+ All Categories
Home > Documents > ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1...

ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1...

Date post: 26-Feb-2021
Category:
Upload: others
View: 5 times
Download: 0 times
Share this document with a friend
100
1 ANALYSIS OF NF1 MUTATION MECHANISMS By REBECCA L. LODA-HUTCHINSON A DISSERTATION PRESENTED TO THE GRADUATE SCHOOL OF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY UNIVERSITY OF FLORIDA 2009
Transcript
Page 1: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

1

ANALYSIS OF NF1 MUTATION MECHANISMS

By

REBECCA L. LODA-HUTCHINSON

A DISSERTATION PRESENTED TO THE GRADUATE SCHOOL

OF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT

OF THE REQUIREMENTS FOR THE DEGREE OF

DOCTOR OF PHILOSOPHY

UNIVERSITY OF FLORIDA

2009

Page 2: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

2

© 2009 Rebecca L. Loda-Hutchinson

Page 3: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

3

To Mrs. Michelle Doyle and Mr. Stephen Sans, whose inspiration and enthusiasm started me on

this path many years ago

Page 4: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

4

ACKNOWLEDGMENTS

First, I need to thank all the families and individuals who participated in these studies;

without them, this work would not have been possible. I want to thank Dr. Peggy Wallace for

taking me on as a student and letting me work in her lab for the past five years. I would not have

made it through this experience without her guidance, understanding, and support. I also want to

thank the members of my supervisory committee, Dr. Daniel Driscoll, Dr. Keith Robertson, and

Dr. Peter Sayeski, for their time, support, and suggestions throughout this process. My thanks

also goes to all the past and present members of the Wallace lab, especially Beth Fisher, for her

patience and smiles as I was learning the ropes; and Michelle Burch, for always having an

answer for me. I also want to thank Dr. Michele Tennant, Dr. Pauline Ng (Fred Hutchinson

Cancer Research Center, Seattle), and Maya Zuhl (University of Maryland Biotechnology

Institute) for their invaluable bioinformatics help.

Finally, I want to thank my family and friends for their support and encouragement no

matter what has come my way these past five years. I especially want to thank Randi Marie,

Deborah, Nicole, and Lauren, who have shared in the graduate school experience with me, for

helping me stay sane, giving me someone to talk to, and reminding me to relax and have fun.

Last, but not least, a huge thanks goes to my husband, Lance, who has seen this process through

with me from applications to a dissertation, and never stopped believing that I could do it.

Page 5: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

5

TABLE OF CONTENTS

page

ACKNOWLEDGMENTS ...............................................................................................................4

LIST OF TABLES...........................................................................................................................7

LIST OF FIGURES .........................................................................................................................8

ABSTRACT.....................................................................................................................................9

CHAPTERS

1 INTRODUCTION ..................................................................................................................11

Neurofibromin and the NF1 Gene ..........................................................................................11

Neurofibromatosis 1 (NF1).....................................................................................................12

Mutations in NF1 ....................................................................................................................15

2 SOMATIC CpG C TO T TRANSITIONS AT NF1 GERMLINE HOTSPOTS....................19

Introduction.............................................................................................................................19

Materials and Methods ...........................................................................................................22

Mutation Detection by PCR and Restriction Digest........................................................22

Methylation Status Analysis............................................................................................23

Results.....................................................................................................................................24

Methylation Status Analysis............................................................................................24

Mutation Detection by PCR and Restriction Digest........................................................26

Discussion...............................................................................................................................26

3 ALTERNATIVE SPLICING OF EXON 23a AND mRNA EDITING .................................32

Introduction.............................................................................................................................32

Material and Methods .............................................................................................................36

Reverse-Transcription and PCR......................................................................................36

Cloning and Sequence Analysis ......................................................................................37

Results.....................................................................................................................................37

Alternative Splicing Patterns ...........................................................................................37

Analysis of RNA Editing.................................................................................................39

Discussion...............................................................................................................................40

4 MISSENSE MUTATION COMPUTATIONAL ANALYSIS OF PATHOGENICITY .......47

Introduction.............................................................................................................................47

Materials and Methods ...........................................................................................................49

Missense Computational Methods ..................................................................................49

Other Databases and Programs........................................................................................49

Page 6: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

6

Data Sets ..........................................................................................................................50

Sequence Input Requirements .........................................................................................50

Splice Analysis ................................................................................................................53

Results.....................................................................................................................................53

Discussion...............................................................................................................................57

5 CONCLUSIONS AND FUTURE DIRECTIONS .................................................................64

Somatic CpG C to T Mutation................................................................................................64

Alternative Splicing of exon23a .............................................................................................66

Computational Analysis Comparison .....................................................................................69

APPENDIX

A CpG C to T Mutation Analysis Data ......................................................................................70

B Exon 23a Alternative Splicing Data .......................................................................................74

C Missense Mutation Computational Analysis Data .................................................................77

LIST OF REFERENCES...............................................................................................................85

BIOGRAPHICAL SKETCH .......................................................................................................100

Page 7: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

7

LIST OF TABLES

Table page

2-1 Primers used in CpG C to T mutation screening. ..............................................................29

3-1 Summary of alternative splicing of exon23a seen in various sample types examined......45

4-1 Summary of computational missense prediction results for 5 data sets (4 controls and

1 unknown). .......................................................................................................................63

A-1 Results of CpG C to T mutation screen using TaqαI restriction enzyme digest................70

B-1 Relative concentrations of Type I v Type II mRNA in blood, tumor and culture

samples...............................................................................................................................74

C-1 Results for each mutation as returned by the various computational methods used. ........78

Page 8: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

8

LIST OF FIGURES

Figure page

1-1. Diagram of neurofibromin’s putative domains..................................................................18

2-1. Methylation-sensitive restriction digest protocol...............................................................29

2-2. Visualization of methylation-specific restriction digest analysis. .....................................30

2-3. Results of bisulfite sequencing. .........................................................................................30

2-4. TaqαI digests of E23.2 PCR products visualized on 8% PAGE........................................31

3-1. Representative gels showing relative concentrations of Type I v Type II mRNA in

various tissue types studied................................................................................................46

3-2. Comparison of alternative splicing of exon23a in primary plexiform tumors v

cultured Schwann cells from the same tumors. .................................................................46

Page 9: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

9

Abstract of Dissertation Presented to the Graduate School

of the University of Florida in Partial Fulfillment of the

Requirements for the Degree of Doctor of Philosophy

ANALYSIS OF NF1 MUTATION MECHANISMS

By

Rebecca L. Loda-Hutchinson

May 2009

Chair: Margaret R. Wallace

Major: Medical Sciences—Genetics

Neurofibromin is a large protein encoded by the NF1 gene, whose best understood

function is as a down-regulator of the RAS signaling pathway, leading to NF1’s classification as

a tumor suppressor gene. NF1 gene mutations, which occur at a rate 10x higher then the genome

average, lead to the autosomal dominant disorder neurofibromatosis 1 (NF1). NF1 has variable

expressivity and is clinically diagnosed using seven diagnostic criteria, of which the key features

are café-au-lait spots, neurofibromas, Lisch nodules, and skin fold freckling. Genetic diagnosis

is difficult, as the gene is very large and very few mutations are recurrent. Additionally, many

NF1 phenotypes are believed to originate with a second mutation in the wild type allele in

specific cell types. To gain insight into mutation mechanisms in NF1, I pursued three projects.

First, I analyzed the rate of somatic C to T mutations at four hotspots for germline mutations.

The methylation status of these sites in somatic cells makes them susceptible to C to T

transitions; however no such mutations were identified in 123 neurofibromas. Next, the

alternative splicing of exon23a, the inclusion of which reduces neurofibromin’s RAS-GAP

function, was examined in various tissue types and tumors. Transcripts containing exon23a

(Type II mRNA) are predominant in most tumors; mRNA lacking exon23a (Type I) is

predominant in blood leukocytes. While previously reported in tumors containing increased

Page 10: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

10

Type II mRNA, no RNA editing was observed in the tumors in this study. Finally, I tested the

accuracy of computational methods at predicting the effects of NF1 missense mutations

(pathogenic versus neutral). These programs are needed clinically since mutations can not be

tested functionally. No program was 100% accurate, but each had advantages in different

situations. This work contributes to the knowledge in NF1, toward a goal of targeted therapies

and improved diagnosis.

Page 11: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

11

CHAPTER 1

INTRODUCTION

Neurofibromin and the NF1 Gene

Neurofibromin is a 220 kD, 2818 amino acid protein expressed in vertebrate and

non-vertebrate animal species, with two homologs in yeast. While it is ubiquitously expressed, it

is found at highest levels in Schwann cells, neurons, and other neural crest-derived cell types

(Daston et al., 1992). Neurofibromin is known to be localized to the cytoplasm and interacts

with microtubules (Gregory et al., 1993), but recent work suggests that it may also travel to the

nucleus (Vandenbroucke et al., 2004) and may localize to different cell compartments based on

stage of development and cell type (Roudebush et al., 1997; Kaufmann et al., 2002; Malminen et

al., 2002; De Schepper et al., 2006). Additionally, four isoforms caused by alternative splicing

have been shown to occur at various developmental stages as well as in specific cell types, which

will be discussed in more detail in Chapter 3 (Nishi et al., 1991; Suzuki et al., 1992; Gutmann et

al., 1993a; reviewed by Skuse and Cappione, 1997). Figure 1-1 shows a diagram of the putative

domains of neurofibromin. While several functional domains have been proposed, including a

microtubule binding domain (Gregory et al., 1993) and a SEC14-like domain (D’Angelo et al.,

2006), the most well-characterized functional domain of neurofibromin is a 360 amino acid

RAS-GTPase activating protein (GAP)-related domain (GRD) (Ballester et al., 1990, Xu et al.,

1990; reviewed by Cichowski and Jacks, 2001). This domain binds activated p21-RAS-GTP and

leads to the hydrolysis of GTP to GDP, inactivating the RAS protein. This inactivation

decreases downstream RAS signaling, and thereby down-regulates cell proliferation and the

inhibition of apoptosis. Since increased RAS signaling is associated with tumorigenesis,

neurofibromin’s inhibitory activity, as well as its gene’s adherence to Knudson’s two-hit

hypothesis (Knudson, 1971), led to its classification as a tumor suppressor (Colman et al., 1995).

Page 12: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

12

It is not surprising then, that mutations in the gene encoding neurofibromin predispose

individuals to developing multiple tumor types. Neurofibromin is encoded by the NF1 gene,

which contains 60 exons that span 280 kb of DNA on chromosome 17q11.2 (Cawthon et al.

1990; Viskochil et al. 1990; Wallace et al. 1990; Li et al. 1995). The GRD is encoded by exons

21 to 27a, and mutations in this region have been studied functionally in yeast. Expressing this

domain alone in yeast lacking the yeast NF1 homologs (Ira1, Ira2) can restore a normal

phenotype (Ballester et al., 1990). While the GRD is the best understood domain, mutations

have been found in all exons/domains of the NF1 gene, which lead to neurofibromatosis 1 (NF1).

Neurofibromatosis 1 (NF1)

Although the reasons are unclear, the mutation rate at the NF1 locus is ten times higher

than the average rate in the human genome, making NF1 one of the most common genetic

disorders (Riccardi, 1992; Hughes, 1994). It is inherited in an autosomal dominant manner, and

occurs in approximately 1 in 3,500 births worldwide (Riccardi, 1992). Occurrence rates do not

vary based on sex or ethnicity. Approximately half of all cases brought to medical attention have

no affected parent and represent a de novo NF1 mutation. Of these de novo mutations, 80% of

those that are not deletions are paternal in origin (Stephens et al., 1992). The physical

manifestations of NF1 are numerous and vary greatly in severity among individuals (variable

expressivity) (Carey and Viskochil, 1999). In addition, even within a family, phenotypic

features and severity may vary, although there is less intra-familial variation than inter-familial

(Easton et al., 1993; Szudek et al., 2002). Because of this, seven diagnostic criteria were

established in 1988 by the National Institutes of Health to aid in clinical diagnosis (reviewed by

Gutmann et al., 1997). These seven diagnostic criteria are: six or more café-au-lait macules

(must be 0.5 cm or larger before puberty, 1.5 cm or larger after puberty), neurofibromas (benign

Schwann cell tumors), optic pathway tumors (benign pilocytic astrocytomas), two or more Lisch

Page 13: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

13

nodules (benign hyperpigmented lesions of the iris that look like freckles), skeletal dysplasia

(typically sphenoid bone dysplasia or tibial dysplasia), skin fold freckling, and having a first

degree relative with NF1. A patient is clinically diagnosed with NF1 if they meet any two or

more of these criteria. Although molecular genetic testing is available, these criteria remain

important in diagnosis as many patients do not have the expensive genetic analysis done

(McClatchey, 2007). NF1 is a progressive disorder, and these phenotypic features may arise or

increase unpredictably over time. Café-au-lait macules are often the earliest phenotype to

develop, usually present by 2 years of age. Tibial dysplasia and skin fold freckling develop in

the first few years of life as well, with the majority of other phenotypes arising by late

childhood/early adolescence (Riccardi, 1992; Williams et al., 2009). In addition to these criteria,

patients with NF1 are also at increased risk for additional complications, including learning

disabilities, short stature, scoliosis, renal artery stenosis, hypertension, macrocephaly, and

increased risk of certain malignancies, such as rhabdomyosarcoma, pheochromocytoma, and

myeloid leukemia (reviewed by Gutmann et al., 1997).

Of the main phenotypic characteristics, neurofibromas generally cause the most trouble

and discomfort to the patients. Discomfort may include itching, pain, and tenderness around

areas of tumor growth, and furthermore patients can suffer socially due to disfigurement

(Riccardi, 1992; Gottfried et al., 2006). Neurofibromas originate from the peripheral nerve

Schwann cells and are clonal in origin (Skuse et al., 1991; Colman et al., 1995). Schwann cells

are one of several cell types that make up the peripheral nerve sheath, which surrounds the nerve

and axons and helps maintain proper nerve function. The healthy peripheral nerve sheath is well

organized, with Schwann cells closely associated with the axon of the nerve, producing the

myelin used to insulate the axon. Disorganization and an increase in all peripheral nerve sheath

Page 14: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

14

cell types are seen in neurofibromas (reviewed by Cichowski and Jacks, 2001). Thus, these

tumors contain primarily clonal, expanded (and somatically mutated, as described below)

Schwann cells, but also contain other cell types normally associated with peripheral nerves, such

as non-clonal Schwann cells, fibroblasts, mast cells, and axons (Serra et al., 2000). The

cutaneous, or dermal type neurofibromas, are derived from Schwann cells associated with nerve

twigs, close to or at the surface of the skin. Some patients develop thousands of these tumors

while others have few or none. They rarely grow beyond one centimeter in diameter and are not

known to become malignant. In contrast, plexiform neurofibromas develop from Schwann cells

associated with larger nerves, are generally deeper in the body, are larger then dermal

neurofibromas, and may involve large areas of the body. It is estimated that 30-40% of NF1

patients develop plexiform tumors, some of which may be asymptomatic nodular masses in the

thorax (Tonsgard et al., 1998). In addition to the risk that a plexiform tumor could grow quite

large and disrupt bone and organs, 10-15% of plexiform neurofibromas will transform into

malignant peripheral nerve sheath tumors (MPNSTs), presumably by the accumulation of genetic

and epigenetic changes (Ferner and Gutmann, 2002). Unfortunately, MPNSTs often become

metastatic and have a poor prognosis unless removed completely by surgery.

How the transformation from benign plexiform neurofibroma to MPNST occurs is unclear.

The development of benign neurofibromas is somewhat better understood and the initiating event

appears to follow Knudson’s two-hit hypothesis of tumor suppressors (1971). Because of the

dominant nature of NF1, the majority of patients are constitutionally heterozygous for an

inactivating mutation in the NF1 gene. In a minority of patients the disease-causing mutation

arose at a point in development shortly after the single cell stage, leading to mosaic distribution

of the mutation (Colman et al. 1996). Regardless, neurofibromas arise when Schwann cells

Page 15: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

15

carrying this mutated NF1 allele develop a second mutation that reduces the function of the

previously wild-type NF1 allele (Serra et al., 1997, 2000; reviewed in MacCollin and Wallace,

2002). When a second, somatic mutation or other inactivating event interferes with the function

of the wild-type NF1 allele in a Schwann cell, neurofibroma development is thought to start or at

least be poised to begin upon additional signals. This somatically-mutated Schwann cell then

undergoes clonal expansion, dividing inappropriately, even without axonal contact. Similarly,

optic pathway tumors initiate from two NF1 mutations in an astrocyte (Bajenaru et al., 2003),

tibial dysplasia is associated with two mutations in osteoblasts (Stevenson et al., 2006), and

café-au-lait macules originate from a melanocyte with two NF1 mutations (DeSchepper et al.,

2007).

Mutations in NF1

The NF1 germline mutation rate is estimated between 1/7,800 and 1/23,000, approximately

10-fold higher than average (reviewed in Gottfried et al. 2006). It is unclear why the NF1 locus

is so susceptible to mutation; the size of the gene alone does not explain this phenomenon

(Friedman 1999; Fahsold et al; 2000). The NF1 germline mutation spectrum is broad, with over

1000 germline mutations identified. Of these, 70-80% are clearly inactivating (frameshift and

nonsense mutations), and none recur in more than 2% of cases (reviewed in Thomson et al.

2002). Approximately 20% are aberrant splicing mutations (Messiaen et al., 2000; Serra et al.,

2001; Wimmer et al., 2007; Pros et al., 2008). The exception to this is a 1.2-1.4 Mb

microdeletion of the region of chromosome 17 that contains the NF1 gene and 14 flanking genes.

This microdeletion was first identified in patients with an early age of tumor onset, high tumor

loads, specific facial features, and mental retardation (Kayes et al., 1992, 1994; Leppig et al.,

1996, 1997; Wu et al., 1995, 1997). Later studies, however, revealed patients with no

outstanding phenotype who also carry microdeletions in the NF1 region (Rasmussen et al., 1998;

Page 16: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

16

Dorschner et al., 2000). This microdeletion is thought to account for 4-7% of NF1 cases, and

approximately 80% of de novo microdeletions are found on the maternally inherited

chromosome 17 (Lázaro et al., 1996; Valero et al., 1997; Rasmussen et al., 1998; Upadhyaya et

al., 1998; Lopez-Correa et al., 2001; Kluwe et al., 2004). These deletions are mediated by

unequal crossover meiotic events in the germline at mis-aligned repetitive regions. Both the 1.2

and 1.4 Mb versions of the microdeletion share a 5’ repeat sequence containing a pseudogene of

JJAZ1, called JAZFP (Forbes et al., 2004). The 1.2 Mb deletion’s 3’ end is in the JJAZ1 gene,

whereas the 1.4 Mb deletion’s breakpoint is distal to JJAZ1 (Raedt et al., 2006). In addition to

the NF1 gene being deleted, 14 other flanking genes are deleted as well, leaving patients

hemizygous for these genes. These deletions can also occur somatically, early in embryogenesis

(mitotic recombination), resulting in a patient who is mosaic for an NF1 microdeletion

(Rasmussen et al., 1998; Kehrer-Sawatzki et al., 2004) and who may show fewer NF1 features.

A much smaller body of knowledge exists for somatic NF1 mutations. Loss of

heterozygosity was first seen in neurofibromas by Colman et al. (1995). Since then, it has

become clear that in most cases the allelic imbalance is due to mitotic recombination, where a

region of the mutant chromosome replaces that of the wild-type chromosome (Serra et al.,

2001a). Loss of heterozygosity has been found in 10-40% of tumors analyzed (Rasmussen et al.,

2000; Serra et al., 2001; Upadhyaya et al., 2004). Sawada et al. (1996) were the first to identify

a specific somatic mutation in the NF1 gene in a tumor from a patient whose germline mutation

was already known. Since then reports of additional somatic mutations have been limited, and

there is still little known about the somatic mutation spectrum (Rasmussen and Wallace, 1998;

Upadhyaya and Cooper, 1998; Eisenbarth et al., 2000; Wiest et al., 2003; Upadhyaya et al.,

2004). The difficulty of identifying these mutations is due not only to the size and complexity of

Page 17: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

17

the NF1 gene, but also to the fact that each tumor is thought to arise from an independent

inactivating event, so that somatic mutations differ not only between individuals, but also

between tumors from the same individual (Colman et al., 1995; Serra et al., 2001b).

Furthermore, the tumors have an admixture of cells and thus only a portion of the cells (the

clonally-expanded Schwann cells) carry the somatic mutation. Further understanding of the

nature and frequency of somatic mutations will provide important information about risk factors,

disease progression and tumorigenesis, as well as help elucidate the pathways that lead to

neurofibroma formation and the transformation to MPNSTs.

This work adds to this body of knowledge by addressing three specific areas of

investigation directly related to the somatic mutation spectrum of NF1. Chapter 2 discusses the

somatic rate of C to T transitions at four sites known to be hotspots for this mutation type in the

germline. Chapter 3 examines alternative splicing of exon23a in NF1-related tumors, its

expression level in various cell types, and its relationship to RNA editing. Chapter 4 addresses

the need for reliable ways to predict the effects of missense mutations (both somatic and

germline) on neurofibromin function, and evaluates several computational methods that make

such predictions.

Page 18: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

18

Figure 1-1. Diagram of neurofibromin’s putative domains. Alternatively spliced exons are

shown, along with the number of amino acids they insert. Well-characterized

domains, with known crystal structure, are shown in red. GRD = GAP-related

domain (exons21-26); Sec14 = SEC14 protein-like domain (exons27b-28). Domains

shown in blue are not as well-characterized. TM = Transmembrane domain

(exon10a2); CSRD = Cystein/serine-rich domain (amino acids 543-909);

TBD = Tubulin-binding domain (amino acids 1095-1194); SB = Syndecan-binding

domains (amino acids 1356-1469 (with in the GRD), 2619-2715);

LRD = Leucine-rich domain (Sec14 domain is within the LRD) (amino acids

1530-1950); NLS = Nuclear localization sequence (exon43).

Page 19: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

19

CHAPTER 2

SOMATIC CpG C TO T TRANSITIONS AT NF1 GERMLINE HOTSPOTS

Introduction

In mammalian DNA, a cytosine immediately 5’ to a guanine, designated CpG, can be

methylated, having a methyl group attached to carbon 5, producing 5-methylcytosine (5mC).

