molecular genetic study of the variability of the tuberous

66
Zentrum für Kinderheilkunde und Jugendmedizin Albert-Ludwigs-Universität Freiburg, Germany MOLECULAR GENETIC INVESTIGATION OF THE VARIABILITY OF THE GTPase ACTIVATING PROTEIN- (GAP-) RELATED DOMAIN OF THE TUBEROUS SCLEROSIS - 2 (TSC2) GENE IN TSC PATIENTS AND HEALTHY SUBJECTS Inaugural - Dissertation zur Erlangung des Medizinischen Doktorgrades der Medizinischen Fakultät der Albert-Ludwigs-Universität Freiburg i. Br. Vorgelegt 2002 von Karin Louise Zügge geb. Gierke geboren in Madison, Wisconsin/U.S.A.

Upload: others

Post on 27-Jan-2022

1 views

Category:

Documents


0 download

TRANSCRIPT

Zentrum für Kinderheilkunde und Jugendmedizin

Albert-Ludwigs-Universität

Freiburg, Germany

MOLECULAR GENETIC INVESTIGATION OF THE VARIABILITY OF THE GTPase

ACTIVATING PROTEIN- (GAP-) RELATED DOMAIN OF THE TUBEROUS

SCLEROSIS - 2 (TSC2) GENE IN TSC PATIENTS AND HEALTHY SUBJECTS

Inaugural - Dissertation

zur

Erlangung des Medizinischen Doktorgrades

der Medizinischen Fakultät

der Albert-Ludwigs-Universität

Freiburg i. Br.

Vorgelegt 2002

von Karin Louise Zügge geb. Gierke

geboren in Madison, Wisconsin/U.S.A.

Dekan: Prof. Dr. med. J. Zentner

1. Gutachter: Prof. Dr. med. L.B. Zimmerhackl

2. Gutachter: Prof. Dr. med. E. Schulz

Jahr der Promotion: 2004

i

Table of Contents 1 Introduction 1-12

1.1 History 1

1.2 Clinical characteristics 3

1.3 Pathogenesis 5

1.4 Genetics 6

1.5 Protein products 7

1.6 The clinical variability predicament 10

1.7 The Freiburg-Heidelberg project 11

1.8 Objectives and research questions 12

2 Materials and Methods 13-22

2.1 Materials 13-15

2.1.1 Apparatus and equipment 13

2.1.2 Buffers and solutions 13

2.1.3 Chemicals, enzymes, kits and size markers 14

2.1.4 Databases and software 14

2.1.5 Media and cell cultures 14

2.1.6 Oligonucleotides and DNA-sequencing 15

2.2 Methods 15-22

2.2.1 Probands 15

2.2.1.1 TSC patients and families 15

2.2.1.2 Healthy population 15

2.2.2 Cell culture and transformation from peripheral blood samples 16

2.2.3 Freezing of cells 16

2.2.4 DNA extraction 17

ii

2.2.5 PCR 18

2.2.6 PCR product analysis 19

2.2.7 SSCP 20

2.2.8 Purification of the PCR product 21

2.2.9 Sequencing 21

2.2.10 Sequence analysis 22

3 Results 23-31

3.1 Patient group 23

3.1.1 Amplification of exons 23

3.1.2 Mutational screening 23

3.1.3 Sequencing results 24

3.1.3.1 Mutations 24

3.1.3.2 Polymorphisms 26

3.1.4 Exon 40 and quality control 26

3.2 Control group of healthy probands 28

3.2.1 Exon amplification 28

3.2.2 Mutational screening 28

3.2.3 Sequencing results and numerical comparison to patient group 29

3.3 The Freiburg-Heidelberg project 30

4 Discussion 32-43

4.1 Patient group 32

4.1.1 PCR products 32

4.1.2 SSCP 32

4.1.3 Sequencing 33

4.1.3.1 Mutations 33

4.1.3.2 Polymorphisms 35

Table of Contents

iii

Table of Contents

4.1.3.3 Quality control and exon 40 36

4.2 Control group of healthy probands 36

4.2.1 Exon amplification and mutational screening 37

4.2.2 Sequencing results and numerical comparison to patient group 37

4.3 Screening the whole gene 38

4.3.1 Number and type of mutations found 38

4.3.2 Genotype-phenotype correlation 39

4.3.3 Clinical variability 40

4.3.4 TSC1 and TSC2 frequency ratios 41

4.4 Conclusions 42

4.5 Future 43

5 Summary 44

6 Zusammenfassung 45

7 Appendix 46-50

8 Abbreviations and symbols 51-52

9 Literature 53-58

10 Acknowledgments 59

11 Curriculum vitae 60

12 Publications 61

1

1 Introduction

Tuberous sclerosis complex (TSC) is not a disease that summons a definite

picture to mind for most of us. It is, however, one of the most common genetic

illnesses with a frequency of approximately 1:6000 (Kwiatkowska et al., 1999; Gomez

et al, 1999).1 A defect in one of two genes, TSC1 or TSC2, causes the disorder. A

wide variety of clinical manifestations and several different names make TSC such a

commonly unfamiliar disease.

TSC is a complex hereditary syndrome with a broad array of clinical

characteristics and thus various faces. The one unifying characteristic of TSC is the

development of benign localized tumors (hamartomas) in various tissues. The basis

of the name tuberous sclerosis is given by the cortical tubers found in the brain.

Further typical manifestations include renal angiomyolipomas and cysts, cardiac

rhabdomyomas, facial angiofibromas and pigment disorders in the skin. The severity

of manifestation is as variable as the tissues affected by the disease. The scale goes

from a few irregularities in the skin to severe multi-organ TSC with epilepsy, behavior

disorders and mental retardation, multiple kidney lesions and renal failure. (Roach et

al., 1998)

1.1 History

Not only do the varying manifestations of TSC make this a hard disease to

characterize, but also the various names given TSC over the centuries illustrate the

complexity of the disorder. The earliest documented description of a likely TSC case

is accredited to von Recklinghausen in 1862. He performed an autopsy on a newborn

child, who died shortly after birth, in which he described multiple heart tumors and

hardened areas in the brain. Bourneville described TSC as a form of epilepsy in Paris

in 1880, and thus TSC is also known as Maladie de Bourneville in France. Pringle

described the skin manifestation of Adenoma sebaceum in 1885, making TSC

synonymous with Pringle’s Disease in England. (Gomez et al., 1999)

These early scholars did not correlate their findings to other manifestations or

make the connection that these might all belong to the same disorder. It was Vogt

1 Some examples of well known inherited disorders include: galactosemia (1:50000), adrenogenital syndrome (1:10000), phenylketonuria (1:7000), hypothyroidism (1:4000), neurofibromatosis (von Recklinghausen, NF1 = 1:4000) and cystic fibrosis (1:2000). (von Harnack, 1994)

2

who first described a trias of facial angiofibromas, epileptic seizures and intellectual

handicap that could also include heart and kidney tumors in 1908. The establishment

of autosomal dominant inheritance was based on the report of a family with affected

members in three generations by H. Berg in 1913. The term “phakomatoses”

(neurocutaneous dysplasias) for diseases such as TSC and neurofibromatosis was

proposed in 1921 by van der Hoeve. Von Hippel-Lindau and Sturge-Weber

syndromes were later included in the list of phakomatoses. Accurate nomenclature

and the term tuberous sclerosis were proposed to replace the poorly chosen epiloia

by Critchley and Earl in 1932. (Gomez et al., 1999)

Over the course of the twentieth century, modern medical technology has

brought great progress in diagnostics. The introduction of computer tomography of

the head in 1973 and the use of ultraviolet light to detect skin lesions has increased

the diagnosis of both patients and asymptomatic family members. Magnetic

resonance imaging is proving to be an even better diagnostic tool for identifying

cortical tubers before they calcify. These methods along with careful clinical

examination allow for a certain diagnosis of TSC -- a decisive factor in the increasing

awareness that this is a relatively common disorder. (Gomez et al., 1999)

Recent events and progress in genetic research promise new possibilities for

molecular diagnostics and hope for future treatment advances. Research impulses

were given and a consensus in diagnostic criteria was established at the 1991 New

York Academy of Science TSC meeting (Roach et al., 1998). The identification of

TSC2 on chromosome 16 in 1993 (European Chromosome 16 Tuberous Sclerosis

Consortium, 1993) and of TSC1 on chromosome 9 in 1997 (van Slegtenhorst et al.,

1997) have provided for new possibilities in molecular screening.

Advances in molecular diagnostic methods are currently being explored to

simplify the as yet very lengthy process of screening all exons of both large genes.

Molecular genetic identification of the disease causing mutation is becoming

increasingly significant for patients and their families, because this information is

essential for adequate genetic counseling. Despite the many current and the hope for

future advances, the clinical diagnosis of TSC remains of central importance for the

individual patient and treatment of TSC continues to emphasize symptomatic

measures.

Introduction

3

1.2 Clinical characteristics

It is difficult to diagnose TSC unless clinicians follow strict diagnostic criteria

and thoroughly examine their patients. Table 1 shows the revised diagnostic criteria

and also demonstrates that almost any organ or system can be affected. A definite

TSC diagnosis requires that the patient exhibit either two major features or one major

and two minor features. (Roach et al., 1998)

Major Features Minor Features Facial angiofibromas or forehead plaque Multiple dental enamel pits Nontraumatic ungual or periungual fibroma

Hamartomatous renal polyps

Hypomelanotic macules (3 or more) Bone cysts Shagreen patch (connective tissue naevus)

Cerebral white matter radial migration lines

Multiple retinal nodular hamartomas Gingival fibromas Cortical tuber Nonrenal hamartoma Subependymal nodule Retinal achromatic patch Subependymal giant cell astrocytoma “Confetti” skin lesions Cardiac rhabdomyoma (single or multiple)

Multiple renal cysts

Lymphangiomyomatosis Renal angiomyolipoma

Table 1: Major and minor features for diagnosis of TSC. Almost every organ or tissue type can be affected by TSC. At least two major features or one major and two minor features are required for definite diagnosis of TSC. (Adapted from Roach et al., 1998)

Yet another varying factor in the complex of tuberous sclerosis is the time at

which the lesions emerge. Typically, cardiac rhabdomyomas arise in the perinatal

period and decline later in life. Cortical tubers also occur in early childhood, which

correlates to the severity of central nervous involvement, epilepsy and intellectual

disability. Various lesions in the skin can occur at very different ages. Hypomelanotic

macules are present at birth or appear in the first months of life and the first

angiofibromas can be seen around three years of age. In contrast, subungual

fibromas can appear much later, between 15 and 60 years of age. Figure 1

demonstrates the time course of tuberous sclerosis lesions. (Kwiatkowski and Short,

1994)

Introduction

4

Age in years

Relative level of expression

Cardiac Rhabdo-myomas

Cortical Tubers Subependymal

Nodules

Facial Lesions

birth appr

Figure 1: Time course of TSC manifestations. Relative levage at manifestation are plotted on an arbitrary scale. Somecardiac rhabdomyomas and cortical tubers) while others subungual fibromas). (Kwiatkowski and Short, 1994)

The course and prognosis of TSC is very dependen

Involvement of the central nervous system (CNS) in partic

serious condition. Although it must be emphasized that n

mentally retarded, over 50 % of children with TSC are autis

with infantile seizures have TSC (Kwiatkowski and Short, 19

symptoms in early childhood are cause for great concern a

quest for a diagnosis in these cases (Hunt and Dennis, 19

possible overinclusion of individuals with severe and profo

(Baker et al., 1998). In contrast, renal involvement may not

death occurs and angiomyolipomas are discovered upon au

about two thirds of all TSC kidneys (Zimmerhackl et al., 1994

Adding to the complexity of tuberous sclerosis are th

cases. Clinically obvious cases are easily detected and

cases are not. Particularly normally intelligent persons with

epidemiological studies (Gillberg et al., 1994). For this reas

TSC to be more common than initially expected, and its curr

is 1:6000 (Kwiatkowska et al., 1999). The asymptomatic

Renal Cysts, Angiomyolipoma

Subungual Fibromas

ox. 50 years

els of expression versus lesions arise early (e.g.

appear later in life (e.g.

t on the organs affected.

ular is a sign of a more

ot all TSC patients are

tic and 50 % of children

94). It is clear that these

nd lead to a determined

87). This may lead to a

und intellectual handicap

be noticed until an early

topsy -- they are found in

).

e asymptomatic and mild

reported, but many mild

TSC may be missed by

on, researchers assume

ent predicted prevalence

cases also explain the

Introduction

5

difficulty in genetic counseling. Parents with a child, who seems to be a sporadic TSC

case, are told that there is a 2 % recurrence risk in subsequent children (Osborne et

al., 1991, Rose et al., 1999). It is hoped that greater awareness of the clinical aspects

of TSC and the intense efforts to fully resolve its genetic mechanisms will help

clinicians more fully advise and treat their patients.

1.3 Pathogenesis

Though TSC is characterized by high variability, the typical manifestation is

the development of hamartomas -- lesions that display abnormal tissue

differentiation. These tumors are rarely malignant, meaning they do not grow

invasively and do not metastasize. Lesions are as a norm localized and the

surrounding tissue usually remains healthy. Rare malignant forms of hamartomas are

known as hamartoblastomas. Localized lesions are responsible for TSC symptoms in

the organs in which they occur, or they may occasionally be asymptomatic.

(summarized in Gomez et al., 1999)

Histological examination of hamartomas shows tissue reminiscent of neuronal

cells. It is assumed that TSC is based upon a disturbance in neuronal cell migration

and differentiation, in particular neural cells derived from the neural crest. For

example, the major feature of hypomelanotic macules in the skin represent a local

pigment disorder in melanocytes, which are derived from the neural crest cells.

(summarized in Rott et al., 1999)

The migration of neural crest cells in various tissues is one possible

explanation for the ability of TSC to occur in almost any organ. TSC lesions have

been often described in the CNS, kidney, skin, heart, lung, retina, bone, and teeth.

Skeletal muscle, peripheral nervous system and thymus involvement have as yet not

been described. Despite these tendencies, there is as yet no common trend in the

emergence of hamartomas that can predict the course of the disease. (summarized

in Rott et al., 1999)

An argument for the theory that TSC lesions arise from a single cell is

demonstrated in the example of angiomyolipomas. A progenitor cell in which one of

the TSC genes is defective can give rise to the three different cell types (blood

vessel, smooth muscle and fat) found in angiomyolipomas (Young and Povey, 1998).

The unicellular origin is confirmed by the finding of non-random X-inactivation in

tumors from female TSC patients (Green et al, 1996).

