You are on page 1of 39

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

net/publication/310050270

Molecular Biology of Human Brain Tumors

Chapter · November 2017


DOI: 10.1007/978-1-59745-458-2_35

CITATIONS READS

0 318

6 authors, including:

Daniel Coluccia Javier Fandino


Luzerner Kantonsspital Kantonspital Aarau AG
51 PUBLICATIONS   1,081 CITATIONS    280 PUBLICATIONS   2,690 CITATIONS   

SEE PROFILE SEE PROFILE

Christian Schneider
Universitätsspital Basel
14 PUBLICATIONS   194 CITATIONS   

SEE PROFILE

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

Book chapter View project

Mild traumatic brain injury View project

All content following this page was uploaded by Daniel Coluccia on 16 August 2019.

The user has requested enhancement of the downloaded file.


Molecular Biology of Human Brain
Tumors 35
Daniel Coluccia, Adrienne Weeks, Javier Fandino,
Christian Schneider, Christian Smith, and James T. Rutka

tumor. For example, brainstem gliomas, medulloblastomas or


35.1 Introduction intracranial ependymomas are usually found in children and
adolescents [1]. Even though children constitute only 7 % of
The Central Brain Tumor Registry of the USA estimates the all CNS tumors, these tumors comprise 25 % of all cancers in
annual incidence of primary brain and central nervous system infants and are the most frequent solid tumor and the second
(CNS) tumors at 7.3/100,000 for malignant and 13.3/100,000 most common neoplasm in their age group after leukemia [1,
for nonmalignant tumors [1]. A true comparison of incidence 5]. The only established environmental risk factor is exposure
numbers across different time periods and countries is diffi- to elevated ionizing radiation [2, 3]. Other environmental car-
cult given the inconsistency of data collection methods and cinogens possibly related to increased brain tumor formation
the diversity in diagnostic criteria. Available diagnosis rates have been examined, such as n-nitroso compounds and poly-
for malignant gliomas spanning the last three decades do not cyclic aromatic hydrocarbons, but no correlation was found
allow the determination of an increase or decline of age [2, 3]. Nonionizing radiation such as cellular phones or smok-
adapted incidence [2, 3], despite remarkable changes of ing have also been extensively evaluated, but no reliable causal
environment and lifestyle in industrial nations. link has been established [2, 3]. Chronic viral infections,
The etiology of most brain tumors is largely unknown. In which are recognized as important risk factors for various can-
gliomas, familial clustering is found in about 5 % and in 1 % cers [6], are also implicated as triggering processes in brain
an autosomal dominant inheritance is assumed [4]. Besides tumors. Studies of persistent immunohistochemical detection
family history, demographic risk factors listed are age and of cytomegalovirus (CMV) in glioblastoma and the benefit of
gender. Most brain tumors such as gliomas occur in older adding antiviral drugs to the standard treatment protocols are
adults with a slight predominance in males [1]. However, this still premature, but importantly remain part of an ongoing
might be different depending on origin and location of the debate into the etiology of some gliomas [7, 8].
The term brain tumor applies to cancers that arise in
different parenchymal layers and tissues from various cellu-
D. Coluccia, M.D. lar origins within the central nervous system. The 2007
Division of Neurosurgery, The Arthur and Sonia Labatt Brain
Tumor Research Centre, The Hospital for Sick Children,
WHO classification identifies seven categories of brain
University of Toronto, Toronto, ON, Canada tumors, depending on the presumed histogenetic origin: (1)
Department of Neurosurgery, Cantonal Hospital of Aarau,
neuroepithelial, (2) cranial and paraspinal nerves, (3) menin-
Aarau, Switzerland geal and mesenchymal, (4) lymphomas and hematopoietic
A. Weeks, Ph.D., M.D., F.R.C.S.C.
neoplasms, (5) germ cell tumors, (6) sellar region, and (7)
Division of Neurosurgery, Dalhousie University, metastases [9].
Halifax, NS, Canada In this chapter we will focus on neuroepithelial tumors
J. Fandino, M.D. derived from central neuroglia, which have an annual inci-
Department of Neurosurgery, Cantonal Hospital of Aarau, dence of 6.6/100,000 and account for the most malignant
Aarau, Switzerland forms of primary brain tumors [1]. We review the basic biol-
C. Schneider, M.D. • C. Smith ogy of human brain tumors with a special focus placed on
J.T. Rutka, M.D., Ph.D., F.R.C.S.C., F.A.C.S., F.A.A.P. (*) the most common primary brain tumors in adults and children
Division of Neurosurgery, The Arthur and Sonia Labatt Brain
Tumor Research Centre, The Hospital for Sick Children,
including the astrocytoma and medulloblastoma (Fig. 35.1).
University of Toronto, Toronto, ON, Canada These and other human brain tumors will be discussed insofar
e-mail: james.rutka@sickkids.ca as their relationships to the familial predisposition syndromes,

© Springer Science+Business Media New York 2017 657


W.B. Coleman, G.J. Tsongalis (eds.), The Molecular Basis of Human Cancer, DOI 10.1007/978-1-59745-458-2_35
658 D. Coluccia et al.

Fig. 35.1 MRI sequences of common primary brain tumors. (a) Sagittal (c) Sagital T1-weighted and gadolinium-enhanced MR image depicting
T1-weighted and gadolinium-enhanced MR image showing a medullo- a slightly and inhomogenously enhancing low grade astrocytoma in the
blastoma in the posterior fossa (arrow) with compression of the pons thalamic and mesencephalic region (arrow); (d) Axial T1-weighted and
and medulla oblongata; (b) Axial T2-weighted MR image showing a gadolinium-enhanced MR image showing a centrally located
space occupying ependymoma (hyperintense signaling, arrow) in the left Glioblastoma (arrow) in the region of basal ganglia resulting in dis-
paramedian posterior fossa resulting in displacement of the brain stem; placement of the pyramidal tracts (brown fringe).

cytogenetic alterations, dysregulation of the cell cycle remarkable differences in prognosis for tumors harboring
machinery, tumor cell migration and invasion, angiogenesis, similar histopathological patterns [10, 11]. Currently, the
developmental signaling pathways, brain tumor stem cells, utilization of some molecular markers informs individual-
and preclinical animal models. ized treatment decisions for patients with the same histo-
The management strategies of human brain tumors rely logical tumor [10]. Future revisions of brain tumors
mostly on histopathological characteristics that are used to classification systems need to implement molecular profiles
predict the clinical behavior and eventual patient prognosis in order to support the differential diagnosis and to be of
of a diagnosed brain tumor. Features including degree of value in predicting the efficacy of treatment options.
anaplasia, mitosis, infiltration, necrosis, and neovasculariza- Despite concerted efforts in the field of neuro-oncology,
tion define the corresponding tumor grade, which ranges brain tumors represent perhaps the most intimidating and
from benign, WHO grade I, to the most malignant, WHO difficult to treat type of cancer. Compared to other neoplasms,
grade IV [9]. The advances in molecular diagnostics over patient with malignant gliomas still face disproportionately
the last 5 years however have led to rapid genetic profiling high rates of morbidity and mortality [12] (Table 35.1).
of many brain tumors, which allow us now to recognize Fortunately, recent advances in molecular biology have
Table 35.1 Clinical characteristics of common primary brain tumors

35
Preferential Annual WHO

Molecular Biology of Human Brain Tumors


Tumor location Age Gender incidence grade Present treatment strategy Prognosis
Medulloblastoma Cerebellar Median: 8 y m: f ~ 1.6:1 0.6/100,000 Grade IV – Maximal safe resection followed Pediatric:
[1, 424–427] 93 % < 15 y (age 1–9) by C (<3 y) and RT (>3 y) – 5-y survival
– High-dose C and autologous • 42 % (infants < 1 y)
stem cell transplantation with or • 72 % (1–9 y, including high- and
without RT to be considered, average-risk (ar))
especially for high-risk group – 10-y survival for ar group:
– Re-resection at progression • 81 %
followed by C with or without Adults:
RT – 5-y survival
• 67 % WNT group correlates with
good prognosis (>90 %
long-term survivals)
Intracranial ependymoma Infratentorial Median: 8 y m: f ~ 1.1:1 0.3/100,000 Grade II – Maximal safe resection followed Pediatric:
[1, 208, 428] (60–70 %), (~50 %) by RT – 5-y survival
supratentorial Grade III – C for Grade III, non-resectable • 66–79 %
(30–40 %) (~50 %) tumors, low EOR and at – 10-y survival
progression • 55–68 %
– Re-resection at progression
followed by RT with or without
C
Pilocytic astrocytoma [1, Cerebellar (in Median: 18 y m: f ~ 1:1 0.9/100,000 (age Grade I – Maximal safe resection Pediatric:
247, 249, 429] children > 50 %), 38 % < 15 y 0–19) (>90 %) – Post-OP: – 10-y survival
optic pathway, 75 % < 20 y • Usually no additional Cerebellar: 100 %
hypothalamus therapy, depending on EOR Overall: 97 %
• RT optional for non- Adults:
resectable tumors and at – 5-y survival
progression Overall: 84 %
• C optional for non-resectable
tumors and at progression
– Re-resection at progression
Diffuse low-grade glioma Frontal (50 %), Median: 45 y m: f ~ 1.3:1 1/100,000 Grade II – Maximal safe resection 5-y survival
(DLGG) temporal (25 %) – A 20–50 % – Post-OP: – >95 % After GTR
– Astrocytoma (A) Insular (10 %) – OD 35 % • C to be considered, – 78 % if Resection < 50 %
– Oligodendroglioma – OA 10–45 % especially for OD, non- 10-y survival
– (OD) resectable DLGG, low EOR – 90 % After GTR
– Oligoastrocytoma and at Progression – 54 % if resection < 50 %
(OA) [107, 430-432] • RT for non-resectable DLGG OD and LOH 1p/19q tend to correlate
and at Progression with better outcome
– Re-resection at progression,
followed by C and/or RT
(continued)

659
Table 35.1 (continued)

660
Preferential Annual WHO
Tumor location Age Gender incidence grade Present treatment strategy Prognosis
High-grade/anaplastic Parietal, frontal, Median: 43 y m: f ~ 1.4:1 0.6/100,000 Grade III – Maximal safe resection followed Median survival
(A) glioma temporal by RT or alkylating C (AA, – 4.6- 6.5 y (AOD, AOA)
– AA AOD and AOA without LOH 5-y survival
– AOD 1p/19q) ~50–65 % (AOD, AOA)
– AOA – Maximum resection followed by ~15–50 % (AA)
[121, 206, 433–435] C with or without RT (AOD and 10-y survival
AOA with LOH 1p/19q) ~20–45 % (AOD, AOA)
– Re-resection and C/RT at ~15–30 % (AA)
progression AOD, IDH1 mutations and LOH
1p/19q correlate with increased
sensitivity to RT/C and better outcome
Glioblastoma [433, 436, Temporal, frontal, Median: 57 y m: f ~ 1.7:1 5/100,000 Grade IV – Maximal safe resection, Median survival
437] parietal followed by RT plus concurrent – 14.6 mt (overall)
TMZ, followed by adjuvant – 23.4 mt with methylated (met)
TMZ (age < 65–70 years) MGMT promoter
– Maximal safe resection, 5-y survival
followed by RT or TMZ with or – 9.8 % (overall)
without RT based on MGMT – 13.8 % with met MGMT
status (age > 65–70 years) IDH1/2 mutations and met MGMT
– Re-resection and C/RT at correlate with better outcome
progression to be considered
y Year(s), mt months, m male, f female, C chemotherapy, RT radiotherapy, EOR extent of resection, GTR gross total resection

D. Coluccia et al.
35 Molecular Biology of Human Brain Tumors 661

provided a wealth of new information that can now be applied


to improving patient treatment strategies and outcome.
Modern medicine is on the threshold of achieving many tan-
gible gains in overall patient survival based on the progress
that has been made in several areas of experimental
neuro-oncology.

35.2 The Familial Central Nervous System


Tumor Syndromes

Most CNS tumors occur in adults and arise spontaneously


with no previous family history or genealogic pattern of
inheritance. In some cases, especially in younger patients,
they may present as a familial syndrome and transmitted
most frequently in an autosomal dominant (AD) pattern of Fig. 35.2 T1-weighted and gadolinium-enhanced MRI sequence
inheritance. Advances in the identification and characteriza- showing a Pilocytic Astrocytoma of the optic pathway in a patient with
tion of gene mutations have permitted insight into the molec- neurofibromatosis type 1.
ular basis of both familial and sporadic CNS tumors.
Many familial CNS cancer syndromes are associated with developed and have reported the identification of more than
mutations in tumor suppressor genes, which mediate func- 95 % of individuals fulfilling clinical NF1 criteria [16].
tions such as DNA repair and cell cycle arrest. Affected indi- The disease is characterized by the presence of multiple
viduals typically inherit a germ-line mutation in one copy of neurofibromas, pilocytic astrocytoma within the optic nerve
the genetic locus and in the cases of non-AD mutations, are pathways, malignant peripheral nerve sheath tumors
associated with somatic loss of the wild-type allele in order (MPNST), and susceptibility to other astrocytomas. The
to promote the initiation of tumor growth. One important dis- optic pathway astrocytoma (OPA) represents the most com-
tinction of the familial syndromes is the predisposition of mon CNS tumor in NF1, affecting up to 20 % of all patients
patients to develop multiple tumors in the CNS as observed and usually arising in children under the age of six [17].
in neurofibromatosis 1 (NF1), neurofibromatosis 2 (NF2), Remarkably, half of the individuals with OPA remain asymp-
von Hippel–Lindau disease, tuberous sclerosis complex, tomatic [17].
Gorlin syndrome, Li–Fraumeni syndrome, Turcot syndrome, NF1 is inherited with high penetrance but the phenotypic
and Cowden syndrome. expression is variable, suggesting a role for other disease-
modifying genes and heterogeneous mutations of the NF1
gene [18–20]. The timing of NF1 gene inactivation or muta-
35.2.1 Neurofibromatosis Type 1 tion is also assumed to interfere with disease manifestation
[20]. The NF1 gene is localized to chromosome 17q in 1987
NF1, also known as von Recklinghausen’s disease, is one of and mapped to the 17q11.2 locus 3 years later [21–23]. The
the most common inherited monogenetic syndromes predis- NF1 gene encodes the protein Neurofibromin and spans at
posing individuals to the formation of CNS tumors, with an least 350 kilobases (kb) and includes over 60 exons. More
incidence of 1 in 3500 individuals and an equal predomi- than 1000 mutations resulting primarily in Neurofibromin
nance in both genders [13] (Fig. 35.2). NF1 is AD, with truncations have been identified [24] and complete gene
marked pleiotropy and variability in clinical expression. deletion is associated with a severe phenotype [25].
Approximately 50 % of newly diagnosed patients have no Neurofibromin functions as a tumor suppressor and is a
previous family history, suggesting the occurrence of de cytoplasmic protein of that is highly expressed in neurons,
novo mutations, with a bias for the paternal germ-line. The Schwann cells, oligodendrocytes, astrocytes, leucocytes and
disease is caused by loss-of-function mutations of the NF1 the adrenal medulla, while its expression in other tissues is
gene. The diagnosis of NF1 is made according to well minimal [26]. Neurofibromin harbors a Guanosine Activating
defined clinical criteria (NIH criteria) [14], which are usually Protein (GAP)-related domain, conferring homology to other
diagnostic by the age of 8 years [15]. Detecting mutations in guanosine triphosphatase (GTP) activating proteins [27].
the NF1 gene is difficult and complex given the large size of Ras is a GTP-binding oncoprotein, which promotes cell
the gene and the lack of established mutation hotspots. growth and differentiation. By accelerating the conversion of
Protocols combining PCR and high-performance liquid active GTP to its inactive, guanosine diphosphatase (GDP)-
chromatography to detect germ-line mutations have been bound form, Neurofibromin acts as a negative regulator of Ras.
662 D. Coluccia et al.

Defects in Neurofibromin are thought to increase cell growth The NF2 gene was mapped to chromosome 22 in 1987
and facilitate tumor formation due to loss of Ras regulation. and further localized on locus 22q12.2 [43]. The gene
In addition, Neurofibromin regulates other proteins includ- responsible for this disorder was cloned in 1993 by two
ing adenylate cyclase in astrocytes, and potentially plays a groups by positional cloning and was notably also found to
role in neural stem cell proliferation, survival, and astroglial be mutated or deleted in the majority of sporadic meningio-
differentiation [28, 29]. mas and schwannomas [44, 45]. The gene spans 17 exons
While most of the neoplasms encountered in NF1 patients distributed approximately over 110 kb, and it encodes for a
are not considered malignant, adults with NF 1 have a protein of 595 amino acids called merlin (for moesin, ezrin,
50–100-fold increased risk of high-grade glioma formation radixin-like) due to its homology to three proteins of the
[30, 31] and a 20 % lifetime risk to develop a malignant ezrin, radixin, moesin (ERM) family [45]. Merlin, alterna-
peripheral nerve sheet tumor (MPNST) [32, 33] which are tively named Schwannomin, is expressed in many normal
assumed to arise from preexisting neurofibromas [32]. To human tissues, including the brain, and functions as a tumor
date, only two genetics alterations have been identified as suppressor [46]. Tumors resulting from dysfunction or loss
progression pathways. Mutation in the TP53 loci and genetic of Merlin activity are mostly benign [47].
alterations in INK4A have been frequently identified in malig- Several studies have shown that Merlin controls cell pro-
nant but not benign forms of neurofibromas [30, 34]. liferation and cell motility by regulating cytoskeletal organi-
Numerous animal models have been employed to explore zation, but the pathways involved in its mediation of growth
the molecular basis of NF1 pathogenesis. NF1 heterozygous suppression are poorly understood [46, 48]. Merlin interacts
mice with selective inactivation of NF1 in Schwann cells with F-actin and negatively regulates cell motility through
develop tumors with histologic and molecular features of cytoskeletal reorganization. Moreover, it may act with CD44
human neurofibromas [35]. The generation of astrocyte- (a hyaluronic acid receptor), β1 integrin (a transmembrane
specific NF1 knockout mouse yielded similar findings, with glycoprotein) and paxillin (a cytoskeletal adaptor protein) to
low-grade optic astrocytomas developing only when regulate cell motility [48]. Merlin deficiency has also been
Neurofibromin expression was ablated in astrocytes of NF1 linked to abnormal activation and overexpression of receptor
heterozygous mice [36]. tyrosine kinases such as EGFR family receptors [49].
Several therapeutic approaches based on blocking the A number of NF2 mouse models have been developed.
dysfunctional neurofibromin pathway have been attempted Homozygous mutation of NF2 results in embryonic lethality,
in animal studies [37]. However, most clinical studies have indicating that its function is essential at early stages of
not extended beyond Phase II trials [38]. The mTOR path- mouse development [50]. Heterozygous NF2 mutant mice
way is regarded as a key mediator of optic pathway glioma develop many neoplasms but do not manifest the classical
development. Currently, an ongoing Phase II study is inves- tumors linked to NF2. However, as observed in NF1 murine
tigating the growth inhibition of NF1 related low-grade glio- models, the model of conditioned Schwann cells for knock-
mas using the mTOR-inhibitor everolimus [39]. out mice show many similarities with NF2 patients [51].
Merlin dysfunction or loss affects various signaling path-
ways, therefore several therapeutic targets have been pro-
35.2.2 Neurofibromatosis Type 2 posed. As schwannomas represent the most common
symptomatic and limiting tumor in NF2 patients, clinical tri-
NF2 is approximately tenfold less common than NF1 and als are primarily directed against schwannoma formation
has distinct phenotypic and genetic features. It is an autoso- [52]. A recent Phase II study in NF2 patients with progres-
mal dominant syndrome that occurs at an incidence of 1 in sive vestibular schwannomas observed some decrease in
25,000–40,000, with a penetrance close to 100 % by the age tumor volume and improvement in hearing in response to the
of 60 [40]. Half of the individuals inherit a germ-line muta- EGRF inhibitor Lapatinib [53].
tion from a parent or acquire a de novo mutation at the stage
of embryogenesis [41]. Bilateral vestibular schwannomas
are pathognomonic of the disorder, which is also character- 35.2.3 von Hippel–Lindau Syndrome
ized by schwannomas of other nerves (cranial, spinal or
peripheral) as well as intracranial and intraspinal meningi- von Hippel–Lindau (VHL) is a rare inherited disease with an
oma and ependymomas. NF2 is characterized by a variable AD pattern of transmission and was first described in the
age at onset, with a majority of patients developing signs in early 1900s by Eugen von Hippel and Arvid Lindau [54, 55].
the second to third decade of life, and thus, NF2 schwanno- It is characterized by the development of benign
mas and meningiomas occur at an earlier age compared to hamartomatous tumors in the CNS and adrenal glands
their sporadic counterparts [41]. NF2 is clinically diagnosed including hemangioblastoma and retina angioma, pheochro-
according to the 2005 Manchester criteria [42]. mocytoma, pancreatic neuroendocrine tumors and inner ear
35 Molecular Biology of Human Brain Tumors 663

tumors as well as malignant tumors of the kidney such as Essentially, any organ system can be involved and common
renal cell carcinoma [56]. Its incidence is estimated at 1 in features of TSC include dermatologic manifestations such as
36,000–45,000 births per year [57, 58]. Up to 20 % of VHL cutaneous angiofibromas, subungual fibromas, cardiac rhab-
patients are estimated to derive from de novo mutations [59]. domyomas, renal angiomyolipomas, intestinal and visceral
The disease shows variable expressivity with almost 95 % cysts, and pulmonary lymphoangiomyomatosis,
penetrance at the age of 60 years [60]. A combination of The diagnostic criteria consists of a set of major and
characteristic clinical features, with or without a positive minor features which appear at distinct points in develop-
family history, permits the establishment of a clinical diag- ment, but most patients have manifestations of the disease
nosis. The most frequent lesion observed is hemangioblas- before 10 years of age [68]. Neurologic disabilities seem to
toma, a mesenchymal solid and cystic tumor mainly localized be intimately related to the presence of cortical tubers, which
in the brain (mostly located in the cerebellum), which occurs occur in approximately 70 % of patients [69].
in 60 to 80 % of all VHL patients [61]. Depending on specific The penetrance of the disease is high but the clinical man-
mutation in the gene locus, VHL is classified as type 1 or ifestations are considerably variable [70]. There is a high rate
type 2, according to the absence or presence of a pheochro- of new mutations, accounting for up to 85 % of cases [71].
mocytoma [56]. TSC results from the mutation of one of two genes, TSC1 on
The gene responsible for this disease is the VHL tumor- chromosomes 9q34 and TSC2 located on 16p13.3. Clinical
suppressor gene, located on chromosome 3p25-26 [57]. and pathological features caused by mutations in these two
Under normoxic conditions, the gene product targets genes are indistinguishable, but the prognosis seems to be
hypoxia-inducible factors (HIF) for polyubiquitination and more severe when the disease is associated with a TSC2
proteosomal degradation. Under hypoxic conditions, the mutation [72]. The TSC1 gene encodes a transcript of 8.6 kb
accumulation of HIF will induce the transcription of hypoxia- called Hamartin, containing 23 exons, whereas the TSC2
regulated genes such as vascular endothelial growth factor gene encodes a transcript of 5.5 kb called Tuberin, contain-
(VEGF), platelet derived growth factor (PDGF) and trans- ing 41 exons. Mutations in these genes can be identified in
forming growth factor (TGF) -α. This explains the high vas- 60–80 % of TSC patients, with a wide spectrum of specific
cularization of VHL-related tumors. Loss of the VHL gene genetic alterations [73]. Hamartin and Tuberin function as a
can also result in aberrant expression of genes that control regulatory complex inhibiting the small GTPase RHEB and
cell cycle progression and invasion [62, 63]. It has been consequently limiting the activity of the mammalian target
shown in renal carcinoma cells that loss of the VHL gene of rapamycin complex 1 (mTORC1), which is an intracellu-
results in the accumulation of hepatocyte growth factor lar regulator of cell growth and metabolism [74]. In a recent
(HGF), which interacts with the β-catenin pathway [64]. Phase I–II open-label study of patients receiving the mTOR-
Loss of VHL protein activity has also been recently corre- inhibitor everolimus, a clinically significant reduction in tumor
lated to the activation of kinases required for Ras-induced volume of TSC related subependymal giant-cell astrocytoma
cell transformation, thus representing a possible target for was shown [75, 76].
molecular )inhibition [65].
Hemangioblastoma is a mesenchymal-derived hamartoma-
tous tumor and patients with VHL-syndrome are not typically 35.2.5 Gorlin Syndrome
linked with glioma manifestation. However, there is some
evidence that VHL expression interferes also with neuroepi- Gorlin syndrome is also known as nevoid basal cell carcinoma
thelial tumor formation via suppression of HIF-1α/VEGF, syndrome (NBCCS) or basal cell nevus syndrome (BCNS)
which is coupled with inhibition of glioma proliferation and and was first described by Goltz and Gorlin in the 1960s.
invasion [66]. It is a rare AD disorder occurring in 1 of 57,000 live births
and characterized by basal cell-carcinoma, skeletal anoma-
lies and a 2–5 % incidence of childhood medulloblastomas
35.2.4 Tuberous Sclerosis Complex (mostly SHH medulloblastomas) [11, 77]. The penetrance is
very high but the clinical expression is variable, with up to
Tuberous sclerosis complex (TSC) is a multisystem autoso- 30 % of patients presenting with de novo mutations [78].
mal dominant disorder, affecting 1 in 6000–10,000 individu- The clinical criteria are based on the association of multiple
als [67]. It was first described by Bourneville in 1880 and is basal cell carcinomas and odontogenic keratocysts [79].
characterized by CNS tumors including cortical hamartomas Other manifestations include the presence of palm or sole
named tubers, subcortical glioneuronal hamartomas, subep- pits, intracranial calcifications in particular along the falx
endymal glial nodules, and subependymal giant cell astrocy- and macrocephaly [77].
tomas. TSC patients often suffer from debilitating neurologic The disease results from a mutation of the human homo-
disorders, including epilepsy, mental retardation, and autism. logue of the Drosophila melanogaster gene PTCH located
664 D. Coluccia et al.

