<<

The Compartmentalization of Folate

in Mammalian Cells

by

Harshila Patel

A thesis submitted to the Faculty of Graduate Studies and Research in partial

fulfillment of the requirements of the degree of Doctor of Philosophy

Department of Biochemistry

McGili University

Montreal, Quebec, Canada

© by Harshila Patel

June, 2004 Library and Bibliothèque et 1+1 Archives Canada Archives Canada Published Heritage Direction du Branch Patrimoine de l'édition

395 Wellington Street 395, rue Wellington Ottawa ON K1A ON4 Ottawa ON K1A ON4 Canada Canada

Your file Votre référence ISBN: 0-494-06334-3 Our file Notre référence ISBN: 0-494-06334-3

NOTICE: AVIS: The author has granted a non­ L'auteur a accordé une licence non exclusive exclusive license allowing Library permettant à la Bibliothèque et Archives and Archives Canada to reproduce, Canada de reproduire, publier, archiver, publish, archive, preserve, conserve, sauvegarder, conserver, transmettre au public communicate to the public by par télécommunication ou par l'Internet, prêter, telecommunication or on the Internet, distribuer et vendre des thèses partout dans loan, distribute and sell th es es le monde, à des fins commerciales ou autres, worldwide, for commercial or non­ sur support microforme, papier, électronique commercial purposes, in microform, et/ou autres formats. paper, electronic and/or any other formats.

The author retains copyright L'auteur conserve la propriété du droit d'auteur ownership and moral rights in et des droits moraux qui protège cette thèse. this thesis. Neither the thesis Ni la thèse ni des extraits substantiels de nor substantial extracts from it celle-ci ne doivent être imprimés ou autrement may be printed or otherwise reproduits sans son autorisation. reproduced without the author's permission.

ln compliance with the Canadian Conformément à la loi canadienne Privacy Act some supporting sur la protection de la vie privée, forms may have been removed quelques formulaires secondaires from this thesis. ont été enlevés de cette thèse.

While these forms may be included Bien que ces formulaires in the document page count, aient inclus dans la pagination, their removal does not represent il n'y aura aucun contenu manquant. any loss of content from the thesis. ••• Canada This thesis is dedicated to my wonderfully supportive parents,

Ashvin and Jyoti ABSTRACT

Folate metabolism is compartmentalized between the cytoplasm and the mitochondria of mammalian cells. Certain folate-dependent are present in both of these compartments, such as methylenetetrahydrofolate dehydrogenases, which are required to interconvert one-carbon substituted tetrahydrofolates. In the cytoplasm, there is a trifunctional NADP-dependent methylenetetrahydrofolate dehydrogenase-cyclohydrolase-synthetase (DCS). Its mitochondrial counterpart is a bifunctional NAD-dependent methylenetetrahydrofolate dehydrogenase-cyclohydrolase (NMDMC), expressed during embryogenesis and in immortalized cells. A comparison of the 3' untranslated region of the NMDMC cDNA with the synthetase domain of the DCS cDNA and gene among different species have revealed significant regions of homology. This suggests that the mammalian mitochondrial NAD-dependent OC evolved from an NADP-dependent DCS precursor through a change in cofactor specificity of the dehydrogenase from NADP to NAD.

Although the folate pathways are compartmentalized, the mitochondrial folate pathway makes an important contribution to total cellular folate metabolism.

Mouse fibroblasts that have a completely inactivated NMDMC gene are auxotrophs. Furthermore, growth of these Nmdmc-/- cell lines is stimulated by supplementation with formate or hypoxanthine. These cell lines also show enhanced incorporation of radioactivity into DNA from formate as compared to , demonstrating that formate is a preferred one-carbon donor for the Nmdmc-/- cell lines. This indicates that NMDMC is required for optimal purine biosynthesis during periods of rapid cellular proliferation su ch as embryogenesis and tumorigenesis. The rescue of these Nmdmc-/- cell lines with NMDMC expression reversed the glycine auxotrophy and the formate to serine incorporation ratio reverted toward the wild type ratio. The rescue of these

Nmdmc-/- cell lines with the NAD-dependent monofunctional dehydrogenase activity also reversed the glycine auxotrophy but these cell lines did not grow as weil as the NMDMC-rescued ce Il lines. This indicates that although the cyclohydrolase activity is not required in the mitochondria, the rate of 10- formylTHF production is not optimal in its absence. Furthermore, when these

Nmdmc-/- cell lines were rescued with the expression of the NADP-dependent

DCS in the mitochondria there was reversai of the glycine auxotrophy as weil, indicating that the NAD cofactor specificity of the mitochondrial methylenetetrahydrofolate dehydrogenase is not absolutely essential to maintain the flux of one-carbon metabolites. RÉSUMÉ

Dans les cellules de mammifères, le métabolisme du folate est compartimenté entre le cytoplasme et les mitochondries. Certaines enzymes qui dépendent du folate sont présentes dans ces deux compartiments, comme par exemple, les déshydrogénases de méthylènetétrahydrofolate nécessaire à l'interconversion des tétrahydrofolates porteurs d'unité de carbone. Le cytoplasme contient une trifonctionnelle, la déshydrogénase de méthylènetétrahydrofolate dépendante du NADP-cyclohydrolase-synthétase

(DCS) dont l'équivalent mitochondrial est une enzyme bifonctionnelle: la déshydrogénase de méthylènetétrahydrofolate dépendante du NAD­ cyclohydrolase (NMDMC). Cette dernière est exprimée durant l'embryogenèse et dans les lignées de cellules transformées. Une comparaison de la région 3' non traduite de l'ADNe de NMDMC avec le domaine de synthétase de l'ADNe et du gène de DCS provenant d'espèces différentes a indiqué des régions significativement homologues. Ceci suggère que l'enzyme mitochondriale de mammifères, la OC dépendante du NAD, aurait évoluée à partir d'un précurseur de la DCS dépendante du NADP ayant subi un changement de cofacteur, passant du NADP au NAD.

Bien que les voies de signalisation du folate soient compartimentées, la voie mitochondriale apporte une contribution importante au métabolisme cellulaire du folate. Les fibroblastes de souris dont le gène NMDMC est complètement inactivé sont auxotrophes pour la glycine. De plus, la croissance de ces cellules I Nmdmc- - est stimulée par la présence de formate ou d'hypoxanthine. Ces cellules montrent également plus d'incorporation de radioactivité à l'ADN lorsqu'on utilise du formate au lieu de la sérine, ce qui démontre l'utilisation préferentielle du formate en tant que donneur d'unités de carbone pour ces cellules. Les résultats indiquent que la NMDMC est nécessaire à la biosynthèse optimale des purines pendant les périodes de prolifération cellulaire rapide telles

I que l'embryogenèse ou la tumorigenèse. Lorsque dans ces cellules Nmdmc- - on restaure la NMDMC par transfection, l'auxotrophie pour la glycine est renversée et le rapport d'incorporation préferentielle du formate par rapport à la sérine revient à la normale. Lorsque l'activité monofonctionnelle de déshydrogénase à

I NAD est restaurée dans les cellules Nmdmc- -, l'auxotrophie pour la glycine est renversée, mais les lignées cellulaires ne prolifèrent pas aussi bien que les

I cellules Nmdmc- - exprimant de nouveau la NMDMC. Ceci indique que, bien que

l'activité de la cyclohydrolase ne soit pas nécessaire dans les mitochondries, le taux de production du 10-formyltétrahydrofolate n'est pas optimal en son absence.

De plus, quand on exprime la DCS dépendante du NADP dans les mitochondries

I des cellules Nmdmc- - par transfection, l'auxotrophie pour la glycine est de

nouveau renversée, indicant que la spécificité pour le NAD de la déshydrogénase

de méthylènetétrahydrofolate mitochondriale n'est pas absolument essentielle au

maintient du flot des métabolites d'unité de carbone. TABLE OF CONTENTS

FOREWORD ...... i

ACKNOWLEDGEMENTS ...... v

PUBLICATION OF THE WORK PRESENTED IN THIS THESIS ...... viii

CONTRIBUTIONS TO ORIGINAL KNOWLEDGE ...... ix

LIST OF FIGURES ...... xii

LIST OF TABLES ...... xiv

LIST OF ABBREVIATIONS ...... xv

CHAPTER ONE: GENERAL INTRODUCTION ...... 1

OISCOVERY OF FOLIC ACIO ...... 2

FOLIC ACIO ...... 2

FOLATE ABSORPTION ...... 4

MECHANISMS OF CELLULAR TRANSPORT ...... 5

FOLATE BINDING PROTEINS ...... 6

Folate Receptor ...... 6

Reduced Folate Carrier ...... 9

LOW PH TRANSPORTER ...... 11

FOLATE POL YGLUTAMYLATION ...... 12

SOURCES OF ONE-CARBON UNITS ...... 13

SERINE ...... 13

GLYCINE ...... 14

HISTIDINE ...... 14 CHOLINE ...... 15

FORMATE ...... 16

OVERVIEW OF FOLATE-MEOIATEO METABOLISM ...... 18

HOMOCYSTEINE REMETHYLATION CYCLE ...... 21

Methionine Regeneration ...... 21

Transsulfuration Of Homocysteine ...... 23

Regulation Of The Methylation Cycle ...... 23

Disruptions ln The Methylation Cycle ...... 25

DE NOVO THYMIDYLATE SYNTHESIS ...... 28

DE NOVO PURINE SYNTHESIS ...... 29

REMOVAL OF ONE-CARBON UNITS ...... 29

ROLE OF FOLIC ACIO IN OISEASE ...... 31

NEURAL TUBE DEFECTS ...... 31

MEGALOBLASTIC ANEMIA ...... 34

CARDIOVASCULAR DISEASE ...... 35

DOWN'S SYNDROME ...... 37

CANCER ...... 37

NEURODEGENERATIVE DISORDERS ...... 39

INTERCONVERSIONS OF ONE-CARBON FOLATES ...... 41

MAMMALIAN NADP-DEPENDENT DCS ...... 42

MAMMALIAN NAD-DEPENDENT DC ...... 46

CATALYTIC PROPERTIES OF D/C DOMAIN ...... 48

YEAST DEHYDROGENASES ...... 49 INSECT DEHYDROGENASES ...... 52

INTERDEPENDENCE OF FOLATE COMPARTMENTS ...... 53

AUXB1 MUTANT ...... 53

GlyA MUTANT ...... 54

GlyB MUTANT ...... 55

FLUX OF ONE-CARBON METABOLISM ...... 57

Yeast ...... 57

Mammals ...... 60

CHAPTER TWO: MAMMALIAN MITOCHONDRIAL METHYLENETETRA­ HYDROFOLATE DEHYDROGENASE-CYCLOHYDROLASE DERIVED FROM A TRIFUNCTIONAL METHYLENETETRAHYDROFOLATE DEHYDROGENASE-CYCLOHYDROLASE-SYNTHETASE ...... 66

PREFACE ...... 67

SUMMARY ...... 68

INTRODUCTION ...... 69

MATE RIALS AND METHODS ...... 71

RESULTS ...... 75

DiSCUSSiON ...... 82

CHAPTER THREE: MAMMALIAN FIBROBLASTS LACKING MITOCHONDRIAL NAD+ -DEPENDENT METHYLENE­ TETRAHYDROFOLATE DEHYDROGENASE-CYCLOHYDROLASE ARE GLYCINE AUXOTROPHS ...... 84

PREFACE ...... 85

SUMMARY ...... 86

INTRODUCTION ...... 87

MATERIALS AND METHODS ...... 90 RESULTS ...... 96

DISCUSSION ...... 110

CHAPTER FOUR: THE ROLE OF MITOCHONDRIAL METHYLENE­ TETRAHYDROFOLATE DEHYDROGENASE-CYCLOHYDROLASE ACTIVITIES IN FORMATE PRODUCTION IN MAMMALIAN FIBROBLASTS ...... 117

PREFACE ...... 118

SUMMARY ...... 119

INTRODUCTION ...... 120

MATERIALS AND METHODS ...... 122

RESULTS ...... 129

DiSCUSSiON ...... 139

CHAPTER FIVE: GENERAL DISCUSSION ...... 144

EVOLUTION OF NMDMC ...... 147

MAMMALIAN MITOCHONDRIAL DEHYDROGENASES ...... 148

METABOLIC ROLE OF NMDMC ...... 152

ANTIFOLATES ...... 153

REPLACEMENT OF DEHYDROGENASE-CYCLOHYDROLASE ACTIVITIES

IN MITOCHONDRIA OF MAMMALIAN CELLS ...... 155

PREFERENCE FOR FORMATE AS ONE-CARBON DONOR ...... 156

FUTURE DIRECTIONS ...... 156

REFERENCES ...... 158 FOREWORD

The chapters 2 and 3 of this thesis include the text of the original papers submiUed for publication. Chapter 4 is a manuscript in preparation. The text of these manuscripts has been included in this thesis in compliance with the Faculty of Graduate Studies and Research "Thesis preparation and submission guidelines". An excerpt fram the text of section l, part C entitled "Manuscript­ based thesis" is cited below:

"As an alternative to the traditional thesis format, the dissertation can consist of a collection of papers of which the student is an author or co-author. These papers must have a cohesive, unitary character making them a report of a single program of research. The structure for the manuscript-based thesis must conform to the following:

1. Candidates have the option of including, as part of the thesis, the text of one or more papers submitted, or to be submiUed, for publication, or the clearly­ duplicated text (not the reprints) of one or more published papers. The text must conform to the "Guidelines for Thesis Preparation" with respect to font size, line spacing and margin sizes and must be bound together as an integral part of the thesis. (Reprints of published papers can be included in the appendices at the end of the thesis.) 2. The thesis must be more than a collection of manuscripts. Ali components must be integrated into a cohesive unit with a logical progression from one chapter to the next. In order to ensure that the thesis has continuity, connecting texts that provide logical bridges preceeding and following each manuscript are mandatory.

3. The thesis must conform to ail other requirements of the "Guidelines for Thesis

Preparation" in addition to the manuscripts.

The thesis must include the following:

1. a table of contents;

2. a brief abstract in both English and French;

3. an introduction which clearly states the rational and objectives of the

research;

4. a comprehensive review of the literature (in addition to that covered in

the introduction to each paper);

5. a final conclusion and summary;

6. a thorough bibliography;

7. Appendix containing an ethic certificate in the case of research involving

human or animal subjects, microorganisms, living cells, other biohazards

and/or radioactive material.

ii 4. As manuscripts for publication are frequently very concise documents, where apprapriate, additional material must be pravided (e.g., in appendices) in sufficient

detail to allow a clear and precise judgement to be made of the importance and originality of the research reported in the thesis.

5. In general, when co-authored papers are included in a thesis the candidate

must have made a substantial contribution to ail papers included in the thesis. In

addition, the candidate is required to make an explicit statement in thesis as to

who contributed to su ch work and to what extent. This statement should appear

in a single section entitled "Contributions of Authors" as a preface to the thesis.

The supervisor must attest to the accuracy of this statement at the doctoral oral

defence. Since the task of the examiners is made more difficult in these cases, it

is in the candidate's interest to clearly specify the responsibilities of ail the authors

of the co-authored papers.

6. When previously published copyright material is presented in a thesis, the

candidate must include signed waivers fram the publishers and submit these to

the Graduate and Postdoctoral Studies Office with the final deposition, if not

submitted previously. The candidate must also include signed waivers fram any

co-authors of unpublished manuscripts.

7. Irrespective of the internai and external examiners reports, if the oral defence

committee feels that the thesis has major omissions with regard to the above

iii guidelines, the candidate may be required to resubmit an amended version of the thesis. See the "Guidelines for Doctoral Oral Examinations," which can be

obtained fram the web (http://www.mcgill.ca/fgsr). Graduate Secretaries of

departments or fram the Graduate and Postdoctoral Studies Office, James

Administration Building, Room 400,389-3990, ext. 00711 or 094220.

8. In no case can a co-author of any component of su ch a thesis serve as an

external examiner for that thesis.

The format specified in this thesis has been appraved by the Department of

Biochemistry, McGili University. The figures and table in each chapter are

numbered relative to the chapter and ail references have been compiled at the

end of thesis. Minimal formatting of sections within the text of the chapters and of

references has been performed to maintain stylistic uniformity thraughout the

thesis.

iv ACKNOWLEDGEMENTS

As 1 reflect back on this amazing journey that began over six years ago, the one thing that really stands out are the people without whom this truly satisfying accomplishment of mine could never have been possible. 1would like to begin by thanking my supervisor, Dr. Robert MacKenzie for giving me the opportunity to work in his lab. 1 have benefited immensely from his expertise and his helpful discussions as weil as from his criticisms, which have enriched my scientific knowledge and given me the confidence to venture into the world of research to establish my own career. 1 profoundly appreciate the time he always had for me no matter how hectic his schedule. His continued passion for research and his attention to detail have impressed upon me the desire to strive to be a better scientist.

1 am very grateful to Narciso Mejia who has taught me so much. From the very first day that 1 entered the lab, his patience has been comforting and his technical skills have been an invaluable asset. His prompt assistance in every aspect of my project has been tremendously appreciated.

1 would also like to thank ail the past and present members of my lab, Dr.

Peter Pawelek, Hai Ping Wu, Saravanan Sundararajan, Dr. Erminia Di Pietro,

Karen Christensen and Uros Kuzmanov for endless sessions of trouble-shooting and filling my days in the lab with su ch entertaining conversations. Also, 1 would like to thank Sofia Moriatis who always made sure there was clean glassware in the lab.

v 1 am very grateful to Dr. Walter Mushynski and Dr. Mark Featherstone for their generous time and helpful advice. 1 would like to thank Carol Miyamoto whose knowledge of molecular biology has been a lifesaver on many occasions for me. 1 am also very appreciative of Ann Brasey and Dr. Laurent Huck for their

help with the french translation of the abstract to this thesis.

1 have been blessed with a family that has never wavered in their support for me as 1 pursued my dream. If there is anyone who feels greater pride than me

at this moment, it is definitely my father whose own love for science was the

inspiration for mine. A very close second is my mother, whose quiet strength

sustains me. 1 am deeply indebted to my sister, Jigisha who was instrumental in

the printing of this thesis and who always finds a way to put up with my many

quirks. 1 can never forget my youngest sister, Avani, whose independent nature is

refreshing to a conservative like me.

Although this degree is technically mine, 1 have to share the achievement

with my husband, Vinod who has been with me every step of this journey. Words

cannot begin to describe the profound appreciation 1 feel for everything he has

do ne for me. 1 am deeply touched by his selflessness and compassion especially

through the toughest of times when he motivated me to stand up and persevere.

During the course of this degree, 1 acquired a new family including two

incredibly caring parents in-Iaw, who treat me like their very own daughter. 1

would particularly like to thank Pravina, Kishore and Nandani. They will never

know how their words of encouragement have helped me to reach the end of this journey.

vi Lastly, 1would also like to sincerely thank my dear friend Wagner Rulli who showed me the importance of maintaining balance in my life.