Introduction

6

Hamartomas also demonstrate loss of heterozygosity, which is evidence for

the tumor suppressor function of the two TSC genes (Carbonara et al., 1994 and

Henske et al., 1996). The loss of both alleles of a gene coding for a tumor suppressor

is required for loss of growth control (Lewin 1998). This loss of the healthy allele

means that TSC behaves recessively at the cellular level. Patients have been

studied, who were heterozygous for TSC1 or TSC2 markers, but whose

angiomyolipoma tissue were homozygous for these markers (Young and Povey,

1998). Depending on when and where the loss of the healthy allele occurs, the

clinical outcome of the disease will therefore vary.

1.4 Genetics

TSC is autosomal dominantly inherited, but more than 50 % of patients with

TSC have sporadic mutations (Verhoef et al., 1999). TSC is genetically

heterogeneous, meaning that a defect in one of two genes (TSC1 and TSC2) and

their products (hamartin and tuberin, respectively) cause the disorder. In 1993 the

TSC2 gene was identified on chromosome 16p13.3 (European Chromosome 16

Tuberous Sclerosis Consortium, 1993) and TSC1 followed in 1997 on chromosome

9q34.3 (van Slegtenhorst et al., 1997). TSC1 spans 45 kb of genomic DNA and

contains 23 exons. The TSC2 genomic DNA is 43 kb long and consists of 41 exons.

The location of TSC2 is particularly interesting. It is located directly next to the

gene affected in polycystic kidney disease, PKD-1, and kidney cysts are also one of

the possible clinical features of TSC (European Chromosome 16 Tuberous Sclerosis

Consortium, 1993). PKD-1 lies directly centromeric to TSC2 and the genes are

separated by only 60 base pairs. Another neighbor of TSC2 on chromosome 16 is

CYLD-1 at 16q12-13, which, when mutated, results in autosomal dominant

cylindromatosis (Verhoef et al., 1998). CYLD causes skin lesions clinically similar to

the typical TSC facial angiofibromas. These diseases, in some cases clinically very

similar to TSC, further complicate the diagnosis of this complex.

Recent initial evidence for the location of genes that might be involved in

autism also revealed linkage to chromosome 16p near TSC2 (International Molecular

Genetic Study of Autism Consortium, 1998). Autism and other psychiatric and

behavioral disorders including attention deficit hyperactivity disorder, aggression and

anxiety are often comorbid with TSC (Smalley, 1998). The chromosomal locations of

TSC1 and TSC2 can be seen in figure 2.

Introduction

7

chromosome 9

TSC 2

PKD 1

chromosome 16

TSC 1

p

q

CYLD 1

p

q

possible location of an autism gene

Figure 2: Chromosomal locations of TSC1 and TSC2. Chromosome 16 is the location not only for TSC2 but also PKD1, CYLD1 and a possible autism gene. TSC1 and TSC2 are emphasized by red text. Neighboring genes are outlined in grey.

Because the manifestations of TSC can be in almost any tissue or organ, the

gene products of TSC1 and TSC2 must be ubiquitous (Cheadle et al., 2000). The

products of TSC1 and TSC2 are hamartin and tuberin, respectively. Since the one

unifying characteristic of TSC is the development of tumors, it has been suggested

that hamartin and tuberin are tumor suppressors (European Chromosome 16

Tuberous Sclerosis Consortium, 1993). Further support for this theory is given in that

both TSC1 (Carbonara et al., 1994 and Henske et al., 1996) and TSC2 (Green et al.,

1994) demonstrate loss of heterozygosity. Loss of heterozygosity is one step in the

process of tumorigenesis in which one of two alleles in a tumor suppressor gene is

lost. If the remaining allele is defective, then the tumor suppressor may be defective

and uncontrolled growth can occur (Knudson, 1993).

1.5 Protein products

The protein products encoded by the TSC genes are hamartin (TSC1, 130 kDa protein, 1164 amino acids) and tuberin (TSC2, 200 kDa protein, 1807 amino acids). Tuberin has been relatively well characterized, but very little is known about hamartin. It has, however, been shown that the two proteins interact and may be part of the same cellular pathway, and both proteins have been shown to function as tumor suppressors (Plank et al., 1998).

Introduction

8

Tuberin is highly conserved among vertebrates including primate, rodent, marsupial and reptile with little homology from fish or nonvertebrate species (European Chromosome 16 Tuberous Sclerosis Consortium, 1993). Tuberin expression is ubiquitous, however, it varies as a function of development (Geist and Gutmann, 1995).

Evidence has been given that tuberin plays a role in cell cycle control. For example, the loss of TSC2 induces quiescent cells to enter the S phase, and prevents active cells from entering G0 (Soucek et al., 1997). Similar results have been produced in a Drosophila model with gigas, the homologue of TSC2. The gigas mutant cells grown were enlarged and repeated S phase without entering M phase (Ito and Rubin, 1999).

Additional evidence that tuberin might be involved in cell cycle control is its

homology to the Ras-related GTPase activating proteins (GAPs) Rap1GAPs

(Wienecke et al., 1995). The Ras family of GAPs are known tumor suppressors. They

regulate the receptors of growth factors, for example. GAPs activate the intrinsic

GTPase of the GTP-binding proteins, keeping them in the inactive GDP-bound state.

These GTPases regulate mitogenic signal transduction pathways and thus are critical

players in controlling the proliferation of many cell types (Lewin, 1998). Mutations in

the genes coding for these regulator proteins lead to the formation of tumors in

various tissues. The homology of TSC2 to this family of tumor suppressor genes

suggests that tuberin has a similar function.

It has furthermore been demonstrated that tuberin possesses specific GAP

activity. When tuberin is incubated with [γ-32P] GTP-bound recombinant Rap1a,

Rap2, Ras and Rho, significant GAP activity toward these proteins is exhibited

(Wienecke et al., 1995). Co-localization of tuberin to Rap1 in the Golgi apparatus

provides additional support for the theory that tuberin plays a role in the catalytic

activity of Rap1 (Wienecke et al., 1996).

Hamartin, the protein product of TSC1, has been less intensely studied than

tuberin, in part because TSC1 was identified four years later. No homology with other

known vertebrate proteins has been shown, but a possible match to a yeast protein

has been detected suggesting that it might participate in eukaryotic cell growth

regulation (van Slegtenhorst et al., 1997).

Hamartin also has no homology to tuberin, but it seems to interact with and is

also localized near tuberin in cytoplasmatic vesicles -- indicating that they are

partners in a common cellular pathway. Although they are shown to interact and co-

Introduction

9

localize, hamartin is also expressed in Eker rat tumor cells (ERC18M) that lack

functional tuberin. This suggests that hamartin is not dependent on functional tuberin.

In spite of this, hamartin may require a part of tuberin for its stability. This means that

the presence of a tuberin fragment might suffice for hamartin to function normally.

(summarized in Plank et al., 1998)

Figure 3 shows a speculative model for the possible roles of hamartin and

tuberin in cell cycle control (Young and Povey, 1998). It demonstrates how a tuberin-

hamartin protein complex bound to an active Rabaptin-5, Rab5 and GTP complex

could intervene with growth factor endocytosis. Evidence of the interaction of tuberin

with Rabaptin-5, GTPase activity for Rab5 and role in endocytosis have been

established (Xiao et al., 1997). This is consistent with tuberin’s homology to GAP

proteins, GAP activity, and the finding that tuberin and hamartin interact with one

another. In this model, GTP is hydrolyzed to GDP in the presence of tuberin and

hamartin, inactivating the stimulating protein complex and growth is inhibited.

Conversely, if there is a lack of functional tuberin or hamartin, the cell cycle is

constitutively activated and pathological growth occurs.

Active, GTP-bound Rab5-

rabaptin-5 complex

Inactive, GDP-bound Rab5

Stimulates endocytosis

of growth factors, etc.

Interaction of tuberin and hamartin with the protein complex leads to hydrolysis of GTP

GDP

Rab5

Rabaptin-5

Rab5

GTP

Tuberin Hamartin

Rabaptin-5

Rab5

GTP

Figure 3: Possible cellular roles of hamartin and tuberin. Tuberin may act as a GTPase activating protein for Rab5. GDP-bound Rab5 is then in the inactive form and does not stimulate cell growth. Tuberin may be required to complex with hamartin and Rabaptin-5 for optimal activity. (According to Young and Povey, 1994)

Introduction

10

Despite these advances in protein characterization, much remains unknown

about the TSC1 and TSC2 protein products. The three dimensional protein structures

have not yet been elucidated, and the exact role of these players in the complex

disease process is unclear. Evidence for the importance of various protein regions is

gradually being revealed, but until significantly more progress is accomplished, the

diagnosis and treatment of TSC remains founded in clinical and symptomatic

measures.

1.6 The clinical variability predicament

Although the exact role of hamartin and tuberin is not known in full detail, it is

clear that either a mutation in TSC1 or TSC2 causes the disease tuberous sclerosis.

This does not, however, explain the broad variation in clinical manifestations of the

disease. Furthermore, most mutational studies find a mutation in TSC1 or TSC2 in

only about two thirds of patients with definite TSC disease (Au et al., 1998, Jones et

al., 1999b, Zhang et al., 1999, Langkau et al., 2002 and others).

Possible explanations for this apparent discrepancy include the existence of a

third allele, the high rate of mosaicism in TSC and the Knudson Two-Hit-Hypothesis.

The existence of a third allele seems at present very improbable, as intense studies

over several years have not yielded a third site for TSC and half of TSC families

show linkage to TSC1 and half to TSC2 (Cheadle et al., 2000). Mosaicism, the

condition in which only a portion of somatic and germ line cells contain a mutation

(Lewin, 1998), is very feasible. Cases of mosaicism in TSC have been described,

and this can greatly complicate clinical and genetic analysis (Kwiatowska et al.,

1999). Knudson’s Two-Hit-Hypothesis describes the process of tumorigenesis and

the possible role of tumor suppressor genes (Knudson, 1971) and also provides a

good explanation for the high clinical variability of TSC (see section 4.3.3, figure 11).

The two hit hypothesis refers to the loss of both alleles of a gene in order for

the sick phenotype to occur. The loss of the first allele can occur very early or be

inherited. Depending on when and where the second loss (second hit) occurs, the

phenotype of the TSC patient can be very variable. If the loss occurs early, for

example during cellular differentiation, then a larger number of cells or even tissues

can be affected. The loss in a later, more differentiated cell, would then affect only

this tissue. Milder forms of TSC can therefore be explained by later occurrence of

second hits.

Introduction

11

Mild forms of TSC may also bias statistics concerning penetrance and

reproductive fitness. Cases of minimal expression and non-penetrance have been

reported, in particular upon meticulous re-examination of seemingly unaffected

parents of a TSC diagnosed child (Webb and Osbourne, 1991). Persons with very

mild forms of TSC have no reduction in reproductive fitness and it has been

postulated that these cases usually involve the TSC1 locus (Langkau et al., 2002).

This might explain why only about 10 % of sporadic cases have been found to

involve TSC1 (Janssen et al., 1994 and Povey et al., 1994), while half of the families

with TSC show mutations in TSC1 and the other half in TSC2 (Cheadle et al., 2000).

Initial reports that TSC lesions may be able to metastasize have recently been

published. Sporadically occurring lymphangioleiomyomatosis of lung was recently

brought into connection with the occurrence of renal angiomyolipomas, proposing a

model involving the spread of smooth muscle cells from the angiomyolipoma to the

lung (Carsillo et al., 2000). Also, a variant epithelial form of angiomyolipoma has

been reported, which demonstrated unusually malignant qualities including

metastases (Martignoni et al., 2000). Although as yet far from proven, a model

involving possible metastasizing of TSC lesions opens new possibilities for explaining

varying localizations of TSC manifestations.

1.7 The Freiburg-Heidelberg project

In the Freiburg-Heidelberg project, the University of Freiburg Children’s

Hospital and the Institute for Human Genetics in Heidelberg cooperated to screen

TSC patients and their families for mutations in the TSC1 and TSC2 genes. An

uncommonly large pool of TSC patient and relative blood samples (over 300) had

been accumulated since 1994. In addition to the advantage of a very large sample

size, the patient information and blood samples were collected from not only several

regions but also several countries thus counteracting any possible regional bias.

Complete screening of all exons of both genes in two centers was a lengthy

process that was only recently completed (Langkau et al., 2002). The detection of

large mutations and the screening of TSC1 and exons 1-19 of TSC2 for small

mutations were pursued in Heidelberg. TSC2 exons 20-41 were examined in

Freiburg. The number of mutations, type, location and size was compared to

previously published studies.

Introduction

12

Introduction

This dissertation deals with the results of the TSC screening project in exons

30, 34, 36 and 38 - 40 of TSC2, including the region of GAP homology of TSC2

(exons 34 to 38). An additional population of healthy probands was subjected to

mutational analysis in exon 40 of TSC2 as a control group to study the variability of

TSC2 in an unaffected population.

1.8 Objectives and research questions

What mutations do the TSC patients demonstrate in the final exons of the

TSC2 gene (the regions of GAP homology)?

The goal of this study was to screen TSC patient DNA for variations and to

analyze the sequence for mutational diagnosis as part of the Freiburg-Heidelberg

Project. It was expected that a relatively large amount of mutations would be found in

the final (5') exons of the TSC2 gene, as a number of these were within the region of

known GAP homology (Wienecke et al., 1995). This region is presumably important

for the function of the protein product (Wienecke et al., 1996), which is assumed to

be defective in TSC patients. The exons examined here were 30, 34, 36 and 38 - 40

of TSC2. These exons were amplified via PCR (polymerase chain reaction),

screened via SSCP (single strand conformational polymorphism), sequenced by

external commission and analyzed by hand.

How much variation is in the healthy TSC2 gene, and how do the types of

variations found compare with the patient group?

An additional group of healthy patients was chosen for mutational analysis in

one exon as the control population. Exon 40 was chosen as the target of this initial

screening. It was an exon in which our group had discovered several variations in the

patient group, and it was the exon chosen for quality control analysis. This means

that all samples of exon 40 were sequenced, allowing for a more complete screening

for better comparison between the patient and healthy populations.

The findings of the patient group were compared with those of previous

studies and it was speculated, that a genotype-phenotype correlation would be

found. These findings were also weighed against those of the healthy population so

as to bring the known mutations in the patient group into proper context. No genetic

analysis of healthy probands has as yet been published, and it was considered

important to examine a control group for this study.

