 (2010).
143. Chakravarty, D. et al. Small-molecule MAPK inhibitors regressions in KRAS mutant cancer models. Cancer 187. Iverson, C. et al. RDEA119/BAY 869766: a potent,
restore radioiodine incorporation in mouse thyroid Cell 23, 121–128 (2013). selective, allosteric inhibitor of MEK1/2 for the
cancers with conditional BRAF activation. J. Clin. 165. Liu, L. et al. The BRAF and MEK inhibitors dabrafenib treatment of cancer. Cancer Res. 69, 6839–6847
Invest. 121, 4700–4711 (2011). and trametinib: effects on immune function and in (2009).
144. Ho, A. L. et al. Selumetinib-enhanced radioiodine combination with immunomodulatory antibodies 188. Schmieder, R. et al. Allosteric MEK1/2 inhibitor
uptake in advanced thyroid cancer. N. Engl. J. Med. targeting PD1, PD‑L1 and CTLA‑4. Clin. Cancer Res. refametinib (BAY 86–9766) in combination
368, 623–632 (2013). 21, 1639–1651 (2015). with sorafenib exhibits antitumor activity in
145. Falchook, G. S. et al. Activity of the oral MEK inhibitor 166. Sturm, O. E. et al. The mammalian MAPK/ERK preclinical murine and rat models of hepatocellular
trametinib in patients with advanced melanoma: a pathway exhibits properties of a negative feedback carcinoma. Neoplasia 15, 1161–1171 (2013).
Phase 1 dose-escalation trial. Lancet Oncol. 13, amplifier. Sci. Signal 3, ra90 (2010). 189. Lee, L. et al. The safety, tolerability, pharmacokinetics,
782–789 (2012). This study uses simplified mathematical models of and pharmacodynamics of single oral doses of
146. Hauschild, A. et al. Dabrafenib in BRAF-mutated negative feedback in the RAF–MEK–ERK cascade CH4987655 in healthy volunteers: target suppression
metastatic melanoma: a multicentre, open-label, coupled with wet-laboratory data to explain using a biomarker. Clin. Cancer Res. 15, 7368–7374
Phase 3 randomised controlled trial. Lancet 380, resistance to inhibition in normality and why BRAF (2009).
358–365 (2012). mutations render the cascade more susceptible to 190. Leijen, S. et al. Phase I dose-escalation study of the
147. Chapman, P. B. et al. Improved survival with inhibition. safety, pharmacokinetics, and pharmacodynamics of
vemurafenib in melanoma with BRAF V600E 167. Heinrich, R., Neel, B. G. & Rapoport, T. A. the MEK inhibitor RO4987655 (CH4987655) in
mutation. N. Engl. J. Med. 364, 2507–2516 Mathematical models of protein kinase signal patients with advanced solid tumors. Clin. Cancer Res.
(2011). transduction. Mol. Cell 9, 957–970 (2002). 18, 4794–4805 (2012).

NATURE REVIEWS | CANCER VOLUME 15 | O CTOBER 2015 | 591

© 2015 Macmillan Publishers Limited. All rights reserved


REVIEWS

191. Kim, K. et al. Blockade of the MEK/ERK signalling 195. Leonowens, C. et al. Concomitant oral and Acknowledgements
cascade by AS703026, a novel selective MEK1/2 intravenous pharmacokinetics of trametinib, The authors thank J. Sebolt-Leopold who provided informa-
inhibitor, induces pleiotropic anti-myeloma activity in vitro a MEK inhibitor, in subjects with solid tumours. tion as a personal communication and apologize to those
and in vivo. Br. J. Haematol. 149, 537–549 (2010). Br. J. Clin. Pharmacol. 78, 524–532 whose work was not included owing to space limitations.
192. Dong, Q. et al. Discovery of TAK‑733, a potent and (2014).
selective MEK allosteric site inhibitor for the treatment 196. Martinez-Garcia, M. et al. First‑in‑human, Phase I Competing interests statement
of cancer. Bioorg. Med. Chem. Lett. 21, 1315–1319 dose-escalation study of the safety, pharmacokinetics, The authors declare competing interests: see Web version for
(2011). and pharmacodynamics of RO5126766, a details.
193. Nakamura, A. et al. Antitumor activity of the selective first‑in‑class dual MEK/RAF inhibitor in patients with
pan-RAF inhibitor TAK‑632 in BRAF inhibitor-resistant solid tumors. Clin. Cancer Res. 18, 4806–4819
melanoma. Cancer Res. 73, 7043–7055 (2013). (2012). DATABASES
194. Adjei, A. A. et al. Multicenter dose-escalation study of 197. Hoeflich, K. P. et al. Intermittent administration of NCT01750281:
the investigational drug TAK‑733, an oral MEK MEK inhibitor GDC‑0973 plus PI3K inhibitor https://clinicaltrials.gov/ct2/show/NCT01750281
inhibitor in patients with advanced solid tumors; GDC‑0941 triggers robust apoptosis and tumor NCT01933932:
preliminary phase 1 result. Oncology PRO [online], growth inhibition. Cancer Res. 72, 210–219 https://clinicaltrials.gov/ct2/show/NCT01933932
http://oncologypro.esmo.org/Meeting-Resources/ (2012).
ESMO-2012/Multicenter-dose-escalation-study-of-the- 198. El‑Khoueiry, A. et al. A first in‑human Phase I study to SUPPLEMENTARY INFORMATION
investigational-drug-TAK-733-an-oral-MEK-inhibitor- evaluate the MEK1/2 inhibitor GDC‑0623 in patients See online article: S1 (figure)
in-patients-pts-with-advanced-solid-tumors- with advanced solid tumors. Mol. Cancer Ther. 12, ALL LINKS ARE ACTIVE IN THE ONLINE PDF
preliminary-phase-1-results (2012). B75 (2013).

592 | O CTOBER 2015 | VOLUME 15 www.nature.com/reviews/cancer

© 2015 Macmillan Publishers Limited. All rights reserved

You might also like