You are on page 1of 17

Handbook of Clinical Neurology, Vol.

134 (3rd series)


Gliomas
M.S. Berger and M. Weller, Editors
© 2016 Elsevier B.V. All rights reserved

Chapter 23

Glioblastoma
HANS-GEORG WIRSCHING1*, EVANTHIA GALANIS2, AND MICHAEL WELLER1
1
Department of Neurology, University Hospital Zurich, Zurich, Switzerland
2
Departments of Oncology and Molecular Medicine, Mayo Clinic, Rochester, MN, USA

Abstract
Glioblastoma is the most common and aggressive primary brain tumor in adults. Defining histopathologic
features are necrosis and endothelial proliferation, resulting in the assignment of grade IV, the highest
grade in the World Health Organization (WHO) classification of brain tumors. The classic clinical term
“secondary glioblastoma” refers to a minority of glioblastomas that evolve from previously diagnosed
WHO grade II or grade III gliomas. Specific point mutations of the genes encoding isocitrate dehydro-
genase (IDH) 1 or 2 appear to define molecularly these tumors that are associated with younger age and
more favorable outcome; the vast majority of glioblastomas are IDH wild-type. Typical molecular
changes in glioblastoma include mutations in genes regulating receptor tyrosine kinase (RTK)/rat sar-
coma (RAS)/phosphoinositide 3-kinase (PI3K), p53, and retinoblastoma protein (RB) signaling. Standard
treatment of glioblastoma includes surgery, radiotherapy, and alkylating chemotherapy. Promoter meth-
ylation of the gene encoding the DNA repair protein, O6-methylguanyl DNA methyltransferase (MGMT),
predicts benefit from alkylating chemotherapy with temozolomide and guides choice of first-line treat-
ment in elderly patients. Current developments focus on targeting the molecular characteristics that drive
the malignant phenotype, including altered signal transduction and angiogenesis, and more recently,
various approaches of immunotherapy.

EPIDEMIOLOGY AND Risk factors


PATHOPHYSIOLOGY Risk factors for the development of glioblastoma other
Glioblastoma is a disease that accounts for 45.6% of pri- than age are poorly defined (Weller et al., 2015a). Males
mary malignant brain tumors, yet the annual incidence are affected more often than females (1.6:1) and whites
of 3.1 per 100 000 is low compared to cancers arising more often than blacks (2:1) (Ostrom et al., 2014a).
from other organs such as breast (171.20 per 100 000) Causes of these asymmetric distributions are elusive.
or prostate (201.40 per 100 000: Ostrom et al., 2014a). A small subset of less than 1% of glioblastomas are asso-
The annual age-adjusted incidence of glioblastoma ciated with hereditary cancer syndromes, including neu-
increases with age from 0.15 per 100 000 in children to rofibromatosis types 1 and 2, Turcot syndrome and Li–
the peak of 15.03 per 100 000 in patients aged 75–84 years Fraumeni syndrome, usually secondary to the diagnosis
(Ostrom et al., 2014a). Survival is inversely associated of World Health Organization (WHO) grade II or III gli-
with age: 5% of all patients diagnosed with glioblastoma omas (Ohgaki et al., 2010). A population-based study
are alive after 5 years, and this measure decreases to 2% including 10 834 patients treated with low-dose radio-
among patients aged 65 years or older (Ostrom therapy of the scalp for tinea capitis of 1–6 Gy in the
et al., 2014a). 1950s confirmed a relative risk (RR) increase for

*Correspondence to: Hans-Georg Wirsching, Department of Neurology, University Hospital Zurich, Frauenklinikstrasse 26,
CH-8091 Zurich, Switzerland. Tel: +41-44-255-5500, Fax: +41-44-255-4507, E-mail: hans-georg.wirsching@usz.ch
382 H-G. WIRSCHING ET AL.
gliomas of 2.6 (95% confidence interval (CI), 0.8–8.6), divide into different, more differentiated cell types).
and the excess RR for malignant tumors followed linear Finally, the capability to form secondary tumors in serial
kinetics and was 1.98/Gy (95% CI, 0.73–4.69) (Ron et al., xenotransplantation assays that resemble the original
1988; Sadetzki et al., 2005). Among long-term survivors tumors is a defining criterion of glioma-initiating cells.
who underwent high-dose irradiation (30–44.9 Gy) for Various markers of glioma-initiating cells have been
primary brain tumors in childhood, the odds ratio was described, including CD133 (prominin) (Singh et al.,
21 for gliomas (n ¼ 40, including n ¼ 9 glioblastomas) 2004), CD15 (stage-specific embryonic antigen-1) (Son
and the excess RR/Gy was 0.33 (95% CI, 0.07–1.71) et al., 2009), CD44 (Pietras et al., 2014), and integrin
(Neglia et al., 2006). The radiation dose of diagnostic alpha 6 (Lathia et al., 2010) but the abundance of any
scans does not suffice as a risk factor (Brada et al., of these markers differs strongly between tumors
1992; Minniti et al., 2005). Mobile phone use has been (Yan et al., 2013). Most progeny of glioma-initiating cells
studied extensively with respect to gliomagenesis, but have features of astrocytes, but differentiation into
no definite association has been reported (Interphone functional endothelial cells and pericytes has been noted,
Study Group, 2010; Ostrom et al., 2014b). No association too (Ricci-Vitiani et al., 2010; Wang et al., 2010; Cheng
with smoking or other cancerogenic agents has been et al., 2013). Glioma-initiating cells have been attributed
reported. Expression of cytomegalovirus (CMV) genes a crucial role in growth, angiogenesis, invasion, and
and interaction of CMV gene products with some of resistance to radiotherapy (Bao et al., 2006) and chemo-
the core pathways that drive the malignant phenotype therapy (Beier et al., 2008; Chen et al., 2012), but, despite
of glioblastoma suggest an oncomodulatory role for promising preclinical data, no drugs specifically target-
CMV, but a firm role for CMV as a glioma-initiating ing glioma-initiating cells have entered clinical practice
agent remains to be confirmed (Dziurzynski et al., to date.
2012; Wick and Platten, 2014). In summary, ionizing irra-
diation of the brain is the only recognized exogenous risk
Tumor location
factor for the development of glioblastoma (Connelly
and Malkin, 2007; Bondy et al., 2008; Ostrom and There is no association of the location of glioblastoma at
Barnholtz-Sloan, 2011). diagnosis with the locations of NSPC, besides the con-
finement of most glioblastomas to the supratentorial
compartment. Reasons likely include excessive dissemi-
Cell of origin
nation of glioblastoma cells throughout the brain even at
The low frequency of glioblastoma and of primary brain early stages of the disease, but development of glioblas-
tumors in general compared to tumors arising from toma from non-NSPC cannot be excluded. Most gliomas
extraneural organs may reflect a high degree of protec- occur in the frontal (25.8%), temporal (19.7%), and pari-
tion of the brain from genotoxic stress. ATP-binding cas- etal (12.2%) lobes, whereas occipital-lobe (3.2%), cere-
sette (ABC) family transporter molecules of the blood– bellar (2.9%), brainstem (4.2%), or spinal cord (4.3%)
brain barrier restrict diffusion of chemical mutagens to gliomas are relatively rare (Ostrom et al., 2014a). Brain-
the brain (El Ali and Hermann, 2011). Furthermore, DNA stem gliomas are rare in adults, but account for the vast
is particularly sensitive to genotoxic stress during repli- majority of pediatric glioblastomas (Ostrom et al., 2015).
cation, but most brain cells have entered a definite post- Survival of patients with frontal-lobe glioblastomas was
mitotic state until adulthood. Experimental data suggest longer compared to patients with temporal- or parietal-
that adult glioblastoma may be derived from the small lobe glioblastoma (median overall survival (OS), 11.4
pool of adult neural stem and progenitor cells (NSPC), vs. 9.1 vs. 9.6 months, log rank p ¼ 0.01) in a cohort of
which are located in the subventricular zone (Sanai 645 adults who were treated within three consecutive tri-
et al., 2004), the subcortical white matter (Nunes et al., als of the Radiation Therapy Oncology Group (RTOG) in
2003), and the dentate gyrus of the hippocampus the pre-temozolomide era (Simpson et al., 1993). This is
(Eriksson et al., 1998). NSPC have retained the ability likely due to an association of frontal-lobe glioblastoma
to enter mitosis and play a major role for the plasticity with other favorable prognostic features, such as youn-
of the brain in learning and memory (Sanai et al., ger age, isocitrate dehydrogenase-1/2 (IDH-1/2) muta-
2005). A subpopulation of glioblastoma cells termed tions (Sturm et al., 2012; Ellingson et al., 2013), and
glioma-initiating (also called glioma stem-like cells) higher rates of gross total resection (Lacroix et al.,
share various features with NSPC, including the location 2001). Approximately half of all glioblastomas in adults
in perivascular and hypoxic niches (Gilbertson and Rich, infiltrate more than one lobe and approximately 5%
2007), self-renewal (i.e., the capability to divide into two grow multifocally (Djalilian et al., 1999). Leptomenin-
progeny, of which at least one resembles the initial cell) geal dissemination of glioblastoma cells is rare and
and multilineage differentiation (i.e., the capability to occurs in late stages of the disease (Herrlinger et al.,
GLIOBLASTOMA 383
2004), although a frequency of up to 14% has been in the particularly impaired elderly population (Roa
reported in one series comprising younger patients et al., 2004; Taphoorn et al., 2005; Keime-Guibert
(median age, 31 years) (Arita et al., 1994). In another et al., 2007; Gallego Perez-Larraya et al., 2011). Person-
series, 5 of 25 autopsied patients with glioblastoma ality changes and mood disorders commonly occur in
had spinal leptomeningeal metastases (Erlich and patients with frontal tumors, and these may be mistaken
Davis, 1978). Associations of leptomeningeal dissemina- for psychogenic disorders or part of the physiologic
tion with younger age, male gender, incomplete tumor aging process and thus delay diagnosis. Sensorimotor
removal, multiple resections, ventricular entry or prox- deficits are the presenting symptom in approximately
imity of the tumor to the ventricular system, and with 20% of all patients with glioblastoma, and approxi-
gains at the 1p36 chromosomal region, have been noted mately 5% of patients present with aphasia due to
(Awad et al., 1986; Grabb et al., 1992; Arita et al., 1994; tumors arising in the speech-dominant hemisphere
Lindsay et al., 2002; Korshunov et al., 2007). Distant (mostly left-sided) (Yuile et al., 2006).
organ metastases arising from glioblastoma have been Epilepsy may mimic aphasia or sensorimotor deficits
reported, predominantly to lung, pleura, lymph nodes, from tissue destruction postictally and occurs more fre-
bone, and liver (Schweitzer et al., 2001), but these cases quently in glioblastomas that affect the temporal lobe
are rare. The reasons for the strong brain tropism of glio- (Chaichana et al., 2009). Epilepsy is the presenting symp-
blastoma cells are not known, but likely include adapta- tom in 24–68% of glioblastomas and develops in 19–38%
tion to the metabolic and immunologic peculiarities of later during the course of the disease (Wick et al., 2005;
the brain microenvironment that precludes survival in Chaichana et al., 2009; van Breemen et al., 2009;
other organs (Fecci et al., 2014; Mashimo et al., 2014). Kerkhof et al., 2013). As a presenting symptom, epilepsy
Furthermore, spreading glioblastoma cells reactivate is associated with longer survival, probably due to an
developmental programs of neuroglial precursor cells association with younger age as well as cortical location
and may follow the same paths along which neuroglial and smaller tumor size, indicating good surgical resect-
precursor cells migrate during development, and extra- ability and possibly diagnosis earlier during the disease
neural migration of neuroglial precursor cells does not course (Yuile et al., 2006; Chaichana et al., 2009). Head-
occur. Due to the restriction of glioblastoma cell spread- aches are the presenting symptom in less than one-third
ing within the borders of the central nervous system, the of all patients with glioblastoma (Yuile et al., 2006), and
transplantation of organs from donors with glioblas- are mostly dull in nature and typically present at night or
toma is per se feasible (Watson et al., 2010; Warrens at awakening. Other signs of increased intracranial pres-
et al., 2012), but only few glioblastoma patients qualify sure, such as nausea, vomiting, dizziness, fatigue, and
as organ donors. This is because the mode of death is neurocognitive slowing, occur increasingly during the
mostly a gradual decline in a home or hospice setting, disease course, but may be present at diagnosis. Rarely,
whereas organ donation requires brain death due to a leptomeningeal dissemination may mimic painful nerve
sudden intracranial event (e.g., hemorrhage), or gradual root compression, myelitis, and other spinal diseases.
death from compromised airway protection in a hospi- Only few patients retain stable disease and remain neu-
tal setting that includes intubation to enable nonheart- rologically largely asymptomatic for years, but the
beating donation (Collignon et al., 2004). majority of patients experience severe impact on quality
of life once first-line treatments have failed (Roa et al.,
2004; Taphoorn et al., 2005; Keime-Guibert et al., 2007;
Clinical course
Gallego Perez-Larraya et al., 2011). In summary, there is
The clinical course of glioblastoma is determined by no typical clinical presentation of glioblastoma. Differ-
tumor location and dynamics of spread within the brain. ences compared to other gliomas include the more rapid
Tissue destruction, edema, and epilepsy likewise con- dynamic and somewhat lower incidence of epilepsy (see
tribute to clinical symptoms, causing rapid deteriora- Chapter 2).
tion in some patients. The clinical term secondary
glioblastoma refers to glioblastomas that are preceded DIAGNOSIS AND MOLECULAR
by the histologic diagnosis of WHO grade II or III CLASSIFICATION
gliomas, but this term is increasingly dispensable with
Imaging
molecular marker-based classifications of gliomas
(Weller et al., 2015a). New-onset seizures or development of neurologic defi-
Despite the invariably fatal prognosis of glioblas- cits are commonly followed by a neurologic workup that
toma, any of the standard treatment options discussed includes magnetic resonance imaging (MRI). Computed
further below may keep stable or even improve quality tomography (CT) with contrast enhancement is less sen-
of life and cognitive functioning for the time being, even sitive in detecting the typical features of glioblastoma.
384 H-G. WIRSCHING ET AL.
Its use is restricted to acute situations, e.g., when hem- blood–brain barrier disruption, central hypointensity on
orrhage is suspected, or when MRI is not available or T2-weighted images as a correlate of necrosis, and perifo-
not possible, e.g., in patients with cardiac pacemakers cal hyperintensity on T2-weighted and fluid-attenuated
or other metallic implants. Prior to biopsies, amino acid inversion recovery (FLAIR) images as a correlate of
positron emission tomography (PET) is increasingly per- edema or noncontrast-enhancing tumor (Fig. 23.1). MR
formed to guide the site of biopsy to metabolic hotspots spectroscopy, T1 contrast subtraction maps, as well as
that may represent sites of higher tumor grade diffusion/perfusion- and susceptibility-weighted MRI
(la Fougere et al., 2011); yet, PET is not part of the stan- sequences have refined radiographic diagnostics and
dard workup of glioblastoma patients (see Chapter 3). enable more reliable discrimination of glioblastoma from
On MRI, glioblastomas appear as masses with con- other contrast-enhancing lesions, including abscesses, pri-
trast enhancement at their margin as a correlate of mary central nervous system lymphomas, and metastases

