Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/275836977

Glioblastoma Multiforme: Pathogenesis and Treatment

Article  in  Pharmacology [?] Therapeutics · May 2015


DOI: 10.1016/j.pharmthera.2015.05.005

CITATIONS READS

120 610

2 authors:

Constantinos Alifieris Dimitris Trafalis


National and Kapodistrian University of Athens National and Kapodistrian University of Athens
14 PUBLICATIONS   150 CITATIONS    108 PUBLICATIONS   712 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Study of Hybrid A & D-modified Steroidal Alkylator Anticancer Activity View project

Alzheimer's disease drug development View project

All content following this page was uploaded by Dimitris Trafalis on 01 January 2019.

The user has requested enhancement of the downloaded file.


Pharmacology & Therapeutics 152 (2015) 63–82

Contents lists available at ScienceDirect

Pharmacology & Therapeutics

journal homepage: www.elsevier.com/locate/pharmthera

Associate editor: M. Panagiotidis

Glioblastoma multiforme: Pathogenesis and treatment


Constantinos Alifieris, Dimitrios T. Trafalis ⁎
Laboratory of Pharmacology, Medical School, University of Athens, Athens, Greece

a r t i c l e i n f o a b s t r a c t

Available online 2 May 2015 Each year, about 5–6 cases out of 100,000 people are diagnosed with primary malignant brain tumors, of which
about 80% are malignant gliomas (MGs). Glioblastoma multiforme (GBM) accounts for more than half of MG
Keywords: cases. They are associated with high morbidity and mortality. Despite current multimodality treatment efforts in-
Glioblastoma cluding maximal surgical resection if feasible, followed by a combination of radiotherapy and/or chemotherapy,
Treatment the median survival is short: only about 15 months. A deeper understanding of the pathogenesis of these tumors
Review
has presented opportunities for newer therapies to evolve and an expectation of better control of this disease.
Pharmacology
Targeted therapy
Lately, efforts have been made to investigate tumor resistance, which results from complex alternate signaling
Stem cell pathways, the existence of glioma stem-cells, the influence of the blood-brain barrier as well as the expression of
06-methylguanine-DNA methyltransferase. In this paper, we review up-to-date information on MGs treatment in-
cluding current approaches, novel drug-delivering strategies, molecular targeted agents and immunomodulative
treatments, and discuss future treatment perspectives.
© 2015 Elsevier Inc. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
2. Pathogenesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
3. Treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
4. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
Conflict of interest statement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77

1. Introduction
Abbreviations: AKT, protein kinase B; BBB, blood-brain barrier; BCNU, carmustin; CCNU,
lomustin; c-MET, hepatocyte growth factor receptor; EGF(R), epidermal growth factor
(receptor); EGFRvIII, epidermal growth factor receptor variant III; ERK, extracellular Each year, about 5–6 cases out of 100,000 people are diagnosed with
signal-regulated kinases; FTI, farnesyltransferase inhibitor; GBM, glioblastoma multiforme; primary malignant brain tumors, of which about 80% are malignant
Gli1, glioma-associated oncogene-1; HDAC, histone deacetylase 1; HER2, human epidermal gliomas (MGs) (Schwartzbaum et al, 2006; Stupp et al., 2010b).
growth factor receptor 2; HGF, hepatocyte growth factor; Hsp, heat-shock protein; IDH1, Gliomas, the most common group of primary brain tumours, include
NADP + -dependent isocitrate dehydrogenase; MAb, monoclonal antibody; MAPK, mito-
gen-activated protein kinases; MG, malignant glioma; MGMT, O6-methylguanine-DNA
astrocytomas, oligodendrogliomas and ependymomas. According to
methyltransferase; miRNA, micro ribonucleic acid; MRI, magnetic resonance imaging; World Health Organization (WHO) malignant gliomas are sub-
MTKI, multitargeted tyrosine kinase inhibitor; mTOR, mammalian target of rapamycin; categorized into grade III/IV tumors such as anaplastic astrocytoma,
OS, overall survival; PARP-1, poly(ADP-ribose)polymerase-1; PCV, procarbazine, lomustine anaplastic oligodendroglioma, anaplastic oligoastrocytoma and ana-
and vincristine; PDGF(R), platelet-derived growth factor (receptor); PFS, progression-free
plastic ependymomas, as well as grade IV/IV tumors, as glioblastoma
survival; PFS-6, six-month progression-free survival; PI3K, phosphotidylinositol 3-kinase;
PTEN, phosphatase and tensin homolog; RAF, rapidly accelerated fibrosarcoma; RAS, rat sar- multiforme (GBM). The WHO grade is assigned based on certain patho-
coma; RT, radiotherapy; SHH, sonic hedgehog; STAT3, signal transducer and activator of logical features, such as nuclear atypia, mitotic activity, vascular prolif-
transcription 3; TKI, tyrosine kinase inhibitor; TMZ, temozolomide; TTP, time-to-progres- eration, necrosis, proliferative potential and features clinical course
sion; VEGF(R), vascular endothelial growth factor (receptor). and treatment outcome (Louis et al., 2007). Its incidence in the United
⁎ Corresponding author at: Lab. Of Pharmacology, Medical School, University of Athens,
75 Mikras Asias Str., 11527-Goudi, Athens, Greece (Building 16, 1st floor). Tel.: +30 210
States is estimated around 3:100,000 while more than 10,000 cases
746 2587; fax: +30 210 746 2504. are diagnosed annually. It constitutes 45.2% of all malignant central ner-
E-mail address: dtrafal@med.uoa.gr (D.T. Trafalis). vous system (CNS) tumors, 80% of all primary malignant CNS tumors

http://dx.doi.org/10.1016/j.pharmthera.2015.05.005
0163-7258/© 2015 Elsevier Inc. All rights reserved.
64 C. Alifieris, D.T. Trafalis / Pharmacology & Therapeutics 152 (2015) 63–82

and approximately 54.4% of all malignant gliomas. Mean age at diagnosis 2. Pathogenesis
is 64 years and it is 1.5 times more common in men than women and 2
times more common in whites compared to blacks (Ostrom et al., The ongoing research on the pathogenesis of malignant gliomas has
2013). The incidence has increased slightly over the past 20 years mostly given opportunities for newer therapies to evolve as well as promises
due to improved radiologic diagnosis and especially in elderly (Fisher for better control of the disease. Efforts are given to understand the de-
et al., 2007). In terms of treatment, grade III tumors and GBM are velopment of tumor resistance (Dean et al., 2005; Furnari et al., 2007).
grouped together and treated similarly. A small subgroup (about 5%) of patients with gliomas, is associated
Clinically, patients with GBM may present with headaches, focal with certain hereditary syndromes (Farrell & Plotkin, 2007) (Table 1).
neurologic deficits, confusion, memory loss, personality changes or All other patients with gliomas represent sporadic cases. An important
with seizures. Diagnosis and treatment response is suggested by aspect of the pathogenesis of gliomas is that malignant transformation
magnetic resonance imaging (MRI) and the use of adjunct technology results from the sequential accumulation of genetic alterations and ab-
such as functional MRI, diffusion-weighted imaging, diffusion tensor normal regulation of growth factor signaling pathways. Aberrant prolif-
imaging, dynamic contrast-enhanced MRI, perfusion imaging, proton erations is thus mediated via molecules such as vascular endothelial
magnetic resonance spectroscopy and positron-emission tomography growth factor (VEGF), epidermal growth factor (EGF), platelet-derived
(Wen & Kesari, 2008). growth factor (PDGF), hepatocyte growth factor (HGF) and loss of
Etiologically, there are known linked risk factors that lead to phosphotensin analogue (PTEN). Downstream cascades in the growth
development of GBM Environmental risk factors include primarily signaling pathways (including PI3K/AKT) may be triggered (Wen &
exposure to therapeutic ionizing radiation and factors such as Kesari, 2008). Also, when those tumors recur, they often show progres-
vinyl chloride or pesticides, smoking, petroleum refining or pro- sion to a higher histologic grade and thus, acquire a different name; this
duction work and employment in synthetic rubber manufacturing actually is a representation of progression along a classification scheme,
(Wrensch et al., 2002). Additional factors such as exposure to res- the natural course of the disease, rather than a new disease (Louis,
idential electromagnetic fields, formaldehyde, diagnostic irradiation 2006). Among the alterations that are most frequently seen in low-
and cell phones have not been proven to lead to GBM. However, regard- grade astrocytomas are mutations affecting p53 and overexpression of
ing cell phone irradiation, a metanalysis released in 2007 did show in- platelet-derived growth factor α (PDGF-α) and its receptor. The transi-
creased incidence among people who used cell phones for at least tion to a higher grade is associated to disruption of RB, p16/CDKNαA
10 years and especially those who had mostly unilateral exposure and 19q tumor suppressor genes (Louis, 2006).
(Hardell et al., 2007).
Currently, maximal surgical resection plus radiotherapy plus con- 2.1. GBM clinical subtypes
comitant and adjuvant temozolomide or carmustin wafers (Gliadel) is
the standard of care in patients younger than 70 years old with newly GBM, is usually described in two different clinical forms, primary
diagnosed GBM. However, recurrence seems to be the rule despite stan- and secondary GBM (Kleihues & Ohgaki, 1999). Primary GBM, is the
dard care. Lately, attention has been given to understand the initial mo- most common form (about 95%) and arises typically de novo, within
lecular pathogenesis of these tumors including alterations in cellular 3–6 months, in older patients. Secondary GBM arises from prior low-
signal transduction pathways, the occurrence of resistance to therapy grade astrocytomas (over 10–15 years) in younger patients. While pri-
and to find methods to penetrate easier the natural blood-brain barrier mary and secondary forms show some molecular differences, the end
(BBB). Despite these efforts to treat however, it remains an incurable result is pretty much the same since the same pathways are affected
disease and the prognosis falls in a poor survival range of 12–15 months and respond similarly to current standard treatment. Primary GBM
(median 14.6 months) and a mean survival rate of only 3.3% at 2 years often has amplified, mutated epidermal-growth factor receptor
and 1.2% at 3 years (Scott et al., 1998; Stupp et al., 2005). (EGFR) which encodes altered EGFR (known as EGFRvIII) whereas sec-
Glioma stem cells contribute to resistance to standard radiotherapy ondary GBM has increased signaling through PDGF-A receptor. Both
via preferential activation of DNA-damage-response pathways; and to types of mutations lead to increased tyrosine kinase receptor (TKR) ac-
standard chemotherapy via O6-methylguanine-DNA methyltransferase tivity and consequently to activation of RAS and PI3K pathways. Again,
(MGMT), the inhibition of apoptosis and the up-regulation of multidrug primary GBM have commonly amplification of MDM2 gene (encodes
resistance genes (Dean et al., 2005). Thus, current efforts are directed for an inhibitor of p53), PTEN mutations and homozygous deletions of
towards personalized treatment through blocking prime signaling CDKN2A whereas secondary GBM usually has more prevalent p53
pathways in gliomagenesis, surpassing acquired resistance and by mutations, IDH1 mutations, MET amplification and overexpression of
penetration of BBB. In this article we review the current concepts as PDGFRA. Finally, progression of low-grade glioma to high-grade is asso-
well as emerging advances in the treatment of GBM with an emphasis ciated with inactivation of the retinoblastoma gene (RB1) and increased
on chemotherapy and targeted agents. activity of human double minute 2 (HDM2) (Kleihues & Ohgaki, 1999;

Table 1
Hereditary risk factors for GBM.

Syndrome Gene Associations

Neurofibromatosis type 1 (NF1) NF1 Neurofibromas


Neurofibromatosis type 2 (NF2) NF2 Schwannomas, ependymomas, meningiomas
Li-Fraumeni syndrome TP53 Sarcomas, breast cancer, leukemia, adrenal cortex
carcinoma, medulloblastomas
Hereditary non-polyposis colorectal cancer (HNPCC/Lynch syndrome), DNA mismatch repair Colorectal adenomas, adenocarcinoma
Turcot syndrome/Brain tumor-polyposis syndrome (BTPS) (MSH2, MLH1, MSH6, PMS1, PMS2)
Multiple endocrine neoplasia type 1 (MEN1) MEN1 Primary hyperparathyroidism, pancreatic
endocrine tumors, pituitary adenomas
Nevoid basal cell carcinoma syndrome (NBCCS), Gorlin-Gotz syndrome PTC Basal cell carcinomas, medulloblastomas,
CH ovarian fibromas
Tuberous sclerosis complex (TSC) TSC1, TSC2 Renal angiomyolipomas, retinal glial hamartomas,
cardiac rhabdomyomas
C. Alifieris, D.T. Trafalis / Pharmacology & Therapeutics 152 (2015) 63–82 65

Louis, 2006). These aberrations primarily lead to abnormal regulation of stem cells. Glioma-associated oncogene-1 (Gli1) –originally thought of
two major cellular systems –the growth factor mediated signaling path- being an oncogene involved in the pathogenesis of GBM- is part of
ways and the cell cycle- and play a role in the increased cell prolifera- SHH pathway. Normally, SHH acts via activation of the Gli1 transcrip-
tion, inhibition of apoptosis, invasion and angiogenesis (Louis ., 2006). tion factor and thus participates in maintenance and proliferation of
Mutation in the nicotinamide adenine dinucleotide phosphate neural stem cells (Clement et al., 2007).
(NADP+)-dependent isocitrate dehydrogenase (IDH1) is conferring fa- Cancer stem cells can be useful as targets by means of immunother-
vorable prognostic value and prediction of response to temozolomide apy or inhibition of important specific pathways, as with signal trans-
in GBM. However, patients with wild type IDH1 anaplastic astrocytomas ducer and activator of transcription 3 (STAT 3) or the activator bone
exhibit worse prognosis than IDH1-mutated glioblastomas, and IDH1 marrow x-linked kinase (BMX) (Nduom et al., 2012).
mutation status accounts for the unfavorable prognostic effect of glio- Targeting of stem cell signaling pathways is an evolving field which
mas (Hartmann et al., 2010). IDH1 mutations were found in less than already shows some promising results. Innovative therapeutic re-
5% of primary GBM and characteristically in greater than 80% of second- sponses as well as preventive approaches may be directed by this type
ary GBM. A recent meta-analysis study associated IDH1 mutations in of cells. Molecular GBM phenotyping has broadened the list of targets
GBM with improved OS (H. Cheng et al., 2013). in tumorigenic pathways. In GBM several tumor suppressor and onco-
genic genes seem to be modulated by miRNAs thus delivery of genes
2.2. GBM molecular subtypes or microRNAs (miRNA) can be another form of treatment (Møller
et al., 2013).
Based on the growing understanding of molecular heterogeneity in
GBM, the Cancer Genome Atlas (TCGA) divided this tumor in molecular
subclasses termed classical, mesenchymal, proneural and neural types 3. Treatment
based on genetic alterations and expression profiles especially of
PDGFRA, EGFR, NF1 and IDH1. Classical GBM is defined by aberrant Currently, the standard approach in managing GBM includes consid-
EGFR amplification, astrocytic cell expression pattern and loss of chro- eration of maximum surgical resection -considering that the entire
mosome 10 whereas IDH1, TP53 or NF1 mutations are not common. tumor cannot be removed because GBM infiltrates surrounding tissue-
The mesenchymal subtype is defined by NF1 and PTEN mutations, a radiation therapy (RT) and medical management/chemotherapy. It
mesenchymal expression profile and less EGFR amplification than in also includes symptomatic treatment of seizures, cerebral edema, infec-
other GBM types. The proneural subtype is characterized by PDGFRA tions, depression, cognitive dysfunction, fatigue and venous thrombo-
focal amplification, TP53 and IDH1 mutations with an oligodenrocytic embolism (Wen et al., 2006a). However, the analysis of symptom
cell expression profile and younger presentation age. Finally, the neural palliation is not in the scope of this review.
subtype, is characterized by normal brain tissue gene expression profile The best therapeutic target is to pursuit a more individualized per-
as well as astrocytic/oligodendrocytic cell markers (Verhaak et al., spective. Thus, treatment depends on several factors such as the time
2010). Primary and secondary GBM may be indistinguishable histolog- of diagnosis, new onset or recurrence, the performance status and the
ically but apparently differ in genetic and epigenetic profiles. Most age of the patient. Most GBM patients will follow a standard treatment
GBM tumors with IDH1 mutations have the proneural gene expression after surgical resection consisting of external beam irradiation five times
pattern but only 30% of preneural GBM has the IDH1 mutation. Thus, a week for six weeks, as well as oral temozolomide daily. Most patients
IDH1 mutation is a more reliable and definitive molecular diagnostic cri- will recur within 6.9 months of initial diagnosis. New agents or treat-
terion of secondary GBM compared to clinical criteria. Secondary GBM ment administration techniques are typically initially tested first in
(almost always the proneural subtype) is a genetically more homoge- the recurrent setting where there are some approved treatment alterna-
nous group regarding IDH1 mutation as compared to primary GBM tives. Table 2 depicts a summary of the current standard options for
(Ohgaki & Kleihues, 2013). This heterogeneity of GBM profiles leads to treatment of newly diagnosed GBM. In patients with either clinical or
different treatment efficacy among patients indicating that therapy radiological progression, various treatment approaches have been
must be personalized to target each patient’s alterations in the molecu- assessed. As part of the multidisciplinary management of high grade gli-
lar level. omas, options include further surgery in selected patients, re-irradiation
if also eligible, systemic chemotherapy, and clinical trial participa-
2.3. Stem cells tion (Butowski et al., 2006; Wen & Kesari, 2008). As most patients
are ineligible for re-operation or re-irradiation, clinical trial partici-
The need for better understanding and managing patients with pation is the best option. Chemotherapy options (given as single
GBM, has led to the search for a defined cell type to target on. The dis- agents) for high-grade gliomas include carboplatin, irinotecan,
covery of multi-potent neural stem cells within the CNS which are capa- carmustine (BCNU), etoposide and procarbazine,lomustine and
ble of proliferation, self-renewal and differentiation, has led to the vincristine (PCV) combination. Exciting preliminary data from clini-
suggestion that a specific transformed CNS cell exists, which has the re- cal phase I/II trials support the development of epidermal growth
spective biological attributes of a somatic stem cell and may behave as a factor receptor (EGFR) antagonists, mammalian target of rapamycin
tumor initiating cell (TIC) (Facchino et al., 2011). These stem cells ex- (mTOR) inhibitors, antiangiogenesis agents and other targeted inhibi-
press surface cluster designation markers such as CD133. Extensive tors (Villano et al., 2009). Even further, investigational approaches
studies, have led to the conclusion that those glioma stem cells promote using convection enhanced delivery, stem cell treatment, immunother-
angiogenesis by producing VEGF and can differentiate into pericytes, apy and gene therapy based on recent pathogenetical discoveries, give
thus they create an optimal microenvironment for survival (L. Cheng excitement and promise for the future.
et al., 2013).
Mutations leading in gliomagenesis can occur in neural stem cells
(Koso et al., 2012). Possibly neural stem cells pass on mutations to
downstream cells such as oligodendrocyte precursor cells (OPC),
which are putative glioma cells. Introducing Neurofibromin-1 and p53 Table 2
Current treatment of newly diagnosed GBM#.
mutations into OPCs produces gliomagenesis in vitro (as occurs in Li-
Fraumeni and Neurofibromatosis-1) (Liu et al., 2011). Sonic hedgehog Maximal surgical resection plus i) RT plus concomitant and andjuvant TMZ
(SHH) signaling pathway, is presumed to be involved in the resistance ii)RT plus concomitant and adjuvant BCNU

(both chemotherapy and radiotherapy) and self renewal of cancer #


: See text for dosages, RT: radiotherapy, TMZ: temozolomide, BCNU: carmustine wafers.
66 C. Alifieris, D.T. Trafalis / Pharmacology & Therapeutics 152 (2015) 63–82

3.1. Chemotherapy for glioblastoma multiforme with metronomic rather than standard schedule as well as with high
average daily dose (N100 mg/m2) (Chen et al., 2013). Patients with pro-
Cytotoxic therapy for GBM has evolved, more due to the approval of moter MGMT methylation treated with RT/TMZ have a median survival
temozolomide -an alkylating agent- for newly diagnosed GBM. Active of 23 months and 5- year OS of 14% versus RT alone (15 months and 5%
agents include also the nitrosureas: carmustine (BCNU) and lomustine respectively) (Stupp et al., 2009).
(CCNU), platinum agents, etoposide, irinotecan and PCV combination The nitrosureas, carmustine (1,2 bis[2-chloreoethyl]-1-nitrosurea
(Table 3). BCNU) and lomustine (CCNU), are two alkylating drugs used in the
Temozolomide (TMZ) is a newer oral alkylating agent that has treatment of GBM and are associated with nausea, vomiting, rash, pul-
excellent penetration into the central nervous system. It is an monary fibrosis (rare) and a delayed myelosuppression and typically
imidazotetrazine derivative of dacarbazine. It has 96-100% bioavailability given in the dosage of 80 mg m−2 day−1 i.v. on days 1–3 every 6–8
and promotes the methylation of the O6 position on guanine (N7-guanine weeks with radiation (Green et al., 1983). Single agent BCNU is used
and N3 adenine). It was approved in 1999 for usage against malignant most often as second line treatment in GBM that has progressed after
gliomas. Major toxicities include nausea and myelosuppression (often TMZ. Carmustine biodegradable wafers (Gliadel) are an implantable
low platelet counts). Usually, an oral 5-HT3 antagonist is given 30–60 depot form of BCNU that are placed in the cavity that is formed after re-
minutes prior to each dose. Dosage is usually started at 75 mg/m2 daily section of newly diagnosed or recurrent tumor. The wafers release top-
concurrently with 6 weeks of regional radiotherapy to the surgical cavity ically BCNU for about 3 weeks. Although they are considered relatively
and followed by 6 adjuvant cycles with maintenance dose at 150 mg/m2 safe they are not used by most centers because of delayed wound
daily p.o. for 5 days for the first cycle and if well tolerated then scaled up healing, intracranial edema, cerebrospinal fluid leakage, intracranial in-
to 200 mg m−2 day−1 for 5 days in 28-day cycles (Chinot et al., 2001; fection and seizures have been reported. Also, alterations in blood-
Villano et al., 2009). Concurrent RT and TMZ results in a median overall brain-barrier make the interpretation of MRI unreliable (Nagpal, 2012).
survival (OS) of 14.6 months and 2-year survival rate of 26.5%. Temozolo- Irinotecan, etoposide and cisplatin have been used in the treatment
mide can be effective for recurrent GBM while its efficacy can be increased of GBM demonstrating modest efficacy as adjuvant chemotherapy.

Table 3
Therapeutic categories.