This is the only nucleotide that can be methylated in mammalian DNA (Bird, 2002). Cytosines

not in the context of CpG are not subject to methylation, with the exception of those found in a

CpNpG context (Clark et al., 1995), and it is believed that the majority of CpG dinucleotides not

in CpG islands are methylated (Cooper and Krawczak, 1993). 5-methylcytosine can be

associated with transcriptionally inactive regions, especially if such methylation occurs in a gene

promoter region where CpGs can be clustered (termed a CpG island) (Bird, 1986) or in other

transcriptional regulatory elements (Cooper and Krawczak, 1993). Cytosine residues in CpG

dinucleotides are susceptible to the spontaneous loss of the amine group at carbon 4 by

hydrophilic attack or by chemical deamination. Once the amine group is lost, a tautomeric shift

can occur. In the unmethylated state this shift produces uracil, while 5mC becomes thymine (a C

to T transition). There are mismatch repair enzymes to recognize and repair both G:U and G:T

mispairs, but G:U mispairing is more efficiently repaired, in part because uracil is not used in the

production of DNA (Cooper and Krawczak, 1993; Lari et al., 2002, 2004). Brown and Jiricny

(1987) found that G:T mispairs (resulting from C to T transitions) were repaired in favor of

guanine 90% of the time. These mispairs were not repaired 2% of the time, and in the remaining

8% of cases, they were corrected in favor of thymine, causing the C to T transitions to become

fixed in the cell’s DNA. These base changes can be neutral (e.g. in non-coding DNA, or a silent

coding-region substitution), or they can cause errors in splicing or coding regions leading to a

premature stop codon, amino acid deletion, or amino acid substitution. No neutral CpG C to T

Page 20: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

20

transitions have been reported in the NF1 gene despite studies that would have detected such

changes in the coding regions and UTRs; no such polymorphisms have been found in the 11 Kb

mRNA. When considering 880 base changes previously reported to be involved in genetic

disorders, Cooper and Krawczak (1993) found that 32.8% were C to T or G to A (resulting from

a C to T transition on the antisense strand) mutations at CpG dinucleotides. One such mutation

has been repeatedly identified as a germline mutation in unrelated neurofibromatosis 1 (NF1)

patients, accounting for 1-2% of mutant alleles. This CpG C to T mutation, R1947X (CGA to

TGA), results in a stop codon in exon31 of the NF1 gene (Horiuchi et al., 1994; Lazaro et al.,

1995; Dublin et al., 1995). There are several other such nonsense germline mutations due to

CpG C to T transitions, each of which has a frequency of 0.5-2%: R416X, R440X, R816X,

R1241X, R1276X, R1362X, R1748X, and R2429X. C to T transitions account for 18-30% of

NF1 germline mutations, most causing nonsense codons, the rest missense mutations (Krkljus et

al., 1998; Fashold et al., 2000; Messiaen et al., 2000). Of these, the majority that were de novo

arose in the paternal genome (Jadayel et al. 1990; Wallace et al. 1991; Krkljus et al. 1998),

which may be related to methylation during spermatogenesis (Driscoll and Migeon, 1990). In

contrast, point mutations at CpG sites account for approximately 50% of the reported germline

mutations in TP53 (Greenblatt et al., 1994), ~30% reported for RB1 (Lohmann et al., 1996), and

54% of germline mutations in NF2 (Baser, 2006). These mutations may be under-represented in

NF1 relative to some other genetic disorders.

In neurofibromatosis 2 (NF2), caused by mutations in the NF2 gene, CpG C to T mutations

are reported to account for 38-52% of somatic mutations (Baser 2006). While there is a strong

body of knowledge on the germline frequency of these mutations in NF1, there are no studies

focused on the rate at which they occur somatically. There are multiple mutation screens that

Page 21: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

21

have identified somatic NF1 CpG mutations (Sawada et al., 1996; Eisenbarth et al., 2000; Serra

et al., 2001b; Wiest et al., 2003; Upadhyaya et al., 2004, 2008; Maertens et al., 2006; Bottillo et

al., 2009). In many of these studies, mutations were not identified in more then half the samples.

Only 142 mutations of any type were identified out of 372 tumors. These studies screened for

mutations in germline C to T transition hotspots, such as the previously mentioned R1947X

mutation. Of the NF1 somatic mutations identified, 7/142 (4.9%) were CpG mutations. Of the 7

CpG somatic mutations found, 5 were at a previously identified germline mutation site

(Eisenbarth et al., 2000; Wiest et al., 2003; Upadhyaya et al., 2004, 2008; Bottillo et al., 2009).

While NF1 contains more CpG containing codons, and specifically CGA codons (which create a

premature stop codon with a C to T transition) than NF2, it does not appear to be as susceptible

to CpG C to T mutations.

Further identification of somatic mutation mechanisms is important since somatic

inactivation of the wild-type NF1 allele in NF1 patients is the initiating step in neurofibroma

tumorigenesis in Schwann cells. It is also possible that sporadic neurofibromas are the result of

two somatic NF1 mutations in a single cell; one such case has been reported, involving

chromosome translocation (Storlazzi et al., 2005). The goal of my work was to test whether

somatic C to T transitions at CpG dinucleotides in the NF1 gene may be a common mechanism

of generating a second hit in Schwann cells. Furthermore, tumors that have any DNA

hypermethylation (which may be present at low levels in some plexiform neurofibromas, based

on promoter studies by Fishbein et al. (2005)) could be at risk for an increase in CpG transitions

since more cytosines are methylated. Further knowledge of the types and frequencies of somatic

mutations such as C to T transitions will be useful for understanding genetic changes

Page 22: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

22

contributing to NF1 tumors. In this work, I will test for somatic presence of each of four

recurrent germline NF1 C to T transition stop mutations in tumor DNA.

Materials and Methods

Mutation Detection by PCR and Restriction Digest

One hundred ninety-seven DNAs were previously prepared by phenol/chloroform

extraction from NF1 patient blood and tumor samples as described in Colman et al. (1993) and

kept in concentrated stocks at 4°C. PCR primers were designed previously for exon10a, 22,

23.2, and 41 of the NF1 gene. These exons contain the mutations R440X, R1241X, R1362X,

and R2429X, respectively. The primers are in the introns flanking each exon, and details are

given in Table 2-1. Diluted DNA samples were used as template (50-100 ng) in polymerase

chain reactions (PCR) under standard conditions. These exon PCR products were then digested

using the TaqαI restriction enzyme in 25 µl reactions incubated at 65°C for 2 hours, with more

enzyme added after 1 hr. The sequence recognized and cut by TaqαI spans the CpG site of

interest in each exon, and any changes in the sequence will result in failure of the enzyme to cut.

These digest reactions were visualized using ethidium bromide staining after electrophoretic

separation on 8% native polyacrylamide gels. With this assay, germline heterozygotes for one of

these mutations show an uncut band plus two smaller bands in the gel. The presence of an uncut

fragment was suggestive of the presence of a mutation and a second digest was used to confirm

the presence of uncut DNA, and controls were used to ensure complete digestion. Table 2-1 lists

the size of both the uncut fragments and the cut fragments produced by TaqαI and methyl-

sensitive digestion for each exon. Numbers of mutation-positive and mutation-negative samples

at each site were determined, as well as total mutations per tumor type.

Page 23: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

23

Methylation Status Analysis

In addition to screening for the four mutations, I also investigated whether those cytosines

were methylated in normal somatic Schwann cells, to ensure that they could be a target for 5meC

deamination. This was first done using methyl-sensitive restriction digest, and then bisulfite

genomic sequencing. A schematic of the methyl-sensitive restriction digest method is given in

Figure 2-1. Genomic DNA was used in restriction digest reactions using methylation-sensitive

restriction enzymes (BstBI for sites in exon10a, 23.2 and 41, incubated at 65°C; AvaI for

exon22, incubated at 37°C). These enzymes recognize and cut the sequence spanning the CpG

sites of interest, but cut only if the cytosine residues are unmethylated. The sequence

surrounding the CpG site of interest in exon10a is 5’GGTTGAACTTCGAAATATGTTT 3’; in

exon41 the sequence is 5’TGAAGAAGTTCGAAGTCGCTGC 3’; in exon23.2 the sequence is

5’CCCTCAACTTCGAAGTGTGTGC 3’. These three sites are cut by BstBI, which recognizes

and cuts the sequence 5’TT/CGAA 3’. The sequence surrounding the CpG site of interest in

exon22 is 5’TGAACTAGCTCGAGTTCTGGTT 3. This site contains the recognition sequence

of AvaI, which is 5’C/YCGRG 3’, where Y is either T or C, and R is either A or G. PCR

reactions were then set up using DNA digested with these enzymes as well as untreated DNA.

The primers listed in Table 2-1 were used in these reactions. This same methylation analysis

was carried out on control samples as well as samples from a plexiform neurofibroma. As a

control, PCR samples were set up using the same samples, and these reactions were digested

with TaqαI restriction enzyme as described above. The TaqαI site (TC/GA) is within the

recognition sites of the methyl-sensitive enzymes, and is cut by TaqαI whether the C of interest

is methylated or not. The presence of digested product in these reactions will confirm that the

recognition sites are not mutated in the samples tested, and are therefore able to be digested by

Page 24: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

24

the methyl-sensitive enzymes. All digestion products were visualized using ethidium bromide

staining after electrophoretic separation on 8% native polyacrylamide gels.

The CpG site in exon23.2 (R1362X) was chosen for analysis by bisulfite sequencing as an

independent measure of somatic methylation. Genomic DNA from cultured normal human

Schwann cells was subjected to sodium bisulfite treatment per the method given in Fishbein et al.

(2005). The treated DNA was then PCR amplified using primers 23-2MEF

(5’GTTAGAATTATTAGAGAGTTTTGAG 3’) and 23-2MER

(5’ATAATCTCTAACTATAAACATACCTAATA 3’) with an annealing temperature of 54°C.

The sequence of these primers was determined assuming the conversion of unmethylated

cytosine to thymine after sodium bisulfite treatment. These primers amplify a 158 bp fragment

spanning exon23-2 and 23 bp of intron23-2. The PCR products were ligated into a plasmid

vector (Topo TA, Invitrogen) and transformed into chemically competent E. coli cells (One Shot

TOP10, Invitrogen) using the manufacturer’s protocol. The cells were plated on LB-ampicillin

agar plates with IPTG and X-gal, and incubated at 37°C overnight. Bacterial colonies (clones)

positive for the insert were identified by blue/white selection, followed by PCR amplification of

picked colonies using the original primers. The PCR products from five of these clones were

sequenced, using cycle sequencing with the ABI Prism Big Dye 3 system and the UF CEG

sequencing core, with the 23-2MEF primer as the sequencing primer.

Results

Methylation Status Analysis

In one set of experiments, the methylation status of all four CpG C to T germline mutation

hotspots was analyzed using methylation-sensitive restriction digest analysis. I examined these

sites in a normal Schwann cell culture, leukocytes from two non-NF1 patients, and one plexiform

neurofibroma. As mentioned above, this analysis used restriction enzymes that will not cut in

Page 25: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

25

the presence of a methylated cytosine. If a cytosine of interest is unmethylated, these enzymes

will cut the genomic DNA, and there will be no PCR product. If methylation is not complete at a

site, there will be uncut band visible after separation by electorphoresis on a PAGE gel, but the

concentration should appear less. As a control, TaqαI digest reactions were also set up, to ensure

that the sites of interest in these samples did not contain any mutations that would prevent either

enzyme from cutting, regardless of methylation status. Representative gels from these

experiments can be seen in Figure 2-2. All four CpG sites are free of mutation in these samples,

as they can be digested by TaqαI (left-most lane of three for each sample). Based on the

presence of roughly equivalent amounts of uncut product in the methylation-sensitive reactions

(right-most lane of three for each sample) when compared to the uncut control reaction (middle

lane for each sample), it appears that the cytosine at each of the four CpG sites is methylated in

all samples. The equivalent amount of product in the uncut control and methyl-sensitive

reactions suggests that these sites are completely methylated, to the level of detection possible

with this method.

To further validate the extent of methylation at these four sites of interest, I chose the CpG

site in exon23.2 for analysis by bisulfite sequencing. In DNA treated with sodium bisulfite,

unmethylated cytosines are converted to uracils by deamination; uracil is not normally present in

DNA (rather it is seen in place of thymine in RNA) and during PCR amplification, thymine takes

the place of the uracil, leading to a C to T change at the unmethylated sites. The cloning of a

sodium bisulfite treated PCR product into a vector allowed for the isolation of a single allele per

bacterial colony. Sequencing was done to reveal the relative number of cytosines versus

thymines at the CpG site of interest, allowing an estimate of the percent methylation at that site.

Figure 2-3 shows the untreated sequence of exon23.2, as well as the chromatogram from the

Page 26: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

26

sequencing of one of the clones. The sequence represented in the chromatogram is underlined in

the untreated sequence for comparison. All 5 clones analyzed contained a cytosine at the CpG

dinucleotide of interest (indicating that it is methylated), while all other cytosines in the fragment

had been replaced by thymine. The agreement between the methylation specific restriction

digest results and the bisulfite sequencing results for the exon23.2 site of interest suggests that

the four germline mutation hotspots are all heavily methylated in somatic tissue. These sites are

thus susceptible to CpG C to T transitions.

Mutation Detection by PCR and Restriction Digest

Somatic mutation detection analysis was carried out on DNA from 63 dermal

neurofibromas and 83 plexiform neurofibromas (Table A-1). Each sample was PCR amplified

with each of the four primer sets listed in Table 2-1 and then subjected to restriction digest by

TaqαI. All 4 CpG sites were analyzed in 117 samples, and 1 to 3 of the CpG sites was analyzed

in the remaining 29 (due to lack of PCR amplification by on or more primer pair). Across all the

samples, 531 CpG dinucleotides were analyzed for C to T transitions. Figure 2-4 shows an

example of the visualization of the digest products on an 8% PAGE, with uncut PCR product

indicating presence of a mutation (normal sequence is digested into two smaller products).

While this analysis was used to screen for somatic CpG C to T mutations, it also revealed two

previously-identified constitutional mutations, one at the exon10a CpG site and the other at the

exon23.2 (Figure 2-2) CpG site. These were the only mutations, germline or somatic, seen at

any of the four CpG sites in the 146 neurofibromas, from 82 independent patients.

Discussion

The methylation of cytosine in a CpG dinucleotide is a common occurrence in the human

genome. It is estimated that 70-80% of all CpG dinucleotides contain 5- methylcytosine (5mC)

(Razin and Riggs, 1980). The percent of isolated CpG sites that are methylated is likely higher,

Page 27: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

27

as this statistic includes all CpG dinucleotides, both isolated and those found in CpG islands, the

latter of which account for the majority of non-methylated CpGs (reviewed by Bird, 2002).

Based on these observations, it was expected that the four CpG sites studied here would be

methylated in normal tissue. The results of my methylation status analysis confirm that all four

of these sites are indeed predominantly, if not completely, methylated in normal Schwann cells

and neurofibromas.

The positive methylation status of these CpG sites makes them susceptible to C to T

transitions due to the spontaneous deamination of 5meC. Since C to T transitions can occur

without need for replication, this mutation mechanism is feasible in Schwann cells, which are

typically quiescent unless stimulated to divide by injury, or occasional divisions to keep up with

nerve growth. Interestingly, no such mutations were identified in 146 tumors analyzed at the

CpG germline mutation hotspots.

All four of these CpG sites have been previously identified as germline mutation hot spots

with a 1-2% recurrence rate (arginine-to-stop), and the estimated combined germline mutation

rate at these four sites is approximately 7% (they account for 7% of all NF1 germline mutations)

based on our lab’s data and published comprehensive studies from other labs. Thus, if the

somatic mutation rate was equivalent, I would have expected to see approximately 37 total

mutations (7% of 531 sites analyzed) among the four sites in the 146 tumors. Since no somatic

mutations were identified, it appears that the rate of somatic CpG C to T transitions is very low

in neurofibromas, (at least at these sites), despite the presence of a methylated cytosine. The

detection of the two known germline mutations by my methods indicated that this apparent

dearth of somatic CpG C to T mutations is not due to a faulty detection method. The level of

mutation seen in my study is consistent with the low level reported in the NF1 mutation

Page 28: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

28

literature. This also suggests that C to T transitions do not play a major role in somatic events in

NF1 tumorigenesis. This is in contrast to the rate of CpG C to T mutations in other, non-NF1

tumors. The rate of somatic CpG C to T mutations in TP53 ranges from 25% in bladder cancer

to nearly 50% in colon cancer (Jones et al., 1991; Greenblatt et al., 1994; Olivier et al., 2002). In

a study of hereditary non-polyposis colorectal cancer, 30.7% (4/13) of somatic APC mutations

were CpG C to T mutations (Huang et al., 2004). The differences in the rate of these mutations

in neurofibromas compared to other tumor types may be due to differences in their ability to

repair this type of mutation. As many cancers exhibit mutations in DNA repair pathways, it may

also be that more malignant tumors are more susceptible to C to T transition mutations, (whereas

neurofibromas are benign) due to reduced ability to repair the mismatches. The status of base

excision repair in neurofibromas has not been analyzed, although it is known that cytogenetic

rearrangements virtually never occur in dermal neurofibromas and less then half the time in

plexiforms, so these tumors tend to have fairly good chromosomal stability (Wallace et al.,

2000). However, NF2 schwannomas are also benign Schwann cell tumors that do not show

chromosomal rearrangements, yet somatic CpG C to T mutations are very frequent. This may be

pointing to basic differences in these tumor types for frequency of 5meC deamination and/or

robustness of excision repair.

Understanding the frequencies and mechanisms of CpG somatic mutation may help predict

whether certain individuals or tumors are more at risk for these, or lead to a specific therapy for

tumors containing such mutations. For example, there are several therapies that have shown

potential in allowing translations through premature stop codons, including gentamicin and

related compounds, and antisense oligonucleotides (reviewed by Ainsworth, 2005; Kulyte et al.,

2005; Pinotti et al., 2006).

Page 29: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

29

Genomic DNA

Digest with methylation-sensitive enzymes

(BstBI for e10a, e41; AvaI for e22, e23.2)

PCR with flanking primers PCR with flanking primers

Digest with TaqαI

Polyacrylamide gel Polyacrylamide gel

Unmethylated C Methylated C not Me-sensitive

No band Uncut band Cut band

Table 2-1. Primers used in CpG C to T mutation screening.

Exon Primer sequence (5’-3’)

Product

size (bp)

Annealing

temp (°C)

TaqαI digest

band sizes (bp)

5’ ACGTAATTTTGTACTTTTTCTTCC 10a

3’ CAATAGAAAGGAGGTGAGATTC 222 60

105

117

5’ TGCTACTCTTTAGCTTCCTAC 22

3’ CCTTAAAAGAAGACAATCAGCC 331 62

87

244

5’ TTTTAAGGAGTGATTTTTGTTATTTG 23.2

3’ CCTTCTTTGATAAAGCATTCTTC 276 55

179

97

5’ TTCATCCTGTTTTAAGTCACACTTG 41

3’ TTGCCTCCATTAGTTGGAAAATTG 273 60

94

179

Figure 2-1. Methylation-sensitive restriction digest protocol.

Page 30: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

30

Figure 2-2. Visualization of methylation-specific restriction digest analysis. a: normal human

Schwann cell culture; b and c: non-NF1 patient blood; d: plexiform neurofibroma

DNA.

GTTAGAACCATCAGAGAGCCTTGAGGAAAACCAGCGGAACCTCCTTCAGATG

ACTGAAAAGTTCTTCCATGCCATCATCAGTTCCTCCTCAGAATTCCCCCCTCA

ACTTCGAAGTGTGTGCCACTGTTTATACCAGTTTATACCAGGTATGCTTACAG

TTAGAGATTAC

Figure 2-3. Results of bisulfite sequencing. The sequence given is the untreated sequence of

exon23.2. Upon treatment with bisulfite followed by sequencing, unmethylated Cs

are converted to Ts, methylated Cs are not converted. All Cs are in blue, the CpG site

of interest is in red. The chromatogram shows the underlined region of sequence

after bisulfite treatment. The only C not converted to T is in the CpG dinuleotide,

indicating it is methylated.

E10a E22

m a b c d m m a b c d m

TaqI + - - + - - + - - + - - + - - + - - + - - + - -

BstBI - - + - - + - - + - - + AvaI - - + - - + - - + - - +

m a b c d m m a b c d m

TaqI + - - + - - + - - + - - + - - + - - + - - + - -

BstBI - - + - - + - - + - - + BstBI - - + - - + - - + - - +

E23.2 E41

Page 31: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

31

U M

Figure 2-4. TaqαI digests of E23.2 PCR products visualized on 8% PAGE. All samples are from

plexiform neurofibromas. The sample indicated by the asterisk is from a patient with

a known constitutional CpG mutation. C= control with known mutation at cut site;

U= uncut sample; M=1 kb marker.

*

*

Page 32: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

32

CHAPTER 3

ALTERNATIVE SPLICING OF EXON 23a AND mRNA EDITING

Introduction

The alternative splicing of exons allows for the production of multiple different transcripts

from a single gene. This usually affects the coding region in between invariant translation start

and stop sites, but this phenomenon can also produce transcripts with different transcription or

translation start and/or end sites. Alternative splicing is often regulated in a spatial,

tissue-specific, and/or temporal manner, and its effects can generally be divided into five

categories: effects on localization of the resulting protein, elimination of the resulting protein’s

function, changes in the resulting protein’s function, creation of a new function of the protein,

and effects at the RNA level, such as transcript stability or efficiency of translation (reviewed by

Smith et al., 1989). While there are many examples of alternative splicing events, there is also

evidence that disregulated alternative splicing plays a role in human cancers (Early et al., 1980;

Nagoshi et al., 1988; Lee and Feinberg, 1997; Stimpfl et al., 2002; Adams et al., 2002). One

array study identified 845 alternative splicing isoforms from throughout the genome that appear

to be tumor associated (Wang et al., 2003). While no NF1 isoforms were included in this data

set (which also did not include Schwann cell tumors), it is known that several common NF1

isoforms normally exist at at least a 10% level relative to the major isoform in the pertinent

tissue. There are three alternatively spliced exons in NF1, as well as a short isoform that lacks

exon11 through most of exon49 (Nishi et al., 1991; Suzuki et al., 1992; Gutmann et al., 1993a;

reviewed by Skuse and Cappione, 1997). The first of these alternatively spliced exons to be

identified was exon23a (Nishi et al., 1991). The inclusion of this exon produces an mRNA

containing 63 additional base pairs from intron23-2, leading to an in-frame insertion of 21 amino

acids in neurofibromin, and is known as the Type II isoform. These additional amino acids are

Page 33: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

33

within the NF1 GAP-related domain (GRD), and have been shown to change the hydrophobicity

and secondary structure of this region of the protein (Nishi and Saya, 1991). While this isoform

is still able to bind RAS and has GAP activity, Andersen et al. (1993) found that cells expressing

the Type II isoform contained much more RAS-GTP (activated RAS) than cells expressing

mostly the Type I isoform (lacking exon23a). This was shown to be due to ten-fold decreased

GAP activity of the protein encoded by NF1 Type II mRNA. It has also been shown that Type II

neurofibromin does not associate with microtubules as Type I does (Gutmann et al., 1995). It

has been suggested that in addition to altering or eliminating known functions of neurofibromin,

the insertion of exon23a may introduce novel functions to the protein. Teinturier et al. (1992)

found that exon23a showed sequence homology with vaccinia virus nucleoside triphosphotase I,

and Andersen et al. (1993) point out similarities between the 21 amino acid insert and common

nuclear localization signals.