Fig. 23.1. Neuroimaging features of glioblastoma. Magnetic resonance imaging (MRI), T1-weighted (A), gadolinium-enhanced
(B), T2-weighted (C), fluid-attenuated inversion recovery (FLAIR, D); (E): 18F-fluoro-ethyl-tyrosine (FET) positron emission
tomography (PET).
GLIOBLASTOMA 385
from nonprimary brain tumors (Kono et al., 2001; Development of gliosarcoma may occur secondary to
Law et al., 2003; Weber et al., 2006; Ellingson et al., glioblastoma and molecularly, mutations of the epithe-
2012; Kickingereder et al., 2014), as discussed in more lial growth factor receptor (EGFR) are infrequent com-
detail in Chapter 3. However, the appearance of glio- pared to glioblastoma. The prognosis may or may not be
blastoma on imaging scans can vary considerably, and worse than for glioblastoma, after adjustment for com-
therefore tissue-based diagnosis is indispensable (Weller mon prognostic factors (hazard ratio (HR), 1.17; 95%
et al., 2014). confidence interval (CI), 1.05–1.31) (Kozak et al., 2009).

Histopathology
GIANT CELL GLIOBLASTOMA
The defining histopathologic features of glioblastoma
are necrosis and microvascular proliferation, and these Giant cell glioblastoma is characterized by multinu-
qualify glioblastoma for the highest grade in the WHO cleated cells with a diameter of more than 500 mm and
classification of primary brain tumors, grade IV lymphocytic infiltration. Giant cell glioblastomas com-
(Louis et al., 2007). Other signs of malignancy that are prise approximately 1% of all glioblastomas. The inci-
present in glioblastoma include anaplasia, high mitotic dence may be higher in young adults (median age at
rates, and invasiveness, but these features are also pre- diagnosis 51 years). The prognosis for patients with giant
sent in anaplastic gliomas, which are assigned WHO cell glioblastoma is better than for glioblastoma, after
grade III (Louis et al., 2007). In addition, immunohisto- adjustment for common prognostic factors (HR, 0.76;
chemical markers are commonly assessed to ascertain 95% CI, 0.59–0.97) (Kozak and Moody, 2009).
the diagnosis of glioblastoma, including glial fibrillary Other histopathologic variants have been suggested,
acidic protein expression to confirm astrocytic lineage including fibrillary glioblastoma, small cell astrocytoma,
differentiation and MIB-1/Ki-67 to aid quantification glioblastoma with oligodendroglial component, glioblas-
of proliferation. An antibody that specifically detects toma with primitive neuroectodermal tumor, and granu-
mutant IDH-1R132H (Capper et al., 2009) has been inte- lar cell astrocytoma (Karsy et al., 2012), but none of
grated into the standard diagnostic workup of suspected these have been accepted as designated variants within
glioblastoma, but it must be recognized that this anti- the current WHO classification of brain tumors because
body does not detect other IDH-1 or IDH-2 mutations. of a lack of distinct molecular and clinical features
A more detailed overview on histopathologic grading (Louis et al., 2007).
of gliomas is provided in Chapter 5. We emphasize
that the traditional histopathologic definition of glio- Molecular markers
blastoma (and other gliomas) is likely to be comple-
MGMT promoter methylation predicts benefit from
mented by molecular approaches with the next edition
temozolomide chemotherapy in newly diagnosed and
of the WHO classification of brain tumors, due to be
probably also recurrent glioblastoma (Hegi et al., 2005;
released in 2016.
Malmstrom et al., 2012; Wick et al., 2012; Weller et al.,
Glioblastoma variants 2015b). MGMT is a DNA repair protein that counteracts
DNA alkylation by chemotherapy; hypermethylation of
Two rare histopathologic variants of glioblastoma are the MGMT promoter results in gene silencing (Weller
defined in the WHO classification of primary brain et al., 2010). Since the discovery of its role in the resistance
tumors, gliosarcoma and giant cell glioblastoma (Louis of glioblastoma to alkylating chemotherapy, MGMT has
et al., 2007). remained the biomarker with the strongest impact on clin-
ical decision making, particularly in elderly glioblastoma
GLIOSARCOMA
patients (Fig. 23.2) (Weller et al., 2014).
Gliosarcoma is characterized by a metaplastic mesenchy- IDH mutations occur in approximately 5–10% of all
mal component that stains for reticulin and may show glioblastomas, are associated with younger age and bet-
signs of fibroblastic, cartilaginous, osseous, smooth ter outcome (Parsons et al., 2008), and rarely occur in
and striated muscle, or adipose cell lineage, and these patients aged 65 years or older. Mutant IDH produces
features are present in addition to characteristics of glio- the oncometabolite 2-hydroxyglutarate (Turcan et al.,
blastoma (Han et al., 2010). Gliosarcomas comprise 2012), which has been linked to a distinct epigenetic pat-
approximately 2% of all patients diagnosed with glio- tern designated glioma CpG island methylator pheno-
blastoma or gliosarcoma and clinical features include type (G-CIMP) (Noushmehr et al., 2010). In contrast,
predilection for location in the temporal lobe, a IDH wild-type glioblastoma is associated with older
meningioma-like macroscopic appearance at surgery, age and poor prognosis; TERT promoter mutation marks
and frequent reports of extracranial metastases. particularly poor outcome in these patients (Labussiere
386 H-G. WIRSCHING ET AL.
MRI
Contrast enhancing, diffusely infiltrating mass

Surgery/Biopsy
Gross total resection, if safely possible

Glioblastoma

Age < 65-70 years Age > 65-70 years

MGMT methylated MGMT unmethylated

TMZ/RT TMZ TMZ or TMZ/RT TMZ Hypofractionated RT

Repeat RT
Mostly stereotactic No firstline RT
1st RT >6 months Recurrent glioblastoma
Young age
Small tumors
Good KPS
Systemic therapies Surgery
contraindicated or Only if gross total resection is safely possible
failed

Combined?

