ArticlePDF AvailableLiterature Review

A WEE1 family business: Regulation of mitosis, cancer progression, and therapeutic target

Authors:

Abstract and Figures

The inhibition of the DNA damage response (DDR) pathway in the treatment of cancer has recently gained interest, and different DDR inhibitors have been developed. Among them, the most promising ones target the WEE1 kinase family, which has a crucial role in cell cycle regulation and DNA damage identification and repair in both nonmalignant and cancer cells. This review recapitulates and discusses the most recent findings on the biological function of WEE1/PKMYT1 during the cell cycle and in the DNA damage repair, with a focus on their dual role as tumor suppressors in nonmalignant cells and pseudo-oncogenes in cancer cells. We here report the available data on the molecular and functional alterations of WEE1/PKMYT1 kinases in both hematological and solid tumors. Moreover, we summarize the preclinical information on 36 chemo/radiotherapy agents, and in particular their effect on cell cycle checkpoints and on the cellular WEE1/PKMYT1-dependent response. Finally, this review outlines the most important pre-clinical and clinical data available on the efficacy of WEE1/PKMYT1 inhibitors in monotherapy and in combination with chemo/radiotherapy agents or with other selective inhibitors currently used or under evaluation for the treatment of cancer patients.
This content is subject to copyright. Terms and conditions apply.
R E V I E W Open Access
A WEE1 family business: regulation of
mitosis, cancer progression, and
therapeutic target
Andrea Ghelli Luserna di Rorà, Claudio Cerchione, Giovanni Martinelli and Giorgia Simonetti
*
Abstract
The inhibition of the DNA damage response (DDR) pathway in the treatment of cancer has recently gained interest,
and different DDR inhibitors have been developed. Among them, the most promising ones target the WEE1 kinase
family, which has a crucial role in cell cycle regulation and DNA damage identification and repair in both
nonmalignant and cancer cells. This review recapitulates and discusses the most recent findings on the biological
function of WEE1/PKMYT1 during the cell cycle and in the DNA damage repair, with a focus on their dual role as
tumor suppressors in nonmalignant cells and pseudo-oncogenes in cancer cells. We here report the available data
on the molecular and functional alterations of WEE1/PKMYT1 kinases in both hematological and solid tumors.
Moreover, we summarize the preclinical information on 36 chemo/radiotherapy agents, and in particular their effect
on cell cycle checkpoints and on the cellular WEE1/PKMYT1-dependent response. Finally, this review outlines the
most important pre-clinical and clinical data available on the efficacy of WEE1/PKMYT1 inhibitors in monotherapy
and in combination with chemo/radiotherapy agents or with other selective inhibitors currently used or under
evaluation for the treatment of cancer patients.
Keywords: WEE1 family kinases, WEE1, PKMYT1, Cell cycle, DNA repair, Pseudo-oncogene, Tumor suppressor
Background
The WEE1 kinase family consists of three serine/threo-
nine kinases sharing conserved molecular structures
and encoded by the following genes: WEE1 (WEE1 G2
Checkpoint Kinase), PKMYT1 (membrane-associated
tyrosine- and threonine-specific cdc2-inhibitory kinase),
and WEE1B (WEE2 oocyte meiosis inhibiting kinase). In
eukaryotic somatic cells, WEE1 and PKMYT1 play a key
role in cell cycle regulation and, in particular, they are
involved in the entry into mitosis [1]. Their role as regula-
tors is crucial during normal cell cycle progression and in
response to DNA damages, as part of the DNA damage
response (DDR) pathways. Similarly, WEE2 regulates cell
cycle progression and, in particular, meiosis [2]. Briefly,
WEE2 plays a dual regulatory role in oocyte meiosis by
preventing premature restart prior to ovulation and per-
mitting metaphase II exit at fertilization [3]. Despite the
identification of WEE2 somatic mutations (1.9% of cases)
and copy number (CN) alterations (22.5% of patients with
CN loss and 22.5% with CN gain) across several cancer
types (https://portal.gdc.cancer.gov), they have not been
functionally linked to tumor development so far. There-
fore, the following sections will be focused on WEE1 and
PKMYT1 kinases that have a well-recognized role in
oncology and hemato-oncology.
WEE1 and PKMYT1 in cell cycle regulation
WEE1 and PKMYT1 act as tumor suppressors in non-
malignant eukaryotic somatic cells. Similarly to other
DDR-related kinases, their main biological function is to
© The Author(s). 2020 Open Access This article is licensed under a Creative Commons Attribution 4.0 International License,
which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give
appropriate credit to the original author(s) and the source, provide 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 Creative Commons
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/.
The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the
data made available in this article, unless otherwise stated in a credit line to the data.
* Correspondence: giorgia.simonetti@irst.emr.it
Biosciences Laboratory (Onco-hematology Unit), Istituto Scientifico
Romagnolo per lo Studio e la Cura dei Tumori (IRST) IRCCS, Via P. Maroncelli
40, 47014 Meldola, FC, Italy
Ghelli Luserna di Rorà et al. Journal of Hematology & Oncology (2020) 13:126
https://doi.org/10.1186/s13045-020-00959-2
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
prevent replication of cells with altered DNA. The main
downstream target of WEE1 family kinases is the
cyclin-dependent kinase 1 (CDK1)-cyclin B1 complex,
also known as mitotic-promoting factor (MPF). WEE1
phosphorylates CDK1 on Tyr15 while PKMYT1 has a
dual activity on Tyr15 and Thr14 [4](Fig.1a). The
phosphorylation of those residues keeps the MPF
complex inhibited until the cell approaches mitosis.
WEE1 is located in the nucleus, while PKMYT1 is
associated with the endoplasmic reticulum and Golgi
apparatus [5,6], and regulates Golgi membrane
reassembly following mitosis [7]. Together, WEE1 and
PKMYT1 ensure that CDK1 remains inactive as it
shuttles into and out of the nucleus [8]. Through its
extra-nuclear localization, PKMYT1 can also promote
CDK1 cytosolic segregation. At the G2/M border, if no
DNA damage has been detected, CDK1 phosphorylation
on Tyr15 and Thr14 is rapidly removed by CDC25C phos-
phatase. In the nucleus, the CDK-activating kinase (CAK)
complex composed by cyclin-dependent kinase 7 (CDK7),
cyclin H1, and MAT1 promotes MPF complex activation
through the phosphorylation of CDK1(Thr161) [9,10].
The active MPF complex is then imported into the
nucleus through phosphorylation of cyclin B1 (Ser126,
Ser128, Ser133, and Ser147) [11]. This event is required to
enter mitosis. The relevance of WEE1 and PKMYT1 regu-
lation of CDK1 has been recently confirmed by in vivo
studies. Indeed, the replacement of the CDK1 inhibitory
phosphorylation sites with non-phosphorylatable amino
acids (CDK1
T14A/Y15F
) was embryonic lethal in mice [12].
Once activated, the MPF complex can phosphorylate
WEE1 and PKMYT1 to promote their inactivation via dif-
ferent cascades [5,13,14]. WEE1 is phosphorylated
(Ser123) by CDK1 at the onset of mitosis, thereby generat-
ing a binding motif for polo like kinase 1 (PLK1) and
casein kinase 2 (CK2), that in turn phosphorylate WEE1
(Ser53 and Ser121, respectively) [14,15]. Together, the
phosphorylation of the three Ser residues serves as a tag
for the degradation of WEE1 by the ubiquitin ligase SCFβ-
TrCP [13]. PKMYT1 is also phosphorylated by CDK1 and
PLK1 and this event promotes its degradation [16]. In
addition to the checkpoint function at the G2/M border,
recent findings highlighted a role of WEE1 in the regula-
tion of replication dynamics during S phase (intra S phase
checkpoint). When cells reach the S phase, replication is
initiated from a large number of replication origins trig-
gered through the activation of the pre-replication com-
plex [17] and following the activation of S phase specific
CDK, primarily CDK2 [18,19]. Similarly to CDK1, CDK2
regulation is controlled through Tyr15 phosphorylation
status, that is balanced by WEE1 (Fig. 1a) and cell division
cycle 25A (CDC25A) activity [20]. Both WEE1 and
CDC25A/C have been shown to modulate unperturbed
replication through regulating CDK1/CDK2 activity.
Monoallelic expression of CDK1
T14A/Y15F
induced replica-
tion stress and S phase arrest in mouse embryonic fibro-
blasts (MEFs), with substantial increase of γH2AX levels,
chromosomal fragmentation, and DDR activation, as a
consequence of intra-S phase DNA damage [12]. More-
over, unscheduled origin firing due to loss of WEE1 leads
to exhaustion of the replication protein A1 (RPA1) pool
and, as a consequence, to death during DNA replication
Fig. 1 WEE1 and PKMYT1 biological functions. aSchematic representation of WEE1 and PKMYT1 involvement in cell cycle checkpoints. WEE1
regulates the activity of both CDK1 and CDK2 kinases (trough phosphorylation of Tyr15) and is involved in the regulation of intra-S, G2/M, and M
phase cell cycle checkpoints. PKMYT1 selectively regulates CDK1 (through phosphorylation of Tyr15 and Thr14) and is plays a role in the G2/M
phase checkpoint. bSchematic representation of the regulation of MUS81-EME1/2 endonuclease complexes by WEE1 during S and G2/M cell
cycle phases. By inhibiting CDK2 or CDK1, WEE1 prevents MUS81 activation and the generation of DNA damages during S phase, and
chromosomes pulverization during G2/M phase
Ghelli Luserna di Rorà et al. Journal of Hematology & Oncology (2020) 13:126 Page 2 of 17
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
(replication catastrophe). The intra S phase activity of
WEE1 is independent from PKMYT1 that is unable to
phosphorylate CDK2 [5]. In addition, WEE1, but not
PKMYT1, contributes to the control of mitosis exit.
Indeed, Wee1-deficient MEFs showed mitotic defects (e.g.,
in the number and position of centrosomes) that induce
arrest in mitosis or, in the majority of cells, mitotic slip-
page [21,22]. At the end of mitosis, WEE1 inhibits CDK1
through phosphorylation of its Tyr15 residue (Fig. 1a).
This event is dependent on the activation of the CTD
phosphatase subunit 1 (FCP1) that dephosphorylates and
activates WEE1 and other crucial component of the spin-
dle assembly checkpoint (SAC) complex [23]. Although
the precise mechanisms that regulate FCP1 activity is still
unknown, it has been showed that FCP1 promotes the
dephosphorylation of crucial SAC components, including
cell division cycle 20 (CDC20) and ubiquitin specific
peptidase 44 (USP44), thus promoting APC/C
Cdc20
activa-
tion and chromosome segregation [2426]. Moreover,
WEE1 directly interacts with APC/C components, includ-
ing fizzy and cell division cycle 20 related 1 (CDH1),
CDC20, cell division cycle 27 (CDC27), and its deletion
enforced APC/C activity, resulting in alterations of the
level of APC/C substrates and mitosis progression at the
expense of genomic stability [21].
WEE1 regulates replication forks and genome
stability
The activity of WEE1 through the cell cycle can explain
its tumor suppressor function, at least in nonmalignant
cells. This observation was confirmed and disentangled
in preclinical studies. Indeed, conditional Wee1 hetero-
zygous deletion in the murine mammary epithelium
caused enhanced proliferation, with cells progressing
into mitosis while still undergoing DNA replication, and
consequent accumulation of DNA damage, resulting in
genomic instability and, ultimately, in tumor develop-
ment [21]. Biological processes such as DNA replication
and homologous recombination involve the formation of
branched DNA structures that physically link chromo-
somes. Such DNA structures needs to be disengaged
prior to entry into mitosis, in order to ensure proper
chromosome segregation. Eukaryotic cells evolved differ-
ent mechanisms to identify and process branched DNA
structures (Y-shape DNA) and the most important one
involves the structure-selective endonuclease MUS81.
MUS81 forms heterodimeric complexes with the non-
catalytic subunits EME1 or EME2 and recognizes Y-
shape DNA structures during DNA replication or during
mitosis (homologous recombination). The activity of
MUS81-EME1/2 complex is crucial to recover stalled
replication forks, during prolonged S phase arrest, and
to reset DNA junction between twin chromatids during
homologous recombination [27]. In unperturbed cells,
WEE1 protects replication forks and prevents the gener-
ation DNA damages and chromosome pulverization
through an indirect inhibition of MUS81 functionality
[28]. Indeed, WEE1 phosphorylates CDK1 and CDK2,
thus preventing the CDK-mediated phosphorylation and
activation of MUS81-EME1/2 complexes [29]. Lack of
WEE1-dependent regulation of MUS81-EME1/2 endonu-
cleases may lead to cleavage of unwanted DNA structure
(excessive replication forks), which would slow down
replication progression and increase genomic instability
[27,28](Fig.1b).
WEE1 and PKMYT1 deregulation in cancer cells
WEE1 and PKMYT1 act like oncogenes
The biological role of WEE1 and PKMYT1 in cancer
cells is not fully understood. Reduced WEE1 expression
has been detected in breast cancer compared with
normal tissues, independently of the tumor grade [21].
However, most findings suggest that both kinases act
like oncogenes rather than tumor suppressors. Indeed,
they are frequently overexpressed in both solid and
hematological tumors and a genome-wide CRISPR
screen of 563 cancer cell lines, showed that they are
essential for the cell viability of almost all cell lines [30].
The dependency of cancer cells on WEE1 family
proteins may be linked to the following mechanisms
(Fig. 2): (i) the high proliferation rate of cancer cells that
follows the activation of driver oncogenes (e.g. RAS,
MYC) needs to be sustained by a strong cell cycle regu-
lation machinery; (ii) cancer cells frequently inactivate
p53, which is a key gatekeeper of G0/G1 and S phases
and, as a consequence, the regulation of cell cycle is
sustained entirely by the G2/M checkpoint; (iii) the
over-expression of DDR-related kinases is fundamental
to maintain a tolerable level of genetic instability, an
intrinsic feature of cancer cells [31,32]. Therefore, we
can speculate that, once the malignant transformation
process has been induced, WEE1 upregulation exerts a
pro-tumorigenic functions by securing a tolerable level
of genomic instability to cancer cells. The following
sections summarize the current knowledge on the mo-
lecular and functional alterations of WEE1 and PKMY
T1 in hematological and solid tumors.