Chemotherapy Alkylating drugs Temozolomide

Nitrosureas (carmustine, lomustine)


Platinum compounds Irinotecan, etoposide, cisplatin
Vinca alkaloids Vincristine (part of PCV regimen)
Other chemotherapeutic agents ANG1005, Dichloroacetate, RTA 744
Targeted Therapy EGFR 1) TKIs (erlotinib, gefitinib, afatinib, dacomitinib)
2) MTKIs (lapatinib, vandetanib, Bay846)
3) MAbs (cetuximab, MAb 806, nimotuzumab, panitumumab, AMG 595)
VEGFR 1) MTKIs (vatalanib, vandetanib, sunitinib, lenvatinib, pazopanib, cediranib, cabozantinib, sorafenib)
2) MAbs (bevacizumab, ramucirumab, IMC-18 F1
3) VEGF trap (aflibercept)
PDGFR 1) MTKIs (imatinib, dasatinib, nilotinib, tandutinib, bosutinib, sunitinib)
2) sTKI (crenolanib)
3) MAbs (IMC-3G3, MEDI-575)
HGFR 1) MTKIs (cabozantinib, vandetanib, crizotinib, amuvatinib)
2) sTKI (SGX523)
3) MAbs (rilotunumab, onartuzumab)
Integrins 1) Cilengitide
2) Lenalidomide, thalidomide
Intracellular pathways RAS/RAF/MEK/MAPK 1) FTIs (Tipifarnib, lonafarnib, manumycin)
2) RAF/MEK/MAPK(sorafenib,TLN-4601)
PI3K/AKT/mTOR 1) PI3K (XL-147, buparlisib)
2) AKT (MK-2206, perifosine)
3) mTOR (temsirolimus, sirolimus, everolimus, ridaforolimus)
4) PI3K/mTOR (XL-765, PI-103, PKI-587)
PKC Tamoxifen, enzastaurin
SHH/Gli1 Cyclopamine, SEN450, vismodegib, rismodegib
NOTCH Gamma Secretase Inhibitors (RO4929097, MK-0752, DAPT)
Other molecular targets HDACs Vorinostat, panobinostat, romidepsin, trichostatin A
Hsp Tanespimycin
Proteasome Bortezomib, marizomib, MG132
TGF-beta Trabedersen, LY2109761
SRC, SFKs dasatinib
Glutamate Talampanel, riluzole
Topoisomerase inhibitors 1) Topoisomerase 1 (captothecin, irinotecan, NKTR-102)
2) Topoisomerase 2 (etoposide, teniposide)
PARP-1 Iniparib, ABT-888
Other agents C-150, aprepitant, artesunate, auranofin, disulfiram, nelfinavir, sertraline, captopril, ketoconazole,
copper gluconate
Stem cell treatment Cell surface molecule targets L1/CAM antagonism
Viral vectors Delta-24-RGD oncolytic adenovirus
Immunotherapy Vaccines 1) Peptide vaccines (Rindopepimut, ITK-1, HSPPV, Vitespen)
2) Gene-modified tumor vaccines
Dendritic cells 1) microglia as therapeutic vectors
2) viral vectors as gene therapy for microglia
JAK2/STAT3 pathway inhibitors WP1193, WP1066
C. Alifieris, D.T. Trafalis / Pharmacology & Therapeutics 152 (2015) 63–82 67

Since the approval of TMZ, they are mostly used as treatment after pro- (Pgp) (De faria et al., 2008). Moreover, cancer stem cells express
gression or recurrence. Since bevacizumab (see below) is increasingly MDR1 at higher levels than differentiated cancer cells, favoring resis-
being used in the recurrent setting, combinations of platinum com- tance to chemotherapy agents such as temozolomide, etoposide, pacli-
pounds with bevacizumab have been used (Carrillo & Munoz, 2012). taxel and carboplatin in undifferentiated cells and thus serving as an
The PCV regimen combining procarbazine, lomustine (CCNU) and adequate reasoning for this high frequency of tumor recurrence with
vincristine is commonly used in the treatment of anaplastic astrocyto- existing treatment (Liu et al., 2005). TMZ also results in induction of
mas and oligodendrogliomas. Procarbazine is also a monoamine oxidase EGFRvIII and EGFR expansion, which can be blunted by the use of
inhibitor and is associated with allergies and hypertensive crisis espe- EGFR inhibitors such as erlotinib (Munoz et al., 2014). Medication
cially in response to certain foods such as those containing tyramine – such as verapamil and cyclosporine A that inhibit p-glycoprotein
aged cheese, nuts, red wine – which patients are advised to avoid con- in vitro, can increase brain retention of chemotherapy drugs such as vin-
suming. Vincristine is a vinca alkaloid (binds to tubulin), is associated cristine (Linnet & Ejsing, 2008).
with a syndrome of jaw pain after the first dose and maybe peripheral
neuropathy. This combination is usually given as CCNU 110 mg/m2 3.3. Drug molecular targets in glioblastoma multiforme
p.o. on day 1, procarbazine 60 mg m−2 day−1 p.o. on days 8–21 and vin-
cristine 1.4 mg/m2 (maximum dose 2 mg) i.v. on days 8 and 29, repeat- The continuously evolving field of understanding the molecular
ed every 6 weeks (Glass et al., 1992). PCV in recurrent GBM has not pathogenesis of GBM prompted to a more rational use of targeted mo-
shown significant results (Schmidt et al., 2006). lecular therapies. Inhibitors that target receptor tyrosine kinases (EGF,
Investigational cytotoxic agents: GRN1005 (ANG1005) is a novel PDGF, VEGF, HGF, IGF receptors) and signal transduction inhibitors
agent that consists of three paclitaxel molecules linked to a novel pep- targeting mTOR, AKT/PI3K and farnesyltransferase are being investigat-
tide – angiopep – that allows it to be transported across the blood- ed. Unfortunatelly, up till now, monotherapy with these agents show
brain barrier through the low-density lipoprotein-related protein 1 re- disappointing results or at most mild-modest efficacy with response
ceptor. In a phase I study, prodrug GRN1005 was tried in patients with that is between 0 to 15% and are unable to prolong the progression
recurrent or progressive malignant gliomas and it was fairly well toler- free survival (Sathornsumetee et al., 2007). Because highly targeted
ated while showed antitumor activity. It penetrated the blood-brain agents were not being successful, the current approach is combined in-
barrier (Drappatz et al., 2013). Dichloroacetate (DCA) an inhibitor of hibition of multiple molecular targets within the same pathway (prox-
the mitochondrial pyruvate dehydrogenase kinase induces apoptosis imal and distal) or between separate-parallel pathways but with the
in vitro and in vivo of GBM cell lines (Michelakis et al., 2010). A phase cost of increased toxicity. Nevertheless, no combination therapy has
II study evaluates DCA in patients with GBM and results are pending shown enhanced clinical benefit over single agents (Wilson et al., 2014)
(NCT00540176). RTA 744 is a close chemical analogue of doxorubicin Tyrosine kinase associated receptors serve in the transmission of sig-
but is more effective in crossing the blood brain barrier and is able to nals from extracellular ligands to cell nucleus and share some common
achieve high concentrations in the CNS tumor tissue in animal models. pathway mechanisms and intracellular signaling. They are composed of
A phase I study of RTA 744 in patients with recurrent high-grade glio- a ligand-binding extracellular domain, a lipophilic transmembrane do-
mas has been completed but results have not been published yet main and an intracellular domain containing a catalytic site. In the ab-
(NCT00526812). sence of their associated ligand, tyrosine kinase receptors remain
unphosphorylated, monomeric and inactive. After the appropriate li-
3.2. Resistance to chemotherapy and radiotherapy gand binds its associated receptor, the receptors undergo dimerization
and then auto-phosphorylation in which the kinase domain of one re-
Many GBMs have intrinsic or acquired resistance to chemotherapy. ceptor phosphorylates one or more intracellular tyrosine residues on
MGMT gene promoter methylation and thus silencing and downregula- the second receptor. Those phosphotyrosine residues in turn, become
tion of the gene is an important mechanism of resistance to temozolo- binding sites and recruit adaptor proteins and activate downstream ef-
mide treatment in GBM. MGMT is a repair gene that removes alkyl fect on molecules which initiate specific signaling cascades with end re-
groups from the O6 position of guanine and in this way it alleviates sult the regulation of gene transcription. Adaptor proteins include
the effects of alkylating drugs (temozolomide or the nitrosureas) growth factor receptor-bound-2/son-of-sevenless (Grb2)/(SOS) effect
(Esteller et al., 2000; Hegi et al., 2005). Methylation of MGMT is found on serine/threonine kinases such as RAS, phosphatidyloinositide-3-ki-
in 30–60% of GBM cases and is associated with a favorable outcome if nase (PI3K) and phospholipase C (PLC). Negative kinase regulators in-
treated with alkylating agents (Hegi et al., 2005). A common mecha- clude PTEN in the PI3K pathway. Important intracellular mediators in
nism that confers this resistance is mediated by MGMT. If the promoter oncogenic biochemical pathways include RAF-MEK-ERK, AKT and
of this gene is methylated, glioma cells are more sensitive to temozolo- mTOR.
mide. One way of surpassing this resistance is with dose-dense (days 1– The alterations in gene synthesis result in cell growth, proliferation,
21 every 28 days) or dose-intense regimens (Clarke et al., 2009), or by migration, angiogenesis and apoptosis. Within tumor cells, dysregula-
direct inhibition of MGMT via O6-benzylguanine in combination with tion can occur with various mechanisms mainly with overexpression
TMZ (Quinn et al., 2009). A second mechanism that gives resistant prop- or mutations of receptors and intracellular domains, activation of
erties to these tumors is that of poly(ADP-ribose)polymerase-1 (PARP- biomolecules, inactivating mutations of their negative regulators and
1). PARP-1 is a critical base-excision-repair gene that when disrupted lead to constitutive activation of signaling pathways and uncontrolled
results in persistence of potentially lethal N7- and N3-purine lesions cellular survival, proliferation and invasion (Arora & Scholar, 2005).
contributing to TMZ cytotoxicity especially when O6-methylguanine ad- C-Kit is another receptor tyrosine kinase that acts similarly to EGFR
ducts are repaired or tolerated. PARP-1 small molecule inhibitors such and PDGFR and is found to be overexpressed in GBM. Binding of ligands
as BSI-201 (iniparib) and ABT-888 are promising either alone or in com- and specifically the stem cell factor to c-Kit, results in receptor dimeriza-
bination with TMZ (Wen & Kesari, 2008; Zhang et al., 2012). In preclin- tion and activation of several transduction pathways including Akt, Src
ical trials, inhibition of another target important to base-excision-repair, family kinases, PI3K, Ras/MAPK and PLC-γ. (Arora & Scholar, 2005).
abasic (AP) endonocluease-1 (APE-1) shows potentiation of TMZ activ- Angiogenic pathway inhibition: GBM is a highly vascularized malig-
ity (Zhang et al., 2012). nancy with marked angiogenesis and high expression of VEGF which is
There is evidence that glioma stem cells contribute to GBM also needed for stem cells to create an optimal functional environment.
chemoresistance. Stem cells from highly chemotherapy resistant grade Antiangiogenic agents, a group of continuously evolving targeted thera-
IV gliomas show expression of multidrug resistance protein-1 (MRP1) peutics, have demonstrated promising results in many trials (Onishi
transporters and grade II gliomas show expression of P-glycoprotein et al., 2013). Older agents like thalidomide and newer anti-VEGF or
68 C. Alifieris, D.T. Trafalis / Pharmacology & Therapeutics 152 (2015) 63–82

anti-VEGFR therapies, as well as anti-HGF and anti-avβ5 integrin mole- (Majumdar et al., 2009). PDGF-C has been shown to induce the
cules are included. VEGF is a member of the platelet-derived growth fac- maturation of blood vessels in GBM experimental models and
tor family. There are numerous ligands for VEGF receptors (VEGFRs) attenuate the response to anti-VEGF treatment (di Tomaso et al.,
including VEGF-A to –E and placenta growth factor. VEGFRs are 2009). Drugs that block the PDGFR include tyrosine kinase inhibi-
expressed on vascular endothelial cells (Ferrara, 2005). Ligand binding tors and MAbs.
to VEGFR-1 (Fms-like tyrosine kinase 1 or Flt-1) induces endothelial Hepatocyte growth factor (HGF) or scatter factor activates the EGF,
cell migration, whereas binding to VEGFR-2 (kinase insert domain- VEGF and HGF pathways. When it binds to the c-MET receptor, it
containing receptor or KDR) stimulates endothelial cell proliferation, activates intracellular signaling cascades similar to those triggered by
increases vascular permeability (this probably accounts for the the EGF and PDGF receptors such as MAPK, PI3K-Akt, V-Src and STAT.
vasogenic edema of GBM seen on MRIs), increases the expression of C-Met signaling has been associated with gliomagenesis, increased cell
proangiogenic factors (matrix metalloproteinase-1, urokinase-type growth, scattering and motility, invasion, resistance to apoptosis and
plasminogen activator and it’s receptor, plasminogen activator angiogenesis (Sierra & Tsao, 2011). It has also been postulated that
inhibitor-1) and has antiapoptotic effects. Binding to VEGFR-3 (Flt-4) c-MET activation may bypass EGFR tyrosine kinase inhibition in sensitive
leads to lymphangiogenesis. Several factors may upregulate VEGF ex- cells by either autophosphorylation of c-MET or transphosphorylation of
pression including hypoxia (the most important factor in tumor cells), other ErbB receptors (‘cross talk’) thus acting as a mechanism of resis-
HGF, PDGF, FGF, EGF, TNF, TGF-β, IL-1, c-KIT, the PI3K-Akt and Ras- tance to anti-EGFR treatment (Sierra & Tsao, 2011). HGF autocrine
MAPK pathways, acidosis, wound healing, endometriosis and embryo- GBM may predict sensitivity of GBM cells to c-MET inhibitors and
genesis (Ferrara, 2005). Neoplasms such as GBM, require their own vas- serum HGF levels may be serve as biomarker for the presence of auto-
cular supply to maintain survival beyond the size of 2 mm. The early crine tumors and responsiveness to c-Met inhibition therapy (Xie et al.,
small tumor extensions can receive the needed vascular supply from 2012). Also, targeting of the c-Met pathway is found to potentiate GBM
the glial vasculature by “co-opting” along the capillaries. To achieve rensposiveness to gamma irradiation (Lal et al., 2005). Preclinical evi-
this, angiogenesis must take place (Jain et al., 2007). However, blocking dence have demonstrated that PTEN loss amplifies c-Met induced GBM
the VEGF pathway provides means of blocking the new vessels with malignancy and it is suggested that trials of the combination of anti-
high permeability (“leaky vessels”) and thus decreasing hypoxia in the HGF blocking plus PTEN restoration or mTOR inhibition should be car-
tumor and consequently they disrupt a vital mechanism of survival in ried out (Li et al., 2009). Elevated serum HGF may be associated with
glioma stem cells, they decrease neoangiogenesis, hypoxia, vascular poor response and a shorter PFS in patients with GBM undergoing
permeability (and vasogenic edema) and increase chemotherapy deliv- first-line RT (Liang et al., 2014).
ery and radiosensitivity (Bao et al., 2006; Jain et al., 2007). Inhibitors of Most important intacellular effector molecules in the pathogenesis
VEGF signaling pathway include small molecule tyrosine inhibitors, of malignant gliomas are biochemically active serine/threonine kinases.
MAbs and VEGF binding agents. These include PI3K, RAS, PLC etc. and either activation of them or inac-
EGFR is the most frequent amplified gene seen in malignant gliomas. tivation of their negative regulators lead to excessive phosphorylation
The EGFR is one of four tyrosine kinase receptors within the ErbB recep- in the intracellular signal pathways and thus contribute to malignancy.
tor family, that is ErbB1 (EGFR/HER1), ErbB2 (HER2/neu), ErbB3 (HER3) RAS-RAF-MEK-MAPK pathway: The Ras gene superfamily encodes
and ErbB4 (HER4). Activation of these receptors result in turn in activa- small GTP (guanine triphosphate)-binding proteins (GBPs) which are
tion of multiple downstream signaling pathways as fore-mentioned, in- found in the inner cell membrane and regulate cellular functions, in-
cluding the Ras/Raf/MEK/ERK1/2-mitogen-activated protein kinase cluding cytoskeletal organization, protein trafficking, cellular differenti-
(MAPK) pathway, the phosphatidylinositol 3’-kinase(PI3K)/Akt/mTOR ation and proliferation as well as secretion of angiogenetic molecules.
pathway and the signal transducer and activator of transcription Ras mutations are not common in malignant gliomas but they rather
pathways (STAT 3 and 5). Ligands for these receptors include EGF and show increased Ras activity due to amplification or mutation of upstream
transforming growth factor α (TGF-α). These pathways affect cell growth factor receptors (Knobbe & Reifenberger, 2004). When Ras is
proliferation, migration, differentiation, inhibition of apoptosis and triggered by either EGFR or PDGFR activation or through independent
upregulation of VEGF expression and thus increased angiogenesis pathway alterations, the mitogenic signal through the mitogen activated
(Mendelsohn & Baselga, 2006). Activation of EGFR can be due to: protein kinase (MAPK also known as ERK) and phosphatidylinositol 3-
a) increased expression or activity of EGFR ligands or cofactors, kinase (PI3K) pathways are initiated. However, translocation of Ras to
b) EGFR gene amplification or increase in translation (50–60% of GBM the cell membrane for activation requires a post-translational alteration
cases), and c) activating mutations of EGFR, like the EGFRvIII mutant which is catalyzed by farnesyltransferase. Farnesylation is the rate-
(which mediates radioresistance and confers a more malignant limiting step in RAS maturation although not unique for Ras (Sebti &
potential than the wild type receptor, found in 50–60% of GBMs) Adjei, 2004). Downstream of Ras, is the pathway Raf-Mek-MAPK path-
(Ye et al., 2012). EGFR signaling alterations are associated with worse way. Activation of MAPK is associated with poor outcome in GBM
clinical prognosis, decreased OS and gliomagenesis (Shinojima et al., (Pelloski et al., 2006).
2003). Currently, EGFR inhibition has been tried by using small Phosphotidylinositol 3-kinase (PI3K) /AKT/mTOR (mammalian tar-
molecule tyrosine kinase inhibitors, monoclonal antibodies, peptide get of rapamycin) pathway is a group of serine/threonine kinases.
vaccines, antisense oligonucleotides and immunotoxin conjugates. PI3K can be activated by other tyrosine kinase receptors (EGFR,
Platelet-derived growth factor (PDGF) and its receptor (PDGFR) are PDGFR, c-Met), active Ras or integrins. It regulates cellular mechanisms
also commonly overactive in about 30% of GBM cases. The PDGF family and when altered can lead to decrease of apoptosis, increase in cell
consists of four members, PDGF-A through –D which signal through the growth and cell proliferation. It is a poor prognostic factor in patients
alpha and beta PDGFR (PDGFR-α; PDGFR-β). Activation of PDGFR trig- with malignant gliomas (Chakravanti et al., 2004). AKT (also known
gers multiple intracellular signaling pathways, much like the EGFR. as protein kinase B), is a downstream effector of PI3K -activated when
They include PI3K, MAPK, Jak family kinase, Src family kinase and phosphorylated- and leads to the same cellular results. mTOR is
phospholipace C-gamma (PLC-γ) (Joensuu et al., 2005). Thus, it pro- downstreamed from AKT and can be activated by RAS pathway also.
motes tumor growth through autocrine stimulation, angiogenesis At least 2 separate mTOR complexes exist. Each complex is composed
through paracrine effects on reciprocal endothelial cells and affects of a common regulatory subunit, the target of rapamycin (TOR) and a
tumor fibroblast regulation (which in turn leads to transvascular trans- downstream-substrate-defining subunit. The latter includes raptor sub-
port of drugs through alterations in intratumoral pressures) (Ostman, unit (in mTOR-complex-1, mTORC1) and rictor subunit (in mTOR-
2004). PDGF expression (but not PDGFR-α) is associated with the complex-2, mTORC2). Rapamycin only inhibits mTORC1. The raptor
tumor grade and proliferative activity of oligodendrogliomas subunit of mTORC1 regulates protein transcription which involves 4E-
C. Alifieris, D.T. Trafalis / Pharmacology & Therapeutics 152 (2015) 63–82 69

BP1 (initiation factor 43 binding protein-1) and p70s6K (ribosomal S6 (Rich et al., 2005). On the other hand, co-expression of EGFRvIII and
kinase). The most important negative regulator of the PI3K/Akt/mTOR PTEN in GBM seems to predict a better response to EGFR inhibition
pathway seems to be PTEN (phosphate and tensin homologue) (fifty times more likely) (Mellinhoff et al., 2005). In a phase II trial
(Sheppard et al., 2012). In GBM the PI3K/AKT/mTOR pathway is active with patients with recurrent GBM given afatinib alone, showed PFS-6
in about 70% of cases and it seems to regulate the proliferation and in 3% of patients but possibly was more effective in patients with
tumorigenic potential in GBM cancer stem cells. Rapamycin seems to EGFRvIII mutations (Eisenstat et al., 2011). A study of afatinib and radio-
inhibit this potential in vitro but not in vivo, hence it may be of limited therapy with or without temozolomide is under way (NCT00977431).
value as monotherapy in GBM (Mendiburu-Eliçabe et al., 2014). Dacomitinib (PF-299804) is a new pan-HER irreversible inhibitor that
Protein kinase C (PKC) is another serine/threonine kinase that regu- shows enhanced action against glioma-specific domain IV EGFR cyste-
lates glioma cell proliferation, angiogenesis and invasion. It is activated ine mutations (Greenall et al., 2014) and is currently been evaluated
by other tyrosine kinase receptors and G-protein coupled receptors. in two phase II studies, one with GBM patients with EGFR amplification
PKC-β mediates VEGFR2 signalling through its associated MEK/MAPK or EGFRvIII expression (NCT01520870) and another in seeking of effica-
pathway (Kreisl et al., 2010). cy in recurrent/relapsed GBM (NCT01112527). As a conclusion, erloti-
Apart from conferring resistance to chemo- and radio-therapy and nib appears to be more effective against malignant gliomas than
increase in survival of stem cells the Sonic hedgehog/Gli1 (SHH/Gli1) gefitinib. Higher levels of EGFR expression can be associated with better
pathway seems to also act in migration and glioma cell invasion response to erlotinib (Haas-Kogan et al., 2005a). Although, EGFR has
(Wang et al., 2010). clearly a great and important role in the pathogenesis of high grade
GBM is characterized by the presence of increased growth factor gliomas, trials up to date with EGFR inhibitors failed to show promising
signaling, hypoxia, angiogenesis and the presence of brain glioma results. Explanations may be the acquisition of new resistance muta-
cancer-stem-cells, conditions in which Notch signaling could prove to tions and/or inability to cross effectively the blood brain barrier (BBB).
be a key. Specifically, regarding stem-cells there is evidence that by Also, proximal block of the EGFR pathway, may be overridden by over-
blocking this pathway in combination with other targeted treatments, active distal effectors (high pAKT) (Haas-Kogan et al., 2005b).
a better control of these neoplasms could be achieved (Stockhausen Tyrosine kinase inhibitors have been developed to block the VEGF
et al., 2012). Gamma secretases, mostly known for their role in the pathway. However, although they are many and primarily antagonize
final step of processing amyloid precursor protein in Alzheimer’s angiogenesis they also have mixed tyrosine kinase blocking effects. In-
disease through the generation of amyloid-beta peptides, also play a hibitors that have primarily VEGFR blocking effects include vatalanib,
key role in Notch signaling pathway(Lin et al., 2010; Stockhausen vandetanib, sunitinib, pazopanib, cediranib and cabozantinib. They
et al., 2012). will be discussed below.
The integrins are heterodimeric transmembrane cell adhesion mole- As in VEGFR pathway, small molecule inhibitors have been devel-
cules and receptors that bind multiple extracellular ligands (e.g. matrix oped against PDGFR. However, they have mixed effects and are not
proteins like collagen, laminins, vitronectins and fibronectins). This selective for PDGFR. Examples of inhibitors with primarily actions
binding activates integrins and thus regulates tumor cell invasion and against PDGFR include imatinib, dasatinib, nilotinib and tandutinib.
metastasis, migration, proliferation, angiogenesis and survival. The They will be discussed below. Crenolanib (CP-868,596), a novel oral
regulation of these effects is mediated via downstream signaling path- selective and potent PDGFR inhibitor, is currently being evaluated in a
ways that signal in parallel to other pathways like EGFR, VEGF and phase II study of adult malignant gliomas (NCT01229644) and in a
PDGFR. There are at least 24 integrin types and are widely expressed phase I study of pediatric/young adult recurrent high grade gliomas
throughout the vasculature. The integrins alphaVbeta3 (αVβ3) and (NCT01393912).
alphaVbeta5 (αVβ5) are highly expressed in malignant glioma tumor Small molecule inhibitors targeting c-MET pathway have been
cells and on endothelial cells at tumor peripheral vessels. Peptide- developed but not all of them show selectivity for the c-MET receptor.
based integrin inhibitors as well as monoclonal antibodies are being in- SGX523 is a highly selective oral c-MET inhibitor that has been shown
vestigated as means of inhibiting this cellular mechanism. Integrin in- to be effective in vitro and in vivo against GBM cell lines. Although
hibitors decrease tumor hypoxia (Tabatabai et al., 2010). Integrin two separate phase I studies with SGX523 were prematurely terminat-
overexpression has been found in glioma stem cell microenvironment ed (NCT00606879, NCT00607399), it may still serve as a guide of inves-
(Lathia et al., 2010). tigating the role of MET in cancer until further clinical evaluation
(Buchanan et al., 2009; Guessous et al., 2010).
3.4. Molecular pathway and kinase inhibitors Farnesyltransferase inhibitors (FTIs) have been tried in the treat-
ment of malignant gliomas. Tipifarnib (R115777), in a phase II trial ex-
EGFR tyrosine kinase inhibitors (TKIs) are small molecules that bind hibited modest efficacy as monotherapy in patients with recurrent
to the ATP-binding site on the tyrosine kinase domain of the receptor malignant gliomas. PFS-6 was 11.9% for GBM patients. Toxicity was tol-
and inhibit the catalytic activity of the kinase. They can also inhibit erated well except the subset of patients that received enzyme inducing
fusion tyrosine kinases by blocking their dimerization. Furthermore, antiepileptic drugs which presented an increased hematological toxicity
they interact with VEGFR and other human epidermal receptors (Cloughesy et al., 2006). R115777 administered before RT in patients
(Yarden & Sliwkowski, 2001). Examples include the competitive antag- with newly diagnosed GBM and residual enhancing disease did not
onists, erlotinib and gefitinib (administered orally), as well as the newer lead to any measurable benefit (Lustig et al., 2008; Nghiempu et al.,
irreversible antagonists afatinib and dacomitinib. 2011). The R115777 plus RT combination with or without TMZ, in
Erlotinib in two separate phase II trials with patients with recurrent newly diagnosed GBM seems to be well tolerated (Nghiempu et al.,
GBM, showed that it is ineffective as monotherapy showing progression 2011). Another phase I/II trial of R115777 with RT in patients with
free survival at 6 months (PFS-6) in 3% of patients and a median PFS of newly diagnosed GBM produced a median OSof 60.4 weeks but time
2 months (Raizer et al., 2009) and PFS-6 in 11.4% vs 24% of control to progression (TTP) was only 18.1 weeks where both FGFR1 and
(treated with TMZ or carmustine) (van den Bent et al., 2009). Gefitinib, αvβ3 integrin expressions being independent bad prognostic factors
in a phase II study with recurrent GBM patients, demonstrated event- (Ducassou et al., 2013). A phase I/II trial of the combination of sorafenib
free survival (EFS) of 13.2% at 6 months (Rich et al., 2004). Systemic combined with erlotinib, tipifanib or temsirolimus in recurrent GBM has
toxicity has been used as a way to predict the clinical efficacy of treat- been completed and results are pending (NCT00335764). Lonafarnib
ment (e.g, rash, diarrhea). For example in a study, diarrhea from gefitin- (SCH66336) against U87 GBM cell lines improved activity of TMZ/RT
ib treatment was associated with better OS whereas EGFR expression (Chaponis et al., 2011). In a phase I study on Ionafarnib with TMZ in
(either EGFRvIII or wild-type) did not correlate with prediction of OS malignant glioma patients demonstrated dose limiting hepatic,
70 C. Alifieris, D.T. Trafalis / Pharmacology & Therapeutics 152 (2015) 63–82