on 9q22.3 [80]. It contains 23 exons and encodes for a trans- absence of Turcot syndrome [88]. In this subgroup, the mean
membrane protein (Ptch), which is a receptor and negative age for occurrence of glioblastoma is lower than in the
regulator for the secreted Hedgehog family proteins such as general population (18 years old versus 40–70 years old).
Sonic Hedgehog (Shh). The type 2 subgroup is characterized by association of WNT
subgroup medulloblastomas with familial adenomatous pol-
yposis (FAP). These patients tend to have a germ-line muta-
35.2.6 Li–Fraumeni Syndrome and TP53 tion in the adenomatous polyposis coli (APC) gene that lies on
Germ-Line Mutations chromosome 5q21 [86]. In the type 2 Turcot syndrome sub-
group, the median age of occurrence of medulloblastoma is 15
Li–Fraumeni syndrome (LFS) is an AD disorder predispos- years, compared to 7 years of age in the general population.
ing patients to an early onset of a wide spectrum of neo-
plasms. The syndrome was defined by Li and Fraumeni after
they noticed a high incidence of cancer among family mem- 35.2.8 Cowden Syndrome and Lhermitte–
bers of children with rhabdomyosarcoma [81]. Often Duclos Disease
described as a very rare syndrome with only several hundred
identified individuals, its effective incidence is probably Cowden syndrome (CS) is a rare AD condition associated
underestimated [82], in part also due to the existence of dif- with multiple benign hamartomatous lesions mostly involv-
ferent classification schemes [83]. The most frequent tumors ing the skin and mucosa, but various organ systems including
observed are breast cancer, sarcomas, osteosarcomas, brain the cerebellum may be affected [89]. CS predisposes patients
tumors, acute leukemia and adrenocortical carcinomas. to malignant tumors, especially of the breast, thyroid, intes-
The major genetic abnormality underlying this syndrome tine, and endometrium [89]. The penetrance of CS is high and
is a germ-line mutation of TP53, which has been identified in approaches over 90 % by the age of 20 years [89]. The syn-
70 % of LFS cases and in 22–44 % of families with the LF drome was first documented in 1963 by Lloyd and Dennis
variant form [84]. Importantly, 50 % of the families with a and named after the first patient [90]. In 1991, more than 70
TP53 germ-line mutation meet the diagnostic criteria of years after its first description [91], it was recognized that
LFS. In the remaining cases of LFS without TP53 mutation, a dysplastic cerebellar gangliocytoma or Lhermitte–Duclos
germ-line mutation of the hCHK2 gene involved in G2 check- disease (LDD) is often a manifestation of CS [92]. The preva-
point control has been identified [84]. TP53 germ-line muta- lence of CS is estimated at 1:200,000, but the frequency is
tions occur most frequently in exons 5–8, with major hotspots likely underestimated given that many non-symptomatic
on codons 245 and 248. In the 143 LFS families reported, the mucocutaneous manifestations related to this disease are also
most frequently identified mutations are point mutations common in the general population and thus may not lead to
(85.3 % of cases). Interestingly, low grade and anaplastic extensive diagnosis [89]. Approximately half of the CS cases
astrocytomas of LFS patients harbor somatic IDH1 mutations are recorded as inherited and half occur spontaneously [93].
at much higher rates compared to sporadic gliomas [85]. The most common germ-line mutation) involves the tumor
suppressor gene PTEN, located at the chromosome 10q23.3
[94, 95]. Early clinical data including nearly 50 CS families
35.2.7 Turcot Syndrome suggest that 85 % of patients with CS harbor a PTEN muta-
tion [96, 97] However this may not be universally adopted
Turcot syndrome is an AD disorder characterized by the given that application of risk assessment criteria from the
occurrence of adenomatous colorectal polyps or colon carci- National Comprehensive Cancer Network (NCCN) results in
nomas and malignant brain tumors, mainly anaplastic astro- many patients diagnosed as CS despite the absence of PTEN
cytomas, as well as glioblastomas and medulloblastomas mutations [98]. Germ-line mutations downstream of PTEN
(95 %) [86]. Turcot described the first cases in 1959, where signaling were found in the PI3K/AKT pathway in up to 10 %
two siblings developed malignant CNS tumors and numer- of CS patients lacking PTEN gene alterations [99]. Not all
ous adenomatous colorectal polyps [87]. patients with LDD show clinical manifestation of CS [100]
There are two major subgroups of Turcot syndrome. The and dysfunction or inactivation of PTEN is found in 67–83 %
type 1 pattern is characterized by an association of glioblasto- of LDD lesions [100, 101]. Selective inactivation of PTEN in
mas and hereditary non-polyposis colorectal carcinomas neural cells of mice results in impaired migration of granular
(HNPCC). This subgroup results from a germ-line mutation cell precursors and formation of cerebellar tumors histopath-
in DNA mismatch repair genes such as hPMS1 (ch 2q32), ologically resembling LDD [102, 103]. Inactivation of PTEN
hPMS2 (ch 7p22), hMSH2 (ch 2p16), or hMLH1 (ch 3p21) is linked with increased levels of PI3K/AKT, which promotes
[86]. These deficiencies lead to DNA replication errors and cell growth and proliferation [102, 103]. Activated AKT
microsatellite instability, which is rare in brain tumors in the regulates various pathways involved in protein synthesis and
35 Molecular Biology of Human Brain Tumors 665

cell growth, including the mTOR pathway, suggesting that


mTOR inhibitors may be effective in reversing cellular hyper- Box 35.1: Primary Versus Secondary GBM
trophy and tumor growth [101]. According to current evidence, there are two different
routes of glioblastoma tumorigenesis, both with differ-
ent genetic alterations but ultimately undistinguishable
35.3 Cytogenetics and Molecular Genetics histology [114]. GBM mostly occurs as a rapidly grow-
of Neuroepithelial Human Brain ing de novo tumor from a glial progenitor or dedifferen-
Tumors tiated cell and affecting patients with a mean age around
60 years. This Primary GBM constitutes 90 % of the
35.3.1 Glioma (Astrocytoma cases, where no evidence of a less malignant precursor
and Oligodendroglioma) lesion can be detected. Secondary GBMs (10 %) affect
typically younger patients around a mean age of 45
Gliomas comprise tumors of astrocytic or oligodendrocytic years. These tumors progress from a preexisting diffuse
histopathological appearance. Whether these tumors have low-grade or anaplastic astrocytoma, are more often
the same origin of tumor-initiating cells from neural progeni- located frontally, and the overall survival after diagno-
tors or arise from dedifferentiated astrodendrocytes/oligo- sis is significantly longer. Histologically, Secondary
dendrocytes is a matter of ongoing debate [104, 105]. The GBMs show less necrosis and frequent regions of oli-
WHO distinguishes between gliomas with more circum- godendroglioma-like appearance when compared to
scribed growth (WHO Grade I) and gliomas with diffuse the Primary form. From a molecular basis, the Primary
infiltration (WHO Grade II: low grade, WHO Grade III–IV: tumors often show loss of chromosome 10, EGFR
high grade) [9]. Diffuse low-grade gliomas are considered a amplification and PTEN mutation, whereas in
precancerous disease, as over time they progress to anaplas- Secondary GBM, TP53 mutations, loss of chromosome
tic, malignant tumors in most instances [106]. 19q and PDGFRA alterations are more frequent.
Gliomas demonstrate the full spectrum of cytogenetic and Classification of GBM origin has been significantly
molecular abnormalities: From numerical and structural improved by the recent identification of mutations in
chromosomal alterations, to gene amplifications and overex- the enzyme gene IDH1/2 (isocitrate dehydrogenase,
pression, deletions and small-scale mutations up to and involved in the citric acid pathway), which are predom-
including epigenetic deregulations. inantly found in Secondary (80 %) but only rarely
(<5 %) in Primary GBMs [132–134]. Consequently,
IDH1/2 mutations are now commonly accepted as
35.3.2 Chromosomal Alterations molecular markers of Secondary GBM.
Over 90 % of secondary GBM diagnosed through
Chromosomal number alterations occur with increased fre- IDH1/2 mutations belong to the proneural subgroup,
quencies in high grade and adult gliomas, and chromosomal which indicates that this tumor subgroup is homoge-
losses are observed more frequently than gains [107]. One of neous compared to the primary form. IDH1/2 mutations
most common numerical autosomal changes seen in high- occur early during progression from low-grade to high-
grade astrocytomas includes gain of chromosome 7, and loss grade glioma given they are detected in over 80 % of
of chromosome 10 [107–109]. Gain of chromosome 7 is grade II and III astrocytoma and oligodendroglial tumors
seen in up to 83 % of adult high-grade astrocytomas, whereas [134, 141]. It is possible IDH1/2 mutations precede
chromosome 10 is reported to be lost in up to 86 % [109, 1p/19q loss in oligodendrogliomas and TP53 mutation in
110]. In some studies, a combined gain of chromosome 7 low-grade astrocytomas [242]. Finally, there is evidence
accompanied by loss of chromosome 10 (7+/10-) occurs in that primary and secondary GBMs derive from different
over 80 % of GBM samples [108, 111]. glial progenitor cells. In support of this, CD133 positive
The occurrence of trisomy 7 and monosomy 10 together cancer stem cells could only be cultured from Primary
has often been correlated with short-term survival, although GBMs, whereas samples from secondary GBMs did not
currently available data does not allow a distinct conclusion have CD133 positive stem cells within [243].
[112]. Gains of chromosome 7 are also among the most com-
mon autosomal changes in low-grade gliomas (LGG), seen
in 57 % of cases [110] and loss of chromosome 10 may be mosomes, seen in approximately 20 % of GBM specimens,
correlated with shorter overall survival [113]. Chromosomal which can further confer oncogene amplification and drug
alterations commonly observed in secondary GBM include resistance [115].
19q loss (see Box 35.1) [114]. An additional numerical chro- Oligodendroglioma (OD) and oligoastrocytomas (OA)
mosomal abnormality is the presence of double minute chro- constitute a subtype of glioma for which the specific loss of
666 D. Coluccia et al.

portions of 1p and 19q (loss of heterozygosity, LOH 1p/19q) early-detected genetic abnormality—seen in about 2/3 of
is associated with important prognostic implications and is precursor low-grade astrocytomas and secondary GBMs
used as a marker to support the histological diagnosis of OD derived from them. TP53 mutations occur in primary GBM
[116, 117]. Approximately 50–80 % of ODs have allelic loss but at a lower frequency (35 %) compared to secondary GBM
of chromosomes 1p and 19q, whereas OA (mixed oligoden- [108, 109, 130]. Additionally, in secondary GBM, 60 % of the
drogliomas, arrangement of cells presenting astrocytic or TP53 mutations are clustered in two hot spot codons (amino
oligodendroglial differentiation) show this co-deletion 40 % acids 248 and 273), whereas in the primary form, TP53 muta-
of the time [118–120]. LOH 1p/19q is frequently observed in tions appear more equally distributed within exon 5–8 [108].
WHO grade II oligodendric tumors compared to those of Mutations in the gene encoding for NADPH-dependent
grade III, though conflicting data exists in this regard [118– isocitrate dehydrogenase (IDH1) were identified from a large-
120]. Notably, oligodendric gliomas of the frontal lobe show scale genomic sequencing analysis of astrocytoma [131].
higher incidence of 1p/19q co-deletion than those of non- Since the initial description, multiple studies have shown that
frontal origin [119]. Low-grade OD (WHO II) treated by sur- IDH1 mutations are common in low-grade diffuse astrocyto-
gical resection and nitrosourea-based chemotherapy without mas, anaplastic astrocytomas and tumors of oligodendroglial
radiotherapy show favorable long-term outcome (10-year origin [132–134]. Notably, there is evidence that IDH1 muta-
overall survival rates over 90 %) irrespective of LOH 1p/19q tions precede TP53 mutations, suggesting an important role
status [121]. Anaplastic pure and mixed oligodendroglio- in tumor initiation [114]. A proportion of low-grade astrocy-
mas (WHO III) identified with 1p/19q co-deletion are gen- tomas that lack IDH1 mutations (e.g., pilocytic astrocytoma,
erally associated with marked higher overall survival and and ependymoma) contain mutations of the related IDH2
chemosensitivity in comparison to non-co-deleted WHO III gene [134]. As IDH1 mutations are very common in second-
oligodendrogliomas, where the addition of chemotherapy to ary (>80 %) but markedly rare in primary GBM (<5 %), a
radiotherapy did not prove beneficial over radiotherapy consensus is that IDH1 mutation can serve as a molecular
alone [120, 122]. marker of secondary GBM, as this form is neither clinically
nor histologically distinguishable from the primary form
[114]. IDH1 status is important to clinicians as patients with
35.3.3 Gene Amplification and Overexpression tumors harboring IDH1 mutations have a survival benefit
[135–139]. IDH mutations also seem to correlate with both
Gene amplification has been described for several target genes TP53 mutational status and 1p/19q deletions, both of which
in astrocytic tumors, and appears to be more common in high- predict survival [137, 140, 141].
grade lesions than low-grade astrocytomas [109, 123]. Genes
continuously found to be amplified in high grade gliomas
include EGFR, MET, PDGFRA, MDM4, MDM2, CCND2, 35.3.5 Epigenetic Alterations
PIK3CA, MYC, CDK4, and CDK6 [124]. Amplification and
overexpression of EGFR occurs frequently and predominantly In addition to aberrant changes in the genome, epigenetic
in primary GBMs (60 % of the samples), but rarely in the sec- alterations have emerged as important factors in gliomagen-
ondary form [114, 125]. In up to 30 % of the cases this gene has esis [142]. Potentially reversible and more amenable to
undergone loss of exon 2–7, resulting in the production of a treatment than numerical or structural changes in chromo-
constitutively active, truncated EGFR receptor (EGFRvIII) somes, studies of epigenetic deregulations is a promising
[109, 126, 127]. At this time, Phase II and III vaccination trials strategy in glioma therapy. GBM cells showing high activity
using rindopepimut to stimulate immune response against the level of the DNA repair enzyme MGMT (O6-methylguanine-
EGFRvIII peptide in GBM patients are ongoing [128, 129] DNA methyltransferase) are more resistant to therapies
including alkylating agent such as temozolomide (TMZ)
[143]. About 40 % of GBM have a methylated MGMT pro-
35.3.4 Gene Mutations and Deletions moter status resulting in epigenetic silencing of the MGMT
gene and thus higher sensitivity to TMZ and distinctive lon-
Tumor suppressor gene silencing events in astrocytomas ger survival with actual standard therapy [143, 144]. Hence,
may be the result of gene deletions, mutations, or promoter MGMT promoter methylation is a strong prognostic and
methylation events. Primary GBMs are characterized by predictive biomarker.
CDKN2A deletion (31-50 %), and PTEN mutation (32 %) The increased awareness of the role of metabolic enzymes
[109, 130, 131]. While CDKN2A deletions occur in primary in astrocytoma pathogenesis has brought about a resurgence
GBM, TP53 mutations appear more common in secondary of interest in the Warburg Effect [145]. The Warburg effect
GBM [109, 114]. In the progression of low-grade tumors to first described in 1956 by biochemist Otto Warburg promotes
secondary GBM, TP53 mutations are the most common and the observation that tumors predominantly use a high rate of
35 Molecular Biology of Human Brain Tumors 667

aerobic glycolysis to meet metabolic demand [146]. Recently,


isoforms of both pyruvate kinase (PK) M1/M2 and hexoki- Box 35.2: Pilocytic Astrocytoma
nase (HK1/HK2) have shown differential expression in glio- Pilocytic Astrocytoma (PA) is a non-infiltrative, slow
blastoma, in which PKM2 and HK2 predominate, promoting growing, and often cystic tumor typically classified as
glycolysis [147, 148]. WHO grade I. It is one of the most common CNS neo-
plasms in childhood accounting for ~20 % of all pediat-
ric brain tumors [1]. Almost three-fourths of the patients
35.3.6 Pediatric Versus Adult Gliomas are younger than 20 years [1]. Approximately half of
PAs are located in the cerebellum, but they can occur
The genetic changes that are found in pediatric astrocytomas throughout the neuraxis such as the optic pathway,
are distinct from those in the adult variant of these tumors. No hypothalamus, and spinal cord (mostly in pediatric
consistent karyotypic abnormalities have been identified in patients) as well as in the cerebral hemispheres (mostly
pediatric low-grade astrocytomas [149, 150]. Even in high- in young adults) [244]. Optic nerve PAs are often asso-
grade gliomas, the number of chromosomal aberrations is gen- ciated with NF1. Provided that the tumor is accessible
erally lower from their adult counterparts with up to 15 % and extensive surgical resection is performed, most
lacking detectably number irregularities [107]. Relatively fre- patients show a favorable outcome with a 10-year sur-
quent compared to adult GBM, some pediatric GBMs have vival rate of over 95 % [244–246]. Until recently, the
gain of chromosome 1q (up to 30 %) and loss of 16p (up to genetic mechanisms underlying this disease were not
24 %), while numerical aberrations of chromosome 7 and 10 extensively investigated. Early genomic analysis of
are relatively rare (<30 %) [151, 152]. Cytogenetic changes tumor samples showed often balanced karyotypes,
recurrently observed in pediatric anaplastic astrocytomas and whereas whole chromosomal gains occurred in about
GBM include gains of 1q, 5q, and loss of 6q, 9q, 10q, 12q, 30 %, frequently affecting chromosomes 5 and 7 [247].
13q, and 22q [152, 153]. In contrast to adult tumors, amplifi- Activating alterations within the MAPK-signaling
cation of the EGFR gene is not common in pediatric astrocy- pathway are considered a hallmark aberration in the
tomas and PDGF-driven signaling is preferentially activated pathogenesis of PA, affecting 80–100 % of analyzed
[151, 154]. Notably, this feature is also prevalent in secondary tumor samples [248, 249]. Consequently it is hypothe-
GBM [155]. Similarly, pediatric astrocytomas rarely exhibit sized that PA represents a single-pathway disease [249].
loss or mutation of PTEN. Aberrant activation of the BRAF The most frequent genetic event leading to MAPK acti-
proto-oncogene (7q34) by copy number alteration or forma- vation is BRAF-KIAAI549 (B-K) fusion, which pro-
tion of an abnormal fusion protein is an important marker of duces a fusion protein that lacks the regulatory domain
pilocytic astrocytoma but not diffuse infiltrating pediatric of BRAF [249, 250]. In addition, various MAPK path-
astrocytomas (see Box 35.2) [156–158]. Aberrant BRAF acti- way-activating fusions )typically involving BRAF are
vation leads to increased signaling through the MAP kinase/ described [249]. The somatic mutation rate is very low
ERK pathways, which results in high levels of mTOR activa- (<0.1/Mb) and BRAF, FGFR1, K/-NRAS, PIK3CA,
tion, and ultimately to increased cell growth [159, 160]. CDKN2A, and NF1 are the main genes found to be
mutated or rearranged [245, 249, 251]. Similar to other
tumor types, the copy number changes and mutation
35.3.7 Core Signaling Pathways and GBM rates positively correlate with patient age and poten-
Subgroups tially with an aggressive clinical course [245, 247,
249]. The distribution of oncogenic hits within the
The understanding of the molecular basis of astrocytoma has MAPK pathway may vary depending on tumor loca-
benefited from large-scale genomic profiling approaches from tion: NF1 mutations are more often detected in optic
networks such as The Cancer Genome Atlas (TCGA) and the PA, while PAs in other locations are dominated by
International Cancer Genome Consortium (ICGC) [109, 130, BRAF activation [248]. BRAF fusions are more com-
131, 161] GBM was the first cancer to be systematically ana- mon in cerebellar PAs and mutations are typically seen
lyzed by TCGA, highlighting the role of oncogenes such as in supratentorial PAs [248]. In adult PAs, BRAF V600E
ERBB2 and PDGFRA and suppressor genes such as NF1 and mutations appear to be very infrequent compared to the
TP53 as well as reinforcing core defects in three signaling pediatric counterpart [245]. As B-K fusions are particu-
pathways involved in the pathogenesis: (1) retinoblastoma larly rare in diffuse astrocytomas (~2 %), the presence
(RB)-signaling (79 %), (2) p53 signaling (90 %), and (3) of this event can be considered an adjuvant diagnostic
PIK(3)K/RAS signaling (88 %) (Fig. 35.3). Large-scale genomic tool for PA [245].
consortia have allowed the stratification of astrocytomas based
668 D. Coluccia et al.