WEE1 and PKMYT1 genetic lesions in cancer
WEE1 and PKMYT1 are rarely mutated in cancer
patients, with an overall mutation frequency of 1.2% and
0.2%, respectively (https://portal.gdc.cancer.gov). The
distribution of somatic mutations is highly heteroge-
neous across cancer types (WEE1: 0.27.6%; PKMYT1:
0.13.6%), with a higher frequency in uterine corpus
endometrial carcinoma (UCEC) and tumors of the
gastrointestinal tract (stomach and colon adenocarcin-
oma, Fig. 3a, b). In particular, WEE1 mutations have
Ghelli Luserna di Rorà et al. Journal of Hematology & Oncology (2020) 13:126 Page 3 of 17
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
been reported in 7.6% of UCEC cases. Moreover, PKMY
T1 lesions have been also detected in 2.7% of diffuse
large B cell lymphoma (DLBC). Conversely, both kinase
genes are rarely mutated in brain lower grade glioma
(LGG), ovarian serous cystadenocarcinoma (OV), pros-
tate adenocarcinoma (PRAD), and sarcoma (SARC), with
a frequency lower than 0.5%. Both genes are mainly tar-
geted by missense mutations that preferentially cluster
in the region encoding the WEE1 kinase domain and its
surroundings (Fig. 3c), suggesting a potential gain of
function effect of the kinase activity. Conversely, the
mutations are scattered throughout the PKMYT1 se-
quence (Fig. 3d). Little is known about the functional
consequences of WEE1 and PKMYT1 mutations. In the
majority of cancer types, the transcript expression in the
mutated cases is higher than the median value of the
entire cohort (https://www.cbioportal.org), supporting
once more an oncogenic function. In pancreatic adeno-
carcinoma (PA) patients and cell lines, an insertion was
identified in the WEE1 poly-T track, which contains the
binding site of the HuR RNA binding protein [33]. The
insertion resulted in decreased WEE1 expression upon
mitomycin-induced DNA damage, which would argue
against a protective effect of the mutation. Copy number
alterations (CNAs) represent a more frequent event
compared with mutations, with the WEE1 gene being
predominantly involved in CN loss (23.7% of cases
versus 7.8% of patients with CN gains), while PKMYT1
showing a higher percentage of CN gain (15.9% versus
12.0% of CN loss, Fig. 3e, f). The predominance of
WEE1 deletion events (6.3% versus 3.25% of cases with
amplification) was also observed in breast cancer, in line
with its reduced expression, as mentioned above [21].
Overall, cancer types showing the highest recurrence (>
10%) of CNAs were OV (27.7%), lung squamous cell
carcinoma (LUSC, 14.8%), uterine carcinosarcoma (UCS,
12.5%), and SARC (11.2%) for WEE1 and OV (18.8%),
bladder urothelial carcinoma (BLCA, 13.7%), and
esophageal carcinoma (ESCA, 10.3%) for PKMYT1.Of
note, OV and LUSC have been classified as tumors with
multiple recurrent chromosomal gains and losses [34],
which may suggest a bystander effect related to chromo-
somal instability in these tumor types, especially in the
case of WEE1 deletion, that is unexpected, based on the
general oncogenic function exerted by the kinase.
WEE1 and PKMYT1 functional role in hematological and
solid tumors
Few studies have analyzed WEE1 and PKMYT1 expres-
sion in hematological malignancies. Our group showed
that WEE1 kinase is highly expressed in acute lympho-
blastic leukemia (ALL) cell lines and primary cells in
comparison with normal hematopoietic cells, and that
PKMYT1 is upregulated in relapsed ALL samples
compared with nonmalignant hematopoietic cells [35].
Moreover, we demonstrated that ALL cells are
dependent on WEE1 functionality for their survival and
proliferation and that PKMYT1 levels may influence the
sensitivity to the WEE1 inhibitor AZD-1775 [35]. Similar
results on the role of WEE1 were obtained in multiple
myeloma (MM), acute myeloid leukemia (AML), chronic
myeloid leukemia (CML), and chronic lymphocyte
leukemia (CLL) [3639]. In AML cells, WEE1 and
PKMYT1 are key gene discriminating between FLT3-
ITD, FLT3-TKD, and NRAS-mutated samples. They
were expressed at lower levels selectively in FLT3-ITD
specimens in comparison with wild-type cells, suggesting
either a tumor suppressor role in the leukemogenic
process or a potential vulnerability n this AML subtype
[40]. Pharmacological WEE1 inhibition alone or in com-
bination with histone deacetylase inhibitors showed
therapeutic potential in FLT3-ITD AML, confirming
their dependency on WEE1 activity [41]. Since FLT3-
ITD AML have intrinsic homologous recombination
repair defects [42]. WEE1 inhibition may exacerbate the
cell genotoxic stress by disrupting multiple cell cycle
checkpoints. WEE1 has been showed to be a valuable
target also for lymphoma patients [43]. In parallel,
Fig. 2 WEE1 family proteins role as tumor suppressors or pseudo-oncogenes in non-malignant and cancer cells
Ghelli Luserna di Rorà et al. Journal of Hematology & Oncology (2020) 13:126 Page 4 of 17
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
PKMYT1 proved to be essential for MM cell line viabil-
ity, since its downregulation strongly decreased cell
growth, while inducing apoptosis [44].
WEE1 and PKMYT1 are also over-expressed in solid
tumors, including hepatocellular carcinoma, colon cancer,
glioblastoma, non-small-cell lung cancer (NSCLS), neuro-
blastoma, and gastric cancers [31,32,4547]. High WEE1
expression has been associated with negative prognostic
factors including lymph node involvement, induction of
metastasis, increased biomarkers of proliferation (CCND1,
Ki67, or CCNA1) and resistance to treatments (radiother-
apy or chemotherapy) [4851]. Elevated PKMYT1 levels
have been associated with tumor progression, a more
aggressive disease, the induction of metastasis at least in
NSCLS patients [45] and, generally, with poor prognosis.
Depending on the cancer subtype, the expression of
WEE1 and PKMYT1 has been linked with the activation
of cellular pathways crucial for the specific disease. In
melanoma cells, WEE1 silencing caused an increase of
phospho p38 protein levels, indicating a role in the
regulation of p38/MAPK pathway activation during p53-
independent DNA damage response [49]. In hepatocellu-
lar carcinoma and colorectal cancers, PKMYT1 regulates
epithelial-mesenchymal transition (EMT), a process rele-
vant to tumor progression, invasion, metastasis, and drug
resistance, through the activation of the beta-catenin/TCF
Fig. 3 WEE1 and PKMYT1 mutations and copy number alterations (CNAs) in cancer. aFrequency of patients with WEE1 or bPKMYT1 gene
mutations across cancers from TCGA cohorts. cDistribution of mutations according to the WEE1 and dPKMYT1 amino acid (aa) sequence and
protein domains (WEE1 transcript ENST00000450114, 646 aa; PKMYT1 transcript ENST00000262300, 499 aa). eFrequency of patients with copy
number gain or loss in WEE1 or fPKMYT1 across cancers (https://portal.gdc.cancer.gov;ACC adrenocortical carcinoma, BLCA bladder urothelial
carcinoma, BRCA breast invasive carcinoma, CESC cervical squamous cell carcinoma and endocervical adenocarcinoma, COAD colon
adenocarcinoma, CHOL cholangiocarcinoma, DLBC diffuse large B cell lymphoma, ESCA esophageal carcinoma, GBM glioblastoma multiforme,
HNSC head and neck squamous cell carcinoma, KICH kidney chromophobe, KIRK kidney renal clear cell carcinoma, KIRP kidney renal papillary cell
carcinoma, LGG brain lower grade glioma, LIHC liver hepatocellular carcinoma, LUAD lung adenocarcinoma, LUSC lung squamous cell carcinoma,
MESO mesothelioma, OV ovarian serous cystadenocarcinoma, PAAD pancreatic adenocarcinoma, PCPG pheochromocytoma and paraganglioma,
PRAD prostate adenocarcinoma, READ rectum adenocarcinoma, SARC sarcoma, SKCM skin cutaneous melanoma, STAD stomach adenocarcinoma,
TGCT testicular germ cell tumors, THYM thymoma, UCS uterine carcinosarcoma, UCEC uterine corpus endometrial carcinoma)
Ghelli Luserna di Rorà et al. Journal of Hematology & Oncology (2020) 13:126 Page 5 of 17
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
signaling [32,46], while PKMYT1 has been reported to
control Notch pathway in NSCLC [45]. In particular,
crucial component of the pathway, including NOTCH1,
p21, and HES1 are downregulated by chemical inhibition
of PKMYT1 [45]. In neuroblastic tumors, PKMYT1 is
required to stabilize MYCN protein, which is a crucial
proto-oncogene for this cancer types [52]. Moreover, in
esophageal squamous cell carcinoma (ESCC) cell lines and
primary cells, the expression of PKMYT1 is associated
with and regulates the activation of the AKY/mTOR path-
way [53](Table1). Taken together, this evidence suggests
a broad role of WEE1/PKMYT1 besides the DNA damage
response pathway that may increase the interest towards
its therapeutic targeting.
Development of WEE1 and PKMYT1 inhibitors
WEE1 and PKMYT1 inhibitors have single agent and
chemo-sensitizer effects
Due to their potential oncogenic role, WEE1 and PKMY
T1 have been investigated as therapeutic targets for
hematological and solid tumors. Several pharmacological
inhibitors have been designed and subsequently vali-
dated in different cancer models. The available literature
highlights a common mechanism of action of WEE1/
PKMYT1 inhibitors in cancer cells either in single agent
or in combination with DNA damaging agents (chemo-
therapy/radiotherapy). WEE1/PKMYT1 kinase inhibition
causes G2/M cell cycle checkpoint override, premature
mitotic entry, and cell death during mitosis, through a
mechanism generally known as mitotic catastrophe
(Fig. 4a). From a biological point of view, the inhibition
of WEE1 kinase causes a significant reduction of phospho-
CDK1 (Tyr15), thus promoting the accumulation of active
CDK1-cyclin B1 complex and, consequently, mitotic entry.
The beginning of mitosis is also associated with a progres-
sive accumulation of DNA damages and the degeneration
in mitotic catastrophe. The sensitivity to WEE1 kinase
inhibitors in relation to TP53 mutational status remains
controversial. Indeed, some studies reported increased sen-
sitivity of TP53 mutant cell lines to WEE1 inhibitors in
comparison to TP53 wild-type ones [62,63], while others
showed no association between p53 functionality and the
effectiveness of WEE1 inhibition [35,64]. These discrepan-
cies may be linked to the intrinsic chromosomal instability
of the analyzed tumors and to additional alterations deregu-
lating the G1 checkpoint in TP53 wild-type cases that may
enhance the sensitivity to WEE1 targeting.
Regarding the role of WEE1 inhibitors as chemo-
sensitizer agents, a large number of studies demonstrated
a synergistic activity between DNA damaging agents
(chemotherapy including doxorubicin, cytarabine, metho-
trexate, cisplatin, clofarabine, etoposide, 5-fluorouracil,
and radiotherapy) and different WEE1/PKMYT1 inhibi-
tors in preclinical models [48,56,6569]. The mechanism
of action of the combination is based on the inhibition of
the DDR pathway following induction of DNA damage in-
duced by the chemotherapy or radiotherapy agents. In this
scenario, cancer cells with damaged DNA fail to arrest cell
cycle, continue to proliferate, and accumulate massive
DNA damage until a point of no return (Fig. 4b). Indeed,
several DNA damaging agent promote the indirect activa-
tion of WEE1 and PKMYT1 kinases, as showed mostly by
the activation of cell cycle checkpoints (S and G2/M
checkpoints) in cancer cells. We summarized in Table 2
the results of preclinical studies in which the effect of
different chemotherapy agents or radiotherapy has been
evaluated in terms of cell cycle perturbation and altered
expression of WEE1 or PKMYT1 following in vitro or
Table 1 WEE1 and PKMYT1 molecular alterations in hematological and solid tumors according to literature
Gene Genetic alteration Disease Effect/prognostic value Reference
Hematological tumors
WEE1 Over-expression ALL; AML; MM; CML;
CLL; DLBCL
Crucial for cell viability of cancer cells (experimentally proven). [3540,43,54]
Copy number Gain AML Biological effect or prognostic value unknown [55]
PKMYT1 Over-expression ALL; MM Crucial for cell viability of cancer cells (experimentally proven). [35,44]
Solid tumors
WEE1 Over-expression GC; MaM; GL; OC; CC Associated with lymph node involvement, induction of
metastasis, increased biomarkers of proliferation (CCND1, Ki67
or CCNA1), resistance to treatment and poor overall survival.
[4851,5660]
Mutation PA Insertion causing decrease WEE1 expression upon DNA
damage
[33]
PKMYT1 Over-expression HC; CC; GLB; NSCLC; N; GS Associated with tumor progression, aggressive disease and
poor overall survival.
[31,32,4547]
Mutation N Biological effect or prognostic value unknown [61]
ALL acute lymphoblastic leukemia, AML acute myeloid leukemia, MM multiple myeloma, CML chronic myeloid leukemia, CLL chronic lymphocyte leukemia, DLBCL
diffuse large B cell lymphoma, GC gastric cancer, MaM malignant melanoma, GL gliomas, OC ovarian cancer, CC colorectal cancer, PA pancreatic adenocarcinoma,
HC hepatocellular carcinoma, GLB glioblastoma, NSCLC non-small-cell lung cancer, Nneuroblastoma
Ghelli Luserna di Rorà et al. Journal of Hematology & Oncology (2020) 13:126 Page 6 of 17
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
in vivo treatment. Taken together, the abovementioned
data prove that WEE1 and PKYMYT1 are ideal targets to
override cell cycle checkpoint regulation and to improve
the efficacy of DNA-damaging agents. In particular,
tumors with a high level of chromosomal instability may
respond to WEE1/PKMYT1 inhibition per se, while cases
with a more stable genomic asset may benefit of the
combination between DNA-damaging agents and WEE1
family kinase inhibitors. The following sections reports
the main preclinical and clinical findings obtained using
small molecules inhibitors of WEE1 and PKMYT1 kinases.
Preclinical studies of WEE1 and PKMYT1 inhibitors
Several targeted compounds showed an inhibitory activ-
ity on WEE1 and PKMYT1 kinases and their efficacy
was proven in a number of tumor types. Table 3shows
the main preclinical studies that used WEE1/PKMYT1
inhibitors in single agent or in combination with chemo/
radiotherapy agents in different tumor types.
PD0166285 is the first reported drug, with an inhibi-
tory activity against WEE1, PKMYT1, and a range of
other kinases including c-Src, EGFR, FGFR1, CHK1, and
PDGFRb [151].
Adavosertib (AZD-1775) is the first highly potent and
selective WEE1 inhibitor. A large number of preclinical
studies evaluated its efficacy in single agent and in
combinatory approaches. Regarding the mechanism of
action, adavosertib induces S and/or G2/M cell cycle
checkpoints override, depending on cancer types, when
used in monotherapy. Cell cycle perturbation is associ-
ated with a progressive accumulation of DNA damages
and by the induction of apoptosis [35,99,119122].
This last event is cell cycle phase-dependent and can
occur (i) as a consequence of S phase checkpoint
Fig. 4 Mechanism of action of WEE1/PKMYT1 inhibitors for the treatment of cancer cells. aSchematic representation of WEE1/PKMYT1 inhibition
as monotherapy. In cancer cells, oncogenes promote high rate of proliferation, replication stress and the over-expression of WEE1/PKMYT1
kinases. In this scenario, cancer cells need WEE1 and PKMYT1 to sustain replication stress and proliferation. The inhibition of WEE1/PKMYT1 results
in the accumulation of DNA damages, the increase of genetic instability and induction of apoptosis. bSchematic representation of WEE1/PKMYT1
inhibition in combination with DNA damaging agents. Cancer cells respond to DNA damages by activating WEE1/PKMYT1 kinases. The inhibition
of WEE1/PKMYT1 enhances the cytotoxicity of DNA damaging agents by inhibiting DNA repair and promoting cell cycle progression even in the
presence of DNA damages. Therefore, cancer cells accumulate massive DNA damages until a point of no return
Ghelli Luserna di Rorà et al. Journal of Hematology & Oncology (2020) 13:126 Page 7 of 17
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
override, when cancer cells start DNA replication even
in the presence of DNA damages (replicative catastro-
phe); (ii) following G2/M phase checkpoint override,
that results in forced entry into mitosis, even in the pres-
ence of DNA damages (mitotic catastrophe).