While the functional effects caused by the inclusion of exon23a have yet to be fully

characterized, Costa et al (2001) found that exon23a knockout mice had an increased incidence

of cognitive deficits, relative to the exon31 knockout mouse (which has no cognitive

impairment). It also appears that this isoforms is developmentally significant, as there is a

switch from the predominant expression of Type I to Type II neurofibromin through

embryogenesis into post-natal life in the majority of tissues in rat, mouse and chick, as well as

differentiating human cell types (Nishi et al., 1991; Baizer et al., 1993; Gutmann et al., 1994,

1995; Huynh et al., 1994; Mantani et al., 1994). Although there are some conflicting reports, the

consensus is that the Type II transcript is typically the predominant transcript, present at levels

greater than or equal to Type I in most adult tissues, with the exception of the adult central

nervous system (Suzuki et al., 1992; Uchida et al., 1992; Teinturier et al., 1992; Baizer et al.,

Page 34: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

34

1993). Two of these studies also indicated that in various non-neurofibromatosis 1 (NF1)

associated cancers the expression of Type II mRNA is preferentially increased (Suzuki et al.,

1992; Teinturier et al., 1992). Ogata et al. (2001) found that a particularly malignant breast

cancer cell line (MDA-MB-231) does not express any type I NF1 mRNA, further supporting a

role of aberrant alternative splicing of exon23a in tumors.

There are two other well-documented NF1 isoforms. One is expressed exclusively in the

central nervous system. This isoform includes exon9br, which encodes 10 additional amino

acids inserted at residue 420 of the protein (Danglot et al., 1995; Geist and Gutmann, 1996). The

other includes exon48a, which is a muscle-specific 18 amino acid insertion near the C-terminal

end of the neurofibromin molecule (Gutmann et al., 1993a). These were not included in this

tumor study since they do not affect RAS-GAP activity, and do not appear to be involved in NF1

tumorigenesis.

In addition to these well-studied isoforms, there have been several reports of multiple rare

novel splice variants of NF1 mRNA (Thomson and Wallace, 2002; Vandenbroucke et al., 2002a,

2002b). While some of these alternative splice events insert or delete intronic or exonic

sequence, many of them involve exon skipping due to splicing at the same sites used in normal

NF1 RNA (Thomson and Wallace, 2002; Vandenbroucke et al., 2002b). Some of these produce

out-of-frame transcripts. Thomson and Wallace (2002) found that the conditions under which

blood samples were drawn, or length of time stored before RNA was isolated, impacts this rare

variant splicing profile of NF1, with the frequency of these increasing over time. However,

relative ratios of the Type I to Type II mRNA are not affected. Some of these novel isoforms

have been shown to occur in a tissue specific manner, implying that they may be functionally

significant despite relatively low levels (Vandenbroucke et al., 2002a, 2002b). It is clear that

Page 35: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

35

these isoforms do exist in vivo, but further study is needed to determine the role of these rare

NF1 isoforms in normal tissues and in NF1 phenotype development.

Another form of post-transcriptional modification that has been shown to occur in NF1

mRNA is RNA editing. Skuse et al. (1996) found that some NF1 mRNA undergoes base

modification editing at position 3916. An in-frame stop codon (R1306X) is created by the

deamination of the transcript’s cytosine at bp 3916 of the mRNA to form a uracil. This

nucleotide is within the region encoding the neurofibromin GRD. It is unclear what effect this

premature stop codon may have on the cells that express the edited mRNA, but if translated it

would produce a truncated protein that would likely be degraded or lack full function.

Additionally, editing can occur in mRNA transcribed from either allele, potentially inactivating

that allele regardless of NF1 gene mutation state. While this editing has been found at low levels

(1.5-2%) in all cell types studied, there has been evidence that NF1 mRNA editing occurs at

somewhat higher levels in tumor cells (Skuse et al., 1996; Cappione et al., 1997; Mukhopadhyay

et al., 2002). Cappione et al. (1997) observed a correlation between increased invasiveness of

tumors with increased levels of mRNA editing, with malignant tumors having higher levels of

editing than plexiform neurofibromas, which in turn had higher levels of editing than dermal

neurofibromas. Importantly, this trend at the NF1 gene does not appear to be due to an overall

increase in general mRNA editing in malignant cells (Cappione et al., 1997). Mukhopadhyay et

al. (2002) evaluated RNA editing in malignant peripheral nerve sheath tumors (MPNSTs) and

found that 76.5% of tumors examined exhibited low levels of mRNA editing (0-2.5%), near the

reliable detection limits of their assay. However, they also identified a distinct subpopulation of

these tumors (23.5%) that exhibited mRNA editing at higher levels (3-12%). Tumors that

exhibited this higher, reproducible level of mRNA editing at position 3916 also exhibited

Page 36: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

36

increased levels of Type II mRNA relative to Type I. The creation of this stop codon in some

transcripts in cells expressing Type II mRNA would further reduce neurofibromin function by

leading to higher levels of activated RAS. These data were suggested to support a connection

between RNA editing, alternative exon23a splicing, and tumorigenesis. The extent of these

post-transcriptional modifications in NF1 tumorigenesis has yet to be validated by multiple

independent groups. I examined alternative splicing of exon23a and mRNA editing in our set of

NF1 tumors and normal Schwann cells to test for a connection between these two

post-transcriptional modifications. It is hoped that these studies will further clarify the roles of

these modifications in NF1 tumorigenesis and provide potential new avenues of treatment, aimed

at altering mRNA expression and post-transcriptional modifications.

Material and Methods

Reverse-Transcription and PCR

RNA was previously isolated from blood (8 non-NF1 patients, 7 NF1 patients), tumor (22

dermal neurofibromas, 21 plexiform neurofibromas, 6 MPNSTs), and culture samples (3 normal

Schwann cell, 4 dermal tumor cultures, 9 plexiform tumor cultures, 2 immortalized plexiform

neurofibroma cell lines) using the Trizol reagent and manufacturer’s protocol (Invitrogen), and

stored at -80°C. Reverse transcription reactions were carried out using Superscript II reverse

transcription kit (Invitrogen) and random hexamer primers. The resulting cDNAs were used in

PCR reactions under standard conditions using primers in exon23.1 and exon24, flanking

exon23a (CAT-H: 5’ ATTGTGATCACATCCTCTGATTGG 3’;

CAT-I: 5’ ATCTAAAATCCCTGCTTCATACGG 3’). Two fragments were amplified, one each

from Type I and Type II mRNA (303 and 366 bp, respectively). The two isoforms were

separated by electrophoresis on 8% native polyacrylamide gels and visualized by ethidium

bromide staining. Based on visual observation of band intensities, the relative ratio of Type I to

Page 37: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

37

Type II NF1 mRNA was noted (I<<II, I<II, I≈II, I>II, etc.) and compared between tumor

samples and controls, and between tumor types. Samples having only (or mostly) type II mRNA

were selected for cloning and sequencing to determine if RNA editing at C3916 was occurring in

these tumors.

Cloning and Sequence Analysis

New RT-PCR products were generated from the cDNA samples identified above, using

primers E21c5’ (5’CACAATGATGGGTGATCAAGG 3’) and

E23.1c3’ (5’CATGTTGCCAATCAGAGGATG 3’), spanning the 3916 edit site (377 bp

fragment from Type II mRNA). The products were ligated into a plasmid vector (Topo TA,

Invitrogen) and transformed into chemically competent E. coli cells (One Shot TOP10,

Invitrogen) using the manufacturer’s protocol. The cells were plated on LB-ampicillin agar

plates with IPTG and X-gal, and incubated at 37°C overnight. Bacterial colonies (clones)

positive for the insert were identified by blue/white selection, followed by PCR amplification

using the original primers. The PCR products from 30 positive clones were sequenced. Cycle

sequencing was done using the ABI Big Dye 3 system and automated sequencers in the UF CEG

sequencing core, with one of the PCR primers as a sequencing primer. Controls were used at all

steps.

Results

Alternative Splicing Patterns

I investigated the relative concentrations of Type I versus Type II mRNA in fresh blood

leukocytes, dermal neurofibromas, plexiform neurofibromas, MPNSTs, and multiple cell

cultures including normal Schwann cells. The use of a single primer pair to amplify both

isoforms helps preserve the ratio present in the original sample. Multiple samples of each tissue

type were analyzed, to determine if there was a common splicing pattern for the given cell type,

Page 38: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

38

and how that compared to the literature. A summary of the results of this analysis is given in

Table 3-1, and a detailed list of samples analyzed can be found in Appendix B (Table B-1).

In leukocytes from non-NF1 patients, Type I mRNA was present at equal to or greater

levels then Type II in 7/7 samples, with the majority showing Type I as the predominant

transcript (5/7). Results from these samples are shown in Figure 3-1A. From left to right in the

figure, lane 1 and 2 show more Type I than Type II, lanes 3 and 4 show approximately equal

levels of the two isoforms, and lanes 5-7 again show more Type I than Type II transcript. In

leukocytes from seven NF1 patients, there was more variability between samples. Two out of

seven showed Type II as the predominant transcript, three out of seven had approximately equal

amounts of the two isoforms, and the remaining two had more Type I transcript than Type II.

Three cultures from normal human Schwann cells were analyzed. All three of these

cultures contained Type II mRNA as the main transcript, with one culture exhibiting a much

higher level of Type II compared to Type I than the others.

Both primary dermal neurofibroma tissue and Schwann cell cultures derived from dermal

tumors were analyzed for their exon23a splicing patterns. Of the 21 primary dermal

neurofibroma samples, there were two tumors from which corresponding cultures were also

analyzed. An additional two dermal tumor-derived cultures were analyzed. The majority of

primary dermal tumors contained predominantly Type II transcript (18/21), with 7 having very

low levels of Type I in comparison. Two dermal tumor samples, UF80T32 and UF505T4, had

Type I as the main transcript. One primary dermal sample, UF80T2, contained no Type I

transcript, and was selected for cloning and sequencing to test for RNA editing. All 4 dermal

tumor-derived cultures contained predominantly Type II transcript. Figure 3-1B shows

experimental results from some representative dermal samples. Lane 1 is a sample with more

Page 39: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

39

Type I mRNA, lanes 2-5 and 7 are samples with predominantly Type II transcript, and lane 6 is

UF80T2 (which has no detectable Type I transcript).

Twenty-five plexiform neurofibromas and nine cultures derived from plexiform tumors

were also analyzed for exon23a splicing. Of the nine cultures analyzed, the corresponding

primary tumor was also analyzed for four. The majority of primary plexiform tumors (23/25)

contained predominantly Type II transcript, three of which had very little Type I in comparison,

and two having only trace amounts of Type I. Of the remaining two tumors, one had about equal

levels of Type I and Type II mRNA, and the other had Type I as the predominant transcript. The

majority of plexiform tumor-derived cultures (8/9) also had Type II as the predominant

transcript, with 3 having barely detectable levels of Type I transcript. The remaining culture,

UF469Tc, contained only Type II transcript and was also selected for RNA editing analysis.

Figure 3-1B shows the experimental results from representative primary plexiform tumors, and

one plexiform-derived culture. Lanes 8-10, and lane 12 show samples with predominantly Type

II transcript, and lane 11 is UF469Tc, with only Type II mRNA.

Finally, 6 MPNSTs were analyzed. All 6 had Type II mRNA as the predominant

transcript, with 3 showing much lower levels of Type I in comparison. Figure 3-1C shows the

experimental results for representative MPNSTs. Lanes 1 and 2 show samples with much more

Type II then Type I transcript, while lanes 3 and 4 show samples with predominantly Type II

mRNA. One MPNST, SNF94.3, had previously been analyzed for exon23a splicing levels, and

was found to contain only Type II transcript. This sample was also chosen for RNA editing

analysis.

Analysis of RNA Editing

Three samples were selected for RNA editing analysis based on their exon23a splicing

patterns. All three samples contained only Type II transcript, based on polyacrylamide gel

Page 40: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

40

electrophoresis, but they were all from different tissue types. UF80T2 is a primary dermal

neurofibroma, UF469Tc is a plexiform neurofibroma-derived Schwann cell culture, and

SNF94.3 is an MPNST. Plasmid vectors containing individual cDNA fragments were cloned in

E. coli and then PCR amplified for sequencing. This allowed for the analysis of the sequence of

a single cDNA molecule at a time, to detect the number of transcripts that were undergoing RNA

editing at C3916. Thirty-five (UF469Tc) to Forty (SNF94.3 and UF80T2) clones were analyzed

for each sample of interest. This number was chosen based on the levels of RNA editing

previously detected in MPNSTs lacking Type I transcript (3-12%) (Mukhopadhyay et al., 2002).

Even at the lowest levels previously seen, at least one in forty clones would be expected to

contain an edited cDNA fragment. No RNA editing was seen in any of these tumor samples,

despite previous reports of increased levels of RNA editing in tumors containing predominantly

Type II transcript.

Discussion

The relative ratios of Type I to Type II NF1 mRNA in a given tissue sample can be

determined using a single set of PCR primers to amplify the fragment of interest. The alternative

splicing pattern of exon23a has previously been studied for many embryonic and adult solid

tissues in mouse and human, but the levels of alternative splicing seen in blood leukocytes were

not well established. I observed that the ratio of Type I to Type II NF1 transcript in non-NF1

blood leukocytes is similar to that reported in the adult human central nervous system, with all

the samples (7/7) showing Type I transcript at equal or greater levels compared to Type II, with 5

of those 7 having Type I as the predominant transcript. This is in contrast to most other human

and mouse postembryonic tissues studied, where Type II predominates. The deviation from this

pattern in leukocytes is interesting. It is known that some alleles with NF1 mutations have

decreased mRNA levels, presumably due to nonsense-mediated decay (Colman et al., 1993;

Page 41: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

41

Hoffmeyer et al., 1995; Pros et al., 2006). However, there should be no such mechanism

operating in normal leukocytes, and it is known that this alternative splicing affects both alleles

(Thomson and Wallace, 2002). Thus, there may be functional reasons in leukocytes and the

central nervous system for a relative lack of Type II transcript, and the presence of an NF1 gene

mutation shouldn’t affect relative ratios of Type I to Type II transcript. Yet in blood leukocytes

from NF1 patients, there is more variation in the relative ratio of Type I to Type II transcript

observed. Only 5/7 samples showed relative ratios similar to those seen in the majority of

non-NF1 bloods, with the majority of those 5 having equal levels of the two transcripts. It

appears that there may be a trend toward inclusion of exon23a in the mRNA from NF1 patient

leukocytes, compared to leukocytes from non-NF1 patients.

To determine if the level of exon23a inclusion varied based on tumor type, I analyzed the

relative ratio of Type I versus Type II mRNA in both dermal and plexiform neurofibroma

samples. There was no significant difference in these relative ratios between the two types of

neurofibromas. The majority of both dermal (18/21) and plexiform (23/25) neurofibromas had

Type II as the predominant transcript. This is not unexpected as cultures of normal human

Schwann cells, the cell type from which neurofibromas are clonally-derived, all contained more

Type II mRNA then Type I. Observation of a relative (but not dramatic) predominance of Type

II transcript suggests a relatively reduced RAS-GAP activity in those cells, which would be

inferred for most normal adult tissues, including, as I have shown here, Schwann cells, based on

their Type I to Type II ratio. However, many of the NF1 tumors had a profile of Type II

transcript at much higher levels relative to Type I, with several tumors, of each variety of

neurofibroma, showing near or complete loss of the Type I transcript. This could result in

sufficiently less RAS-GAP activity relative to the native Schwann cell, which could contribute to

Page 42: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

42

tumorigenesis. If true, then a mechanism to shift the ratio back toward more equal levels of the

two transcripts could potentially decrease tumorigenic potential.

Previous reports had indicated that sample collection conditions and other environmental

factors may influence the level of alternative splicing in NF1 mRNA (Thomson and Wallace,

2002). To determine if culturing condition altered the level of exon23a inclusion, I examined the

relative ratios of Type I to Type II NF1 mRNA in several primary tumor and corresponding

tumor Schwann cell culture pairs. Six such pairs were analyzed: two were dermal neurofibromas

and their corresponding cultured Schwann cells, and four were plexiform neurofibromas and

their corresponding Schwann cell cultures. All 6 primary tumors analyzed contained

predominantly Type II mRNA, and there was no obvious change in the relative ratio of the two

isoforms compared to the corresponding cultures. The result of one of these comparisons is

shown in Figure 3-2 (plexiform neurofibromas). These results suggest that there is an inherent

Type I:Type II control mechanism in these tumor Schwann cells that is not susceptible to

influences of tissue culture, and it appears that the relative ratio of Type I to Type II mRNA seen

in a tumor cell culture can be a good estimate of the ratio seen in the primary tumor.

It was previously reported that MPNSTs containing little or no Type I transcript may

exhibit RNA editing (Mukhopadhyay et al., 2002). All six of the MPNSTs analyzed here

showed Type II as the predominant transcript and one (SNF94.3) that had been previously

analyzed showed complete loss of the Type I transcript by ethidium bromide visualization. I

analyzed this MPNST, as well as one dermal neurofibroma and one plexiform

neurofibroma-derived culture, both of which also showed loss of the Type I transcript, for the

presence of RNA editing. Thirty-five to forty clones from each sample were sequenced to check

for RNA editing at C3916, but this edit was not seen in any of the three samples with little to no

Page 43: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

43

Type I transcript. Mukhopadhyay et al. (2002) reported levels of RNA editing in their MPNSTs

of interest at 3-12%. If RNA editing occurred to a similar extent in all tumors showing a near or

complete loss of the Type I transcript, I would have expected to see 1-4 clones out of 40

undergoing RNA editing. The absence of editing in my samples could have been due simply to

chance. However, it could imply that any connection seen between increased levels of Type II

mRNA and RNA editing is not universal, even in MPNSTs of this type, or that the amount of

such editing is quite low, in which case its functional significance would seem minimal.

While the creation of a premature stop codon in NF1 mRNA by RNA editing has a clear

negative effect on neurofibromin function, the effect of having predominantly Type II transcript

may also have a negative impact, as the protein encoded by this transcript has a reduced

RAS-GAP activity. The first MPNST that I studied for RNA editing has one of the germline

NF1 microdeletions, but no somatic mutation has been found in the remaining allele despite

sequencing the entire mRNA open reading frame and the 3’ UTR. This raises the interesting

possibility that the alternative splicing in this tumor substitutes as a mutation, producing a

hypomorphic allele. The cells thus would lack sufficient NF1 RAS-GAP activity to prevent

tumorigenesis. This suggests that there may be a threshold of neurofibromin activity that keeps a

cell from becoming tumorigenic. This is an important notion that could have relevance for future

therapies. Alternatively, there could be one of several epigenetic changes (other than RNA

editing) that could constitute the second hit in this MPNST, and possibly other NF1 tumors.

These include histone modifications, changes in microRNA effects, and changes in downstream

regulatory effects (Schmegner et al., 2005; Ling et al., 2006; Martinez and Schackert, 2007;

Shelton et al., 2008; Bartels and Tsongalis, 2009). The latter is evidenced by studies by Hawes

et al. (2007) that found that different mouse strains have different levels of Nf1 expression. This

Page 44: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

44

implies that different background levels of transcription factors can have an effect on NF1

expression levels (Zhu et al., 2004).

The observation of different Type I to Type II ratios in two dermal neurofibromas from the

same individual (UF80T2 and UF80T32; UF505T4 and UF505T7) suggests that this ratio is

specific to each tumor and is not heavily controlled by systemic factors. This is consistent with

the fact that each neurofibroma has a different NF1 somatic mutation (and possibly other genetic

or epigenetic alterations) and is therefore independent.

Page 45: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

45

Table 3-1. Summary of alternative splicing of exon23a seen in various sample types examined.

mRNA levels

Cell Type

Total

sample # No Type I Type I < II Type I ≈ II Type I > II

Normal Schwann cell culture 3 3 (1+)

Non-NF1 patient blood 7 2 5

Non-NF1 patient fibroblasts 1 1

NF1 patient blood 7 2 3 2

Primary dermal tumor 21 1 18 (5+, 2*) 2

Dermal tumor cell culture 4 4 (2+)

Primary plexiform tumor 25 23 (3+, 2*) 1 1

Plexiform tumor cell culture 9 1 8 (1+, 3*)

Immortalized plexiform cell

lines 2 2

MPNSTs 7 1 6 (3+)

+: Much more Type II; *: Barely detectable Type I in comparison to Type II.

Page 46: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

46

A

B

C

Figure 3-1. Representative gels showing relative concentrations of Type I v Type II mRNA in

various tissue types studied. Upper band represents Type II mRNA, lower band Type

I. A) leukocytes from 7 non-NF1 patients. B) Primary tissue samples from dermal

(lanes 1-7) and plexiform (lanes 8-12) tumors. C) Cultured (left two lanes) and

primary tissue (right two lanes) from MPNSts.

Figure 3-2. Comparison of alternative splicing of exon23a in primary plexiform tumors v

cultured Schwann cells from the same tumors. Tumor culture sample is loaded first,

corresponding primary tumor sample in the lane to the right. T= primary tumor,

C= tumor culture.