1 have always felt like the last six years have been a marathon and now 1 have finally reached the finish line. 1 can honestly say that despite ail of the struggles, 1 have no regrets about the path 1 have chosen as ail of the sacrifices seem to fade next to those sweet moments of victory.

vii PUBLICATION OF THE WORK PRESENTED IN THIS THESIS

"Mammalian Mitochondrial Methylenetetrahydrofolate Dehydrogenase­

Cyclohydrolase Derived From A Trifunctional Methylenetetrahydrofolatedehydro­ genase-Cyclohydrolase-Synthetase" by Harshila Patel, Karen E. Christensen,

Narciso R. Mejia and Robert E. MacKenzie in the Archives Of Biochemistry and

Biophysics (2002), vol. 403, pages 145-148.

"Mammalian Fibroblasts Lacking Mitochondrial NAD+ -Dependent

Methylenetetrahydrofolate Dehydrogenase-Cyclohydrolase Are Glycine

Auxotrophs" by Harshila Patel, Erminia Di Pietro and Robert E. MacKenzie in the

Journal of Biological Chemistry (2003), vol. 278, pages 19436-19441.

viii CONTRIBUTIONS TO ORIGINAL KNOWLEDGE

This thesis deals with the compartmentalization of folate metabolism in mammalian cells. This was studied by examining the roles of the cytoplasmic

NADP-dependent methylenetetrahydrofolate dehydrogenase-methenyl­ tetrahydrofolate cyclohydrolase-formyltetrahydrofolate synthetase (DCS) and the mitochondrial NAD-dependent methylenetetrahydrofolate dehydrogenase­ methenyltetrahydrofolate cyclohydrolase (NMDMC) in mouse fibroblasts.

1. The cDNA encoding the mou se cytoplasmic DCS was isolated from an

embryonic mouse kidney cDNA library. This cDNA codes for a protein that

has 87% amine acid homology to the human DCS protein.

2. The gene encoding the mou se cytoplasmic DCS was isolated. This gene

spans approximately 65 kilobases. It contains 28 exons and 27 introns. Ali

of the exon/intron splice junctions follow the GT/AG rule.

3. A comparison of the 3' untranslated region (UTR) of the NMDMC with the

synthetase region of the murine DCS cDNA and gene revealed several

regions of greater than 50% identity. Similar comparisons with the human

and Drosophila sequences revealed similar homologies. The residual

synthetase sequences in the 3' UTR of the NMDMC cDNA demonstrates

that in higher eukaryotes, the bifunctional NMDMC was derived from a

ix trifunctional NADP-dependent DCS precursor with a change in cofactor

2 specificity from NADP to NAD and a requirement for Mg + and inorganic

phosphate.

4. A series of growth studies of the Nmdmc-/- cell lines showed that although

these mutant cell lines do not require serine, they are glycine auxotrophs.

The glycine auxotrophy is interpreted to demonstrate that the NMDMC is

the only methylenetetrahydrofolate dehydrogenase expressed in

mitochondria of these cells.

5. The growth of the Nmdmc-/- cell lines on complete medium with dialyzed

serum is stimulated by approximately 2-fold by the addition of formate or

hypoxanthine but not thymidine, supporting a role in purine synthesis for

the NMDMC.

6. Radiolabeling experiments with C4C]-labeled one-carbon donors

demonstrate a 3 to 10-fold enhanced incorporation of radioactivity into

DNA from formate as compared to serine by Nmdmc-/- cell lines and

supports the same metabolic role of NMDMC.

7. The rescue of Nmdmc-/- celllines with the expression of the NMDMC cDNA

alleviates the glycine auxotrophy.

x 8. The rescue of Nmdmc-/- cell lines with the expression of the NAD­

dependent monofunctional dehydrogenase activity of the NMDMC protein

alleviates the glycine auxotrophy as weil but these cell lines do not grow as

weil as the NMDMC-rescued cell lines which suggests that although the

cyclohydrolase activity is not essential in mammalian mitochondria, the rate

of 10-formylTHF production is not optimal in its absence.

9. The rescue of Nmdmc-/- cell lines with the expression of the NADP­

dependent DCS protein in the mitochondria also reverses the glycine

auxotrophy which suggests that the NAD cofactor specificity of the

mitochondrial DC is the not absolutely required for the maintenance of the

flux of one-carbon metabolites.

xi LIST OF FIGURES

CHAPTER ONE

Figure 1. The structure of tetrahydrofolate ...... 3

Figure 2. The one-carbon donors and products of the folate metabolic

pathways ...... 17

Figure 3. The compartmentalization of folate metabolism in mammalian

cells ...... 43

Figure 4. The compartmentalization of folate metabolism in yeast...... 51

CHAPTER TWO

Figure 1. The gene for the murine cytoplasmic DCS ...... 76

Figure 2. Schematic diagram of the homologousregions of the NADP­

dependent DCS cDNAs and genes and the 3'-untranslated

regions of the NAD-dependent mt-OC for mouse, human

and Drosophila ...... 79

Figure 3. Details of the homolgies found between the synthetase

region of the NADP-dependent DCS and the 3'-untranslated

region of the mt-OC ...... 81

xii CHAPTER THREE

Figure 1. Genotypic analysis ...... 97

Figure 2. Embryonic fibroblasts stained with Mitotracker Red CMXRos

to detect mitochondria with an intact membrane potential ...... 99

Figure 3. Growth of fibroblasts in defined medium containing 10%

redialyzed fetal bovine serum ...... 101

Figure 4. Incorporation of radiolabeled one-carbon donors into total DNA

of wild type and mutant fibroblasts ...... 104

Figure 5. Phenotypic comparison of embryos from pregnant dams

supplemented with glycine or formate ...... 109

Figure 6. Folate-dependent activities in the cytoplasm and mitochondria

of mammalian cells ...... 114

CHAPTER FOUR

Figure 1. Western blot analysis of rescued NMDMC null mutant

fibroblasts ...... 130

Figure 2. Subcellular localization of NMDMC protein ...... 132

Figure 3. Subcellular localization of NADP-dependent DCS protein ...... 134

Figure 4. Growth of rescued celllines in glycine-free medium

containing 10% redialyzed fetal bovine serum ...... 136

xiii LIST OF TABLES

CHAPTER TWO

Table 1. The intron-exon boundaries of the murine DCS gene ...... 77

Table 2. Details of the homologies found between the 3'-untranslated

region of the mt-DC cDNA and the synthetase region of the

NADP-dependent DCS ...... 80

CHAPTER THREE

Table 1. Summary of the properties of established fibroblast cell

lines ...... 102

Table 2. Incorporation of radiolabeled precursors into the bases of

DNA ...... 106

Table 3. Effects of supplements on embryonic phenotype ...... 108

CHAPTER FOUR

Table 1. Incorporation of radiolabeled precursors into total DNA ...... 138

xiv LIST OF ABBREVIATIONS

AD Alzheimer's disease AICAR aminoimidazole-4-carboxamide ribonucleotide ALDH aldehyde dehydrogenase BHMT betaine-homocysteine methyltransferase Bp basepair( s) C methenyltetrahydrofolate cyclohydrolase CBS cystathionine ~-synthase CHD coronary heart disease CHO Chinese Hamster Ovary CSHMT cytoplasmic serine hydroxymethyltransferase o methylenetetrahydrofolate dehydrogenase OC methylenetetrahydrofolate dehydrogenase-methenyltetrahydrofolate cyclohydrolase DCS methylenetetrahydrofolate dehydrogenase-methenyltetrahydrofolate cyclohydrolase-formyltetrahydrofolate synthetase or Mthfd1 DO dimethylglycine dehydrogenase DDATHF 5,10-dideazatetrahydrofolate DHF dihydrofolate DHFR dihydrofolate reductase dTMP deoxythymidine monophosphate dUMP deoxyuridine monophosphate E embryonic day EF elongation factor FIGLU formiminoglutamic acid FLDH formaldehyde dehydrogenase fMET formylmethionine FMT methionyl-tRNA formyltransferase FPGS folylpolyglutamate synthetase FR folate receptor FTCD formiminotransferase-cyclodeaminase FTD 1O-formylTHF dehydrogenase GAR glycinamide ribonucleotide GCS glycine cleavage system GCV glycine decarboxylase complex GNMT glycine N-methyltransferase GPI glycosylphosphatidylinositiol hMFT human mitochondrial folate transporter IF transformed fibroblasts kb kilobasepair(s) KO knockout LB Luria Bertani MPTP 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine xv MS synthase mSHMT mitochondrial serine hydroxymethyltransferase mt mitochondrial MTHFR methylenetetrahydrofolate reductase MTHFS methenyltetrahydrofolate synthetase NAD nicotinamide adenine dinucleotide NADP nicotinamide adenine dinucleotide phosphate NMDA N-methyl-D aspartate NMDMC NAD-dependent methylenetetrahydrofolate dehydrogenase­ cyclohydrolase or Mthfd2 NMR nuclear magnetic resonance NTD neural tube defect PD Parkinson's disease Pi inorganic phosphate PLP pyridoxal-5' -phosphate RFC reduced folate carrier S 10-formyltetrahydrofolate synthetase SAH S-adenosylhomocysteine SAM S-adenosylmethionine SD sarcosine dehydrogenase SF spontaneously immortalized fibroblasts SHMT serine hydroxylmethyltransferase SRE serum responsive element THF tetrahydrofolate tRNA transfer RNA TS thymidylate synthase UTR untranslated region wt wild type

xvi CHAPTER ONE

GENERAL INTRODUCTION

1 OISGOVERY OF FOLIG AGIO

The importance of folic acid was initially demonstrated in 1931 by Lucy

Wilis. Her studies revealed that yeast extract contained a critical factor, which was effective in treating macrocytic anemia that frequently manifested during late

pregnancy. This factor was determined to be folic acid.

FOLIG AGIO

Folic acid is a water-soluble B9 vitamin that is converted to its active form, tetrahydrofolate (THF), by two sequential reductions catalyzed by dihydrofolate

reductase (OH FR). The structure of THF consists of a pteridine ring linked to p­

aminobenzoic acid by a methylene carbon. The p-aminobenzoic acid has a varying number of glutamates linked to it via a series of y-glutamyl bonds.

Furthermore, one-carbon substituents can be attached to THF at its N5 or N10

positions or they can be bridged between these two positions (Fig. 1). The major forms of cellular folates are 10-formyITHF, 5-formyITHF, 5-methylTHF and 5,10-

methyleneTHF although there are also low concentrations of 5-formiminoTHF and

5,10-methenyITHF (Schirch 1997). Mammals cannot synthesize folates and must

depend on a variety of dietary sources such as leafy green vegetables, including

spinach, brussel sprouts and turnip greens, citrus fruits, potatoes and beans

(Lucock 2000). In fact, the name "folic acid" is derived from the Latin word

"folium" which means "Ieat".

2 FIGURE 1. The structure of tetrahydrofolate.

3 c Ir-::I:------.I

o N N o ::I: ::I: ::I: 0 0-0-0-0-0 1 1 1 ::I: ::I:Z o 1 o ~ ~ 0 O-::I:-O-O-Oo 1 ::I:Z 1 o o

::I: Z ~ Z ::I: o Z

::I:z-.!I Z ::I:N FOLATE ABSORPTION

Dietary sources of folate exist predominantly as 5-methylTHF polyglutamate which is readily oxidized and, consequently, a significant amount of folate can be destroyed during cooking (Steinberg 1984). However, this folate form is relatively stable under the mildly acidic conditions in the gastric environ ment. Folates exist as anions at intraluminal pH. Since the dietary folates

are polyglutamylated, they must be hydrolysed to folylmonoglutamates by

pteroylpolyglutamate hydrolase or "conjugase", which is located in saliva,

intestinal juice and the mucosal brush border membrane. The monoglutamates

are transported across the enterocyte brush border membrane by a saturable pH­

dependent mechanism whereby the folate is exchanged for a hydroxyl anion.

Folate absorption occurs throughout the length of the small intestine. The 5-

methylTHF monoglutamate is the plasma form of the vitamin.

Once absorbed, the 5-methylTHF monoglutamate is released into the

portal circulation (Steinberg 1984). At this point, a significant portion of the folate

is transported to the liver and the remaining folate is taken up by the peripheral

tissues. In the liver, the majority of the methylTHF is not polyglutamated but it

actually undergoes a recirculation process. This unique transport pathway is

known as the folate enterohepatic cycle whereby the folate is released in bile

secretions, to the small intestine where it is reabsorbed and then a portion is

distributed to peripheral tissues. Once the folate is transported to peripheral

tissues, converts the 5-methylTHF to THF, the preferred

4 substrate for folylpolyglutamate synthetase (FPGS), an enzyme that adds residues to THF (Cook et al. 1987, Schirch and Strong 1989). This

polyglutamylated THF is a biologically active intracellular form of folic acid that is able to participate in the various cellular folate-mediated processes.

There are endogenous pools of folate monoglutamates that provide a sufficient supply to critical tissues during conditions of folate de privation

(Steinberg 1984). These pools reside in the enterohepatic cycle as weil as

storage tissues, namely the liver and kidney. Senescent erythrocytes represent

another source of folates. These cells have a limited life span and are constantly

being generated by the highly proliferative bone marrow. They possess

significant folate pools, which allow them to make a significant contribution to folate homeostasis. When these erythrocytes die, their folate is salvaged by the

reticuloendothelial system. The folate is largely captured by the liver, from which

point it is redistributed to peripheral tissues via the enterohepatic cycle (Steinberg

1984).

MECHANISMS OF CELLULAR TRANSPORT

Folate monoglutamates are hydrophilic, bivalent anions present in the

serum at nanomolar concentrations. Consequently, there is a need for

concentrative transport systems to fulfill cellular folate requirements. The

predominant mechanisms for folate delivery from the extracellular space to the

cytoplasm are receptor- and carrier-mediated processes.

5 FOLATE BINDING PROTEINS

The folate binding proteins constitute a major class that is further divided into categories: soluble forms and membrane-associated forms, which are known as folate receptors (FRs) (Antony 1996). The soluble folate binding proteins are derived from FRs as a result of either cleavage of their glycosylphosphatidylinositol membrane anchor or proteolysis. These forms are structurally related but they have different functions.

The soluble FRs probably have multiple functions, such as providing a convenient mechanism for concentrating folate compounds, protecting the bound, reduced folates from oxidation and serving as storage proteins to conserve folates

(Antony 1996).

Folate Receptor

Folate receptors are anchored to the ce Il membrane by glycosylphosphatidylinositol (GPI) linkages. Although the different isoforms vary in their affinities for the various folate derivatives, FRs generally exhibit high­ affinity binding with a preference for folie acid and 5-methylTHF over other reduced folates but they bind methotrexate, a folate analogue, poorly. There are four known human isoforms that are expressed in specifie tissues and cellular populations. FRa is expressed predominantly in epithelial cells (Elwood 1989).

FR~ is expressed in various tissues (Ratnam et al. 1989). FRy is a secretory

6 form due to the lack of signal for GPI attachment, which is predominantly

expressed in hematopoietic tissues (Shen et al. 1995). FR'y is a truncated form of

FRy that is most likely generated as a result of gene polymorphism or by

alternative splicing. FRa is a novel isoform that was discovered by a "database

mining" strategy utilizing the amine acid sequence of the FRa (Spiegelstein et al.

2000). However, this study concluded that FRa had undetectable expression in

numerous selected tissues. This isoform may be under developmental regulation

or a highly restricted spatial and/or temporal expression pattern or it may simply

be a pseudogene (Spiegelstein et al. 2000).

Both the FRa and FR~ isoforms are expressed in fetal and adult tissues.

Although, FR~ exhibits more widespread expression in normal tissues as

compared to FRa, the latter isoform was found to be abundantly expressed in the

placenta (Ross et al. 1994). This was confirmed by a similar expression pattern

of the murine homolog, Folbp1 (Barber et al. 1999). It plays a critical raie in the

two-step maternal-to-fetal transplacental folate transport. The first step consists

of the concentrative component pravided by the placental FRa, located on the

maternally-facing chorionic surface. The FRa binds the circulating 5-methylTHF

to generate a placental folate pool. There is a graduai release of 5-methylTHF

fram this pool, which contributes to the folate that is carried along a downhill

concentration gradient to the fetus (Antony 1996).

Folate transport into kidneys also occurs by a FR-mediated pracess (Birn

et al. 1993). After glomerular filtration, the luminal folate binds the FR located in

7 the brush border membrane of proximal renal tubular cells and is internalized.

This process represents an efficient mechanism to prevent the loss of folate

compounds in the urine.

Folate transport by the FR has been suggested to occur by potocytosis

(Anderson et al. 1992). This process involves caveolae, which are invaginated

vesicles located on the cell surface. The caveolae contain c1usters of FRs that

bind folate when the compartment is open. Subsequently, the caveolae close and

the compartment is acidified by an H+ pump. At low pH, the folate dissociates

from its receptor and is transported into the cytoplasm where it is

polyglutamylated and consequently, sequestered inside the cell. Subsequently,

the FR returns to the cell surface where it can mediate another cycle of folate

internalization (Rothberg et al. 1990). It is controversial whether the folate

reaches the cytoplasm by diffusion across the membrane through water-filled

channels or by a specifie carrier, which acts in conjunction with the FR, namely

the reduced folate carrier (RFC1), which will be discussed shortly. Furthermore,

the entire concept of potocytosis and its role in receptor-mediated folate transport

has been challenged as weil, due to a lack of direct evidence demonstrating the

presence of FR in caveolae in the absence of c1ustering agents. 1nstead , it is

argued that FR-mediated transport actually occurs by bulk-membrane

endocytosis (Mayor et al. 1994).

There are three known folate receptors in mice. These three proteins,

Folbp1 (Brigle et al. 1991), Folbp2 and Folbp3 (Spiegelstein et al. 2001) are

actually the orthologs of human FRa, FRf3 and FRe, respectively. To date, only

8 the first two genes have been individually inactivated in mice in order to determine their importance. The disruption of the Folbp2 had no effect (Piedrahita et al.

1999). However, the inactivation of the Folbp1 resulted in significant congenital

malformations by embryonic day 10. There was also arrested development of the

embryos in utero, and ultimately death due to development of a neural tube

defect. The phenotype can be rescued with maternai folinic acid supplementation

(Piedrahita et al. 1999). It appears that Folbp1 is essential for providing an

adequate level of folate to developing embryos. These results also suggest that a

defect in the FRa may contribute to the occurrence of neural tube defects in

humans.

Reduced Folate Carrier

The major folate transport system is known as the reduced folate carrier

(RFC1). The RFC1 is an integral membrane glycoprotein that is a member of the

Major Facilitator superfamily of transport carriers. It consists of 12

transmembrane domains with the N- and C-termini and the large loop between

the 6th and 7th transmembrane domains directed to the cytoplasm (Dixon et al.