In combination strategies, adavosertib was able to
enhance the cytotoxicity of chemo/radiotherapy agents,
by inducing cell cycle checkpoint override, inhibition of
DNA damage repair, and induction of apoptosis [35,37,
38,92,121,127129]. The chemo-sensitizer efficacy of
DDR inhibitors has been linked to drug scheduling [94,
152,153]. Recently in pancreatic adenocarcinoma cells,
it has been reported that the efficacy of a triple regimen
combining gemcitabine, CHK1, and WEE1 inhibitors is
strictly dependent on the timing of drug administration.
Indeed, the maximum effect of the combination is
Table 2 Effects of standard of care chemo/radiotherapy agents on cell cycle checkpoints activation
Chemotherapy agents/radiotherapy Intra S checkpoint G2/M checkpoint WEE1 and/or PKMYT1 experimentally
proven involvement in cancer model
Actinomycin No Yes [70,71] WEE1 upregulation [71]
Azacitidine No Yes [72]NA
Bleomycin No Yes [73]NA
Carboplatin No Yes [74]NA
Cisplatin No Yes [75,76] WEE1 inhibition enhanced cytotoxicity [76,77]
Cyclophosphamide NA NA WEE1 upregulation [78]
Cytarabine Yes [79,80] Yes [79,80] WEE1 upregulation [38,81]
Clofarabine Yes [35] No WEE1 inhibition enhanced cytotoxicity [35]
Daunorubicin Yes [82] Yes [82]NA
Decitabine No Yes [83]NA
Docetaxel No Yes [84,85]NA
Doxorubicin No Yes [86] WEE1 upregulation [86]; WEE1 inhibition
enhanced cytotoxicity [35]
Epirubicin No Yes [87,88] WEE1 inhibition enhanced cytotoxicity [89]
Epothilone No Yes [90]NA
Etoposide No Yes [91] WEE1 inhibition enhanced cytotoxicity [67]
Fluorouracil Yes [92] No WEE1 inhibition enhanced cytotoxicity [92]
Fludarabine Yes [80]No NA
Gemcitabine Yes [80] No WEE1 upregulation [93]; WEE1 inhibition
enhanced cytotoxicity [94]
Hydroxyurea Yes [95] No WEE1 inhibition enhanced cytotoxicity [96]
Idarubicin No Yes [97]NA
Irinotecan No Yes [98] WEE1 inhibition enhanced cytotoxicity [99]
Mechlorethamine Yes [100] Yes [100]NA
Mercaptopurine Yes [101]No NA
Methotrexate Yes [102] No WEE1 inhibition enhanced cytotoxicity [103]
Mitoxantrone No Yes [104] WEE1 inhibition enhanced cytotoxicity [78]
Oxaliplatin No Yes [105] WEE1 inhibition enhanced cytotoxicity [106]
Paclitaxel No Yes [107] WEE1 inhibition enhanced cytotoxicity [108]
Pemetrexed Yes [109] No WEE1 inhibition enhanced cytotoxicity [110]
Radiotherapy (ionizing radiation) No Yes [111] WEE1 inhibition enhanced cytotoxicity [68]
Teniposide Yes [112] Yes [112]NA
Thioguanine Yes [113] Yes [113] WEE1 inhibition enhanced cytotoxicity [65]
Topotecan No Yes [114] WEE1 inhibition enhanced cytotoxicity [115]
Vinblastine No Yes [116]NA
Vincristine No Yes [117] WEE1 inhibition enhanced cytotoxicity [118]
Ghelli Luserna di Rorà et al. Journal of Hematology & Oncology (2020) 13:126 Page 8 of 17
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
obtained when gemcitabine and CHK1 inhibitors are
administered simultaneously (thus inducing replicative
stress) and adavosertib is added at a later time [94].
Moreover, strong synergism has been observed by
combining adavosertib with small molecules, including
DDR-related inhibitors (CDK2 [89], CDK4-6 [149],
Table 3 Preclinical studies evaluating the effect of WEE1 inhibitors in monotherapy or in combination with chemotherapy/
radiotherapy in cancer
Inhibitor Treatment Cancer model Main biological effect References
PD0166285 M GBM-astrocytoma -G2/M checkpoint override
-Forced mitotic entry
[51]
Adavosertib M MM, ALL, AML TNBC, DLBCL, MCL -G2/M checkpoint override
-Forced mitotic entry
-Mitotic catastrophe
-Replicative catastrophe
[35,99,119122]
PD0166285 +R GBM-astrocytoma -Mitotic catastrophe
-Inhibition of DNA repair
[51]
Adavosertib +R CC, LC, BC, PC, OC, DLBCL, ES -Increased DNA damage
-Induction of apoptosis
-Mitotic catastrophe
[78,99,123126]
Adavosertib +C AML, ALL, MM, BC, CC, GC, DLBCL -S or G2/M checkpoint override
-Increased DNA damaged
-Induction of apoptosis
[35,37,38,76,92,99,
121,127129]
Adavosertib +HDAC i AML, HNSCC -Replication stress
-Replicative catastrophe
-Increased DNA damage
-Inhibition of DNA repair
[41,130,131]
Adavosertib +ATR i AML, DLBCL, MCL, BC -Replication stress
-Replicative catastrophe
-Increased DNA damaged
-Inhibition of DNA repair
[132135]
Adavosertib +mTOR i AML, ALL, OC, NSCLC -Inhibition of DNA repair [136139]
Adavosertib +CHK1 i MCL, DLBCL, ALL, AML -Replication stress
-Increased DNA damage
-Replicative catastrophe
[103,140142]
Adavosertib +BCL2i/MCL-1 i DLBCL -Force mitotic entry
-Increase DNA damage
-INDUCTION of apoptosis
[143]
Adavosertib +PARP1 i NSCLC, AML, ALL -G2/M checkpoint override
-Replication stress
-Increased DNA damage
-Inhibition of DNA repair
[126,144146]
Adavosertib +AURORA A i HNSCC -Forced mitotic entry
-Mitotic catastrophe
[147]
Adavosertib +CDK2 i BC -Replication stress
-Replicative catastrophe
[89]
Adavosertib +SIRT1 i LC -Inhibition of DNA repair [148]
Adavosertib +CDK4-6 i S -Replication stress [149]
Adavosertib +BCR-ABL1 i ALL -Inhibition of DNA repair -G2/M
checkpoint override
[35]
Adavosertib +Proteasome i MM -G2/M checkpoint override
-Forced mitotic entry
-Inhibition of DNA repair
[36]
Adavosertib +BET i NSCLC -Inhibition of DNA repair
-Forced mitotic entry
-Mitotic catastrophe
[150]
Mmonotherapy, Rradiotherapy, Cchemotherapy, ALL acute lymphoblastic leukemia, AML acute myeloid leukemia, MM multiple myeloma, DLBCL diffuse large B
cell lymphoma, MCL mantle cell lymphoma, GC gastric cancer, GL gliomas, OC ovarian cancer, CC colorectal cancer, PC pancreatic cancer, ES esophageal cancer, HC
hepatocellular carcinoma, GLB glioblastoma, NSCLC non-small-cell lung cancer, Nneuroblastoma, Ssarcomas, LC lung cancer, BC breast cancer, HNSCC head and
neck squamous cell carcinoma, TNBC triple negative breast cancer
Ghelli Luserna di Rorà et al. Journal of Hematology & Oncology (2020) 13:126 Page 9 of 17
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
CHK1 [103,140142], ATM [132135], AURORA A
[147], PARP1 [144], SIRT1 [148] inhibitors), histone
deacetylase (HDAC) inhibitors [41,130,131], tyrosine
kinase inhibitors (BCR-ABL1 inhibitors [35]), anti-
apoptotic protein inhibitors (BCL2 and MCL1 inhibitors
[143]), mTOR inhibitor [136139], and proteasome in-
hibitors [36].
We have recently reported synergistic effects of adavo-
sertib in combination with different tyrosine kinase
inhibitors in both BCR-ABL1-positive and -negative
ALL cell lines and primary cells. Interestingly, strong
synergism was found in BCR-ABL1-negative ALL cell
lines treated with adavosertib in combination with bosu-
tinib isomer. In the study, we speculated that the strong
cytotoxic effect of the combination was due to the con-
comitant inhibition of WEE1 and PKMYT1 kinases [35].
Indeed, no selective inhibitor has been currently devel-
oped to target its functionality. However, several known
tyrosine kinase inhibitors have an inhibitory off-target
effect on PKMYT1. Among them, compounds com-
monly used for the treatment of BCR-ABL1-positive
CML and ALL, as dasatinib and bosutinib (and a struc-
tural isomer of bosutinib [154,155]) were shown to
inhibit PKMYT1 activity.
Overall, the data suggest that WEE1/PKMYT1 inhib-
ition is a suitable pharmacological target for combin-
ation strategies in cancer. The broad spectrum of
activities exerted by the two kinases, and especially by
WEE1, across the cell cycle, makes them good candi-
dates for a number of diverse therapeutic combinations.
WEE1 inhibitors from bench to bedside
Several clinical studies are currently evaluating the effi-
cacy of adavosertib on different aggressive and advanced
tumors (Table 4).
The results of phase I trials showed that adavosertib is
well tolerated both in single agent and in combination.
Depending on the study, the maximum tolerated dose
(MTD) was established between 150 and 225 mg orally
twice per day for 2.5 days per 2 weeks [156158]. The
most common adverse events reported in the abovemen-
tioned studies were fatigue, nausea, vomiting, diarrhea,
and hematologic toxicity. Moreover, correlative studies
performed on tumor biopsies confirmed in vivo the mech-
anism of action of adavosertib. Indeed, immunohisto-
chemistry analyses showed a reduction of phospho-CDK1
(Tyr15) and an increase of DNA damages (phospho-
γH2AX) in cancer cells [156,157].
The phase II studies confirmed that adavosertib sensi-
tizes cancer patients to different chemotherapy agents.
Interestingly, adavosertib showed efficacy when com-
bined with carboplatin in TP53-mutated ovarian cancer
patients, refractory or resistant to first-line platinum-
based chemotherapy [159]. Similar results were reported
in platinum-resistant primary ovarian cancer patients
after treatment with the combination of adavosertib and
a single chemotherapeutic agent (carboplatin, paclitaxel,
gemcitabine, or pegylated liposomal doxorubicin) [160].
Primary resistance and predictive markers of
response to WEE1/PKMYT1 based therapies
Several DDR inhibitors have proved their efficacy against
different cancer types in the preclinical and clinical set-
tings [161165]. Among them, WEE1 inhibitor seems to
be the most effective ones, also favored by a relative low
off-target toxicity. However, despite the number of stud-
ies and the promising results, few predictive markers of
response have been identified. Recently, cyclin E level
has been linked to the efficacy of adavosertib in breast
cancer models [89], with cyclin E-high cells, that gener-
ally show elevated chromosome instability, being more
sensitive compared with cyclin E-low ones. Despite the
reported low levels of WEE1 expression in breast cancer,
chromosome instability, that has also prognostic poten-
tial mainly in grade 2 tumors [89], may explain the
effectiveness of WEE1 inhibitors, as supported by the
predictive role of cyclin E. Our group and others showed
that high PKMYT1 expression associates with reduced
sensitivity to adavosertib, indicating a potential compen-
satory effect [35,166]. Moreover, high-throughput
proteomic profiling demonstrated that small cell lung
cancer and ovarian cancer models with primary resist-
ance to adavosertib express high levels of AKT/mTOR
pathway molecules and phosphorylated S6 ribosomal
protein [137,138]. In acute leukemia models, the sensi-
tivity to adavosertib has been recently linked to HDAC
and MYC regulation. Indeed, by generating adavosertib-
resistant models, the researchers found that resistant
acute leukemia cell lines are dependent on increased
HDAC activity for their survival, partly due to increased
KDM5A function. In addition, gene expression analyses
demonstrated a HDAC-dependent expression of MYC in
the adavosertib-resistant cell lines [167]. These observa-
tions support the success of preclinical studies combining
WEE1 and HDAC [41,130,131] or bromodomain inhibi-
tors [150].
Conclusion
Thanks to a constantly growing amount of preclinical and
clinical data, our knowledge on cancer biology is increasing
and, consequently, the list of cancer hallmarks has been
progressively expanding. Recent findings demonstrated that
cancer cells are characterized by functional and molecular
alterations in crucial genes involved in the DDR pathway,
which is fundamental for cell cycle regulation, DNA dam-
ages recognition, and repair. Functional alterations of DDR-
gene have a deep impact on tumor progression and on the
clinical outcome of cancer patients. Indeed, the efficacy of
Ghelli Luserna di Rorà et al. Journal of Hematology & Oncology (2020) 13:126 Page 10 of 17
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Table 4 Clinical trials evaluating WEE1/PKMYT1 inhibitor in monotherapy or in combination for cancer therapy
Study ID Study title Tumor Interventions Status Phase
NCT02610075 Phase Ib Study to Determine MTD of AZD1775 Monotherapy in
Patients With Locally Advanced or Metastatic Solid Tumours.