gastrointestinal (vomiting, diarrhea, anorexia), renal, thrombotic and FKBP12. In a phase I trial of everolimus and temozolomide in combina-
constitutional toxicities (headaches) (Desjardins et al., 2011). In tion with radiotherapy was found to be well tolerated in newly
progressing GBM, intensifying TMZ treatment and adding lonafarnib a diagnosed GBM patients (Sarkaria et al., 2011). A phase II study of con-
PFS-6 of 33% and partial response of 27% in previously non-responding current TMZ/RT/bevacizumab followed by bevacizumab/everolimus
patients was achieved (Gilbert et al., 2006). In another phase I/Ib following surgical resection in newly diagnosed GBM patients, demon-
study of lonafarnib and dose-dense schedule of TMZ in recurrent/TMZ strated PFS of 11.3 months and median OS of 13.9 months. This efficient
refractory GBM, toxicity was acceptable and PFS-6 was 38% and median regimen was associated with everolimus related dose limiting toxicities
survival 13.7 months (Yust-Katz et al., 2013). Manumycin, a FTI, has like fatigue, pneumonitis and stomatitis. Bevacizumab related side-
been found to induce reactive osxygen species (ROS) and glioma cell ap- effects were also observed. However, the PFS is favorable to standard
optosis, inhibits STAT3 and telomerase activity (Dixit et al., 2009) as wall TMZ/RT historical reports (Hainsworth et al., 2012). Ridaforolimus
as targets the IL1β-Ras-HIF-1α axis (Sharma et al., 2012). (Deferolimus, AP23573, MK-8669) is a selective mTOR inhibitor. In
TLN-4601, a novel Ras-MAPK inhibitor, has been studied on first pro- a phase II study on progressive or recurrent malignant gliomas,
gression status of GBM in a phase II study. Although it was well tolerated ridaforolimus given perisurgically was found to be able to cross the
(except two patients who discontinued the medication due to toxic- blood-brain barrier and with histological evidence it achieved mTOR in-
ities) at the proposed dosage, it was largely ineffective as monotherapy hibition with an expected, well tolerated and no dose-limiting toxicity
(Mason et al., 2012). Sorafenib, is an oral MTKI that also targets B-Raf profile (Reardon et al., 2012b). A phase I safety study of ridaforolimus
and C-Raf (Guan & He, 2011). It is discussed under MTKI section. is completed in progressive/recurrent gliomas (NCT00087451). XL765
Selective PI3K inhibitors include XL-147, buparlisib and BKM120. is a dual inhibitor of PI3K/mTOR that has been studied in vivo and
XL147 (SAR245408) is a selective and reversible class I PI3K inhibitor in vitro against GBM cells and demonstrated efficacy both alone and in
that has been used in an exploratory trial in patients with recurrent combination with TMZ (Prasad et al., 2011). Two ongoing phase I
GBM who were surgical candidates (NCT01240460). Buparlisib studies use XL765, (a) in combination with TMZ with or without RT in
(NVP-BKM120) is a selective PI3K inhibitor demonstrated antitumor malignant gliomas (NCT00704080) and (b) along with XL147 in
activity in U87 glioma cells (Koul et al., 2012). BKM120 is currently surgical candidates with recurrent GBM (NCT01240460). PI-103 is a
under investigation in a phase II study with recurrent GBM patients potent, cell permeable, ATP-competitive class I PI3K inhibitor and
(NCT01339052), a phase I dose escalation trial along with TMZ/RT in mTORC1/2 inhibitor. It was found to be efficient against malignant glio-
newly diagnosed GBM (NCT01473901) and a phase I/II study in combi- ma xenografts (Fan et al., 2006). PKI-587 (PF-05212384) is a dual PI3K/
nation with bevacizumab in patients with relapsed/recurrent GBM mTOR inhibitor that was found to be effective against glioma U87 cell
(NCT01349660). lines (Mallon et al., 2011) and is now being investigated in a safety
GBM patients who carry mutant PIK3R1 alleles may benefit from phase I study. Two novel ATP-competitive mTOR inhibitors CC214-1
targeted inhibition of AKT since signaling through PI3K promotes tu- and CC214-2 suppress rapamycin resistant mTORC1 signalling, block
morigenesis (Quayle et al., 2012). MK-2206, an allosteric AKT inhibitor, mTORC2 especially in EGFRvIII activated GBMs. It also seems that autoph-
is currently undergoing phase I/II trials (Hirai et al., 2010). Perifosine agy helps cancer cells to overcome mTOR inhibition (Gini et al., 2013).
(KRX-0401) is a novel Akt inhibitor. It has been evidenced to inhibit Enzastaurin (LY317615), a potent PKC inhibitor has been trialled
multiple intracellular signaling pathways in glial progenitors and coop- against malignant gliomas. A phase I/II study of enzastaurin monother-
erates with TMZ to arrest glioma cell proliferation in vivo (Momota apy in malignant gliomas demonstrated 25% radiographic response
et al., 2005). It does not seem to be able to enhance the radiosensitivity with PFS-6 of 7% (in GBM) and of 6% (in anaplastic glioma). Glycogen
of human glioma cells (De la Pena et al., 2006). Also, perifosine in synthase kinase-3 in peripheral blood mononuclear cells was identified
combination with temsirolimus shows synergism in inducing apo- as a potential marker for drug activity (Kreisl et al., 2010). However, a
ptosis and decrease proliferation in PTEN-intact and PTEN-deficient later phase III study comparing enzastaurin with lomustine in recurrent
PDGF-driven murine GBM cells (Pitter et al., 2011). Currently an intracranial GBM did not show superior efficacy although it was less
ongoing phase I/II study seeks to investigate the combination of hematologically toxic (Wick et al., 2010). A phase II study of TMZ plus
temsirolimus plus perifosine in recurrent or progressive malignant glio- enzastaurin during and after RT in newly diagnosed GBM was well tol-
mas (NCT01051557) and a phase II study investigates perifosine mono- erated and exhibited a median OS of 74 weeks which was positively cor-
therapy in recurrent/progressive malignant gliomas (NCT00590954). related with MGMT methylation status and a median PFS of 36 weeks
mTOR inhibitors used in malignant gliomas are either selective (e.g. (Butowski et al., 2011). Enzastaurin before and concurrently with RT,
temsirolimus, sirolimus, everolimus, ridaforolimus) or dual PI3k/mTOR followed by enzastaurin mainentance therapy in the newly diagnosed
inhibitors (XL-765, PI-103, PKI-587). Temsirolimus (CCI-779) is a spe- MGMT unmethylated GBM, demonstrated median OS of 15 months
cific mTOR inhibitor. In a phase II trial of temsirolimus against recurrent and PFS-6 of 53.6% (less than primary planned outcome) while it was
GBM, radiographic response in 36% of patients, PFS-6 7.8%, median well tolerated (Wick et al., 2013). A phase II trial of enzastaurin and
OS4.4 months, median TTP 2.3 months (responders: 5.4 months vs bevacizumab in recurrent GBM is underway (NCT00586508). Tamoxi-
non-responders: 1.9 months) were observed. It was generally well tol- fen, an antiestrogen, has been long described to act also as a PKC inhib-
erated (Galanis et al., 2005). In a second phase II study of temsirolimus itor (Baltuch et al., 1993). Generally, tamoxifen has not been proved to
on recurrent GBM no efficacy was demonstrated, though 50% of patients be effective in the treatment of malignant gliomas (Spence et al.,
were initially stable. Once again the toxicity profile was well tolerated 2004). In a phase II trial of RT plus high dose tamoxifen in GBM the me-
(Chang et al., 2005). However, temsirolimus in combination with dian survival time was 9.7 months with a lower than expected throm-
TMZ/RT showed increased risk of infectious complications (Sarkaria boembolic incidence, a fact attributed to the PKC inhibitory effects of
et al., 2010). Various combinations have been either in ongoing trials tamoxifen (Robins et al., 2006).
or in completed ones. Sirolimus (rapamycin, AY-22989, WY-090217), Cyclopamine, is a specific SHH signaling pathway antagonist of
is a selective mTOR inhibitor. It was found to achieve a partial radio- Smoothened (SMO). In vitro studies show that it leads to depletion of
graphic response in 6% and stable disease in 38% of patients when com- stem-like cells in GBM (Bar et al., 2007). Also, it produces a significant
bined with gefitinib, in a phase I study. Dose limiting toxicity included but incomplete GBM xenograft cell regression (Sarangi et al., 2009).
mucositis, diarrhea, thrombocytopenia and hypertriglyceridemia SEN450, a novel Smoothened receptor antagonist, seems to be cytostat-
(Reardon et al., 2006). In a phase II trial, combination of elrotinib and ic on its own and it further reduces tumor volume after TMZ pretreat-
sirolimus in patients with recurrent GBM was ineffective and demon- ment in vivo and in vitro (Ferruzzi et al., 2012). Vismodegib (GDC-
strated a dismal PFS-6 of 3.1% but was generally well tolerated 0449, HhAntag691), is a potent novel and specific SHH pathway inhibi-
(Reardon et al., 2010). Everolimus (RAD001) is an mTOR inhibitor of tor. It has been granted FDA approval for basal cell carcinoma (Meiss &
C. Alifieris, D.T. Trafalis / Pharmacology & Therapeutics 152 (2015) 63–82 71

Zeiser, 2014). Vismodegib is being investigated in a phase II study of Increased thromboembolic risk was also found in a phase I study of
patients with recurrent GBM that are operatable (NCT00980343). lenalidomide plus RT in newly-diagnosed GBM (Drappatz et al., 2009).
Erismodegib (LDE225), a novel oral SHH pathway inhibitor and a selec-
tive Smoothened antagonist, suppresses epithelial-mesenchymal tran- 3.5. Multitargeted tyrosine kinase inhibitors (MTKIs)
sition and self-renewal of GBM cells through modulation of miRNA-
200, -128 and -21 (Fu et al., 2013). Many tyrosine kinase inhibitors can antagonize more than one type
Gamma secretase inhibitors (GSIs) are investigated in malignant of kinase, thus it is somewhat arbitrary to classify them. However, most
gliomas to block the Notch mathway (Lin et al., 2010; Stockhausen MTKIs do tend to have a more predominant inhibitory action against a
et al., 2012). RO4929097 is an oral GSI that sensitizes GBM cells to specific type of kinase.
TMZ in vivo and in vitro (Hiddingh et al., 2014). It is under evaluation MTKIs with effect on EGFR TKI include two reversible inhibitors
in recurrent/progressive GBM (phase II, NCT01122901). MK-0752 is a (lapanitib and vandetanib) as well as an irreversible inhibitor
moderately potent gamma secretase inhibitor. It has been used in a (Bay846). Lapatinib (GW572016) is a reversible MTKI with dual action
phase I study of children wth refractory CNS malignancies and was on EGFR/ERbB-1 and Her-2/ERbB2 family of receptors. In a phase I/II
well tolerated (Fouladi et al., 2011). DAPT (GSI-IX) is a novel GSI that trial in recurrent GBM, did not show any significant efficacy and when
suppresses GBM growth in xenografts via uncoupling of tumor vessel given with anti-epileptic medication, its clearance increased about
density from vessel function as evidenced by poor perfusion and aggra- 1,000%. Neither EGFRvIII excpression, nor PTEN loss could predict a
vated hypoxia (Zou et al., 2013). Recently, a study showed that alpha favorable subtype (Thiesen et al., 2010). It has been used along with
secretase inhibitors (ASIs) may be an alternative to GSIs in GBM since temozolomide in a phase I trial without any considerable efficacy
they decrease Notch activity and tumor size and increase survival time (Karavasilis et al., 2013). Lapatinib has also been investigated in combi-
(compared to GSIs) in GBM cells (Floyd et al., 2012). nation with pazopanib and is being investigated along with TMZ and ra-
Cilengitide (EMD121974) is a synthetic Arg-Gly-Asp pentapeptide diation therapy in patients with newly diagnosed GBM (NCT01591577).
(RGD) that recognizes the RGD ligand binding site on the integrin recep- Vandetanib (ZD6474), is a reversible MTKI with action against EGFR,
tors αVβ3 and αVβ5 and competitively blocks integrin ligand binding VEGFR-2, -3 and RET. In tumor models, it achieved higher reduction of
(A.A. Reardon et al., 2008a). A phase II study of cilengitide in patients GBM tumor volume when combined with TMZ than as monotherapy
with recurrent GBM, demonstrated a PFS-6 of 15% and median OSof (Jo et al., 2012). It has already been through phase I and II trials in pa-
9.9 months. It was generally well tolerated (D.A. Reardon et al., tients with recurrent GBM that received fractionated radiosurgery
2008b). A phase I/IIa study on cilengitide and TMZ/RT followed by the (Fields et al., 2012; Kreisl et al., 2012) and it seems to be effective against
administration of cilengitide and TMZ as maintenance was conducted gliomas that express the EGFRvIII mutant, too (Yiin et al., 2010).
in newly diagnosed GBM, according to MGMT status of the patients. Re- Autophagy protects glioblastoma cells from the proapoptotic effects of
sults showed PFS-6 in 69% and PFS at 12 months in 33% of patients, me- vandetanib and thus it might contribute significantly to the resistance
dian PFS at 8 months, OS of 68% (at one year) and 35% (at 2 years) and against it (Shen et al., 2013). Bay846 is a novel irreversible MTKI against
median OS of 16.1 months. PFS and OS were longer in patients with EGFR and Her2 and shows efficacy in malignant glioma brain tumor
MGMT promoter methylation. The combination was well tolerated cells. Bay846 was more efficient compared to lapanitib and its sensitiv-
(Stupp et al., 2010a). A similar phase II study to verify the safety of ity was associated with wild type PTEN in conjuction with the expres-
cilengitide plus standard chemoradiation (TMZ/RT) also in newly diag- sion of EGFRvIII (Longo et al., 2012).
nosed GBM was performed. The suggested dose was 2000 mg and in MTKIs with effevt on VEGFR include cabozanitinib, cediranib, E7080,
this group, the median survival was 20.8 months, and regarding pazopanib, sorafenib, sunitinib, vandetanib and vatalanib. Cabozantinib
MGMT methylation status median survival was 30 months in patients (XL184), is a MTKI that inhibits VEGF2, c-MET, AXL, TIE2, Flt-3 and RET.
with methylated MGMT and 17.4 months in the unmethylated status In a phase II trial as monotherapy, in the cohort that received the 175 mg
subgroup. It was generally well tolerated (Nabors et al., 2012). In daily dose it showed a response rate of 21% and in the cohort with the
NABTC 03-02 phase II trial evidence of the efficacy of cilengitide was better tolerated 125 mg daily dose showed PFS-6 in 25% of patients.
found to be modest as monotherapy in recurrent GBM while it was XL184 also appeared to have modest activity in patients with prior
adequately delivered into the tumor (Gilbert et al., 2012). Completed antiangiogenic treatment. It was noted that only a small number of
trials with cilengitide include, (a) cilengitide, TMZ and RT in newly diag- patients developed distant or diffuse disease when the neoplasm
nosed GBM and MGMT promoter methylation (phase III – CENTRIC progressed, indicating antiangiogenic plus anti-invasive effects of
trial- NCT00689221), (b) cilengitide, TMZ, RT in newly diagnosed XL184 (Wen, 2010). Cabozantinib, is probably a promising agent that
GBM and unmethylated MGMT promoter (phase II – CORE trial – suppresses angiogenesis, tumor growth and metastasis in neoplasms
NCT00813943), and (c) Cilengitide in patients undergoing surgery for with dysregulated VEGFR and MET signaling (Yakes et al., 2011). More-
recurrent or progressive GBM (NCT00112866). Results are pending. over, two phase II studies on cabozantinib in recurrent (NCT00704288)
Thalidomide has been considered as an older antiangiogenic agent and newly diagnosed (NCT00960492) GBM are under way. Cediranib,
(Wen & Kesari, 2008). It has been associated specifically with a decrease an oral MTKI, antagonizes all VEGFRs (especially VEGFR2), c-Kit and
in the expression of adhesion molecules and especially integrins and PDGFR. In a phase II study, used as monotherapy in patients with recur-
thus is postulated to inhibit cell migration with antiangiogenic and rent GBM, it exhibited encouraging results including PFS-6 in 26% of pa-
anti-inflammatory (decrease in leukocyte migration) activity tients, median OS of 7 months and median PFS of 3.5 months. More than
(McCarty, 1997). It is also widely known for its teratogenic potential. half of patients showed radiographic improvement. However, dose re-
Generally, thalidomide has shown to be ineffective in malignant glio- duction or drug interruption was required in over half of the patients
mas alone or in various combinations. In a phase II study of thalidomide due to grade ¾ toxicities such as hypertension (12.9%), diarrhea and fa-
plus irinotecan in newly diagnosed GBM showed limited efficacy com- tigue (Batchelor et al., 2010). Later, a randomized phase III study, com-
pared to the standard approach while venous thromboembolism was pared cediranib as monotherapy or in combination with lomustine to
common (Fadul et al., 2008). In a phase II study on thalidomide com- lomustine monotherapy (with cediranib-matched placebo, as control)
bined with RT in newly diagnosed GBM did not improve survival was performed. Cediranib was used at lower doses relative to the previ-
(Alexander et al., 2013). Studies with the more potent analogue of tha- ous phase II study (30 mg/day p.o. as monotherapy or 20 mg/day p.o.
lidomide, lenalidomide are under way including a combination of when combined with lomustine) and lomustine was administrated
lenalidomide along with RT in newly diagnosed GBM (NCT00165477). p.o. at 110 mg/m2. Both monotherapies and combinations were given
In a phase I trial, lenalidomide monotherapy did not seem to be efficient every 6 weeks. The PFS-6 in the ceriranib monotherapy arm was
with the added risk of thromboembolic disease (Fine et al., 2007). lower than the one observed in the previous phase II study (16%). PFS-
72 C. Alifieris, D.T. Trafalis / Pharmacology & Therapeutics 152 (2015) 63–82

