ArticlePDF AvailableLiterature Review

CDK7 inhibitors as anticancer drugs

Authors:

Abstract and Figures

Cyclin-dependent kinase 7 (CDK7), along with cyclin H and MAT1, forms the CDK-activating complex (CAK), which directs progression through the cell cycle via T-loop phosphorylation of cell cycle CDKs. CAK is also a component of the general transcription factor, TFIIH. CDK7-mediated phosphorylation of RNA polymerase II (Pol II) at active gene promoters permits transcription. Cell cycle dysregulation is an established hallmark of cancer, and aberrant control of transcriptional processes, through diverse mechanisms, is also common in many cancers. Furthermore, CDK7 levels are elevated in a number of cancer types and are associated with clinical outcomes, suggestive of greater dependence on CDK7 activity, compared with normal tissues. These findings identify CDK7 as a cancer therapeutic target, and several recent publications report selective CDK7 inhibitors (CDK7i) with activity against diverse cancer types. Preclinical studies have shown that CDK7i cause cell cycle arrest, apoptosis and repression of transcription, particularly of super-enhancer-associated genes in cancer, and have demonstrated their potential for overcoming resistance to cancer treatments. Moreover, combinations of CDK7i with other targeted cancer therapies, including BET inhibitors, BCL2 inhibitors and hormone therapies, have shown efficacy in model systems. Four CDK7i, ICEC0942 (CT7001), SY-1365, SY-5609 and LY3405105, have now progressed to Phase I/II clinical trials. Here we describe the work that has led to the development of selective CDK7i, the current status of the most advanced clinical candidates, and discuss their potential importance as cancer therapeutics, both as monotherapies and in combination settings. ClinicalTrials.gov Identifiers: NCT03363893; NCT03134638; NCT04247126; NCT03770494.
Content may be subject to copyright.
CDK7 inhibitors as anticancer drugs
Georgina P. Sava
1
&Hailing Fan
1
&R. Charles Coombes
1
&Lakjaya Buluwela
1
&Simak Ali
1
#The Author(s) 2020
Abstract
Cyclin-dependent kinase 7 (CDK7), along with cyclin H and MAT1, forms the CDK-activating complex (CAK), which directs
progression through the cell cycle via T-loop phosphorylation of cell cycle CDKs. CAK is also a component of the general transcription
factor, TFIIH. CDK7-mediated phosphorylation of RNA polymerase II (Pol II) at active gene promoters permits transcription. Cell
cycle dysregulation is an established hallmark of cancer, and aberrant control of transcriptional processes, through diverse mechanisms,
is also common in many cancers. Furthermore, CDK7 levels are elevated in a number of cancer types and are associated with clinical
outcomes, suggestive of greater dependence on CDK7 activity, compared with normal tissues. These findings identify CDK7 as a
cancer therapeutic target, and several recent publications report selective CDK7 inhibitors (CDK7i) with activity against diverse cancer
types. Preclinical studies have shown that CDK7i cause cell cycle arrest, apoptosis and repression of transcription, particularly of super-
enhancer-associated genes in cancer, and have demonstrated their potential for overcoming resistance to cancer treatments. Moreover,
combinations of CDK7i with other targeted cancer therapies, including BET inhibitors, BCL2 inhibitors and hormone therapies, have
shown efficacy in model systems. Four CDK7i, ICEC0942 (CT7001), SY-1365, SY-5609 and LY3405105, have now progressed to
Phase I/II clinical trials. Here we describe the work that has led to the development of selective CDK7i, the current status of the most
advanced clinical candidates, and discuss their potential importance as cancer therapeutics, both as monotherapies and in combination
settings. ClinicalTrials.gov Identifiers: NCT03363893; NCT03134638; NCT04247126; NCT03770494.
Keywords CDK7 .CDK inhibitors .Cell cycle .Transcription .Cancer therapy .Combination therapy
1 Introduction
Cyclin-dependent kinase 7 (CDK7), along with cyclin H and
MAT1, comprises the CDK-activating kinase (CAK), which
provides the T-loop phosphorylation required for activation of
CDKs 1,2, 4 and 6, which drive cell cycle progression
(Table 1, Fig. 1a)[14]. CAK also has a role in the regulation
of transcription, as a component of the general transcription
factor TFIIH. At active gene promoters, CDK7 phosphory-
lates the C-terminal domain (CTD) of RNA polymerase II
(Pol II), at serine 5 (Ser5), to facilitate transcription initiation
(Table 1,Fig.1b)[57]. CDK7 also phosphorylates CDK9,
which in turn phosphorylates the Pol II CTD at Ser2, to drive
transcription elongation [8]. The activities of a variety of tran-
scription factors, including p53 [9,10], retinoic acid receptor
[1113], oestrogen receptor [14,15] and androgen receptor
[16,17], are also regulated by CDK7-mediated phosphoryla-
tion (Table 1).
Because of its dual role in regulating the cell cycle
and transcription, CDK7 has been studied as an antican-
cer drug target, and a number of selective inhibitors of
CDK7 have been developed and investigated as cancer
therapies. Preclinical studies have revealed that cancer
cells can be preferentially targeted by transcriptional in-
hibition, at least in part because they are more reliant
than normal cells on high levels of super-enhancer (SE)-
driven transcription [18,19] mediated by specific onco-
genic drivers, such as RUNX1 in acute lymphoblastic
lymphoma (ALL) [20] and N-MYC in neuroblastoma
[21]. To date, four selective CDK7 inhibitors,
ICEC0942 [22], SY-1365 [23], SY-5609 [24,25]and
LY340515 [26], have progressed to Phase I/II clinical
trial for the treatment of advanced solid malignancies.
In this review we outline the role of CDK7 in both normal
and tumour cells and the rationale for inhibiting CDK7 in
cancer. We also discuss the development of selective CDK7
inhibitors, their mechanism of action in cancer and their po-
tential for use in combination therapies.
*Simak Ali
simak.ali@imperial.ac.uk
1
Division of Cancer, Department of Surgery & Cancer, Imperial
College London, Hammersmith Hospital Campus, London, UK
https://doi.org/10.1007/s10555-020-09885-8
Published online: 8 May 2020
Cancer and Metastasis Reviews (2020) 39:805–823
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
P
P
P
P
P
P
P
P
CDK7
Cyclin H
MAT1
CAK
CDK4/6
Cyclin D
CDK2
Cyclin E
Cyclin A
CDK2
CDK1
Cyclin A
CDK1
Cyclin B
CDK7
Cyclin H
MAT1
TFIIH
Pol II
P
P
P
P
Pol II
P
P
P
P
PCDK9
Cyclin T
a
b
PKCι
CK2
CDK8
Cyclin C
P
P
P
CDK7
Cyclin H
MAT1
Fig. 1 Overview of the regulation of CAK and the role of CDK7 in
regulating the cell cycle (a) and transcription (b). CAK = CDK
activatingkinase, CDK = cyclin-dependent kinase, CK2 = protein kinase
CK2, G1 = gap phase 1, G2 = gap phase 2, M = mitosis, P = phosphate,
PKCι= protein kinase C iota, Pol II = RNA polymerase II, S = synthesis,
TFIIH = transcription factor II H
Table 1 CDK7 substrates
Substrate Residue(s) Possible role(s) Refs
Cell cycle CDK1 Threonine 161 T-loop activation and cyclin binding [1,2]
CDK2 Threonine 160 T-loop activation [1]
CDK4 Threonine 172 T-loop activation [3]
CDK6 Threonine 177 T-loop activation [3]
CDK9 Threonine 186 T-loop activation [4]
Basal transcription RNA Pol II Serine 5 and Serine 7 Transcription initiation (Ser5); Unknown (Ser7) [57]
TFIIB Serine 65 Promotion of transcription [8]
MED1 Threonine 1457 Recruitment to chromatin [9]
Transcription factors AR Serine 515 Activation and turnover [10,11]
E2F1 Serine 403 and Threonine 433 Degradation [12]
ERSerine 118 Activation and turnover [13,14]
Ets1 Threonine 38 Recruitment of coactivators [15]
p53 Serine 33 and a residue between 311 and 393 Enhanced DNA binding (Ser33) [16,17]
PPARSerine 112 Activation [18]
PPARγ2 Serine 12/21 Activation [18]
RARSerine 77 Activation [19,20]
RARγSerine 77/79 Activation [21]
YAP/TAZ Serine 128/90 Prevention of degradation [22]
806 Cancer Metastasis Rev (2020) 39:805–823
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
2 CDK7 function
2.1 CAK structure and regulation
CDK7 is a 346 amino acid kinase, having a predicted
molecular mass of 39 kDa, with an N-terminal cyclin
H-binding region and a C-terminal MAT1 binding re-
gion [27]. A single crystal structure has been reported
forCDK7boundtoATP,intheinactiveconformation,
the structure being similar to that of the inactive con-
formation of ATP-bound CDK2 [28]. Cyclin H binding
is obligatory for CDK7 kinase activity, whilst the addi-
tion of MAT1 stabilises the trimeric CAK complex and
anchors it to TFIIH [27]. In addition, cyclin H and
MAT1bindinghavebeenshowntoregulateCDK7sub-
strate specificity, with the trimeric CDK7-cyclin H-
MAT1 complex having greater kinase activity for Pol
II, in comparison to CDK7-cyclin H, which preferential-
ly phosphorylates CDK2 [2729].
The T-loop of CDK7 can be phosphorylated at two posi-
tions, threonine 170 (Thr170) and Ser164, enhancing both its
kinase activity and ability to bind cyclin H [6]. Furthermore,
T-loop phosphorylation of CDK7 seems to direct substrate
specificity, with Thr170 phosphorylation stimulating activity
towards Pol II over CDK2 [29]. In vitro, CDK1 and CDK2
can phosphorylate CDK7 and as substrates of CDK7 them-
selves; this hints at the possibility of a reinforcement activa-
tion loop between these CDKs [30]. In addition, protein ki-
nase C iota (PKCι), acting downstream of PI3K signalling,
can phosphorylate CDK7 at Thr170 (Fig. 1a)[3135].
Regulation of CAK activity may also be mediated through
phosphorylation of cyclin H. CK2 can activate CAK in vitro,
via phosphorylation of cyclin H at Thr315 (Fig. 1a)[36],
whereas CDK8 has been shown to negatively regulate tran-
scription initiation, via phosphorylation of cyclin H at Ser5
and Ser304 (Fig. 1a)[37]. Furthermore, CDK7 complexed
with cyclin H and/or the trimeric CAK can phosphorylate
cyclin H in vitro. This autophosphorylation reduces activity
of CDK7-cyclin H but has no apparent effect on CDK7-cyclin
H-MAT1 activity. This suggests that MAT1 binding aids
maintenance of the transcriptional activity of CAK by
preventing regulation by cyclin H phosphorylation [28].
An additional means of CDK7 regulation has been ob-
served in mouse neural progenitor cells, where the
microRNA (miRNA) miR-210 regulates cell cycle progres-
sion by modulating expression levels of CDK7 [38]. This
raises the possibility that there may be additional miRNAs
that regulate CAK expression and activity in other cellular
contexts. There is clearly more to be discovered with regard
to the regulation of CDK7 and CAK activity and the identifi-
cation of players acting upstream of CDK7 could potentially
provide additional means by which to manipulate CDK7
activity.
2.2 CDK7 in the cell cycle
CDK7 controls the cell cycle by phosphorylating the cell cycle
CDKs 1, 2, 4 and 6 in their T-loops, to promote their activities
(Fig. 1a)[1]. Both CDK1 and CDK2 are activated by CDK7-
mediated T-loop phosphorylation, at Thr161 and Thr160,
respectively (Table 1)[2,2022,39]. Inhibiting CDK7 during
G1 prevents CDK2 activation and delays S phase, whilst in-
hibition of CDK7 during S/G2 prevents CDK1 activation and
mitotic entry [2,22]. Whilst CDK7 can phosphorylate CDK2
prior to its binding to cyclin, and is not strictly requiredfor the
formation of CDK2-cyclin complexes, CDK7 phosphorylates
CDK1 in concert with cyclin B binding and is required for the
stabilisation of CDK1-cyclin B complexes [2,40].
Full commitment to the cell cycle is controlled at the re-
striction point, through phosphorylation of retinoblastoma
(RB) by CDK4/6-cyclin D, in response to mitogens (Fig.
1a). CDK7 phosphorylates both CDK4 and CDK6 in their
T-loops, at Thr172 and Thr177 (Table 1), respectively, and
CDK7 inhibition prevents their RB kinase activity, halting
G1 progression [3,4]. Although expression levels of the
CAK components remain constant throughout the cell cycle,
T-loop phosphorylation of CDK7 increases when cells are
released from serum starvation [3]. Therefore, a mitogen-
induced cascade of CDK T-loop phosphorylation regulates
progression through G1 [3].
Unlike cyclin-bound CDK2, which remains phosphorylat-
ed for up to 12 hours after CDK7 inhibition, CDK4 and CDK6
activity is rapidly lost following CDK7 inhibition [3]. This
difference is likely due to structural differences between the
complexes; the T-loop of CDK2 is protected from dephos-
phorylation by cyclin binding, whereas the T-loops of cyclin
D-bound CDK4/6 remain exposed to phosphatases [3]. As a
result, CDK7 activity is required to maintain CDK4/6 activity
during G1 whilst being required only for initial activation of
CDK1 and CDK2 during S/G2 [3].
2.3 CDK7 in transcription
CDK7 regulates gene expression, as a component of the gen-
eral transcription factor complex, TFIIH (Fig. 1b). TFIIH is
composed of two distinct sub-complexes: the core complex,
which contains two DNA helicases, xeroderma pigmentosum
type B (XPB) and xeroderma pigmentosum type D (XPD),
along with five other structural and regulatory proteins, and
the CAK complex. CAK is recruited to the core TFIIH com-
plex via a reversible interaction between the ARCH domain of
XPD and the latch domain of MAT1 [41,42]. TFIIH is re-
cruited by TFIIE to active gene promoters, where it joins the
other assembled general transcription factors (TFs), and Pol II,
in the preinitiation complex (PIC) [27]. The composition of
TFIIH and the structure of the PIC have recently been
reviewed by Rimel and Taatjes [27].
807Cancer Metastasis Rev (2020) 39:805–823
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
After DNA is unwound at the transcription start site (TSS)
by XPB [43], Pol II must be released from the PIC to initiate
transcription, in a CDK7-regulated process termed promoter
escape [5]. The CTD of mammalian RPB1, the largest subunit
of Pol II, contains 52 repeats of a heptad sequence,
conformingto the consensus Y1-S2-P3-T4-S5-P6-S7, the res-
idues of which can be sequentially phosphorylated to regulate
Pol II activity throughout the transcription cycle [44]. Whilst
unphosphorylated, Pol II remains anchored to the PIC, via an
interaction with the mediator complex (another PIC compo-
nent) [5]. CDK7 phosphorylates Ser5 and Ser7 of the Pol II
CTD at gene promoters [6,7]; Ser5 phosphorylation facilitates
the release of Pol II from mediator, allowing Pol II to escape
the PIC and initiate transcription (Table 1, Fig. 1b)[5,45]. The
precise function of CDK7-directed Ser7 phosphorylation is as
yet unclear, but evidence suggests that Ser7 phosphorylation
may promote the transcription and post-transcriptional pro-
cessing of small nuclear RNA transcripts, by facilitating an
interaction between the integrator complex and Pol II [46].
After promoter escape, Pol II generally generates a tran-
script of around 2080 bases, before halting progress, in a
process known as promoter-proximal pausing, which likely
functions as a checkpoint to ensure the establishment of a
range of co-transcriptional processes [6,47]. CDK7 is re-
quired for the recruitment of two complexes, the DRB sensi-
tivity inducing factor (DSIF) and the negative elongation fac-
tor (NELF), both of which are required to establish the
promoter-proximal pause [6,8,4850]. For the release of
paused Pol II and commencement of the productive elonga-
tion phase of transcription, the activity of CDK9, as a compo-
nent of the positive transcription elongation factor (P-TEFb),
is required [8]. Like the cell cycle CDKs, for full functionality,
CDK9 must undergo T-loop phosphorylation by CDK7
(Table 1, Fig. 1b)[8]. Therefore, CDK7 plays a role in both
establishing the promoter-proximal pause and in release from
the pause, and inhibition of CDK7 has been shown to increase
the amount of Pol II paused at promoter-proximal regions [6,
51]. Active CDK9 phosphorylates the Pol II CTD, on Ser2,
promoting transcriptional elongation [52]; therefore, there is
an indirect requirement for CDK7 activity after Pol II pause
release.
CDK7 also regulates further transcriptional processes; for
example, CTD phosphorylation by CDK7 allows the co-
transcriptional interaction of Pol II with enzymes that add
the 5-monomethyl-guanosine cap to nascent RNA transcripts
[50]. Additionally, CDK7 is necessary for appropriate tran-
scription termination, with read-through transcription ob-
served upon CDK7 inhibition [6]. CDK12 and CDK13 are
also involved in regulating transcription by phosphorylating
the Pol II CTD during elongation [53]. In vitro, CDK12 can
phosphorylate Ser2, Ser5 and Ser7 [54], whereas CDK13 can
phosphorylate Ser2 and Ser5 [55]. Like the previously
discussed CDKs, T-loop phosphorylation is necessary for
CDK12/13 activation and is likely mediated by CDK7 [54,
56]; thus, it is probable that additional transcriptional sub-
strates of CDK7, and further roles in transcriptional regula-
tion, remain to be identified.
Genetic targeting of Mat1 or Cdk7 in mice is early embry-
onic lethal and cells cultured from embryos of these animals
fail to enter S phase [57,58]. The activities of Cdks 2, 4 and 6
are reduced in mouse embryonic fibroblasts (MEFs) with
Cdk7 knockout, indicating that Cdk7 has an essential role in
cell proliferation [58]. Cdk7 targeting in adult animals results
in phenotypically normal low-proliferating tissues, such as the
liver, kidney or cerebellum. However, in rapidly dividing ep-
ithelial tissues, Cdk7 expression is retained due to tissue re-
newal sustained by stem cells with incomplete Cdk7 knock-
out. This eventually leads to stem cell exhaustion and prema-
ture ageing [58]. Interestingly, MEFs lacking Cdk7 expression
have unaltered Pol II CTD Ser5 phosphorylation and a largely
unchanged gene expression program, indicating that Cdk7 is
dispensable for de novo transcription [58]. This raises the pos-
sibility that another Pol II CTD kinase can compensate for a
lack of Cdk7.