13

2 Materials and Methods

2.1 Materials

2.1.1 Apparatus and equipment

Analytical balance Sartorius

Automatic shaker Desaga, Heidelberg

Centrifuge Sorvall - Du Pont

Electrophoresis apparatus Amersham Pharmacia

Ice machine Ziegra

Incubators Heraeus Hera Cell, Heraeus Function Line

Microscope Zeiss

Microwave Siemens

Mini-centrifuge Eppendorf

Multiphor (SSCP) Amersham Pharmacia

Phast system, gels and buffer strips Amersham Pharmacia

Photography of DNA gels Mitsubishi P96, Biocapt software

Speedvac Savant

Thermocycler (PCR) Techne Progene

Vortexer Heidolph

Water bath GFL

2.1.2 Buffers and solutions

20X TBE (DNA-gels) 0.5 M Tris 0.5 M boric acid 0.1 M EDTA adjust volume to 500 ml with ddH20 autoclave to dissolve

DNA loading buffer 50 % glycerol 1 mM EDTA 0.25 % bromphenol blue adjust volume to 20 ml with ddH20

SSCP developing buffer 250 ml 2.5% Na2CO3 100 µl formaldehyde

SSCP loading buffer 95 % formamid 5 % Na-EDTA

14

SSCP staining buffer 200 ml 0.1 % AgNO3 50 µl formaldehyde

2.1.3 Chemicals, enzymes, kits and size markers

Chemicals Difco, Eurobio, Merck, Riedel-de-Häen

DNA extraction kit Qiagen

DNA ladder (size standard marker) Amersham Pharmacia

DNA purification kit for PCR products Qiagen

Taq DNA-Polymerase + 10X Buffer Promega

2.1.4 Databases and software

HUSAR (search for primers and exons) http://genius.embnet.dkfz-heidelberg.de

Genome database (search for primers) http://www.gdb.org

Human Gene Mutation Database, Cardiff-Rotterdam database (search for published mutations)

http://www.uwcm.ac.uk

Aladdin Expander (to expand compacted sequence data)

Aladdin Systems Inc., Watsonville, CA, USA http://www.aladdinsys.com

Chromas (to view sequences) Technelysium Pty Ltd, Helensvale, Australia http://www.technelysium.com.au/chromas.html

2.1.5 Media and cell cultures

Culture bottles, centrifuge vials Falcon

Cryovials Nalgene

DMSO Sigma

EBV culture supernatant Kindly provided by Prof. Hildebrandt, ephrolabor II, Univ. Kinderklinik, Freiburg N

FCS (fetal calf serum) Biochrom

Fecoll Plaque Amersham Pharmacia

RPMI (medium without FCS, with 10 % FCS, and with 20 % FCS)

Seromed

Sterile filters Millipore

Materials and Methods

15

2.1.6 Oligonucleotides and DNA-sequencing

HPLC purified oligonucleotides Gene Scan (Big Biotech)

Primer sources • Literature (Northrup et al., genome data base)• Sequence search (HUSAR database) and

Gerd Wiegele

Sequencing Seqlab (Göttingen) and CORE-Facility (Freiburg) 2.2 Methods 2.2.1 Probands

The source of DNA for this study was from two groups of probands; one

population of TSC patients and their families, and one healthy population.

2.2.1.1 TSC patients and families

Blood samples from TSC patients have been collected in Freiburg and

Heidelberg since 1994. Patients were German, Austrian or from the Czech Republic.

It was requested that each attending physician send a completed TSC clinical

evaluation form and also blood samples from all available relatives. Of the more than

300 patients in our pool, the first 86 patients and their families were selected for this

study whose physician had completed a clinical evaluation and a definitive TSC

diagnosis was made. 70 of these 86 patients were completely screened in both

Freiburg and Heidelberg.

2.2.1.2 Healthy population

Healthy genomic DNA samples were provided for analysis by Dr. Klaus

Deichmann from the department of Zentrale Klinische Forschung in Freiburg. DNA

was isolated from the peripheral blood of various healthy volunteers (mostly

students), who represented a relatively heterogeneous population, as they were not

all from the same geographical region or family (K. Deichmann, personal

communication). 25 specimens from these healthy DNA samples were randomly

selected for the control group in this study.

Materials and Methods

16

2.2.2 Cell culture and transformation from peripheral blood samples

Peripheral blood mononuclear cells are transformed with EBV in order to

increase the rate of cell divisions and therefore facilitate cultivation. A greater cell

count is achieved and thereby a greater amount of DNA can be harvested.

Procedure:

4-6 ml of RPMI was added to 2-3 ml of peripheral blood (collected in Na-

heparin or NH4-heparin vials) and mixed well. The blood-medium mixture was gently

layered over 9 ml Ficoll-Plaque in a 50 ml Falcon tube and centrifuged at 1500-1800

rpm for 30 min. (without break). Approximately 6 ml of supernatant were then slowly

aspirated and discarded. The white leukocyte ring was then cautiously aspirated and

saved in a 50 ml Falcon tube. 10 ml RPMI with 10 % FCS were added to the

leukocytes and recentrifuged at 1500-1800 rpm for 10 min. The supernatant was

discarded and the pellet was resuspended in RPMI with 10 % FCS. This was again

centrifuged at 1500-1800 rpm for 10 min. and the supernatant was again discarded.

5 ml supernatant of an L922-EBV culture was harvested and sterile filtered. 2

ml of this and 2 ml of RPMI with 20 % FCS were added to the cell pellet. The pellet

was resuspended and incubated in culture flasks. This portion of the procedure was

always prepared to yield a double amount.

After 24-48 h of incubation (at 37°C, 7.5 % CO2), 1 ml of medium was carefully

aspirated and 3 ml RPMI were added to the culture. The medium was renewed once

every week as follows: 2 ml of cell suspension were removed and 3 ml of RPMI with

20 % FCS were added. After the indicating buffer in the medium changed color from

red to yellow, RPMI with 10 % FCS was used and renewed as needed.

2.2.3 Freezing of cells Procedure for cultured, transformed leukocytes:

The medium was carefully aspirated from two full culture flasks and discarded.

The cells were transferred to a 50 ml Falcon tube, which was subsequently filled with

RPMI (10 % FCS) and centrifuged at 1200 rpm for 10 min. The supernatant was

discarded and 7 ml of cold FCS with 10 % sterile DMSO were added to the pellet and

mixed well. This cell suspension was distributed into 4 cryovials, placed in a cooled

cryo-box and frozen at -70 °C for 24 h. The cryovials were then transferred to a

liquid nitrogen tank for storage. Materials and Methods

17

Procedure for whole blood:

1.8 ml of each sample of peripheral blood collected in either Na-heparin or

NH4-heparin was set aside before transformation with EBV. This was added to a

cryovial containing 200 µl of sterile DMSO and mixed well. The cryovials were then

placed in a cooled cryo-box and frozen at -70 °C. After 24 h they were transferred to

a liquid nitrogen tank for storage.

2.2.4 DNA extraction

Leukocyte suspensions (or alternatively whole blood in DMSO) frozen in liquid

nitrogen were defrosted at room temperature and DNA was extracted according to

the Qiagen Genomic Tip Protocol:

Leukocyte suspensions were poured into a 15 ml Falcon tube and centrifuged

at 4500 rpm for 5-10 min at 20 °C. The supernatant was discarded. The cells were

resuspended with 2 ml ice cold PBS and recentrifuged. This step was repeated one

more time, then the leukocytes were resuspended with 2 ml of C1 buffer and 6 ml of

cold ddH20 and incubated on ice for 10 min.

Cells from whole blood samples in DMSO were poured into 15 ml Falcon

tubes, directly centrifuged, resuspended with 2 ml of C1buffer and 6 ml of cold ddH20

and then incubated on ice for 10 min.

Continued procedure for either leukocyte suspensions or cells from whole blood samples in DMSO:

After lysis, the samples were centrifuged at 3500 rpm for 15 min. at 4 °C, and

the supernatant was discarded. The pellet was resuspended with 1 ml cold C1 buffer

and 3 ml ddH2O, and then recentrifuged. This time the pellet was resuspended with 5

ml of G2 buffer, and 95 µl of QIAGEN protease or Proteinase K was added. This

mixture was incubated in a water bath at 50 °C over night.

On the next day, the QIAGEN genomic tip 100/G columns were equilibrated

with 4 ml of QBT buffer. Each incubated sample was vortexed to resuspend and was

immediately applied to a column. The column was then washed twice with 7.5 ml of

QC buffer. DNA was eluted into 15 ml Falcon tubes with 5 ml of QF buffer that was

warmed to 50 °C in the water bath. The DNA was then precipitated with 3.5 ml

isopropanol, vortexted to mix well, and centrifuged at 8000 rpm and 4 °C for 30-35

min. The supernatant was discarded and the pellet was washed with 2 ml of ice cold Materials and Methods

18

70 % ethanol, vortexed to resuspend and centrifuged again at 8000 rpm and 4 °C for

15 min. The supernatant was discarded and the tubes were allowed to dry for

approx. 15 min. The extracted DNA was resuspended in 500 µl ddH2O, allowed to

dissolve over night at 55 °C and stored at -20 °C in Eppendorf tubes.

2.2.5 PCR

PCR (polymerase chain reaction) was employed to selectively amplify each of

the exons in the TSC-2 gene. The amount of amplified product is so large that the

amount of template DNA is in comparison infinitely small and consequently

negligible. Nucleotides and other chemicals in each sample are also very small and

can be considered insignificant. Therefore each of the PCR products for each exon

could be analyzed via the SSCP method as described below.

First, sense and antisense primers were found for each exon (see table 2

below). These were chosen to be in a nearby intron sequence bordering each exon.

For best results the primers must have similar annealing temperatures (nearly the

same guanosine-cytidine nucleoside content) and should not be complementary to

each other to prevent the formation of primer-dimers. A standard PCR procedure

follows. In this basic protocol, the annealing temperature was varied according to the

annealing temperature of the primers; the annealing time, elongation time and

number of cycles were also varied to obtain an optimal product.

Procedure: reaction mix: 5 µl 10 X reaction buffer (contains 25 mM MgCl2)

5 % DMSO 200 µM dNTP mix 0.3 µM sense primer 0.3 µM antisense primer 1 Unit Taq polymerase 0.5 µg template DNA adjust to 50 µl with ddH2O

All ingredients except the template DNA were added to PCR vials and kept on

ice. The template DNA was added to the prepared vials as a final step -- the vials

were then briefly centrifuged and immediately placed in the programmed

thermocycler. As a negative control one sample was prepared with 5 µl ddH2O

instead of template DNA.

Materials and Methods

19

Thermocycler program: DNA melting (1 cycle): 5 min. at 95 °C

Amplification (30 cycles): • 30 sec. melting at 95 °C

• 30 sec. annealing temperatures depending on primer optimum

• 1 min. elongation at 72 °C

Elongation (1 cycle): 5 min. at 72 °C

TSC-2 Exon

source sense primer (5' 3')

antisense primer (5' 3')

amplification conditions

product length

30 Northrup et al. aga tgg gta agg gga ggt ac

ttg gca cgg gct ctg ctc c

62 °C 296 bp

34 Northrup et al. tgg cca ggc cct cac ctg ggt

ctg gag ctg tgt ggc tcg gg

64 °C, 45 sec. elongation

170 bp

36 Northrup et al. ttg tct gcc tca ggg atc ag

cat cca gtc ctg cta ccc c

60 °C 379 bp

38 G. Wiegele (sense), Northrup et al. (antisense)

cag gtg tct agc agt gca ac

gct gca gtg tgg cgc caa g

64 °C, 45 sec. elongation

205 bp

39 Northrup et al. gct gca gtg tgg cgc caa g

tgc cac ctg ata gtg agc tc

61 °C 212 bp

40 G. Wiegele agg cgt gag ctg gtg gga cag

tta ctg ccc aag ccg cct ctg

62 °C 281 bp

Table 2: Primers used. The primers used to amplify each exon studied, their source, sense and antisense sequences, amplification conditions used and the expected product length are listed above.

2.2.6 PCR product analysis

Aliquots of PCR products were resolved on 2 % agarose gels to verify their

amount and purity.

Procedure:

1 g of agarose in 50 ml TBE was boiled, cooled carefully, poured into the gel

form with well comb and allowed to polymerize for about one hour. 5 µl of PCR

product was mixed with 1 µl 5X loading buffer and pipetted into each of the gel

sample wells. The DNA ladder from Amersham was used as a size reference marker

and positive control; 1 µl ladder + 1µl 5X loading buffer + 3 µl TBE were mixed and

Materials and Methods

20

pipetted into one well. As a negative control, the negative PCR control (ddH2O in

stead of template DNA) was also applied to the gel; 5 µl product mixed with 1 µl 5X

loading buffer and pipetted into one well. 100 V were applied to the gel for 1 h.

The gel was then developed in a shaking bath of ethidium bromide for approx.

30 min. and a photograph was taken in UV-light to visualize bands of DNA.

Impure products were discarded and the PCR was repeated at a higher

annealing temperature and/or with fewer cycles. This step was repeated until single

bands were visualized. If no product was achieved, the annealing temperature was

reduced and/or more cycles were run. If pure product could not be achieved, a

different set of primers was used to try to amplify a pure product.

2.2.7 SSCP

SSCP (single strand conformational polymorphism) is a technique used to

detect small mutations, and is based on the migration of single stranded DNA

fragments in native gradient polyacrylamide gel electrophoresis. Individual variances

in a DNA sequence can be detected due to the varying banding patterns, which are

caused by the various DNA segment lengths and the formation of secondary

structures in the single stranded state. Varying banding patterns can be detected for

sequence variations as small as single base pair mutations. (summarized in Orita et

al., 1989)

Procedure:

The Phast system, ready-to-use gradient polyacrylamide gels and native buffer

strips from Amersham Pharmacia were used for gel electrophoresis.

Ready-to-use gels with buffer strips were placed in the Phast apparatus according

to the manufacturer’s instructions and pre run at 250 V and 100 Vh to equilibrate.

Meanwhile, 3 µl of PCR product were mixed with 3 µl of SSCP loading buffer and

denatured at 95 °C for 5 minutes in a thermocycler and immediately cooled on ice.

After the pre run, 4.5 µl of this denatured DNA sample was loaded onto the gel. The

electrophoresis was then run at 250 V and 300-600 Vh depending on the size and

nature of the PCR product. The length of the run (in Vh) and the temperature (5-15

°C) was varied for each exon, so that the bands could be visualized near the center

of the gradient gel for optimal separation. Conditions for each exon can be seen in

Materials and Methods

21

table 3. Exon 40 SSCPs were performed on the healthy population, and those for the

patient group were performed by Gerd Wiegele.