6 was 34.5% in the combination arm and 24.5% in the control arm cell factor (SCF)/c-Kit, and PDGFR-α/β. Its toxicity profile includes
(Ahluwalia, 2011). The efficacy of cediranib in gliomas, seems to be myelosuppression, aplastic anemia, GI effects (nausea, vomiting, diar-
limited by effective efflux at the brain-blood barrier via P-glycoprotein rhea, GI hemorrhage), liver toxicity (elevated transaminases/ bilirubin),
(P-gp) and breast cancer resistance protein (Bcrp) (Wang et al., 2012). respiratory effects (cough, dyspnea, upper respiratory tract infections,
Lenvatinib (E7080), an MTKI active against VEGF, PDGF and FGF recep- chest pain, pulmonary fibrosis), skin manifestations (erythema
tors, suppresses tumor cell migration and invasion (Glen et al., 2011). It multiforme/ Stevens-Johnson syndrome, pruritus, angioedema), ner-
is currently being investigated in a phase I/II study of the combination of vous system effects (headache, migraines, central nervous system hem-
E7080 plus E7050 (a dual c-MET and VEGFR-2 TKI) in patients with re- orrhage, peripheral neuropathy, syncopy), fatigue, pyrexia, blurred
current GBM (NCT01433991). Pazopanib (GW786034) is an MTKI vision, conjunctivitis, ascites, fluid retention-edema, arthralgias/myal-
against VEGFR-1 through -3, PDGFRα/β and c-Kit. In a phase II study gias and electrolyte disturbances. It is metabolized through CYP 3A4
on patients with recurrent glioblastoma, it was given at 800 mg/daily and subjective to drug interactions when this enzyme is induced (Gan
in 4 week cycles. It was generally ineffective with a response rate of et al., 2009). In a 2005 phase II study, imanitib plus hydroxyurea in
just 6%, PFS-6 of 3%, median PSF of 3 months and median survival of adults with recurrent GBM showed 9% radiographic response, PFS-6 in
35 weeks. However, radiographic response was observed. It was well 27% of patients and median PFS 14.4 weeks, in a median follow up of
tolerated with a toxicity profile similar to that of other anti VEGFR 58 weeks. It was generally well tolerated and effective in some patients
agents (Iwamoto et al., 2010b). A phase II study on pazopanib in combi- (Reardon et al., 2005). However, in a later phase I/II study, imatinib
nation with lapatinib in adult patients with relapsed malignant glioma mesylate when used in recurrent malignant gliomas as monotherapy
has demonstrated limited therapeutic results (Reardon et al., 2013). Su- was largely ineffective with a PFS-6 of 3% for GBM patients. CYP3A4 in-
nitinib, another MTKI, is antagonizing VEGFR-1, -2, -3, c-Kit, PDGFRα/β, ducers substantially decreased its effectiveness (Wen et al., 2006b). In a
Flt-3 and colony stimulating factor 1 receptor 9 (CSF-1R). It is approved phase II study the efficacy of imatinib monotherapy in patients with re-
for gastrointestinal stromal tumors (GISTs) and advanced renal cancer current malignant gliomas was tested. Imatinib was generally well tol-
(Gan et al., 2009). In a phase I study on 25 patients with recurrent erated in the doses that it was administrated and demonstrated limited
high-grade gliomas treated with sunitinib plus irinotecan, limited antitumor effects with a PFS-6 of 16% in GBM (Raymond et al., 2008). In
results were demonstrated with radiographic response was noted in another phase II study imatinib was given as neoadjuvant imatinib be-
only in one patient and PFS -6 months was 25% (Reardon et al., fore either definitive surgery or re-biopsy in GBM cases. Median surviv-
2011b). In a phase II trial of patients with recurrent high grade glioma, al was found to be 6.2 months. Imatinib was found in the histological
although radiographic response was noted, clinical reponse was specimens but it was not associated with reduction in either prolifera-
achieved in none with OS of 3.8 months (Neyns et al., 2011). In a pro- tion or other survival mechanisms (according to biochemical evidence
spective phase II study, sunitinib monotherapy in patients with recurrent of Ki67, AKT, MAPK or p27 level changes) (Razis et al., 2009). During
malignant gliomas (GBM or anaplastic astrocytoma) was disappointing the same year, a multicentre phase II study of the combination of ima-
and ineffective since it showed zero response rates, PFS-6 months of tinib plus hydroxyurea in patients with progressive GBM was not
16.7% and median OS of 12.6 months in GBM (Pan et al., 2012). Vatalanib found to be associated with clinically important antitumor activity but
(PTK787/ZK222584)), is an oral VEGFR, PDGFR and c-Kit inhibitor. In a was well tolerated. PFS-6 was 10.6%, median OS was 26 weeks and ra-
phase I study on vatalanib plus imatinib plus hydroxyurea in recurrent diographic response was noted in 3.4% of cases. 7% of patients had either
malignant glioma patients, it was found to be well tolerated at grade 3 fatigue, neutropenia, thrombocytopenia (Reardon et al., 2009a).
1000 mg b.i.d. (Reardon et al., 2009b). In a 2010 phase I/II study on RT Since the combination of hydroxyurea and imatinib had shown mixed
plus concomitant/adjuvant TMZ plus vatalanib/ZK222584 in patients results, in 2010, a randomized multicenter open label phase III study
with newly diagnosed GBM, the combination was found to be well toler- was carried out comparing the combination of hydroxyurea plus ima-
ated and safe. However, the study was discontinued because the industry tinib versus hydroxyurea alone in patients with recurrent progressive
that was developing vatalanib, chose not to carry on and withdrew it GBM resistant to TMZ. No clinically important differences were found
(Brandes et al., 2010). in comparing the two arms thus implying the ineffectiveness of imatin-
MTKIs with effect on PDGFR include bosutinib, cediranib, dasatinib, ib in recurrent GBM. The median PFS was similar in both arms
imatinib, nilotinib, pazopanib, sorafenib, sunitinib, tandutinib. Bosutinib (6 weeks), and PFS-6 was 5% for the combination arm and 7% for the
(SKI-606), inhibits PDGFR, Src/Srk and BCR/ABL tyrosine kinase used in monotherapy arm. In the arm with imatinib, toxicities were well tolerat-
the treatment of chronic myeloid leukemia (Rassi & Khoury, 2013) and ed (Dresemann et al., 2010). Nilotinib, another PDGFR, BCR-ABL and c-
is currently being evaluated in a phase II study of patients with recur- Kit inhibitor, is used in chronic myeloid leukemia (Jabbour &
rent GBM (NCT01331291). Dasatinib, is an oral MTKI that affects BCR- Kantarjian, 2012) and is currently being investigated in a phase II study
ABL, the Src family, c-Kit, EPHA2 and PDGFR-β. It binds to both active of patients with recurrent malignant gliomas (NCT01140568).
and inactive ABL kinase domains. It induces apoptosis, autophagic cell Tandutinib (MLN518), a PDGFR, Flt-3 and c-Kit inhibitor (Lehky et al.,
death of GBM cells, inhibits their invasion, alter cell adhesion and 2011), has been evaluated in a phase II trial of the combination
motility-migration of GBM cells in vitro. In vitro findings are being tandutinib plus bevacizumab in recurrent malignant gliomas. The trial
followed by ongoing preclinical trials in GBM including, (a) dasatinib is completed and the results are pending (NCT00667394). A second
plus RT/TMZ in newly diagnosed GBM, and (b) dasatinib monotherapy study on tandutinib combination in recurrent or progressive GBM is
in recurrent GBM (Manmeet et al., 2010). Dasatinib plus lomustine in also completed (NCT00379080).
recurrent GBM was marked by significant hematological toxicity and MTKIs with action against HGF/SF include cabozantinib, crizotinib and
suboptimal efficacy (Franceschi et al., 2012). A phase I trial of dasatinib MP470. Crizotinib (PF02341066), a dual c-MET, ALK inhibitor, has been
plus erlotinib in patients with recurrent malignant gliomas, showed that tried in patients with both c-MET and ALK positive histology. Recently,
it is a safe combination on a daily continuous dosing schedule. However there was a report of rapid radiographic and clinical improvement after
no radiographic response was observed and the PFS-6 was 2%. No grade treatment with crizotinib in MET-amplified recurrent GBM (Chi et al.,
3–4 toxicities were observed and most commonly, dose limiting toxic- 2012). Amuvatinib (MP470), a novel oral potent c-Met, c-Kit and PDGFR in-
ities were diarrhea and fatigue (Reardon et al., 2012a). Dasatinib, does hibitor has been shown to radiosensitize GBM cells both in vivo and in vitro
not appear to have any effect in combination with bevacizumab after (Welsh et al., 2009). Three different safety phase I trials have already been
bevacizumab failure in recurrent heavily pretreated GBM (Lu- completed in healthy volunteers (Choy et al., 2012).
Emerson et al., 2011). P-gp and Bcrp seem to cause active efflux of Sorafenib is an oral MTKI with targets such as VEGFR-2, -3, Flt-3,
dasatinib from the brain hence decreasing efficacy (Agarwal et al., PDGFR-β, c-Kit and it also inhibits B-Raf and C-Raf. These inhibitory ef-
2012). Imatinib is an oral MTKI that is effective against BCR-ABL, stem fects decrease tumor cells and angiogenesis. This drug is commonly
C. Alifieris, D.T. Trafalis / Pharmacology & Therapeutics 152 (2015) 63–82 73

used in hepatocellular carcinoma (Guan & He, 2011). In vitro results of The 26S proteasome is a protein complex that degrades ubiquinated
sorafenib plus nutritional factors such as vitamin K1, has demonstrated proteins. The role of the ubiquitin-proteasome pathway is to regulate
enhancement of growth inhibition and apoptosis of malignant glioma intracellular concentrations of specific proteins and maintain cellular
cells by blocking the Raf/MEK/ERK pathway (Du et al., 2012). A phase homeostasis as it is related to regulation of the cell cycle and transcrip-
I trial on concurrent RT and TMZ followed by TMZ and sorafenib in tion, cell signaling and apoptosis by degrading regulatory proteins such
newly diagnosed GBM did not appear to improve efficacy compared as p53 and cyclin dependent kinases or CDKs (Mani & Gelmann, 2005).
to historical standard therapy results (Hainsworth et al., 2010). In a Bortezomib (PS-341) is a selective reversible inhibitor of the 26S protea-
Phase II study, patients with recurrent GBM were treated with sorafenib some. In a phase I of recurrent malignant glioma patients, some clinical
plus daily TMZ. However, PFS-6 was only 9.4% (only one patient efficacy was observed. The overall response rate was 3%, PFS-6 was 21%
achieved partial response) and this was attributed to drug interactions for patients without exposure to enzyme inducing antiepileptics-EIAED
with CYP3A inducing antiepileptic drugs that lowered sorafenib expo- and 9% for those exposed, and median PFS was 1.9 months (exposed)
sure in these patients (Reardon et al., 2011c). In a phase I/II trial with pa- versus 2.6 months (for non-exposed). Median OS was 5.8 months (in
tients with recurrent GBM, the combination of temsirolimus plus EIAED exposed patients) versus 6.1 months (in EIAED non-exposed pa-
sunitinib was discontinued due to zero PFS-6 (Wen et al., 2009). A tients) (Phurphanich et al., 2010). A phase II trial of the combination of
phase II trial of erlotinib along with sorafenib in patients with recurrent vorinostat plus bortezomib in recurrent GBM was disappointing and
or progressive GBM did not have acceptable results as there were phar- demonstrated an OS of 3.2 months and median TTP of 1.5 months
macokinetic interactions between the two drugs (Peereboom et al., (Friday et al., 2012). Bortezomib has shown to overcome MGMT-
2013). A phase I/II trial of sorafenib combined with erlotinib, tipifanib related resistance of GBM cells to TMZ in vitro by inhibiting MGMT
or temsirolimus in recurrent GBM has been completed (NCT00335764). with scheduling of administration being critical (bortezomib should
be given prior to TMZ) (Vlachostergios et al., 2013). Bortezomib acts
3.6. Other molecular targets and drugs synergistically along with HDAC inhibitors to eliminate GBM stem-
like cells (Asklund et al., 2012). Marizomib (NPI-0052) is a new pro-
The importance of epigenetic alterations in cancer is well appreciat- teasome inhibitor that has been tested in models for glioma (Potts
ed. Histone deacetylases (HDAC) are involved in chromatin structure et al., 2011). The proteasome inhibitor MG132 produces apoptosis
that organize DNA and regulate transcription of genes and are altered of GBM cells and acts as a chemosensitizer in vitro (Zanotto-Filho
in malignant gliomas. HDAC inhibitors have been used in an attempt et al., 2012).
to halt growth arrest and to promote cancer cell apoptosis. Histone Tranforming-growth factor beta (TGF-β) and in particular the β2
acetylases (HATs) and histone hyperacetylation can also degrade isoform, is considered a key factor in malignant glioma progression
HIF-1a and decrease VEGF and thus can lead to antiangiogenic effects (TGF-β2 was originally described as “glioblastoma-derived T-cell sup-
(Galanis et al., 2009). Treatment with HDACs reduces proliferation of pressor factor”) and is associated to the depressed immune status of pa-
glioblastoma-derived stem cells and induces their differentiation tients with GBM. High plasma and tissue levels of TGF-β2 are associated
(Alvarez et al., 2015). Vorinostat (SAHA) is an oral quinolone-based with poor prognosis and advanced disease (Hau et al., 2011). It seems to
HDAC inhibitor that in a phase II study as monotherapy in recurrent also have an essential role in regulation of glioma-stem cells through
GBM demonstrated modest activity with median OS of 5.7 months Smad-dependent induction of LIF and the subsequent activation of
and PFS-6 was achieved in 9 of the first 57 patients (total of 66). Toxic- JAK-STAT pathway (Penuelas et al., 2009). Trabedersen (AP 12009)
ities included hematologic grade 3 or worse (mainly thrombocytope- an antisense oligonucleotide that specifically inhibits TGF-β2
nia) as well as grade 3 or worse nonhematologic toxicities (fatigue, mRNA was tried as monotherapy (administered intratumorally by
dehydration, hypernatremia) (Galanis et al., 2009). Vorinostat in combi- convection enhanced delivery) versus standard chemotherapy
nation with TMZ in high grade gliomas seems to be well tolerated (Lee (TMZ or PCV) in a recently randomized phase IIb study, of patients
et al., 2012). Many trials with vorinostat are ongoing including a phase with recurrent or refractory malignant gliomas. At 10 μM, it achieved
I/II trial of vorinostat plus TMZ/RT in newly diagnosed GBM a median survival of 39.1 months compared with 21.7 months for che-
(NCT00731731) and in various combinations such as with isotretinoin motherapy (Bogdahn et al., 2011). LY2109761 is a TGF-beta receptor I
plus TMZ or bevacizumab plus TMZ/RT. However, it seems that inhibition kinase inhibitor that seems to act as a radiosensitizer and prolongs sur-
of HDAC potentiates the evolution of TMZ resistance by MGMT overex- vival in vitro and in vivo (Zhang et al., 2011) and may be used as an ad-
pression in vitro (Kitange et al., 2012). Panobinostat (LBH589) another junct to RT/TMZ in GBM (Zhang, Herion, et al., 2011).
HDAC inhibitor has been tested in patients with recurrent high-grade gli- SRC and SRC-family kinases (SFKs) mediate downstream signaling
omas in combination with bevacizumab in a phase I trial, while a phase II pathways from several growth factor receptors and they are frequently
study is ongoing (Drappatz et al., 2012). Romidepsin (FR901228, activated in glioblastoma (Wick et al., 2011). It is supported that in-
Depsipeptide), another quinolone-based HDAC inhibitor has been evalu- creased invasiveness associated with antiangiogenic treatment with
ated in a phase I/II trial of patients with recurrent malignant glioma. It VEGF inhibitors is mediated through increased SFK signaling and thus
was largely ineffective with a median PFS was 8 weeks, PFS-6 was 3% justifying the combination of dasatinib with bevacizumab (Huveldt
and median survival duration was 34 weeks (Iwamoto et al., 2011). et al., 2013), however without much success [See dasatinib).
Trichostatin A (TSA) is an HDAC inhibitor that promotes apoptosis on Glutamate was until recently unrecognized as significant in the
GBM cells and potentiates innate immune response against cancer cells pathogenesis of the gliomas. Glutamate is produced as a byproduct of
in vitro and in vivo (Höring et al., 2013). glutathione synthesis and is released by the x(c) (−) cystine glutamate
Heat shock proteins (Hsp) and specifically Hsp90 are molecular receptors in glioma cells. It is involved by either binding on peritumoral
chaperons that assist in stabilizing client proteins such as protein Glu receptors or by paracrine/autocrine effects in: a) in seizure induc-
kinases and several transcription factors. Hsp90 seem to be vital in main- tion, b) excitotoxicity, c) Akt/MAPK activation and cell invasion through
taining malignant transformation and in increasing the growth, survival AMPA receptors, d) inducing focal adhesion kinases, critical for growth
and invasiveness of cancer cells. It stabilizes surface expression of factor regulation and integrin-stimulated cell motility/invasion (de
EGFRvIII and EGFRvIV mutants and probably maintains invasiveness of Groot & Sontheimer, 2011). Talampanel, a well tolerated oral alpha-
GBM (Pines et al., 2010). Tanespimycin (17-AAG) is a potent Hsp90 in- amino-3-hydroxy-5-methyl-4-isoxazolepropionic (AMPA) receptor
hibitor. It exhibits in vitro inhibition of GBM cells and synergistic effects blocker was tested in a phase II trial in patients with newly diagnosed
with radiation. It can also attenuate glioma stem cell radioresistance but GBM in addition to standard RT/TMZ. When compared with historical
not with TMZ (Sauvageot et al., 2009). Acquired resistance is found results with standard treatment, the median survival was 20.3 versus
during prolonged exposure of tanespimycin (Gaspar et al., 2009). 14.6 months and 2 month-survival was 41.7% versus 26.5% favoring
74 C. Alifieris, D.T. Trafalis / Pharmacology & Therapeutics 152 (2015) 63–82

the addition of talampanel (Grossman et al., 2009). In a phase II trial of oxidative damage whereas cathepsin B promotes GBM growth and inva-
talampanel monotherapy against recurrent malignant gliomas was sion, thus proposing auranofin as a GBM treatment adjunct (Kast, 2010).
disappointing but well tolerated with most common toxicities were Disulfiram is an oral aldehyde dehydrogenase (ALD) inhibitor and has
fatigue, dizziness, ataxia and respectively results for GBM/anaplastic strong anti-GBM effects. In vitro effects include P-glycoprotein inactiva-
glioma were PFS-6: 4.6%/0%, median PFS: 5.9 weeks/8.9 weeks and tion, MGMT repair function inhibition, superoxide dismutase inhibition
OS: 13 weeks/14 months (Iwamoto et al., 2010a). Riluzole, a glutamate which secondarily inhibits proteosome, induction of oxidative stress, in-
release inhibitor, in combination with LY341495, a group II mGluR hibition of DNA methyltransferase, reduction of DNA replication and re-
inhibitor, seems to act synergistically to inhibit the growth of GBM duction of NF-kB pathway (Kast & Halatsch, 2012). Nelfinavir, a
in vitro (Yelskaya et al., 2013). protease inhibitor used in the treatment of HIV infection, exhibits anti-
Topoisomerase I inhibitors camptothecin, irinotecan and novel GBM potential by binding to and limiting the chaperone function of
NKTR-102 as well as topoisomerase II inhibitors etoposide and teniposide HSP90 (Kast & Halatsch, 2012). Sertraline, an oral selective serotonine re-
show in vitro efficacy against glioblastoma cell lines (Tamura et al., uptake inhibitor (SSRI), demonstrates downregulation of AKT together
2012). Irinotecan and topotecan have been tested on GBM in various with TMZ and imatinib in vitro (Tzadok et al., 2010). Captopril, an angio-
combinations including TMZ and bevacizumab with irinotecan and tensin conversion enzyme inhibitor (ACEI) inhibits glioblastoma cell syn-
have shown some promising results (Sasine et al., 2010). NKTR-102 is thesis of MMP-2 and MMP-9 thus strongly inhibiting GBM growth and
a topoisomerase I inhibitor polymer conjugate that releases irinotecan invasion (Kast & Halatsch, 2012). Furthermore, GBM cells express
following administration and is further metabolized to SN-38, is cur- AT-1R (67%) and AT-2R (53%) and this high expression is associated
rently being tested in a phase II study of bevacizumab resistant GBM with a poor prognosis (Arrieta et al., 2008). Coordinated undermining
(NCT01663012). Etoposide in a recent meta-analysis was found to im- of survival paths (CUSP9) project is investigating aprepitant, artesunate,
prove OS more than irinotecan in patients with high grade gliomas auranofin, captopril, disulfiram, nelfinavir, sertraline, ketoconazole and
(Leonard & Wolff, 2013). copper gluconate as adjuvant to low dose TMZ in patients with recur-
JAK2/STAT3 pathway, signal transducers and activators of transcrip- rent GBM (Kast et al., 2013).
tion 3, inhibitors have recently risen as approaches for cancer immuno-
therapy. STAT3 is constitutively active in some GBM cells, helping tumor 3.8. Monoclonal antibodies
cells resist apoptosis and enhance local immunosuppression. The non-
receptor tyrosine kinase BMX activates STAT3 signaling to maintain Monoclonal antibodies (MAbs), are targeting immunoglobulins that
self renewal and tumorigenic potential of glioma stem cells and it can specifically recognize cell surface proteins/receptors as antigens, espe-
serve also as a therapeutic target (Guryanova et al., 2011). WP1193 cially targeted on the surface of tumor cells. They are mainly classified
in vitro (Sai et al., 2012) and WP1066 in GBM patients (Hussain et al., on whether they are unconjugated or conjugated (to protein toxins/
2007), are both effective by reducing glioma cell growth as well as re- cytotoxic agents), or radioimmunoconjugates (to radioisotopes). Mabs
ducing glioma stem cell lines. See also manumycin, a FTI. may be of murine, human, primate or chimeric (murine variable region
Double stranded RNA (dsRNA) is an example of pattern recognition plus human constant region) origin (Harris, 2004). Examples of anti
agonists. In a phase II study of the immune modulator polyinosinic- EGFR chimeric MAbs are cetuximab and MAb 806, whereas examples
polycytifylic (poly-ICLC) acid in patients with recurrent anaplastic of humanized MAb are nimotuzumab and panitumumab.
glioma, produced discouraging results: 11% radiographic response, Cetuximab, is an unconjugated chimeric murine-human IgG1 MAb
PFS-6 24% and median survival of 43 weeks (Butowski et al., 2009). A which binds to the extracellular domain of the EGFR, thus inhibiting
phase II study of the poly-ICLC plus standard TMZ/RT in newly diag- the binding of EGF to its receptor. It has a stronger affinity to EGFR
nosed GBM patients, demonstrated improved results relative to histor- (plus the EGFRvIII) than EGF and TGF-α. It also downregulates EGFR,
ical data of standard regimens (Rosenfeld et al., 2010). inhibits cell growth and metastasis, induces apoptosis, inhibits VEGF
production and causes antibody dependent cell-mediated cytotoxicity
3.7. Other experimental therapeutics (Fukai et al., 2008). In recurrent high grade glioma, cetuximab as mono-
therapy has demonstrated PFS-6 in 9.2% of patients, median OS(OS) of
C-150, a novel curry spice curcumin nanoformulative derivative, is a 5 months and average time to progression (TTP) of 2 months. Although
multitargeted agent that reduces the transcription activation of NFkB, patients where stratified according to EGFR amplification, no correla-
inhibits PKC-alpha, VEGF, Cyclin D1, BCL-XL and promotes differentia- tion to response was found (Neyns et al., 2009). Again in the recurrent
tion of glioma-initiating cells by inducing autophagy in vitro and GBM setting, the combination of bevacizumab, irinotecan and
shows efficacy in vivo (Langone et al., 2013; Zhuang et al., 2012). cetuximab showed PFS-6 in 30% of patients, median OS 7.2 months
When curcumin is coupled to a glioblastoma-directed antibody, it po- and radiographic response of 34% (Hasselbalch et al., 2010). Amplifica-
tentiates its antitumor activity (Langone et al., 2013). Curcumin in- tion of EGFR and the type of EGFR mutation (EGFRvIII or EGFRvIV) may
creases GBM cell sensitivity to TMZ by generating reactive oxygen determine the outcome in GBM patients treated with cetuximab.
species and by disrupting the AKT/mTOR pathway (Yin et al., 2014). Patients with EGFR amplification and lack of EGFRvIII had better OS
Niclosamide is an anthelminthic medication that has demonstrated (p = 0.12) than patients with EGFRvIII expression (p = 0.08) (Lv
anti-tumor effects in vitro, probably mediated through inhibition et al., 2012). No trials have been publiced with cetuximab as treatment
of WNT/CTNNB1-, Notch-, mTOR-, and NF-kB signaling pathways in newly diagnosed high grade gliomas up to this date. MAb 806, seems
(Wieland et al., 2013). Aprepitant is an oral neurokinin-1 receptor to enhance the efficacy of ionizing radiation in glioma xenografts ex-
(NK-1R) antagonist used in treating nausea and vomiting. Substance P pressing the EGFRvIII mutation and can inhibit the growth of U87MG
is a natural ligand of NK-1R and thus blockage of this pathway seems (Johns et al., 2012). Nimotuzumab, a humanized MAb with preferential
to inhibit proliferation, induce apoptosis, antiangiogenic and anti- binding to high-EGFR-density tissues (e.g tumors, sparing normal
magration effects in preclinical trials (Muñoz & Coveñas, 2012). tissue) has been tested in pediatric recurrent high grade gliomas
Artesunate, an antimalarial drug, induces oxidative DNA damage, DNA (MackDonald et al., 2011). Radiotherapy plus nimotuzumab or placebo
double-strand breakage and the ATM/ATR damage response in cancer in anaplastic astrocytoma and GBM demonstrated excellent safety pro-
cells (Berdelle et al., 2011) and potentiates erlotinib’s EGFR mediated cy- file and significant survival bebefit with mean survival of 31.06 months
totoxicity (Efferth et al., 2004). Auranofin, a medication used to treat in nimotuzumab treated group versus 21.07 for the control group
rheumatoid arthritis, inhibits cytoprotective thioredoxin reductase and (Solomón et al., 2013). A Chinese Phase II study of nimotuzumab in
cathepsin B. Thioredoxin reductase is upregulated by TNF-alpha- combination with TMZ/RT demonstrated good safety and tolerability
induced apoptosis (ATIA) protein in glioblastoma and protects from without prolonging survival compared to standard therapy (Wang
C. Alifieris, D.T. Trafalis / Pharmacology & Therapeutics 152 (2015) 63–82 75