2.4 CDK7 as a regulator of transcription factor activity
Alongside its critical role in directing transcription by Pol II,
CDK7 phosphorylates a number of TFs, functioning to either
promote their activities and/or regulate their degradation
(Table 1). The activity of retinoic acid receptor (RAR)is
promoted by XPD-dependent phosphorylation of Ser77 by
CDK7 [11,13]. Likewise, the activity of RARγis also mod-
ulated by phosphorylation by TFIIH-incorporated CDK7 [12].
CDK7, as part of TFIIH, mediates ligand-dependent phos-
phorylation of oestrogen receptor (ER)atSer118[14,
15], regulating the activity and turnover of the TF [59,60].
Phosphorylation by CDK7, at Ser515 in the transcription ac-
tivation function of androgen receptor (AR), has also been
reported [16,17]. Additionally, CDK7 can phosphorylate
p53 in a MAT1-dependent fashion, at both the C-terminus
(between residues 311 and 393) [10] and the N-terminus, at
Ser33 [9], the former of which has been shown to stimulate
p53 binding to DNA. Evidence that CDK7 phosphorylates
Ets1 [61], peroxisome proliferator-activated receptors
(PPARs) [62] and E2F1 has also been demonstrated, the latter
functioning to trigger E2F1 degradation [63](Table1).
Recently, the stabilisation of the transcriptional regulators
YAP/TAZ was shown to be mediated by CDK7, with phos-
phorylation of YAP at Ser128 and TAZ at Ser90, preventing
their ubiquitination and degradation [64]. At present we have
an incomplete understanding of the role CDK7 plays in regu-
lating the activities of sequence-specific transcriptional regu-
lators. Further knowledge in this area may be helpful in
informing the use of CDK7 inhibitors in specific cellular
contexts.
808 Cancer Metastasis Rev (2020) 39:805–823
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
2.5CDK7inDNArepair
TFIIH plays a key role in the nucleotide excision repair (NER)
pathway [27], which repairs single-stranded DNA damage,
particularly that caused by ultraviolet light. TFIIH is recruited
to damaged DNA, where the NER protein, xeroderma
pigmentosum group A (XPA), catalyses the release of CAK
from the core TFIIH complex, allowing NER to proceed [65].
After DNA repair, CAK reassociates with TFIIH, and the
complex resumes its role in transcription [65]. Inhibition of
CDK7 kinase activity improves NER efficiency, suggesting
that CDK7 negatively regulates NER, directly or indirectly,
via phosphorylation of an as yet unidentified substrate(s) [66].
3 CDK7 in cancer
3.1 CDK7 expression in tumours
Two decades ago, immunohistochemical analyses on a range
of tumour types indicated that CDK7 expression is elevated in
tumour cells compared with their normal counterparts [67].
Since then, numerous studies have provided support for this
finding [6873]. In oestrogen receptor-positive (ER+) breast
cancer, CDK7, cyclinH and MAT1 are overexpressed and are
co-regulated at the mRNA level [68]. Expression of the CAK
components positively correlates with ER expression and
Ser118 phosphorylation, as well as with improved patient out-
comes [68]. Conversely, in triple-negative breast cancer
(TNBC), CDK7 expression is correlated with poor prognosis
[74]. In addition, associations between CDK7 and reduced
survival have been observed in gastric cancer [69,70], ovarian
cancer [75], oral squamous cell carcinoma (OSCC) [71], he-
patocellular carcinoma [72] and glioblastoma [73]. For
OSCC, animal studies have also revealed a potential role for
CDK7 in disease development [71].
These findings raise the possibility that tumours with in-
creased expression of CDK7 may be more sensitive to CDK7
inhibition, particularly in the case ofER+ breast cancer, where
the CDK7-activated nuclear receptor, ER, drives tumour
progression.
3.2 Transcriptional addiction in cancer
Common molecular features of cancer, such as mutation, copy
number changes and genomic rearrangements, can either di-
rectly or indirectly impact gene expression profiles that drive
cancer. For instance, a BRAF mutation in melanoma causes a
cascade of signalling events that ultimately leads to an altered
transcriptional profile and a distinct gene expression signature
[76,77]. Mutations in TF genes are also common in cancer
[78,79]. Across all cancer types, the most frequently mutated
gene (TP53)encodesfortheTFp53[80], and the most
frequently amplified gene, MYC [81], also encodes for a TF.
Other TFs are critical in specific tumour types. For example,
ERactivity drives the majority of breast cancer, and thera-
pies that target ER,liketamoxifen[82] and fulvestrant [83]
are used in the treatment of ER+ breast cancer. Mutation,
rearrangements and deregulated expression of genes encoding
chromatin remodelling and histone modification enzymes,
such as EZH2 and ARID1A, are also frequent in cancer [78,
79]. These aberrations alter the accessibility of gene regulato-
ry regions, ultimately leading to downstream changes in gene
expression.
Recently, clusters of enhancers, termed super-enhancers
(SE), that control the expression of genes integral for cell
identity and function have been defined [84]. Deregulation
of the SE landscape is common in cancer and leads to dramatic
changes in gene expression and high transcriptional outputs,
which maintain the oncogenic cell state (Fig. 2). As a result,
cancer cells become transcriptionally addicted, requiring
higher levels of transcription than normal cells to sustain
growth [19]. The phenomenon of transcriptional addiction
suggests that cancer cells may be more responsive than normal
cells to transcriptional inhibition and provides a strong
basis for targeting transcriptional kinases, including CDK7,
in cancer (Fig. 2)[18]. Furthermore, oncogenic TFs, like
MYC, have proven notoriously difficult to target directly with
small molecules; therefore, the ability to target the general
transcription machinery to reduce their transcriptional output
is an attractive prospect.
4 Development of CDK7 inhibitors
4.1 Pan-CDK inhibitors
Early efforts to develop CDK inhibitors yielded relatively un-
selective compounds, with activities against multiple CDKs,
often including CDK7 [85]. The first CDK inhibitor to enter
clinical trial was the semi-synthetic flavone derivative,
alvocidib (flavopiridol; Fig. 3a), which inhibits CDK1, 2, 4,
6,7and9(Table2)[8790]. Between 2008 and 2014,
alvocidib was evaluated in more than 60 clinical trials for
numerous tumour types [91]. Limited clinical activity was
seen in the majority of trials, however, modest responses
against chronic lymphocytic leukaemia (CLL) [92,93]and
mantle cell lymphoma [94] were shown. Currently, alvocidib,
marketed as a CDK9 inhibitor, is being trialled by Tolero
Pharmaceuticals for the treatment of acute myeloid leukaemia
(AML) (Clinicaltrials.gov identifiers: NCT03298984;
NCT03969420; NCT02520011). Another early pan-CDK in-
hibitor, the purine-based seliciclib (roscovitine; Fig. 3a),
which inhibits CDK1, 2, 5, 7, and 9 (Table 2)[9597], was
also assessed in clinical trials for a variety of tumour types but,
likewise, showed limited clinical activity [91,98].
809Cancer Metastasis Rev (2020) 39:805–823
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Attempts to develop CDK inhibitors with improved selec-
tivity for CDK1 and CDK2 led to a second generation of
multi-target CDK inhibitors, including the aminothiazole-
based compound, SNS-032 (Fig. 3a), which potently inhibits
CDK2, 7 and 9 (Table2)[85,91,99,100]. Although SNS-032
has been trialled for the treatment of advanced lymphoid [101]
and advanced solid malignancies [102], the drug has not
progressed further than Phase I [91].
The inability of these early CDK inhibitors to selectively
target individual CDK family members probably contributed
to their failure in the clinic. As several CDK proteins are
critical for the function of normal tissues, the promiscuity of
Alvocidib Seliciclib SNS-032
YKL-5-124
SY-1365
ICEC0942
LDC4297
LY3405105
BS-181
a
b
THZ2
THZ1
Fig. 3 Chemical structures of selected inhibitors that target CDK7. Chemical structures of non-specific inhibitors of CDK7 (a) and selective inhibitors of
CDK7 (b). (The chemical structures of QS1189 and SY-5609 have not been disclosed)
gene
mRNA
enhancer
Super-enhancer
Normal
CDK7iCDK7i
Cancer
CDK7i CDK7i
CDK7i
ab
Fig. 2 Super-enhancer-driven genederegulation in cancer can be targeted
by CDK7 inhibitors. The super-enhancer landscape in normal cells (a)
becomes deregulated in cancer (b), leading to altered gene expression.
CDK7 inhibitors preferentially reduce gene expression driven by super-
enhancers in cancer cells compared with normal cells (A and B). CDK7i
= CDK7 inhibitor
810 Cancer Metastasis Rev (2020) 39:805–823
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
these compounds likely limits their ability to discern cancer
cells from normal cells, resulting in a narrow therapeutic win-
dow and associated toxicities, which include fatigue, diar-
rhoea, nausea and hyperglycaemia [91,102105]. In addition,
their lack of specificity makes it difficult to decipher which
CDKs are inhibited in vivo, and which are most important for
their underlying mechanism of action [91]. This paucity of
knowledge limits the potential to develop these pan-CDK in-
hibitors further as targeted therapies.
More recent efforts have focused on further improving the
selectivity of CDK inhibitors, with selective CDK4/6 inhibi-
tors proving the biggest success story to date. Three CDK4/6
inhibitors have been approved for the treatment of hormone
receptor (HR)-positive metastatic breast cancer: palbociclib
(PD0332991; Ibrance), ribociclib (LEE011; Kisquali) and
abemaciclib (LY2835219; Verzenio), in combination with
aromatase inhibitors or fulvestrant [106]. More than one hun-
dred clinical trials for CDK4/6 inhibitors in breast, but also in
other cancers, including glioma, sarcoma, lung, pancreatic,
head and neck, colorectal, prostate and ovarian cancer, are
actively recruiting patients or about to initiate. The success
of these selective CDK4/6 inhibitors is encouraging and pro-
vides some confidence that selective inhibitors of other CDKs
may prove similarly successful.
4.2 CDK7-specific inhibitors
A number of selective small molecule inhibitors of CDK7
have been developed. These include the pyrazolopyrimidine
derivatives, BS-181 [39] and ICEC0942 [22,107], and the
pyrazolotriazine derivatives, LDC4297 [108] and QS1189
[109](Fig.3b,Table3). These are type I inhibitors that bind
reversibly to the ATP-binding site of CDK7. ATP-competitive
covalent inhibitors of CDK7 have also been developed, in-
cluding the pyrimidine based THZ1 [20] and SY-1365 [23]
and the pyrrolidinopyrazole based YKL-5-124 [130](Fig.3b,
Table 3).
The first example of a highly selective CDK7 inhibitor was
BS-181, which is structurally related to the pan-CDK inhibitor
roscovitine (Fig. 3b, Table 3)[39]. BS-181 reduced phosphor-
ylation of CDK7 targets and impaired cancer cell line and
xenograft tumour growth, establishing CDK7 as a putative
cancer drug target [39]. Although in vivo activity was demon-
strated, poor bioavailability and insufficient cell permeability
precluded the development of BS-181 as a clinical candidate
[39].
Efforts to develop BS-181 analogues which retain CDK7
selectivity, but have improved drug-like properties, led to the
first orally bioavailable CDK7 inhibitor, ICEC0942 (CT7001;
Fig. 3b, Table 3)[22,107]. Although crystal structures of
CDK7 bound to ICEC0942 could not be obtained, a crystal
structure of CDK2 in complex with ICEC0942 was solved
[107]. Using this structure as a starting point, modelling stud-
ies revealed aspartate 155 (Asp155) as a residue that is likely
key in determining the selective binding of ICEC0942 to
CDK7 [107]. ICEC0942 potently inhibited the growth of a
panel of cancer cell lines and of ER+ breast cancer xenografts,
and its favourable absorption, distribution, metabolism, and
excretion (ADME) and pharmacokinetic (PK) properties
made ICEC0942 a promising clinical candidate [22]. The drug
was licenced to Carrick Therapeutics and is now in Phase I/II
clinical trial for advanced solid malignancies, with focused
cohorts of breast and prostate cancer patients (Table 4).
A number of covalent CDK7 inhibitors have also been
developed, the first being THZ1 (Fig. 3b, Table 3), which
targets a cysteine residue (Cys312) on a C-terminal extension
just outside the ATP-binding site of CDK7 [20,132] and has
strong activity in many cancer types (Table 3)[20,21,71,75,
114124]. However, THZ1 also covalently links to CDK12
and CDK13, at Cys1039 and Cys1017, respectively,
inhibiting their activity [20,132]. THZ1 has been widely
employed as a tool to interrogate CDK7 function [50,51];
however, it was recently shown that its anti-transcriptional
and antitumour activities are reliant on inhibition of CDK12
Table 2 Characteristics of selected multi-target CDK7 inhibitors (see ref [86])
Name(s) Company
a
IC
50
(nM)
b
Development
phase reached
Alvocidib (flavopiridol) Tolero Pharmaceuticals
(Sanofi-Aventis)
CDK1-CycB= 41; CDK2-CycA = 100; CDK4-CycD= 65;
CDK6-CycD= ~100; CDK7-CycH = ~300; CDK9-CycT = 6
Phase II
Seliciclib (roscovotine;
CYC202)
Cyclacel (ManRos
Therapeutics)
CDK1-CycB= 2700; CDK2-CycE = 100; CDK4-CycD1 > 10,000;
CDK6-CycD1> 100,000; CDK7-CycH = 490; CDK9-CycT = 600
Phase II
SNS-032 Sunesis (Bristol-Myers
Squibb)
CDK1-CycB= 480; CDK2-CycA = 38; CDK4-CycD= 92;5
CDK6-CycD>1000; CDK7-CycH = 62; CDK9-CycT = 4
Phase I
a
Current developer (previous developer in brackets)
b
Data from in vitro kinase assays for CDKs 1, 2, 4, 6, 7, 9 and 12 have been listed,where available. Where data are available for a CDK in multiple cyclin
complexes, the complex with the lowest IC
50
is presented
811Cancer Metastasis Rev (2020) 39:805–823
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Table 3 Characteristics of key CDK7 inhibitors in development
Name(s) Company Type of
inhibitor
IC
50
(nM)
a
Activity in preclinical models Combination agents tested Current development
phase (clinical Trial
ID)
BS-181 [39] - Non-covalent CDK1-CycB = 8100; CDK2-CycE = 880;
CDK4-CycD1 = 33,000; CDK6-CycD1 = 47,000;
CDK7-CycH-Mat1 = 21; CDK9-CycT = 4200
ER+ breast cancer [39], gastric cancer
[110], papillary thyroid cancer [111]
-
ICEC0942 [22]
(CT7001)
CarrickTherapeutics Non-covalent CDK1-CycA1 = 1800; CDK2-CycA1 = 620;
CDK4-CycD1 = 49,000; CDK6-CycD1 = 34,000;
CDK7-CycH-MAT1= 40; CDK9-CycT1 = 1200
ER+ breast cancer [22], AML [112] Fulvestrant, tamoxifen [22]Phase
I/II (NCT03363893)
LY3405105 [26] Eli Lilly and Company - CDK1-CycB1 = 20,000; CDK2-CycE1 = 20,000;
CDK4-CycD1 = 2830; CDK6-CycD1 = 8079;
CDK7-CycH-Mat1 = 92.8; CDK9-CycT1 = 6320;
CDK12-CycK = 14,780
- - Phase I (NCT03770494)
LDC4297 [108] Lead Discovery Center
GmbH
Non-covalent CDK1-CycB = 54; CDK2-CycE = 6.4; CDK4-CycD 1000;
CDK6-CycD > 1000; CDK7-CycH-MAT1 < 5;
CDK9-CycT = 1711
HCMV antiviral activity [113]- -
QS1189 [109] Qurient Therapeutics Non-covalent CDK1-CycE1 = 690; CDK2-CycE1 = 270;
CDK4-CycD1 = 3700; CDK6-CycD1 = 6200;
CDK7-CycH-MAT1= 15; CDK9-CycK = 710;
CDK12-CycK = 570
Mantle cell lymphoma, Burkitts
lymphoma, DLBCL [109]
--
SY-5609 [24] Syros Pharmaceuticals Non-covalent CDK2-CycE1 = 2900
c
; CDK7-CycH-MAT1 = 0.06
b
;
CDK9-CycT1 = 970
c
; CDK12-CycK = 770 nM
c
ER+ breast cancer [25], ovarian cancer
[24], TNBC [24]
Fulvestrant [25]-
SY-1365 [23] SyrosPharmaceuticals Covalent CDK2-CycE1 = 2117; CDK7-CycH-MAT1 = 84;
CDK9-CycT1 = 914; CDK12-CycK = 204
AML [23] Venetoclax [23] Phase I (NCT03134638)
THZ1 [85] (SY-079) Syros Pharmaceuticals Covalent CDK7-CycH = 3.2 T-ALL [20], neuroblastoma [21], SCLC
[114], OSCC [71], PTCL [115],
ovarian cancer [75], DIPG [116], HGG
[117], melanoma [118], hepatocellular
carcinoma [119], thyroid cancer [120],
pancreatic cancer [121], cervical cancer
[122], TNBC [123], multiple myeloma
[124]
Fulvestrant [125], JQ1 [75,116,126],
panobinostat [116],
carfilzomib/bortezomib [124],
venetoclax [124]/navitoclax [127], 5-
-fluorouracil, nutlin-3 [128]
-
THZ2 [123] Syros Pharmaceuticals Covalent CDK1-CycB = 97; CDK2-CycA = 222; CDK7-CycH = 14;
CDK9-CycT = 194
TNBC [123], gastric cancer [129]- -
YKL-5-124 [130] Syros Pharmaceuticals Covalent CDK7-CycH-MAT1 = 9.7; CDK2-CycA = 1300;
CDK9-CycT1 = 3020
Mantle cell lymphoma [130] anti-PD-1+chemotherapy [131]-
a
Data from in vitro kinase assays for CDKs 1, 2, 4, 6, 7, 9 and 12 have been listed, where available. Where data are available for CDKs in complex with multiple cyclins, the complex with the lowest IC
50
is
presented.