Exon Temperature Run length30 5 °C 380 Vh 34 5 °C 250 Vh 36 5 °C 580 Vh 38 5 °C 380 Vh 39 15 °C 320 Vh 40 5 °C 500 Vh

Table 3: SSCP conditions for each exon.

Following electrophoresis, the gels were stained in the Phast staining apparatus

in several steps:

1. 20 % TCA (20 °C, 5 min.)

2. 0,3 % glutaraldehyd (50 °C, 5 min.)

3. ddH2O (50 °C, 2 min.), twice

4. 0,1% AgNO3 with 37% formaldehyde (40 °C, 8 min.)

5. ddH2O (30 °C, 30 s), twice

6. 2,5 % Na2CO3 with 37% formaldehyde (30 °C, 30 s, then 30 °C, 4 min.)

7. 5 % acetic acid (50 °C, 2 min.)

8. 10 % acetic acid (50 °C, 3 min.)

9. 10 % glycerin (50 °C, 3 min.)

The gels were then dried at room temperature and stored in glass slide covers at 4

°C in the dark.

2.2.8 Purification of the PCR product

PCR products that showed variations in banding patterns were purified before

sequencing with the Qiaquick DNA purification kit for PCR products. The procedure

for microcentrifuge use was followed, and the DNA was eluted in 50 µl ddH2O.

2.2.9 Sequencing PCR products were sequenced in Göttingen at Seqlab or at the CORE facility

in Freiburg. The procedure for Seqlab was established in our research group and

was very simple. The use of the CORE facility was later initiated and was somewhat

more labor intensive. The procedures for each facility are described below.

Materials and Methods

22

Materials and Methods

Procedure for Seqlab, Göttingen:

Purified PCR products, sense and antisense primers were sent to Seqlab in

Göttingen for sequencing.

Procedure for CORE facility: PCR reaction mix: 5 µl of purified PCR product 1 µl of only either sense or antisense Primer 4 µl of Big Dye Mix (provided by CORE) Final volume adjusted to 20 µl Thermocycler program (25 cycles): 96 °C, 30 s 50 °C, 15 s 60 °C, 60 s

The reaction products were immediately cooled on ice after thermocycling.

Each product was placed in an autoclaved 1.5 ml Eppendorf tube for precipitation as

follows:

80 µl of 75 % isopropanol was added to 20 µl of product, vortexed to mix well

and let stand at room temperature for 15 min. The product was then centrifuged in a

microcentrifuge for 20 min at maximum speed. The supernatant was carefully

discarded. 250 µl of 75 % isopropanol was added to the pellet. This was again

vortexed to mix and then centrifuged at maximum speed for 5 min. The supernatant

was again carefully discarded and the pellet was dried for 10 to 15 min. in a vacuum

centrifuge. This product was stored at - 20 °C until it was sent to CORE.

2.2.10 Sequence analysis

Sequencing results were received electronically as compacted .sea files.

These files were expanded with the Aladdin expander program and visualized using

the Chromas program. The sequences were printed and interpreted by hand -- each

exon and splice region was read and compared to the wild-type sequence.

Sequences that were difficult to interpret were resequenced (depending on

occasional difficulty with a freshly produced and purified PCR product) or

reinvestigated in the Humangenetisches Institut in Heidelberg.

23

3 Results 3.1 Patient group 3.1.1 Amplification of exons

Exons 30, 34, 36, 38, 39, and 40 were amplified from genomic DNA for each

of the patients via PCR, according to methods described in section 2. The PCR

products were then separated on an agarose gel to determine their purity. The gel

depicted in figure 4 is an illustration of typical products -- exon 30, in this example. It

shows single, sharp bands at the size expected for the PCR product. The positive

control was the DNA size standard (lane one) and no bands appear in the lane of the

negative control (final lane). For examples of agarose gel results for each exon, see

figure 12 in the appendix. Some gels were photographed as negative images for

better visualization of faint bands. For the conditions used in each exon, see table 2

in the methods section 2.2.5.

300 bp

1 2 3 4 5 6 7 8 Figure 4: Sample PCR result, exon 30. Lane 1: DNA size standard (positive control). Lanes 2-7: PCR products (expected size = 296 bp, see table 2 in section 2.2.5). Lane 8: negative control.

3.1.2 Mutational screening

The exons for each of the patients were then submitted to the SSCP method

for mutational screening as described in section 2.2.7. Variations in the SSCP

banding pattern could signify a deviation in DNA sequence for these samples. If a

sample differed from the others in SSCP, it was sent out for sequence analysis.

Figure 5 shows one SSCP gel result for exon 34. Here a relatively simple pattern of

two main bands with the upper faintly doubled, and a few slighter bands visible at

higher concentrations is the rule. Differing samples are easily detected in this case.

For SSCP gel examples with normal and variant banding patterns for each exon, see

figure 14 in the appendix. For SSCP conditions in each exon, see table 3 in section

2.2.7.

24

143138128124122121117 111

extra bands

Figure 5: Sample SSCP result, exon 34. Conditions: 5 °C and 250 Vh run length. Patients 111 and 117 show variations in banding pattern.

Selecting samples for sequencing was often simple, when the SSCP banding

patterns showed obvious deviation. This can easily be seen in figure 5 in exon 34. In

some cases, however, complex banding patterns were difficult to screen. Exon 36,

for example, required careful scrutiny to see a missing band in patient 154 (figure 14,

appendix). At times a number of patients were manually selected for sequencing,

when a large group of seemingly identical patterns were revealed.

3.1.3 Sequencing results 3.1.3.1 Mutations

Both the sense and antisense sequences for each exon sample sent in were

interpreted by hand. That is, each exon and splice region was read and carefully

compared to the wild type sequences. Shown here are two sequence examples in

figure 6. One is of patient 111 in exon 34, whose SSCP result can be seen in figure 5

(section 3.1.2). The deletion of 4 nucleotides (here ACAA at position 4562-4565)

results in a shift in the reading frame (figure 6A and B). Figure 6B clearly shows a

double sequence beginning with nucleotide position 4562. The other example is of a

small mutation in exon 38 in patient 109, in which thymine is substituted for cytosine

at position 5042 (figure 6C and D). A double peak at this position is displayed in

figure 6D. The protein product in this case has an amino acid point mutation in which

proline is substituted for leucine at amino acid 1675.

Results

25

A. Exon 34:

B. Patient 111: 4562-4565 del ACAA C. Exon 38: D. Patient 109: C5042T

4562

5042

Nucleotides 5008-5086: ggccagt tcaactttgt ccacgtgatc gtcaccccgc tggactacga gtgcaacctg gtgtccctgc agtgcaggaa ag

Nucleotides 4512-4587: tttcgtgtt cctgcagctc taccattccc ccttctttgg cgacgagtca aacaagccaa tcctgctgcc caatgag

Figure 6: Sequence result examples for two patients. A: A portion of the wild-type text sequence of exon 34, highlighting the deletion found in patient 111 in red. B: Portion of the sequencing result of patient 111, exon 34. A double sequence beginning with nucleotide 4562 indicates a frame shift mutation. C: Part of the text sequence of exon 38, highlighting the substitution found in patient 109 in red. D: Portion of the sequencing result of patient 109, exon 38. A double peak at nucleotide position 5042 represents the change in nucleotide sequence from C to T.

Complete results for all patients in the exons studied here are also

summarized in table 4 (in this section). The wild-type and variant sequences for each

exon are shown in figures 15 and 16 in the appendix. For all mutational findings in

the Freiburg-Heidelberg project, see Langkau, et al. (2002).

In most cases the variations found correlate clearly with the SSCP results. At

times, however, the sequence did not match the expectations based on what was

seen in SSCP. SSCPs of Exon 30, for example, displayed two types of banding

patterns (figure 14, appendix). Several samples of these patterns were chosen for

sequence analysis, but no variations in sequence were found. Results

26

Exon Pat. Nr. Pat. name DNA variation Protein variation 34 117 M.S. 4521 del C frame shift 34 139 I.J. 4545-4547 del CTT in-frame deletion 34 111 M.R. 4562-4565 del ACAA frame shift 36 154 N.P. C4695G no change 38 109 T.T. C5042T P1675L 38 115 Z.C. C5042T P1675L 39 155 A.T.H. -2bp: A C (splice) possible exon skipping 40 46 L.K. C5245T R1743W 40 99 V.S. 5188-89 del CA frame shift 40 80 G.J. 5256-5273 del 18bp in-frame deletion 40 113 T.H.* 5256-5273 del 18bp in-frame deletion 40 114 M.H.* 5256-5273 del 18bp in-frame deletion

Table 4: Summary of mutational findings. Mutations found in exons 30, 34, 36, and 38-40 of TSC2 are listed here according to nucleotide position and the type of change is documented in the column protein variation. The mutation in patients marked * is in a pair of monozygotic twins and has been previously described in our research group by Nicola Martin (Martin et al., in press).

3.1.3.2 Polymorphisms

Numerous smaller sequence variations were found in several patients, and

these were defined as polymorphisms. A large number was found in Exon 40 in

particular. This is because all samples were sequenced entirely in exon 40 in addition

to the SSCP screening method as a quality control measure. The polymorphisms

found here are summarized in table 5. For a complete list of those found in the

Freiburg-Heidelberg project, see Langkau et al. (2002).

Table 5: Summary of polymorphisms found. Polymorphisms found are listed here according to frequency.

Exon DNA variation Protein variation Frequency40 T5220C no change 13/70 40 -10bp: A C (splice) possibly exon skipping 4/70 40 both T5220C and -10bp: A C (splice) none, poss. exon skip 9/70 40 T5180G M1721R 7/70

3.1.4 Exon 40 and quality control

Exon 40 was entirely sequenced in addition to SSCP analysis to compare the

results of our screening method with actual sequence results. Interestingly, not all

variations found in the sequences were seen in SSCP gels. Quite a large number (26

of 36) were missed. Most of these, however, were smaller changes or

polymorphisms (15 of 26). These results are summarized in table 6.

Results

27

Patient number

Variant SSCP

Variant sequence

Both Type of Variation Protein change

6 r T5180G M1721R 9 r T5220C silent 46 r -10 bp A-->C, T5220C splice, silent 18 r T5180G M1721R 23 r 28 r r yes T5220C silent 39 r T5220C silent 46 r r yes C5245T R1730W 50 r T5220C silent 61 r T5180G M1721R 64 r 68 r r yes T5220C silent 72 r r yes T5220C silent 76 r T5180G M1721R 80 r 5256-5273 del 18bp in-frame deletion 83 r -10 bp A-->C splice 93 r 97 r T5220C silent 99 r 5188-89 del CA frame shift 103 r r yes -10 bp A-->C splice 105 r T5220C silent 106 r r yes -10 bp A-->C splice 109 r -10 bp A-->C, T5220C splice, silent 113 r r yes 5256-5273 del 18bp in-frame deletion 114 r r yes 5256-5273 del 18bp in-frame deletion 115 r T5220C silent 121 r T5220C silent 132 r r yes -10 bp A-->C, T5220C splice, silent 134 r r yes -10 bp A-->C, T5220C splice, silent 135 r -10 bp A-->C, T5220C splice, silent 145 r -10 bp A-->C, T5220C splice, silent 146 r r yes -10 bp A-->C splice 149 r 151 r -10 bp A-->C, T5220C splice, silent 154 r -10 bp A-->C, T5220C splice, silent 155 r T5180G M1721R 156 r r yes -10 bp A-->C, T5220C splice, silent 158 r T5180G M1721R 179 r T5180G, T5220C M1721R, silent 182 r T5220C silent Totals:

16

36

12

Table 6: Comparison of SSCP and sequence analysis in exon 40. The variations found by SSCP and by sequence analysis are listed in comparison in this table. 4 patients had variant SSCPs and no sequence deviations, 24 Patients had variant sequences, but normal SSCP results. 15 of these 24 missed variations are silent or splice mutations. Results

28

Mutational screening of exon 40 in the patient group was performed by Gerd

Wiegele and the column of SSCP results shown in table 6 are based upon his

analysis of the samples.

3.2 Control group of healthy probands

A group of 25 healthy probands was screened for mutations in exon 40 of

TSC2. Since this exon appeared to have the highest density of mutations it was

chosen as the best candidate for comparison between healthy and affected

individuals. (See figure 9, section 3.3.) Another advantage was that the patient group

was entirely sequenced in exon 40, so that complete results were available for

comparison. (See section 3.1.4.) The control group was therefore also entirely

sequenced in exon 40 in addition to SSCP analysis.

3.2.1 Exon amplification

Exon 40 was amplified from genomic DNA for each of the healthy probands

via PCR according to methods described in section 2.2.5. These DNA samples were

kindly provided by Dr. Klaus Deichmann. The PCR products were then separated on

an agarose gel to determine their purity. These results are analogous to those for the

patient population in section 3.1.1. For an example of PCR results for the healthy

population, see figure 13, in the appendix.

3.2.2 Mutational screening

Once pure PCR products were obtained, these were separated on an SSCP

gel to determine, if there were variations in banding patterns that could signify a

variation in DNA sequence for these samples. Figure 7 shows an example of an

SSCP gel result from the healthy population. A typical, simple banding pattern was

revealed for exon 40 (analogous to the results demonstrated by Gerd Wiegele in the

patient population), with two lighter upper bands and a triplet of stronger bands

below. Variations are marked in the figure, and the sequence results verify these

findings (figure 8).

Results

29

151411 13129 10 7

extra bands

Figure 7: SSCP exon 40, healthy population. Conditions: 5 °C and 500 Vh run length. Probands 11 and 14 demonstrate varying banding patterns.

3.2.3 Sequencing results and numerical comparison to patient group

An example of a variant sequence in the healthy population is displayed in

figure 8. The wild type sequence of exon 40 can be read in figure 15 in the appendix.

Healthy proband number 10 from figure 7, for example, showed no variation in exon

40. However, the equally healthy probands 11 and 14 both show the variation

T5220C.

t

Fishpo

in

va

us

va

star

gure 8: A portion of the exon 40 sequence in proband number 11. ows a variation from the wild type, represented here as a doublesition 5220.

A surprisingly high rate of variation was found in the group of he

general, and this is demonstrated in table 7. 8 out of 25 or 32% d

rying sequence. 7 of the 8 varying sequences showed two sepa

ually both the splice site -10 bp A C and the silent T5220C. Th

riation is similar in the TSC-patient population.

