Skip to main content
Erschienen in: Experimental Hematology & Oncology 1/2024

Open Access 01.12.2024 | Review

Mechanistic insights and the clinical prospects of targeted therapies for glioblastoma: a comprehensive review

verfasst von: Yating Shen, Dexter Kai Hao Thng, Andrea Li Ann Wong, Tan Boon Toh

Erschienen in: Experimental Hematology & Oncology | Ausgabe 1/2024

Abstract

Glioblastoma (GBM) is a fatal brain tumour that is traditionally diagnosed based on histological features. Recent molecular profiling studies have reshaped the World Health Organization approach in the classification of central nervous system tumours to include more pathogenetic hallmarks. These studies have revealed that multiple oncogenic pathways are dysregulated, which contributes to the aggressiveness and resistance of GBM. Such findings have shed light on the molecular vulnerability of GBM and have shifted the disease management paradigm from chemotherapy to targeted therapies. Targeted drugs have been developed to inhibit oncogenic targets in GBM, including receptors involved in the angiogenic axis, the signal transducer and activator of transcription 3 (STAT3), the PI3K/AKT/mTOR signalling pathway, the ubiquitination-proteasome pathway, as well as IDH1/2 pathway. While certain targeted drugs showed promising results in vivo, the translatability of such preclinical achievements in GBM remains a barrier. We also discuss the recent developments and clinical assessments of targeted drugs, as well as the prospects of cell-based therapies and combinatorial therapy as novel ways to target GBM. Targeted treatments have demonstrated preclinical efficacy over chemotherapy as an alternative or adjuvant to the current standard of care for GBM, but their clinical efficacy remains hindered by challenges such as blood-brain barrier penetrance of the drugs. The development of combinatorial targeted therapies is expected to improve therapeutic efficacy and overcome drug resistance.
Hinweise

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Introduction

Glioblastoma (GBM) is the most common and lethal malignancy that makes up about 75% of all adult gliomas [1, 2]. The prognosis for GBM is poor, with a median survival of 12–15 months and a five-year survival rate of 6.9% [3]. The Stupp protocol, which comprises of concomitant radiation and chemotherapy, temozolomide (TMZ) to destroy cancer cells, is used as the first line of therapy for GBM. The Stupp protocol has been the standard of treatment since 2005 after reported to have improved two-year survival from 10.4 to 26.5% as compared to radiation alone [4]. However, long-term treatment success is limited as more than 90% of GBM patients still develop resistance and suffer relapse [5], highlighting the limitations of this standard regimen. This has prompted studies aimed at understanding mechanisms that drive rapid growth and proliferation of GBM, as well as modulation of the tumour microenvironment to support tumour growth, invasion, and resistance. While GBM remains highly lethal and difficult to treat, understanding this mechanism will pave the way for the development of better targeted therapeutics.
Traditionally, gliomas are classified based on histological features. Advancements in molecular profiling have allowed for a more accurate classification based on genetic and epigenetic characteristics. This refinement in patient stratification paves the way for better disease management. A study by Wen and Kesari described pathogenetic hallmarks of primary GBMs including amplified epidermal growth factor receptor (EGFR), loss of heterogeneity of chromosomal 10q, and deletion of phosphatase and tensin homologue on chromosome 10 (PTEN) and p16 [6]. Following the identification of these key genetic aberrations in GBM, Eckel-Passow et al. investigated the significance and association to survival of three markers for defining molecular groups in gliomas, namely chromosome arms 1p and 19q, IDH mutations, and TERT promoter mutations. TERT promoter mutations are the most common (74%) in GBM patients, distinguishing GBM from low grade gliomas. Patients with glioma who solely have TERT mutations have poorer prognosis than those who also have IDH mutations and/or 1p/19q codeletion [7].
Building on the basis of above-mentioned landmark findings, the current version of the WHO Classification of Tumours of the Central Nervous System 5th edition (WHO CNS 5) classifies GBM as diffused astrocytic glioma, characterised by isocitrate dehydrogenase (IDH)-WT with the presence of other key genetic alterations, such as mutations in TERT promoter, amplification of EGFR, and concurrent gain of chromosome 7 and loss of chromosome 10 [8]. Phillips et al. further categorized GBM into three subgroups based on differential expression of particular markers, namely the proneural, proliferative, and mesenchymal subtypes [9]. Following this landmark study, Verhaak et al. performed a more in-depth analysis of genomic data from the TCGA-GBM cohort and classified GBM into four subtypes - Proneural, Neural, Classical and Mesenchymal, of which the neural subtype was later confirmed to be contaminated with normal neural cells and thus removed from the classification [10, 11]. Since then, there has been a surge in interest on the utilization of targeted drugs against genomic aberrations identified in respective subtypes, including EGFR amplification in the classical subtype, NF-kB hyperactivation in the mesenchymal subtype and PDGFR mutation in the proneural subtype.
In this review, we discuss the therapeutic potential and clinical assessment of drugs targeting common oncogenic pathways in GBM. Additionally, we also describe the recent advancement in cell-based therapies against GBM. Finally, we explore the prospects of combinatorial therapies in improving the clinical outcomes for GBM patients.

Targeted therapy

Targeted therapy aims at treating cancers by using drugs or monoclonal antibodies that target specific genes or proteins that are critical for the growth or survival of the cancers. This contrasts with conventional cytotoxic chemotherapy that negatively affects both normal and cancer cells. Targeted therapies minimize the off-target toxicities in normal cells and have gained traction over the past decades. A notable example of clinically-approved targeted therapy is bevacizumab, a vascular endothelial growth factor receptors (VEGFR) inhibitor, approved by the U.S. Food and Drug Administration (FDA) for treatment of progressive recurrent GBM [12]. Following, we examine various therapeutic approaches targeting the angiogenic axis, the JAK/STAT pathway, IDH1/2, and the protein clearance system in GBM (Fig. 1).

Targeting angiogenic axis in GBM

GBM tumours frequently exhibit extensive abnormal vasculature necessary for rapid tumorigenesis. Studies have therefore investigated the role of growth factor receptors involved in angiogenesis, such as the platelet-derived growth factor receptor (PDGFR), vascular endothelial growth factor receptor (VEGFR), and epidermal growth factor receptor (EGFR), contributing to the development of GBM as well as the therapeutic potential of these targets (Fig. 1A).
PDGF signalling is conventionally involved in the growth and differentiation of cells of mesenchymal origin [13]. In embryonic development, PDGFs are known to play a significant role in encouraging development of the central nervous system by supporting oligodendrocyte precursor growth [14]. PDGFR activation has numerous downstream effects including but not limited to activation of Src, SHP-2 tyrosine phosphatase, Phospholipase Cγ, Ras, and various STAT proteins [15] (Fig. 1A). PDGFR activation positively regulates cell proliferation and survival, as well as actin reorganization and cellular migration. Importantly, increased activity of the PDGFR signalling pathway has been observed in high grade gliomas. Glioma cell lines and primary GBM tissues have demonstrated overexpression of both PDGFs and PDGFRs [16]. This is particularly prevalent in the proneural subtype of GBM, which exhibits a high rate of focal PDGFRA amplification (35%), making PDGFRA amplification a specific hallmark of the proneural GBM [11]. Furthermore, expression of PDGFRβ has been observed in the endothelial cells of GBM but not in the vessels of normal human brain, demonstrating its potential as a GBM-specific targeted therapy [17].
To date, a few anti-PDGFR pharmacological agents have been developed and demonstrated efficacy in mitigating GBM tumorigenesis. Anti-PDGFRα antibody, olaratumab, was found to have promising results in GBM xenografts, significantly inhibiting tumour growth in vivo with a concomitant reduction in PDGFR phosphorylation [18, 19]. Alternatively, small molecule inhibitor against both PDGFR⍺ and PDGFRβ, CP-673,451, inhibits tumour growth by inducing terminal differentiation of GBM cells into neural-like cells [20, 21]. Another inhibitor of PDGFR⍺ and PDGFRβ is crenolanib (CP-868,596), which has demonstrated brain penetration as well as in vivo inhibition of PDGFR phosphorylation in murine models and in high grade glioma patients [22, 23]. Avapritinib and ripretinib are PDGFR inhibitors developed and approved for the treatment of PDGFR⍺ mutant gastrointestinal stromal tumour (GIST) [20, 21]. Unlike other PDGFR inhibitors, avapritinib and ripretinib are more potent as they selectively bind to the activation loop of PDGFR⍺ [24]. Among these two inhibitors, avapritinib is a more promising inhibitor for GBM for its high CNS penetrance as compared to ripretinib [22, 23]. Currently, only olaratumab has successfully progressed to clinical trials for GBM (Table 1). Despite its therapeutic potential in pre-clinical models of GBM, olaratumab showed minimal clinical efficacy in GBM patients, with a median overall survival (OS) of 34.3 weeks compared to the VEGFR inhibitor, ramucirumab (49.5 weeks) [25]. Though no trial has been conducted on the use of crenolanib and avapritinib in GBM patients specifically, trials conducted on high-grade glioma patients suggested that crenolanib has limited additional benefits [26], while avapritinib demonstrated more promising response [22, 27]. More clinical data are therefore needed for the aforementioned PDGFR inhibitors in order to assess their efficacy as prospective inhibitors for GBM.
Table 1
Clinical studies of recent targeted therapies for GBM, as single agent or in combination with standard of care. (Data from http://​clinicaltrials.​gov was searched until Nov 2023)
Target
Drug
Adjuvant Therapy
Phase
Patient Group
NCT number
Safety profile1
RANO criteria2
PFS2
OS2
PDGFR
Olaratumab
2
rGBM
NCT00895180
(Completed)
-
Yes
PFS-6: 7.5%
Median: 34.3 weeks
VEGF
Bevacizumab
TMZ, RT
3
ndGBM
NCT05271240
(Recruiting)
-
Yes
Ongoing
Ongoing
Bevacizumab
TMZ, RT
3
rGBM
NCT02761070
(Active, not recruiting)
-
No
Ongoing
Ongoing
Bevacizumab
TMZ
2
GBM
NCT02898012
(Completed)
-
Yes
Median: 15.3 weeks
Median: 23.9 weeks
Vatalanib
TMZ, RT
1/2
ndGBM
NCT00128700
(Completed)
Well tolerated
Yes
Median: 7.2 months
Median: 16.2 months-
EGFR
Erlotinib
1
rGBM
NCT01257594
(Completed)
Unreported
-
-
-
Afatinib
TMZ, RT
1
ndGBM
NCT00977431
(Completed)
MTD: 30 mg with daily TMZ/RT and 40 mg with RT alone
-
-
-
Afatinib
TMZ
2
rGBM
NCT00727506
(Completed)
-
Yes
Median: 0.99 months, 1.53 months with TMZ
PSF-6: 3.0% and 10.3% with TMZ
Unreported
Afatinib
2
GBM
NCT00875433
(Completed)
-
No
Median: 10.6 weeks
Median: 39.6 weeks
Dacomitinib
2
rGBM
NCT01112527
(Completed)
-
No
Unreported
Unreported
Dacomitinib
2
rGBM
NCT01520870
(Completed)
-
Yes
Median: 2.7 months
PFS-6: 10.6%
Median: 7.4 months
Neratinib
2
GBM
NCT02977780
(Recruiting)
-
Yes
Median: 4.7 months
Median: 14.9 months
Cetuximab
RT
2
rGBM
NCT02800486
(Recruiting)
-
No
Ongoing
Ongoing
Cetuximab
1/2
ndGBM
NCT02861898
(Recruiting)
 
Yes
Ongoing
Ongoing
Cetuximab
RT, TMZ
2
ndGBM
NCT01044225
(Terminated, non-drug related)
-
No
Unreported
Unreported
STAT3
WP1066
1
rGBM
NCT01904123
(Completed)
MFD: 8 mg/kg
-
-
-
WP1066
RT
2
ndGBM
NCT05879250
(Not yet recruiting)
-
Yes
Ongoing
Ongoing
PI3K
Buparlisib
RT, TMZ
1
rGBM
NCT01473901
(Completed)
Not tolerable
-
-
-
Buparlisib
2
rGBM
NCT01339052
(Completed)
-
Yes
Median: 1.7 months
PFS-6: 8%
Median: 9.8 months
Sonolisib
2
GBM
NCT01259869
(Completed)
-
Yes
Median:
PFS-6: 17%
Unreported
mTOR
Everolimus
TMZ, RT
1/2
ndGBM
NCT01062399
(Completed)
Combining with TMZ increased toxicity
Yes
Median: 8.2 months
Median: 16.5 months
Sirolimus
1/2
GBM
NCT00047073
(Completed)
Tolerable
No
Unreported
Unreported
Metformin
TMZ
2
ndGBM
NCT04945148
(Not yet recruiting)
-
Yes
Ongoing
Ongoing
Vistusertib
1
rGBM
NCT02619864
(Completed)
Tolerable
-
-
-
AZD8055
1
rGBM
NCT01316809
(Completed)
Unreported
-
-
-
Onatasertib
1/2
Advanced solid tumours
NCT01177397
(Completed)
Tolerable
Yes
PFS-6: 0%
Unreported
RMC-5552
1/1b
rGBM
NCT05557292
(Active, not recruiting)
Ongoing
-
-
-
Sapanisertib
1
rGBM
NCT02133183
(Active, not recruiting)
Unreported
-
-
-
PI3K/mTOR
Paxalisib
RT, TMZ
2
ndGBM
NCT03522298
(Active, not recruiting)
-
Yes
Median: 8.4 months
Median: 15.7%
Voxtalisib
RT, TMZ
1
GBM
NCT00704080
(Completed)
MTD: 90 mg q.d. and 40 mg b.i.d. with TMZ
-
-
-
Proteasome
Bortezomib
TMZ
2
rGBM
NCT03643549
(Recruiting)
-
Yes
Ongoing
Median: 12.7 months for unmethylated, 21.7 months for methylated
Bortezomib
RT, TMZ
2
ndGBM
NCT00998010
(Completed)
-
Yes
Median: 6.2 months
Median: 19.1 months
Marizomib
RT, TMZ
1/2
ndGBM
NCT02903069
(Completed)
Tolerable
Yes
Unreported
Median: 14.8 months
Marizomib
RT, TMZ
3
ndGBM
NCT03345095
(Completed)
-
Yes
Median: 6.34 months
Median: 16.13 months
Disulfiram (+ Copper gluconate)
RT, TMZ
1
GBM
NCT01907165
(Completed)
Safe but can cause reversible neurological toxicities
-
-
-
Disulfiram (+ Copper gluconate)
TMZ
2
rGBM
NCT03034135
(Completed)
-
Yes
Median: 1.7 months
Median: 7.1 months
Disulfiram (+ Copper gluconate)
TMZ
2
Unmethylated GBM
NCT03363659
(Terminated, non-drug related)
-
No
PFS-6: 8.3%
69.2% after 1 year, 15.4% after 2 years
Disulfiram (+ Copper gluconate)
RT, TMZ
1/2
ndGBM
NCT02715609
(Active, not recruiting)
MTD: 375 mg/kg
Yes
Ongoing
Ongoing
PARP
Olaparib
TMZ
1
rGBM
NCT01390571
(Completed)
Tolerable
-
-
-
Olaparib/ Pamiparib
TMZ, RT
1/2
rGBM
NCT04614909
(Recruiting)
Ongoing
No
Ongoing
Ongoing
Olaparib
TMZ, RT
1/2
GBM
NCT03212742
(Recruiting)
Tolerable
No
Ongoing
Ongoing
Olaparib
Early 1
rGBM
NCT05432518
(Recruiting)
Ongoing
-
-
-
Olaparib
2
rGBM
NCT03212274
(Active, not recruiting)
-
Yes
Ongoing
Ongoing
Niraparib
2
rGBM
NCT05297864
(Active, not recruiting)
-
Yes
Ongoing
Ongoing
Niraparib
RT
2
rGBM
NCT04221503
(Active, not recruiting)
-
Yes
Ongoing
Ongoing
Niraparib
RT
2
rGBM
NCT04715620
(Unknown status)
-
No
Median: 157 days
PFS-6: 36%
Median: Ongoing
OS-6: 82.61%
Niraparib
RT
Early 1
ndGBM
NCT05076513
(Recruiting)
Tolerable
-
-
-
Niraparib
TMZ
Early 1
rGBM
NCT01294735
(Completed)
MTD: 40 mg niraparib + 150mg/m2 TMZ
-
-
-
BSI-201
TMZ
1/2
ndGBM
NCT00687765
(Completed)
Unreported
No
Unreported
Unreported
BGB-290
TMZ
1/2
rGBM
NCT03914742
(Completed)
Unreported
Yes
Unreported
Unreported
BGB-290
TMZ
1
GBM
NCT03749187
(Recruiting)
Ongoing
-
-
-
Veliparib
TMZ, RT
2
ndGBM
NCT03581292
(Active, not recruiting)
-
No
Ongoing
Ongoing
Veliparib
TMZ, RT
1/2
ndGBM
NCT01514201
(Completed)
Tolerable
No
PFS-3: 2.9%
OS-3: 5.3%
Fluzoparil
TMZ
2
rGBM
NCT04552977
(Unknown status)
-
No
Unreported
Unreported
NMS-03305293
TMZ
1/2
rGBM
NCT04910022
(Recruiting)
Ongoing
Yes
Ongoing
Ongoing
GBM, glioblastoma; rGBM, recurrent glioblastoma; ndGBM, newly diagnosed glioblastoma; RT, radiotherapy; TMZ, temozolomide; MTD, maximum tolerated dose; RANO, Response assessment in neuro-oncology criteria; PFS-3/6, 3/6 months progression free survival; OS-3/6, 3/6 year overall survival
1Only in Phase 1 trials; 2Only in Phase 2 and Phase 3 trials
VEGF is a key signalling protein that mediates angiogenesis in response to hypoxia. Activation of VEGF-receptor 2 (VEGFR2) upon binding of VEGF ligand induces the activation of oncogenic PI3K/AKT, MAPK and angiopoietin-Tie pathways [28, 29] (Fig. 1A). In high grade gliomas, VEGFs are often highly expressed, leading to disorganization of tumour vasculature and leaky blood-brain barrier (BBB) [28, 29]. Bevacizumab (BEV), a recombinant humanized monoclonal antibody against all isoforms of VEGFs, is the most frequently investigated VEGF inhibitor, with more than 100 clinical trials evaluating its efficacy singly or in combination for both primary and recurrent GBM patients. After promising results in two Phase 2 trials [30], the FDA granted BEV expedited approval in 2009 for use as a single drug in the treatment of recurrent GBM [31]. In the BRAIN trial, which investigated BEV as a single agent and in combination with irinotecan, the reported six-month progression free survival (PFS6) was 42% in single-agent BEV and 50% in BEV plus irinotecan, with OS of 37% and 35% in respective arms [30]. The results indicate that, although both regimens exceeded benchmark records, combinatorial treatment was comparable with single-agent treatment [30]. Additionally, trials of BEV in combination with other treatment modalities in primary GBM also showed disappointing results [3234]. Importantly, subsequent studies pointed out that the favourable progression free survival (PFS) observed after BEV treatment may be attributed to pseudoresponse, which gave rise to the brief improvement in radiologic response rather than true tumour shrinkage [35]. Hence, the potential anti-tumour effect of BEV in GBM warrants further validation.
Subsequently, more studies investigated the efficacy of BEV in combination with other chemotherapies (Table 2). Notably, there has been an surge in clinical trials investigating the value of BEV in combination with immunomodulatory agents, such as PD-1 antagonists. Several studies have demonstrated that anti-angiogenic agents can reprogramme the immunosuppressive tumour microenvironment to an immunosupportive one, potentially sensitizing patients to immunotherapies in combination with antiangiogenic therapies [36]. In a recent case report, BEV in combination with pembrolizumab improved the overall survival (OS) of a GBM patient with extracranial metastases to 27 months, higher than the median OS of 11–17 months [37]. However, the response to these combinations in larger clinical trials exhibited only a modest improvement in PFS6 between cohorts who received BEV monotherapies (6.7%) and BEV in combination with pembrolizumab (26.0%) [38]. This suggests that clinical response to these combination therapies is still largely varied and warrants the identification of predictive biomarkers for patient stratification towards BEV with immunomodulatory agents. Subsequent trials are therefore investigating the potential of BEV with various anti-PD-1 agents while simultaneously altering BEV dosing schedule and performing biomarker correlative studies (Table 2).
Table 2
Clinical studies of recent targeted combinatorial therapies in GBM. (Data from http://​clinicaltrials.​gov was searched until Nov 2023)
Targets
Drug combination
Phase
Patient Group
NCT number
Safety profile1
RANO criteria2
PFS2
OS2
VEGF, PD-1
Bevacizumab + Nivolumab
2
GBM
NCT03452579
(Active, not recruiting)
-
Yes
Unreported
OS-1: 46.2% for standard bevacizumab, 37.7% for low dose bevacizumab
VEGF, PD-1
Bevacizumab + Nivolumab
2
GBM
NCT03743662
(Active, not recruiting)
-
Yes
Ongoing
Ongoing
VEGF, PD-1
Bevacizumab + Camrelizumab
2
rGBM
NCT04952571
(Suspended, non-drug related)
-
No
Suspended
Suspended
VEGF, PD-1
Bevacizumab + Pembrolizumab
2
rGBM
NCT03661723
(Active, not recruited)
-
Yes
Median: 4 months
PFS-6: 10.6%
Median: 8.6 months
OS-6: 56.7%
OS-12: 16.6%
VEGF, PD-1
Bevacizumab + Tislelizumab
2
rGBM
NCT05540275
(Not yet recruiting)
-
Yes
Ongoing
Ongoing
VEGF, PD-1
Bevacizumab + Sintilimab
2
rGBM
NCT05502991
(Not yet recruiting)
-
Yes
Ongoing
Ongoing
VEGF, PD-1
Bevacizumab + Durvalumab
2
rGBM
NCT02336165
(Completed)
-
Yes
Median: 16 weeks
PFS-6: 15.2%
Median: 19.3 weeks
OS-6: 36.4%
VEGF, T-cell
Bevacizumab + GX-17
2
rGBM
NCT05191784
(Active, not recruiting)
-
Yes
Ongoing
Ongoing
VEGF, endoglin
Bevacizumab + TRC105
2
GBM
NCT01564914
(Completed)
-
Yes
Median: 1.81 months
Median: 5.75 months
VEGF, endoglin
Bevacizumab + TRC105
1/2
rGBM
NCT01648348
(Completed)
Tolerable, MTD: 10 mg/kg bevacizumab + 10 mg/kg TRC105
Yes
Median: 2.9 months
PFS-6: 25%
Median: 9.7%
VEGF, PI3K
Bevacizumab + BKM120
1/2
rGBM
NCT01349660
(Completed)
Tolerable at 60 mg/day BKM120
Yes
Median: 2.8 months for prior bevacizumab, 5.3 months for bevacizumab naïve
Median: 6.6 months for prior bevacizumab, 10.8 months for bevacizumab naïve
VEGF, mTOR
Bevacizumab + Everolimus
2
GBM
NCT00805961
(Completed)
-
Yes
Median: 11.3 months
Median: 13.9 months
VEGF, mTOR
Bevacizumab + nab-sirolimus
2
ndGBM
NCT03463265
(Completed)
-
Yes
Median: 3.1 months
PFS-6: 37.5%
PFS-12: 12.5%
Median: 6.8 months
OS-12: 25%
VEGF, mTOR
Bevacizumab + Sapanisertib
1
rGBM
NCT02142803
(Active, not recruiting)
MTD/RP2D: 5 mg/day
-
-
-
VEGF, angiopoietin1/2
Bevacizumab + Trebananib
1/2
rGBM
NCT01290263
(Completed)
Tolerable
Yes
Median: 285 days
PFS-6: 24.3%
Median: 108 days
VEGF, angiopoietin1/2
Bevacizumab + Trebananib
2
rGBM
NCT01609790
(Completed)
-
Yes
Median: 4.2 months
PFS-6: 22.6%
Median: 7.5 months
VEGF, proteasome
Bevacizumab + Marizomib
1/2
GBM
NCT02330562
(Completed)
Tolerable, RP2D: MRZ 0.8 mg/m2
Yes
Median: 3.9 months
Median: 10.4 months
VEGF, CDK4/6
Bevacizumab + Abemaciclib
1
rGBM
NCT04074785
(Active, not recruiting)
Ongoing
-
-
-
VEGF, multi-kinases
Bevacizumab + Ponatinib
2
rGBM
NCT02478164
(Completed)
-
Yes
Median: 28 days
PFS-3: 0%
Median: 98 days
VEGF, FASN
Bevacizumab + ASC40
3
rGBM
NCT05118776
(Recruiting)
-
Yes
Ongoing
Ongoing
VEGF, topoisomerase I
Bevacizumab + Irinotecan
1
rGBM
NCT05201326
(Recruiting)
Ongoing
-
-
-
VEGF, angiogenesis
Bevacizumab + VB-111
3
rGBM
NCT02511405
(Completed)
-
Yes
Median: 3.4 months
Median: 6.8 months
EGFR, glucose
Osimertinib + Fludeoxyglucose F-18
2
rGBM
NCT03732352
(Active, not recruiting)
-
Yes
Ongoing
Ongoing
EGFR, Osmolality
Cetuximab + Mannitol
1/2
GBM
NCT02861898
(Recruiting)
Ongoing
Yes
Ongoing
Ongoing
EGFR, multi-kinases
Erlotinib + Sorafenib
2
rGBM
NCT00445588
(Completed)
-
No
PFS-6: 14%
Median: 5.7 months
PI3K, c-MET
Buparlisib + INC280
1/2
rGBM
NCT01870726
(Terminated, no clear activity)
RP2D not declared due to drug-drug interactions
Yes
Combination arms not initiated
Combination arms not initiated
PI3K
Buparlisib + Carboplatin / Lomustine
1
rGBM
NCT01934361
(Completed)
MTD: 100 mg/day with carboplatin, not determined with lomustine
-
-
-
PI3K
Paxalisib + Metformin + Ketogenic diet
2
rGBM and unmethylated ndGBM
NCT05183204
(Recruiting)
-
Yes
Ongoing
Ongoing
PI3K, PD-1
Pictilisib + MK-3475
1/2
GBM
NCT02430363
(Unknown status)
Unreported
No
Unreported
Unreported
AKT, mTORC1
Perifosine + Temsirolimus
1
rGBM
NCT02238496
(Completed)
Tolerable at 115 mg/week temsirolimus with 100 mg/day perifosine
-
-
-
AKT, PD-L1
Ipatasertib + Atezolizumab
1/2
GBM
NCT03673787
(Unknown status)
RP2D: 400 mg/day ipatasertib with 1200 mg atezolizumab every 3 weeks
No
Unreported
Unreported
mTORC1, CDK4/6
Everolimus + Ribociclib
Early 1
rGBM
NCT03834740
(Completed)
Unreported
Yes
Unreported
Unreported
mTORC1, CDK4/6
Everolimus + Ribociclib
1
Glioma
NCT03355794
(Completed)
Tolerable, RP2D: 120mg/m2 ribociclib with 1.2mg/m2 everolimus
-
-
-
mTORC1, VEGF
Everolimus + Bevacizumab + TMZ
2
GBM
NCT00805961
(Completed)
-
Yes
Median: 11.3 months
Median: 13.9 months
mTORC1, EGFR
Temsirolimus + Erlotinib
1/2
Glioma
NCT00112736
(Completed)
MTD: 15 mg temsirolimus with 150 mg erlotinib
Yes
PFS-6: 13.95%
Unreported
mTORC1, EGFR
Sirolimus + Erlotinib
2
rGBM
NCT00672243
(Completed)
-
Yes
Median: 6.9 weeks
PFS-6: 3.1%
Median: 33.8 weeks
mTORC1, multi-kinases
Everolimus + Gefitinib
1/2
GBM
NCT00085566
(Completed)
Unreported
No
Unreported
Unreported
mTORC1, multi-kinases
Temsirolimus + Sorafenib
1/2
rGBM
NCT00329719
(Completed)
Considerable grade 3 + toxicity
MTD: 200 mg sorafenib twice daily with 20 mg temsirolimus weekly
Yes
Median: 2.6 months for VEGFi-naïve group, 1.9 months for VEGFi-treated group
Median: 17.1% for VEGFi-naïve group, 9.8% for VEGFi-treated group
Median: 6.3 months for VEGFi-naïve group, 3.9 months for VEGFi-treated group
mTORC1, multikinase
Sirolimus + Vandetanib
1
rGBM
NCT00821080
(Completed)
Tolerable, MTD: 200 mg vandetanib with 2 mg sirolimus
-
PFS-6: 15.8%
-
mTORC1, BCR-ABL tyrosine kinase, ribonucleotide reductase
Everolimus + Imatinib + Hydroxyurea
1
rGBM
NCT00613132
(Completed)
Unreported
-
-
-
mTORC1, AKT, 𝛾-secretase
Ridaforolimus + MK-2206 + MK-0752
1
rGBM
NCT01295632
(Completed)
Unreported
-
-
-
mTOR, proteasome
Nab-sirolimus + Marizomib
2
ndGBM
NCT03463265
(Completed)
-
Yes
Median: 1.7 months
PFS-6: 10%
PFS-12: 0%
Median: 6.7 months
OS-12: 0%
Proteasome, VEGF
Bortezomib + Bevacizumab
2
rGBM
NCT00611325
(Completed)
-
Yes
Median: 2 months for EIAED arm, 2.5 months for non-EIAED arm
PFS-6: 25% for EIAED arm, 28.6% for non-EIAED arm
Median: 8 months for EIAED arm, 6 months for non-EIAED arm
Proteasome, VEGF
Bortezomib + Bevacizumab + TMZ
1
rGBM
NCT01435395
(Completed)
Unreported
-
-
-
Proteasome, histone deacetylases
Bortezomib + Vorinostat
2
rGBM
NCT00641706
(Completed)
-
Yes
Median: 1.5 months for no surgery arm, 4.2 months for surgery arm
PFS-6: 0% for no surgery arm, 29% for surgery arm
Median: 3.2 months for no surgery arm, 8.7 months for surgery arm
Proteasome
Disulfiram + Copper + TMZ + Lomustine + PCV
2/3
rGBM
NCT02678975
(Completed)
-
Yes
Median: 2.3 months
Median: 5.5 months
OS-6: 44%
PARP
Talazoparib + Carboplatin
2
rGBM
NCT04740190
(Unknown status)
-
Yes
Unreported
Unreported
NSC
5-FU
1
rGBM
NCT01172964
(Completed)
Tolerable
-
-
-
NSC
5-FU, Leucovorin
1
rGBM
NCT02015819
(Completed)
Tolerable, RP2D: 150 × 106 CD-NSCs
-
-
-
NSC
Irinotecan Hydrochloride
1
rGBM
NCT02192359
(Active, not recruiting)
Unreported
-
-
-
NSC
Adenovirus, TMZ
1
ndGBM
NCT03072134
(Completed)
Tolerable, RP2D: 1.50 × 108 NSCs loading 1.875 × 1011 viral particles
Yes
Median: 9.1 months
Median: 18.4 months
MSC
5-FU
1/2
rGBM
NCT04657315
(Completed)
Tolerable
Yes
Median: >4 months
Median: not reached
MSC
Adenovirus
1
rGBM
NCT03896568
(Recruiting)
Ongoing
-
-
-
GBM, glioblastoma; rGBM, recurrent glioblastoma; ndGBM, newly diagnosed glioblastoma; RT, radiotherapy; TMZ, temozolomide; PCV, combination of procarbazine, lomustine, and vincristine; MTD, maximum tolerated dose; RP2D, recommended Phase 2 dose; RANO criteria, Response assessment in neuro-oncology criteria; PFS-3/6/12, 3/6/12 months progression free survival; OS-1/6/12, 1/3/6 year overall survival; EIAED, enzyme-inducing anti-epileptic drug
1Only in Phase 1 trials; 2Only in Phase 2 and Phase 3 trials
EGFR belongs to the ErbB receptor family of trans-membrane receptor tyrosine kinases. Aberrant EGFR activation has been observed in multiple cancers including gliomas, commonly driven by hyperactivating mutations and gene amplification. EGFR amplification leads to high protein expression of EGFR that activates a multitude of signalling cascades (including activation of PI3K/AKT, RAS/MAPK, and JAK/STAT pathways) contributing to tumorigenesis and progression [39, 40] (Fig. 1A). EGFR amplification is detected in more than 57% of primary GBM patients [41], with a predilection in the classical GBM subtype and only infrequently in other subtypes [11]. Besides amplification of EGFR and mutant ligands such as constitutively active EGFRvIII, EGFR can also be amplified in extra-chromosomal DNA (ecDNA) as amplicons that are capable of random integration into chromosomes and unequal segregation to daughter cells [42]. As such, ecDNA harbouring EGFRvIII would result in uncontrolled increase in oncogene copy number and intratumoral heterogeneity, preventing adequate drug targeting. There are several clinical trials exploring anti-EGFR strategy in GBM, such as dacomitinib, gefitinib, erlotinib, neratinib, and cetuximab (Table 1and Table 2). However, anti-EGFR targeted therapies have been less successful than expected due to acquired drug resistance and tumour heterogeneity, with only marginal increase in clinical benefits as monotherapy for GBM patients [4348].
While activation of compensatory oncogenic signalling pathways such as the PI3K and MET pathways contributes to resistance against EGFR inhibitors in GBM, a significant limitation in current EGFR inhibitors is the specificity of its mechanism of action [49]. EGFR mutations commonly found in GBM are predominantly in the extracellular domain, limiting the efficacy of current first- and second-generation EGFR inhibitors such as dacomitinib, erlotinib and afatinib [40, 47, 50]. Though third-generation EGFR inhibitors, such as osimertinib, have been developed with improved binding to specific mutant EGFR, acquired mutations (C797S, G724S, L718Q) hinder the binding of inhibitors to EGFR and continue to limit the efficacy of third-generation EGFR inhibitors [51]. Unfortunately, even with the development of potent EGFR inhibitors, the oncogenic function of EGFR may still be retained in GBM cells through EGFR-PDGFRA receptor heterodimerization, suggesting the need for combinatorial treatment to effectively target EGFR activity [51].
To address the above-mentioned challenges of using EGFR inhibitors in GBM, novel therapeutics have been developed. Among which is ABT-414, an antibody-drug conjugate (ADC) that preferentially binds to overexpressed EGFR or EGFRvIII, thus conferring selective cytotoxic effect of monomethyl auristatin F to EGFR-amplified cells independent of EGFR signalling [52]. Despite exhibiting promising efficacy against both wildtype EGFR and EGFRvIII in cell lines and patient-derived xenografts [53], ABT-414 did not confer survival benefits in newly diagnosed or recurrent GBM patients in the INTELLANCE 2 trial [54, 55]. This is owing to the preferential loss of EGFR-amplification in resistant clones, which allows them to escape ABT-414 binding [56]. Another explanation is that ADCs are inefficient in penetrating the BBB into large tumours such as GBM [57, 58], which again highlights BBB penetrance as one of the biggest challenges in targeting GBM. In addition to harbouring EGFR mutations, cancer cells may also evade treatment by constitutively activating downstream effectors in an EGFR-independent manner. One such resistance mechanism in GBM is KRAS-driven hyperactivation of the MAPK signalling pathway, which can be facilitated by DDR1 overexpression [59]. To overcome KRAS-driven resistance to EGFR inhibitors, co-inhibition of EGFR and DDR1/BCR-ABL has previously demonstrated synergistic efficacy in retarding cell growth and inducing apoptosis in tumouroids of patient-derived recurrent GBM [60]. Overall, while targeting EGFR-driven tumorigenesis is a potentially effective therapeutic strategy, the unmet need for inhibitors with greater specificity towards GBM-associated EGFR mutations limits their current clinical actionability in GBM patients, and they should be administered in combination with other therapeutic agents to combat GBM resistance.