TMZ Lomustine BEV


MGMT is prognostic MGMT is prognostic MGMT is prognostic
PD on TMZ dose-intensified Suitable for most patients Benefit may be higher in
TMZ-free interval standard dose Adequate hematologic elderly patients with poor KPS
Adequate function required Recent hemorrhage/thromboembolism
hematologic function required are contraindications
Availability is restricted in the EU

Fig. 23.2. Therapeutic approach to glioblastoma. MRI, magnetic resonance imaging; MGMT, O6-methylguanyl DNA methyl-
transferase gene promoter; TMZ, temozolomide 150–200 mg/m2 on 5/28 days; TMZ/RT, 30  2 Gy ¼ 60 Gy with daily concom-
itant temozolomide at 75 mg/m2; hypofractionated RT, radiotherapy at 15  2.66 Gy ¼ 40 Gy; BEV, bevacizumab 10 mg/kg body
weight every 2 weeks; KPS, Karnofsky performance score; PD, progressive disease.

et al., 2014). Despite identical histopathologic appear- EGFR to drive cellular transformation and the glioma-
ance, IDH wild-type and mutant glioblastomas consti- initiating cell phenotype via signal transducer and activa-
tute two distinct molecular and prognostic entities that tor of transcription (STAT) 3 signaling (Fan et al., 2013),
are likely to be separated in future classifications of and the EGFRvIII deletion mutation may confer poor
brain tumors, as outlined in more detail in Chapter 5. prognosis compared to EGFR-amplified glioblastoma
About half of all glioblastomas with EGFR amplifi- without co-occurrence of EGFRvIII (Shinojima et al.,
cation (i.e., 20–25% of all IDH wild-type glioblastomas) 2003; Heimberger et al., 2005). Promising results with
harbor a deletion within the extracellular ligand-binding vaccination targeting EGFRvIII in phase II clinical trials
domain of EGFR (designated delta-EGFR or EGFRvIII) have been reported (Sampson et al., 2010; Hegi et al.,
that yields ligand-independent receptor activity (Aldape 2012; Reardon et al., 2015).
et al., 2004). EGFRvIII is commonly expressed only in
a subset of EGFR-amplified cells (Nishikawa et al.,
Molecular signatures
2004), but preclinical and clinical data support a role
for EGFRvIII-positive cells as drivers of disease pro- Integrated large-scale analyses of genetic, epigenetic,
gression. The EGFRvIII protein interacts with wild-type and expression data increasingly complement the
GLIOBLASTOMA 387
understanding of the biology and yield continuous proneural glioblastoma, whereas G34, K27, and RTK
refinement of the subclassification of glioblastoma I are associated with poor prognosis that is comparable
beyond histologic grading. In 2006 three gene expression to the two nonproneural subtypes (Sturm et al., 2012).
subtypes were defined by cluster analysis of 35 genes Meanwhile, the gene expression-based glioblastoma sub-
that were correlated the strongest with survival in types have also been detected on the single cell level
76 patients with newly diagnosed anaplastic astrocytoma within the same tumors, and an association of these gene
or glioblastoma and termed proneural, mesenchymal, expression profiles with stemness-related genes and
and proliferative (Phillips et al., 2006). The proneural glioma-initiating cells was suggested (Patel et al.,
subtype was associated with prolonged survival, youn- 2014). To date no therapeutic implications for clinical
ger age, and anaplastic histology and lacked alterations routine are based on these global characterizations of
of phosphatase and tensin homolog on chromosome the molecular background of individual tumors. The
10 (PTEN) or EGFR. The proliferative and mesenchymal molecular subclassification of glioblastoma is discussed
subtypes expressed genes that were associated with pro- in more detail in Chapter 6.
liferation or angiogenesis, respectively, and survival was
poor. These analyses were extended to larger datasets TREATMENT AND FOLLOW-UP
and complemented by mutation analyses in subsequent
Surgery
years. In 2008, three altered key signaling pathways were
determined by The Cancer Genome Atlas project (2008), Tissue is required to establish the diagnosis of glioblas-
namely receptor RTK/RAS/PI3K (88%), p53 (87%), and toma, as outlined above. It can be obtained by stereotac-
retinoblastoma protein (78%), and in the same year, tic or open biopsy, or by microsurgical resection of the
IDH-1/2 mutations were discovered as a marker that tumor. Biopsies are mostly performed as stereotactic
is strongly associated with secondary glioblastoma, biopsies in patients with multifocal disease and in tumors
younger age, and better outcome (Parsons et al., 2008; deemed unresectable due to their location in so-called
Yan et al., 2009). In 2010, the initially suggested gene “eloquent” areas, i.e., areas that cannot be resected with-
expression-based subclassification of glioblastoma was out causing major disability. However, biopsies may not
further refined using 601 genes and sequence data yield enough tissue for molecular analyses and bear the
from 91 patients, yielding four subtypes designated pro- risk of sampling errors. The therapeutic value of micro-
neural, neural, classic, and mesenchymal, based on sim- surgical resection has been demonstrated best in a ran-
ilarities with known gene expression profiles (Verhaak domized trial exploring the use of the fluorescent
et al., 2010). label, 5-aminolevulinc acid, to facilitate gross total resec-
IDH-mutant glioblastomas constitute a subgroup of tion (Stummer et al., 2006). It is further supported by
proneural glioblastomas associated with G-CIMP numerous retrospective cohort studies, establishing
(Noushmehr et al., 2010). Of note, there is no association gross total resection as standard of care whenever
of gene expression-based subtypes of G-CIMP with deemed feasible (Fig. 23.2). (Sanai et al., 2011; Weller
MGMT promoter methylation (Bady et al., 2012). By et al., 2014) Feasibility of this concept has also been dem-
2012, epigenetic profiling yielded a classification of onstrated for elderly patients aged 65 years or older:
six subtypes, four of which clustered within the pro- Extent of resection was a prespecified, independent
neural gene expression subtype: two distinct histone prognostic factor in a survival model that controlled
H3 (H3F3A) mutations, G34 and K27, were detected for age, extent of resection, histology, MGMT status,
in a large proportion of pediatric glioblastomas and these and study treatment among 371 elderly patients with ana-
mutations as well as IDH mutations were mutually plastic astrocytoma (n ¼ 40) or glioblastoma (n ¼ 331)
exclusive. These three subtypes share an association with randomized to receive postoperative radiotherapy or
mutations of the tumor suppressor gene TP53. A fourth temozolomide chemotherapy within the NOA-08 trial
proneural subgroup, designated RTK I, was character- (Wick et al., 2012). Further, among 342 patients with
ized by upregulation of platelet-derived growth factor newly diagnosed glioblastoma aged 60 or older who
receptor alpha (PDGFRA) and frequent deletions of were randomized to two different radiotherapy regi-
the CDKN2A gene encoding cyclin-dependent kinase mens or temozolomide, biopsy was associated with infe-
inhibitor 2A (Sturm et al., 2012). The nonproneural sub- rior OS compared to surgical resection (HR, 1.50; 95%
types were designated RTK II/classic, which is charac- CI, 1.17–1.92) in multivariate analyses controlling for
terized by frequent EGFR amplifications and age, type of surgery, study treatment, and WHO perfor-
CDKN2A deletions, and the mesenchymal subtype. mance score (Malmstrom et al., 2012). The use of
These six subclasses were associated with distinct tumor 5-aminolevulinic acid (see above) (Stummer et al.,
locations, age distributions, and prognosis. Only IDH- 2006, 2008) or intraoperative imaging by ultrasound
mutant glioblastomas retain the good prognosis of or MRI may aid the surgeon to obtain the goal of an
388 H-G. WIRSCHING ET AL.
optimal resection (Rygh et al., 2008; Kubben et al., 2011), radiotherapy was likely that standard radiotherapy was
while nuclear imaging with 18F-DOPA is currently in clin- not completed by a substantial fraction of patients
ical testing (NCT02020720). (Malmstrom et al., 2012).