b
K
d
determined by SPR.
c
K
i
determined by activity assay
812 Cancer Metastasis Rev (2020) 39:805–823
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Table 4 Summary of clinical trials investigating CDK7 inhibitors in cancer
Drug name(s) Clinical trial
ID
Dates Administration Trial type Trial design No. of
patients
Combination
agent
Results
ICEC0942
(CT7001)
NCT03363893 Nov
2017March
2021
Orally once daily Modular
Phase I/II
Module
1A
Dose-escalation/safety advanced solid tumours 39 MBAD = 120 mg
once daily
Module
1B
Refine dose-escalation/safetyup to 4 cohorts: MTD =360 mg
once daily
Locally advanced or metastatic TNBC Up to 50
Castrate-resistant prostate cancer Up to 25
Additional cohorts (may include ovarian and SCLC) Up to 25
Module
2
Phase Ib/II safety and efficacy
Locally advanced or metastatic HR+HER2breast
cancer
Up to 75 Fulvestrant
LY3405105 NCT03770494 Jan 2019May
2022
Orally Phase Ia/Ib Safety advanced or metastatic solid tumours Up to 215
SY-1365 NCT03134638 May 2017Nov
2019
Intravenously
once/twice weekly
Phase I
(2 parts)
Part 1 Dose-escalation/safety advanced solid tumours ~ 35
Part 2 Refine safety and test efficacy5 cohorts:
Ovarian cancer treated with 3 prior lines of therapy ~ 24
Relapsed ovarian cancer with previous platinum
therapy
~ 24 Carboplatin
Primary platinum refractory ovarian cancer ~ 12
Biopsy-accessible advanced solid tumours 2030
HR+metastatic breast cancer post CDK4/6 + aroma-
tase inhibitor treatment
~ 12 Fulvestrant
SY-5609 NCT04247126 Jan 2020Jun
2021
Orally Phase I Dose-escalation select advanced solid tumours 60
HER2 human epidermal growth factor receptor 2, HR hormone receptor, MBAD minimum biologically active dose, MTD maximum tolerated dose, SCLC small cell lung cancer, TNBC triple-negative
breast cancer
813Cancer Metastasis Rev (2020) 39:805–823
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
and CDK13, in addition to CDK7 [130]. Consequently, YKL-
5-124 was developed, with a strategy that combined the co-
valent warhead of THZ1 with the pyrrolidinopyrazole core of
the PAK4 inhibitor, PF-3758309 [130]. Like THZ1, YKL-5-
124 covalently links to Cys312 of CDK7 but does not affect
the activities of CDK12 and 13 (Fig. 3b, Table 3)[130]. An
analogue of THZ1, with altered regiochemistry of the acryl-
amide and increased in vivo stability has also been developed
and was designated THZ2 (Fig. 3b, Table 3)[123].
To improve on the potency, selectivity and metabolic sta-
bility of THZ1, the THZ1-derived CDK7 inhibitor, SY-1365,
was developed by Syros Pharmaceuticals as a candidate for
clinical development [23]. SY-1365 entered Phase I clinical
trial for the treatment of advanced solid tumours, with planned
expansion cohorts focusing on ovarian cancer and breast can-
cer (Fig. 3b, Table 3and 4). However, Syros Pharmaceuticals
recently announced discontinuation of the clinical develop-
ment of SY-1365 and the prioritisation of a new, orally avail-
able CDK7 inhibitor, SY-5609 (Table 3), with greater selec-
tivity and potency for CDK7 [24,25]. SY-5609 has
antitumour activity in preclinical models of ovarian cancer
[24,25], TNBC [24,25] and ER+ breast cancer, in combina-
tion with fulvestrant [25], and sustained tumour regressions
were associated with alterations in the RB pathway [25]. A
Phase I trial, in patients with select advanced solid tumours,
began in early 2020 (Table 4).
Another CDK7 inhibitor, LY3405105, developed byEli Lilly,
is also undergoing clinical testing for advanced or metastatic
solid cancers [26](Fig.3b,Table3and 4). Little information
on LY3405105 has been released; however, selectivity data from
the corresponding patent (WO2019099298) is listed in Table 3.
5 Inhibiting CDK7 in cancer
Due to the importance of CDKs in regulating cell prolifera-
tion, and the deregulation of CDK pathways in many cancer
types, CDKs have long been considered important targets for
the design of cancer therapeutics [39]. The early pan-CDK
inhibitors, alvocidib and seliciclib, cause cell cycle arrest
and apoptosis, as well as altered expression of genes in these
pathways [85]. Seliciclib was also shown to reduce Pol II
CTD phosphorylation and Pol II-dependent transcription in
myeloma cells [133]. Whilst it is likely that some cellular
actions of these inhibitors are mediated through CDK7, their
lack of selectivity made it difficult to distinguish CDK7-
specific effects and ultimately led to the numerous side effects
that resulted in their failure in the clinic. The recently devel-
oped, highly specific inhibitors of CDK7 have been instru-
mental in revealing the potential of CDK7 as a cancer drug
target. Xenograft studies in mice showed that CDK7 inhibitors
are well tolerated and effective at reducing tumour growth
in vivo [2123,39,114].
A number of reversible and covalent inhibitors of
CDK7 have been tested on large panels of cancer cell
lines [20,22,23,39].ScreeningofICEC0942against
the NCI-60 cancer cell line panel demonstrated a medi-
an GI
50
value of 250 nM [22]. Of over 1000 cancer cell
lines tested with THZ1, around half had a GI
50
value
under 200 nM [20]. Whilst CDK7 inhibitors are potent
at impairing the growth of many cancer cell lines,
representing a variety of tumour types, it is clear that
some cell lines respond more favourably than others.
Responses of 386 human cell lines, encompassing 26
cancer types, to the covalent CDK7 inhibitor SY-1365,
revealed varied responses ranging from cytostatic to
highly cytotoxic [23]. Expression levels of the anti-
apoptotic protein BCL-XL were predictive of SY-1365
response, with low BCL-XL expression associated with
high SY-1365 sensitivity [133]. It is clear that additional
features associated with response to CDK7 inhibition
remain to be discovered and this knowledge will likely
be beneficial for their future clinical success.
5.1 Effects on cell cycle progression
As CDK7 directs cell cycle progression, via the activation of
other CDK proteins, it is unsurprising that CDK7 inhibitors
reduce phosphorylation of cell cycle CDKs and consequently
cause cell cycle arrest [21,22,39,108,109,130]. CDK7
inhibition with ICEC0942 blocks progression at all stages of
the cell cycle and is associated with a reduction in phosphor-
ylation of CDK1, CDK2 and RB [22]. YKL-5-124 primarily
causes G1 arrest, with a reduction in CDK2 phosphorylation
[130], whereas THZ1 and QS1189 both arrest cells at G2/M
[21]. Interestingly, the extent and timing of cell cycle arrest
upon treatment with an individual CDK7 inhibitor can vary
among cell lines; A549 lung cancer cells, treated with
LDC4297, arrested in G1, whereas HCT116 colon cancer
cells exhibited a G2/M delay, only after an extended incuba-
tion period [108]. Again, it is apparent that factors influencing
the effect of CDK7 inhibition on cell cycle progression across
different cancers remain to be identified and these may have
an important impact on the clinical use of these inhibitors. In
addition to cell cycle arrest, apoptosis is observed following
CDK7 inhibition, in numerous cancer types, including solid
tumours [2123,39,108,114] and haematological malignan-
cies [20,109]. YKL-5-124 is unique among the current CDK7
inhibitors, in that it causes cell cycle arrest at both G1 and G2/
M in the apparent absence of apoptosis [130].
5.2 Effects on transcription
The majority of CDK7 inhibitors, again, with YKL-5-124
being an exception, reduce Pol II CTD phosphorylation and
cause widespread alterations in Pol II-mediated transcription
814 Cancer Metastasis Rev (2020) 39:805–823
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
[20,22,23,39,108,109,130]. As CDK7 is dispensable for
global transcription [58], it makes sense that rather than reduc-
ing gene expression globally, only subsets of genes are down-
regulated by CDK7 inhibitors. Pathway analyses suggest that
cell cycle and DNA damage repair pathways are enriched in
gene sets altered by CDK7 inhibition [23,109,114].
Additionally, smaller subsets of genes are actually upregulated
following CDK7 inhibition [21,108]. For example, around
half of the 2% of genes whose expression was affected by
short-term LDC4297 treatment were upregulated [108].
mRNAs have differing half-lives, and it has been noted that
those with short half-lives are preferentially downregulated by
CDK7 inhibitors [108]. One explanation for the counterintui-
tive observation that some genes are upregulated following
CDK7 inhibition is that these represent a subset that are neg-
atively regulated by genes whose transcripts have short half-
lives. It is also possible that CDK7 plays a direct and crucial
role in regulating the transcription of specific subsets of genes,
possibly also repressing the expression of some genes, either
in a direct or an indirect fashion. One clue that may go some
way to explaining how the transcription of certain genes may
escape CDK7 inhibition comes from the work of Shandilya
et al. [134]. They showed that CDK7 phosphorylates the gen-
eral transcription factor, TFIIB (Table 1), and this is required
for the transcription of certain genes, but not for transcription
of p53 target genes, which escape CDK7 inhibition.
Numerous studies have shown that SE-associated genes
are preferentially downregulated in cancer cells treated
with CDK7 inhibitors (Fig. 2)[20,21,7173,75,114,
116,118,123,135,136]. In T cell acute lymphoblastic
leukaemia (ALL), expression of the oncogenic TF,
RUNX1, is driven by a large SE and is disproportionately
repressedbyTHZ1treatment[20]. Similarly, neuroblasto-
mas driven by MYCN amplification, which promotes the
formation of aberrant SEs, are selectively sensitive to
CDK7 inhibition [21]. As MYC TFs are frequently upreg-
ulated in cancer but have proven difficult to target direct-
ly, blocking MYC expression and/or targeting the tran-
scription machinery downstream of MYC, by inhibiting
CDK7, is an attractive strategy. Similarly, oncogenic
ETS TFs, which are also notoriously difficult to target,
are reduced by THZ1 treatment in prostate cancer [137].
Other cancers in which THZ1-mediated downregulation
of SE-associated genes has been demonstrated include
ovarian cancer [75], melanoma [118] and small cell lung
cancer [114]. These studies have been integral for
uncovering the aforementioned phenomenon of transcrip-
tional addiction in cancer (Fig. 2). Furthermore, targeting
specific oncogenic transcriptional programs provides
some explanation for the observation that cancer cells
are more vulnerable to transcriptional inhibition than nor-
mal cells.
5.3 CDK7 inhibitors to treat drug-resistant cancers
CDK7 inhibition represents a novel strategy to treat cancers
with de novo or acquired resistance to other drugs, where
further treatment options are limited. Mutations in the ESR1
gene are common in advanced ER+ breast cancer, causing
oestrogen-independent receptor activation and hormone ther-
apy resistance [125]. CDK7 is an essential gene in both ER-
wild-type and ER-mutant breast cancer, and ER-mutant
MCF7 cells, that are partially resistant to anti-oestrogens, are
sensitive to CDK7 inhibition [125,138]. Furthermore, activat-
ing Ser118 phosphorylation of mutant ER is inhibited by
THZ1 [125,138]. This suggests that CDK7 inhibitors may
be effective at treating advanced, ER-mutant breast cancer.
THZ1 has also been shown to overcome HER2 inhibitor re-
sistance in breast cancer [139] and venetoclax resistance in
mantle cell lymphoma [109], and to inhibit castration-
resistant prostate cancer [137].
A number of CDK4/6 inhibitors are clinically approved for
use in combination with endocrine therapies for the treatment
of ER+ breast cancer, however, resistance to these drugs is an
emerging problem [140]. Breast cancer cells with acquired
resistance to palbociclib remain sensitive to THZ1 [140], sug-
gesting that CDK7 inhibitors may be useful following the
onset of resistance to drugs that target other CDKs.
Current clinical trials of CDK7 inhibitors are aimed at pa-
tients with advanced or metastatic cancer; therefore, a majority
of these will have received other lines of therapy prior to their
recruitment. The SY-1365 trial was designed to target specific
cohorts of patients with resistance to prior treatments, includ-
ing platinum resistance in ovarian cancer and CDK4/6 inhib-
itor plus aromatase inhibitor resistance in HR+ breast cancer
(Table 4). The prospect of using CDK7 inhibitors to overcome
resistance to prior treatments in cancer is exciting, and it is
crucial that information garnered from preclinical studies of
CDK7 inhibitors in the context of acquired drug-resistance
continues to be considered during clinical trial design.
5.4 Combination treatment strategies
Cancers are frequently treated with two or more therapeutic
agents simultaneously. Compared with single-agent treat-
ments, these combination therapies often have enhanced effi-
cacy, can delay the onset of resistance and may allow lower
doses of individual drugs to be used, thus reducing toxicity.
Multiple studies have investigated the potential of combining
CDK7 inhibitors with other anticancer drugs [22,23,116,
124127].
As ER is the key transcriptional driver of ER+ breast cancer
and is activated by CDK7, CDK7 inhibitors have been
assessed in combination with anti-oestrogens in this context
(Table 3)[22,125]. In the ER+ breast cancer cell line, MCF7,
treatment with ICEC0942, plus either tamoxifen or
815Cancer Metastasis Rev (2020) 39:805–823
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
fulvestrant, caused greater growth inhibition than either agent
alone [22]. The combinatorial action of ICEC0942 with ta-
moxifen was also verified in mice bearing MCF7 xenografts
[22]. In addition, ER-mutant breast cancer cells are especially
sensitive to the combination treatment of THZ1 and
fulvestrant [125]. ICEC0942 is now being assessed in combi-
nation with fulvestrant in clinical trials (Table 4).
It has been noted that some CDK7 inhibitors reduce ex-
pression of the anti-apoptotic BCL2 family member, MCL1
[23,124,127]; therefore, it is plausible that CDK7 inhibitors
will work synergistically with apoptotic agents. CDK7 inhib-
itors have been investigated in combination with the
apoptosis-inducing BH3-mimetics, venetoclax (ABT-199)
[23,124] and navitoclax (ABT-263) (Table 3)[127].
Venetoclax, which inhibits anti-apoptotic BCL2 to cause ap-
optosis, is approved for the treatment of CLL, small lympho-
cytic leukaemia (SLL) and AML [23]. The combination of
SY-1365 and venetoclax is synergistic in AML cell lines and
xenografts [23]. Although the current clinical trials of CDK7
inhibitorsare focused on solid tumours, these results highlight
the potential for CDK7 inhibitors to combat blood cancers,
particularly if combined with BH3-mimetics. CDK7 inhibi-
tion also synergises with p53-activating agents, including the
chemotherapeutic, 5-fluorouracil, to induce apoptosis in colo-
rectal cancer cells [128].
BET inhibitors, compounds that target the
bromodomain extra-terminal (BET) family of proteins,
have garnered much recent interest as potential cancer ther-
apeutics and are being assessed clinically across a range of
cancers. The preclinical and clinical advancement of BET
inhibitors in cancer therapy has recently been reviewed
[141]. BET inhibitors are exemplified by the tool com-
pound JQ1, which, like THZ1, preferentially represses
SE-driven transcription and can be used to overcome on-
cogenic transcriptional addiction in cancer [19]. For this
reason, THZ1 has been tested alongside JQ1 (Table 3),
and this combination synergistically inhibits the growth
of diffuse intrinsic pontine glioma (DIPG) [116], ovarian
cancer [75] and neuroblastoma [126]. THZ1 has also been
investigated alongside the histone deacetylase (HDAC) in-
hibitor, panobinostat (Table 3), for the treatment of the
universally fatal paediatric cancer DIPG [116]. Like JQ1
and THZ1, panobinostat supresses transcription of SE-
associated genes in DIPG and acts synergistically with
CDK7 inhibition to suppress growth of DIPG-derived cell
lines [116]. However, the development of CDK7 inhibitors
with adequate brain penetrance would be required before
this combination therapy could be achieved in the clinic.
Recent work has highlighted the potential of CDK7 inhib-
itors to be used in combination with immunotherapies. The
CDK7 inhibitor, YKL-5-124, was shown to elicit immune
response signalling in small cell lung cancer (SCLC), activat-
ing anti-tumourigenic T cells [131]. In immunocompetent
mouse models of SCLC, the combination of YKL-5-124 and
anti-PD-1 immune checkpoint inhibition increased overall
survival in comparison with either treatment alone, and was
further enhanced by the addition of chemotherapeutics (cis-
platin and etoposide) [131].
Overall, there is much potential for CDK7 inhibitors to be
used in combination with other drugs for cancer therapy. We
anticipate that screening of CDK7 inhibitors in combination
with other compounds, and large-scale genomic perturbation
studies, may pave the way for the identification of additional
co-targeting strategies.
5.5 Resistance to CDK7 inhibitors
The emergence of resistance to cancer treatment, including
targeted therapies, remains a major issue, and it is possible
that even if CDK7 inhibitors prove a clinical success, resis-
tance may develop in some patients. To gain understanding of
potential resistance mechanisms, cell lines with acquired re-
sistance to both ICEC0942 and THZ1 have been developed
[142,143].
ATP-binding cassette transporters (ABC-transporters) are a
well-characterised mechanism of multidrug resistance in can-
cer and, when upregulated, can mediate the ATP-dependent
efflux of drugs that are substrates. Upregulation of ABCB1,
also known as p-glycoprotein, mediates resistance to THZ1 in
neuroblastoma and lung cancer [142] and resistance to both
THZ1 and ICEC0942 in breast cancer cell lines [143].
Increased expression of another ABC-transporter, ABCG2,
results in resistance to THZ1 [142], but not to ICEC0942
[143]. With no clinical data available as yet, it remains to be
seen whether ABC-transporter upregulation will arise in the
context of CDK7 inhibitor resistance in patients. Despite nu-
merous clinical trials, ABC-transport inhibitors have thus far
proven unsuccessful in the clinic, mainly due to issues with
potency and toxicity [144]. However, should future develop-
ments be made in this area, there may be scope to use
CDK7 inhibitors in combination with ABC-transport inhibi-
tors. Another way to overcome this mechanism of resistance is
the development of CDK7 inhibitors that are not substrates for
multidrug resistance transporters [145]. It is likely that other,
as yet unidentified, mechanisms of resistance will also be im-
portant in a clinical setting.