et al., 2014). Panitumumab (ABX-EGF), a human monoclonal antibody of 10.5 months (Nagane et al., 2012). However, most patients progress
specific for HER-1, has been used in combination with irinotecan and die from the disease within 6–9 months (Kang et al., 2008).
for GBM but results have not been published (NCT01017653) Patients with progression after bevacizumab failure, have no standard op-
(Berezowska & Schlegel, 2011). AMG 595, an anti-EGFRvIII-DM1 tions for salvage treatment. Given the lack of effective salvage regimens,
immunoconjugate human monoclonal antibody is currently being eval- possible options include treatment, with bevacizumab again until a sec-
uated in a phase I study patients with recurrent GBM expressing mutant ond or further recurrence in patients with high performance status and
EGFRvIII (NCT01475006). small non-rapidly enlarging tumors or combinations, with targeted ther-
Anti-VEGFR MAbs include bevacizumab, ramucirumab (IMC-1121B) apies that inhibit possible resistance mechanisms such as other growth
and icrucumab (IMC 18 F1). Bevacizumab, is an unconjugated recombi- factors, or circulating progenitor cells (Krashraw & Lassman, 2010).
nant humanized murine MAb which binds to and neutralizes VEGF-A, Other combinations of bevacizumab have been tested such as
preventing the activation of VEGF receptor tyrosine kinases VEGFR-1 plus etoposide or TMZ (Reardon et al., 2011a), plus erlotinib
and -2. Given i.v., it has an important toxicity profile. Common (Sathornsumetee et al., 2010) and bevacizumab with dose-dense TMZ
manifestations of toxicity include hypertension, headache, gastrointes- (Gilbert et al., 2010). However, response rate, PFS-6 and OS where
tinal effects (abdominal pain, nausea, vomiting, diarrhea, anorexia, con- similar to the bevacizumab monotherapy rates seen in BRAIN study.
stipation), proteinuria, leucopenia, weakness, stomatitis, epistaxis, Ramucirumab (IMC-1121B) (NCT00895180) and icrucumab (IMC-
dyspnea, upper respiratory tract infection and impaired wound healing. 18 F1) are newer monoclonal antibodies that target the VEGF receptors
Some more rare but severe side-effects are central nervous system hem- VEGFR-2 and VEGFR-1 respectively. Trials on GBM are under way but
orrhage, venous and arterial thromboembolic events, bowel perfora- not published up to this date (Hsu & Wakelee, 2009).
tion, left ventricular dysfunction, infusion related reactions and Aflibercept (Hsu & Wakelee, 2009) (also known as VEGF trap), is a
nephrotic syndrome (De Fazio et al., 2012). In the newly diagnosed set- recombinantly produced soluble peptide-antibody fusion that scav-
ting of GBM, since the concern of serious toxicity is great enough, assess- enges VEGF ligand and placental growth factor. In patients with GBM,
ment of risk versus benefit is important. When patients with MGMT two separate phase II studies have been published (Gomez-Manzano
methylation is treated with RT and TMZ without bevacizumab, about et al., 2008; de Groot et al., 2011). Generally it showed minimal activity
50% survive 2 years (Stupp et al., 2009). So it is of concern whether and specifically demonstrated PFS-6 in 7.7% of patients, response rate of
the addition of bevacizumab will prolong this OS or shorten it due to 18% and median PFS of 3 months (Gomez-Manzano et al., 2008). Toxic-
toxicity. Also, there is concern that addition of bevacizumab promotes ity led to discontinuation of the treatment in a quarter of patients which
a more invasive and aggressive tumor hence shortening survival com- included hypertension, lymphopenia, fatigue and central nervous
pared to RT/TMZ alone. In a way to avoid the risk, a stratification system system toxicity (cerebral ischemia).
has been employed to design GBM trials. The recursive partitioning MAbs against PDGFR-α, include IMC-3G3 and MEDI-575. MEDI-575
technique (RPA) can be used. Preliminary data demonstrated that is investigated in ongoing phase I/II studies in recurrent GBM patients
there is a trend towards longer survival (when compared to standard (NCT01268566) and results are pending. IMC-3G3, has shown inhibi-
RT/TMZ treatment) for patients with low RPA class (e.g. poor perfor- tion of growth (69%) in U118 GBM cell lines and has been tried in
mance status and older age) but not for patients with a high RPA class patients with ovarian cancer (see NCT00895180) (Shah et al., 2010).
when bevacizumab is added. However, a longer PFS-6 was achieved Rilotumumab (AMG102), is a fully human MAb that effectively
with no association to RPA class (Curran et al., 1993; Lai et al., 2009). blocks c-Met phosphorylation, thus inhibiting c-Met activated signaling
In a phase II study which compared the addition of bevacizumab to pathways (Giordano, 2009). It has been shown to radiosensitize GBM
standard RT/TMZ versus standard treatment alone, showed PFS-6 in cells in vivo and in vitro (Buchanan et al., 2011). In a phase II study, pa-
77.5% versus 51.6% of patients respectively and median PFS 17 months tients with recurrent GBM treated with rilotunumab. Two cohorts were
versus 7 months respectively (Gruber et al., 2009). A randomized designed depending on the rilotunumab dose (10 and 20 mg/kg) and
phase III trial of the addition of bevacizumab or placebo in newly diag- respectively median OS was 6.5 months versus 5.4 months and PFS
nosed GBM, did not improve the OS (median 15.7 vs 16.1 months re- was 4.1 weeks versus 4.3 weeks. No significant statistical difference
spectively) though it improved the PFS (10.7 vs 7.3 months) but was attained regarding the pretreatment status with bevacizumab.
increased symptom burden and the rate of adverse effects led to an at- The results were disappointing and the most common adverse effects
tenuation in quality of life (Gilbert et al., 2014). were fatigue, headache and peripheral edema (Wen et al., 2011). Two
In the recurrent setting, bevacizumab has been evaluated either more phase II studies are currently underway evaluating the combina-
alone or in combination with other agents such as irinotecan, etoposide, tion of AMG102 and bevacizumab in patients with recurrent malignant
carboplatin and TMZ. The phase II BRAIN study, evaluated the role of glioma (NCT01113398) as well as AMG102 monotherapy in advanced
bevacizumab alone or in combination with irinotecan in patients with malignant gliomas (NCT00427440). Onartuzumab (MetMAb or OA-
recurrent GBM. Monotherapy versus combination showed respectively 5D5), is a novel human, monovalent (one-armed) 5D5 (OA-5D5) anti-
PFS-6 in 42.6% versus 50.3% of patients, objective response 28.2% versus c-Met antibody that was found to be effective in reducing proliferation
37.8% and median OS of 9.2 versus 8.7 months. Both groups of patients and microvessel density, as well as in increasing apoptosis in U87
had superior radiographic response rates, median PFS and PFS-6 in rela- GBM cells (c-Met and HGF/scatter factor positive cell lines) (Martens
tion to historical controls (Friedman et al., 2009). However, this study et al., 2006). A phase II study of onartuzumab in combination with
was a non-comparative study, so it wasn’t designed to compare the bevacizumab versus bevacizumab alone or onartuzumab alone in pa-
two arms. A second study, tested patients with recurrent GBM treated tients with recurrent GBM is in progress (NCT01632228).
with bevacizumab and it showed a response rate of 35%, PFS-6 in 29%
of patients and a median OS of 31 weeks, decreases in cerebral edema, 3.9. Stem cells
lower needs for corticosteroids and improved neurologic function.
Patients that experienced later progression, did not show any improve- It has been noted earlier that glioma stem cells have proliferative, in-
ment when irinotecan was added (Kreisl et al., 2009). After the BRAIN vasive capabilities and promote angiogenesis. EGFR inhibitors (that re-
study, there was still dispute on whether monotherapy of bevacizumab duce the CD133+ population of stem cells), PI3K/Akt/mTOR pathway
was of any real benefit but after the second forementioned study, inhibitors, GSIs, anti-angiogenetic agents (e.g bevacizumab or cediranib
FDA rushed into an accelerated approval in 2009 and since then, to target the vascular niche of glioma stem cells), Notch pathway inhib-
bevacizumab is the approved monotherapy for recurrent GBM. Clinical itors, Hsp90 inhibitors, Sonic hedgehog pathway inhibitors have all
benefit was also demonstrated in Japanese patients with PFS-6 in been used to block glioma stem cell population and perhaps enhance ra-
33.9% of patients, median PFS of 3.3 months and median overall survival diosensitivity. Cell surface molecules on glioma stem cells may be used
76 C. Alifieris, D.T. Trafalis / Pharmacology & Therapeutics 152 (2015) 63–82

as pharmacological targets. L1CAM specifically is a glycoprotein that chimeric antigen receptors (CAR), can recognize antigens with HLA-
contains a cytoplasmic tail, a transmembrane domain and its extracellu- independent mechanisms. CARs are fusion molecules of MAb specific
lar domain that is preferentially found in glioma stem cells compared to for a tumor-antigen and the ζ-chain of the T-cell-receptor (TCR). CAR-
non-stem cells of GBM tumors. It can either bind with extracellular ma- modified T-cells have been manufactured against IL-13 receptor alpha
trix proteins, EGFR, FGFR, α5β1, avb3, neurophilin-1 or with another 2 (IL13Ra2) (Krebs et al., 2014) and EGFRvIII (Choi et al., 2014). A
L1CAM molecule. L1CAM acts as a mediator of molecular signals and reg- phase I/II trial of autologous CARs against EGFRvIII in glioblastoma is
ulates survival, growth, adhesion, invasion and migration glioma stem in development (NCT01454596). A phase I trial of postsurgical tumoral
cells. Some of its effects may be due to upregulation of Olig2 that sup- cavity injection of IL13Ra2 in GBM has been completed (NCT01082926).
presses p21WAF/CIP1 (Schmidt & Maness, 2008). L1CAM expression is in- Active Immunotherapy – Vaccine strategies: Although we have a
creased with mutations of TP53 or expression of TGF-beta (Tsuzuki vast arsenal of agents to attack the EGFRs in GBM, a large amount of
et al., 1998). L1CAM also contributes in the chemoresistance of GBM EGFRs still cannot be blocked due to the receptor’s ubiquitous and
cells and stem cells (Held-Feindt et al., 2012). Blocking L1CAM, seems high level expression. A peptide vaccine, rindopepimut (CDX-110), is
to be able to suppress tumor growth and prolong survival of tumor directed against the novel exon 1–8 junction produced by the EGFRvIII
cells in vivo (Bao et al., 2008). mutation. Recent phase II studies have shown EGFRvIII specific immune
Since malignant gliomas contain many genetic alterations, various response and significantly prolonged TTP and OS in patients with newly
viral vectors have been employed to transmit therapeutic genes into diagnosed GBM compared to standard care. It represents a promising
their neoplasmatic cells and target those alterations. The replication field of therapy in patients with GBM (Del Vecchio et al., 2012). A
oncolytic adenovirus Delta-24-RGD is capable of engaging the complete phase III study in patients with newly diagnosed GBM is under way
cellular system within the glioma cells for its own replication and sur- (NCT01480479). ITK-1 is a personalized peptide vaccine given in HLA-
vival. Even resistant glioma stem cells seem to be vulnerable, suggesting A24 positive recurrent or progressive GBM patients (Terasaki et al.,
autophagy (upregulation of Atg5 in the tumor cells) followed by lysis of 2011).
the malignant cell (Jiang et al., 2007). A phase I/II trial in patients with Heat shock protein peptide vaccine (HSPPV) uses peptides com-
recurrent GBM (NCT01582516) and a phase I trial for patients with plexed with Hsp proteins that are effectively used by APCs (See et al.,
recurrent malignant gliomas (NCT00805376) are ongoing. 2011). Vitespen a heat shock protein (gp96)-peptide complex purified
from autologous tumors has been used in phase I/II trials with minimal
3.10. Immunotherapy side-effects and clinical responses in early-stage disease (Wood &
Mulders, 2009).
Immunotherapy is a promising approach that has demonstrated an Other peptide vaccines for malignant gliomas include peptides
ability to specifically eliminate cancer cells without damaging the sur- against melanoma/testis antigens, viral antigens, cytokine receptors
rounding healthy tissue. It includes passive and active immunotherapy and differentiation antigens (Mitchell & Sampson, 2009).
strategies. Active immunotherapy creates a long-term immune re- Gene-modified tumor vaccines work in a different way by inducing
sponse against the tumor (propably effective against recurrences) and GBM cells to be more immunogenic, express new antigens and thus
passive immunotherapy involves the transfer of immediately and give more targets for the immune system to fight against. Using autolo-
short-acting effective molecules. Tumor-specific antigens are processed gous glioma cells modified to secrete either granulocyte/macrophage
by antigen presenting cells (APCs) and recognized by lymphocytes. colony factor (GM-CSF) or TGFβ2 antisense vectors (Fakhrai et al.,
GBM has the ability to modulate the activity of APCs and this is hypoth- 2006) or IL-4 (Okada et al., 2007), response has been seen in small trials
esized from vaccination studies using tumor lysates that have been of progressive or recurrent GBM patients.
unsuccesful (Jackson et al., 2013). Evidence shows that microglia is controlled by glioma cells resulting
Passive immunotherapy utilizes monoclonal antibodies, adoptive in their support for growth and thus, instead of destroying them they
cell transfer and cytokine-mediated therapies. Targeted monoclonal an- help by contributing to tumor progression. This may be mediated
tibodies that were discussed in the above sections can be lethal to tumor through colony-stimulating factors, interleukin-6, TGF-β2 and mono-
cells by immune-mediated mechanisms as well as by their specific- cyte chemotactic protein-1(MCP-1) that are being produced by glioma
targeted effects. MAbs bind on cancer cell surfaces in a major histocom- cells (Charles et al., 2011). In turn, microglial cells support the survival,
patibility complex (MHC) unrestricted manner and induce GBM cell proliferation and invasion of glioma cells via secretion of factors such as
death by both immune and non-immune mechanisms. Nivolumab, a EGF, VEGF and stress-inducible-protein-1 (STI1) (Alves et al., 2011).
fully human MAb against programmed death receptor-1 (PD-1), a neg- Dendritic cell (antigen-presenting) vaccines are made from dendrit-
ative checkpoint regulator with immunosuppressive capabilities has ic cells extracted from the patient, then exposed to the antigens of GBM
been approved for treatment in unresectable malignant melanoma in cells and the same (now “primed”) cells are infused back into the pa-
Japan (Deeks, 2014). Ipilimumab is a MAb that binds to cytotoxic tient. In turn, they activate the CD8+ (cytotoxic) T cells and present
T-lymphocyte (CTL)-associated antigen-4 (CTLA-4/CD152) and en- GBM antigen which is now being recognized by the host immune sys-
hances T-cell activation thus allowing CTLs to destroy the cancer cells. tem. Tumor antigens that are being used are HER2, TRP2, Epha 2,
It has been approved for use in malignant melanoma (Karimkhari GP100 and KI3ra (Lesniak, 2011). Glioma stem cells contribute to resis-
et al., 2014). Nivolumab efficacy in recurrent GBM is being investigated tance to chemoradiotherapy, thus introducing dendritic cells against
in a phase IIb trial as monotherapy or in combination with ipilimumab these stem cells (using specific antigens such as CD133 and nestin)
versus bevacizumab (NCT02017717). would be an important therapeutic option. In a non-randomized
Adoptive cell transfer uses effector cells that are activated ex vivo phase II study, anti-EGFRvIII dendritic cell tumor vaccine addition to
and then tranfered to a patient to attack GBM cells. Effector cells can standard first line treatment in GBM patients with good performance
be lymphocyte-activated killer (LAK) cells, cytotoxic T lymphocytes status, when compared to a matched control group from MD Anderson,
(CTLs), natural-killer (NK) cells and γδ-T-cells. Results generally it resulted in an increase of median OS from 15 to 26 months and an in-
have been mixed with good tolerability but the OS was not significantly crease in median PFS from 6.3 months to 14.2 months (hazard ration
improved (Dillman et al., 2009). Newer CTLs that are more specific 2.4; P = o.o13 in favor of the vaccination group). The PFS-6 rate after
to tumor-associated antigens like CD133 which show activity vaccination was 67% (95% CI, 40–83%) and after diagnosis was 94%
against GBM stem cells (Hua et al., 2011) or virus-specific CTLs (e.g. (95%, 67–99%, n = 18). The development of antibodies or delayed
against CMV proteins pp65 and IE1-72 or nucleic acids) are intensively type hypersensitivity had a significant effect on OS. When the tumor re-
being investigated against GBM (Nair et al., 2014) (NCT00693095, lapsed, EGFRvIII expression was lost in 82% (95% CI, 48–97%). In subcat-
NCT01109095). Genetically modified T-cells with novel artificial egory analysis, outcomes were improved in both O-MGMT methylated
C. Alifieris, D.T. Trafalis / Pharmacology & Therapeutics 152 (2015) 63–82 77

and unmethylated patients. These results demonstrate that in EGFRvIII- Batchelor, T. T., Duda, D. G., di Tomaso, E., Ancukiewicz, M., Plotkin, S. R., Gerstner, E., et al.
(2010). Phase II study of cediranib, an oral pan-vasular endothelial growth factor
positive GBM, outcomes are improved (Sampson et al., 2010). receptor tyrosine kinase inhibitor, in patients with recurrent glioblastoma. J Clin
Other treatment strategies include microglia as gene or factor thera- Oncol 28(17), 2817–2823 (Epub 2010 May 10).
peutic vectors into the central nervous system and glioma tumors. Berdelle, N., Nikolova, T., Quiros, S., Efferth, T., & Kaina, B. (2011). Artesunate induces
oxidative DNA damage, sustained DNA double-strand breaks, and the ATM/ATR dam-
Examples include microglia carrying the LacZ gene (Sawada et al., age response in cancer. Mol Cancer Ther 10(12), 2224–2233.
1998), inducible thymidine kinase gene or MRI contrast agents (Ribot Berezowska, S., & Schlegel, J. (2011). Targeting ErbB receptors in high-grade glioma. Curr
et al., 2007) or biodegradable nano-vehicles carrying DNA or RNA Pharm Des 17, 2468–2487.
Bhat, N. R., & Fan, F. (2002). Adenovirus infection induces microglial activation: involve-
silencer fragmants into the glioma tumor (Carbon nanotubes) (Zhao ment of mitogen-activated protein kinase pathways. Brain Res 948, 93–101.
et al., 2011). Microglial cell gene therapy through viral vectors of either Bogdahn, U., Hau, P., Stockhammer, G., Venkataramana, N. K., Mahapatra, A. K., Suri, A.,
adenovirus (Bhat & Fan, 2002) or lentivirus (the latter using an et al. (2011). Targeted therapy for high-grade glioma with the TGF-β2 inhibitor
trabedersen: results of a randomized and controlled phase IIb study. Neuro Oncol
enhanced GFP gene) (Balcaitis et al., 2005) have also been tried.
13(1), 132–142.
Intratumoral administration of oncolytic myxoma virus with rapamycin Brandes, A. A., Stupp, R., Hau, P., Lacombe, D., Gorlia, T., Tosoni, A., et al. (2010). EORTC
was another approach (Lun et al., 2012). Although phase III data are study 26041-22041: phase I/II study on concomitant and adjuvant temozolomide
lacking, preclinical and phaseI/II trial results, stongly state that immuno- (TMZ) and radiotherapy (RT) with PTK787/XK222584 (PTK/ZK) in newly diagnosed
glioblastoma. Eur J Cancer 46(2), 348–354.
therapy is an area with promising results and offer a new weapon Buchanan, S. G., Hendle, J., Lee, P. S., Smith, C. R., Bounaud, P. Y., Jessen, K. A., et al. (2009).
against this lethal tumor. SGX523 is an exquisitely selective, ATP-competitive inhibitor of the MET receptor
tyrosine kinase with antitumor activity in vivo. Mol Cancer Ther 8(12), 3181–3190.
Buchanan, I. M., Scott, T., Tandle, A. T., Burgan, W. E., Burgess, T. L., Tofilon, P. J., et al. (2011).
4. Conclusion Radiosensitization of glioma cells by modulation of Met signaling with the hepatocyte
growth factor neutralizing antibody, AMG102. J Cell Mol Med 15(9), 1999–2006.
Butowski, N., Chang, S. M., Lamborn, K. R., Polley, M. Y., Pieper, R., Costello, J. F., et al.
Research in GBM treatment is ongoing, vast, and rapidly evolving. (2011). Phase II and pharmacogenomics study of enzastaurin plus temozolomide
However, even all this data and progression of molecular science during and after radiation therapy in patients with newly diagnosed glioblastoma
multiforme and gliosarcoma. Neuro Oncol 13(12), 1331–1338.
it has not been able to battle effectively this tumor. No matter how
Butowski, N., Lamborn, K. R., Lee, B. L., Prados, M. D., Cloughesy, T., DeAngelis, L. M., et al.
many different targets are discovered and molecules to aim them are (2009). A North American brain tumor consortium phase II study of Poly-ICLC for
enginered, the end result is that we have made only a little progress for- adult patients with recurrent anaplastic gliomas. J Neurooncol 91(2), 183–189.
Butowski, N. A., Sneed, P. K., & Chang, S. M. (2006). Diagnosis and treatment of recurrent
ward in improving overall survival. However, as is seen in this review,
high-grade astrocytoma. J Clin Oncol 24, 1273–1280.
every step in the way, new lessons drive us forward and small details Carrillo, J. A., & Munoz, C. A. (2012). Alternative chemotherapeutic agents: nitrosureas,
help us bypass the presenting obstacles. New targets, novel antagonists, cisplatin, irinotecan. Neurosurg Clin N Am 23(2), 297–306.
more effective drug penetration through the BBB and new methods of Chakravanti, A., Zhai, G., Suzuki, Y., Sarkesh, S., Black, P. M., Muzikansky, A., et al. (2004).
The prognostic significance of phosphatidylinositol 3-kinase pathway activation in
manipulating the immune response are all puzzles to be solved in the human gliomas. J Clin Oncol 22, 1926–1933.
near future. Chang, S. M., Wen, P., Cloughesy, T., Greenberg, H., Schiff, D., Conrad, C., et al. (2005).
Phase II study of CCI-779 in patients with recurrent glioblastoma multiforme. Invest
New Drugs 23(4), 357–361.
Conflict of interest statement Chaponis, D., Barnes, J. W., Dellagatta, J. L., Kesari, S., Fast, E., Sauvageot, C., et al. (2011).
Lonafanib (SCH66336) improves the activity of temozolomide and radiation for
orthotopic malignant gliomas. J Neurooncol 104(1), 179–189.
The authors declare that there are no conflicts of interest. Charles, N. A., Holland, E. C., Gilbertson, R., Glass, R., & Kettermann, H. (2011). The brain
tumor microenvironment. Glia 59, 1169–1180.
Chen, C., Xu, T., Chen, J., & Wu, S. (2013). The efficacy of temozolomide for recurrent
References glioblastoma multiforme. Eur J Neurol 20, 223–230.
Cheng, L., Huang, Z., Zhou, W., Wu, Q., Donnola, S., Liu, J. K., et al. (2013). Glioblastoma
Agarwal, S., Mittapali, R. K., Zellmer, D. M., Gallardo, J. L., Donelson, R., Seiler, C., et al. stem cells generate vascular pericytes to support vessel function and tumor growth.
(2012). Active efflux of dasatinib from the brain limits efficacy against murine Cell 153, 139–152.
glioblastoma: broad implications for the clinical use of molecularly-targeted agents. Cheng, H., Yue, W., Xie, C., Zhang, R., Hu, S., & Wang, Z. (2013). IDH1 mutation is associ-
Mol Cancer Ther 11(10), 2183–2192. ated with improved overall survival in patients with glioblastoma: a meta-analysis.
Ahluwalia, M. S. (2011). 2010 Society for Neuro-Oncology Annual meeting: a report of Tumour Biol 34(6), 3555-3359.
selected studies. Expert Rev Anticancer Ther 11(2), 161–163. Chi, A. S., Batchelor, T. T., Kwak, E. L., Clark, J. W., Wang, D. L., Wilner, K. D., et al. (2012).
Alexander, B. M., Wang, M., Yung, W. K., Fine, H. A., Donahue, B. A., Tremont, I. W., et al. Rapid radiographic and clinical improvement after treatment of a MET-amplified
(2013). A phase II study of conventional radiation therapy and thalidomide for recurrent glioblastoma with a mesenchymal-epithelial transition inhibitor. J Clin
supratentorial, newly diagnosed glioblastoma (RTOG 9806). J Neurooncol 111, 33–39. Oncol 30(3), e30–e33 (Epub Dec 12).
Alvarez, A. A., Field, M., Bushnev, S., Longo, M. A., & Sugaya, K. (2015). The effects of his- Chinot, O. L., Honore, S., Dufour, H., Barrie, M., Figarella-Branger, D., Muracciole, X., et al.
tone deacetylase inhibitors on glioblastoma derived cells. J Mol Neurosci 55, 7–20. (2001). Safety and efficacy of temozolomide in patients with recurrent anaplastic
Alves, T. R., Lima, F. R., Kahn, S. A., Lobo, D., Dubois, L. G., Soletti, R., et al. (2011). Glioblas- oligodendrogliomas after standard radiotherapy and chemotherapy. J Clin Oncol
toma cells: a heterogenous and fatal tumor interacting with the parenchyma. Life Sci 19(9), 2449–2455.
89, 532–539. Choi, B. D., Suryadevara, C. M., Gedeon, P. C., Herndon Ii, J. E., Sanchez-Perez, L., Bigner, D.
Arora, A., & Scholar, E. M. (2005). Role of tyrosine kinase inhibitors in cancer therapy. J D., et al. (2014). Intracerebral delivery of a third generation EGFRvIII-specific chimeric
Pharmacol Exp Ther 315(3), 971–979. antigen receptor is efficacious against human glioma. J Clin Neurosci 21(1), 189–190.
Arrieta, O., Pineda-Olvera, B., Guevara-Salazar, P., Hernandez-Pedro, N., Morales-Espinosa, Choy, G., Joshi-Hangal, R., Oganesian, A., Fine, G., Rasmussen, S., Collier, J., et al. (2012).
D., Ceron-Lizarraga, T. L., et al. (2008). Expression of AT1 and AT2 angiotensin recep- Safety, tolerability, and pharmacokinetics of amuvatinib from three phase 1 clinical
tors in astrocytomas is associated with poor prognosis. Br J Cancer 99, 160–166. studies in healthy volunteers. Cancer Chemother Pharmacol 70(1), 183–190.
Asklund, T., Kvarnbrink, S., Holmlund, C., Bergenheim, T., Henriksson, R., & Hedman, H. Clarke, J. L., Iwamoto, F. M., Sul, J., Panageas, K., Lassman, A. B., DeAngelis, L. M., et al.
(2012). Synergistic killing of glioblastoma stem-like cells by bortezomib and HDAC (2009). Randomized phase II trial of chemoradiotherapy followed by either dose-
inhibitors. Anticancer Res 32(7), 2407–2413. dense or metronomic temozolomide for newly diagnosed glioblastoma. J Clin Oncol
Balcaitis, S., Weinstein, J. R., Li, S., Chamberlain, J. S., & Moller, T. (2005). Lentiviral trans- 27(23), 3861–3867.
duction of microglial cells. Glia 50, 48–55. Clement, V., Sanchez, P., de Tribolet, N., Radovanovic, I., & Ruiz i Altaba, A. (2007).
Baltuch, G. H., Couldwell, W. T., Villemure, J. G., & Yong, V. W. (1993). Protein kinase C HEDGEHOG-GLI1 signaling regulates human glioma growth, cancer stem cell self-
inhibitors suppress cell growth in established and low-passage glioma cell lines. A renewal, and tumorigenicity. Curr Biol 17, 165–172.
comparison between staurosporine and tamoxifen. Neurosurgery 33(3), 495–501. Cloughesy, T. F., Wen, P. Y., Robins, H. I., Chang, S. M., Groves, M. D., Fink, K. L., et al.
Bao, S., Wu, Q., Li, Z., Sathornsumetee, S., Wang, H., McLendon, R. E., et al. (2008). (2006). Phase II trial of tipifarnib in patients with recurrent malignant glioma either
Targeting cancer stem cells through L1CAM suppresses glioma growth. Cancer Res receiving or not receiving enzyme-inducing anti-epileptic drugs: a North American
68(15), 6043–6048. Brain Tumor Consortium Study. J Clin Oncol 24(22), 3651–3656.
Bao, S., Wu, Q., Sathornsumetee, S., Hao, Y., Li, Z., Hjelmeland, A. B., et al. (2006). Stem cell- Curran, W. J., Jr., Scott, C. B., Horton, J., Nelson, J. S., Weinstein, A. S., Fischbach, A. J., et al.
like glioma cells promote tumor angiogenesis through vascular endothelial growth (1993). Recursive partitioning analysis of prognostic factors in three Radiation
factor. Cancer Res 66(16), 7843–7848. Therapy Oncology Group malignant glioma trials. J Natl Cancer Inst 85, 704–710.
Bar, E. E., Chaudhry, A., Lin, A., Fan, X., Schreck, K., Matsui, W., et al. (2007). Cyclopamine- De Faria, G. P., De Oliveira, J. A., Oliveira, J. G. P., Romano, S. D., Neto, V. M., & Maia, R. C.
mediated hedgehog pathway inhibition depletes stem-like cancer cells in glioblasto- (2008). Differences in the expression pattern of P-glycoproteins and MRP1 in low-
ma. Stem Cells 25(10), 2524–2533. grade and hogh-grade gliomas. Cancer Invest 26, 883–889.
78 C. Alifieris, D.T. Trafalis / Pharmacology & Therapeutics 152 (2015) 63–82