1 2 3 4 5 6 7

1 2 3 4 5 6 7 8 9 10 11 12

1 2 3 4

C T C T

Page 47: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

47

CHAPTER 4

MISSENSE MUTATION COMPUTATIONAL ANALYSIS OF PATHOGENICITY

Introduction

Missense mutations are single base substitutions that result in an altered mRNA codon,

leading to an amino acid substitution at the protein level. In neurofibromatosis 1 (NF1),

missense mutations account for 10-20% of disease-causing germline lesions (reviewed by

Thomson and Wallace, 2002). Missense mutations a priori may be pathogenic or represent

neutral polymorphisms. For this reason, determining the pathogenicity of such mutations is

critical and is a major challenge in molecular diagnosis. Cooper and Krawczak (1993) outline

eight points of evidence that may indicate a missense mutation is pathogenic: if the mutation is in

an important structural of functional region of the gene; if the mutation alters an highly

conserved codon; if there are multiple unrelated reports of the mutation in patients; if there is no

observation of the mutation in healthy individuals; if the mutation segregates with the disease

phenotype within a family; if the mutant protein produced in vitro has the same properties and

characteristics as protein produced in vivo; and if introduction of wild-type protein can rescue

the disease phenotype in patients or culture. The first 6 of these points can be useful in

determining the pathogenicity of NF1 missense mutations, but the general difficulty of this

process is exacerbated by the fact that most NF1 mutations cannot be tested functionally in the

lab, eliminating the final 2 points of evidence. Thus, the finding of an NF1 missense mutation in

a person lacking sufficient diagnostic criteria can be a clinical dilemma. Missense mutations

should also be tested for a cryptic splicing effect, to best understand pathogenesis. When there

are no splicing errors, and if there is no useful information from the family or literature regarding

a novel mutation’s effects, other methods must be used to predict the mutation's pathogenicity.

Page 48: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

48

It has been estimated that over 50% of gene lesions known to cause hereditary disorders in

humans are missense mutations (Cooper et al., 1998), and predicting the effects of these

mutations on their corresponding proteins as well as their contribution to disease can be difficult

as well. Previously computational-based methods were designed to aid in the understanding of

the importance of specific amino acids in protein structure and function. These original methods

did not take into account all information specific to the protein of interest, and were not generally

designed to predict the effects that a missense mutation will have on its protein (Henikoff and

Henikoff, 1992; Ng and Henikoff, 2001). Rather, they provided information about the likelihood

of finding a particular amino acid at a particular position based on ortholog sequence alignments.

Additionally, while some studies have used these likelihoods to extrapolate pathogenicity, this

use has not been experimental validated. In recent years, however, several new computational

methods have been developed for the purpose of predicting the effect of a missense mutation,

aided partially by the increasing amount of genomic sequence information available, and taking

into account more factors specific to the protein, such as biochemical properties of the amino

acids in specific regions of the protein. The first of these new program was SIFT (Sorting

Intolerant From Tolerant amino acid substitutions (Ng and Henikoff, 2001, 2002)), which uses

protein sequence orthologs to predict the tolerance of a particular amino acid at a particular site.

SIFT has been compared to traditional substitution matrices in the accuracy of predicting the

effect of amino acid substitutions in LacI, HIV-I protease, and bacteriophage T4 lysozyme, and

in all cases was more accurate overall (Ng and Henikoff, 2001). As new programs have been

developed, their predictions have been compared to those from SIFT as a test of each program’s

efficiency and accuracy. These programs include MAPP (Multivariate Analysis of Protein

Polymorphism (Stone and Sidow, 2005)) and SNPs3D (Yue and Moult, 2006). These two latter

Page 49: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

49

programs have been shown to be more accurate in their predictions than SIFT, using small, well

characterized proteins for analysis, with functional analyses available to confirm the results. In

comparison to these previously analyzed proteins, the NF1 protein product, neurofibromin, is a

much larger, 2818-amino acid ubiquitously-expressed peptide (plus or minus a few alternative

exons) (DeClue et al., 1991; Marchuk et al., 1991). With its large size and complex nature,

neurofibromin was examined here as a robust test of the efficiency and accuracy of these new

prediction programs, not only in comparison to SIFT, but to each other. This work has been

submitted for publication (Loda-Hutchinson et al., 2009).

Materials and Methods

Missense Computational Methods

Three freely available programs were chosen based on their reported performance, as well

as applicability to the analysis of neurofibromin (lack of structural data, etc.). The three

programs chosen were SIFT: Sorts Intolerant From Tolerant

(http://blocks.fhcrc.org/sift/SIFT.html (Ng and Henikoff, 2001; 2002)),

SNPs3D: (http://www.snps3d.org/ (Yue and Moult, 2006)), and MAPP: Multivariate Analysis of

Protein Polymorphism (http://mendel.stanford.edu/SidowLab/downloads/MAPP/MAPP.html

(Stone and Sidow, 2005)).

Other Databases and Programs

The following databases and programs were also used.

For sequence acquisition and analysis:

• NCBI

• http://snpper.chip.org

• SwissPROT/TrEMBL (http://www.expasy.ch/sprot/)

Page 50: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

50

For sequence alignment:

• CLUSTAL (www.clustal.org)

For ortholog data:

• Inparanoid (http://inparanoid.sbc.su.se/cgi-bin/index.cgi (O’Brien et al. 2005))

To generate phylogenetic trees:

• SEMPHY-Structural EM Phylogenetic Reconstruction

(http://compbio.cs.huji.ac.il/semphy/ (Ninio et al. 2006))

For prediction of novel splice sites created by mutations:

• NNSPLICE version 0.9 (www.fruitfly.org/seq_tools/splice.html (Reese et al. 1997))

Data Sets

Compiled from a literature search, the Human Gene Mutation Database (www.hgmd.org),

and unpublished laboratory findings, the three data sets included missense mutations of known

effect, mutations created by site-directed mutagenesis in neurofibromin’s isolated GAP Related

Domain (GRD) and tested for RAS-GTP activity, and missense mutations of unknown effect.

For mutations known to create cryptic splice sites, the amino acid residue that would normally

result from the new codon was used for analysis. Sample size: known germline neutral n=8,

known germline pathogenic n=18, neutral in site-directed mutagenesis n=7, pathogenic in

site-directed mutagenesis n=5, Known pathogenic due to splicing error n=8, and unknown n=39.

References listed for these mutations do not necessarily include every report. I reported these

references to our best ability from the literature and the Human Gene Mutation Database

(www.hgmd.org).

Sequence Input Requirements

For SIFT, the human neurofibromin amino acid sequence (gi|4557793|ref|NP_000258.1|

neurofibromin [Homo sapiens]) was provided as a reference, and the program assembled a

Page 51: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

51

multiple sequence alignment (MSA) from similar sequences found in the SwissPROT/TrEMBLE

database. In total, 43 sequences were aligned by the SIFT program, and saved for further use.

All “-” in the returned alignment were converted to “X” to allow proper analysis by SIFT when

the sequences were re-entered (Ng, personal communication). Not all sequences were

considered at all mutation sites (due to high sequence similarity); SIFT returned data as to the

number of sequences used in the analysis of each mutation, but not on which sequences these

were. Listed below is the identification information provided by SIFT for these sequences.

• NF1_RAT Neurofibromin (Neurofibromatosis-related protein NF-1)

• NF1_HUMAN Neurofibromin (Neurofibromatosis-related protein)

• Q5SYI1_MOUSE (Q5SYI1) Neurofibromatosis 1

• NF1_MOUSE Neurofibromin (Neurofibromatosis-related protein)

• Q5SYI2_MOUSE (Q5SYI2) Neurofibromatosis 1

• Q9YGV2_FUGRU (Q9YGV2) Neurofibromatosis type 1

• Q59DT9_DROME (Q59DT9) CG8318-PD, isoform D

• Q9VBJ2_DROME (Q9VBJ2) CG8318-PB, isoform B

• O01399_DROME (O01399) Neurofibromin

• O01398_DROME (O01398) Neurofibromin

• O01397_DROME (O01397) Neurofibromin

• Q7QBJ9_ANOGA (Q7QBJ9) ENSANGP00000003216 (Fragment)

• Q7PGW6_ANOGA (Q7PGW6) ENSANGP00000025084 (Fragment)

• Q8IMS2_DROME (Q8IMS2) CG8318-PC, isoform C

• Q4T1K3_TETNG (Q4T1K3) Chromosome 16 SCAF10562, whole genome shotgun

• Q4T1K5_TETNG (Q4T1K5) Chromosome 16 SCAF10562, whole genome shotgun

• Q8WZ-6_NEUCR (Q8WZ-6) Related to NEUROFIBROMIN

• Q6CMT2_KLULA (Q6CMT2) Kluyveromyces lactis strain NRRL Y-1140

• IRA2_YEAST Inhibitory regulator protein IRA2

• Q6FJ13_CANGA (Q6FJ13) Candida glabrata strain CBS138 chromosome M

• Q757I8_ASHGO (Q757I8) AER025Cp

• Q3TYD2_MOUSE (Q3TYD2) Visual corte- cDNA, RIKEN full-length enriched

• Q59F-3_HUMAN (Q59F-3) Neurofibromin variant (Fragment)

• Q5SYH9_MOUSE (Q5SYH9) Neurofibromatosis 1 (Fragment)

• Q8CCE8_MOUSE (Q8CCE8) 15 days embryo male testis cDNA, RIKEN

• Q4HTV9_GIBZE (Q4HTV9) Hypothetical protein

• Q14931_HUMAN (Q14931) NF1 N-isoform-e-on11

• Q4R3N5_MACFA (Q4R3N5) Testis cDNA clone: QtsA-15713, similar to human

• Q7RWZ8_NEUCR (Q7RWZ8) Hypothetical protein

• Q8BQG3_MOUSE (Q8BQG3) 9 days embryo whole body cDNA, RIKEN

• Q95U43_DROME (Q95U43) GH08833p

Page 52: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

52

• Q5C2A5_SCHJA (Q5C2A5) SJCHGC09051 protein (Fragment)

• Q62595_RATLE (Q62595) Neurofibromatosis protein type 1 (Fragment)

• Q62596_RATLE (Q62596) Neurofibromatosis protein type 1 (Fragment)

• Q8UVE4_9FALC (Q8UVE4) Neurofibromatosis type 1 (Fragment)

• NF1_CHICK Neurofibromin (Neurofibromatosis-related protein

• Q55CR5_DICDI (Q55CR5) Hypothetical protein

• Q62597_RATLE (Q62597) Neurofibromatosis protein type 1 (Fragment)

• Q8SSU4_DICDI (Q8SSU4) Similar to Dictyostelium discoideum (Slime

• Q7Z3J5_HUMAN (Q7Z3J5) Hypothetical protein DKFZp686J1293 (Fragment)

• Q5C284_SCHJA (Q5C284) SJCHGC08175 protein (Fragment)

• P79186_9PRIM (P79186) Neurofibromin (Fragment)

• P79796_HYLCO (P79796) Neurofibromin (Fragment)

SNPs3D required “NF1” to be entered as a “gene ID” for SNP analysis, individual

mutations were then entered in the “Your SNP” box for analysis, and data was returned

regarding the sequences chosen for use in the prediction for that specific mutation site. The

number and species of the chosen sequences varied between mutation sites, ranging from 5 to 8,

and links were provided to each sequence’s NCBI entry.

Finally, MAPP requires a strong ortholog set for best results, and I tested several different

sets (mammals only: Homo Sapien, Bos taurus (cow), Canis familiaris (dog), Macaca mulatta

(rhesus macaque), Monodelphis domestica (opossum), Pan troglodytes (chimp), Rattus

norvegicus (rat), Mus musculus (mouse); mammals, amphibians, and birds: mammalian

sequences, Gallus gallus (chicken), Xenopus tropicalis (frog); mammals, amphibians, birds, and

fish: previously listed sequences, Takifugu rubripes (puffer fish), Gasterosteus aculeatus (3

spine stickleback), Tetraodon nigroviridis; mammals, amphibians, birds, fish, and insects:

Previously listed sequences, Anopheles gambiae (mosquito), Apis mellifera (honeybee),

Drosophila melanogaster (fruit fly)). Inparanoid was used to determine the best sequences to

use, and these were arranged into several different multiple sequences alignments (MSAs) by

Clustal. In addition, a phylogenetic tree is required for MAPP analysis. Using an MSA from

Clustal, SEMPHY was employed to produce a phylogenetic tree, and both the MSA and the

Page 53: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

53

corresponding tree were used in MAPP analysis. No specific mutations are analyzed by MAPP,

but rather the results for all possible substitutions at all sites are returned. Differences in

terminology used in reporting the results reflect those used by each program.

Splice Analysis

To analyze the possibility that mutations in our data sets could produce abnormal splicing

as their pathogenic mechanism rather than an amino acid substitution, the sequence of the entire

exon in which the mutation occurred, as well as the last 10 nucleotides of the 5’ and first 10

nucleotides of the 3’ intron were used for analysis by the NNSPLICE algorithm. Prediction

parameters were: Organism: Human or other, search for both 5’ and 3’ splice sites, no reverse

strand included, and a minimum score of 0.4 for both site types (out of a maximum 1.0).

Results

Table 4-1 shows the summary of the results, while Table C-1 (Appendix C) gives detailed

results for each individual mutation analyzed. In both tables, “tolerated” and “no significant

impairment (NSI)” are the terms used to infer a non-pathogenic (neutral) prediction by the

respective programs. “Pathogenic” (SIFT, SNPs3D) and “deleterious” (MAPP) are used for

those missense mutations predicted to substantially alter protein structure and thus likely to be

associated with disease. For SNPs3D and MAPP, “Failed” indicates inability of the program to

make a prediction because there was insufficient data about that residue in other species. SIFT

also returned some predictions as pathogenic with “low confidence” (“low conf” in Table 4-1

and “(!)” in Table C-1). These were residues with prediction scores very close to the

tolerated/pathogenic threshold.

The first two data sets, germline mutations with known effect and mutations from

site-directed mutagenesis studies (control columns 1-5 of Table 4-1), were used to gauge the

accuracy of the various programs in predicting the functional effects of missense mutations on

Page 54: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

54

neurofibromin. Each of these two data sets can be further broken into two subsets, neutral

mutations (present in an unaffected relative, or did not alter RAS-GAP activity) and pathogenic

mutations (proven de novo mutations, or found in two or more independent patients but not

controls, or shown to have affected RAS-GAP activity by in vitro methods). The “Unknown”

mutations set included mutations identified in our lab or reported in the literature, lacking

conclusive data about pathogenicity beyond initial identification in an NF1 patient.

For the eight mutations known to be neutral (control column 2, Table 4-1), accuracy varied

by program, as well as by the sequences used in making the predictions. When using the 8

mammalian ortholog sequences, SIFT correctly predicted only 1/8 (12.5%) of the known neutral

germline mutations. When SIFT was allowed to compile a set of sequences from the available

databases (43 total), the number of correct neutral germline predictions increased to 50%. The

results were similar for MAPP: 1/8 (12.5%) of the known neutral germline mutations were

correctly predicted when the 8 mammalian ortholog sequences were used. However, adding the

additional sequences to the ortholog set (amphibian, chicken, fish, and insects) only increased

MAPP’s correct neutral predictions to 25%. The SNPs3D program itself chooses which

sequences to consider at each mutation site, and correctly predicted 4/8 (50%) of the known

neutral germline mutations. SNPs3D analysis of the remaining four known neutral mutations

failed due to inadequate data at those sites.

For the 18 known pathogenic germline mutations (control column 1, Table 4-1), the

sequences included in the analysis also affected the accuracy of the predictions. When only

using the 8 mammalian ortholog set, SIFT predicted all 18 mutations to be pathogenic; however,

the program reported low confidence in all these predictions. When using the 43 sequences,

SIFT accurately predicted 13/18 known pathogenic germline mutations (72.2%), with only two

Page 55: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

55

given a rating of low confidence. The differences seen between the various MAPP analyses are

not as great as those seen for the known neutral germline mutations. When MAPP used the 8

mammalian ortholog sequences, 17/18 mutations were predicted correctly (94.0%), and this only

decreased to 77.8% (14/18) when the entire ortholog set was used. SNPs3D accurately predicted

12/18 (66.7%) of the known pathogenic germline mutations.

For the remaining two data subsets (site-directed mutagenesis data, control columns 4-5,

Table 4-1), the number of ortholog sequences used did not affect the accuracy of the predictions

by SIFT and MAPP. SNPs3D called all of these mutations “pathogenic”. In the case of the 7

mutations found to be neutral in the site-directed mutagenesis RAS-GAP studies, all three

programs failed to predict any of them correctly. In contrast, all three programs correctly

predicted a pathogenic effect for the 5 mutations found to alter RAS-GAP activity in the

site-directed mutagenesis studies (100% accuracy)(control column 4, Table 4-1). To the best of

my knowledge, these mutations have not been reported in NF1 patients yet, and are less likely to

occur since some have 2 or more bases altered in the codons.

Overall, among the 46 control mutations, SIFT (8 mammals) accurately predicted the

effects of 69.5% of the mutations, and SIFT (43 sequences) was accurate for 54.3% of the

mutations. While SNPs3D accurately predicted the effects of 58.7% of the control mutations,

analysis did not return a result (failed) for 9 of the mutations due to insufficient sequence data at

those sites. Because in these cases the program needs more information to ensure an accurate

prediction, it is slightly misleading to include these samples in the “inaccurate” category. If only

taking into account those control mutations for which a prediction was returned (either correct or

incorrect) the accuracy of SNPs3D is 73% (27/37). The accuracy of MAPP varied slightly based

on the ortholog sequences used in analysis. Using the MAPP predictions made when the

Page 56: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

56

mammalian, amphibian, chicken, and fish neurofibromin ortholog sequences were used in

analysis (the most accurate), MAPP had a 60.9% accuracy rate (28/46) in predicting the known

effects of the control mutations.

All three programs (all predictions under all conditions, not including failures) were in

agreement, correctly calling 15 germline pathogenic control mutations (out of 18). However,

only one known germline neutral mutation was correctly called neutral by all three programs.

Among the 39 “Unknown” missense mutations, SNPs3D predicted that 16 would be

pathogenic, SIFT (43 sequences) predicted 31, SIFT (8 sequences) predicted 37, MAPP (8

sequences) predicted 36, and the remaining three MAPP analyses predicted 37 or fewer to be

pathogenic (inversely related to number of homolog sequences). In total, 14 different

“Unknown” mutations were predicted to be neutral by one or more programs. Of those 14, 6

were predicted to be neutral in only one analysis, and 8 were predicted to be neutral in two or

more of the analyses.

Seven of the mutations were already known to create cryptic splice sites. This is a pitfall

for predicting missense pathogenesis, since a “neutral” result could be inaccurate if the point

change actually induced cryptic splicing. NNSPLICE predicted 2 of the 7 reported cryptic splice

sites (29% accuracy). As missense mutations, SIFT predicted 2 of these as pathogenic, 5 as

tolerated. SNPs3D predicted 2 to be pathogenic, 1 to be tolerated, and the analysis of the

remaining 4 failed due to insufficient sequence data. The most accurate MAPP run (mammal,

amphibian, chicken, and fish orthologs) predicted 4 to be deleterious, 2 to have no significant

impairment, and the analysis of 1 failed. Thus, the 3 computational methods would have missed

the true pathogenicity of these mutations half or more of the time. Interestingly, NNSPLICE

found two mutations in the “Unknown” mutations set that it predicted to create cryptic splice

Page 57: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

57

sites, one a splice acceptor (score of 0.95) and one a splice donor (score of 0.56). Both of these

mutations happened to be predicted to be pathogenic/deleterious by all three programs.

To express the accuracy of these tests in terms of specificity (probability that a test calls a

mutation neutral when it really is neutral) and sensitivity (probability that a test calls a mutation

pathogenic when it really is pathogenic), data from control columns 1 and 2 (Table 4-1) were

used (real NF1 patient data, no splice errors). For SNPs3D, sensitivity was 92.3% (12/13),

specificity was 100% (4/4) and failure rate was 34.6% (9/26). For SIFT (8 mammals), sensitivity

was 100% (18/18), and specificity was 12.5% (1/8). For SIFT (43 orthologs), sensitivity was

72.2% (13/18), and specificity was 50% (4/8). For MAPP (8 mammals), sensitivity was 94.4%

(17/18), and specificity was 12.5% (1/8). For MAPP (mammals + amphibians + birds),

sensitivity was 94.4% (17/18), and specificity was 12.5% (1/8). For MAPP (mammals +

amphibians + birds + fish), sensitivity was 94.1% (16/17), and specificity was 25% (2/8), with

one failed analysis. For MAPP (mammals + amphibians + birds + fish + insects), sensitivity was

82.4% (14/17), and specificity was 25% (2/8), with one failed analysis.

Discussion

I compared the ability of three freely-available programs with various parameters to predict

NF1 missense mutation pathogenicity, and each had pros and cons in ease-of-use and in

accuracy. MAPP requires the most preparation and the ability to use a Java program. The user

must assemble multiple sequence alignments and generate phylogenetic trees from these MSAs

prior to using MAPP to make predictions. The program returns results regarding substitution of

all 22 amino acids for every amino acid position in your sequence (provided there is enough

information in the MSA) which prevents the need to re-run analysis for future mutations of

interest. However, if some of the orthologs used in the MSA have different numbers of encoded

amino acids (or insertions/deletions relative to other orthologs), finding the amino acid of interest

Page 58: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

58

can be difficult due to altered numbering systems. SIFT can also provide a similar analysis for

later use, and this output is simpler to search through then the MAPP results. SIFT also allows

for several different options in the amount of preparation needed before analysis. The program

will except a gene ID number or a protein sequence and produce a MSA, or align a group of

sequences already collected. Finally, an MSA can also be submitted for analysis. If reusing an

MSA previously produced by SIFT, every "-" used as a place holder in partial sequences must be

replaced with "X" for analysis to proceed correctly. The length of time an analysis takes varies

based on amount of input given, with analysis of an MSA taking the least amount of time. In all

cases, a list of mutations for analysis is also entered, and all are analyzed simultaneously. If

more then one mutation occurs at the same amino acid, they must be analyzed separately, as the

program only returns specific results (number of sequences considered, scores, etc.) for one. In

the case of SNPs3D, all mutations are analyzed individually. Once the correct protein sequence

is found, individual mutations are entered one at a time and the program selects the sequences to

use in analysis. These will vary by position somewhat, but the user cannot choose which to

include or exclude. Though analysis of large numbers of mutations can be time consuming since

each mutation is entered separately, the SNPs3D results are quickly returned since the program

only aligned a small number of sequences (5-8) over a short sequence surrounding the site of

your mutation. The results are also well explained through links, and easy to interpret. These, in

addition to accuracy data (discussed below), are important considerations for future users.