1994). The transmembrane a-helices of the RFC 1 are thought to form channels

which permit the passage of folates (Brigle et al. 1995).

ln contrast to the FRs, the RFC1 exhibits low-affinity binding with varied

substrate specificity such that it binds reduced folates and methotrexate with a

higher affinity than folic acid (Henderson 1990). An early study (Goldman 1971)

9 proposed that the RFC1 is a bi-directional transporter, which allows for the internalization of folates against an electrochemical concentration gradient (uphill transport) as a consequence of the downhill counterflow of intracellular anions across the cell membrane. These anions are actually organic phosphates synthesized and present only in the cells and they drive the uphill folate transport.

Furthermore, this folate uptake process is inhibited by organic and inorganic anions present in the extracellular medium. However, this anion-exchange model of folate transport was reexamined by another study (Yang et al. 1984) which argued that under physiological conditions, the concentration of inorganic counter-anion required to activate folate influx was considerably higher than that

normally found within the intracellular water of intact cells.

The RFC1 protein is ubiquitously located in ail tissues and cell types. The

presence of multiple RFC1 transcripts due to multiple transcriptional starts and variable splicing give the possibility to generate tissue or cell li ne-specifie RFC1

proteins (Sirotnak and Tolner 1999). The RFC1 has been shown to mediate folate absorption across the brush border membrane of the small intestine (Chiao et al. 1997).

The importance of the RFC 1 was clearly demonstrated by its targeted

inactivation in mice (Zhao et al. 2001). The mutation is embryonic lethal due to a

1 failure of the hematopoietic organs. The RFC1- - null embryos die prior to

1 embryonic day 9.5, which is comparable to when the Folbp1- - mutant embryos

died (Piedrahita et al. 1999), which indicates that these two folate transport systems play precise roles in embryonic development as the presence of one is

10 unable to compensate for the other.

Although receptor- and carrier-mediated systems may occur together in some cell types, these two transport pro cesses are quite independent of each other. According to an experiment (Spinella et al. 1995) in which RFC 1-defective mouse leukemic cells were transfected with FRa cDNA, FR-mediated transport was shown to differ from RFC1-mediated transport on the basis of energy, ion and pH dependence. The contribution of each transport system to net translocation of folates is unclear and depends on the expression level of the genes involved. Specifically, when FRa is expressed to a sufficient level, it may

be a significant route for folates at physiological concentrations but as the levels of folates increase to pharmacological concentrations, FRa becomes a very minor

contributor to folate transport. At this point the RFC1 becomes the predominant folate transport route (Spinella et al. 1995). In general, carrier-mediated

processes are much more efficient than receptor-mediated transport of folates as

a result of the slow net cycling rate of the FR (Spinella et al. 1995).

LOW PH TRANSPORTER

Another folate transport pathway has been documented in mouse leukemic

L 1210 cells (Sierra and Goldman 1998). This low pH transporter has

characteristics that are distinct from the RFC1. This energy-dependent

transporter functions optimally at pH 5.5-6.5. It has a much higher affinity for folic

11 acid than RFC1. It is minimally sensitive to the anionic composition of the medium. This transporter can mediate the influx of certain folates at physiological

pH but it does so at a markedly lower rate than RFC1. These results indicate

that the low pH transporter operates by a different mechanism than RFC1.

FOLATE POL YGLUTAMYLATION

As was discussed before, folylmonoglutamates are the transport forms of folate. FPGS catalyzes the addition of glutamate residues by forming a peptide

bond between a-amino group of the incoming glutamate and the y-carboxyl of the

glutamate already linked to THF in an ATP-dependent reaction to generate folylpolyglutamates. The reason why the cell expends this energy is because

there are many advantages to this modification of folates (Shane 1989). Firstly, folylpolyglutamates are more efficient substrates for the majority of the enzymes

in one-carbon metabolism. Specifically, the glutamate portion enhances the

catalytic specificity, which is demonstrated by folylpolyglutamates showing a lower

Km for the enzyme than the corresponding monoglutamate derivatives (Matthews

et al. 1982) but the extension of the glutamate chain also decreases the Vmax

(Atkinson et al. 1997). In addition, they regulate the specifie enzyme activities

and the flux of one-carbon units through different metabolic pathways. Also, folylpolyglutamates can allow channeling of substrates in multifunctional enzymes

(Shane 1989). Finally, folylpolyglutamates assist in cellular retention of folates as

12 they are not transported across the cell membrane.

ln mammals, there are mitochondrial and cytoplasmic isoforms of FPGS, which are encoded by a single gene. The isoforms are generated as a result of alternate transcription start sites (Freemantle et al. 1995, Chen et al. 1996). The specifie activity of the mitochondrial FPGS is higher than that of the cytosolic isoform, which may account for the longer folylpolyglutamate derivatives in the mitochondria (Lin and Shane 1994).

Hexa- and heptaglutamates are the predominant intracellular forms of folate in mammalian cells because they are poor substrates for FPGS (Cook et al.

1987). It has also been shown that only low levels of FPGS activity are needed to metabolize folates to folylpolyglutamates of sufficient glutamate chain length to be retained by mammalian cells. However, in order to generate the long chain derivatives that are usually present in cells, there is a requirement for higher

FPGS levels.

SOURCES OF ONE-CARBON UNITS

There are several one-carbon donors that provide their one-carbon units at various entry points into the folate pathways (Fig. 2).

SERINE

The major source of one-carbon units in mammalian cells is carbon 3 of

13 serine (Fig. 2), which can be obtained from the diet or originate from the glycolytic pathway. This one-carbon unit is transferred to THF by serine hydroxymethyltransferase (SHMT) to generate glycine and 5,1 O-methyleneTHF.

GLYCINE

The mitochondrial glycine cleavage system (GCS) is a complex of four

proteins including a P protein, which is a pyridoxal phosphate enzyme. This complex utilizes an important source of one-carbon units, glycine (Fig. 2). In this reaction, carbon 2 of glycine is transferred to THF forming 5,1 O-methyleneTHF, and carbon dioxide. This pathway is the predominant route of glycine degradation in the liver and kidney of mammalian cells (Yoshida and Kikuchi

1973).

HISTIDINE

Histidine (Fig. 2) serves as a significant donor of one-carbon units

exclusively in the mammalian kidney and liver. It is catabolised to formiminoglutamic acid (FIGLU) by the sequential actions of histidase,

and imidazolepropionate amino hydrolase. The formimino group of FIGLU, which

is derived from the carbon 2 position of the imidazole ring of histidine, is transferred to THF by 5-formiminotransferase to produce 5-formiminoTHF and glutamic acid. The formimino group of 5-formiminoTHF is deaminated by 5-

14 formiminoTHF cyclodeaminase to yield 5,1 O-methenyITHF, which can serve as an entry point into the cytoplasmic folate pathway (Schalinske and Steele 1996).

The latter two enzyme activities actually constitute a bifunctional protein, formiminotransferase-cyclodeaminase (FTCO) (Orury et al. 1975). It has been shown in rat liver with the use of an in vivo tracer kinetic technique that 10- formylTHF receives 60% of its formyl group from histidine oxidation whereas methylTHF obtains 21 % of its methyl groups from the reduction of the histidine carbon through the one-carbon pool (Schalinske and Steele 1996).

CHOLINE

Choline (Fig. 2) is obtained from the diet or it can be synthesized through sequential methylations of phosphatidylethanolamine. Hepatic choline metabolism generates dimethylglycine (Fig. 2). One of the methyl groups of dimethylglycine is transferred to THF by the flavoprotein, dimethylglycine dehydrogenase (DO), which yields 5,1 O-methyleneTHF and sarcosine in the mitochondria (Wittwer and Wagner 1981). Sarcosine (Fig. 2) can also transfer its methyl group to THF via another flavoprotein, sarcosine dehydrogenase (SO) yielding glycine and 5,1 O-methyleneTHF (Wittwer and Wagner 1981). Sarcosine can also be generated in the cytoplasm by glycine N-methyltransferase (GNMT).

The GNMT catalyzes the methylation of glycine by S-adenosylmethionine (SAM), which results in the formation of sarcosine.

15 FORMATE

Although formate is detectable in numerous physiological fluids of

mammals, the actual sources of formate remain to be determined. Formate (Fig.

2) can be derived from the oxidation of formaldehyde by cytoplasmic formaldehyde dehydrogenase (FLDH), mitochondrial aldehyde dehydrogenase

(ALDH) and peroxisomal catalase (MacKenzie 1984). Subsequently, formate can

serve as a one-carbon donor upon activation to 1O-formylTH F by the ATP­

dependent 10-formylTHF synthetase.

16 FIGURE 2. The one-carbon donors and products of the folate metabolic pathways.

Abbreviations are: AICART, aminoimidazole-4-carboxamide ribonucleotide transformylase; ALDH, aldehyde dehydrogenase; BHMT, betaine-homocysteine methyltransferase; C, methenylTHF cyclohydrolase; CD, formiminoTHF cyclodeaminase; D, methyleneTHF dehydrogenase; DO, dimethylglycine dehydrogenase; FDH, 10-formylTHF dehydrogenase; FLDH, formaldehyde dehydrogenase; FMT, methionyl-tRNA formyltransferase; FT, formiminoglutamate:THF formiminotransferase; GART, glycinamide ribonucleotide transformylase; GCS, glycine cleavage system; GNMT, glycine N­ methyltransferase; MS, methionine synthase; MTHFR, methyleneTHF reductase; MTHFS, methenylTHF synthetase; S, 10-formylTHF synthetase; SD, sarcosine dehydrogenase; SHMT, serine hydroxymethyltransferase; TS, thymidylate synthase. Adapted from MacKenzie, 1984.

17 histidine choline 1 1 formiminoglutamate dimethylglycine serine glycine FT THF tGNMT sarcosine 1 5-formiminoTHF formaldehyde THF purines NADPH FDH CO FLDH 9lYCinej CD 1 2 IGCS NADP ALDH t D 1 C 7 S 1 5,10-methyleneTHF ( r ~ ) 5,10-methenylTHF ( ) 10-formylTHF E ::;a a:c: < ') formate NAD(P) NAD(P)H II GARTri ~ FMT ADP ATP THF MTHFS SHMT

IT / formylmethionyl-tRNA

5_formYITHFA~::i!

FT 5-methylTHF thymidylate r homocysteine formylglutamate

THF

dimethylglycine

methionine OVERVIEW OF FOLATE-MEDIATED METABOLISM

ln eukaryotic cells, the major pathways of folate-mediated one-carbon

metabolism are distributed between the cytoplasm and the mitochondria. Several

of the folate-dependent enzymes are present in both of these compartments.

Folate is also present in the nucleus but it does not make any significant

contributions to total cellular folate concentrations (Appling, D.R. 1991).

Serine hydroxymethyltransferase (SHMT) is a pyridoxal phosphate­

dependent enzyme that catalyzes the reversible interconversion of serine and

THF for glycine and 5,1 O-methyleneTHF (Fig. 2). There is a cytoplasmic as weil

as a mitochondrial isoform of SHMT (Garrow et al. 1993, Stover et al. 1997, Girgis

et al. 1998). The cSHMT gene displays tissue-specifie expression with

particularly high levels in the kidney and liver, implying that it is not a house­

keeping gene. In contrast, the mSHMT expression appears to be more ubiquitous

(Girgis et al. 1998).

The cSHMT also catalyzes the irreversible ATP-dependent conversion of

5,1 O-methenyITHF, in the presence of glycine, to 5-formylTHF (Stover and

Schirch 1990). The 5-formylTHF remains bound to the enzyme and behaves as

an effective slow-binding inhibitor of SHMT. The metabolic role of 5-formylTHF

has not been established but it comprises up to -10% of the total folate pool in

most organisms (Horne et al. 1989). It is also the most stable naturally occurring form of reduced folate. The D,L-mixture of 5-formylTHF is known as folinic acid or

clinically as leucovorin, a therapeutic drug used to rescue patients from

18 methotrexate toxicity by elevating cellular folate levels.

MethenylTHF synthetase (MTHFS) is the only known enzyme that

metabolizes 5-formylTHF to produce 5,1 O-methenylTHF in an ATP-dependent

reaction (Fig. 2). Consequently, MTHFS prevents the accumulation of 5- formylTHF in mammalian cells (Holmes and Appling 2002). Although in most

mammalian ceillines, MTHFS is exclusively found in the cytoplasm, some activity

has been detected in the mitochondria isolated from human cells, whereby it

accounts for 15% of the total MTHFS activity (Bertrand et al. 1995).

The physiological role of MTHFS is uncertain. One study suggested that

MTHFS acts as a detoxifying agent to prevent the inhibition of de novo purine

synthesis which is mediated by the direct inhibition on AICAR transformylase by

the polyglutamate forms of 5-formylTHF (Bertrand and Jolivet 1989). A recent

study (Anguera et al. 2003) concerning MTHFS has demonstrated that it

regulates intracellular folate concentrations by accelerating the rate of folate

catabolism. They propose two possible mechanisms for MTHFS-mediated folate

catabolism. First, MTHFS may have an additional catalytic activity whereby it can

oxidize reduced folate coenzymes. Alternatively, MTHFS may accelerate folate

catabolism by changing the folate distribution, resulting in the accumulation of

more labile folate derivatives (Anguera et al. 2003).

A study exploring the role of 5-formylTHF in cultured human

neuroblastoma (Girgis et al. 1997) has proposed that it mediates the flow of one­

carbon units in the cytoplasm though either the homocysteine remethylation or

serine synthesis pathways. The overexpression of MTHFS in these cells cultured

19 in exogenous glycine depletes 5-formylTHF levels. This results in the activation of cSHMT, which in combination with the excess glycine supplied in the medium,

enhances the flow of 5,1 O-methyleneTHF in the direction of cSHMT instead of

MTHFR. This results in the depletion of 5-methylTHF pools and a subsequent

impairment of homocysteine remethylation, elevated serine levels and an

increased methionine requirement for maximal cell growth in cultured human

neuroblastoma (Girgis et al. 1997).

5,10-methyleneTHF represents an important branch point in folate

metabolism: 1) it can be metabolized irreversibly to methylTHF, which is the entry

point into the cellular methylation cycle, 2) it can donate its methyl group to

deoxyuridine monophosphate (dUMP) to form deoxythymidine monophosphate

(dTMP) and dihydrofolate (OHF) in the cytoplasm or 3) it can be interconverted to

5,10-methenyITHF by methyleneTHF dehydrogenase in either compartment.

The interconversions of THF that carry one-carbon substituents are carried

out by three different enzyme activities: 5,1 O-methyleneTHF dehydrogenase (0),

5,10-methenyITHF cyclohydrolase (C) and 10-formylTHF synthetase (S). They

catalyze the following reactions:

o methyleneTHF + NAOP <=> methenylTHF + NAOPH C

methenylTHF + H20 <=> formylTHF S formylTHF + AOP + Pi <=> THF + ATP + HCOOH

20 These three enzyme activities will be discussed comprehensively in a subsequent section of this introductory chapter.

HOMOCYSTEINE REMETHYLATION CYCLE

One of the key components of the cellular methylation cycle is homocysteine. This metabolite is at the branch point of two critical pathways: it can undergo de novo methionine regeneration by either folate-dependent methionine synthase (MS) or folate-independent betaine-homocysteine methyltransferase (BHMT) or it can undergo transsulfuration to cystathionine.

Methionine Regeneration

ln this pathway, 5,1 O-methyleneTHF is irreversibly reduced to 5-methylTHF by methyleneTHF reductase (MTHFR) in the cytoplasm (Fig. 2) (Green et al.

FASEB J. 1988). Subsequently, the methyl group of 5-methylTHF is transferred to homocysteine by vitamin B12-dependent methionine synthase (MS) to generate

THF and methionine, an essential amine acid in mammalian cells. Methionine is then converted to S-adenosylmethionine (SAM), which serves as the universal methyl donor for numerous cellular compounds, which include DNA, RNA, proteins, membrane lipids and neurotransmitters (Banerjee and Matthews 1990).

As a by-product of these methylation reactions, SAM is converted to S­ adenosylhomocysteine (SAH), which is then hydrolyzed back to homocysteine.

21 Alternatively, betaine-homocysteine methyltransferase (BHMT) catalyzes the transfer of the methyl group of betaine, which is derived from the catabolism of choline, to homocysteine to form dimethylglycine and methionine (Fig. 2). This

is a cytoplasmic enzyme that is specifie to the liver and kidney in mammals

(Sunden et al. 1997).

Since BHMT has a narrow expression pattern as compared to MS, which is

expressed in ail tissues, it was investigated just how much BHMT contributes to

methionine regeneration in mammals. It was shown that in rats maintained on a

methionine-deficient diet, there is a slight increase in BHMT activity (Park and

Garrow 1999). However, a source of methyl donor (i.e. betaine, choline) is

required to significantly induce BHMT activity. In addition, the magnitude of this

induction is proportional to the am ou nt of methyl donor in the diet and the seve rit y

of the methionine deficiency (Park and Garrow 1999). Consequently, the role of

BHMT is to conserve the backbone of homocysteine under conditions of

methionine deficiency and prevent its degradation to cystathionine by the

transsulfuration pathway. This implies that MS is the preferred pathway for the

regeneration of methionine, in contrast to an earlier hypothesis that predicted that

the activities of BHMT and MS contributed equally to the remethylation of

homocysteine (Finkelstein and Martin 1984).

22 Transsulfuration Of Homocysteine

ln this pathway, homocysteine condenses with serine to form cystathionine in a committed step catalyzed by the pyridoxal-5'-phosphate (PLP)-containing enzyme, cystathionine ~-synthase (CSS), which can only be detected in the liver and kidney (Fig. 2). Cystathionine is hydrolyzed to and a-ketoglutarate by another PLP-containing enzyme, y-cystathionase. The excess cysteine is oxidized to taurine and eventually to inorganic sulfates (Selhub 1999). The transsulfuration pathway allows for the catabolism of excess homocysteine in order to provide protection against the toxic effects of this metabolite, which will be discussed shortly.

Regulation Of The Methylation Cycle

These pathways are regulated by the SAM/SAH ratio, which reflects the overall methylating ability of the cell. This is achieved by the level of homocysteine being allosterically regulated by SAM whereby SAM inhibits

MTHFR but stimulates CSS and the transsulfuration pathway (Sanerjee and

Matthews 1990). The inhibition of MTHFR by SAM provides feedback regulation that protects against a folate methyl trap and ensures that during methionine repletion, folate-activated one-carbon units are spared for DNA precursor synthesis. Furthermore, SAH impedes the binding of SAM to MTHFR but it does not itself inhibit or activate MTHFR.

23 There is another component to this regulation of homocysteine. It involves glycine N-methyltransferase (GNMT), abundant in the liver, which catalyzes the

transfer of a methyl group to glycine from SAM to produce sarcosine and SAH

(Fig. 2). This enzyme is inhibited by 5-methylTHF pentaglutamate (Yeo et al.