Fig. 35.3 Core defects in


three signaling pathways
involved in the pathogenesis
of Glioblastoma. Overall
alteration rate is summarized
for the TP53 pathway
(eluding apoptosis), PI(3)K/
RAS pathway (increasing
proliferation), and RB1
pathway (avoiding cell cycle
checkpoints).

on genetic profiles giving valuable clinical information. The ther correlated with IDH1 mutations and in extended analysis
first subgrouping analysis on high-grade astrocytomas was also often found in low-grade gliomas [109, 163]. This glioma
reported by Philips et al. differentiating three distinct GBM phenotype is an expression subtype and termed G-CIMP (CpG
clusters: proneural, proliferative, and mesenchymal, with the island methylator phenotype). G-CIMP positive gliomas affect
proneural subgroup predicting better survival [162]. A recent younger patients, are probably more common in secondary
TCGA based study classified GBM into four molecular sub- GBM, and are associated with improved outcome [163]. In
groups: classical, proneural, neural, and mesenchymal [155]. anaplastic oligodendroglioma the G-CIMP phenotype also
Genetic alterations in EGFR, PDGFRA/IDH1 and NF1 char- consistently exhibit IDH1-mutations, MGMT promoter meth-
acterize the classical, proneural, and mesenchymal subgroups ylation, and LOH 1p/19q and is used as a predictor for better
respectively [155]. Classical GBMs are further delineated by survival [164].
amplification of chromosome 7 paired with loss of chromo-
some 10, while often lacking TP53 mutations [155]. No dis-
tinctive genetic alteration is known that distinguishes the 35.3.8 Medulloblastoma
Neural from the other subtypes [124]. It still has to be eluci-
dated to what extent the subgroup defining molecular profile Medulloblastomas are considered embryonal tumors arising
may affect clinical outcome and sensitivity to specific therapeu- from the dorsal brainstem or the cerebellum. Progenitor
tic agents. DNA methylation profiling of 272 TCGA glioblas- cells from the cochlear nucleus, dorsal brainstem, and neu-
tomas revealed a subset of tumors with a hypermethylation ron precursors from the external granular layer of the cere-
phenotype that was assigned to the proneural subgroup, fur- bellar germinal zone are the proposed cells of origin [165].
35 Molecular Biology of Human Brain Tumors 669

Medulloblastomas exhibit a high amount of heterogeneity at 35.3.11 Gene Mutations and Deletions
histomorphological and subcellular level resulting in vari-
able clinical behavior. Histological subtypes comprise classi- Various mutations of tumor suppressor genes important in
cal (70–80 %), nodular desmoplastic/extensive nodular developmental signaling pathways have been described in
(16 %) and large cell anaplastic (10 %), while all are classi- medulloblastoma. Mutations in PTCH1, SMOH, Gli2, and
fied as WHO Grade IV [9, 166]. Over the past few years SUFU lead to loss of inhibitory factors and thus promote con-
novel molecular subclassifications have been established that stitutive activation of SHH pathways [181–184]. Germ-line
put histopathologic differentiations into perspective. APC mutations as found in familial adenomatous polyposis
(FAP) syndrome or Turcot syndrome predispose to medullo-
blastoma formation [86]. The APC-protein is encoded on
35.3.9 Chromosomal Alterations chromosome 5 and is part of a protein complex inhibiting the
WNT signaling pathway. Loss of APC function results in
Almost all medulloblastoma samples show various numerical intranuclear accumulation of ß-catenin, which activates tran-
gains or losses of chromosomal regions. The most common scription factors and promotes the pathogenesis of medullo-
specific cytogenetic abnormality seen in this tumor type is iso- blastoma. Overall, mutations involving the WNT pathway
chromosome 17q (iso17q), which occurs in about 50 % of occur in approximately 10 % of sporadic medulloblastoma
medulloblastomas [167, 168]. The breakpoint is often in the cases, including activating mutations in the CTNNB1 gene
proximal p-arm (17p11), producing a dicentromeric isochro- [185–188]. In addition, TP53 mutations have been found in
mosome. Loss of one arm of chromosome 17p (25–35 %) is 10 % of medulloblastoma cases (16 % in WNT, 21 % in SHH,
often a consequence of isochromosome 17q formation, but virtually absent in Group 3 and 4) and are considered a risk
may also occur in isolation [169, 170]. In certain cases, smaller factor for poor outcome in the SHH subgroup [189].
deletions are seen, with the minimal deleted region occurring at
17p13.3 [171]. The frequent involvement of 17p in medullo-
blastoma has resulted in efforts to identify the targeted putative 35.3.12 Epigenetic Alterations
tumor suppressor gene (TSG). REN(KCTD11) maps to
17p13.2, and is a candidate TSG for medulloblastoma on chro- Epigenetic alterations such as modification of histone or
mosome 17p [172]. Aberrations of chromosomal regions have DNA methylation status are important factors in medullo-
also been evaluated as prognostic markers, whereas iso17q and blastoma pathogenesis and mostly recorded in Group 3 and 4
losses of 10q and 17p are correlated with a poor prognosis and medulloblastoma [190, 191]. Large-scale genomic analyses
loss of chromosome 6 (monosomy 6) with good outcome have identified OTX2, a developmentally regulated transcrip-
across all subgroups [173]. Loss of chromosome 9q, which tion factor important for early brain morphogenesis, as a fre-
contains the suppressor gene locus PTCH 1, promotes activa- quently occurring focal genetic gain in medulloblastoma
tion of Sonic Hedgehog (SHH) signaling pathway and is used (21 %) [168, 192, 193]. Overexpression of OTX2 results in
as a co-determinant of the SHH molecular subgroup [11]. impaired demethylation activity of histone modifying pro-
teins. Consequently, levels of trimethylated histone H3 lysine
27 (H3K27me3) are increased which in turn leads to specific
35.3.10 Gene Amplification chromatin condensation [194]. Medulloblastoma patients
and Overexpression with OTX2 gene amplification (mostly restricted to subgroup
3 and 4) exhibited poorer survival and were more likely to
Gene amplifications are relatively infrequent in medulloblas- have anaplastic tumors [193]. There is also epigenetic dereg-
toma (<10 %), and typically involve the MYC and MYCN ulation of DNA methylation as an underlying mechanism of
proto-oncogenes [166]. Such amplifications have been medulloblastoma formation. Subgroup specific patterns of
observed in the context of double minute chromosomes DNA methylation have already been proposed as a biomarker
[174]. Amplification and/or overexpression of MYC-family for improving survival prediction in non-WNT medulloblas-
genes are regularly observed in the large cell anaplastic vari- toma [195, 196].
ant of medulloblastoma, and correlate with poor clinical out-
come [175, 176]. In addition to gene amplifications, certain
genes are aberrantly overexpressed in medulloblastoma such 35.3.13 Subgroups
as ERBB2 [177]. Increased ERBB2 and ERBB4 expression
correlates with higher risk of metastasis and is associated Advances in next generation DNA sequencing and ultrahigh-
with a poor prognosis [178–180]. resolution genetic mapping over recent years have enabled
670 D. Coluccia et al.

Fig. 35.4 Medulloblastoma Subgroups. Infographic illustrating the incidence, gender ratio, survival, histology, cytogenetic alterations, and common
driver genes for each group. LCA large cell anaplastic, MBEN medulloblastoma with extensive nodularity.

remarkable insight into some of the underlying molecular Group 3 [199]. These molecular subgroups typically do not
interactions leading to the observed heterogeneity in clinical correlate with the histological phenotypes. The histopatho-
behavior. According to current consensus among investiga- logical classic type occurs in all four medulloblastoma sub-
tors, four main subgroups can be distinguished: WNT, SHH, groups. The nodular desmoplastic or extensive nodular type
Group 3, and Group 4 [11, 197, 198] (Fig. 35.4). The WNT falls within the SHH subtype, although it should be men-
and the SHH group are marked by mutations in the WNT and tioned that the SHH subgroup is the only subgroup that may
SHH pathway, respectively. For Group 3 and 4, no single include all of the known major histological variants. Large
pathway alteration has been identified and the diagnosis cell anaplastic tumors are also found in all four subgroups
relies primarily on transcriptional profile clusters. However, but the majority will be classified to Group 3. The WNT sub-
recent studies detected prevalent drivers in particular for group has mostly classic histology and accounts for 10 % of
35 Molecular Biology of Human Brain Tumors 671

all medulloblastomas and has the best prognosis with >90 % chemotherapy, whereas patient with tumors subgrouped into
long-term survivals. Alterations resulting in increased activity more aggressive lesions may benefit from exhaustive thera-
of the WNT pathway results from sporadic somatic mutations peutic strategies.
of CTNNB1 encoding ß-catenin or less common in germ-line
mutations of the APC gene as in Turcot or FAP syndrome.
Monosomy 6 and TP53 mutation are often found. The SHH 35.3.14 Ependymoma
subgroup (30 %) has a good to intermediate prognosis; almost
half of the cases exhibit desmoplastic/nodular histology. Ependymomas originate from the wall of the ventricular
In sporadic SHH tumors, somatic mutation in PTCH, system with possible manifestations along the entire cranio-
SMO, and SUFU leads to increased SHH activity. In addi- spinal axis of the CNS [202]. Radial glia cells are presumed
tion, amplifications of GLI1 and GLI2 can trigger this path- to constitute the progenitor cells [203]. In children, over
way. Deletion of chromosome 9q, which includes the PTCH 90 % of ependymomas occur intracranially with 70 % aris-
gene, is primarily limited to SHH tumors. Hereditary SHH ing in the posterior fossa, whereas in adults the spinal mani-
medulloblastomas are found in Gorlin syndrome and are festation is more common [204–206]. Based on
characterized by germ-line mutation of the tumor suppressor histopathological features, the WHO divides ependymomas
gene PTCH1. MYCN amplification and mutations at TP53 in grade I–III and recognizes four histological types: subep-
are regularly observed. Recently, small molecules targeting endymoma (grade I), myxopapillary (grade I, usually spi-
smoothened (SMO), a positive regulator of the SHH path- nal), classic (grade II), and anaplastic (grade III) [9, 202].
way, showed rapid (although transient due to acquired drug However, as for most tumors, molecular characteristics will
resistance) regression of tumor volume in a subset of patients have to complement and specify future grading systems.
with SHH medulloblastoma [200, 201]. Group 3 medullo- Recent studies indicate that ependymomas from different
blastomas (25 %) primarily demonstrate classic histology but compartments of the CNS display distinct genetic signa-
have a high incidence of large cell anaplastic types. They are tures, reflecting their unique origins, despite appearing his-
almost never observed in adults and harbor the worst progno- tologically homogeneous [203, 207].
sis of all subtypes. Metastases are frequently found and cou-
pled to poor prognosis. MYC and OTX2 amplification are
common features and some tumors show loss of both 5q and 35.3.15 Chromosomal Alterations
10q and gain of chromosome 1q. Recently, somatic genomic
rearrangements resulting in GFI1 and GFI1B activation have Cytogenetic changes are seen in a significant proportion of
been found to occur in approximately one-third of Group 3 ependymomas and are overall more common in adults [206,
patients; thus enhancer hijacking is now considered a promi- 208, 209]. About 30 % of sporadic ependymomas have
nent mechanism driving Group 3 medulloblastoma [199]. monosomy 22, and deletions or translocations involving
Group 4 is the most common medulloblastoma subgroup chromosome 22 are also observed [210]. The NF2 gene is
(35 %), usually of classic histology and with an intermediate located on chromosome 22 (22q12), and NF2 patients are
prognosis. Isochromosome 17q is the major cytogenic altera- predisposed to develop spinal ependymomas in addition to
tion observed (60–80 % of Group 4 samples). Amplification of other CNS tumors [211]. In fact, NF2 mutations are almost
OTX2, CDK6, and MYCN are also frequently associated with exclusively found in spinal ependymomas [212], thus another
this subgroup, whereas structural variants associated with tumor suppressor gene on chromosome 22 might be involved
GFI1 and GFI1B activation are detected in 5–10 % [199]. in cranial ependymomas [210, 213]. Gain of chromosome 1q
Molecular classification will have a large impact on future is the most common genomic aberration in pediatric intra-
treatment recommendations and strategies. Given the four cranial ependymomas, possibly associated with anaplastic
subgroups show distinct clinical behavior that does not cor- histology [206, 214]. Besides numeric chromosomal altera-
relate with histological features, the use of histopathology tions within the ependymoma genome, translocations are
alone is considered inadequate for the classification of frequently reported to affect chromosomes 1, 11, and 22
medulloblastoma. It remains to be seen how the revealed [208]. The assumption that tumor location-specific genomic
diversity in molecular patterns and clinical outcome will alterations exist in ependymoma has been encouraged by
influence the WHO Grade IV classification for medulloblas- recent revelation of chromothripsis within chromosome
toma. The current treatment modality of surgery, radiation 11q13.1 [215]. In more than two-thirds of supratentorial
and chemotherapy is highly toxic to the developing brain. ependymoma samples, a fusion of the oncogene RELA to an
Complication and morbidity observed in patients with uncharacterized gene involving 11q13.1 occurs, which is
tumors subgrouped into good prognosis, such as the WNT absent in posterior fossa tumors. The aberrant RELA-fusion
subgroup, could correspond to a higher extent with iatro- proteins modulate activation of the transcription factor NF-
genic interventions than with the natural course of disease. kB. The largest molecular analysis of ependymoma to date,
These patients may benefit from a reduction in radiation and including 583 tissue samples, demonstrated that posterior
672 D. Coluccia et al.

fossa ependymomas can be classified in to two subgroups, ation by administration of established epigenetic drugs that
Group A and B, which differ in their genomics, transcrip- target either DNA or histone methylation [219].
tomics, location and clinical outcome [209]. Group A epen-
dymomas affect younger patients, are more laterally located
and exhibit more frequent recurrence and metastasis with an 35.4 Cell Cycle Dysregulation
overall poorer prognosis compared with Group B. The and Mitogenic Factors in Human
genome in Group A is more balanced, shows increased Brain Tumors
occurrence of chromosome 1q gains, whereas in Group B the
genome is highly unstable with frequent segmental gains of The cell cycle can be conceptualized as a circle divided by a
chromosomes 9, 15, 18, as well as losses of chromosomes 6 number of discrete phases during which nuclear content
and 22 [209]. changes in preparation for cell division (Fig. 35.5). In healthy
neural tissue, with advancing differentiation and age, most cells
will become quiescent and do not express cell cycle genes due
35.3.16 Gene Amplification, Overexpression, to the presence of cell-type specific repression mechanisms. In
Mutations, and Deletions human brain tumors, the intrinsic machinery of the cell cycle is
the target of mutations or functional derangements leading to
In ependymoma and especially in children, the overall release of repression mechanisms and rapid, uncontrolled pro-
genetic mutation rate appears to be lower than astrocytic liferation—a hallmark of cancer.
tumors. Expression profiling by microarray analysis on
pediatric ependymomas have identified a subset of genes
abnormally expressed in this tumor, including increased 35.4.1 General Mechanisms of Cell Cycle
expression of EGFR, WNT5A, TP53, and many cell cycle, Dysregulation
cell adhesion, angiogenesis, and cell proliferation genes,
with downregulated genes including SCHIP-1 and EB1 Cyclins (A-E) are a group of proteins whose expression levels
[213]. EGFR overexpression has been recorded as a prog- are differentially and tightly regulated during the cell cycle.
nostic marker in intracranial grade II ependymomas corre- When bound to their respective cyclin dependent kinases
lating with reduced progression free and overall survival (CDKs), an activated complex promotes transcription of
[216, 217]. Amplification and rearrangement of the c-erb B products required for a specific stage in the cell cycle [220].
gene has been identified, as well as rearrangements of the The activity of the various CDKs are coupled to both binding
MEN1 gene (11q13) [212, 218]. A large genomic analysis of the cyclins and a series of phosphorylations and dephos-
of ependymoma including tumors from 207 samples dem- phorylations on specific residues by cyclin-dependent kinase
onstrated focal amplification of genes involved in stem cell activating kinases (CAKs) and cyclin-dependent kinase
proliferation, pluripotency, and neuronal differentiation inhibitors (CKIs) [221, 222].
such as THAP11, PSPH, EPHB2, KCNN1, RAB3A, One of the most important cell cycle pathways known to
PTPRN2, PCDH cluster, and NOTCH1 [207]. This same be perturbed in human brain tumors, especially astrocytomas,
study identified focal deletions of known tumor suppressor is the cyclin D-CDK4-RB-E2F pathway, (RB-pathway)
genes (PTEN, INK4A/ARF) and novel genes such as STAG1 (Fig. 35.5). Here, cyclin D forms an activated complex with
and TNRC6B [207]. CDK4 and/or 6. This active complex partially phosphorylates
a previously hypophosphorylated retinoblastoma protein, RB
[221, 222]. Partial phosphorylation of RB liberates E2F-DP
35.3.17 Epigenetic Alteration such that it can act to upregulate target genes involved in the
continued regulation of cell cycle, such as cyclin E [221, 222].
Epigenetic modifiers involved in the formation of ependy- Thus, the phosphorylation of RB is a key regulatory step of
momas have recently been identified. The methylation status the G1–S checkpoint.
of CpG islands occurs more frequently in posterior fossa Whereas the cyclins and CDKs could be termed accelerators
Group A (PFA) tumors, leading to transcriptional silencing of the cell cycle, inhibitor of cyclin-dependent Kinase 4 (INK4)
of the histone methyltransferase Polycomb repressive com- and Cip/Kip could be termed the brakes of the cell cycle. INK4
plex 2 (PRC2), which in turn results in repressive effect on is a family of four proteins A-D, p16INK4a, p15INK4B, p18INK4C, and
expression of cell differentiation genes [219]. This epigene- p19INK4D. INK4A is encoded from a gene on 9q21. Interestingly,
tic mechanism predicts for the more aggressive behavior of this gene also encodes for a novel second protein, using an alter-
Group A tumors. Mice xenograft models implanted with native reading frame in exon 1 and splicing it to the common
human PFA showed a decreased tumor volume and longer exons 2 and 3. This protein is called ARF (p14ARF), alternative
survival after that PRC2 was reactivated through demethyl- reading frame, and is also involved in cycle arrest with p53.
35 Molecular Biology of Human Brain Tumors 673

Fig. 35.5 Dysregulation in cell cycle pathways involved in the pathogenesis of astrocytomas.

The Cip/Kip family of CDK inhibitors (p21CIP1, p27Kip1, cell cycle passage and promoting apoptosis. Several cell
p57Kip2) can bind and inhibit the G1/S cyclin-CDK complexes, cycle proteins interact as negative or positive regulators with
and the mitotic cyclin B/A complexes. They can also act as p53. One such mechanism of regulation of p53 is via the
both positive and negative regulators of the cell cycle. In early Murine Double Minute-2 protein (MDM2) (Fig. 35.4),
G1 phase Cip/Kip family members bind and aid in the forma- which is able to repress p53 activity both at the transcrip-
tion of cyclin D/CDK4/6 complexes without inhibiting kinase tional and protein levels [227–229]. MDM2 binds and con-
activity and therefore contribute to the formation of hyper- ceals the p53 transactivating domain and blocks the basal
phosphorylated RB and cell cycle progression [223]. Once RB transcriptional machinery for p53. In response to DNA dam-
is partially phosphorylated and transcription of Cyclin E and age both MDM2 and p53 are phosphorylated resulting in
Cdk 2 are upregulated, the Cip/Kip members then bind with conformational changes that incapacitate binding [230, 231].
higher affinity to Cyclin E/CDK 2 complexes and inhibit CDK This allows p53 to be stabilized in the nucleus where it is
2 activity and halt cell cycle progression (Fig. 35.5). able to upregulate genes involved in either cell cycle arrest
or apoptosis. The ability of MDM2 to block the p53 pathway
is reversed by ARF which binds to and sequesters MDM2 in
35.4.2 The Importance of p53 and Cell Cycle the nucleolus away from p53, as also impairs MDM2’s ubiq-
Control uitin ligase activity [232–234]. This allows for stabilization
of p53 levels and downstream activation of p21 and cell
The TP53 gene encodes the most well studied tumor sup- cycle arrest.
pressor, p53 and to date has been implicated to play a role in
almost 50 % of human cancers, including human brain
tumors [224]. Mutations in the TP53 gene can lead to mis- 35.4.3 Mitogenic Signals That Impact on Brain
sense or loss of protein expression, whereas the missense Tumor Growth
mutation is associated with earlier cancer onset and more
aggressive tumor profile [225, 226]. P53 has a multitude of The decision of the cell to divide is provided by both internal
stress-induced effects on cellular homeostasis by blocking stimuli and external conditions. External conditions favorable
674 D. Coluccia et al.

to cell division, growth, and survival are conveyed via cell PDGFRA is a surface tyrosine kinase receptor repeatedly
membrane tyrosine kinase receptors that typically dimerize found to be altered in both high and low-grade human astro-
upon ligand binding, phosphorylate their cytoplasmic tails cytomas [109, 241]. As a growth factor it interacts with the
and activate downstream signaling molecules. One key MAPK/RAS pathway, which is implicated in primary GBM
growth activator is the small GTPase Ras. It activates multi- formation, however the role of PDGFR in tumor initiation or
ple mitogenic pathways: RAF-MEF-MAPK, CDC42-RAC- progression has yet to be elucidated.
Rho, and PI3K-PTEN-AKT [235]. These downstream The tumor suppressor PTEN acts as a negative regulator of
signaling molecules subsequently activate cyclin/CDK com- the PI3K-AKT growth control pathway. PTEN mutations may
plexes triggering growth, demonstrating these pathways are be found in 30 % of primary GBMs and up to 8 % of secondary
commonly targeted during tumor mutagenesis. GBMs [114], however these numbers are inconsistent among
Epidermal growth factor receptor (EGFR), platelet the different study groups to allow conclusive interpretation
derived growth factor receptor (PDGFR), fibroblast growth regarding order in succession of tumorigenesis.
factor receptor (FGFR), C-Met, and C-kit are all tyrosine
kinases implicated in the formation of astrocytomas and
other neurological tumors. 35.5 Brain Tumor Migration and Invasion

Brain invasion by infiltrative tumor cells is one of the most


35.4.4 Analysis of Cell Cycle and Mitogenic sinister histopathological hallmarks of malignant brain
Pathway Disturbances in Brain Tumors: tumors. Brain tumor invasion is a complex cellular phenom-
Low and High Grade Astrocytomas enon including permissive changes in the extracellular milieu
and intracellular biomechanical systems. The spread of brain
Our understanding of the pathways involved in tumor initia- tumor cells within the cerebral tissue involves changes in
tion and progression has been enhanced through molecular cell/cell and cell/extracellular matrix (ECM) adhesion,
genetic analyses of cell cycle and mitogenic stimulation enzyme degradation of the ECM by proteases, and cell motil-
pathways. Genetic alterations of TP53 are present in low ity into normal brain parenchyma [252, 253]. Specifically,
grade, anaplastic and GBMs, and the genetic aberrations are cell migration requires a complex balance between extracel-
higher for the secondary GBM (up to 80 %) compared to the lular cues and responsive intracellular signals that lead to
primary GBM (around 30 %) suggesting a role in astrocy- dynamic regulation of the interactions between actin micro-
toma initiation [114, 123]. filaments, microtubules, intermediate filaments and associ-
Mutations in the RB pathway of INK4A, CDK4, and RB ated adhesions [254, 255]. The driving force for cell
represent about 80 % of GBMs and only a small portion of movement is normally provided by dynamic reorganization
low-grade astrocytomas [109] [236]. There are also differ- of the actin cytoskeleton, which directs protrusions at the
ences between primary and secondary GBMs. Deletion of the front of the cell and retraction at the rear [254, 255]. For a
CDKN2A locus leading to reduced CDK4 inhibition is found cell to move, it must establish polarity resulting in leading
in over 50 % of primary and 20 % of secondary GBMs [114, and trailing edges with directionalized forces. One of the
130, 131]. This locus encodes both INK4A and ARF and there- first steps in cell migration is the formation of actin-rich
fore disrupts both the p53 and RB pathways indirectly, per- structures called lamellipodia at the leading edge of the
haps contributing to the rapid rate of growth attributed to motile cell [256, 257]. Subsequent attachment of these pro-
primary GBMs. Alterations of RB itself are found in up to trusive structures to the substratum followed by tension
14 % of GBMs [130, 131]. CDK4 amplification is present in across the cell, generated by myosin motor proteins, will
14 % of GBMs. Other mutations found to arise in GBMs have lead to the contractile force ultimately required for cell body
been identified in CDK6, cyclin D, and cyclin E [237, 238]. translocation [256, 257].
EGFR is amplified or altered in up to 57 % of glioblasto- Ultimately, this interplay of contractile elements with
mas, mainly detected in the primary and rarely in secondary adhesive matrix increases metabolic demand and triggers
GBM, which is why this alteration might not be involved enzymes for enhanced glycolysis. As shown recently, GBM
early in tumor initiation [109, 239]. A mouse model express- cells overexpress glucose-6-phosphatase (G6P) to overcome
ing the mutated EGFR alone was insufficient to cause astro- glycolytic inhibition [258]. By downregulating glucose-6-
cytomas, but expressed in a background of TP53 +/− or phosphatase in human derived GBM cells the migration and
Ink4A/ARF−/− astrocytoma formation occurred rapidly invasion rate was markedly reduced [258].
[240]. This suggests that EGFR may be responsible for Migration along the tight extracellular space requires also
tumor progression but require additional mutations in other considerable adaptation in volume and shape of glioma cells
genes for tumor initiation. [253, 259], which is mediated by modulation of Cl− and K+
35 Molecular Biology of Human Brain Tumors 675