Chemoradiotherapy Chemotherapy in elderly patients


Daily concomitant temozolomide at 75 mg/m2 as an Postoperative chemotherapy with temozolomide is a
adjunct to radiotherapy (30  2 Gy ¼ 60 Gy of the valid alternative to radiotherapy in elderly patients with
involved field) followed by up to six cycles of temozolo- glioblastomas with MGMT promoter hypermethylation,
mide at 150–200 mg/m2 on 5 of out 28 days prolonged but not in patients with tumors without MGMT promoter
median OS compared to radiotherapy alone (14.6 versus methylation (Fig. 23.2) (Weller et al., 2014). This was
12.1 months; HR, 0.63; 95% CI, 0.52–0.75, p < 0.001) demonstrated in parallel by two phase III clinical trials
(Stupp et al., 2005), but benefit from temozolomide in elderly patients: (1) the Nordic trial, which randomized
was mainly restricted to patients with MGMT promoter patients aged 60 years or older to receive temozolomide
methylation (Hegi et al., 2005). Patients aged older than dosed at 150–200 mg/m2 on 5 of out 28 days or one of
70 years were not included in this trial, and the subgroup two radiotherapy regimens (discussed above) (Wick
of patients aged 66–70 years appeared not to benefit et al., 2012); and (2) the NOA-08 trial, which randomized
from combined chemoradiotherapy in post hoc analyses patients aged 65 years or older with glioblastoma (89%)
(HR, 0.78; 95% CI, 0.50–1.25; p ¼ 0.29) (Laperriere et al., or anaplastic astrocytoma (11%) to receive standard
2013). The recently completed National Cancer Institute radiotherapy or temozolomide at a dose-dense schedule
of Canada CE6 phase III trial will define the efficacy of of 100 mg/m2 given on days 1–7 every other week
combined chemoradiotherapy versus radiotherapy alone (1 week on/1 week off) (Wick et al., 2012). In the Nordic
in elderly patients with newly diagnosed glioblastoma trial, temozolomide (n ¼ 93) and hypofractionated radio-
and good clinical performance (NCT00482677), but to therapy (n ¼ 98) were similarly effective (HR, 0.82; 96%
date combined chemoradiotherapy is the standard of CI, 0.63–1.06), adjusted for age, type of surgery (biopsy
care only for fit elderly patients (Fig. 23.2) (Weller versus resection), and WHO performance score among
et al., 2014). the intention-to-treat population (Malmstrom et al.,
2012). Efficacy analyses of the NOA-08 trial included
Radiotherapy in elderly patients 373 patients who received at least one dose of treatment
among a total of 412 randomized and 584 screened
Postoperative radiotherapy has activity in elderly
patients. After adjustment for age, histologic diagnosis,
patients with glioblastoma, as demonstrated by a trial
extent of resection, and MGMT promoter methylation,
that randomized 81 patients with glioblastoma and 2
temozolomide was noninferior to radiotherapy (HR,
patients with anaplastic astrocytoma aged 70 years or
1.09; 95% CI, 0.84–1.42) (Wick et al., 2012). Benefit from
older to either radiotherapy with best supportive care,
temozolomide was predicted by MGMT promoter meth-
or best supportive care alone (median OS, 29.1 vs.
ylation in both trials and radiotherapy was favorable
16.9 weeks; p ¼ 0.002) (Keime-Guibert et al., 2007).
among patients with tumors without MGMT promoter
These results were underpinned by a population-based
methylation. Dose-intensified temozolomide regimens
retrospective review of almost 3000 patients with glio-
became largely obsolete, because of greater toxicity,
blastoma aged 71–98 years (median age 76.9 years) after
but similar activity was noted compared to standard
adjusting for tumor size, location, type of surgery, and
temozolomide dosing at 150–200 mg/m2 on 5 out of
demographics (HR, 0.43; 95% CI, 0.38–0.49) (Scott et al.,
28 days in two phase III trials that directly compared
2011). Radiotherapy in elderly patients with glioblastoma
standard and dose-intensified temozolomide in patients
is commonly administered as a hypofractionated proto-
with newly diagnosed and recurrent glioblastoma (Brada
col (Weller et al., 2014), based on a randomized trial in
et al., 2010; Gilbert et al., 2013).
95 patients aged 60 years or older that compared stan-
dard radiotherapy (30  2 Gy ¼ 60 Gy) versus hypofrac-
Antiangiogenic therapy
tionated radiotherapy (15  2.66 Gy ¼ 40 Gy), which
demonstrated comparable activity of both regimens Angiogenesis is a defining characteristic of glioblastoma
(HR, 0.90; 95% CI, 0.60–1.35; p ¼ 0.61) (Roa et al., that is driven by vascular endothelial growth factor
2004). Among the intention-to-treat population of the (VEGF) (Batchelor et al., 2014) and other mediators,
randomized Nordic trial, standard radiotherapy was and the concept of targeting blood vessels to induce
even inferior to hypofractionated radiotherapy with “tumor starvation” has rendered VEGF an intensely
10  3.4 Gy ¼ 34 Gy (OS 7.0 vs. 5.2 months; p ¼ 0.02). studied therapeutic target in a plethora of tumors beyond
Notably, one reason for the inferiority of standard glioblastoma (Folkman, 1971; Ellis and Hicklin, 2008;
GLIOBLASTOMA 389
Grothey and Galanis, 2009). The anti-VEGF antibody age-stratified survival differences. Ongoing research is
bevacizumab is the antiangiogenic drug best studied in focusing on the identification of molecular markers
glioblastoma. It is commonly dosed at 10 mg/kg body predicting benefit from antiangiogenic therapies; to
weight intravenously every other week and is well toler- date, no such marker has been defined, but recently
ated by most patients. Arterial hypertension is the most association of the proneural subtype with benefit
common side-effect, and rare severe complications that from bevacizumab was reported in the AVAGlio trial
are associated with bevacizumab include hemorrhages, (Sandmann et al., 2015). In summary, the available evi-
thromboembolic events, complications of wound heal- dence currently does not support the administration of
ing, congestive heart failure, and gastrointestinal antiangiogenic therapy as a first-line postoperative
perforations. therapy in glioblastoma (Batchelor et al., 2014).
Two placebo-controlled phase III trials in patients
with newly diagnosed glioblastoma (RTOG 0825 and Response assessment
AVAGlio) noted prolonged progression-free survival
MRI is the method of choice to assess response to therapy
with the addition of bevacizumab to standard combined
in glioblastoma patients and is commonly performed
chemoradiotherapy (Chinot et al., 2014; Gilbert et al.,
every 2–3 months. Response is categorized as complete
2014). Less corticosteroid use, preserved general condi-
response, partial response, stable disease, and progressive
tion, and preserved quality of life were additional bene-
disease. Contrast enhancement on T1-weighted sequences
fits from bevacizumab reported in the AVAGlio trial
has been the standard measure to assess response in
(Chinot et al., 2014; Taphoorn et al., 2015), whereas the
glioblastoma for almost two decades (Macdonald et al.,
investigators of the RTOG 0825 trial noted stronger
1990), but was revisited due to increasing recognition
impairment of quality of life and neurocognitive decline
of “pseudoresponse” resulting from antiangiogenic
under bevacizumab (Gilbert et al., 2014). Reasons for
treatments, and of “pseudoprogression” resulting from
these conflicting data are speculative, but differences
radiotherapy or chemoradiotherapy (Wen et al., 2010).
in the response assessment within both trials may be
Pseudoresponse refers to a rapid decrease of contrast--
explanatory. The AVAGlio trial included measures to
enhancing lesions that may occur within hours of antian-
detect “pseudoresponse,” i.e., a sustained reduction in
giogenic treatment as a result of blood–brain barrier
contrast enhancement due to normalization of the
normalization. Pseudoprogression refers to a subacute,
blood–brain barrier (see below and Chapter 3), whereas
transient increase in the size of contrast-enhancing lesions
in the RTOG 0825 trial, early tumor progression may
after radiotherapy and appears to occur more frequently
have been missed due to the definition of progressive
under more intense combined chemoradiotherapy
disease solely based on contrast enhancement
(Brandsma et al., 2008). Pseudoprogression was identi-
(Batchelor et al., 2014). To date, three phase III trials
fied on the first MRI scans after radiotherapy in 9–31%
have failed to demonstrate an OS benefit from antian-
of patients with glioblastoma and contrast-enhancing
giogenic therapies in newly diagnosed glioblastoma,
anaplastic gliomas (de Wit et al., 2004; Brandsma et al.,
including the AVAGlio and RTOG 0825 trials of bevaci-
2008), but the precise frequency of pseudoprogression
zumab and a trial of the integrin inhibitor cilengitide
is difficult to estimate, because chemotherapy is usually
(Chinot et al., 2014; Gilbert et al., 2014; Stupp et al.,
continued when pseudoprogression is suspected, thus
2014a) The diverging results of improved progression-
precluding a differentiation of pseudoprogression from
free survival versus a lack of OS benefit in the AVAGlio
delayed treatment responses. The pathophysiology of
and RTOG 0825 trials may be attributed to substantial
pseudoprogression has not been clarified, but likely
crossover from the placebo-controlled arm to receive
constitutes a pronounced inflammatory response to
bevacizumab at progression (AVAGlio: 31.3%, RTOG
treatment yielding edema and disruption of the blood–
0825: 48.3%), or due to misdiagnosis of pseudoresponse,
brain barrier, which present as an increase in contrast
but this remains speculative (Batchelor et al., 2014). Fur-
enhancement on MRI scans. In most patients, pseudopro-
thermore, early reports suggested that particularly
gression is oligosymptomatic and transient, but severe
elderly patients in poor general condition might benefit
cases that correspond to treatment-induced necrosis
from bevacizumab (Nghiemphu et al., 2009; Lai et al.,
may require further treatment, as discussed in more detail
2011), but these patients were underrepresented in the
in Chapters 3 and 13.
AVAGlio and RTOG 0825 trials, because good general
condition was an inclusion criterion (Chinot et al.,
Recurrent glioblastoma
2014; Gilbert et al., 2014). A tendency to improved OS
in the bevacizumab group of the AVAGlio trial has Escape from antitumor therapy is caused by adaptations
indeed been noted with increasing age (Chinot et al., of the molecular machinery of tumor cells, mostly due
2014), but the trial was not powered to identify to selection of resistant, genetically distinct clones.
390 H-G. WIRSCHING ET AL.