As clinical trials progress, analyses of tumour features
enriched in subsets of patients that are intrinsically
resistant to CDK7 inhibitors, or those who acquire resistance
after initially responding well, should help shed light on mech-
anisms of CDK7 inhibitor resistance. Alongside these, further
preclinical models of CDK7 inhibitor resistance, including 3D
cell culture and in vivo models, may prove informative.
Ultimately, the identification of mechanisms of CDK7 inhib-
itor resistance should aim to aid the identification of patients
816 Cancer Metastasis Rev (2020) 39:805–823
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
who will derive most benefit from these drugs, helping to
advance their clinical progress.
6 Conclusions
CDK7 has a dual role in driving the cell cycle and transcrip-
tion, is upregulated in a variety of cancers and has emerged as
a promising cancer therapeutic target. At least ten selective
inhibitors of CDK7, with activity against a wide range of
cancer types, have been developed, their antitumour action
likely mediated both through cell cycle arrest and inhibition
of oncogenic transcriptional programs. In the preclinical set-
ting, these inhibitors have demonstrated potential to overcome
treatment-resistant cancer, both asmonotherapies, and in com-
bination with other cancer drugs. To date, four CDK7 inhibi-
tors have progressed to Phase I/II clinical trial for the treatment
of advanced solid malignancies. Whilst ABC-transporters can
mediate resistance to some CDK7 inhibitors, additional fac-
tors that influence tumour response to CDK7 inhibition are yet
to be identified. Further efforts to elucidate mechanisms of
response, and to define patient selection strategies, will help
to facilitate the clinical utility of CDK7 inhibitors.
Funding information The authorswork is generously funded by Cancer
Research UK (grant C37/A18784). Additional support was provided by
the Imperial Experimental Cancer Medicine Centre, the Imperial NIHR
Biomedical Research Centre and the Cancer Research UK Imperial
Centre. The views expressed are those of the authors and not necessarily
those of the NHS, the NIHR or the Department of Health.
Compliance with ethical standards
Conflict of interest RCC and SA are named as inventors on CDK7
inhibitor patents and own shares in Carrick Therapeutics.
Open Access This article is licensed under a Creative Commons
Attribution 4.0 International License, which permits use, sharing, adap-
tation, distribution and reproduction in any medium or format, as long as
you give appropriate credit to the original author(s) and the source, pro-
vide a link to the Creative Commons licence, and indicate if changes were
made. The images or other third party material in this article are included
in the article's CreativeCommons licence, unless indicated otherwise in a
credit line to the material. If material is not included in the article's
Creative Commons licence and your intended use is not permitted by
statutory regulation or exceeds the permitted use, you will need to obtain
permission directly from the copyright holder. To view a copy of this
licence, visit http://creativecommons.org/licenses/by/4.0/.
References
1. Schachter, M. M., & Fisher, R. P. (2013). The CDK-activating
kinase Cdk7. Cell Cycle, 12(20), 32393240. https://doi.org/10.
4161/cc.26355.
2. Larochelle, S., Merrick, K. A., Terret, M.-E., Wohlbold, L.,
Barboza, N. M., Zhang, C., Shokat, K. M., Jallepalli, P. V., &
Fisher, R. P. (2007). Requirements for Cdk7 in the assembly of
Cdk1/Cyclin B and activation of Cdk2 revealed by chemical ge-
netics in human cells. Molecular Cell, 25(6), 839850. https://doi.
org/10.1016/j.molcel.2007.02.003.
3. Schachter, M. M., Merrick, K. A., Larochelle, S., Hirschi, A.,
Zhang, C., Shokat, K. M., Rubin, S. M., & Fisher, R. P. (2013).
A Cdk7-Cdk4 T-loop phosphorylation Cascade promotes G1 pro-
gression. Molecular Cell, 50(2), 250260. https://doi.org/10.
1016/j.molcel.2013.04.003.
4. Bisteau, X., Paternot, S., Colleoni, B., Ecker, K., Coulonval, K.,
De Groote, P., et al. (2013). CDK4 T172 phosphorylation is cen-
tral in a CDK7-dependent bidirectional CDK4/CDK2 interplay
mediated by p21 phosphorylation at the restriction point. PLoS
Genetics, 9(5), e1003546. https://doi.org/10.1371/journal.pgen.
1003546.
5. Wong, K. H., Jin, Y., & Struhl, K. (2014). TFIIH phosphorylation
of the pol II CTD stimulates mediator dissociation from the
Preinitiation complex and promoter escape. Molecular Cell,
54(4), 601612. https://doi.org/10.1016/j.molcel.2014.03.024.
6. Glover-Cutter, K., Larochelle, S., Erickson, B., Zhang, C., Shokat,
K., Fisher, R. P., et al. (2009). TFIIH-associated Cdk7 kinase
functions in phosphorylation of C-terminal domain Ser7 residues,
promoter-proximal pausing, and termination by RNA polymerase
II. Molecular and Cellular Biology, 29(20), 54555464. https://
doi.org/10.1128/mcb.00637-09.
7. Akhtar, M. S., Heidemann, M., Tietjen, J. R., Zhang, D. W.,
Chapman, R. D., Eick, D., & Ansari, A. Z. (2009). TFIIH kinase
places bivalent marks on the carboxy-terminal domain of RNA
polymerase II. Molecular Cell, 34(3), 387393. https://doi.org/10.
1016/j.molcel.2009.04.016.
8. Larochelle, S., Amat, R., Glover-Cutter, K., Sansó, M., Zhang, C.,
Allen, J. J., et al. (2012). Cyclin-dependent kinase control of the
initiation-to-elongation switch of RNA polymerase II. Nature
Structural & Molecular Biology, 19, 1108, https://doi.org/10.
1038/nsmb.2399 https://www.nature.com/articles/nsmb.2399#
supplementary-information.
9. Ko, L. J., Shieh, S. Y., Chen, X., Jayaraman, L., Tamai, K., Taya,
Y., Prives, C., & Pan, Z. Q. (1997). p53 is phosphorylated by
CDK7-cyclin H in a p36MAT1-dependent manner. Molecular
and Cellular Biology, 17(12), 72207229. https://doi.org/10.
1128/mcb.17.12.7220.
10. Lu, H., Fisher, R. P., Bailey, P., & Levine, A. J. (1997). The
CDK7-cycH-p36 complex of transcription factor IIH phosphory-
lates p53, enhancing its sequence-specific DNA binding activity
in vitro. Molecular and Cellular Biology, 17(10), 59235934.
https://doi.org/10.1128/mcb.17.10.5923.
11. Keriel, A., Stary, A., Sarasin, A., Rochette-Egly, C., & Egly, J.-M.
(2002). XPD mutations prevent TFIIH-dependent transactivation
by nuclear receptors and phosphorylation of RARα.Cell, 109(1),
125135. https://doi.org/10.1016/S0092-8674(02)00692-X.
12. Bastien, J., Adam-Stitah, S., Riedl, T., Egly, J.-M., Chambon, P.,
& Rochette-Egly, C. (2000). TFIIH interacts with the retinoic acid
receptor γand phosphorylates its AF-1-activating domain through
cdk7. Journal of Biological Chemistry, 275(29), 2189621904.
https://doi.org/10.1074/jbc.M001985200.
13. Rochette-Egly, C., Adam, S., Rossignol, M., Egly, J.-M., &
Chambon, P. (1997). Stimulation of RARαactivation function
AF-1 through binding to the general transcription factor TFIIH
and phosphorylation by CDK7. Cell, 90(1), 97107. https://doi.
org/10.1016/S0092-8674(00)80317-7.
14. Chen, D., Riedl, T., Washbrook, E., Pace, P. E., Coombes, R. C.,
Egly, J.-M., & Ali, S. (2000). Activation of estrogen receptor αby
S118 phosphorylation involves a ligand-dependent interaction
with TFIIH and participation of CDK7. Molecular Cell, 6(1),
127137. https://doi.org/10.1016/S1097-2765(05)00004-3.
817Cancer Metastasis Rev (2020) 39:805–823
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
15. Chen, D., Washbrook, E., Sarwar, N., Bates, G. J., Pace, P. E.,
Thirunuvakkarasu, V., Taylor, J., Epstein, R. J., Fuller-Pace, F.
V., Egly, J. M., Coombes, R. C., & Ali, S. (2002).
Phosphorylation of human estrogen receptor αat serine 118 by
two distinct signal transduction pathways revealed by
phosphorylation-specific antisera. Oncogene, 21(32), 4921
4931. https://doi.org/10.1038/sj.onc.1205420.
16. Lee, D. K., Duan, H. O., & Chang, C. (2000). From Androgen
receptor to the general transcription factor TFIIH: identification of
cdk activating kinase (CAK) as an androgen receptor NH2- termi-
nal associated coactivator. Journal of Biological Chemistry,
275(13), 93089313. https://doi.org/10.1074/jbc.275.13.9308.
17. Chymkowitch, P., Le May, N., Charneau, P., Compe, E., & Egly,
J.-M. (2011). The phosphorylation of the androgen receptor by
TFIIH directs the ubiquitin/proteasome process. The EMBO
Journal, 30(3), 468479. https://doi.org/10.1038/emboj.2010.
337.
18. Galbraith, M. D., Bender, H., & Espinosa, J. M. (2019).
Therapeutic targeting of transcriptional cyclin-dependent kinases.
Transcription, 10(2), 118136. https://doi.org/10.1080/
21541264.2018.1539615.
19. Bradner, J. E., Hnisz, D., & Young, R. A. (2017). Transcriptional
addiction in cancer. Cell, 168(4), 629643. https://doi.org/10.
1016/j.cell.2016.12.013.
20. Kwiatkowski, N., Zhang, T., Rahl, P. B., Abraham, B. J., Reddy,
J., Ficarro, S. B., Dastur, A., Amzallag, A., Ramaswamy, S., Tesar,
B., Jenkins, C. E., Hannett, N. M., McMillin, D., Sanda, T., Sim,
T., Kim, N. D., Look, T., Mitsiades, C. S., Weng, A. P., Brown, J.
R., Benes, C.H., Marto, J. A., Young, R. A., & Gray, N. S. (2014).
Targeting transcription regulation in cancer with a covalent CDK7
inhibitor. Nature, 511(7511), 616620. https://doi.org/10.1038/
nature13393.
21. Chipumuro, E., Marco, E., Christensen, C. L., Kwiatkowski, N.,
Zhang, T., Hatheway, C. M., Abraham, B. J., Sharma, B., Yeung,
C., Altabef, A., Perez-Atayde, A., Wong, K. K., Yuan, G. C., Gray,
N. S., Young, R. A., & George, R. E. (2014). CDK7 inhibition
suppresses super-enhancer-linked oncogenic transcription in
MYCN-driven cancer. Cell, 159(5), 11261139. https://doi.org/
10.1016/j.cell.2014.10.024.
22. Patel, H., Periyasamy, M., Sava, G. P., Bondke, A., Slafer, B. W.,
Kroll, S. H. B., Barbazanges, M., Starkey, R., Ottaviani, S.,
Harrod, A., Aboagye, E. O., Buluwela, L., Fuchter, M. J.,
Barrett, A. G. M., Coombes, R. C., & Ali, S. (2018). ICEC0942,
an orally bioavailable selectiveinhibitor of CDK7 for cancertreat-
ment. Molecular Cancer Therapeutics, 17(6), 11561166. https://
doi.org/10.1158/1535-7163.Mct-16-0847.
23. Hu, S., Marineau, J. J., Rajagopal, N., Hamman, K. B., Choi, Y. J.,
Schmidt, D. R., et al. (2019). Discovery and characterization of
SY-1365, a selective, covalent inhibitor of CDK7. Cancer
Research, canres.0119.2019, https://doi.org/10.1158/0008-5472.
Can-19-0119.
24. Hu, S., Marineau, J., Hamman, K., Bradley, M., Savinainen, A.,
Alnemy, S., et al. (2019). Abstract 4421: SY-5609, an orally avail-
able selective CDK7 inhibitor demonstrates broad anti-tumor ac-
tivity in vivo.Cancer Research, 79(13 Supplement), 44214421.
https://doi.org/10.1158/1538-7445.Am2019-4421.
25. Johannessen, L. H., Hu, S., Ke, N., D'Ippolito, A., Rajagopal, N.,
Marineau, J., et al. (2019). Abstract C091: Preclinical evaluation
of PK, PD, and antitumor activity of the oral, non-covalent,potent
and highly selective CDK7 inhibitor, SY-5609, provides rationale
for clinical development in multiple solid tumor indications.
Molecular Cancer Therapeutics, 18(12 Supplement), C091
C091. https://doi.org/10.1158/1535-7163.Targ-19-c091.
26. Coates, D. A., Montero, C., Patel, B. K. R., Remick, D. M., &
Yadav, V. (2019). Compounds useful for inhibiting CDK7. United
States: Eli Lilly and Company. IN, US: Indianapolis.
27. Rimel, J. K., & Taatjes, D. J. (2018). The essential and multifunc-
tional TFIIH complex. Protein Science, 27(6), 10181037. https://
doi.org/10.1002/pro.3424.
28. Lolli, G., Lowe, E. D., Brown, N. R., & Johnson, L. N. (2004).
The crystal structure of human CDK7 and its protein recognition
properties. Structure, 12(11), 20672079. https://doi.org/10.1016/
j.str.2004.08.013.
29. Larochelle, S., Chen, J., Knights, R., Pandur, J., Morcillo, P.,
Erdjument-Bromage, H., Tempst, P., Suter, B., & Fisher, R. P.
(2001). T-loop phosphorylation stabilizes the CDK7-cyclin H-
MAT1 complex in vivo and regulates its CTD kinase activity.
The EMBO Journal, 20(14), 37493759. https://doi.org/10.1093/
emboj/20.14.3749.
30. Garrett, S., Barton, W. A., Knights, R., Jin, P., Morgan, D. O., &
Fisher, R. P. (2001). Reciprocal activation by Cyclin-dependent
kinases 2 and 7 is directed by substrate specificity determinants
outside the T loop. Molecular and Cellular Biology, 21(1), 8899.
https://doi.org/10.1128/mcb.21.1.88-99.2001.
31. Desai, S. R., Pillai, P. P., Patel, R. S., McCray, A. N., Win-Piazza,
H. Y., & Acevedo-Duncan, M. E. (2011). Regulation of Cdk7
activity through a phosphatidylinositol (3)-kinase/PKC-ι-mediat-
ed signaling cascade in glioblastoma. Carcinogenesis, 33(1), 10
19. https://doi.org/10.1093/carcin/bgr231.
32. Pillai, P., Desai, S., Patel, R., Sajan, M., Farese, R., Ostrov, D., &
Acevedo-Duncan, M. (2011). A novel PKC-ιinhibitor abrogates
cell proliferation and induces apoptosis in neuroblastoma. The
International Journal of Biochemistry & Cell Biology, 43(5),
784794. https://doi.org/10.1016/j.biocel.2011.02.002.
33. Acevedo-Duncan, M., Patel, R., Whelan, S., & Bicaku, E. (2002).
Human glioma PKC-ιand PKC-βII phosphorylate cyclin-
dependent kinase activating kinase during the cell cycle. Cell
Proliferation, 35(1), 2336. https://doi.org/10.1046/j.1365-2184.
2002.00220.x.
34. Ni, S., Chen, L., Li, M., Zhao, W., Shan, X., Wu, M., Cheng, J.,
Liang, L., Wang, Y., Jiang, W., Zhang, J., & Ni, R. (2016). PKC
iota promotes cellular proliferation by accelerated G1/S transition
via interaction with CDK7 in esophageal squamous cell carcino-
ma. Tumor Biology, 37(10), 1379913809. https://doi.org/10.
1007/s13277-016-5193-9.
35. Ghezzi, C., Wong, A., Chen, B. Y., Ribalet, B., Damoiseaux, R., &
Clark, P. M. (2019). A high-throughput screen identifies that
CDK7 activates glucose consumption in lung cancer cells.
Nature Communications, 10(1), 5444. https://doi.org/10.1038/
s41467-019-13334-8.
36. Schneider, E., Kartarius, S., Schuster, N., & Montenarh, M.
(2002). The cyclin H/cdk7/Mat1 kinase activity is regulated by
CK2 phosphorylation of cyclin H. Oncogene, 21(33), 50315037.
https://doi.org/10.1038/sj.onc.1205690.
37. Akoulitchev, S., Chuikov, S., & Reinberg, D. (2000). TFIIH is
negatively regulated by cdk8-containing mediator complexes.
Nature, 407(6800), 102106. https://doi.org/10.1038/35024111.
38. Abdullah, A. I., Zhang, H., Nie, Y., Tang, W., & Sun, T. (2016).
CDK7 and miR-210 co-regulate cell-cycle progression of neural
progenitors in the developing Neocortex. Stem Cell Reports, 7(1),
6979. https://doi.org/10.1016/j.stemcr.2016.06.005.
39. Ali,S.,Heathcote,D.A.,Kroll,S.H.B.,Jogalekar,A.S.,
Scheiper, B., Patel, H., Brackow, J., Siwicka, A., Fuchter, M. J.,
Periyasamy, M., Tolhurst, R. S., Kanneganti, S. K., Snyder, J. P.,
Liotta, D. C., Aboagye, E. O., Barrett, A. G. M., & Coombes, R.
C. (2009). The development of a selective cyclin-dependent ki-
nase inhibitor that shows antitumor activity. Cancer Research,
69(15), 62086215. https://doi.org/10.1158/0008-5472.Can-09-
0301.
40. Merrick, K. A., Larochelle, S., Zhang, C., Allen, J. J., Shokat, K.
M., & Fisher, R. P. (2008). Distinct activation pathways confer
cyclin-binding specificity on Cdk1 and Cdk2 in human cells.
818 Cancer Metastasis Rev (2020) 39:805–823
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Molecular Cell, 32(5), 662672. https://doi.org/10.1016/j.molcel.
2008.10.022.