The accuracy of these programs is in part based on the gene sequences used in making the

predictions. As can be seen in the results from the multiple MAPP analyses, the inclusion of

more diverse sequences (in this case amphibians, birds, and fish) can improve the reliability of

predictions (additional known neutral mutation predicted correctly compared to mammals-only

Page 59: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

59

results). This is also the case when comparing the results for the control sets returned by SIFT

when 8 mammalian sequences are used for analysis (where nearly all results were returned with

“low confidence”) versus the 43 sequences gathered by SIFT from the SwissPROT/TrEMBLE

database (very few “low confidence”). The program that used the fewest sequences (5-8) was

SNPs3D. Interestingly, for those mutations where SNPs3D was able to return a result, it appears

to be the most accurate. However, a major weakness was that SNPs3D analysis failed to return a

prediction for 34.6% of the controls (50% of the known neutral mutations, 27.7% of known

pathogenic mutations), preventing the accuracy of the program from being fully established.

While including sequences from more divergent species provides insight into what

mutations and amino acid substitutions are tolerated, this increased diversity may represent

divergence of function rather than tolerance for mutations. If that is the case, some mutations

may erroneously be predicted to be tolerated. In contrast to this idea, however, all three

programs showed error in favor of pathogenicity in the NF1 analysis. Across all analyses of the

control sets, (not including splicing errors), 68.9% of the predictions made for a known neutral

mutation (germline and those from site-directed mutagenesis) were “pathogenic/deleterious”,

while only 13.1% of the predictions made for a known pathogenic mutation (germline and those

from site-directed mutagenesis) were “neutral”. In the control sets (not including splice errors),

known neutral mutations were incorrectly predicted to be pathogenic 4 times as often as known

pathogenic mutations were predicted to be neutral. This suggests that the diversity included in

my sequences may not represent divergence of function. It also indicates that these programs

have an inherently higher false-positive rate than false-negative rate (higher sensitivity, lower

specificity)

Page 60: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

60

The poor performance of these programs in correctly predicting the effects of the known

neutral mutations could partially be a result of how a mutation is defined as neutral. It has been

estimated that the majority of missense mutations in the human genome are slightly deleterious

(Kryukov et al, 2007). It is possible that, while my known neutral mutations do not alter the

protein structure/function sufficiently to lead to a clinical-diagnosed NF1 phenotypes, the

biochemical effect is enough to be picked up by these prediction programs. Such mutations

could be considered NF1 hypomorphs, not completely neutral but not absolutely pathogenic.

Such mutations have not yet been proven in NF1, but it is theoretically possible (e.g. a mutation

alters RAS-GAP activity but isn’t seen in NF1 patients, or is found in individuals who only meet

one NF1 diagnostic criteria). Hypomorphic alleles would also confound the type of analysis

done here, and the interpretation of the outcome.

The inaccuracy of these programs in predicting neutral NF1 mutations does call into

question the number of false-pathogenic predictions that might be contained in the results for my

“unknown” data set. In my analyses of the control sets (across all programs and analyses),

88/262 pathogenic predictions were incorrect (33.6%). This number varies based on the specific

program, but it is possible that there are some mutations in my “unknown” set that may actually

be neutral despite being predicted to be pathogenic. None of the substitutions in this data set are

reported in dbSNP (http://www.ncbi.nlm.nih.gov/SNP/), which decreases the likelihood that my

unknowns are non-disease-related polymorphisms. This is consistent with the expectation that

novel NF1 mutations in NF1 patients are likely to be pathogenic. In the analyses of the control

sets, 66.7% of the neutral predictions made were incorrect. If this is approximately the rate of

incorrect neutral predictions within the “unknown” set as well, it can be estimated that

Page 61: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

61

approximately 9/16 of the unknowns predicted to be neutral are actually pathogenic. More

confidence can be placed in the “unknowns” predicted to be neutral by more than one analysis.

My results indicate that a prediction by any of these programs, whether it is a prediction of

pathogenic or neutral, cannot be taken alone in determining the effects of a given missense

mutation. As these programs were not designed to replace functional studies, it is

understandable that they seem to consider mutations pathogenic until proven neutral. SNPs3D

had the highest control accuracy but also the highest failed rate. SIFT predicted pathogenic

mutations at the expense of a high false-positive rate. MAPP called some known pathogenic

mutations “neutral”, but also had false-positives. No single program stood out as superior

overall. All three had high sensitivity (with at least one version of analysis). Only SNPs3D had

high specificity (100% of the germline neutral mutations, the others were ≤50%), but that

analysis also had a 50% failure rate. My SIFT accuracy and sensitivity results are consistent

with SIFT data for other monogenic situations (sensitivity 80-90%), although SIFT had a worse

specificity score in NF1 analysis (compared to 67-74% in other studies)(Mathe et al., 2006; Chan

et al., 2007). As indicated in other studies, it is considered beneficial to run multiple programs,

with greater faith in the results agreed upon by more than one analysis (Mathe et al., 2006; Chan

et al., 2007; Valdmanis et al., 2008). These programs are often updated and so the accuracy may

improve with newer versions. In addition, there are other programs becoming available and

refined (some of which are based on amino acid biochemistry rather than MSA), such as

PolyPhen. When choosing a program (or set of programs) one should consider the nature of the

question to be answered, the sequence data available, failure rate, as well as which type of errors

are more important to avoid. If one is looking to catch all possible pathogenic mutations, for

example to test further in functional studies, the program used will likely differ from one chosen

Page 62: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

62

to correctly identify as many possible neutral mutations as possible. We also found that a

missense analysis should be complimented by a splice site analysis (computational or in the lab

with RT-PCR) to attempt to find mutations that cause exon skipping or cryptic splicing rather

than an amino acid substitution. The data reported here will be useful for individuals considering

computational methods for testing pathogenicity of missense mutations, particularly in large

genes such as NF1.

Page 63: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

63

Table 4-1. Summary of computational missense prediction results for 6 data sets (5 controls and

1 unknown).

Controls (n=46)

Data set:

Program

Known

(germline)

Pathogenic

(n=18)

Known

(germline)

Neutral

(n=8)

Known

pathogenic

due to

altering

splicing

(n=8)

Altered GAP

function in

Site-directed

Mutagenesis

(n=5)

Neutral in

Site-directed

Mutagenesis

(n=7)

Unknown

(n=39)

SNPs3D 12 pathogenic

1 tolerated

5 failed

4 tolerated

4 failed

2 pathogenic

2 tolerated

4 failed

5 pathogenic 7 pathogenic 16 pathogenic

5 tolerated

18 failed

SIFT (43

sequences)

13 pathogenic

-2 low conf

5 tolerated

4 pathogenic

-1 low conf

4 tolerated

3 pathogenic

-1 low conf

5 tolerated

5 pathogenic 7 pathogenic 31 pathogenic

-16 low conf

8 tolerated

SIFT (8

mammalian

sequences

18 pathogenic

-18 low conf

7 pathogenic

-7 low conf

1 tolerated

8 pathogenic

-8 low conf

5 pathogenic

-5 low conf

7 pathogenic

-7 low conf

37 pathogenic

-37 low conf

2 tolerated

MAPP (8

mammalian

sequences)

17 deleterious

1 NSI

7 deleterious

1 NSI

6 deleterious

2 NSI

5 deleterious 7 deleterious 36 deleterious

3 NSI

MAPP

(mammals,

amphibians,

and birds)

17 deleterious

1 NSI

7 deleterious

1 NSI

5 deleterious

3 NSI

5 deleterious 7 deleterious 37 deleterious

2 NSI

MAPP *

(Mammals,

amphibians,

birds, and

fish)

16 deleterious

1 NSI

1 failed

6 deleterious

2 NSI

5 deleterious

2 NSI

1 failed

5 deleterious 7 deleterious 35 deleterious

4 NSI

MAPP

(Mammals,

amphibians,

birds, fish,

and insects)

14 deleterious

3 NSI

1 failed

6 deleterious

2 NSI

3 deleterious

4 NSI

1 failed

5 deleterious 7 deleterious 33 deleterious

6 NSI

*the most accurate MAPP analysis; NSI= no significant impairment.

Page 64: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

64

CHAPTER 5

CONCLUSIONS AND FUTURE DIRECTIONS

More then 20 years after the identification of the gene and protein responsible for

neurofibromatosis 1(NF1), there is still much that is unclear about NF1 mutations mechanisms,

how these mutations affect neurofibromin function, and how this relates to the heterogeneous

phenotypes of NF1. Additional knowledge of the mutation mechanisms occurring both in the

germline and somatic cells, as well as elucidating which mutation mechanisms do not play a

major role in NF1 progression, may contribute to developing targeted therapies and better

diagnosis. In this chapter, I reiterate the major findings of my work and discuss possible future

directions.

Somatic CpG C to T Mutations

From my work described here, C to T transition mutations at CpG dinucleotides are clearly

not a common somatic mutation mechanism in the NF1 gene. While only four such sites were

analyzed, they were chosen based on the fact that they are hotspots for these same mutations

occurring in the germline, are scattered across the gene, and are not known to be involved in

exon skipping. These same sites are also examined in screens for somatic NF1 mutations. It is

however still possible that other NF1 CpG sites are more susceptible to C to T transitions in

somatic cells, and that these mutations could play a role in NF1 tumorigenesis. Compared to

some genes, such as TP53, RB1, and NF2, germline NF1 C to T mutations are somewhat less

common, and somatic C to T mutations are much rarer. The difference in somatic C to T

mutation rates between NF1 (in neurofibromas), TP53 (in bladder and colon cancers), and APC

(in colon cancer) could be due to defects of the base excision repair pathway in more malignant

tumors, leading to a decreased ability of cells to recognize and correctly repair a G:T mispair.

Several reports have shown that neurofibromas, which are not malignant, show no evidence of

Page 65: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

65

microsatellite instability (MSI), which is often an indication of mutation in the DNA repair

pathways (Luijten et al., 2000; Upadhyaya et al., 2004; in contrast: Ottini et al., 1995). However,

malignant peripheral nerve sheath tumors (MPNSTs), which can develop from a plexiform

neurofibroma, have been shown to exhibit MSI in 30-45% of cases (Upadhyaya et al., 2004;

Kobayashi et al., 2006). Additionally, Wang et al. (2003) found that half of 10 non-NF1 human

cancer cells lines containing MSI had point mutations in NF1, whereas no mutations in NF1

were found in non-NF1 human cancer cell lines with functional mismatch repair pathways.

These data suggest that MPNSTs, with a higher occurrence of MSI (and likely repair pathway

mutations), may be more susceptible to CpG C to T transitions in NF1 than neurofibromas. This

could be determined by a similar mutation screen as the one I performed, with a large cohort of

MPNSTs.

If future studies were able to identify C to T transitions at CpG sites in neurofibromas, it

would be of interest to see if this is a common mutation mechanism within an individual.

Multiple neurofibromas from the same patient could help answer this question. If individuals

were found to be generally susceptible to such mutations, (e.g. less-than-adequate DNA excision

repair mechanisms, or exposed to mutagens), perhaps they could benefit from therapies being

designed to compensate for mutations causing premature stop codons, which are often caused by

C to T transitions at CpG sites. An interesting report described two siblings with some features

of NF1 (café-au-lait spots and skin-fold freckling) but also non-NF1-related cancer at an early

age; these children were found to be homozygous for a MSH2 mutation, suggesting that

pigmentary NF1 features may be mimicked by deficiency in mismatch repair (Toledano et al.,

2008). Full characterization of the NF1 somatic mutation repertoire may indicate whether

Page 66: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

66

certain types of mutations besides C to T transitions predominate, and whether these are related

to mutagen exposure.

Of interest, we do not yet know when somatic mutations occur in the Schwann cells that

initiate neurofibromas. The peripheral nervous system growth (and Schwann cell replication) is

predominantly done by adulthood, with Schwann cells going quiescent. These only become

active and enter mitosis when there is a nerve injury. It is possible that two-hit Schwann cells

could lie dormant until an environmental trigger such as injury or endocrine change starts clonal

cell expansion. There is no evidence for outright mutation at other genes in neurofibromas.

Conversely, perhaps second hits occur randomly throughout life and are followed shortly by

clonal expansion. Dermal neurofibromas rarely appear before adolescence, but plexiform tumors

can appear anytime in life (including infancy). So these two tumor types may have some basic

differences relative to somatic mutation occurrence that we do not yet understand. However, I

saw no CpG mutations in large sets of both tumor types, so the occurrence of this mutation type

does not appear different. CpG C to T transitions can occur in the germline (paternal), or any

time after the egg is fertilized throughout life. Knowing the timing of the second hit could also

useful for future therapies/preventions.

Alternative Splicing of exon23a

While it appears that the alternative splicing of exon23a is developmentally significant, the

role that inclusion of exon23a might play in tumorigenesis is still unclear. Analysis of RNA

transcripts from multiple cell types provided an estimate of the level of inclusion of exon23a

occurring in these cells (the relative ratio of Type I to Type II transcript). Leukocytes produce

predominantly Type I transcript, with a slight trend towards increased levels of Type II

compared to Type I mRNA in leukocytes from NF1 patients versus non-NF1 patients, although

the numbers were low. There were no strong differences seen in the relative ratios of Type I to

Page 67: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

67

Type II transcript seen in dermal versus plexiform neurofibroma samples, or between primary

tumor samples and their corresponding cultures. There also does not appear to be a difference in

the ratio of Type I to Type II mRNA when samples are analyzed by gender. However, there was

nearly twice the number of female individuals in this study as males. It would be of interest to

analyze samples from more male patients to rule out gender-specific trends in the alternative

splicing of exon23a. Additionally, I found several cases where two tumors from the same

individual had very different Type I to Type II ratios, indicating that these ratios are tumor-, not

individual-specific. Development of a real-time PCR protocol to distinguish levels of Type I and

Type II mRNA, as has been used to quantify skipping of exon37 in NF1 mRNA (Vandenbroucke

et al., 2001), would be useful in quantifying more subtle differences in the ratios of Type I to

Type II mRNA between primary tumors and their corresponding cultures, between tumor types,

or between genders. However, this would require a large sample set since the ratios would fall

into a greater number of bins.

Despite some differences in the amount of alternative splicing of exon23a between tumors,

there is a trend toward increased levels of exon23a inclusion in NF1 tumors compared to normal

Schwann cells, making this mechanism a potential target for therapy. Before such therapies can

be developed it must be shown that having (virtually) only Type II transcript can substitute as a

somatic mutation. To help determine if the reduced GAP activity of Type II neurofibromin is

“pathogenic” one of two experiments could prove informative. A construct containing the Type

II GAP domain driven by a low-level promoter could be introduced to an NF1 null cell with an

abnormal phenotype (e.g. increased invasiveness on Matrigel, increased passage number prior to

senescence, growth-factor independent) to see if the Type II GAP has sufficient activity to rescue

the phenotype. Alternatively, Schwann cells hemizygous for NF1 (e.g. from a patient with a

Page 68: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

68

germline large deletion) could be forced, through recently developed splicing therapies, to

express only Type II transcript to determine if this causes any changes in phenotype

(development of tumorigenic characteristics). Additionally, both the trans- and cis-acting

elements involved in controlling the splicing of exon23a must be identified. These elements

would be the targets of therapies designed to decrease the amount of exon23a inclusion, most

likely in Schwann cells if targeted correctly. Zhu et al. (2008) have identified some of these

factors that appear to regulate the inclusion of exon23a in neurons. Hu proteins are mRNA

binding proteins that are proposed to bind to AU-rich regions on either side of exon23a and

inhibit inclusion of the exon. As this interaction is in neurons, further studies need to be carried

out to identify splicing elements in Schwann cells. Interestingly, mice lacking exon23a have a

high rate of cognitive deficits (e.g. water maze memory test) where as exon31 knockout mice (an

out of frame deletions) do not show these deficits (Silva et al., 2001). Perhaps increased

inclusion of exon23a in more NF1 transcripts could be used to address this problem, which

affects approximately half of children with NF1. Such therapies are currently being developed to

alter the alternative splicing of exon10 of tau, a gene indicated in several progressive dementia

disorders (reviewed by Zhou et al., 2008). Interestingly, alternative splicing of exon10 in tau is

developmentally regulated in the central nervous system, as is exon23 alternative splicing in

NF1, although the pattern is not the same (Gao et al., 2000).

RNA editing at C3916 did not appear to be associated with loss of the Type I transcript in

tumors in this study, in contrast to published results. I saw no evidence of any RNA editing in

our samples, including MPNSTs. A larger study, repeating the analysis of the published samples

as well as new samples, may shed light on this inconsistency. It may also be of interest to

determine which of these changes is occurring first, in the tumors reported to exhibit both loss of

Page 69: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

69

the Type I transcript and RNA editing. My study used a qualification of complete or nearly

complete loss of the Type I transcript to identify samples to be evaluated for RNA editing, and

since none was found, it appears that this shift in alternative splicing alone does not lead to

increased RNA editing.

Computational Analysis Comparison

Recent advances in bioinformatics, as well as the vast amount of data being generated by

high-throughput techniques has greatly improved the methods available to predict the effects of

missense mutations on a protein of interest. In the study of NF1 and neurofibromin, this is

especially important as there is no way to study these mutations functionally. Using the limited

set of NF1 missense mutations with known effect to test the accuracy of these new prediction

programs, it appears that each may have its advantages in a given situation. It would be

advantageous to use multiple programs when possible. These programs are already being used

to supplement molecular diagnostic data, and so it is important to determine the best tools for the

job. This is a rapidly growing and evolving field, and it will be key to continue to increase the

control data set as more NF1 mutations with known effect are identified, to test the abilities of

these programs and ensure that best sequence data is being used. A clinical diagnosis may hinge

on data provided by programs such as these.

Page 70: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

70

APPENDIX A

CPG C TO T MUTATION ANALYSIS DATA

Table A-1. Results of CpG C to T mutation screen using TaqαI restriction enzyme digest.

Mutation at CPG site in:

Sample ID Exon 10a Exon 22 Exon 23.2 Exon 41

Dermal tumors

UF80 T1 N N

UF80 T2 N N

UF80 T4 N N

UF80 T6 N N N N

UF80 T8 N N

UF80 T11 N N

UF80 T12 N N

UF80 T12c N

UF80 T23 N N N N

UF80 T32 N N N N

UF113 T3 N N N

UF113 T4 N N N

UF113 T5 N N N N

UF113 T6 N N N

UF113 T7 N N N N

UF113 T9 N N N N

UF113 T10 N N N

UF113 T11 N N N

UF113 T12 N

UF113 T13 N

UF113 T14 N N

UF113 T15 N N

UF113 T16 N N N

UF113 T17 N N

UF113 T18 N N

UF113 T19 N N

UF113 T20 N N

UF113 T21 N N

UF113 T22 N N N

UF113 T23 N N

UF113 T24 N N

UF233 T1 N N N N

UF233 T2 N N N N

UF287 T1 N N N N

UF287 T2 N N N N

UF327 T1 N N N N

N= no, Y=yes, C=constitutional

Page 71: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

71

Tabel A-1. Continued

Mutation at CPG site in:

Sample ID Exon 10a Exon 22 Exon 23.2 Exon 41

Dermal tumors

UF328 T4 N N N N

UF328 T6 N N N N

UF328 T7 N N N N

UF328 T14 N N N N

UF417A N N N N

UF417D N N N N

UF431 T2 N N N N

UF474 T3 N N N N

UF486 T N N N N

UF486 T2 N N N N

UF505 T3 N N N N

UF505 T4 N N N N

UF509 T2 N N N N

UF510 T1 N N N N

UF512 T3 N N N N

UF532 T1 N N N N

UF552 T3 N N N N

UF705 T1 N N N N

UF743 T1 N N N N

UF831 T1 N N N N

UF831 T2 N N N N

UF831 T2c N N N N

UF835 T1 N N N N

UF1150 T N N N N

UF1345 T1 N N N N

UF1346 T1 N N N N

AW T2 N N N N

Plexiform tumors

UF158 T3 N N N N

UF158 T4 N N N N

UF181 T1 N N N N

UF303 T N N N N

UF310 T N N N N

UF327 T2 N N N N

UF340 T1 N N N N

UF344 T2 N N N N

UF346 T1 N N N N

UF356 T1 N N N N

UF362 T N N N N

N= no, Y=yes, C=constitutional

Page 72: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

72

Table A-1. Continued

Mutation at CPG site in:

Sample ID Exon 10a Exon 22 Exon 23.2 Exon 41

Plexiform tumors

UF375 T1 N N N N

UF375 T2 N N N N

UF378 T1 N N N N

UF378 T2a N N N N

UF378 T2b N N N N

UF386 T1 N N N N

UF387 T2 N N N N

UF389 T1 N N N N

UF420 T1 Y/C N N N

UF428 T1 N N N N

UF429 T N N N N

UF440 T1 N N N ?

UF440 Tc N N N N

UF450 T1 N N N N

UF452 T1 N N Y/C N

UF454 T1 N N N N

UF454 T2 N N N N

UF454 T3 N N N N

UF454 T4 N N N N

UF454 T5 N N N N

UF454 T6 N N N N

UF456 T1 N N N N

UF456 T3 N N N N

UF468 T N N

UF469 T1 N N N N

UF475 T1 N N N N

UF495 T N N N N

UF499 T N N N N

UF504 T N N N N

UF511 T1 N N N N

UF511 T2 N N N N

UF526 T1 N N N N

UF526 T2 N N N N

UF537 T1 N N N N

UF549 T1 N N N N

UF550 T1 N N N N

UF554 T1 N N N

UF554 Tc N N N N

UF555 T N N N N

UF562 T N N N N

N= no, Y=yes, C=constitutional

Page 73: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

73

Table A-1. Continued

Mutation at CPG site in:

Sample ID Exon 10a Exon 22 Exon 23.2 Exon 41

Plexiform tumors

UF572 T1 N N N N

UF572 Tc N N N

UF573 T1 N N N N

UF573 T2 N N N N

UF573 T3 N N N N

UF573 T4 N

UF593 T1 N N N N

UF609 T N N N N

UF622 T N N N N

UF632 T1 N N N N

UF746 T N N N N

UF787 T N N N N

UF836 T N N N N

UF836 Tc N N N N

UF860 T N N N N

UF1072 T2 N N N N

UF1093 T N N N N

UF1151 T N N N N

UF1160 T N N N N

UF1169 T N N N N

UF1207 T N N N N

UF1243 T1 N N N N

UF1243 Tc N N N N

UF1258 T N N N N

UF1296 T N N N N

UF1296 Tc N N N N

UF1308 T N N N N

UF1308 Tc N N N N

UF1371 Tdr ? N N N

UF1371 Tsw N N N N

UF1371 Tc dr N N N N

UF1371 Tc sw N N N N

N= no, Y=yes, C=constitutional

Page 74: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

74

APPENDIX B

EXON 23a ALTERNATIVE SPLICING DATA

Table B-1. Relative concentrations of Type I v Type II mRNA in blood, tumor and culture

samples.