S AZD1775 C 1
NCT03668340 AZD1775 in Women With Recurrent or Persistent Uterine
Serous Carcinoma
S AZD1775 R 2
NCT02482311 Safety, Tolerance, PK, and Anti-tumour Activity of AZD1775
Monotherapy in Patients With Advanced Solid Tumours
S AZD 1775 C 1
NCT02207010 A Phase 0 Study of AZD1775 in Recurrent GBM Patients S AZD1775 NA 1
NCT03315091 Phase I Study to Assess the Effect of Food on AZD1775
Pharmacokinetics in Patients With Advanced Solid Tumours
S AZD1775 C 1
NCT01748825 AZD1775 for Advanced Solid Tumors S/H AZD1775 ANR 1
NCT02511795 AZD1775 Combined With Olaparib in Patients With Refractory
Solid Tumors
S AZD1775 + Olaparib C 1
NCT03313557 AZD1775 Continued Access Study to Assess Safety and
Tolerability for Patients Enrolled in AZD1775 Clinical
Pharmacology Studies
S AZD1775 C 1
NCT02593019 Phase II, Single-arm Study of AZD1775 Monotherapy in
Relapsed Small Cell Lung Cancer Patients
S AZD1775 NA 2
NCT02688907 Phase II, Single-arm Study of AZD1775 Monotherapy in Relapsed
Small Cell Lung Cancer Patients With MYC Family Amplification
or CDKN2A Mutation Combined With TP53 Mutation
S AZD1775 T 2
NCT02087176 A Placebo Controlled Study Comparing AZD1775 + Docetaxel
Versus Placebo + Docetaxel to Treat Lung Cancer
S AZD1775 + Docetaxel T 2
NCT03012477 CISPLATIN + AZD-1775 In Breast Cancer S AZD1775 + Cisplatin ANR 2
NCT02341456 Phase Ib Study AZD1775 in Combination With Carboplatin and
Paclitaxel in Adult Asian Patients With Solid Tumours
S AZD1775 + Carboplatin
or Paclitaxel
C1
NCT02791919 Wee1 Kinase Inhibitor AZD1775 and Combination Chemotherapy
in Treating Children, Adolescents and Young Adults With
Relapsed or Refractory Acute Myeloid Leukemia
H AZD1775 + Cytarabine
or Filgrastim
or Fludarabine Phosphate
W1
NCT02513563 AZD1775 Plus Carboplatin-Paclitaxel in Squamous Cell Lung Cancer S AZD1775 + Carboplatin
or Paclitaxel
R2
NCT03718143 AZD1775 in Advanced Acute Myeloid Leukemia, Myelodysplastic
Syndrome and Myelofibrosis
H AZD1775 + Cytarabine T 2
NCT02585973 Dose-escalating AZD1775 + Concurrent Radiation + Cisplatin for
Intermediate/High Risk HNSCC
S AZD1775 + Cisplatin +
Radiation
R1
NCT02087241 Ph II Trial of Carboplatin and Pemetrexed With or Without
AZD1775 for Untreated Lung Cancer
S AZD1775 + pemetrexed
or carboplatin
T2
NCT02381548 Phase I Trial of AZD1775 and Belinostat in Treating Patients With
Relapsed or Refractory Myeloid Malignancies or Untreated Acute
Myeloid Leukemia
H AZD1775 + Belinostat T 1
NCT03333824 Effects of AZD1775 on the PK Substrates for CYP3A, CYP2C19,
CYP1A2 and on QT Interval in Patients With Advanced Cancer
S AZD1775 C 1
NCT02906059 Study of Irinotecan and AZD1775, a Selective Wee 1 Inhibitor, in
RAS or BRAF Mutated, Second-line Metastatic Colorectal Cancer
S AZD1775 + Irinotecan R 1
NCT02037230 Dose Escalation Trial of AZD1775 and Gemcitabine (+Radiation)
for Unresectable Adenocarcinoma of the Pancreas
S AZD1775 + Gemcitabine+
Radiation Therapy
C 1,2
NCT02617277 Safety, Tolerability and Pharmacokinetics of AZD1775
(Adavosertib) Plus MEDI4736 (Durvalumab) in Patients With
Advanced Solid Tumours
S AZD1775 + Durvalumab ANR 1
NCT02666950 WEE1 Inhibitor AZD1775 With or Without Cytarabine in Treating
Patients With Advanced Acute Myeloid Leukemia or
Myelodysplastic Syndrome
H AZD1775 + Cytarabine C 2
NCT01047007 A Dose Escalation Study of MK1775 in Combination With 5-FU or
5-FU/CDDP in Patients With Advanced Solid Tumor (1775-005)
S AZD1775 + 5-FU or
5-FU/CDDP
T1
NCT01164995 Study With Wee-1 Inhibitor MK-1775 and Carboplatin to Treat S AZD1775 + carboplatin NA 2
Ghelli Luserna di Rorà et al. Journal of Hematology & Oncology (2020) 13:126 Page 11 of 17
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Table 4 Clinical trials evaluating WEE1/PKMYT1 inhibitor in monotherapy or in combination for cancer therapy (Continued)
Study ID Study title Tumor Interventions Status Phase
p53 Mutated Refractory and Resistant Ovarian Cancer
NCT02448329 Study of AZD1775 in Combination With Paclitaxel, in Advanced
Gastric Adenocarcinoma Patients Harboring TP53 Mutation as a
Second-line Chemotherapy
S AZD1775 + paclitaxel R 2
NCT02508246 WEE1 Inhibitor MK-1775, Docetaxel, and Cisplatin Before Surgery
in Treating Patients With Borderline Resectable Stage III-IVB
Squamous Cell Carcinoma of the Head and Neck
S AZD1775 + Cisplatin +
Docetaxel
C1
NCT03253679 AZD1775 in Treating Patients With Advanced Refractory Solid
Tumors With CCNE1 Amplification
S AZD1775 R 2
NCT01076400 A Study of MK-1775 in Combination With Topotecan/Cisplatin in
Participants With Cervical Cancer (MK-1775-008)
S AZD1775 + Topotecan or
Cisplatin
T 1,2
NCT02196168 Cisplatin With or Without WEE1 Inhibitor MK-1775 in Treating
Patients With Recurrent or Metastatic Head and Neck Cancer
S AZD1775 +Cisplatin T 2
NCT02101775 Gemcitabine Hydrochloride With or Without WEE1 Inhibitor MK-
1775 in Treating Patients With Recurrent Ovarian, Primary
Peritoneal, or Fallopian Tube Cancer
S AZD1775 + Gemcitabine ANR 2
NCT03028766 WEE1 Inhibitor With Cisplatin and Radiotherapy: A Trial in Head
and Neck Cancer
S AZD1775 + Cisplatin +
Radio therapy
ANR 1
NCT01357161 A Study of MK-1775 in Combination With Paclitaxel and
Carboplatin Versus Paclitaxel and Carboplatin Alone for
Participants With Platinum-Sensitive Ovarian Tumors With the
P53 Gene Mutation (MK-1775-004)
S AZD1775 + paclitaxel +
carboplation
C2
NCT03284385 Testing AZD1775 in Advanced Solid Tumors That Have a
Mutation Called SETD2
S AZD1775 R 2
NCT00648648 A Dose Escalation Study of MK-1775 in Combination With Either
Gemcitabine, Cisplatin, or Carboplatin in Adults With Advanced
Solid Tumors (MK-1775-001)
S AZD1775 + Gemcitabine
or Cisplatin or Carboplatin
C1
NCT02194829 Paclitaxel Albumin-Stabilized Nanoparticle Formulation and
Gemcitabine Hydrochloride With or Without WEE1 Inhibitor MK-
1775 in Treating Patients With Previously Untreated Pancreatic
Cancer That Is Metastatic or Cannot Be Removed by Surgery
S AZD-1775 + Gemcitabine
+ paclitaxel
ANR 1,2
NCT02576444 Olaparib Combinations S AZD1775 + olaparib ANR 2
NCT04197713 Testing the Sequential Combination of the Anti-cancer Drugs
Olaparib Followed by Adavosertib (AZD1775) in Patients With
Advanced Solid Tumors With Selected Mutations and PARP
Resistance, STAR Study
S AZD1775 + olaparib ANR 1
NCT01922076 Adavosertib and Local Radiation Therapy in Treating Children
With Newly Diagnosed Diffuse Intrinsic Pontine Gliomas
S AZD1775 + Radiation
Therapy
ANR 1
NCT03579316 Adavosertib With or Without Olaparib in Treating Patients With
Recurrent Ovarian, Primary Peritoneal, or Fallopian Tube Cancer
S AZD1775 + olaparib R 2
NCT02095132 Adavosertib and Irinotecan Hydrochloride in Treating Younger
Patients With Relapsed or Refractory Solid Tumors
S AZD1775 + Irinotecan or
Irinotecan Hydrochloride
R 1,2
NCT03345784 Adavosertib, External Beam Radiation Therapy, and Cisplatin in
Treating Patients With Cervical, Vaginal, or Uterine Cancer
S AZD1775 +Cisplatin +
Radiation (External Beam
Radiation Therapy)
R1
NCT01849146 Adavosertib, Radiation Therapy, and Temozolomide in Treating
Patients With Newly Diagnosed or Recurrent Glioblastoma
S AZD1775 + Radiation
Therapy + Temozolomide
R1
NCT02937818 A Phase II, Study to Determine the Preliminary Efficacy of Novel
Combinations of Treatment in Patients With Platinum Refractory
Extensive-Stage Small-Cell Lung Cancer
S AZD1775 + carboplatin ANR 2
NCT02546661 Open-Label, Randomised, Multi-Drug, Biomarker-Directed, Phase
1b Study in Pts w/ Muscle Invasive Bladder Cancer
S AZD1775 + Durvalumab ANR 1
NCT02659241 Adavosertib Before Surgery in Treating Patients With Advanced
High Grade Ovarian, Fallopian Tube, or Primary Peritoneal Cancer
S AZD1775 R 1
NCT02272790 Adavosertib Plus Chemotherapy in Platinum-Resistant Epithelial
Ovarian, Fallopian Tube, or Primary Peritoneal Cancer
S AZD1775 + Paclitaxel or
Carboplatin or
ANR 2
Ghelli Luserna di Rorà et al. Journal of Hematology & Oncology (2020) 13:126 Page 12 of 17
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
standard of care chemo/radiotherapy regimens depends on
the generation of DNA damages in proliferating malignant
cells. In this scenario, the overexpression or uncontrolled
activation of DDR pathways has been showed to protect
cancer cells from the therapeutic effect of DNA damaging
agents. Moreover, a large number of preclinical studies
highlighted that cancer cells depend on the functionality of
DDR pathways in order to survive, to tolerate the replica-
tive stress induced by the high proliferative rate and to
sustain the intrinsic genetic instability. For these reasons,
selective inhibitors have been developed in order to exploit
cancer cellsdependency on DDR-gene functionality. Pre-
clinical data has proven the efficacy of DDR inhibition in
different kinds of hematological and solid tumors, both as
monotherapy and in combination with a wide number of
DNA damaging agents. Among DDR inhibitors, the most
effectiveoncearethosetargetingPARP1andWEE1family
kinases. The effectiveness of PARP1 inhibitors is however
dependent on homologous recombination (HR) repair
deficiency while WEE1 family kinases inhibitors seems to
have a widespread efficacy independently from a specific
the genetic background. Indeed, cancer cells seem to be
strictly dependent on the functionality of WEE1/PKMYT1
kinases to survive, especially those with alterations targeting
the G1 checkpoint. WEE1/PKMYT1 kinases are involved
in different biological processes and they seem to play di-
verse roles in nonmalignant and in cancer cells. Indeed,
they control cell cycle regulation and genetic stability in
nonmalignant cells and for these reasons act as tumor
suppressor genes. Conversely, their ability of promote DNA
damages repair and cell cycle control makes them act as
pseudo-oncogenes in cancer cells. Several molecular studies
showed that malignant cells have high expression level of
WEE1 and PKMYT1, which has become a good prognostic
biomarker for chemo/radiotherapy regimens. However, we
currently lack information regarding predictive markers of
response to WEE1/PKMYT1 inhibitors. Large preclinical
and clinical studies should be conducted in order to identify
specific molecular backgrounds in which the use of WEE1/
PKYMT1 inhibitors may be recommended. The identifica-
tion of molecular vulnerabilities in cancer patients will be
fundamental to design novel therapeutic regimens using
WEE1/PKMYT1 inhibitors in a chemo/radiotherapy-free,
synthetic lethality-based approach.
Abbreviations
ALL: Acute lymphoblastic leukemia; AML: Acute myeloid leukemia; BC: Breast
cancer; C: Chemotherapy; CC: Colorectal cancer; CN: Copy number;
CNA: Copy number alteration; CLL: Chronic lymphocyte leukemia;
CML: Chronic myeloid leukemia; DDR: DNA damage response; DLBCL: Diffuse
large B cell lymphoma; ES: Esophageal cancer; GC: Gastric cancer;
GL: Gliomas; GLB: Glioblastoma; HC: Hepatocellular carcinoma;
HR: Homologous recombination; HNSCC: Head and neck squamous cell
carcinoma; LC: Lung cancer; M: Monotherapy; MaM: Malignant melanoma;
MCL: Mantle cell lymphoma; MM: Multiple myeloma; MPF: Mitotic promoting
factor; N: Neuroblastoma; NSCLC: Non-small-cell lung cancer; OC: Ovarian
cancer; PC: Pancreatic cancer; R: Radiotherapy; S: Sarcomas; SAC: Spindle
assembly checkpoint; TNBC: Triple-negative breast cancer
Acknowledgements
Not applicable.
Authorscontributions
AGLDR, CC, and GS drafted the first version of the manuscript and created
the figures. GM critically revised the manuscript for important intellectual
content. All authors read and approved the final manuscript.
Funding
This work was supported by ERA-Per-Med (reference number: ERAPERME
D2018-275).
Availability of data and materials
Not applicable.
Ethics approval and consent to participate
Not applicable.
Consent for publication
All authors read and approved the final manuscript.
Competing interests
GM has competing interests with Novartis, BMS, Roche, Pfizer, ARIAD, and
MSD. The other authors declare that they have no competing interests.
Table 4 Clinical trials evaluating WEE1/PKMYT1 inhibitor in monotherapy or in combination for cancer therapy (Continued)
Study ID Study title Tumor Interventions Status Phase
Gemcitabine or pegylated
liposomal doxorubicin
NCT02813135 European Proof-of-Concept Therapeutic Stratification Trial of
Molecular Anomalies in Relapsed or Refractory Tumors
S/H AZD1775 + carboplatin R 1,2
NCT03330847 To Assess Safety and Efficacy of Agents Targeting DNA Damage
Repair With Olaparib Versus Olaparib Monotherapy.
S AZD1775 + olaparib R 2
NCT01827384 MPACT Study to Compare Effects of Targeted Drugs on Tumor
Gene Variations
S AZD1775 + carboplatin R 2
NCT02465060 Targeted Therapy Directed by Genetic Testing in Treating
Patients With Advanced Refractory Solid Tumors, Lymphomas,
or Multiple Myeloma (The MATCH Screening Trial)
S/H AZD1775 R 2
Ssolid tumor, Hhematological tumor, Ccompleted, Rrecruiting, Wwithdraw, ANR active not recruiting, Tterminated, NA status unknown (last update 04/22/2020)
Ghelli Luserna di Rorà et al. Journal of Hematology & Oncology (2020) 13:126 Page 13 of 17
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Received: 18 June 2020 Accepted: 2 September 2020
References
1. Schmidt M, Rohe A, Platzer C, et al. Regulation of G2/M transition by
inhibition of WEE1 and PKMYT1 Kinases. Molecules. 2017;22:2045.
2. Solc P, Schultz RM, Motlik J. Prophase I arrest and progression to metaphase
I in mouse oocytes: Comparison of resumption of meiosis and recovery
from G2-arrest in somatic cells. Mol Hum Reprod. 2010;16:65464.
3. Nakanishi M, Ando H, Watanabe N, et al. Identification and characterization
of human Wee1B, a new member of the Wee1 family of Cdk-inhibitory
kinases. Genes Cells. 2000;5(10):83947.
4. Mueller PR, Coleman TR, Kumagai A, Dunphy WG. Myt1: A membrane-
associated inhibitory kinase that phosphorylates Cdc2 on both threonine-14
and tyrosine-15. Science. 1995;270(5233):8690.
5. Booher RN, Holman PS, Fattaey A. Human Myt1 is a cell cycle-regulated
kinase that inhibits Cdc2 but not Cdk2 activity. J Biol Chem. 1997;272(35):
223006.
6. Liu F, Stanton JJ, Wu Z, Piwnica-Worms H. The human Myt1 kinase
preferentially phosphorylates Cdc2 on threonine 14 and localizes to the
endoplasmic reticulum and Golgi complex. Mol Cell Biol. 1997;17(2):57183.
7. Nakajima H, Yonemura S, Murata M, et al. Myt1 protein kinase is
essential for Golgi and ER assembly during mitotic exit. J Cell Biol.
2008;181(1):89103.
8. Chow JPH, Poon RYC, Ma HT. Inhibitory phosphorylation of cyclin-
dependent kinase 1 as a compensatory mechanism for mitosis exit. Mol Cell
Biol. 2011;31(7):147891.
9. Solomon MJ, Harper JW, Shuttleworth J. CAK, the p34cdc2 activating kinase,
contains a protein identical or closely related to p40MO15. EMBO J. 1993;
12(8):313342.
10. Lolli G, Johnson LN. CAK-Cyclin-dependent activating kinase: a key kinase in
cell cycle control and a target for Drugs? Cell Cycle. 2005;4(4):5727.
11. Walsh S, Margolis SS, Kornbluth S. Phosphorylation of the cyclin B1
cytoplasmic retention sequence by mitogen-activated protein kinase and
Plx. Mol Cancer Res. 2003;1(4):2809.