41. Abdulrahman, W., Iltis, I., Radu, L., Braun, C., Maglott-Roth, A.,
Giraudon, C., Egly, J. M., & Poterszman, A. (2013). ARCH do-
main of XPD, an anchoring platform for CAK that conditions
TFIIH DNA repair and transcription activities. Proceedings of
the National Academy of Sciences, 110(8), E633E642. https://
doi.org/10.1073/pnas.1213981110.
42. Luo, J., Cimermancic, P., Viswanath, S., Ebmeier, C. C., Kim, B.,
Dehecq, M., Raman, V., Greenberg, C. H., Pellarin, R., Sali, A.,
Taatjes, D. J., Hahn, S., & Ranish, J. (2015). Architecture of the
human and yeast general transcription and DNA repair factor
TFIIH. Molecular Cell, 59(5), 794806. https://doi.org/10.1016/
j.molcel.2015.07.016.
43. Tirode, F., Busso, D., Coin, F., & Egly, J.-M. (1999).
Reconstitution of the transcription factor TFIIH: assignment of
functions for the three enzymatic subunits, XPB, XPD, and
cdk7. Molecular Cell, 3(1), 8795. https://doi.org/10.1016/
S1097-2765(00)80177-X.
44. Bataille, A. R., Jeronimo, C., Jacques, P.-É., Laramée, L., Fortin,
M.-È., Forest, A., Bergeron, M., Hanes, S. D., & Robert, F. (2012).
A universal RNA polymerase II CTD cycle is orchestrated by
complex interplays between kinase, phosphatase, and isomerase
enzymes along genes. Molecular Cell, 45(2), 158170. https://doi.
org/10.1016/j.molcel.2011.11.024.
45. Jeronimo, C., & Robert, F. (2014). Kin28 regulates the transient
association of mediator with core promoters. Nature Structural &
Molecular Biology, 21(5), 449455. https://doi.org/10.1038/
nsmb.2810.
46. Egloff, S., O'Reilly, D., Chapman, R. D., Taylor, A., Tanzhaus, K.,
Pitts, L., Eick, D., & Murphy, S. (2007). Serine-7 of the RNA
polymerase II CTD is specifically required for snRNA gene ex-
pression.Science, 318(5857), 17771779. https://doi.org/10.1126/
science.1145989.
47. Glover-Cutter, K., Kim, S., Espinosa, J., & Bentley, D. L. (2007).
RNA polymerase II pauses and associates with pre-mRNA pro-
cessing factors at both ends of genes. Nature Structural &
Molecular Biology, 15,71.https://doi.org/10.1038/nsmb1352
https://www.nature.com/articles/nsmb1352#supplementary-
information.
48. Kwak, H., & Lis, J. T. (2013). Control of transcriptional elonga-
tion. Annual Review of Genetics, 47,483508. https://doi.org/10.
1146/annurev-genet-110711-155440.
49. Yamaguchi, Y., Takagi, T., Wada, T., Yano, K., Furuya, A.,
Sugimoto, S., Hasegawa, J., & Handa, H. (1999). NELF, a
multisubunit complex containing RD, cooperates with DSIF to
repress RNA polymerase II elongation. Cell, 97(1), 4151.
https://doi.org/10.1016/S0092-8674(00)80713-8.
50. Nilson, K. A., Guo, J., Turek, M. E., Brogie, J. E., Delaney, E.,
Luse, D. S., & Price, D. H. (2015). THZ1 reveals roles for Cdk7 in
co-transcriptional capping and pausing. Molecular Cell, 59(4),
576587. https://doi.org/10.1016/j.molcel.2015.06.032.
51. Ebmeier, C.C., Erickson, B., Allen, B. L., Allen, M. A., Kim, H.,
Fong, N., Jacobsen, J. R., Liang, K., Shilatifard,A., Dowell, R. D.,
Old, W. M., Bentley, D. L., & Taatjes, D. J. (2017). Human TFIIH
kinase CDK7 regulates transcription-associated chromatin modi-
fications. Cell Reports, 20(5), 11731186. https://doi.org/10.1016/
j.celrep.2017.07.021.
52. Bacon, C. W., & DOrso, I. (2019). CDK9: A signaling hub for
transcriptional control. Transcription, 10(2), 5775. https://doi.
org/10.1080/21541264.2018.1523668.
53. Bartkowiak, B., Liu, P., Phatnani, H. P., Fuda, N. J., Cooper, J. J.,
Price, D. H., Adelman, K., Lis, J. T., & Greenleaf, A. L. (2010).
CDK12 is a transcription elongation-associated CTD kinase, the
metazoan ortholog of yeast Ctk1. Genes & Development, 24(20),
23032316. https://doi.org/10.1101/gad.1968210.
54. sken, C. A., Farnung, L., Hintermair, C., Merzel Schachter, M.,
Vogel-Bachmayr, K., Blazek, D., et al. (2014). The structure and
substrate specificity of human Cdk12/Cyclin K. Nature
Communications, 5, 3505, https://doi.org/10.1038/ncomms4505
https://www.nature.com/articles/ncomms4505#supplementary-
information.
55. Liang, K., Gao, X., Gilmore, J. M., Florens, L., Washburn, M. P.,
Smith, E., & Shilatifard, A. (2015). Characterization of human
cyclin-dependent kinase 12 (CDK12) and CDK13 complexes in
C-terminal domain phosphorylation, gene transcription, and RNA
processing. Molecular and Cellular Biology, 35(6), 928938.
https://doi.org/10.1128/mcb.01426-14.
56. Greifenberg, A. K., Hönig, D., Pilarova, K., Düster, R.,
Bartholomeeusen, K., Bösken, C. A., Anand, K., Blazek, D., &
Geyer, M. (2016). Structural and functional analysis of the
Cdk13/Cyclin K complex. Cell Reports, 14(2), 320331. https://
doi.org/10.1016/j.celrep.2015.12.025.
57. Rossi, D. J., Londesborough, A., Korsisaari, N., Pihlak, A.,
Lehtonen, E., Henkemeyer, M., & Mäkelä, T. P. (2001). Inability
to enter S phase and defective RNA polymerase II CTD phosphor-
ylation in mice lacking Mat1. The EMBO Journal, 20(11), 2844
2856. https://doi.org/10.1093/emboj/20.11.2844.
58. Ganuza, M., Sáiz-Ladera, C., Cañamero, M., Gómez, G.,
Schneider,R.,Blasco,M.A.,Pisano,D.,Paramio,J.M.,
Santamaría, D., & Barbacid, M. (2012). Genetic inactivation of
Cdk7 leads tocell cycle arrest and induces premature aging due to
adult stem cell exhaustion. The EMBO Journal, 31(11), 2498
2510. https://doi.org/10.1038/emboj.2012.94.
59. Ali, S., Metzger, D., Bornert, J. M., & Chambon, P. (1993).
Modulation of transcriptional activation by ligand-dependent
phosphorylation of the human oestrogen receptor A/B region.
The EMBO Journal, 12(3), 11531160. https://doi.org/10.1002/j.
1460-2075.1993.tb05756.x.
60. Valley, C. C., Métivier, R., Solodin, N. M., Fowler, A. M.,
Mashek, M. T., Hill, L., et al. (2005). Differential regulation of
estrogen-inducible proteolysis and transcription by the estrogen
receptor αNterminus.Molecular and Cellular Biology, 25(13),
54175428. https://doi.org/10.1128/mcb.25.13.5417-5428.2005.
61. Drané, P., Compe, E., Catez, P., Chymkowitch, P., & Egly, J.-M.
(2004). Selective regulation of vitamin D receptor-responsive
genes by TFIIH. Molecular Cell, 16(2), 187197. https://doi.org/
10.1016/j.molcel.2004.10.007.
62. Compe, E., Drané, P., Laurent, C., Diderich, K., Braun, C.,
Hoeijmakers, J. H. J., et al. (2005). Dysregulation of the peroxi-
some proliferator-activated receptor target genes by XPD muta-
tions. Molecular and Cellular Biology, 25(14), 60656076.
https://doi.org/10.1128/mcb.25.14.6065-6076.2005.
63. Vandel, L., & Kouzarides, T. (1999). Residues phosphorylated by
TFIIH are required for E2F-1 degradation during S-phase. The
EMBO Journal, 18(15), 42804291. https://doi.org/10.1093/
emboj/18.15.4280.
64. Cho, Y. S., Li, S., Wang, X., Zhu, J., Zhuo, S., Han, Y., Yue, T.,
Yang, Y., & Jiang,J. (2019). CDK7 regulates organ size and tumor
growth by safeguarding the hippo pathway effector Yki/yap/Taz in
the nucleus. Genes & Development, 34,5371. https://doi.org/10.
1101/gad.333146.119.
65. Coin, F., Oksenych, V., Mocquet, V., Groh, S., Blattner, C., &
Egly, J. M. (2008). Nucleotide excision repair driven by the dis-
sociation of CAK from TFIIH. Molecular Cell, 31(1), 920.
https://doi.org/10.1016/j.molcel.2008.04.024.
66. Araújo, S. J., Tirode, F., Coin, F., Pospiech, H., Syväoja, J. E.,
Stucki, M., Hübscher, U., Egly, J. M., & Wood, R. D. (2000).
Nucleotide excision repair of DNA with recombinant human pro-
teins: definition of the minimal set of factors, active forms of
TFIIH, and modulation by CAK. Genes & Development, 14(3),
349359. https://doi.org/10.1101/gad.14.3.349.
819Cancer Metastasis Rev (2020) 39:805–823
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
67. Bartkova, J., Zemanova, M., & Bartek, J. (1996). Expression of
CDK7/CAK in normal and tumour cells of diverse histogenesis,
cell-cycle position and differentiation. International Journal of
Cancer, 66(6), 732737. https://doi.org/10.1002/(sici)1097-
0215(19960611)66:6<732::Aid-ijc4>3.0.Co;2-0.
68. Patel, H., Abduljabbar, R., Lai, C.-F., Periyasamy, M., Harrod, A.,
Gemma, C., Steel, J. H., Patel, N., Busonero, C., Jerjees, D.,
Remenyi, J., Smith, S., Gomm, J. J., Magnani, L., Gyorffy, B.,
Jones, L. J., Fuller-Pace, F., Shousha, S., Buluwela, L., Rakha, E.
A., Ellis, I. O., Coombes, R. C., & Ali, S. (2016). Expression of
CDK7, cyclin H, and MAT1 is elevated in breast cancer and is
prognostic in estrogen receptorpositive breast cancer. Clinical
Cancer Research, 22(23), 59295938. https://doi.org/10.1158/
1078-0432.Ccr-15-1104.
69. Wang, Q., Li, M., Zhang, X., Huang, H., Huang, J., Ke, J., Ding,
H., Xiao, J., Shan, X., Liu, Q., Bao, B., & Yang, L. (2016).
Upregulation of CDK7 in gastric cancer cell promotes tumor cell
proliferation and predicts poor prognosis. Experimental and
Molecular Pathology, 100(3), 514521. https://doi.org/10.1016/j.
yexmp.2016.05.001.
70. Naseh, G., Mohammadifard, M., & Mohammadifard, M. (2016).
Upregulation of cyclin-dependent kinase 7 and matrix
metalloproteinase-14 expression contribute to metastatic proper-
ties of gastric cancer. IUBMB Life, 68(10), 799805. https://doi.
org/10.1002/iub.1543.
71. Jiang, L., Huang, R., Wu, Y., Diao, P., Zhang, W., Li, J., Li, Z.,
Wang, Y., Cheng, J., & Yang, J. (2019). Overexpression ofCDK7
is associated with unfavourable prognosis in oral squamous cell
carcinoma. Pathology, 51(1), 7480. https://doi.org/10.1016/j.
pathol.2018.10.004.
72. Tsang, F. H.-C., Law, C.-T., Tang, T.-C. C., Cheng, C. L.-H., Chin,
D. W.-C., Tam, W.-S. V., Wei, L., Wong, C. C. L., Ng, I. O. L., &
Wong, C. M. (2019). Aberrant super-enhancer landscape in hu-
man hepatocellular carcinoma. Hepatology, 69(6), 25022517.
https://doi.org/10.1002/hep.30544.
73. Meng, W., Wang, J., Wang, B., Liu, F., Li, M., Zhao, Y., Zhang,
C., Li, Q., Chen, J., Zhang, L., Tang, Y., & Ma, J. (2018). CDK7
inhibition is a novel therapeutic strategy against GBM both
in vitro and in vivo. Cancer Management and Research, 10,
57475758. https://doi.org/10.2147/CMAR.S183696.
74. Li, B., Ni Chonghaile, T., Fan, Y., Madden, S. F., Klinger, R.,
O'Connor, A. E., Walsh, L., O'Hurley, G., Mallya Udupi, G.,
Joseph,J.,Tarrant,F.,Conroy,E.,Gaber,A.,Chin,S.F.,
Bardwell, H. A., Provenzano, E., Crown, J., Dubois, T., Linn, S.,
Jirstrom, K., Caldas, C., O'Connor, D. P., & Gallagher, W. M.
(2017). Therapeutic rationale to target highly expressed CDK7
conferring poor outcomes in triple-negative breast cancer.
Cancer Research, 77(14), 38343845. https://doi.org/10.1158/
0008-5472.Can-16-2546.
75. Zhang, Z., Peng, H., Wang, X., Yin, X., Ma, P., Jing, Y., Cai, M.
C., Liu, J., Zhang, M., Zhang, S., Shi, K., Gao, W. Q., di, W., &
Zhuang, G. (2017). Preclinical efficacy and molecular mechanism
of targeting CDK7-dependent transcriptional addiction in ovarian
cancer. Molecular Cancer Therapeutics, 16(9), 17391750.
https://doi.org/10.1158/1535-7163.Mct-17-0078.
76. Pavey, S., Johansson, P., Packer, L., Taylor, J., Stark, M.,Pollock,
P. M., Walker, G. J., Boyle, G. M., Harper, U., Cozzi, S. J.,
Hansen, K., Yudt, L., Schmidt, C., Hersey, P., Ellem, K. A. O.,
O'Rourke, M. G. E., Parsons, P. G., Meltzer, P., Ringnér, M., &
Hayward, N. K. (2004). Microarray expression profiling in mela-
noma reveals a BRAF mutation signature. Oncogene, 23(23),
40604067. https://doi.org/10.1038/sj.onc.1207563.
77. Johansson, P., Pavey, S., & Hayward, N. (2007). Confirmation of a
BRAF mutation-associated gene expression signature in melano-
ma. Pigment Cell Research, 20(3), 216221. https://doi.org/10.
1111/j.1600-0749.2007.00375.x.
78. Kandoth, C., McLellan, M. D., Vandin, F., Ye, K., Niu, B., Lu, C.,
Xie, M., Zhang, Q., McMichael, J. F., Wyczalkowski, M. A.,
Leiserson, M. D. M., Miller, C. A., Welch, J. S., Walter, M. J.,
Wendl, M. C., Ley, T. J., Wilson, R. K., Raphael, B. J., & Ding, L.
(2013). Mutational landscape and significance across 12 major
cancer types. Nature, 502(7471), 333339. https://doi.org/10.
1038/nature12634.
79. Lawrence, M. S., Stojanov, P., Mermel, C. H., Robinson, J. T.,
Garraway, L. A., Golub, T. R., Meyerson, M., Gabriel, S. B.,
Lander, E. S., & Getz, G. (2014). Discovery and saturation anal-
ysis of cancer genes across 21 tumour types. Nature, 505(7484),
495501. https://doi.org/10.1038/nature12912.
80. Lane, D., & Levine, A. (2010). p53 research: the past thirty years
and the next thirty years. Cold Spring Harbor Perspectives in
Biology, 2(12). https://doi.org/10.1101/cshperspect.a000893.
81. Beroukhim, R., Mermel, C. H.,Porter, D., Wei, G., Raychaudhuri,
S., Donovan, J., Barretina, J., Boehm, J. S., Dobson, J., Urashima,
M., Mc Henry, K. T., Pinchback, R. M., Ligon, A. H., Cho, Y. J.,
Haery, L., Greulich, H., Reich, M., Winckler, W., Lawrence, M.
S., Weir, B.A., Tanaka, K. E., Chiang, D. Y.,Bass, A. J., Loo,A.,
Hoffman, C., Prensner, J., Liefeld, T., Gao, Q., Yecies, D.,
Signoretti, S., Maher, E., Kaye, F. J., Sasaki, H., Tepper, J. E.,
Fletcher, J. A., Tabernero, J., Baselga, J., Tsao, M. S.,
Demichelis, F., Rubin, M. A., Janne, P. A., Daly, M. J., Nucera,
C., Levine, R. L., Ebert, B. L., Gabriel, S., Rustgi, A. K.,
Antonescu, C. R., Ladanyi, M., Letai, A., Garraway, L. A.,
Loda, M., Beer, D. G., True, L. D., Okamoto, A., Pomeroy, S.
L., Singer, S., Golub, T. R., Lander, E. S., Getz, G., Sellers, W.
R., & Meyerson, M. (2010). The landscape of somatic copy-
number alteration across human cancers. Nature, 463(7283),
899905. https://doi.org/10.1038/nature08822.
82. Clemons, M., Danson, S., & Howell, A. (2002). Tamoxifen
(Nolvadex): a review: Antitumour treatment. Cancer
Treatment Reviews, 28(4), 165180. https://doi.org/10.1016/
S0305-7372(02)00036-1.
83. Nathan, M. R., & Schmid, P. (2017). A review of fulvestrant in
breast cancer. Oncology and Therapy, 5(1), 1729. https://doi.org/
10.1007/s40487-017-0046-2.
84. Hnisz, D., Abraham, B. J., Lee, T. I., Lau, A., Saint-André, V.,
Sigova, A. A., Hoke, H. A., & Young, R. A. (2013). Super-
enhancers in the control of cell identity and disease. Cell,
155(4), 934947. https://doi.org/10.1016/j.cell.2013.09.053.
85. Whittaker, S. R., Mallinger, A., Workman, P., & Clarke, P. A.
(2017). Inhibitors of cyclin-dependent kinases as cancer therapeu-
tics. Pharmacology & Therapeutics, 173,83105. https://doi.org/
10.1016/j.pharmthera.2017.02.008.
86. Conroy, A., Stockett, D. E., Walker, D., Arkin, M. R., Hoch, U.,
Fox, J. A., & Hawtin, R. E. (2009). SNS-032 is a potent and
selective CDK 2, 7 and 9 inhibitor that drives target modulation
in patient samples. Cancer Chemotherapy and Pharmacology,
64(4), 723732. https://doi.org/10.1007/s00280-008-0921-5.