Targeting signal transducer and activator of transcription 3 (STAT3) in GBM

Signal Transducer and Activator of Transcription 3 (STAT3) is another potential target of therapy for GBM. Acting downstream of multiple kinases and growth factor receptors, including but not limited to PDGFR, EGFR and IL-6, STAT3 is an important mediator of gliomagenesis through its role in modulating cancer cell survival, invasiveness, and immune evasion (Fig. 1B).
STAT3 is a transcription factor which modulates transcription of numerous genes involved in cell-cycle regulation and anti-cell death activity of the JAK/STAT3 pathway [61]. STAT3 is found to be constitutively active in 90% of human GBM tumours [62]. Persistent STAT3 activation occurs when there is hyperactivation in the upstream signalling cascade or defective downstream regulation, thus resulting in upregulation of several major oncogenic signalling pathways and contributing to tumorigenesis of multiple cancers, including GBM [63]. In GBM, STAT3 is reported to directly promote cell survival by enhancing expression of Bcl-2-like protein 1, driving the inhibition of cell death and promoting tumour cell proliferation [64]. Constitutively active STAT3 confers resistance to apoptosis by enhancing transcription of anti-apoptotic regulators including Bcl-2, Bcl-XL and Mcl-1 (Fig. 1B) [62], whereas inhibition of STAT3 selectively induced apoptosis in WP1066-treated GBM cells by downregulating expression of anti-apoptotic genes and restoring BAX activity [65]. Furthermore, oncogenic STAT3 also confers resistance to autophagy by suppressing pro-autophagic pathways including Bcl-2/Beclin-1 and AMP-activated protein kinase ⍺ (AMPK⍺)/Unc-51-like kinase 1 (ULK1) signalling in GBM cells [66, 67]. Given the pivotal role of STAT3, targeting STAT3-dependent apoptosis and autophagy might be a promising strategy for sensitizing GBM cells to therapy-induced cell deaths.
STAT3 is also implicated in promoting cellular differentiation, namely in assisting the epithelial-mesenchymal transition (EMT) of radioresistant GBM by upregulating the expression of EMT markers such as MMPs, Rho and Rac [68, 69] (Fig. 1B). This promotes the migratory and invasive properties of glioma stem cells (GSCs), and are maintained via STAT3-mediated upregulation of the Notch pathway [70]. Importantly, activation of STAT3 signalling has been found to induce a switch from the less aggressive proneural to the more aggressive mesenchymal tumour subtype associated with chemoradiotherapy-resistance and recurrence in GBM [71]. Recently, it has been shown that depleting insulin-like growth factor binding protein 2 (IGFBP2) can lead to the sensitisation of STAT3-low expressing cells to STAT3 inhibitors, suggesting that targeting both the STAT3 and IGF-1R/IGFBP2 signalling axis is a promising therapeutic strategy for GBM [72]. To further contribute to the invasiveness of GSCs, constitutively active STAT3 can also enhance VEGF-mediated angiogenesis by promoting VEGF expression (Fig. 1B), facilitating neovascularisation and intracranial extension in GBM [7375].
Therapeutics have thus been developed to target STAT3 as single agent or in combinations, and their potential use as anti-neoplastic drugs have been explored in preclinical settings. One class of inhibitors directly binds to and impedes STAT3 function. For instance, STA-21 targets the SH2 domain of STAT3, preventing STAT3 dimerization and suppressing stem cell properties in GSCs [76, 77]. Similar inhibitors that have been tested in preclinical studies include inhibitors of STAT3 phosphorylation, LLL12, an STA-21 analogue, LLL3, and an inhibitor of STAT3 dimerization, STX-0119 [7880]. However, all these inhibitors have only demonstrated anti-cancer properties in preclinical models of GBM and have yet to advance into clinical studies.
Another therapeutic approach is to block STAT3 activation by attenuating the upstream signalling pathway in the form of JAK inhibitors. AG490 is a JAK2 inhibitor that has demonstrated efficacy in mitigating STAT3 activity through downregulation of STAT-regulated genes MMP2 and MMP9 in GBM cell lines, albeit exhibiting limited anticancer effect in vivo [62, 65, 81, 82]. WP1066, a more potent second-generation analogue of AG490, demonstrated compelling in vivo anti-tumour effect against GBM as well as in patient-derived GSCs [82, 83]. Following promising preclinical results, a Phase 1 trial was conducted to determine the safety of WP1066 in patients with recurrent GBM [84] (Table 1). However, given that subjects in this Phase 1 trial were heavily pre-treated recurrent GBM patients, WP1066 monotherapy did not significantly improve patients’ PFS as patients may have developed several mechanisms of general drug resistance [84]. Moving forward, the potential therapeutic effect of WP1066 and radiotherapy in treating newly diagnosed GBM patients will be evaluated in a planned Phase 2 trial, at the maximum tolerated dose/maximum feasible dose of 8 mg/kg identified in the Phase 1 trial [84].
A more recent approach to target STAT3 is through the use of oligonucleotide therapeutics. This approach includes antisense oligonucleotides (ASO) such as AZD9150, GQ-ODN and decoy oligonucleotide [8591]. Though oligonucleotide therapeutics have achieved success in other cancer types and proceeded on to clinical trials (NCT01839604) [85, 86, 89], BBB penetration and drug delivery still remain a challenge for brain tumours such as GBM. Thus, while nucleic acid therapeutics show promise in targeting STAT3 in cancer, clinical inhibition of STAT3 in GBM is still dependent on the development of effective pharmacological agents which can cross the BBB.

Targeting the PI3K/AKT/mTOR signalling pathway in GBM

Given its involvement in tumour development and progression, the phosphoinositide 3-kinase (PI3K)/AKT/mammalian target of rapamycin (mTOR) pathway has emerged as a promising therapeutic target in GBM. The PI3K/AKT/mTOR signalling network is known to be activated in nearly 90% of GBM patients [11], making it a potentially beneficial therapeutic target. This pathway is crucial for cell survival, proliferation, and angiogenesis, all of which contribute to the aggressive nature of GBM. By targeting this pathway, it is possible to disrupt the aberrant signalling cascades that drive tumour growth and improve treatment outcomes.
Several mechanisms contribute to the hyperactivation of the PI3K/AKT/mTOR pathway in GBM. As discussed in the previous section, upstream receptor amplification, such as EGFR and PDGFR, is a typical mechanism leading to PI3K/AKT/mTOR pathway hyperactivation in GBM. These amplifications enhance ligand binding and, as a result, activate the PI3K/Akt/mTOR signalling cascade.
Phosphatase and Tensin Homolog (PTEN), which is located on the q arm of chromosome 10 (10q), functions as a negative regulator of the PI3K/Akt/mTOR signalling pathway (Fig. 1C), and is deleted in 90% of primary GBM due to a loss of heterozygosity of chromosome 10 [11, 92] [11, 92]. PTEN function can also be lost due to homozygous and hemizygous deletion of the gene [11], as well as poor stability of the mutant protein [93]. PTEN deficiency or dysfunction results in sustained stimulation of PI3K/AKT/mTOR signalling in either instances, resulting in poorer prognosis in PTEN-deficient GBM patients [94, 95].
In addition to PTEN loss, mutations in the PI3K complex can contribute to PI3K/AKT/mTOR pathway hyperactivation in GBM. Activating somatic mutations in phosphatidylinositol-4,5-bisphosphate 3-kinase catalytic subunit alpha (PIK3CA) or phosphoinositide-3-kinase regulatory subunit 1 (PIK3R1) genes are frequent alterations that disrupt the conformation of the p110⍺ catalytic subunit, resulting in constitutive PI3K activation [96, 97]. To target hyperactivation of PI3K, pan-PI3K inhibitors, isoform-selective PI3K inhibitors and dual PI3K/mTOR inhibitors have been developed. Alpelisib, a p110⍺-selective inhibitor licensed by the FDA for the treatment of PIK3CA-mutated breast cancer, has demonstrated preferential inhibition of proneural GSCs [98]. Thus far, only a few pan-PI3K inhibitors (buparlisib, pilaralisib and sonolisib) and dual PI3K/mTOR inhibitors (paxalisib, dactolisib, voxtalisib, PQR309) have been evaluated in clinical trials for the treatment of GBM (Table 1and Table 2), among which, buparlisib is the most commonly studied. Despite demonstrating strong efficacy in vitro and greater BBB permeability, such positive results did not translate to clinical efficacy when buparlisib was investigated as a single agent and in combination with the standard radio-chemotherapy [99101]. Paxalisib, a dual PI3K/mTOR inhibitor specifically developed for the treatment of GBM, has demonstrated favourable safety profile and promising efficacy as a first-line treatment in a Phase 2 trial, and is being further investigated in patients with newly diagnosed or recurrent GBM as a part of the AGILE GBM trial [102, 103]. However, recent preclinical studies have revealed that tumour cells in GBM and other cancers may circumvent PI3K inhibition by inducing insulin feedback as a resistance mechanism to reactivate PI3K-mTOR signalling, suggesting that PI3K inhibitors may need to be coupled with anti-hyperglycemic therapies such as metformin to increase treatment efficacy [104, 105]. Correspondingly, a clinical trial is therefore underway to assess the clinical value of this strategy by combining paxalisib with metformin and a ketogenic diet (NCT05183204) (Table 2).
While less common and less characterized, hyperactivation of AKT and mTOR can occur downstream of the pathway in GBM without significant mutations [106]. Upstream signalling components, such as RTKs and PI3K can be dysregulated, resulting in increased activation of AKT or mTOR without necessity for direct mutations in these genes. As a critical central regulator of multiple oncogenic signals, aberrant activity of mTOR and downstream effectors including S6K and 4EBP1 is significantly higher in GBM in comparison to low grade glioma (Fig. 1C), implying that AKT and mTOR effector activity may be emerging as novel prognostic markers of glioma malignancy [107, 108].
The only AKT-specific inhibitor that has been tested in GBM patients is the allosteric inhibitor, perifosine. However, like most other investigational drugs, efficacy of perifosine was not observed in GBM patients [109]. In recent years, more AKT inhibitors are developed and tested pre-clinically as candidate drugs for GBM therapy. MK2206, a new allosteric inhibitor of AKT, was found to potentially sensitize GBM spheroids to TMZ treatment and radiotherapy, warranting further investigations into its clinical prospects for GBM patients [110].
On the other hand, as downstream effectors of AKT, mTOR complexes are more attractive as therapeutic targets for clinical investigation. Among mTOR inhibitors, the most commonly investigated drugs in GBM are sirolimus and its analogues, including everolimus and temsirolimus, which are collectively known as rapalogues (Fig. 1C). However, they have shown limited efficacy in clinical trials as single agents or in combination with the current standard of care (Table 1). Sirolimus only exhibited anti-GBM effects in PTEN-deficient GBM patients as a single therapeutic agent, and has demonstrated limited additional benefit when combined with EGFR inhibitor, erlotinib [111, 112]. Notably, rapalogues preferentially inhibit mTORC1 as opposed to mTORC2 [113]. On the contrary, vistusertib, an inhibitor of both mTORC complexes, demonstrated therapeutic effects in sensitizing stem-like GBM cells to radiation both in vitro and in vivo [114]. Such promising preclinical results have encouraged the conduct of an ongoing Phase 1 trial in recurrent GBM (NCT02619864). Another dual mTORC1/2 inhibitor, onatasertib has demonstrated potential antitumour activity in patients with advanced solid tumours including GBM [115]. AZD8055, dual mTORC1/2 inhibitor, additionally warranted further clinical investigation (NCT01316809) after demonstrating the induction of autophagy and autophagy-regulated Notch1 degradation in GBM cell lines [116]. Importantly, AZD8055 also demonstrated synergistic inhibitory effect and improved survival with TMZ in orthotropic xenografts [117]. These collectively suggest that dual inhibition of both mTOR complexes is essential in targeting the PI3K/AKT/mTOR axis in GBM as opposed to only targeting mTORC1 which could result in compensatory activation of mTORC2 and hence limited clinical efficacy.
Studies have additionally investigated how GBM cells can circumvent therapy resistance induced by the PI3K/AKT/mTOR pathway. Specifically, Vehlow and team found that AKT increased pro-survival signalling in GBM by associating with DDR1, 14-3-3 and Beclin-1 in a complex [118]. Inhibition of DDR1 in the complex suppressed the pro-survival AKT and mTOR signalling pathway and resensitized GBM cells to radio- and chemotherapy by activating autophagy [118]. Concurrent attenuation of the PI3K and EGFR signalling axis through DDR1 and EGFR inhibitors respectively may also revert KRAS-induced hyperactivation in recurrent GBM [60]. These studies highlight DDR1 as a promising and upcoming target in inhibiting the PI3K axis in GBM. While DDR1 inhibition is a novel therapeutic strategy against recurrent GBM in vitro, further validation of its in vivo and clinical efficacy is warranted, and the development of more DDR1-specific inhibitors holds promise as targeted therapies against GBM.

Targeting IDH1/2-associated vulnerabilities in GBM

IDH1/2 mutation status is a prognostic marker used to differentiate between astrocytoma and GBM as IDH mutations are associated with more optimistic prognosis. Patients harbouring IDH-mutant gliomas exhibit better survival than IDH wildtype gliomas in GBM (31 months vs. 15 months) and anaplastic astrocytoma (65 months vs. 20 months) [119]. Despite its association with lower-grade gliomas, IDH mutations are often observed in secondary GBM (73%) as well [120], suggesting that lower-grade IDH-mutant gliomas are prone to malignant progression and recurrence as higher-grade gliomas [121]. Targeting vulnerabilities in IDH-mutant gliomas is thus a viable strategy to mitigate the risks of recurrence in patients.
IDH1/2 mutation is associated with the induction of multiple mechanisms that drive gliomagenesis. IDH1 plays a major role in metabolic pathways by metabolizing isocitrate to produce alpha-ketoglutarate (⍺-KG). IDH mutation in glioma often occurs at arginine residues that are responsible for isocitrate binding (R132 for IDH1, R140 or R172 for IDH2) [122]. Heterozygous missense mutations, which are the major forms of IDH1 mutations observed in IDH-mutant gliomas, generate mutant IDH1 that converts isocitrate into the oncometabolite D-2-hydroxyglutarate (D-2-HG) in place of ⍺-KG [123]. Accumulation of D-2-HG in glioma cells triggers multiple aberrant cellular processes, including epigenetic modifications which contribute to metastasis progression through induction of oncogenes activation and silencing of tumour suppressor genes [124, 125].
A key epigenetic characteristic of IDH-mutant glioma is the presence of hypermethylation. Hypermethylation in IDH-mutant GBM cells is mostly mediated by DNA methylation and histone methylation. Due to structural similarity, oncometabolite D-2-HG competitively inhibits ⍺-KG-dependent ten-eleven translocation (TET) demethylase, decreasing the functional activity of TET demethylase which is responsible for reversing DNA methylation [126128] (Fig. 1D). Consequently, DNA methylation may be extended in glioma cells to establish glioma CpG island methylator phenotype (G-CIMP) [129], and potentially leads to irreversible epigenetic alterations which commit glioma cells to oncogenesis, such as the silencing of tumour suppressive microRNA, miR-148a [130], as well as activation of oncogene, PDGFRA [131]. Furthermore, D-2-HG inhibition of TET demethylase in IDH-mutant glioma cells has also been shown to maintain stemness. Inhibition of TET demethylase in IDH1-mutant astrocytes resulted in upregulation of stem cell marker, Nestin [129], whereas restoration of TET2 expression in GBM cells upregulated genes crucial for neural differentiation, such as brain fatty acid-binding protein (BFABP) and Mash1 [132]. Taken together, D-2-HG-mediated hypermethylation in IDH-mutants contributes to gliomagenesis by triggering downstream epigenetic alterations of oncogenes and impairments in cellular differentiation.
One clinical significance of IDH-mutant-induced histone methylation is the silencing of MGMT, a gene involved in DNA damage repair. Methylation at two key differentially methylated regions (DMRs) of the MGMT promoter, DMR1 and DMR2, has been identified as contributing factors of MGMT silencing in GBM patients [133, 134], which accounts for approximately 45% of patients [135]. While the mechanism of methylation at DMR2 has yet to be elucidated, the positive association between IDH1 mutation and high MGMT promoter methylation has been shown in xenografts and GBM patients [136, 137]. This is further supported by the better therapeutic effects of TMZ for IDH1-WT glioma patients with MGMT promoter methylation over IDH1-mutants [138]. Given that the G-CIMP is highly enriched in IDH1-mutant proneural GBM [139], as well as the positive association between MGMT methylation and G-CIMP in GBM tumours [41], the role of IDH1 mutations in MGMT silencing may be related to IDH-mutant-induced histone methylation. These findings show that IDH1 mutation and high MGMT methylation may offer prognostic value as paired predictive biomarkers.
In IDH-mutant GBM cells, the oncometabolite D-2-HG interferes with the homologous recombination (HR) pathway by masking H3K9 trimethylation signal through hypermethylation, preventing the recruitment of homology-dependent repair factors to the site of double strand break (DSB), including ATM and histone demethylase KDM4B [140, 141]. This results in a state of ‘BRCAness’, where the IDH-mutant GBM cells exhibit a phenotype similar to BRCA-mutant cancer cells with impaired HR pathway. Similarly, IDH-mutant glioma cells demonstrate synthetic lethality with PARP inhibition, and are hence vulnerable to PARP inhibitors [140, 141]. (Fig. 1D). By inhibiting the DNA repair pathway, PARP-inhibitors enhance cytotoxicity of chemotherapy by inducing excessive accumulation of DNA damage which results in elevated apoptosis observed in PARP-inhibitor treated IDH1 mutant GBM cells [142].
Currently, clinically investigated PARP inhibitors in GBM include olaparib, niraparib, BSI-201, BGB-290, veliparib, fluozoparil, and NMS-03305293 (Table 1and Table 2). Olaparib was granted FDA approval for the treatment of BRCA-mutated advanced breast, ovarian and pancreatic cancer, and was recently investigated in the application for GBM. In the Phase 1 OPARATIC trial, olaparib was detected at radiosensitizing concentrations in all recurrent GBM tumour specimens, demonstrating its ability to cross the BBB [143]. Promising therapeutic profile observed encouraged ongoing trials to further investigate the clinical potential of olaparib in IDH1/2-mutant GBM [144] as well as tumour protein P53 (TP53) mutant GBM (NCT05432518). While olaparib did not meet the pre-specified response-based threshold to proceed with a Phase 3 trial in IDH-mutant glioma patients, a subset of patients demonstrated prolonged stable disease, suggesting that olaparib may still be clinically useful as a form of maintenance therapy for a subset of GBM patients [144]. An alternative PARP inhibitor, niraparib was found to exhibit better tumour exposure and sustainability than olaparib, and was tolerable when administered in combination with TMZ, though the two drugs demonstrated no synergy [145, 146]. Previous studies have also shown that PARP inhibitors exhibit radiosensitizing effects by inhibiting the base excision repair pathway [147150]. The combination of niraparib and radiotherapy is thus being evaluated in three ongoing studies (NCT04221503, NCT04715620, NCT05076513) (Table 1). In the Phase 1 trial for GBM patients, the combination demonstrated promising efficacy and favourable safety characteristics [151].
However, tumour-suppressing effect of PARP inhibitor is low in IDH1/2-wildtype gliomas. In IDH1/2-wildtype cells, functional BRCA1/2 can be recruited to the site of DNA damage to carry out HR and allow cells to bypass the damaged site [152]. To confirm this finding, an ongoing clinal study is investigating the efficacy of the combination of PARP inhibitor NMS-03305293 and TMZ in IDH-wildtype GBM (NCT04910022). While the trial is currently ongoing, the results would provide justification for stratifying patients according to the mutation status of IDH when utilizing PARP inhibition as a therapeutic strategy for GBM.
Given that the neomorphic activity of IDH promotes oncogenesis in IDH-mutant GBM cells through the production of oncometabolite D-2-HG, IDH inhibitors such as ivosidenib and vorasidenib have also been developed to block the downstream production of D-2-HG. Notably, IDH inhibitors have shown promising clinical efficacy in only IDH-mutant low-grade glioma (LGG) patients. Therapeutic benefits of vorasidenib and ivosidenib in treatment IDH-mutant LGGs have been demonstrated in multiple trials, especially in the INDIGO trial, in which vorasidenib significantly improved PFS in LGG patients [153156]. Importantly, promising results from the INDIGO trial has granted vorasidenib priority review by the FDA. On the other hand, the only reported clinical case of an IDH-mutant GBM patient treated with an IDH inhibitor is a recurrent GBM patient treated with ivosidenib in a Phase 1 clinical trial [153, 157]. Given the high proportion of IDH-mutated diffuse astrocytoma, CNS WHO grade 4, in secondary GBM, it would be advantageous to potentially extend the IDH inhibitor trial cohort to GBM patients so as to evaluate the therapeutic value of IDH inhibitors in secondary GBM.

Targeting protein clearance in GBM: proteasomes

Proteasomes are responsible for the degradation of unwanted or damaged intracellular proteins, as well as the regulation of proteins that are involved in cell cycle and apoptosis. Proteasomes carry out the degradation of proteins through protein ubiquitination and proteolysis, together with other components in the ubiquitination-proteasome pathway (UPP) [158]. Oncogenic addiction to high proteasome levels has been observed in a multiple malignancies as an adaptation to high protein homeostasis in rapidly proliferating cancer cells [159]. UPP is also exploited in cancer cells to downregulate tumour-suppressor proteins such as p21, p27 and p53, as well as to activate oncogenic targets including NF-𝝹B, essentially contributing to carcinogenesis [160, 161]. The selective potency of proteasome inhibition in cancer cells supports the dependency on proteasome functions in cancer cells, implying that proteasome inhibitors may be used as a targeted therapeutic agent [162]. Inhibition of the UPP in GBM cells results in a plethora of biological effects, including the stimulation of oxidative stress, ER stress, receptor-mediated cell death and cell cycle arrest [163168] (Fig. 2). This knowledge has led to bortezomib being the first proteasome inhibitor to be approved for the treatment of multiple myeloma [169]. The success of bortezomib has spurred many studies to investigate the potential application of proteasome inhibition in other malignancies, including GBM.
Most proteasome inhibitors target the β5 catalytic subunit of the 20S core particle (CP) of the 26S proteasome, which is responsible for chymotrypsin-like proteolytic activities (Fig. 2). Among these proteasome inhibitors, bortezomib is the most investigated drug in clinical studies. In one study, bortezomib was found to be more beneficial for MGMT-methylated patients compared to unmethylated patients in terms of PFS (24.7 months vs. 5.1 months) and overall survival (OS) (61 months vs. 16.4 months) [170]. However, promising results from this study were contradicted by other in vivo studies, in which bortezomib exhibited limited BBB penetration, and thus had limited efficacy in animal models [171]. Consequently, an alternative 20S proteasome inhibitor with greater BBB penetrance, marizomib, was developed [171, 172]. Marizomib exerts antitumour effects in GBM by inducing caspase 9-dependent apoptosis [173], and it could potentially achieve synergistic effect with TNF-related apoptosis-inducing ligand (TRAIL) receptor agonist [174]. Despite having promising pre-clinical results and passing two clinical trials, the efficacy of marizomib in the Phase 3 trial was disappointing, though it could be attributed to the lack of patient stratification for MGMT promoter methylation status in the trial [175177] (Table 1and Table 2).
Disulfiram is another proteasome inhibitor that has potential therapeutic effect in treating GBM. Disulfiram is an FDA-approved acetaldehyde dehydrogenase inhibitor originally approved for the treatment of alcoholism. It was later discovered that disulfiram can additionally inhibit chymotrypsin-like activity by forming a proteasomal-inhibitory complex with tumour cellular copper, initiating apoptosis in copper-rich cancer cells [178, 179] (Fig. 2). Further functional analysis indicated that disulfiram inhibits protein turnover in GBM cells by targeting the p97/NPL4 pathway, which is essential for the processing of ubiquitylated proteins [180182]. Clinical trials are therefore underway to investigate the efficacy of disulfiram together with copper gluconate to recreate the copper-rich environment necessary for the inhibition of the chymotrypsin-like proteases in GBM patients (Table 1and Table 2). While treatment with disulfiram, copper gluconate and TMZ exhibited manageable safety profiles and preliminary clinical benefits, combination therapy including other alkylating agents such as lomustine induced severe adverse events in patients, limiting the use of disulfiram in combination with alternative standard of care for GBM patients apart from TMZ [183186].

Targeting fusion genes in GBM

Recent advances in sequencing technologies, such as fluorescence in situ hybridization and Next-Generation Sequencing, have paved the way for robust characterization of the genomic landscape in GBM, resulting in the discovery of novel oncogenic fusion genes. Genomic and molecular studies have discovered many fusion genes in GBM, including fibroblast growth factor receptor (FGFR) fusions, anaplastic lymphoma kinase (ALK) fusions, EGFR fusions, and neurotrophic tyrosine receptor kinase (NTRK) fusions [187191]. FGFR fusions are the most common and well-studied fusion gene in GBM, accounting for up to 8.33% of GBM patients [190192]. EGFR fusions are the second most common fusion in GBM (4%) [188]. ALK fusions are more prevalent in paediatric GBMs, with only 1.9% found in adult GBMs [193]. NTRK fusion on the other hand are relatively more uncommon in GBM (1.2%) [193]. To date, several clinical cases have reported effective application of inhibitors for treatment of gliomas harboring fusion genes, including lorlatinib for the treatment of SPECC1L-ALK fusion harbouring paediatric high-grade glioma and larotrectinib against EML4-NTRK3 positive recurrent GBM [194, 195]. As FGFR-TACC fusions, specifically FGFR3-TACC3, are the most common gene fusions reported in GBM, the discussions in this section will be focused on the therapeutic value of targeting FGFR3-TACC3 in GBM.
FGFR3-TACC3 originates from tandem duplication of the FGFR3 and TACC3 genes on 4p16.3 [191]. The coiled-coil domain at the C-terminus of TACC3 in the oncogenic chimeric protein facilitates kinase transphosphorylation and localization of FGFR3-TACC3 to the mitotic spindle, where it disrupts chromosomal segregation, resulting in chromosome instability (CIN) and aneuploidy [189]. Physiologically, expression of FGFR3 is negatively regulated by the binding of miR-99a to the 3’-UTR of FGFR3 transcripts [191]. This regulation is absent in FGFR3-TACC3 positive GBM cells due to truncation of the 3’-UTR-containing C-terminal in the fusion transcript [191]. Without posttranscriptional regulation by miR-99a, FGFR3-TACC3 fusion is overexpressed in GBM cells, resulting in production of the hyperactive chimeric oncoprotein. Interestingly, the fusion genes and either IDH1/2 mutations or EGFR amplification were found to be mutually exclusive [190, 191]. This finding has refined the selection criteria for multiple clinical trials investigating therapeutic effects of FGFR inhibitors in GBM patients.
To date, a handful of FGFR inhibitors are being investigated in clinical studies to target FGFR-TACC-positive GBM. Erdafitinib (JNJ-42,756,493), a pan-FGFR selective inhibitor, has demonstrated survival benefits in mice bearing FGFR-TACC gliomas [189]. Clinically, erdafitinib showed antitumour activity in recurrent GBM in two Phase 1 studies [189, 190, 196]. Hence, a Phase 2 trial studying erdafitinib on IDH-wild type gliomas with FGFR-TACC gene fusion is now ongoing (NCT05859334). Another FGFR inhibitors, pemigatinib, has demonstrated promising results in the FIGHT-207 trial in solid tumours including GBM, and is now being applied in the GBM-focused FIGHT-209 trial [197, 198]. Infigratinib is also a promising FGFR inhibitor that achieved partial response or stable disease in 34.6% of recurrent GBM patients [199]. Numerous other FGFR inhibitors that have been investigated in GBM are fexagratinib, ponatinib, anlotinib and futibatinib, reflecting the clinical potential of FGFR inhibitors in GBM [200203]. Taken together, targeting oncogenic fusion genes in GBM provides a personalized treatment avenue that holds great promise, especially for the subset of gliomas with druggable kinase fusion, despite their low occurrence rate.

Emerging strategies to target cancer cells in GBM

Apart from targeted pharmacological agents, alternate modalities for targeting GBM tumours have been investigated. This includes the use of oncolytic viruses which have been shown to exhibit antineoplastic properties. Recently, emerging technologies have been developed to treat the malignancy, such as leveraging on cell-based therapies, trafficking of immune cells to the tumour, and using pulsed ultrasound to transiently disrupt the BBB, thereby enhancing the delivery of targeted therapies which would otherwise be unable to penetrate the BBB [204, 205]. This section focuses on the developments and emergence of alternative treatment modalities to specifically target the receptors presented on cancer cells in GBM tumours. Specifically, progress in neural and mesenchymal stem cell therapy, chimeric antigen receptor (CAR) T-cell therapy, and oncolytic virotherapy will be discussed (Fig. 3).