T5220C

This sequence peak: T C at

althy probands

emonstrated a

rate variations,

is frequency of

Results

30

Exon Mutation TSC patients Healthy probands 40 -10 bp A C 19 % (13/70) 28 % (7/25) 40 T5220C 37 % (26/70) 32 % (8/25)

Table 7: Frequent mutations in both populations. Two mutations were found with similar frequency in the patient and healthy populations.

These variations, however, are both polymorphisms. The change T5220C has

no consequence at the amino acid level, and the splice site -10bp A C may cause

exon skipping. Interestingly, there were a few, more significant mutations that were

found only in TSC patients. One of these changes is also a frequent polymorphism --

the amino acid change M1721R. These results are summarized in table 8.

Exon Nucleotide change Amino acid change TSC Patients 40 T5180G M1721R 10 % (7/70) 40 5256-5273 del 18 bp in-frame deletion 4 % (3/70) 40 5188-5189 del CA frame shift unique 40 T5245C R1743W unique

Table 8: Mutations found only in TSC patients. Four mutations were found in only TSC patients. The first two changes listed here were relatively frequent mutations.

3.3 The Freiburg-Heidelberg project

70 patients (69 unrelated) were screened in all 41 exons of TSC2 and all 23

exons of TSC1 as part of the Freiburg-Heidelberg project. This intense screening

effort resulted in the detection of 29 definitive sporadic mutations. Not included in

these 29 are 4 familial cases, 9 unclassified sequence variations and 16

polymorphisms. Altogether 58 variations from the wild-type sequence were identified

(83 %). A few large defects were among the 29 definitive mutations, including one

contiguous gene deletion syndrome (affecting both TSC2 and PKD1), and a 7.5 kb

intragenic deletion in TSC2. These results concentrate on TSC2 for several reasons.

In particular, because the six exons studied in this thesis are in the TSC2 gene.

Moreover, mutations in TSC2 have been found to be more frequent (Jones et al.,

1999 and Zhang et al., 1999), and in our study more mutations were found in TSC2

than in TSC1 (TSC2:TSC1 approximately 3:1) (Langkau et al., 2002). The number,

location, and type of variations found in the TSC2 gene are displayed in figure 9.

Results

31

Results

Legend

Nr

Splice

Missense

Insertion

Deletion

Nonsense

Unclassified

Polymorphism

35 40 30 25 20 15 10

2

27

10

2

1 4 1 4 10

1

2

5

Chromosome 16

qp

13.3 13.2 13.1 12 11.2 11.1 11.1 11.2 12.1 13 21 22 23 24 12.2

TSC2

Figure 9: Smaller Mutations in TSC2. The number and type of each variation is plotted to its location in the TSC2 gene in the figure above. The exon scale is arbitrary -- exons containing a mutation are spaced further apart for better visualisation. Exon number is indicated with an arrow. Mutation type is designated by a symbol and polymorphisms are denoted by the number of patients that showed the variation (see legend box in figure).

32

4 Discussion 4.1 Patient group 4.1.1 PCR products

The screening of TSC2 for mutations ensued exon by exon. Each exon was

individually examined for variations that could be caused by mutations in the

sequence. The basis of this method required the amplification of each exon via PCR

as described in sections 2 and 3 above. This was in most cases the rate-limiting step

in the lengthy process of obtaining a sequence. Some PCR products were more

difficult to obtain than others. In order to achieve a pure product, a set of primers was

chosen and an initial run was performed. The conditions were varied as described in

2.2.5 until a single, sharp band, no byproducts and a negative control were visualized

in an agarose gel.

In some cases, extreme variations in number of cycles and temperatures

became necessary to achieve a pure product. A new set of primers was required in

two cases (exons 34 and 38) after intense efforts with the first pair of primers failed to

yield acceptable results. It was particularly difficult to achieve quality products in exon

38. Almost none of the agarose gels showed clean bands for each patient. Usually

only half of the products were acceptable, when a run of PCR batches was

performed in exon 38. Consequently, many repeated runs were required to achieve

satisfactory products for each patient.

4.1.2 SSCP

The SSCP method is based on the migration of single stranded DNA

fragments in gel electrophoresis, as explained in section 2.2.7. The banding pattern

shown is unique for each DNA sequence. The amplified exons from each patient

were applied to SSCP gels for mutational screening as described in sections 2.2.7

and 3.2.2. Variations in the principal pattern visualized suggest variations in DNA

sequence, and each varying sample was sequenced.

Some exons exhibited simple patterns, which made the decision for samples

that required sequencing obvious. Exon 34, for example, shows a pattern with only 3

major bands, and figure 5 in section 3.1.2 demonstrates easily recognizable

variations from the norm. Other exons displayed complex banding, which made the

selection of varying patterns difficult. This is the case for exon 36, in which diffuse

banding was shown and patients were at times manually selected for sequencing

Discussion

33

(see appendix, figure 14). Exon 30 showed two main patterns of banding, and a

selection of patients from each of the two patterns was sequenced, revealing no

sequence variations (figure 14, appendix).

The use of SSCP for mutational screening is dependant upon the examination

of gels by a trained eye, and is therefore in certain cases not extraordinarily

objective. This is a definite drawback to the SSCP method. Subtle variations could be

overlooked and the diagnosis of a mutation consequently missed. The detection of

mutations in only about two thirds of patients could be in part explained by

weaknesses in SSCP screening.

Analysis of the sensitivity of SSCP vary somewhat from study to study. Orita et

al. (1989) state that it is possible to detect single nucleotide differences in a DNA

sample. Vidal-Puig and Moller (1994) report a 95 % (18 of 19) rate of mutation

detection via SSCP and the Phast System. In contrast, more recent studies

comparing SSCP with other detection methods, including direct sequencing found an

only 85 % rate of detection in a large pool of samples (88 of 103) (Jones et al.,

1999a). These studies concur, however, in that SSCP is a cost-effective and

technically simple method for screening a large patient pool prior to sequencing. The

results of this dissertation are in agreement on this point. We detected a total of 58

variations in 70 patients (83 %).

4.1.3 Sequencing 4.1.3.1 Mutations

9 variant sequences that were defined as mutations were found in 12 patients.

These results are summarized in table 4 (section 3.1.3.1), and the varying sections of

these sequences are shown in detail in figure 16 (appendix). Most of the sequence

findings correlate with a varying banding pattern in SSCP. In a few cases, however,

these results did not match up. Exon 30, for example, showed no deviations in

sequence despite seemingly different patterns in SSCP. The two patterns seen

(figure 14, appendix) were simply two normal banding versions.

As discussed in the previous section (4.1.2), it is a limitation of the SSCP

method that a level of uncertainty exists in screening DNA samples for mutations

before sequencing. The sequencing itself, on the other hand, is merely dependent

upon the quality of the DNA sample, which was purified according to methods

described in 2.2.8. If the resulting sequence was readable, a good DNA sample

Discussion

34

(PCR product) was achieved. The SSCP method was used as a pre-screening, to

significantly reduce the cost of our project and the time needed to sift through a large

patient pool.

As for the types of mutations found, these were also very diverse. Deletions,

insertions, silent base pair changes, missense mutations, and splice site variances

were all included in the findings (see table 4, section 3.1.3.1 and figure 9, section

3.3). The consequences of these mutations lie at the amino acid level. That is not

always easy to determine. For example, a change in a single nucleotide found in

patient 154 in exon 36 has no significance at the amino acid level. This is due to the

degenerate nature of the genetic code -- the fact that several different triplets can

code for a single amino acid. Often single nucleotide changes will not cause a

change in the amino acid. However, the single nucleotide change in exon 40 in

patient 46 causes a missense variation in the tuberin protein (R1743W), because a

significant nucleotide was substituted (here C5245T).

Small deletions may seem relatively harmless at first consideration. However,

depending at what position and how many nucleotides are affected, the change can

be quite serious. Patients 80, 113 and 114 all have the same 18 base pair in-frame

deletion. In contrast, patient 99 has a mere two base pair deletion that changes the

entire reading frame for exon 40. An insertion can also cause a shift in reading frame.

One insertion was found in the Freiburg-Heidelberg project (see figure 9, section

3.3), but none in the exons examined for this thesis (see table 4, section 3.1.3.1).

Another interesting type of sequence variation that was quite frequently

detected is the so-called splice variation. These deviations from the wild type

sequence occur in the intron region up to 10 nucleotides before exon begin, and can

affect the splicing of exons in the production of mRNA. This can result in exon

skipping, if the change is not homologous enough to the wild type sequence

(Krawczak et al., 1992). The finding that 13 patients showed the change A C at

-10bp in exon 40 (see table 5 in section 3.1.3.2) is an example of a very common

splice site variation. According to Krawczak et al. (1992), a pyrimidine is predicted to

be at this position with an 85 % probability. However, according to the Cardiff-

Rotterdam reference sequence (European Chromosome 16 Tuberous Sclerosis

Consortium, 1993), a purine (adenine, A) belongs in this position. This was

substituted with a pyrimidine (cytosine, C) in these patients. Interestingly, this

particular change is also quite frequent in the healthy control population (see section

4.2), which is evidence against the pathogenicity of this variation.

Discussion

35

Cheadle et al. (2000) consider a mutation to be pathogenic, when the

following criteria for missense, in-frame deletions and splice mutations outside the

two invariant donor and acceptor bases are met:

• The mutation was found in a sporadic TSC patient or a founding family member and

• the mutation was not present in the unaffected parents and

• the TSC patient DNA sample had been completely screened for other changes in all exons of TSC1 and TSC2.

Nonsense (nucleotide change resulting in a stop codon) and frame shift mutations,

on the other hand, are seen as clearly pathogenic because of protein truncation or

other significant disturbance in the amino acid sequence.

Despite these reasonable guidelines, what a mutation means at the protein

level is still not completely clear. How the amino acid sequence is affected is known,

but what a change in this sequence means for the protein function is not known. Until

the protein structure and function has been more clearly elucidated, the mutations

found can not be put into their complete context.

4.1.3.2 Polymorphisms

As described in section 3.1.3.2, a significant number of smaller changes were

discovered in the patient group and were classified as polymorphisms (see table 5 in

section 3.1.3.2). This classification was not simple, because the term polymorphism

seems to imply a benign change. This is not always the rule: the coexistence of

several variations of a sequence in a population is called polymorphism. More

precisely, an allele is defined as polymorph, when it is found to have a frequency

greater than 1 % in the population. This definition includes such examples of

polymorphism in healthy individuals as the blood groups and the genes that code for

the MHC (major histocompatibilty complex). (summarized in Lewin, 1998)

This definition does not, however, exclude frequent stable variations that can

be categorized as “sick”. A classic example of this is the polymorphism in the

hemoglobin gene that causes the disease sickle cell anemia. In wild type

hemoglobin, glutamic acid occupies amino acid position 6 of the β chain, while sickle

cell hemoglobin has valine at this position (von Harnack, 1994). This common

change satisfies the frequency regulation for the definition polymorphism. Although it

is small, its location dictates a structural change in hemoglobin and a sick phenotype.

Discussion

36

One exon in which the most polymorphisms were detected was exon 40 (see

table 5 in section 3.1.3.2). The reason for this is undoubtedly the complete screening

of this exon, as discussed in the following section.

4.1.3.3 Quality control and exon 40

The comparison of SSCP and sequence analysis in exon 40 is summarized in table 6 in section 3.1.4. In 4 patients, a variant SSCP was shown, but no sequence

deviation was found. Conversely, 26 patients had variant sequences but apparently

normal SSCP results. 15 of these 26 missed variations are silent or splice mutations.

This could indicate that most of the larger mutations were detected by SSCP and

less significant changes are more likely to be missed. This assumption is clearly

contradicted by the fact that 9 of the 10 sequence variations detected via SSCP are

also silent or splice mutations.

The SSCP method should detect changes as small as a single nucleotide

(Orita et al., 1989), and this is not contradicted by our results. Also, SSCP provides

for an inexpensive and relatively expeditious way to screen a large patient group.

However, the ability to detect mutations with a great degree of certainty was not

substantiated here. Hindrances that account for the discrepancies between SSCP

and sequence results are discussed in section 4.1.2.

4.2 Control group of healthy probands

The mutations found in the patient group lacked meaning without further

knowledge of the protein product and its functionality in TSC patients. As we know,

some smaller changes in DNA sequence do not always lead to amino acid changes.

Some amino acid changes are also not crucial to protein function. The size of a

mutation, therefore, does not always directly correlate to its gravity. This can be

decided when more about the protein is known.

One simple way to lend context to the mutational screening results was to

examine the variability of healthy TSC2. For this purpose, a group of 25 healthy

probands was screened for mutations in exon 40 of TSC2. Exon 40 was chosen,

because this exon was completely sequenced and would have the most complete

results. A control study of healthy exon 40 was of particular interest also because this

was an exon in which several seemingly significant mutations were found in our

patient group.

Discussion

37

DNA samples from 25 random, non-related, healthy probands for this control

group were kindly provided by Dr. Klaus Deichmann. As described in section 2.2.1.2,

these healthy samples were randomly selected from a large group of donor DNA.

The source of the DNA was healthy volunteers from various geographical regions, so

no hereditary or regional bias could be assumed.

4.2.1 Exon amplification and mutational screening

Exon 40 was amplified from genomic DNA for each of the healthy probands

via PCR. The procedure for exon 40 was already established in our patient group,

and it was not problematic to achieve satisfactory products. Mutational screening via

SSCP was also straightforward, due to the simple banding pattern displayed by exon

40 (see figure 13 in the appendix and figure 7 in section 3.2.2). All samples were

sequenced so as not to miss any minor variances.

4.2.2 Sequencing results and numerical comparison to patient group

As described in section 3.2.3, the group of healthy probands demonstrated an

unexpectedly high rate of variation. Two mutations were found to have a similar

frequency in both populations (see table 7). However, four mutations were found only

in the patient population (see table 8). Figure 8 in section 3.2.3 shows an example of

a variant sequence in the healthy population and the wild type sequence of exon 40

can be read in figure 15 in the appendix.

This is the key to placing known mutations in proper context. It is difficult to

say whether a mutation is disease causing until we know the full extent of natural

variation in the healthy population or can directly compare the wild type with the

mutated protein. Certain healthy genes in humans are by nature highly variant. One

example is the gene that codes for the variable portion of antibody L-chain (Lewin,

1998). In this case, variation is necessary to achieve the goal of specificity for many

various antigens with which our immune system is confronted. Other genes must be

strictly nonvariant, because their variation could cause a serious disease or be fatal.

The question is, then, how variant is TSC2 normally? Once we completely answer

this, then much more can be said about the mutations found in the patient population.