Fig. 35.6 Mechanisms involved in the migration and invasion of Glioblastoma (GBM) cells along the perivascular space and white matter tracts.

channels. The Na+-K+-Cl− co-transporter (NKCC1) is highly vessels and meninges resembles the movement observed in
expressed in glioma cells, which accumulates intracellular neural progenitor cells during tissue maturation [253].
Cl− leading to opening of Cl−channels (ClC3) and resulting There have been numerous reviews describing many of
in volume shrinkage [260]. In recent studies, glioma migra- the extracellular factors relating to invading brain tumor cells
tion and invasion could be decreased after inhibiting NKCC1 such as the proteolytic degradation by various proteases
or ClC3 channels [261, 262]. (MMPs, cathepsins, serine proteases, urokinase plasminogen
In contrast to solid tumors of other organs, malignant activator receptor), and ECM deposition and adhesion facili-
glioma cells usually do not spread through an intravascular tated by matrix receptors [252, 264, 265]. The increased
route but rather within the perivascular zone. Moreover, lym- release of glutamate by glioma cells additionally acts as an
phatic channels do not exist in the brain. Despite their pro- autocrine and paracrine ligand supporting Ca2+ triggered cell
found parenchymal invasiveness, metastasis outside the migration and invasion [266]. Indeed, blocking intracellular
brain is negligible in primary malignant gliomas [263], glutamate uptake through sulfasalazine resulted in reduced
which is likely due to the blood–brain barrier and an unsuit- tumor spreading in a rodent model [266]. There are also
able environment for growth in other organ systems. The many well characterized genetic lesions in brain tumors
migration is mainly directed into two compartments: the which have been shown to affect astrocytoma progression
perivascular space and the brain parenchyma, which then [254, 267, 268], but the understanding of the intracellular
results in a spreading of glioma satellites at distant sites from and molecular mechanisms that mediate brain tumor inva-
the main tumor mass [253] (Fig. 35.6). The alignment sion is still in its infancy. These molecular pathways include
towards existing structures such as white matter tracts, blood signaling regulated by (1) the non-receptor tyrosine kinases
676 D. Coluccia et al.

focal adhesion kinase (FAK) and proline-rich tyrosine kinase mation. The most important mitogen orchestrating vascular
2 (Pyk2), (2) members of the Rho family of small GTPases, neogenesis in brain tumors is vascular endothelial growth
(RhoA and Rac1) including signaling by lysophosphatidic factor, VEGF.
acid (LPA) and sphingosine-1-phosphate (S1P), and (3) the
nuclear factor-kB (NF-kB) family of transcription factors.
The Rho family of small GTPases is essential in cellular 35.6.1 Mechanisms of Angiogenesis
migration pathways. RhoA activation mediates formation in Human Astrocytomas
of stress fibers and focal adhesions in the trailing edge of
the cells and leads to a rounded amoeboid form of move- A distinguishing histopathological feature of astrocytoma is
ment, while Rac1 directs actin assembly in lamellipodia at endothelial proliferation, which can be 40 times higher than
the leading edge of migrating cells which results in cells in normal brain [278]. VEGF overexpression is central to the
moving in an elongated mesenchymal manner [269, 270]. neovascularization observed in these high-grade lesions such
Attempts to reduce cell motility by blocking migration as GBM [279–281]. Monoclonal antibodies directed against
pathways have to take into account that cells can switch VEGF in GBM mice xenografts and withdrawal of VEGF
between these different modes of migration. CDC42 is a from implanted GBM cell lines in mice resulted in vessel
well-studied member of the Rho family and, while not regression underscoring the importance of VEGF in these
essential for cellular movement, aids in regulating filopodia lesions [282, 283].
formation and cellular polarity [271, 272]. Inhibition of VEGF is a 34–45 kDa glycosylated homodimeric protein
Rac1 with siRNA leads to decreased lamellipodia forma- that acts as a specific mitogen and chemotactic factor for
tion and cellular movement [273, 274]. Using a dominant endothelial cells [284]. There are five other VEGF homologs,
negative Rac1 inhibitor induced death on GBM cells but VEGF-A, VEGF-B, VEGF-C, VEGF-D, VEGF-E, and pla-
not normal astrocytes [275]. Additionally, it was shown that cental growth factor (PGF) [285]. VEGF, itself, is composed
downstream inhibition of Rho-Kinase (downstream of RhoA of five splice variants, which have equivalent effects in
activation) lead to increase motility through activation of regards to angiogenesis [286]. Expression of VEGF results in
Rac1 [276]. Therefore, it may be a balance between RhoA increased permeability of the microvasculature, endothelial
and Rac1 that regulates cellular migration and invasion in production of proteases to break down basement membrane,
astrocytoma cells. endothelial survival, and endothelial proliferation [287].
The greatest limitation of brain tumor invasion research to The effects of VEGF are exerted via the tyrosine kinase
date has been the lack of a reliable and reproducible in vivo receptors vascular endothelial growth factor receptor one and
animal model of brain invasion. Current applications of two, VEGFR1-2, and neuropilin-1 (Nrp-1). Binding of VEGF
existing animal models are currently not optimized or char- to VEGFR-2 promotes endothelial proliferation, whereas
acterized for use in brain tumor invasion research. Induction binding to VEFGR-1 regulates endothelial cells [288].
of invasion in vivo, tracking infiltrative cells and quantitation VEGF-2 is considered a core driver in complete activation of
of brain invasion are the most critical considerations of a live the angiogenic cascade and thus progression of neovascular-
animal model of brain tumor invasion. ization observed in high grade astrocytomas [280].
Central to the expression of VEGF is hypoxia. Increased
transcription of VEGF mRNA in vitro can be efficiently
35.6 Brain Tumor Angiogenesis induced by hypoxic conditions and upon restoration of oxygen
levels, basal VEGF expression is returned [289]. Rapidly pro-
Angiogenesis is defined as the formation of new blood ves- liferating tumors, such as GBMs, quickly outstrip their blood
sels from the sprouting of previously existing vessels. In the supply creating a hypoxic and necrotic environment. Indeed,
fetal brain, angiogenesis begins during embryogenesis when VEGF expression of cells at the rim of necrosis shows 50
soluble angiogenic factors are secreted to cause capillary times basal VEGF expression [290, 291]. This is mediated
sprouting from the extracerebral plexus and neovasculariza- through hypoxia inducible factor, HIF-1 [292].
tion. In contrast, the adult brain shows very little angiogenic HIF-1 is a heterodimer consisting of an α and β subunit.
activity due to expression of antiangiogenic factors. Under normal physiological conditions the α subunit is ubiq-
Solid tumors beyond 2 mm in diameter require recruit- uitinated by the tumor suppressor von Hippel–Lindau, VHL
ment of blood vessels for continued growth [277]. In patho- [293]. As a result of hypoxia, HIF-1α interacts with the
logical states such as tumors, an angiogenic signaling switch hypoxia-response element leading to the increased expression
occurs that results in a cascade of events including enzymatic and upregulation of pro-angiogenesis genes, such as VEGF
degradation of basement membrane proteins, endothelial [292]. HIF-1 also cooperates in the stabilization of the VEGF
migration, endothelial proliferation, and tubular vessel for- protein and increasing VEGF secretion [294]. VEGF expression
35 Molecular Biology of Human Brain Tumors 677

can also be upregulated by fibroblast growth factor (FGF),


PDGF, EGF, tumor necrosis factor α, (TNF-α), transforming
growth factor β (TGF-β), and interleukins 1 and 6 [287].
Angiopoietins are a four-member family of secreted ligands
that bind to the endothelial specific tyrosine kinase Tie-2.
The two best understood angiopoietins are angiopoietin 1
(Ang-1), and angiopoietin 2 (Ang-2), which act antagonisti-
cally at the Tie-2 receptor. In the presence of VEGF, Ang-2
promotes vessel destabilization in preparation for neovascu-
larization [295]. The combination of VEGF and Ang-2 is
found at the hypoxic rim of GBMs [296], whereas Ang-1
stimulation of Tie-2 results in vessel stabilization and is
found at the tumor periphery [287].
Bevacizumab, a human monoclonal antibody against
VEGF-A, leads to a marked reduction in tumor size and pro-
longation of progression-free and overall survival in GBM
patients in promising preclinical and clinical studies [297,
298]. Two highly awaited and large Phase III randomized
trials have been conducted recently (RTOG-0825 and
AVAGlio trial) [299, 300]; however neither study demon-
strated any benefit in overall survival, albeit a slight increase
in progression-free survival was observed.

35.7 Developmental Signaling Pathways


and Human Brain Tumors

During normal neuroembryogenesis multiple signaling path-


ways are active to allow fast proliferation and increased
migration of neural progenitor cells. The phase of differen-
tiation and maturation requires a precisely adapted regula-
tion of expression and silencing of the involved signaling
networks. In the past decade, an increasing number of genes
and pathways involved in this maturation process have been
discovered and characterized. As advanced molecular analy-
sis in human brain tumors recurrently revealed mutations
and other alterations in these genes, it is assumed that such
alterations are involved in the early phase of brain cancer
initiation and eventually enable tumor cells to acquire prolif-
Fig. 35.7 Notch signaling pathway.
eration and migration proprieties similarly found in primitive
neural cells. Here we discuss three developmental signaling
cascades with proven associations with formation of human growth factor-like ligand domain. When engaged with ligand
brain tumors. from the Jagged or Delta families, the extracellular compo-
nent gets cleaved by ADAM metalloproteases, which then
leads to a second cleavage at the transmembrane region by a
35.7.1 The Notch Signaling Pathway γ-secretase. This second proteolytic event releases the Notch
intracellular domain which promotes transcription of Hes
The Notch signaling pathway mediates cell–cell signaling family genes, among others [301] (Fig. 35.7).
and is a key developmental pathway regulating cell decisions The functions of the Notch genes (Notch1-4) are time-
regarding cell survival, maintenance of stem cell self- dependent and cellular context-dependent, and thus, they may
renewal, proliferation, differentiation, and apoptosis [301– act as an oncogene or a tumor suppressor [304–306].
303]. The Notch protein forms a transmembrane receptor, Activation of Notch signaling is known in various human neo-
whereas the extracellular region contains an epidermal plasms such as cervical carcinomas, T-cell leukemia, pancreatic
678 D. Coluccia et al.

cancer, and mucoepidermoid carcinoma [307, 308]. In the for pattering the midline structures in brain and spinal cord as
nervous system, Notch regulates stem cells in developing and well as controlling of axonal spreading and stem cell prolifera-
modeling the shape and cytoarchitecture of brain and spinal tion [330]. Proper regulation of SHH is particularly critical for
cord [309, 310]. Notch upregulation was observed in glioma normal development of the cerebellum, explaining the fre-
cell lines and primary human glioblastomas resulting in stem quent implication of this pathway in medulloblastomas
cell maintenance and tumor proliferation [311, 312]. when one or more genes of this pathway are dysregulated
Specifically, the self-renewal of stem cell-like GBM cells is [331, 332]. Through analysis of gene expression in medul-
sustained in the perivascular stem cell niche by endothelial loblastomas, it has been shown that tumors with mutations
cells, which provide Notch ligands, promoting radioresis- in SHH pathway express high levels of Gli1, as well as N-myc
tance [313, 314]. Notch inactivation by γ-secretase inhibitors and C-myc, cyclin D1 and D2 [333]. Interestingly, sporadic
(GSI) resulted in increased radiosensitivity through regula- Gorlin’s syndrome and medulloblastomas are often character-
tion of the PI3K/Akt pathway [314]. GSI induced suppres- ized by inactivation of Ptch1 gene or activation of SMO [80,
sion of Notch can also enhance the antitumoral effect of 184, 334]. Bmi1, a polycomb group gene, is involved in the
temozolomide (TMZ) [315]. From the four GBM subgroups, clonal expansion of granule cell precursors, and its overex-
the proneural type particularly shows high Notch pathway pression is linked to that of Ptch and is ultimately required for
activation and sensitivity to GSI in the mouse model [316]. SHH driven tumorigenesis [331, 332]. A recent genome
Notch signaling deregulation also occurs in medulloblas- sequencing of SHH pathway medulloblastoma showed that
tomas, primarily within the SHH subgroup, where Notch 2 the most frequently detected gene mutations were found in
amplification and overexpression correlated with tumor pro- PTCH1 (45 %), SMO (14 %), and SUFU (8 %) [335].
gression and poor prognosis [317–319]. Interestingly, inhibit- Interestingly, while PTCH1 mutations were found at roughly
ing the Notch pathway in medulloblastoma cell lines using equal distribution in infants, children, and adults, SMO altera-
inhibitors of GSI could decrease the number of CD133+ cells tions were markedly more frequent in adults, and SUFU
and block in vivo engraftments [320]. The role of Notch in almost exclusively in infants under 3 years [335]. On the other
medulloblastoma genesis and how the Notch pathway inter- hand, children older than 3 years exhibited strong MYCN and
acts with the SHH pathway remains to be elucidated given GLI2 amplification together with TP53 mutations (possibly in
that recent studies demonstrate that Notch signaling may not correlation with Li–Fraumeni syndrome), while all of these
be crucial for SHH-driven medulloblastomas [321–323]. findings were rare in infants and adults [335].
Several studies reported an upregulation of the Notch SHH pathway is best studied for the pathogenesis of medul-
pathway in ependymomas, supporting its potential role in the loblastoma, but there is increasing data revealing involvement
pathogenesis of these tumors [203, 324, 325]. Taylor et al. in glioma incurrence. Amplification and overexpression of
identified genetically distinct ependymomas by patterns of Gli1 has been associated with some low- and high-grade astro-
gene expression that recapitulated those of radial glial cells cytomas [336, 337], and is assumed that the SHH pathway
in corresponding regions of the CNS [203]. Specifically, it might have a stronger role in glioma initiation compared to
was shown that supratentorial tumors expressed markedly other developmental pathways such as Notch or WNT [338].
elevated levels of Notch family members. Furthermore, it has been demonstrated that CD133+ glioblas-
toma cancer stem cells (CSC) are promoted by endothelial
cells through SHH pathways and inhibition of SHH by SMO
35.7.2 The Sonic Hedgehog Signaling Pathway knockdown in vitro and in vivo results in reduction of CSC
phenotype-appearance [337]. High expression of SHH and
The SHH signaling pathway plays a primary role in mor- Gli1 in glioma patient samples were correlated with both
phogenic development of different organs during embryo- increasing WHO grade and worse prognosis [337].
genesis [326]. The secreted Sonic Hedgehog ligand, encoded Therapeutic approaches targeting the SHH pathway have
on chromosome 7q36, binds and inactivates the transmem- mainly focused on SMO inhibition [335, 339]. Preclinical
brane receptor Patched (PTCH), which results in releasing and clinical studies appeared initially promising with evi-
the inhibition of Smoothened (SMO) [304, 327]. SMO sub- dence of marked reduction in tumor growth; however, point
sequently acts to activate transcription factors of the GLI mutations in SMO and GLI2 amplifications eventually lead
family to express target genes, while SUFU (Suppressor of to drug inefficacy [340, 341]. Recent xenograft models
Fused) regulates Gli (Gli1–4) proteins to inhibit the SHH revealed that SHH medulloblastomas with PTCH1 muta-
pathway [327, 328] A model of SHH signaling is shown in tions were responsive to SMO-inhibition, while tumors with
Fig. 35.8. SUFU or MYNC mutations, which are often found in infants,
Dysregulation of the SHH pathway is implicated in cancers showed resistance [335]. The addition of PI3K inhibitors to
of different tissues, including the brain, lung, mammary gland, SMO antagonist therapy may be a strategy to delay the
prostate, and skin [329]. In the nervous system, SHH is crucial development of resistance [340, 341].
35 Molecular Biology of Human Brain Tumors 679

Fig. 35.8 Sonic Hedgehog


(SHH) signaling pathway.

35.7.3 The Wingless Pathway differentiation [345–347]. The canonical WNT pathway is
also a critical regulator of stem cells properties, and its acti-
The Wingless (WNT) proteins form a family of secretory mol- vation has been associated with various cancers [348].
ecules that regulate cell–cell interactions during embryogene- Evidence for the involvement of the WNT signaling path-
sis. To date, 19 WNT genes and more than 15 associated way in brain tumors has come from studies of Turcot’s syn-
receptors have been identified in the human genome [304, drome, characterized by the development of medulloblastomas
342]. WNT binds to the G-protein-coupled receptor Frizzled, and astrocytomas resulting from a germ-line mutation in the
which results in intracellular signaling carried out by the phos- APC gene [188]. Since then it has been established that WNT
phoprotein Dishevelled (DSH) and leads to a release of inhibi- signaling influences the proliferation and the renewal of neu-
tion of the transcription factor β-catenin [343]. With inactive ral stem cells and progenitors [347]. There is increasing evi-
WNT signaling, the cytoplasmic concentration of β-catenin is dence that aberrant activation of WNT signaling, resulting in
regulated and decreased by a destruction complex, which accumulation or increased activity of β-catenin, will lead to
mainly is formed by Axin, APC, GCK-3, and CKI [344]. the development of sporadic brain tumors, Notably, pediatric
A current model of WNT signaling is shown in Fig. 35.9. medulloblastomas harbor mutations of β-catenin, APC, and
Through its interactions with other families of secreted Axin [188, 329]. These point mutations are found in approxi-
growth factors, including EGFR, FGF, TGF-β, and the mately 10–15 % of medulloblastomas but are mutually
Hedgehog proteins, WNT signaling can regulate diverse pro- exclusive. For instance, 4 % of sporadic medulloblastomas
cesses including cell proliferation, migration, polarity and harbor missense mutations in APC, and 1–5 % have Axin1
680 D. Coluccia et al.

Fig. 35.9 Wingless/Int 1


(WNT) signaling pathway.

point mutations [186, 349]. Interestingly, there is evidence


of cross-talk between the WNT and SHH signaling path- 35.8 Stem Cells in Brain Tumors
ways as SUFU has been shown to regulate the activity of
β-catenin [350]. SUFU mutations in medulloblastomas may The cancer stem cell (CSC) hypothesis was initially derived
lead to the activation of the two pathways: WNT and SHH from the concept of CSCs giving rise to the hematopoietic
[351]. The role of WNT signaling in brain tumor formation malignancies. Since then the CSC hypothesis attempts to
has primarily been studied in the context of medulloblas- explain the development of solid tumors, such as brain
toma, but there is increasing evidence that WNT signaling tumors. Furthermore, the CSC hypothesis provides impor-
may also modulate the early phase of tumorigenesis in some tant concepts for future brain tumor treatment. The failure or
gliomas [346, 347, 352]. only short lived beneficial effect of chemotherapy drugs and
So far, agents targeted against the WNT pathway in radiotherapy in addition to rapid tumor recurrence for most
medulloblastoma cells or tumors have only been explored in malignant brain cancers lead to the assumption that actual
preclinical setting [353, 354]; however given that medullo- therapies might not adequately target a subset of dormant or
blastoma patients harboring mutations involving WNT sig- slow cycling cells with abilities of self-renewal and adaptation
naling show favorable outcome compared to other subgroups [105]. Brain tumor stem cells are hypothesized to reside
[198], WNT targeted strategies seem to be of lower priority. within a nutritive perivascular niche [355] (Fig. 35.10).
35 Molecular Biology of Human Brain Tumors 681

Fig. 35.10 Glioblastoma


stem cells and perivascular
niche. GBM glioblastoma
multiforme.

As in normal tissue, tumors are organized in a cellular and mas, ependymomas and medulloblastomas, skin malignancies,
functional hierarchy based on stem cells. Neural stem cells and colon cancer [203, 367–373].
(NSC) are defined as cells that have the ability to perpetuate Evidence supporting the CSC hypothesis in brain tumors,
themselves through self-renewal (symmetric division) and to include finding that these cells have indefinite potential for
generate mature cells of a particular tissue through differen- self-renewal, proliferation, differentiation, and have clono-
tiation (asymmetric division) [356, 357]. These cells have genic potential based on neurosphere assays. Several studies
been isolated from the embryonic brain of humans and per- demonstrate a close link between developmental biology,
sist in the adult in the subventricular zone of the frontal lat- stem cells and cancer cells. In vitro, CSCs have been shown
eral ventricles and dentate gyrus of the hippocampus [105]. to express molecular markers of neural precursors as Sox2,
Different markers have been proposed to characterize the Bmi1, Musashi, Notch, Emx2, Pax6, and Jagged1 [374].
NSC population, but none have proven to be specific enough These findings suggest that CSCs may have been initially
to ensure the identification of a pure population. The expres- derived from normal stem cells. Several proteins have been
sion of Nestin is a widely used marker to delineate NSCs, as proposed as markers of the CDC population in brain tumors
is the cell surface marker CD133. Additionally, some NSCs including CD133, CD15, α-6 integrin, and ATP-Binding
have characteristics of glial cells and will express on their cassette transporters (ABCT) [371, 375–377]. However,
surface SSEA-1, CD44, α-6 integrin, BLBP, GLAST, RC2, other studies have shown that the negative populations of
and GFAP markers [358–361]. these markers can also give rise to tumors and therefore these
Many genes that have been identified as regulators of neu- markers may only enrich for CSC and not define CSC [378].
ral stem cells properties are also known to be involved in Importantly, it is not clear yet if cancer-initiating events
brain tumorigenesis [361]. For instance, Bmi-1 is important occur in the NSC, progenitor cells, or terminally differenti-
for cell proliferation and maintenance, and PTEN acts as a ated cells, which reacquire a de-differentiated state [379].
tumor suppressor and important negative regulator of cellu- Recent studies have suggested the microenvironment is
lar proliferation [362, 363]. The Notch, Sonic Hedgehog, important in defining the CSC. In melanoma, for example,
and Wingless pathways have also been implicated in regulat- the complete obliteration of the immune response of the
ing NSCs [348, 364–366]. Importantly, deregulation of these xenograft host lead to a larger fraction of the tumor popula-
pathways is also strongly implicated in the initiation and tion exhibiting CSC properties [380].
progression of CNS cancers when they are deregulated. Using a high-throughput drug screen, Diamandis et al.
To date, CSCs have been identified in many solid tumors, demonstrated that the functional ground state of neural stem
including breast carcinoma, CNS tumors including astrocyto- cells may not only be dependent on developmental signaling
682 D. Coluccia et al.