Stem cell therapy in GBM: neural stem cells (NSC) / mesenchymal stem cells (MSC)

The discovery of neural stem cells (NSCs) and mesenchymal stem cells (MSCs) has led to the development of novel cell-based therapies to overcome the challenge of BBB penetration and achieve better management of GBM. NSCs are specific groups of multipotent stem cells found in the brain, with the potential to self-renew and differentiate into astrocytes, neurons, and oligodendrocytes [206]. Advancement in somatic cell reprogramming has further allowed NSCs to be generated through transdifferentiation, allowing the conversion of human skin fibroblast into “human-induced Neural Stem Cells (hi-NSCs)” [207, 208] This technique has further been applied in fibroblasts derived from GBM patients, permitting the establishment of patient-specific hi-NSCs [209]. On the other hand, MSCs are multipotent stem cells responsible for the generation of differentiated cells of the mesenchymal lineage and can be found in multiple tissue types including the bone marrow, adipose, muscle and umbilical cord [210].
Apart from their ability to cross the BBB, NSCs and MSCs are attractive treatment modalities due to their unique glioma-tropic migration capabilities [211213]. Both NSCs and MSCs express cell surface markers and secrete cytokines that facilitate the chemotactic migration through normal tissues and preferentially home to the tumour (Fig. 3A). Specifically, NSCs and MSCs express higher levels of chemokine receptors CXCR1, CXCR2, CXCR4 and CCR2 to facilitate the chemotactic migration of the cells to GBM tumours exhibiting elevated levels of IL-8, SDF-1 and MCP-1 [214216].
hi-NSCs and MSCs can be manipulated to achieve specific aims, including but not limited to immunomodulation and drug delivery. For instance, hi-NSCs have been engineered to maintain an anti-tumour microenvironment by secreting inflammatory mediators including IL-7, IL-12 and IL-23, thereby recruiting immune cells to inhibit the growth of gliomas (Fig. 3A) [211, 217, 218]. Several studies have also evaluated the therapeutic effect of TRAIL-producing NSCs in targeting GBM tumours. TRAIL-producing hi-NSCs induced cell death selectively in GBM cells via caspase-mediated apoptosis, successfully mitigating GBM development and extended median survival in human GBM xenografts (Fig. 3A) [212, 219221].
Given their ability to penetrate the BBB, NSCs and MSCs can also serve as vectors in enzyme/prodrug-based and oncolytic virus-based drug delivery systems (Fig. 3A). The enzyme/prodrug-based system can be achieved by modifying the NSCs/MSCs to produce enzymes that convert prodrugs into their active therapeutic forms. An example of such a system is the FDA-approved cytosine deaminase (CD)-expressing clonal human NSC line, HB1.F3.CD, engineered to home to gliomas and convert prodrug 5-fluorocytosine (5-FC) to the active chemotherapeutic 5-fluorouracil (5-FU) [222]. Clinical assessment of this system demonstrated specific NSC and 5-FU localization in the brain tumour and was well tolerated by patients [223]. The safety profile of 5-FU-releasing NSCs in combination with leucovorin in high-grade gliomas was further established, and its efficacy is currently under investigation [224]. Carboxylesterase-releasing NSCs are also currently in clinical trial to investigate their ability to sensitize recurrent high-grade gliomas to irinotecan hydrochloride by converting the prodrug irinotecan into its active metabolite, SN-38 (NCT02192359) (Table 2). Development and clinical assessments of similar enzyme/prodrug-based systems in MSCs are also currently underway (NCT04657315) [225].
The application of NSCs and MSCs as carriers has also recently been extended into the delivering of adenovirus to overcome shortcomings in virotherapy such as the activation of host immunological response and poor biodistribution [226]. The most notable applications are NSCs loaded with CRAd-S-pk7 (NSC-CRAd-S-pk7), a glioma-restricted oncolytic adenovirus, which could improve viral biodistribution in mice brains, enhance inhibition of GBM tumour growth and improve median survival by 50% compared to viral treatment alone [227229]. The NSC-CRAd-S-pk7 system exhibited favourable safety profiles in 12 newly diagnosed malignant glioma patients, with a median progression-free survival of 9.1 months and OS of 18.4 months [230]. Synergy between NSC-CRAd-S-pk7 with standard of care treatments was also observed in in vivo models of GBM, increasing survival by approximately 46% [228]. The safety profile of this approach was also determined, with a reported clearance of NSC and viral components from various regions of patients’ brains 4–24 months after treatment, suggesting that multiple administration of the NSCs is tolerable in patients whilst simultaneously achieving greater therapeutic benefit [230]. Following the promising development of NSC-CRAd-S-pk7, MSCs loaded with oncolytic adenoviruses have since been engineered and clinical trials are currently undergoing to evaluate their value as prospective treatment modalities against GBM (NCT03896568, NCT04758533). Taken together, the evidence has suggested that hi-NSCs and MSCs are promising and novel therapeutic delivery vectors that provide specific drug localization, improving clinical response in GBM patients.

Chimeric antigen receptor (CAR) T-cell therapy in GBM

Recently, the use of chimeric antigen receptor (CAR) T-cells as a form of cell-based immunotherapy and treatment strategy for GBM patients has been gaining interest. To develop CAR T-cells, chimeric antigen receptors, comprising of extracellular domains which recognise specific epitopes frequently presented on the tumours and intracellular domains necessary for T-cell activation, are engineered into isolated T-cells. This allows for the CAR T-cells to home specifically to the tumour sites to exert their cytotoxic effects, bypassing the need for MHC antigen presentation.
To date, several CAR T-cells targeting various antigens frequently reported in GBM tumours have been generated and are being evaluated as viable treatment modalities in the clinics. Notably, the most common tumour-associated antigens are EGFRvIII and interleukin 13-receptor α 2 (IL-13Rα2) (Fig. 3B). However, clinical trials of CAR T-cells targeting either EGFRvIII or IL-13Rα2 while tolerated in patients, did not present durable response in patients [231233]. In particular, only one patient demonstrated a complete response to IL-13Rα2-targeting CAR T-cell therapy before tumour recurrence [234]. Hence, there have been recent identification of alternative target antigens for which CAR T-cells have been developed against, including HER2 and B7-H3 (Fig. 3B) [235238]. Interestingly, recent preclinical investigations have identified the natural compound, chlorotoxin, as a novel antigen which exhibits preferential binding to GBM tumours compared to normal brain tissues, mediated by the expression of MMP-2 on tumour cells [239]. Chlorotoxin-directed CAR T-cells have demonstrated effective and specific targeting of GBM cancer cells in vivo, culminating in a Phase 1 trial (NCT05627323) [239].
However, while the development of GBM-targeting CAR T-cells have progressed on to the clinical phase, there still remains several challenges which limit the durability of patient response to these therapies. Hence, recent efforts have been directed at engineering novel CAR T-cell modalities which can overcome these issues. To overcome tumour escape and enhance tumour specificity, bivalent and trivalent CAR T-cells co-targeting two or three tumour-associated antigens have exhibited greater tumour coverage, mitigating tumour growth more efficaciously in vivo [240, 241]. Importantly, the development of IL-8 receptor-modified CD70 CAR T-cells led to greater tumour trafficking and persistence, resulting in a Phase 1 clinical trial for newly diagnosed GBM patients (NCT05353530) [242]. Interestingly, engineering bispecific CAR T-cells presenting antibodies which combine the binding domains of two antigens, such as IL-13Rα2 with either HER2, increased the antitumour effects of the T-cells [243]. Furthermore, they were shown to be superior to bivalent CAR T-cells, suggesting that these advanced bispecific CAR T-cells may hold more promise in treating GBM patients in the clinics [243]. Designing novel SynNotch-CAR T-cells which can exhibit spatially controlled activation has significantly improved the specificity and durability of CAR T-cell therapy [244, 245]. In GBM specifically, Choe et al. engineered SynNotch-CAR T-cells against EGFRvIII or myelin oligodendrocyte glycoprotein (MOG) as priming antigens, ensuring that the T cells localised to EGFRvIII- or MOG-expressing GBM tumours [244]. Subsequent expression of the IL-13Rα2/EphA2 CAR upon binding to the priming antigens induced spatially controlled and tumour-specific T-cell cytotoxicity in patient-derived in vivo models of GBM [244]. Finally, CAR T-cell therapies can be combined with immunotherapies to prevent exhaustion of T-cells, ensuring that the antitumour effects are sustained and durable. Following promising preclinical results, two clinical trials have been established to investigate the therapeutic value of co-delivering immune checkpoint inhibitors with EGFRvIII-targeting and IL-13Rα2-targeting CAR T-cells in GBM patients (NCT03726515, NCT04003649) [246]. Unfortunately however, combining EGFRvIII-targeting CAR T-cells with pembrolizumab did not confer clinical efficacy in patients, suggesting that greater efforts have to be focused on enhancing CAR T-cell therapy for GBM therapy [247].

Oncolytic virotherapy in GBM

Oncolytic virotherapy is a strategy commonly used to target GBM tumours. Oncolytic viruses typically targets tumour cells through two complementary mechanisms. Firstly, they directly induce oncolytic lysis of the tumour cells following infection, and the new virus particles proceed to infect and lyse neighbouring target cells (Fig. 3C). Secondly, in the process of viral infection and tumour cell lysis, several cytokines, viral pathogen-associated molecular patterns (PAMPs) and disease-associated molecular patterns (DAMPs) are released into the tumour microenvironment. This release promotes immune cell infiltration and activation, mounting an immune response against the tumours(Fig. 3C). To date, numerous oncolytic virotherapies have been engineered from various strains of viruses and are being trialled in GBM patients, including commonly used herpes simplex virus type 1 (HSV-1) and adenoviruses, as well as polio-rhinovirus chimeras, parvoviruses, reoviruses (NCT00528684) and Newcastle disease virus [230, 248255].
To further improve the efficacy of viral-based therapies against GBM, recent advances have been made in engineering viral strains with more complex systems. For instance, a HSV-based oncolytic virus, CAN-3110 (formerly designated rQNestin34.5v.2), was engineered to express the viral gene, ICP34.5, to promote viral replication and oncolysis of the tumour cells specifically by placing the gene under the transcriptional control of a nestin promoter [256258]. This ensured that expression of ICP34.5 was spatially confined to specifically GBM tumours which exhibit high expression of nestin and not in normal brain tissues where nestin is not expressed [259]. Given the increased specificity and enhanced potency of the virus strain, CAN-3110 is currently in a Phase 1 clinical trial (NCT03152318) [254].
Additionally, with the advent of immunotherapy in oncology, viral strains which promote immune cell trafficking and infiltration into GBM tumours have demonstrated promise in several preclinical studies. Oncolytic viruses expressing proinflammatory cytokines and antigens, such as IL-12 (HSV-1 M032) and CXCL11 (oAd-CXCL11), as well as viral particles expressing antibodies against immunosuppressive receptors including CD47 (OV-αCD47-G1) were able to elicit a strong tumour immune response and enhance the therapeutic efficacy of CAR T-cells in vivo [260263]. Promising preclinical evidence have thus led to the assessment of HSV-1 M032, as well as IL-12 and anti-PD-1 co-expressing MVR-C5252 in clinical trials (NCT05095441) [260]. Notably, by designing viral particles with anti-EGFR cetuximab and CCL5 chimeric receptors (OV-Cmab-CCL5), Tian and colleagues were additionally able to promote the specific infiltration of various immune cells to EGFR-positive GBM tumours [264]. Collectively, these studies demonstrated the value of engineering novel viral particles to further improve the efficacy and specificity of oncolytic virotherapy in GBM.

Combination therapies in GBM

Due to occurrence of several clonal subpopulations inside a single tumour, GBM cells frequently undergo clonal evolution in response to therapies, acquiring mutations that were not present at the time of diagnosis [265]. As a result, recurrent GBM cells frequently develop resistance to administered therapies. Furthermore, the complexity of the signalling pathway network in GBM cells also contributes to treatment resistance through a variety of processes, including acquisition of gain-of-function mutations and activation of compensatory oncogenic pathways. Therefore, combination therapeutic strategies targeting several molecular pathways serve as a potential approach to overcome drug resistance in GBM.
A common strategy for developing combination therapies in GBM involves the concurrent inhibition of the VEGF signalling pathway and a secondary oncogenic pathway triggered in response to VEGFR inhibition. Studies have reported upregulation of the TGF-β–CD105–Smad pathway as an alternative angiogenic pathway that contributes to BEV resistance in GBM [266268]. This suggests that targeting the ligand, endoglin (CD105), may be a potential therapeutic approach. However, neither clinical trials evaluating the combination of BEV and anti-endoglin agent, TRC105, received positive responses in patients [269, 270]. Concurrent inhibition of VEGFR and the PI3K/AKT/mTOR pathway is an alternative strategy as previous studies have shown that the aberrant activation of PI3K/AKT/mTOR signalling stimulates excessive secretion of VEGF [271, 272]. However, such combinations did not exhibit promising clinical value for GBM patients. In a Phase 1/2 study of BEV with PI3K inhibitor, buparlisib, the combination was poorly tolerated at low doses of buparlisib, and it failed the trial with unsatisfactory PFS of only 4 months and an overall response rate (ORR) of 26% [273]. Similarly, simultaneous inhibition of VEGFR and mTOR via BEV and everolimus did not yield promising results, with no significant improvement in patients’ PFS and OS [274]. Ongoing trials are thus underway to evaluate the efficacy of BEV in combination with new generation mTOR inhibitors such as nab-sirolimusand sapanisertib, which are predicted to yield more promising results [275, 276]. More clinical trials are ongoing to evaluate the efficacy of BEV with other potential therapeutic agents including the emerging PD-1 inhibitors, angiopoietin1/2 inhibitors, proteasome inhibitors, and multi-kinase inhibitors (Table 2).
The PI3K/AKT/mTOR is another crucial signalling pathway that is often hyperactive and complexed with multiple feedback loops. In one study, MK-2206 suppression of p-AKT was observed to cause aberrant activation of mTOR and radiation resistance in PTEN-deficient GBM cells, implying that the efficacy of AKT inhibitors may be limited by the negative feedback loop that increases mTORC1 activity in GBM cells [277]. This suggests that dual-inhibition of AKT and mTOR may be a more effective approach in targeting GBM cells. This has led to ongoing clinical trials investigating the efficacy of perifosine in combination with temsirolimus in recurrent malignant gliomas following patient tolerance in a Phase 1 trial [278] (Table 2). Combination therapies with mTOR inhibitors and other drugs are also undergoing investigations. The combination of everolimus and CDK4/6 inhibitor, ribociclib, was tested in two clinical trials, both suggesting that the drug combination was well-tolerated and may achieve therapeutic effects [279, 280]. Multiple trials have also looked into the potency of mTOR inhibitors in combination with multi-kinases inhibitors, although initial assessment with first generation kinase inhibitors did not yield clinical benefit for patients. Everolimus in combination with gefitinib has failed a Phase 1/2 trial with unsatisfactory antitumour activity [281], while the combination of temsirolimus and sorafenib received considerable grade 3 + toxicities in a Phase 1/2 trial [282]. On the contrary, a Phase 1 trial which evaluated the combination of sirolimus and vandetanib has demonstrated that the two drugs can be co-administered safely, suggesting that only specific pairs of mTOR inhibitors and multi-kinase inhibitors would present satisfactory safety profiles in patients [283].
The therapeutic effects of proteasome inhibitors have been investigated in many clinical trials as a single agent. Based on strong preclinical evidence of disulfiram as a therapeutic agent in combination with TMZ [284286], the safety profile of disulfiram in combination with Stupp protocol and copper was investigated in two clinical trials [184, 287] (Table 1). In contrary to preclinical results, though the drug combination was found to be well-tolerated in both newly diagnosed and recurrent GBM patients, it conferred minimal proteasome inhibition, and no overall benefits over the control group [183, 287]. The efficacy of this drug combination is being further investigated in two ongoing trials [184, 288] (Table 1). Previous preclinical study has also demonstrated that inhibition of proteasomes could induce HIF1α and VEGF production in malignant GSCs, thus suggesting that co-inhibition of the proteasome and VEGF could achieve better tumour inhibition [289]. However, the efficacy of BEV in combination with either bortezomib or marizomib did not yield meaningful benefits in two independent trials (NCT00611325) [176].
Despite much effort to develop combination therapies for GBM patients, there is still no FDA-approved combination targeted therapy for GBM patients, with few generating promising results. There is thus a need to develop and identify novel combinations which may have therapeutic potential in GBM. In addition to the complex network of feedback mechanisms that drive drug resistance and compensatory pathways in GBM, the incomplete knowledge in underlying molecular interactions makes it challenging to identify novel optimal drug combinations from a pool of potential drug candidates. To overcome the challenges in conventional drug combination design, various models and algorithms have been developed to predict potential synergistic drug combinations using smaller datasets. One such platform is the quadratic phenotypic optimization platform (QPOP) which identifies the best drug combination for each patient via second-order linear regression analysis without prior knowledge of the molecular mechanism of the drugs [290]. Other computational techniques designed to predict drug synergism include the Feedback System Control (FSC) [291], Markov chain-based models [292] and the drug combination network (DCN) [293]. Interestingly, recent efforts to identify novel drug combinations in GBM have been successful in identifying novel drug combinations and repurposing FDA-approved drugs for the inhibition of GBM. The computational platform SynergySeq integrates transcriptional data with perturbagen-induced transcriptional signatures to identify the novel synergistic combination of BRD4 inhibitor, JQ1, and aurora kinase inhibitor, alisertib, in mitigating GBM growth in vivo [294]. The study additionally identified the combination of FDA-approved gemcitabine and imatinib for the treatment of GBM, offering novel combination therapies for GBM, which may otherwise not be investigated [294]. While such combinations require further validation and clinical assessment, this study supports the use of computational tools to identify promising combination therapies for GBM.
Although combination therapy is a promising approach to treat GBM, there are many challenges in finding and testing novel drug combinations. Apart from the potential toxicity that arises from the use of several drugs, many clinical trials may have failed due to the lack of patient stratification. As tumour heterogeneity is one of the most important hallmarks of GBM, it is expected that GBM patients display varied drug sensitivity, and their responses to the same targeted treatment is bound to be diverse. It is therefore imperative to adopt selection strategies such as molecular-based selection or pathway-based stratification in clinical trials to identify specific patient cohorts who can attain clinical benefit from the respective targeted therapies [295, 296]. Furthermore, with advancements in imaging techniques, imaging-based biomarkers can potentially be developed to investigate target engagement and treatment response as a strategy to stratify patient sensitivity [297300].

Conclusion

In conclusion, the identification of key genetic mutations and their roles in oncogenesis in GBM has paved the way for robust research and drug discoveries in the field of targeted therapies and has presented a positive outlook in improving clinical benefit for GBM patients. However much still has to be done to significantly improve patient response. Given the heterogeneous nature of GBM, a future challenge is the prioritization of target combinations to overcome therapy resistance arising from cross-talks between various signalling pathways. Additionally, endeavours to stratify patients according to their molecular characteristics will greatly improve the identification of patient cohorts who exhibit greater sensitivity to corresponding targeted therapies and combinations. Future efforts to develop therapeutic strategies concurrent with the incorporation of specific molecular and imaging biomarkers will significantly improve the treatment outcome of GBM patients in the clinics.

Acknowledgements

All figures were created on BioRender.com.

Declarations

Competing interests

The authors declare no competing interests.
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.