12. Szmyd R, Niska-Blakie J, Diril MK, et al. Premature activation of Cdk1 leads to
mitotic events in S phase and embryonic lethality. Oncogene. 2019;38(7):
9981018.
13. Watanabe N, Arai H, Nishihara Y, et al. M-phase kinases induce phospho-
dependent ubiquitination of somatic Wee1 by SCFbeta-TrCP. Proc Natl
Acad Sci U S A. 2004;101(13):441924.
14. Toyoshima-Morimoto F, Taniguchi E, Shinya N, et al. Polo-like kinase 1
phosphorylates cyclin B1 and targets it to the nucleus during prophase.
Nature. 2001;410(6825):21520.
15. Van Vugt MATM, Brás A, Medema RH. Polo-like kinase-1 controls recovery
from a G2 DNA damage-induced arrest in mammalian cells. Mol Cell. 2004;
15(5):799811.
16. Nakojima H, Toyoshima-Morimoto F, Taniguchi E, Nishida E. Identification of
a consensus motif for PlK (Polo-like kinase) phosphorylation reveals Myt1 as
a Plk1 substrate. J Biol Chem. 2003;278(28):2527780.
17. Takisawa H, Mimura S, Kubota Y. Eukaryotic DNA replication: from pre-
replication complex to initiation complex. Curr Opin Cell Biol. 2000;12(6):
6906.
18. Heller RC, Kang S, Lam WM, et al. Eukaryotic origin-dependent DNA
replication in vitro reveals sequential action of DDK and S-CDK kinases. Cell.
2011;146(1):8091.
19. Labib K. How do Cdc7 and cyclin-dependent kinases trigger the initiation of
chromosome replication in eukaryotic cells? Genes Dev. 2010;24(12):120819.
20. Gu Y, Rosenblatt J, Morgan DO. Cell cycle regulation of CDK2 activity by
phosphorylation of Thr160 and Tyr15. EMBO J. 1992;11(11):39954005.
21. Vassilopoulos A, Tominaga Y, Kim HS, et al. WEE1 murine deficiency induces
hyper-activation of APC/C and results in genomic instability and
carcinogenesis. Oncogene. 2015;34(23):302335.
22. Ghelli Luserna Di Rorà A, Martinelli G, Simonetti G. The balance between
mitotic death and mitotic slippage in acute leukemia: a new therapeutic
window? J Hematol Oncol. 2019;12(1):123.
23. Visconti R, Grieco D. Fighting tubulin-targeting anticancer drug toxicity and
resistance. Endocr Relat Cancer. 2017;24(9):T10717.
24. Visconti R, Palazzo L, Della Monica R, Grieco D. Fcp1-dependent
dephosphorylation is required for M-phase-promoting factor inactivation at
mitosis exit. Nat Commun. 2012;3:894.
25. Musacchio A, Salmon ED. The spindle-assembly checkpoint in space and
time. Nat Rev Mol Cell Biol. 2007;8(5):37993.
26. Visconti R, Palazzo L, Pepe A, et al. The end of mitosis from a phosphatase
perspective. Cell Cycle. 2013;12(1):179.
27. Martín Y, Domínguez-Kelly R, Freire R. Novel insights into maintaining
genomic integrity: Wee1 regulating Mus81/Eme1. Cell Div. 2011;6:21.
28. Domínguez-Kelly R, Martín Y, Koundrioukoff S, et al. Wee1 controls genomic
stability during replication by regulating the Mus81-Eme1 endonuclease. J
Cell Biol. 2011;194(4):56779.
29. Duda H, Arter M, Gloggnitzer J, et al. A mechanism for controlled
breakage of under-replicated chromosomes during mitosis. Dev Cell.
2016;39(6):74055.
30. Asquith CRM, Laitinen T, East MP. PKMYT1: a forgotten member of the
WEE1 family. Nat Rev Drug Discov. 2020;19(3):157.
31. Liu Y, Qi J, Dou Z, et al. Systematic expression analysis of WEE family kinases
reveals the importance of PKMYT1 in breast carcinogenesis. Cell Prolif. 2020;
53(2):e12741.
32. Jeong D, Kim H, Kim D, et al. Protein kinase, membrane-associated tyrosine/
threonine 1 is associated with the progression of colorectal cancer. Oncol
Rep. 2018;39(6):282936.
33. Lal S, Cozzitorto JA, Blanco F, et al. 988 Sequence alterations in the WEE1
non-coding region is a facilitator and marker for pancreatic tumorigenesis.
Gastroenterology. 2014;S-1034.
34. Ciriello G, Miller ML, Aksoy BA, et al. Emerging landscape of oncogenic
signatures across human cancers. Nat Genet. 2013;45(10):112733.
35. Ghelli Luserna Di Rorà A, Beeharry N, Imbrogno E, et al. Targeting WEE1 to
enhance conventional therapies for acute lymphoblastic leukemia. J
Hematol Oncol. 2018;11(1):99.
36. Barbosa RSS, Dantonio PM, Guimarães T, et al. Sequential combination of
bortezomib and WEE1 inhibitor, MK-1775, induced apoptosis in multiple
myeloma cell lines. Biochem Biophys Res Commun. 2019;519(3):597604.
37. Van Linden AA, Baturin D, Ford JB, et al. Inhibition of Wee1 sensitizes cancer
cells to antimetabolite chemotherapeutics in vitro and in vivo, independent
of p53 functionality. Mol Cancer Ther. 2013;12(12):267584.
38. Porter CC, Kim J, Fosmire S, et al. Integrated genomic analyses identify
WEE1 as a critical mediator of cell fate and a novel therapeutic target in
acute myeloid leukemia. Leukemia. 2012;26(6):126676.
39. Johnston HE, Carter MJ, Larrayoz M, et al. Proteomics profiling of CLL versus
healthy B-cells identifies putative therapeutic targets and a subtype-
independent signature of spliceosome dysregulation. Mol Cell Proteomics.
2018;17(4):77691.
40. Neben K, Schnittger S, Brors B, et al. Distinct gene expression patterns
associated with FLT3- and NRAS-activating mutations in acute myeloid
leukemia with normal karyotype. Oncogene. 2005;24(9):15808.
41. Zhou L, Zhang Y, Chen S, et al. A regimen combining the Wee1 inhibitor
AZD1775 with HDAC inhibitors targets human acute myeloid leukemia cells
harboring various genetic mutations. Leukemia. 2015;29(4):80718.
42. Fan J, Li L, Small D, Rassool F. Cells expressing FLT3/ITD mutations exhibit
elevated repair errors generated through alternative NHEJ pathways:
implications for genomic instability and therapy. Blood. 2010;116(24):5298305.
43. De Jong MRW, Visser L, Huls G, et al. Identification of relevant drugable
targets in diffuse large B-cell lymphoma using a genome-wide unbiased
CD20 guilt-by association approach. PLoS One. 2018;13(2):e0193098.
44. Bolomsky A, Gruber F, Stangelberger K, et al. Preclinical validation studies
support causal machine learning based identification of novel drug targets
for high-risk multiple myeloma. Blood. 2018;132(Supplement 1):3210.
45. Sun QS, Luo M, Zhao HM, Sun H. Overexpression of PKMYT1 indicates the
poor prognosis and enhances proliferation and tumorigenesis in non-small
cell lung cancer via activation of Notch signal pathway. Eur Rev Med
Pharmacol Sci. 2019;23(10):42109.
46. Liu L, Wu J, Wang S, et al. PKMYT1 promoted the growth and motility of
hepatocellular carcinoma cells by activating beta-catenin/TCF signaling. Exp
Cell Res. 2017;358(2):20916.
47. Wang XM, Li QY, Ren LL, et al. Effects of MCRS1 on proliferation, migration,
invasion, and epithelial mesenchymal transition of gastric cancer cells by
interacting with Pkmyt1 protein kinase. Cell Signal. 2019;59:17181.
48. Kim HY, Cho Y, Kang HG, et al. Targeting the WEE1 kinase as a molecular
targeted therapy for gastric cancer. Oncotarget. 2016;7(31):4990216.
49. Magnussen GI, Holm R, Emilsen E, et al. High expression of Wee1 is
associated with poor disease-free survival in Malignant Melanoma: Potential
for targeted therapy. PLoS One. 2012;7(6):e38254.
Ghelli Luserna di Rorà et al. Journal of Hematology & Oncology (2020) 13:126 Page 14 of 17
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
50. Mueller S, Hashizume R, Yang X, et al. Targeting wee1 for the treatment of
pediatric high-grade gliomas. Neuro Oncol. 2014;16(3):35260.
51. Mir SE, De Witt Hamer PC, Krawczyk PM, et al. In silico analysis of kinase
expression identifies WEE1 as a gatekeeper against mitotic catastrophe in
glioblastoma. Cancer Cell. 2010;18(3):24457.
52. Chayka O, DAcunto CW, Middleton O, et al. Identification and
pharmacological inactivation of the MYCN gene network as a therapeutic
strategy for neuroblastic tumor cells. J Biol Chem. 2015;290(4):2198212.
53. Zhang Q, Zhao X, Zhang C, et al. Overexpressed PKMYT1 promotes tumor
progression and associates with poor survival in esophageal squamous cell
carcinoma. Cancer Manag Res. 2019;11:781324.
54. Tibes R, Bogenberger JM, Chaudhuri L, et al. RNAi screening of the kinome
with cytarabine in leukemias. Blood. 2012;119(12):286372.
55. Simonetti G, Padella A, do Valle IF, et al. Aneuploid acute myeloid leukemia
exhibits a signature of genomic alterations in the cell cycle and protein
degradation machinery. Cancer. 2018;125:114.
56. Caretti V, Hiddingh L, Lagerweij T, et al. WEE1 kinase inhibition enhances
the radiation response of diffuse intrinsic pontine gliomas. Mol Cancer Ther.
2013;12(2):14150.
57. Music D, Dahlrot RH, Hermansen SK, et al. Expression and prognostic value
of the WEE1 kinase in gliomas. J Neurooncol. 2016;127(2):3819.
58. Egeland EV, Flatmark K, Nesland JM, et al. Expression and clinical
significance of Wee1 in colorectal cancer. Tumor Biol. 2016;37(9):1213340.
59. Slipicevic A, Holth A, Hellesylt E, et al. Wee1 is a novel independent
prognostic marker of poor survival in post-chemotherapy ovarian carcinoma
effusions. Gynecol Oncol. 2014;135(1):11824.
60. Shu C, Wang Q, Yan X, Wang J. Whole-genome expression microarray
combined with machine learning to identify prognostic biomarkers for
high-grade glioma. J Mol Neurosci. 2018;64(4):491500.
61. Novak EM, Halley NS, Gimenez TM, et al. BLM germline and somatic PKMY
T1 and AHCY mutations: genetic variations beyond MYCN and prognosis in
neuroblastoma. Med Hypotheses. 2016;97:225.
62. Ku BM, Bae Y-H, Koh J, et al. Mutational status of TP53 defines the efficacy
of Wee1 inhibitor AZD1775 in KRAS -mutant non-small cell lung cancer.
Oncotarget. 2017;8(40):6752637.
63. Bauman JE, Chung CH. CHK it out! Blocking WEE kinase routs TP53 mutant
cancer. Clin Cancer Res. 2014;20(16):41735.
64. Kreahling JM, Foroutan P, Reed D, et al. Wee1 inhibition by MK-1775 leads
to tumor inhibition and enhances efficacy of gemcitabine in human
sarcomas. PLoS One. 2013;8(3):e57523.
65. Ford JB, Baturin D, Burleson TM, et al. AZD1775 sensitizes T cell acute
lymphoblastic leukemia cells to cytarabine by promoting apoptosis over
DNA repair. Oncotarget. 2015;6(29):2800110.
66. Webster PJ, Littlejohns AT, Gaunt HJ, et al. AZD1775 induces toxicity through
double-stranded DNA breaks independently of chemotherapeutic agents in
p53-mutated colorectal cancer cells. Cell Cycle. 2017;16(22):217682.
67. Kahen E, Yu D, Harrison DJ, et al. Identification of clinically achievable
combination therapies in childhood rhabdomyosarcoma. Cancer
Chemother Pharmacol. 2016;78(2):31323.
68. Lee YY, Cho YJ, Shin SW, et al. Anti-tumor effects of Wee1 kinase inhibitor
with radiotherapy in human cervical cancer. Sci Rep. 2019;9(1):15394.
69. PosthumaDeBoer J, Würdinger T, Graat HCA, et al. WEE1 inhibition sensitizes
osteosarcoma to radiotherapy. BMC Cancer. 2011;11(1):156.
70. Xu H, Krystal GW. Actinomycin D decreases Mcl-1 expression and acts
synergistically with ABT-737 against small cell lung cancer cell lines. Clin
Cancer Res. 2010;16(17):4392400.
71. Hayashi Y, Fujimura A, Kato K, et al. Nucleolar integrity during interphase
supports faithful Cdk1 activation and mitotic entry. Sci Adv. 2018;4(6):
eaap7777.
72. Alexander VM, Roy M, Steffens KA, et al. Azacytidine induces cell cycle arrest
and suppression of neuroendocrine markers in carcinoids. Int J Clin Exp
Med. 2010;3(2):95102.
73. Uchida R, Yokota S, Matsuda D, et al. Habiterpenol, a novel abrogator of
bleomycin-induced G2 arrest in Jurkat cells, produced by Phytohabitans
suffuscus 3787-5. J Antibiot (Tokyo). 2014;67(11):77781.
74. Zhang Z, Zhang H, Hu Z, et al. Synergy of 1,25-dihydroxyvitamin D3 and
carboplatin in growth suppression of SKOV-3 cells. Oncol Lett. 2014;8(3):
134854.
75. Sarin N, Engel F, Kalayda GV, et al. Cisplatin resistance in non-small cell lung
cancer cells is associated with an abrogation of cisplatin-induced G2/M cell
cycle arrest. PLoS One. 2017;12(7):e0181081.
76. Chen D, Lin X, Gao J, et al. Wee1 Inhibitor AZD1775 Combined with
cisplatin potentiates anticancer activity against gastric cancer by increasing
DNA damage and cell apoptosis. Biomed Res Int. 2018;2018:5813292.
77. Zheng H, Shao F, Martin S, et al. WEE1 inhibition targets cell cycle
checkpoints for triple negative breast cancers to overcome cisplatin
resistance. Sci Rep. 2017;7:43517.
78. de Jong MRW, Langendonk M, Reitsma B, et al. WEE1 inhibition synergizes
with CHOP chemotherapy and radiation therapy through induction of
premature mitotic entry and DNA damage in diffuse large B-cell lymphoma.
Ther Adv Hematol. 2020;11:2040620719898373.
79. Ma J, Li X, Su Y, et al. Mechanisms responsible for the synergistic
antileukemic interactions between ATR inhibition and cytarabine in acute
myeloid leukemia cells. Sci Rep. 2017;7:41950.
80. Shi Z, Azuma A, Sampath D, et al. S-phase arrest by nucleoside analogues
and abrogation of survival without cell cycle progression by 7-
hydroxystaurosporine. Cancer Res. 2001;61(3):106572.
81. Garcia TB, Fosmire SP, Porter CC. Increased activity of both CDK1 and CDK2
is necessary for the combinatorial activity of WEE1 inhibition and cytarabine.
Leuk Res. 2018;64:303.