87. Kaur, G., Stetler-Stevenson, M., Sebers, S., Worland, P., Sedlacek,
H., Myers, C., Czech, J.,Naik, R., & Sausville, E. (1992). Growth
inhibition with reversible cell cycle arrest of carcinoma cells by
flavone L86-8275. JNCI: Journal of the National Cancer
Institute, 84(22), 17361740. https://doi.org/10.1093/jnci/84.22.
1736.
88. Losiewicz, M. D., Carlson, B. A., Kaur, G., Sausville, E. A., &
Worland, P. J. (1994). Potent inhibition of Cdc2 kinase activity by
the flavonoid L86-8275. Biochemical and Biophysical Research
Communications, 201(2), 589595. https://doi.org/10.1006/bbrc.
1994.1742.
89. Carlson, B. A., Dubay, M. M., Sausville, E. A., Brizuela, L., &
Worland, P. J. (1996). Flavopiridol induces G1 arrest with inhibi-
tion of cyclin-dependent kinase (CDK) 2 and CDK4 in human
breast carcinoma cells. Cancer Research, 56(13), 29732978.
820 Cancer Metastasis Rev (2020) 39:805–823
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
90. Chen, R., Keating, M. J., Gandhi, V., & Plunkett, W. (2005).
Transcription inhibition by flavopiridol: mechanism of chronic
lymphocytic leukemia cell death. Blood, 106(7), 25132519.
https://doi.org/10.1182/blood-2005-04-1678.
91. Asghar, U., Witkiewicz, A. K., Turner, N. C., & Knudsen, E. S.
(2015). The history and future of targeting cyclin-dependent ki-
nases in cancer therapy. [review article]. Nature Reviews Drug
Discovery, 14, 130. https://doi.org/10.1038/nrd4504 https://
www.nature.com/articles/nrd4504#supplementary-information.
92. Byrd, J. C., Lin, T. S., Dalton, J. T., Wu, D., Phelps, M. A.,
Fischer, B., Moran, M., Blum, K. A., Rovin, B., Brooker-
McEldowney, M., Broering, S., Schaaf, L. J., Johnson, A. J.,
Lucas, D. M., Heerema, N. A., Lozanski, G., Young, D. C.,
Suarez,J.R.,Colevas,A.D.,&Grever,M.R.(2006).
Flavopiridol administered using a pharmacologically derived
schedule is associated with marked clinical efficacy in refractory,
genetically high-risk chronic lymphocytic leukemia. Blood,
109(2), 399404. https://doi.org/10.1182/blood-2006-05-020735.
93. Byrd, J. C., Peterson, B. L., Gabrilove,J., Odenike, O. M., Grever,
M. R., Rai, K., Larson, R. A., & Cancer and Leukemia Group B.
(2005). Treatment of relapsed chronic lymphocytic leukemia by
72-hour continuous infusion or 1-hour bolus infusion of
flavopiridol: results from cancer and leukemia group B study
19805. Clinical Cancer Research, 11(11), 41764181. https://
doi.org/10.1158/1078-0432.Ccr-04-2276.
94. Kouroukis, C. T., Belch, A., Crump, M., Eisenhauer, E.,
Gascoyne, R. D., Meyer, R., Lohmann, R., Lopez, P., Powers,
J., Turner, R., Connors, J. M., & National Cancer Institute of
Canada Clinical Trials Group. (2003). Flavopiridol in untreated
or relapsed mantle-cell lymphoma: results of a phase II study of
the National Cancer Institute of Canada Clinical Trials Group.
Journal of Clinical Oncology, 21(9), 17401745. https://doi.org/
10.1200/jco.2003.09.057.
95. Meijer, L., Borgne, A., Mulner, O., Chong, J. P. J., Blow, J. J.,
Inagaki, N., Inagaki, M.,Delcros, J. G., & Moulinoux, J. P. (1997).
Biochemical and cellular effects of roscovitine, a potent and se-
lective inhibitor of the cyclin-dependent kinases cdc2, cdk2 and
cdk5. European Journal of Biochemistry, 243(12), 527536.
https://doi.org/10.1111/j.1432-1033.1997.t01-2-00527.x.
96. Whittaker, S. R., Walton, M. I., Garrett, M. D., & Workman, P.
(2004). The cyclin-dependent kinase inhibitor CYC202 (R-
Roscovitine) inhibits retinoblastoma protein phosphorylation,
causes loss of Cyclin D1, and activates the mitogen-activated pro-
tein kinase pathway. Cancer Research, 64(1), 262272. https://
doi.org/10.1158/0008-5472.Can-03-0110.
97. McClue, S. J., Blake, D., Clarke, R., Cowan, A., Cummings, L.,
Fischer, P. M., MacKenzie, M., Melville, J., Stewart, K., Wang, S.,
Zhelev, N., Zheleva, D., & Lane, D. P. (2002). In vitro and in vivo
antitumor properties of the cyclin dependent kinase inhibitor
CYC202 (R-roscovitine). International Journal of Cancer,
102(5), 463468. https://doi.org/10.1002/ijc.10738.
98. Benson, C., White, J., Bono, J. D., O'Donnell, A., Raynaud, F.,
Cruickshank, C., McGrath, H., Walton, M., Workman, P., Kaye,
S., Cassidy, J., Gianella-Borradori, A., Judson, I., & Twelves, C.
(2007). A phaseI trial of the selectiveoral cyclin-dependent kinase
inhibitor seliciclib (CYC202; R-Roscovitine), administered twice
daily for 7 days every 21 days. British Journal of Cancer, 96(1),
2937. https://doi.org/10.1038/sj.bjc.6603509.
99. Misra, R. N., Xiao, H.-y., Kim, K. S., Lu, S., Han, W.-C., Barbosa,
S. A., et al. (2004). N-(Cycloalkylamino)acyl-2-aminothiazole in-
hibitors of cyclin-dependent kinase 2. N-[5-[[[5-(1,1-
Dimethylethyl)-2-oxazolyl]methyl]thio]-2-thiazolyl]-4-
piperidinecarboxamide (BMS-387032), a highly efficacious and
selective antitumor agent. Journal of Medicinal Chemistry, 47(7),
17191728. https://doi.org/10.1021/jm0305568.
100. Nuwayhid, S., Stockett, D., Hyde, J., Aleshin, A., Walker, D. H.,
& Arkin, M. R. SNS-032 is a potent and selective inhibitor of
Cdk2, 7 and 9 and induces cell death by inhibiting cell cycle
progression and the expression of antiapoptotic proteins. In Proc
Am Assoc Cancer Res, 2006 (Vol. 47, pp. 491).
101. Tong, W.-G., Chen, R., Plunkett, W., Siegel, D., Sinha, R.,Harvey,
R. D., Badros, A. Z., Popplewell, L., Coutre, S., Fox, J. A.,
Mahadocon, K., Chen, T., Kegley, P., Hoch, U., & Wierda, W.
G. (2010). Phase I and pharmacologic study of SNS-032, a potent
and selective Cdk2, 7, and 9 inhibitor, in patients with advanced
chronic lymphocytic leukemia and multiple myeloma. Journal of
Clinical Oncology, 28(18), 30153022. https://doi.org/10.1200/
jco.2009.26.1347.
102. Heath, E. I., Bible, K., Martell, R. E., Adelman, D. C., & LoRusso,
P. M. (2008). A phase 1 study of SNS-032 (formerly BMS-
387032), a potent inhibitor of cyclin-dependent kinases 2, 7 and
9 administered as a single oral dose and weekly infusion in pa-
tients with metastatic refractory solid tumors. Investigational New
Drugs, 26(1), 5965. https://doi.org/10.1007/s10637-007-9090-3.
103. Aklilu, M., Kindler, H. L., Donehower, R. C., Mani, S., & Vokes,
E. E. (2003). Phase II study of flavopiridol in patients with ad-
vanced colorectal cancer. Annals of Oncology, 14(8), 12701273.
https://doi.org/10.1093/annonc/mdg343.
104. Burdette-Radoux, S., Tozer, R. G., Lohmann, R. C., Quirt, I.,
Ernst, D. S., Walsh, W., Wainman, N., Colevas, A. D., &
Eisenhauer, E. A. (2004). Phase II trial of flavopiridol, a cyclin
dependent kinase inhibitor, in untreated metastatic malignant mel-
anoma. Investigational New Drugs, 22(3), 315322. https://doi.
org/10.1023/B:DRUG.0000026258.02846.1c.
105. Le Tourneau, C., Faivre, S., Laurence, V., Delbaldo, C., Vera, K.,
Girre, V., et al. (2010). Phase I evaluation of seliciclib (R-
roscovitine), a novel oral cyclin-dependent kinase inhibitor, in
patients with advanced malignancies. European Journal of
Cancer, 46(18), 32433250. https://doi.org/10.1016/j.ejca.2010.
08.001.
106. Marra, A., & Curigliano, G. (2019). Are all cyclin-dependent ki-
nases 4/6 inhibitors created equal? npj Breast Cancer, 5(1), 27.
https://doi.org/10.1038/s41523-019-0121-y.
107. Hazel, P., Kroll, S. H. B., Bondke, A., Barbazanges, M., Patel, H.,
Fuchter, M. J., Coombes, R. C., Ali, S., Barrett, A. G. M., &
Freemont, P. S. (2017). Inhibitor selectivity for cyclin-dependent
kinase 7: a structural, thermodynamic, and Modelling study.
ChemMedChem, 12(5), 372380. https://doi.org/10.1002/cmdc.
201600535.
108. Kelso, T. W., Baumgart, K., Eickhoff, J., Albert, T., Antrecht, C.,
Lemcke, S., et al. (2014). Cyclin-dependent kinase 7 controls
mRNA synthesis by affecting stability of preinitiation complexes,
leading to altered gene expression, cell cycle progression, and
survival of tumor cells. Molecular and Cellular Biology, 34(19),
36753688. https://doi.org/10.1128/mcb.00595-14.
109. Choi, Y. J., Kim, D. H., Yoon, D. H., Suh, C., Choi, C.-M., Lee, J.
C., Hong, J. Y., & Rho, J. K. (2019). Efficacy of the novel CDK7
inhibitor QS1189 in mantle cell lymphoma. Scientific Reports,
9(1), 7193. https://doi.org/10.1038/s41598-019-43760-z.
110. Wang, B. Y., Liu, Q. Y., Cao, J., Chen, J. W., & Liu, Z. S. (2016).
Selective CDK7 inhibition with BS-181 suppresses cell prolifera-
tion and induces cell cycle arrest and apoptosis in gastric cancer.
Drug Design, Development and Therapy, 10, 11811189. https://
doi.org/10.2147/dddt.S86317.
111. Gong, Y., Yang, J., Liu, F., Li, Z., Gong, R., & Wei, T. (2018).
Cyclin-dependent kinase 7 is a potential therapeutic target in pap-
illary thyroid carcinoma. Journal of Biological Regulators and
Homeostatic Agents, 32(6), 13611368.
112. Clark, K., Ainscow, E., Peall, A., Thomson, S., Leishman, A.,
Elaine, S., et al. (2017). CT7001, a novel orally bio-available
821Cancer Metastasis Rev (2020) 39:805–823
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
CDK7 inhibitor, is highly active in in-vitro and in-vivo models of
AML. Blood, 130(Suppl 1), 26452645.
113. Hutterer, C., Eickhoff, J., Milbradt, J., Korn, K., Zeittrager, I.,
Bahsi, H., et al. (2015). A novel CDK7 inhibitor of the
pyrazolotriazine class exerts broad-spectrum antiviral activity at
nanomolar concentrations. Antimicrobial Agents and
Chemotherapy, 59(4), 20622071. https://doi.org/10.1128/aac.
04534-14.
114. Christensen, C. L., Kwiatkowski, N., Abraham, B. J., Carretero, J.,
Al-Shahrour, F., Zhang, T., et al. (2014). Targeting transcriptional
addictions in small cell lung cancer with a covalent CDK7 inhib-
itor. Cancer Cell, 26(6), 909922. https://doi.org/10.1016/j.ccell.
2014.10.019.
115. Cayrol, F., Praditsuktavorn, P., Fernando, T. M., Kwiatkowski, N.,
Marullo, R., Calvo-Vidal, M. N., Phillip, J., Pera, B., Yang, S. N.,
Takpradit, K., Roman, L., Gaudiano, M., Crescenzo, R., Ruan, J.,
Inghirami,G., Zhang, T., Cremaschi, G.,Gray, N. S., & Cerchietti,
L. (2017). THZ1 targeting CDK7 suppresses STAT transcriptional
activity and sensitizes T-cell lymphomas to BCL2 inhibitors.
Nature Communications, 8(1), 14290. https://doi.org/10.1038/
ncomms14290.
116. Nagaraja, S., Vitanza, N. A., Woo, P. J., Taylor, K. R., Liu, F.,
Zhang, L., et al. (2017). Transcriptional dependencies in diffuse
intrinsic pontine glioma. Cancer Cell, 31(5), 635652.e636.
https://doi.org/10.1016/j.ccell.2017.03.011.
117. Greenall, S. A., Lim,Y. C., Mitchell, C. B., Ensbey, K. S., Stringer,
B. W., Wilding, A. L., O'Neill, G. M., McDonald, K. L., Gough,
D. J., Day, B. W., & Johns, T. G. (2017). Cyclin-dependent kinase
7 is a therapeutic target in high-grade glioma. Oncogenesis, 6(5),
e336. https://doi.org/10.1038/oncsis.2017.33.
118. Eliades, P., Abraham, B. J., Ji, Z., Miller, D. M., Christensen, C.
L., Kwiatkowski, N., Kumar, R., Njauw, C. N., Taylor, M., Miao,
B., Zhang, T., Wong, K.K., Gray, N. S., Young, R. A., & Tsao, H.
(2018). High MITF expression is associated with super-enhancers
and suppressed by CDK7 inhibition in melanoma. Journal of
Investigative Dermatology, 138(7), 15821590. https://doi.org/
10.1016/j.jid.2017.09.056.
119. Zhong, L., Yang, S., Jia, Y., & Lei, K. (2018). Inhibition of cyclin-
dependent kinase 7 suppresses human hepatocellular carcinoma
by inducing apoptosis. Journal of Cellular Biochemistry,
119 (12), 97429751. https://doi.org/10.1002/jcb.27292.
120. Cao, X., Dang, L., Zheng, X., Lu, Y., Lu, Y., Ji, R., Zhang, T.,
Ruan, X., Zhi, J., Hou, X., Yi, X., Li, M. J., Gu, T., Gao, M.,
Zhang, L., & Chen, Y. (2019). Targeting super-enhancer-driven
oncogenic transcription by CDK7 inhibition in anaplastic thyroid
carcinoma. Thyroid, 29(6), 809823. https://doi.org/10.1089/thy.
2018.0550.
121. Lu, P., Geng, J., Zhang, L., Wang, Y., Niu, N., Fang, Y., Liu, F.,
Shi, J., Zhang, Z. G., Sun, Y. W., Wang, L. W., Tang, Y., & Xue, J.
(2019). THZ1 reveals CDK7-dependent transcriptional addictions
in pancreatic cancer. Oncogene, 38(20), 39323945. https://doi.
org/10.1038/s41388-019-0701-1.
122. Zhong, S., Zhang, Y., Yin, X., & Di, W. (2019). CDK7 inhibitor
suppresses tumor progression through blocking the cell cycle at
the G2/M phase and inhibiting transcriptional activity in cervical
cancer. OncoTargets and Theraphy, 12,21372147. https://doi.
org/10.2147/ott.S195655.
123. Wang, Y., Zhang, T., Kwiatkowski, N., Abraham, B. J., Lee, T. I.,
Xie, S., Yuzugullu, H., von, T., Li, H., Lin, Z., Stover, D. G., Lim,
E., Wang, Z. C., Iglehart, J.D., Young, R. A., Gray, N. S., & Zhao,
J. J. (2015). CDK7-dependent transcriptional addiction in triple-
negative breast cancer. Cell, 163(1), 174186. https://doi.org/10.
1016/j.cell.2015.08.063.
124. Zhang, Y., Zhou, L., Bandyopadhyay, D., Sharma, K., Allen, A. J.,
Kmieciak, M., & Grant, S. (2019). The covalent CDK7 inhibitor
THZ1 potently induces apoptosis in multiple myeloma cells
in vitro and in vivo.Clinical Cancer Research, 25,61956205.
https://doi.org/10.1158/1078-0432.Ccr-18-3788.
125. Harrod, A., Fulton, J., Nguyen, V. T. M., Periyasamy, M., Ramos-
Garcia, L., Lai, C. F., et al. (2016). Genomic modelling of the
ESR1 Y537S mutation for evaluating function and new therapeu-
tic approaches for metastatic breast cancer. Oncogene, 36, 2286,
https://doi.org/10.1038/onc.2016.382 https://www.nature.com/
articles/onc2016382#supplementary-information.
126. Durbin, A. D., Zimmerman, M. W., Dharia, N. V., Abraham, B. J.,
Iniguez, A. B., Weichert-Leahey, N., He, S., Krill-Burger, J. M.,
Root, D. E., Vazquez, F., Tsherniak, A., Hahn, W. C., Golub, T. R.,
Young, R. A., Look, A. T., & Stegmaier, K.(2018). Selective gene
dependencies in MYCN-amplified neuroblastoma include the core
transcriptional regulatory circuitry. Nature Genetics, 50(9), 1240
1246. https://doi.org/10.1038/s41588-018-0191-z.
127. Huang, T., Ding, X., Xu, G., Chen, G., Cao,Y., Peng, C., Shen, S.,
Lv, Y., Wang, L., & Zou, X. (2019). CDK7 inhibitor THZ1 in-
hibits MCL1 synthesis and drives cholangiocarcinoma apoptosis
in combination with BCL2/BCL-XL inhibitor ABT-263. Cell
Death & Disease, 10(8), 602. https://doi.org/10.1038/s41419-
019-1831-7.