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Literatur
1.
Zurück zum Zitat Ostrom QT, Patil N, Cioffi G, Waite K, Kruchko C, Barnholtz-Sloan JS. CBTRUS Statistical Report: primary brain and other Central Nervous System tumors diagnosed in the United States in 2013–2017. Neuro Oncol. 2020;22(12 Suppl 2):iv1–96.PubMedPubMedCentralCrossRef Ostrom QT, Patil N, Cioffi G, Waite K, Kruchko C, Barnholtz-Sloan JS. CBTRUS Statistical Report: primary brain and other Central Nervous System tumors diagnosed in the United States in 2013–2017. Neuro Oncol. 2020;22(12 Suppl 2):iv1–96.PubMedPubMedCentralCrossRef
2.
3.
Zurück zum Zitat Ostrom QT, Price M, Neff C, Cioffi G, Waite KA, Kruchko C, et al. CBTRUS Statistical Report: primary brain and other Central Nervous System tumors diagnosed in the United States in 2015–2019. Neuro Oncol. 2022;24(Suppl 5):v1–95.PubMedPubMedCentralCrossRef Ostrom QT, Price M, Neff C, Cioffi G, Waite KA, Kruchko C, et al. CBTRUS Statistical Report: primary brain and other Central Nervous System tumors diagnosed in the United States in 2015–2019. Neuro Oncol. 2022;24(Suppl 5):v1–95.PubMedPubMedCentralCrossRef
4.
Zurück zum Zitat Stupp R, Mason WP, van den Bent MJ, Weller M, Fisher B, Taphoorn MJB, et al. Radiotherapy plus concomitant and adjuvant temozolomide for glioblastoma. N Engl J Med. 2005;352(10):987–96.PubMedCrossRef Stupp R, Mason WP, van den Bent MJ, Weller M, Fisher B, Taphoorn MJB, et al. Radiotherapy plus concomitant and adjuvant temozolomide for glioblastoma. N Engl J Med. 2005;352(10):987–96.PubMedCrossRef
5.
Zurück zum Zitat Chen W, Wang Y, Zhao B, Liu P, Liu L, Wang Y, et al. Optimal therapies for recurrent glioblastoma: a bayesian network Meta-analysis. Front Oncol. 2021;11:641878.PubMedPubMedCentralCrossRef Chen W, Wang Y, Zhao B, Liu P, Liu L, Wang Y, et al. Optimal therapies for recurrent glioblastoma: a bayesian network Meta-analysis. Front Oncol. 2021;11:641878.PubMedPubMedCentralCrossRef
7.
Zurück zum Zitat Eckel-Passow JE, Lachance DH, Molinaro AM, Walsh KM, Decker PA, Sicotte H, et al. Glioma groups based on 1p/19q, IDH, and TERT promoter mutations in tumors. N Engl J Med. 2015;372(26):2499–508.PubMedPubMedCentralCrossRef Eckel-Passow JE, Lachance DH, Molinaro AM, Walsh KM, Decker PA, Sicotte H, et al. Glioma groups based on 1p/19q, IDH, and TERT promoter mutations in tumors. N Engl J Med. 2015;372(26):2499–508.PubMedPubMedCentralCrossRef
8.
Zurück zum Zitat Louis DN, Perry A, Wesseling P, Brat DJ, Cree IA, Figarella-Branger D, et al. The 2021 WHO classification of tumors of the Central Nervous System: a summary. Neuro Oncol. 2021;23(8):1231–51.PubMedPubMedCentralCrossRef Louis DN, Perry A, Wesseling P, Brat DJ, Cree IA, Figarella-Branger D, et al. The 2021 WHO classification of tumors of the Central Nervous System: a summary. Neuro Oncol. 2021;23(8):1231–51.PubMedPubMedCentralCrossRef
9.
Zurück zum Zitat Phillips HS, Kharbanda S, Chen R, Forrest WF, Soriano RH, Wu TD, et al. Molecular subclasses of high-grade glioma predict prognosis, delineate a pattern of disease progression, and resemble stages in neurogenesis. Cancer Cell. 2006;9(3):157–73.PubMedCrossRef Phillips HS, Kharbanda S, Chen R, Forrest WF, Soriano RH, Wu TD, et al. Molecular subclasses of high-grade glioma predict prognosis, delineate a pattern of disease progression, and resemble stages in neurogenesis. Cancer Cell. 2006;9(3):157–73.PubMedCrossRef
10.
Zurück zum Zitat Wang Q, Hu B, Hu X, Kim H, Squatrito M, Scarpace L, et al. Tumor evolution of Glioma-Intrinsic Gene Expression Subtypes Associates with immunological changes in the Microenvironment. Cancer Cell. 2017;32(1):42–56. e6.PubMedPubMedCentralCrossRef Wang Q, Hu B, Hu X, Kim H, Squatrito M, Scarpace L, et al. Tumor evolution of Glioma-Intrinsic Gene Expression Subtypes Associates with immunological changes in the Microenvironment. Cancer Cell. 2017;32(1):42–56. e6.PubMedPubMedCentralCrossRef
11.
Zurück zum Zitat Verhaak RG, Hoadley KA, Purdom E, Wang V, Qi Y, Wilkerson MD, et al. Integrated genomic analysis identifies clinically relevant subtypes of glioblastoma characterized by abnormalities in PDGFRA, IDH1, EGFR, and NF1. Cancer Cell. 2010;17(1):98–110.PubMedPubMedCentralCrossRef Verhaak RG, Hoadley KA, Purdom E, Wang V, Qi Y, Wilkerson MD, et al. Integrated genomic analysis identifies clinically relevant subtypes of glioblastoma characterized by abnormalities in PDGFRA, IDH1, EGFR, and NF1. Cancer Cell. 2010;17(1):98–110.PubMedPubMedCentralCrossRef
12.
Zurück zum Zitat Cohen MH, Shen YL, Keegan P, Pazdur R. FDA drug approval summary: bevacizumab (avastin) as treatment of recurrent glioblastoma multiforme. Oncologist. 2009;14(11):1131–8.PubMedCrossRef Cohen MH, Shen YL, Keegan P, Pazdur R. FDA drug approval summary: bevacizumab (avastin) as treatment of recurrent glioblastoma multiforme. Oncologist. 2009;14(11):1131–8.PubMedCrossRef
14.
Zurück zum Zitat Ng F, Boucher S, Koh S, Sastry KS, Chase L, Lakshmipathy U, et al. PDGF, TGF-beta, and FGF signaling is important for differentiation and growth of mesenchymal stem cells (MSCs): transcriptional profiling can identify markers and signaling pathways important in differentiation of MSCs into adipogenic, chondrogenic, and osteogenic lineages. Blood. 2008;112(2):295–307.PubMedCrossRef Ng F, Boucher S, Koh S, Sastry KS, Chase L, Lakshmipathy U, et al. PDGF, TGF-beta, and FGF signaling is important for differentiation and growth of mesenchymal stem cells (MSCs): transcriptional profiling can identify markers and signaling pathways important in differentiation of MSCs into adipogenic, chondrogenic, and osteogenic lineages. Blood. 2008;112(2):295–307.PubMedCrossRef
15.
Zurück zum Zitat Funa K, Sasahara M. The roles of PDGF in Development and during neurogenesis in the normal and diseased nervous system. J Neuroimmune Pharmacol. 2014;9(2):168–81.PubMedCrossRef Funa K, Sasahara M. The roles of PDGF in Development and during neurogenesis in the normal and diseased nervous system. J Neuroimmune Pharmacol. 2014;9(2):168–81.PubMedCrossRef
16.
Zurück zum Zitat Lokker NA, Sullivan CM, Hollenbach SJ, Israel MA, Giese NA. Platelet-derived growth factor (PDGF) autocrine signaling regulates survival and mitogenic pathways in glioblastoma cells: evidence that the novel PDGF-C and PDGF-D ligands may play a role in the development of brain tumors. Cancer Res. 2002;62(13):3729–35.PubMed Lokker NA, Sullivan CM, Hollenbach SJ, Israel MA, Giese NA. Platelet-derived growth factor (PDGF) autocrine signaling regulates survival and mitogenic pathways in glioblastoma cells: evidence that the novel PDGF-C and PDGF-D ligands may play a role in the development of brain tumors. Cancer Res. 2002;62(13):3729–35.PubMed
17.
Zurück zum Zitat Plate KH, Breier G, Weich HA, Risau W. Vascular endothelial growth factor is a potential tumour angiogenesis factor in human gliomas in vivo. Nature. 1992;359(6398):845–8.PubMedCrossRef Plate KH, Breier G, Weich HA, Risau W. Vascular endothelial growth factor is a potential tumour angiogenesis factor in human gliomas in vivo. Nature. 1992;359(6398):845–8.PubMedCrossRef
18.
Zurück zum Zitat Lane R, Cilibrasi C, Chen J, Shah K, Messuti E, Mazarakis NK, et al. PDGF-R inhibition induces glioblastoma cell differentiation via DUSP1/p38(MAPK) signalling. Oncogene. 2022;41(19):2749–63.PubMedPubMedCentralCrossRef Lane R, Cilibrasi C, Chen J, Shah K, Messuti E, Mazarakis NK, et al. PDGF-R inhibition induces glioblastoma cell differentiation via DUSP1/p38(MAPK) signalling. Oncogene. 2022;41(19):2749–63.PubMedPubMedCentralCrossRef
19.
Zurück zum Zitat Roberts WG, Whalen PM, Soderstrom E, Moraski G, Lyssikatos JP, Wang HF, et al. Antiangiogenic and antitumor activity of a selective PDGFR tyrosine kinase inhibitor, CP-673,451. Cancer Res. 2005;65(3):957–66.PubMedCrossRef Roberts WG, Whalen PM, Soderstrom E, Moraski G, Lyssikatos JP, Wang HF, et al. Antiangiogenic and antitumor activity of a selective PDGFR tyrosine kinase inhibitor, CP-673,451. Cancer Res. 2005;65(3):957–66.PubMedCrossRef
20.
Zurück zum Zitat Blay J-Y, Serrano C, Heinrich MC, Zalcberg J, Bauer S, Gelderblom H, et al. Ripretinib in patients with advanced gastrointestinal stromal tumours (INVICTUS): a double-blind, randomised, placebo-controlled, phase 3 trial. Lancet Oncol. 2020;21(7):923–34.PubMedPubMedCentralCrossRef Blay J-Y, Serrano C, Heinrich MC, Zalcberg J, Bauer S, Gelderblom H, et al. Ripretinib in patients with advanced gastrointestinal stromal tumours (INVICTUS): a double-blind, randomised, placebo-controlled, phase 3 trial. Lancet Oncol. 2020;21(7):923–34.PubMedPubMedCentralCrossRef
21.
Zurück zum Zitat Jones RL, Serrano C, von Mehren M, George S, Heinrich MC, Kang Y-K, et al. Avapritinib in unresectable or metastatic PDGFRA D842V-mutant gastrointestinal stromal tumours: long-term efficacy and safety data from the NAVIGATOR phase I trial. Eur J Cancer. 2021;145:132–42.PubMedPubMedCentralCrossRef Jones RL, Serrano C, von Mehren M, George S, Heinrich MC, Kang Y-K, et al. Avapritinib in unresectable or metastatic PDGFRA D842V-mutant gastrointestinal stromal tumours: long-term efficacy and safety data from the NAVIGATOR phase I trial. Eur J Cancer. 2021;145:132–42.PubMedPubMedCentralCrossRef
22.
Zurück zum Zitat Mayr L, Trissal M, Schwark K, Labelle J, Groves A, Furtner-Srajer J, et al. Abstract 5719: clinical response to the PDGFRα inhibitor avapritinib in high-grade glioma patients. Cancer Res. 2023;83(7Supplement):5719.CrossRef Mayr L, Trissal M, Schwark K, Labelle J, Groves A, Furtner-Srajer J, et al. Abstract 5719: clinical response to the PDGFRα inhibitor avapritinib in high-grade glioma patients. Cancer Res. 2023;83(7Supplement):5719.CrossRef
23.
Zurück zum Zitat Teuber A, Schulz T, Fletcher BS, Gontla R, Mühlenberg T, Zischinsky ML, et al. Avapritinib-based SAR studies unveil a binding pocket in KIT and PDGFRA. Nat Commun. 2024;15(1):63.PubMedPubMedCentralCrossRef Teuber A, Schulz T, Fletcher BS, Gontla R, Mühlenberg T, Zischinsky ML, et al. Avapritinib-based SAR studies unveil a binding pocket in KIT and PDGFRA. Nat Commun. 2024;15(1):63.PubMedPubMedCentralCrossRef
24.
Zurück zum Zitat Martin-Broto J, Moura DS. New drugs in gastrointestinal stromal tumors. Curr Opin Oncol. 2020;32(4):314–20.PubMedCrossRef Martin-Broto J, Moura DS. New drugs in gastrointestinal stromal tumors. Curr Opin Oncol. 2020;32(4):314–20.PubMedCrossRef
25.
Zurück zum Zitat Blakeley JON, Fisher JD, Lieberman FS, Lupo J, Nabors LB, Crane J, et al. Imaging biomarkers of ramucirumab and olaratumab in patients with recurrent glioblastoma. JCO. 2013;31(15suppl):2044.CrossRef Blakeley JON, Fisher JD, Lieberman FS, Lupo J, Nabors LB, Crane J, et al. Imaging biomarkers of ramucirumab and olaratumab in patients with recurrent glioblastoma. JCO. 2013;31(15suppl):2044.CrossRef
26.
Zurück zum Zitat Tinkle CL, Broniscer A, Chiang J, Campagne O, Huang J, Orr BA et al. Phase I study using crenolanib to target PDGFR kinase in children and young adults with newly diagnosed DIPG or recurrent high-grade glioma, including DIPG. Neuro-Oncology Adv. 2021;3(1). Tinkle CL, Broniscer A, Chiang J, Campagne O, Huang J, Orr BA et al. Phase I study using crenolanib to target PDGFR kinase in children and young adults with newly diagnosed DIPG or recurrent high-grade glioma, including DIPG. Neuro-Oncology Adv. 2021;3(1).
27.
Zurück zum Zitat Mayr L, Trissal M, Schwark K, Labelle J, Kong S, Groves A, et al. DDDR-22. TRANSLATION OF THE PDGFRA/KIT INHIBITOR AVAPRITINIB FOR PEDIATRIC HIGH-GRADE GLIOMA. Neuro Oncol. 2022;24(Supplement7):vii103–vii.CrossRef Mayr L, Trissal M, Schwark K, Labelle J, Kong S, Groves A, et al. DDDR-22. TRANSLATION OF THE PDGFRA/KIT INHIBITOR AVAPRITINIB FOR PEDIATRIC HIGH-GRADE GLIOMA. Neuro Oncol. 2022;24(Supplement7):vii103–vii.CrossRef
28.
Zurück zum Zitat Kargiotis O, Rao JS, Kyritsis AP. Mechanisms of angiogenesis in gliomas. J Neurooncol. 2006;78(3):281–93.PubMedCrossRef Kargiotis O, Rao JS, Kyritsis AP. Mechanisms of angiogenesis in gliomas. J Neurooncol. 2006;78(3):281–93.PubMedCrossRef
29.
Zurück zum Zitat Rubenstein JL, Kim J, Ozawa T, Zhang M, Westphal M, Deen DF, et al. Anti-VEGF antibody treatment of Glioblastoma Prolongs Survival but results in increased vascular cooption. Neoplasia. 2000;2(4):306–14.PubMedPubMedCentralCrossRef Rubenstein JL, Kim J, Ozawa T, Zhang M, Westphal M, Deen DF, et al. Anti-VEGF antibody treatment of Glioblastoma Prolongs Survival but results in increased vascular cooption. Neoplasia. 2000;2(4):306–14.PubMedPubMedCentralCrossRef
30.
Zurück zum Zitat Friedman HS, Prados MD, Wen PY, Mikkelsen T, Schiff D, Abrey LE, et al. Bevacizumab alone and in combination with irinotecan in recurrent glioblastoma. J Clin Oncol. 2009;27(28):4733–40.PubMedCrossRef Friedman HS, Prados MD, Wen PY, Mikkelsen T, Schiff D, Abrey LE, et al. Bevacizumab alone and in combination with irinotecan in recurrent glioblastoma. J Clin Oncol. 2009;27(28):4733–40.PubMedCrossRef
31.
Zurück zum Zitat Kreisl TN, Kim L, Moore K, Duic P, Royce C, Stroud I, et al. Phase II trial of single-Agent Bevacizumab followed by Bevacizumab Plus Irinotecan at Tumor Progression in Recurrent Glioblastoma. JCO. 2009;27(5):740–5.CrossRef Kreisl TN, Kim L, Moore K, Duic P, Royce C, Stroud I, et al. Phase II trial of single-Agent Bevacizumab followed by Bevacizumab Plus Irinotecan at Tumor Progression in Recurrent Glioblastoma. JCO. 2009;27(5):740–5.CrossRef
32.
Zurück zum Zitat Lai A, Tran A, Nghiemphu PL, Pope WB, Solis OE, Selch M, et al. Phase II study of bevacizumab plus temozolomide during and after radiation therapy for patients with newly diagnosed glioblastoma multiforme. J Clin Oncol. 2011;29(2):142–8.PubMedCrossRef Lai A, Tran A, Nghiemphu PL, Pope WB, Solis OE, Selch M, et al. Phase II study of bevacizumab plus temozolomide during and after radiation therapy for patients with newly diagnosed glioblastoma multiforme. J Clin Oncol. 2011;29(2):142–8.PubMedCrossRef
33.
Zurück zum Zitat Chinot OL, Wick W, Mason W, Henriksson R, Saran F, Nishikawa R, et al. Bevacizumab plus radiotherapy-temozolomide for newly diagnosed glioblastoma. N Engl J Med. 2014;370(8):709–22.PubMedCrossRef Chinot OL, Wick W, Mason W, Henriksson R, Saran F, Nishikawa R, et al. Bevacizumab plus radiotherapy-temozolomide for newly diagnosed glioblastoma. N Engl J Med. 2014;370(8):709–22.PubMedCrossRef
34.
Zurück zum Zitat Gilbert MR, Dignam JJ, Armstrong TS, Wefel JS, Blumenthal DT, Vogelbaum MA, et al. A randomized trial of bevacizumab for newly diagnosed glioblastoma. N Engl J Med. 2014;370(8):699–708.PubMedPubMedCentralCrossRef Gilbert MR, Dignam JJ, Armstrong TS, Wefel JS, Blumenthal DT, Vogelbaum MA, et al. A randomized trial of bevacizumab for newly diagnosed glioblastoma. N Engl J Med. 2014;370(8):699–708.PubMedPubMedCentralCrossRef
35.
Zurück zum Zitat Hygino da Cruz LC Jr., Rodriguez I, Domingues RC, Gasparetto EL, Sorensen AG. Pseudoprogression and pseudoresponse: imaging challenges in the assessment of posttreatment glioma. AJNR Am J Neuroradiol. 2011;32(11):1978–85.PubMedPubMedCentralCrossRef Hygino da Cruz LC Jr., Rodriguez I, Domingues RC, Gasparetto EL, Sorensen AG. Pseudoprogression and pseudoresponse: imaging challenges in the assessment of posttreatment glioma. AJNR Am J Neuroradiol. 2011;32(11):1978–85.PubMedPubMedCentralCrossRef
36.
Zurück zum Zitat Fukumura D, Kloepper J, Amoozgar Z, Duda DG, Jain RK. Enhancing cancer immunotherapy using antiangiogenics: opportunities and challenges. Nat Rev Clin Oncol. 2018;15(5):325–40.PubMedPubMedCentralCrossRef Fukumura D, Kloepper J, Amoozgar Z, Duda DG, Jain RK. Enhancing cancer immunotherapy using antiangiogenics: opportunities and challenges. Nat Rev Clin Oncol. 2018;15(5):325–40.PubMedPubMedCentralCrossRef
37.
Zurück zum Zitat Yang G, Fang Y, Zhou M, Li W, Dong D, Chen J, et al. Case report: the effective response to pembrolizumab in combination with bevacizumab in the treatment of a recurrent glioblastoma with multiple extracranial metastases. Front Oncol. 2022;12:948933.PubMedPubMedCentralCrossRef Yang G, Fang Y, Zhou M, Li W, Dong D, Chen J, et al. Case report: the effective response to pembrolizumab in combination with bevacizumab in the treatment of a recurrent glioblastoma with multiple extracranial metastases. Front Oncol. 2022;12:948933.PubMedPubMedCentralCrossRef
38.
Zurück zum Zitat Nayak L, Molinaro AM, Peters K, Clarke JL, Jordan JT, de Groot J, et al. Randomized Phase II and Biomarker Study of Pembrolizumab plus Bevacizumab versus Pembrolizumab alone for patients with recurrent glioblastoma. Clin Cancer Res. 2021;27(4):1048–57.PubMedCrossRef Nayak L, Molinaro AM, Peters K, Clarke JL, Jordan JT, de Groot J, et al. Randomized Phase II and Biomarker Study of Pembrolizumab plus Bevacizumab versus Pembrolizumab alone for patients with recurrent glioblastoma. Clin Cancer Res. 2021;27(4):1048–57.PubMedCrossRef
39.
Zurück zum Zitat Eskilsson E, Røsland GV, Solecki G, Wang Q, Harter PN, Graziani G, et al. EGFR heterogeneity and implications for therapeutic intervention in glioblastoma. Neuro Oncol. 2018;20(6):743–52.PubMedCrossRef Eskilsson E, Røsland GV, Solecki G, Wang Q, Harter PN, Graziani G, et al. EGFR heterogeneity and implications for therapeutic intervention in glioblastoma. Neuro Oncol. 2018;20(6):743–52.PubMedCrossRef
40.
Zurück zum Zitat An Z, Aksoy O, Zheng T, Fan Q-W, Weiss WA. Epidermal growth factor receptor and EGFRvIII in glioblastoma: signaling pathways and targeted therapies. Oncogene. 2018;37(12):1561–75.PubMedPubMedCentralCrossRef An Z, Aksoy O, Zheng T, Fan Q-W, Weiss WA. Epidermal growth factor receptor and EGFRvIII in glioblastoma: signaling pathways and targeted therapies. Oncogene. 2018;37(12):1561–75.PubMedPubMedCentralCrossRef
41.
Zurück zum Zitat Brennan CW, Verhaak RG, McKenna A, Campos B, Noushmehr H, Salama SR, et al. The somatic genomic landscape of glioblastoma. Cell. 2013;155(2):462–77.PubMedPubMedCentralCrossRef Brennan CW, Verhaak RG, McKenna A, Campos B, Noushmehr H, Salama SR, et al. The somatic genomic landscape of glioblastoma. Cell. 2013;155(2):462–77.PubMedPubMedCentralCrossRef
42.
Zurück zum Zitat Turner KM, Deshpande V, Beyter D, Koga T, Rusert J, Lee C, et al. Extrachromosomal oncogene amplification drives tumour evolution and genetic heterogeneity. Nature. 2017;543(7643):122–5.PubMedPubMedCentralCrossRef Turner KM, Deshpande V, Beyter D, Koga T, Rusert J, Lee C, et al. Extrachromosomal oncogene amplification drives tumour evolution and genetic heterogeneity. Nature. 2017;543(7643):122–5.PubMedPubMedCentralCrossRef
43.
Zurück zum Zitat Uhm JH, Ballman KV, Wu W, Giannini C, Krauss JC, Buckner JC, et al. Phase II evaluation of gefitinib in patients with newly diagnosed Grade 4 astrocytoma: Mayo/North Central Cancer Treatment Group Study N0074. Int J Radiat Oncol Biol Phys. 2011;80(2):347–53.PubMedCrossRef Uhm JH, Ballman KV, Wu W, Giannini C, Krauss JC, Buckner JC, et al. Phase II evaluation of gefitinib in patients with newly diagnosed Grade 4 astrocytoma: Mayo/North Central Cancer Treatment Group Study N0074. Int J Radiat Oncol Biol Phys. 2011;80(2):347–53.PubMedCrossRef
44.
Zurück zum Zitat Gallego O, Cuatrecasas M, Benavides M, Segura PP, Berrocal A, Erill N, et al. Efficacy of erlotinib in patients with relapsed gliobastoma multiforme who expressed EGFRVIII and PTEN determined by immunohistochemistry. J Neurooncol. 2014;116(2):413–9.PubMedCrossRef Gallego O, Cuatrecasas M, Benavides M, Segura PP, Berrocal A, Erill N, et al. Efficacy of erlotinib in patients with relapsed gliobastoma multiforme who expressed EGFRVIII and PTEN determined by immunohistochemistry. J Neurooncol. 2014;116(2):413–9.PubMedCrossRef
45.
Zurück zum Zitat Raizer JJ, Abrey LE, Lassman AB, Chang SM, Lamborn KR, Kuhn JG, et al. A phase II trial of erlotinib in patients with recurrent malignant gliomas and nonprogressive glioblastoma multiforme postradiation therapy. Neuro Oncol. 2010;12(1):95–103.PubMedCrossRef Raizer JJ, Abrey LE, Lassman AB, Chang SM, Lamborn KR, Kuhn JG, et al. A phase II trial of erlotinib in patients with recurrent malignant gliomas and nonprogressive glioblastoma multiforme postradiation therapy. Neuro Oncol. 2010;12(1):95–103.PubMedCrossRef
46.
Zurück zum Zitat Reardon DA, Nabors LB, Mason WP, Perry JR, Shapiro W, Kavan P, et al. Phase I/randomized phase II study of afatinib, an irreversible ErbB family blocker, with or without protracted temozolomide in adults with recurrent glioblastoma. Neuro Oncol. 2015;17(3):430–9.PubMed Reardon DA, Nabors LB, Mason WP, Perry JR, Shapiro W, Kavan P, et al. Phase I/randomized phase II study of afatinib, an irreversible ErbB family blocker, with or without protracted temozolomide in adults with recurrent glioblastoma. Neuro Oncol. 2015;17(3):430–9.PubMed
47.
Zurück zum Zitat Sepulveda-Sanchez JM, Vaz MA, Balana C, Gil-Gil M, Reynes G, Gallego O, et al. Phase II trial of dacomitinib, a pan-human EGFR tyrosine kinase inhibitor, in recurrent glioblastoma patients with EGFR amplification. Neuro Oncol. 2017;19(11):1522–31.PubMedPubMedCentralCrossRef Sepulveda-Sanchez JM, Vaz MA, Balana C, Gil-Gil M, Reynes G, Gallego O, et al. Phase II trial of dacomitinib, a pan-human EGFR tyrosine kinase inhibitor, in recurrent glioblastoma patients with EGFR amplification. Neuro Oncol. 2017;19(11):1522–31.PubMedPubMedCentralCrossRef
48.
Zurück zum Zitat Hasselbalch B, Lassen U, Hansen S, Holmberg M, Sorensen M, Kosteljanetz M, et al. Cetuximab, bevacizumab, and irinotecan for patients with primary glioblastoma and progression after radiation therapy and temozolomide: a phase II trial. Neuro Oncol. 2010;12(5):508–16.PubMedPubMedCentral Hasselbalch B, Lassen U, Hansen S, Holmberg M, Sorensen M, Kosteljanetz M, et al. Cetuximab, bevacizumab, and irinotecan for patients with primary glioblastoma and progression after radiation therapy and temozolomide: a phase II trial. Neuro Oncol. 2010;12(5):508–16.PubMedPubMedCentral
49.
Zurück zum Zitat Pan PC, Magge RS. Mechanisms of EGFR Resistance in Glioblastoma. Int J Mol Sci. 2020;21(22). Pan PC, Magge RS. Mechanisms of EGFR Resistance in Glioblastoma. Int J Mol Sci. 2020;21(22).
50.
Zurück zum Zitat Hu C, Leche CA 2nd, Kiyatkin A, Yu Z, Stayrook SE, Ferguson KM, et al. Glioblastoma mutations alter EGFR dimer structure to prevent ligand bias. Nature. 2022;602(7897):518–22.PubMedPubMedCentralCrossRef Hu C, Leche CA 2nd, Kiyatkin A, Yu Z, Stayrook SE, Ferguson KM, et al. Glioblastoma mutations alter EGFR dimer structure to prevent ligand bias. Nature. 2022;602(7897):518–22.PubMedPubMedCentralCrossRef
51.
Zurück zum Zitat Chakravarty D, Pedraza AM, Cotari J, Liu AH, Punko D, Kokroo A, et al. EGFR and PDGFRA co-expression and heterodimerization in glioblastoma tumor sphere lines. Sci Rep. 2017;7(1):9043.PubMedPubMedCentralCrossRef Chakravarty D, Pedraza AM, Cotari J, Liu AH, Punko D, Kokroo A, et al. EGFR and PDGFRA co-expression and heterodimerization in glioblastoma tumor sphere lines. Sci Rep. 2017;7(1):9043.PubMedPubMedCentralCrossRef
52.
Zurück zum Zitat von Achenbach C, Silginer M, Blot V, Weiss WA, Weller M. Depatuxizumab Mafodotin (ABT-414)-induced Glioblastoma Cell Death requires EGFR overexpression, but not EGFR(Y1068) phosphorylation. Mol Cancer Ther. 2020;19(6):1328–39.CrossRef von Achenbach C, Silginer M, Blot V, Weiss WA, Weller M. Depatuxizumab Mafodotin (ABT-414)-induced Glioblastoma Cell Death requires EGFR overexpression, but not EGFR(Y1068) phosphorylation. Mol Cancer Ther. 2020;19(6):1328–39.CrossRef
53.
Zurück zum Zitat Phillips AC, Boghaert ER, Vaidya KS, Mitten MJ, Norvell S, Falls HD, et al. ABT-414, an antibody-drug Conjugate Targeting a Tumor-Selective EGFR Epitope. Mol Cancer Ther. 2016;15(4):661–9.PubMedCrossRef Phillips AC, Boghaert ER, Vaidya KS, Mitten MJ, Norvell S, Falls HD, et al. ABT-414, an antibody-drug Conjugate Targeting a Tumor-Selective EGFR Epitope. Mol Cancer Ther. 2016;15(4):661–9.PubMedCrossRef
54.
Zurück zum Zitat Van Den Bent M, Eoli M, Sepulveda JM, Smits M, Walenkamp A, Frenel JS, et al. INTELLANCE 2/EORTC 1410 randomized phase II study of Depatux-M alone and with temozolomide vs temozolomide or lomustine in recurrent EGFR amplified glioblastoma. Neuro Oncol. 2020;22(5):684–93.PubMedCrossRef Van Den Bent M, Eoli M, Sepulveda JM, Smits M, Walenkamp A, Frenel JS, et al. INTELLANCE 2/EORTC 1410 randomized phase II study of Depatux-M alone and with temozolomide vs temozolomide or lomustine in recurrent EGFR amplified glioblastoma. Neuro Oncol. 2020;22(5):684–93.PubMedCrossRef
55.
Zurück zum Zitat Lassman AB, Pugh SL, Wang TJC, Aldape K, Gan HK, Preusser M, et al. Depatuxizumab mafodotin in EGFR-amplified newly diagnosed glioblastoma: a phase III randomized clinical trial. Neuro Oncol. 2023;25(2):339–50.PubMedCrossRef Lassman AB, Pugh SL, Wang TJC, Aldape K, Gan HK, Preusser M, et al. Depatuxizumab mafodotin in EGFR-amplified newly diagnosed glioblastoma: a phase III randomized clinical trial. Neuro Oncol. 2023;25(2):339–50.PubMedCrossRef
56.
Zurück zum Zitat Ahluwalia M, Narita Y, Muragaki Y, Gan H, Merrell R, van den Bent M, et al. OS1.2 Stability of EGFR amplification in glioblastoma is differentially impacted based on therapeutic pressure. Neuro Oncol. 2018;20(suppl3):iii217–iii.PubMedCentralCrossRef Ahluwalia M, Narita Y, Muragaki Y, Gan H, Merrell R, van den Bent M, et al. OS1.2 Stability of EGFR amplification in glioblastoma is differentially impacted based on therapeutic pressure. Neuro Oncol. 2018;20(suppl3):iii217–iii.PubMedCentralCrossRef
57.
Zurück zum Zitat Gan HK, Parakh S, Lassman AB, Seow A, Lau E, Lee ST, et al. Tumor volumes as a predictor of response to the anti-EGFR antibody drug conjugate depatuxizumab mafadotin. Neurooncol Adv. 2021;3(1):vdab102.PubMedPubMedCentral Gan HK, Parakh S, Lassman AB, Seow A, Lau E, Lee ST, et al. Tumor volumes as a predictor of response to the anti-EGFR antibody drug conjugate depatuxizumab mafadotin. Neurooncol Adv. 2021;3(1):vdab102.PubMedPubMedCentral
58.
Zurück zum Zitat Marin BM, Porath KA, Jain S, Kim M, Conage-Pough JE, Oh JH, et al. Heterogeneous delivery across the blood-brain barrier limits the efficacy of an EGFR-targeting antibody drug conjugate in glioblastoma. Neuro Oncol. 2021;23(12):2042–53.PubMedPubMedCentralCrossRef Marin BM, Porath KA, Jain S, Kim M, Conage-Pough JE, Oh JH, et al. Heterogeneous delivery across the blood-brain barrier limits the efficacy of an EGFR-targeting antibody drug conjugate in glioblastoma. Neuro Oncol. 2021;23(12):2042–53.PubMedPubMedCentralCrossRef
59.
Zurück zum Zitat Ma Y-S, Wu Z-J, Bai R-Z, Dong H, Xie B-X, Wu X-H, et al. DRR1 promotes glioblastoma cell invasion and epithelial-mesenchymal transition via regulating AKT activation. Cancer Lett. 2018;423:86–94.PubMedCrossRef Ma Y-S, Wu Z-J, Bai R-Z, Dong H, Xie B-X, Wu X-H, et al. DRR1 promotes glioblastoma cell invasion and epithelial-mesenchymal transition via regulating AKT activation. Cancer Lett. 2018;423:86–94.PubMedCrossRef
60.
Zurück zum Zitat Gupta K, Jones JC, Farias VA, Mackeyev Y, Singh PK, Quiñones-Hinojosa A, et al. Identification of synergistic drug combinations to Target KRAS-Driven Chemoradioresistant Cancers utilizing Tumoroid models of colorectal adenocarcinoma and recurrent glioblastoma. Front Oncol. 2022;12:840241.PubMedPubMedCentralCrossRef Gupta K, Jones JC, Farias VA, Mackeyev Y, Singh PK, Quiñones-Hinojosa A, et al. Identification of synergistic drug combinations to Target KRAS-Driven Chemoradioresistant Cancers utilizing Tumoroid models of colorectal adenocarcinoma and recurrent glioblastoma. Front Oncol. 2022;12:840241.PubMedPubMedCentralCrossRef
61.
Zurück zum Zitat Carpenter RL, Lo H-W. STAT3 target genes relevant to human cancers. Cancers (Basel). 2014;6(2):897–925.PubMedCrossRef Carpenter RL, Lo H-W. STAT3 target genes relevant to human cancers. Cancers (Basel). 2014;6(2):897–925.PubMedCrossRef
62.
Zurück zum Zitat Rahaman SO, Harbor PC, Chernova O, Barnett GH, Vogelbaum MA, Haque SJ. Inhibition of constitutively active Stat3 suppresses proliferation and induces apoptosis in glioblastoma multiforme cells. Oncogene. 2002;21(55):8404–13.PubMedCrossRef Rahaman SO, Harbor PC, Chernova O, Barnett GH, Vogelbaum MA, Haque SJ. Inhibition of constitutively active Stat3 suppresses proliferation and induces apoptosis in glioblastoma multiforme cells. Oncogene. 2002;21(55):8404–13.PubMedCrossRef
63.
Zurück zum Zitat Kim JE, Patel M, Ruzevick J, Jackson CM, Lim M. STAT3 activation in Glioblastoma: biochemical and therapeutic implications. Cancers (Basel). 2014;6(1):376–95.PubMedCrossRef Kim JE, Patel M, Ruzevick J, Jackson CM, Lim M. STAT3 activation in Glioblastoma: biochemical and therapeutic implications. Cancers (Basel). 2014;6(1):376–95.PubMedCrossRef
64.
Zurück zum Zitat Kouri FM, Jensen SA, Stegh AH. The role of Bcl-2 family proteins in therapy responses of malignant astrocytic gliomas: Bcl2L12 and Beyond. ScientificWorldJournal. 2012;2012:838916.PubMedPubMedCentralCrossRef Kouri FM, Jensen SA, Stegh AH. The role of Bcl-2 family proteins in therapy responses of malignant astrocytic gliomas: Bcl2L12 and Beyond. ScientificWorldJournal. 2012;2012:838916.PubMedPubMedCentralCrossRef
65.
Zurück zum Zitat Iwamaru A, Szymanski S, Iwado E, Aoki H, Yokoyama T, Fokt I, et al. A novel inhibitor of the STAT3 pathway induces apoptosis in malignant glioma cells both in vitro and in vivo. Oncogene. 2007;26(17):2435–44.PubMedCrossRef Iwamaru A, Szymanski S, Iwado E, Aoki H, Yokoyama T, Fokt I, et al. A novel inhibitor of the STAT3 pathway induces apoptosis in malignant glioma cells both in vitro and in vivo. Oncogene. 2007;26(17):2435–44.PubMedCrossRef
66.
Zurück zum Zitat Bhattacharya S, Yin J, Yang C, Wang Y, Sims M, Pfeffer LM, et al. STAT3 suppresses the AMPKalpha/ULK1-dependent induction of autophagy in glioblastoma cells. J Cell Mol Med. 2022;26(14):3873–90.PubMedPubMedCentralCrossRef Bhattacharya S, Yin J, Yang C, Wang Y, Sims M, Pfeffer LM, et al. STAT3 suppresses the AMPKalpha/ULK1-dependent induction of autophagy in glioblastoma cells. J Cell Mol Med. 2022;26(14):3873–90.PubMedPubMedCentralCrossRef
67.
Zurück zum Zitat Li H, Chen L, Li JJ, Zhou Q, Huang A, Liu WW, et al. miR-519a enhances chemosensitivity and promotes autophagy in glioblastoma by targeting STAT3/Bcl2 signaling pathway. J Hematol Oncol. 2018;11(1):70.PubMedPubMedCentralCrossRef Li H, Chen L, Li JJ, Zhou Q, Huang A, Liu WW, et al. miR-519a enhances chemosensitivity and promotes autophagy in glioblastoma by targeting STAT3/Bcl2 signaling pathway. J Hematol Oncol. 2018;11(1):70.PubMedPubMedCentralCrossRef
68.
Zurück zum Zitat Lin J-C, Tsai J-T, Chao T-Y, Ma H-I, Liu W-H. The STAT3/Slug Axis enhances Radiation-Induced Tumor Invasion and Cancer Stem-like Properties in Radioresistant Glioblastoma. Cancers (Basel). 2018;10(12):512.PubMedCrossRef Lin J-C, Tsai J-T, Chao T-Y, Ma H-I, Liu W-H. The STAT3/Slug Axis enhances Radiation-Induced Tumor Invasion and Cancer Stem-like Properties in Radioresistant Glioblastoma. Cancers (Basel). 2018;10(12):512.PubMedCrossRef
69.
Zurück zum Zitat Dave B, Mittal V, Tan NM, Chang JC. Epithelial-mesenchymal transition, cancer stem cells and treatment resistance. Breast Cancer Res. 2012;14(1):202.PubMedPubMedCentralCrossRef Dave B, Mittal V, Tan NM, Chang JC. Epithelial-mesenchymal transition, cancer stem cells and treatment resistance. Breast Cancer Res. 2012;14(1):202.PubMedPubMedCentralCrossRef
70.
Zurück zum Zitat Garner JM, Fan M, Yang CH, Du Z, Sims M, Davidoff AM, et al. Constitutive Activation of Signal Transducer and activator of transcription 3 (STAT3) and nuclear factor κB Signaling in Glioblastoma Cancer Stem cells regulates the Notch Pathway. J Biol Chem. 2013;288(36):26167–76.PubMedPubMedCentralCrossRef Garner JM, Fan M, Yang CH, Du Z, Sims M, Davidoff AM, et al. Constitutive Activation of Signal Transducer and activator of transcription 3 (STAT3) and nuclear factor κB Signaling in Glioblastoma Cancer Stem cells regulates the Notch Pathway. J Biol Chem. 2013;288(36):26167–76.PubMedPubMedCentralCrossRef
71.
Zurück zum Zitat Segerman A, Niklasson M, Haglund C, Bergstrom T, Jarvius M, Xie Y, et al. Clonal variation in drug and Radiation Response among Glioma-initiating cells is linked to Proneural-Mesenchymal transition. Cell Rep. 2016;17(11):2994–3009.PubMedCrossRef Segerman A, Niklasson M, Haglund C, Bergstrom T, Jarvius M, Xie Y, et al. Clonal variation in drug and Radiation Response among Glioma-initiating cells is linked to Proneural-Mesenchymal transition. Cell Rep. 2016;17(11):2994–3009.PubMedCrossRef
72.
Zurück zum Zitat Tan MSY, Sandanaraj E, Chong YK, Lim SW, Koh LWH, Ng WH, et al. A STAT3-based gene signature stratifies glioma patients for targeted therapy. Nat Commun. 2019;10(1):3601.PubMedPubMedCentralCrossRef Tan MSY, Sandanaraj E, Chong YK, Lim SW, Koh LWH, Ng WH, et al. A STAT3-based gene signature stratifies glioma patients for targeted therapy. Nat Commun. 2019;10(1):3601.PubMedPubMedCentralCrossRef
73.
Zurück zum Zitat de Groot J, Liang J, Kong LY, Wei J, Piao Y, Fuller G, et al. Modulating antiangiogenic resistance by inhibiting the signal transducer and activator of transcription 3 pathway in glioblastoma. Oncotarget. 2012;3(9):1036–48.PubMedPubMedCentralCrossRef de Groot J, Liang J, Kong LY, Wei J, Piao Y, Fuller G, et al. Modulating antiangiogenic resistance by inhibiting the signal transducer and activator of transcription 3 pathway in glioblastoma. Oncotarget. 2012;3(9):1036–48.PubMedPubMedCentralCrossRef
74.
Zurück zum Zitat Papadakis AI, Paraskeva E, Peidis P, Muaddi H, Li S, Raptis L, et al. eIF2alpha kinase PKR modulates the hypoxic response by Stat3-dependent transcriptional suppression of HIF-1alpha. Cancer Res. 2010;70(20):7820–9.PubMedCrossRef Papadakis AI, Paraskeva E, Peidis P, Muaddi H, Li S, Raptis L, et al. eIF2alpha kinase PKR modulates the hypoxic response by Stat3-dependent transcriptional suppression of HIF-1alpha. Cancer Res. 2010;70(20):7820–9.PubMedCrossRef
75.
Zurück zum Zitat Schaefer LK, Ren Z, Fuller GN, Schaefer TS. Constitutive activation of Stat3alpha in brain tumors: localization to tumor endothelial cells and activation by the endothelial tyrosine kinase receptor (VEGFR-2). Oncogene. 2002;21(13):2058–65.PubMedCrossRef Schaefer LK, Ren Z, Fuller GN, Schaefer TS. Constitutive activation of Stat3alpha in brain tumors: localization to tumor endothelial cells and activation by the endothelial tyrosine kinase receptor (VEGFR-2). Oncogene. 2002;21(13):2058–65.PubMedCrossRef
76.
Zurück zum Zitat Sherry MM, Reeves A, Wu JK, Cochran BH. STAT3 is required for proliferation and maintenance of multipotency in glioblastoma stem cells. Stem Cells. 2009;27(10):2383–92.PubMedCrossRef Sherry MM, Reeves A, Wu JK, Cochran BH. STAT3 is required for proliferation and maintenance of multipotency in glioblastoma stem cells. Stem Cells. 2009;27(10):2383–92.PubMedCrossRef
77.
Zurück zum Zitat Song H, Wang R, Wang S, Lin J. A low-molecular-weight compound discovered through virtual database screening inhibits Stat3 function in breast cancer cells. Proc Natl Acad Sci U S A. 2005;102(13):4700–5.PubMedPubMedCentralCrossRef Song H, Wang R, Wang S, Lin J. A low-molecular-weight compound discovered through virtual database screening inhibits Stat3 function in breast cancer cells. Proc Natl Acad Sci U S A. 2005;102(13):4700–5.PubMedPubMedCentralCrossRef
78.
Zurück zum Zitat Ball S, Li C, Li PK, Lin J. The small molecule, LLL12, inhibits STAT3 phosphorylation and induces apoptosis in medulloblastoma and glioblastoma cells. PLoS ONE. 2011;6(4):e18820.PubMedPubMedCentralCrossRef Ball S, Li C, Li PK, Lin J. The small molecule, LLL12, inhibits STAT3 phosphorylation and induces apoptosis in medulloblastoma and glioblastoma cells. PLoS ONE. 2011;6(4):e18820.PubMedPubMedCentralCrossRef