82. Al-Aamri HM, Ku H, Irving HR, et al. Time dependent response of
daunorubicin on cytotoxicity, cell cycle and DNA repair in acute
lymphoblastic leukaemia. BMC Cancer. 2019;19(1):179.
83. Shang D, Han T, Xu X, Liu Y. Decitabine induces G2/M cell cycle arrest by
suppressing p38/NF-κB signaling in human renal clear cell carcinoma. Int J
Clin Exp Pathol. 2015;8(9):111408.
84. Singh SK, Banerjee S, Acosta EP, et al. Resveratrol induces cell cycle arrest
and apoptosis with docetaxel in prostate cancer cells via a p53/p21WAF1/
CIP1 and p27KIP1 pathway. Oncotarget. 2017;8(10):1721628.
85. Morse DL, Gray H, Payne CM, Gillies RJ. Docetaxel induces cell death
through mitotic catastrophe in human breast cancer cells. Mol Cancer Ther.
2005;4(10):1495504.
86. Vera J, Raatz Y, Wolkenhauer O, et al. Chk1 and Wee1 control genotoxic-
stress induced G2-M arrest in melanoma cells. Cell Signal. 2015;27(5):95160.
87. Wu CL, Ping SY, Yu CP, Yu DS. Tyrosine kinase receptor inhibitor-targeted
combined chemotherapy for metastatic bladder cancer. Kaohsiung J Med
Sci. 2012;28(4):194203.
88. Senthebane DA, Jonker T, Rowe A, et al. The role of tumor
microenvironment in chemoresistance: 3D extracellular matrices as
accomplices. Int J Mol Sci. 2018;19(10):2861.
89. Chen X, Low KH, Alexander A, et al. Cyclin E overexpression sensitizes triple-
negative breast cancer to Wee1 kinase inhibition. Clin Cancer Res. 2018;
24(24):6594610.
90. Seung HL, Seung MS, Dong JS, et al. Epothilones induce human colon
cancer SW620 cell apoptosis via the tubulin polymerization-independent
activation of the nuclear factor-κB/IκB kinase signal pathway. Mol Cancer
Ther. 2007;6(10):278697.
91. Zhang R, Zhu L, Zhang L, et al. PTEN enhances G2/M arrest in etoposide-
treated MCF-7 cells through activation of the ATM pathway. Oncol Rep.
2016;35(5):270714.
92. Pitts TM, Simmons DM, Bagby SM, et al. Wee1 inhibition enhances the anti-
tumor effects of capecitabine in preclinical models of triple-negative breast
cancer. Cancers. 2020;12(3):719.
93. Rajeshkumar NV, De Oliveira E, Ottenhof N, et al. MK-1775, a potent Wee1
inhibitor, synergizes with gemcitabine to achieve tumor regressions,
selectively in p53-deficient pancreatic cancer xenografts. Clin Cancer Res.
2011;17(9):2799806.
94. Koh SB, Wallez Y, Dunlop CR, et al. Mechanistic distinctions between CHK1
and WEE1 inhibition guide the scheduling of triple therapy with
gemcitabine. Cancer Res. 2018;78(11):305466.
95. Xu B, Sun Z, Liu Z, et al. Replication stress induces micronuclei comprising
of aggregated DNA double-strand breaks. PLoS One. 2011;6(4):e18618.
96. Aarts M, Sharpe R, Garcia-Murillas I, et al. Forced mitotic entry of S-phase
cells as a therapeutic strategy induced by inhibition of WEE1. Cancer Discov.
2012;2(6):52439.
97. Morgan MA, Onono FO, Spielmann HP, et al. Modulation of anthracycline-
induced cytotoxicity by targeting the prenylated proteome in myeloid
leukemia cells. J Mol Med. 2012;90(2):14961.
98. Subhash VV, Tan SH, Yeo MS, et al. ATM expression predicts veliparib
and irinotecan sensitivity in gastric cancer by mediating p53-
independent regulation of cell cycle and apoptosis. Mol Cancer Ther.
2016;15(12):308796.
Ghelli Luserna di Rorà et al. Journal of Hematology & Oncology (2020) 13:126 Page 15 of 17
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
99. Yin Y, Shen Q, Tao R, et al. Wee1 inhibition can suppress tumor proliferation
and sensitize p53 mutant colonic cancer cells to the anticancer effect of
irinotecan. Mol Med Rep. 2018;17(2):33449.
100. Jan YH, Heck DE, Laskin DL, Laskin JD. Sulfur mustard analog
mechlorethamine (Bis(2-chloroethyl)methylamine) modulates cell cycle
progression via the DNA damage response in human lung epithelial A549
cells. Chem Res Toxicol. 2019;32(6):112333.
101. Mahbub A, Le Maitre C, Haywood-Small S, et al. Dietary polyphenols
influence antimetabolite agents: Methotrexate, 6-mercaptopurine and 5-
fluorouracil in leukemia cell lines. Oncotarget. 2017;8(62):10487793.
102. Costantini DL, Villani DF, Vallis KA, Reilly RM. Methotrexate, paclitaxel, and
doxorubicin radiosensitize HER2-amplified human breast cancer cells to the
auger electron-emitting radiotherapeutic agent 111In-NLS-trastuzumab. J
Nucl Med. 2010;51(3):47783.
103. Di Rorà AGL, Bocconcelli M, Ferrari A, et al. Synergism through WEE1 and
CHK1 inhibition in acute lymphoblastic leukemia. Cancers (Basel). 2019;
11(11):1654.
104. Guerriero E, Sorice A, Capone F, et al. Vitamin C effect on mitoxantrone-
induced cytotoxicity in human breast cancer cell lines. PLoS One. 2014;
9(12):e115287.
105. Voland C, Bord A, Péleraux A, et al. Repression of cell cycle-related proteins
by oxaliplatin but not cisplatin in human colon cancer cells. Mol Cancer
Ther. 2006;5(9):214957.
106. Lal S, Zarei M, Chand SN, et al. WEE1 inhibition in pancreatic cancer cells is
dependent on DNA repair status in a context dependent manner. Sci Rep.
2016;6:33323.
107. George J, Banik NL, Ray SK. Molecular mechanisms of taxol for induction of
cell death in glioblastomas. In: Ray S, editor. Glioblastoma. New York:
Springer; 2010. https://doi.org/10.1007/978-1-4419-0410-2_14.
108. Lewis CW, Jin Z, Macdonald D, et al. Prolonged mitotic arrest induced by
Wee1 inhibition sensitizes breast cancer cells to paclitaxel. Oncotarget. 2017;
8(43):7370522.
109. Chen KC, Yang TY, Wu CC, et al. Pemetrexed induces S-phase arrest and
apoptosis via a deregulated activation of Akt signaling pathway. PLoS One.
2014;9(5):e97888.
110. Hirai H, Arai T, Okada M, et al. MK-1775, a small molecule Wee1 inhibitor,
enhances antitumor efficacy of various DNA-damaging agents, including 5-
fluorouracil. Cancer Biol Ther. 2010;9(7):51422.
111. Maier P, Hartmann L, Wenz F, Herskind C. Cellular pathways in response to
ionizing radiation and their targetability for tumor radiosensitization. Int J
Mol Sci. 2016;17(1):102.
112. Li J, Chen W, Zhang P, Li N. Topoisomerase II trapping agent teniposide
induces apoptosis and G2/M or S phase arrest of oral squamous cell
carcinoma. World J Surg Oncol. 2006;4:41.
113. Wotring LL, Roti Roti JL. Thioguanine-induced S and G2 blocks and
their significance to the mechanism of cytotoxicity. Cancer Res. 1980;
40(5):145862.
114. Nguyen D, Zajac-Kaye M, Rubinstein L, et al. Poly(ADP-ribose) polymerase
inhibition enhances p53-dependent and -independent DNA damage
responses induced by DNA damaging agent. Cell Cycle. 2011;10(23):407482.
115. Shumway SD, Kubica JL, Guertin AD, et al. Abstract 2969: a Wee1 kinase
inhibitor, MK-1775, sensitizes cervical carcinoma cell lines to cisplatin and
topotecan. Cancer Res. 2011;71(8 Supplement):2969.
116. Brandl MB, Pasquier E, Li F, et al. Computational analysis of image-based
drug profiling predicts synergistic drug combinations: applications in triple-
negative breast cancer. Mol Oncol. 2014;8(8):154860.
117. Tu Y, Cheng S, Zhang S, et al. Vincristine induces cell cycle arrest and
apoptosis in SH-SY5Y human neuroblastoma cells. Int J Mol Med. 2013;
31(1):1139.
118. Visconti R, Della Monica R, Palazzo L, et al. The Fcp1-Wee1-Cdk1 axis affects
spindle assembly checkpoint robustness and sensitivity to antimicrotubule
cancer drugs. Cell Death Differ. 2015;22(9):155160.
119. Zhu JY, Cuellar RA, Berndt N, et al. Structural basis of Wee kinases
functionality and inactivation by diverse small molecule inhibitors. J Med
Chem. 2017;60(18):786375.
120. Restelli V, Chilà R, Lupi M, et al. Characterization of a mantle cell lymphoma
cell line resistant to the Chk1 inhibitor PF-00477736. Oncotarget. 2015;6(35):
3722940.
121. Qi W, Xie C, Li C, et al. CHK1 plays a critical role in the anti-leukemic activity
of the wee1 inhibitor MK-1775 in acute myeloid leukemia cells. J Hematol
Oncol. 2014;7(1):53.
122. Young LA, OConnor LO, de Renty C, et al. Differential activity of ATR and
Wee1 inhibitors in a highly sensitive subpopulation of DLBCL linked to
replication stress. Cancer Res. 2019;79(14):376275.
123. Bridges KA, Hirai H, Buser CA, et al. MK-1775, a novel wee1 kinase inhibitor,
radiosensitizes p53-defective human tumor cells. Clin Cancer Res. 2011;
17(17):563848.
124. Ma H, Takahashi A, Sejimo Y, et al. Targeting of carbon ion-induced G2
checkpoint activation in lung cancer cells using Wee-1 inhibitor MK-1775.
Radiat Res. 2016;185(2):e52.
125. Lindenblatt D, Terraneo N, Pellegrini G, et al. Combination of lutetium-177
labelled anti-L1CAM antibody chCE7 with the clinically relevant protein
kinase inhibitor MK1775: a novel combination against human ovarian
carcinoma. BMC Cancer. 2018;18(1):922.
126. Parsels LA, Karnak D, Parsels JD, et al. PARP1 Trapping and DNA replication
stress enhance radiosensitization with combined WEE1 and PARP inhibitors.
Mol Cancer Res. 2018;16(2):22232.
127. Caldwell JT, Edwards H, Buck SA, et al. Targeting the wee1 kinase for
treatment of pediatric Down syndrome acute myeloid leukemia. Pediatr
Blood Cancer. 2014;61(10):176773.
128. Xing L, Lin L, Yu T, et al. A novel BCMA PBD-ADC with ATM/ATR/WEE1
inhibitors or bortezomib induce synergistic lethality in multiple myeloma.
Leukemia. 2020. https://doi.org/10.1038/s41375-020-0745-9.
129. Tibes R, Ferreira Coutinho D, Tuen MT, et al. DNA damage repair
interference By WEE1 inhibition with AZD1775 overcomes combined
azacitidine and Venetoclax resistance in acute myeloid leukmeia (AML).
Blood. 2019;134(Supplement_1):2559.
130. Qi W, Zhang W, Edwards H, et al. Synergistic anti-leukemic interactions
between panobinostat and MK-1775 in acute myeloid leukemia ex vivo.
Cancer Biol Ther. 2015;16(12):178493.
131. Tanaka N, Patel AA, Tang L, et al. Replication stress leading to apoptosis
within the S-phase contributes to synergism between vorinostat and
AZD1775 in HNSCC harboring high-risk TP53 mutation. Clin Cancer Res.
2017;23(21):654154.
132. Qi W, Xu X, Wang M, et al. Inhibition of Wee1 sensitizes AML cells to ATR
inhibitor VE-822-induced DNA damage and apoptosis. Biochem Pharmacol.
2019;164:27382.
133. Restelli V, Lupi M, Chila R, et al. DNA damage response inhibitor
combinations exert synergistic antitumor activity in aggressive B-cell
lymphomas. Mol Cancer Ther. 2019;18(7):125564.
134. Bukhari AB, Lewis CW, Pearce JJ, et al. Inhibiting Wee1 and ATR kinases
produces tumor-selective synthetic lethality and suppresses metastasis. J
Clin Invest. 2019;129(3):132944.
135. Jin J, Fang H, Yang F, et al. Combined inhibition of ATR and WEE1 as a
novel therapeutic strategy in triple-negative breast cancer. Neoplasia
(United States). 2018;20(5):47888.
136. Weisberg E, Nonami A, Chen Z, et al. Identification of Wee1 as a novel
therapeutic target for mutant RAS-driven acute leukemia and other
malignancies. Leukemia. 2014;29(1):2737.
137. Li F, Guo E, Huang J, et al. mTOR inhibition overcomes primary and
acquired resistance to Wee1 inhibition by augmenting replication stress in
epithelial ovarian cancers. Am J Cancer Res. 2020;10(3):90824.
138. Sen T, Tong P, Diao L, et al. Targeting AXL and mTOR pathway overcomes
primary and acquired resistance to WEE1 inhibition in small-cell lung
cancer. Clin Cancer Res. 2017;23(20):623954.
139. Hai J, Liu S, Bufe L, et al. Synergy of WEE1 and mTOR inhibition in mutant
KRAS-driven lung cancers. Clin Cancer Res. 2017;23(22):69937005.
140. Chila R, Basana A, Lupi M, et al. Combined inhibition of Chk1 and Wee1 as a
new therapeutic strategy for mantle cell lymphoma. Oncotarget. 2015;6(5):
3394408.
141. Restelli V, Vagni M, Arribas AJ, et al. Inhibition of CHK1 and WEE1 as a new
therapeutic approach in diffuse large B cell lymphomas with MYC
deregulation. Br J Haematol. 2018;181(1):12933.
142. Chaudhuri L, Vincelette ND, Koh BD, et al. CHK1 and WEE1 inhibition
combine synergistically to enhance therapeutic efficacy in acute myeloid
leukemia ex vivo. Haematologica. 2014;99(4):68896.
143. De Jong MRW, Langendonk M, Reitsma B, et al. WEE1 inhibition enhances
anti-apoptotic dependency as a result of premature mitotic entry and DNA
damage. Cancers (Basel). 2019;11(11):1743.
144. Fang Y, McGrail DJ, Sun C, et al. Sequential therapy with PARP and WEE1
inhibitors minimizes toxicity while maintaining efficacy. Cancer Cell. 2019;
35(6):851867.e7.
Ghelli Luserna di Rorà et al. Journal of Hematology & Oncology (2020) 13:126 Page 16 of 17
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
145. Lallo A, Frese KK, Morrow CJ, et al. The combination of the PARP inhibitor
olaparib and the WEE1 Inhibitor AZD1775 as a new therapeutic option for
small cell lung cancer. Clin Cancer Res. 2018;24(20):515364.
146. Garcia TB, Snedeker JC, Baturin D, et al. A small-molecule inhibitor of WEE1,
AZD1775, synergizes with olaparib by impairing homologous
recombination and enhancing DNA damage and apoptosis in acute
leukemia. Mol Cancer Ther. 2017;16(10):205868.