Cell type Sample # No Type I Type I < II Type I ≈ II TypeI > II

Normal pm 97.3 X

Schwann pm 97.4 X+

cells pm 02.3 X

Dermal UF80T2 X

tumors UF80T4 X

UF80T5 X

UF80T6 X

UF80T18 X+

UF80T31 X

UF80T32 X

UF328T4 X

UF328T6 X

UF328T11 X

UF389T1 X

UF470T1 X

UF505T4 X

UF505T7 X*

UF526T1 X*

UF526T2 X+

UF1312T1 X+

UF1312T2 X

UF1313T1 X+

UF1313T1 sc-/- X+

UF1313T2 X

Dermal UF328T8c X

cultures UF470T1c X

UF470T2c X+

UF1313T2c X+

Immortal PNF95.11b P23 X

cell lines PNF95.11b P24 X

+: Much more Type II; *: Barely detectable Type I in comparison to Type II

Page 75: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

75

Table B-1. Continued

Cell type Sample # No Type I Type I < II Type I ≈ II TypeI > II

Plexiform PNF95.11b2 X

tumors PNF95.11a X

PNF95.6 X*

UF428T1 X+

UF429T(GD) X

UF429T X

UF450T1 X

UF450T1(P) X

UF532T1 X*

UF548T4 X

UF548T5 X

UF548T6 X

UF609T X

UF622T1 X

UF746T1A X

UF746T1B X

UF746T1C X

UF836TA X+

UF836TB X

UF860T X

UF1151T X

UF1160T X

UF1201T1 X+

UF1258T X

UF1371 DR X

Plexiform UF440Tc X*

cultures UF469Tc X

UF554T1c X*

UF609Tc X*

UF746TcA X

UF746TcB X

UF1243Tc X

UF1258Tc X+

UF1371 DR -/- X

+: Much more Type II; *: Barely detectable Type I in comparison to Type II

Page 76: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

76

Table B-1. Continued

Cell type Sample # No Type I Type I < II Type I ≈ II TypeI > II

NF blood UF80 X

UF389 X

UF450 X

UF470 X

UF746 X

UF836 X

UF1160 X

mPNSTs SNF 02.2 X+

SNF94.3 X

SNF 96.2 X+

UF158T X

UF344T1 X+

UF459T1 X

UF860T X

Control UF86 X

blood UF91 X

UF563 X

UF733 X

T80G GG X

PW Fresh X

RET Fresh X

fibroblasts UF1104 X+

Timed UF328fresh X

blood UF328 1 day X

UF328 3 day X

+: Much more Type II; *: Barely detectable Type I in comparison to Type II

Page 77: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

77

APPENDIX C

MISSENSE MUTATION COMPUTATIONAL ANALYSIS DATA

Page 78: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

78

Table C-1. Results for each mutation as returned by the various computational methods used.

Mutation

and

location

Codon

change Reference Type SNPs3D SIFT

SIFT

8

mammal

MAPP

mammals

MAPP

mam/am/

ck

MAPP

mam/am/

ck/fsh

MAPP

mam/am/

ck/fs/ins

Splice

Site

L549P

exon11 CTG-CCG

Fahsold et al

2000; Wallace

lab,

unpublished P failed Tol Patho (!) Del Del Del Del no

K505E

exon10b AAG-GAG Park et al 1998 P failed Tol Patho (!) Del Del Del NSI no

L508P

exon10b CTT-CCT

Wallace lab,

unpublished P failed Patho (!) Patho (!) Del Del Del Del no

S665F

exon12b TCC-TTC

Mattocks et al

2004; Fahsold

et al 2000 P failed Tol Patho (!) Del Del N/A N/A no

T780K

exon15 ACA-AAA

Han et al 2001;

Fahsold et al

2000 P Patho Patho Patho (!) Del Del Del Del no

W784R

exon15 TGG-CGG

Upadhyaya et

al 2008 P Patho Patho Patho (!) Del Del Del Del no

L844F

exon16 CTT-TTT

Mattocks et al

2004; Girodon

& Bouduret

2000 P Patho Patho Patho (!) Del Del Del Del no

L847P

exon16 CTT-CCT

Fahsold et al

2000; Wallace

lab,

unpublished P Patho Patho Patho (!) Del Del Del Del no

L898P

exon15 CTG-CCG

Fahsold et al.

2000; Maynard

et al 1997 P Patho Patho Patho (!) Del Del Del Del no

Page 79: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

79

Table C-1. Continued

Mutation

and

location

Codon

change Reference Type SNPs3D SIFT

SIFT

8

mammal

MAPP

mammals

MAPP

mam/am/

ck

MAPP

mam/am/

ck/fsh

MAPP

mam/am/

ck/fs/ins

Splice

Site

M968R

exon17 ATG-AGG

DeLuca et al

2003 P Patho Patho Patho (!) Del Del Del Del no

R1204W

exon21 CGG-TGG

Pros et al 2008;

Ars et al 2000 P Patho Patho Patho (!) Del Del Del Del no

R1276Q

exon22 CGA-CAA

Jeong et al

2006; Fahsold

et al 2000 P Tol Patho Patho (!) Del Del Del Del no

R1391S

exon24 AGA-AGT

Gutmann et al

1993b D/M Patho Patho Patho (!) Del Del Del Del no

K1419Q

exon24 AAG-CAG

Mattocks et al

2004;

Upadhyaha et

al 1997 P Patho Patho Patho (!) Del Del Del NSI no

K1423E

exon24 AAG-GAG

Li et al 1992;

Xu and

Gutmann 1997 D/M Patho Patho Patho (!) Del Del Del Del no

K1423Q

exon24 AAG-CAG

Li et al 1992;

Xu and

Gutmann 1997 D/M Patho Patho Patho (!) Del Del Del Del no

K1423S

exon24 AAG-TCG

Gutmann et al

1993b D/M Patho Patho Patho (!) Del Del Del Del no

L1425P

exon25 CTT-CCT

Peters et al

1999 P Patho Patho Patho (!) Del Del Del Del no

Q1426R

exon25 CAG-CGT

Gutmann et al

1993b D Patho Patho Patho (!) Del Del Del Del no

N1430R

exon25

AAT-AGA

or AGG

Wallace lab,

unpublished P Patho Patho Patho (!) Del Del Del Del no

S1468G

exon26 AGT-GGT

Mattocks et al

2004;

Upadhyaha et

al 1997 P Patho Tol Patho (!) NSI NSI NSI NSI no

Page 80: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

80

Table C-1. Continued

Mutation

and

location

Codon

change Reference Type SNPs3D SIFT

SIFT

8

mammal

MAPP

mammals

MAPP

mam/am/

ck

MAPP

mam/am/

ck/fsh

MAPP

mam/am/

ck/fs/ins

Splice

Site

G1498E

exon26 GGG-GAG

Pros et al

2008; Ars et

al. 2003 P Patho Tol Patho (!) Del Del Del Del no

L2317P

exon38 CTT-CCT

Wu et al 1999;

Wallace lab,

unpublished P failed Patho Patho (!) Del Del Del Del no

D176E

exon4b GAT-GAA

Wallace lab,

unpublished N Tol Tol Patho (!) Del Del NSI Del no

R765H

exon14 CGC-CAC

Wallace lab,

unpublished N failed Tol Patho (!) Del Del Del Del no

S858C

exon16 TCT-TGT

Wallace lab,

unpublished N Tol Tol Tol NSI NSI NSI NSI no

R873C

exon16 CGT-TGT Mattocks 2004 N Tol Patho (!) Patho (!) Del Del Del Del no

N1229S

exon21 AAT-AGT Ars et al 2003 N Tol Tol Patho (!) Del Del Del NSI no

E1264Y

exon22 GAA-TAC

Gutmann et al

1993b N Patho Patho Patho (!) Del Del Del Del no

A1281R

exon22 GCC-CGC

Gutmann et al

1993b N Patho Patho Patho (!) Del Del Del Del no

F1389H

exon24 TTC-CAC

Poullet et al

1994 N Patho Patho Patho (!) Del Del Del Del no

P1395I

exon24 CCT-ATT

Gutmann et al

1993b N Patho Patho Patho (!) Del Del Del Del no

A1396G

exon24 GCC-GGC

Poullet et al

1994 N Patho Patho Patho (!) Del Del Del Del no

P1400R

exon24 CCG-CGG

Gutmann et al

1993b N Patho Patho Patho (!) Del Del Del Del no

N1430M

exon25 AAT-ATG

Gutmann et al

1993b N Patho Patho Patho (!) Del Del Del Del no

Page 81: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

81

Table C-1. Continued

Mutation

and

location

Codon

change Reference Type SNPs3D SIFT

SIFT

8

mammal

MAPP

mammals

MAPP

mam/am/

ck

MAPP

mam/am/

ck/fsh

MAPP

mam/am/

ck/fs/ins

Splice

Site

R1809C

exon29 CGC-TGC Ars et al 2003 N failed Patho Patho (!) Del Del Del Del no

R1825W

exon29 CGG-TGG Ars et al 2003 N failed Patho Patho (!) Del Del Del Del no

A2058D

exon33 GCT-GAT Mattocks 2004 N failed Patho Patho (!) Del Del Del Del no

D186V

exon4b GAT-GTT

Zatkova et al.

2004 C.S. Patho Patho (!) Patho (!) Del Del Del Del no

Y489C

exon10b TAT-TGT

Messiaen et al

1999 C.S. Tol Tol Patho (!) Del Del Del NSI 0.97 A

G629R

exon12b GGG-AGG Ars et al 2000 CS failed Tol Patho (!) Del NSI N/A N/A 0.98 A

G922S

exon16 GGT-AGT Ars et al 2000 CS failed Tol Patho (!) NSI NSI NSI NSI no

V1093M

exon19b GTG-ATG Ars et al. 2003 C.S. Tol Tol Patho (!) Del Del Del NSI no

S1479G

exon26 AGT-GGT

Wallace lab,

unpublishes;

Upadhyaya et

al. 1997

(missense) C.S. Patho Tol Patho (!) NSI NSI NSI Del no

R1849Q

border

exon 29

+ 30 CGG-CAG Ars et al. 2000 C.S. failed Patho Patho (!) Del Del Del Del no

K2286N

exon37

AAG-AAC

or AAT

Messiaen et al

2000 CS failed Patho Patho (!) Del Del Del NSI no

Page 82: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

82

Table C-1. Continued

Mutation

and

location

Codon

change Reference Type SNPs3D SIFT

SIFT

8

mammal

MAPP

mammals

MAPP

mam/am/

ck

MAPP

mam/am/

ck/fsh

MAPP

mam/am/

ck/fs/ins

Splice

Site

H31R

exon2 CAT-CGT

Mattocks et al

2004 ? failed Tol Patho (!) Del Del Del Del no

S82F

exon3 TCT-TTT

Kluwe et al

2002 ? failed Patho (!) Patho (!) Del Del Del Del no

C93Y

exon3 TGT-TAT

Messiaen et al

2000 ? failed Patho (!) Patho (!) NSI Del Del Del no

L145P

exon4a CTC-CCC

Mattocks et al

2004 ? Patho Patho (!) Patho (!) Del Del Del Del no

C187Y

exon 4b TGT-TAT

Messiaen et al

2000 ? Tol Patho (!) Patho (!) Del Del Del NSI no

L194R

exon4b CTG-CCG

De Luca et al.

2005 ? Patho Patho (!) Patho (!) Del Del Del Del no

C324R

exon7 TGT-CGT

Mattocks et al

2004 ? Patho Patho (!) Patho (!) Del Del Del Del no

E337V

exon7 GAA-GTA

Mattocks et al

2004 ? Tol Patho (!) Patho (!) Del Del Del NSI no

D338G

exon7 GAT-CGT

Upadhyaya

1997 ? Patho Patho (!) Patho (!) Del Del Del Del 0.56 D

R440P

exon10a CGA-TGA

Wallace lab,

unpublished ? Patho Patho (!) Patho (!) Del Del Del Del no

Q519P

exon10c CAA-CCA

Upadhyaya et

al 2004 ? failed Patho (!) Patho (!) Del Del Del Del no

L532P

exon10c CTG-CCG

Mattocks et al

2004 ? Patho Patho (!) Patho (!) Del Del Del Del no

S574R

border

exon 11

& 12a AGC-CGC

Mattocks et al

2004 ? Tol Patho (!) Patho (!) Del Del NSI NSI no

L578P

exon12a CTT-CCT

Jeong et al

2004 ? Patho Patho (!) Patho (!) Del Del Del Del no

Page 83: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

83

Table C-1. Continued

Mutation

and

location

Codon

change Reference Type SNPs3D SIFT

SIFT

8

mammal

MAPP

mammals

MAPP

mam/am/

ck

MAPP

mam/am/

ck/fsh

MAPP

mam/am/

ck/fs/ins

Splice

Site

I581S

exon12a ATC-AGC Lee et al. 2006 ? Failed Patho (!) Patho (!) Del Del Del Del 0.95 A

R815M

exon16 AGG-ATG

Wallace lab,

unpublished ? Patho Patho (!) Patho (!) Del Del NSI NSI no

L1015R

exon18 CTG-CGG

Kluwe et al

2003 ? Patho Patho Patho (!) Del Del Del Del no

M1073V

exon19b ATG-GTG

Mattocks et al

2004 ? Tol Tol Patho (!) Del Del Del Del no

G1166D

border

exon 20

& 21 GGC-GAC

Purandare et al

1994 ? failed Tol Patho (!) Del Del Del NSI no

L1196R

exon21 CTT-CGT

Mattocks et al

2004 ? failed Patho Patho (!) Del Del Del Del no

R1204G

exon21 CGG-GGG

Krkljus et al

1998 ? Patho Patho Patho (!) Del Del Del Del no

R1276G

exon22 CGA-GGA

Mattocks et al

2004 ? Patho Patho Patho (!) Del Del Del Del no

R1325G

exon23-1 AGG-GGG Lee et al. 2006 ? Patho Patho Patho (!) Del Del Del Del no

P1400S

exon24 CCG-TCG

Xu and

Gutmann 1997 ?/M Patho Patho Patho (!) Del Del Del Del no

K1419R

exon24 AAG-AGG

Purandare et al

1994 ? Tol Patho Patho (!) NSI NSI NSI NSI no

N1430I

exon25 AAT-ATT Wallace lab,

unpublished ? Patho Patho Patho (!) Del Del Del Del no

N1430T

exon25 AAT-ACT

De Luca et al.

2005 ? Patho Patho Patho (!) Del Del Del Del no

V1432L

exon25 GTT-CTT

De Luca et al.

2005 ? Patho Patho Patho (!) Del Del Del Del no

Page 84: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

84

Table C-1. Continued

Mutation

and

location

Codon

change Reference Type SNPs3D SIFT

SIFT

8

mammal

MAPP

mammals

MAPP

mam/am/

ck

MAPP

mam/am/

ck/fsh

MAPP

mam/am/

ck/fs/ins

Splice

Site

R1590W

exon27b CGG-TGG

Upadhyaha et

al 1997 ? failed Patho Patho (!) Del Del Del Del no

V1621R

exon28 GTA-ATA

Jeong et al

2006 ? failed Patho Patho (!) Del Del Del Del no

D1623V

exon28 GAC-GTC

Wallace lab,

unpublished ? failed Patho Patho (!) Del Del Del Del no

A1764S

exon29 GCT-TCT Han et al 2001 ? failed Patho Tol Del Del Del Del no

T1787M

exon29 ACG-ATG Lee et al. 2006 ? failed Tol Tol NSI NSI Del Del no

L1812P

exon29 CTG-CCG

Wallace lab,

unpublished ? failed Tol Patho (!) Del Del Del Del no

C1909R

exon30 TGT-CGT Lee et al. 2006 ? failed Tol Patho (!) Del Del Del Del no

L1932P

exon31 CTG-CCG

Cawthon et al

1990 ? failed Patho Patho (!) Del Del Del Del no

R2129S

exon34 AGA-AGC

Upadhyaya et

al 2004 ? failed Tol Patho (!) Del Del Del Del no

Y2171D

exon34 TAT-GAT

Upadhyaya et

al 1992 ? failed Tol Patho (!) Del Del NSI Del no

S2739Y

exon48 TCT-TAT

Wallace lab,

unpublished ? failed Patho (!) Patho (!) Del Del Del Del no

Under “Type” P= Pathogenic, CS= Cryptic Splice, D/M= Detrimental, affects Microtubule binding, N= Neutral, and ?= Unknown.

For results returned by programs Patho=Pathogenic, Tol= Tolerated, Del= Deleterious, NSI= No Significant Impairment. For

results reported in the SIFT columns, (!)= a warning of low confidence in the prediction was returned by the program. Under

“Splice Site” A= Acceptor, D= Donor.

Page 85: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

85

LIST OF REFERENCES

Adams, M., Jones, J.L., Walker, R.A., Pringle, J.H., Bell, S.C., 2002. Changes in tenascin-C

isoform expression in invasive and preinvasive breast disease. Cancer Res. 62, 3289-97.

Ainsworth, C., 2005. Nonsense mutations: running the red light. Nature 438, 726-8.

Andersen, L.B., Ballester, R., Marchuk, D.A., Chang, E., Gutmann, D.H., Saulino, A.M.,

Camonis, J., Wigler, M., Collins, F.S., 1993. A conserved alternative splice in the von

Recklinghausen neurofibromatosis (NF1) gene produces two neurofibromin isoforms, both

of which have GTPase-activating protein activity. Mol. Cell Biol. 13, 487-495.

Ars, E., Kruyer, H., Morell, M., Pros, E., Serra, E., Ravella, A., Estivill, X., Lazaro, C., 2003.

Recurrent mutations in the NF1 gene are common among neurofibromatosis type 1

patients. J. Med. Genet. 40, e82.

Ars, E., Serra, E., García, J., Kruyer, H., Gaona, A., Lázaro, C., Estivill, X., 2000. Mutations

affecting mRNA splicing are the most common molecular defects in patients with

neurofibromatosis type 1. Hum. Mol. Genet. 9, 237-47. Erratum in: Hum Mol Genet 9,

659.

Baizer, L., Ciment, G., Hendrickson, S.K., Schafer, G.L., 1993. Regulated expression of the

neurofibromin type I transcript in the developing chicken brain. J. Neurochem. 61, 2054-

60.

Bajenaru, M.L., Hernandez, M.R., Perry, A., Zhu, Y., Parada, L.F., Garbow, J.R., Gutmann,

D.H., 2003. Optic nerve glioma in mice requires astrocyte Nf1 gene inactivation and Nf1

brain heterozygosity. Cancer Res. 63, 8573-7.

Ballester, R., Marchuk, D., Boguski, M., Saulino, A., Letcher, R., Wigler, M., Collins, F., 1990.

The NF1 locus encodes a protein functionally related to mammalian GAP and yeast IRA

proteins. Cell 63, 851-9.

Bartels, C.L., Tsongalis, G.J., 2009. MicroRNAs: novel biomarkers for human cancer. Clin.

Chem. 55, 623-31.

Baser, M.E., Contributors to the International NF2 Mutation Database, 2006. The distribution of

constitutional and somatic mutations in the neurofibromatosis 2 gene. Hum. Mutat. 27,

297-306.

Bird, A., 2002. DNA methylation patterns and epigenetic memory. Genes Dev. 16, 6–21.

Bird, A.P., 1986. CpG-rich islands and the function of DNA methylation. Nature 321, 209-13.

Bottillo, I., Ahlquist, T., Brekke, H., Danielsen, S.A., van den Berg, E., Mertens, F., Lothe, R.A.,

Dallapiccola, B., 2009. Germline and somatic NF1 mutations in sporadic and NF1-

associated malignant peripheral nerve sheath tumours. J. Pathol. 217, 693-701.

Page 86: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

86

Brown, T.C., Jiricny, J., 1987. A specific mismatch repair event protects mammalian cells from

loss of 5-methylcytosine. Cell 50, 945-50.

Cappione, A.J., French, B.L., Skuse, G.R., 1997. A potential role for NF1 mRNA editing in the

pathogenesis of NF1 tumors. Am. J. Hum. Genet. 60, 305-12.

Carey, J.C., Viskochil, D.H., 1999. Neurofibromatosis type 1: A model condition for the study of

the molecular basis of variable expressivity in human disorders. Am. J. Med. Genet. 89, 7-

13.

Cawthon, R.M., Weiss, R., Xu, G.F., Viskochil, D., Culver, M., Stevens, J., Robertson, M.,

Dunn, D., Gesteland, R., O'Connell, P., et al., 1990. A major segment of the

neurofibromatosis type 1 gene: cDNA sequence, genomic structure, and point mutations.

Cell 62, 193-201. Erratum in: Cell 1990 62, following 608.

Cichowski, K., Jacks, T., 2001. NF1 tumor suppressor gene function: narrowing the GAP. Cell

104, 593-604.

Chan, P.A., Duraisamy, S., Miller, P.J., Newell, J.A., McBride, C., Bond, J.P., Raevaara, T.,

Ollila, S., Nystrom, M., Grimm, A.J., Christodoulou, J., Oetting, W.S., Greenblatt, M.S.,

2007. Interpreting missense variants: comparing computational methods in human disease

genes CDKN2A, MLH1, MSH2, MECP2, and tyrosinase (TYR). Hum. Mutat. 28, 683-

693.

Clark, S.J., Harrison, J., Frommer, M., 1995. CpNpG methylation in mammalian cells. Nat.

Genet. 10, 20-7.