79.
Zurück zum Zitat Fuh B, Sobo M, Cen L, Josiah D, Hutzen B, Cisek K, et al. LLL-3 inhibits STAT3 activity, suppresses glioblastoma cell growth and prolongs survival in a mouse glioblastoma model. Br J Cancer. 2009;100(1):106–12.PubMedPubMedCentralCrossRef Fuh B, Sobo M, Cen L, Josiah D, Hutzen B, Cisek K, et al. LLL-3 inhibits STAT3 activity, suppresses glioblastoma cell growth and prolongs survival in a mouse glioblastoma model. Br J Cancer. 2009;100(1):106–12.PubMedPubMedCentralCrossRef
80.
Zurück zum Zitat Ashizawa T, Miyata H, Iizuka A, Komiyama M, Oshita C, Kume A, et al. Effect of the STAT3 inhibitor STX-0119 on the proliferation of cancer stem-like cells derived from recurrent glioblastoma. Int J Oncol. 2013;43(1):219–27.PubMedCrossRef Ashizawa T, Miyata H, Iizuka A, Komiyama M, Oshita C, Kume A, et al. Effect of the STAT3 inhibitor STX-0119 on the proliferation of cancer stem-like cells derived from recurrent glioblastoma. Int J Oncol. 2013;43(1):219–27.PubMedCrossRef
81.
Zurück zum Zitat Senft C, Priester M, Polacin M, Schröder K, Seifert V, Kögel D, et al. Inhibition of the JAK-2/STAT3 signaling pathway impedes the migratory and invasive potential of human glioblastoma cells. J Neurooncol. 2011;101(3):393–403.PubMedCrossRef Senft C, Priester M, Polacin M, Schröder K, Seifert V, Kögel D, et al. Inhibition of the JAK-2/STAT3 signaling pathway impedes the migratory and invasive potential of human glioblastoma cells. J Neurooncol. 2011;101(3):393–403.PubMedCrossRef
82.
Zurück zum Zitat Zheng Q, Han L, Dong Y, Tian J, Huang W, Liu Z, et al. JAK2/STAT3 targeted therapy suppresses tumor invasion via disruption of the EGFRvIII/JAK2/STAT3 axis and associated focal adhesion in EGFRvIII-expressing glioblastoma. Neuro Oncol. 2014;16(9):1229–43.PubMedPubMedCentralCrossRef Zheng Q, Han L, Dong Y, Tian J, Huang W, Liu Z, et al. JAK2/STAT3 targeted therapy suppresses tumor invasion via disruption of the EGFRvIII/JAK2/STAT3 axis and associated focal adhesion in EGFRvIII-expressing glioblastoma. Neuro Oncol. 2014;16(9):1229–43.PubMedPubMedCentralCrossRef
83.
Zurück zum Zitat Stechishin OD, Luchman HA, Ruan Y, Blough MD, Nguyen SA, Kelly JJ, et al. On-target JAK2/STAT3 inhibition slows disease progression in orthotopic xenografts of human glioblastoma brain tumor stem cells. Neuro Oncol. 2013;15(2):198–207.PubMedCrossRef Stechishin OD, Luchman HA, Ruan Y, Blough MD, Nguyen SA, Kelly JJ, et al. On-target JAK2/STAT3 inhibition slows disease progression in orthotopic xenografts of human glioblastoma brain tumor stem cells. Neuro Oncol. 2013;15(2):198–207.PubMedCrossRef
84.
Zurück zum Zitat Groot J, Ott M, Wei J, Kassab C, Fang D, Najem H, et al. A first-in-human phase I trial of the oral p-STAT3 inhibitor WP1066 in patients with recurrent malignant glioma. CNS Oncol. 2022;11(2):CNS87.PubMedPubMedCentralCrossRef Groot J, Ott M, Wei J, Kassab C, Fang D, Najem H, et al. A first-in-human phase I trial of the oral p-STAT3 inhibitor WP1066 in patients with recurrent malignant glioma. CNS Oncol. 2022;11(2):CNS87.PubMedPubMedCentralCrossRef
85.
Zurück zum Zitat Hong D, Kurzrock R, Kim Y, Woessner R, Younes A, Nemunaitis J, et al. AZD9150, a next-generation antisense oligonucleotide inhibitor of STAT3 with early evidence of clinical activity in Lymphoma and Lung Cancer. Sci Transl Med. 2015;7(314):314ra185.PubMedPubMedCentralCrossRef Hong D, Kurzrock R, Kim Y, Woessner R, Younes A, Nemunaitis J, et al. AZD9150, a next-generation antisense oligonucleotide inhibitor of STAT3 with early evidence of clinical activity in Lymphoma and Lung Cancer. Sci Transl Med. 2015;7(314):314ra185.PubMedPubMedCentralCrossRef
86.
Zurück zum Zitat Reilley MJ, McCoon P, Cook C, Lyne P, Kurzrock R, Kim Y, et al. STAT3 antisense oligonucleotide AZD9150 in a subset of patients with heavily pretreated lymphoma: results of a phase 1b trial. J Immunother Cancer. 2018;6(1):119.PubMedPubMedCentralCrossRef Reilley MJ, McCoon P, Cook C, Lyne P, Kurzrock R, Kim Y, et al. STAT3 antisense oligonucleotide AZD9150 in a subset of patients with heavily pretreated lymphoma: results of a phase 1b trial. J Immunother Cancer. 2018;6(1):119.PubMedPubMedCentralCrossRef
87.
Zurück zum Zitat Jing N, Zhu Q, Yuan P, Li Y, Mao L, Tweardy DJ. Targeting signal transducer and activator of transcription 3 with G-quartet oligonucleotides: a potential novel therapy for head and neck cancer. Mol Cancer Ther. 2006;5(2):279–86.PubMedCrossRef Jing N, Zhu Q, Yuan P, Li Y, Mao L, Tweardy DJ. Targeting signal transducer and activator of transcription 3 with G-quartet oligonucleotides: a potential novel therapy for head and neck cancer. Mol Cancer Ther. 2006;5(2):279–86.PubMedCrossRef
88.
Zurück zum Zitat Casas G, Perche F, Midoux P, Pichon C, Malinge J-M. DNA minicircles as novel STAT3 decoy oligodeoxynucleotides endowed with anticancer activity in triple-negative breast cancer. Mol Therapy - Nucleic Acids. 2022;29:162–75.PubMedCrossRef Casas G, Perche F, Midoux P, Pichon C, Malinge J-M. DNA minicircles as novel STAT3 decoy oligodeoxynucleotides endowed with anticancer activity in triple-negative breast cancer. Mol Therapy - Nucleic Acids. 2022;29:162–75.PubMedCrossRef
89.
Zurück zum Zitat Sen M, Thomas SM, Kim S, Yeh JI, Ferris RL, Johnson JT, et al. First-in-human trial of a STAT3 decoy oligonucleotide in head and neck tumors: implications for cancer therapy. Cancer Discov. 2012;2(8):694–705.PubMedPubMedCentralCrossRef Sen M, Thomas SM, Kim S, Yeh JI, Ferris RL, Johnson JT, et al. First-in-human trial of a STAT3 decoy oligonucleotide in head and neck tumors: implications for cancer therapy. Cancer Discov. 2012;2(8):694–705.PubMedPubMedCentralCrossRef
90.
Zurück zum Zitat Zhang X, Zhang J, Wang L, Wei H, Tian Z. Therapeutic effects of STAT3 decoy oligodeoxynucleotide on human lung cancer in xenograft mice. BMC Cancer. 2007;7:149.PubMedPubMedCentralCrossRef Zhang X, Zhang J, Wang L, Wei H, Tian Z. Therapeutic effects of STAT3 decoy oligodeoxynucleotide on human lung cancer in xenograft mice. BMC Cancer. 2007;7:149.PubMedPubMedCentralCrossRef
91.
Zurück zum Zitat Leong PL, Andrews GA, Johnson DE, Dyer KF, Xi S, Mai JC, et al. Targeted inhibition of Stat3 with a decoy oligonucleotide abrogates head and neck cancer cell growth. Proc Natl Acad Sci USA. 2003;100(7):4138–43.PubMedPubMedCentralCrossRef Leong PL, Andrews GA, Johnson DE, Dyer KF, Xi S, Mai JC, et al. Targeted inhibition of Stat3 with a decoy oligonucleotide abrogates head and neck cancer cell growth. Proc Natl Acad Sci USA. 2003;100(7):4138–43.PubMedPubMedCentralCrossRef
92.
Zurück zum Zitat Karlbom AE, James CD, Boethius J, Cavenee WK, Collins VP, Nordenskjold M, et al. Loss of heterozygosity in malignant gliomas involves at least three distinct regions on chromosome 10. Hum Genet. 1993;92(2):169–74.PubMedCrossRef Karlbom AE, James CD, Boethius J, Cavenee WK, Collins VP, Nordenskjold M, et al. Loss of heterozygosity in malignant gliomas involves at least three distinct regions on chromosome 10. Hum Genet. 1993;92(2):169–74.PubMedCrossRef
93.
Zurück zum Zitat Yang JM, Schiapparelli P, Nguyen HN, Igarashi A, Zhang Q, Abbadi S, et al. Characterization of PTEN mutations in brain cancer reveals that pten mono-ubiquitination promotes protein stability and nuclear localization. Oncogene. 2017;36(26):3673–85.PubMedPubMedCentralCrossRef Yang JM, Schiapparelli P, Nguyen HN, Igarashi A, Zhang Q, Abbadi S, et al. Characterization of PTEN mutations in brain cancer reveals that pten mono-ubiquitination promotes protein stability and nuclear localization. Oncogene. 2017;36(26):3673–85.PubMedPubMedCentralCrossRef
94.
Zurück zum Zitat Yang Y, Shao N, Luo G, Li L, Zheng L, Nilsson-Ehle P, et al. Mutations of PTEN gene in gliomas correlate to tumor differentiation and short-term survival rate. Anticancer Res. 2010;30(3):981–5.PubMed Yang Y, Shao N, Luo G, Li L, Zheng L, Nilsson-Ehle P, et al. Mutations of PTEN gene in gliomas correlate to tumor differentiation and short-term survival rate. Anticancer Res. 2010;30(3):981–5.PubMed
95.
Zurück zum Zitat Srividya MR, Thota B, Shailaja BC, Arivazhagan A, Thennarasu K, Chandramouli BA, et al. Homozygous 10q23/PTEN deletion and its impact on outcome in glioblastoma: a prospective translational study on a uniformly treated cohort of adult patients. Neuropathology. 2011;31(4):376–83.PubMedCrossRef Srividya MR, Thota B, Shailaja BC, Arivazhagan A, Thennarasu K, Chandramouli BA, et al. Homozygous 10q23/PTEN deletion and its impact on outcome in glioblastoma: a prospective translational study on a uniformly treated cohort of adult patients. Neuropathology. 2011;31(4):376–83.PubMedCrossRef
96.
Zurück zum Zitat Cancer Genome Atlas Research N. Comprehensive genomic characterization defines human glioblastoma genes and core pathways. Nature. 2008;455(7216):1061–8.CrossRef Cancer Genome Atlas Research N. Comprehensive genomic characterization defines human glioblastoma genes and core pathways. Nature. 2008;455(7216):1061–8.CrossRef
97.
Zurück zum Zitat Gallia GL, Rand V, Siu IM, Eberhart CG, James CD, Marie SK, et al. PIK3CA gene mutations in pediatric and adult glioblastoma multiforme. Mol Cancer Res. 2006;4(10):709–14.PubMedCrossRef Gallia GL, Rand V, Siu IM, Eberhart CG, James CD, Marie SK, et al. PIK3CA gene mutations in pediatric and adult glioblastoma multiforme. Mol Cancer Res. 2006;4(10):709–14.PubMedCrossRef
98.
Zurück zum Zitat Eckerdt FD, Bell JB, Gonzalez C, Oh MS, Perez RE, Mazewski C, et al. Combined PI3Kalpha-mTOR targeting of glioma stem cells. Sci Rep. 2020;10(1):21873.PubMedPubMedCentralCrossRef Eckerdt FD, Bell JB, Gonzalez C, Oh MS, Perez RE, Mazewski C, et al. Combined PI3Kalpha-mTOR targeting of glioma stem cells. Sci Rep. 2020;10(1):21873.PubMedPubMedCentralCrossRef
99.
Zurück zum Zitat de Gooijer MC, Zhang P, Buil LCM, Çitirikkaya CH, Thota N, Beijnen JH, et al. Buparlisib is a brain penetrable pan-PI3K inhibitor. Sci Rep. 2018;8(1):10784.PubMedPubMedCentralCrossRef de Gooijer MC, Zhang P, Buil LCM, Çitirikkaya CH, Thota N, Beijnen JH, et al. Buparlisib is a brain penetrable pan-PI3K inhibitor. Sci Rep. 2018;8(1):10784.PubMedPubMedCentralCrossRef
100.
Zurück zum Zitat Wen PY, Rodon JA, Mason W, Beck JT, DeGroot J, Donnet V et al. Phase I, open-label, multicentre study of buparlisib in combination with temozolomide or with concomitant radiation therapy and temozolomide in patients with newly diagnosed glioblastoma. ESMO Open. 2020;5(4). Wen PY, Rodon JA, Mason W, Beck JT, DeGroot J, Donnet V et al. Phase I, open-label, multicentre study of buparlisib in combination with temozolomide or with concomitant radiation therapy and temozolomide in patients with newly diagnosed glioblastoma. ESMO Open. 2020;5(4).
101.
Zurück zum Zitat Wen PY, Touat M, Alexander BM, Mellinghoff IK, Ramkissoon S, McCluskey CS, et al. Buparlisib in patients with recurrent glioblastoma harboring phosphatidylinositol 3-Kinase pathway activation: an Open-Label, Multicenter, Multi-arm, Phase II Trial. J Clin Oncol. 2019;37(9):741–50.PubMedPubMedCentralCrossRef Wen PY, Touat M, Alexander BM, Mellinghoff IK, Ramkissoon S, McCluskey CS, et al. Buparlisib in patients with recurrent glioblastoma harboring phosphatidylinositol 3-Kinase pathway activation: an Open-Label, Multicenter, Multi-arm, Phase II Trial. J Clin Oncol. 2019;37(9):741–50.PubMedPubMedCentralCrossRef
102.
Zurück zum Zitat Cloughesy TF, Alexander BM, Berry DA, Colman H, Groot JFd, Ellingson BM, et al. GBM AGILE: a global, phase 2/3 adaptive platform trial to evaluate multiple regimens in newly diagnosed and recurrent glioblastoma. JCO. 2022;40(16suppl):TPS2078–TPS.CrossRef Cloughesy TF, Alexander BM, Berry DA, Colman H, Groot JFd, Ellingson BM, et al. GBM AGILE: a global, phase 2/3 adaptive platform trial to evaluate multiple regimens in newly diagnosed and recurrent glioblastoma. JCO. 2022;40(16suppl):TPS2078–TPS.CrossRef
103.
Zurück zum Zitat Wen PY, Groot JFd, Battiste J, Goldlust SA, Garner JS, Friend J, et al. Paxalisib in patients with newly diagnosed glioblastoma with unmethylated MGMT promoter status: final phase 2 study results. JCO. 2022;40(16suppl):2047.CrossRef Wen PY, Groot JFd, Battiste J, Goldlust SA, Garner JS, Friend J, et al. Paxalisib in patients with newly diagnosed glioblastoma with unmethylated MGMT promoter status: final phase 2 study results. JCO. 2022;40(16suppl):2047.CrossRef
104.
Zurück zum Zitat Hopkins BD, Pauli C, Du X, Wang DG, Li X, Wu D, et al. Suppression of insulin feedback enhances the efficacy of PI3K inhibitors. Nature. 2018;560(7719):499–503.PubMedPubMedCentralCrossRef Hopkins BD, Pauli C, Du X, Wang DG, Li X, Wu D, et al. Suppression of insulin feedback enhances the efficacy of PI3K inhibitors. Nature. 2018;560(7719):499–503.PubMedPubMedCentralCrossRef
105.
Zurück zum Zitat Noch EK, Palma LN, Yim I, Bullen N, Qiu Y, Ravichandran H, et al. Insulin feedback is a targetable resistance mechanism of PI3K inhibition in glioblastoma. Neuro Oncol. 2023;25(12):2165–76.PubMedCrossRef Noch EK, Palma LN, Yim I, Bullen N, Qiu Y, Ravichandran H, et al. Insulin feedback is a targetable resistance mechanism of PI3K inhibition in glioblastoma. Neuro Oncol. 2023;25(12):2165–76.PubMedCrossRef
106.
107.
Zurück zum Zitat Korkolopoulou P, Levidou G, El-Habr EA, Piperi C, Adamopoulos C, Samaras V, et al. Phosphorylated 4E-binding protein 1 (p-4E-BP1): a novel prognostic marker in human astrocytomas. Histopathology. 2012;61(2):293–305.PubMedCrossRef Korkolopoulou P, Levidou G, El-Habr EA, Piperi C, Adamopoulos C, Samaras V, et al. Phosphorylated 4E-binding protein 1 (p-4E-BP1): a novel prognostic marker in human astrocytomas. Histopathology. 2012;61(2):293–305.PubMedCrossRef
108.
Zurück zum Zitat Li XY, Zhang LQ, Zhang XG, Li X, Ren YB, Ma XY, et al. Association between AKT/mTOR signalling pathway and malignancy grade of human gliomas. J Neurooncol. 2011;103(3):453–8.PubMedCrossRef Li XY, Zhang LQ, Zhang XG, Li X, Ren YB, Ma XY, et al. Association between AKT/mTOR signalling pathway and malignancy grade of human gliomas. J Neurooncol. 2011;103(3):453–8.PubMedCrossRef
109.
Zurück zum Zitat Kaley TJ, Panageas KS, Mellinghoff IK, Nolan C, Gavrilovic IT, DeAngelis LM, et al. Phase II trial of an AKT inhibitor (perifosine) for recurrent glioblastoma. J Neurooncol. 2019;144(2):403–7.PubMedPubMedCentralCrossRef Kaley TJ, Panageas KS, Mellinghoff IK, Nolan C, Gavrilovic IT, DeAngelis LM, et al. Phase II trial of an AKT inhibitor (perifosine) for recurrent glioblastoma. J Neurooncol. 2019;144(2):403–7.PubMedPubMedCentralCrossRef
110.
Zurück zum Zitat Narayan RS, Fedrigo CA, Brands E, Dik R, Stalpers LJA, Baumert BG, et al. The allosteric AKT inhibitor MK2206 shows a synergistic interaction with chemotherapy and radiotherapy in glioblastoma spheroid cultures. BMC Cancer. 2017;17(1):204.PubMedPubMedCentralCrossRef Narayan RS, Fedrigo CA, Brands E, Dik R, Stalpers LJA, Baumert BG, et al. The allosteric AKT inhibitor MK2206 shows a synergistic interaction with chemotherapy and radiotherapy in glioblastoma spheroid cultures. BMC Cancer. 2017;17(1):204.PubMedPubMedCentralCrossRef
111.
Zurück zum Zitat Cloughesy TF, Yoshimoto K, Nghiemphu P, Brown K, Dang J, Zhu S, et al. Antitumor activity of rapamycin in a phase I trial for patients with recurrent PTEN-deficient glioblastoma. PLoS Med. 2008;5(1):e8.PubMedPubMedCentralCrossRef Cloughesy TF, Yoshimoto K, Nghiemphu P, Brown K, Dang J, Zhu S, et al. Antitumor activity of rapamycin in a phase I trial for patients with recurrent PTEN-deficient glioblastoma. PLoS Med. 2008;5(1):e8.PubMedPubMedCentralCrossRef
112.
Zurück zum Zitat Reardon DA, Desjardins A, Vredenburgh JJ, Gururangan S, Friedman AH, Herndon JE, et al. Phase 2 trial of erlotinib plus sirolimus in adults with recurrent glioblastoma. J Neurooncol. 2010;96(2):219–30.PubMedCrossRef Reardon DA, Desjardins A, Vredenburgh JJ, Gururangan S, Friedman AH, Herndon JE, et al. Phase 2 trial of erlotinib plus sirolimus in adults with recurrent glioblastoma. J Neurooncol. 2010;96(2):219–30.PubMedCrossRef
113.
Zurück zum Zitat Kim DH, Sarbassov DD, Ali SM, King JE, Latek RR, Erdjument-Bromage H, et al. mTOR interacts with raptor to form a nutrient-sensitive complex that signals to the cell growth machinery. Cell. 2002;110(2):163–75.PubMedCrossRef Kim DH, Sarbassov DD, Ali SM, King JE, Latek RR, Erdjument-Bromage H, et al. mTOR interacts with raptor to form a nutrient-sensitive complex that signals to the cell growth machinery. Cell. 2002;110(2):163–75.PubMedCrossRef
114.
Zurück zum Zitat Kahn J, Hayman TJ, Jamal M, Rath BH, Kramp T, Camphausen K, et al. The mTORC1/mTORC2 inhibitor AZD2014 enhances the radiosensitivity of glioblastoma stem-like cells. Neuro Oncol. 2014;16(1):29–37.PubMedCrossRef Kahn J, Hayman TJ, Jamal M, Rath BH, Kramp T, Camphausen K, et al. The mTORC1/mTORC2 inhibitor AZD2014 enhances the radiosensitivity of glioblastoma stem-like cells. Neuro Oncol. 2014;16(1):29–37.PubMedCrossRef
115.
Zurück zum Zitat Bendell JC, Kelley RK, Shih KC, Grabowsky JA, Bergsland E, Jones S, et al. A phase I dose-escalation study to assess safety, tolerability, pharmacokinetics, and preliminary efficacy of the dual mTORC1/mTORC2 kinase inhibitor CC‐223 in patients with advanced solid tumors or multiple myeloma. Cancer. 2015;121(19):3481–90.PubMedCrossRef Bendell JC, Kelley RK, Shih KC, Grabowsky JA, Bergsland E, Jones S, et al. A phase I dose-escalation study to assess safety, tolerability, pharmacokinetics, and preliminary efficacy of the dual mTORC1/mTORC2 kinase inhibitor CC‐223 in patients with advanced solid tumors or multiple myeloma. Cancer. 2015;121(19):3481–90.PubMedCrossRef
116.
Zurück zum Zitat Tao Z, Li T, Ma H, Yang Y, Zhang C, Hai L, et al. Autophagy suppresses self-renewal ability and tumorigenicity of glioma-initiating cells and promotes Notch1 degradation. Cell Death Dis. 2018;9(11):1063.PubMedPubMedCentralCrossRef Tao Z, Li T, Ma H, Yang Y, Zhang C, Hai L, et al. Autophagy suppresses self-renewal ability and tumorigenicity of glioma-initiating cells and promotes Notch1 degradation. Cell Death Dis. 2018;9(11):1063.PubMedPubMedCentralCrossRef
117.
Zurück zum Zitat Luchman HA, Stechishin OD, Nguyen SA, Lun XQ, Cairncross JG, Weiss S. Dual mTORC1/2 blockade inhibits glioblastoma brain tumor initiating cells in vitro and in vivo and synergizes with temozolomide to increase orthotopic xenograft survival. Clin Cancer Res. 2014;20(22):5756–67.PubMedCrossRef Luchman HA, Stechishin OD, Nguyen SA, Lun XQ, Cairncross JG, Weiss S. Dual mTORC1/2 blockade inhibits glioblastoma brain tumor initiating cells in vitro and in vivo and synergizes with temozolomide to increase orthotopic xenograft survival. Clin Cancer Res. 2014;20(22):5756–67.PubMedCrossRef
118.
Zurück zum Zitat Vehlow A, Klapproth E, Jin S, Hannen R, Hauswald M, Bartsch J-W, et al. Interaction of Discoidin Domain Receptor 1 with a 14-3-3-Beclin-1-Akt1 Complex modulates Glioblastoma Therapy Sensitivity. Cell Rep. 2019;26(13):3672–e837.PubMedCrossRef Vehlow A, Klapproth E, Jin S, Hannen R, Hauswald M, Bartsch J-W, et al. Interaction of Discoidin Domain Receptor 1 with a 14-3-3-Beclin-1-Akt1 Complex modulates Glioblastoma Therapy Sensitivity. Cell Rep. 2019;26(13):3672–e837.PubMedCrossRef
119.
Zurück zum Zitat Han S, Liu Y, Cai SJ, Qian M, Ding J, Larion M, et al. IDH mutation in glioma: molecular mechanisms and potential therapeutic targets. Br J Cancer. 2020;122(11):1580–9.PubMedPubMedCentralCrossRef Han S, Liu Y, Cai SJ, Qian M, Ding J, Larion M, et al. IDH mutation in glioma: molecular mechanisms and potential therapeutic targets. Br J Cancer. 2020;122(11):1580–9.PubMedPubMedCentralCrossRef
120.
Zurück zum Zitat Nobusawa S, Watanabe T, Kleihues P, Ohgaki H. IDH1 mutations as molecular signature and predictive factor of secondary glioblastomas. Clin Cancer Res. 2009;15(19):6002–7.PubMedCrossRef Nobusawa S, Watanabe T, Kleihues P, Ohgaki H. IDH1 mutations as molecular signature and predictive factor of secondary glioblastomas. Clin Cancer Res. 2009;15(19):6002–7.PubMedCrossRef
121.
123.
Zurück zum Zitat Gupta R, Flanagan S, Li CCY, Lee M, Shivalingham B, Maleki S, et al. Expanding the spectrum of IDH1 mutations in gliomas. Mod Pathol. 2013;26(5):619–25.PubMedCrossRef Gupta R, Flanagan S, Li CCY, Lee M, Shivalingham B, Maleki S, et al. Expanding the spectrum of IDH1 mutations in gliomas. Mod Pathol. 2013;26(5):619–25.PubMedCrossRef
124.
Zurück zum Zitat Ceccarelli M, Barthel FP, Malta TM, Sabedot TS, Salama SR, Murray BA, et al. Molecular profiling reveals biologically discrete subsets and pathways of progression in diffuse glioma. Cell. 2016;164(3):550–63.PubMedPubMedCentralCrossRef Ceccarelli M, Barthel FP, Malta TM, Sabedot TS, Salama SR, Murray BA, et al. Molecular profiling reveals biologically discrete subsets and pathways of progression in diffuse glioma. Cell. 2016;164(3):550–63.PubMedPubMedCentralCrossRef
125.
Zurück zum Zitat Turcan S, Makarov V, Taranda J, Wang Y, Fabius AWM, Wu W, et al. Mutant-IDH1-dependent chromatin state reprogramming, reversibility, and persistence. Nat Genet. 2018;50(1):62–72.PubMedCrossRef Turcan S, Makarov V, Taranda J, Wang Y, Fabius AWM, Wu W, et al. Mutant-IDH1-dependent chromatin state reprogramming, reversibility, and persistence. Nat Genet. 2018;50(1):62–72.PubMedCrossRef
127.
Zurück zum Zitat Xu W, Yang H, Liu Y, Yang Y, Wang P, Kim S-H, et al. Oncometabolite 2-hydroxyglutarate is a competitive inhibitor of α-ketoglutarate-dependent dioxygenases. Cancer Cell. 2011;19(1):17–30.PubMedPubMedCentralCrossRef Xu W, Yang H, Liu Y, Yang Y, Wang P, Kim S-H, et al. Oncometabolite 2-hydroxyglutarate is a competitive inhibitor of α-ketoglutarate-dependent dioxygenases. Cancer Cell. 2011;19(1):17–30.PubMedPubMedCentralCrossRef
128.
Zurück zum Zitat Xu W, Yang H, Liu Y, Yang Y, Wang P, Kim SH, et al. Oncometabolite 2-hydroxyglutarate is a competitive inhibitor of alpha-ketoglutarate-dependent dioxygenases. Cancer Cell. 2011;19(1):17–30.PubMedPubMedCentralCrossRef Xu W, Yang H, Liu Y, Yang Y, Wang P, Kim SH, et al. Oncometabolite 2-hydroxyglutarate is a competitive inhibitor of alpha-ketoglutarate-dependent dioxygenases. Cancer Cell. 2011;19(1):17–30.PubMedPubMedCentralCrossRef
129.
Zurück zum Zitat Turcan S, Rohle D, Goenka A, Walsh LA, Fang F, Yilmaz E, et al. IDH1 mutation is sufficient to establish the glioma hypermethylator phenotype. Nature. 2012;483(7390):479–83.PubMedPubMedCentralCrossRef Turcan S, Rohle D, Goenka A, Walsh LA, Fang F, Yilmaz E, et al. IDH1 mutation is sufficient to establish the glioma hypermethylator phenotype. Nature. 2012;483(7390):479–83.PubMedPubMedCentralCrossRef
130.
Zurück zum Zitat Li S, Chowdhury R, Liu F, Chou AP, Li T, Mody RR, et al. Tumor-suppressive miR148a is silenced by CpG island hypermethylation in IDH1-mutant gliomas. Clin Cancer Res. 2014;20(22):5808–22.PubMedPubMedCentralCrossRef Li S, Chowdhury R, Liu F, Chou AP, Li T, Mody RR, et al. Tumor-suppressive miR148a is silenced by CpG island hypermethylation in IDH1-mutant gliomas. Clin Cancer Res. 2014;20(22):5808–22.PubMedPubMedCentralCrossRef
131.
Zurück zum Zitat Flavahan WA, Drier Y, Liau BB, Gillespie SM, Venteicher AS, Stemmer-Rachamimov AO, et al. Insulator dysfunction and oncogene activation in IDH mutant gliomas. Nature. 2016;529(7584):110–4.PubMedCrossRef Flavahan WA, Drier Y, Liau BB, Gillespie SM, Venteicher AS, Stemmer-Rachamimov AO, et al. Insulator dysfunction and oncogene activation in IDH mutant gliomas. Nature. 2016;529(7584):110–4.PubMedCrossRef
132.
Zurück zum Zitat Garcia MG, Carella A, Urdinguio RG, Bayon GF, Lopez V, Tejedor JR, et al. Epigenetic dysregulation of TET2 in human glioblastoma. Oncotarget. 2018;9(40):25922–34.PubMedPubMedCentralCrossRef Garcia MG, Carella A, Urdinguio RG, Bayon GF, Lopez V, Tejedor JR, et al. Epigenetic dysregulation of TET2 in human glioblastoma. Oncotarget. 2018;9(40):25922–34.PubMedPubMedCentralCrossRef
133.
Zurück zum Zitat Bady P, Sciuscio D, Diserens AC, Bloch J, van den Bent MJ, Marosi C, et al. MGMT methylation analysis of glioblastoma on the Infinium methylation BeadChip identifies two distinct CpG regions associated with gene silencing and outcome, yielding a prediction model for comparisons across datasets, tumor grades, and CIMP-status. Acta Neuropathol. 2012;124(4):547–60.PubMedPubMedCentralCrossRef Bady P, Sciuscio D, Diserens AC, Bloch J, van den Bent MJ, Marosi C, et al. MGMT methylation analysis of glioblastoma on the Infinium methylation BeadChip identifies two distinct CpG regions associated with gene silencing and outcome, yielding a prediction model for comparisons across datasets, tumor grades, and CIMP-status. Acta Neuropathol. 2012;124(4):547–60.PubMedPubMedCentralCrossRef
134.
Zurück zum Zitat Malley DS, Hamoudi RA, Kocialkowski S, Pearson DM, Collins VP, Ichimura K. A distinct region of the MGMT CpG island critical for transcriptional regulation is preferentially methylated in glioblastoma cells and xenografts. Acta Neuropathol. 2011;121(5):651–61.PubMedCrossRef Malley DS, Hamoudi RA, Kocialkowski S, Pearson DM, Collins VP, Ichimura K. A distinct region of the MGMT CpG island critical for transcriptional regulation is preferentially methylated in glioblastoma cells and xenografts. Acta Neuropathol. 2011;121(5):651–61.PubMedCrossRef
135.
Zurück zum Zitat Hegi ME, Diserens AC, Gorlia T, Hamou MF, de Tribolet N, Weller M, et al. MGMT gene silencing and benefit from temozolomide in glioblastoma. N Engl J Med. 2005;352(10):997–1003.PubMedCrossRef Hegi ME, Diserens AC, Gorlia T, Hamou MF, de Tribolet N, Weller M, et al. MGMT gene silencing and benefit from temozolomide in glioblastoma. N Engl J Med. 2005;352(10):997–1003.PubMedCrossRef
136.
Zurück zum Zitat van den Bent MJ, Dubbink HJ, Sanson M, van der Lee-Haarloo CR, Hegi M, Jeuken JW, et al. MGMT promoter methylation is prognostic but not predictive for outcome to adjuvant PCV chemotherapy in anaplastic oligodendroglial tumors: a report from EORTC Brain Tumor Group Study 26951. J Clin Oncol. 2009;27(35):5881–6.PubMedPubMedCentralCrossRef van den Bent MJ, Dubbink HJ, Sanson M, van der Lee-Haarloo CR, Hegi M, Jeuken JW, et al. MGMT promoter methylation is prognostic but not predictive for outcome to adjuvant PCV chemotherapy in anaplastic oligodendroglial tumors: a report from EORTC Brain Tumor Group Study 26951. J Clin Oncol. 2009;27(35):5881–6.PubMedPubMedCentralCrossRef
137.
Zurück zum Zitat Kadiyala P, Carney SV, Gauss JC, Garcia-Fabiani MB, Haase S, Alghamri MS et al. Inhibition of 2-hydroxyglutarate elicits metabolic reprogramming and mutant IDH1 glioma immunity in mice. J Clin Invest. 2021;131(4). Kadiyala P, Carney SV, Gauss JC, Garcia-Fabiani MB, Haase S, Alghamri MS et al. Inhibition of 2-hydroxyglutarate elicits metabolic reprogramming and mutant IDH1 glioma immunity in mice. J Clin Invest. 2021;131(4).
138.
Zurück zum Zitat Wick W, Meisner C, Hentschel B, Platten M, Schilling A, Wiestler B, et al. Prognostic or predictive value of MGMT promoter methylation in gliomas depends on IDH1 mutation. Neurology. 2013;81(17):1515–22.PubMedCrossRef Wick W, Meisner C, Hentschel B, Platten M, Schilling A, Wiestler B, et al. Prognostic or predictive value of MGMT promoter methylation in gliomas depends on IDH1 mutation. Neurology. 2013;81(17):1515–22.PubMedCrossRef
139.
Zurück zum Zitat Noushmehr H, Weisenberger DJ, Diefes K, Phillips HS, Pujara K, Berman BP, et al. Identification of a CpG island methylator phenotype that defines a distinct subgroup of glioma. Cancer Cell. 2010;17(5):510–22.PubMedPubMedCentralCrossRef Noushmehr H, Weisenberger DJ, Diefes K, Phillips HS, Pujara K, Berman BP, et al. Identification of a CpG island methylator phenotype that defines a distinct subgroup of glioma. Cancer Cell. 2010;17(5):510–22.PubMedPubMedCentralCrossRef
140.
Zurück zum Zitat Sule A, Van Doorn J, Sundaram RK, Ganesa S, Vasquez Juan C, Bindra Ranjit S. Targeting IDH1/2 mutant cancers with combinations of ATR and PARP inhibitors. NAR Cancer. 2021;3(2):zcab018.PubMedPubMedCentralCrossRef Sule A, Van Doorn J, Sundaram RK, Ganesa S, Vasquez Juan C, Bindra Ranjit S. Targeting IDH1/2 mutant cancers with combinations of ATR and PARP inhibitors. NAR Cancer. 2021;3(2):zcab018.PubMedPubMedCentralCrossRef
141.
Zurück zum Zitat Sulkowski PL, Corso CD, Robinson ND, Scanlon SE, Purshouse KR, Bai H, et al. 2-Hydroxyglutarate produced by neomorphic IDH mutations suppresses homologous recombination and induces PARP inhibitor sensitivity. Sci Transl Med. 2017;9(375):eaal2463.PubMedPubMedCentralCrossRef Sulkowski PL, Corso CD, Robinson ND, Scanlon SE, Purshouse KR, Bai H, et al. 2-Hydroxyglutarate produced by neomorphic IDH mutations suppresses homologous recombination and induces PARP inhibitor sensitivity. Sci Transl Med. 2017;9(375):eaal2463.PubMedPubMedCentralCrossRef
142.
Zurück zum Zitat Lu Y, Kwintkiewicz J, Liu Y, Tech K, Frady LN, Su Y-T, et al. Chemosensitivity of IDH1-Mutated gliomas due to an impairment in PARP1-Mediated DNA repair. Cancer Res. 2017;77(7):1709–18.PubMedPubMedCentralCrossRef Lu Y, Kwintkiewicz J, Liu Y, Tech K, Frady LN, Su Y-T, et al. Chemosensitivity of IDH1-Mutated gliomas due to an impairment in PARP1-Mediated DNA repair. Cancer Res. 2017;77(7):1709–18.PubMedPubMedCentralCrossRef
143.
Zurück zum Zitat Halford SE, Cruickshank G, Dunn L, Erridge S, Godfrey L, Herbert C, et al. Results of the OPARATIC trial: A phase I dose escalation study of olaparib in combination with temozolomide (TMZ) in patients with relapsed glioblastoma (GBM). American Society of Clinical Oncology; 2017. Halford SE, Cruickshank G, Dunn L, Erridge S, Godfrey L, Herbert C, et al. Results of the OPARATIC trial: A phase I dose escalation study of olaparib in combination with temozolomide (TMZ) in patients with relapsed glioblastoma (GBM). American Society of Clinical Oncology; 2017.
144.
Zurück zum Zitat Fanucci K, Pilat MJP, Shah R, Boerner SA, Li J, Durecki DE, et al. Multicenter phase 2 trial of the PARP inhibitor (PARPi) olaparib in recurrent IDH1 and IDH2-mutant contrast-enhancing glioma. American Society of Clinical Oncology; 2022. Fanucci K, Pilat MJP, Shah R, Boerner SA, Li J, Durecki DE, et al. Multicenter phase 2 trial of the PARP inhibitor (PARPi) olaparib in recurrent IDH1 and IDH2-mutant contrast-enhancing glioma. American Society of Clinical Oncology; 2022.
145.
Zurück zum Zitat Sun K, Mikule K, Wang Z, Poon G, Vaidyanathan A, Smith G, et al. A comparative pharmacokinetic study of PARP inhibitors demonstrates favorable properties for niraparib efficacy in preclinical tumor models. Oncotarget. 2018;9(98):37080–96.PubMedPubMedCentralCrossRef Sun K, Mikule K, Wang Z, Poon G, Vaidyanathan A, Smith G, et al. A comparative pharmacokinetic study of PARP inhibitors demonstrates favorable properties for niraparib efficacy in preclinical tumor models. Oncotarget. 2018;9(98):37080–96.PubMedPubMedCentralCrossRef
146.
Zurück zum Zitat Kurzrock R, Galanis E, Johnson DR, Kansra V, Wilcoxen K, McClure T, et al. A phase I study of niraparib in combination with temozolomide (TMZ) in patients with advanced cancer. JCO. 2014;32(15suppl):2092.CrossRef Kurzrock R, Galanis E, Johnson DR, Kansra V, Wilcoxen K, McClure T, et al. A phase I study of niraparib in combination with temozolomide (TMZ) in patients with advanced cancer. JCO. 2014;32(15suppl):2092.CrossRef
147.
Zurück zum Zitat Dungey FA, Löser DA, Chalmers AJ. Replication-dependent radiosensitization of human glioma cells by inhibition of poly(ADP-Ribose) polymerase: mechanisms and therapeutic potential. Int J Radiat Oncol Biol Phys. 2008;72(4):1188–97.PubMedCrossRef Dungey FA, Löser DA, Chalmers AJ. Replication-dependent radiosensitization of human glioma cells by inhibition of poly(ADP-Ribose) polymerase: mechanisms and therapeutic potential. Int J Radiat Oncol Biol Phys. 2008;72(4):1188–97.PubMedCrossRef
148.
Zurück zum Zitat Carruthers RD, Ahmed SU, Ramachandran S, Strathdee K, Kurian KM, Hedley A, et al. Replication stress drives constitutive activation of the DNA damage response and radioresistance in glioblastoma stem-like cells. Cancer Res. 2018;78(17):5060–71.PubMedPubMedCentralCrossRef Carruthers RD, Ahmed SU, Ramachandran S, Strathdee K, Kurian KM, Hedley A, et al. Replication stress drives constitutive activation of the DNA damage response and radioresistance in glioblastoma stem-like cells. Cancer Res. 2018;78(17):5060–71.PubMedPubMedCentralCrossRef
149.
Zurück zum Zitat Albert JM, Cao C, Kim KW, Willey CD, Geng L, Xiao D, et al. Inhibition of poly(ADP-ribose) polymerase enhances cell death and improves tumor growth delay in irradiated lung cancer models. Clin Cancer Research: Official J Am Association Cancer Res. 2007;13(10):3033–42.CrossRef Albert JM, Cao C, Kim KW, Willey CD, Geng L, Xiao D, et al. Inhibition of poly(ADP-ribose) polymerase enhances cell death and improves tumor growth delay in irradiated lung cancer models. Clin Cancer Research: Official J Am Association Cancer Res. 2007;13(10):3033–42.CrossRef
150.
Zurück zum Zitat Bao S, Wu Q, McLendon RE, Hao Y, Shi Q, Hjelmeland AB, et al. Glioma stem cells promote radioresistance by preferential activation of the DNA damage response. Nature. 2006;444(7120):756–60.PubMedCrossRef Bao S, Wu Q, McLendon RE, Hao Y, Shi Q, Hjelmeland AB, et al. Glioma stem cells promote radioresistance by preferential activation of the DNA damage response. Nature. 2006;444(7120):756–60.PubMedCrossRef
151.
Zurück zum Zitat Jiang W, Wang Z. CTNI-63. A STUDY OF NIRAPARIB, COMBINED WITH RADIOTHERAPY IN PATIENTS WITH RECURRENT GLIOBLASTOMA. Neuro Oncol. 2022;24(Supplement7):vii87–vii.CrossRef Jiang W, Wang Z. CTNI-63. A STUDY OF NIRAPARIB, COMBINED WITH RADIOTHERAPY IN PATIENTS WITH RECURRENT GLIOBLASTOMA. Neuro Oncol. 2022;24(Supplement7):vii87–vii.CrossRef
152.
Zurück zum Zitat Helleday T. The underlying mechanism for the PARP and BRCA synthetic lethality: Clearing up the misunderstandings. Mol Oncol. 2011;5(4):387–93.PubMedPubMedCentralCrossRef Helleday T. The underlying mechanism for the PARP and BRCA synthetic lethality: Clearing up the misunderstandings. Mol Oncol. 2011;5(4):387–93.PubMedPubMedCentralCrossRef
153.
Zurück zum Zitat Mellinghoff IK, Ellingson BM, Touat M, Maher E, Fuente MIDL, Holdhoff M, et al. Ivosidenib in Isocitrate dehydrogenase 1–Mutated Advanced Glioma. JCO. 2020;38(29):3398–406.CrossRef Mellinghoff IK, Ellingson BM, Touat M, Maher E, Fuente MIDL, Holdhoff M, et al. Ivosidenib in Isocitrate dehydrogenase 1–Mutated Advanced Glioma. JCO. 2020;38(29):3398–406.CrossRef
154.
Zurück zum Zitat Mellinghoff IK, Lu M, Wen PY, Taylor JW, Maher EA, Arrillaga-Romany I, et al. Vorasidenib and Ivosidenib in IDH1-mutant low-grade glioma: a randomized, perioperative phase 1 trial. Nat Med. 2023;29(3):615–22.PubMedPubMedCentralCrossRef Mellinghoff IK, Lu M, Wen PY, Taylor JW, Maher EA, Arrillaga-Romany I, et al. Vorasidenib and Ivosidenib in IDH1-mutant low-grade glioma: a randomized, perioperative phase 1 trial. Nat Med. 2023;29(3):615–22.PubMedPubMedCentralCrossRef
155.
Zurück zum Zitat Peters KB, Alford C, Heltemes A, Savelli A, Landi DB, Broadwater G et al. Use, access, and initial outcomes of off-label ivosidenib in patients with IDH1 mutant glioma. Neuro-Oncology Pract. 2023. Peters KB, Alford C, Heltemes A, Savelli A, Landi DB, Broadwater G et al. Use, access, and initial outcomes of off-label ivosidenib in patients with IDH1 mutant glioma. Neuro-Oncology Pract. 2023.
156.