147. Lee JW, Parameswaran J, Sandoval-Schaefer T, et al. Combined aurora
kinase A (AURKA) and WEE1 inhibition demonstrates synergistic antitumor
effect in squamous cell carcinoma of the head and neck. Clin Cancer Res.
2019;25(11):343042.
148. Chen G, Zhang B, Xu H, et al. Suppression of Sirt1 sensitizes lung cancer
cells to WEE1 inhibitor MK-1775-induced DNA damage and apoptosis.
Oncogene. 2017;36(50):686372.
149. Francis AM, Alexander A, Liu Y, et al. CDK4/6 inhibitors sensitize Rb-positive
sarcoma cells to Wee1 kinase inhibition through reversible cell-cycle arrest.
Mol Cancer Ther. 2017;16(9):175164.
150. Takashima Y, Kikuchi E, Kikuchi J, et al. Bromodomain and extraterminal
domain inhibition synergizes with WEE1-inhibitor AZD1775 effect by
impairing nonhomologous end joining and enhancing DNA damage in
nonsmall cell lung cancer. Int J Cancer. 2020;146(4):111424.
151. Panek RL, Lu GH, Klutchko SR, et al. In vitro pharmacological
characterization of PD 166285, a new nanomolar potent and broadly
active protein tyrosine kinase inhibitor. J Pharmacol Exp Ther. 1997;
283(3):143344.
152. Duan L, Perez RE, Hansen M, et al. Increasing cisplatin sensitivity by
scheduledependent inhibition of AKT and Chk1. Cancer Biol Ther. 2014;
15(12):160012.
153. Blackwood E, Epler J, Yen I, et al. Combination drug scheduling defines a
window of opportunityfor chemopotentiation of gemcitabine by an orally
bioavailable, selective ChK1 inhibitor, GNE-900. Mol Cancer Ther. 2013;
12(10):196880.
154. Levinson NM, Boxer SG. Structural and spectroscopic analysis of the kinase
inhibitor bosutinib and an isomer of bosutinib binding to the Abl tyrosine
kinase domain. PLoS One. 2012;7(4):e29828.
155. Beeharry N, Banina E, Hittle J, et al. Re-purposing clinical kinase inhibitors to
enhance chemosensitivity by overriding checkpoints. Cell Cycle. 2014;13(14):
217291.
156. Mendez E, Rodriguez CP, Kao MC, et al. A phase I clinical trial of AZD1775 in
combination with neoadjuvant weekly docetaxel and cisplatin before
definitive therapy in head and neck squamous cell carcinoma. Clin Cancer
Res. 2018;24(12):27408.
157. Do K, Wilsker D, Ji J, et al. Phase I study of single-agent AZD1775 (MK-1775),
a wee1 kinase inhibitor, in patients with refractory solid tumors. J Clin
Oncol. 2015;33(30):340915.
158. Leijen S, Van Geel RMJM, Pavlick AC, et al. Phase I study evaluating WEE1
inhibitor AZD1775 as monotherapy and in combination with gemcitabine,
cisplatin, or carboplatin in patients with advanced solid tumors. J Clin
Oncol. 2016;34(36):437180.
159. Leijen S, Van Geel RMJM, Sonke GS, et al. Phase II study of WEE1 inhibitor
AZD1775 plus carboplatin in patientswith tp53-mutated ovarian cancer
refractory or resistant to first-line therapy within 3 months. J Clin Oncol.
2016;34(36):435461.
160. Moore KN, Chambers SK, Hamilton EP, et al. Adavosertib with
chemotherapy (CT) in patients (pts) with platinum-resistant ovarian cancer
(PPROC): an open label, four-arm, phase II study. J Clin Oncol. 2019;5(_
suppl):5513.
161. Yap TA, Plummer R, Azad NS, Helleday T. The DNA damaging revolution:
PARP inhibitors and beyond. Am Soc Clin Oncol Educ B. 2019;39:18595.
162. Forment JV, OConnor MJ. Targeting the replication stress response in
cancer. Pharmacol Ther. 2018;188:15567.
163. Fu S, Wang Y, Keyomarsi K, Meric-Bernstein F. Strategic development of
AZD1775, a Wee1 kinase inhibitor, for cancer therapy. Expert Opin Investig
Drugs. 2018;27(9):74151.
164. Qiu Z, Oleinick NL, Zhang J. ATR/CHK1 inhibitors and cancer therapy.
Radiother Oncol. 2018;126(3):45064.
165. Ghelli Luserna Di Rora A, Iacobucci I, Martinelli G. The cell cycle checkpoint
inhibitors in the treatment of leukemias. J Hematol Oncol. 2017;10(1):77.
166. Lewis CW, Bukhari AB, Xiao EJ, et al. Upregulation of MyT1 promotes
acquired resistance of cancer cells to WEE1 inhibition. Cancer Res. 2019;
79(23):597185.
167. Garcia TB, Uluisik RC, van Linden AA, et al. Increased HDAC activity and c-
MYC expression mediate acquired resistance to WEE1 inhibition in acute
leukemia. Front Oncol. 2020;10:296.
PublishersNote
Springer Nature remains neutral with regard to jurisdictional claims in
published maps and institutional affiliations.
Ghelli Luserna di Rorà et al. Journal of Hematology & Oncology (2020) 13:126 Page 17 of 17
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
... WEE1 is a tyrosine kinase that functions through inhibition of both CDK1 and CDK2 activity by phosphorylation at tyrosine 15 (Y15). It is involved in several stages of cell cycle regulation, including intra-S, G2/M and M (23). Additionally, WEE1 regulates the G1/S cell cycle transition and protects stalled replication forks through inhibiting CDK2 activity (24,25). ...
... Compared with control cells, DNA fibers in OV90 Cyclin E1 cells were shorter, indicating Cyclin E1 overexpression alone increased baseline replication stress. Azenosertib treatment reduced the length of DNA fibers in both cell lines ( Fig. 2D; Supplementary Fig. S2F), supporting previous reports that WEE1 inhibition resulted in replication stress (23). The significantly shorter length of DNA fibers in OV90 Cyclin E1 cells compared to control cells suggests a model wherein azenosertib treatment exacerbates existing replication stress beyond threshold levels, ultimately leading to cell death. ...
Preprint
Full-text available
Upregulation of Cyclin E1 and subsequent activation of CDK2 accelerates cell-cycle progression from G1 to S phase and is a common oncogenic driver in gynecological malignancies. WEE1 kinase counteracts the effects of Cyclin E1/CDK2 activation by regulating multiple cell cycle checkpoints. Here we characterized the relationship between Cyclin E1/CDK2 activation and sensitivity to the selective WEE1 inhibitor azenosertib. We found that ovarian cancer cell lines with high levels of endogenous Cyclin E1 expression or forced overexpression were exquisitely sensitive to azenosertib and these results extended to in vivo models of ovarian and uterine serous carcinoma. Models with high Cyclin E1 expression showed higher baseline levels of replication stress and enhanced cellular responses to azenosertib treatment. We found azenosertib synergized with different classes of chemotherapy and described distinct underlying mechanisms. Finally, we provided early evidence from an ongoing phase I study demonstrating clinical activity of monotherapy azenosertib in patients with Cyclin E1/CDK2 activated ovarian and uterine serous carcinomas.
... An additional gene implicated in human bladder cancer, the tumor suppressor WEE1, is located in the disease associated chr21:31405230-33951574 region. WEE1 regulates the G2/M checkpoint through an activation loop with cell division kinases, CDK1 and CDK2, within the MAPK signaling pathway 56,57 . Activating BRAF mutations, which are found in the majority of canine iUC tumors, increase cellular levels of WEE1 protein 58 . ...
... This mutation is in the N-terminal regulatory domain of WEE1, adjacent to the nuclear export signal and is predicted to affect the function of the gene. In humans bladder urothelial carcinomas WEE1 is frequently dysregulated through copy number loss (13.7%) 57 . ...
Article
Full-text available
Naturally occurring canine invasive urinary carcinoma (iUC) closely resembles human muscle invasive bladder cancer in terms of histopathology, metastases, response to therapy, and low survival rate. The heterogeneous nature of the disease has led to the association of large numbers of risk loci in humans, however most are of small effect. There exists a need for new and accurate animal models of invasive bladder cancer. In dogs, distinct breeds show markedly different rates of iUC, thus presenting an opportunity to identify additional risk factors and overcome the locus heterogeneity encountered in human mapping studies. In the association study presented here, inclusive of 100 Shetland sheepdogs and 58 dogs of other breeds, we identify a homozygous protein altering point mutation within the NIPAL1 gene which increases risk by eight-fold (OR = 8.42, CI = 3.12–22.71), accounting for nearly 30% of iUC risk in the Shetland sheepdog. Inclusion of six additional loci accounts for most of the disease risk in the breed and explains nearly 75% of the phenotypes in this study. When combined with sequence data from tumors, we show that variation in the MAPK signaling pathway is an overarching cause of iUC susceptibility in dogs.
... Similarly, a combination of PARPi with WEE1 inhibitors has exhibited encouraging results in a recent phase II study [98]. WEE1 is a kinase that regulates the cell cycle by inhibiting CDK1 in response to DNA damage, thereby preventing DNA replication [99]. When WEE1 is inhibited, the cell cycle progresses without checkpoints, leading to DNA damage accumulation and cell death [99]. ...
... WEE1 is a kinase that regulates the cell cycle by inhibiting CDK1 in response to DNA damage, thereby preventing DNA replication [99]. When WEE1 is inhibited, the cell cycle progresses without checkpoints, leading to DNA damage accumulation and cell death [99]. The WEE1 inhibitor adavosertib was evaluated alone or in combination with olaparib in 80 patients with PARPi-resistant OC [98]. ...
Article
Full-text available
Epithelial ovarian cancer (EOC) is one of the most aggressive forms of gynaecological malignancies. Survival rates for women diagnosed with OC remain poor as most patients are diagnosed with advanced disease. Debulking surgery and platinum-based therapies are the current mainstay for OC treatment. However, and despite achieving initial remission, a significant portion of patients will relapse because of innate and acquired resistance, at which point the disease is considered incurable. In view of this, novel detection strategies and therapeutic approaches are needed to improve outcomes and survival of OC patients. In this review, we summarize our current knowledge of the genetic landscape and molecular pathways underpinning OC and its many subtypes. By examining therapeutic strategies explored in preclinical and clinical settings, we highlight the importance of decoding how single and convergent genetic alterations co-exist and drive OC progression and resistance to current treatments. We also propose that core signalling pathways such as the PI3K and MAPK pathways play critical roles in the origin of diverse OC subtypes and can become new targets in combination with known DNA damage repair pathways for the development of tailored and more effective anti-cancer treatments.
Preprint
Full-text available
Mature T-cell lymphomas and leukemias (mTCL) comprise a clinically and genetically heterogeneous group of lymphoid malignancies. Most subtypes of peripheral T-cell lymphomas and leukemic T-cell malignancies show an aggressive clinical course and poor prognosis. Thus, these diseases urgently require novel therapeutic strategies. Taking advantage of recent progress deciphering the genetic basis of mTCL, we generated a comprehensive database of genetic alterations from >1 800 patients with mTCL and utilized bioinformatic methodology developed to support treatment decisions in molecular tumorboards to identify novel potential therapeutics. To assess the in vitro activity of potential therapeutics, broad drug screening was performed in molecularly characterized cell lines of mTCL. Notably, the cell cycle regulator WEE1 was identified as a novel therapeutic target in mTCL. Indeed, WEE1 kinase inhibitors potently induced replication stress, premature mitotic entry, accumulation of DNA damage and induction of apoptosis in mTCL cell lines. Exploring potential drug combination strategies through mechanistic studies, we identified strong synergistic effects of combined WEE1 and JAK inhibition in JAK/STAT driven preclinical models as well as in primary patient samples of T-cell prolymphocytic leukemia (T-PLL). In summary, our results identified combinatorial effects of WEE1 and JAK inhibition in genetically defined subtypes of mTCL.
Article
Full-text available
Resistance to therapy commonly develops in patients with high-grade serous ovarian carcinoma (HGSC) and triple-negative breast cancer (TNBC), urging the search for improved therapeutic combinations and their predictive biomarkers. Starting from a CRISPR knockout screen, we identified that loss of RB1 in TNBC or HGSC cells generates a synthetic lethal dependency on casein kinase 2 (CK2) for surviving the treatment with replication-perturbing therapeutics such as carboplatin, gemcitabine, or PARP inhibitors. CK2 inhibition in RB1-deficient cells resulted in the degradation of another RB family cell cycle regulator, p130, which led to S phase accumulation, micronuclei formation, and accelerated PARP inhibition–induced aneuploidy and mitotic cell death. CK2 inhibition was also effective in primary patient-derived cells. It selectively prevented the regrowth of RB1-deficient patient HGSC organoids after treatment with carboplatin or niraparib. As about 25% of HGSCs and 40% of TNBCs have lost RB1 expression, CK2 inhibition is a promising approach to overcome resistance to standard therapeutics in large strata of patients.
Article
Objectives GL-V9 exhibited anti-tumour effects on various types of tumours. This study aimed to verify if GL-V9 synergized with oxaliplatin in suppressing colorectal cancer (CRC) and to explore the synergistic mechanism. Methods The synergy effect was tested by MTT assays and the mechanism was examined by comet assay, western blotting and immunohistochemistry (IHC). Xenograft model was constructed to substantiated the synergy effect and its mechanism in vivo. Results GL-V9 was verified to enhance the DNA damage effect of oxaliplatin, so as to synergistically suppress colon cancer cells in vitro and in vivo. In HCT-116 cells, GL-V9 accelerated the degradation of Wee1 and induced the abrogation of cell cycle arrest and mis-entry into mitosis, bypassing the DNA damage response caused by oxaliplatin. Our findings suggested that GL-V9 binding to HSP90 was responsible for the degradation of Wee1 and the vulnerability of colon cancer cells to oxaliplatin. Functionally, overexpression of either HSP90 or WEE1 annulled the synergistic effect of GL-V9 and oxaliplatin. Conclusions Collectively, our findings revealed that GL-V9 synergized with oxaliplatin to suppress CRC and displayed a promising strategy to improve the efficacy of oxaliplatin.
Article
Full-text available
Pancreatic cancer is a malignancy with high mortality. In addition to the few symptoms until the disease reaches an advanced stage, the high fatality rate is attributed to its rapid development, drug resistance and lack of appropriate treatment. In the selection and research of therapeutic drugs, gemcitabine is the first-line drug for pancreatic cancer. Solving the problem of gemcitabine resistance in pancreatic cancer will contribute to the progress of pancreatic cancer treatment. Long non coding RNAs (lncRNAs), which are RNA transcripts longer than 200 nucleotides, play vital roles in cellular physiological metabolic activities. Currently, our group and others have found that some lncRNAs are aberrantly expressed in pancreatic cancer cells, which can regulate the process of cancer through autophagy and Wnt/β-catenin pathways simultaneously and affect the sensitivity of cancer cells to therapeutic drugs. This review presents an overview of the recent evidence concerning the node of lncRNA for the cross-talk between autophagy and Wnt/β-catenin signaling in pancreatic cancer, together with the practicability of lncRNAs and the core regulatory factors as targets in therapeutic resistance.