Not only the amount of variation, but significance of location in the TSC2 gene

can be much better determined, once a healthy population is thoroughly examined. It

can now be more accurately assumed, that the four mutations found only in the

Discussion

38

patient group are disease causing. The two frequent variations (polymorphisms) that

were found in both groups then have little consequence for TSC disease.

Polymorphisms in general, however, can not be dismissed as irrelevant. One

of the four mutations found only in the patient population occurred with a high

frequency as well. This is the nucleotide change T5180G, which results in the amino

acid change M1721R and is found in 10 % of our patients.

4.3 Screening the whole gene 4.3.1 Number and type of mutations found

Concerted screening of both TSC genes in this project yielded the detection of

29 sporadic mutations, 4 familial mutations and 9 unclassified variations in 69 TSC

patients. These results correspond to a 61 % (42/69) detection rate, which matches

literature values (Au et al., 1998, Zhang et al., 1999, among others). Nonetheless,

this is quite a discrepancy, when one considers that each of the patients included in

our study was carefully screened and received a definite TSC diagnosis. This is one

reason why it is particularly difficult to correlate phenotype with genotype in tuberous

sclerosis.

TSC is also characterized by a high rate of somatic mosaicism, which offers a

feasible explanation for finding mutations in only 60 % of patients (Yates et al., 1997).

In addition to mosaicism, an eventual third gene involved in TSC could explain the

failure to find a mutation in a number of the patients examined. However, this

possibility is becoming increasingly unlikely due to the concerted efforts of linkage

studies over the years (Cheadle et al., 2000). The likelihood of another locus is even

further reduced by the finding that half of the TSC families show linkage to TSC1 and

the other half to TSC2 (Cheadle et al., 2000).

To further complicate matters, there are no hot spots that can be concentrated

on in the study of this disease. As figure 9 in the results section shows, there is no

clustering of mutations in a specific region of the TSC2 gene. Of particular interest is

that mutations occur in regions of known GAP-related domains (exons 34-38) and

coiled-coil regions (exons 10 and 26) as well as in other regions. See figure 10 for a

summary of these findings.

Discussion

39

0

1

2

3

4

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41

Exon

coiled-coil domain N

umbe

r

coiled-coil domain

GAP-related domain

Figure 10: Variations by exon and region of the TSC2 gene. The number of definitive mutations found in the Freiburg-Heidelberg project (Langkau et al., 2002) are plotted by exon in the figure below. GAP-related and coiled-coil domains are highlighted. The mutations found are relatively evenly distributed across the entire gene. 4.3.2 Genotype-phenotype correlation

So far, no genotype-phenotype correlation has been convincingly

demonstrated for TSC. Part of the problem is the highly varying clinical

manifestations of the disease combined with the lack of a mutation pattern. As

mentioned above, we found no clustering of mutations in a particular region of the

TSC2 gene, which would provide clues for a functionally important area. Au et al.

(1998) clinically examined patients and compared these results with the type of

mutation found and concluded that the germ line mutation type does not influence the

clinical phenotype. These results are supported by other investigations as well

(Kwiatkowska et al., 1999; van Slegtenhorst et al., 1999; Mayer et al., 1999; Niida et

al., 1999).

There are few possible exceptions to this lack of correlation. In the case of a

contiguous gene deletion syndrome involving PKD1 and TSC2, which are very close

neighbors on chromosome 16, PKD1 may play a role in the etiology of renal cystic

disease in TSC patients as part of a one-step double second hit event (Cheadle et

Discussion

40

al., 2000). There is also evidence for a relationship between pulmonary

Lymphangiomyomatosis and TSC2 (Carsillo et al., 2000). Jones et al., (1997)

reported the finding that patients with TSC2 are significantly more often mentally

retarded than patients with TSC1 mutations, but this is not supported by other studies

(van Slegtenhorst et al., 1999; Niida et al., 1999; Yamashita et al., 2000).

Examples of inherited diseases with a more obvious connection between

genetic defect and clinical manifestation include cystic fibrosis and Duchenne

muscular dystrophy. The most common mutation causing cystic fibrosis, delta F508,

results in a more easily degradable protein product. Duchenne muscular dystrophy is

the consequence of a dystrophin protein deficiency. The size of the faulty area of the

gene correlates directly with the lack of dystrophin and the severity of the disease.

(summarized in von Harnack, 1994)

4.3.3 Clinical variability

If it is not yet possible to definitely correlate genetic defect and clinical

phenotype in TSC, then other theories for the very broad manifestations of TSC

disease must be explored. Despite the variable expressivity of TSC, it is thought to

be highly penetrant (Rose et al., 1999). This means that a defective TSC gene (TSC1

or TSC2) very definitely results in TSC disease. A complication of high penetrance is

the case of mosaicism, in which not all cells are affected by the mutation and,

consistent with a dosage effect, less severe clinical phenotypes result -- emulating

non-penetrance. The very nature of the tumor suppressor genes hamartin and

tuberin and the timing of their loss of function is the underlying reason for the varying

clinical outcome of TSC.

Cell growth and differentiation is complex, but highly regulated. A mistake

must be made in the normal process in order for tumorous tissue to occur.

Depending on when and where a mistake happens, the outcome can be highly

variable. (Young and Povey, 1998). This process of tumorigenesis and the role

played by tumor suppressor genes can be described by Knudson’s Two-Hit

Hypothesis (Knudson, 1971). This theory also provides explanation for the high

clinical variability of TSC.

According to Knudson, we start with two healthy alleles of a tumor suppressor

gene (such as TSC-2, for example). At some point a random event can occur,

causing a mutation in one of the alleles (usually a germ line mutation). The remaining

healthy allele is dominant and the phenotype is still healthy. Later, a random second

Discussion

41

hit can affect the second allele, for example a loss of heterozygosity, resulting in a

somatic mutation and a sick phenotype. A seemingly healthy parent could carry a

germ line mutation and pass this on to his children. When a second hit occurs in the

child, no healthy allele is available to provide a healthy gene product, and the result

is a TSC patient. This theory is diagrammed in figure 11 below.

Knudson Hypothesis

second hit first hit

somatic mutation germ line mutationnormal

Figure 11: The Knudson Two-Hit Hypothesis. We initially have two alleles of a normal gene. A random event causing a mutation in one of the alleles is called the first hit. If this is a germ line mutation, the phenotype is healthy, but the first hit can be inherited. If another random event occurs -- a second hit -- both alleles are affected and a sick phenotype results. 4.3.4 TSC1 and TSC2 frequency ratios

It has been suggested, however, that the ascertainment of TSC patients for

studies may be biased toward severe cases of TSC (Jones et al., 1997) and that

milder cases may be missed by epidemiological studies (Gillberg et al., 1994). This

could mean that the true number and phenotype of TSC patients is underestimated,

to further complicate a possible correlation to genotype.

An important result of this potential for under representation is that it may

further complicate genetic counseling. Genetic mosaicism may account for many mild

cases and for cases in which affected children are born to apparently unaffected

parents (Cheadle et al., 2000). Rose et al. (1999) examined families in which two

TSC children have clinically healthy parents and found germ line mosaicism in 5 of

Discussion

42

the 7 families. Based upon these results, a conservative recurrence risk for

apparently unaffected parents with one affected child is estimated to be 2 to 3 %

(Rose et al., 1999).

According to Langkau et al. (2002) and Cheadle et al. (2000), mildly affected

cases of TSC do not have a reduced reproductive fitness -- that is they are just as

likely to have offspring as healthy persons. It is assumed that mild forms of TSC are

based on mutations in TSC1 (Jones et al., 1997; Langkau et al., 2002), because

most sporadic cases involve TSC2 although half of TSC families show linkage to

TSC1 and the other half to TSC2. The explanation for this is that mild (TSC1) cases

are not clinically obvious and lie outside the ascertainment window. Our patient

population, for example, required strict fulfillment of diagnostic criteria for inclusion in

this study. Because of this, many mild cases may have been missed and that is why

so few TSC1 mutations were represented here and in other studies (Jones et al.,

1997; Cheadle et al., 2000; Langkau et al., 2002).

4.4 Conclusions

The mutations detected in this study were in number and type similar to

previous studies. No clustering at a particular region of the TSC2 gene was

observed. No genotype-phenotype correlation has been decisively made, and this

may be influenced by the widely varying clinical presentation of TSC. This variability

makes it a difficult disease to characterize, and may also account for missing milder

cases. Clinical variability can be explained by mosaicism (somatic and germ line) and

the Knudson Two Hit Hypothesis. The potential for metastasizing TSC lesions has

been initially examined, but is not proven and requires further study (Martignoni et al.,

2000).

Until more is known about the protein structures and functions of hamartin and

tuberin, the context of discovered mutations remains incomplete. The screening of a

healthy population has proven helpful in bringing mutations and polymorphisms

found into perspective until more is known about the protein products of the TSC

genes. This initial screen should be broadened to include all exons. An SSCP

example of various banding patterns observed in healthy populations would be

helpful in this pre-screening of samples, in particular those cases with complex

banding.

Discussion

43

Discussion

4.5 Future Screening: As screening efforts advance a clustering of mutations may be found to allow

for a concentrated effort on a smaller group of exons. SSCP as a screening method

is inexpensive, uncomplicated and relatively reliable. In combination with other

methods (such as DHPLC or PTT), an increased detection sensitivity has been

observed (Jones et al., 1999a). Other methods are more complex and can be used in

specialized centers with greater human resources. The eventual development of

DNA chips for diagnostics in the future would provide a possibility for diagnostic

screening (Young and Povey, 1998). As yet the time consuming and expensive

process is a hindrance to this goal.

Therapy:

As yet the therapy of TSC is symptomatic. Antiepileptic drugs for reducing

seizure frequency and early support are key factors in combating mental retardation

and behavioral complications. These symptoms are of great concern to patients and

parents. The surgical removal of hemorrhages and fibromas when possible is

important for the management of the disease as well. (Gutierrez et al., 1998)

Causal therapy in the form of gene therapy is far in the future. Local

application or substitution of missing proteins may become possible once more is

known about hamartin and tuberin. This may be especially difficult in TSC because

both proteins involved are ubiquitous -- reducing the possibility for a cure via gene

therapy. The concentration of efforts in localized lesions may therefore be necessary.

Cancer research:

The implications of TSC in cancer research in general will be exciting in the

years to come. Ubiquitous proteins that are phylogenetically conserved and play a

role in cell cycle control are undeniably interesting to study. The understanding of the

tumor suppressor process in this disease may provide valuable insight for research

into this and other diseases involving tumors.

44

5 Summary

Tuberous sclerosis complex is an inherited disorder characterized by the development of benign tumors in various tissues. These tumors can affect the central nervous system, skin, kidney, heart and almost any organ. The frequency is currently estimated to be 1:6000 and the variation in severity of disease is great. Epilepsy, mental retardation, renal failure and reduced life expectancy are all possible, although a number of mildly affected persons lead normal lives with no reduction in reproductive fitness.

Two genes and their products have been identified as causing TSC when they are defective: TSC1 on chromosome 9q34.3 codes for the protein hamartin and TSC2 on chromosome 16p13.3 codes for tuberin. Tuberin shows homology to GTPase activating proteins (GAPs) and hamartin and tuberin may interact with one another to control the cell cycle, suggesting a tumor suppressor function for these proteins.

Efforts to screen patient populations for mutations in both genes in this study and others have been successful in identifying a large number of mutations. However, no clustering of mutations has been shown that might allow for concentration of efforts on a particular region of either gene. Sporadic mutations in TSC1 are less frequent than in TSC2 despite the fact that half of the TSC families show linkage to TSC1 and the other half to TSC2. It has been postulated that TSC1 patients are more mildly affected and may have a higher reproductive fitness than TSC2 patients.

The wide distribution of mutations and the high variability in clinical manifestation of TSC further complicate the quest for genotype-phenotype correlation. If the structure and function of the protein products were clearly established, then the consequence of the great number of mutations found thus far would be known. In order to lend some context to the mutations in the patient population, a group of healthy probands was screened as a control group in exon 40 of TSC2.

29 sporadic mutations, 4 familial cases, 9 unclassified variations and 16 polymorphisms making a total of 58 variations from the wild-type sequence were found in the 70 patients examined for this project. 8 out of 25 healthy probands also demonstrated variations in sequence, which were classified as polymorphisms. The presence of these variations in the healthy population helps define them as non-disease causing.

Future screening efforts may include methods of even higher sensitivity and the control of a healthy population until the gene products are better characterized. As screening efforts improve, diagnostic methods will evolve and greatly support genetic counseling. And as more is known about the function of tuberin and hamartin the closer we will come to the possibility of a causal therapy for TSC.

45

6 Zusammenfassung

Tuberöse Sklerose Komplex (TSC) ist eine vererbbare Erkrankung, die durch das Auftreten von gutartigen Tumoren gekennzeichnet wird. Die Tumoren können sich im zentralen Nervensystem, der Haut, der Niere, dem Herz und fast jedem anderen Organ entwickeln. Derzeit wird eine Häufigkeit von 1:6000 angenommen, wobei die klinischen Erscheinungen bei den einzelnen Betroffenen stark variieren. Es gibt schwere Verläufe mit Epilepsie, geistiger Behinderung, Niereninsuffizienz und verminderter Lebens-erwartung. Gleichzeitig sind auch Patienten bekannt, die ein normales Leben führen.

Zwei Gene und ihre jeweiligen Proteinprodukte sind identifiziert worden, die TSC verursachen, wenn sie Defekte aufweisen: TSC1 auf Chromosom 9q34.3 kodiert für das Protein Hamartin und TSC2 auf Chromosom 16p13.3 für Tuberin. Tuberin weist Sequenzhomologien zu bekannten GTPase-aktivierenden Proteinen (GAP) auf. Hamartin und Tuberin interagieren und üben möglicherweise Einfluß auf den Zellzyklus aus, was auf eine Tumorsuppressorfunktion dieser Proteine hinweist.

Durch genetisches Screening von Patientengruppen wurden eine Vielzahl von Mutationen in dieser und in anderen Arbeiten gefunden. Jedoch wurde keine Anhäufung von Mutationen in einer bestimmten Genregion festgestellt, auf die sich die Forschung konzentrieren könnte. Obwohl die Hälfte aller TSC Familien genetische Kopplung zu TSC1 und die andere zu TSC2 zeigen, sind sporadische Mutationen in TSC1 seltener als in TSC2. Es ist vermutet worden, daß TSC1 Patienten mildere Verläufe und bessere relative Fortpflanzungs-Fitness als TSC2 Patienten aufweisen.