Zurück zum Zitat Ribeiro A, Ney D, INNV-11. USE OF IVOSIDENIB IN THE TREATMENT OF LOW GRADE GLIOMAS. Neuro Oncol. 2023;25(Supplement5):v158–v.CrossRef Ribeiro A, Ney D, INNV-11. USE OF IVOSIDENIB IN THE TREATMENT OF LOW GRADE GLIOMAS. Neuro Oncol. 2023;25(Supplement5):v158–v.CrossRef
157.
Zurück zum Zitat Tejera D, Kushnirsky M, Gultekin SH, Lu M, Steelman L, de la Fuente MI. Ivosidenib, an IDH1 inhibitor, in a patient with recurrent, IDH1-mutant glioblastoma: a case report from a phase I study. CNS Oncol. 2020;9(3):CNS62.PubMedPubMedCentralCrossRef Tejera D, Kushnirsky M, Gultekin SH, Lu M, Steelman L, de la Fuente MI. Ivosidenib, an IDH1 inhibitor, in a patient with recurrent, IDH1-mutant glioblastoma: a case report from a phase I study. CNS Oncol. 2020;9(3):CNS62.PubMedPubMedCentralCrossRef
158.
Zurück zum Zitat Konstantinova IM, Tsimokha AS, Mittenberg AG. Role of proteasomes in cellular regulation. Int Rev Cell Mol Biol. 2008;267:59–124.PubMedCrossRef Konstantinova IM, Tsimokha AS, Mittenberg AG. Role of proteasomes in cellular regulation. Int Rev Cell Mol Biol. 2008;267:59–124.PubMedCrossRef
159.
Zurück zum Zitat Walerych D, Lisek K, Sommaggio R, Piazza S, Ciani Y, Dalla E, et al. Proteasome machinery is instrumental in a common gain-of-function program of the p53 missense mutants in cancer. Nat Cell Biol. 2016;18(8):897–909.PubMedCrossRef Walerych D, Lisek K, Sommaggio R, Piazza S, Ciani Y, Dalla E, et al. Proteasome machinery is instrumental in a common gain-of-function program of the p53 missense mutants in cancer. Nat Cell Biol. 2016;18(8):897–909.PubMedCrossRef
161.
Zurück zum Zitat Love IM, Shi D, Grossman SR. p53 ubiquitination and proteasomal degradation. Methods Mol Biol. 2013;962:63–73.PubMedCrossRef Love IM, Shi D, Grossman SR. p53 ubiquitination and proteasomal degradation. Methods Mol Biol. 2013;962:63–73.PubMedCrossRef
162.
Zurück zum Zitat Orlowski RZ, Kuhn DJ. Proteasome inhibitors in cancer therapy: lessons from the first decade. Clin Cancer Res. 2008;14(6):1649–57.PubMedCrossRef Orlowski RZ, Kuhn DJ. Proteasome inhibitors in cancer therapy: lessons from the first decade. Clin Cancer Res. 2008;14(6):1649–57.PubMedCrossRef
163.
Zurück zum Zitat Boccellato C, Kolbe E, Peters N, Juric V, Fullstone G, Verreault M, et al. Marizomib sensitizes primary glioma cells to apoptosis induced by a latest-generation TRAIL receptor agonist. Cell Death Dis. 2021;12(7):647.PubMedPubMedCentralCrossRef Boccellato C, Kolbe E, Peters N, Juric V, Fullstone G, Verreault M, et al. Marizomib sensitizes primary glioma cells to apoptosis induced by a latest-generation TRAIL receptor agonist. Cell Death Dis. 2021;12(7):647.PubMedPubMedCentralCrossRef
164.
Zurück zum Zitat Fan WH, Hou Y, Meng FK, Wang XF, Luo YN, Ge PF. Proteasome inhibitor MG-132 induces C6 glioma cell apoptosis via oxidative stress. Acta Pharmacol Sin. 2011;32(5):619–25.PubMedPubMedCentralCrossRef Fan WH, Hou Y, Meng FK, Wang XF, Luo YN, Ge PF. Proteasome inhibitor MG-132 induces C6 glioma cell apoptosis via oxidative stress. Acta Pharmacol Sin. 2011;32(5):619–25.PubMedPubMedCentralCrossRef
165.
Zurück zum Zitat Jane EP, Premkumar DR, Pollack IF. Bortezomib sensitizes malignant human glioma cells to TRAIL, mediated by inhibition of the NF-kappaB signaling pathway. Mol Cancer Ther. 2011;10(1):198–208.PubMedPubMedCentralCrossRef Jane EP, Premkumar DR, Pollack IF. Bortezomib sensitizes malignant human glioma cells to TRAIL, mediated by inhibition of the NF-kappaB signaling pathway. Mol Cancer Ther. 2011;10(1):198–208.PubMedPubMedCentralCrossRef
166.
Zurück zum Zitat Manton CA, Johnson B, Singh M, Bailey CP, Bouchier-Hayes L, Chandra J. Induction of cell death by the novel proteasome inhibitor marizomib in glioblastoma in vitro and in vivo. Sci Rep. 2016;6:18953.PubMedPubMedCentralCrossRef Manton CA, Johnson B, Singh M, Bailey CP, Bouchier-Hayes L, Chandra J. Induction of cell death by the novel proteasome inhibitor marizomib in glioblastoma in vitro and in vivo. Sci Rep. 2016;6:18953.PubMedPubMedCentralCrossRef
167.
Zurück zum Zitat Penaranda Fajardo NM, Meijer C, Kruyt FA. The endoplasmic reticulum stress/unfolded protein response in gliomagenesis, tumor progression and as a therapeutic target in glioblastoma. Biochem Pharmacol. 2016;118:1–8.PubMedCrossRef Penaranda Fajardo NM, Meijer C, Kruyt FA. The endoplasmic reticulum stress/unfolded protein response in gliomagenesis, tumor progression and as a therapeutic target in glioblastoma. Biochem Pharmacol. 2016;118:1–8.PubMedCrossRef
168.
Zurück zum Zitat Yin D, Zhou H, Kumagai T, Liu G, Ong JM, Black KL, et al. Proteasome inhibitor PS-341 causes cell growth arrest and apoptosis in human glioblastoma multiforme (GBM). Oncogene. 2005;24(3):344–54.PubMedCrossRef Yin D, Zhou H, Kumagai T, Liu G, Ong JM, Black KL, et al. Proteasome inhibitor PS-341 causes cell growth arrest and apoptosis in human glioblastoma multiforme (GBM). Oncogene. 2005;24(3):344–54.PubMedCrossRef
169.
Zurück zum Zitat Kane RC, Bross PF, Farrell AT, Pazdur R, Velcade. U.S. FDA approval for the treatment of multiple myeloma progressing on prior therapy. Oncologist. 2003;8(6):508–13.PubMedCrossRef Kane RC, Bross PF, Farrell AT, Pazdur R, Velcade. U.S. FDA approval for the treatment of multiple myeloma progressing on prior therapy. Oncologist. 2003;8(6):508–13.PubMedCrossRef
170.
Zurück zum Zitat Kong XT, Nguyen NT, Choi YJ, Zhang G, Nguyen HN, Filka E, et al. Phase 2 study of Bortezomib Combined with Temozolomide and Regional Radiation Therapy for Upfront treatment of patients with newly diagnosed Glioblastoma Multiforme: Safety and Efficacy Assessment. Int J Radiat Oncol Biol Phys. 2018;100(5):1195–203.PubMedPubMedCentralCrossRef Kong XT, Nguyen NT, Choi YJ, Zhang G, Nguyen HN, Filka E, et al. Phase 2 study of Bortezomib Combined with Temozolomide and Regional Radiation Therapy for Upfront treatment of patients with newly diagnosed Glioblastoma Multiforme: Safety and Efficacy Assessment. Int J Radiat Oncol Biol Phys. 2018;100(5):1195–203.PubMedPubMedCentralCrossRef
171.
Zurück zum Zitat Huehnchen P, Springer A, Kern J, Kopp U, Kohler S, Alexander T, et al. Bortezomib at therapeutic doses poorly passes the blood–brain barrier and does not impair cognition. Brain Commun. 2020;2(1):fcaa021.PubMedPubMedCentralCrossRef Huehnchen P, Springer A, Kern J, Kopp U, Kohler S, Alexander T, et al. Bortezomib at therapeutic doses poorly passes the blood–brain barrier and does not impair cognition. Brain Commun. 2020;2(1):fcaa021.PubMedPubMedCentralCrossRef
172.
Zurück zum Zitat Di K, Lloyd GK, Abraham V, MacLaren A, Burrows FJ, Desjardins A, et al. Marizomib activity as a single agent in malignant gliomas: ability to cross the blood-brain barrier. Neuro Oncol. 2016;18(6):840–8.PubMedCrossRef Di K, Lloyd GK, Abraham V, MacLaren A, Burrows FJ, Desjardins A, et al. Marizomib activity as a single agent in malignant gliomas: ability to cross the blood-brain barrier. Neuro Oncol. 2016;18(6):840–8.PubMedCrossRef
173.
Zurück zum Zitat Manton CA, Johnson B, Singh M, Bailey CP, Bouchier-Hayes L, Chandra J. Induction of cell death by the novel proteasome inhibitor marizomib in glioblastoma in vitro and in vivo. Sci Rep. 2016;6(1):18953.PubMedPubMedCentralCrossRef Manton CA, Johnson B, Singh M, Bailey CP, Bouchier-Hayes L, Chandra J. Induction of cell death by the novel proteasome inhibitor marizomib in glioblastoma in vitro and in vivo. Sci Rep. 2016;6(1):18953.PubMedPubMedCentralCrossRef
174.
Zurück zum Zitat Boccellato C, Kolbe E, Peters N, Juric V, Fullstone G, Verreault M, et al. Marizomib sensitizes primary glioma cells to apoptosis induced by a latest-generation TRAIL receptor agonist. Cell Death Dis. 2021;12(7):1–11.CrossRef Boccellato C, Kolbe E, Peters N, Juric V, Fullstone G, Verreault M, et al. Marizomib sensitizes primary glioma cells to apoptosis induced by a latest-generation TRAIL receptor agonist. Cell Death Dis. 2021;12(7):1–11.CrossRef
175.
Zurück zum Zitat Bota DA, Kesari S, Piccioni DE, Aregawi DG, Roth P, Stupp R, et al. A phase 1, multicenter, open-label study of marizomib (MRZ) with temozolomide (TMZ) and radiotherapy (RT) in newly diagnosed WHO grade IV malignant glioma (glioblastoma, ndGBM): dose-escalation results. American Society of Clinical Oncology; 2018. Bota DA, Kesari S, Piccioni DE, Aregawi DG, Roth P, Stupp R, et al. A phase 1, multicenter, open-label study of marizomib (MRZ) with temozolomide (TMZ) and radiotherapy (RT) in newly diagnosed WHO grade IV malignant glioma (glioblastoma, ndGBM): dose-escalation results. American Society of Clinical Oncology; 2018.
176.
Zurück zum Zitat Bota DA, Mason W, Kesari S, Magge R, Winograd B, Elias I, et al. Marizomib alone or in combination with bevacizumab in patients with recurrent glioblastoma: phase I/II clinical trial data. Neuro-Oncology Adv. 2021;3(1):vdab142.CrossRef Bota DA, Mason W, Kesari S, Magge R, Winograd B, Elias I, et al. Marizomib alone or in combination with bevacizumab in patients with recurrent glioblastoma: phase I/II clinical trial data. Neuro-Oncology Adv. 2021;3(1):vdab142.CrossRef
177.
Zurück zum Zitat Roth P, Gorlia T, Reijneveld JC, De Vos FYFL, Idbaih A, Frenel J-S, et al. EORTC 1709/CCTG CE. 8: a phase III trial of marizomib in combination with temozolomide-based radiochemotherapy versus temozolomide-based radiochemotherapy alone in patients with newly diagnosed glioblastoma. Wolters Kluwer Health; 2021. Roth P, Gorlia T, Reijneveld JC, De Vos FYFL, Idbaih A, Frenel J-S, et al. EORTC 1709/CCTG CE. 8: a phase III trial of marizomib in combination with temozolomide-based radiochemotherapy versus temozolomide-based radiochemotherapy alone in patients with newly diagnosed glioblastoma. Wolters Kluwer Health; 2021.
178.
Zurück zum Zitat Yu Z, Wang F, Milacic V, Li X, Cui QC, Zhang B, et al. Evaluation of copper-dependent proteasome-inhibitory and apoptosis-inducing activities of novel pyrrolidine dithiocarbamate analogues. Int J Mol Med. 2007;20(6):919–25.PubMed Yu Z, Wang F, Milacic V, Li X, Cui QC, Zhang B, et al. Evaluation of copper-dependent proteasome-inhibitory and apoptosis-inducing activities of novel pyrrolidine dithiocarbamate analogues. Int J Mol Med. 2007;20(6):919–25.PubMed
179.
Zurück zum Zitat Daniel KG, Gupta P, Harbach RH, Guida WC, Dou QP. Organic copper complexes as a new class of proteasome inhibitors and apoptosis inducers in human cancer cells. Biochem Pharmacol. 2004;67(6):1139–51.PubMedCrossRef Daniel KG, Gupta P, Harbach RH, Guida WC, Dou QP. Organic copper complexes as a new class of proteasome inhibitors and apoptosis inducers in human cancer cells. Biochem Pharmacol. 2004;67(6):1139–51.PubMedCrossRef
180.
Zurück zum Zitat Skrott Z, Mistrik M, Andersen KK, Friis S, Majera D, Gursky J, et al. Alcohol-abuse drug disulfiram targets cancer via p97 segregase adapter NPL4. Nature. 2017;552(7684):194–9.PubMedPubMedCentralCrossRef Skrott Z, Mistrik M, Andersen KK, Friis S, Majera D, Gursky J, et al. Alcohol-abuse drug disulfiram targets cancer via p97 segregase adapter NPL4. Nature. 2017;552(7684):194–9.PubMedPubMedCentralCrossRef
181.
Zurück zum Zitat Hothi P, Martins TJ, Chen L, Deleyrolle L, Yoon J-G, Reynolds B, et al. High-Throughput Chemical screens identify disulfiram as an inhibitor of human glioblastoma stem cells. Oncotarget. 2012;3(10):1124–36.PubMedPubMedCentralCrossRef Hothi P, Martins TJ, Chen L, Deleyrolle L, Yoon J-G, Reynolds B, et al. High-Throughput Chemical screens identify disulfiram as an inhibitor of human glioblastoma stem cells. Oncotarget. 2012;3(10):1124–36.PubMedPubMedCentralCrossRef
182.
Zurück zum Zitat Liu P, Brown S, Goktug T, Channathodiyil P, Kannappan V, Hugnot JP, et al. Cytotoxic effect of disulfiram/copper on human glioblastoma cell lines and ALDH-positive cancer-stem-like cells. Br J Cancer. 2012;107(9):1488–97.PubMedPubMedCentralCrossRef Liu P, Brown S, Goktug T, Channathodiyil P, Kannappan V, Hugnot JP, et al. Cytotoxic effect of disulfiram/copper on human glioblastoma cell lines and ALDH-positive cancer-stem-like cells. Br J Cancer. 2012;107(9):1488–97.PubMedPubMedCentralCrossRef
183.
Zurück zum Zitat Huang J, Chaudhary R, Cohen AL, Fink K, Goldlust S, Boockvar J, et al. A multicenter phase II study of temozolomide plus disulfiram and copper for recurrent temozolomide-resistant glioblastoma. J Neurooncol. 2019;142(3):537–44.PubMedCrossRef Huang J, Chaudhary R, Cohen AL, Fink K, Goldlust S, Boockvar J, et al. A multicenter phase II study of temozolomide plus disulfiram and copper for recurrent temozolomide-resistant glioblastoma. J Neurooncol. 2019;142(3):537–44.PubMedCrossRef
184.
Zurück zum Zitat Huang J, DeWees T, Campian JL, Chheda MG, Ansstas G, Tsien C, et al. A TITE-CRM phase I/II study of disulfiram and copper with concurrent radiation therapy and temozolomide for newly diagnosed glioblastoma. American Society of Clinical Oncology; 2019. Huang J, DeWees T, Campian JL, Chheda MG, Ansstas G, Tsien C, et al. A TITE-CRM phase I/II study of disulfiram and copper with concurrent radiation therapy and temozolomide for newly diagnosed glioblastoma. American Society of Clinical Oncology; 2019.
185.
Zurück zum Zitat Kelley KC, Grossman KF, Brittain-Blankenship M, Thorne KM, Akerley WL, Terrazas MC, et al. A phase 1 dose-escalation study of disulfiram and copper gluconate in patients with advanced solid tumors involving the liver using S-glutathionylation as a biomarker. BMC Cancer. 2021;21(1):510.PubMedPubMedCentralCrossRef Kelley KC, Grossman KF, Brittain-Blankenship M, Thorne KM, Akerley WL, Terrazas MC, et al. A phase 1 dose-escalation study of disulfiram and copper gluconate in patients with advanced solid tumors involving the liver using S-glutathionylation as a biomarker. BMC Cancer. 2021;21(1):510.PubMedPubMedCentralCrossRef
186.
Zurück zum Zitat Werlenius K, Kinhult S, Solheim TS, Magelssen H, Lofgren D, Mudaisi M, et al. Effect of Disulfiram and Copper Plus Chemotherapy vs Chemotherapy Alone on survival in patients with recurrent glioblastoma: a Randomized Clinical Trial. JAMA Netw Open. 2023;6(3):e234149.PubMedPubMedCentralCrossRef Werlenius K, Kinhult S, Solheim TS, Magelssen H, Lofgren D, Mudaisi M, et al. Effect of Disulfiram and Copper Plus Chemotherapy vs Chemotherapy Alone on survival in patients with recurrent glioblastoma: a Randomized Clinical Trial. JAMA Netw Open. 2023;6(3):e234149.PubMedPubMedCentralCrossRef
187.
Zurück zum Zitat Zheng Z, Liebers M, Zhelyazkova B, Cao Y, Panditi D, Lynch KD, et al. Anchored multiplex PCR for targeted next-generation sequencing. Nat Med. 2014;20(12):1479–84.PubMedCrossRef Zheng Z, Liebers M, Zhelyazkova B, Cao Y, Panditi D, Lynch KD, et al. Anchored multiplex PCR for targeted next-generation sequencing. Nat Med. 2014;20(12):1479–84.PubMedCrossRef
188.
Zurück zum Zitat Frattini V, Trifonov V, Chan JM, Castano A, Lia M, Abate F, et al. The integrated landscape of driver genomic alterations in glioblastoma. Nat Genet. 2013;45(10):1141–9.PubMedPubMedCentralCrossRef Frattini V, Trifonov V, Chan JM, Castano A, Lia M, Abate F, et al. The integrated landscape of driver genomic alterations in glioblastoma. Nat Genet. 2013;45(10):1141–9.PubMedPubMedCentralCrossRef
189.
Zurück zum Zitat Singh D, Chan JM, Zoppoli P, Niola F, Sullivan R, Castano A, et al. Transforming fusions of FGFR and TACC genes in human glioblastoma. Science. 2012;337(6099):1231–5.PubMedPubMedCentralCrossRef Singh D, Chan JM, Zoppoli P, Niola F, Sullivan R, Castano A, et al. Transforming fusions of FGFR and TACC genes in human glioblastoma. Science. 2012;337(6099):1231–5.PubMedPubMedCentralCrossRef
190.
Zurück zum Zitat Di Stefano AL, Fucci A, Frattini V, Labussiere M, Mokhtari K, Zoppoli P, et al. Detection, characterization, and inhibition of FGFR–TACC fusions in IDH Wild-type glioma. Clin Cancer Res. 2015;21(14):3307–17.PubMedPubMedCentralCrossRef Di Stefano AL, Fucci A, Frattini V, Labussiere M, Mokhtari K, Zoppoli P, et al. Detection, characterization, and inhibition of FGFR–TACC fusions in IDH Wild-type glioma. Clin Cancer Res. 2015;21(14):3307–17.PubMedPubMedCentralCrossRef
191.
Zurück zum Zitat Parker BC, Annala MJ, Cogdell DE, Granberg KJ, Sun Y, Ji P, et al. The tumorigenic FGFR3-TACC3 gene fusion escapes miR-99a regulation in glioblastoma. J Clin Invest. 2013;123(2):855–65.PubMedPubMedCentral Parker BC, Annala MJ, Cogdell DE, Granberg KJ, Sun Y, Ji P, et al. The tumorigenic FGFR3-TACC3 gene fusion escapes miR-99a regulation in glioblastoma. J Clin Invest. 2013;123(2):855–65.PubMedPubMedCentral
192.
Zurück zum Zitat Wu Y-M, Su F, Kalyana-Sundaram S, Khazanov N, Ateeq B, Cao X, et al. Identification of targetable FGFR gene fusions in diverse cancers. Cancer Discov. 2013;3(6):636–47.PubMedPubMedCentralCrossRef Wu Y-M, Su F, Kalyana-Sundaram S, Khazanov N, Ateeq B, Cao X, et al. Identification of targetable FGFR gene fusions in diverse cancers. Cancer Discov. 2013;3(6):636–47.PubMedPubMedCentralCrossRef
193.
Zurück zum Zitat Blandin AF, Giglio R, Graham MS, Garcia G, Malinowski S, Woods JK, et al. ALK amplification and rearrangements are recurrent targetable events in congenital and adult glioblastoma. Clin Cancer Res. 2023;29(14):2651–67.PubMedPubMedCentralCrossRef Blandin AF, Giglio R, Graham MS, Garcia G, Malinowski S, Woods JK, et al. ALK amplification and rearrangements are recurrent targetable events in congenital and adult glioblastoma. Clin Cancer Res. 2023;29(14):2651–67.PubMedPubMedCentralCrossRef
194.
Zurück zum Zitat Bagchi A, Orr BA, Campagne O, Dhanda S, Nair S, Tran Q, et al. Lorlatinib in a child with ALK-Fusion-positive high-Grade Glioma. N Engl J Med. 2021;385(8):761–3.PubMedPubMedCentralCrossRef Bagchi A, Orr BA, Campagne O, Dhanda S, Nair S, Tran Q, et al. Lorlatinib in a child with ALK-Fusion-positive high-Grade Glioma. N Engl J Med. 2021;385(8):761–3.PubMedPubMedCentralCrossRef
195.
Zurück zum Zitat Drilon A, Siena S, Ou SI, Patel M, Ahn MJ, Lee J, et al. Safety and Antitumor Activity of the Multitargeted Pan-TRK, ROS1, and ALK inhibitor entrectinib: combined results from two phase I trials (ALKA-372-001 and STARTRK-1). Cancer Discov. 2017;7(4):400–9.PubMedPubMedCentralCrossRef Drilon A, Siena S, Ou SI, Patel M, Ahn MJ, Lee J, et al. Safety and Antitumor Activity of the Multitargeted Pan-TRK, ROS1, and ALK inhibitor entrectinib: combined results from two phase I trials (ALKA-372-001 and STARTRK-1). Cancer Discov. 2017;7(4):400–9.PubMedPubMedCentralCrossRef
196.
Zurück zum Zitat Tabernero J, Bahleda R, Dienstmann R, Infante JR, Mita A, Italiano A, et al. Phase I dose-escalation study of JNJ-42756493, an oral pan-fibroblast growth factor receptor inhibitor, in patients with Advanced Solid tumors. J Clin Oncol. 2015;33(30):3401–8.PubMedCrossRef Tabernero J, Bahleda R, Dienstmann R, Infante JR, Mita A, Italiano A, et al. Phase I dose-escalation study of JNJ-42756493, an oral pan-fibroblast growth factor receptor inhibitor, in patients with Advanced Solid tumors. J Clin Oncol. 2015;33(30):3401–8.PubMedCrossRef
197.
Zurück zum Zitat Rodon J, Damian S, Furqan M, Garcia-Donas J, Imai H, Italiano A, et al. Abstract CT016: clinical and translational findings of pemigatinib in previously treated solid tumors with activating FGFR1-3 alterations in the FIGHT-207 study. Cancer Res. 2023;83(8Supplement):CT016–CT.CrossRef Rodon J, Damian S, Furqan M, Garcia-Donas J, Imai H, Italiano A, et al. Abstract CT016: clinical and translational findings of pemigatinib in previously treated solid tumors with activating FGFR1-3 alterations in the FIGHT-207 study. Cancer Res. 2023;83(8Supplement):CT016–CT.CrossRef
198.
Zurück zum Zitat Ahluwalia M, Franceschi E, Veronese L, Oliveira N, Li X, van den Bent M, MULTICENTER STUDY OF PEMIGATINIB IN PATIENTS WITH PREVIOUSLY TREATED GLIOBLASTOMA OR OTHER PRIMARY CENTRAL NERVOUS SYSTEM TUMORS WITH ACTIVATING FGFR. CTNI-39. FIGHT-209: A PHASE 2, OPEN-LABEL. ALTERATIONS Neuro Oncol. 2022;24(Supplement7):1–3. Ahluwalia M, Franceschi E, Veronese L, Oliveira N, Li X, van den Bent M, MULTICENTER STUDY OF PEMIGATINIB IN PATIENTS WITH PREVIOUSLY TREATED GLIOBLASTOMA OR OTHER PRIMARY CENTRAL NERVOUS SYSTEM TUMORS WITH ACTIVATING FGFR. CTNI-39. FIGHT-209: A PHASE 2, OPEN-LABEL. ALTERATIONS Neuro Oncol. 2022;24(Supplement7):1–3.
199.
Zurück zum Zitat Lassman A, Sepúlveda-Sánchez J, Cloughesy T, Gil-Gil J, Puduvalli V, Raizer J, et al. ACTR-33. INFIGRATINIB (BGJ398) IN PATIENTS WITH RECURRENT GLIOMAS WITH FIBROBLAST GROWTH FACTOR RECEPTOR (FGFR) ALTERATIONS: A MULTICENTER PHASE II STUDY. Neuro Oncol. 2019;21(Supplement6):vi20–vi.PubMedCentralCrossRef Lassman A, Sepúlveda-Sánchez J, Cloughesy T, Gil-Gil J, Puduvalli V, Raizer J, et al. ACTR-33. INFIGRATINIB (BGJ398) IN PATIENTS WITH RECURRENT GLIOMAS WITH FIBROBLAST GROWTH FACTOR RECEPTOR (FGFR) ALTERATIONS: A MULTICENTER PHASE II STUDY. Neuro Oncol. 2019;21(Supplement6):vi20–vi.PubMedCentralCrossRef
200.
Zurück zum Zitat Picca A, Stefano ALD, Savatovsky J, Ducray F, Chinot O, Moyal EC-J, et al. CTNI-33. TARGET TRIAL: A PHASE I/II OPEN-LABEL MULTICENTER STUDY TO ASSESS SAFETY, TOLERABILITY, AND CLINICAL EFFICACY OF AZD4547 IN PATIENTS WITH RELAPSED/REFRACTORY FGFR FUSION POSITIVE GLIOMA. Neuro Oncol. 2023;25(Supplement5):v81–v.CrossRef Picca A, Stefano ALD, Savatovsky J, Ducray F, Chinot O, Moyal EC-J, et al. CTNI-33. TARGET TRIAL: A PHASE I/II OPEN-LABEL MULTICENTER STUDY TO ASSESS SAFETY, TOLERABILITY, AND CLINICAL EFFICACY OF AZD4547 IN PATIENTS WITH RELAPSED/REFRACTORY FGFR FUSION POSITIVE GLIOMA. Neuro Oncol. 2023;25(Supplement5):v81–v.CrossRef
201.
Zurück zum Zitat Wang Y, Liang D, Chen J, Chen H, Fan R, Gao Y, et al. Targeted therapy with Anlotinib for a patient with an oncogenic FGFR3-TACC3 Fusion and recurrent glioblastoma. Oncologist. 2021;26(3):173–7.PubMedCrossRef Wang Y, Liang D, Chen J, Chen H, Fan R, Gao Y, et al. Targeted therapy with Anlotinib for a patient with an oncogenic FGFR3-TACC3 Fusion and recurrent glioblastoma. Oncologist. 2021;26(3):173–7.PubMedCrossRef
202.
Zurück zum Zitat Meric-Bernstam F, Bahleda R, Hierro C, Sanson M, Bridgewater J, Arkenau H-T, et al. Futibatinib, an irreversible FGFR1–4 inhibitor, in patients with Advanced Solid tumors harboring FGF/FGFR aberrations: a phase I dose-expansion study. Cancer Discov. 2022;12(2):402–15.PubMedCrossRef Meric-Bernstam F, Bahleda R, Hierro C, Sanson M, Bridgewater J, Arkenau H-T, et al. Futibatinib, an irreversible FGFR1–4 inhibitor, in patients with Advanced Solid tumors harboring FGF/FGFR aberrations: a phase I dose-expansion study. Cancer Discov. 2022;12(2):402–15.PubMedCrossRef
203.
Zurück zum Zitat Lee EQ, Muzikansky A, Reardon DA, Dietrich J, Nayak L, Duda DG, et al. Phase II trial of ponatinib in patients with bevacizumab-refractory glioblastoma. JCO. 2018;36(15suppl):2032.CrossRef Lee EQ, Muzikansky A, Reardon DA, Dietrich J, Nayak L, Duda DG, et al. Phase II trial of ponatinib in patients with bevacizumab-refractory glioblastoma. JCO. 2018;36(15suppl):2032.CrossRef
204.
Zurück zum Zitat Carpentier A, Canney M, Vignot A, Reina V, Beccaria K, Horodyckid C, et al. Clinical trial of blood-brain barrier disruption by pulsed ultrasound. Sci Transl Med. 2016;8(343):343re2.PubMedCrossRef Carpentier A, Canney M, Vignot A, Reina V, Beccaria K, Horodyckid C, et al. Clinical trial of blood-brain barrier disruption by pulsed ultrasound. Sci Transl Med. 2016;8(343):343re2.PubMedCrossRef
205.
Zurück zum Zitat Idbaih A, Canney M, Belin L, Desseaux C, Vignot A, Bouchoux G, et al. Safety and feasibility of repeated and transient blood-brain barrier disruption by Pulsed Ultrasound in patients with recurrent glioblastoma. Clin Cancer Res. 2019;25(13):3793–801.PubMedCrossRef Idbaih A, Canney M, Belin L, Desseaux C, Vignot A, Bouchoux G, et al. Safety and feasibility of repeated and transient blood-brain barrier disruption by Pulsed Ultrasound in patients with recurrent glioblastoma. Clin Cancer Res. 2019;25(13):3793–801.PubMedCrossRef
206.
Zurück zum Zitat Benmelouka AY, Munir M, Sayed A, Attia MS, Ali MM, Negida A, et al. Neural stem cell-based therapies and Glioblastoma Management: current evidence and Clinical challenges. Int J Mol Sci. 2021;22(5):2258.PubMedPubMedCentralCrossRef Benmelouka AY, Munir M, Sayed A, Attia MS, Ali MM, Negida A, et al. Neural stem cell-based therapies and Glioblastoma Management: current evidence and Clinical challenges. Int J Mol Sci. 2021;22(5):2258.PubMedPubMedCentralCrossRef
207.
Zurück zum Zitat Ring KL, Tong LM, Balestra ME, Javier R, Andrews-Zwilling Y, Li G, et al. Direct reprogramming of mouse and human fibroblasts into multipotent neural stem cells with a single factor. Cell Stem Cell. 2012;11(1):100–9.PubMedPubMedCentralCrossRef Ring KL, Tong LM, Balestra ME, Javier R, Andrews-Zwilling Y, Li G, et al. Direct reprogramming of mouse and human fibroblasts into multipotent neural stem cells with a single factor. Cell Stem Cell. 2012;11(1):100–9.PubMedPubMedCentralCrossRef
208.
Zurück zum Zitat Matsui T, Takano M, Yoshida K, Ono S, Fujisaki C, Matsuzaki Y, et al. Neural stem cells directly differentiated from partially reprogrammed fibroblasts rapidly acquire gliogenic competency. Stem Cells. 2012;30(6):1109–19.PubMedCrossRef Matsui T, Takano M, Yoshida K, Ono S, Fujisaki C, Matsuzaki Y, et al. Neural stem cells directly differentiated from partially reprogrammed fibroblasts rapidly acquire gliogenic competency. Stem Cells. 2012;30(6):1109–19.PubMedCrossRef
209.
Zurück zum Zitat Buckley A, Hagler SB, Lettry V, Bagó JR, Maingi SM, Khagi S, et al. Generation and Profiling of Tumor-Homing Induced neural stem cells from the skin of Cancer patients. Mol Ther. 2020;28(7):1614–27.PubMedPubMedCentralCrossRef Buckley A, Hagler SB, Lettry V, Bagó JR, Maingi SM, Khagi S, et al. Generation and Profiling of Tumor-Homing Induced neural stem cells from the skin of Cancer patients. Mol Ther. 2020;28(7):1614–27.PubMedPubMedCentralCrossRef
210.
Zurück zum Zitat Zomer HD, Vidane AS, Gonçalves NN, Ambrósio CE. Mesenchymal and induced pluripotent stem cells: general insights and clinical perspectives. Stem Cells Cloning. 2015;8:125–34.PubMedPubMedCentral Zomer HD, Vidane AS, Gonçalves NN, Ambrósio CE. Mesenchymal and induced pluripotent stem cells: general insights and clinical perspectives. Stem Cells Cloning. 2015;8:125–34.PubMedPubMedCentral
211.
Zurück zum Zitat Ehtesham M, Kabos P, Kabosova A, Neuman T, Black KL, Yu JS. The use of interleukin 12-secreting neural stem cells for the treatment of intracranial glioma. Cancer Res. 2002;62(20):5657–63.PubMed Ehtesham M, Kabos P, Kabosova A, Neuman T, Black KL, Yu JS. The use of interleukin 12-secreting neural stem cells for the treatment of intracranial glioma. Cancer Res. 2002;62(20):5657–63.PubMed
212.
Zurück zum Zitat Bagó JR, Alfonso-Pecchio A, Okolie O, Dumitru R, Rinkenbaugh A, Baldwin AS, et al. Therapeutically engineered induced neural stem cells are tumour-homing and inhibit progression of glioblastoma. Nat Commun. 2016;7(1):10593.PubMedPubMedCentralCrossRef Bagó JR, Alfonso-Pecchio A, Okolie O, Dumitru R, Rinkenbaugh A, Baldwin AS, et al. Therapeutically engineered induced neural stem cells are tumour-homing and inhibit progression of glioblastoma. Nat Commun. 2016;7(1):10593.PubMedPubMedCentralCrossRef
213.
Zurück zum Zitat Doucette T, Rao G, Yang Y, Gumin J, Shinojima N, Bekele BN, et al. Mesenchymal stem cells display tumor-specific tropism in an RCAS/Ntv-a glioma model. Neoplasia. 2011;13(8):716–25.PubMedPubMedCentralCrossRef Doucette T, Rao G, Yang Y, Gumin J, Shinojima N, Bekele BN, et al. Mesenchymal stem cells display tumor-specific tropism in an RCAS/Ntv-a glioma model. Neoplasia. 2011;13(8):716–25.PubMedPubMedCentralCrossRef
214.
Zurück zum Zitat Egea V, von Baumgarten L, Schichor C, Berninger B, Popp T, Neth P, et al. TNF-α respecifies human mesenchymal stem cells to a neural fate and promotes migration toward experimental glioma. Cell Death Differ. 2011;18(5):853–63.PubMedCrossRef Egea V, von Baumgarten L, Schichor C, Berninger B, Popp T, Neth P, et al. TNF-α respecifies human mesenchymal stem cells to a neural fate and promotes migration toward experimental glioma. Cell Death Differ. 2011;18(5):853–63.PubMedCrossRef
215.
Zurück zum Zitat Kim DS, Kim JH, Lee JK, Choi SJ, Kim JS, Jeun SS, et al. Overexpression of CXC chemokine receptors is required for the superior glioma-tracking property of umbilical cord blood-derived mesenchymal stem cells. Stem Cells Dev. 2009;18(3):511–9.PubMedCrossRef Kim DS, Kim JH, Lee JK, Choi SJ, Kim JS, Jeun SS, et al. Overexpression of CXC chemokine receptors is required for the superior glioma-tracking property of umbilical cord blood-derived mesenchymal stem cells. Stem Cells Dev. 2009;18(3):511–9.PubMedCrossRef
216.
Zurück zum Zitat Magge SN, Malik SZ, Royo NC, Chen HI, Yu L, Snyder EY, et al. Role of monocyte chemoattractant protein-1 (MCP-1/CCL2) in migration of neural progenitor cells toward glial tumors. J Neurosci Res. 2009;87(7):1547–55.PubMedCrossRef Magge SN, Malik SZ, Royo NC, Chen HI, Yu L, Snyder EY, et al. Role of monocyte chemoattractant protein-1 (MCP-1/CCL2) in migration of neural progenitor cells toward glial tumors. J Neurosci Res. 2009;87(7):1547–55.PubMedCrossRef
217.
Zurück zum Zitat Gunnarsson S, Bexell D, Svensson A, Siesjö P, Darabi A, Bengzon J. Intratumoral IL-7 delivery by mesenchymal stromal cells potentiates IFNgamma-transduced tumor cell immunotherapy of experimental glioma. J Neuroimmunol. 2010;218(1–2):140–4.PubMedCrossRef Gunnarsson S, Bexell D, Svensson A, Siesjö P, Darabi A, Bengzon J. Intratumoral IL-7 delivery by mesenchymal stromal cells potentiates IFNgamma-transduced tumor cell immunotherapy of experimental glioma. J Neuroimmunol. 2010;218(1–2):140–4.PubMedCrossRef
218.
Zurück zum Zitat Yuan X, Hu J, Belladonna ML, Black KL, Yu JS. Interleukin-23-expressing bone marrow-derived neural stem-like cells exhibit antitumor activity against intracranial glioma. Cancer Res. 2006;66(5):2630–8.PubMedCrossRef Yuan X, Hu J, Belladonna ML, Black KL, Yu JS. Interleukin-23-expressing bone marrow-derived neural stem-like cells exhibit antitumor activity against intracranial glioma. Cancer Res. 2006;66(5):2630–8.PubMedCrossRef
219.
Zurück zum Zitat Satterlee AB, Dunn DE, Lo DC, Khagi S, Hingtgen S. Tumoricidal stem cell therapy enables killing in novel hybrid models of heterogeneous glioblastoma. Neuro Oncol. 2019;21(12):1552–64.PubMedPubMedCentralCrossRef Satterlee AB, Dunn DE, Lo DC, Khagi S, Hingtgen S. Tumoricidal stem cell therapy enables killing in novel hybrid models of heterogeneous glioblastoma. Neuro Oncol. 2019;21(12):1552–64.PubMedPubMedCentralCrossRef
220.
Zurück zum Zitat Satterlee AB, Dunn DE, Valdivia A, Malawsky D, Buckley A, Gershon T, et al. Spatiotemporal analysis of induced neural stem cell therapy to overcome advanced glioblastoma recurrence. Mol Therapy - Oncolytics. 2022;26:49–62.CrossRef Satterlee AB, Dunn DE, Valdivia A, Malawsky D, Buckley A, Gershon T, et al. Spatiotemporal analysis of induced neural stem cell therapy to overcome advanced glioblastoma recurrence. Mol Therapy - Oncolytics. 2022;26:49–62.CrossRef
221.
Zurück zum Zitat Bagó JR, Okolie O, Dumitru R, Ewend MG, Parker JS, Werff RV, et al. Tumor-homing cytotoxic human induced neural stem cells for cancer therapy. Sci Transl Med. 2017;9(375):eaah6510.PubMedPubMedCentralCrossRef Bagó JR, Okolie O, Dumitru R, Ewend MG, Parker JS, Werff RV, et al. Tumor-homing cytotoxic human induced neural stem cells for cancer therapy. Sci Transl Med. 2017;9(375):eaah6510.PubMedPubMedCentralCrossRef
223.
Zurück zum Zitat Portnow J, Synold TW, Badie B, Tirughana R, Lacey SF, D’Apuzzo M, et al. Neural stem cell–based Anticancer Gene Therapy: A First-in-human study in recurrent high-Grade Glioma patients. Clin Cancer Res. 2017;23(12):2951–60.PubMedCrossRef Portnow J, Synold TW, Badie B, Tirughana R, Lacey SF, D’Apuzzo M, et al. Neural stem cell–based Anticancer Gene Therapy: A First-in-human study in recurrent high-Grade Glioma patients. Clin Cancer Res. 2017;23(12):2951–60.PubMedCrossRef
224.
Zurück zum Zitat Portnow J, Badie B, Suzette Blanchard M, Kilpatrick J, Tirughana R, Metz M, et al. Feasibility of intracerebrally administering multiple doses of genetically modified neural stem cells to locally produce chemotherapy in glioma patients. Cancer Gene Ther. 2021;28(3–4):294–306.PubMedCrossRef Portnow J, Badie B, Suzette Blanchard M, Kilpatrick J, Tirughana R, Metz M, et al. Feasibility of intracerebrally administering multiple doses of genetically modified neural stem cells to locally produce chemotherapy in glioma patients. Cancer Gene Ther. 2021;28(3–4):294–306.PubMedCrossRef
225.
Zurück zum Zitat Tu GXE, Ho YK, Ng ZX, Teo KJ, Yeo TT, Too H-P. A facile and scalable in production non-viral gene engineered mesenchymal stem cells for effective suppression of temozolomide-resistant (TMZR) glioblastoma growth. Stem Cell Res Ther. 2020;11(1):391.PubMedPubMedCentralCrossRef Tu GXE, Ho YK, Ng ZX, Teo KJ, Yeo TT, Too H-P. A facile and scalable in production non-viral gene engineered mesenchymal stem cells for effective suppression of temozolomide-resistant (TMZR) glioblastoma growth. Stem Cell Res Ther. 2020;11(1):391.PubMedPubMedCentralCrossRef
226.
Zurück zum Zitat Kim JW, Murphy J, Chang AL, Spencer DA, Kane JR, Kanojia D et al. 19 - All Aboard: Mesenchymal Stem/Stromal Cells as Cell Carriers for Virotherapy. In: Bolontrade MF, García MG, Bolontrade MF, García MG, editors. Mesenchymal Stromal Cells as Tumor Stromal Modulators. Boston2017. pp. 475 – 99. Kim JW, Murphy J, Chang AL, Spencer DA, Kane JR, Kanojia D et al. 19 - All Aboard: Mesenchymal Stem/Stromal Cells as Cell Carriers for Virotherapy. In: Bolontrade MF, García MG, Bolontrade MF, García MG, editors. Mesenchymal Stromal Cells as Tumor Stromal Modulators. Boston2017. pp. 475 – 99.
227.
Zurück zum Zitat Thaci B, Ahmed AU, Ulasov IV, Tobias AL, Han Y, Aboody KS, et al. Pharmacokinetic study of neural stem cell-based cell carrier for oncolytic virotherapy: targeted delivery of the therapeutic payload in an orthotopic brain tumor model. Cancer Gene Ther. 2012;19(6):431–42.PubMedPubMedCentralCrossRef Thaci B, Ahmed AU, Ulasov IV, Tobias AL, Han Y, Aboody KS, et al. Pharmacokinetic study of neural stem cell-based cell carrier for oncolytic virotherapy: targeted delivery of the therapeutic payload in an orthotopic brain tumor model. Cancer Gene Ther. 2012;19(6):431–42.PubMedPubMedCentralCrossRef
228.
Zurück zum Zitat Tobias AL, Thaci B, Auffinger B, Rincón E, Balyasnikova IV, Kim CK, et al. The timing of neural stem cell-based virotherapy is critical for optimal therapeutic efficacy when Applied with Radiation and Chemotherapy for the treatment of Glioblastoma. Stem Cells Transl Med. 2013;2(9):655–66.PubMedPubMedCentralCrossRef Tobias AL, Thaci B, Auffinger B, Rincón E, Balyasnikova IV, Kim CK, et al. The timing of neural stem cell-based virotherapy is critical for optimal therapeutic efficacy when Applied with Radiation and Chemotherapy for the treatment of Glioblastoma. Stem Cells Transl Med. 2013;2(9):655–66.PubMedPubMedCentralCrossRef
229.
Zurück zum Zitat Thaci B, Auffinger B, Ahmed AU, Tobias AL, Ulasov IV, Lesniak MS. Radiation and Temozolomide Chemotherapy do not inhibit the Oncolytic Adenovirus Loaded Carrier cells to effectively deliver their payload. Mol Ther. 2012;20(1supplement):S200. Thaci B, Auffinger B, Ahmed AU, Tobias AL, Ulasov IV, Lesniak MS. Radiation and Temozolomide Chemotherapy do not inhibit the Oncolytic Adenovirus Loaded Carrier cells to effectively deliver their payload. Mol Ther. 2012;20(1supplement):S200.
230.