Table 35.2 Pros and cons of common animal models of brain tumors
Brain tumor in vivo model Pros Cons
Orthotopic xenograft Cell lines can be genetically Brain tumor cell lines generally
manipulated prior to transplantation in show genotypic deviation and
order to explore specific genetic events homogeneity compared to the
Large numbers of animals can be original patient tumor (applies
generated in relatively short time and less to biopsy xenografts, at the
at low breading costs cost of longer engraftment time
Relative consistent histology and and more variable tumor
chromosomal profile supports histology)
standardization Mostly demonstrate an expansive
growth with only limited
infiltration into the perivascular
space and occurrence of necrosis
or microvascular alterations
(better results with sphere cultures
compared to monolayer cell lines)
Inhibited immunological
tumor-host response
Chemically induced model Immune competent animals can be Genotype and phenotype of the
used tumor is not reproducible
Tumors show relative high degree of Histologically present more like
invasive growth gliosarcomas or “glioma-like
tumors”

Genetically engineered Tumors arise de novo and in situ Time point of tumor initiation
mouse model (GEMM) within immune competent animals cannot be controlled
Induction of defined genetic alterations Biological differences of murine
allow to study the role of oncogenes and human malignancies
and tumor suppressors in Complex and expensive breading
tumorigenesis
Intact immune system and distinct
molecular characterization of induced
brain tumors improves validation of
drug testing
Transplantable, chemically induced, and genetically engineered mouse models (GEMM)

pathways such as Notch, SHH, and WNT, but also on path- we describe transplantable, chemically induced, and genetically
ways relating to neural transmission [381]. engineered models [242] (Table 35.2).
Although the cancer stem cell hypothesis remains some-
what controversial it offers a unique prospective to examine
and understand brain tumor biology. The properties required 35.9.1 Transplantable Allograft
for a population of cells to become tumorigenic requires a and Xenograft Models
dynamic interplay between genetic aberrations, tumor hier-
archy, and microenvironment. Classic modeling involves transplantation of cultured rodent
(allograft) or human (xenograft) cell lines under the skin
(subcutaneous) or in the brain (orthotropic) of immunodefi-
35.9 Animal Models of Brain Tumors cient/immunonaive rodents [382, 383]. These rodent models
are highly reproducible in terms of tumor formation, growth,
The ideal animal model of a human brain tumor recapitulates and survival pattern. The cell lines can be genetically manipu-
the molecular, genetic, histopathologic, and clinical features lated prior to transplantation in order to explore specific
that are found in the human tumor. While none of the current genetic events. However, these tumors may not faithfully
animal models can fully achieve these requirements, these mod- recapitulate the original tumor histology, and pathobiology
els are critically important as they can be used to investigate [384]. Established brain tumor cell lines generally show
tumor biology, screen novel molecular targets and pave the way genotypic deviation and homogeneity compared to the origi-
towards preclinical trials. In the past 10 years, great progress has nal patient tumor [385]. One strategy to overcome this is the
been made in establishing several relevant and reproducible ani- use of orthotopic serial passaging of primary human brain
mal models of human brain tumors that are based on the molec- tumors in mice, as these xenografts often retain the original
ular genetic alterations that characterize the human tumors. Here tumor phenotype. A common drawback of tumor transplants
35 Molecular Biology of Human Brain Tumors 683

is that they mostly demonstrate an expansive growth with tenance, although mouse tumors that phenotypically resem-
only limited infiltration into the perivascular space of the ble human tumors still behave differently with respect to
brain and rare occurrence of necrosis or microvascular altera- their genetics and therapeutic response [391].
tions [242]. The advancement in cell selection and prepara- The generation of transgenic mice involves inserting a
tion of growth media now allows for cultivating cancer DNA construct (usually an oncogene) downstream of a pro-
stem-like cells which grow as spheres [242, 368]. These moter into a fertilized oocyte, which is then implanted into the
sphere cultures closely retain the genetic profile of the patient mouse [392, 393]. All cells of the resulting transgenic off-
tumor and are highly tumorigenic, invasive, and angiogeneic spring homogenously contain the new genetic material.
[242, 386]. However, EGFR amplification is rather rare in Conditional expression can be conferred using tissue/develop-
neurosphere cultures and some cell populations lose glial and mental stage specific promoters, drug response elements such
gain mesenchymal features in a process known as mesenchy- as tetracycline or doxycycline, or recombination techniques
mal drift [387]. The retention of the initial chromosomal pro- (Cre/lox) [394, 395]. Strategies employing conditional expres-
file and tissue architecture of the tumor can be improved sion allow the study of genes that would otherwise result in
when biopsy tissue is directly cultivated as multicellular embryonic lethality, and confer tissue and stage specificity.
organotypic spheroids and implanted into immunodeficient Knock-out mice or loss-of-function models traditionally
rodents [388]. Biopsy xenograft models usually maintain involve using homologous recombination to substitute a
their heterogeneity and show invasiveness, necrotic areas and mutated gene with a selection marker (i.e., antibiotic resis-
frequently display EGFR amplification [242]. However given tant gene) for a wild type gene. The cells of interest are then
that the tumor histology differs between animals and initial selected for and placed into embryonic stem cells, which are
engraftment may take up to 1 year, experimental planning then transferred to mouse embryos and generate chimeric
becomes problematic [242]. mice, which can be mated to generate a homozygous mouse
with the inactivated gene [393, 396]. These mice as well as
transgenic mice can be mated for a combination of genetic
35.9.2 Induction of Brain Tumors modifications.
via Embryonic Mutagenesis Somatic cell transfer techniques allow the infection of
only a population of cells in an effort to more closely model
The second classical method of modeling is inducing tumor the spontaneous genetic events responsible for the majority
formation in the developing brain of embryonic rodents with of brain tumors. These techniques employ the use of viral
early exposure to mutagenic alkylating agents (chemically vectors for the delivery of genetic information. One strategy
induced model) [389, 390]. These tumors, mostly estab- takes advantage of the Moloney Murine Leukemia Virus
lished in rats (e.g., 9L, C6), show some degree of invasion (MMLV) and its wild-type helper virus to deliver genes to
and necrosis, but histologically present more like gliosarco- dividing cells [397]. A second more specific system uses the
mas or glioma-like tumors [242]. An advantage over trans- replication incompetent avian leukosis virus (ALV) to only
planted tumors is that immunocompetent animals can be deliver genes to cells engineered to express the viral receptor
used, which mimics the situation normally found in humans. tv-a [398, 399]. The ability to study secondary mutations in
However, when cell lines from chemically induced tumors these systems is related to infection efficiency, as well, the
are used as an allograft transplant, immunogenic tumor inclusion of large genes in the vectors decreases viral effi-
rejection may still occur and interfere with the interpretation ciency. However, multiple viral constructs can be coinfected
of the effect of experimental antitumoral agents. An essential to study multiple genetic hits.
disadvantage of chemically induced tumors in embryonic
animals is that the genotype and phenotype of the tumor is
not reproducible, given the inciting genetic event is unknown 35.9.4 Models of Astrocytoma
and reveals little about the molecular etiology of brain
tumors in humans. Alterations within the three core signaling pathways (p53,
RB, RTK/RAS) as characterized by the TCGA are often used
to induce astrocytoma formation in GEM models [391].
35.9.3 Genetically Engineered Mouse Transgenic mice with constitutively active H-RAS driven by
Models (GEMM) a GFAP promoter (astrocyte specific) develop Grade II–IV
astrocytomas in a dose dependent manner [400]. High H-RAS
GEMMs are created by germ-line modifications or somatic expressors develop tumors characteristic of GBM and moder-
cell gene transfer techniques. An advantage over xenografts ate expressors develop low-grade or anaplastic astrocytomas
is that the tumors arise de novo and in situ within immune [393, 400].
competent animals. GEMMs have allowed further insight Mice with tumors characteristic of GBMs can be created
into the molecular events behind tumor initiation and main- by breeding mice with TP53 deletions with mice with Nf1
684 D. Coluccia et al.

(the mouse form of the human NF1) deletions [401]. Nf1, as tumors increased the latency to tumor formation [407].
in humans, is a RAS-GAP protein that downregulates RAS Recently, oligodendroglioma cells from fresh human tissue
activation, therefore, its loss would result in RAS activation. samples with co-deletion of chromosomes 1p and 19q have
Transgenic mice with v-src expression driven by a GFAP been successfully cultivated as permanent cell lines (BT054
promoter develop low grade or anaplastic astrocytomas and BT088). BT088 was shown to engraft in the brain of
14.4 % of the time [393, 402]. The src receptor interacts with immunocompromised mice, and demonstrated extensive
the EGFR and PDGFR signaling pathways. infiltration and histology resembling anaplastic oligoden-
Anaplastic astrocytomas were modeled in mice engineered droglioma [408].
to express the SV40 T-antigen late in development. SV40
inactivates RB and RB family member p107 and p103. These
mice develop anaplastic astrocytomas within 10 months in 35.9.6 Models of Medulloblastoma
almost 100 % of cases [393, 403]. Interestingly, mutating RB
alone does not result in astrocytomas suggesting redundancy Mice heterozygous for mouse PTCH deletion develop
between RB family members. tumors similar to human medulloblastoma at a rate of 15 %
Using the RCAS/tv-a retroviral somatic cell transfer tech- by 10 months [409]. This incidence can be increased to 98 %
nique to over express K-RAS and AKT in nestin positive by 12 weeks if the PTCH +/- mice are combined with p53
CNS progenitor cells results in the formation of tumors char- null mice [409]. Using the RCAS retroviral and inducing
acteristic of GBM [404]. Neither K-RAS nor AKT overex- Sonic Hedgehog mis-expression in the cerebellum of fetal or
pression alone formed tumors. AKT over expression may be newborn mice resulted in medulloblastoma formation at
analogous to PTEN deletions in human GBM as the inhibi- high rates [410, 411]. The co-overexpression of C-MYC in
tory effects of PTEN are immediately upstream of AKT acti- these mice enhanced the effect [393]. In murine transgenic
vation [393]. Marumoto et al. transduced a small number of models, the overexpression of MYC in concert with loss of
GFAP+cells heterozygous for TP53 with lentivirus harboring TP53 resulted in medulloblastomas that expressed markers
activated oncogenes H-RAS and AKT in the cortex, subven- typical of Group 3 medulloblastoma [412]. Interestingly,
tricular zone and hippocampus of mice [405]. Tumors that while trying to create astrocytoma models in GFAP express-
recapitulated human GBM arose only from transduced cells ing cells of mice by deletion of both TP53 and RB, tumors
in the subventricular zone or hippocampus [405]. phenotypically similar to medulloblastoma developed in the
Deletions and alterations in tumor suppressor or onco- cerebellum [413]. The WNT subgroup could be recapitu-
genes can be combined in order to generate mouse models of lated in mice with activating mutations in the WNT pathway
GBM subgroups. For example, knocking down p53 and acti- effect protein CTNNB1 [414]. These WNT pathway tumors
vating RAS signaling resulted in tumors harboring a molecu- in the mouse model revealed that this subgroup of medullo-
lar profile resembling the mesenchymal GBM subtype [104]. blastoma arise from regions in the dorsal brainstem as
Furthermore, the first orthotopic xenograft model of anaplas- opposed to other genetically distinct medulloblastoma
tic oligodendroglioma with mutations in IDH1 and tumor subgroups such as SHH which arise from the cerebellar
suppressor genes FUB1 and CIC (both involved in MYC hemispheres [414].
expression and RTK signaling repression) has been reported
recently [406].
35.9.7 Models of Ependymoma

35.9.5 Models of Oligodendroastrocytoma For a long time, preclinical models of ependymoma were
lacking [415]. Johnson et al. introduced the first mouse
Overactive PDGFR has been implicated in the pathogenesis model in 2010 using genetic engineering techniques [207].
of human oligodendroastrocytoma tumors and as such After transducing the oncogene EPHB2 in neural stem cells
murine models of this tumor have focused on this abnormal- that were INK4A/ARF−/− the mice developed tumors that
ity. Holland et al. used the RCAS retroviral system to were histologically indistinguishable from human ependy-
express oncogenic PDGFB in either nestin or GFAP express- momas at a rate of 50 %. Advances in optimum medium
ing cells [407]. The nestin group acquired mainly low grade preparation have permitted ependymoma cell culture derived
oligodendroastrocytomas approximately 60 % of the time, directly from primary patient material and to create ortho-
whereas the GFAP group developed mixed oligoastrocyto- topic xenograft mice models for posterior fossa ependy-
mas or oligodendroastrocytomas 40 % of the time [407]. moma [416]. Recently, an ependymoma stem cell line
When the oncogenic PDGFB mice were crossed with mice (DKFZ-EP1NS) derived from a patient with metastasizing
lacking the INK4A/ARF gene, the proportion of anaplastic supratentorial anaplastic ependymoma has been cultivated as
35 Molecular Biology of Human Brain Tumors 685

neurospheres and used to establish the first orthotopic animal 15. DeBella K, Szudek J, Friedman JM. Use of the national institutes
of health criteria for diagnosis of neurofibromatosis 1 in children.
model for supratentorial ependymoma harboring homozy- Pediatrics. 2000;105:608–14.
gous deletion for CDKN2A [417]. 16. Valero MC, Martin Y, Hernandez-Imaz E, et al. A highly sensitive
genetic protocol to detect NF1 mutations. J Mol Diagn. 2011;13:
113–22.
17. Listernick R, Ferner RE, Liu GT, Gutmann DH. Optic pathway
35.10 Conclusion gliomas in neurofibromatosis-1: controversies and recommenda-
tions. Ann Neurol. 2007;61:189–98.
Continued pursuits in understanding the molecular and 18. Easton DF, Ponder MA, Huson SM, Ponder BA. An analysis of
genetic events in brain tumor initiation, maintenance and variation in expression of neurofibromatosis (NF) type 1 (NF1): evi-
dence for modifying genes. Am J Hum Genet. 1993;53:305–13.
progression will lead to improved therapies in the future. It 19. Carey JC, Viskochil DH. Neurofibromatosis type 1: A model con-
will one day be possible to analyze the molecular biology of dition for the study of the molecular basis of variable expressivity
a particular patient’s brain tumor and tailor individual thera- in human disorders. Am J Med Genet. 1999;89:7–13.
pies based on these findings. The goal of improving patient 20. Gutmann DH, Parada LF, Silva AJ, Ratner N. Neurofibromatosis
type 1: modeling CNS dysfunction. J Neurosci. 2012;32:
outcomes in this devastating disease will only be accom- 14087–93.
plished through the continued application of the latest tech- 21. Marchuk DA, Saulino AM, Tavakkol R, et al. cDNA cloning of the
nologies to the field. type 1 neurofibromatosis gene: complete sequence of the NF1
gene product. Genomics. 1991;11:931–40.
22. Cawthon RM, O’Connell P, Buchberg AM, et al. Identification
and characterization of transcripts from the neurofibromatosis 1
References region: the sequence and genomic structure of EVI2 and mapping
of other transcripts. Genomics. 1990;7:555–65.
1. CBTRUS statistical report: primary brain and central nervous 23. Xu GF, O’Connell P, Viskochil D, et al. The neurofibromatosis
system tumors diagnosed in the United States in 2005–2009 2011 type 1 gene encodes a protein related to GAP. Cell. 1990;62:
(cited 2014 July); www.cbtrus.org. 599–608.
2. Ohgaki H, Kleihues P. Epidemiology and etiology of gliomas. 24. Abulencia A, Acosta D, Adelman J, et al. Measurement of the tt
Acta Neuropathol. 2005;109:93–108. production cross section in pp collisions at square root of s = 1.96
3. Ostrom QT, Bauchet L, Davis FG, et al. The epidemiology of gli- TeV. Phys Rev Lett. 2006;97:082004.
oma in adults: a “state of the science” review. Neuro Oncol. 25. Jett K, Friedman JM. Clinical and genetic aspects of neurofibro-
2014;16:896–913. matosis 1. Genet Med. 2010;12:1–11.
4. Malmer B, Iselius L, Holmberg E, et al. Genetic epidemiology of 26. Gutmann DH, Wood DL, Collins FS. Identification of the neurofi-
glioma. Br J Cancer. 2001;84:429–34. bromatosis type 1 gene product. Proc Natl Acad Sci U S A.
5. Institute NC. Cancer Stat Rev 1975–2011. (July 2014). http://seer. 1991;88:9658–62.
cancer.gov. 27. Xu GF, Lin B, Tanaka K, et al. The catalytic domain of the neuro-
6. de Martel C, Ferlay J, Franceschi S, et al. Global burden of can- fibromatosis type 1 gene product stimulates ras GTPase and com-
cers attributable to infections in 2008: a review and synthetic plements ira mutants of S. cerevisiae. Cell. 1990;63:835–41.
analysis. Lancet Oncol. 2012;13:607–15. 28. Dasgupta B, Gutmann DH. Neurofibromatosis 1: closing the GAP
7. Dziurzynski K, Chang SM, Heimberger AB, et al. Consensus on between mice and men. Curr Opin Genet Dev. 2003;13:20–7.
the role of human cytomegalovirus in glioblastoma. Neuro Oncol. 29. Dasgupta B, Gutmann DH. Neurofibromin regulates neural stem
2012;14:246–55. cell proliferation, survival, and astroglial differentiation in vitro
8. Soderberg-Naucler C, Rahbar A, Stragliotto G. Survival in patients and in vivo. J Neurosci. 2005;25:5584–94.
with glioblastoma receiving valganciclovir. N Engl J Med. 30. Kourea HP, Cordon-Cardo C, Dudas M, Leung D, Woodruff
2013;369:985–6. JM. Expression of p27(kip) and other cell cycle regulators in
9. Louis DN, Ohgaki H, Wiestler OD, et al. The 2007 WHO classifi- malignant peripheral nerve sheath tumors and neurofibromas: the
cation of tumours of the central nervous system. Acta Neuropathol. emerging role of p27(kip) in malignant transformation of neurofi-
2007;114:97–109. bromas. Am J Pathol. 1999;155:1885–91.
10. Weller M, Stupp R, Hegi ME, et al. Personalized care in neuro- 31. Packer RJ, Gutmann DH, Rubenstein A, et al. Plexiform neurofi-
oncology coming of age: why we need MGMT and 1p/19q testing bromas in NF1: toward biologic-based therapy. Neurology.
for malignant glioma patients in clinical practice. Neuro Oncol. 2002;58:1461–70.
2012;14 Suppl 4:iv100–8. 32. Korf BR. Malignancy in neurofibromatosis type 1. Oncologist.
11. Northcott PA, Korshunov A, Pfister SM, Taylor MD. The clinical 2000;5:477–85.
implications of medulloblastoma subgroups. Nat Rev Neurol. 33. Evans DG, Baser ME, McGaughran J, et al. Malignant peripheral
2012;8:340–51. nerve sheath tumours in neurofibromatosis 1. J Med Genet.
12. Dempfle A, Wudy SA, Saar K, et al. Evidence for involvement of 2002;39:311–4.
the vitamin D receptor gene in idiopathic short stature via a 34. Menon AG, Anderson KM, Riccardi VM, et al. Chromosome 17p
genome-wide linkage study and subsequent association studies. deletions and p53 gene mutations associated with the formation of
Hum Mol Genet. 2006;15:2772–83. malignant neurofibrosarcomas in von Recklinghausen neurofibro-
13. Friedman JM. Epidemiology of neurofibromatosis type 1. Am matosis. Proc Natl Acad Sci U S A. 1990;87:5435–9.
J Med Genet. 1999;89:1–6. 35. Cichowski K, Shih TS, Schmitt E, et al. Mouse models of tumor devel-
14. National Institutes of Health Consensus Development Conference. opment in neurofibromatosis type 1. Science. 1999;286:2172–6.
Neurofibromatosis. Conference statement. Arch Neurol. 1988; 36. Ward BA, Gutmann DH. Neurofibromatosis 1: from lab bench to
45:575–8. clinic. Pediatr Neurol. 2005;32:221–8.
686 D. Coluccia et al.