Die breite Streuung der Mutationen und die hohe klinische Variabilität der Erkrankung sorgt für Herausforderungen in der Suche nach genotyp-phänotyp Korrela-tionen. Wäre die Struktur der Proteine aufgeklärt, würde man die Konsequenzen der Vielzahl der bisher gefundenen Mutationen eher erkennen können. Um die bekannten Sequenzvariationen in einen größeren Zusammenhang zu stellen, wurde in dieser Arbeit Exon 40 des TSC2 Gens zusätzlich in einer Gruppe gesunder Probanden untersucht.

In den 70 Patienten dieses Projektes sind zusammen 58 Abweichungen von der Wildtyp-Sequenz gefunden worden: 29 sporadische Mutationen, 4 familäre Fälle, 9 unklassifizierte Variationen und 16 Polymorphismen. 8 von 25 gesunde Probanden wiesen ebenfalls Sequenzveränderungen auf, die als Polymorphismen klassifiziert wurden. Diese Veränderungen in der gesunden Gruppe können somit als nicht krankheitsverursachend eingestuft werden.

Genetische Screeningmethoden, die noch höhere Sensitivität bieten, werden sicherlich in der Zukunft weiterentwickelt. Bis die Strukturen der TSC-Proteine bekannt werden, werden aber intensivere Untersuchungen in gesunden Populationen notwendig bleiben. Diagnostische Möglichkeiten für betroffene Familien werden ebenso mit der Zeit verfeinert und werden die genetische Beratung weiter unterstützen. Je mehr über die Funktion von Hamartin und Tuberin bekannt wird, desto näher rückt die Möglichkeit einer kausalen Therapie für TSC.

46

7 Appendix

Exon 30

Exon 34

Exon 36

Exon 38

Exon 39

Exon 40

300 bp

300 bp

200 bp

200 bp

400 bp

200 bp

Figure 12: Examples of PCR results for each exon. For PCR conditions and expected product size, see table 2 in section 2.2.5. Some gels were photographed as negative images to show faint bands more clearly (here in exon 38, for example).

300 bp

1 2 3 4 5 6 7 8 9 10 11 Figure 13: Agarose gel result from healthy population, exon 40. Lane 1: DNA size standard (and positive control). Lanes 2 – 9: probands 7, 9, 10, 11, 12, 13, 14, and 15. Lane 10: empty. Lane 11: negative control. For PCR conditions and product size, see table 2 in section 2.2.5.

47

Exon 30 Exon 34

Exon 36 Exon 38

Exon 39 Exon 40

111 117 121 122 124 128 138 143

extra bands

pattern 1

pattern 2 63 21 117 23 72 77 28 33

• Conditions: 5 °C and 380 Vh run length• Patients 21, 23, 28, 33, 117 sequenced

with no variations revealed.

• Conditions: 5 °C and 250 Vh run length • Patients 111 and 117 show variations

See results section 3.2.2 (control group) for SSCP of exon 40 (figure 7).

138 139 143 146 151 152 154 155

• Conditions: 5 °C and 580 Vh run length • Patient 154 selected for sequencing

missing band

• Conditions: 5 °C and 380 Vh run length • Patients 109 and 115 show variations

extra bands

double band

17 101 103 104 105 138 151 155

105 106 109 111 113 114 115 116

• Conditions: 15 °C and 320 Vh run length • Patient 155 shows variant banding

Figure 14: Examples of SSCP results for each exon. Comments for each exon shown here are below each gel; table 4 in section 3.1.3.1 lists the sequencing results.

Appendix

48

Exon 30:

Exon 34:

Exon 36:

Exon 38:

Exon 39:

Exon 40:

N

ucleotides 5179-5277: (Intron ctac gtccccag...) atg gcctcacagg tgcatcatag ccgctccaac cccaccgata tctacccctc caagtggatt gcccggctcc gccacatcaa gcggctccgc cagcgg

N

ucleotides 5087-5178: (Intron accaccaagt ctccccag...) START ac atggagggcc ttgtggacac cagcgtggcc aagatcgtgt ctgaccgcaa cctgcccttc gtggcccgcc agatggccct gcacgcaaat

N

ucleotides 5008-5086: ggccagt tcaactttgt ccacgtgatc gtcaccccgc tggactacga gtgcaacctg gtgtccctgc agtgcaggaa ag

N

ucleotides 4681-4867: agcaaca gcgagctcgc catcctgtcc aatgagcatg gctcctacag gtacacggag ttcctgacgg gcctgggccg gctcatcgag ctgaaggact gccagccgga caaggtgtac ctgggaggcc tggacgtgtg tggtgaggac ggccagttca cctactgctg gcacgatgac atcatgcaag

N

ucleotides 4512-4587: tttcgtgtt cctgcagctc taccattccc ccttctttgg cgacgagtca aacaagccaa tcctgctgcc caatgag

N

ucleotides 3629-3832: ggaac accagctggc tgatgagcct ggagaacccg ctcagccctt tctcctcgga catcaacaac atgcccctgc aggagctgtc taacgccctc atggcggctg agcgcttcaa ggagcaccgg gacacagccc tgtacaagtc actgtcggtg ccggcagcca gcacggccaa accccctcct ctgcctcgct ccaacacag

Figure 15: Wild-type sequence for each exon. The normal text sequence for each exon is printed below. The intron sequences shortly before exons 39 and 40 are included for reference to frequent polymorphisms (39: -2 bp A C and 40: -10 bp A C).

Appendix

49

Figure 16: Variant sequences. Variant sequences for each exon are displayed in the figure below. The patient number is noted and the nucleotide position is marked. For a complete list of changes found, see table 4 in section 3.1.3.1 (mutations) and table 5 in section 3.1.3.2 (polymorphisms).

Exon 34, Patient 111: 4562-4565 del ACAA, frame shift Exon 34, Patient 117: 4521 del C, frame shift Exon 34, Patient 139: 4545-4547 del CTT, Exon 36, Patient 154: C4695G, frame shift silent change Exon 38, Patient 109: C5042T, missense Exon 38, Patient 115: C5042T, P1675L missense P1675L (…continued on next page…) Appendix

50425042

start

4695

4545

4521

start

4562

50

Appendix

(Figure 16: Variant sequences, continued): Exon 39, Patient 155: -2 A C, splice Exon 40, Patient 99: 5188-5189 del CA,

frame shift

Exon 40, Patient 46: T5220C, silent (polymorphism found in 22 of 70 patients and 8 of 25 controls) and C5245T, missense R1743W (unique) Exon 40, Patient 80: 5256-5273 in-frame deletion of 18 bp (same mutation as in patients 113 and 114)

Exon 40, Patient 121: -10 A C, splice (possible exon skipping, polymorphism found in 13 of 70 patients and 7 of 25 controls)

Exon 40, Patient 128: T5180G, missense M1721R (polymorphism found in7 of 70 patients)

start

5188

start

-2A C

start

5180

-10 A C start

52455220

5256

51

8 Abbreviations and Symbols A Adenosine (nucleoside, as in AMP)

AgNO3 Silver nitrate

approx. Approximately

AMP Adenosine monophospate, a nucleotide

ATP Adenosine triphospate

bp Base pairs

C Cytidine (nucleoside, as in CMP)

°C Degrees Celsius

CMP Cytidine monophosphate, a nucleotide

CNS Central nervous system

CORE Sequence facility of the Universitätsklinikum Freiburg

CTP Cytidine triphosphate

CYLD Familial cylindromatosis

ddH2O Deionized distilled water

DHPLC Denaturing high performance liquid chromatography

DMSO Dimethyl sulfoxide

DNA Desoxyribonucleic acid

dNTP Desoxyribonucleic acid triphosphate (dATP, dCTP, dGTP, dTTP)

EBV Epstein-Barr virus

EDTA Ethylene diamine tetraamino acid

F Phenylalanine

FCS Fetal calf serum

FISH Fluorescent in-situ hybridization

g Gram(s)

G Guanosine (nucleoside, as in GMP)

G0 Part of the interphase portion of the cell cycle (gap - or interruption in NA synthesis). D

GAP GTPase activating protein

GDP Guanosine diphosphate

GMP Guanosine monophosphate, a nucleotide

GTP Guanosine triphosphate

h Hour(s)

H2O Water

52

Abbreviations

HPLC High performance liquid chromatography MgCl2 Magnesium chloride µg Microgram(s) MHC Major histocompatibility complex

min. Minute(s) ml Milliliter(s) µl Microliter(s) mM Millimolar µM Micromolar Na Sodium

Na2CO3 Sodium carbonate NH4 Ammonium PBS Phosphate buffered saline PCR Polymerase chain reaction PKD Polycystic kidney disease PTT Protein truncation test

rpm Rotations per minute RPMI Cell culture medium (Roswell Park Memorial Institute) s Second(s) S Part of the interphase portion of the cell cycle (synthesis of DNA) SSCP Single strand conformational polymorphism

T Thymidine (nucleoside, as in TMP) TBE Buffer consisting of: Tris, boric acid and EDTA TCA Tricarboxylic acid TMP Thymidine monophosphate, a nucleotide TTP Thymidine triphosphate TSC Tuberous sclerosis complex

TSC1 TSC gene in humans, on chromosome 9q34 TSC2 TSC gene in humans, on chromosome 16p13 tsc1 or tsc2 TSC genes in non-humans (e.g. Eker rat model) U Uridine (nucleoside, as in UMP) UMP Uridine monophosphate, a nucleotide

UV Ultra violet V Volt(s) Vh Volt-hour(s)

53

9 Literature

Au KS, Rodriguez JA, Finch JL, Volcik KA, Roach ES, Delgado MR, Rodrigues E Jr,

and Northrup H (1998) Germ-line mutational analysis of the TSC2 gene in 90

tuberous-sclerosis patients. Am J Hum Genet 62:286-294

Baker P, Piven J, Yutaka S (1998) Autism and tuberous sclerosis complex:

prevalence and clinical features. J Autism Dev Disord 28(4):279-285

Becker WM and Deamer DW (1991) The World of the Cell. Benjamin/Cummings

Publishing Company, Inc., Redwood City, CA/USA

Carbonara C, Longa L, Grosso E, Borrone C, Garre MG, Brisigotti M, Bigone N (1994)

9q34 loss of heterozygosity in tuberous sclerosis astrocytoma suggests a growth

suppressor-like activity also for the TSC1 gene. Hum Mol Genet 3:1829-1832

Carsillo T, Astrinidis A, Henske EP (2000) Mutations in the tuberous sclerosis complex

gene TSC are a cause of sporadic pulmonary lymphangioleiomyomatosis. Proc Natl

Acad Sci USA 97(11):6085-6090

Cheadle JP, Reeve MP, Sampson JR, Kwiatkowski DJ (2000) Molecular genetic

advances in tuberous sclerosis. Hum Genet 107:97-114

Crino PB, Henske EP (1999) New developments in the neurobiology of the tuberous

sclerosis complex. Neurology 53(7):1384-1390

European Chromosome 16 Tuberous Sclerosis Consortium (1993) Identification and

characterization of the tuberous sclerosis gene on chromosome 16. Cell 75:1305-1313

Geist R, Gutmann DH (1995) The tuberous sclerosis 2 gene is expressed at high

levels in the cerebellum and developing spinal cord. Cell Growth Differ 6(11):1477-

1483

Gillberg Ca, Gillberg Ch, Ahlsen G (1994) Autistic behavior and attention deficits in

tuberous sclerosis: a population based study. Dev Med Child Neurol 36:50-56

Gomez MR, Sampson JR, Whittemore VH (1999) Tuberous Sclerosis Complex.

Oxford University Press, New York-Oxford

Berg H (1913) Z Ges Neurol Psychiatr 19:528-539 (as cited in Gomez, 1999)

Critchley M, Earl CJC (1932) Brain 55:311-346 (as cited in Gomez, 1999)

Vogt H (1908) Z Erforsch Behandl Jugendl Schwachsinns 2:1-12 (as cited in

Gomez, 1999)

Von Recklinghausen F (1862) Monatsschr Geburtskd 20:1-2 (as cited in Gomez,

1999)

54

Green AJ, Smith M, Yates JR (1994) Loss of heterozygosity on chromosome 16p133

in hamartomas from tuberous sclerosis patients. Nat Genet 6:193-196

Green AJ, Sepp T, Yates JRW (1996) Clonality of tuberous sclerosis hamartomas

shown by non-random X-chromosome inactivation. Hum Genet 97:240-243

Gutierrez G, Smalley SL, Tanguay PE (1998) Autism in tuberous sclerosis complex. J

Autism Dev Disord 28(2):97-103

Harms V (1988) Biomathematik, Statistik und Dokumentation. Harms Verlag, Kiel

Henry KW, Yuan X, Koszewaki NJ, Onda H, Kwiatkowski DJ, and Noonan DJ (1998)

Tuberous sclerosis gene 2 product modulates transcription mediated by steroid

hormone receptor family members. J Biol Chem 273:20535-20539

Henske EP, Scheithauer BW, Short MP, Wollman R, Nahmias J, Hornigold N, van

Slegtenhorst M, Welsh CT, Kwiatkowski D (1996) Allelic loss is frequent in tuberous

sclerosis kidney lesions but rare in brain lesions. Am J Hum Genet 59:400-406

Hunt A, Dennis J (1987) Psychiatric disorder among children with tuberous sclerosis.

Dev Med Child Neurol 29:190-198

Hunt A, Stores G (1994) Sleep disorder and epilepsy in children with tuberous

sclerosis: A questionnaire-based study. Dev Med Child Neurol 36:108-115

International Molecular Genetic Study of Autism Consortium (1998) A full genome

screen for autism with evidence for linkage to a region on chromosome 7q. Hum

Mol Genet 7(3):571-8

Ito N and Rubin G (1999) Gigas, a Drosophila homolog of tuberous sclerosis gene

product-2, regulates the cell cycle. Cell 96:529-539

Janssen B, Sampson J, van der Est M, Deelen W, Verhoef S, Daniels I, Hesseling A,

Brook-Carter P, Nellist M, Lindhout D, Sandkuijl L, Halley D (1994) Refined

localization of TSC1 by combined analysis of 9q34 and 16p13 data in 14 tuberous

sclerosis families. Hum Genet 94(4):437-40

Jin F, Wienecke R, Guang-Hui X, Maize JC, DeClue JE, Yeung RS (1996)

Suppression of tumorigenicity by the wild-type tuberous sclerosis 2 (Tsc2) gene and

its C-terminal region. Proc Natl Acad Sci USA 93:9154-9159

Jones AC, Austin J, Hansen N, Hoogendoorn B, Oefner PJ, Cheadle JP, O’Donovan

MC (1999a) Optimal temperature selection for mutation detection by denaturing

HPLC and comparison to single-stranded conformation polymorphism and

heteroduplex analysis. Clin Chem 45(8):1133-1140

Literature

55

Jones AC, Daniells CE, Snell RG, Tachatake M, Idziaszczyk SA, Krawczak M,

Sampson JR, Cheadle JP (1997) Molecular genetic and phenotypic analysis reveals

differences between TSC1 and TSC2 associated familial and sporadic tuberous

sclerosis. Hum Mol Genet 6(12):2155-2161

Jones AC, Shyamsundar MM, Thomas MW, Maynard J, Idziaszczyk S, Tomkins S,

Sampson JR, Cheadle JP (1999b) Comprehensive mutation analysis of TSC1 and

TSC2 – and phenotypic correlations in 150 families with tuberous sclerosis. Am J

Hum Genet 64:1305-1315

Knudson AG (1971) Mutation and cancer: statistical study of retinoblastoma. Proc Natl

Acad Sci USA 68:820-823

Knudson AG (1993) Antioncogenes and human cancer. Proc Natl Acad Sci USA

90:10914-10921

Krawczak M, Reiss J, Cooper DN (1992) The mutational spectrum of single base-pair

substitutions in mRNA splice junctions of human genes: causes and consequences.