Zurück zum Zitat Fares J, Ahmed AU, Ulasov IV, Sonabend AM, Miska J, Lee-Chang C, et al. Neural stem cell delivery of an oncolytic adenovirus in newly diagnosed malignant glioma: a first-in-human, phase 1, dose-escalation trial. Lancet Oncol. 2021;22(8):1103–14.PubMedPubMedCentralCrossRef Fares J, Ahmed AU, Ulasov IV, Sonabend AM, Miska J, Lee-Chang C, et al. Neural stem cell delivery of an oncolytic adenovirus in newly diagnosed malignant glioma: a first-in-human, phase 1, dose-escalation trial. Lancet Oncol. 2021;22(8):1103–14.PubMedPubMedCentralCrossRef
231.
Zurück zum Zitat Burger MC, Zhang C, Harter PN, Romanski A, Strassheimer F, Senft C, et al. CAR-Engineered NK cells for the treatment of Glioblastoma: turning Innate effectors into Precision Tools for Cancer Immunotherapy. Front Immunol. 2019;10:2683.PubMedPubMedCentralCrossRef Burger MC, Zhang C, Harter PN, Romanski A, Strassheimer F, Senft C, et al. CAR-Engineered NK cells for the treatment of Glioblastoma: turning Innate effectors into Precision Tools for Cancer Immunotherapy. Front Immunol. 2019;10:2683.PubMedPubMedCentralCrossRef
232.
Zurück zum Zitat O’Rourke DM, Nasrallah MP, Desai A, Melenhorst JJ, Mansfield K, Morrissette JJD, et al. A single dose of peripherally infused EGFRvIII-directed CAR T cells mediates antigen loss and induces adaptive resistance in patients with recurrent glioblastoma. Sci Transl Med. 2017;9(399):eaaa0984.PubMedPubMedCentralCrossRef O’Rourke DM, Nasrallah MP, Desai A, Melenhorst JJ, Mansfield K, Morrissette JJD, et al. A single dose of peripherally infused EGFRvIII-directed CAR T cells mediates antigen loss and induces adaptive resistance in patients with recurrent glioblastoma. Sci Transl Med. 2017;9(399):eaaa0984.PubMedPubMedCentralCrossRef
233.
Zurück zum Zitat Brown CE, Badie B, Barish ME, Weng L, Ostberg JR, Chang WC, et al. Bioactivity and Safety of IL13Rα2-Redirected chimeric Antigen receptor CD8 + T cells in patients with recurrent glioblastoma. Clin Cancer Res. 2015;21(18):4062–72.PubMedPubMedCentralCrossRef Brown CE, Badie B, Barish ME, Weng L, Ostberg JR, Chang WC, et al. Bioactivity and Safety of IL13Rα2-Redirected chimeric Antigen receptor CD8 + T cells in patients with recurrent glioblastoma. Clin Cancer Res. 2015;21(18):4062–72.PubMedPubMedCentralCrossRef
234.
Zurück zum Zitat Brown CE, Alizadeh D, Starr R, Weng L, Wagner JR, Naranjo A, et al. Regression of Glioblastoma after chimeric Antigen receptor T-Cell therapy. N Engl J Med. 2016;375(26):2561–9.PubMedPubMedCentralCrossRef Brown CE, Alizadeh D, Starr R, Weng L, Wagner JR, Naranjo A, et al. Regression of Glioblastoma after chimeric Antigen receptor T-Cell therapy. N Engl J Med. 2016;375(26):2561–9.PubMedPubMedCentralCrossRef
235.
Zurück zum Zitat Ahmed N, Brawley V, Hegde M, Bielamowicz K, Kalra M, Landi D, et al. HER2-Specific Chimeric Antigen Receptor-Modified Virus-Specific T Cells for Progressive Glioblastoma: a phase 1 dose-escalation Trial. JAMA Oncol. 2017;3(8):1094–101.PubMedPubMedCentralCrossRef Ahmed N, Brawley V, Hegde M, Bielamowicz K, Kalra M, Landi D, et al. HER2-Specific Chimeric Antigen Receptor-Modified Virus-Specific T Cells for Progressive Glioblastoma: a phase 1 dose-escalation Trial. JAMA Oncol. 2017;3(8):1094–101.PubMedPubMedCentralCrossRef
236.
Zurück zum Zitat Shen L, Li H, Bin S, Li P, Chen J, Gu H, et al. The efficacy of third generation anti–HER2 chimeric antigen receptor T cells in combination with PD1 blockade against malignant glioblastoma cells. Oncol Rep. 2019;42(4):1549–57.PubMed Shen L, Li H, Bin S, Li P, Chen J, Gu H, et al. The efficacy of third generation anti–HER2 chimeric antigen receptor T cells in combination with PD1 blockade against malignant glioblastoma cells. Oncol Rep. 2019;42(4):1549–57.PubMed
237.
Zurück zum Zitat Tang X, Wang Y, Huang J, Zhang Z, Liu F, Xu J, et al. Administration of B7-H3 targeted chimeric antigen receptor-T cells induce regression of glioblastoma. Signal Transduct Target Therapy. 2021;6(1):125.CrossRef Tang X, Wang Y, Huang J, Zhang Z, Liu F, Xu J, et al. Administration of B7-H3 targeted chimeric antigen receptor-T cells induce regression of glioblastoma. Signal Transduct Target Therapy. 2021;6(1):125.CrossRef
238.
Zurück zum Zitat Tang X, Zhao S, Zhang Y, Wang Y, Zhang Z, Yang M, et al. B7-H3 as a novel CAR-T therapeutic target for Glioblastoma. Mol Ther Oncolytics. 2019;14:279–87.PubMedPubMedCentralCrossRef Tang X, Zhao S, Zhang Y, Wang Y, Zhang Z, Yang M, et al. B7-H3 as a novel CAR-T therapeutic target for Glioblastoma. Mol Ther Oncolytics. 2019;14:279–87.PubMedPubMedCentralCrossRef
239.
Zurück zum Zitat Farkas S, Cioca D, Murányi J, Hornyák P, Brunyánszki A, Szekér P et al. Chlorotoxin binds to both matrix metalloproteinase 2 and neuropilin 1. J Biol Chem. 2023;299(9). Farkas S, Cioca D, Murányi J, Hornyák P, Brunyánszki A, Szekér P et al. Chlorotoxin binds to both matrix metalloproteinase 2 and neuropilin 1. J Biol Chem. 2023;299(9).
240.
Zurück zum Zitat Bielamowicz K, Fousek K, Byrd TT, Samaha H, Mukherjee M, Aware N, et al. Trivalent CAR T cells overcome interpatient antigenic variability in glioblastoma. Neuro Oncol. 2017;20(4):506–18.PubMedCentralCrossRef Bielamowicz K, Fousek K, Byrd TT, Samaha H, Mukherjee M, Aware N, et al. Trivalent CAR T cells overcome interpatient antigenic variability in glioblastoma. Neuro Oncol. 2017;20(4):506–18.PubMedCentralCrossRef
241.
Zurück zum Zitat Hegde M, Corder A, Chow KK, Mukherjee M, Ashoori A, Kew Y, et al. Combinational targeting offsets antigen escape and enhances effector functions of adoptively transferred T cells in glioblastoma. Mol Ther. 2013;21(11):2087–101.PubMedPubMedCentralCrossRef Hegde M, Corder A, Chow KK, Mukherjee M, Ashoori A, Kew Y, et al. Combinational targeting offsets antigen escape and enhances effector functions of adoptively transferred T cells in glioblastoma. Mol Ther. 2013;21(11):2087–101.PubMedPubMedCentralCrossRef
242.
Zurück zum Zitat Jin L, Tao H, Karachi A, Long Y, Hou AY, Na M, et al. CXCR1- or CXCR2-modified CAR T cells co-opt IL-8 for maximal antitumor efficacy in solid tumors. Nat Commun. 2019;10(1):4016.PubMedPubMedCentralCrossRef Jin L, Tao H, Karachi A, Long Y, Hou AY, Na M, et al. CXCR1- or CXCR2-modified CAR T cells co-opt IL-8 for maximal antitumor efficacy in solid tumors. Nat Commun. 2019;10(1):4016.PubMedPubMedCentralCrossRef
243.
Zurück zum Zitat Hegde M, Mukherjee M, Grada Z, Pignata A, Landi D, Navai SA, et al. Tandem CAR T cells targeting HER2 and IL13Rα2 mitigate tumor antigen escape. J Clin Investig. 2016;126(8):3036–52.PubMedPubMedCentralCrossRef Hegde M, Mukherjee M, Grada Z, Pignata A, Landi D, Navai SA, et al. Tandem CAR T cells targeting HER2 and IL13Rα2 mitigate tumor antigen escape. J Clin Investig. 2016;126(8):3036–52.PubMedPubMedCentralCrossRef
244.
Zurück zum Zitat Choe JH, Watchmaker PB, Simic MS, Gilbert RD, Li AW, Krasnow NA, et al. SynNotch-CAR T cells overcome challenges of specificity, heterogeneity, and persistence in treating glioblastoma. Sci Transl Med. 2021;13:591.CrossRef Choe JH, Watchmaker PB, Simic MS, Gilbert RD, Li AW, Krasnow NA, et al. SynNotch-CAR T cells overcome challenges of specificity, heterogeneity, and persistence in treating glioblastoma. Sci Transl Med. 2021;13:591.CrossRef
245.
Zurück zum Zitat Roybal KT, Rupp LJ, Morsut L, Walker WJ, McNally KA, Park JS, et al. Precision Tumor Recognition by T cells with Combinatorial Antigen-Sensing circuits. Cell. 2016;164(4):770–9.PubMedPubMedCentralCrossRef Roybal KT, Rupp LJ, Morsut L, Walker WJ, McNally KA, Park JS, et al. Precision Tumor Recognition by T cells with Combinatorial Antigen-Sensing circuits. Cell. 2016;164(4):770–9.PubMedPubMedCentralCrossRef
246.
Zurück zum Zitat Yin Y, Boesteanu AC, Binder ZA, Xu C, Reid RA, Rodriguez JL, et al. Checkpoint blockade reverses Anergy in IL-13Rα2 Humanized Scfv-based CAR T cells to treat murine and canine gliomas. Mol Ther Oncolytics. 2018;11:20–38.PubMedPubMedCentralCrossRef Yin Y, Boesteanu AC, Binder ZA, Xu C, Reid RA, Rodriguez JL, et al. Checkpoint blockade reverses Anergy in IL-13Rα2 Humanized Scfv-based CAR T cells to treat murine and canine gliomas. Mol Ther Oncolytics. 2018;11:20–38.PubMedPubMedCentralCrossRef
247.
Zurück zum Zitat Bagley SJ, Binder ZA, Lamrani L, Marinari E, Desai AS, Nasrallah MP et al. Repeated peripheral infusions of anti-EGFRvIII CAR T cells in combination with pembrolizumab show no efficacy in glioblastoma: a phase 1 trial. Nat Cancer. 2024. Bagley SJ, Binder ZA, Lamrani L, Marinari E, Desai AS, Nasrallah MP et al. Repeated peripheral infusions of anti-EGFRvIII CAR T cells in combination with pembrolizumab show no efficacy in glioblastoma: a phase 1 trial. Nat Cancer. 2024.
248.
Zurück zum Zitat Geletneky K, Hajda J, Angelova AL, Leuchs B, Capper D, Bartsch AJ, et al. Oncolytic H-1 Parvovirus shows Safety and signs of immunogenic activity in a first phase I/IIa glioblastoma trial. Mol Ther. 2017;25(12):2620–34.PubMedPubMedCentralCrossRef Geletneky K, Hajda J, Angelova AL, Leuchs B, Capper D, Bartsch AJ, et al. Oncolytic H-1 Parvovirus shows Safety and signs of immunogenic activity in a first phase I/IIa glioblastoma trial. Mol Ther. 2017;25(12):2620–34.PubMedPubMedCentralCrossRef
249.
Zurück zum Zitat van Putten EHP, Kleijn A, van Beusechem VW, Noske D, Lamers CHJ, de Goede AL, et al. Convection enhanced delivery of the Oncolytic Adenovirus Delta24-RGD in patients with recurrent GBM: a phase I clinical Trial Including Correlative studies. Clin Cancer Res. 2022;28(8):1572–85.PubMedPubMedCentralCrossRef van Putten EHP, Kleijn A, van Beusechem VW, Noske D, Lamers CHJ, de Goede AL, et al. Convection enhanced delivery of the Oncolytic Adenovirus Delta24-RGD in patients with recurrent GBM: a phase I clinical Trial Including Correlative studies. Clin Cancer Res. 2022;28(8):1572–85.PubMedPubMedCentralCrossRef
250.
Zurück zum Zitat Thompson EM, Landi D, Brown MC, Friedman HS, McLendon R, Herndon JE 2, et al. Recombinant polio-rhinovirus immunotherapy for recurrent paediatric high-grade glioma: a phase 1b trial. Lancet Child Adolesc Health. 2023;7(7):471–8.PubMedCrossRef Thompson EM, Landi D, Brown MC, Friedman HS, McLendon R, Herndon JE 2, et al. Recombinant polio-rhinovirus immunotherapy for recurrent paediatric high-grade glioma: a phase 1b trial. Lancet Child Adolesc Health. 2023;7(7):471–8.PubMedCrossRef
251.
Zurück zum Zitat Nassiri F, Patil V, Yefet LS, Singh O, Liu J, Dang RMA, et al. Oncolytic DNX-2401 virotherapy plus pembrolizumab in recurrent glioblastoma: a phase 1/2 trial. Nat Med. 2023;29(6):1370–8.PubMedPubMedCentralCrossRef Nassiri F, Patil V, Yefet LS, Singh O, Liu J, Dang RMA, et al. Oncolytic DNX-2401 virotherapy plus pembrolizumab in recurrent glioblastoma: a phase 1/2 trial. Nat Med. 2023;29(6):1370–8.PubMedPubMedCentralCrossRef
252.
Zurück zum Zitat Desjardins A, Gromeier M, Herndon JE 2nd, Beaubier N, Bolognesi DP, Friedman AH, et al. Recurrent glioblastoma treated with recombinant poliovirus. N Engl J Med. 2018;379(2):150–61.PubMedPubMedCentralCrossRef Desjardins A, Gromeier M, Herndon JE 2nd, Beaubier N, Bolognesi DP, Friedman AH, et al. Recurrent glioblastoma treated with recombinant poliovirus. N Engl J Med. 2018;379(2):150–61.PubMedPubMedCentralCrossRef
253.
Zurück zum Zitat Todo T, Ito H, Ino Y, Ohtsu H, Ota Y, Shibahara J, et al. Intratumoral oncolytic herpes virus G47∆ for residual or recurrent glioblastoma: a phase 2 trial. Nat Med. 2022;28(8):1630–9.PubMedPubMedCentralCrossRef Todo T, Ito H, Ino Y, Ohtsu H, Ota Y, Shibahara J, et al. Intratumoral oncolytic herpes virus G47∆ for residual or recurrent glioblastoma: a phase 2 trial. Nat Med. 2022;28(8):1630–9.PubMedPubMedCentralCrossRef
254.
Zurück zum Zitat Ling AL, Solomon IH, Landivar AM, Nakashima H, Woods JK, Santos A, et al. Clinical trial links oncolytic immunoactivation to survival in glioblastoma. Nature. 2023;623(7985):157–66.PubMedPubMedCentralCrossRef Ling AL, Solomon IH, Landivar AM, Nakashima H, Woods JK, Santos A, et al. Clinical trial links oncolytic immunoactivation to survival in glioblastoma. Nature. 2023;623(7985):157–66.PubMedPubMedCentralCrossRef
255.
Zurück zum Zitat Freeman AI, Zakay-Rones Z, Gomori JM, Linetsky E, Rasooly L, Greenbaum E, et al. Phase I/II trial of intravenous NDV-HUJ oncolytic virus in recurrent Glioblastoma Multiforme. Mol Ther. 2006;13(1):221–8.PubMedCrossRef Freeman AI, Zakay-Rones Z, Gomori JM, Linetsky E, Rasooly L, Greenbaum E, et al. Phase I/II trial of intravenous NDV-HUJ oncolytic virus in recurrent Glioblastoma Multiforme. Mol Ther. 2006;13(1):221–8.PubMedCrossRef
256.
Zurück zum Zitat Chiocca EA, Nakashima H, Kasai K, Fernandez SA, Oglesbee M. Preclinical toxicology of rQNestin34.5v.2: an oncolytic herpes virus with transcriptional regulation of the ICP34.5 Neurovirulence Gene. Mol Ther Methods Clin Dev. 2020;17:871–93.PubMedPubMedCentralCrossRef Chiocca EA, Nakashima H, Kasai K, Fernandez SA, Oglesbee M. Preclinical toxicology of rQNestin34.5v.2: an oncolytic herpes virus with transcriptional regulation of the ICP34.5 Neurovirulence Gene. Mol Ther Methods Clin Dev. 2020;17:871–93.PubMedPubMedCentralCrossRef
257.
Zurück zum Zitat Kanai R, Zaupa C, Sgubin D, Antoszczyk SJ, Martuza RL, Wakimoto H, et al. Effect of gamma34.5 deletions on oncolytic herpes simplex virus activity in brain tumors. J Virol. 2012;86(8):4420–31.PubMedPubMedCentralCrossRef Kanai R, Zaupa C, Sgubin D, Antoszczyk SJ, Martuza RL, Wakimoto H, et al. Effect of gamma34.5 deletions on oncolytic herpes simplex virus activity in brain tumors. J Virol. 2012;86(8):4420–31.PubMedPubMedCentralCrossRef
258.
Zurück zum Zitat Peters C, Paget M, Tshilenge KT, Saha D, Antoszczyk S, Baars A et al. Restriction of replication of Oncolytic Herpes Simplex Virus with a deletion of gamma34.5 in Glioblastoma Stem-Like cells. J Virol. 2018;92(15). Peters C, Paget M, Tshilenge KT, Saha D, Antoszczyk S, Baars A et al. Restriction of replication of Oncolytic Herpes Simplex Virus with a deletion of gamma34.5 in Glioblastoma Stem-Like cells. J Virol. 2018;92(15).
259.
Zurück zum Zitat Kambara H, Okano H, Chiocca EA, Saeki Y. An oncolytic HSV-1 mutant expressing ICP34.5 under control of a nestin promoter increases survival of animals even when symptomatic from a brain tumor. Cancer Res. 2005;65(7):2832–9.PubMedCrossRef Kambara H, Okano H, Chiocca EA, Saeki Y. An oncolytic HSV-1 mutant expressing ICP34.5 under control of a nestin promoter increases survival of animals even when symptomatic from a brain tumor. Cancer Res. 2005;65(7):2832–9.PubMedCrossRef
260.
Zurück zum Zitat Patel DM, Foreman PM, Nabors LB, Riley KO, Gillespie GY, Markert JM. Design of a Phase I Clinical Trial to evaluate M032, a genetically Engineered HSV-1 expressing IL-12, in patients with Recurrent/Progressive Glioblastoma Multiforme, anaplastic astrocytoma, or Gliosarcoma. Hum Gene Ther Clin Dev. 2016;27(2):69–78.PubMedPubMedCentralCrossRef Patel DM, Foreman PM, Nabors LB, Riley KO, Gillespie GY, Markert JM. Design of a Phase I Clinical Trial to evaluate M032, a genetically Engineered HSV-1 expressing IL-12, in patients with Recurrent/Progressive Glioblastoma Multiforme, anaplastic astrocytoma, or Gliosarcoma. Hum Gene Ther Clin Dev. 2016;27(2):69–78.PubMedPubMedCentralCrossRef
261.
Zurück zum Zitat Xu B, Tian L, Chen J, Wang J, Ma R, Dong W, et al. An oncolytic virus expressing a full-length antibody enhances antitumor innate immune response to glioblastoma. Nat Commun. 2021;12(1):5908.PubMedPubMedCentralCrossRef Xu B, Tian L, Chen J, Wang J, Ma R, Dong W, et al. An oncolytic virus expressing a full-length antibody enhances antitumor innate immune response to glioblastoma. Nat Commun. 2021;12(1):5908.PubMedPubMedCentralCrossRef
262.
Zurück zum Zitat Wang G, Zhang Z, Zhong K, Wang Z, Yang N, Tang X, et al. CXCL11-armed oncolytic adenoviruses enhance CAR-T cell therapeutic efficacy and reprogram tumor microenvironment in glioblastoma. Mol Ther. 2023;31(1):134–53.PubMedCrossRef Wang G, Zhang Z, Zhong K, Wang Z, Yang N, Tang X, et al. CXCL11-armed oncolytic adenoviruses enhance CAR-T cell therapeutic efficacy and reprogram tumor microenvironment in glioblastoma. Mol Ther. 2023;31(1):134–53.PubMedCrossRef
263.
Zurück zum Zitat Friedman GK, Bernstock JD, Chen D, Nan L, Moore BP, Kelly VM, et al. Enhanced sensitivity of patient-derived Pediatric High-Grade Brain Tumor xenografts to Oncolytic HSV-1 Virotherapy correlates with Nectin-1 expression. Sci Rep. 2018;8(1):13930.PubMedPubMedCentralCrossRef Friedman GK, Bernstock JD, Chen D, Nan L, Moore BP, Kelly VM, et al. Enhanced sensitivity of patient-derived Pediatric High-Grade Brain Tumor xenografts to Oncolytic HSV-1 Virotherapy correlates with Nectin-1 expression. Sci Rep. 2018;8(1):13930.PubMedPubMedCentralCrossRef
264.
Zurück zum Zitat Tian L, Xu B, Chen Y, Li Z, Wang J, Zhang J, et al. Specific targeting of glioblastoma with an oncolytic virus expressing a cetuximab-CCL5 fusion protein via innate and adaptive immunity. Nat Cancer. 2022;3(11):1318–35.PubMedPubMedCentralCrossRef Tian L, Xu B, Chen Y, Li Z, Wang J, Zhang J, et al. Specific targeting of glioblastoma with an oncolytic virus expressing a cetuximab-CCL5 fusion protein via innate and adaptive immunity. Nat Cancer. 2022;3(11):1318–35.PubMedPubMedCentralCrossRef
265.
Zurück zum Zitat Wang J, Cazzato E, Ladewig E, Frattini V, Rosenbloom DIS, Zairis S, et al. Clonal evolution of Glioblastoma under Therapy. Nat Genet. 2016;48(7):768–76.PubMedPubMedCentralCrossRef Wang J, Cazzato E, Ladewig E, Frattini V, Rosenbloom DIS, Zairis S, et al. Clonal evolution of Glioblastoma under Therapy. Nat Genet. 2016;48(7):768–76.PubMedPubMedCentralCrossRef
266.
Zurück zum Zitat Bastos AGP, Carvalho B, Silva R, Leitão D, Linhares P, Vaz R, et al. Endoglin (CD105) and proliferation index in recurrent glioblastoma treated with anti-angiogenic therapy. Front Oncol. 2022;12:910196.PubMedPubMedCentralCrossRef Bastos AGP, Carvalho B, Silva R, Leitão D, Linhares P, Vaz R, et al. Endoglin (CD105) and proliferation index in recurrent glioblastoma treated with anti-angiogenic therapy. Front Oncol. 2022;12:910196.PubMedPubMedCentralCrossRef
267.
Zurück zum Zitat Zhang X, Zhang Y, Jia Y, Qin T, Zhang C, Li Y, et al. Bevacizumab promotes active biological behaviors of human umbilical vein endothelial cells by activating TGFβ1 pathways via off-VEGF signaling. Cancer Biol Med. 2020;17(2):418–32.PubMedPubMedCentralCrossRef Zhang X, Zhang Y, Jia Y, Qin T, Zhang C, Li Y, et al. Bevacizumab promotes active biological behaviors of human umbilical vein endothelial cells by activating TGFβ1 pathways via off-VEGF signaling. Cancer Biol Med. 2020;17(2):418–32.PubMedPubMedCentralCrossRef
268.
Zurück zum Zitat Burghardt I, Ventura E, Weiss T, Schroeder JJ, Seystahl K, Zielasek C, et al. Endoglin and TGF-β signaling in glioblastoma. Cell Tissue Res. 2021;384(3):613–24.PubMedPubMedCentralCrossRef Burghardt I, Ventura E, Weiss T, Schroeder JJ, Seystahl K, Zielasek C, et al. Endoglin and TGF-β signaling in glioblastoma. Cell Tissue Res. 2021;384(3):613–24.PubMedPubMedCentralCrossRef
269.
Zurück zum Zitat Ahluwalia MS, Rogers LR, Chaudhary RT, Newton HB, Seon BK, Jivani MA, et al. A phase 2 trial of TRC105 with bevacizumab for bevacizumab refractory glioblastoma. JCO. 2016;34(15suppl):2035.CrossRef Ahluwalia MS, Rogers LR, Chaudhary RT, Newton HB, Seon BK, Jivani MA, et al. A phase 2 trial of TRC105 with bevacizumab for bevacizumab refractory glioblastoma. JCO. 2016;34(15suppl):2035.CrossRef
270.
Zurück zum Zitat Galanis E, Anderson SK, Butowski NA, Hormigo A, Schiff D, Tran DD, et al. NCCTG N1174: phase I/comparative randomized phase (Ph) II trial of TRC105 plus bevacizumab versus bevacizumab in recurrent glioblastoma (GBM) (Alliance). JCO. 2017;35(15suppl):2023.CrossRef Galanis E, Anderson SK, Butowski NA, Hormigo A, Schiff D, Tran DD, et al. NCCTG N1174: phase I/comparative randomized phase (Ph) II trial of TRC105 plus bevacizumab versus bevacizumab in recurrent glioblastoma (GBM) (Alliance). JCO. 2017;35(15suppl):2023.CrossRef
271.
Zurück zum Zitat Koul D, Shen R, Kim Y-W, Kondo Y, Lu Y, Bankson J, et al. Cellular and in vivo activity of a novel PI3K inhibitor, PX-866, against human glioblastoma. Neuro Oncol. 2010;12(6):559–69.PubMedPubMedCentralCrossRef Koul D, Shen R, Kim Y-W, Kondo Y, Lu Y, Bankson J, et al. Cellular and in vivo activity of a novel PI3K inhibitor, PX-866, against human glioblastoma. Neuro Oncol. 2010;12(6):559–69.PubMedPubMedCentralCrossRef
272.
Zurück zum Zitat Land SC, Tee AR. Hypoxia-inducible factor 1alpha is regulated by the mammalian target of rapamycin (mTOR) via an mTOR signaling motif. J Biol Chem. 2007;282(28):20534–43.PubMedCrossRef Land SC, Tee AR. Hypoxia-inducible factor 1alpha is regulated by the mammalian target of rapamycin (mTOR) via an mTOR signaling motif. J Biol Chem. 2007;282(28):20534–43.PubMedCrossRef
273.
Zurück zum Zitat Hainsworth JD, Becker KP, Mekhail T, Chowdhary SA, Eakle JF, Wright D, et al. Phase I/II study of bevacizumab with BKM120, an oral PI3K inhibitor, in patients with refractory solid tumors (phase I) and relapsed/refractory glioblastoma (phase II). J Neurooncol. 2019;144(2):303–11.PubMedCrossRef Hainsworth JD, Becker KP, Mekhail T, Chowdhary SA, Eakle JF, Wright D, et al. Phase I/II study of bevacizumab with BKM120, an oral PI3K inhibitor, in patients with refractory solid tumors (phase I) and relapsed/refractory glioblastoma (phase II). J Neurooncol. 2019;144(2):303–11.PubMedCrossRef
274.
Zurück zum Zitat Hainsworth JD, Shih KC, Shepard GC, Tillinghast GW, Brinker BT, Spigel DR. Phase II study of concurrent radiation therapy, temozolomide, and bevacizumab followed by bevacizumab/everolimus as first-line treatment for patients with glioblastoma. Clin Adv Hematol Oncol. 2012;10(4):240–6.PubMed Hainsworth JD, Shih KC, Shepard GC, Tillinghast GW, Brinker BT, Spigel DR. Phase II study of concurrent radiation therapy, temozolomide, and bevacizumab followed by bevacizumab/everolimus as first-line treatment for patients with glioblastoma. Clin Adv Hematol Oncol. 2012;10(4):240–6.PubMed
275.
Zurück zum Zitat Nayak L, Hays JL, Muzikansky A, Gaffey SC, Do KT, Puduvalli VK, et al. A phase I study of MLN0128 and bevacizumab in patients with recurrent glioblastoma and other solid tumors. JCO. 2016;34(15suppl):2013.CrossRef Nayak L, Hays JL, Muzikansky A, Gaffey SC, Do KT, Puduvalli VK, et al. A phase I study of MLN0128 and bevacizumab in patients with recurrent glioblastoma and other solid tumors. JCO. 2016;34(15suppl):2013.CrossRef
276.
Zurück zum Zitat Kesari S, Juarez T, Carrillo J, Truong J, Nguyen M, Heng A, et al. RBTT-01. A PHASE 2 TRIAL WITH ABI-009 (NAB-SIROLIMUS) AS SINGLE-AGENT AND COMBINATIONS IN RECURRENT HIGH-GRADE GLIOMA (rHGG) AND IN NEWLY DIAGNOSED GLIOBLASTOMA (ndGBM). Neuro Oncol. 2019;21(Supplement6):vi218–9.PubMedCentralCrossRef Kesari S, Juarez T, Carrillo J, Truong J, Nguyen M, Heng A, et al. RBTT-01. A PHASE 2 TRIAL WITH ABI-009 (NAB-SIROLIMUS) AS SINGLE-AGENT AND COMBINATIONS IN RECURRENT HIGH-GRADE GLIOMA (rHGG) AND IN NEWLY DIAGNOSED GLIOBLASTOMA (ndGBM). Neuro Oncol. 2019;21(Supplement6):vi218–9.PubMedCentralCrossRef
277.
Zurück zum Zitat Djuzenova CS, Fiedler V, Memmel S, Katzer A, Sisario D, Brosch PK, et al. Differential effects of the akt inhibitor MK-2206 on migration and radiation sensitivity of glioblastoma cells. BMC Cancer. 2019;19(1):299.PubMedPubMedCentralCrossRef Djuzenova CS, Fiedler V, Memmel S, Katzer A, Sisario D, Brosch PK, et al. Differential effects of the akt inhibitor MK-2206 on migration and radiation sensitivity of glioblastoma cells. BMC Cancer. 2019;19(1):299.PubMedPubMedCentralCrossRef
278.
Zurück zum Zitat Kaley TJ, Panageas KS, Pentsova EI, Mellinghoff IK, Nolan C, Gavrilovic I, et al. Phase I clinical trial of temsirolimus and perifosine for recurrent glioblastoma. Ann Clin Transl Neurol. 2020;7(4):429–36.PubMedPubMedCentralCrossRef Kaley TJ, Panageas KS, Pentsova EI, Mellinghoff IK, Nolan C, Gavrilovic I, et al. Phase I clinical trial of temsirolimus and perifosine for recurrent glioblastoma. Ann Clin Transl Neurol. 2020;7(4):429–36.PubMedPubMedCentralCrossRef
279.
Zurück zum Zitat Tien A-C, Li J, Bao X, Derogatis A, Kim S, Mehta S, et al. A phase 0 trial of Ribociclib in Recurrent Glioblastoma patients incorporating a Tumor Pharmacodynamic- and pharmacokinetic-guided expansion cohort. Clin Cancer Res. 2019;25(19):5777–86.PubMedPubMedCentralCrossRef Tien A-C, Li J, Bao X, Derogatis A, Kim S, Mehta S, et al. A phase 0 trial of Ribociclib in Recurrent Glioblastoma patients incorporating a Tumor Pharmacodynamic- and pharmacokinetic-guided expansion cohort. Clin Cancer Res. 2019;25(19):5777–86.PubMedPubMedCentralCrossRef
280.
Zurück zum Zitat DeWire M, Fuller C, Campagne O, Lin T, Pan H, Young-Poussaint T, et al. A phase I and surgical study of ribociclib and everolimus in children with recurrent or refractory malignant brain tumors: a Pediatric Brain Tumor Consortium Study. Clin cancer Research: Official J Am Association Cancer Res. 2021;27(9):2442–51.CrossRef DeWire M, Fuller C, Campagne O, Lin T, Pan H, Young-Poussaint T, et al. A phase I and surgical study of ribociclib and everolimus in children with recurrent or refractory malignant brain tumors: a Pediatric Brain Tumor Consortium Study. Clin cancer Research: Official J Am Association Cancer Res. 2021;27(9):2442–51.CrossRef
281.
Zurück zum Zitat Rathkopf DE, Larson SM, Anand A, Morris MJ, Slovin SF, Shaffer DR, et al. Everolimus Combined with Gefitinib in patients with metastatic castration-resistant prostate Cancer: phase I/II results and signaling pathway implications. Cancer. 2015;121(21):3853–61.PubMedCrossRef Rathkopf DE, Larson SM, Anand A, Morris MJ, Slovin SF, Shaffer DR, et al. Everolimus Combined with Gefitinib in patients with metastatic castration-resistant prostate Cancer: phase I/II results and signaling pathway implications. Cancer. 2015;121(21):3853–61.PubMedCrossRef
282.
Zurück zum Zitat Schiff D, Jaeckle KA, Anderson SK, Galanis E, Giannini C, Buckner JC, et al. Phase I/II trial of Temsirolimus and Sorafenib in treatment of patients with recurrent glioblastoma: North Central Cancer Treatment Group Study/Alliance N0572. Cancer. 2018;124(7):1455.PubMedCrossRef Schiff D, Jaeckle KA, Anderson SK, Galanis E, Giannini C, Buckner JC, et al. Phase I/II trial of Temsirolimus and Sorafenib in treatment of patients with recurrent glioblastoma: North Central Cancer Treatment Group Study/Alliance N0572. Cancer. 2018;124(7):1455.PubMedCrossRef
283.
Zurück zum Zitat Chheda MG, Wen PY, Hochberg FH, Chi AS, Drappatz J, Eichler AF, et al. Vandetanib plus Sirolimus in adults with recurrent glioblastoma: results of a phase I and dose expansion cohort study. J Neurooncol. 2015;121(3):627–34.PubMedCrossRef Chheda MG, Wen PY, Hochberg FH, Chi AS, Drappatz J, Eichler AF, et al. Vandetanib plus Sirolimus in adults with recurrent glioblastoma: results of a phase I and dose expansion cohort study. J Neurooncol. 2015;121(3):627–34.PubMedCrossRef
284.
Zurück zum Zitat Skaga E, Skaga IØ, Grieg Z, Sandberg CJ, Langmoen IA, Vik-Mo EO. The efficacy of a coordinated pharmacological blockade in glioblastoma stem cells with nine repurposed drugs using the CUSP9 strategy. J Cancer Res Clin Oncol. 2019;145(6):1495–507.PubMedPubMedCentralCrossRef Skaga E, Skaga IØ, Grieg Z, Sandberg CJ, Langmoen IA, Vik-Mo EO. The efficacy of a coordinated pharmacological blockade in glioblastoma stem cells with nine repurposed drugs using the CUSP9 strategy. J Cancer Res Clin Oncol. 2019;145(6):1495–507.PubMedPubMedCentralCrossRef
285.
Zurück zum Zitat Lun X, Wells JC, Grinshtein N, King JC, Hao X, Dang N-H, et al. Disulfiram when combined with copper enhances the therapeutic effects of Temozolomide for the treatment of Glioblastoma. Clin Cancer Research: Official J Am Association Cancer Res. 2016;22(15):3860–75.CrossRef Lun X, Wells JC, Grinshtein N, King JC, Hao X, Dang N-H, et al. Disulfiram when combined with copper enhances the therapeutic effects of Temozolomide for the treatment of Glioblastoma. Clin Cancer Research: Official J Am Association Cancer Res. 2016;22(15):3860–75.CrossRef
286.
Zurück zum Zitat Triscott J, Lee C, Hu K, Fotovati A, Berns R, Pambid M, et al. Disulfiram, a drug widely used to control alcoholism, suppresses the self-renewal of glioblastoma and over-rides resistance to temozolomide. Oncotarget. 2012;3(10):1112–23.PubMedPubMedCentralCrossRef Triscott J, Lee C, Hu K, Fotovati A, Berns R, Pambid M, et al. Disulfiram, a drug widely used to control alcoholism, suppresses the self-renewal of glioblastoma and over-rides resistance to temozolomide. Oncotarget. 2012;3(10):1112–23.PubMedPubMedCentralCrossRef
287.
Zurück zum Zitat Huang J, Campian JL, Gujar AD, Tsien C, Ansstas G, Tran DD, et al. Final results of a phase I dose-escalation, dose-expansion study of adding disulfiram with or without copper to adjuvant temozolomide for newly diagnosed glioblastoma. J Neurooncol. 2018;138(1):105–11.PubMedCrossRef Huang J, Campian JL, Gujar AD, Tsien C, Ansstas G, Tran DD, et al. Final results of a phase I dose-escalation, dose-expansion study of adding disulfiram with or without copper to adjuvant temozolomide for newly diagnosed glioblastoma. J Neurooncol. 2018;138(1):105–11.PubMedCrossRef
288.
Zurück zum Zitat Jakola AS, Werlenius K, Mudaisi M, Hylin S, Kinhult S, Bartek J Jr., et al. Disulfiram repurposing combined with nutritional copper supplement as add-on to chemotherapy in recurrent glioblastoma (DIRECT): study protocol for a randomized controlled trial. F1000Res. 2018;7:1797.PubMedPubMedCentralCrossRef Jakola AS, Werlenius K, Mudaisi M, Hylin S, Kinhult S, Bartek J Jr., et al. Disulfiram repurposing combined with nutritional copper supplement as add-on to chemotherapy in recurrent glioblastoma (DIRECT): study protocol for a randomized controlled trial. F1000Res. 2018;7:1797.PubMedPubMedCentralCrossRef
289.
Zurück zum Zitat Bota DA, Alexandru D, Keir ST, Bigner D, Vredenburgh J, Friedman HS. Proteasome inhibition with bortezomib induces cell death in GBM stem-like cells and temozolomide-resistant glioma cell lines, but stimulates GBM stem-like cells’ VEGF production and angiogenesis: Laboratory investigation. J Neurosurg. 2013;119(6):1415–23.PubMedPubMedCentralCrossRef Bota DA, Alexandru D, Keir ST, Bigner D, Vredenburgh J, Friedman HS. Proteasome inhibition with bortezomib induces cell death in GBM stem-like cells and temozolomide-resistant glioma cell lines, but stimulates GBM stem-like cells’ VEGF production and angiogenesis: Laboratory investigation. J Neurosurg. 2013;119(6):1415–23.PubMedPubMedCentralCrossRef
290.
Zurück zum Zitat Rashid M, Toh TB, Hooi L, Silva A, Zhang Y, Tan PF et al. Optimizing drug combinations against multiple myeloma using a quadratic phenotypic optimization platform (QPOP). Sci Transl Med. 2018;10(453). Rashid M, Toh TB, Hooi L, Silva A, Zhang Y, Tan PF et al. Optimizing drug combinations against multiple myeloma using a quadratic phenotypic optimization platform (QPOP). Sci Transl Med. 2018;10(453).
291.
Zurück zum Zitat Nowak-Sliwinska P, Weiss A, Ding X, Dyson PJ, van den Bergh H, Griffioen AW, et al. Optimization of drug combinations using Feedback System Control. Nat Protoc. 2016;11(2):302–15.PubMedCrossRef Nowak-Sliwinska P, Weiss A, Ding X, Dyson PJ, van den Bergh H, Griffioen AW, et al. Optimization of drug combinations using Feedback System Control. Nat Protoc. 2016;11(2):302–15.PubMedCrossRef
292.
Zurück zum Zitat Ma S, Dang D, Wang W, Wang Y, Liu L. Concentration optimization of combinatorial drugs using Markov chain-based models. BMC Bioinformatics. 2021;22(1):451.PubMedPubMedCentralCrossRef Ma S, Dang D, Wang W, Wang Y, Liu L. Concentration optimization of combinatorial drugs using Markov chain-based models. BMC Bioinformatics. 2021;22(1):451.PubMedPubMedCentralCrossRef
293.
Zurück zum Zitat Zou J, Ji P, Zhao Y-L, Li L-L, Wei Y-Q, Chen Y-Z, et al. Neighbor communities in drug combination networks characterize synergistic effect. Mol Biosyst. 2012;8(12):3185–96.PubMedCrossRef Zou J, Ji P, Zhao Y-L, Li L-L, Wei Y-Q, Chen Y-Z, et al. Neighbor communities in drug combination networks characterize synergistic effect. Mol Biosyst. 2012;8(12):3185–96.PubMedCrossRef
294.
Zurück zum Zitat Stathias V, Jermakowicz AM, Maloof ME, Forlin M, Walters W, Suter RK, et al. Drug and disease signature integration identifies synergistic combinations in glioblastoma. Nat Commun. 2018;9(1):5315.PubMedPubMedCentralCrossRef Stathias V, Jermakowicz AM, Maloof ME, Forlin M, Walters W, Suter RK, et al. Drug and disease signature integration identifies synergistic combinations in glioblastoma. Nat Commun. 2018;9(1):5315.PubMedPubMedCentralCrossRef
295.
Zurück zum Zitat Alhalabi OT, Fletcher MNC, Hielscher T, Kessler T, Lokumcu T, Baumgartner U, et al. A novel patient stratification strategy to enhance the therapeutic efficacy of dasatinib in glioblastoma. Neuro Oncol. 2022;24(1):39–51.PubMedCrossRef Alhalabi OT, Fletcher MNC, Hielscher T, Kessler T, Lokumcu T, Baumgartner U, et al. A novel patient stratification strategy to enhance the therapeutic efficacy of dasatinib in glioblastoma. Neuro Oncol. 2022;24(1):39–51.PubMedCrossRef
296.
Zurück zum Zitat Weiss T, Weller M. Pathway-based stratification of glioblastoma. Nat Rev Neurol. 2021;17(5):263–4.PubMedCrossRef Weiss T, Weller M. Pathway-based stratification of glioblastoma. Nat Rev Neurol. 2021;17(5):263–4.PubMedCrossRef
297.
Zurück zum Zitat Chevaleyre C, Novell A, Tournier N, Dauba A, Dubois S, Kereselidze D, et al. Efficient PD-L1 imaging of murine glioblastoma with FUS-aided immunoPET by leveraging FcRn-antibody interaction. Theranostics. 2023;13(15):5584–96.PubMedPubMedCentralCrossRef Chevaleyre C, Novell A, Tournier N, Dauba A, Dubois S, Kereselidze D, et al. Efficient PD-L1 imaging of murine glioblastoma with FUS-aided immunoPET by leveraging FcRn-antibody interaction. Theranostics. 2023;13(15):5584–96.PubMedPubMedCentralCrossRef
298.
Zurück zum Zitat Maynard J, Hart P. The opportunities and Use of Imaging to measure Target Engagement. SLAS Discov. 2020;25(2):127–36.PubMedCrossRef Maynard J, Hart P. The opportunities and Use of Imaging to measure Target Engagement. SLAS Discov. 2020;25(2):127–36.PubMedCrossRef
299.
Zurück zum Zitat Menke-van der Houven, van Oordt CW, McGeoch A, Bergstrom M, McSherry I, Smith DA, Cleveland M, et al. Immuno-PET imaging to assess Target Engagement: experience from (89)Zr-Anti-HER3 mAb (GSK2849330) in patients with solid tumors. J Nucl Med. 2019;60(7):902–9.CrossRef Menke-van der Houven, van Oordt CW, McGeoch A, Bergstrom M, McSherry I, Smith DA, Cleveland M, et al. Immuno-PET imaging to assess Target Engagement: experience from (89)Zr-Anti-HER3 mAb (GSK2849330) in patients with solid tumors. J Nucl Med. 2019;60(7):902–9.CrossRef
300.
Zurück zum Zitat Pantel AR, Gitto SB, Makvandi M, Kim H, Medvedv S, Weeks JK, et al. [18F]FluorThanatrace ([18F]FTT) PET imaging of PARP-Inhibitor drug-target Engagement as a biomarker of response in Ovarian Cancer, a pilot study. Clin Cancer Res. 2023;29(8):1515–27.PubMedCrossRef Pantel AR, Gitto SB, Makvandi M, Kim H, Medvedv S, Weeks JK, et al. [18F]FluorThanatrace ([18F]FTT) PET imaging of PARP-Inhibitor drug-target Engagement as a biomarker of response in Ovarian Cancer, a pilot study. Clin Cancer Res. 2023;29(8):1515–27.PubMedCrossRef
Metadaten
Titel
Mechanistic insights and the clinical prospects of targeted therapies for glioblastoma: a comprehensive review
verfasst von
Yating Shen
Dexter Kai Hao Thng
Andrea Li Ann Wong
Tan Boon Toh
Publikationsdatum
01.12.2024
Verlag
BioMed Central
Erschienen in
Experimental Hematology & Oncology / Ausgabe 1/2024
Elektronische ISSN: 2162-3619
DOI
https://doi.org/10.1186/s40164-024-00512-8