37. Gutmann, D.H. (2014) Eliminating barriers to personalized medi- 60. Catapano D, Muscarella LA, Guarnieri V, et al. Hemangioblastomas
cine: Learning from neurofibromatosis type 1. Neurology. of central nervous system: molecular genetic analysis and clinical
38. Gutmann DH, Blakeley JO, Korf BR, Packer RJ. Optimizing bio- management. Neurosurgery. 2005;56:1215–21.
logically targeted clinical trials for neurofibromatosis. Expert 61. Wanebo JE, Lonser RR, Glenn GM, Oldfield EH. The natural his-
Opin Investig Drugs. 2013;22:443–62. tory of hemangioblastomas of the central nervous system in
39. Study of RAD001 (everolimus) for children with NF1 and patients with von Hippel-Lindau disease. J Neurosurg.
chemotherapy-refractory radiographic progressive low grade glio- 2003;98:82–94.
mas. 2014 (cited 2014 July); http://clinicaltrials.gov. 62. Pause A, Lee S, Lonergan KM, Klausner RD. The von Hippel-
40. Evans DG. Neurofibromatosis type 2 (NF2): a clinical and molec- Lindau tumor suppressor gene is required for cell cycle exit upon
ular review. Orphanet J Rare Dis. 2009;4:16. serum withdrawal. Proc Natl Acad Sci U S A. 1998;95:993–8.
41. Asthagiri AR, Parry DM, Butman JA, et al. Neurofibromatosis 63. Koochekpour S, Jeffers M, Wang PH, et al. The von Hippel-
type 2. Lancet. 2009;373:1974–86. Lindau tumor suppressor gene inhibits hepatocyte growth factor/
42. Evans DG, Baser ME, O'Reilly B, et al. Management of the scatter factor-induced invasion and branching morphogenesis in
patient and family with neurofibromatosis 2: a consensus confer- renal carcinoma cells. Mol Cell Biol. 1999;19:5902–12.
ence statement. Br J Neurosurg. 2005;19:5–12. 64. Peruzzi B, Athauda G, Bottaro DP. The von Hippel-Lindau
43. Rouleau GA, Wertelecki W, Haines JL, et al. Genetic linkage of tumor suppressor gene product represses oncogenic beta-catenin
bilateral acoustic neurofibromatosis to a DNA marker on chromo- signaling in renal carcinoma cells. Proc Natl Acad Sci U S A.
some 22. Nature. 1987;329:246–8. 2006;103:14531–6.
44. Rouleau GA, Merel P, Lutchman M, et al. Alteration in a new gene 65. An J, Liu H, Magyar CE, et al. Hyperactivated JNK is a therapeu-
encoding a putative membrane-organizing protein causes neuro- tic target in pVHL-deficient renal cell carcinoma. Cancer Res.
fibromatosis type 2. Nature. 1993;363:515–21. 2013;73:1374–85.
45. Trofatter JA, MacCollin MM, Rutter JL, et al. A novel moesin-, 66. Chen L, Han L, Zhang K, et al. VHL regulates the effects of
ezrin-, radixin-like gene is a candidate for the neurofibromatosis miR-23b on glioma survival and invasion via suppression of HIF-
2 tumor suppressor. Cell. 1993;72:791–800. 1alpha/VEGF and beta-catenin/Tcf-4 signaling. Neuro Oncol.
46. Cooper J, Giancotti FG. Molecular insights into NF2/Merlin 2012;14:1026–36.
tumor suppressor function. FEBS Lett. 2014;588:2743–52. 67. Kwiatkowski DJ. Tuberous sclerosis: from tubers to mTOR. Ann
47. Lim SH, Ardern-Holmes S, McCowage G, de Souza P. Systemic Hum Genet. 2003;67:87–96.
therapy in neurofibromatosis type 2. Cancer Treat Rev. 68. Jozwiak S, Schwartz RA, Janniger CK, Bielicka-Cymerman
2014;40:857–61. J. Usefulness of diagnostic criteria of tuberous sclerosis complex
48. McClatchey AI. Merlin and ERM proteins: unappreciated roles in in pediatric patients. J Child Neurol. 2000;15:652–9.
cancer development? Nat Rev Cancer. 2003;3:877–83. 69. Goodman M, Lamm SH, Engel A, et al. Cortical tuber count: a
49. Ammoun S, Cunliffe CH, Allen JC, et al. ErbB/HER receptor acti- biomarker indicating neurologic severity of tuberous sclerosis
vation and preclinical efficacy of lapatinib in vestibular schwan- complex. J Child Neurol. 1997;12:85–90.
noma. Neuro Oncol. 2010;12:834–43. 70. Short MP, Richardson EP, Haines JL, Kwiatkowski DJ. Clinical,
50. McClatchey AI, Saotome I, Ramesh V, Gusella JF, Jacks T. The neuropathological and genetic aspects of the tuberous sclerosis
Nf2 tumor suppressor gene product is essential for extraembry- complex. Brain Pathol. 1995;5:173–9.
onic development immediately prior to gastrulation. Genes Dev. 71. Weiner DM, Ewalt DH, Roach ES, Hensle TW. The tuberous scle-
1997;11:1253–65. rosis complex: a comprehensive review. J Am Coll Surg.
51. Giovannini M, Robanus-Maandag E, van der Valk M, et al. 1998;187:548–61.
Conditional biallelic Nf2 mutation in the mouse promotes mani- 72. Sancak O, Nellist M, Goedbloed M, et al. Mutational analysis of
festations of human neurofibromatosis type 2. Genes Dev. the TSC1 and TSC2 genes in a diagnostic setting:
2000;14:1617–30. genotype--phenotype correlations and comparison of diagnostic
52. Subbiah V, Slopis J, Hong DS, et al. Treatment of patients with DNA techniques in Tuberous Sclerosis Complex. Eur J Hum
advanced neurofibromatosis type 2 with novel molecularly tar- Genet. 2005;13:731–41.
geted therapies: from bench to bedside. J Clin Oncol. 73. Jones AC, Shyamsundar MM, Thomas MW, et al. Comprehensive
2012;30:e64–8. mutation analysis of TSC1 and TSC2-and phenotypic correlations
53. Karajannis MA, Legault G, Hagiwara M, et al. Phase II trial of in 150 families with tuberous sclerosis. Am J Hum Genet.
lapatinib in adult and pediatric patients with neurofibromatosis 1999;64:1305–15.
type 2 and progressive vestibular schwannomas. Neuro Oncol. 74. Huang J, Dibble CC, Matsuzaki M, Manning BD. The TSC1-
2012;14:1163–70. TSC2 complex is required for proper activation of mTOR com-
54. Hippel E. Ueber eine sehr seltene Erkrankung der Netzhaut. plex 2. Mol Cell Biol. 2008;28:4104–15.
Albrecht von Graefes Archiv für Ophthalmologie. 75. Krueger DA, Care MM, Holland K, et al. Everolimus for subepen-
1904;59:83–106. dymal giant-cell astrocytomas in tuberous sclerosis. N Engl
55. Lindau A. Zur Frage der Angiomatosis Rentinae und Ihrer J Med. 2010;363:1801–11.
Hirnkomplikationen. Acta Ophtalmologica. 1926;4:193–226. 76. Krueger DA, Care MM, Agricola K, et al. Everolimus long-term
56. Robinson CM, Ohh M. The multifaceted von Hippel-Lindau safety and efficacy in subependymal giant cell astrocytoma.
tumour suppressor protein. FEBS Lett. 2014;588:2704–11. Neurology. 2013;80:574–80.
57. Latif F, Tory K, Gnarra J, et al. Identification of the von Hippel- 77. Evans DG, Farndon PA, Burnell LD, Gattamaneni HR, Birch
Lindau disease tumor suppressor gene. Science. JM. The incidence of Gorlin syndrome in 173 consecutive cases of
1993;260:1317–20. medulloblastoma. Br J Cancer. 1991;64:959–61.
58. Maddock IR, Moran A, Maher ER, et al. A genetic register for von 78. Jones EA, Sajid MI, Shenton A, Evans DG. Basal cell carcinomas
Hippel-Lindau disease. J Med Genet. 1996;33:120–7. in gorlin syndrome: a review of 202 patients. J Skin Cancer.
59. Richards FM, Payne SJ, Zbar B, et al. Molecular analysis of de 2011;2011:217378.
novo germ-line mutations in the von Hippel-Lindau disease gene. 79. Gorlin RJ. Nevoid basal cell carcinoma syndrome. Dermatol Clin.
Hum Mol Genet. 1995;4:2139–43. 1995;13:113–25.
35 Molecular Biology of Human Brain Tumors 687

80. Hahn H, Wicking C, Zaphiropoulous PG, et al. Mutations of the 102. Backman SA, Stambolic V, Suzuki A, et al. Deletion of Pten in
human homolog of Drosophila patched in the nevoid basal cell mouse brain causes seizures, ataxia and defects in soma size
carcinoma syndrome. Cell. 1996;85:841–51. resembling Lhermitte-Duclos disease. Nat Genet. 2001;29:
81. Li FP, Fraumeni Jr JF. Rhabdomyosarcoma in children: epidemio- 396–403.
logic study and identification of a familial cancer syndrome. J Natl 103. Kwon CH, Zhu X, Zhang J, et al. Pten regulates neuronal soma
Cancer Inst. 1969;43:1365–73. size: a mouse model of Lhermitte-Duclos disease. Nat Genet.
82. Achatz MI, Olivier M, Le Calvez F, et al. The TP53 mutation, 2001;29:404–11.
R337H, is associated with Li-Fraumeni and Li-Fraumeni-like syn- 104. Friedmann-Morvinski D, Bushong EA, Ke E, et al. Dedifferentiation
dromes in Brazilian families. Cancer Lett. 2007;245:96–102. of neurons and astrocytes by oncogenes can induce gliomas in mice.
83. Gonzalez KD, Noltner KA, Buzin CH, et al. Beyond Li Fraumeni Science. 2012;338:1080–4.
syndrome: clinical characteristics of families with p53 germ-line 105. Stiles CD, Rowitch DH. Glioma stem cells: a midterm exam.
mutations. J Clin Oncol. 2009;27:1250–6. Neuron. 2008;58:832–46.
84. Bell DW, Varley JM, Szydlo TE, et al. Heterozygous germ line 106. Capelle L, Fontaine D, Mandonnet E, et al. Spontaneous and ther-
hCHK2 mutations in Li-Fraumeni syndrome. Science. apeutic prognostic factors in adult hemispheric World Health
1999;286:2528–31. Organization Grade II gliomas: a series of 1097 cases: clinical
85. Watanabe T, Vital A, Nobusawa S, Kleihues P, Ohgaki H. article. J Neurosurg. 2013;118:1157–68.
Selective acquisition of IDH1 R132C mutations in astrocytomas 107. Sturm D, Bender S, Jones DT, et al. Paediatric and adult glioblas-
associated with Li-Fraumeni syndrome. Acta Neuropathol. toma: multiform (epi)genomic culprits emerge. Nat Rev Cancer.
2009;117:653–6. 2014;14:92–107.
86. Hamilton SR, Liu B, Parsons RE, et al. The molecular basis of 108. Ohgaki H, Dessen P, Jourde B, et al. Genetic pathways to glioblas-
Turcot’s syndrome. N Engl J Med. 1995;332:839–47. toma: a population-based study. Cancer Res. 2004;64:6892–9.
87. Turcot J, Despres JP, St Pierre F. Malignant tumors of the central 109. Brennan CW, Verhaak RG, McKenna A, et al. The somatic
nervous system associated with familial polyposis of the colon: genomic landscape of glioblastoma. Cell. 2013;155:462–77.
report of two cases. Dis Colon Rectum. 1959;2:465–8. 110. Hirose Y, Sasaki H, Miwa T, et al. Whole genome analysis from
88. van Meir E, de Tribolet N. Microsatellite instability in human microdissected tissue revealed adult supratentorial grade II-III
brain tumors. Neurosurgery. 1995;37:1231–2. gliomas are divided into clinically relevant subgroups by genetic
89. Hobert JA, Eng C. PTEN hamartoma tumor syndrome: an over- profile. Neurosurgery. 2011;69:376–90.
view. Genet Med. 2009;11:687–94. 111. Hirose Y, Aldape K, Bollen A, et al. Chromosomal abnormalities
90. Lloyd 2nd KM, Dennis M. Cowden’s disease. A possible new subdivide ependymal tumors into clinically relevant groups. Am
symptom complex with multiple system involvement. Ann Intern J Pathol. 2001;158:1137–43.
Med. 1963;58:136–42. 112. Hirose Y, Sasaki H, Abe M, et al. Subgrouping of gliomas on the
91. Lhermitte D. Sur un ganglioneurome diffus du cortex du cervelet. basis of genetic profiles. Brain Tumor Pathol. 2013;30:203–8.
Bull Assoc Fr Etud Cancer. 1920;9:99–107. 113. Houillier C, Mokhtari K, Carpentier C, et al. Chromosome 9p and
92. Padberg GW, Schot JD, Vielvoye GJ, Bots GT, de Beer FC. 10q losses predict unfavorable outcome in low-grade gliomas.
Lhermitte-Duclos disease and Cowden disease: a single phakoma- Neuro Oncol. 2010;12:2–6.
tosis. Ann Neurol. 1991;29:517–23. 114. Ohgaki H, Kleihues P. The definition of primary and secondary
93. Eng C. PTEN: one gene, many syndromes. Hum Mutat. glioblastoma. Clin Cancer Res. 2013;19:764–72.
2003;22:183–98. 115. Sanborn JZ, Salama SR, Grifford M, et al. Double minute chro-
94. Myers MP, Pass I, Batty IH, et al. The lipid phosphatase activity of mosomes in glioblastoma multiforme are revealed by precise
PTEN is critical for its tumor suppressor function. Proc Natl Acad reconstruction of oncogenic amplicons. Cancer Res. 2013;73:
Sci U S A. 1998;95:13513–8. 6036–45.
95. Orloff MS, Eng C. Genetic and phenotypic heterogeneity in the 116. Reifenberger J, Reifenberger G, Liu L, et al. Molecular genetic
PTEN hamartoma tumour syndrome. Oncogene. 2008;27:5387–97. analysis of oligodendroglial tumors shows preferential allelic
96. Liaw D, Marsh DJ, Li J, et al. Germ-line mutations of the PTEN deletions on 19q and 1p. Am J Pathol. 1994;145:1175–90.
gene in Cowden disease, an inherited breast and thyroid cancer 117. Bettegowda C, Agrawal N, Jiao Y, et al. Mutations in CIC and
syndrome. Nat Genet. 1997;16:64–7. FUBP1 contribute to human oligodendroglioma. Science.
97. Marsh DJ, Coulon V, Lunetta KL, et al. Mutation spectrum and 2011;333:1453–5.
genotype-phenotype analyses in Cowden disease and Bannayan- 118. Yip S, Butterfield YS, Morozova O, et al. Concurrent CIC muta-
Zonana syndrome, two hamartoma syndromes with germ-line tions, IDH mutations, and 1p/19q loss distinguish oligodendro-
PTEN mutation. Hum Mol Genet. 1998;7:507–15. gliomas from other cancers. J Pathol. 2012;226:7–16.
98. Tan MH, Mester J, Peterson C, et al. A clinical scoring system for 119. Ren X, Cui X, Lin S, et al. Co-deletion of chromosome 1p/19q
selection of patients for PTEN mutation testing is proposed on the and IDH1/2 mutation in glioma subsets of brain tumors in Chinese
basis of a prospective study of 3042 probands. Am J Hum Genet. patients. PLoS One. 2012;7:e32764.
2011;88:42–56. 120. Cairncross G, Wang M, Shaw E, et al. Phase III trial of chemora-
99. Orloff MS, He X, Peterson C, et al. Germ-line PIK3CA and AKT1 diotherapy for anaplastic oligodendroglioma: long-term results of
mutations in Cowden and Cowden-like syndromes. Am J Hum RTOG 9402. J Clin Oncol. 2013;31:337–43.
Genet. 2013;92:76–80. 121. Iwadate Y, Matsutani T, Hasegawa Y, et al. Favorable long-term
100. Zhou XP, Marsh DJ, Morrison CD, et al. Germ-line inactivation of outcome of low-grade oligodendrogliomas irrespective of 1p/19q
PTEN and dysregulation of the phosphoinositol-3-kinase/Akt status when treated without radiotherapy. J Neurooncol.
pathway cause human Lhermitte-Duclos disease in adults. Am 2011;102:443–9.
J Hum Genet. 2003;73:1191–8. 122. van den Bent MJ, Brandes AA, Taphoorn MJ, et al. Adjuvant pro-
101. Abel TW, Baker SJ, Fraser MM, et al. Lhermitte-Duclos disease: carbazine, lomustine, and vincristine chemotherapy in newly
a report of 31 cases with immunohistochemical analysis of the diagnosed anaplastic oligodendroglioma: long-term follow-up of
PTEN/AKT/mTOR pathway. J Neuropathol Exp Neurol. EORTC brain tumor group study 26951. J Clin Oncol.
2005;64:341–9. 2013;31:344–50.
688 D. Coluccia et al.

123. Bourne TD, Schiff D. Update on molecular findings, management 144. Wick W, Weller M, van den Bent M, et al. MGMT testing-the
and outcome in low-grade gliomas. Nat Rev Neurol. 2010; challenges for biomarker-based glioma treatment. Nat Rev Neurol.
6:695–701. 2014;10:372–85.
124. Dunn GP, Rinne ML, Wykosky J, et al. Emerging insights into the 145. Yan H, Bigner DD, Velculescu V, Parsons DW. Mutant metabolic
molecular and cellular basis of glioblastoma. Genes Dev. enzymes are at the origin of gliomas. Cancer Res. 2009;69:9157–9.
2012;26:756–84. 146. Warburg O. On the origin of cancer cells. Science.
125. Watanabe K, Tachibana O, Sata K, et al. Overexpression of the 1956;123:309–14.
EGF receptor and p53 mutations are mutually exclusive in the 147. David CJ, Chen M, Assanah M, Canoll P, Manley JL. HnRNP pro-
evolution of primary and secondary glioblastomas. Brain Pathol. teins controlled by c-Myc deregulate pyruvate kinase mRNA
1996;6:217–23. discussion 223–214. splicing in cancer. Nature. 2010;463:364–8.
126. Ekstrand AJ, James CD, Cavenee WK, et al. Genes for epidermal 148. Wolf A, Agnihotri S, Micallef J, et al. Hexokinase 2 is a key medi-
growth factor receptor, transforming growth factor alpha, and epi- ator of aerobic glycolysis and promotes tumor growth in human
dermal growth factor and their expression in human gliomas glioblastoma multiforme. J Exp Med. 2011;208:313–26.
in vivo. Cancer Res. 1991;51:2164–72. 149. Sanoudou D, Tingby O, Ferguson-Smith MA, Collins VP,
127. Hegi ME, Rajakannu P, Weller M. Epidermal growth factor recep- Coleman N. Analysis of pilocytic astrocytoma by comparative
tor: a re-emerging target in glioblastoma. Curr Opin Neurol. genomic hybridization. Br J Cancer. 2000;82:1218–22.
2012;25:774–9. 150. Zattara-Cannoni H, Gambarelli D, Lena G, et al. Are juvenile pilo-
128. Contreras CM, Azamar-Arizmendi G, Saavedra M, Hernandez- cytic astrocytomas benign tumors? A cytogenetic study in 24
Lozano M. A five-day gradual reduction regimen of chlormadi- cases. Cancer Genet Cytogenet. 1998;104:157–60.
none reduces premenstrual anxiety and depression: a pilot study. 151. Bax DA, Mackay A, Little SE, et al. A distinct spectrum of copy
Arch Med Res. 2006;37:907–13. number aberrations in pediatric high-grade gliomas. Clin Cancer
129. El Imam M, Omran M, Nugud F, et al. Obstructive uropathy in Res. 2010;16:3368–77.
Sudanese patients. Saudi J Kidney Dis Transpl. 2006;17:415–9. 152. Paugh BS, Qu C, Jones C, et al. Integrated molecular genetic pro-
130. Cancer Genome Atlas Research Network. Comprehensive filing of pediatric high-grade gliomas reveals key differences with
genomic characterization defines human glioblastoma genes and the adult disease. J Clin Oncol. 2010;28:3061–8.
core pathways. Nature. 2008;455:1061–8. 153. Rickert CH, Strater R, Kaatsch P, et al. Pediatric high-grade astro-
131. Parsons DW, Jones S, Zhang X, et al. An integrated genomic anal- cytomas show chromosomal imbalances distinct from adult cases.
ysis of human glioblastoma multiforme. Science. Am J Pathol. 2001;158:1525–32.
2008;321:1807–12. 154. Sung T, Miller DC, Hayes RL, et al. Preferential inactivation of
132. Balss J, Meyer J, Mueller W, et al. Analysis of the IDH1 codon the p53 tumor suppressor pathway and lack of EGFR amplifica-
132 mutation in brain tumors. Acta Neuropathol. 2008;116: tion distinguish de novo high grade pediatric astrocytomas from
597–602. de novo adult astrocytomas. Brain Pathol. 2000;10:249–59.
133. Nobusawa S, Watanabe T, Kleihues P, Ohgaki H. IDH1 mutations 155. Verhaak RG, Hoadley KA, Purdom E, et al. Integrated genomic
as molecular signature and predictive factor of secondary glioblas- analysis identifies clinically relevant subtypes of glioblastoma
tomas. Clin Cancer Res. 2009;15:6002–7. characterized by abnormalities in PDGFRA, IDH1, EGFR, and
134. Yan H, Parsons DW, Jin G, et al. IDH1 and IDH2 mutations in NF1. Cancer Cell. 2010;17:98–110.
gliomas. N Engl J Med. 2009;360:765–73. 156. Bar EE, Lin A, Tihan T, Burger PC, Eberhart CG. Frequent gains
135. Dubbink HJ, Taal W, van Marion R, et al. IDH1 mutations in low- at chromosome 7q34 involving BRAF in pilocytic astrocytoma.
grade astrocytomas predict survival but not response to temozolo- J Neuropathol Exp Neurol. 2008;67:878–87.
mide. Neurology. 2009;73:1792–5. 157. Jones DTW, Kocialkowski S, Liu L, et al. Tandem duplication
136. Sanson M, Marie Y, Paris S, et al. Isocitrate dehydrogenase 1 producing a novel oncogenic BRAF fusion gene defines the
codon 132 mutation is an important prognostic biomarker in glio- majority of pilocytic astrocytomas. Cancer Res. 2008;68:8673–7.
mas. J Clin Oncol. 2009;27:4150–4. 158. Jones DTW, Kocialkowski S, Liu L, et al. Oncogenic RAF1 rear-
137. van den Bent MJ, Dubbink HJ, Marie Y, et al. IDH1 and IDH2 rangement and a novel BRAF mutation as alternatives to
mutations are prognostic but not predictive for outcome in ana- KIAA1549:BRAF fusion in activating the MAPK pathway in
plastic oligodendroglial tumors: a report of the European pilocytic astrocytoma. Oncogene. 2009;28:2119–23.
Organization for Research and Treatment of Cancer Brain Tumor 159. Chen YH, Gutmann DH. The molecular and cell biology of pedi-
Group. Clin Cancer Res. 2010;16:1597–604. atric low-grade gliomas. Oncogene. 2014;33:2019–26.
138. Weller M, Felsberg J, Hartmann C, et al. Molecular predictors of 160. Davies H, Bignell GR, Cox C, et al. Mutations of the BRAF gene
progression-free and overall survival in patients with newly diag- in human cancer. Nature. 2002;417:949–54.
nosed glioblastoma: a prospective translational study of the 161. Hudson TJ, Anderson W, Artez A, et al. International network of
German Glioma Network. J Clin Oncol. 2009;27:5743–50. cancer genome projects. Nature. 2010;464:993–8.
139. Wick W, Hartmann C, Engel C, et al. NOA-04 randomized phase 162. Phillips HS, Kharbanda S, Chen R, et al. Molecular subclasses of
III trial of sequential radiochemotherapy of anaplastic glioma with high-grade glioma predict prognosis, delineate a pattern of disease
procarbazine, lomustine, and vincristine or temozolomide. J Clin progression, and resemble stages in neurogenesis. Cancer Cell.
Oncol. 2009;27:5874–80. 2006;9:157–73.
140. Ichimura K, Pearson DM, Kocialkowski S, et al. IDH1 mutations 163. Noushmehr H, Weisenberger DJ, Diefes K, et al. Identification of
are present in the majority of common adult gliomas but rare in a CpG island methylator phenotype that defines a distinct sub-
primary glioblastomas. Neuro Oncol. 2009;11:341–7. group of glioma. Cancer Cell. 2010;17:510–22.
141. Watanabe T, Nobusawa S, Kleihues P, Ohgaki H. IDH1 mutations 164. van den Bent MJ, Gravendeel LA, Gorlia T, et al. A hypermethyl-
are early events in the development of astrocytomas and oligoden- ated phenotype is a better predictor of survival than MGMT meth-
drogliomas. Am J Pathol. 2009;174:1149–53. ylation in anaplastic oligodendroglial brain tumors: a report from
142. Kondo Y, Katsushima K, Ohka F, Natsume A, Shinjo K. Epigenetic EORTC study 26951. Clin Cancer Res. 2011;17:7148–55.
dysregulation in glioma. Cancer Sci. 2014;105:363–9. 165. Wang J, Wechsler-Reya RJ. The role of stem cells and progenitors
143. Hegi ME, Diserens AC, Gorlia T, et al. MGMT gene silencing and in the genesis of medulloblastoma. Exp Neurol. 2012;260:69–73.
benefit from temozolomide in glioblastoma. N Engl J Med. 166. Kool M, Korshunov A, Remke M, et al. Molecular subgroups of
2005;352:997–1003. medulloblastoma: an international meta-analysis of transcriptome,
35 Molecular Biology of Human Brain Tumors 689