Colman, S.D., Rasmussen, S.A., Vu, T.H., Abernathy, C.R., Wallace, M.R., 1996. Somatin

mosaicism in a patient with neurofibromatosis type 1. Am. J. Hum. Genet. 58, 484-490.

Colman S.D., Williams C.A., Wallace M.R., 1995. Benign neurofibromas in type 1

neurofibromatosis (NF1) show somatic deletions of the NF1 gene. Nat. Genet. 11, 90-2.

Colman, S.D., Collins, F.S., Wallace, M.R., 1993. Characterization of a single base-pair deletion

in neurofibromatosis type 1. Hum. Mol. Genet. 2, 1709-11.

Cooper, D.N., Ball, E.V., Krawczak, M., 1998. The human gene mutation database. Nucleic

Acids Res. 26, 285-7.

Cooper, D.N., Krawczak, M., 1993. Human Gene Mutation. Oxford: BIOS Scientific Publishers.

Costa, R.M., Yang, T., Huynh, D.P., Pulst, S.M., Viskochil, D.H., Silva, A.J., Brannan, C.I.,

2001. Learning deficits, but normal development and tumor predisposition, in mice lacking

exon 23a of Nf1. Nat. Genet. 27, 399-405.

D'Angelo, I., Welti, S., Bonneau, F., Scheffzek, K., 2006. A novel bipartite phospholipid-binding

module in the neurofibromatosis type 1 protein. EMBO. Rep. 7, 174-9.

Page 87: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

87

Danglot, G., Régnier, V., Fauvet, D., Vassal, G., Kujas, M., Bernheim, A., 1995.

Neurofibromatosis 1 (NF1) mRNAs expressed in the central nervous system are

differentially spliced in the 5' part of the gene. Hum. Mol. Genet. 5, 915-20.

Daston, M.M., Scrable, H., Nordlund, M., Sturbaum, A.K., Nissen, L.M., Ratner, N., 1992. The

protein product of the neurofibromatosis type 1 gene is expressed at highest abundance in

neurons, Schwann cells, and oligodendrocytes. Neuron 8, 415-28.

DeClue, J.E., Cohen, B.D., Lowy, D.R., 1991. Identification and characterization of the

neurofibromatosis type 1 protein product. Proc. Natl. Acad. Sci. U S A. 88, 9914-8.

De Luca, A., Bottillo, I., Sarkozy, A., Carta, C., Neri, C., Bellacchio, E., Schirinzi, A., Conti, E.,

Zampino, G., Battaglia, A., Majore, S., Rinaldi, M.M., Carella, M., Marino, B., Pizzuti, A.,

Digilio, M.C., Tartaglia, M., Dallapiccola, B., 2005. NF1 gene mutations represent the

major molecular event underlying neurofibromatosis-Noonan syndrome. Am. J. Hum.

Genet. 77, 1092-101. Epub 2005 Oct 26.

De Luca, A., Buccino, A., Gianni, D., Mangino, M., Giustini, S., Richetta, A., Divona, L.,

Calvieri, S., Mingarelli, R., Dallapiccola, B., 2003. NF1 gene analysis based on DHPLC.

Hum. Mutat. 21, 171-2.

De Schepper, S., Maertens, O., Callens, T., Naeyaert, J.M., Lambert, J., Messiaen, L., 2008.

Somatic mutation analysis in NF1 café au lait spots reveals two NF1 hits in the

melanocytes. J. Invest. Dermatol. 128, 1050-3.

De Schepper, S., Boucneau, J.M., Westbroek, W., Mommaas, M., Onderwater, J., Messiaen, L.,

Naeyaert, J.M., Lambert, J.L., 2006. Neurofibromatosis type 1 protein and amyloid

precursor protein interact in normal human melanocytes and colocalize with melanosomes.

J Invest. Dermatol. 3, 653-9.

Dorschner, M.O., Sybert, V.P., Weaver, M., Pletcher, B.A., Stephens, K., 2000. NF1

microdeletion breakpoints are clustered at flanking repetitive sequences. Hum. Mol. Genet.

9, 35-46.

Driscoll, D.J., Migeon, B.R., 1990. Sex difference in methylation of single-copy genes in human

meiotic germ cells: implications for X chromosome inactivation, parental imprinting, and

origin of CpG mutations. Somat. Cell Mol. Genet. 16, 267-282.

Dublin, S., Riccardi, V.M., Stephens, K., 1995. Methods for rapid detection of a recurrent

nonsense mutation and documentation of phenotypic features in neurofibromatosis type 1

patients. Hum. Mutat. 5, 81-85.

Early, P., Rogers, J., Davis, M., Calame, K., Bond, M., Wall, R., Hood, L., 1980. Two mRNAs

can be produced from a single immunoglobulin mu gene by alternative RNA processing

pathways. Cell 20, 313-9.

Page 88: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

88

Easton, D.F., Ponder, M.A., Huson, S.M., Ponder, B.A., 1993. An analysis of variation in

expression of neurofibromatosis (NF) type 1 (NF1): evidence for modifying genes. Am. J.

Hum. Genet. 53, 305-13.

Eisenbarth, I., Beyer, K., Krone, W., Assum, G., 2000. Toward a survey of somatic mutation of

the NF1 gene in benign neurofibromas of patients with neurofibromatosis type 1. Am. J.

Hum. Genet. 66, 393-401.

Fahsold, R., Hoffmeyer, S., Mischung, C., Gille, C., Ehlers, C., Kucukceylan, N., Abdel-Nour,

M., Gewies, A., Peters, H., Kaufmann, D., Buske, A., Tinschert, S., Nurnberg, P., 2000.

Minor lesion mutational spectrum of the entire NF1 gene does not explain its high

mutability but points to a functional domain upstream of the GAP-related domain. Am. J.

Hum. Genet. 66, 790-818.

Ferner, R.E., Gutmann, D.H., 2002. International consensus statement on malignant peripheral

nerve sheath tumors in neurofibromatosis 1. Cancer Res. 62, 1573-1577.

Fishbein, L., Eady, B., Sanek, N., Muir, D., Wallace, M.R., 2005. Analysis of somatic NF1

promoter methylation in plexiform neurofibromas and Schwann cells. Cancer Genet.

Cytogenet. 157, 181-186.

Forbes, S.H., Dorschner, M.O., Le, R., Stephens, K., 2004. Genomic context of paralogous

recombination hotspots mediating recurrent NF1 region microdeletion. Genes

Chromosomes Cancer 41, 12-25.

Friedman, J.M., 1999. Clinical Genetics, in: Friedman, J.M., Gutmann, D.H., MacCollin, M.,

Riccardi, V.M. (eds.), Neurofibromatosis: Phenotype, Natural History, and Pathogenesis.

Baltimore, Johns Hopkins University Press.

Gao, Q.S., Memmott, J., Lafyatis, R., Stamm, S., Screaton, G., Andreadis, A., 2000. Complex

regulation of tau exon 10, whose missplicing causes frontotemporal dementia. J.

Neurochem. 74, 490-500.

Geist, R.T., Gutmann, D.H., 1996. Expression of a developmentally-regulated neuron-specific

isoform of the neurofibromatosis 1 (NF1) gene. Neurosci. Lett. 211, 85-8.

Girodon-Boulandet, E., Pantel, J., Cazeneuve, C., Gijn, M.V., Vidaud, D., Lemay, S., Martin, J.,

Zeller, J., Revuz, J., Goossens, M., Amselem, S., Wolkenstein, P., 2000. NF1 gene analysis

focused on CpG-rich exons in a cohort of 93 patients with neurofibromatosis type 1. Hum.

Mutat. 16, 274-5.

Gottfried, O.N., Viskochil, D.H., Fults, D.W., Couldwell, W.T., 2006. Molecular, genetic, and

cellular pathogenesis of neurofibromas and surgical implications. Neurosurgery 58, 1-16.

Greenblatt, M.S., Bennett, W.P., Hollstein, M., Harris, C.C., 1994. Mutations in the p53 tumor

suppressor gene: clues to cancer etiology and molecular pathogenesis. Cancer Res. 54,

4855-78.

Page 89: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

89

Gregory, P.E., Gutmann, D.H., Mitchell, A., Park, S., Boguski, M., Jacks, T., Wood, D.L., Jove,

R., Collins, F.S., 1993. Neurofibromatosis type 1 gene product (neurofibromin) associates

with microtubules. Somat. Cell Mol. Genet. 19, 265-74.

Gutmann, D.H., Aylsworth, A., Carey, J.C., Korf, B., Marks, J., Pyeritz, R.E., Rubenstein, A.,

Viskochil, D., 1997. The diagnostic evaluation and multidisciplinary management of

neurofibromatosis 1 and neurofibromatosis 2. JAMA 278, 51-7.

Gutmann, D.H., Geist, R.T., Wright, D.E., Snider, W.D., 1995. Expression of the

neurofibromatosis 1 (NF1) isoforms in developing and adult rat tissues. Cell Growth Dif.

6, 315-323.

Gutmann, D.H., Cole, J.L., Collins, F.S., 1994. Modulation of neurofibromatosis type 1 gene

expression during in vitro myoblast differentiation. J. Neuro. Res. 37, 398-405.

Gutman, D.H., Andersen, L.B., Cole, J.L., Swaroop, M., Collins, F.S., 1993a. An alternatively-

spliced mRNA in the carboxy terminus of the neurofibromatosis type 1 (NF1) gene is

expressed in muscle. Hum. Mol. Genet. 2, 989-92.

Gutmann, D.H., Boguski, M., Marchuk, D., Wigler, M., Collins, F.S., Ballester, R., 1993b.

Analysis of the neurofibromatosis type 1 (NF1) GAP-related domain by site-directed

mutagenesis. Oncogene 8, 761-9.

Gutmann, D.H., Wood, D.L., Collins, F.S., 1991. Identification of the neurofibromatosis type 1

gene product. Proc. Natl. Acad. Sci. U S A 88, 9658-62.

Han, S.S., Cooper, D.N., Upadhyaya, M.N., 2001. Evaluation of denaturing high performance

liquid chromatography (DHPLC) for the mutational analysis of the neurofibromatosis type

1 (NF1) gene. Hum. Genet. 109, 487-97. Epub 2001 Oct 11.

Hawes, J.J., Tuskan, R.G., Reilly, K.M., 2007. Nf1 expression is dependent on strain

background: implications for tumor suppressor haploinsufficiency studies. Neurogenetics

8, 121-30.

Henikiff, S., Henikoff, J.G., 1992. Amino acid substitution matrices from protein blocks. Proc.

Natl. Acad. Sci. U S A 89, 10915-10919.

Horiuchi, T., Hatta, N., Matsumoto, M., Ohtsuka, H., Collins, F.S., Kobayashi, Y., Fujita, S.,

1994. Nonsense mutations at Arg-1947 in two cases of familial neurofibromatosis type 1 in

Japanese. Hum. Genet. 93, 81-83.

Huang, J., Zheng, S., Jin, S.H., Zhang, S.Z., 2004. Somatic mutations of APC gene in

carcinomas from hereditary non-polyposis colorectal cancer patients. World J.

Gastroenterol. 10, 834-6.

Huynh, D.P., Nechiporuk, T., Pulst, S.M., 1994. Alternative transcripts in the mouse

neurofibromatosis type 2 (NF2) gene are conserved and code for schwannomins with

distinct C-terminal domains. Hum. Mol. Genet. 3, 1075-9.

Page 90: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

90

Jadayel, D., Fain, P., Upadhyaya, M., Ponder, M.A., Huson, S.M., Carey, J., Fryer, A., Mathew,

C.G., Barker, D.F., Ponder, B.A., 1990. Paternal origin of new mutations in von

Recklinghausen neurofibromatosis. Nature 343, 558-559.

Jeong, S.Y., Park, S.J., Kim, H.J., 2006. The spectrum of NF1 mutations in Korean patients with

neurofibromatosis type 1. J. Korean Med. Sci. 21, 107-12.

Jones, P.A., Buckley, J.D., Henderson, B.E., Ross, R.K., Pike, M.C., 1991. From gene to

carcinogen: a rapidly evolving field in molecular epidemiology. Cancer Res. 51, 3617-

3620.

Kaufmann, D., Müller, R., Kenner, O., Leistner, W., Hein, C., Vogel, W., Bartelt, B., 2002. The

N-terminal splice product NF1-10a-2 of the NF1 gene codes for a transmembrane segment.

Biochem. Biophys. Res. Commun. 294, 496-503.

Kayes, L.M., Burke, W., Riccardi, V.M., Ehrlich, P., Rubenstein, A., Stephens, K., 1994.

Deletions spanning the neurofibromatosis 1 gene: identification and phenotype of five

patients. Am. J. Hum. Genet. 54, 424-436.

Kayes, L.M., Riccardi, V.M., Burke, W., Bennett, R.L., Stephens, K., 1992. Large de novo DNA

deletion in a patient with sporadic neurofibromatosis 1, mental retardation, and

dysmorphism. J Med. Genet. 29, 686-690.

Kehrer-Sawatzki, H., Kluwe, L., Sandig, C., Kohn, M., Wimmer, K., Krammer, U., Peyrl, A.,

Jenne, D.E., Hansmann, I., Mautner, V.F., 2004. High frequency of mosaicism among

patients with neurofibromatosis type 1 (NF1) with microdeletions caused by somatic

recombination of the JJAZ1 gene. Am. J. Hum. Genet. 75, 410-23. Epub 2004 Jul 15.

Kluwe, L., Siebert, R., Gesk, S., Friedrich, R.E., Tinschert, S., Kehrer-Sawatzki, H., Mautner,

V.F., 2004. Screening 500 unselected neurofibromatosis 1 patients for deletions of the NF1

gene. Hum. Mutat. 23, 111-116.

Kluwe, L., Friedrich, R.E., Peiper, M., Friedman, J., Mautner, V.F., 2003. Constitutional NF1

mutations in neurofibromatosis 1 patients with malignant peripheral nerve sheath tumors.

Hum. Mutat. 22, 420.

Kluwe, L., Friedrich, R.E., Korf, B., Fahsold, R., Mautner, V.F., 2002. NF1 mutations in

neurofibromatosis 1 patients with plexiform neurofibromas. Hum. Mutat. 19, 309.

Knudson, A.G. Jr., 1971. Mutation and cancer: statistical study of retinoblastoma. Proc. Natl.

Acad. Sci. U.S.A. 68, 820-823.

Kobayashi, C., Oda, Y., Takahira, T., Izumi, T., Kawaguchi, K., Yamamoto, H., Tamiya, S.,

Yamada, T., Oda, S., Tanaka, K., Matsuda, S., Iwamoto, Y., Tsuneyoshi, M., 2006.

Chromosomal aberrations and microsatellite instability of malignant peripheral nerve

sheath tumors: a study of 10 tumors from nine patients. Cancer Genet. Cytogenet. 165, 98-

105.

Page 91: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

91

Krkljus, S., Abernathy, C.R., Johnson, J.S., Williams, C.A., Driscoll, D.J., Zori, R., Stalker, H.J.,

Rasmussen, S.A., Collins, F.S., Kousseff, B.G., Baumbach, L., Wallace, M.R., 1998.

Analysis of CpG C-to-T mutations in neurofibromatosis type 1. Mutations in brief no. 129.

Online. Hum. Mutat. 11, 411.

Kryukov, G., Pennacchio, L., Sunyaev, S., 2007. Most rare missense alleles are deleterious in

humans: implications for complex disease and association studies. Am. J. Hum. Genet. 80,

727-39. Epub 2007 Mar 8.

Kulyte, A., Dryselius, R., Karlsson, J., Good, L., 2005. Gene selective suppression of nonsense

termination using antisense agents. Biochem. Biophys. Acta. 1730, 165-72.

Lari, S.U., Famulski, K., Al-Khodairy, F., 2004. Multiplicity of strand incision at G:T base

mismatches in DNA by human cell extracts. Biochemistry 43, 6691-7.

Lari, S.U., Al-Khodairy, F., Paterson, M.C., 2002. Substrate specificity and sequence preference

of G:T mismatch repair: incision at G:T, O6-methylguanine:T, and G:U mispairs in DNA

by human cell extracts. Biochemistry 29, 9248-55.

Lazaro, C., Gaona, A., Ainsworth, P., Tenconi, R., Vidaud, D., Kruyer, H., Ars, E., Volpini, V.,

Estivill, X., 1996. Sex differences in mutational rate and mutational mechanism in the NF1

gene in neurofibromatosis type 1 patients. Hum. Genet. 98, 696-699.

Lazaro, C., Kruyer, H., Gaona, A., Estivill, X., 1995. Two further cases of mutation R1947X in

the NF1 gene: screening for a relatively common recurrent mutation. Hum. Genet. 3, 361-

363.

Leppig, K.A., Kaplan, P., Viskochil, D., Weaver, M., Orterberg, J., Stephens, K., 1997. Familial

neurofibromatosis 1 gene deletions: cosegregation with distinctive facial features and early

onset of cutaneous neurofibromas. Am. J. Med. Genet. 73, 197-204.

Leppig, K.A., Viskochil, D., Neil, S., Rubenstein, A., Johnson, V.P., Zhu, X.L., Brothman, A.R.,

Stephens, K., 1996. The detection of contiguous gene deletions at the neurofibromatosis 1

locus with fluorescence in situ hybridization. Cytogenet. Cell Genet. 72, 95-98.

Lee, M.J., Su, Y.N., You, H.L., Chiou, S.C., Lin, L.C., Yang, C.C., Lee, W.C., Hwu, W.L.,

Hsieh, F.J., Stephenson, D.A., Yu, C.L., 2006. Identification of forty-five novel and

twenty-three known NF1 mutations in Chinese patients with neurofibromatosis type 1.

Hum. Mutat. 27, 832.

Lee, M.P., Feinberg, A.P., 1997. Aberrant splicing but not mutations of TSG101 in human breast

cancer. Cancer Res. 57, 3131-4.

Li, Y., O’Connell, P., Breidenbach, H.H., Cawthon, R., Stevens, J., Xu, G., Neil, S., Robertson,

M., White, R., Viskochil, D., 1995. Genomic organization of the neurofibromatosis 1 gene

(NF1). Genomics 25, 9-18.

Page 92: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

92

Li, Y., Bollag, G., Clark, R., Stevens, J., Conroy, L., Fults, D., Ward, K., Friedman, E.,

Samowitz, W., Robertson, M., Bradley, P., McCormick, F., White, R., Cawthon, R., 1992.

Somatic mutations in the neurofibromatosis 1 gene in human tumors. Cell 69, 275-81.

Ling, J.Q., Li, T., Hu, J.F., Vu, T.H., Chen, H.L., Qiu, X.W., Cherry, A.M., Hoffman, A.R.,

2006. CTCF mediates interchromosomal colocalization between Igf2/H19 and Wsb1/Nf1.

Science 312, 269-72.

Loda-Hutchinson, R.L., Driscoll, D.J., Wallace, M.R., 2009. Comparison of computational

methods in predicting pathogenicity of NF1 missense mutations.

Lohmann, D.R., Brandt, B., Höpping, W., Passarge, E., Horsthemke, B., 1996. The spectrum of

RB1 germ-line mutations in hereditary retinoblastoma. Am. J. Hum. Genet. 58, 940-9.

Lopez-Correa, C., Dorschner, M., Brems, H., Lazaro, C., Clementi, M., Upadhyaya, M., Dooijes,

D., Moog, U., Kehrer-Sawatzki, H., Rutkowski, J.L., Fryns, J.P., Marynen, P., Stephens,

K., Legius, E., 2001. Recombinations hotspot in NF1 microdeletion patients. Hum. Mol.

Genet. 10, 1387-1392.

Luijten, M., Redeker, S., van Noesel, M.M., Troost, D., Westerveld, A., Hulsebos, T.J., 2000.

Microsatellite instability and promoter methylation as possible causes of NF1 gene

inactivation in neurofibromas. Eur. J. Hum. Genet. 8, 939-45.

MacCollin, M., Wallace, M.R., 2002. Molecular genetics of NF1 and NF2, in: Cowell, J.K. (ed.),

Molecular Genetics of Cancer, Second ed. Oxford BIOS Scientific Publishers, 95-113.

Maertens, O., Brems, H., Vandesompele, J., De Raedt, T., Heyns, I., Rosenbaum, T., De

Schepper, S., De Paepe, A., Mortier, G., Janssens, S., Speleman, F., Legius, E., Messiaen,

L., 2006. Comprehensive NF1 screening on cultured Schwann cells from neurofibromas.

Hum. Mutat. 27, 1030-40.

Malminen, M., Peltonen, S., Koivunen, J., Peltonen, J., 2002. Functional expression of NF1

tumor suppressor protein: association with keratin intermediate filaments during the early

development of human epidermis. BMC Dermatol. 2,10.

Mantani, A., Wakasugi, S., Yokota, Y., Abe, K., Ushio, Y., Yamamura, K., 1994. A novel

isoform of the neurofibromatosis type-1 mRNA and a switch of isoforms during murine

cell differentiation and proliferation. Gene 148, 245-51.

Martinez, R., Schackert, G., 2007. Epigenetic aberrations in malignant gliomas: an open door

leading to better understanding and treatment. Epigenetics 2,147-50.

Mathe, E., Olivier, M., Kato, S., Ishioka, C., Hainaut, P., Tavtigian, S.V., 2006. Computational

approaches for predicting the biological effect of p53 missense mutations: a comparison of

three sequence analysis based tools. Nucleic Acids Res. 34, 1317-1325.

Page 93: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

93

Mattocks, C., Baralle, D., Tarpey, P., ffrench-Constant, C., Bobrow, M., Whittaker, J., 2004.

Automated comparative sequence analysis identifies mutations in 89% of NF1 patients and

confirms a mutation cluster in exons 11-17 distinct from the GAP related domain. J. Med.

Genet. 41, e48.

Maynard, J., Krawczak, M., Upadhyaya, M., 1997. Characterization and significance of nine

novel mutations in exon 16 of the neurofibromatosis type 1 (NF1) gene. Hum. Genet. 99,

674-6.

McClatchey, A,I., 2007. Neurofibromatosis. Annu. Rev. Pathol. 2, 191-216.

Messiaen, L.M., Callens, T., Mortier, G., Beysen, D., Vandenbroucke, I., Van Roy, N.,

Speleman, F., Paepe, A.D., 2000. Exhaustive mutation analysis of the NF1 gene allows

identification of 95% of mutations and reveals a high frequency of unusual splicing

defects. Hum. Mutat. 15, 541-55.