128. Kalan, S., Amat, R., Schachter, M. M., Kwiatkowski, N.,
Abraham, B. J., Liang, Y., Zhang, T., Olson, C. M., Larochelle,
S., Young, R. A., Gray, N. S., & Fisher, R. P. (2017). Activation of
the p53 transcriptional program sensitizes cancer cells to Cdk7
inhibitors. Cell Reports, 21(2), 467481. https://doi.org/10.1016/
j.celrep.2017.09.056.
129. Huang, J.-R., Qin, W.-M., Wang, K., Fu, D.-R., Zhang, W.-J.,
Jiang, Q.-W., Yang, Y., Yuan, M. L., Xing, Z. H., Wei, M. N.,
Li, Y., & Shi, Z. (2018). Cyclin-dependent kinase 7 inhibitor
THZ2 inhibits the growth of human gastric cancer in vitro and
in vivo. American Journal of Translational Research, 10(11),
36643676.
130. Olson, C. M., Liang, Y., Leggett, A., Park, W. D., Li, L., Mills, C.
E., Elsarrag, S. Z., Ficarro, S. B., Zhang, T., Düster, R., Geyer, M.,
Sim, T., Marto, J. A., Sorger, P. K., Westover, K. D., Lin, C. Y.,
Kwiatkowski, N., & Gray, N. S. (2019). Development of a selec-
tive CDK7 covalent inhibitor reveals predominant cell-cycle phe-
notype. Cell Chemical Biology, 26,792803.e10. https://doi.org/
10.1016/j.chembiol.2019.02.012.
131. Zhang, H., Christensen, C. L., Dries, R., Oser, M. G., Deng, J.,
Diskin, B., et al. (2020). CDK7 inhibition potentiates genome
instability triggering anti-tumor immunity in small cell lung can-
cer. Cancer Cell, 37(1), 3754.e39. https://doi.org/10.1016/j.ccell.
2019.11.003.
132. Zhang, T., Kwiatkowski, N., Olson, C. M., Dixon-Clarke, S. E.,
Abraham, B. J., Greifenberg, A. K., et al. (2016). Covalent
targeting of remote cysteine residues to develop CDK12 and
CDK13 inhibitors. Nature Chemical Biology, 12, 876. https://
doi.org/10.1038/nchembio.2166 https://www.nature.com/articles/
nchembio.2166#supplementary-information.
133. MacCallum, D. E., Melville, J., Frame, S., Watt, K., Anderson, S.,
Gianella-Borradori, A., Lane, D. P., & Green, S. R. (2005).
Seliciclib (CYC202, R-Roscovitine) induces cell death in multiple
myeloma cells by inhibition of RNA polymerase IIdependent
transcription and down-regulation of Mcl-1. Cancer Research,
65(12), 53995407. https://doi.org/10.1158/0008-5472.Can-05-
0233.
134. Shandilya, J., Wang, Y., & Roberts, S. G. E. (2012). TFIIB de-
phosphorylation links transcription inhibition with the p53-
dependent DNA damage response. Proceedings of the National
Academy of Sciences, 109(46), 1879718802. https://doi.org/10.
1073/pnas.1207483109.
135. Wong, R. W. J., Ngoc, P. C. T., Leong, W. Z., Yam, A. W. Y.,
Zhang, T., Asamitsu, K., Iida, S., Okamoto, T., Ueda, R., Gray, N.
S., Ishida, T., & Sanda, T. (2017). Enhancer profiling identifies
822 Cancer Metastasis Rev (2020) 39:805–823
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
critical cancer genes and characterizes cell identity in adult T-cell
leukemia. Blood, 130(21), 23262338. https://doi.org/10.1182/
blood-2017-06-792184.
136. Sharifnia, T., Wawer, M. J., Chen, T., Huang, Q.-Y., Weir, B. A.,
Sizemore, A., Lawlor, M. A., Goodale, A., Cowley, G. S.,
Vazquez, F., Ott, C. J., Francis, J. M., Sassi, S., Cogswell, P.,
Sheppard, H. E., Zhang, T., Gray, N. S., Clarke, P. A., Blagg, J.,
Workman, P., Sommer, J., Hornicek,F., Root, D. E., Hahn, W. C.,
Bradner, J. E., Wong, K. K., Clemons,P. A., Lin, C. Y., Kotz, J. D.,
& Schreiber, S. L. (2019). Small-molecule targeting of brachyury
transcription factor addiction in chordoma. Nature Medicine,
25(2), 292300. https://doi.org/10.1038/s41591-018-0312-3.
137. Rasool, R. U., Natesan, R., Deng, Q., Aras, S., Lal, P., Sander
Effron, S., et al. (2019). CDK7 inhibition suppresses castration-
resistant prostate cancer through MED1 inactivation. Cancer
Discovery, CD-19-0189. https://doi.org/10.1158/2159-8290.Cd-
19-0189.
138. Jeselsohn, R., Bergholz, J. S., Pun, M., Cornwell, M., Liu, W.,
Nardone, A., et al. (2018). Allele-specific chromatin recruitment
and therapeutic vulnerabilities of ESR1 activating mutations.
Cancer Cell, 33(2), 173186.e175. https://doi.org/10.1016/j.
ccell.2018.01.004.
139. Sun, B., Mason, S., Wilson, R. C., Hazard, S. E., Wang, Y., Fang,
R., Wang, Q., Yeh, E. S., Yang, M., Roberts, T. M., Zhao, J. J., &
Wang, Q. (2019). Inhibition of the transcriptional kinase CDK7
overcomestherapeutic resistance in HER2-positive breast cancers.
Oncogene, 39,5063. https://doi.org/10.1038/s41388-019-0953-
9.
140. Martin, L.-A., Pancholi, S., Ribas, R., Gao, Q., Simigdala, N.,
Nikitorowicz-Buniak, J., et al. (2017). Abstract P3-03-09: resis-
tance to palbociclib depends on multiple targetable mechanisms
highlighting the potential of drug holidays and drug switching to
improve therapeutic outcome. Cancer Research, 77(4
Supplement), P3-03-09-P03-03-09. https://doi.org/10.1158/1538-
7445.Sabcs16-p3-03-09.
141. Alqahtani, A., Choucair, K., Ashraf, M., Hammouda, D. M.,
Alloghbi, A., Khan, T., Senzer, N., & Nemunaitis, J. (2019).
Bromodomain and extra-terminal motif inhibitors: a review of
preclinical and clinical advances in cancer therapy. Future
Science OA, 5(3), FSO372. https://doi.org/10.4155/fsoa-2018-
0115.
142. Gao, Y., Zhang, T., Terai, H., Ficarro, S. B., Kwiatkowski, N.,
Hao, M.-F., et al. (2018). Overcoming resistance to the THZ series
of covalent transcriptional CDK inhibitors. Cell Chemical
Biology, 25(2), 135142.e135. https://doi.org/10.1016/j.
chembiol.2017.11.007.
143. Sava, G. P., Fan, H., Fisher, R. A., Lusvarghi, S., Pancholi, S.,
Ambudkar, S. V., Martin, L. A., Charles Coombes, R., Buluwela,
L., & Ali, S. (2019). ABC-transporter upregulation mediates re-
sistance tothe CDK7 inhibitors THZ1 and ICEC0942. Oncogene,
39,651663. https://doi.org/10.1038/s41388-019-1008-y.
144. Robey, R. W., Pluchino, K. M., Hall, M. D., Fojo, A. T., Bates, S.
E., & Gottesman, M. M. (2018). Revisiting the role of ABC trans-
porters in multidrug-resistant cancer. Nature Reviews Cancer,
18(7), 452464. https://doi.org/10.1038/s41568-018-0005-8.
145. Kaliszczak, M., Patel, H., Kroll, S. H. B., Carroll, L., Smith, G.,
Delaney,S.,Heathcote,D.A.,Bondke,A.,Fuchter,M.J.,
Coombes, R. C., Barrett, A. G. M., Ali, S., & Aboagye, E. O.
(2013). Development of a cyclin-dependent kinase inhibitor de-
void of ABC transporter-dependent drug resistance. British
Journal of Cancer, 109(9), 23562367. https://doi.org/10.1038/
bjc.2013.584.
PublishersnoteSpringer Nature remains neutral with regard to jurisdic-
tional claims in published maps and institutional affiliations.
823Cancer Metastasis Rev (2020) 39:805–823
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com
... Cyclin-dependent kinase 7 (CDK7) plays two primary roles in regulating the cell cycle and transcription factors [6]. During cell cycle progression, CDK7 activates CDK1 and CDK2 during the S/G2 phases and CDK4/6 during the G1 phase via phosphorylation [7]. ...
... During transcription, CDK7 phosphorylates serine 5 (Ser5) and Ser7 in the RNA polymerase II (Pol II)-C-terminal domain (CTD) and CDK9, which induces Ser2 phosphorylation of the Pol II CTD [6]. CDK7/9 regulate cell division, gene transcription, and other important biological processes in normal cells; however, they are overexpressed in most carcinomas [8][9][10]. ...
... The development of therapies targeting CDK7 is ongoing for various cancers [6,13,14]. However, early CDK7 inhibitors were not initially CDK7-specific but instead acted as multi-CDK inhibitors. ...
Article
Full-text available
Triple-negative breast cancer (TNBC) accounts for approximately 15–20% of all breast cancer types, indicating a poor survival prognosis with a more aggressive biology of metastasis to the lung and a short response duration to available therapies. Ibulocydine (IB) is a novel (cyclin-dependent kinase) CDK7/9 inhibitor prodrug displaying potent anti-cancer effects against various cancer cell types. We performed in vitro and in vivo experiments to determine whether IB inhibits metastasis and eventually overcomes the poor drug response in TNBC. The result showed that IB inhibited the growth of TNBC cells by inducing caspase-mediated apoptosis and blocking metastasis by reducing MMP-9 expression in vitro. Concurrently, in vivo experiments using the metastasis model showed that IB inhibited metastasis of MDA-MB-231-Luc cells to the lung. Collectively, these results demonstrate that IB inhibited the growth of TNBC cells and blocked metastasis by regulating MMP-9 expression, suggesting a novel therapeutic agent for metastatic TNBC.
... Additionally, CDK7 activates the CDK9/cyclin T1 complex (TFEB) to promote the elongation process [21]. While CDK7 and CDK9 pathways have recently been targeted in cancer treatment [22,23], further elucidation is required regarding their involvement in the paraptotic program. ...
Article
Full-text available
Background Paraptosis is a programmed cell death characterized by cytoplasmic vacuolation, which has been explored as an alternative method for cancer treatment and is associated with cancer resistance. However, the mechanisms underlying the progression of paraptosis in cancer cells remain largely unknown. Methods Paraptosis-inducing agents, CPYPP, cyclosporin A, and curcumin, were utilized to investigate the underlying mechanism of paraptosis. Next-generation sequencing and liquid chromatography-mass spectrometry analysis revealed significant changes in gene and protein expressions. Pharmacological and genetic approaches were employed to elucidate the transcriptional events related to paraptosis. Xenograft mouse models were employed to evaluate the potential of paraptosis as an anti-cancer strategy. Results CPYPP, cyclosporin A, and curcumin induced cytoplasmic vacuolization and triggered paraptosis in cancer cells. The paraptotic program involved reactive oxygen species (ROS) provocation and the activation of proteostatic dynamics, leading to transcriptional activation associated with redox homeostasis and proteostasis. Both pharmacological and genetic approaches suggested that cyclin-dependent kinase (CDK) 7/9 drive paraptotic progression in a mutually-dependent manner with heat shock proteins (HSPs). Proteostatic stress, such as accumulated cysteine-thiols, HSPs, ubiquitin-proteasome system, endoplasmic reticulum stress, and unfolded protein response, as well as ROS provocation primarily within the nucleus, enforced CDK7/CDK9–Rpb1 (RNAPII subunit B1) activation by potentiating its interaction with HSPs and protein kinase R in a forward loop, amplifying transcriptional regulation and thereby exacerbating proteotoxicity leading to initiate paraptosis. The xenograft mouse models of MDA-MB-231 breast cancer and docetaxel-resistant OECM-1 head and neck cancer cells further confirmed the induction of paraptosis against tumor growth. Conclusions We propose a novel regulatory paradigm in which the activation of CDK7/CDK9–Rpb1 by nuclear proteostatic stress mediates transcriptional regulation to prime cancer cell paraptosis. Graphical Abstract
... Proteolysis-targeting chimeras (PROTACs) can selectively hijack BRD4, CDKs and TFs into the ubiquitin-proteasome system to elicit its degradation, resulting to interruption of SE-driven transcriptional program. CRISPR/Cas9-mediated genetic perturbation can directly targeting individual components within SEs an integral component of TF IIH (TFIIH), mediating the phosphorylation of RNA Pol II C-terminal domain (CTD) heptapeptide repeats at serine 5 (Ser5) and serine 7 (Ser7) residues to initiate transcription [167]. In addition, CDK7 phosphates to CDK9, a part of the positive transcription elongation factor b (P-TEFb), which then phosphorylates serine 2 (Ser2) residues within the RNA Pol II CTD to promote transcriptional elongation [168]. ...
Article
Full-text available
Metastasis remains the principal cause of cancer-related lethality despite advancements in cancer treatment. Dysfunctional epigenetic alterations are crucial in the metastatic cascade. Among these, super-enhancers (SEs), emerging as new epigenetic regulators, consist of large clusters of regulatory elements that drive the high-level expression of genes essential for the oncogenic process, upon which cancer cells develop a profound dependency. These SE-driven oncogenes play an important role in regulating various facets of metastasis, including the promotion of tumor proliferation in primary and distal metastatic organs, facilitating cellular migration and invasion into the vasculature, triggering epithelial-mesenchymal transition, enhancing cancer stem cell-like properties, circumventing immune detection, and adapting to the heterogeneity of metastatic niches. This heavy reliance on SE-mediated transcription delineates a vulnerable target for therapeutic intervention in cancer cells. In this article, we review current insights into the characteristics, identification methodologies, formation, and activation mechanisms of SEs. We also elaborate the oncogenic roles and regulatory functions of SEs in the context of cancer metastasis. Ultimately, we discuss the potential of SEs as novel therapeutic targets and their implications in clinical oncology, offering insights into future directions for innovative cancer treatment strategies.
... Preclinical investigations have consistently revealed robust anti-tumour effects across a range of cancer types, for example osteosarcoma [24], multiple myeloma [25], and breast carcinoma [26]. SY-1365, a modified variant of THZ1 and a CDK7 inhibitor, started its phase I clinical trial for the treatment of ovarian and breast carcinomas in 2017 [27,28]. ...
Article
Full-text available
Chordomas are very rare malignant neoplasms of the bone occurring almost exclusively along the spine. As the tumours are thought to arise from notochordal remnants, the vast majority of chordomas express the TBXT gene, resulting in detectable nuclear amounts of its gene product brachyury. This T-Box transcription factor is commonly recognised as being essential in chordoma cells, and limiting TBXT expression is thought to be the key factor in controlling this tumour. Although the tumour is rare, distinct molecular differences and vulnerabilities have been described with regard to its location and the progression status of the disease, rendering it mandatory for novel cell lines to reflect all relevant chordoma subtypes. Here, we describe a novel chordoma cell line arising from the pleural effusion of a disseminated, poorly differentiated chordoma. This cell line, U-CH22, represents a highly aggressive terminal chordoma and, therefore, fills a relevant gap within the panel of available cell culture models for this orphan disease. CDK7 and CDK9 inhibition was lately identified as being effective in reducing viability in four chordoma cell lines, most likely due to a reduction in brachyury levels. In this study, we determined the capability of the CDK7 inhibitor THZ1 and the CDK1/2/5/9 inhibitor dinaciclib to reduce TBXT expression at mRNA and protein levels in a broad range of nine cell lines that are models of primary, recurrent, and metastasised chordoma of the clivus and the sacrum.
... Cell cycle dysregulation, which is an established hallmark of cancer, has great significance for cancer treatment [32]. Flow cytometry results showed that many cells were in the G0/G1 after SP treatment. ...
... Two oncogenes promoting progression from G2 to M phase (CDK7 and CDC25B) show a similar mismatch of expression profiles after CRNDE silencing. The expression of the former is significantly reduced, which is consistent with the observed decrease in the expression of cyclin H (the CDK7 protein forms a complex with this cyclin, which promotes cell division [35]). However, at the same time, the concentration of the CDC25B gene transcript increases significantly, which in turn should exert an opposite, pro-proliferative physiological effect [36]. ...
Article
Full-text available
CRNDE is considered an oncogene expressed as long non-coding RNA. Our previous paper is the only one reporting CRNDE as a micropeptide-coding gene. The amino acid sequence of this micropeptide (CRNDEP) has recently been confirmed by other researchers. This study aimed at providing a mass spectrometry (MS)-based validation of the CRNDEP sequence and an investigation of how the differential expression of CRNDE(P) influences the metabolism and chemoresistance of ovarian cancer (OvCa) cells. We also assessed cellular localization changes of CRNDEP, looked for its protein partners, and bioinformatically evaluated its RNA-binding capacities. Herein, we detected most of the CRNDEP sequence by MS. Moreover, our results corroborated the oncogenic role of CRNDE, portraying it as the gene impacting carcinogenesis at the stages of DNA transcription and replication, affecting the RNA metabolism, and stimulating the cell cycle progression and proliferation, with CRNDEP being detected in the centrosomes of dividing cells. We also showed that CRNDEP is located in nucleoli and revealed interactions of this micropeptide with p54, an RNA helicase. Additionally, we proved that high CRNDE(P) expression increases the resistance of OvCa cells to treatment with microtubule-targeted cytostatics. Furthermore, altered CRNDE(P) expression affected the activity of the microtubular cytoskeleton and the formation of focal adhesion plaques. Finally, according to our in silico analyses, CRNDEP is likely capable of RNA binding. All these results contribute to a better understanding of the CRNDE(P) role in OvCa biology, which may potentially improve the screening, diagnosis, and treatment of this disease.
... A compelling candidate in this context is CDK7, a cyclin-dependent kinase-activating kinase that is integral to cell cycle progression and gene transcription [7]. Targeting CDK7 via various modalities, such as small interfering RNA (siRNA) or pharmacological inhibitors, has yielded encouraging antineoplastic results [8,9]. Recent advancements have ushered in a cadre of CDK7 inhibitors into phase I/II clinical trials for breast cancer [10]. ...