genetic aberrations, and clinical data of WNT, SHH, Group 3, 187. Kang DE, Soriano S, Xia X, et al. Presenilin couples the paired
and Group 4 medulloblastomas. Acta Neuropathol. 2012;123: phosphorylation of beta-catenin independent of axin: implications
473–84. for beta-catenin activation in tumorigenesis. Cell.
167. Bigner SH, Burger PC, Wong AJ, et al. Gene amplification in 2002;110:751–62.
malignant human gliomas: clinical and histopathologic aspects. 188. Zurawel RH, Chiappa SA, Allen C, Raffel C. Sporadic medullo-
J Neuropathol Exp Neurol. 1988;47:191–205. blastomas contain oncogenic beta-catenin mutations. Cancer Res.
168. Northcott PA, Nakahara Y, Wu X, et al. Multiple recurrent genetic 1998;58:896–9.
events converge on control of histone lysine methylation in medul- 189. Zhukova N, Ramaswamy V, Remke M, et al. Subgroup-specific
loblastoma. Nat Genet. 2009;41:465–72. prognostic implications of TP53 mutation in medulloblastoma.
169. Giordana MT, Migheli A, Pavanelli E. Isochromosome 17q is a J Clin Oncol. 2013;31:2927–35.
constant finding in medulloblastoma. An interphase cytogenetic 190. Hovestadt V, Jones DT, Picelli S, et al. Decoding the regulatory
study on tissue sections. Neuropathol Appl Neurobiol. landscape of medulloblastoma using DNA methylation sequenc-
1998;24:233–8. ing. Nature. 2014;510:537–41.
170. Pan E, Pellarin M, Holmes E, et al. Isochromosome 17q is a nega- 191. Remke M, Ramaswamy V, Taylor MD. Medulloblastoma molecu-
tive prognostic factor in poor-risk childhood medulloblastoma lar dissection: the way toward targeted therapy. Curr Opin Oncol.
patients. Clin Cancer Res. 2005;11:4733–40. 2013;25:674–81.
171. McDonald JD, Daneshvar L, Willert JR, et al. Physical mapping of 192. Di C, Liao S, Adamson DC, et al. Identification of OTX2 as a
chromosome 17p13.3 in the region of a putative tumor suppressor medulloblastoma oncogene whose product can be targeted by all-
gene important in medulloblastoma. Genomics. 1994;23:229–32. trans retinoic acid. Cancer Res. 2005;65:919–24.
172. Ferretti E, De Smaele E, Di Marcotullio L, Screpanti I, Gulino 193. Adamson DC, Shi Q, Wortham M, et al. OTX2 is critical for the
A. Hedgehog checkpoints in medulloblastoma: the chromosome maintenance and progression of Shh-independent medulloblasto-
17p deletion paradigm. Trends Mol Med. 2005;11:537–45. mas. Cancer Res. 2010;70:181–91.
173. Shih DJ, Northcott PA, Remke M, et al. Cytogenetic prognostica- 194. Bunt J, Hasselt NA, Zwijnenburg DA, et al. OTX2 sustains a
tion within medulloblastoma subgroups. J Clin Oncol. bivalent-like state of OTX2-bound promoters in medulloblastoma
2014;32:886–96. by maintaining their H3K27me3 levels. Acta Neuropathol.
174. Aldosari N, Wiltshire RN, Dutra A, et al. Comprehensive molecular 2013;125:385–94.
cytogenetic investigation of chromosomal abnormalities in human 195. Hovestadt V, Remke M, Kool M, et al. Robust molecular sub-
medulloblastoma cell lines and xenograft. Neuro Oncol. grouping and copy-number profiling of medulloblastoma from
2002;4:75–85. small amounts of archival tumour material using high-density
175. Eberhart CG, Kratz J, Wang Y, et al. Histopathological and molec- DNA methylation arrays. Acta Neuropathol. 2013;125:913–6.
ular prognostic markers in medulloblastoma: c-myc, N-myc, 196. Schwalbe EC, Williamson D, Lindsey JC, et al. DNA methylation
TrkC, and anaplasia. J Neuropathol Exp Neurol. 2004;63:441–9. profiling of medulloblastoma allows robust subclassification and
176. Eberhart CG, Kratz JE, Schuster A, et al. Comparative genomic improved outcome prediction using formalin-fixed biopsies. Acta
hybridization detects an increased number of chromosomal altera- Neuropathol. 2013;125:359–71.
tions in large cell/anaplastic medulloblastomas. Brain Pathol. 197. Taylor MD, Northcott PA, Korshunov A, et al. Molecular sub-
2002;12:36–44. groups of medulloblastoma: the current consensus. Acta
177. Gilbertson RJ, Clifford SC, MacMeekin W, et al. Expression of Neuropathol. 2012;123:465–72.
the ErbB-neuregulin signaling network during human cerebellar 198. Northcott PA, Korshunov A, Witt H, et al. Medulloblastoma
development: implications for the biology of medulloblastoma. comprises four distinct molecular variants. J Clin Oncol.
Cancer Res. 1998;58:3932–41. 2011;29:1408–14.
178. Lee CJ, Chan WI, Scotting PJ. CIC, a gene involved in cerebellar 199. Northcott PA, Lee C, Zichner T, et al. Enhancer hijacking activates
development and ErbB signaling, is significantly expressed in GFI1 family oncogenes in medulloblastoma. Nature.
medulloblastomas. J Neurooncol. 2005;73:101–8. 2014;511:428–34.
179. Gilbertson R, Hernan R, Pietsch T, et al. Novel ERBB4 juxtamem- 200. Rudin CM, Hann CL, Laterra J, et al. Treatment of medulloblas-
brane splice variants are frequently expressed in childhood medul- toma with hedgehog pathway inhibitor GDC-0449. N Engl J Med.
loblastoma. Genes Chromosomes Cancer. 2001;31:288–94. 2009;361:1173–8.
180. Gajjar A, Hernan R, Kocak M, et al. Clinical, histopathologic, and 201. Kieran MW. Targeted treatment for Sonic Hedgehog-dependent
molecular markers of prognosis: toward a new disease risk strati- medulloblastoma. Neuro Oncol. 2014;16:1037–47.
fication system for medulloblastoma. J Clin Oncol. 202. Kleihues P, Louis DN, Scheithauer BW, et al. The WHO classifi-
2004;22:984–93. cation of tumors of the nervous system. J Neuropathol Exp Neurol.
181. Erez A, Ilan T, Amariglio N, et al. GLI3 is not mutated commonly 2002;61:215–25.
in sporadic medulloblastomas. Cancer. 2002;95:28–31. 203. Taylor MD, Poppleton H, Fuller C, et al. Radial glia cells are
182. Pietsch T, Koch A, Wiestler OD. Molecular genetic studies in candidate stem cells of ependymoma. Cancer Cell.
medulloblastomas: evidence for tumor suppressor genes at the 2005;8:323–35.
chromosomal regions 1q31-32 and 17p13. Klin Padiatr. 204. Dolecek TA, Propp JM, Stroup NE, Kruchko C. CBTRUS statisti-
1997;209:150–5. cal report: primary brain and central nervous system tumors diag-
183. Reifenberger J, Wolter M, Weber RG, et al. Missense mutations in nosed in the United States in 2005-2009. Neuro Oncol. 2012;14
SMOH in sporadic basal cell carcinomas of the skin and primitive Suppl 5:v1–49.
neuroectodermal tumors of the central nervous system. Cancer 205. Ruda R, Gilbert M, Soffietti R. Ependymomas of the adult: molec-
Res. 1998;58:1798–803. ular biology and treatment. Curr Opin Neurol. 2008;21:754–61.
184. Taylor MD, Liu L, Raffel C, et al. Mutations in SUFU predispose 206. Kilday JP, Rahman R, Dyer S, et al. Pediatric ependymoma: bio-
to medulloblastoma. Nat Genet. 2002;31:306–10. logical perspectives. Mol Cancer Res. 2009;7:765–86.
185. Dahmen RP, Koch A, Denkhaus D, et al. Deletions of AXIN1, a 207. Johnson RA, Wright KD, Poppleton H, et al. Cross-species
component of the WNT/wingless pathway, in sporadic medullo- genomics matches driver mutations and cell compartments to
blastomas. Cancer Res. 2001;61:7039–43. model ependymoma. Nature. 2010;466:632–6.
186. Huang H, Mahler-Araujo BM, Sankila A, et al. APC mutations in 208. Mack SC, Taylor MD. The genetic and epigenetic basis of ependy-
sporadic medulloblastomas. Am J Pathol. 2000;156:433–7. moma. Childs Nerv Syst. 2009;25:1195–201.
690 D. Coluccia et al.

209. Witt H, Mack SC, Ryzhova M, et al. Delineation of two clinically 232. Kamijo T, Weber JD, Zambetti G, et al. Functional and physical
and molecularly distinct subgroups of posterior fossa ependy- interactions of the ARF tumor suppressor with p53 and Mdm2.
moma. Cancer Cell. 2011;20:143–57. Proc Natl Acad Sci U S A. 1998;95:8292–7.
210. Huang B, Starostik P, Kuhl J, Tonn JC, Roggendorf W. Loss of 233. Honda R, Yasuda H. Association of p19(ARF) with Mdm2 inhib-
heterozygosity on chromosome 22 in human ependymomas. Acta its ubiquitin ligase activity of Mdm2 for tumor suppressor p53.
Neuropathol. 2002;103:415–20. EMBO J. 1999;18:22–7.
211. Plotkin SR, O'Donnell CC, Curry WT, et al. Spinal ependymomas 234. Pomerantz J, Schreiber-Agus N, Liegeois NJ, et al. The Ink4a
in neurofibromatosis type 2: a retrospective analysis of 55 patients. tumor suppressor gene product, p19Arf, interacts with MDM2 and
J Neurosurg Spine. 2011;14:543–7. neutralizes MDM2's inhibition of p53. Cell. 1998;92:713–23.
212. Lamszus K, Lachenmayer L, Heinemann U, et al. Molecular 235. Besson A, Yong VW. Mitogenic signaling and the relationship to
genetic alterations on chromosomes 11 and 22 in ependymomas. cell cycle regulation in astrocytomas. J Neurooncol. 2001;
Int J Cancer. 2001;91:803–8. 51:245–64.
213. Suarez-Merino B, Hubank M, Revesz T, et al. Microarray analysis 236. Nakamura M, Yonekawa Y, Kleihues P, Ohgaki H. Promoter
of pediatric ependymoma identifies a cluster of 112 candidate hypermethylation of the RB1 gene in glioblastomas. Lab Invest.
genes including four transcripts at 22q12.1-q13.3. Neuro Oncol. 2001;81:77–82.
2005;7:20–31. 237. Costello JF, Plass C, Arap W, et al. Cyclin-dependent kinase 6
214. Ward S, Harding B, Wilkins P, et al. Gain of 1q and loss of 22 are (CDK6) amplification in human gliomas identified using two-
the most common changes detected by comparative genomic dimensional separation of genomic DNA. Cancer Res. 1997;
hybridisation in paediatric ependymoma. Genes Chromosomes 57:1250–4.
Cancer. 2001;32:59–66. 238. Riemenschneider MJ, Buschges R, Wolter M, et al. Amplification
215. Parker M, Mohankumar KM, Punchihewa C, et al. C11orf95- and overexpression of the MDM4 (MDMX) gene from 1q32 in a
RELA fusions drive oncogenic NF-kappaB signalling in ependy- subset of malignant gliomas without TP53 mutation or MDM2
moma. Nature. 2014;506:451–5. amplification. Cancer Res. 1999;59:6091–6.
216. Mendrzyk F, Korshunov A, Benner A, et al. Identification of gains 239. Ohgaki H, Kleihues P. Genetic alterations and signaling pathways
on 1q and epidermal growth factor receptor overexpression as in the evolution of gliomas. Cancer Sci. 2009;100:2235–41.
independent prognostic markers in intracranial ependymoma. Clin 240. Holland EC, Hively WP, Gallo V, Varmus HE. Modeling muta-
Cancer Res. 2006;12:2070–9. tions in the G1 arrest pathway in human gliomas: overexpression
217. Senetta R, Miracco C, Lanzafame S, et al. Epidermal growth fac- of CDK4 but not loss of INK4a-ARF induces hyperploidy in cul-
tor receptor and caveolin-1 coexpression identifies adult supraten- tured mouse astrocytes. Genes Dev. 1998;12:3644–9.
torial ependymomas with rapid unfavorable outcomes. Neuro 241. Huang H, Colella S, Kurrer M, et al. Gene expression profiling of
Oncol. 2011;13:176–83. low-grade diffuse astrocytomas by cDNA arrays. Cancer Res.
218. Diedrich U, Soja S, Behnke J, Zoll B. Amplification of the c-erbB 2000;60:6868–74.
oncogene is associated with malignancy in primary tumours of 242. Huszthy PC, Daphu I, Niclou SP, et al. In vivo models of primary
neuroepithelial tissue. J Neurol. 1991;238:221–4. brain tumors: pitfalls and perspectives. Neuro Oncol. 2012;14:
219. Mack SC, Witt H, Piro RM, et al. Epigenomic alterations define 979–93.
lethal CIMP-positive ependymomas of infancy. Nature. 243. Beier D, Hau P, Proescholdt M, et al. CD133(+) and CD133(-) glio-
2014;506:445–50. blastoma-derived cancer stem cells show differential growth charac-
220. Dirks PB, Hubbard SL, Murakami M, Rutka JT. Cyclin and teristics and molecular profiles. Cancer Res. 2007;67:4010–5.
cyclin-dependent kinase expression in human astrocytoma cell 244. Burkhard C, Di Patre PL, Schuler D, et al. A population-based
lines. J Neuropathol Exp Neurol. 1997;56:291–300. study of the incidence and survival rates in patients with pilocytic
221. Malumbres M, Barbacid M. To cycle or not to cycle: a critical astrocytoma. J Neurosurg. 2003;98:1170–4.
decision in cancer. Nat Rev Cancer. 2001;1:222–31. 245. Theeler BJ, Ellezam B, Sadighi ZS, et al. Adult pilocytic astrocy-
222. Ivanchuk SM, Rutka JT. The cell cycle: accelerators, brakes, and tomas: clinical features and molecular analysis. Neuro Oncol.
checkpoints. Neurosurgery. 2004;54:692–9. 2014;16:841–7.
223. Malumbres M. Physiological relevance of cell cycle kinases. 246. Ogiwara H, Bowman RM, Tomita T. Long-term follow-up of
Physiol Rev. 2011;91:973–1007. pediatric benign cerebellar astrocytomas. Neurosurgery.
224. Kandoth C, McLellan MD, Vandin F, et al. Mutational landscape 2012;70:40–7.
and significance across 12 major cancer types. Nature. 2013; 247. Jones DT, Ichimura K, Liu L, et al. Genomic analysis of pilocytic
502:333–9. astrocytomas at 0.97 Mb resolution shows an increasing tendency
225. Muller PA, Vousden KH. Mutant p53 in cancer: new functions and toward chromosomal copy number change with age. J Neuropathol
therapeutic opportunities. Cancer Cell. 2014;25:304–17. Exp Neurol. 2006;65:1049–58.
226. Zerdoumi Y, Aury-Landas J, Bonaiti-Pellie C, et al. Drastic effect 248. Jones DT, Gronych J, Lichter P, Witt O, Pfister SM. MAPK path-
of germ-line TP53 missense mutations in Li-Fraumeni patients. way activation in pilocytic astrocytoma. Cell Mol Life Sci.
Hum Mutat. 2013;34:453–61. 2012;69:1799–811.
227. Stewart CL, Soria AM, Hamel PA. Integration of the pRB and p53 249. Jones DT, Hutter B, Jager N, et al. Recurrent somatic alterations
cell cycle control pathways. J Neurooncol. 2001;51:183–204. of FGFR1 and NTRK2 in pilocytic astrocytoma. Nat Genet.
228. Malkin D. The role of p53 in human cancer. J Neurooncol. 2013;45:927–32.
2001;51:231–43. 250. Pfister S, Janzarik WG, Remke M, et al. BRAF gene duplication
229. Marine JC, Lozano G. Mdm2-mediated ubiquitylation: p53 and constitutes a mechanism of MAPK pathway activation in low-
beyond. Cell Death Differ. 2010;17:93–102. grade astrocytomas. J Clin Invest. 2008;118:1739–49.
230. Shieh SY, Ikeda M, Taya Y, Prives C. DNA damage-induced 251. Horbinski C, Nikiforova MN, Hagenkord JM, Hamilton RL,
phosphorylation of p53 alleviates inhibition by MDM2. Cell. Pollack IF. Interplay among BRAF, p16, p53, and MIB1 in pediat-
1997;91:325–34. ric low-grade gliomas. Neuro Oncol. 2012;14:777–89.
231. Siliciano JD, Canman CE, Taya Y, et al. DNA damage induces 252. Newton HB. Molecular neuro-oncology and the development of
phosphorylation of the amino terminus of p53. Genes Dev. targeted therapeutic strategies for brain tumors. Part 3: brain
1997;11:3471–81. tumor invasiveness. Expert Rev Anticancer Ther. 2004;4:803–21.
35 Molecular Biology of Human Brain Tumors 691

253. Cuddapah VA, Robel S, Watkins S, Sontheimer H. A neurocentric 277. Brem S. The role of vascular proliferation in the growth of brain
perspective on glioma invasion. Nat Rev Neurosci. tumors. Clin Neurosurg. 1976;23:440–53.
2014;15:455–65. 278. Plate KH, Breier G, Risau W. Molecular mechanisms of develop-
254. Raftopoulou M, Etienne-Manneville S, Self A, Nicholls S, Hall mental and tumor angiogenesis. Brain Pathol. 1994;4:207–18.
A. Regulation of cell migration by the C2 domain of the tumor 279. Plate KH, Breier G, Weich HA, Risau W. Vascular endothelial
suppressor PTEN. Science. 2004;303:1179–81. growth factor is a potential tumour angiogenesis factor in human
255. Ridley AJ, Schwartz MA, Burridge K, et al. Cell migration: inte- gliomas in vivo. Nature. 1992;359:845–8.
grating signals from front to back. Science. 2003;302:1704–9. 280. Jain RK, di Tomaso E, Duda DG, et al. Angiogenesis in brain
256. Beadle C, Assanah MC, Monzo P, et al. The role of myosin II in tumours. Nat Rev Neurosci. 2007;8:610–22.
glioma invasion of the brain. Mol Biol Cell. 2008;19:3357–68. 281. Hicklin DJ, Ellis LM. Role of the vascular endothelial growth fac-
257. Terakawa Y, Agnihotri S, Golbourn B, et al. The role of drebrin in tor pathway in tumor growth and angiogenesis. J Clin Oncol.
glioma migration and invasion. Exp Cell Res. 2013;319:517–28. 2005;23:1011–27.
258. Abbadi S, Rodarte JJ, Abutaleb A, et al. Glucose-6-phosphatase is 282. Yuan F, Chen Y, Dellian M, et al. Time-dependent vascular regres-
a key metabolic regulator of glioblastoma invasion. Mol Cancer sion and permeability changes in established human tumor xeno-
Res. 2014;12:1547–59. grafts induced by an anti-vascular endothelial growth factor/
259. Watkins S, Sontheimer H. Hydrodynamic cellular volume changes vascular permeability factor antibody. Proc Natl Acad Sci U S A.
enable glioma cell invasion. J Neurosci. 2011;31:17250–9. 1996;93:14765–70.
260. Garzon-Muvdi T, Schiapparelli P, ap Rhys C, et al. Regulation of 283. Benjamin LE, Keshet E. Conditional switching of vascular endo-
brain tumor dispersal by NKCC1 through a novel role in focal thelial growth factor (VEGF) expression in tumors: induction of
adhesion regulation. PLoS Biol. 2012;10:e1001320. endothelial cell shedding and regression of hemangioblastoma-
261. Haas BR, Sontheimer H. Inhibition of the sodium-potassium- like vessels by VEGF withdrawal. Proc Natl Acad Sci U S A.
chloride cotransporter isoform-1 reduces glioma invasion. Cancer 1997;94:8761–6.
Res. 2010;70:5597–606. 284. Leung DW, Cachianes G, Kuang WJ, Goeddel DV, Ferrara
262. Cuddapah VA, Sontheimer H. Molecular interaction and func- N. Vascular endothelial growth factor is a secreted angiogenic
tional regulation of ClC-3 by Ca2+/calmodulin-dependent protein mitogen. Science. 1989;246:1306–9.
kinase II (CaMKII) in human malignant glioma. J Biol Chem. 285. Harrigan MR. Angiogenic factors in the central nervous system.
2010;285:11188–96. Neurosurgery. 2003;53:639–60.
263. Preusser M, de Ribaupierre S, Wohrer A, et al. Current concepts 286. Jansen M, de Witt Hamer PC, Witmer AN, Troost D, van Noorden
and management of glioblastoma. Ann Neurol. 2011;70:9–21. CJF. Current perspectives on antiangiogenesis strategies in the
264. Giese A, Westphal M. Glioma invasion in the central nervous sys- treatment of malignant gliomas. Brain Res Brain Res Rev.
tem. Neurosurgery. 1996;39:235–50. 2004;45:143–63.
265. Hsieh WT, Yeh WL, Cheng RY, et al. Exogenous endothelin-1 287. Kargiotis O, Rao JS, Kyritsis AP. Mechanisms of angiogenesis in
induces cell migration and matrix metalloproteinase expression in gliomas. J Neurooncol. 2006;78:281–93.
U251 human glioblastoma multiforme. J Neurooncol. 2014; 288. Hiratsuka S, Minowa O, Kuno J, Noda T, Shibuya M. Flt-1 lacking
118:257–69. the tyrosine kinase domain is sufficient for normal development
266. Lyons SA, Chung WJ, Weaver AK, Ogunrinu T, Sontheimer and angiogenesis in mice. Proc Natl Acad Sci U S A.
H. Autocrine glutamate signaling promotes glioma cell invasion. 1998;95:9349–54.
Cancer Res. 2007;67:9463–71. 289. Shweiki D, Itin A, Soffer D, Keshet E. Vascular endothelial growth
267. Guha A, Mukherjee J. Advances in the biology of astrocytomas. factor induced by hypoxia may mediate hypoxia-initiated angio-
Curr Opin Neurol. 2004;17:655–62. genesis. Nature. 1992;359:843–5.
268. Rao RD, James CD. Altered molecular pathways in gliomas: an 290. Weindel K, Moringlane JR, Marme D, Weich HA. Detection and
overview of clinically relevant issues. Semin Oncol. 2004;31: quantification of vascular endothelial growth factor/vascular
595–604. permeability factor in brain tumor tissue and cyst fluid: the key to
269. Sahai E, Marshall CJ. Differing modes of tumour cell invasion angiogenesis? Neurosurgery. 1994;35:439–48.
have distinct requirements for Rho/ROCK signalling and extracel- 291. Plate KH, Breier G, Weich HA, Mennel HD, Risau W. Vascular
lular proteolysis. Nat Cell Biol. 2003;5:711–9. endothelial growth factor and glioma angiogenesis: coordinate
270. Seol HJ, Chang JH, Yamamoto J, et al. Overexpression of CD99 induction of VEGF receptors, distribution of VEGF protein and
Increases the Migration and Invasiveness of Human Malignant possible in vivo regulatory mechanisms. Int J Cancer.
Glioma Cells. Genes Cancer. 2012;3:535–49. 1994;59:520–9.
271. Reymond N, Im JH, Garg R, et al. Cdc42 promotes transendothe- 292. Forsythe JA, Jiang BH, Iyer NV, et al. Activation of vascular endo-
lial migration of cancer cells through beta1 integrin. J Cell Biol. thelial growth factor gene transcription by hypoxia-inducible fac-
2012;199:653–68. tor 1. Mol Cell Biol. 1996;16:4604–13.
272. Ridley AJ, Hall A. Distinct patterns of actin organization regulated 293. Semenza GL. HIF-1 and tumor progression: pathophysiology and
by the small GTP-binding proteins Rac and Rho. Cold Spring therapeutics. Trends Mol Med. 2002;8:62–7.
Harb Symp Quant Biol. 1992;57:661–71. 294. May D, Itin A, Gal O, et al. Ero1-L alpha plays a key role in a HIF-
273. Nobes CD, Hall A. Rho GTPases control polarity, protrusion, 1-mediated pathway to improve disulfide bond formation and
and adhesion during cell movement. J Cell Biol. 1999;144: VEGF secretion under hypoxia: implication for cancer. Oncogene.
1235–44. 2005;24:1011–20.
274. Bouzahzah B, Albanese C, Ahmed F, et al. Rho family GTPases 295. Lobov IB, Brooks PC, Lang RA. Angiopoietin-2 displays VEGF-
regulate mammary epithelium cell growth and metastasis through dependent modulation of capillary structure and endothelial cell
distinguishable pathways. Mol Med. 2001;7:816–30. survival in vivo. Proc Natl Acad Sci U S A. 2002;99:11205–10.
275. Senger DL, Tudan C, Guiot M-C, et al. Suppression of Rac activ- 296. Koga K, Todaka T, Morioka M, et al. Expression of angiopoietin-
ity induces apoptosis of human glioma cells but not normal human 2 in human glioma cells and its role for angiogenesis. Cancer Res.
astrocytes. Cancer Res. 2002;62:2131–40. 2001;61:6248–54.
276. Salhia B, Rutten F, Nakada M, et al. Inhibition of Rho-kinase 297. Friedman HS, Prados MD, Wen PY, et al. Bevacizumab alone and
affects astrocytoma morphology, motility, and invasion through in combination with irinotecan in recurrent glioblastoma. J Clin
activation of Rac1. Cancer Res. 2005;65:8792–800. Oncol. 2009;27:4733–40.
692 D. Coluccia et al.