1999). The purpose of GNMT is believed to serve as an alternate pathway for the

conversion of SAM to SAH in order to preserve the SAM/SAH ratio. The

regulatory mechanism of the inhibition of GNMT by 5-methylTHF links the

availability of preformed methyl groups from dietary methionine to the de novo

synthesis of methyl groups via the folate pools.

When there is a high dietary intake of methionine, it is rapidly converted to

SAM (Selhub and Miller 1992). The elevated SAM levels will cause an inhibition

of MTHFR, thereby suppressing the production of 5-methyITHF. This alleviates

the inhibition of GNMT. The high SAM concentrations also activate CSS, which

causes homocysteine to be shunted through the transsulfuration pathway.

Conversely, when there is a reduction in dietary methionine, there are inadequate

SAM levels to inhibit MTHFR. This leads to the production of 5-methylTHF, which

inhibits GNMT and thus conserves SAM but the synthesis of cystathionine will be

inhibited.

There are additional methods of maintaining low homocysteine levels. The

equilibrium of conversion of SAH to homocysteine catalyzed by SAH hydrolase

actually favors the formation of SAH and is driven forward only by removal of

products (Selhub and Miller 1992). Excess homocysteine can also be exported

from the cell and disposed into the blood and urine. This mechanism prevents

24 cell toxicity by maintaining low intracellular concentrations of homocysteine but it results in hyperhomocysteinemia and homocystinuria.

Disruptions ln The Methylation Cycle

Under conditions of either a genetic defect in one of the enzymes of

homocysteine metabolism (i.e. methyleneTHF reductase, methionine synthase or

cystathionine synthase) or a nutritional deficiency of one or more of the vitamins

that participate in homocysteine metabolism (i.e. folate, vitamin B12 or vitamin B6),

the most common manifestation of this disruption is hyperhomocysteinemia

(reviewed by Selhub 1999).

Methylene THF Reductase Defect

A defect in MTHFR causes a decrease in intracellular SAM concentrations.

It also impedes the production of 5-methyITHF, which will activate GNMT and further decrease the SAM levels. This ultimately results in the elevated synthesis

of homocysteine.

ln humans, the two most common mutant MTHFR alleles are C677T and

A 1298C (reviewed by Botto and Yang 2000). Patients that are homozygous for

the C677T allele usually have slightly increased blood homocysteine levels if they

have an inadequate folate status. Patients that are homozygous for the A 1298C

allele have normal serum homocysteine levels but individuals that are compound

25 heterozygous for the A 1298C and C677T alleles display a similar phenotype as

C677T homozygotes.

ln mice, the inactivation of this gene exhibits a similar phenotype as to

what is seen in humans, whereby Mthf(l- and Mthftl- mice have significantly

elevated homocysteine levels (Chen et al. 2001). Furthermore, these mice also

have low SAM and/or high SAH levels. The elevated SAH levels promote feedback inhibition of SAM-dependent methylation reactions, resulting in a

reduced methylation capacity, which manifests as global hypomethylation in these

Mthfr-deficient mice. The disruption in SAM regulation due to the loss of MTHFR

activity is responsible for the high homocysteine levels.

Cystathionine j3-Synthase Defect

I l The Cbs- - and Cbs+ - mice exhibit severe and moderate homocysteinemia,

respectively (Watanabe et al. 1995). In the case of a CSS defect, the

homocysteine is shunted towards the remethylation pathway. This results in an

elevation of the intracellular SAM levels, which continues until there is feedback

inhibition on MTHFR. At this point, there is inhibition of the remethylation pathway

in addition to the initial disruption of the transsulfuration pathways, which leaves

neither pathway to metabolize the accumulating homocysteine levels.

26 Methionine Synthase Defect

A defect in MS or a deficiency in its cofactor, cobalamin will result in an increase in 5-methylTHF but the decrease in SAM synthesis will lead to an accumulation of homocysteine. This was clearly demonstrated by the targeted

1 disruption of the Ms gene in mice (Swanson et al. 2001). The Ms+ - mice have slightly elevated plasma homocysteine and methionine levels but they are able to survive. However, the complete inactivation of the Ms gene results in embryonic lethality. Even BHMT cannot compensate for the MS deficiency because BHMT

1 is not expressed until the liver is formed and the Ms- - embryos die prior to this stage of development (Swanson et al. 2001 ).

Although the toxicity of the accumulation of homocysteine could account for the lethal phenotype of the MS deficiency in mice, there are other possible mechanisms. One of these is the folate methyl trap. This phenomenon usually arises as a result of an inhibition of MS or a deficiency in cobalamin, whereby there is an accumulation of 5-methylTHF (Shane and Stokstad 1985, Banerjee and MaUhews 1990). This 5-methylTHF is trapped due to the irreversibility of the

MTHFR activity and there is no regeneration of THF by the MS activity. The 5- methylTHF is a poor substrate for FPGS. This results in shorter chain forms of folate, which are no longer retained intracellularly and hence, there is a decrease in total intracellular folate levels. This is known as the folate methyl trap because ail the available THF is trapped as 5-methyITHF. The depletion of methyleneTHF, which is essential for de novo thymidylate synthesis is the primary cause of

27 megaloblastic anemia associated with either folate or cobalamin deficiency in humans.

Glycine N-Methyltransferase Defect

Humans that have defective glycine N-methyltransferase (GNMT) activity due to two missense mutations, are unable to metabolize methionine normally and have elevated plasma methionine and SAM levels (Luka et al. 2002). They also show evidence of mild liver disease with hepatomegaly and modest elevations of liver transaminases. These authors speculate that a possible reason for this phenotype is that the increased SAM levels may lead to an accumulation of one or more methylated products or it may cause a depletion of hepatic ATP.

DE NOVa THYMIDYLATE 8YNTHE818

As mentioned previously, 5,1 O-methyleneTHF is also essential for de novo thymidylate synthesis. Its methylene group is transferred to deoxyuridine monophosphate (dUMP) to generate DHF and deoxythymidine monophosphate

(dTMP) in a reaction catalyzed by thymidylate synthase (TS) (Fig. 2). The synthesis of deoxynucleotides, which requires ribonucleotide reductase and T8 is considered to be the rate limiting step in DNA synthesis (Lucock 2000). The DHF produced by T8 must be reduced to THF by DHFR so that it may participate

28 again in one-carbon transfer reactions. In rapidly praliferating cells, the elevated synthesis of thymidylate raises the levels of dihydrofolate. The dihydrafolate plays a regulatory raie by inhibiting MTHFR and thus allowing 5,1 O-methyleneTHF to be diverted to nucleic acid synthesis at the expense of methionine formation

(Matthews and Baugh 1980).

DE NOVa PURINE SYNTHESIS

ln the cytoplasm, folates play an essential role in de nova purine synthesis whereby the formyl group of 10-formylTHF is required for two separate steps: aminoimidazole-4-carboxamide ribonucleotide (AI CAR) transformylase catalyzes the transfer of a one-carbon unit to AICAR, and glycinamide ribonucleotide (GAR) transformylase catalyzes the transfer of a one-carbon unit to GAR (Fig. 2). They

become carbon atoms 2 and 8, respectively, of the purine backbone.

REMOVAL OF ONE-CARBON UNITS

Mammalian liver has been shown to contain high concentrations of 10- formylTHF dehydragenase (FDH). Consequently, it is logical to assume that it must play an important raie in cellular metabolism. It is a cytoplasmic enzyme that catalyzes the irreversible NADP-dependent oxidation of 10-formylTHF to carbon dioxide and THF, which serves as a mechanism for the removal of excess one-carbon units (Fig. 2).

29 The NEUT2 trait was one of the mutations detected in the F1 offspring of y­ irradiated male mice (Giometti et al. 1994). A preliminary analysis of NEUT2 homozygous mice revealed that these mice displayed a 35% reduction in their liver cell total folate pool. This was accompanied by a 2.5-fold elevation in the level of 10-formylTHF and a 4-fold decrease in the level of THF (Champion et al.

1994). The study concluded that this phenotype was due to a complete inactivation of the 10-formylTHF dehydrogenase enzyme activity. Therefore, it would appear that the physiological role of 10-formylTHF dehydrogenase is to remove excess 10-formyITHF, not required for purine synthesis, from the one­ carbon pool and regenerate THF, presumably to make it available for other reactions of folate metabolism. However, a subsequent study indicated that these

NEUT2 homozygous mice had an additional loss of urocanase activity, the second reaction in the folate-dependent histidine catabolite pathway (Cook 2001).

These authors propose that the lack of urocanase activity might have reduced the seve rit y of the 10-formylTHF dehydrogenase defect in these mice as evidenced by the viability of the NEUT2 mice.

There is some evidence that suggests that 10-formylTHF dehydrogenase may play a crucial role in removing formate during methanol toxicity. In rats, the retina, optic nerve and brain possess a high level of 10-formylTHF dehydrogenase activity, which protects these tissues against the accumulation of formate produced during the metabolism of methanol (Neymeyer and Tephly

1994). These same tissues are known to be targets of formate toxicity in humans, which are more affected by methanol poisoning because they possess lower

30 levels of 1O-formylTHF dehydrogenase.

ROLE OF FOLIG AGIO IN DISEASE

Since folate is critically important for cell function, division and differentiation, it is not surprising that an inadequate folate status has been implicated in numerous health disorders. There is irrefutable evidence that preconceptual folate supplementation can significantly reduce the risk of birth defects, particularly neural tube defects. A deficiency in folic acid has also been associated with megaloblastic anemia, cardiovascular disease, Down's syndrome and neurodegenerative disorders. The role of folic acid in the prevention of each of these medical conditions will be examined in this section.

NEURAL TUBE DEFEGTS

Folate metabolism is critical for the entire period of embryonic development. It is particularly important during neurulation, which comprises the formation of the neural tube, from which originates the complete central nervous system and a portion of the peripheral nervous system. This process is initiated by the induction of the fiat neural plate, which comes together to create the neural folds, which fuse to become the neural tube. Neural tube defects (NTD) are second only to heart defects as the most prevalent congenital malformation affecting on average one in 500 babies worldwide (Dolk et al. 1991 ).

31 Periconceptual folic acid supplementation can prevent the incidence of neural tube defects by more than 70% (Czeizel and Dudas 1992). Consequently, ail women contemplating pregnancy are recommended to increase their folate intake

(Refsum 2001). This also prompted the Food and Drug Administration of the

United States to issue a mandate requiring ail enriched cereal-grain foods to be fortified with folic acid, which was in effect by 1998 (Quinlivan and Gregory 2003).

Since the formation of the neural tube is a multifactorial process, it

suggests that NTD are also of a multifactorial origin, involving both genetic and

environmental factors contributing to the onset of NTD. Although there are known

causes of NTD which include single-gene mutations and chromosomal

abnormalities, greater than 90% of NTD cases have no known etiology

(Rosenquist and Finnell 2001). The mechanisms by which folate mediates a

protective effect against neural tube defects are still undetermined. One

hypothesis suggests that maternai folate supplementation simply overcomes

borderline folate levels in the fetus. The cells of highly vulnerable areas of the

embryo, such as the neural epithelium, may be extremely sensitive to maternai folate concentration. Lower concentrations might impede folate transport or

reduce the flux of folates through a particular pathway. Collectively, this may

affect metabolism of folic acid in the embryo, which could lead to developmental

abnormalities of the neural tube. The studies involving the Folbp1 knockout

mouse support this hypothesis (Piedrahita et al. 1999). The phenotype of the

Folbpr/- embryos consisted of gross abnormalities of the neural tube, which is in

accordance with the localization of this gene almost exclusively in the neural

32 epithelium and its responsibility in supplying the embryonic folate requirement.

However, there is no known polymorphism of the human FRa gene (Van der Put et al. 2001).

There are several studies that have demonstrated elevated levels of homocysteine in women with babies that have NTD. A defect in certain candidate enzymes in folate metabolism may be associated with the development of NTD, particularly the enzymes involved in the remethylation and transsulfuration of homocysteine (Van der Put et al. 2001). A disturbance in any of these pathways could result in the elevated levels of homocysteine that are documented in mothers with babies with NTD. It was observed that the frequency of the C677T mutation in the Mthfr gene, which is linked to elevated plasma homocysteine concentrations, is high in babies with NTD and their mothers in some populations.

Consequently, this mutation was the first identified genetic risk factor for NTD.

There are observations to indicate that elevated homocysteine levels may also be a factor in the development of neural tube defects, despite normal folate levels (Rosenquist and Finnell 2001). Homocysteine is an antagonist of the N­ methyl-D aspartate (NMDA) receptor. A general characteristic of NMDA antagonists is that they are able to promote abnormal neural tube development.

ln fact, homocysteine can act synergistically with other NMDA antagonists to further disrupt neural tube development. This effect can be reversed by NMDA agonists.

33 MEGALOBLASTIC ANEMIA

Since folate is a critical component of thymidine and purine biosynthesis, it plays an essential role in cell division. It is especially important in rapidly proliferating ce Ils su ch as the hematopoietic tissue of bone marrow. Folate deficiency results in the prolongation of the synthesis phase of cell division and a delay in germ cell maturation. In the case of bone marrow, this leads to abnormally large red blood cell precursors, known as megaloblasts, which are unable ta divide properly. This causes inefficient erythropoiesis whereby the delivery of new erythrocytes into circulation is severely compromised. The primary cause of megaloblastic anemia appears to be a depletion in methyleneTHF which is required for thymidylate synthesis (8anerjee and

Matthews 1990). This condition is often associated with pregnancy as a result of the increased requirement for folate. It has been shown that folate supplementation is able to treat and prevent megaloblastic anemia (Fishman et al.

2000).

A deficiency in vitamin 8 12 (cobalamin) can also produce megaloblastic anemia as a result of the folate methyl trap. Although the prevalent cause of megaloblastic anemia is folate deficiency, the treatment of this condition with folate alone can mask concomitant vitamin 812 deficiency. Consequently, it is recommended that megaloblastic anemia be treated with both vitamins (Fishman et al. 2000).

34 CARDIOVASCULAR DISEASE

Cardiovascular disease, specifically coronary heart disease (CHD), is one of the prominent causes of death in industrialized countries. In CHD, there is obstruction of the coronary arteries, which supply blood to the heart. This blockage usually begins with the formation of artherosclerotic plaques. The growth of these plaques can eventually lead to thrombosis, the formation of blood clots and obstruct the blood flow through the arteries (De Bree et al. 2002).

It has been suggested that one of the contributing factors to CHD is an increased concentration of plasma homocysteine. One of the earliest studies to support this hypothesis involved the observation that an infant with homocystinuria due to CBS deficiency had similar arterial damage as another

infant that also had homocystinuria due to MS deficiency (McCully 1969). Severe

MTHFR and MS deficiencies occur less frequently than CBS deficiency but they

also result in homocystinuria. Irrespective of the actual enzyme defect, it was

reported that that there was a high prevalence of premature vascular disease,

even constituting the major cause of death among the untreated patients (De

Bree et al. 2002).

There have been several mechanisms proposed to explain how

hyperhomocysteinemia increases the risk of coronary heart disease based on the

atherogenic and thrombogenic effects of elevated concentrations of

homocysteine. However, there are several problems with extrapolating the

results of these studies to apply them to human subjects. The main one is that

35 the majority of these studies use concentrations of homocysteine that are too high to actually reflect conditions of hyperhomocysteinemia in humans. Also, the

effects observed in some of these studies may not be limited to homocysteine but

may be accounted for by other sulfur-containing amino acids, such as cysteine

(McKinley 2000, Verhaar et al. 2002). Consequently, the association between

elevated homocysteine levels and an increased risk for cardiovascular disease

appears to be correlative.

There was an interesting observation made among patients receiving

vitamin supplementation for CSS deficiency who showed a significant

improvement in their vascular prognosis despite still possessing higher than

normal levels of homocysteine (Ashfield-Watt et al. 2001). This suggested that

there might be another homocysteine-independent mechanism involved in

reducing the risk for cardiovascular disease.

Folate is known to reliably reduce plasma homocysteine levels and thus

may indirectly prevent cardiovascular disease. There have been studies which

demonstrate a protective role of folate against cardiovascular disease. Oral folate

supplementation can restore endothelial function in patients with

hypercholesterolemia but with normal plasma homocysteine levels (Verhaar et al.

2001). The mechanism proposed is that the antioxidant properties of folate can

preserve the functional integrity of the vascular endothelium, which is essential for

the prevention of atherosclerosis and thrombosis.

36 DOWN'S SYNDROME

Down's syndrome is a genetic defect occurring with a frequency of approximately one in 600 live births. It results from the presence of three copies of chromosome 21 due to abnormal chromosomal segregation during meiosis but the mechanism predisposing to altered recombination is unknown. One study has found that the presence of the C677T MTHFR mutation on one or both alleles resulted in a pronounced increase in the risk of having a baby with Down's syndrome (James et al. 1999). In addition, they also observed a significantly higher plasma homocysteine concentration in the mothers of babies with Down's syndrome, which is indicative of abnormal folate and methyl metabolism (James et al. 1999). This is consistent with studies that have demonstrated that genomic hypomethylation is associated with chromosomal instability and abnormal segregation. These preliminary results indicating possible maternai risk factors for Down's syndrome still need to be confirmed with further studies.

CANCER

Folate also seems to have a potential role as a cancer preventative agent.

A diminished folate status has been associated with various types of cancer, particularly colorectal cancer (reviewed by Kim 1999). Folate depletion enhances carcinogenesis by disrupting the balance between the partitioning of cellular folates towards the methylation and nucleotide synthesis pathways. This is

37 supported by the interesting observation that folate-depleted men who are homozygous for the MTHFR C677T mutation have half the risk of colorectal cancer compared with men who are either wild type or heterazygous for the same mutation. It has been suggested that the cancer-pratective effect of the MTHFR

C677T mutation is related to the increased availability of 5,1 O-methyleneTHF, which can now support nucleotide synthesis (Choi and Mason 2002).

Genomic hypomethylation is an early sign of carcinogenesis but gene­ specifie hypomethylation may be more relevant due to damage at critical loci within DNA in carcinogenesis. This is demonstrated by the "hypermutable region" of the p53 tumor suppressor gene, which is particularly susceptible to hypomethylation by folate depletion. RNA methylation is required to maintain

RNA stability and facilitate its transport acrass the nuclear membrane (Choi and

Mason 2002). It has been recently shown that folate deficiency can cause hypomethylation of some RNA species as weil, notably small nuclear RNA, which is essential for the maturation of messenger RNA.