Messiaen, L.M., Callens, T., Roux, K.J., Mortier, G.R., De Paepe, A., Abramowicz, M., Pericak-

Vance, M.A., Vance, J.M., Wallace, M.R., 1999. Exon 10b of the NF1 gene represents a

mutational hotspot and harbors a recurrent missense mutation Y489C associated with

aberrant splicing. Genet. Med. 1, 248-53.

Mukhopadhyay, D., Anant, S., Lee, R.M., Kennedy, S., Viskochil, D., Davidson, N.O., 2002. C--

>U editing of neurofibromatosis 1 mRNA occurs in tumors that express both the type II

transcript and apobec-1, the catalytic subunit of the apolipoprotein B mRNA-editing

enzyme. Am. J. Hum. Genet. 70, 38-50.

Nagoshi, R.N., McKeown, M., Burtis, K.C., Belote, J.M., Baker, B.S., 1988. The control of

alternative splicing at genes regulating sexual differentiation in D. melanogaster. Cell 53,

229-36.

Ng, P.C., Henikoff, S., 2001. Predicting deleterious amino acid substitutions. Genome Res. 11,

863-874.

Ninio, M., Privman, E., Pupko, T., Friedman, N., 2007. Phylogeny reconstruction: increasing the

accuracy of pairwise distance estimation using Bayesian inference of evolutionary rates.

Bioinformatics 23, e136-41.

Nishi, T., Lee, P.S., Oka, K., Levin, V.A., Tanase, S., Morino, Y., Saya, H., 1991. Differential

expression of two types of the neurofibromatosis type 1 (NF1) gene transcripts related to

neuronal differentiation. Oncogene. 6, 1555-1559.

Nishi, T., Saya, H., 1991. Neurofibromatosis type 1 (NF1) gene: implication in neuroectodermal

differentiation and genesis of brain tumors. Cancer Metastasis Rev. 10, 301-10.

O'Brien, K.P., Remm, M., Sonnhammer, E.L., 2005. Inparanoid: a comprehensive database of

eukaryotic orthologs. Nucleic Acids Res. 33(Database issue), D476-80.

Page 94: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

94

Ogata, H., Sato, H., Takatsuka, J., De Luca, L.M., 2001. Human breast cancer MDA-MB-231

cells fail to express the neurofibromin protein, lack its type I mRNA isoform and show

accumulation of P-MAPK and activated Ras. Cancer Letters. 172, 159-164.

Olivier, M., Eeles, R., Hollstein, M., Khan, M.A., Harris, C.C., Hainaut, P., 2002. The IARC

TP53 database: new online mutation analysis and recommendations to users. Hum. Mutat.

19, 607-614.

Ottini, L., Esposito, D.L., Richetta, A., Carlesimo, M., Palmirotta, R., Verí, M.C., Battista, P.,

Frati, L., Caramia, F.G., Calvieri, S., et al., 1995. Alterations of microsatellites in

neurofibromas of von Recklinghausen's disease. Cancer Res. 55, 5677-80.

Park, V.M., Pivnick, E.K., 1998. Neurofibromatosis type 1 (NF1): a protein truncation assay

yielding identification of mutations in 73% of patients. J. Med. Genet. 35, 813-20.

Peters, H., Hess, D., Fahsold, R., Schülke, M., 1999. A novel mutation L1425P in the GAP-

region of the NF1 gene detected by temperature gradient gel electrophoresis (TGGE).

Mutation in brief no. 230. Online. Hum. Mutat. 13, 337.

Pinotti, M., Rizzotto, L., Chuansumrit, A., Mariani, G., Bernardi, F., International Factor VII

Deficiency Study Group, 2006. Gentamicin induces sub-therapeutic levels of coagulation

factor VII in patients with nonsense mutations. J. Thromb. Haemost. 4, 1828-30.

Poullet, P., Lin, B., Esson, K., Tamanoi, F., 1994. Functional significance of lysine 1423 of

neurofibromin and characterization of a second site suppressor which rescues mutations at

this residue and suppresses RAS2Val-19-activated phenotypes. Mol. Cell Biol. 14, 815-21.

Pros, E., Gómez, C., Martín, T., Fábregas, P., Serra, E., Lázaro, C., 2008. Nature and mRNA

effect of 282 different NF1 point mutations: focus on splicing alterations. Hum. Mutat. 29,

E173-E193. [Epub ahead of print]

Purandare, S.M., Lanyon, W.G., Connor, J.M., 1994. Characterisation of inherited and sporadic

mutations in neurofibromatosis type-1. Hum. Mol. Genet. 3, 1109-15.

Raedt, T.D., Stephens, M., Heyns, I., Brems, H., Thijs, D., Messiaen, L., Stephens, K., Lazaro,

C., Wimmer, K., Kehrer-Sawatzki, H., Vidaud, D., Kluwe, L., Marynen, P., Legius, E.,

2006. Conservation of hotspots for recombination in low-copy repeats associated with the

NF1 microdeletion. Nat. Genet. 38, 1419-23. Epub 2006 Nov 19.

Rasmussen, S.A., Colman, S.D., Ho, V.T., Abernathy, C.R., Arn, P.H., Weiss, L., Schwartz, C.,

Saul, R.A., Wallace, M.R., 1998. Constitutional and mosaic large NF1 gene deletions in

neurofibromatosis type 1. J. Med. Genet. 35, 468-471.

Rasmussen, S.A., Wallace, M.R., 1998. Somatic mutations of the NF1 gene in type 1

neurofibromatosis and cancer, in: Upadhyaya, M., Cooper, D.N. (eds.), Neurofibromatosis

Type 1: from Genotype to Phenotype, Oxford: BIOS Scientific Publishers, 153-165.

Page 95: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

95

Rasmussen, S.A., Overman, J., Thomson, S.A., Colman, S.D., Abernathy, C.R., Trimpert, R.E.,

Moose, R., Virdi, G., Roux, K., Bauer, M., Rojiani, A.M., Maria, B.L., Muir, D., Wallace,

M.R., 2000. Chromosome 17 loss-of-heterozygosity studies in benign and malignant

tumors in neurofibromatosis type 1. Genes Chromosomes Cancer 28, 425-31.

Razin, A. and Riggs, A. D., 1980. DNA methylation and gene function. Science 210, 604–610.

Reese, M.G., Eeckman, F.H., Kulp, D., Haussler, D., 1997. Improved splice site detection in

Genie. J. Comput. Biol. 4, 311-23.

Riccardi, V.M., 1992. The prenatal diagnosis of NF-1 and NF-2. J. Dermatol. 19, 885-91.

Roudebush, M., Slabe, T., Sundaram, V., Hoppel, C.L., Golubic, M., Stacey, D.W., 1997.

Neurofibromin colocalizes with mitochondria in cultured cells. Exp. Cell Res. 236, 161-72.

Sawada, S., Florell, S., Purandare, S.M., Ota, M., Stephens, K., Viskochil, D., 1996.

Identification of NF1 mutations in both allales of a dermal neurofibroma. Nat. Genet. 14,

110-112.

Serra, E., Rosenbaum, T., Nadal, M., Winner, U., Ars, E., Estivill, X., Lázaro, C., 2001a. Mitotic

recombination effects homozygosity for NF1 germline mutations in neurofibromas. Nat.

Genet. 28, 294-6. Erratum in: Nat Genet 2001 29, 100.

Serra, E., Ars, E., Ravella, A., Sanchez, A., Puig, S., Rosenbaum, T., Estivill, X., Lazaro, C.,

2001b. Somatic NF1 mutational spectrum in benign neurofibromas: mRNA splice defects

are common among point mutations. Hum. Genet. 108, 416-429

Serra, E., Rosenbaum, T., Winner, U., Aledo, R., Ars, E., Estivill, X., Lenard, H., Lazaro, C.,

2000. Schwann cells harbor the somatic NF1 mutation in neurofibromas: evidence of two

different Schwann cell subpopulations. Hum. Mol. Genet. 9, 3055-3064.

Serra, E., Puig, S., Otero, D., Gaona, A., Kruyer, H., Ars, E., Estivill, X., Lazaro, C., 1997.

Confirmation of a double-hit model for the NF1 Gene in benign neurofibromas. Am. J.

Hum. Genet. 61, 512-519.

Shelton, B.P., Misso, N.L., Shaw, O.M., Arthaningtyas, E., Bhoola, K.D., 2008. Epigenetic

regulation of human epithelial cell cancers. Curr. Opin. Mol. Ther. 10, 568-78.

Skuse, G.R., Cappione, A.J., 1997. RNA processing and clinical variability in neurofibromatosis

type I (NF1). Hum. Mol. Genet. 6, 1707-1712.

Skuse, G.R., Cappione, A.J., Sowden, M., Metheny, L.J., Smith, H.C., 1996. The

neurofibromatosis type I messenger RNA undergoes base-modification RNA editing.

Nucleic Acids Res. 24, 478-85.

Skuse, G.R., Kosciolek, B.A., Rowley, P.T., 1991. The neurofibroma in von Recklinghausen

neurofibromatosis has a unicellular origin. Am. J. Hum. Genet. 49, 600-607.

Page 96: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

96

Smith, C.W., Patton, J.G., Nadal-Ginard, B., 1989. Alternative splicing in the control of gene

expression. Annu. Rev. Genet. 23, 527-77.

Stephens, K., Kayes, L., Riccardi, V.M., Rising, M., Sybert, V.P., Pagon, R.A., 1992.

Preferential mutation of the neurofibromatosis type 1 gene in paternally derived

chromosomes. Hum. Genet. 88, 279-82.

Stevenson, D.A., Zhou, H., Ashrafi, S., Messiaen, L.M., Carey, J.C., D'Astous, J.L., Santora,

S.D., Viskochil, D.H., 2006. Double inactivation of NF1 in tibial pseudarthrosis. Am. J.

Hum. Genet. 79, 143-8. Epub 2006 May 10.

Stimpfl, M., Tong, D., Fasching, B., Schuster, E., Obermair, A., Leodolter, S., Zeillinger, R.,

2002. Vascular endothelial growth factor splice variants and their prognostic value in

breast and ovarian cancer. Clin. Cancer Res. 8, 2253-9.

Stone, E.A., Sidow, A., 2005. Physicochemical constraint violation by missense substitutions

mediates impairment of protein function and disease severity. Genome Res. 15, 978-86.

Epub 2005 Jun 17.

Storlazzi, C.T., Von Steyern, F.V., Domanski, H.A., Mandahl, N., Mertens, F., 2005. Biallelic

somatic inactivation of the NF1 gene through chromosomal translocations in a sporadic

neurofibroma. Int. J. Cancer 117,1055-7.

Suzuki, H., Takahashi, K., Kubota, Y., Shibahara, S., 1992. Molecular cloning of a cDNA

coding for neurofibromatosis type 1 protein isoform lacking the domain related to ras

GTPase-activating protein. Biochem. Biophys. Res. Commun. 187, 984-990.

Szudek, J., Joe, H., Friedman, J.M., 2002. Analysis of intrafamilial phenotypic variation in

neurofibromatosis 1 (NF1). Genet. Epidemiol. 23, 150-64.

Teinturier, C., Danglot, G., Slim, R., Pruliere, D., Launay, J.M., Bernheim, A., 1992. The

neurofibromatosis 1 gene transcripts expressed in peripheral nerve and neurofibromas bear

the additional exon located in the GAP domain. Biochem. Biophys. Res. Commun. 188,

851-857.

Thomson, S.A., Fishbein, L., Wallace, M.R., 2002. NF1 mutations and molecular testing. J.

Child. Neurol. 17, 555-561.

Toledano, H., Goldberg, Y., Kedar-Barnes, I., Baris, H., Porat, R.M., Shochat, C., Bercovich, D.,

Pikarsky, E., Lerer, I., Yaniv, I., Abeliovich, D., Peretz, T., 2008. Homozygosity of MSH2

c.1906G-->C germline mutation is associated with childhood colon cancer, astrocytoma

and signs of Neurofibromatosis type I. Fam. Cancer Dec 20.

Tonsgard, J.H., Kwak, S.M., Short, M.P., Dachman, A.H., 1998. CT imaging in adults with

neurofibromatosis-1: frequent asymptomatic plexiform lesions. Neurology 50, 1755-60.

Page 97: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

97

Uchida, T., Matozaki, T., Suzuki, T., Matsuda, K., Wada, K., Nakano, O., Konda, Y., Nishisaki,

H., Nagao, M., Sakamoto, C., et al., 1992. Expression of two types of neurofibromatosis

type 1 gene transcripts in gastric cancers and comparison of GAP activities. Biochem.

Biophys. Res. Commun. 187, 332-9.

Upadhyaya, M., Spurlock, G., Monem, B., Thomas, N., Friedrich, R.E., Kluwe, L., Mautner, V.,

2008. Germline and somatic NF1 gene mutations in plexiform neurofibromas. Hum.

Mutat. 29, E103-E111

Upadhyaya, M., Han, S., Consoli, C., Majounie, E., Horan, M., Thomas, N.S., Potts, C.,

Griffiths, S., Ruggieri, M., von Deimling, A., Cooper, D.N., 2004. Characterization of the

somatic mutational spectrum of the neurofibromatosis type 1 (NF1) gene in

neurofibromatosis patients with benign and malignant tumors. Hum. Mutat. 23, 134-46.

Upadhyaya, M., Cooper, D.N., 1998. The mutational spectrum in neurofibromatosis 1 and its

underlying mechanisms, in: Upadhyaya, M., Cooper, D.N., (eds.), Neurofibromatosis Type

1: from Genotype to Phenotype, Oxford: BIOS Scientific Publishers, 65-88.

Upadhyaya, M., Ruggieri, M., Maynard, J., Osborn, M., Hartog, C., Mudd, S., Penttinen, M.,

Cordeiro, I., Ponder, M., Ponder, B.A., Krawczak, M., Cooper, D.N., 1998. Gross

deletions of the neurofibromatosis type 1 (NF1) gene are predominantly of maternal origin

and commonly associated with a learning disability, dysmorphic features and

developmental delay. Hum. Genet. 102, 591-597.

Upadhyaya, M., Maynard, J., Osborn, M., Harper, P.S., 1997. Six novel mutations in the

neurofibromatosis type 1 (NF1) gene. Hum. Mutat. 10, 248-50

Upadhyaya, M., Osborn, M.J., Maynard, J., Kim, M.R., Tamanoi, F., Cooper, D.N., 1997.

Mutational and functional analysis of the neurofibromatosis type 1 (NF1) gene. Hum.

Genet. 99, 88-92.

Upadhyaya, M., Maynard, J., Osborn, M., Huson, S.M., Ponder, M., Ponder, B.A., Harper, P.S.,

1995. Characterisation of germline mutations in the neurofibromatosis type 1 (NF1) gene.

J. Med. Genet. 32, 706-10.

Upadhyaya, M., Shen, M., Cherryson, A., Farnham, J., Maynard, J., Huson, S.M., Harper, P.S.,

1992. Analysis of mutations at the neurofibromatosis 1 (NF1) locus. Hum. Mol. Genet. 1,

735-40.

Valero, M.C., Pascual-Castroviejo, I., Velasco, E., Moreno, F., Hernandez-Chico, C., 1997.

Identification of de novo deletions at the NF1 gene: no preferential paternal origin and

phenotypic analysis of patients. Hum. Genet. 99, 720-726.

Vandenbroucke, I., Van Oostveldt, P., Coene, E., De Paepe, A., Messiaen, L., 2004.

Neurofibromin is actively transported to the nucleus. FEBS Lett. 560, 98-102.

Vandenbroucke, I.I., Vandesompele, J., Paepe, A.D., Messiaen, L., 2001. Quantification of splice

variants using real-time PCR. Nucleic Acids Res. 29, E68-8.

Page 98: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

98

Viskochil, D., Buchberg, A.M., Xu, G., Cawthon, R.M., Stevens, J., Wolff, R.K., Culver, M.,

Carey, J.C., Copeland, N.G., Jenkins, N.A., White, R., O'Connell, P., 1990. Deletions and

a translocation interrupt a cloned gene at the neurofibromatosis type 1 locus. Cell 62, 87-

192.

Wallace, M.R., Rasmussen, S.A., Lim, I.T., Gray, B.A., Zori, R.T., Muir, D., 2000. Culture of

cytogenetically abnormal schwann cells from benign and malignant NF1 tumors. Genes

Chromosomes Cancer 27,117-23.

Wallace, M.R., Andersen, L.B., Saulino, A.M., Gregory, P.E., Glover, T.W., Collins, F.S., 1991.

A de novo Alu insertion results in neurofibromatosis type 1. Nature 353, 864-866.

Wallace, M.R., Marchuk, D.A., Andersen, L.B., Letcher, R., Odeh, H.M., Saulino, A.M.,

Fountain, J.W., Brereton, A., Nicholson, J., Mitchell, A.L., Brownstein, B.H., Collins, F.S.,

1990. Type 1 neurofibromatosis gene: identification of a large transcript disrupted in three

NF1 patients. Science 249, 181-186.

Wang, Q., Montmain, G., Ruano, E., Upadhyaya, M., Dudley, S., Liskay, R.M., Thibodeau, S.N.,

Puisieux, A., 2003. Neurofibromatosis type 1 gene as a mutational target in a mismatch

repair-deficient cell type. Hum. Genet. 112, 117-23.

Wang, Z., Lo, H.S., Yang, H., Gere, S., Hu, Y., Buetow, K.H., Lee, M.P., 2003. Computational

analysis and experimental validation of tumor-associated alternative RNA splicing in

human cancer. Cancer Res. 633, 655-7.

Wiest, V., Eisenbarth, I., Schmegner, C., Krone, W., Assum, G., 2003. Somatic NF1 mutation

spectra in a family with neurofibromatosis type 1: toward a theory of genetic modifiers.

Hum. Mut. 22, 423-427.

Williams, V.C., Lucas, J., Babcock, M.A., Gutmann, D.H., Korf, B., Maria, B.L., 2009.

Neurofibromatosis type 1 revisited. Pediatrics 123, 124-33.

Wimmer, K., Roca, X., Beiglbock, H., Callens, T., Etzler, J., Rao, A.R., Krainer, A.R., Fonatsch,

C., Messiaen, L., 2007. Extensive in silico analysis of NF1 splicing defects uncovers

determinants for splicing outcome upon 5' splice-site disruption. Hum. Mutat. 28, 599-612.

Wu, R., López-Correa, C., Rutkowski, J.L., Baumbach, L.L., Glover, T.W., Legius, E., 1999.

Germline mutations in NF1 patients with malignancies. Genes Chromosomes Cancer 26,

376-80.

Wu, B.L., Schneider, G.H., Korf, B.R., 1997. Deletion of the entire NF1 gene causing distinct

manifestations in a family. Am. J. Med. Genet. 69, 98-101.

Wu, B.L., Austin, M.A., Schneider, G.H., Boles, R.G., Korf, B.R., 1995. Deletion of the entire

NF1 gene detected by the FISH: four deletion patients associated with severe

manifestations. Am. J. Med. Genet. 59, 528-535.

Page 99: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

99

Xu, G.F., O'Connell, P., Viskochil, D., Cawthon, R., Robertson, M., Culver, M., Dunn, D.,

Stevens, J., Gesteland, R., White, R., et al., 1990. The neurofibromatosis type 1 gene

encodes a protein related to GAP. Cell 62, 599-608.

Xu, H., Gutmann, D.H., 1997. Mutations in the GAP-related domain impair the ability of

neurofibromin to associate with microtubules. Brain Res. 759, 149-52.

Yue, P., Melamud, E., Moult, J., 2006. SNPs3D: Candidate Gene and SNP selection for

Association Studies. BMC Bioinformatics 7, 166.

Yue, P., Moult, J., 2006. Identification and analysis of deleterious human SNPs. J. Mol. Biol.

356,1263-74. Epub 2005 Dec 27.

Zatkova, A., Messiaen, L., Vandenbroucke, I., Wieser, R., Fonatsch, C., Krainer, A.R., Wimmer,

K., 2004. Disruption of exonic splicing enhancer elements is the principal cause of exon

skipping associated with seven nonsense or missense alleles of NF1. Hum. Mutat. 24, 491-

501.

Zhou, J., Yu, Q., Zou, T., 2008. Alternative splicing of exon 10 in the tau gene as a target for

treatment of tauopathies. BMC Neurosci. 9(Suppl 2), S10.

Zhu, C., Saberwal, G., Lu, Y., Platanias, L.C., Eklund, E.A., 2004. The interferon consensus

sequence-binding protein activates transcription of the gene encoding neurofibromin 1. J.

Biol. Chem. 279, 50874-85.

Zhu, H., Hinman, M.N., Hasman, R.A., Mehta, P., Lou, H., 2008. Regulation of neuron-specific

alternative splicing of neurofibromatosis type 1 pre-mRNA. Mol. Cell Biol. 28, 1240-51

Page 100: ANALYSIS OF NF1 MUTATION MECHANISMSufdcimages.uflib.ufl.edu/UF/E0/02/43/64/00001/loda_r.pdf · NF1 gene mutations, which occur at a rate 10x higher then the genome ... (must be 0.5

BIOGRAPHICAL SKETCH

Rebecca Loda-Hutchinson grew up in northern Nevada, where her interest in genetics

began in her high school biology classes. After graduating valedictorian of the class of 2000, she

attended the University of Nevada, Reno, majoring in biology. During her time at UNR, she

participated in several summer research opportunities, the first of which was in the laboratory of

Dr. John Postlethwait at the University of Oregon, studying armor formation in threespine

stickleback. The second was in the laboratory of Dr. Lee Weber and Dr. Eileen Hickey at UNR.

Dr. Weber was her honor’s thesis advisor, overseeing her investigation of using SNPs in the

HSP30 gene to detect hybrid populations of cutthroat trout. She graduated Summa cum Laude

with a bachelor’s degree in biology in 2004. Rebecca’s love of learning and desire to add to the

knowledge of others had led her to apply to graduate school, and, in the fall of 2004, she began

her studies at the University of Florida. While all her previous research involved fish, Rebecca

was excited to make the transition to studying human genetics, and joined the laboratory of Dr.

Margaret R. Wallace, where her doctoral research focused on Neurofibromatosis 1. Now that

she has graduated with her PhD, Rebecca isn’t sure which direction her career path will take, but

is looking forward to having a little more time to spend with her husband, Lance, and their two

rescue kitties.


Recommended