298. Kreisl TN, Kim L, Moore K, et al. Phase II trial of single-agent beva- 320. Fan X, Matsui W, Khaki L, et al. Notch pathway inhibition
cizumab followed by bevacizumab plus irinotecan at tumor progres- depletes stem-like cells and blocks engraftment in embryonal
sion in recurrent glioblastoma. J Clin Oncol. 2009;27:740–5. brain tumors. Cancer Res. 2006;66:7445–52.
299. Chinot OL, Wick W, Mason W, et al. Bevacizumab plus 321. Wang Q, Li H, Liu N, et al. Correlative analyses of notch signaling
radiotherapy-temozolomide for newly diagnosed glioblastoma. N with resveratrol-induced differentiation and apoptosis of human
Engl J Med. 2014;370:709–22. medulloblastoma cells. Neurosci Lett. 2008;438:168–73.
300. Gilbert MR, Dignam JJ, Armstrong TS, et al. A randomized trial 322. Hatton BA, Villavicencio EH, Pritchard J, et al. Notch signaling is
of bevacizumab for newly diagnosed glioblastoma. N Engl J Med. not essential in Sonic Hedgehog-activated medulloblastoma.
2014;370:699–708. Oncogene. 2010;29:3865–72.
301. Kopan R, Ilagan MX. The canonical Notch signaling pathway: 323. Julian E, Dave RK, Robson JP, Hallahan AR, Wainwright BJ.
unfolding the activation mechanism. Cell. 2009;137:216–33. Canonical Notch signaling is not required for the growth of
302. Artavanis-Tsakonas S, Rand MD, Lake RJ. Notch signaling: cell Hedgehog pathway-induced medulloblastoma. Oncogene. 2010;
fate control and signal integration in development. Science. 29:3465–76.
1999;284:770–6. 324. Palm T, Figarella-Branger D, Chapon F, et al. Expression profiling
303. Lasky JL, Wu H. Notch signaling, brain development, and human of ependymomas unravels localization and tumor grade-specific
disease. Pediatr Res. 2005;57:109. tumorigenesis. Cancer. 2009;115:3955–68.
304. Samsioe A, Feinstein R, Saade G, et al. Intrauterine death, fetal 325. Puget S, Grill J, Valent A, et al. Candidate genes on chromosome
malformation, and delayed pregnancy in Ljungan virus-infected 9q33-34 involved in the progression of childhood ependymomas.
mice. Birth Defects Res B Dev Reprod Toxicol. 2006;77:251–6. J Clin Oncol. 2009;27:1884–92.
305. Purow BW, Sundaresan TK, Burdick MJ, et al. Notch-1 regulates 326. Villavicencio EH, Walterhouse DO, Iannaccone PM. The Sonic
transcription of the epidermal growth factor receptor through p53. Hedgehog-Patched-Gli pathway in human development and disease.
Carcinogenesis. 2008;29:918–25. Am J Hum Genet. 2000;67:1047–54.
306. Radtke F, Raj K. The role of Notch in tumorigenesis: oncogene or 327. Varjosalo M, Taipale J. Hedgehog signaling. J Cell Sci.
tumour suppressor? Nat Rev Cancer. 2003;3:756–67. 2007;120:3–6.
307. Sjolund J, Manetopoulos C, Stockhausen M-T, Axelson H. The 328. Lee Y, Kawagoe R, Sasai K, et al. Loss of suppressor-of-fused
Notch pathway in cancer: differentiation gone awry. Eur J Cancer. function promotes tumorigenesis. Oncogene. 2007;26:6442–7.
2005;41:2620–9. 329. Dahmane N, Sanchez P, Gitton Y, et al. The Sonic Hedgehog-Gli
308. Weng AP, Aster JC. Multiple niches for Notch in cancer: context pathway regulates dorsal brain growth and tumorigenesis.
is everything. Curr Opin Genet Dev. 2004;14:48–54. Development. 2001;128:5201–12.
309. Kageyama R, Ohtsuka T, Hatakeyama J, Ohsawa R. Roles of 330. Shahi MH, Rey JA, Castresana JS. The Sonic Hedgehog-GLI1
bHLH genes in neural stem cell differentiation. Exp Cell Res. signaling pathway in brain tumor development. Expert Opin Ther
2005;306:343–8. Targets. 2012;16:1227–38.
310. Hatakeyama J, Sakamoto S, Kageyama R. Hes1 and Hes5 regulate 331. Leung C, Lingbeek M, Shakhova O, et al. Bmi1 is essential for
the development of the cranial and spinal nerve systems. Dev cerebellar development and is overexpressed in human medullo-
Neurosci. 2006;28:92–101. blastomas. Nature. 2004;428:337–41.
311. Kanamori M, Kawaguchi T, Nigro JM, et al. Contribution of 332. Wang X, Venugopal C, Manoranjan B, et al. Sonic Hedgehog
Notch signaling activation to human glioblastoma multiforme. regulates Bmi1 in human medulloblastoma brain tumor-initiating
J Neurosurg. 2007;106:417–27. cells. Oncogene. 2012;31:187–99.
312. Purow BW, Haque RM, Noel MW, et al. Expression of Notch-1 333. Pomeroy SL, Tamayo P, Gaasenbeek M, et al. Prediction of central
and its ligands, Delta-like-1 and Jagged-1, is critical for glioma nervous system embryonal tumour outcome based on gene expres-
cell survival and proliferation. Cancer Res. 2005;65:2353–63. sion. Nature. 2002;415:436–42.
313. Zhu TS, Costello MA, Talsma CE, et al. Endothelial cells create a 334. Bale AE, Yu KP. The hedgehog pathway and basal cell carcino-
stem cell niche in glioblastoma by providing NOTCH ligands that mas. Hum Mol Genet. 2001;10:757–62.
nurture self-renewal of cancer stem-like cells. Cancer Res. 335. Kool M, Jones DT, Jager N, et al. Genome sequencing of SHH
2011;71:6061–72. medulloblastoma predicts genotype-related response to smooth-
314. Wang J, Wakeman TP, Lathia JD, et al. Notch promotes radioresis- ened inhibition. Cancer Cell. 2014;25:393–405.
tance of glioma stem cells. Stem Cells. 2010;28:17–28. 336. Kinzler KW, Bigner SH, Bigner DD, et al. Identification of an
315. Gilbert CA, Daou MC, Moser RP, Ross AH. Gamma-secretase amplified, highly expressed gene in a human glioma. Science.
inhibitors enhance temozolomide treatment of human gliomas by 1987;236:70–3.
inhibiting neurosphere repopulation and xenograft recurrence. 337. Yan GN, Yang L, Lv YF, et al. Endothelial cells promote stem-like
Cancer Res. 2010;70:6870–9. phenotype of glioma cells through activating the Hedgehog path-
316. Saito N, Fu J, Zheng S, et al. A high Notch pathway activation pre- way. J Pathol. 2014;234:11–22.
dicts response to gamma secretase inhibitors in proneural subtype of 338. Takezaki T, Hide T, Takanaga H, et al. Essential role of the
glioma tumor-initiating cells. Stem Cells. 2014;32:301–12. Hedgehog signaling pathway in human glioma-initiating cells.
317. Fan X, Mikolaenko I, Elhassan I, et al. Notch1 and notch2 have Cancer Sci. 2011;102:1306–12.
opposite effects on embryonal brain tumor growth. Cancer Res. 339. Hadden MK. Hedgehog pathway inhibitors: a patent review
2004;64:7787–93. (2009--present). Expert Opin Ther Pat. 2013;23:345–61.
318. Hallahan AR, Pritchard JI, Hansen S, et al. The SmoA1 mouse 340. Yauch RL, Dijkgraaf GJ, Alicke B, et al. Smoothened mutation
model reveals that notch signaling is critical for the growth and confers resistance to a Hedgehog pathway inhibitor in medullo-
survival of Sonic Hedgehog-induced medulloblastomas. Cancer blastoma. Science. 2009;326:572–4.
Res. 2004;64:7794–800. 341. Buonamici S, Williams J, Morrissey M, et al. Interfering with
319. Ingram WJ, McCue KI, Tran TH, Hallahan AR, Wainwright resistance to smoothened antagonists by inhibition of the PI3K
BJ. Sonic Hedgehog regulates Hes1 through a novel mechanism pathway in medulloblastoma. Sci Transl Med. 2010;2:51ra70.
that is independent of canonical Notch pathway signalling. 342. Niehrs C. The complex world of WNT receptor signalling. Nat
Oncogene. 2008;27:1489–500. Rev Mol Cell Biol. 2012;13:767–79.
35 Molecular Biology of Human Brain Tumors 693

343. Klaus A, Birchmeier W. Developmental signaling in myocardial 367. Al-Hajj M, Wicha MS, Benito-Hernandez A, Morrison SJ, Clarke
progenitor cells: a comprehensive view of Bmp- and Wnt/beta- MF. Prospective identification of tumorigenic breast cancer cells.
catenin signaling. Pediatr Cardiol. 2009;30:609–16. Proc Natl Acad Sci U S A. 2003;100:3983–8.
344. Li VS, Ng SS, Boersema PJ, et al. Wnt signaling through inhibi- 368. Galli R, Binda E, Orfanelli U, et al. Isolation and characterization
tion of beta-catenin degradation in an intact Axin1 complex. Cell. of tumorigenic, stem-like neural precursors from human glioblas-
2012;149:1245–56. toma. Cancer Res. 2004;64:7011–21.
345. Anastas JN, Moon RT. WNT signalling pathways as therapeutic 369. Gregorieff A, Pinto D, Begthel H, et al. Expression pattern of Wnt
targets in cancer. Nat Rev Cancer. 2013;13:11–26. signaling components in the adult intestine. Gastroenterology.
346. Paul I, Bhattacharya S, Chatterjee A, Ghosh MK. Current under- 2005;129:626–38.
standing on EGFR and Wnt/beta-catenin signaling in glioma and 370. Hemmati HD, Nakano I, Lazareff JA, et al. Cancerous stem cells
their possible crosstalk. Genes Cancer. 2013;4:427–46. can arise from pediatric brain tumors. Proc Natl Acad Sci U S A.
347. Holland JD, Klaus A, Garratt AN, Birchmeier W. Wnt signaling in 2003;100:15178–83.
stem and cancer stem cells. Curr Opin Cell Biol. 2013;25: 371. Singh SK, Clarke ID, Hide T, Dirks PB. Cancer stem cells in nervous
254–64. system tumors. Oncogene. 2004;23:7267–73.
348. Reya T, Clevers H. Wnt signalling in stem cells and cancer. 372. Tumbar T, Guasch G, Greco V, et al. Defining the epithelial stem
Nature. 2005;434:843–50. cell niche in skin. Science. 2004;303:359–63.
349. Baeza N, Masuoka J, Kleihues P, Ohgaki H. AXIN1 mutations but 373. Yuan X, Curtin J, Xiong Y, et al. Isolation of cancer stem cells from
not deletions in cerebellar medulloblastomas. Oncogene. adult glioblastoma multiforme. Oncogene. 2004;23:9392–400.
2003;22:632–6. 374. Dirks P. Bmi1 and cell of origin determinants of brain tumor phe-
350. Meng X, Poon R, Zhang X, et al. Suppressor of fused negatively notype. Cancer Cell. 2007;12:295–7.
regulates beta-catenin signaling. J Biol Chem. 2001;276:40113–9. 375. Read T-A, Fogarty MP, Markant SL, et al. Identification of CD15
351. Taylor MD, Zhang X, Liu L, et al. Failure of a medulloblastoma- as a marker for tumor-propagating cells in a mouse model of
derived mutant of SUFU to suppress WNT signaling. Oncogene. medulloblastoma. Cancer Cell. 2009;15:135–47.
2004;23:4577–83. 376. Ward RJ, Lee L, Graham K, et al. Multipotent CD15+ cancer stem
352. Rathod SS, Rani SB, Khan M, Muzumdar D, Shiras A. Tumor cells in patched-1-deficient mouse medulloblastoma. Cancer Res.
suppressive miRNA-34a suppresses cell proliferation and tumor 2009;69:4682–90.
growth of glioma stem cells by targeting Akt and Wnt signaling 377. Uchida N, Buck DW, He D, et al. Direct isolation of human central
pathways. FEBS Open Bio. 2014;4:485–95. nervous system stem cells. Proc Natl Acad Sci U S A.
353. Cimmino F, Scoppettuolo MN, Carotenuto M, et al. Norcantharidin 2000;97:14720–5.
impairs medulloblastoma growth by inhibition of Wnt/beta- 378. Prestegarden L, Enger PO. Cancer stem cells in the central ner-
catenin signaling. J Neurooncol. 2012;106:59–70. vous system—a critical review. Cancer Res. 2010;70:8255–8.
354. Kahn M. Can we safely target the WNT pathway? Nat Rev Drug 379. Passegue E, Jamieson CHM, Ailles LE, Weissman IL. Normal and
Discov. 2014;13:513–32. leukemic hematopoiesis: are leukemias a stem cell disorder or a
355. Calabrese C, Poppleton H, Kocak M, et al. A perivascular niche reacquisition of stem cell characteristics? Proc Natl Acad Sci U S
for brain tumor stem cells. Cancer Cell. 2007;11:69–82. A. 2003;100 Suppl 1:11842–9.
356. Reya T, Morrison SJ, Clarke MF, Weissman IL. Stem cells, cancer, 380. Quintana E, Shackleton M, Sabel MS, et al. Efficient tumour forma-
and cancer stem cells. Nature. 2001;414:105–11. tion by single human melanoma cells. Nature. 2008;456:593–8.
357. Temple S. Stem cell plasticity—building the brain of our dreams. 381. Diamandis P, Wildenhain J, Clarke ID, et al. Chemical genetics
Nat Rev Neurosci. 2001;2:513–20. reveals a complex functional ground state of neural stem cells. Nat
358. Son MJ, Woolard K, Nam DH, Lee J, Fine HA. SSEA-1 is an Chem Biol. 2007;3:268–73.
enrichment marker for tumor-initiating cells in human glioblas- 382. Shapiro WR, Basler GA, Chernik NL, Posner JB. Human brain
toma. Cell Stem Cell. 2009;4:440–52. tumor transplantation into nude mice. J Natl Cancer Inst.
359. Anido J, Saez-Borderias A, Gonzalez-Junca A, et al. TGF-beta 1979;62:447–53.
receptor inhibitors target the CD44(high)/Id1(high) glioma- 383. Kobayashi N, Allen N, Clendenon NR, Ko LW. An improved rat
initiating cell population in human glioblastoma. Cancer Cell. brain-tumor model. J Neurosurg. 1980;53:808–15.
2010;18:655–68. 384. Finkelstein SD, Black P, Nowak TP, et al. Histological character-
360. Lathia JD, Gallagher J, Heddleston JM, et al. Integrin alpha 6 istics and expression of acidic and basic fibroblast growth factor
regulates glioblastoma stem cells. Cell Stem Cell. genes in intracerebral xenogeneic transplants of human glioma
2010;6:421–32. cells. Neurosurgery. 1994;34:136–43.
361. Nicolis SK. Cancer stem cells and “stemness” genes in neuro- 385. Li A, Walling J, Kotliarov Y, et al. Genomic changes and gene
oncology. Neurobiol Dis. 2007;25:217–29. expression profiles reveal that established glioma cell lines are
362. Gargiulo G, Cesaroni M, Serresi M, et al. In vivo RNAi screen for poorly representative of primary human gliomas. Mol Cancer Res.
BMI1 targets identifies TGF-beta/BMP-ER stress pathways as 2008;6:21–30.
key regulators of neural- and malignant glioma-stem cell homeo- 386. Gunther HS, Schmidt NO, Phillips HS, et al. Glioblastoma-
stasis. Cancer Cell. 2013;23:660–76. derived stem cell-enriched cultures form distinct subgroups
363. Dirks PB. Brain tumor stem cells: the cancer stem cell hypothesis according to molecular and phenotypic criteria. Oncogene.
writ large. Mol Oncol. 2010;4:420–30. 2008;27:2897–909.
364. Uchida H, Arita K, Yunoue S, et al. Role of Sonic Hedgehog sig- 387. Schulte A, Gunther HS, Phillips HS, et al. A distinct subset of
naling in migration of cell lines established from CD133-positive glioma cell lines with stem cell-like properties reflects the tran-
malignant glioma cells. J Neurooncol. 2011;104:697–704. scriptional phenotype of glioblastomas and overexpresses CXCR4
365. Turchi L, Debruyne DN, Almairac F, et al. Tumorigenic potential as therapeutic target. Glia. 2011;59:590–602.
of miR-18A* in glioma initiating cells requires NOTCH-1 signal- 388. Bjerkvig R, Tonnesen A, Laerum OD, Backlund EO. Multicellular
ing. Stem Cells. 2013;31:1252–65. tumor spheroids from human gliomas maintained in organ culture.
366. Kim KH, Seol HJ, Kim EH, et al. Wnt/beta-catenin signaling is a J Neurosurg. 1990;72:463–75.
key downstream mediator of MET signaling in glioblastoma stem 389. Kleihues P, Lantos PL, Magee PN. Chemical carcinogenesis in the
cells. Neuro Oncol. 2013;15:161–71. nervous system. Int Rev Exp Pathol. 1976;15:153–232.
694 D. Coluccia et al.

390. Swenberg JA, Koestner A, Wechsler W. The induction of tumors 404. Holland EC. A mouse model for glioma: biology, pathology, and
of the nervous system in rats with intravenous methylnitrosourea therapeutic opportunities. Toxicol Pathol. 2000;28:171–7.
(MNU). J Neuropathol Exp Neurol. 1971;30:122. 405. Marumoto T, Tashiro A, Friedmann-Morvinski D, et al.
391. Simeonova I, Huillard E. In vivo models of brain tumors: roles of Development of a novel mouse glioma model using lentiviral vec-
genetically engineered mouse models in understanding tumor tors. Nat Med. 2009;15:110–6.
biology and use in preclinical studies. Cell Mol Life Sci. 406. Klink B, Miletic H, Stieber D, et al. A novel, diffusely infiltrative
2014;71:4007–26. xenograft model of human anaplastic oligodendroglioma with
392. Aguzzi A, Brandner S, Isenmann S, Steinbach JP, Sure mutations in FUBP1, CIC, and IDH1. PLoS One. 2013;8:e59773.
U. Transgenic and gene disruption techniques in the study of neu- 407. Dai C, Celestino JC, Okada Y, et al. PDGF autocrine stimulation
rocarcinogenesis. Glia. 1995;15:348–64. dedifferentiates cultured astrocytes and induces oligodendroglio-
393. Hesselager G, Holland EC. Using mice to decipher the molecular mas and oligoastrocytomas from neural progenitors and astrocytes
genetics of brain tumors. Neurosurgery. 2003;53:685–94. in vivo. Genes Dev. 2001;15:1913–25.
394. Rajewsky K, Gu H, Kuhn R, et al. Conditional gene targeting. 408. Kelly JJ, Blough MD, Stechishin OD, et al. Oligodendroglioma
J Clin Invest. 1996;98:600–3. cell lines containing t(1;19)(q10;p10). Neuro Oncol. 2010;12:
395. Fomchenko EI, Holland EC. Mouse models of brain tumors and 745–55.
their applications in preclinical trials. Clin Cancer Res. 409. Wetmore C, Eberhart DE, Curran T. The normal patched allele is
2006;12:5288–97. expressed in medulloblastomas from mice with heterozygous
396. Thomas KR, Capecchi MR. Site-directed mutagenesis by gene germ-line mutation of patched. Cancer Res. 2000;60:2239–46.
targeting in mouse embryo-derived stem cells. Cell. 410. Weiner HL, Bakst R, Hurlbert MS, et al. Induction of medullo-
1987;51:503–12. blastomas in mice by Sonic Hedgehog, independent of Gli1.
397. Uhrbom L, Hesselager G, Nister M, Westermark B. Induction of Cancer Res. 2002;62:6385–9.
brain tumors in mice using a recombinant platelet-derived growth 411. Wechsler-Reya RJ, Scott MP. Control of neuronal precursor
factor B-chain retrovirus. Cancer Res. 1998;58:5275–9. proliferation in the cerebellum by Sonic Hedgehog. Neuron.
398. Holland EC, Hively WP, DePinho RA, Varmus HE. A constitu- 1999;22:103–14.
tively active epidermal growth factor receptor cooperates with 412. Kawauchi D, Robinson G, Uziel T, et al. A mouse model of the
disruption of G1 cell-cycle arrest pathways to induce glioma-like most aggressive subgroup of human medulloblastoma. Cancer
lesions in mice. Genes Dev. 1998;12:3675–85. Cell. 2012;21:168–80.
399. Srivastava S, Zou ZQ, Pirollo K, Blattner W, Chang EH. Germ- 413. Marino S, Vooijs M, van Der Gulden H, Jonkers J, Berns
line transmission of a mutated p53 gene in a cancer-prone family A. Induction of medulloblastomas in p53-null mutant mice by
with Li-Fraumeni syndrome. Nature. 1990;348:747–9. somatic inactivation of Rb in the external granular layer cells of
400. Ding H, Roncari L, Shannon P, et al. Astrocyte-specific expression the cerebellum. Genes Dev. 2000;14:994–991004.
of activated p21-ras results in malignant astrocytoma formation 414. Gibson P, Tong Y, Robinson G, et al. Subtypes of medulloblas-
in a transgenic mouse model of human gliomas. Cancer Res. toma have distinct developmental origins. Nature. 2010;468:
2001;61:3826–36. 1095–9.
401. Reilly KM, Loisel DA, Bronson RT, McLaughlin ME, Jacks 415. Witt H, Korshunov A, Pfister SM, Milde T. Molecular approaches
T. Nf1;Trp53 mutant mice develop glioblastoma with evidence of to ependymoma: the next step(s). Curr Opin Neurol. 2012;
strain-specific effects. Nat Genet. 2000;26:109–13. 25:745–50.
402. Weissenberger J, Steinbach JP, Malin G, et al. Development and 416. Guan S, Shen R, Lafortune T, et al. Establishment and character-
malignant progression of astrocytomas in GFAP-v-src transgenic ization of clinically relevant models of ependymoma: a true
mice. Oncogene. 1997;14:2005–13. challenge for targeted therapy. Neuro Oncol. 2011;13:748–58.
403. Xiao A, Wu H, Pandolfi PP, Louis DN, Van Dyke T. Astrocyte 417. Milde T, Kleber S, Korshunov A, et al. A novel human high-risk
inactivation of the pRb pathway predisposes mice to malignant ependymoma stem cell model reveals the differentiation-inducing
astrocytoma development that is accelerated by PTEN mutation. potential of the histone deacetylase inhibitor Vorinostat. Acta
Cancer Cell. 2002;1:157–68. Neuropathol. 2011;122:637–50.

View publication stats

You might also like