Folate deficiency has been shown to generate breaks in the phosphodiester backbone of DNA, which result in chramosomal breaks and elevate the risk for cancer (Choi and Mason 2002). It is known that 5,10- methyleneTHF is a cofactor for thymidylate synthesis fram deoxyuridine monophosphate in the reaction catalyzed by thymidylate synthase. Folate deficiency decreases thymidylate synthesis, which increases the misincorporation of uracil bases into DNA. A repair glycosylase excises the uracil fram the DNA strand and creates strand breaks in the DNA. Furthermore, these breaks in DNA

38 can also be induced by deamination of non-methylated cytosine to uracil and the subsequent removal of the uracil.

The disruption of DNA repair is another mechanism by which folate depletion mediates carcinogenesis. In a study with folate-deficient rodents, there was an impairment in the DNA excision repair machinery, partly due to an imbalance in the cellular pool of deoxyribonucleotides (Choi and Mason 2002).

Since DNA methylation is critical for DNA strand discrimination during

postreplication mismatch repair, it may be affected by site-selective DNA hypomethylation resulting from folate deficiency.

The timing of folate chemoprevention is very important. Preliminary evidence from a murine model of colon cancer indicates that moderate folate supplementation strongly protects against development of intestinal and colorectal tumors but this is contingent upon the treatment being implemented prior to the establishment of microscopie neoplastic foci (Song et al. 1999). However, if folate supplementation commences after this point, there is actually promotion of tumor

development because the demand for additional folate for the accelerated DNA

replication and cell division in neoplastic cells is being satisfied.

NEURODEGENERATIVE DISORDERS

Recent epidemiological studies have associated folate deficiency and the

resulting elevated homocysteine levels with neurodegenerative diseases.

Alzheimer's disease (AD) is one of the most common neurodegenerative

39 disorders that usually affect the elderly population. It is believed to occur as a

result of the accumulation of the neurotoxic self-aggregating amyloid ~-peptide. It

has been observed that folate levels are significantly lower in the cerebrospinal fluid of patients with AD (Miller 2002, Mattson and Shea 2003). Furthermore, folate deficiency and increased homocysteine appear to interact synergistically to

exacerbate the effects of other established neurotoxic agents that contribute to

the development of AD.

There have been a few mechanisms postulated to outline the detrimental

effects of homocysteine on AD. Impaired DNA repair and increased oxidative

stress result in DNA damage, which ultimately activates apoptosis (Miller 2002,

Mattson and Shea 2003). It is also possible that homocysteine may mediate its

neurotoxic effects by damaging cerebral blood vessels.

Parkinson's disease (PD) is another prominent neurodegenerative disease

among the elderly. It is a motor disease resulting from a progressive

degeneration of dopaminergic neurons in the substantia nigra. A series of studies

involving various models of PD suggest that folate deficiency sensitizes

dopaminergic neurons to the effects of environmental neurotoxins, su ch as MPTP

(1-methyl-4-phenyl-1 ,2,3,6-tetrahydropyridine), rotenone and excess iron, which

have been implicated in the pathogenesis of PD (Miller 2002, Mattson and Shea

2003). Folate deficiency possibly contributes to the severity of PD through a

homocysteine-mediated pathogenetic mechanism. Although there is no

conclusive evidence, low folate levels are implicated as a serious risk factor for

PD.

40 INTERCONVERSIONS OF ONE-CARBON FOLATES

Proper folate-mediated metabolism requires that sufficient amounts of the vitamin be present in the diet and that cells and tissues are able to take it up and

convert it to metabolically active forms. Obviously a folate deficiency will be

detrimental to ail cells of an organism, but certain cells and tissues might exhibit

different responses. In general, one-carbon tetrahydrofolates are used to support

either methyl- or formyl- transfer reactions. 5,1 O-methyleneTHF represents an

entry point to either thymidylate synthesis or the methylation cycle while 10-

formylTHF represents an entry point to purine biosynthesis. Consequently, these

pathways are in competition for the pool of one-carbon donors and it is important

to understand how the necessary balance is achieved to meet the requirements

for ail the products of folate metabolism. 1 will describe the enzymes that are

responsible for the interconversion of these key methyleneTHF and formylTHF

intermediates in the following section.

As was stated earlier, there are three separate enzyme activities, 5,10-

methyleneTHF dehydrogenase (0), 5,1 O-methenylTHF cyclohydrolase (C) and

10-formylTHF synthetase (S), which interconvert the one-carbon substituents of

THF. They can exist as several different combinations in nature. In prokaryotes,

the 0 and C activities can exist as separate monofunctional proteins or together

as a bifunctional protein. There are certain prokaryotes that, in addition to a

bifunctional OC protein, also possess a monofunctional S protein (Le. Clostridium

thermoaceticum and Clostridium acidiurici). In eukaryotes, the three activities are

41 usually present as a trifunctional protein (reviewed by MacKenzie 1984).

While the monofunctional and bifunctional D proteins can be either NAD- or

NADP-dependent, the trifunctional D proteins are NADP-dependent.

MAMMALIAN NADP-DEPENDENT DeS

The NADP-dependent DeS is expressed ubiquitously in various mammalian tissues and cell lines (both normal and transformed) (Mejia and

MacKenzie 1985). This enzyme is localized to the cytoplasm in mammalian cells

(Fig. 3) (Mejia and MacKenzie 1988). Another study has suggested that there is also a mitochondrial NADP-dependent DeS (Barlowe and Appling 1988). The trifunctional DeS protein exists as a dimer (Mejia and MacKenzie 1985). The human trifunctional DeS gene maps to chromosome 14q24 (Rozen et al. 1989)

and an intronless pseudogene maps to the X chromosome at Xp11 (Italiano et al.

1991 ).

42 FIGURE 3. The compartmentalization of folate metabolism in mammalian cells.

43 Methyl-Acceptor Acceptor

S-adenosylhomocysteine> S-adenosylmethionine< Methylation l Serine • ) Serine Cycle " Methionine THF ~ __ 1,.--= ) THF l SHMT SHMT Homocysteine ~ ~Glycine • ) Glycine • ~ 5-methylTHF ~_ :::: MethyleneTHF GCS'-.. MethyleneTHF Thymidylate .. - NADP :i NAD!i Dehydrogenase Pi Dehydrogenase NADPH NADH MethenylTHF MethenylTH F

Cyclohydrolase l lCyc/ohydrolase

Purines ....----- FormylTHF FormylTHF

ADP+Pi~ynthetase ? /' \ ,:!ethionyltRNA './ \OrmYltranSferase ATP THF ? Formate .. ----- .:. -1- --- ~Formate Met-tRNNmet

Cytoplasm Mitochondria The DCS is expressed in various tissues but kidney and liver contain the highest levels of expression (Thigpen et al. 1990, Peri and MacKenzie 1991). It was proposed that the regulation of the trifunctional enzyme is predominantly at the pretranslational level (Thigpen et al. 1990). At the transcriptional level, the

DeS gene is not inducible by mitogenic agents in mouse fibroblasts implying that it is a house-keeping gene whose expression is influenced by tissue-specifie factors (Peri and MacKenzie 1991).

The mammalian DCS is composed of two functional domains: an amino­ terminal domain (33 kD) consisting of the 0 and C activities (Tan and MacKenzie

1977) and a carboxy-terminal domain (67 kD) consisting of the S activity (Tan and

MacKenzie 1979). The mammalian DCS protein consists of 935 amine acids.

Residues 292-310 constitute the linker region that connects the D/C and S domains (Hum and MacKenzie 1991). It is a hydrophilic region of disordered conformation that is flanked by a-helices, making it susceptible to proteolysis.

This sequence is not conserved among different species containing trifunctional

DCS proteins, which is consistent with its role as a linker domain. Furthermore, the two domains are able to fold independently of each other (Hum et al. 1988,

Hum and MacKenzie 1991). The S domain is more highly conserved among different species than the D/C domain (Thigpen et al. 1990).

There has been significant work done on the mutational analysis of the different activities of the trifunctional DCS protein in mammals. The first known study was an Ade-E mutant in Chinese Hamster Ovary cells. This mutant possessed diminished catalytic activities of ail three activities and demonstrated

44 purine auxotrophy (Mascisch and Rozen 1991).

Recently, the studies involving the site-directed mutagenesis of the human

OC301 have proven to be quite informative. The X-ray crystallographic model of

OC301 (Allaire et al. 1998) was used to identify residue R173 as a residue likely to be involved in the binding of the 2'-phosphate of NAOP. When this residue was mutated to a glutamate, the dehydrogenase activity was undetectable due to the drastically reduced ability to bind NAOP but there was little effect on the cyclohydrolase activity (Pawelek et al. 2000). This R173E mutant essentially represents a monofunctional cyclohydrolase protein.

Another study (Schmidt et al. 2000) indicated that the K56 residue was

important in cyclohydrolase catalysis. A subsequent study (Sundararajan and

MacKenzie 2002) explored the role of residues surrounding K56, specifically

0100 and 0125. The K560/0100K double mutant had no detectable

cyclohydrolase activity but retained over two-thirds of the wild type

dehydrogenase activity. This indicates that the function of 0100 is to activate K56 for cyclohydrolase catalysis. Furthermore, neither dehydrogenase nor

cyclohydrolase activity is detectable in a series of 0125 mutants, indicating that

this residue is required for the binding of tetrahydrofolate derivatives.

45 MAMMALIAN NAD-DEPENDENT DC

The only known mammalian NAD-dependent 0 protein is the bifunctional

OC (NMOMC). It is unique in that it requires magnesium as weil as inorganic phosphate for activity, which is not demonstrated by the bacterial NAO-dependent o proteins (Mejia and MacKenzie 1985, Mejia et al. 1986). This protein is expressed in mammalian embryonic tissues and immortalized or transformed cell lines (Mejia and MacKenzie 1985, Mejia et al. 1986). However, NMOMC activity is not detected in normal adult differentiated cells and tissues. The NAO­ dependent OC is located in the mitochondria (Fig. 3) (Mejia and MacKenzie

1988). There are two NMOMC mRNAs, which are due to alternative polyadenylation signais (Belanger and MacKenzie 1989).

The regulation of NMOMC expression occurs predominantly at the transcriptionallevel, whereby 5'-regulatory sequences and trans-activating factors are most likely involved in its expression (Belanger and MacKenzie 1989). One particular sequence in the 5'-flanking region of the NMDMC gene is homologous to the serum responsive element (SRE), which is situated in the promoter region of the c-fos gene. Consequently, it was demonstrated that the NMOMC promoter does respond to stimulation by serum (Belanger et al. 1991).

The metabolic role of NMOMC has been the focus of ongoing investigations. The fact that it exhibits NAD specificity and that the intracellular

[NAO]/[NAOH] is very high has elicited speculation that it must shift the interconversion of one-carbon folates toward 10-formylTHF and ultimately

46 formate, which could exit into the cytoplasm and supply one-carbon units for purine biosynthesis. This would explain its expression to be limited to rapidly growing cells such as immortalized and embryonic cells (Mejia and MacKenzie

1985), which have an increased demand for purines.

Its expression in embryonic cell lines implies a role in growth and

development. This is supported by analysis of the NMDMC knockout mice, which

exhibit an embryonic lethal phenotype (Di Pietro et al. 2002). It has also been

suggested that the NMDMC plays a role in mitochondrial biogenesis by

supporting mitochondrial protein biosynthesis. Methionyl-tRNA formyltransferase

catalyzes the transfer of the formyl group from 10-formylTHF to the initiator

fMet methionyl-tRNA to produce met _tRNA . The fMet group permits the initiation factor IF-2 to recognize the initiator tRNA which initiates protein synthesis. It also

blocks the binding of the elongating factor EF-Tu and prevents the initiator met­

tRNAfMet from being misincorporated into the elongating protein. However, recent

studies involving the disruption of the gene encoding methionyl-tRNA formyltransferase (FMT1), which is required for the synthesis of formylmethionine

fMET tRNA , have demonstrated that this formylation step is not essential for the

initiation of protein synthesis in the mitochondria of Saccharomyces cerevisiae (Li

et al. 2000). This formyltransferase activity also appears not to be essential in

certain bacterial species such as Pseudomonas aeruginosa (Newton et al. 1999).

It seems unlikely that NMDMC is required in this process since the null mutant cell

lines have intact and functional mitochondria (Di Pietro et al. 2002, Patel et al.

2003).

47 CATALYTIC PROPERTIES OF D/C DOMAIN

The pools of 5,1 O-methyleneTHF and 10-formylTHF are maintained near equilibrium in the cytoplasm. Therefore, alterations in the cellular requirements for either reactant can efficiently draw from the pool of the other reactant, which would allow for efficient use of both serine and formate as one-carbon donors to the folate pool (Pelletier and MacKenzie 1995).

It was shown that NADP, a substrate for the 0 activity, inhibits the C activity. This implied that a single folate binding site exists per monomer of the

D/C domain (Smith and MacKenzie 1985, Pelletier and MacKenzie 1995). It was later demonstrated that the O/C domain actually contained a "shared" catalytic site (Pelletier and MacKenzie 1994). In addition, the O/C domain channels methenylTHF from the 0 to the C in the forward direction at -50% efficiency.

However, in the reverse direction, methenylTHF is channeled at practically

-100% efficiency with the reverse C activity being rate-limiting (Pawelek and

MacKenzie 1998). This serves to prevent the methenylTHF generated from 10- formylTHF from being released and hydrolyzed back to 10-formyITHF.

These results suggest that the C activity is not required for the forward direction because methenylTHF can be hydrolyzed nonenzymatically to produce

10-formyITHF. However, the C activity is essential for the reverse direction so as to channel ail of the methenylTHF to the 0 activity for the production of methyleneTHF. Therefore, the NAOP-dependent OCS is optimized to catalyze the reverse reaction in the cytoplasm of mammalian cells.

48 The mammalian mitochondrial OC protein is also able to channel in the forward direction at -50% efficiency. However, the C activity of the NAD­ dependent OC is kinetically independent of the 0 activity, in that NAD does not inhibit the C activity of the bifunctional OC (Rios-Orlandi and MacKenzie 1988). In addition, the kinetic properties of the mitochondrial OC protein suggests that its cyclohydrolase activity is not optimized to catalyze the reverse reaction as is the case for the cytoplasmic NADP-dependent OC domain (Pawelek and MacKenzie

1998). This provides further support that the metabolic role of the mitochondrial

OC is to provide 10-formyITHF, which can be ultimately converted to formate and exit the cytoplasm to supply one-carbon units for purine biosynthesis.

YEAST DEHYDROGENASES

Yeast have a trifunctional NADP-dependent DCS protein in both the cytoplasm and the mitochondria, encoded by the ADE3 and MIS1 genes, respectively (Fig. 4). Disruption of the MIS1 gene has no detectable effect on cell growth nor is it required to support the formylation of methionyl-tRNAfMET for mitochondrial protein synthesis (Shannon and Rabinowitz 1988). However, deletion of the ADE3 gene results in adenine auxotrophy. Interestingly, one study

(Barlowe and Appling Mol. Cell Biol. 1990) reported that as long as there was expression of a full-Iength cytoplasmic DCS, point mutations that affected one or

more of the catalytic activities of this trifunctional protein did not result in adenine auxotrophy. Their results suggested that adequate cytoplasmic 10-formylTHF

49 must be produced by another yeast cytoplasmic dehydrogenase but efficient utilization of the 10-formylTHF for purine synthesis requires the ADE3 gene product as a structural component. However, a subsequent study (Song and

Rabinowitz 1993) challenged this hypothesis by demonstrating that the catalytic activity of the cytoplasmic DCS was indeed involved in purine biosynthesis. They also showed that expression of either the DC or S domain alone of this protein was enough to complement the adenine requirement in ADE3 mutant strains.

50 FIGURE 4. The compartmentalization of folate metabolism in yeast.

51 MethYI-):eptor Aztor

S-adenosylhomocysteine S-adenosylmethionine Methylation t Serine • ) Serine Cycle .. Methionine THF ~ r-- ) THF l SHMT SHMT Homocysteine 1 I-----"Glycine • ) Glycine ~ ~ 5-methylTHF ~ __ -:: MethyleneTHF GCS'-. MethyleneTHF Thymidylate .. -- ~ADP:i NADP:i Dehydrogenase Dehydrogenase NADPH NADPH MethenylTHF MethenylTHF

lCyclohydrolase lCyclohydrolase

Purines ~ ----- FormylTHF FormylTHF ADP+Pi ADP+Pi:k: Synthetase Synthe~ase~ ATP ATP THF THF Formate • ) Formate Met-tRNNmet

Cytoplasm Mitochondria Inspite of the controversy surrounding the actual role of the cytoplasmic

DCS, a second dehydrogenase was found in the yeast cytoplasm. This protein is the NAD-dependent monofunctional 0 protein, the only one known in eukaryotes

(Barlowe and Appling Biochemistry 1990). It is encoded by the MTD1 gene.

Disruption of this gene in yeast strain CBY6, which expresses full-Iength, but catalytically inactive cytoplasmic DCS, results in adenine auxotrophy (West et al.

1993). Further analysis of the yeast cytoplasmic 0 activities has shown that the monofunctional NAD-dependent 0 is interchangeable with the trifunctional NADP­ dependent DCS when flow of one-carbon units is in the oxidative direction but the former does not participate significantly when the flux is in the reductive direction

(West et al. 1996).

Collectively, these results allow some important conclusions. The C activity is required to catalyze the overall reaction in vivo, which was also shown in mammalian cells. Furthermore, the role of the NAD-dependent 0 activity is in the oxidation of cytoplasmic one-carbon units.

INSECT DEHYDROGENASES

Both the NADP-dependent DCS and the NAD-dependent OC are detectable at ail stages of insect development but they are differentially expressed

(Tremblay et al. 1995). The NAD-dependent OC is expressed at low levels in adult tissues except for the testes, whereas the NADP-dependent DCS remains at relatively high levels. As in mammals, NMDMC is tightly regulated.

52 The distribution of dehydrogenases in insect cell lines differs from that in mammalian cells. There are undetectable levels of the NADP-dependent DCS in insect cell lines. Instead, there is a NAD-dependent OC in the cytoplasm, which has similar properties to its mammalian homolog. However, it is a truncated protein lacking a mitochondrial targeting sequence because translation is initiated at a downstream AUG start site (Tremblay et al. 1995). The purpose of the

NMDMC localization to the cytoplasm appears to be to provide one-carbon units for purine synthesis. This is a general characteristic of ail insect cell lines analyzed but the reason for this is unknown.

INTERDEPENDENCE OF FOLATE COMPARTMENTS

Although it is known that there are two compartments of folate metabolism

in mammalian cells, it has remained unclear how they share function. The analysis of various spontaneously selected mutants in the folate pathways has

aided in the establishment of the interdependence of these two compartments.

AUXB 1 MUTANT

The Chinese Hamster Ovary (CHO) auxB1 mutant cell line lacks both

cytoplasmic and mitochondrial FPGS activities (McBurney and Whitmore 1974).