Therefore, therapeutic options at progression depend on clinical trials have compared different radiotherapy
previously administered therapies, but tumor character- schedules in previously irradiated glioblastoma patients.
istics, availability, and local preferences are factors that Systemic treatments are the mainstay of therapy for
influence treatment choices, too. No standard of care for recurrent glioblastoma (Fig. 23.2) and include single-
the treatment of recurrent glioblastoma has been defined agent or combination treatments with nitrosoureas (in
and the efficacy of available treatment options at pro- particular, CCNU/lomustine), temozolomide, and beva-
gression is commonly limited (Weller et al., 2013). cizumab (Weller et al., 2013). All of these options are well
Repeat surgery may improve survival in some tolerated by most patients. Lomustine dosed at
patients, but this has not been validated in prospective 90–110 mg/m2 every 6 weeks is commonly utilized as a
controlled trials. Data supporting second surgery at pro- control in clinical trials for recurrent glioblastoma and
gression suggest that gross total resection, but not progression-free survival rates at 6 months are in the
incomplete resection, is beneficial (Fig. 23.2) (Bloch et al., range of 19–25% (Wick et al., 2010; Batchelor et al.,
2012; Suchorska et al., 2015). Other factors associated 2013). In patients who appeared to derive benefit from
with favorable outcome after repeat surgery apply to first-line temozolomide, rechallenge with temozolomide
approximately 1 in 4 patients with recurrent glioblastoma is a viable option after a temozolomide-free interval
and include age <70 years, Karnofsky performance (Weller et al., 2015b; Brada et al., 2010). MGMT pro-
score >80% and small tumor volume (<50 cm3), moter methylation is associated with prolonged survival
whereas involvement of functionally relevant at temozolomide rechallenge (Weller et al., 2015b). Dose-
(“eloquent”) brain structures is associated with poor intensified temozolomide regimens have been studied
postoperative survival (Park et al., 2010). The option extensively, but standard dosing at 150–200 mg/m2 on
of repeat surgery should thus only be considered if the 5/28 days may be preferred (Weller et al., 2014): more
risk for surgical complications is low, and if the general toxicity, but no survival benefit from dose intensifica-
condition of patients is estimated to remain good enough tion, was noted in two phase III trials directly comparing
postoperatively to allow for additional systemic therapy standard and dose-intensified temozolomide in glioblas-
(or repeat radiotherapy in selected cases, discussed toma (Brada et al., 2010; Gilbert et al., 2013).
below). Surgery should not replace systemic therapy, Accelerated approval by the US Food and Drug
because infiltrating cells beyond the site of resection fre- Administration (FDA) and in various other countries
quently cannot be captured by neuroimaging, and these was obtained for bevacizumab for the treatment of
cells can drive further tumor progression rapidly, even recurrent glioblastoma based on two uncontrolled phase
when gross total resection is confirmed radiographically. II trials based on durable response rate (Friedman et al.,
Hypofractionated radiotherapy is the treatment of 2009; Kreisl et al., 2009), but two trials of direct or indi-
choice in elderly patients with tumors with MGMT pro- rect small-molecule inhibitors of VEGF signaling,
moter methylation who received first-line postoperative cediranib and enzastaurin, in recurrent glioblastoma
treatment with temozolomide alone, but the role of failed to prolong survival (Wick et al., 2010; Batchelor
repeat radiotherapy is less clear for all other patients et al., 2013).
(Fig. 23.2). Some activity of repeat radiotherapy was In the randomized noncomparative phase II BELOB
deduced from uncontrolled trials and retrospective data trial, a combination of lomustine with bevacizumab at
analyses, but these included mostly patients with favor- first recurrence of glioblastoma was well tolerated and
able prognostic factors due to smaller tumor volumes, appeared to confer additive effects: median postrecur-
younger age, and better general condition (Dhermain rence survival was 8 months with lomustine alone,
et al., 2004; Butowski et al., 2006). Repeat radiotherapy 8 months with bevacizumab alone, and 12 months with
should only be considered when the recurrence has a combination of lomustine and bevacizumab (Taal
occurred after a relatively long interval from first radio- et al., 2014). A variety of other chemotherapy agents,
therapy (at least >6 months) and when chemotherapy is such as single-agent carboplatin, or procarbazine, have
contraindicated or has failed. Even in these settings, the also been used in the treatment of recurrent glioblastoma
unclear efficacy of repeat radiotherapy needs to be patients, but given the lack of randomized data, their
weighed against the risk for complications that increases benefit versus supportive care is unclear and at best
with increasing cumulative radiation exposure of normal modest. The application of “tumor-treating fields” via
brain, as discussed in more detail in Chapter 13. Recom- skin electrodes was equally active as physician’s best
mendations regarding dose or radiotherapy technique choice in a phase III trial in patients with recurrent glio-
vary based on the available evidence (Butowski et al., blastoma (Stupp et al., 2012) and was approved by the
2006). Most commonly, 30–36 Gy are applied as frac- FDA, the European Medicines Agency, and in various
tionated stereotactic radiotherapy with or without inten- other countries as a device based on these results
sity modulation (Combs et al., 2007), but no prospective (Stupp et al., 2014b).
GLIOBLASTOMA 391
In summary, therapeutic options at glioblastoma pro- cells (mostly dendritic cells) from peripheral blood, puls-
gression and effects on outcome are limited. Patients ing these cells in vitro, and returning them to the patient.
recruited to controlled trials of recurrent glioblastoma Dendritic cells will then activate T-cell clones and thereby
represent a subselection of patients with more favorable boost specific immune responses (Thomas et al., 2012).
prognosis, whereas best supportive care may be ade- The efficacy of a platform technology utilizing auto-
quate in a substantial fraction of patients who are heavily logous dendritic cells pulsed with the patient’s tumor
impaired already at first recurrence. Symptomatic treat- lysate (DCVax) is currently being evaluated in a double-
ment and psycho-oncologic interventions are therefore a blind phase III trial in patients with unresectable or
key aspect of the treatment of glioblastoma patients that recurrent glioblastoma (NCT00045968). In a phase II trial
is increasingly moving into the focus in individual in patients with newly diagnosed glioblastoma, auto-
patients with disease progression (Weller et al., 2014), logous dendritic cells pulsed with epitopes from six
as discussed in more detail in Chapters 18 and 19. glioblastoma- and glioma-initiating cell-associated pro-
teins (ICT-107) as an adjunct to standard chemoradiother-
apy prolonged progression-free survival by 2.4 months
EXPERIMENTAL APPROACHES
(HR, 0.54; p ¼ 0.006) compared to controls who received
AND OUTLOOK
unpulsed dendritic cells (Wen et al., 2014a). The efficacy
Progress in the molecular characterization of glioblas- of this concept will be explored in a phase III trial in the
toma has led to the identification of a plethora of novel near future. However, immune escape is one of the
treatment targets. Consequently, the traditional hallmarks of glioblastoma, and cytotoxic therapies and
histology-based WHO classification is increasingly com- steroids are likely counterproductive in supporting
plemented by assessment of molecular markers, includ- immune-mediated antitumor responses (Weller et al.,
ing IDH and MGMT status. Taking the heterogeneity of 2015a). Utilizing humanized monoclonal antibodies tar-
glioblastoma into account, a more personalized geting the programmed death protein (PD)1 (nivolumab,
approach to the treatment of glioblastoma is warranted. pembrolizumab) or the cytotoxic T-lymphocyte antigen
A major challenge of molecular marker diagnostics that (CTLA) 4 (ipilimumab) to counteract tumor cell-mediated
may need to be resolved is intratumoral and temporal suppression of the adaptive immune response has proven
heterogeneity of the molecular profile of glioblastoma successful in malignant melanoma (Hodi et al., 2010;
(Sottoriva et al., 2013; Ozawa et al., 2014; Patel Ribas et al., 2013; Wolchok et al., 2013) and nonsmall
et al., 2014). cell lung cancer (Brahmer et al., 2015; Rizvi et al.,
The concept of immunotherapy in treating glioblas- 2015). These agents are currently evaluated in various
toma has been explored for decades, with little or no suc- early trials and one phase III trial in recurrent glioblas-
cess (Wilson, 1979), but has moved back to the focus of toma (NCT02017717); the rationale for these treatments
attention with the identification of promising targets, is supported by the association of low expression of the
technologic progress, and success of early clinical trials. PD1 ligand on blood monocytes with improved survival
The term “tumor vaccination” refers to the activation of among patients with newly diagnosed glioblastoma
the patient’s adaptive immune response against tumor- treated with an autologous heat shock protein peptide
specific antigens, in analogy to classic vaccination vaccine as an adjunct to standard chemoradiotherapy
against virus antigens (discussed in more detail in (Bloch et al., 2015).
Chapter 10). Other emerging targets for vaccination include the
One prominent example of a putative target for vac- tumor-specific epitope of mutant IDH, which evoked
cination is the EGFRvIII deletion mutant, which harbors an adaptive immune response in preclinical studies
a unique epitope that does not occur in normal cells (Schumacher et al., 2014), and the herpesvirus cytomeg-
(Humphrey et al., 1990). Vaccination against this epitope alovirus antigen pp65, which triggered dendritic cell-
with the 13-amino-acid peptide rindopepimut was dem- mediated immune response against glioblastoma cells
onstrated to be effective in triggering an immunologic in a preliminary study of 12 patients (Mitchell et al.,
response against EGFRvIII-positive cells in single-arm 2015). Of note, the immune response in this setting