Hum Genet 90(1-2):41-54

Kwiatkowska J, Wigowska-Sowinska J, Napierala D, Slomski R, Kwiatkowski D (1999)

Mosaicism in tuberous sclerosis as a potential cause of the failure of molecular

diagnosis. N Engl J Med 340(9):703-707

Kwiatkowski D and Short P (1994) Tuberous Sclerosis. Arch Dermatol 130:348-354

Langkau N, Martin N, Brandt R, Zügge K, Quast S, Wiegele G, Jauch A, Rehm M,

Kuhl A, Mack-Vetter M, Zimmerhackl LB, Janssen B (2002) TSC1 and TSC2

mutations in tuberous sclerosis, the associated phenotypes and a model to explain

observed TSC1/TSC2 frequency ratios. Eur J Pediatr 161:393-402.

Lewin B (1998) Molekularbiologie der Gene. Spektrum Akademischer Verlag,

Heidelberg-Berlin

Martignoni G, Pea M, Rigaud G, Manfrin E, Colato C, Zamboni G, Scarpa A,

Tardanico R, Roncalli M, Bonetti F (2000) Renal angiomyolipoma with epithelioid

sarcomatous transformation and metastases: demonstration of the same genetic

defects in the primary and metastatic lesions. Am J Surg Pathol 24:889-894

Martin N, Gierke K, Langkau N, Brandt R, Quast S, Wiegele G, Jauch A, Rehm M,

Kuhl A, Mack-Vetter M, Janssen B and Zimmerhackl LB. Discordant clinical

manifestations in monozygotic twins with the identical mutation in the TSC2 gene.

(accepted for publication by Clin Genet)

Literature

56

Mayer K, Ballhausen W, Rott HD (1999) Mutation screening of the entire coding

regions of the TSC1 and the TSC2 gene with the protein truncation test (PTT)

identifies frequent splicing defects. Hum Mutat 14(5):401-11

Niida Y, Lawrence-Smith N, Banwell A, Hammer E, Lewis J, Beauchamp RL, Sims K,

Ramesh V, Ozelius L (1999) Analysis of both TSC1 and TSC2 for germline

mutations in 126 unrelated patients with tuberous sclerosis. Hum Mutat 14(5):412-

22

Orita M, Iwahana H, Kanazawa H, Hayashi K, Sekiya T (1989) Detection of

polymorphisms of human DNA by gel electrophoresis as single-strand-conformation

polymorphisms. Proc Natl Acad Sci USA 83(3):2766-2770

Osborne JP, Fryer A, Webb D (1991) Epidemiology of tuberous sclerosis. Ann NY

Acad Sci 615:125-127

Plank TL, Yeung RS, Henske EP (1998) Hamartin, the product of the tuberous

sclerosis 1 (TSC1) gene, interacts with tuberin and appears to be localized to

cytoplasmatic vesicles. Cancer Res 58:4766-4770

Povey S, Burley MW, Attwood J, Benham F, Hunt D, Jeremiah SJ et al. (1994) Two

loci for tuberous sclerosis: one on 9q34 and one on 16p13. Ann Hum Genet 58:107-

127

Roach ES, DiMario FJ, Kandt RS, Northrup H (1999) Tuberous sclerosis consensus

conference: recommendations for diagnostic evaluation. J Child Neurol 14(6):401-

407

Rose VM, Au KS, Pollom G, Roach ES, Prashner HR, Northrup H (1999) Germ-line

mosaicism in tuberous sclerosis: how common? Am J Hum Genet 64:986-992

Rott HD, Ballhausen W, Mayer K (1999) Klinik und Genetik der tuberösen Sklerose.

Pädiat Prax 56:233-244

Smalley SL (1998) Autism and tuberous sclerosis. J Autism Dev Disord 28(5):407-14

Soucek T, Pusch O, Wienecke R, DeClue JE, Hentstschläger M (1997) Role of the

tuberous sclerosis gene-2 product in cell cycle control. J of Biol Chem

272(46):29301-29308

Vanderhooft SL, Francis JS, Pagon RA, Smith LT, Sybert VP (1996) Prevalence of

hypopigmented macules in a healthy population. J Pediatr 129(3):355-361

Literature

57

Van Slegtenhorst M, de Hoogt R, Hermans C, Nellist M, Janssen B, Verhoef S,

Lindhout D, van den Ouwenland A, Halley D, Young J, Burley M, Jeremiah S,

Woodward K, Nahmias J, Fox M, Ekong R, Osborne J, Wolfe J, Povey S, Snell R,

Cheadle J, Jones A, Tachataki M, Ravine D, Sampson J, Reeve M, Richardson P,

Wilmer R, Munro C,Hawkins T, Sepp T, Ali J, Ward S, Green A, Yates J,

Kwiatkowska J, Henske E, Short M, Haines J, Jozwiak S, and Kwiatkowski D (1997)

Identification of the tuberous sclerosis gene TSC1 on chromosome 9q34. Science

277:805-808

Van Slegtenhorst M, Nellist M, Nagelkerken B, Cheadle J, Snell R, van den

Ouwenland A, Reuser A, Sampson J, Halley D, van der Sluijs P (1998) Interaction

between hamartin and tuberin, the TSC1 and TSC2 gene products. Hum Mol Genet

7(6):1053-1057

Van Slegtenhorst M, Verhoef S, Tempelaars A, Bakker L, Wang Q, Wessels M,

Bakker R, Nellist M, Lindhout D, Halley D, van den Ouwenland A, (1999) Mutational

spectrum of the TSC1 gene in a cohort of 225 tuberous sclerosis complex patients:

no evidence for genotype-phenotype correlation. J Med Genet 36:285-289

Verhoef S, Bakker L, Tempelaars AMP, Hesseling-Janssen ALW, Mazurczak T,

Jozwiak S, Fois A, Bartalini G, Zonnenberg BA, van Essen AJ, Lindhout D, Halley

DJJ, van den Oweland AMW (1999) High rate of mosaicism in tuberous sclerosis

complex. Am J Hum Gen 64:1632-1637

Verhoef S, Schrander-Stumpel CTRM, Vuzevski VD, Tempelaars AMP, Jansen ALJ,

Malfeyt GAM, Ceelen TL, Lindhout D, Halley DJJ, van den Oweland AMW (1998)

Familial cylindromatosis mimicking tuberous sclerosis complex and confirmation of

the cylindromatosis locus, CYLD-1, in a large family. J Med Genet 35(10):841-5

Vidal-Puig A and Moller DE (1994) Comparative sensitivity of alternative single-strand

conformational polymorphism (SSCP) methods. Biotechniques 17(3):490-2, 494,

496.

Von Harnack, ed. (1994) Kinderheilkunde. 9 Aufl. Springer, Berlin-Heidelberg

Webb DW, Osborne JP, (1991) Non-penetrance in tuberous sclerosis. J Med Genet

28:417-419

Wienecke R, König A, De Clue JE (1995) Identification of tuberin, the tuberous

sclerosis-2 product. J Biol Chem 270(27):16409-16414

Literature

58

Literature

Wienecke R, Maize JC Jr, Shoarinejad F, Vass WC, Reed J, Bonifacino JS, Resau JH,

de Gunzburg J, Yeung RS, DeClue JE (1996) Co-localization of the TSC2 product

tuberin with its target Rap1 in the Golgi apparatus. Oncogene 13:913-923

Xiao GH, Shoarinejad F, Jin F, Golemis EA, Yeung RS (1997) The tuberous sclerosis

2 gene product, tuberin, functions as a Rab5 GTPase activating protein (GAP) in

modulating endocytosis. J Biol Chem 272:6097-6100

Yamashita Y, Ono J, Okada S, Wataya-Kaneda M, Yoshikawa K, Nishizawa M,

Hirayama Y, Kobayashi E, Seyama K, Hino O (2000) Analysis of all exons of TSC1

and TSC2 genes for germ line mutations in Japanese patients with tuberous

sclerosis: report of 10 mutations. Am J Med Genet 90(2):123-6

Yates JRW, van Bakel I, Sepp T, Payne SJ, Webb DW, Nevin NC, Green AJ (1997)

Female germ line mosaicism in tuberous sclerosis confirmed by molecular genetic

analysis. Hum Mol Genet 6(13):2265-2269

Young J and Povey S (1998) The genetic basis of tuberous sclerosis. Mol Med Today

July:313-319

Zhang H, Nanba E, Yamamoto T, Ninomiya H, Ohno K, Mizuguchi M, Takeshita K

(1999) Mutational analysis of TSC1 and TSC2 genes in Japanese patients with

tuberous sclerosis complex. J Hum Genet 44:391-396

Zimmerhackl LB, Rehm M, Kaufmehl K, Kurlemann G, Brandis M (1994) Renal

involvement in tuberous sclerosis complex: a retrospective survey. Pediatr Nephrol

8(4):451-7

59

10 Acknowledgments

I would first like to thank Prof. Dr. L.B. Zimmerhackl for allowing me to be part

of this research project, for his immense time investment and excellent advising

throughout the entire dissertation.

I am likewise very indebted to my predecessor, Dr. Nicola Martin, for her

patient explanations, fruitful discussions and thorough training in methods.

The entire Nephrolabor and all members of our research group deserve many

thanks for their teamwork and for the very valuable, pleasant and educational

experience we shared. I am particularly grateful to the technical assistants in the

Nephrolabor, Bärbel Höfflin, Simone Joos, Monika Kramer and Monika Mack-Vetter

who organized materials and were in charge of all the cell cultures.

A great deal of credit for an enjoyable and productive cooperation on this

project is owed to Dr. Bart Janssen and coworkers from the Heidelberg Institut für

Humangenetik.

I also thank Dr. Angela Gerber, Andrea Lohrmann, Dr. Nicola Martin, Dr.

Jochen Zügge, and Dr. Ulrich Zügge for their indispensable efforts in reading,

correcting and commenting on this thesis.

I truly appreciate the work of Dr. Hennighausen and Prof. Dr. Schulz for their

support as second referees of this thesis.

Last but far from least, I would like to thank my family for their constant

encouragement and support, without which I never would have been able to finish

this project.

60

11 Curriculum vitae

Birth date and place

January 25th, 1973 in Madison, Wisconsin / U.S.A. Maiden name: Gierke

Current position

July 1st, 2004 - present

Ärztin im Praktikum (first year resident) Department of Anesthesiology Hospital zum Heiligen Geist, Frankfurt

University education

July 2nd, 2004 USMLE Step 2

J une 9th, 2004 Drittes Staatsexamen (3rd state examination)

October 2002 - June 2004

J ohann-Wolfgang-von-Goethe-Universität, Frankfurt Practical year, Universitätsklinikum April 2003 - March 2004

Elective subject: pediatrics Rotations in transplant and thoracic surgery, University

of Wisconsin - Madison, April - June 2003 Zweites Staatsexamen (2nd state examination) March 2003

October 1996 - September 2002

Albert-Ludwigs-Universität, Freiburg Thesis in the nephrology laboratory of the Universitäts

Kinderklinik, advisor: Prof. Dr. L.B. Zimmerhackl, October 1999 - July 2002

USMLE Step 1, March 2001 Erstes Staatsexamen (1st state examination), August 1999 Physikum (preclinical examination), March 1998

1991 - 1996 University of Wisconsin - Madison / U.S.A.

Bachelor of Science (B.S.) with honors in biochemistry and German literature, May 1996

Honors thesis in biochemistry, 1995 - 1996 Study abroad year at the Albert-Ludwigs-Universität,

Freiburg, 1994 - 1995

61

12 Publications

(in chronological order) Martin N, Zügge K, Langkau N, Brandt R, Friebel D, Janssen B and Zimmerhackl LB

(2003) Discordant clinical manifestations in monozygotic twins with the identical

mutation in the TSC2 gene. Clin Genet May 63(5):427-30.

Lankau N, Martin N, Brandt R, Zügge K, Quast S, Wiegele G, Jauch A, Rehm M,

Kuhl A, Mack-Vetter M, Zimmerhackl LB and Janssen B (2002) TSC1 and TSC2

mutations in tuberous sclerosis, the associated phenotypes and a model to

explain observed TSC1/TSC2 frequency ratios. Eur J Pediatr July 161:393-402.

Gierke K, Martin N, Langkau N, Brandt R, Quast S, Wiegele G, Jauch A, Rehm M,

Mack-Vetter M, Janssen B and Zimmerhackl LB (2001) Tuberous Sclerosis

Complex (TSC): Mutational analysis in patients and healthy probands – the

challenges of genotype-phenotype correlation. Arbeitsgemeinschaft für

Pädiatrische Nephrologie in Wien, 22.-24. März.

Gierke K, Martin N, Langkau N, Brandt R, Quast S, Wiegele G, Jauch A, Rehm M,

Mack-Vetter M, Janssen B and Zimmerhackl LB (2000) Mutational Analysis and

Genotype-Phenotype Correlation in Tuberous Sclerosis Complex (TSC) – the

Freiburg-Heidelberg Project. Third Symposium of the Upper Rhine University

Children‘s Hospitals in Basel, June 30th.

Spangler NJ, Meyers MR, Gierke KL, Kerby RL, Roberts GP, Ludden PW (1998)

Substitution of valine for histidine 265 in carbon monoxide dehydrogenase from

Rhodospirillum rubrum affects activity and spectroscopic states. J Biol Chem Feb

13; 273(7): 4059-64.