De Fazio, S., Russo, E., Ammendola, M., Donato Di paola, E., & De Sarro, G. (2012). Efficacy Floyd, D. H., Kefas, B., Seleverstov, O., Mykhaylyk, O., Dominguez, C., Comeau, L., et al.
and safety of bevacizumab in glioblastomas. Curr Med Chem 19(7), 972–981. (2012). Alpha secretase inhibition reduces human glioblastoma stem cell growth
de Groot, J. F., Lamborn, K. R., Chang, S. M., Gilbert, M. R., Cloughesy, T. F., Aldape, K., et al. in vitro and in vivo by inhibiting Notch. Neuro Oncol 14(10), 1215–1226.
(2011). Phase II study of aflibercept in recurrent malignant glioma: a North American Fouladi, M., Stewart, C. F., Olson, J., Wagner, L. M., Onar-Thomas, A., Kocak, M., et al.
Brain Tumor Consortium study. J Clin Oncol 29(19), 2689–2698. (2011). Phase I trial of MK-0752 in children with refractory CNS malignancies: a
de Groot, J., & Sontheimer, H. (2011). Glutamate and the biology of gliomas. Glia 59(8), pediatric consortium study. J Clin Oncol 29(26), 3529–3534.
1181–1189. Franceschi, E., Stupp, R., van den Bent, M. J., van Herpen, C., Laigle, Donadey F., Gorlia, T.,
De la Pena, L., Burgan, W. E., Carter, D. J., Hollingshead, M. G., Satyamitra, M., Camphausen, et al. (2012). EORTC 26083 phase I/II trial of dasatinib in combination with CCNU in
K., et al. (2006). Inhibition of Akt by the alkylphospholipid perifosine does not en- patients with recurrent glioblastoma. Neuro Oncol 14, 1503–1510.
hance the radiosensitivity of human glioma cells. Mol Cancer Ther 5(6), 1504–1510. Friday, B. B., Anderson, S. K., Buckner, J., Yu, C., Giannini, C., Geoffroy, F., et al. (2012). Phase
Dean, M., Fojo, T., & Bates, S. (2005). Tumour stem cells and drug resistance. Nat Rev II trial of vorinostat in combination with bortezomib in recurrent glioblastoma: a
Cancer 5, 275–284. north central cancer treatment group study. Neuro Oncol 14(2), 215–221.
Deeks, E. D. (2014). Nivolumab: a review of its use in patients with malignant melanoma. Friedman, H. S., Prados, M. D., Wen, P. Y., Mikkelsen, T., Schiff, D., Abrey, L. E., et al. (2009).
Drugs 74(11), 1233–1239. Bevacizumab alone and in combination with irinotecan in recurrent glioblastoma. J
Del Vecchio, C. A., Li, G., & Wong, A. J. (2012). Targeting EGF receptor variant III: tumor Clin Oncol 27, 4733–4740.
specific peptide vaccination for malignant gliomas. Expert Rev Vaccines 11(2), Fu, J., Rodova, M., Nanta, R., Meeker, D., Van Veldhuizen, P. J., Srivastava, R. K., et al. (2013).
133–144. NPV-LDE-225 (Erismodegib) inhibits mesenchymal transition and self-renewal of
Desjardins, A., Reardon, D. A., Peters, K. B., Threatt, K. B., Coan, A. D., Herndon, J. E., et al. glioblastoma initiating cells by regulating miR-21, miR-128, and miR-200. Neuro
(2011). A phase I trial of the farnesyltransferase inhibitor, SCH 66336, with temozo- Oncol 15, 691–706.
lomide for patients with malignant glioma. J Neurooncol 105(3), 601–606. Fukai, J., Nishio, K., Itakura, T., & Koizumi, F. (2008). Antitumor activity of cetuximab
di Tomaso, E., London, N., Fuja, D., Logie, J., Tyrrell, J. A., Kamoun, W., et al. (2009). PDGF-C against malignant glioma cells overexpressing EGFR deletion mutant variant III.
induces maturation of blood vessels in a model of glioblastoma and attenuates the Cancer Sci 99(10), 2062–2069.
response to anti-VEGF treatment. PLoS One 4(4) (Apr 8, Epub ahead of print). Furnari, F. B., Fenton, T., Bachoo, R., Mukasa, A., Stommel, J. M., Stegh, A., et al. (2007).
Dillman, R. O., Duma, C. M., Ellis, R. A., Cornforth, A. N., Schiltz, P. M., Sharp, S. L., et al. Malignant astrocytic glioma: genetics, biology and paths to treatment. Genes Dev
(2009). Intralesional lymphokine-activated killer cells as adjuvant therapy for primary 21(21), 2683–2710.
glioblastoma. J Immunother 32(9), 914–919. Galanis, E., Buckner, J. C., Maurer, M. J., Kreisberg, J. I., Ballman, K., Boni, J., et al. (2005).
Dixit, D., Sharma, V., Ghosh, S., Koul, N., Mishra, P. K., & Sen, E. (2009). Manumycin inhibits Phase II trial of temsirolimus (CCI-779) in recurrent glioblastoma multiforme: a
STAT3, telomerase activity, and growth of glioma cells by elevating intracellular reac- North Central Cancer Treatment Group Study. J Clin Oncol 23(23), 5294–5304.
tive oxygen species generation. Free Radic Biol Med 47(4), 364–374. Galanis, E., Jaeckle, K. A., Maurer, M. J., Reid, J. M., Ames, M. M., Hardwick, J. S., et al.
Drappatz, J., Brenner, A., Wong, E. T., Eichler, A., Schiff, D., Groves, M. D., et al. (2013). Phase (2009). Phase II trial of vorinostat in recurrent glioblastoma multiforme: a north
I study of GRN1005 in recurrent malignant glioma. Clin Cancer Res 19, 1567–1576. central cancer treatment group study. J Clin Oncol 27(12), 2052–2058.
Drappatz, J., Lee, E. Q., Hammond, S., Grimm, S. A., Norden, A. D., Beroukhim, R., et al. Gan, H. K., Seruga, B., & Knox, J. J. (2009). Sunitinib in solid tumors. Expert Opin Investig
(2012). Phase I study of panobinostat in combination with bevacizumab for recurrent Drugs 18(6), 821–834.
high grade glioma. J Neurooncol 107(1), 133–138. Gaspar, N., Sharp, S. Y., Pacey, S., Jones, C., Walton, M., Vassal, G., et al. (2009). Acquired
Drappatz, J., Wong, E. T., Schiff, D., Kesari, D., Batchelor, T. T., Doherty, L., et al. (2009). A resistance to 17-allylamino-17-demethoxygeldanamycin (17-AAG, tanespimycin)
pilot safety study of lenalidomide and radiotherapy for patients with newly in glioblastoma cells. Cancer Res 69(5), 1966–1975.
diagnosed glioblastoma multiforme. Int J Radiat Oncol Biol Phys 73(1), 222–227. Gilbert, M. R., Dignam, J. J., Armstrong, T. S., Wefel, J. S., Blumenthal, D. T., Vogelbaum, M.
Dresemann, G., Weller, M., Rosenthal, M. A., Wedding, U., Wagner, W., Engel, E., et al. A., et al. (2014). A randomized trial of bevacizumab for newly diagnosed glioblastoma.
(2010). Imatinib in combination with hydroxyurea versus hydroxyurea alone as N Engl J Med 370, 699–708.
oral therapy in patients with progressive pretreated glioblastoma resistant to stan- Gilbert, M. R., Gaupp, P., Liu, V., Conrad, C., Colman, H., Groves, M., et al. (2006). A phase I
dard dose temozolomide. J Neurooncol 96(3), 393–402. study of temozolomide (TMZ) and the farnesyltransferase inhibitor (FTI), lonafarnib
Du, W., Zhou, J. R., Wang, D. L., Gong, K., & Zhang, Q. J. (2012). Vitamin K1 enhances (Sarazar, SCH66336) in recurrent glioblastoma.2006 ASCO Annual Meeting
sorafenib-induced growth inhibition and apoptosis of human malignant glioma Procedings Part I. J Clin Oncol 24(18S), 1556.
cells by blocking the Raf/MEK/ERK pathway. World J Surg Oncol 10(1), 60. Gilbert, M. R., Kuhn, J., Lamborn, K. R., Lieberman, F., Wen, P. Y., Mehta, M., et al.
Ducassou, A., Uro-Coste, E., Verelle, P., Filleron, T., Benouaich-Amiel, A., Lubrano, V., et al. (2012). Cilengitide in patients with recurrent glioblastoma: the results of
(2013). αvβ3 integrin and fibroblast growth factor receptor 1 (FGFR1): prognostic NABTC 03-02, a phase II trial with measures of treatment delivery. J Neurooncol
factors in a phase I-II clinical trial associating continuous administration of tipifarnib 106(1), 147–153.
with radiotherapy for patients with newly diagnosed glioblastoma. Eur J Cancer Gilbert, M. R., Wang, M., Aldape, K., Sorensen, A., Mikkelsen, T., Bokstein, F., et al. (2010).
49(9), 2161–2169. RTOG 0635: A randomized phase II trial of bevacizumab with either irinotecan (CPT)
Efferth, T., Ramirez, T., Gebhart, E., & Haltsch, M. E. (2004). Combination treatment of glio- or dose-dense temozolomide (TMZ) in recurrent glioblastoma (GBM). Neuro Oncol
blastoma multiforme cell lines with the antimalarial artesunate and the epidermal 12(Suppl. 4), iv36–iv57.
growth factor receptor tyrosine kinase inhibitor OSI-774. Biochem Pharmacol 67, Gini, B., Zanca, C., Guo, D., Matsutani, T., Masui, K., Ikegami, S., et al. (2013). The mTOR
1689–1700. kinase inhibitors, CC214-1 and CC214-2, preferentially block the growth of EGFRvIII-
Eisenstat, D. D., Nabors, L. B., Mason, W. P., Perry, J., Shapiro, W., Kavan, P., et al. (2011). A activated glioblastomas. Clin Cancer Res 19, 5722–5732.
phase II study of daily afatinib (BIBW 2992) with or without temozolomide (21/28 days) Giordano, S. (2009). Rilotumumab, a mAb against human hepatocyte growth factor for
in the treatment of patients with recurrent glioblastoma. Chicago: ASCO (3-7 Jun). the treatment of cancer. Curr Opin Mol Ther 11(4), 288–455.
Esteller, M., Garcia-Foncillas, J., Andion, E., Goodman, S. N., Hidalgo, O. F., Vanaclocha, V., Glass, J., Hochberg, F. H., Gruber, M. L., Louis, D. N., Smith, D., & Rattner, B. (1992). The
et al. (2000). Inactivation of the DNA-repair gene MGMT and the clinical response treatment of oligodendrogliomas and mixed oligodendroglioma-astrocytomas with
of gliomas to alkylating agents. N Engl J Med 343, 1350–1354. PCV chemotherapy. J Neurosurg 75(5), 741–745.
Facchino, S., Abdouh, M., & Bernier, G. (2011). Brain cancer stem cells: Current status on Glen, H., Mason, S., Patel, H., Macleod, K., & Brunton, V. G. (2011). E7080, a multi-targeted
glioblastoma multiforme. Cancers 3, 1777–1797. tyrosine kinase inhibitor suppresses tumor cell migration and invasion. BMC Cancer
Fadul, C. E., Kingman, L. S., Meyer, L. P., Cole, B. F., Eskey, C. J., Rhodes, C. H., et al. (2008). A 11, 309.
phase II study of thalidomide and irinotecan for treatment of glioblastoma Gomez-Manzano, C., Holash, J., Fueyo, J., Xu, J., Conrad, C. A., Aldape, K. D., et al. (2008).
multiforme. J Neurooncol 90(2), 229–235. VEGF Trap induces antiglioma effect at different stages of disease. Neuro Oncol
Fakhrai, H., Mantil, J. C., Liu, L., Nicholson, G. L., Murphy-Satter, C. S., Ruppert, J., et al. 10(6), 940–945.
(2006). Phase I clinical trial of TGF-beta antisense-modified tumor cell vaccine in pa- Green, S. B., Byar, D. P., Walker, M. D., Pistenmaa, D. A., Alexander, E., Jr., Batzdorf, U., et al.
tients with advanced gliioma. Cancer Gene Ther 13(12), 1052–1060. (1983). Comparisons of carmustine, procarbazine, and high dose methylprednisolone
Fan, Q. W., Knight, Z. A., Goldenberg, D. D., Yu, W., Mostov, K. E., Stokoe, D., et al. (2006). A as additions to surgery and radiotherapy for the treatment of malignant glioma.
dual PI3 kinase/mTOR inhibitor reveals emergent efficacy in glioma. Cancer Cell 9(5), Cancer Treat Rep 67(2), 121–132.
341–349. Greenall, S. A., Donoghue, J. F., Gottardo, N. G., Johns, T. G., & Adams, T. E. (2014). Glioma-
Farrell, C. J., & Plotkin, S. R. (2007). Genetic causes of brain tumors: neurofibromato- specific Domain IV EGFR cysteine mutations promote ligand-induced covalent recep-
sis, tuberous sclerosis, von Hippel-Lindau and other syndromes. Neurol Clin 25, tor dimerization and display enhanced sensitivity to dacomitinib in vivo. Oncogene
925–946. 106 (Epub ahead of print).
Ferrara, N. (2005). VEGF as a therapeutic target in cancer. Oncology 69(Suppl. 3), 11–16. Grossman, S. A., Ye, X., Chamberlain, M., Mikkelsen, M., Batchelor, T., Desideri, S., et al.
Ferruzzi, P., Mennillo, F., De Rossa, A., Giordano, C., Rossi, M., Benedetti, G., et al. (2012). In (2009). Talampanel with standard radiation and temozolomide in patients with
vitro and in vivo characterization of a novel Hedgehog signaling antagonist in human newly diagnosed glioblastoma multiforme: a multicenter phase II trial. J Clin Oncol
glioblastoma cell lines. Int J Cancer 131(2), E33–E44. 27(25), 4155–4166.
Fields, E. C., Damek, D., Gaspar, L. E., Liu, A. K., Kavanaqh, B. D., Waziri, A., et al. (2012). Gruber, M. L., Raza, S., Gruber, D., & Narayana, A. (2009). Bevacizumab in combination
Phase I dose escalation trial of vandetanib with fractionated radiosurgery in patients with radiotherapy plus concomitant and adjuvant temozolomide for newly diag-
with recurrent malignant gliomas. Int J Radiat Oncol Biol Phys 82(1), 51–57 (Epub nosed glioblastoma: update progression-free survival, overall survival, and toxicity.
2010 Oct 29). J Clin Oncol 27, 2017 (Meeting abstracts).
Fine, H. A., Kim, L., Albert, P. S., Duic, J. P., Ma, H., Zhang, W., et al. (2007). A phase I trial of Guan, Y. S., & He, Q. (2011). Sorafenib: activity and clinical application in patients with
lenalidomide in patients with recurrent primary central nervous system tumors. Clin hepatocellular carcinoma. Expert Opin Pharmacother 12(2), 303–313.
Cancer Res 13(23), 7101–7106. Guessous, F., Zhang, Y., diPierro, C., Marcinkiewicz, L., Sarkaria, J., Schiff, D., et al. (2010).
Fisher, J. L., Schwartzbaum, J. A., Wresch, M., & Wiemels, J. L. (2007). Epidemiology of An orally available c-Mat kinase inhibitor potently inhibits brain tumor malignancy
brain tumors. Neurol Clin 25, 867–890. and growth. Anticancer Agents Med Chem 10(1), 28–35.
C. Alifieris, D.T. Trafalis / Pharmacology & Therapeutics 152 (2015) 63–82 79