phase II trials, in patients with newly diagnosed glioblas- was further triggered by tetanus/diphtheria toxoid,
toma who received rindopepimut in addition to standard which may thus be further evaluated as a booster of
chemoradiotherapy (Sampson et al., 2010; Schuster immune responses against glioblastoma cells (Mitchell
et al., 2015). Results from the double-blinded ACT IV et al., 2015).
phase III trial evaluating the efficacy of rindopepimut Other approaches include viruses used both as vectors
are currently awaited (NCT01480479). for somatic gene therapy by targeting molecular path-
Besides utilizing peptides, adaptive immune responses ways that mediate malignancy and as oncolytic viruses
can be evoked by isolating a patient’s antigen-presenting that provoke an inflammatory host response or kill
392 H-G. WIRSCHING ET AL.
glioma cells directly by excessive replication, with uncer- As outlined above, glioma-initiating cells are thought
tain success in clinical trials (Kaufmann and Chiocca, to constitute a subpopulation that mediate recurrence
2014). These concepts are discussed in more detail in and resistance to standard chemoradiotherapy, but to
Chapter 11. date, attempts to target signaling cascades deemed spe-
Other means to target the molecular machinery that cific for glioma-initiating cells have been disappointing
drives malignancy in glioblastoma are small molecules in early clinical trials (Thomas et al., 2014). Other emerg-
that interfere with enzyme functions. IDH point muta- ing treatments include low-intensity, intermediate-
tions are biologically relevant, early glioma-initiating frequency electric fields applied to the head of patients
events that are present in approximately 10% of glioblas- with the use of an electrode device. Although the exact
tomas (Parsons et al., 2008) and thus considered a prime mechanism of action is under active investigation and
therapeutic target. A small-molecule inhibitor of mutant thought to be mediated via mitotic arrest, results of an
IDH function has shown promising preclinical results interim analysis of a phase III trial evaluating this
that require clinical exploration (Rohle et al., 2013). option as an adjunct to standard chemoradiotherapy in
Amplification of EGFR is another central glioma- newly diagnosed glioblastoma were promising (Stupp
associated molecular signature that is mutually exclusive et al., 2014b).
with IDH mutations and occurs in approximately half of In summary, the exponentially growing understand-
all IDH wild-type glioblastomas (Cancer Genome Atlas ing of the molecular mechanisms that drive glioblastoma
Research, 2008; Ozawa et al., 2014). To date, various has not yet translated into survival benefit. Thorough
small-molecule inhibitors targeting EGFR in glioblas- patient selection for clinical trials and probably combina-
toma yielded disappointing results as single-agent ther- tion treatments, oriented at the molecular profile of indi-
apies, but notably, none of these trials selected for vidual tumors, may ultimately improve outcome.
patients with EGFR amplification (Hegi et al., 2012)
and their blood–brain barrier penetration was uniformly REFERENCES
suboptimal. Furthermore, various other RTK share parts
of the downstream signaling pathway with EGFR and Aldape KD, Ballman K, Furth A et al. (2004).
Immunohistochemical detection of EGFRvIII in high
therefore, other RTK may compensate for blocking of
malignancy grade astrocytomas and evaluation of prognos-
EGFR. Yet, combination therapies have not been tic significance. J Neuropathol Exp Neurol 63: 700–707.
explored in cohorts that were preselected based on Arita N, Taneda M, Hayakawa T (1994). Leptomeningeal
molecular characteristics (Hegi et al., 2012). Two dissemination of malignant gliomas. Incidence, diagnosis
early-phase trials in unselected patients with recurrent and outcome. Acta Neurochir (Wien) 126: 84–92.
glioblastoma explored combination treatments of the Awad I, Bay JW, Rogers L (1986). Leptomeningeal metastasis
EGFR inhibitor erlotinib or of the multi-RTK inhibitor from supratentorial malignant gliomas. Neurosurgery 19:
sorafenib in combination with temsirolimus (CCI-779), 247–251.
an inhibitor of the downstream RTK convergence mol- Bady P, Scuiscio D, Diserens AC et al. (2012). MGMT meth-
ecule mammalian target of rapamycine (mTOR), with ylation analysis of glioblastoma on the Infinium methyla-
disappointing results (Lee et al., 2012; Wen et al., tion BeadChip identifies two distinct CpG regions
associated with gene silencing and outcome, yielding a pre-
2014b). Inhibition of mTOR-utilizing temsirolimus was
diction model for comparisons across datasets, tumor
also explored as an adjunct to standard chemoradiother- grades, and CIMP-status. Acta Neuropathol 124: 547–560.
apy in patients with newly diagnosed glioblastoma and Bao S, Wu Q, McLendon RE et al. (2006). Glioma stem cells
an unmethylated MGMT promoter, but preliminary promote radioresistance by preferential activation of the
results are disappointing, too (Wick et al., 2014). DNA damage response. Nature 444: 756–760.
An important conclusion from these and other clini- Batchelor TT, Reardon DA, de Groot JF et al. (2013). Phase III
cal trials utilizing inhibitors of RTK signaling was that randomized trial comparing the efficacy of cediranib as
proving the presence of molecular targets in individual monotherapy, and in combination with lomustine, versus
tumors is necessary in order to accurately assess efficacy lomustine alone in patients with recurrent glioblastoma.
of these targeted agents. Examples of ongoing clinical J Clin Oncol 31: 3212–3218.
trials that apply such molecular entry controls include Batchelor TT, Reardon DA, de Groot JF et al. (2014).
Antiangiogenic therapy for glioblastoma: current status
a phase II trial in newly diagnosed glioblastoma with
and future prospects. Clin Cancer Res 20: 5612–5619.
activating fibroblast growth factor fusion proteins and Beier D, Rohrl S, Pillai DR et al. (2008). Temozolomide pref-
activating mutations (NCT01975701), which are present erentially depletes cancer stem cells in glioblastoma.
in less than 5% of all glioblastomas, or trials in recurrent Cancer Res 68: 5706–5715.
glioblastoma utilizing an inhibitor of PI3K alone or Bloch O, Han SJ, Cha S et al. (2012). Impact of extent of resec-
in combinations with different antiangiogenic agents tion for recurrent glioblastoma on overall survival: clinical
(NCT01339052, NCT01870726, NCT01349660). article. J Neurosurg 117: 1032–1038.
GLIOBLASTOMA 393
Bloch O, Raizer JJ, Lim M et al. (2015). Newly diagnosed glio- El Ali A, Hermann DM (2011). ATP-binding cassette trans-
blastoma patients treated with an autologous heat shock porters and their roles in protecting the brain.
protein peptide vaccine: PD-L1 expression and response Neuroscientist 17: 423–436.
to therapy. J Clin Oncol (suppl): abstr 2011. Ellingson BM, Zaw T, Cloughesy TF et al. (2012).
Bondy ML, Scheurer ME, Malmer B et al. (2008). Brain tumor Comparison between intensity normalization techniques
epidemiology: consensus from the Brain Tumor for dynamic susceptibility contrast (DSC)-MRI estimates
Epidemiology Consortium. Cancer 113: 1953–1968. of cerebral blood volume (CBV) in human gliomas.
Brada M, Ford D, Ashley S et al. (1992). Risk of second brain J Magn Reson Imaging 35: 1472–1477.
tumour after conservative surgery and radiotherapy for Ellingson BM, Lai A, Harris RJ et al. (2013). Probabilistic
pituitary adenoma. Br Med J 304: 1343–1346. radiographic atlas of glioblastoma phenotypes. AJNR
Brada M, Stenning S, Gabe R et al. (2010). Temozolomide ver- Am J Neuroradiol 34: 533–540.
sus procarbazine, lomustine, and vincristine in recurrent Ellis LM, Hicklin DJ (2008). VEGF-targeted therapy: mecha-
high-grade glioma. J Clin Oncol 28: 4601–4608. nisms of anti-tumour activity. Nat Rev Cancer 8: 579–591.
Brahmer J, Reckamp KL, Baas P et al. (2015). Nivolumab Eriksson PS, Perfilieva E, Bjork-Eriksson T et al. (1998).
versus docetaxel in advanced squamous-cell non-small- Neurogenesis in the adult human hippocampus. Nat Med
cell lung cancer. N Engl J Med 373: 123–135. 4: 1313–1317.
Brandsma D, Stalpers L, Taal W et al. (2008). Clinical fea- Erlich SS, Davis RL (1978). Spinal subarachnoid metastasis
tures, mechanisms, and management of pseudoprogression from primary intracranial glioblastoma multiforme.
in malignant gliomas. Lancet Oncol 9: 453–461. Cancer 42: 2854–2864.
Butowski NA, Sneed PK, Chang SM (2006). Diagnosis and Fan QW, Cheng CK, Gustafson WC et al. (2013). EGFR phos-
treatment of recurrent high-grade astrocytoma. J Clin phorylates tumor-derived EGFRvIII driving STAT3/5 and
Oncol 24: 1273–1280. progression in glioblastoma. Cancer Cell 24: 438–449.
Cancer Genome Atlas Research (2008). Comprehensive geno- Fecci PE, Heimberger AB, Sampson JH (2014). Immunotherapy
mic characterization defines human glioblastoma genes for primary brain tumors: no longer a matter of privilege. Clin
and core pathways. Nature 455: 1061–1068. Cancer Res 20: 5620–5629.
Capper D, Zentgraf H, Balss J et al. (2009). Monoclonal anti- Folkman J (1971). Tumor angiogenesis: therapeutic implica-
body specific for IDH1 R132H mutation. Acta Neuropathol tions. N Engl J Med 285: 1182–1186.
118: 599–601. Friedman HS, Prados MD, Wen PY et al. (2009). Bevacizumab
Chaichana KL, Parker SL, Olivi A et al. (2009). Long-term alone and in combination with irinotecan in recurrent glio-
seizure outcomes in adult patients undergoing primary blastoma. J Clin Oncol 27: 4733–4740.
resection of malignant brain astrocytomas. Clinical article. Gallego Perez-Larraya J, Ducray F, Chinot O et al. (2011).
J Neurosurg 111: 282–292. Temozolomide in elderly patients with newly diagnosed
Chen J, Li Y, Yu TS et al. (2012). A restricted cell population glioblastoma and poor performance status: an ANOCEF
propagates glioblastoma growth after chemotherapy. phase II trial. J Clin Oncol 29: 3050–3055.
Nature 488: 522–526. Gilbert MR, Wang M, Aldape KD et al. (2013). Dose-dense
Cheng L, Huang Z, Zhou W et al. (2013). Glioblastoma stem temozolomide for newly diagnosed glioblastoma: a ran-
cells generate vascular pericytes to support vessel function domized phase III clinical trial. J Clin Oncol 31: 4085–4091.
and tumor growth. Cell 153: 139–152. Gilbert MR, Dignam JJ, Armstrong TS et al. (2014).
Chinot OL, Wick W, Mason M et al. (2014). Bevacizumab plus A randomized trial of bevacizumab for newly diagnosed
radiotherapy-temozolomide for newly diagnosed glioblas- glioblastoma. N Engl J Med 370: 699–708.
toma. N Engl J Med 370: 709–722. Gilbertson RJ, Rich JN (2007). Making a tumour’s bed: glio-