Preprint
Full-text available
Viruses of eukaryotic algae have become an important topic due to their roles in nutrient cycling and top-down control of algal blooms. Omics-based studies have identified a boon of genomic and transcriptional potential among the Nucleocytoviricota , a phylum of large dsDNA viruses which have been shown to infect algal and non-algal eukaryotes. However, little is still understood regarding the infection cycle of these viruses, particularly in how they take over a metabolically active host and convert it into a virocell state. Of particular interest are the roles light and the diel cycle in virocell development. Yet despite such a large proportion of Nucleocytoviricota infecting phototrophs, little work has been done to tie infection dynamics to the presence, and absence, of light. Here, we examine the role of the diel cycle on the physiological and transcriptional state of the pelagophyte Aureococcus anophagefferens while undergoing infection by Kratosvirus quantuckense . Our observations demonstrate how infection by the virus interrupts the diel growth and division of this cell line, and that infection further complicates the system by enhancing export of cell biomass.
Article
Full-text available
Triple-negative breast cancer (TNBC) is an aggressive subtype defined by lack of hormone receptor expression and non-amplified HER2. Adavosertib (AZD1775) is a potent, small-molecule, ATP-competitive inhibitor of the Wee1 kinase that potentiates the activity of many DNA-damaging chemotherapeutics and is currently in clinical development for multiple indications. The purpose of this study was to investigate the combination of AZD1775 and capecitabine/5FU in preclinical TNBC models. TNBC cell lines were treated with AZD1775 and 5FU and cellular proliferation was assessed in real-time using IncuCyte® Live Cell Analysis. Apoptosis was assessed via the Caspase-Glo 3/7 assay system. Western blotting was used to assess changes in expression of downstream effectors. TNBC patient-derived xenograft (PDX) models were treated with AZD1775, capecitabine, or the combination and assessed for tumor growth inhibition. From the initial PDX screen, two of the four TNBC PDX models demonstrated a better response in the combination treatment than either of the single agents. As confirmation, two PDX models were expanded for statistical comparison. Both PDX models demonstrated a significant growth inhibition in the combination versus either of the single agents. (TNBC012, p < 0.05 combo vs. adavosertib or capecitabine, TNBC013, p < 0.01 combo vs. adavosertib or capecitabine.) An enhanced anti-proliferative effect was observed in the adavosertib/5FU combination treatment as measured by live cell analysis. An increase in apoptosis was observed in two of the four cell lines in the combination when compared to single-agent treatment. Treatment with adavosertib as a single agent resulted in a decrease in p-CDC2 in a dose-dependent manner that was also observed in the combination treatment. An increase in γH2AX in two of the four cell lines tested was also observed. No significant changes were observed in Bcl-xL following treatment in any of the cell lines. The combination of adavosertib and capecitabine/5FU demonstrated enhanced combination effects both in vitro and in vivo in preclinical models of TNBC. These results support the clinical investigation of this combination in patients with TNBC, including those with brain metastasis given the CNS penetration of both agents.
Article
Full-text available
WEE1 is a cell cycle and DNA damage response kinase that is emerging as a therapeutic target for cancer. AZD1775 is a small molecule inhibitor of WEE1, currently in early phase clinical trials as a single agent and in combination with more conventional anti-neoplastic agents. As resistance to kinase inhibitors is frequent, we sought to identify mechanisms of resistance to WEE1 inhibition in acute leukemia. We found that AZD1775 resistant cell lines are dependent upon increased HDAC activity for their survival, in part due to increased KDM5A activity. In addition, gene expression analyses demonstrate HDAC dependent increase in MYC expression and c-MYC activity in AZD1775 treated resistant cells. Overexpression of c-MYC confers resistance to AZD1775 in cell lines with low baseline expression. Pharmacologic inhibition of BRD4, and thereby c-MYC, partially abrogated resistance to AZD1775. Thus, acquired resistance to WEE1 inhibition may be reversed by HDAC or BRD4 inhibition in leukemia cells.
Article
Full-text available
To target mechanisms critical for multiple myeloma (MM) plasma cell adaptations to genomic instabilities and further sustain MM cell killing, we here specifically trigger DNA damage response (DDR) in MM cells by a novel BCMA antibody-drug conjugate (ADC) delivering the DNA cross-linking PBD dimer tesirine, MEDI2228. MEDI2228, more effectively than its anti-tubulin MMAF-ADC homolog, induces cytotoxicity against MM cells regardless of drug resistance, BCMA levels, p53 status, and the protection conferred by bone marrow stromal cells and IL-6. Distinctly, prior to apoptosis, MEDI2228 activates DDRs in MM cells via phosphorylation of ATM/ATR kinases, CHK1/2, CDK1/2, and H2AX, associated with expression of DDR-related genes. Significantly, MEDI2228 synergizes with DDR inhibitors (DDRi s) targeting ATM/ATR/WEE1 checkpoints to induce MM cell lethality. Moreover, suboptimal doses of MEDI2228 and bortezomib (btz) synergistically trigger apoptosis of even drug-resistant MM cells partly via modulation of RAD51 and accumulation of impaired DNA. Such combination further induces superior in vivo efficacy than monotherapy via increased nuclear γH2AX-expressing foci, irreversible DNA damages, and tumor cell death, leading to significantly prolonged host survival. These results indicate leveraging MEDI2228 with DDRi s or btz as novel combination strategies, further supporting ongoing clinical development of MEDI2228 in patients with relapsed and refractory MM.
Article
Full-text available
Background Diffuse large B-cell lymphoma (DLBCL) is a heterogeneous disease, characterized by high levels of genomic instability and the activation of DNA damage repair pathways. We previously found high expression of the cell cycle regulator WEE1 in DLBCL cell lines. Here, we investigated the combination of the WEE1 inhibitor, AZD1775, with cyclophosphamide, doxorubicin, vincristine and prednisone (CHOP) and radiation therapy (RT), with the aim of improving first-line treatment. Methods Cell viability experiments were performed to determine synergistic combinations. Levels of DNA damage were established using flow cytometry for γH2AX and protein analysis for DNA damage response proteins CHK1 and CHK2. Flow cytometry analysis for cell cycle and pH3 were performed to determine cell cycle distribution and premature mitotic entry. Results Treatment with either RT or CHOP led to enhanced sensitivity to AZD1775 in several DLBCL cell lines. Treatment of cells with AZD1775 induced unscheduled mitotic progression, resulting in abnormal cell cycle distribution in combination with RT or CHOP treatment. In addition, a significant increase in DNA damage was observed compared with CHOP or RT alone. Of the single CHOP components, doxorubicin showed the strongest effect together with AZD1775, reducing viability and increasing DNA damage. Conclusion In conclusion, the combination of RT or CHOP with AZD1775 enhances sensitivity to WEE1 inhibition through unscheduled G2/M progression, leading to increased DNA damage. Based on these results, WEE1 inhibition has great potential together with other G2/M arresting or DNA damaging (chemo) therapeutic compounds and should be further explored in clinical trials.
Article
Full-text available
Objectives: Many cancer cells depend on G2 checkpoint mechanism regulated by WEE family kinases to maintain genomic integrity. The PKMYT1 gene, as a member of WEE family kinases, participates in G2 checkpoint surveillance and probably links with tumorigenesis, but its role in breast cancer remains largely unclear. Materials and methods: In this study, we used a set of bioinformatic tools to jointly analyse the expression of WEE family kinases and investigate the prognostic value of PKMYT1 in breast cancer. Results: The results indicated that PKMYT1 is the only frequently overexpressed member of WEE family kinases in breast cancer. KM plotter data suggests that abnormally high expression of PKMYT1 predicts poor prognosis, especially for some subtypes, such as luminal A/B and triple-negative (TNBC) types. Moreover, the up-regulation of PKMYT1 was associated with HER2-positive (HER2+), basal-like (Basal-like), TNBC statuses and increased classifications of Scarff, Bloom and Richardson (SBR). Co-expression analysis showed PKMYT1 has a strong positive correlation with Polo-like kinase 1 (PLK1), implying they may cooperate in regulating cancer cell proliferation by synchronizing rapid cell cycle with high quality of genome maintenance. Conclusions: Collectively, this study demonstrates that overexpression of PKMYT1 is always found in breast cancer and predicts unfavourable prognosis, implicating it as an appealing therapeutic target for breast carcinoma.
Article
Full-text available
Mitosis is the process whereby an eukaryotic cell divides into two identical copies. Different multiprotein complexes are involved in the fine regulation of cell division, including the mitotic promoting factor and the anaphase promoting complex. Prolonged mitosis can result in cellular division, cell death, or mitotic slippage, the latter leading to a new interphase without cellular division. Mitotic slippage is one of the causes of genomic instability and has an important therapeutic and clinical impact. It has been widely studied in solid tumors but not in hematological malignancies, in particular, in acute leukemia. We review the literature data available on mitotic regulation, alterations in mitotic proteins occurring in acute leukemia, induction of prolonged mitosis and its consequences, focusing in particular on the balance between cell death and mitotic slippage and on its therapeutic potentials. We also present the most recent preclinical and clinical data on the efficacy of second-generation mitotic drugs (CDK1-Cyclin B1, APC/CCDC20, PLK, Aurora kinase inhibitors). Despite the poor clinical activity showed by these drugs as single agents, they offer a potential therapeutic window for synthetic lethal combinations aimed to selectively target leukemic cells at the right time, thus decreasing the risk of mitotic slippage events.
Article
Objective: The protein kinase, membrane‑associated tyrosine/threonine 1 (PKMYT1) has been implicated as an important factor promoting the tumorigenesis of hepatocellular carcinoma and colorectal cancer. The current study was designed to explore the functional role of PKMYT1 in non-small cell lung cancer (NSCLC) cell behaviors and to investigate the possible molecular mechanisms. Patients and methods: The expression levels of PKMYT1 in NSCLC tissues and cell lines were determined by quantitative Real Time-Polymerase Chain Reaction (qRT-PCR). The clinical and prognostic significance of PKMYT1 in 153 cases of NSCLC was determined. We also evaluated the effects of KMYT1 on NSCLC cell proliferation, migration, and invasion in vitro. Western blot was performed to assure whether PKMYT1 affected the Notch signal pathway and MET pathway. Results: We observed that PKMYT1 expression was significantly up-regulated in both NSCLC tissues and cell lines. Higher expression of PKMYT1 was associated with clinical stage and lymph nodes metastasis. Clinical survival assays demonstrated that patients with high PKMYT1 expression had a shorter overall survival time than those with low PKMYT1 expression. Moreover, the multivariate analysis confirmed that increased expression of PKMYT1 was an independent predictor of overall survival. Functionally, knockdown of PKMYT1 in the NSCLC cell lines A549 and H1299 suppressed NSCLC cells proliferation, invasion and migration, and promoted apoptosis. In addition, the down-regulation of PKMYT1 resulted in the inhibition of EMT in NSCLC cells. Further mechanistic studies revealed that when PKMYT1 was silenced, the expression levels of Notch1, p21, and Hes1 were respectively downregulated, suggesting that PKMYT1 could promote the activity of the Notch signal pathway. Conclusions: PKMYT1 plays a significant role in NSCLC aggressiveness and clinical outcome, and may serve as a promising therapeutic target for this disease.
Article
Epithelial ovarian cancer is characterized by universal TP53 mutations, which result in G1/S checkpoint deficiencies. Therefore, it is hypothesized that the abrogation of the G2/M checkpoint with Wee1 inhibitor might preferentially sensitize TP53-defective ovarian cancer cells. Given the extremely high molecular diversity in ovarian cancer, one approach to improving the clinical efficacy is to identify drug combinations that either broaden the applicable spectrum or circumvent resistance. Here, through a high-throughput unbiased proteomic profiling (RPPA), we found the complementary activated mTOR pathway contributes greatly to Wee1 inhibitor resistance. A combination of Wee1 and mTOR inhibits synergistically inhibiting tumor growth in ovarian cancer cell lines and patient-derived xenograft that closely mimic the heterogeneity of patient tumors. Mechanistically, dual Wee1/mTOR inhibition induced massive DNA replication stress, leading to fork stalling and DNA damage. Moreover, we found that the addition of nucleotide metabolic substrate dNTPs alleviated replication stress, restored the cell cycle and reduced apoptosis to some extent, supporting dNTPs depletion is necessary for the synergy between Wee1 and mTOR inhibits. These results suggest that our study opening up a wider therapeutic window of Wee1 inhibitor for the treatment in epithelial ovarian cancers.
Article
Acute myeloid leukemia (AML) has remained one of the most treatment resistant and deadliest cancers. The survival of AML blast cells is controlled by the balance of anti- and pro-apoptotic proteins. Recently approved Bcl-2 targeted therapy of AML with the Bcl-2 specific inhibitor Venetoclax in combinations has improved patients outcomes. However, a priori and developing resistance to venetoclax combinations with hypomethylating agents (HMA) azacitidine and decitabine challenge this treatment. As such, novel therapies to overcome venetoclax-HMA resistance are urgently needed. We have identified a combination of DNA damage repair interference by WEE1 inhibition with AZD1775, combined with low dose cytarabine (AraC) as an effective strategy to overcome combined venetoclax-azacitidine resistance (VAR). AZD1775 with low dose AraC induced massive apoptosis (by Annexin V and cleaved caspase-3) and almost completely reduced viability and clonogenic growth of primary AML cells. To delineate the molecular mechanism of the synergistic effect of AZD1775/AraC we performed RNAseq analysis of single agent or the combination of AZD1775+AraC in AML cell lines and primary CD34+ selected AML patient cells with the goal to identify deferentially regulated genes indicating a mechanistic underpinning of the potent activity. Only 2 genes were deferentially regulated across cell lines and CD34+ selected cells under AZD1775+AraC treatment: one of these is NR4A1, an orphan nuclear receptor, which we went on to validate as a potential downstream target of Wee1 inhibition. The inactivation of NR4A1 in mice was previously shown to induce AML and to maintain leukemia stem cells. Using qPCR we confirmed that the expression of NR4A1 is upregulated after AZD1775/AraC combo treatment in human leukemic cells. We then demonstrated that activators of NR4A1 (cytosporone B and pPhOCH3) reduce viability of leukemic cells, while NR4A1 inhibitor pPhOH was able to abolish the effect of AZD1775/AraC combo treatment increasing leukemic cell viability]. To investigate the involvement of mitochondria in the effect of AZD1775/AraC treatment we performed the expression of mitochondrial genes and pathway analyses in RNAseq data and found that mitochondrial gene expression, including many genes involved in apoptosis, has most dramatic changes in the combo treatment if compared to the single agents. Subsequently, we have examined the expression of the main BCL-2 family apoptotic genes by qPCR and western blot analysis. We found that AZD1775/AraC induces the expression of Bim isoforms, whereas Bcl-2, Mcl-1 and Bcl-Xl were largely unaffected. NR4A1 was previously shown to translocate to mitochondria, release Bim from Bcl-2 protein binding, as well as convert Bcl-2 to an extreme potent pro-apoptotic form. Finally, we generated several additional VAR cell lines and cells with subclones and demonstrated that AZD1775/AraC combination treatment is able to overcome VAR in almost every clone. Our results show that DNA damage repair interference with Wee1 inhibition has the potential to overcome VAR through a novel mechanisms of AZD1775 increasing NR4A1, freeing pro-apoptotic Bim irrespective of anti-apoptotic Bcl-2 proteins leading to massive apoptotic cell death in AML cells. The precise molecular mechanisms and the involvement of NR4A1 in this phenomenon will be presented at the meeting. Our findings will help to develop new therapeutic strategies in AML treatment and a trial of AZD1775 + AraC in AML is currently ongoing. Disclosures No relevant conflicts of interest to declare.