Guryanova, O., Wu, Q., Cheng, L., Lathia, J. D., Huang, Z., Yang, J., et al. (2011). Non-receptor Joensuu, H., Puputti, M., Sihto, H., Tynninen, O., & Nupponen (2005). Amplification of
tyrosine kinase BMX maintains self-renewal and tumorigenic potential of glioblasto- genes encoding KIT, PDGFRα and VGFR2 receptor tyrosine kinases is frequent in glio-
ma stem cells by activating STAT3. Cancer Cell 19(4), 498–511. blastoma multiforme. J Pathol 207, 224–231.
Haas-Kogan, D. A., Prados, M. D., Lamborn, K. R., Tihan, T., Berger, M. S., & Stokoe (2005). Johns, T. G., McKay, M. J., Cvrljevic, A. N., Gan, H. K., Taylor, C., Xu, H., et al. (2012). MAb
Biomarkers to predict response to epidermal growth factor receptor inhibitors. Cell 806 enhances the efficacy of ionizing radiation in glioma xenografts expressing the
Cycle 4(10), 1369–1372. de2-7 epidermal growth factor receptor. Int J Radiat Oncol Biol Phys 78(2), 572–578.
Haas-Kogan, D. A., Prados, M. D., Tihan, T., Eberhard, D. A., Jelluma, N., Arvold, N. D., et al. Kang, T. Y., Jin, T., Elinzano, H., & Peereboom, D. (2008). Irinotecan and bevacizumab in
(2005). Epidermal growth factor receptor protein kinase B/Akt, and glioma response progressive primary brain tumors, an evaluation of efficacy and safety. J Neurooncol
to erlotinib. J Natl Cancer Inst 97(12), 880–887. 89(1), 113–118.
Hainsworth, J. D., Ervin, T., Friedman, E., Priego, V., Murphy, P. B., Clark, B. L., et al. (2010). Karavasilis, V., Kotoula, V., Pentheroudakis, G., Televantou, D., Lambaki, S., Chrisafi, S., et al.
Concurrent radiotherapy and temozolomide followed by temozolomide and sorafenib (2013). A phase I study of temozolomide and lapatinib combination in patients with
in the first-line treatment of patients with glioblastoma multiforme. Cancer 116(15), recurrent high-grade gliomas. J Neurol 260, 1469–1480.
3663–3669. Karimkhari, C., Gonzalez, R., & Dellavalle, R. P. (2014). A review of novel therapies for
Hainsworth, J. D., Shih, K. C., Shepard, G. C., Tillinghast, G. W., Brinker, B. T., & Spigel, D. R. melanoma. Am J Clin Dermatol 15, 323–337.
(2012). Phase II study of concurrent radiation therapy, temozolomide, and bevacizumab, Kast, R. E. (2010). Glioblastoma invasion, cathepsin B, and the potential for both to be
followed by bevacizumab/everolimus as first-line treatment for patients with glio- inhibited by auranofin, an old anti-rhematoid arthritis drug. Cen Eur Neurosurg
blastoma. Clin Adv Hematol Oncol 10(4), 240–246. 71(3), 139–142.
Hardell, L., Carlberg, M., Soderqvist, F., Mild, K. H., & Morgan, L. L. (2007). Long-term use of Kast, R. E., Boockvar, J. A., Bruning, A., Cappello, F., Chang, W. W., Cvek, B., et al. (2013). A
cellular phones and brain tumors: increased risk associated with use for N or = conceptually new treatment approach for relapsed glioblastoma: Coordinated
10 years. Occup Environ Med 64, 626–632. Undermining of survival paths with nine repurposed drugs (CUSP9) by the Interna-
Harris, M. (2004). Monoclonal antibodies as therapeutic agents for cancer. Lancet Oncol 5, tional Initiative for Accelerated Improvement of Glioblastoma Care. Oncotarget 4(4),
292–302. 502–503.
Hartmann, C., Hentschel, B., Wick, W., Capper, D., Felsberg, J., Simon, M., et al. (2010). Pa- Kast, R. E., & Halatsch, M. E. (2012). Matrix metalloproteinase-2 and -9 in glioblastoma: a
tients with IDH1 wild type anaplastic astrocytomas exhibit worse prognosis than trio of old drugs-captopril, disulfiram and nelfinavir-are inhibitors with potential as
IDH1-mutated glioblastomas, and IDH1 mutation status accounts for the unfavorable adjunctive treatments in glioblastoma. Arch Med Res 43(3), 243–247.
prognostic effect of higher age: implications for classification of gliomas. Acta Kitange, G. J., Mladek, A. C., Carlson, B. L., Schroeder, M. A., Pokorny, J. L., Cen, L., et al.
Neuropathol 120, 707–718. (2012). Inhibition of histone deacetylation potentiates the evolution of acquired
Hasselbalch, B., Lassen, U., Hansen, S., Homberg, M., Sørensen, M., Kosteljanetz, M., et al. temozolomide resistance linked to MGMT upregulation in glioblastoma xenografts.
(2010). Cetuximab, bevacizumab, and irinotecan for patients with primary glioblas- Clin Cancer Res 18, 4070–4079.
toma and progression after radiation therapy and temozolomide: a phase II trial. Kleihues, P., & Ohgaki, H. (1999). Primary and secondary glioblastomas: from concept to
Neuro Oncol 12(5), 508–516. clinical diagnosis. Neuro Oncol 1, 44–51.
Hau, P., Jachimczak, P., Schlaier, J., & Bogdahn, U. (2011). TGF-β2 signalling in high grade Knobbe, C. B., & Reifenberger, G. (2004). Mutation analysis of the RAS pathway genes
gliomas. Curr Pharm Biotechnol 12(12), 2150–2157. NRAS, HRAS, KRAS and BRAF in glioblastomas. Acta Neuropathol 108, 467–470.
Hegi, M. E., Diserens, A. C., Gorlia, T., Hamou, M. F., de Tribolet, N., Weller, M., et al. (2005). Koso, H., Takeda, H., Yew, C., Ward, J., Nariai, N., Ueno, K., et al. (2012). Transposon muta-
MGMT gene silencing and benefit from temozolomide in glioblastoma. N Engl J Med genesis identifies genes that transform neural stem cells into glioma-initiating cells.
352, 997–1003. Proc Natl Acad Sci U S A 109, E2998–E3007.
Held-Feindt, J., Schmeiz, S., Hattermann, K., Mentlein, R., Mehdorn, H. M., & Sebens, S. Koul, D., Fu, J., Shen, R., LaFortune, T. A., Wang, S., Tiao, N., et al. (2012). Antitumor activity
(2012). The neural adhesion molecule L1CAM confers chemoresistance in human of NVP-BKM120 – a selective pan class I PI3 kinase inhibitor showed differential
glioblastomas. Neurochem Int 61(7), 1183–1191. forms of cell death based on p53 status of glioma cells. Clin Cancer Res 18(1),
Hiddingh, L., Tannous, B. A., Teng, J., Tops, B., Jeuken, J., Hulleman, E., et al. (2014). EFEMP1 184–195.
induces γ-secretase/Notch-mediated temozolomide resistance in glioblastoma. Krashraw, M., & Lassman, A. B. (2010). Advances in the treatment of malignant gliomas.
Oncotarget 5(2), 363–374. Curr Oncol Rep 12, 26–33.
Hirai, H., Sootome, H., Nakatsuru, Y., Miyama, K., Taguchi, S., Tsujioka, K., et al. (2010). Krebs, S., Chow, K. K., Yi, Z., Rodriguez-Cruz, T., Hegde, M., Gerken, C., et al. (2014). T cells
MK-2206, an allosteric Akt inhibitor, enhances antitumor efficacy by standard che- redirected to interleukin-13Ra2 with interleukin-13-mutein-chimeric antigen recep-
motherapeutic agents ormolecular targeted drugs in vitro and in vivo. Mol Cancer tors have anti-glioma activity but also recognize interleukin-13Ra1. Cytotherapy
Ther 9, 1956–1967. 16(8), 1121–1131.
Höring, E., Podlech, O., Silkenstedt, B., Rota, I. A., Adamopoulou, E., & Naumann, U. (2013). Kreisl, T. N., Kotliarova, S., Butman, J. A., Albert, P. S., Kim, L., Musib, L., et al. (2010). A
The histone deacetylase inhibitor trichostatin A promotes apoptosis and antitumor phase I/II trial of enzastaurin in patients with recurrent high-grade gliomas. Neuro
immunity in glioblastoma cells. Anticancer Res 33, 1351–1360. Oncol 12(2), 181–189.
Hsu, J. Y., & Wakelee, H. A. (2009). Monoclonal antibodies targeting vascular endothelial Kreisl, T. N., Kim, L., Moore, K., Duic, P., Royce, C., Stroud, I., et al. (2009). Phase II trial of
growth factor: current status and future challenges in cancer therapy. BioDrugs single-agent bevacizumab plus irinotecan at tumor progression in recurrent glioblas-
23(5), 289–304. toma. J Clin Oncol 27(5), 740–745.
Hua, W., Yao, Y., Chu, Y., Zhong, P., Sheng, X., Xiao, B., et al. (2011). CD133+ tumor stem- Kreisl, T., McNeil, K., Sul, J., Iwamoto, F., Shih, J., & Fine, H. A. (2012). A phase I/II trial of
like cell-associated antigen may elicit highly intense immune responses against vandetanib for patients with recurrent malignant glioma. Neuro Oncol 14, 1519–1526.
human malignant glioma. J Neurooncol 105(2), 149–157. Lai, A., Nghiemphu, P., Green, R., Spier, L., Peak, S., Phurphanich, S., et al. (2009). Phase II
Hussain, S. F., Kong, L. Y., Jordan, J., Conrad, C., Madden, T., Fokt, I., et al. (2007). A novel trial of bevacizumab in combination with temozolomide and regional radiation ther-
small molecule inhibitor of signal transducers and activators of transcription 3 apy for up-front treatment of patients with newly diagnosed glioblastoma
reverses immune tolerance in malignant glioma patients. Cancer Res 67(20), multiforme. J Clin Oncol 27, 2000 (Meeting abstracts).
9630–9636. Lal, B., Xia, S., Abounader, R., & Laterra, J. (2005). Targeting the c-Met pathway
Huveldt, D., Lewis-Tuffin, L. J., Carlson, B. L., Schroeder, M. A., Rodriguez, F., Giannini, C., potentiates glioblastoma responses to gamma-irradiation. Clin Cancer Res
et al. (2013). Targeting Src family kinases inhibits bevacizumab-induced glioma cell 11(12), 4479–4486.
invasion. PLoS One 8(2), e56505. Langone, P., Debata, P. R., Inigo, Jdel R., Dolai, S., Mukherjee, S., Halat, P., et al. (2013).
Iwamoto, F. M., Kreisl, T. N., Kim, L., Duic, J. P., Butman, J. A., Albert, P. S., et al. (2010). Coupling to a glioblastoma-directed antibody potentiates anti-tumor activity of
Phase II trial of talampanel, a glutamate receptor inhibitor, for adults with recurrent curcumin. Int J Cancer 135, 710–719.
malignant gliomas. Cancer 117(7), 1776–1782. Lathia, J. D., Gallagher, J., Heddleston, J. M., Wang, J., Eyler, C. E., Macswords, J., et al. (2010).
Iwamoto, F. M., Lamborn, K. R., Kuhn, J. G., Wen, P. Y., Yung, W. K., Gilbert, M. R., et al. Integrin alpha 6 regulates glioblastoma stem cells. Cell Stem Cell 6(5), 421–432.
(2011). A phase I/II trial of the histone deacetylase inhibitor romidepsin for adults Lee, E. Q., Puduvalli, V. K., Reid, J. M., Kuhn, J. M., Lamborn, K. R., Cloughesy, T. F., et al.
with recurrent malignant glioma: North American Brain Tumor Consortium Study (2012). Phase I Study of Vorinostat in combination with Temozolomide in patients
03-03. Neuro Oncol 13(5), 509–516. with High-Grade Gliomas: North American Brain Tumor Consortium Study 04-03.
Iwamoto, F. M., Lamborn, K. R., Robins, H. I., Mehta, M. P., Chang, S. M., Butowski, N. 2012. Clin Cancer Res 18(21), 6032–6039.
A., et al. (2010). Phase II trial of pazopanib (GW786034), an oral multi-targeted Lehky, T. J., Iwamoto, F. M., Kreisl, T. N., Floeter, T. N., & Fine, H. A. (2011). Neuromuscular
angiogenesis inhibitor, for adults with recurrent glioblastoma (North American junction toxicity with tandutinib induces a myasthenic-like syndrome. Neurology 76,
Brain Tumor Consortium Study 06-02). Neuro Oncol 12(8), 855–861. 236–241.
Jabbour, E., & Kantarjian, H. (2012). Chronic myeloid leukemia: 2012 update on diagnosis, Leonard, A., & Wolff, J. E. (2013). Etoposide improves survival in high-grade glioma: a
monitoring, and management. Am J Hematol 87, 1037–1045. meta-analysis. Anticancer Res 33(8), 3307–3315.
Jackson, C., Ruzevic, J., Brem, H., & Lim, M. (2013). Vaccine strategies for glioblastoma: Lesniak, M. S. (2011). Immunotherapy for glioblastoma: the devil is in the details. J Clin
progress and future directions. Immunotherapy 5, 155–167. Oncol 29(22), 3105.
Jain, R. K., di Tomaso, E., Duda, D. G., Loeffler, J. S., Sorensen, A. G., & Batchelor, T. T. (2007). Li, Y., Guessous, F., DiPierro, C., Zhang, Y., Mudrick, T., Fuller, L., et al. (2009). Interactions
Angiogenesis in brain tumours. Nat Rev Neurosci 8(8), 610–622. between PTEN and the c-Met pathway in glioblastoma and implications for therapy.
Jiang, H., Gomez-Manzano, C., Aoki, H., Alonso, M. M., Kondo, S., McCormick, F., et al. Mol Cancer Ther 8(2), 376–385.
(2007). Examination of the therapeutic potential of Delta-24-RGD in brain tumor Liang, Q. L., Mo, Z. Y., Wang, P., Li, X., Liu, Z. X., & Zhou, Z. M. (2014). The clinical value of
cells: role of autophagic cell death. J Natl Cancer Inst 99, 1410–1414. serum hepatocyte growth factor levels in patients undergoing primary radiotherapy
Jo, M. Y., Kim, Y. G., Kim, Y., Lee, S. J., Kim, M. H., Joo, K. M., et al. (2012). Combined therapy for glioma: effect on progression-free survival. Med Oncol 31(9), 122.
of temozolomide and ZD6474 (vandetanib) effectively reduces glioblastoma tumor Lin, J., Zhang, X. M., Yang, J. C., Ye, Y. B., & Luo, S. Q. (2010). γ-secretase inhibitor-I
volume through anti-angiogenic and anti-proliferative mechanisms. Mol Med Rep enhances radiosensitivity of glioblastoma cell lines by depleting CD133+ tumor
6(1), 88–92. cells. Arch Med Res 41(7), 519–529.
80 C. Alifieris, D.T. Trafalis / Pharmacology & Therapeutics 152 (2015) 63–82

Linnet, K., & Ejsing, T. (2008). A review on the impact of P-glycoprotein on the penetration Nair, S. K., Sampson, J. H., & Mitchell, D. A. (2014). Immunological targeting of cytomega-
of drugs into the brain. Focus on psychotropic drugs. Eur Neuropsychopharmacol 18, lovirus for glioblastoma therapy. OncoImmunology 25(3), e29289.
157–169. Nduom, E., Hadjipanayis, C., & Meir, E. (2012). Glioblastoma cancer stem-like cells –
Liu, G., Akasaki, Y., Khong, H. T., Wheeler, C. J., Das, A., Black, K. L., et al. (2005). Cytotoxic T implications for pathogenesis and treatment. Cancer J 18(1), 100–106.
cell targeting of TRP-2 sensitizes human malignant glioma to chemotherapy. Neyns, B., Sadones, J., Chaskis, C., Dujardin, M., Everaert, H., Lv, S., et al. (2011). Phase II
Oncogene 24(33), 5226–5234. study of sunitinib malate in patients with recurrent high-grade glioma. J Neurooncol
Liu, C., Sage, J., Miller, M., Verhaak, R., Hippenmeyer, S., Vogel, H., et al. (2011). Mosaic 103(3), 491–501.
analysis with double markers reveals tumor cell of origin in glioma. Cell 146, Neyns, B., Sadones, J., Joosens, E., Bouttens, F., Verbeke, L., Baurain, J. F., et al. (2009). Strat-
209–221. ified phase II trial of cetuximab in patients with recurrent high-grade glioma. Ann
Longo, S. L., Padalino, D. J., McGillis, S., Petersen, K., Schirok, H., Politz, O., et al. (2012). Oncol 20(9), 1596–1603.
Bay846, a new irreversible small molecule inhibitor of EGFR and Her2, is highly effec- Nghiempu, P. L., Wen, P. Y., Lamborn, K. R., Drappatz, J., Robins, H. I., Fink, K., et al. (2011).
tive against malignant glioma brain tumor models. Invest New Drugs 30(6), 2161–2172. A phase I trial of tipifarnib with radiation therapy, with and without temozolomide,
Louis, D. N. (2006). Molecular pathology of malignant gliomas. Annu Rev Pathol 1, 97. for patients with newly diagnosed glioblastoma. Int J Radiat Oncol Biol Phys 81(5),
Louis, D. N., Ohgaki, H., Wiestler, O. D., Cavenee, W. K., Burger, P. C., Jouvet, A., et al. 1422–1427.
(2007). The 2007 WHO classification of tumours of the central nervous system. Ohgaki, H., & Kleihues, P. (2013). The definition of primary and secondary glioblastoma.
Acta Neuropathol 114(2), 97–109. Clin Cancer Res 19(4), 764–772.
Lu-Emerson, C., Norden, A. D., Drappatz, J., Quant, E. C., Beroukhim, R., Ciampa, A. S., et al. Okada, H., Lieberman, F. S., Walter, K. A., Lunsford, L. D., Kondziolka, D. S., Bejjani, G. K.,
(2011). Retrospective study of dasatinib for recurrent glioblastoma after bevacizumab et al. (2007). Autologous glioma cell vaccine admixed with interleukin-4 gene
failure. J Neurooncol 104(1), 287–291. transfected fibroblasts in the treatment of patients with malignant gliomas. J Transl
Lun, X., Alain, T., Zemp, F. J., Zhou, H., Rahman, M. M., Hamilton, M. G., et al. (2012). Myx- Med 5, 67.
oma virus virotherapy for glioma in immunocompetent animal models: optimizing Onishi, M., Kurozumi, K., Ichikawa, T., & Date, I. (2013). Mechanisms of tumor develop-
administration routes and synergy with rapamycin. Cancer Res 70, 598–608. ment and anti-angiogenic therapy in glioblastoma multiforme. Neurol Med Chir
Lustig, R., Mikkelsen, T., Lesser, G., Grossman, S., Ye, X., Desideri, S., et al. (2008). Phase II (Tokyo) 53, 755–763.
preradiation R115777 in newly diagnosed GBM with residual enhancing disease. Ostman, A. (2004). PDGF receptors-mediators of autocrine tumor growth and regulators
Neuro Oncol 10(6), 1004–1009. of tumor vasculature and stroma. Cytokine Growth Factor Rev 15(4), 4899–4907.
Lv, S., Teugels, E., Sadones, J., De Brakeleer, S., Duerinck, J., Du Four, S., et al. (2012). Cor- Ostrom, Q. T., Gittleman, H., Farah, H., Ondracek, A., Chen, Y., Wolinsky, Y., et al. (2013).
relation of EGFR, IDH1 and PTEN status with the outcome of patients with recurrent CBTRUS statistical report: Primary brain and central nervous system tumors diag-
glioblastoma treated in a phase II clinical trial with the EGFR-blocking monoclonal nosed in the United States in 2006–2010. Neuro Oncol 15(Suppl 2), ii1–ii56.
antibody cetuximab. Int J Oncol 41(3), 1029–1035. Pan, E., Yu, D., Yue, B., Potthast, L., Chowdhary, S., Smith, P., et al. (2012). A prospective
MackDonald, T. J., Aguilera, D., & Kramm, C. M. (2011). Treatment of high-grade glioma in phase II single-institution trial of sunitinib for recurrent malignant glioma. J
children and adolescents. Neuro Oncol 13(10), 1049–1058. Neurooncol 110(1), 111–118.
Majumdar, K., Radotra, B. D., Vasishta, R. K., & Pathak, A. (2009). Platelet-derived growth Peereboom, D. M., Ahluwalia, M. S., Ye, X., Supko, J. G., Hildebrand, S. L., Phupbanich, S.,
factor expression correlates with tumor grade and proliferative activity in human et al. (2013). NABTT 0502: a phase II and pharmacokinetic study of erlotinib and
oligodendrogliomas. Surg Neurol 72(1), 54–60. sorafenib for patients with progressive or recurrent glioblastoma multiforme. Neuro
Mallon, R., Feldberg, L. R., Lucas, J., Chaudhary, I., Dehnhardt, C., Santos, E. D., et al. (2011). Oncol 15, 490–496.
Antitumor efficacy of PKI-587, a highly potent dual PI3K/mTOR kinase inhibitor. Clin Pelloski, C. E., Lin, E., Zhang, L., Yung, W. K., Colman, H., Liu, J. L., et al. (2006). Prognostic
Cancer Res 17(10), 3193–3203. associations of activated mitogen-activated protein kinase and Akt pathways in
Mani, A., & Gelmann, E. P. (2005). The ubiquitin-proteasome pathway and its role in can- glioblastoma. Clin Cancer Res 12, 3935–3941.
cer. J Clin Oncol 23(21), 4776–4789. Penuelas, S., Anido, J., Prieto-Sanchez, R. M., Folch, G., Barba, I., Cuartas, I., et al. (2009).
Manmeet, A. S., de Groot, J., Wei, L. M., & Gladson, C. L. (2010). Targeting SRC in glioblas- TGF-beta increases glioma-initiating cell self renewal through the induction of LIF
toma tumors and brain metastases: Rationale and preclinical studies. Cancer Lett in human glioblastoma. Cancer Cell 15(4), 315–327.
298(2), 139–149. Phurphanich, S., Supko, J. G., Carson, K. A., Grossman, S. A., Burt, Nabors L., Mikkelsen, T.,
Martens, T., Schmidt, N. O., Eckerich, C., Fillbrandt, R., Merchant, M., Schwall, R., et al. et al. (2010). Phase 1 clinical trial of bortezomib in adults with recurrent malignant
(2006). A novel one-armed anti-c-Met antibody inhibits glioblastoma growth glioma. J Neurooncol 100(1), 95–103.
in vivo. Clin Cancer Res 12(20 Pt 1), 6144–6152. Pines, G., Huang, P. H., Zwang, Y., White, F. M., & Yarden, Y. (2010). EGFRvIV: a previously
Mason, W. P., Belanger, K., Nicholas, G., Vallières, I., Mathieu, D., Kavan, P., et al. (2012). A uncharacterized oncogenic mutant reveals a kinase autoinhibitory mechanism.
phase II study of the Ras-MAPK signaling pathway inhibitor TLN-4601 in patients Oncogene 29(43), 5850–5860.
with glioblastoma at first progression. J Neurooncol 107(2), 343–349. Pitter, K. L., Galban, C. J., Galban, S., Tehrani, O. S., Li, F., Charles, N., et al. (2011). Perifosine
McCarty, M. F. (1997). Thalidomide may impede cell migration in primates by down- and CCI 779 co-operate to induce cell death and decrease proliferation in PTEN-intact
regulating integrin beta chains: potential therapeutic utility in solid malignancies, and PTEN-deficient PDGF-driven murine glioblastoma. PLoS One 6(1), e14545.
proliferative retinopathy, inflammatory disorders, neointimal hyperplasia, and osteo- Potts, C. B., Albitar, X. M., Anderson, C. K., Baritaki, S., Berkers, C., Bonavida, B., et al. (2011).
porosis. Med Hypotheses 49(2), 123–131. Marizomib, a proteasome inhibitor for all seasons: preclinical profile and a frame-
Meiss, F., & Zeiser, R. (2014). Vismodegib. Recent Results Cancer Res 201, 405–417. work for clinical trials. Curr Cancer Drug Targets 11(3), 254–284.
Mellinhoff, I. K., Wang, M. Y., Vivanco, I., Haas-Kogan, D. A., Zhu, S., Dia, E. Q., et al. (2005). Prasad, G., Sottero, T., Yang, X., Mueller, S., James, C. D., Weiss, W. A., et al. (2011). Inhibi-
Molecular determinants of the response of glioblastomas to EGFR kinase inhibitors. N tion of PI3K/mTOR pathways in glioblastoma and implications in combination thera-
Engl J Med 353(19), 2012–2024. py with temozolomide. Neuro Oncol 13(4), 384–392.
Mendelsohn, J., & Baselga, J. (2006). Epidermal growth factor receptor targeting in cancer. Quayle, S. N., Lee, J. Y., Cheung, L. W., Ding, L., Wiedemeyer, R., Dewan, R. W., et al. (2012).
Semin Oncol 33(4), 369–385. Somatic mutations of PIK3R1 promote gliomagenesis. PLoS One 7, e49466.
Mendiburu-Eliçabe, M., Gil-Ranedo, J., & Izquierdo, M. (2014). Efficacy of rapamycin Quinn, J. A., Jiang, S. X., Reardon, D. A., Desjardins, A., Vredenburgh, J. J., Rich, J. N., et al.
against glioblastoma cancer stem cells. Clin Transl Oncol 16(5), 495–502. (2009). Phase II trial of temozolomide plus O6-benzylguanine in adults with recur-
Michelakis, E. D., Sutendra, G., Dromparis, P., Webster, L., Haromy, A., Niven, E., et al. rent, temozolomide-resistant malignant glioma. J Clin Oncol 27(8), 1262–1267.
(2010). Metabolic modulation of glioblastoma with dichloroacetate. Sci Transl Med Raizer, J. J., Abrey, L. E., Lassman, A. B., Chang, S. M., Lamborn, K. R., Kuhn, J. G., et al.
2(31), 31ra34. (2009). A phase II trial of erlotinib in patients with recurrent malignant gliomas
Mitchell, D. A., & Sampson, J. H. (2009). Toward effective immunotherapy for the treat- and nonprogressive glioblastoma multiforme postradiation therapy. Neuro Oncol
ment of malignant brain tumors. Neurotherapeutics 6(3), 527–538. 22(1), 133–142.
Møller, H. G., Rasmussen, A., Andersen, H. H., Johnsen, K. B., Henriksen, M., & Duroux, M. Rassi, F. E., & Khoury, H. J. (2013). Bosutinib: a SRC-ABL tyrosine kinase inhibitor for treat-
(2013). A systematic review of MicroRNA in glioblastoma multiforme: Micro- ment of chronic myeloid leukemia. Pharmacogenomics and Pers Med 6, 57–62.
modulators in the mesenchymal mode of migration and invasion. Mol Neurobiol 47, Raymond, E., Brandes, A. A., Dittrich, C., Fumoleau, P., Coudert, B., Clement, P. M., et al.
131–144. (2008). Phase II study of imatinib in patients with recurrent gliomas of various histol-
Momota, H., Nerio, E., & Holland, E. C. (2005). Perifosine inhibits multiple signaling path- ogies: a European Organisation for Research and Treatment of Cancer Brain Tumor
ways in glial progenitors and cooperates with temozolomide to arrest cell prolifera- Group Study. J Clin Oncol 26(28), 4659–4665.
tion in gliomas in vivo. Cancer Res 65(16), 7429–7435. Razis, E., Selviaridis, P., Labropoulos, S., Norris, J. L., Zhu, M. J., Song, D. D., et al. (2009).
Muñoz, M., & Coveñas, R. (2012). NK-1 receptor antagonists: a new generation of antican- Phase II study of neoadjuvant imatinib in glioblastoma: evaluation of clinical and
cer drugs. Mini Rev Chem 12, 593–599. molecular effects of the treatment. Clin Cancer Res 15(19), 6258–6266.
Munoz, J. L., Rodriguez-Cruz, V., Greco, S. J., Nagula, V., Scotto, K. W., & Rameshwar, P. Reardon, D. A., Desjardins, A., Peters, K., Gururangan, S., Sampson, J., Rich, J. N., et al.
(2014). Temozolomide induces the production of epidermal growth factor to (2011). Phase II study of metronomic chemotherapy with bevacizumab for recurrent
regulate MDR1 expression in glioblastoma cells. Mol Cancer Ther 13(10), glioblastoma after progression on bevacizumab therapy. J Neurooncol 103(2),
2399–2411. 371–379.
Nabors, L. B., Mikkelsen, T., Hegi, M. E., Ye, X., Batchelor, T., Lesser, G., et al. (2012). A safety Reardon, D. A., Desjardins, A., Vredenburgh, J. J., Gururangan, S., Friedman, A. H., Herndon,
run-in and randomized phase 2 study of cilengitide combined with chemoradiation J. E., II, et al. (2010). Phase 2 trial of erlotinib plus sirolimus in adults with recurrent
for newly diagnosed glioblastoma (NABTT 0306). Cancer 118, 5601–5607. glioblastoma. J Neurooncol 96(2), 219–230.
Nagane, M., Nishikawa, R., Narita, Y., Kobayashi, H., Takano, S., Shinoura, N., et al. (2012). Reardon, D. A., Dresemann, G., Taillibert, S., Campone, M., van den Bent, M., Clement, P.,
Phase II study of single-agent bevacizumab in Japanese patients with recurrent et al. (2009). Multicentre phase II studies evaluating imatinib plus hydroxyurea in
malignant glioma. Jpn J Clin Oncol 42(10), 887–895. patients with progressive glioblastoma. Br J Cancer 101(12), 119–2004.
Nagpal, S. (2012). The role of BCNU polymer wafers (Gliadel) in the treatment of Reardon, D. A., Egorin, M. J., Desjardins, A., Vredenburgh, J. J., Beumer, J. H., Lagattuta, T.,
malignant glioma. Neurosurg Clin N Am 23(2), 289–295. et al. (2009). Phase I pharmacokinetic study of the vascular endothelial growth factor
C. Alifieris, D.T. Trafalis / Pharmacology & Therapeutics 152 (2015) 63–82 81