Collignon FP, Holland EC, Feng S (2004). Organ donors with blastoma stem cells and the vascular niche. Nat Rev Cancer
malignant gliomas: an update. Am J Transplant 4: 15–21. 7: 733–736.
Combs SE, Debus J, Schulz-Ertner D (2007). Radiotherapeutic Grabb PA, Albright AL, Pang D (1992). Dissemination of
alternatives for previously irradiated recurrent gliomas. supratentorial malignant gliomas via the cerebrospinal
BMC Cancer 7: 167. fluid in children. Neurosurgery 30: 64–71.
Connelly JM, Malkin MG (2007). Environmental risk factors Grothey A, Galanis E (2009). Targeting angiogenesis: pro-
for brain tumors. Curr Neurol Neurosci Rep 7: 208–214. gress with anti-VEGF treatment with large molecules.
de Wit MC, de Bruin HG, Eijkenboom W et al. (2004). Nat Rev Clin Oncol 6: 507–518.
Immediate post-radiotherapy changes in malignant glioma Han SJ, Yang I, Tihan T et al. (2010). Primary gliosarcoma: key
can mimic tumor progression. Neurology 63: 535–537. clinical and pathologic distinctions from glioblastoma with
Dhermain F, de Crevoisier R, Parker F et al. (2004). Role of implications as a unique oncologic entity. J Neurooncol 96:
radiotherapy in recurrent gliomas. Bull Cancer 91: 883–889. 313–320.
Djalilian HR, Shah MV, Hall WA (1999). Radiographic inci- Hegi ME, Diserens AC, Gorlia T et al. (2005). MGMT gene
dence of multicentric malignant gliomas. Surg Neurol 51: silencing and benefit from temozolomide in glioblastoma.
554–557. discussion 557–558. N Engl J Med 352: 997–1003.
Dziurzynski K, Chang SM, Heimberger AB et al. (2012). Hegi ME, Rajakannu P, Weller M (2012). Epidermal growth
Consensus on the role of human cytomegalovirus in factor receptor: a re-emerging target in glioblastoma.
glioblastoma. Neuro Oncol 14: 246–255. Curr Opin Neurol 25: 774–779.
394 H-G. WIRSCHING ET AL.
Heimberger AB, Hlatky R, Suki D et al. (2005). Prognostic Labussiere M, Di Stefano AL, Gleize V et al. (2014). TERT pro-
effect of epidermal growth factor receptor and EGFRvIII moter mutations in gliomas, genetic associations and clinico-
in glioblastoma multiforme patients. Clin Cancer Res 11: pathological correlations. Br J Cancer 111: 2024–2032.
1462–1466. Lacroix M, Abi-Said D, Fourney DR et al. (2001).
Herrlinger U, Forschler H, Kuker W et al. (2004). A multivariate analysis of 416 patients with glioblastoma
Leptomeningeal metastasis: survival and prognostic fac- multiforme: prognosis, extent of resection, and survival.
tors in 155 patients. J Neurol Sci 223: 167–178. J Neurosurg 95: 190–198.
Hodi FS, O’Day SJ, McDermott MF et al. (2010). Improved Lai A, Tran A, Nghiemephu PL et al. (2011). Phase II study of
survival with ipilimumab in patients with metastatic mela- bevacizumab plus temozolomide during and after radiation
noma. N Engl J Med 363: 711–723. therapy for patients with newly diagnosed glioblastoma
Humphrey PA, Wong AJ, Vogelstein B et al. (1990). Anti- multiforme. J Clin Oncol 29: 142–148.
synthetic peptide antibody reacting at the fusion junction of Laperriere N, Weller M, Stupp R et al. (2013). Optimal man-
deletion-mutant epidermal growth factor receptors in human agement of elderly patients with glioblastoma. Cancer
glioblastoma. Proc Natl Acad Sci U S A 87: 4207–4211. Treat Rev 39: 350–357.
Interphone Study Group (2010). Brain tumour risk in relation to Lathia JD, Gallagher J, Heddleston JM et al. (2010). Integrin
mobile telephone use: results of the INTERPHONE inter- alpha 6 regulates glioblastoma stem cells. Cell Stem Cell 6:
national case-control study. Int J Epidemiol 39: 675–694. 421–432.
Karsy M, Gelbman M, Shah P et al. (2012). Established and Law M, Yang S, Wang H et al. (2003). Glioma grading: sen-
emerging variants of glioblastoma multiforme: review of sitivity, specificity, and predictive values of perfusion MR
morphological and molecular features. Folia Neuropathol imaging and proton MR spectroscopic imaging compared
50: 301–321. with conventional MR imaging. AJNR Am J Neuroradiol
Kaufmann JK, Chiocca EA (2014). Glioma virus therapies 24: 1989–1998.
between bench and bedside. Neuro Oncol 16: 334–351. Lee EQ, Kuhn J, Lamborn KR et al. (2012). Phase I/II study of
Keime-Guibert F, Chinot O, Taillandier L et al. (2007). sorafenib in combination with temsirolimus for recurrent
Radiotherapy for glioblastoma in the elderly. N Engl glioblastoma or gliosarcoma: North American Brain
J Med 356: 1527–1535. Tumor Consortium study 05-02. Neuro Oncol 14: 1511–1518.
Kerkhof M, Dielemans JC, van Breemen MS et al. (2013). Lindsay A, Holthouse D, Robbins P et al. (2002). Spinal lep-
Effect of valproic acid on seizure control and on survival tomeningeal metastases following glioblastoma multi-
in patients with glioblastoma multiforme. Neuro Oncol forme treated with radiotherapy. J Clin Neurosci 9:
15: 961–967. 725–728.
Kickingereder P, Wiestler B, Sahm F et al. (2014). Primary Louis DN, Ohgaki H, Wiestler OD et al. (2007). The 2007
central nervous system lymphoma and atypical glioblas- WHO classification of tumours of the central nervous sys-
toma: multiparametric differentiation by using diffusion-, tem. Acta Neuropathol 114: 97–109.
perfusion-, and susceptibility-weighted MR imaging. Macdonald DR, Cascino TL, Schold Jr SC et al. (1990).
Radiology 272: 843–850. Response criteria for phase II studies of supratentorial
Kono K, Inoue Y, Nakayama K et al. (2001). The role of malignant glioma. J Clin Oncol 8: 1277–1280.
diffusion-weighted imaging in patients with brain tumors. Malmstrom A, Gronberg BH, Marosi C et al. (2012).
AJNR Am J Neuroradiol 22: 1081–1088. Temozolomide versus standard 6-week radiotherapy ver-
Korshunov A, Sycheva R, Golanov A et al. (2007). Gains at the sus hypofractionated radiotherapy in patients older than
1p36 chromosomal region are associated with symptomatic 60 years with glioblastoma: the Nordic randomised, phase
leptomeningeal dissemination of supratentorial glioblasto- 3 trial. Lancet Oncol 13: 916–926.
mas. Am J Clin Pathol 127: 585–590. Mashimo T, Pichumani K, Vemireddy V et al. (2014). Acetate
Kozak KR, Moody JS (2009). Giant cell glioblastoma: a glio- is a bioenergetic substrate for human glioblastoma and
blastoma subtype with distinct epidemiology and superior brain metastases. Cell 159: 1603–1614.
prognosis. Neuro Oncol 11: 833–841. Minniti G, Traish D, Ashley S et al. (2005). Risk of second
Kozak KR, Mahadevan A, Moody JS (2009). Adult gliosar- brain tumor after conservative surgery and radiotherapy
coma: epidemiology, natural history, and factors associated for pituitary adenoma: update after an additional 10 years.
with outcome. Neuro Oncol 11: 183–191. J Clin Endocrinol Metab 90: 800–804.
Kreisl TN, Kim L, Moore K et al. (2009). Phase II trial of Mitchell DA, Batich KA, Gunn MD et al. (2015). Tetanus tox-
single-agent bevacizumab followed by bevacizumab plus oid and CCL3 improve dendritic cell vaccines in mice and
irinotecan at tumor progression in recurrent glioblastoma. glioblastoma patients. Nature 519: 366–369.
J Clin Oncol 27: 740–745. Neglia JP, Robison LL, Stovall M et al. (2006). New primary
Kubben PL, ter Meulen KJ, Schijns OE et al. (2011). neoplasms of the central nervous system in survivors of
Intraoperative MRI-guided resection of glioblastoma mul- childhood cancer: a report from the Childhood Cancer
tiforme: a systematic review. Lancet Oncol 12: 1062–1070. Survivor Study. J Natl Cancer Inst 98: 1528–1537.
La Fougere C, Suchorska B, Bartenstein P et al. (2011). Nghiemphu PL, Liu W, Lee Y et al. (2009). Bevacizumab and
Molecular imaging of gliomas with PET: opportunities chemotherapy for recurrent glioblastoma: a single-
and limitations. Neuro Oncol 13: 806–819. institution experience. Neurology 72: 1217–1222.
GLIOBLASTOMA 395
Nishikawa R, Sugiyama T, Narita Y et al. (2004). Ricci-Vitiani L, Pallini R, Biffoni M et al. (2010). Tumour vas-
Immunohistochemical analysis of the mutant epidermal cularization via endothelial differentiation of glioblastoma
growth factor, deltaEGFR, in glioblastoma. Brain stem-like cells. Nature 468: 824–828.
Tumor Pathol 21: 53–56. Rizvi NA, Mazieres J, Planchard D et al. (2015). Activity and
Noushmehr H, Weisenberger DJ, Diefes K et al. (2010). safety of nivolumab, an anti-PD-1 immune checkpoint
Identification of a CpG island methylator phenotype that inhibitor, for patients with advanced, refractory squamous
defines a distinct subgroup of glioma. Cancer Cell 17: non-small-cell lung cancer (CheckMate 063): a phase 2,
510–522. single-arm trial. Lancet Oncol 16: 257–265.
Nunes MC, Roy NS, Keyoung HM et al. (2003). Identification Roa W, Brasher PM, Bauman G et al. (2004). Abbreviated
and isolation of multipotential neural progenitor cells from course of radiation therapy in older patients with glioblas-
the subcortical white matter of the adult human brain. Nat toma multiforme: a prospective randomized clinical trial.
Med 9: 439–447. J Clin Oncol 22: 1583–1588.
Ohgaki H, Kim YH, Steinbach JP (2010). Nervous system Rohle D, Popovici-Muller J, Palaskas N et al. (2013). An inhib-
tumors associated with familial tumor syndromes. Curr itor of mutant IDH1 delays growth and promotes differen-
Opin Neurol 23: 583–591. tiation of glioma cells. Science 340: 626–630.
Ostrom QT, Barnholtz-Sloan JS (2011). Current state of our Ron E, Modan B, Boice Jr JD et al. (1988). Tumors of the brain
knowledge on brain tumor epidemiology. Curr Neurol and nervous system after radiotherapy in childhood. N Engl
Neurosci Rep 11: 329–335. J Med 319: 1033–1039.
Ostrom QT, Gittleman H, Liao P et al. (2014a). CBTRUS sta- Rygh OM, Selbekk T, Torp SH et al. (2008). Comparison of
tistical report: primary brain and central nervous system navigated 3D ultrasound findings with histopathology in
tumors diagnosed in the United States in 2007–2011. subsequent phases of glioblastoma resection. Acta
Neuro Oncol 16 (Suppl 4): iv1–iv63. Neurochir (Wien) 150: 1033–1041. discussion 1042.
Ostrom QT, Bauchet L, Davis FG et al. (2014b). The epidemi- Sadetzki S, Chetrit A, Freedman L et al. (2005). Long-term
ology of glioma in adults: a “state of the science” review. follow-up for brain tumor development after childhood
Neuro Oncol 16: 896–913. exposure to ionizing radiation for tinea capitis. Radiat
Ostrom QT, de Blank PM, Kruchko C et al. (2015). Alex’s Res 163: 424–432.
Lemonade Stand Foundation infant and childhood primary Sampson JH, Heimberger AB, Archer GE et al. (2010).
brain and central nervous system tumors diagnosed in the Immunologic escape after prolonged progression-free sur-
United States in 2007–2011. Neuro Oncol 16 (Suppl 10): vival with epidermal growth factor receptor variant III pep-