These cells have a diminished level of intracellular folates due to an inability to synthesize folylpolyglutamates, which leads to the efflux of monoglutamates. In

53 addition, these cells are auxotrophic for glycine, purines and thymidine. When the

AUXB1 cells are transfected with the human cytoplasmic FPGS, they no longer require thymidine and purines for growth but they are still glycine auxotrophs

(Garrow et al. 1992). This implies that there was a restoration of only the cytoplasmic folate pools. However, when the AUXB1 cells were transfected with the mitochondrial FPGS, they were no longer auxotrophic for thymidine, purines or glycine, which implies a restoration of both the cytoplasmic and mitochondrial folate pools (Freemantle et al. 1995, Chen et al. 1996).

These studies demonstrate that the mitochondria are able to contribute to

cytoplasmic folate pools. In a subsequent study, folylpolyglutamates synthesized

in the mitochondria were shown to be released into the cytosol. However, folylpolyglutamates cannot enter the mitochondria. Although the mitochondrial folate efflux may be a significant factor in expanding the cytosolic folate pool, this

actual rate of efflux would be insignificant as compared to the rate of one-carbon fluxes that occur in one-carbon metabolism in mammalian cells (Kim and Shane

1994, Lin and Shane 1994).

GlyA MUTANT

The CHO glyA mutant cell line is deficient in mSHMT activity but retains

cSHMT activity (Chasin et al. 1974, Pfendner and Pizer 1980). These mutant

cells are glycine auxotrophs. This implies that the mitochondria are the primary

site for glycine synthesis whereas the cytoplasm is responsible for serine,

54 thymidylate, purine and methionine biosynthesis (Chasin et al. 1974, Pfendner and Pizer 1980, Narkewicz et al. 1996).

Interestingly, the disruption of the SHM1 and the SHM2 genes, which encode for the yeast mSHMT and cSHMT respectively, alone or together, does not result in an observable phenotype. The additional disruption of the yeast

GL Y1 gene, which encodes for aldolase, is required for the cells to exhibit glycine auxotrophy (McNeil et al. 1994). Consequently, unlike the mammalian system, yeast glycine synthesis is not confined to the mitochondria

but requires a significant contribution from the cSHMT. It would appear that caution must be taken when comparing the mammalian and yeast folate metabolic pathways as there are obvious differences.

GlyB MUTANT

The CHO glyB mutant cell line is also auxotrophic for glycine (Taylor and

Hanna 1982). These mutant cells have normal mFPGS and mSHMT activities as weil as normal cytoplasmic folate metabolism. However, there is a low content of

mitochondrial folylpolyglutamates. The glycine auxotrophy is alleviated by folinic

acid treatment. This suggests that the inability to accumulate mitochondrial folylpolyglutamates may be due to a defect in the transport of folate into the

mitochondrial matrix from the cytoplasm.

The human mitochondrial folate transporter (hMFT), which was shown to transport folates into the mitochondria, was recently cloned (Titus and Moran

55 2000). This ATP-independent mitochondrial inner membrane protein transporter consists of six transmembrane domains and three interspersed loops facing the mitochondrial matrix. The hMFT was shown to complement the glycine

auxotrophy and the lack of mitochondrial folate accumulation in the gly8 cells.

There has also been evidence to demonstrate the colocalization of a folate transport protein to the mitochondrial membrane of human T -cell CCRF-CEM

Iymphoblastic leukemia cells (Trippett et al. 2001). It is uncertain whether this

protein is actually RFC1 or another related protein; the latter case would imply the

presence of more than one folate carrier protein in the mitochondria of

mammalian cells.

The cloning of the hMFT gene has provided evidence that reduced folates

can cross between the mitochondria and the cytoplasm but this exchange of folates does not occur at a metabolically significant rate to support the one-carbon

requirements of the folate compartments. It has been demonstrated that at

physiological pH, serine and glycine exist primarily as zwitterions and these

neutral amino acids are able to permeate the mitochondrial inner membrane

(Cybulski and Fisher 1976). Therefore, it would appear that the one-carbon

donors, namely serine, glycine and formate are the only components of one­

carbon metabolism that are transported between these two folate compartments.

56 FLUX OF ONE-CARBON METABOLISM

Once the interdependence of the mitochondrial and cytoplasmic folate pathways had been established, it became relevant to determine the direction of the flux of folate metabolites in these two compartments in eukaryotic cells. There are several factors that influence this directionality of the flow of the folate

pathways. Firstly, the cellular concentration of folate-binding proteins exceeds that of folate derivatives (Schirch and Strong 1989), which implies that the

concentration of free folates is low. This suggests that there is a competition for the restricted supply of one-carbon units carried by folate derivatives between the

serine-to-glycine interconversion and methionine regeneration pathways as weil

as purine and thymidylate synthesis. Furthermore, some of the folate-dependent

reactions are reversible whereas others are irreversible. These irreversible

reactions are important in regulating the availability of folate derivatives to the

other pathways. It is also thought that the modulation of the length of the

glutamate chain of folate may regulate the flow of folate metabolites through the various folate pathways.

Yeast

One of the most extensively studied eukaryotic folate systems is that of

Saccharomyces cerevisiae. There have been numerous studies to establish the

preferred flux of one-carbon metabolites in both the mitochondria and the

57 cytoplasm in yeast. Early NMR experiments have shown that serine is metabolized to formate in the mitochondria. This formate exits into the cytoplasm

and contributes to at least 25% of the one-carbon units supplied to purine

synthesis via the synthetase activity of the cytoplasmic DCS (Pasternack et al.

1994). Subsequent studies have indicated that there are two pools of 5,10-

methyleneTHF in the mitochondria (McNeil et al. 1996). The mSHMT, encoded

by the SHM1 gene, contributes to the methyleneTHF pool that is used to generate

the formate that provides the one-carbon units for purine synthesis whereas the

GCV (glycine c\eavage system) contributes to the other methyleneTHF pool,

which is used to synthesize serine in the mitochondria. However, when there is a

disruption of the mSHMT activity, GCV is observed to make a significant

contribution to formate production in the mitochondria. When both the SHM1 and

the MIS1 genes, the latter encoding the mitochondrial DCS, are simultaneously

inactivated there is no effect on the growth of these cells. These studies suggest

that although the mitochondrial folate pathway makes a contribution to

cytoplasmic folate metabolism, it is not absolutely required.

Further characterization of yeast mutant strains have demonstrated that

unlike mammalian cells, the disruption of mitochondrial SHMT activity does not

result in glycine auxotrophy (McNeil et al. 1994). The generation of glycine

auxotrophs in yeast requires the simultaneous disruption of the SHM2 gene,

encoding the cytoplasmic SHMT activity, the GL Y1 gene, encoding the threonine

aldolase activity as weil as the SHM1 gene. In fact, as compared to the

mitochondrial SHMT, the cytoplasmic SHMT makes a more significant

58 contribution to glycine synthesis (McNeil et al. 1994). The mitochondrial SHMT functions predominantly in serine cleavage to ultimately produce formate, which can supply one-carbon units for cytoplasmic folate metabolism and the cytoplasmic SHMT functions mostly in the catalysis of serine synthesis.

Unlike mammalian cells, yeast are highly adaptable to their environ ment.

This was demonstrated in a subsequent study (Kastanos et al. 1997), which revealed that although the two SHMT isozymes usually operate in opposing directions, their directional flux is influenced by the nutritional status of the cells.

When serine is available as the primary source of one-carbon units, the cytoplasmic SHMT is the main source of glycine and one-carbon units for purine synthesis (McNeil et al. 1994, Kastanos et al. 1997). However, under normal conditions, when both serine and glycine are available in the medium, the mitochondrial SHMT makes a significant contribution to purine synthesis.

A recent study (Piper et al. 2000) that investigated the regulation of the balance of one-carbon metabolism in yeast has proposed a model whereby this control is mediated by the levels of 5,1 O-methyleneTHF in the cytoplasm. In this model, this folate intermediate modulates the transcriptional control of the expression of the GCV genes encoding the glycine decarboxylase complex.

Under normal conditions, cytoplasmic 5,1 O-methyleneTHF is mainly involved in methionine regeneration in order to provide methyl groups for the various cellular methylation reactions whereas the one-carbon units derived from the mitochondrial folate pathway are directed towards purine synthesis. Therefore, the cytoplasmic 5,1 O-methyleneTHF represses the GCV genes because there is

59 no requirement for alternate sources of one-carbon units from this glycine decarboxylase complex pathway. When the levels of cytoplasmic 5,10- methyleneTHF are diminished, either due to one-carbon starvation or the

presence of excess of glycine, which forms part of an inhibitory complex with

8HMT, the cell shifts to mitochondrial glycine catabolism via the GCV genes to

supplement its requirements for one-carbon units.

Mammals

The determination of the flux of folate metabolites in mammalian cells has

involved studies that have taken different approaches. The earlier experiments

investigated the effects of the polyglutamate chain length on the regulation of

one-carbon metabolism. There have been a series of studies performed on four

mammalian enzymes, 8HMT, D, T8 and MTHFR, which ail represent a branch

point in the folate pathway because they ail use methyleneTHF as a substrate.

8HMT and D catalyze reversible reactions that are thought to be maintained at

equilibrium and therefore, neither is believed to play an important role in

determining the direction of the flow of one-carbon units. However, T8 and

MTHFR catalyze irreversible reactions and therefore, they are thought to compete

for methyleneTHF which would be critical in deciding the flux of one-carbon units

through their respective pathways. These experiments looked at the specificities

of each of these enzymes for the polyglutamate chain length of methyleneTHF in

an attempt to predict the flux through the competing pathways. 8HMT (Matthews

60 et al. 1982) and 0 (Ross et al. 1984) exhibit a much broader specificity for polyglutamate chain length than MTHFR and TS (Lu et al. 1984).

Cellular concentrations of methyleneTHF have also been proposed to regulate the flux between nucleotide synthesis and methionine regeneration.

Increased cellular levels of methyleneTHF lead to increased flux of this folate derivative into thymidylate and purine synthetic pathways (Green et al.

Biochemistry 1988). When there are limiting concentrations of methyleneTHF, its

polyglutamate chain length will become an important determinant of the flux.

Although mammalian cells usually contain long chain polyglutamates, folytriglutamates are sufficient to support glycine synthesis. MethyleneTHF with

longer chain lengths is preferentially reduced to methylTHF because MTHFR demonstrates the strongest preference for long chain polyglutamates as

compared to the other enzymes in folate metabolism (Matthews and Baugh 1980,

Lu et al. 1984, Lowe et al. 1993).

A recent study on MCF-7 cells (Fu et al. 2001) evaluated the roles of the

cytoplasmic and mitochondrial SHMT isozymes in determining the metabolic flux.

The advantage of using this human breast cancer cell line is that there is no

glycine cleavage system so carbon 2 of glycine does not act as a donor of one­

carbon units. This allows the focus to be on the contribution of carbon 3 of serine towards de novo purine and thymidylate synthesis.

This study consisted of NMR analysis to follow the incorporation of label from one-carbon units provided by either serine or formate, into purines and

thymidylate. These results showed that carbon 3 of serine contributes

61 approximately 95% of the one-carbon units in the total folate pool. Subsequently, there was an evaluation of the portion of carbon 3 of serine that is incorporated into purines and thymidylate as a result of either the cytoplasmic SHMT or formate production by the mitochondria. The approach taken was a competition experiment involving both one-carbon donors. It was seen that unlabeled formate heavily competed with the labeled serine but there was no reciprocal effect with the reverse situation. The availability of formate reduces the need to convert serine to formate in the mitochondria. The important conclusion is that virtually ail of the one-carbon units required for purine biosynthesis are derived from formate generated by the mitochondrial serine metabolism via the SHMT. The mitochondrial pathway also makes an important contribution of one-carbon units for thymidylate synthesis (Fu et al. 2001).

However, it is possible that these isotopie enrichment studies that were used to determine the incorporation of radiolabeled one-carbon donors may be overestimating the mitochondrial contribution to cellular function especially in light of two subsequent studies that also examined the direction of the flux of one­ carbon metabolites in mammalian cells.

These studies (Oppenheim et al. 2001, Herbig et al. 2002) looked specifically at the role of the cytoplasmic SHMT in regulating folate metabolism in

MCF-7 cells. Evidence from previous experiments has concluded that SAM synthesis has a higher metabolic priority than de novo thymidylate synthesis.

This study proposes that cSHMT is functionally poised to regulate the flux of one­ carbon units to either homocysteine remethylation, thymidylate or purine

62 biosynthesis by regulating 5,10-methyleneTHF pools. Their results indicate that an increase in cSHMT activity enhanced de novo thymidine biosynthesis, most likely by making 5,1 O-methyleneTHF available for thymidylate synthesis rather than methionine regeneration (Oppenheim et al. 2001, Herbig et al. 2002).

One of these studies (Herbig et al. 2002) also examined the effects of

cellular glycine on cSHMT activity. Glycine modulates the distribution of folate

derivatives, specifically by reducing SAM synthesis and limiting the availability of

methyl groups for the cellular methylation reactions. This indicates that glycine

increases the effectiveness of cSHMT to compete with MTHFR for 5,10-

methyleneTHF and promotes serine synthesis. In addition, cSHMT has a high

affinity for 5-methylTHF with resultant inhibition of the activity. Consequently, this

results in the depletion of cellular SAM levels and subsequent inhibition of

methionine regeneration in a glycine-independent manner.

It was also documented that a glycine-dependent increase in cSHMT

activity also stimulates labeled formate incorporation (via cytoplasmic

methyleneTHF) into methionine, which is most likely due to a decrease in the

demand for this pathway to provide one-carbon units for thymidylate synthesis.

Instead, TS appears to incorporate methyleneTHF provided by serine via the forward cSHMT reaction. Although cSHMT can supply one-carbon units to both

thymidylate synthesis and methionine regeneration, glycine enhances the

contribution of one-carbon units by cSHMT to thymidylate synthesis (Herbig et al.

2002).

It has been documented that iron deficiency can alter folate metabolism

63 and influence folate status, leading to hematological disorders that are usually caused by folate deficiency. However, the mechanism by which iron deficiency affects folate metabolism has not yet been determined. A recent study

(Oppenheim et al. 2001) demonstrated that iron sequestration may be involved in regulating folate metabolism based on the observation that the levels of cSHMT protein are elevated by an increase in HCF protein, a component of the iron storage protein, ferritin.

Therefore, it would appear that in mammalian cells, the flow of one-carbon units in the mitochondria proceeds from 5,1 O-methyleneTHF to 10-formyITHF, known as the forward direction whereas it proceeds from formate to 5,10- methyleneTHF, known as the reverse direction in the cytoplasm. This situation is similar to that in the yeast.

As has been discussed in this general introduction to my thesis, the importance of an adequate dietary folate status is evident in light of the various medical disorders that result as a consequence of folate deficiency or an impediment in the folate pathways. Some governments have recognized the benefits of folic acid and have made recommendations about daily doses of folic acid as weil as implementing the fortification of cereal-grain foods with folic acid.

However, the cellular and molecular mechanisms that exactly define the protective effects of folic acid have yet to be fully elucidated.

An essential first step that needs to be taken is to increase our comprehension of the folate-dependent pathways, which is the basis of my thesis.

Specifically, my research project has consisted of determining the flux of folate

64 metabolites in both the mitochondria and the cytoplasm of mammalian cells.

Furthermore, it has also been critical to establish the extent of the interdependence of the folate pathways between these compartments as has been demonstrated in the yeast model of folate metabolism. One key difference between these two eukaryotic folate systems is the mammalian model contains an NAD-dependent bifunctional OC protein with no attached synthetase activity in the mitochondria instead of an NADP-dependent trifunctional DCS protein as in yeast although the advantage of this evolvement remains to be determined.

Furthermore, unlike the yeast model of folate metabolism, it is not known whether the mammalian mitochondria are able to generate formate as there is no known mammalian mitochondrial synthetase protein. Although another group (Barlowe and Appling 1988) had proposed that a mitochondrial DCS protein existed in mammalian mitochondria, no such protein was ever isolated. However, this may have been solved with the recent cloning of a human cDNA that encodes a putative mitochondrial DCS protein (Prasannan et al. 2003, Sugiura et al. 2004).

These objectives have been accomplished by analyzing mammalian cell lines that are lacking activities of key enzymes in these folate pathways and investigating the consequences on folate metabolism. The two central enzymes in the mammalian folate system that are the focus of this thesis are the mitochondrial bifunctional NAD-dependent dehydrogenase-cyclohydrolase and the cytoplasmic trifunctional NADP-dependent dehydrogenase-cyclohydrolase­ synthetase.

65 CHAPTER TWO

MAMMALIAN MITOCHONDRIAL METHYLENETETRAHYDROFOLATE

DEHYDROGENASE-CYCLOHYDROLASE DERIVED FROM A

TRIFUNCTIONAL METHYLENETETRAHYDROFOLATE

DEHYDROGENASE-CYCLOHYDROLASE-SYNTH ET ASE

66 PREFACE

ln this chapter we have defined the evolutionary relationship between the mitochondrial NAD-dependent OC and the cytoplasmic NADP-dependent DCS by comparing their DNA sequences among different species.

The following chapter has been published under the title "Mammalian

Mitochondrial Methylenetetrahydrofolate Dehydrogenase-Cyclohydrolase Derived

From A Trifunctional Methylenetetrahydrofolate Dehydrogenase-Cyclohydrolase­

Synthetase" by Harshila Patel, Karen E. Christensen, Narciso Mejia and Robert E.

MacKenzie in Archives of Biochemistry and Biophysics (2002), vol. 403, pages

145-148.

Karen E. Christensen and the candidate contributed equally to this publication.

This publication does not include a section entitled "Materials and

Methods" and ail the methods are briefly described in the figure legends. In order to keep the same format as the following chapters and to avoid extensive description in the figure legends, 1 have included a section for "Materials and

Methods" which describe in more details the experimental procedures.

The references have been compiled in alphabetical order at the end of the thesis.

67 SUMMARY

We have isolated the cDNA and the gene encoding the murine cytoplasmic methylenetetrahydrofolate dehydrogenase-methenyltetrahydrofolate cyclohydro­ lase-formyltetrahydrofolate synthetase (DCS). Comparison of these sequences with the 3'-untranslated region of the mitochondrial NAD+-dependent methylenetetrahydrofolate dehydrogenase-cyclohydrolase (mt-OC) revealed areas of significant homology. Both exon and intron sequences of the synthetase domain of DCS are homologous to sequences in the untranslated region of mt­

OC. A similar comparison between the mt-OC and the DCS sequences of humans as weil as Drosophila supports the conclusion that in higher eukaryotes the bifunctional mt-OC replaced a trifunctional precursor through inactivation of the synthetase domain. The mt-OC should be considered in models of one­ carbon folate fluxes in mammals.