Article
Full-text available
Cyclin-dependent kinase 7 (CDK7) serves as a pivotal regulator in orchestrating cellular cycle dynamics and gene transcriptional activity. Elevated expression levels of CDK7 have been ubiquitously documented across a spectrum of malignancies and have been concomitantly correlated with adverse clinical outcomes. This review delineates the biological roles of CDK7 and explicates the molecular pathways through which CDK7 exacerbates the oncogenic progression of breast cancer. Furthermore, we synthesize the extant literature to provide a comprehensive overview of the advancement of CDK7-specific small-molecule inhibitors, encapsulating both preclinical and clinical findings in breast cancer contexts. The accumulated evidence substantiates the conceptualization of CDK7 as a propitious therapeutic target in breast cancer management.
Article
Full-text available
The cell cycle is tightly regulated to ensure controlled cell proliferation. Dysregulation of the cell cycle machinery is a hallmark of cancer that leads to unchecked growth. This review comprehensively analyzes key molecular regulators of the cell cycle and how they contribute to carcinogenesis when mutated or overexpressed. It focuses on cyclins, cyclin‐dependent kinases (CDKs), CDK inhibitors, checkpoint kinases, and mitotic regulators as therapeutic targets. Promising strategies include CDK4/6 inhibitors like palbociclib, ribociclib, and abemaciclib for breast cancer treatment. Other possible targets include the anaphase‐promoting complex/cyclosome (APC/C), Skp2, p21, and aurora kinase inhibitors. However, challenges with resistance have limited clinical successes so far. Future efforts should focus on combinatorial therapies, next‐generation inhibitors, and biomarkers for patient selection. Targeting the cell cycle holds promise but further optimization is necessary to fully exploit it as an anti‐cancer strategy across diverse malignancies.
Article
Full-text available
Hippo signaling controls organ size and tumor progression through a conserved pathway leading to nuclear translocation of the transcriptional effector Yki/Yap/Taz. Most of our understanding of Hippo signaling pertains to its cytoplasmic regulation, but how the pathway is controlled in the nucleus remains poorly understood. Here we uncover an evolutionarily conserved mechanism by which CDK7 promotes Yki/Yap/Taz stabilization in the nucleus to sustain Hippo pathway outputs. We found that a modular E3 ubiquitin ligase complex CRL4DCAF12 binds and targets Yki/Yap/Taz for ubiquitination and degradation, whereas CDK7 phosphorylates Yki/Yap/Taz at S169/S128/S90 to inhibit CRL4DCAF12 recruitment, leading to Yki/Yap/Taz stabilization. As a consequence, inactivation of CDK7 reduced organ size and inhibited tumor growth, which could be reversed by restoring Yki/Yap activity. Our study identifies an unanticipated layer of Hippo pathway regulation, defines a novel mechanism by which CDK7 regulates tissue growth, and implies CDK7 as a drug target for Yap/Taz-driven cancer.
Article
Full-text available
Elevated glucose consumption is fundamental to cancer, but selectively targeting this pathway is challenging. We develop a high-throughput assay for measuring glucose consumption and use it to screen non-small-cell lung cancer cell lines against bioactive small molecules. We identify Milciclib that blocks glucose consumption in H460 and H1975, but not in HCC827 or A549 cells, by decreasing SLC2A1 (GLUT1) mRNA and protein levels and by inhibiting glucose transport. Milciclib blocks glucose consumption by targeting cyclin-dependent kinase 7 (CDK7) similar to other CDK7 inhibitors including THZ1 and LDC4297. Enhanced PIK3CA signaling leads to CDK7 phosphorylation, which promotes RNA Polymerase II phosphorylation and transcription. Milciclib, THZ1, and LDC4297 lead to a reduction in RNA Polymerase II phosphorylation on the SLC2A1 promoter. These data indicate that our high-throughput assay can identify compounds that regulate glucose consumption and that CDK7 is a key regulator of glucose consumption in cells with an activated PI3K pathway. Many cancer cells have increased glucose consumption compared to normal cells, a feature that can be exploited therapeutically. Here, the authors carry out a chemical screen and identify compounds that selectively blocks glucose metabolism in non-small-cell lung cancer cell lines.
Article
Full-text available
The CDK7 inhibitors (CDK7i) ICEC0942 and THZ1, are promising new cancer therapeutics. Resistance to targeted drugs frequently compromises cancer treatment. We sought to identify mechanisms by which cancer cells may become resistant to CDK7i. Resistant lines were established through continuous drug selection. ABC-transporter copy number, expression and activity were examined using real-time PCR, immunoblotting and flow cytometry. Drug responses were measured using growth assays. ABCB1 was upregulated in ICEC0942-resistant cells and there was cross-resistance to THZ1. THZ1-resistant cells upregulated ABCG2 but remained sensitive to ICEC0942. Drug resistance in both cell lines was reversible upon inhibition of ABC-transporters. CDK7i response was altered in adriamycin- and mitoxantrone-resistant cell lines demonstrating ABC-transporter upregulation. ABCB1 expression correlated with ICEC0942 and THZ1 response, and ABCG2 expression with THZ2 response, in a panel of cancer cell lines. We have identified ABCB1 upregulation as a common mechanism of resistance to ICEC0942 and THZ1, and confirmed that ABCG2 upregulation is a mechanism of resistance to THZ1. The identification of potential mechanisms of CDK7i resistance and differences in susceptibility of ICEC0942 and THZ1 to ABC-transporters, may help guide their future clinical use.
Article
Full-text available
Metastatic castration-resistant prostate cancer (CRPC) is a fatal disease, primarily resulting from the transcriptional addiction driven by androgen receptor (AR). First-line CRPC treatments typically target AR signaling, but are rapidly bypassed, resulting in only a modest survival benefit with antiandrogens. Therapeutic approaches that more effectively block the AR-transcriptional axis are urgently needed. Here, we investigated the molecular mechanism underlying the association between the transcriptional coactivator MED1 and AR as a vulnerability in AR-driven CRPC. MED1 undergoes CDK7-dependent phosphorylation at T1457 and physically engages AR at superenhancer sites, and is essential for AR-mediated transcription. In addition, a CDK7-specific inhibitor, THZ1, blunts AR-dependent neoplastic growth by blocking AR/MED1 corecruitment genome-wide, as well as reverses the hyperphosphorylated MED1-associated enzalutamide-resistant phenotype. In vivo, THZ1 induces tumor regression of AR-amplified human CRPC in a xenograft mouse model. Together, we demonstrate that CDK7 inhibition selectively targets MED1-mediated, AR-dependent oncogenic transcriptional amplification, thus representing a potential new approach for the treatment of CRPC. Significance Potent inhibition of AR signaling is critical to treat CRPC. This study uncovers a driver role for CDK7 in regulating AR-mediated transcription through phosphorylation of MED1, thus revealing a therapeutically targetable potential vulnerability in AR-addicted CRPC. See related commentary by Russo et al., p. 1490. This article is highlighted in the In This Issue feature, p. 1469
Article
Full-text available
The harnessing in clinical practice of cyclin-dependent kinases 4/6 inhibitors, namely palbociclib, ribociclib, and abemaciclib, has substantially changed the therapeutic approach for hormone receptor-positive metastatic breast cancer (BC). Phase II–III clinical trials evaluating the addition of these agents to standard endocrine therapy reported consistent improvements in response rates and progression-free survival as well as manageable toxicity profiles and excellent impact on patients’ quality of life. Hence, pivotal trials provided comparable results among different cyclin-dependent kinases 4/6 inhibitors, there is an increasing interest in finding substantial differences in order to implement their use in clinical practice. The aim of this paper is to summarize the current evidences raised from preclinical and clinical studies on cyclin-dependent kinases 4/6 inhibitors in BC, focusing on differences in terms of pharmacological properties, toxicity profile, and patients’ quality of life.
Article
Full-text available
Resistance of breast cancer to human epidermal growth factor receptor 2 (HER2) inhibitors involves reprogramming of the kinome through HER2/HER3 signaling via the activation of multiple tyrosine kinases and transcriptional upregulation. The heterogeneity of induced kinases prevents kinase targeting by a single kinase inhibitor and presents a major challenge to the treatment of therapeutically recalcitrant HER2-positive breast cancers (HER2+ BCs). As a result, there is a critical need for effective treatment that attacks the aberrant kinome activation associated with resistance to HER2-targeted therapy. Here, we describe a novel treatment strategy that targets cyclin-dependent kinase 7 (CDK7) in HER2 inhibitor-resistant (HER2iR) breast cancer. We show that both HER2 inhibitor-sensitive (HER2iS) and HER2iR breast cancer cell lines exhibit high sensitivity to THZ1, a newly identified covalent inhibitor of the transcription regulatory kinase CDK7. CDK7 promotes cell cycle progression through inhibition of transcription, rather than via direct phosphorylation of classical CDK targets. The transcriptional kinase activity of CDK7 is regulated by HER2, and by the receptor tyrosine kinases activated in response to HER2 inhibition, as well as by the downstream SHP2 and PI3K/AKT pathways. A low dose of THZ1 displayed potent synergy with the HER2 inhibitor lapatinib in HER2iR BC cells in vitro. Dual HER2 and CDK7 inhibition induced tumor regression in two HER2iR BC xenograft models in vivo. Our data support the utilization of CDK7 inhibition as an additional therapeutic avenue that blocks the activation of genes engaged by multiple HER2iR kinases.
Article
Full-text available
Cholangiocarcinoma (CCA) is a fatal disease without effective targeted therapy. We screened a small-molecule library of 116 inhibitors targeting different targets of the cell cycle and discovered several kinases, including Cyclin-dependent kinase 7 (CDK7) as vulnerabilities in CCA. Analysis of multiple CCA data sets demonstrated that CDK7 was overexpressed in CCA tissues. Further studies demonstrated that CDK7 inhibitor THZ1 inhibited cell viability and induced apoptosis in CCA cells. We also showed that THZ1 inhibited CCA cell growth in a xenograft model. RNA-sequencing followed by Gene ontology analysis showed a striking impact of THZ1 on DNA-templated transcriptional programs. THZ1 downregulated CDK7-mediated phosphorylation of RNA polymerase II, indicative of transcriptional inhibition. A number of oncogenic transcription factors and survival proteins, like MCL1, FOSL1, and RUNX1, were repressed by THZ1. MCL1, one of the antiapoptotic BCL2 family members, was significantly inhibited upon THZ1 treatment. Accordingly, combining THZ1 with a BCL2/BCL-XL inhibitor ABT-263 synergized in impairing cell growth and driving apoptosis. Our results demonstrate CDK7 as a potential target in treating CCA. Combinations of CDK7 inhibition and BCL2/BCL-XL inhibition may offer a novel therapeutic strategy for CCA.
Conference Paper
p>Introduction: CDK7 is a key regulator of transcription and cell cycle progression and has been implicated in multiple tumor types driven by aberrant transcriptional control (e.g. MYC- , ESR1- activation) and/or aberrant cell cycle control (e.g. RB1 , CCNE1, CDKN2A alterations). SY-5609 is an oral, potent, and highly selective CDK7 inhibitor that is advancing through IND-enabling studies to support initiation of a planned Phase 1 oncology trial in early 2020. Here we report on the relationship between pharmacokinetics (PK), pharmacodynamics (PD), and tumor growth inhibition (TGI) in xenograft models of tumor types with transcriptional and/or cell cycle aberrations including high grade serous ovarian cancer (HGSOC), small cell lung cancer (SCLC), triple negative breast cancer (TNBC), and estrogen receptor positive breast cancer (ER+BC). Methods: The relationship between SY-5609 PK, PD, and TGI was evaluated in xenograft models of HGSOC (OVCAR3) and TNBC (HCC70) using once daily (QD) or twice daily (BID) dosing via oral gavage. SY-5609 TGI was evaluated as a single agent (SA) QD in patient-derived xenograft (PDX) models of HGSOC (n=3), SCLC (n=4), TNBC (n=4), and in combination with once weekly fulvestrant in ER+BC PDX models selected in vivo for resistance to the CDK4/6 inhibitor palbociclib (n=1) or resistance to both palbociclib and fulvestrant (n=1). Results: SY-5609 plasma exposure was dose proportional and did not accumulate after repeated therapeutic doses. Dose-dependent transcriptional responses in xenograft tissue were observed within 4 hours of SY-5609 dosing and were sustained for 24 hours. TGI was dose-dependent, with tumor regressions observed at doses significantly below the maximum tolerated dose (MTD). Similar TGI was seen when the same daily dose was administered either QD or BID suggesting that the effect was driven by overall exposure or minimum concentration. In HGSOC, SCLC, and TNBC PDX models, SA SY-5609 induced >50% TGI in all models tested (11/11), with 7/11 (64%) demonstrating robust anti-tumor activity (≥90% TGI or regression): 3/3 HGSOC, 2/4 SCLC, and 2/4 TNBC. In a palbociclib-resistant ER+BC PDX model, the combination of SY-5609 and fulvestrant induced significant TGI (89%), with no evident tumor regrowth up to 21 days after dosing cessation, distinguishing the observed effects from SY-5609 SA or fulvestrant SA. In a palbociclib and fulvestrant double-resistant ER+ BC PDX model, SY-5609 SA resulted in 33% TGI and fulvestrant SA had no activity. In contrast, the combination of SY-5609 and fulvestrant demonstrated significant TGI (68%; p<0.0001 vs fulvestrant SA), suggesting re-sensitization to fulvestrant. Conclusions: We have characterized PK, PD, and anti-tumor activity of the CDK7 inhibitor SY-5609 in a series of xenograft models. Exposure in plasma is dose-proportional and does not accumulate at therapeutic dose levels. SY-5609 induces dose-dependent transcriptional responses in tumor xenograft tissue and shows robust TGI, including regressions, in PDX models derived from multiple solid tumor indications. These results highlight the broad potential for SY-5609 across a variety of solid tumor types, including treatment resistant ER+BC, and support the development of SY-5609 in early phase clinical trials. Citation Format: Liv H Johannessen, Shanhu Hu, Nan Ke, Anthony D9Ippolito, Nisha Rajagopal, Jason Marineau, Anneli Savinainen, William Zamboni, Graeme Hodgson. Preclinical evaluation of PK, PD, and antitumor activity of the oral, non-covalent, potent and highly selective CDK7 inhibitor, SY-5609, provides rationale for clinical development in multiple solid tumor indications [abstract]. In: Proceedings of the AACR-NCI-EORTC International Conference on Molecular Targets and Cancer Therapeutics; 2019 Oct 26-30; Boston, MA. Philadelphia (PA): AACR; Mol Cancer Ther 2019;18(12 Suppl):Abstract nr C091. doi:10.1158/1535-7163.TARG-19-C091</p
Article
Cyclin-dependent kinase 7 (CDK7) is a central regulator of the cell cycle and gene transcription. However, little is known about its impact on genomic instability and cancer immunity. Using a selective CDK7 inhibitor, YKL-5-124, we demonstrated that CDK7 inhibition predominately disrupts cell-cycle progression and induces DNA replication stress and genome instability in small cell lung cancer (SCLC) while simultaneously triggering immune-response signaling. These tumor-intrinsic events provoke a robust immune surveillance program elicited by T cells, which is further enhanced by the addition of immune-checkpoint blockade. Combining YKL-5-124 with anti-PD-1 offers significant survival benefit in multiple highly aggressive murine models of SCLC, providing a rationale for new combination regimens consisting of CDK7 inhibitors and immunotherapies.
Conference Paper
Previously, we reported on a series of highly potent, selective, and non-covalent CDK7 inhibitors that demonstrated antiproliferative activity against triple-negative breast cancer (TNBC) and ovarian cancer (OVA) cell lines and tumor growth inhibition in cell line-derived (CDX) and patient-derived (PDX) mouse xenograft models. Here, we report on the in vitro and profile of our development candidate, SY-5609. Methods: Kinase inhibition assays at both Km and 2 mM [ATP] were used to assess inhibition of CDK2, CDK7, CDK9, and CDK12. SPR was used to determine the Kd, kon, and koff binding characteristics of SY-5609 to immobilized CDK7/Cyclin H dimer. CDK7 compound occupancy was determined using a biotinylated small molecule probe to pull down free CDK7 following incubating of HL60 cells with SY-5609. Inhibition of tumor cell line growth was assessed following 72 hrs of incubation with SY-5609. Flow cytometry was used to assess apoptosis and cell cycle modulation after 48 hrs of treatment. Effects on DNA damage and repair were assessed by immunofluorescence staining for γH2AX and RAD51 proteins. To assess effects, mice were implanted subcutaneously and randomized for treatment when tumors reached 150-200 mm³ and dosed orally for 3 weeks by both QD and BID dosing regimens. Collected tumor tissue samples were analyzed for protein levels of MCL1, pCDK2, MYC, and RNA Pol II CTD pSer5 by western blot. Results: SY-5609 bound CDK7/Cyclin H with a Kd of 0.059 nM and occupied CDK7 in HL60 cells with an EC50 of 33 nM. Cell growth inhibition EC50 values were 6-17 nM in a panel of solid tumor cell lines. Selectivity of SY-5609 over CDK12, CDK9, and CDK2 was 2492-, 2508-, and 8068-fold, respectively. SY-5609 led to induction of apoptosis, cell cycle arrest, and inhibition of DNA damage repair in tumor cell lines. Dose-dependent tumor growth inhibition was observed in a panel of CDX and PDX solid tumor models with both QD and BID dosing of SY-5609 with resulting decreases in direct (pCDK2, RNA Pol II CTD pSer5) and indirect (MCL1, MYC) protein biomarkers. In summary, we describe SY-5609, an orally available, potent, and selective CDK7 inhibitor that drives strong responses in CDX and PDX tumor models. These data support the rationale for advancing SY-5609 into IND-enabling studies. Citation Format: Shanhu Hu, Jason Marineau, Kristin Hamman, Michael Bradley, Anneli Savinainen, Sydney Alnemy, Nisha Rajagopal, David Orlando, Claudio Chuaqui, Eric Olson. SY-5609, an orally available selective CDK7 inhibitor demonstrates broad anti-tumor activity [abstract]. In: Proceedings of the American Association for Cancer Research Annual Meeting 2019; 2019 Mar 29-Apr 3; Atlanta, GA. Philadelphia (PA): AACR; Cancer Res 2019;79(13 Suppl):Abstract nr 4421.