x1–x36. tide vaccination in patients with newly diagnosed
Ozawa T, Riester M, Cheng YK et al. (2014). Most human glioblastoma. J Clin Oncol 28: 4722–4729.
non-GCIMP glioblastoma subtypes evolve from a common Sanai N, Tramontin AD, Quinones-Hinojosa A et al. (2004).
proneural-like precursor glioma. Cancer Cell 26: 288–300. Unique astrocyte ribbon in adult human brain contains neu-
Park JK, Hodges T, Arko L et al. (2010). Scale to predict sur- ral stem cells but lacks chain migration. Nature 427:
vival after surgery for recurrent glioblastoma multiforme. 740–744.
J Clin Oncol 28: 3838–3843. Sanai N, Alvarez-Buylla A, Berger MS (2005). Neural stem
Parsons DW, Jones S, Zhang X et al. (2008). An integrated cells and the origin of gliomas. N Engl J Med 353: 811–822.
genomic analysis of human glioblastoma multiforme. Sanai N, Polley MY, McDermott MW et al. (2011). An extent
Science 321: 1807–1812. of resection threshold for newly diagnosed glioblastomas.
Patel AP, Tirosh I, Trombetta JJ et al. (2014). Single-cell J Neurosurg 115: 3–8.
RNA-seq highlights intratumoral heterogeneity in primary Sandmann T, Bourgon R, Garcia J et al. (2015). Patients with
glioblastoma. Science 344: 1396–1401. proneural glioblastoma may derive overall survival benefit
Phillips HS, Kharbanda S, Chen R et al. (2006). Molecular sub- from the addition of bevacizumab to first-line radiotherapy
classes of high-grade glioma predict prognosis, delineate a and temozolomide: retrospective analysis of the AVAglio
pattern of disease progression, and resemble stages in neu- trial. J Clin Oncol 33: 2735–2744.
rogenesis. Cancer Cell 9: 157–173. Schumacher T, Bunse L, Pusch S et al. (2014). A vaccine tar-
Pietras A, Katz AM, Ekstrom EJ et al. (2014). Osteopontin- geting mutant IDH1 induces antitumour immunity. Nature
CD44 signaling in the glioma perivascular niche enhances 512: 324–327.
cancer stem cell phenotypes and promotes aggressive Schuster J, Lai RK, Recht LD et al. (2015). A phase II, multi-
tumor growth. Cell Stem Cell 14: 357–369. center trial of rindopepimut (CDX-110) in newly diag-
Reardon DA, Schuster JM, Tran DD et al. (2015). 107 ReACT: nosed glioblastoma: the ACT III study. Neuro Oncol 17:
overall survival from a randomized phase II study of rindo- 854–861.
pepimut (CDX-110) plus bevacizumab in relapsed glio- Schweitzer T, Vince GH, Herbold C et al. (2001). Extraneural
blastoma. Neurosurgery 62 (Suppl 1): 198–199. metastases of primary brain tumors. J Neurooncol 53:
Ribas A, Kefford R, Marshall M et al. (2013). Phase III ran- 107–114.
domized clinical trial comparing tremelimumab with Scott J, Tsai YY, Chinnaiyan P et al. (2011). Effectiveness of
standard-of-care chemotherapy in patients with advanced radiotherapy for elderly patients with glioblastoma. Int
melanoma. J Clin Oncol 31: 616–622. J Radiat Oncol Biol Phys 81: 206–210.
396 H-G. WIRSCHING ET AL.
Shinojima N, Tada K, Shiraishi S et al. (2003). Prognostic Taphoorn MJ, Henriksson R, Bottomley M et al. (2015). Health-
value of epidermal growth factor receptor in patients with related quality of life in a randomized phase III study of
glioblastoma multiforme. Cancer Res 63: 6962–6970. bevacizumab, temozolomide, and radiotherapy in newly
Simpson JR, Horton J, Scott C et al. (1993). Influence of loca- diagnosed glioblastoma. J Clin Oncol 33: 2166–2175.
tion and extent of surgical resection on survival of patients Thomas AA, Ernstoff MS, Fadul CE (2012). Immunotherapy
with glioblastoma multiforme: results of three consecutive for the treatment of glioblastoma. Cancer J 18: 59–68.
Radiation Therapy Oncology Group (RTOG) clinical trials. Thomas AA, Brennan CW, DeAngelis LM et al. (2014).
Int J Radiat Oncol Biol Phys 26: 239–244. Emerging therapies for glioblastoma. JAMA Neurol 71:
Singh SK, Hawkins C, Clarke ID et al. (2004). Identification of 1437–1444.
human brain tumour initiating cells. Nature 432: 396–401. Turcan S, Rohle D, Goenka A et al. (2012). IDH1 mutation is
Son MJ, Woolard K, Nam DH et al. (2009). SSEA-1 is an sufficient to establish the glioma hypermethylator pheno-
enrichment marker for tumor-initiating cells in human glio- type. Nature 483: 479–483.
blastoma. Cell Stem Cell 4: 440–452. van Breemen MS, Rijisman RM, Taphoorn MJ et al. (2009).
Sottoriva A, Spiteri I, Piccirillo SG et al. (2013). Intratumor Efficacy of anti-epileptic drugs in patients with gliomas
heterogeneity in human glioblastoma reflects cancer and seizures. J Neurol 256: 1519–1526.
evolutionary dynamics. Proc Natl Acad Sci U S A 110: Verhaak RG, Hoadley KA, Purdom E et al. (2010). Integrated
4009–4014. genomic analysis identifies clinically relevant subtypes of
Stummer W, Pichlmeier U, Meinel T et al. (2006). glioblastoma characterized by abnormalities in PDGFRA,
Fluorescence-guided surgery with 5-aminolevulinic acid IDH1, EGFR, and NF1. Cancer Cell 17: 98–110.
for resection of malignant glioma: a randomised controlled Wang R, Chadalavada K, Wilshire J et al. (2010).
multicentre phase III trial. Lancet Oncol 7: 392–401. Glioblastoma stem-like cells give rise to tumour endothe-
Stummer W, Reulen HJ, Meinel T et al. (2008). Extent of lium. Nature 468: 829–833.
resection and survival in glioblastoma multiforme: identi- Warrens AN, Birch R, Collett D et al. (2012). Advising poten-
fication of and adjustment for bias. Neurosurgery 62: tial recipients on the use of organs from donors with pri-
564–576. discussion 564–576. mary central nervous system tumors. Transplantation 93:
Stupp R, Mason WP, van den Bent MJ et al. (2005). 348–353.
Radiotherapy plus concomitant and adjuvant temozolo- Watson CJ, Roberts R, Wright KA et al. (2010). How safe is it
mide for glioblastoma. N Engl J Med 352: 987–996. to transplant organs from deceased donors with primary
Stupp R, Wong ET, Kanner AA et al. (2012). NovoTTF-100A intracranial malignancy? An analysis of UK Registry data.
versus physician’s choice chemotherapy in recurrent glio- Am J Transplant 10: 1437–1444.
blastoma: a randomised phase III trial of a novel treatment Weber MA, Zoubaa S, Schlieter M et al. (2006). Diagnostic
modality. Eur J Cancer 48: 2192–2202. performance of spectroscopic and perfusion MRI for dis-
Stupp R, Hegi MF, Gorlia T et al. (2014a). Cilengitide com- tinction of brain tumors. Neurology 66: 1899–1906.
bined with standard treatment for patients with newly diag- Weller M, Stupp R, Reifenberger G et al. (2010). MGMT pro-
nosed glioblastoma with methylated MGMT promoter moter methylation in malignant gliomas: ready for person-
(CENTRIC EORTC 26071-22072 study): a multicentre, alized medicine? Nat Rev Neurol 6: 39–51.
randomised, open-label, phase 3 trial. Lancet Oncol 15: Weller M, Cloughesy T, Perry JR et al. (2013). Standards of
1100–1108. care for treatment of recurrent glioblastoma – are we there
Stupp R, Wong E, Scott C et al. (2014b). Interim analysis of the yet? Neuro Oncol 15: 4–27.
EF-14 trial: a prospective, multi-center trial of NovoTTF- Weller M, van den Bent M, Hopkins K et al. (2014). EANO
100A together with temozolomide compared to temozolo- guideline for the diagnosis and treatment of anaplastic gli-
mide alone in patients with newly diagnosed GBM. Neuro omas and glioblastoma. Lancet Oncol 15: e395–e403.
Oncol 16: v167. Weller M, Wick W, Aldape K et al. (2015a). Glioma. Nat Rev
Sturm D, Witt H, Hovestadt V et al. (2012). Hotspot mutations Dis Primers.
in H3F3A and IDH1 define distinct epigenetic and biolog- Weller M, Tabatabai G, Kastner B et al. (2015b). MGMT pro-
ical subgroups of glioblastoma. Cancer Cell 22: 425–437. moter methylation is a strong prognostic biomarker for
Suchorska B, Weller M, Tabatabai G et al. (2015). Complete benefit from dose-intensified temozolomide rechallenge
resection of contrast-enhancing tumor volume is associated in progressive glioblastoma: the DIRECTOR trial. Clin
with improved survival in recurrent glioblastoma results Cancer Res 21: 2057–2064.
from the DIRECTOR trial. Neurosurgery 62 (Suppl 1): 209. Wen PY, Macdonald DR, Reardon DA et al. (2010). Updated
Taal W, Oosterkamp HM, Walenkamp AM et al. (2014). response assessment criteria for high-grade gliomas:
Single-agent bevacizumab or lomustine versus a combina- response assessment in neuro-oncology working group.
tion of bevacizumab plus lomustine in patients with recur- J Clin Oncol 28: 1963–1972.
rent glioblastoma (BELOB trial): a randomised controlled Wen P, Reardon D, Phuphanich S et al. (2014a). A randomized
phase 2 trial. Lancet Oncol 15: 943–953. double blind placebo-controlled phase 2 trial of dendritic
Taphoorn MJ, Stupp R, Coens C et al. (2005). Health-related cell (DC) vaccine ICT-107 following standard treatment
quality of life in patients with glioblastoma: a randomised in newly diagnosed patients with GBM. Neuro Oncol 16:
controlled trial. Lancet Oncol 6: 937–944. v8–v22.
GLIOBLASTOMA 397
Wen PY, Chang SM, Lamborn KR et al. (2014b). Phase I/II Wick W, Gorlia T, van den Bent M et al. (2014). Radiation
study of erlotinib and temsirolimus for patients with recur- therapy and concurrent plus adjuvant temsirolimus (CCI-
rent malignant gliomas: North American Brain Tumor 779) versus chemoirradiation with temozolomide in newly
Consortium trial 04-02. Neuro Oncol 16: 567–578. diagnosed glioblastoma without methylation of the MGMT
Wick W, Platten M (2014). CMV infection and glioma, a gene promoter. J Clin Oncol 32: 5s.
highly controversial concept struggling in the clinical Wilson CB (1979). Current concepts in cancer: brain tumors.
arena. Neuro Oncol 16: 332–333. N Engl J Med 300: 1469–1471.
Wick W, Menn O, Meisner C et al. (2005). Pharmacotherapy Wolchok JD, Kluger H, Callahan MK et al. (2013). Nivolumab
of epileptic seizures in glioma patients: who, when, why plus ipilimumab in advanced melanoma. N Engl J Med 369:
and how long? Onkologie 28: 391–396. 122–133.
Wick W, Puduvalli VK, Chamberlain MC et al. (2010). Phase Yan H, Parsons DW, Jing G et al. (2009). IDH1 and IDH2
III study of enzastaurin compared with lomustine in the mutations in gliomas. N Engl J Med 360: 765–773.
treatment of recurrent intracranial glioblastoma. J Clin Yan K, Yang K, Rich JN (2013). The evolving landscape of
Oncol 28: 1168–1174. glioblastoma stem cells. Curr Opin Neurol 26: 701–707.
Wick W, Platten M, Meisner C et al. (2012). Temozolomide Yuile P, Dent O, Cook R et al. (2006). Survival of glioblastoma
chemotherapy alone versus radiotherapy alone for malig- patients related to presenting symptoms, brain site and
nant astrocytoma in the elderly: the NOA-08 randomised, treatment variables. J Clin Neurosci 13: 747–751.
phase 3 trial. Lancet Oncol 13: 707–715.

You might also like