Weitere Artikel der Ausgabe 1/2024

Experimental Hematology & Oncology 1/2024 Zur Ausgabe

Erhebliches Risiko für Kehlkopfkrebs bei mäßiger Dysplasie

29.05.2024 Larynxkarzinom Nachrichten

Fast ein Viertel der Personen mit mäßig dysplastischen Stimmlippenläsionen entwickelt einen Kehlkopftumor. Solche Personen benötigen daher eine besonders enge ärztliche Überwachung.

15% bedauern gewählte Blasenkrebs-Therapie

29.05.2024 Urothelkarzinom Nachrichten

Ob Patienten und Patientinnen mit neu diagnostiziertem Blasenkrebs ein Jahr später Bedauern über die Therapieentscheidung empfinden, wird einer Studie aus England zufolge von der Radikalität und dem Erfolg des Eingriffs beeinflusst.

Erhöhtes Risiko fürs Herz unter Checkpointhemmer-Therapie

28.05.2024 Nebenwirkungen der Krebstherapie Nachrichten

Kardiotoxische Nebenwirkungen einer Therapie mit Immuncheckpointhemmern mögen selten sein – wenn sie aber auftreten, wird es für Patienten oft lebensgefährlich. Voruntersuchung und Monitoring sind daher obligat.

Costims – das nächste heiße Ding in der Krebstherapie?

28.05.2024 Onkologische Immuntherapie Nachrichten

„Kalte“ Tumoren werden heiß – CD28-kostimulatorische Antikörper sollen dies ermöglichen. Am besten könnten diese in Kombination mit BiTEs und Checkpointhemmern wirken. Erste klinische Studien laufen bereits.

Update Onkologie

Bestellen Sie unseren Fach-Newsletter und bleiben Sie gut informiert.