68 INTRODUCTION

Folate metabolism has been associated with several disease conditions, including neural tube defects in infants and cardiovascular disease in adults

(Lucock 2000). Tetrahydrofolates (THF) are used metabolically in two major roles, to support methylation reactions (methionine, thymidylate) and to provide formyl groups used in purine biosynthesis. The dehydrogenase-cyclohydrolase activities interconvert methyleneTHF and formylTHF used for these two purposes respectively. The role of mitochondria in generating one-carbon tetrahydrofolates is an area of active investigation. Barlowe and Appling (Barlowe and Appling

1988) suggested that mammalian mitochondria produce formate that is used by the cytosolic DCS as a significant source of one-carbon units. Genetic, biochemical and metabolic studies in yeast support such a model that incorporates the products of two nuclear genes: ADE3, encoding a cytosolic DCS, and MIS1, encoding a mitochondrial DCS. In this model, mitochondrial methyleneTHF is converted to formylTHF by the mitochondrial OC activities and the formyl group is released by the "reverse" mitochondrial synthetase activity producing THF, ATP and formate. The formate exits the mitochondria and in the cytosol, the three DCS activities are proposed to work in the opposite direction to convert formate to formylTHF and methyleneTHF. The model nicely explains the existence of DCS isozymes in both compartments and is supported by metabolic studies in yeast (Pasternack et al. 1994).

This model has been widely accepted and has also been applied to the

69 interpretation of metabolic studies on humans (Gregory et al. 2000). However, no trifunctional DCS protein has yet been purified from mammalian mitochondria.

Only a single gene (and a pseudogene on X) encoding a cytosolic DCS has been reported in mammals (Rozen et al. 1989). At the time of writing no candidate mitochondrial DCS was found by protein homology searches of the human genome database (National Center for Biotechnology Information). The results presented in this research report strengthen the hypothesis that the mammalian mitochondrial bifunctional NAD-dependent OC enzyme is the homolog of the yeast mitochondrial trifunctional MIS1 enzyme.

70 MATERIALS AND METHODS

Restriction enzymes and T 4 DNA ligase were purchased from New

England Biolabs. Ali the cDNA probes were labeled by the random primer

method using [a-32 P]dCTP as the labeled nucleotide and the Multiprime DNA

labeling System, which were ail purchased from Amersham Biosciences.

Isolation of the Murine Cytoplasmic DCS cDNA-A AZAPII library prepared from an embryonic mouse kidney tissue was provided by Dr. Michel Tremblay.

The titer of the library was determined to be 4.7 x 107 plaque forming units/ml. A

1000-fold diluted library was infected into XL-1 Blue cells and plated on LB plates.

The plaques were transferred to Hybond-C nitrocellulose membranes (Amersham

Biosciences). The membranes were treated with denaturing solution (1.5 M NaCI,

0.5 M NaOH) for 7 min, neutralizing solution (1.5 M NaCI, 0.5 M Tris, 1 mM

EDTA) for 3 min, washing solution (2x SSC) for 3 min and placed under UV light

for 5 min to cross link the DNA transferred from the plaques. The membranes

were prehybridized at 42 Oc for 4 h in a buffer containing 40% formamide, 4x

Denhardt solution, 5x SSPE, 0.1 % SOS, 10 IJg/ml denatured salmon sperm DNA

and 50 mM sodium phosphate (pH 6.8). Then, the membranes were hybridized in

the same buffer containing 2 x 105 cpm/ml 32P-labeled 1965-bp Nar1-Hindlll

fragment of the human cytoplasmic DCS cDNA for 16 h at 42 oC. The hybridized

membranes were washed three times for 30 min each in washing solution #1 (2x

SSC, 0.1% SOS) at 42 oC, once for 30 min in washing solution #2 (0.1x SSC,

0.1% SOS) at 55 oC and twice for 30 min in washing solution #3 (0.1x SSC) at

71 room temperature. The filters were exposed with autoradiography. The positive

plaques that hybridized to the probe were selected and subjected to second and

third rounds of screening until pure plaques were obtained.

The recombinant phagemids are contained within the pBluescript SK+

plasmid, which itself is encompassed within the two arms of the ÀlAPIi cloning

vector. Consequently, the positive recombinants were subjected to an in vivo

excision protocol whereby the ExAssist helper phage expels the pBluescript SK+

plasmid containing positive clones from the ÀlAPIi vector. The SK+ plasmid then

recircularizes and propagates in the SOLR cells. The plasmid DNA was isolated from the recombinant SOLR ce Ils by the Wizard Miniprep Kit.

A total of 14 positive clones were obtained and they were subjected to

EcoRI restriction digestion to determine which clones gave a similar restriction

pattern to the human cytoplasmic DeS cDNA. The longest clone, 14aEB, was

completely sequenced by automated sequencing performed by Sheldon

Biotechnologies. The clone was approximately 3 kb in length and it comprised

essentially the full-Iength mouse DCS cDNA but it lacked 24 nucleotides from the

3' end of the coding region. The missing portion of the mouse DCS cDNA was

determined upon isolation of the corresponding gene.

Isolation and Mapping of the Mouse DeS Gene-We sent a central 1617-

bp EcoRI fragment of the mouse DCS cDNA to Genome Systems Inc. to screen a

mouse 129/SvJ genomic library constructed in the pBeloBACl1 vector. Their

screening efforts resulted in the isolation of four genomic clones, which were

approximately 100 kb in length. Three of these Genome Systems (GS) clones,

72 GS21429, GS21430 and GS21431 were digested with BamHI, EcoRI, Hindlll

Pstl, Sacl and Xbal restriction enzymes, in triplicate and analyzed by Southern

analysis. The digested genomic DNA fragments from each set were separated by

aga rose gel electrophoresis and vacuum transferred onto three separate Hybond

N+ membranes (Amersham Biosciences). The membranes were prehybridized at

42 Oc for 4 h in a butter containing 40% formamide, 4x Denhardt solution, 5x

SSPE, 0.1 % SDS, 10 ~g/ml denatured salmon sperm DNA and 50 mM sodium

phosphate (pH 6.8). Then, each membrane was hybridized in the sa me butter

containing 2 x 105 cpm/ml of one of three 32P-labeled probes for 16 h at 42 oC.

The three probes consisted of three mouse DCS cDNA EcoRI fragments which

encompass the whole cDNA sequence: 5' end probe (371 bp), central probe

(1617 bp) and 3' end probe (847 bp). The hybridized membranes were washed

once in washing solution #1 (1x SSC, 0.1% SDS) as the temperature was allowed

to increase from 42 oC until 65 oC, once again in washing solution #1 for 30 min at

65 oC and once in washing solution #2 (0.1x SSC, 0.1 % SDS) for 30 min at 65 oC.

The filters were exposed with autoradiography. Approximately 120 positive

mouse DCS genomic clones were identified. The GS DCS clones were digested

again with the same six restriction enzymes and separated by gel electrophoresis.

The positive restricted fragments were purified from the agarose gel and shotgun

ligated into the pBluescript SK+ vector. Each ligation was transformed into DH5a

cells and plated on LB ampicillin plates. The resulting colonies were transferred

onto Hybond-C nitrocellulose membranes using the same protocol that was used

for the screening of the embryonic mouse kidney AZAPII library for the mouse

73 Des cDNA. The positive colonies were innoculated and the Qiagen Miniprep

Spin Kit was used to isolate the plasmid DNA from each bacterial culture. Each

DeS genomic subclone was ligated into the pBluescript SK+ vector, mapped by restriction enzymes and unique clones were sequenced by automated sequencing (Sheldon Biotechnology). The sequence of each DeS genomic subclone was compared with the DeS cDNA sequence in order to determine the boundaries of the intron-exon junctions and the restriction map of the genomic clones was used to assemble the sequence of the DeS gene.

74 RESULTS

We have isolated the cDNA for the murine cytoplasmic DCS. The murine

DCS encoded by the cDNA has 87% amino acid homology with the human DCS that this laboratory has previously expressed and characterized (Hum and

MacKenzie 1991). The murine DCS gene isolated using this cDNA is shown diagrammatically in Fig. 1. The intron-exon boundaries of the DCS gene are shown in Table 1. The DC domain is encoded by exons 1 through 10 and the synthetase domain is encoded by exons 10 through 27. The two domains are connected by a linker sequence (Hum and MacKenzie 1991) found in exon 10.

75 FIGURE 1. The gene for the murine cytoplasmic DeS.

The gene was isolated from GenomeSystems 129Sv/J genomic BAC clones by Southern analysis using the DCS cDNA, isolated from a ,A,ZAPII embryonic mou se kidney cDNA library (a generous gift from Dr. M. Tremblay), to identify gene fragments. Subclones were subjected to automated sequencing (Sheldon Biotechnology Center). The gene consists of 28 exons and spans -65 kb. The protein is divided into the OC and S domains by the linker region found in exon 10. The cDNA sequence is deposited under GenBank Accesion Number AF364580 and the gene sequences are deposited under GenBank Accession Numbers AF364581-AF364592.

76 I~

c co E o "'C Cf)

c CO E o "'C Ü --o TABLE 1

The intron-exon boundaries of the murine DeS gene. ("Crystal" shareware font: Jerry Fitzpatrick, www.SottwareRenovation.com)

Exon Exon Intron-Exon Boundaries Intron Number Size 5' 3' Size (b~} {b~) 1 100a ... TCTGCgtaagtacttgccagtgata 14580a 2 85 caccttttctttccctctagGCAAA ... TGCAGgtattggagcatcttattcc 9340a 3 59 gcatgtatttattattccagGTTGG ... AAGAGgtaagccaagacagggatgc 690 4 54 TcttgtgtttatttctgcagATTGG ... CCGAGgtgagcgtgtggatgcacat 169 5 137 aactgtggcactacttttagGTGTT ... GATGGgtgagtgctccccgttggct 112 6 101 tcctaatttgactgttttagGTTGA ... GGCAGgtaagagtgagggaaaccag 2160a 7 137 tgtgctggcttctcttccagGGGTG ... AGGAGgtaggcaatccagggtagcg 932 8 112 tagatccctcccctccttagGTAAA ... TCCAGgtgagtggtgccagacagag 4900a 9 128 ctctccctttcctgatccagATGAT ... TGCAGgtatggcgcagaaagttctg 421 10 98 gtccttctccctttgcccagAGCAC ... CCAAGgtaggcgtgaaatgttctcg 156 11 174 ccatccttgtgtttttaaagTGACA ... ACTGGgtatgtgtgttctgtggcca 1107 12 137 tgtctctctcttctttgaagAATTA ... AAAAGgtactgtgccagagcgcctg 2050a 13 47 gctgttgctctgtcccctagGTGGC ... AAGAGgtaaaaccgctcagggtttg 485a 14 108 tggcttccgctcttttgtagTTTAA ... ACAAGgtagaatgccgagggtgtgg 109 15 75 ttttcttctctttctttaagGCTCT ... TACGGgtaagcttttgcctactaga 3120a 16 103 ttaatttcctttcttttcagAGGCT ... GAGAGgtatgtacccagcccaggct 2786 17 77 gccctccctttctgccatagTGTTG ... GCACGgtaactgttcatctccagta 791 a 18 141 ctttatctctctgcccacagGCCCA ... ACCTGgtgagtgacaggcagggagg 1553 19 69 ttcgtgggcttgtcttgcagGGCGT ... TAGAGgtgagctggctgactcctgg 606 20 112 tgacgtcatggtctccacagGGCAC ... CGTAGgtgagttctcctgcatcctc 97 21 140 gttgctcctctcccttgcagTGACG ... CCACGgtgaggaccagggactggga 3560a 22 42 ctttctatcctatcttctagGTCAC ... AAGAGgtaactacaattatttccga 3470a 23 101 catttgctgttgtttgacagGACCT ... TTCAAgtaagtgtgaaatatagaga 796 24 178 cgtgaagccctcctgtgcagGACAG ... TCAAGgtaggtttgcttgttctgtg 1647 25 108 ccttctcttttgtacactagCTCTC ... AGCAGgtaggtgtttacacccgagc 439 26 153 attccctttttcccacccagGGCTT ... GAACGgtaagtgaaggctgtaagag 2493 27 94 aatgctgttttgcctcacagATGAG ... AGCAGgtgagctgctactgggcagg 2910a 28 215a cattctgcttttctttccagATCTT ... a Approximate value

77 Previous work from our laboratory showed that the murine mt-OC has a long 3'-untranslated region (UTR) (Belanger and MacKenzie 1989). We compared the synthetase region of the murine DCS cDNA and gene with the 3'­

UTR of the mt-OC cDNA. The results are shown schematically in Fig. 2 with the homologies summarized in Table 2. Fig. 3 shows the homologous regions found

in the mouse sequences. Several regions, including two sequences found in

intron 21 of the synthetase portion of the DCS gene, showed greater than 50%

identity with regions in the mt-OC cDNA. This indicates that there are residual synthetase sequences in the 3'-UTR of the mt-OC cDNA. Comparison of the human DCS gene and cDNA with the mt-OC cDNA showed similar homologies.

One interesting observation is that part of exon 10 and a contiguous portion of

intron 10 of the human DCS are found in the 3'-UTR of the corresponding mt-OC.

While the homologies between the Drosophila cDNAs are not as extensive, they

also show a similar relationship. These observations suggest that a trifunctional

DCS is the precursor of mt-OC.

78 FIGURE 2. Schematic diagram of the homologous regions.

The schematic diagram of the homologous regions of the NADP-dependent DCS cDNAs and genes and the 3' untranslated regions of the NAD-dependent mt-DC for mouse, human and Drosophila. Sequences for the mouse DCS are reported in this paper (Fig. 1). Other sequences were obtained trom GenBank: murine mt­ DC cDNA (J04627); human mt-DC cD NA (X16396); human DCS cDNA (J04031); human DeS gene (NT_025892.6: 1850220-18601848) Drosophila mt-DC cDNA (L07958) and Drosophila DCS (AF082097). Solid boxes are numbered and indicate exons. Homologous regions carry letter designations and are shown as shaded boxes.

79 DCS gene 20 21 D E 22

Mouse D/C s DCScDNA B A C

D/C DCcDNA A BC D E

DCS gene 9 lOB C 11

Human D/C s DCScDNA B DA E

D/C 11_1 fif~4 DC cDNA ABC D E

D/C U s 1 i DCS cDNA Drosophila A B

1 D/C ~ DCcDNA AB - lOObp TABLE 1\

Details of the homologies found between the 3'-untranslated region of the mt-OC cDNA and the synthetase region of the NA DP-dependent DCS. Letter designations refer to the corresponding regions from Fig. 2 for each species. Sequences were compared using LALIGN (http://www.ch.embnet.org/software/LALlGNtorm.html).

Region Sequences tram Length Homology Exon(s) Intran Base pairs % Mouse 21,22, A 131 58.0 23 B 10 80 57.5 C 24,25 139 57.6 0 21 159 57.9 E 21 105 59.0 Human A 24 72 59.7 B 10 53 62.3 C 10 95 53.7 0 23 50 66.0 E 27 227 50.9 Drosophila A 4 42 61.9 B 5 78 60.0

80 FIGURE 3. Details of the homologies.

The details of the homologies found between the synthetase region of the NADP­ dependent DCS (top sequence) and the 3' untranslated region of the mt-OC (bottom sequence). Letter designations refer to corresponding regions for the mouse from Fig. 2. The cDNA sequences are numbered fram the start codons.

81 Region A 2103 cagggcccttaagatgc--acgggggtggccccacggtcaccgctggactgcctcttcccaaggcttacacagaagaggacctgg-acctggttgaaaagggcttca catggctctacagaacctcaccaggatg--ccca---tca--gttgtcctgcctttgacctggtgatgcacaggagacgtcctcacacc-----gagagtggatgct 1180 2230 gtaacttgaggaaacaaatcgaaa g-aacttcatcaaaaaaaaaaaaa 1297

Region B 879 955 cttcctacagaaatttaagccag-ggaagtgga--caattcagtataacaagctgaacctcaagactcctgtcccaagtg...... cttagtagaggaaatgaagggaaaggaagagaaagcagcacagggtttcaa-cggcactcccttactc-tctcgggagtg...... 1402 1480

Region C 2343 tgctgtca--agtgcacccactgggcagaagggggc-cagg-----gggccttgg-----ccctggctcaggctgtccagagagcgtcacaggc-ccccagcagctt tgcagccaggagtgctccgattgccctgaggagagctcagctggctgggcctagctgtccccctgtggcgggtcgtcagggaagggttagagtctccccacttgctg 1479 2466 ccagctcctc-tatgacctcaagctctcaatt tta--tcctcctgtgagcttgtcctatctatt 1615 Region D gttaaatttctgttccttaggcatcggtattgtaacttag-agaac---aagacctcccttgtctacagaatagaatctatt-tcttggt--gcagtggcat-acat gttatcctcctgtgagcttgtcctatctattttaattttctagaatcgaaaaagctca-ttttataaatgat-gtgtgtattatcacgtttggcttttgcagcacat 1585 ttgttcttata--tagca-attctcaagtgttttggattttt-ttttctttg ttaaacttttaagtttcagactgttaagagagttgtgtgtctgttttctttg 1741

Region E ttggatttataaataactggttgacactgtttttttttttttgttggttttgttgttgttg---ttgttgtttaatatcaattagggccaagcacaagcctgtgg...... ttgggttaatataagact--tagtc---ggattttcttattaaagggttttcttgcagttgggttttttgtttgtttttaaataaagctcatctgtcgggtgtgg...... 1895 1994 DISCUSSION

MethyleneTHF dehydrogenases are found in nature as monofunctional (0), bifunctional (OC), and in eukaryotes, trifunctional (DCS) enzymes. Synthetases are usually monofunctional except in eukaryotes (MacKenzie 1984). It had been shown that mt-OC is a 34-kDa protein lacking synthetase activity (Mejia et al.

2 1986) that retains Mg + -dependent, low residual dehydrogenase activity with

NADP (Yang and MacKenzie 1993). It has been suggested that the protein changed from NADP to NAD specificity by replacing the 2'-P of NADP with

2 inorganic phosphate using Mg + (Yang and MacKenzie 1993). The shift from

NADP to NAD specificity for the methyleneTHF dehydrogenase in mitochondria dramatically shifts the equilibrium between methyleneTHF and formylTHF toward the latter (Pelletier and MacKenzie 1995). It seems clear from our results that the mammalian mt-OC arose not from a bifunctional OC precursor, but through the loss of the synthetase activity of a trifunctional NADP-dependent DCS.