receptor tyrosine kinase inhibitor vatalanib (PTK787) plus imatinib and hydroxyurea Schwartzbaum, J. A., Fisher, J. L., Aldape, K. D., & Wrensch, M. (2006). Epidemiology and
for malignant glioma. Cancer 115(10), 2188–2198. molecular pathology of glioma. Nat Clin Pract Neurol 2(9), 494–503.
Reardon, D. A., Egorin, M. J., Quinn, J. A., Rich, J. N., Gururangan, S., Vredenburgh, J. J., et al. Scott, C. B., Scarantino, C., Urtasun, R., Movsas, B., Jones, C. U., Simpson, J. R., et al. (1998).
(2005). Phase II study of imatinib mesylate plus hydroxyurea in adults with recurrent Validation and predictive power of Radiation Therapy Oncology Group (RTOG) recur-
glioblastoma multiforme. J Clin Oncol 23(36), 9359–9368. sive partitioning analysis classes for malignant glioma patients: a report using RTOG
Reardon, D. A., Fink, K. L., Mikkelsen, T., Cloughesy, T. F., O’Neill, A., Plotkin, S., et al. (2008). 90-96. Int J Radiat Oncol Biol Phys 40, 51–55.
Randomized phase II study of cilengitide, an integrin-targeting arginine-glycine- Sebti, S. M., & Adjei, A. A. (2004). Farnesyltransferase inhibitors. Semin Oncol 31(1), 28–39.
aspartic acid peptide, in patients with recurrent glioblastoma multiforme. J Clin See, A. P., Pradilla, G., Yang, I., Han, S., Parsa, A. T., & Lim, M. (2011). Heat shock
Oncol 26(34), 5610–5617. protein-peptide complex in the treatment of glioblastoma. Expert Rev Vaccines
Reardon, D. A., Groves, M. D., Wen, P. Y., Nabors, L., Mikkelsen, T., Rosenfeld, S., et al. 10, 721–731.
(2013). A phase I/II trial of pazopanib in combination with lapatinib in adult patients Shah, G. D., Loizos, N., Youssoufian, H., Schwartz, J. D., & Rowinsky, E. K. (2010). Rationale
with relapsed malignant glioma. Clin Cancer Res 19, 900–908. for the development of IMC-3G3, a fully human immunoglobulin subclass 1 monoclo-
Reardon, A. A., Nabors, B. L., Stupp, R., & Mikkelsen, T. (2008). Cilengitide: an integrin- nal antibody targeting the platelet-derived growth factor receptor alpha. Cancer
targeting arginine-glycine-aspartic acid peptide with promising activity for glioblas- 116(Suppl. 4), 1018–1026.
toma multiforme. Expert Opin Investig Drugs 17(8), 1225–1235. Sharma, V., Shaheen, S. S., Dixit, D., & Sen, E. (2012). Farnesyltransferase inhibitor
Reardon, D. A., Quinn, J. A., Vredenburgh, J. J., Gururangan, S., Friedman, A. H., Desjardins, manumycin targets IL1β-Ras-HIF-1α axis in tumor cells of diverse origin.
A., et al. (2006). Phase 1 trial of gefitinib plus sirolimus in adults with recurrent Inflammation 35, 516–519.
malignant glioma. Clin Cancer Res 12(3), 860–868. Shen, J., Zheng, H., Ruan, J., Fang, W., Li, A., Tian, G., et al. (2013). Autophagy inhibition in-
Reardon, D. A., Vredenburgh, J. J., Coan, A., Desjardins, A., Peters, K. B., Gururangan, S., et al. duces enhanced proapoptotic effects of ZD6474 in glioblastoma. Br J Cancer 109,
(2011). Phase I study of sunitinib and irinotecan for patients with recurrent malig- 164–171.
nant glioma. J Neurooncol 105(3), 621–627. Sheppard, K., Kinross, K. M., Solomon, B., Pearson, R. B., & Phillips, W. A. (2012). Targeting
Reardon, D. A., Vredenburgh, J. J., Desjardins, A., Peters, K., Gururangan, S., Sampson, J. H., PI3 kinase/AKT/mTOR signaling in cancer. Crit Rev Oncog 17(1), 69–95.
et al. (2011). Effect of CYP3A-inducing anti-epileptics on sorafenib exposure: results Shinojima, N., Tada, K., Shiraishi, S., Kamiryo, T., Kochi, M., Nakamura, H., et al. (2003).
of a phase II study of sorafenib plus daily temozolomide in adults with recurrent Prognostic value of epidermal growth factor receptor in patients with glioblastoma
glioblastoma. J Neurooncol 101(1), 57–66. multiforme. Cancer Res 63(20), 6962–6970.
Reardon, D. A., Vredenburgh, J. J., Desjardins, A., Peters, K. B., Sathornsumetee, S., Threatt, Sierra, J. R., & Tsao, M. S. (2011). c-MET as a potential therapeutic target and biomarker in
S., et al. (2012). Phase 1 trial of dasatinib plus erlotinib in adults with recurrent cancer. Ther Adv Med Oncol 3(Suppl. 1), s21–s35.
malignant glioma. J Neurooncol 108(3), 499–506. Solomón, M., Selva, J., Figueredo, J., Vaquer, J., Toledo, C., Quintanal, N., et al. (2013). Radio-
Reardon, D. A., Wen, P. Y., Alfred Yung, W. K., Berk, L., Narasimhan, N., Turner, C. D., et al. therapy plus nimotuzumab or placebo in the treatment of high grade glioma
(2012). Ridaforolimus for patients with progressive or recurrent malignant glioma: a patients: results from a randomized, double blind trial. BMC Cancer 13, 299.
perisurgical, sequential, ascending-dose trial. Cancer Chemother Pharmacol 69(4), Spence, A. M., Peterson, R. A., Schanhorst, J. D., Silbergeld, D. L., & Rostomily, R. C. (2004).
849–860. Phase II study of concurrent continuous Temozolomide (TMZ) and Tamoxifen (TMX)
Ribot, E., Bouzier-Sore, A. K., Bouchaud, V., Miraux, S., Delville, M. H., Franconi, J. M., et al. for recurrent malignant astrocytic gliomas. J Neurooncol 70, 91–95.
(2007). Microglia used as vehicles for both inducible thymidine kinase gene therapy Stockhausen, M. T., Kristoffersen, K., & Poulsen, H. S. (2012). Notch signaling and brain
and MRI contrast agents for glioma therapy. Cancer Gene Ther 14, 724–737. tumors. Adv Exp Med Biol 727, 289–304.
Rich, J. N., Reardon, D. A., Peery, T., Dowell, J. M., Quinn, J. A., Penne, K. L., et al. (2004). Stupp, R., Hegi, M. E., Mason, W. P., van den Bent, M. J., Taphoorn, M. J., Janzer, R. C., et al.
Phase II trial of gefitinib in recurrent glioblastoma. J Clin Oncol 22(1), 133–142. (2009). Effects of radiotherapy with concomitant and adjuvant temozolomide versus
Rich, J. N., Reardon, D. A., Quinn, J. A., Vredenburgh, J. J., Desjardins, A., Sathornsumetee, S., radiotherapy alone on survival in glioblastoma in a randomized phase III study:
et al. (2005). A phase I trial of gefitinib (ZD1 839) plus rapamycin for patients with 5-year analysis of the EORTC-NCIC trial. Lancet Oncol 10, 459–466.
recurrent malignant glioma. J Clin Oncol 23 (130-S-S). Stupp, R., Hegi, M. E., Neyns, B., Goldbrunner, R., Schlegel, U., Clement, P. M., et al. (2010).
Robins, H. I., Won, M., Seiferheld, W. F., Schultz, C. J., Choucair, A. K., Brachman, D. G., et al. Phase I/IIa study of cilengitide and temozolomide with concomitant radiotherapy
(2006). Phase 2 trial of radiation plus high-dose tamoxifen for glioblastoma followed by cilengitide and temozolomide maintenance therapy in patients with
multiforme: RTOG protocol BR-0021. Neuro Oncol 8(1), 47–52. newly diagnosed glioblastoma. J Clin Oncol 28(16), 2712–2718.
Rosenfeld, M. R., Chamberlain, M. C., Grossman, S. A., Peereboom, D. M., Lesser, G. J., Stupp, R., Mason, W. P., van den Bent, M. J., Weller, M., Fisher, B., Taphoorn, M. J., et al.
Batchelor, T. T., et al. (2010). A multi-institution phase II study of poly-ICLC and radio- (2005). Radiotherapy plus concomitant and adjuvant temozolomide for glioblastoma.
therapy with concurrent and adjuvant temozolomide in adults with newly diagnosed N Engl J Med 352, 987–996.
glioblastoma. Neuro Oncol 12(10), 1071–1077. Stupp, R., Tonn, J. C., Brada, M., & Pentheroudakis, G. (2010). High-grade malignant
Sai, K., Wang, S., Balasubramaniyan, V., Conrad, C., Lang, F. F., Aldape, K., et al. (2012). In- glioma: ESMO clinical practice guidelines for diagnosis, treatment and follow up.
duction of cell-cycle arrest and apoptosis in glioblastoma stem-like cells by WP1193, Ann Oncol 5(21), 190–193.
a novel small molecule inhibitor of the JAK2/STAT3 pathway. J Neurooncol 107(3), Tabatabai, G., Weller, M., Nabors, B., Picard, M., Reardon, D., Mikkelsen, T., et al. (2010).
487–501. Targeting integrins in malignant glioma. Target Oncol 5, 175–181.
Sampson, J. H., Heimberger, A. B., Archer, G. E., Aldape, K. D., Friedman, A. H., Gilbert, M. R., Tamura, N., Hirano, K., Kishino, K., Hashimoto, K., Amano, O., Shimada, J., et al. (2012).
et al. (2010). Immonulogic escape after prolonged progression-free survival with Analysis of type of cell death induced by topoisomerase inhibitor SN-38 in human
epidermal growth factor receptor variant III peptide vaccination in patients with oral squamous cell carcinoma lines. Anticancer Res 32(11), 4823–4832.
newly diagnosed glioblastoma. J Clin Oncol 28(31), 422–429. Terasaki, M., Shibui, S., Narita, Y., Fujimaki, T., Aoki, T., Kajiwara, K., et al. (2011). Phase I
Sarangi, A., Valadez, J. G., Rush, S., Abel, T. W., Thompson, R. C., Cooper, M. K., et al. (2009). trial of a personalized peptide vaccine for patients positive for human leukocyte an-
Targeted inhibition of the Hedgehog pathway in established malignant glioma cells tigen – A24 with recurrent or progressive glioblastoma multiforme. J Clin Oncol 29(3),
enhances survival. Oncogene 28(39), 3468–3476. 337–344.
Sarkaria, J. N., Galanis, E., Wu, W., Dietz, A. B., Kaufmann, T. J., Gustafson, M. P., et al. Thiesen, B., Stewart, C., Tsao, M., Kamel-Reid, S., Schaiquevich, P., Mason, W., et al. (2010).
(2010). Combination of temsirolimus (CCI-779) with chemoradiation in newly A phase I/II trial of GW572016 (lapatinib) in recurrent glioblastoma multiforme: clin-
diagnosed glioblastoma multiforme (GBM) (NCCTG trial N027D) is associated with ical outcomes, pharmacokinetics and molecular correlation. Cancer Chemother
increased infectious risks. Clin Cancer Res 16(22), 5573–5580. Pharmacol 65(2), 353–361.
Sarkaria, J. N., Galanis, E., Wu, W., Peller, P. J., Giannini, C., Brown, P. D., et al. (2011). Tsuzuki, T., Izumoto, S., Ohnisi, T., Hiraga, S., Arita, N., & Hayakawa, T. (1998). Neural cell
North Central Cancer Treatment group Phase I trial N057K of everolimus adhesion molecule L1 in gliomas: correlation with TGF-beta and p53. J Clin Pathol
(RAD001) and temozolomide in combination with radiation therapy in patients 51(1), 13–17.
with newly diagnosed glioblastoma multiforme. Int J Radiat Oncol Biol Phys Tzadok, S., Beery, E., Israeli, M., Uziel, O., Lahav, M., Fenig, E., et al. (2010). In vitro novel
81(2), 468–475. combinations of psychotropics and anti-cancer modalities in U87 human glioblasto-
Sasine, J. P., Savaraj, N., & Feun, L. G. (2010). Topoisomerase I inhibitors in the treatment of ma cells. Int J Oncol 37, 1043–1051.
primary CNS malignancies: an update on recent trends. Anticancer Agents Med Chem van den Bent, M. J., Brandes, A. A., Rampling, R., Kouwenhoven, M. C., Kros, J. M.,
10(9), 683–696. Carpentier, A. F., et al. (2009). Randomized trial of erlotinib versus temozolomide
Sathornsumetee, S., Desjardins, A., Vredenburgh, J. J., McLendon, R. E., Marcello, J., or carmustine in recurrent glioblastoma: EORTC brain tumor group study 26034. J
Herndon, J. E., et al. (2010). Phase II trial of bevacizumab and erlotinib in patients Clin Oncol 27(8), 1268–1274.
with recurrent malignant glioma. Neuro Oncol 12(12), 1300–1310. Verhaak, R. G., Hoadley, K. A., Purdom, E., Wang, V., Qi, Y., Wilkerson, M. D., et al. (2010).
Sathornsumetee, S., Reardon, D. A., Desjardins, A., Quinn, J. A., Vredenburgh, J. J., & Rich, J. Intergrated genomic analysis identifies clinically relevant subtypes of glioblastoma
N. (2007). Molecularly targeted therapy for malignant glioma. Cancer 110, 13–24. characterized by abnormalities in PDGFRA, IDH1, EGFR and NF1. Cancer Cell 17(1),
Sauvageot, C. M., Weatherbee, J. L., Kesari, S., Winters, S. E., Barnes, J., Dellagatta, J., et al. 98–1100.
(2009). Efficacy of the HSP90 inhibitor 17-AAG in human glioma cell lines and tu- Villano, J. L., Seery, T. E., & Bressler, L. R. (2009). Temozolomide in malignant gliomas:
morigenic glioma stem cells. Neuro Oncol 11(2), 109–121. current use and future targets. Cancer Chemother Pharmacol 64(4), 647–655.
Sawada, M., Imai, F., Suzuki, H., Hayakawa, M., Kanno, T., & Nagatsu, T. (1998). Brain Vlachostergios, P. J., Hatzidaki, E., Befani, C. D., Liakos, P., & Papandreou, C. N. (2013).
specific gene expression by immortalized microglial cell-mediated gene transfer in Bortezomib overcomes MGMT-related resistance of glioblastoma cell lines to temo-
the mammalian brain. FEBS Lett 433, 37–40. zolomide in a schedule-dependent manner. Invest New Drugs 31, 1169–1181.
Schmidt, F., Fischer, J., Herrlinger, U., Dietz, K., Dichgans, J., & Weller, M. (2006). PCV che- Wang, T., Agarwal, S., & Elmquist, W. F. (2012). Brain distribution of cediranib is limited
motherapy for recurrent glioblastoma. Neurology 66, 587–589. by active efflux at the blood-brain barrier. J Pharmacol Exp Ther 341(2), 386–395.
Schmidt, R. S., & Maness, P. F. (2008). L1 and CAM adhesion molecules as signaling Wang, K., Pan, L., Che, X., Cui, D., & Li, C. (2010). Sonic Hedgehog/GLI1 signalling pathway
coreceptors in neuronal migration and process outgrowth. Curr Opin Neurobiol inhibition restricts cell migration and invasion in human gliomas. Neurol Res 32(9),
18(3), 245–250. 975–980.
82 C. Alifieris, D.T. Trafalis / Pharmacology & Therapeutics 152 (2015) 63–82

Wang, Y., Pan, L., Sheng, X. F., Chen, S., & Dai, J. Z. (2014). Nimotuzumab, a humanized Yakes, F. M., Chen, J., Tan, J., Yamaguchi, K., Shi, Y., Yu, P., et al. (2011). Cabozantinib
monoclonal antibody specific for the EGFR, in combination with temozolomide and (XL184), a novel MET and VEGFR2 inhibitor, simultaneously suppresses metastasis,
radiation therapy for newly diagnosed glioblastoma multiforme: First results in angiogenesis, and tumor growth. Mol Cancer Ther 10(12), 2298–2308.
Chinese patients. Asia Pac J Clin Oncol, http://dx.doi.org/10.1111/ajco.12166 (Epub Yarden, Y., & Sliwkowski, M. X. (2001). Untangling the ErbB signaling network. Nat Rev
ahead of print). Mol Cell Biol 2(2), 127–137.
Welsh, J. W., Mahadevan, D., Ellsworth, R., Cooke, L., Bearss, D., & Stea, B. (2009). The Ye, F., Gao, Q., & Cai, M. J. (2012). Therapeutic targeting of EGFR in malignant gliomas.
c-Met receptor tyrosine kinase inhibitor MP470, radiosensitizes glioblastoma cells. Expert Opin Ther Targets 14(3), 303–316.
Radiat Oncol 4, 69. Yelskaya, Z., Carrillo, V., Dubisz, E., Gulzar, H., Morgan, D., & Mahajan, S. S. (2013). Syner-
Wen, P. Y. (2010). American Society of Clinical Oncology 2010: report of selected studies gistic inhibition of survival, proliferation, and migration of U87 cells with a combina-
from the CNS tumor section. Expert Rev Anticancer Ther 10(9), 1367–1369. tion of LY341495 and Iressa. PLoS One 8(5), e64588.
Wen, P. Y., Cloughesy, T., Kuhn, J., Lamborn, K., Abrey, L. E., Lieberman, F., et al. (2009). Yiin, J. J., Hu, B., Schornack, P. A., Sengar, R. S., Liu, K. W., Feng, H., et al. (2010). ZD6474, a
Phase I/II study of sorafenib and temsirolimus for patients with recurrent glioblasto- multitargeted inhibitor for receptor tyrosine kinases, suppresses growth of gliomas
ma (GBM) (NABTC 05-02). ASCO Meet Abstr 27(15S), 2006. expressing an epidermal growth factor receptor mutant, EGFRvIII, in the brain. Mol
Wen, P. Y., & Kesari, S. (2008). Malignant gliomas in adults. N Engl J Med 359, 492–507. Cancer Ther 9(4), 929–941.
Wen, P., Schiff, D., Cloughesy, T., Raizer, J. J., Laterra, J., Smitt, M., et al. (2011). A phase II Yin, H., Zhou, Y., Wen, C., Zhou, C., Zhang, W., Hu, X., et al. (2014). Curcumin sensitizes
study evaluating the efficacy and safety of AMG 102 (rilotumumab) in patients glioblastoma to temozolomide by simultaneously generating ROS and disrupting
with recurrent glioblastoma. Neuro Oncol 13, 437–446. AKT/mTOR signallling. Oncol Rep 32(4), 1610–1616.
Wen, P. Y., Schiff, D., Kesari, S., Drappatz, J., Gigas, D., & Doherty, L. (2006). Medical man- Yust-Katz, S., Liu, D., Yuan, Y., Liu, V., Kang, S., Groves, M., et al. (2013). Phase 1/1b study of
agement of patients with brain tumors. J Neurooncol 80, 131–132. lonafarnib and temozolomide in patients with recurrent or TMZ refractory glioblasto-
Wen, P. Y., Yung, W. K., Lamborn, K. R., Dahia, P. L., Wang, Y., Peng, B., et al. (2006). Phase ma. Cancer 2013(119), 2747–2753.
I/II study of imatinib mesylate for recurrent malignant gliomas: North American Zanotto-Filho, A., Braganhol, E., Battastini, A. M., & Moreira, J. C. (2012). Proteasome inhib-
Brain Tumor Consortium Study 99–08. Clin Cancer Res 2(16), 4899–4907. itor MG132 induces selective apoptosis in glioblastoma cells through inhibition of
Wick, W., Puduvalli, V. K., Chamberlain, M. C., van den Bent, M. J., Carpentier, A. F., Cher, L. PI3K/Akt and NFkappaB pathways, mitochondrial dysfunction, and activation of
M., et al. (2010). Phase III study of enzastaurin compared with lomustine in the treat- p38-JNK1/2 signaling. Invest New Drugs 30, 2252–2262.
ment of recurrent intracranial glioblastoma. J Clin Oncol 28(7), 1168–1174. Zhang, M., Herion, T. W., Timke, C., Han, N., Hauser, K., Weber, K. J., et al. (2011). Trimodal
Wick, W., Steibach, J. P., Plattern, M., Hartmann, C., Wenz, F., von Deimling, A., et al. gioblastoma treatment consisting of concurrent radiotherapy, temozolomide, and the
(2013). Enzastaurin before and concomitant with radiation therapy, followed by novel TGF-β receptor kinase inhibitor LY2109761. Neoplasia 13(6), 537–549.
enzastaurin maintenance therapy, in patients with newly diagnosed glioblastoma Zhang, M., Kleber, S., Röhring, M., Timke, C., Han, N., Tuettenberg, J., et al. (2011). Blockade
without MGMT promoter hypermethylation. Neuro Oncol 15, 1405–1412. of TGF-β signaling by the TGFRβR-I kinase inhibitor LY2109761 enhances radiation
Wick, W., Weller, M., Weiler, M., Batchelor, T., Yung, A. W., & Platten, M. (2011). Pathway in- response and prolongs survival in glioblastoma. Cancer Res 71(23), 7155–7167.
hibition: emerging molecular targets for treating glioblastoma. Neuro Oncol 13(6), Zhang, J., Stevens, M. F., & Bradshaw, T. D. (2012). Temozolomide: mechanisms of action,
566–579. repair and resistance. Curr Mol Pharmacol 5(1), 102–114.
Wieland, A., Trageser, D., Gogolok, S., Reinartz, R., Hofer, H., Keller, M., et al. (2013). Anticancer Zhao, D., Alizadeh, D., Zhang, L., Liu, W., Farrukh, O., Manuel, E., et al. (2011). Carbon nano-
effects of niclosamide in human glioblastoma. Clin Cancer Res 19(15), 4124–4136. tubes enhance CpG uptake and potentiate antiglioma immunity. Clin Cancer Res 17,
Wilson, T. A., Karajannis, M. A., & Harter, D. H. (2014). Glioblastoma multiforme: State of 771–782.
the art and future therapeutics. Surg Neurol Int 5, 64. Zhuang, W., Long, L., Zheng, B., Ji, W., Yang, Q., Zhang, Q., et al. (2012). Curcumin promotes
Wood, C. G., & Mulders, P. (2009). Vitespen: a preclinical and clinical review. Future Oncol differentiation of glioma-initiating cells by inducing autophagy. Cancer Science 103,
5, 763–774. 684–690.
Wrensch, M., Minn, Y., Chew, T., Bondy, M., & Berger, M. S. (2002). Epidemiology of Zou, Y., Cao, Y., Yue, Z., & Liu, J. (2013). Gamma secretase inhibitor DAPT suppresses
primary brain tumors: current concepts and review of the literature. Neuro Oncol 4, glioblastoma growth via uncoupling of tumor vessel density from vessel function.
278–299. Clin Exp Med 13(4), 271–278.
Xie, Q., Bradley, R., Kang, L., Koeman, J., Ascierto, M. L., Worschech, A., et al. (2012). Hepa-
tocyte growth factor (HGF) autocrine activation predicts sensitivity to MET inhibition
in glioblastoma. Proc Natl Acad Sci U S A 109(2), 570–575.

View publication stats

You might also like