The role of mt-OC is still not entirely c1ear. Its expression during embryonic development and in transformed cells, but not in adult differentiated tissues (Mejia and MacKenzie 1985), must be explained. While we originally proposed a role in producing formylTHF for formylation of met-tRNA it would probably not require a shift in nucleotide specificity to support this quantitatively minor role. Significantly, recent results from studies in yeast showed that formylation is not required for mitochondrial protein synthesis (Li et al. 2000), suggesting that it might not be required in mammalian mitochondria either. However, because of the cofactor

82 specificity change to NAD, mt-OC is actually weil designed to participate in formate generation by mitochondria. The lack of a synthetase domain associated with the mt-OC rules out this activity as part of the process but it is important to note that the lack of mitochondrial trifunctional DCS does not rule out the

possibility of the production of formate by an entirely different protein in the

mitochondria. It is possible that a more efficient process for releasing formate from formylTHF has developed in mammalian mitochondria. If mt-OC is indeed

the only methyleneTHF dehydrogenase activity in mitochondria, then the "formate

cycle" that produces one-carbon units for the cytosol from mitochondrial

methyleneTHF will operate predominantly under conditions or in tissues where it

is expressed at significant levels and would be particularly important during

em bryogenesis.

83 CHAPTER THREE

MAMMALIAN FIBROBLASTS LACKING MITOCHONDRIAL

NAD+ -DEPENDENT METHYLENETETRAHYDROFOLATE

DEHYDROGENASE-CYCLOHYDROLASE ARE

GLYCINE AUXOTROPHS

84 PREFACE

ln this chapter, 1 have characterized NMDMC null mutant fibroblasts in order to determine the contribution of the mitochondrial folate pathway to cellular function in mammalian cells.

The following chapter has been published under the title "Mammalian

Fibroblasts Lacking Mitochondrial NAD+ -Dependent Methylenetetrahydrofolate

Dehydrogenase-Cyclohydrolase Are Glycine Auxotrophs" by Harshila Patel,

Erminia Di Pietro and Robert MacKenzie in the Journal of Biological Chemistry

(2003), vol. 278, pages 19436-19441.

Erminia Di Pietro and the candidate contributed equally to this publication.

The work presented was conducted by the candidate except for Figures 1, 2, 5 and Table III, which were provided by Erminia Di Pietro.

The references have been compiled in alphabetical order at the end of the thesis.

85 SUMMARY

Primary fibroblasts established from embryos of NAD-dependent mitochondrial methylenetetrahydrofolate dehydrogenase-cyclohydrolase

(NMDMC) knockout mice were spontaneously immortalized or transformed with

SV40 Large T antigen. Mitotracker Red CMXRos staining of the cells indicates the presence of intact mitochondria with a membrane potential. The nmdmc(-I-) cells are auxotrophic for glycine demonstrating that NMDMC is the only methylenetetrahydrofolate dehydrogenase normally expressed in the mitochondria of these cell lines. Growth (increase in cell number) of null mutant but not wild type cells on complete medium with dialyzed serum is stimulated about 2-fold by added formate or hypoxanthine. Radiolabeling experiments demonstrated a 3-10x enhanced incorporation of radioactivity into DNA from formate relative to serine by nmdmc(-I-) cells. The generation of one-carbon units by mitochondria in nmdmc(-I-) cells is completely blocked, and the cytoplasmic folate pathways alone are insufficient for optimal purine synthesis. The results demonstrate a metabolic role for NMDMC in supporting purine biosynthesis.

Despite the recognition of these metabolic defects in the mutant cell lines, the phenotype of nmdmc(-I-) embryos that begin to die at E 13.5 is not improved when pregnant dams are given a glycine-rich diet or daily injections of sodium formate.

86 INTRODUCTION

Folate-dependent enzymes are found in the mitochondria as weil as the

cytoplasm of eukaryotic cells. Isozymes of certain folate-dependent enzymes are

present in both compartments, and a number of observations demonstrated that the folate-dependent pathways in mitochondria contribute to total cellular folate

metabolism. For example, serine hydroxymethyltransferase, encoded by two

different nuclear genes, is located in each compartment where it catalyzes the

interconversion of serine and tetrahydrofolate (THF) with glycine and

methylenetetrahydrofolate. A Chinese hamster ovary cell line that is missing

mitochondrial serine hydroxymethyltransferase (glyA) was shown to be a glycine

auxotroph despite the fact that it retains the cytoplasmic isoform of the enzyme

(Chasin et al. 1974). A similar phenotype was seen in the auxB1 mutant ceilline

that lacks the ability to make folate polyglutamates. Replacing the missing folylpoly-y-glutamate synthetase in the cytoplasm of this cell line reverses the

requirement for thymidine and purines, but the enzyme must be targeted to

mitochondria to overcome its requirement for glycine (Garrow et al. 1992, Lin et

al. 1993, Lin and Shane 1994, Chen et al. 1996).

Methylenetetrahydrofolate dehydrogenase-cyclohydrolase activities are

located in both cellular compartments. In yeast, both cytoplasmic and

mitochondrial isoenzymes occur as trifunctional NADP-dependent

dehydrogenase-cyclohydrolase-synthetase (DCS) proteins (Shannon and

Rabinowitz 1986). The D and C activities interconvert methyleneTHF and 10-

87 formylTHF and the S converts formate to formylTHF in an ATP-dependent

reaction. Dean Appling's group (Pasternack et al. 1994, West et al. 1996,

Kastanos et al. 1997) has proposed a rational model wherein mitochondria use

the synthetase activity "in reverse" to produce formate, ATP and THF from

formylTHF. The formate is then recaptured as an important source of active one­

carbon units by the activities of the cytoplasmic DCS (Pasternack et al. 1994,

West et al. 1996, Kastanos et al. 1997, Piper et al. 2000). In mammals, the

cytoplasmic DCS is ubiquitously expressed in ail cells and tissues (Mejia and

MacKenzie 1985, Smith et al. 1990, Peri and MacKenzie 1991, Thigpen et al.

1990), whereas a mitochondrial version (Mejia and MacKenzie 1988), NAO+­

dependent OC (NMDMC), is expressed during embryogenesis and is detectable

in ail immortalized cells (Mejia and MacKenzie 1985, Peri and MacKenzie 1991,

Mejia and MacKenzie 1988). This bifunctional enzyme was shown to have Mg2+

and inorganic phosphate-dependent methyleneTHF dehydrogenase activity

(Mejia et al. 1985, Mejia et al. 1986). 8ased on its kinetic properties, NMOMC

was proposed to have derived from an NADP+-dependent precursor (Yang and

MacKenzie 1993, Pawelek and MacKenzie 1996). More recent evidence

comparing the nucleotide sequence of the synthetase domain of the cytoplasmic

DCS with the sequence of the 3'-untranslated region of the NMOMC cDNA

supported the conclusion that its putative precursor was in fact an NADP+­

dependent DCS (Patel et al. 2002). The metabolic advantage for the

mitochondrial NMOMC enzyme to lose its synthetase activity is not obvious. This

feature raises the question as to whether mammalian cells expressing the

88 NMDMC enzyme export formate from their mitochondria to the cytoplasm and if so, what is the mechanism for the generation of formate?

Although the metabolic role of the NMDMC is not weil understood, deletion of the gene in mice has been shown to be embryonic lethal beyond E13.5 (Di

Pietro et al. 2002). At E12.5, the embryos appear smaller and pale when

compared with their wild type littermates. They also demonstrate a failure of the

liver to develop and to take over hematopoiesis from the yolk sac, despite the

presence of hematopoietic precursor cells (Di Pietro et al. 2002). It was suggested earlier that the role of the enzyme was to provide formylTHF for the

production of formylmethionyl-tRNA used in mitochondrial protein synthesis or to

produce one-carbon units for purine synthesis (Mejia and MacKenzie 1988).

Gene deletion studies indicated that the role of formylTHF in protein synthesis is

not essential in yeast (Li et al. 2000), and examination of an NMDMC null mouse

cell line indicated that mitochondrial protein synthesis is also not impaired when

compared with a wild type control (Di Pietro et al. 2002). In this report, we

establish and characterize immortalized cell lines and examine their nutritional

requirements to further characterize the metabolic consequences caused by

inactivation of the nmdmc gene.

89 MATERIALS AND METHODS

Embryonic Fibroblast Ce" Unes-The targeted inactivation of NMDMC in

ES cells and the generation of NMDMC knockout mice were described previously

(Di Pietro et al. 2002). Heterozygous mice were generated from two

independently targeted ES cell lines. To obtain primary embryonic fibroblast cell

lines, E9.5 and E11.5 embryos were isolated from matings of heterozygous mice.

Embryos were minced and trypsinized for 10-30 min. at 37 oC, and single cell

suspensions were resuspended in Dulbecco's modified Eagle's medium

containing 10-15% fetal bovine serum, supplemented with 1x non-essential amino

acids, 1x , 1x penicillin/streptomycin (ail from Invitrogen), as weil as 50

jJg/ml uridine (Sigma). This complete medium was supplemented with 100 jJM

sodium formate obtained from Sigma in ail studies where formate was not a variable. Spontaneously immortalized cells (SF) were obtained by continuous

passage in culture. Transformed cells (IF) were derived by infection of primary

cultures with a recombinant retrovirus expressing the SV40 large T antigen for 2 h

at 37 Oc in serum-free medium (Di Pietro et al. 2002). Following infection,

medium containing serum was added to the cells to obtain a final concentration of

15% fetal bovine serum. DNA was isolated from embryonic fibroblast cell lines

using the Qiagen genomic DNA purification kit. Genotypes of embryos and cell

lines were determined by Southern blot analysis as described previously (Di

Pietro et al. 2002). The presence of the sequence encoding SV40 large T antigen

in transformed cells was confirmed by PCR. Genomic DNA from fibroblasts (100

90 ng) was used with PCR primers TAgL, 5'-AAGTTCAGCCTGTCCAAG-3' and

TAgR, 5'-GTTTGCCACCTGGGTTAAG-3' to amplify a 638-bp fragment of the

SV40 large T antigen coding region. PCR conditions were as follows: 200 nM each of 5'- and 3'-primers, 250 !JM dNTPS, 1x Qiagen Taq butter, 2 mM MgCb, and 2.5 units Qiagen HotStar Taq polymerase in a final volume of 50 !JI with 1 cycle at 95 oC for 15 min; 30 cycles at 94 Oc for 1 min, 55 oC for 30 s, and 72 oC for 1 min; and 1 cycle at 72 oC for 10 min. PCR amplification products were analyzed by aga rose gel electrophoresis.

Rho 0 Cell Une-The pO cell and the 1438 parent cell line from which it was derived were generous gifts from Dr. Eric Shoubridge. The cells were maintained in Dulbecco's modified Eagle's medium supplemented with 10% fetal

bovine serum, 1x penicillin/streptomycin, 1x glutamine, and 50 ug/ml uridine.

Mitochondrial Staining with Mitotracker -selective Probes­

Medium containing Mitotracker Red CMXRos (Molecular Probes) in a range of 10-

500 nM was added to cells grown on tissue culture cham ber slides (Nunc Lab­

Tek) and incubated at 37 oC for 30-45 min. After incubation, the medium

containing the Mitotracker dye was replaced with fresh medium, and the stained

cells were observed by fluorescence microscopy.

Cell Growth Studies-The various cell lines used in these experiments were grown in defined medium that resembled Dulbecco's modified Eagle's

medium except that glycine, serine and methionine were added separately so that

one or more of these amine acids could be omiUed in order to determine the

nutritional requirements of cell lines. Our standard medium also contained 1x

91 glutamine, 1x penicillin/streptomycin, 50 !Jg/ml uridine, 100 !JM sodium formate and 1x ITS-S (insu lin, transferrin, sodium selenite and ethanolamine) from

Invitrogen. Formate was omitted when required. Another component of the medium, dialyzed fetal bovine serum (Invitrogen), was further dialyzed (50 ml against 2 liters of phosphate-buffered saline, pH 7.2) for 16 h with a change of buffer halfway during the dialysis. The cells were adapted to the redialyzed fetal bovine serum in a four-step process. Cells were plated at a density of 2.5 x 104 cells/well in 6-well plates and 24 h later, the medium was replaced with medium lacking glycine, serine, methionine, or formate. The cells were counted by the trypan blue exclusion method at 24-h intervals.

Precursor Incorporation into DNA-Cells were plated at a density of 5 x 104 cells/well in 24-well plates and grown for 24 h in defined medium but with glycine, serine or formate at 10% of the normal concentration. The growth medium was changed with the addition of 300 !JI of medium with 2 x 106 cpm of the respective

[1- 14C]glycine (97 !JM, 32 mCi/mmol); [2- 14C]glycine (90 !JM, 31 mCi/mmol); [3-

14C]serine (100 !JM, 32 mCi/mmol) or C4C]formate (67 !JM, 47.5 mCi/mmol) per weil. As a control for DNA synthesis, the cells were incubated with [3H]thymidine

(40 !JM, 25 Ci/mmol). In ail cases, experiments are compared where the difference in uptake of thymidine was ::;1 0% to ensure equal rates of cell growth.

The cells were allowed to grow for 24 h in the presence of each radiolabeled

precursor (Amersham Biosciences and Moravek Biochemicals) and were washed twice with 500 !JI of cold phosphate-buffered saline, pH 7.2. They were Iysed directly in the 24-well plates upon addition of a proteinase K solution (100 mM

92 Tris-HCI, pH 7.5, 100 mM NaCI, 50 mM EDTA, pH 8.0, 1% SDS, 1% 2- mercaptoethanol and 200 !Jg/ml proteinase K) for 1 h with moderate shaking at 37 oC. The Iysates were transferred to 1.5-ml Eppendorf tubes and extracted twice with one volume of phenol:chloroform:isoamyl alcohol (25:24: 1 ). The supernatants were precipitated with one volume of isopropanol followed by centrifugation at 4 oC for 15 min. The DNA pellets were resuspended in 100 !JI of

Tris-EDTA, pH 8.0. The DNA was transferred to a nylon membrane using a Bio­

Rad slot blot apparatus and washed three times with 200 !JI of Tris-EDTA, pH 8.0.

Each individual sample was cut out from the membrane with a sterile blade and placed into a scintillation vial containing 10 ml of Scintisafe (Fisher). This experiment was done in triplicate per cell line per condition, and the values for the amount of radioactivity incorporated into DNA, determined by liquid scintillation counting, were averaged.

Precursor Incorporation into Bases-Cells were plated at a density of 1 x

105 cells/well in 6-well plates, grown, and DNA isolated as described for the precursor incorporation into DNA. The cells were labeled in the presence of 2 ml of medium containing 4 x 106 cpm of the respective [1_14C]glycine (58 !JM, 17 mCi/mmol); [2_ 14C]glycine (58 !JM, 17 mCi/mmol); [3- 14C]serine 58.5 !JM, 17 mCi/mmol) or C4C]formate (27.5 !JM, 36 mCi/mmol). The DNA pellets were resuspended in 8 !JI of Tris-EDTA, pH 8.0 and transferred to 200-!J1 PCR tubes containing 30 !Jg of single-stranded salmon sperm DNA. Perchloric acid (2 !JI of

7.5 N) was added to each sample, and the DNA was subjected to hydrolysis at

105 Oc for 40 min. The hydrolyzed DNA samples were separated into their

93 individual bases by ascending paper chromatography, using a solvent consisting of isopropanol:HCI:H20 (130:33:37). An equimolar mixture of the four bases of

DNA was included in each chromatogram to serve as markers. The chromatograms were dried in a fume hood and examined under short wave UV

Iight, and each spot (representing the individual bases) was marked with a pencil.

The spots were carefully cut out and placed in a scintillation vial containing 10 ml of Scintisafe. The amount of radioactivity incorporated into each base was determined by liquid scintillation counting. This experiment was do ne in duplicate per cellline per condition, and the values were averaged.

Supplementation of Mouse Diet with Glycine-A custom research diet TD

02123 (Harlan Teklad) was formulated that added 6% glycine (60 g/kg) to the standard rodent diet 2018 that contains 0.8% glycine (Harlan Teklad). Both female and male heterozygous mice were started on the custom research diet 1 week before the first attempt at mating. Supplementation was continued until pregnant females were sacrificed. Embryos were isolated from females, phenotypes were recorded, and yolk sacs were collected for genotypic analysis.

As a control, embryos from matings of heterozygous mice fed on the standard rodent diet (2018) were also analyzed. Genotypes of embryos were determined as described previously (Di Pietro et al. 2002).

Injection of Pregnant Females with Sodium Formate-Pregnant dams on the standard mou se diet were injected daily with 100 !Jg/g of body weight of sodium formate in 100 !JI of phosphate-buffered saline, pH 7.2 from day E4.5 through day E6.5 with the dose increased ta 200 !Jg/g body weight from E7.5

94 through E13.5. Mice were euthanized, embryos were isolated, phenotypes were recorded, and yolk sacs were collected for genotypic analysis as described previously (Di Pietro et al. 2002).

95 RESULTS

Derivation of NMDMC Null Mutant Cell Unes-The details of the generation of NMDMC knockout mice have been described previously (Di Pietro et al. 2002). Primary embryonic fibroblasts derived from E9.5 and E11.5 embryos isolated from the matings of nmdmc(+I-) mice were spontaneously immortalized

(SF cells) or transformed with the SV40 Large T Antigen (IF cells), and genotypes were determined by Southern blot analysis using the E2/161 probe as described previously (Di Pietro et al. 2002). As shown in Fig. 1A, the probe hybridizes to a single 5-kb DNA fragment in BamHI digests of DNA from wild type embryonic fibroblasts and to an 11-kb fragment in digests from null mutant embryonic fibroblasts. Northern blot analysis (Fig. 1B) confirmed the absence of NMDMC mRNA in null mutant cells. Null mutant fibroblasts characterized in earlier experiments were found to contain no detectable NMDMC enzyme activity (Di

Pietro et al. 2002).

96 FIGURE 1. Genotypic analysis.

A, Southern blot analysis of DNA isolated from established mouse embryonic fibroblasts. BamHI-digested DNA (10 J..Ig) was hybridized with probe E2/161. Fragment sizes are indicated in kilobases. KO, knockout; wt, wild type. B, Northern blot analysis of total RNA isolated from (+/+) and (-/-) established mou se embryonic fibroblasts. Total RNA (50 J..Ig) was hybridized with a probe produced from the fuillength NMDMC cDNA.

97 ..... 0'1 1 1 ,.;;"' ';If(' 0" 0" ...... j tD ...... 6 ...... y y ...... » SF22(+/+) SF74(-/-) IF121 (-/-) IF130(-/-) SF132(-/-) IF132(-/-) SF134(+/+) IF139(+/+) Analysis of Mitochondria-We previously showed that mitochondria of nmdmc(-I-) fibroblasts are structurally intact by staining with the mitochondrion­ selective dye, Mitotracker Green FM, which is taken up by mitochondria independent of a membrane potential. As a further measure of mitochondrial function, we stained fibroblasts with Mitotracker Red CMXRos, which requires a mitochondrial membrane potential for uptake. There are no differences in the staining between wild type and null mutant fibroblasts even at the very low (10 nM) concentrations of dye used (Fig. 2). Secause of the location of the NMDMC enzyme in mitochondria, and the possibility that the nmdmc(-I-) cells would show a mitochondrial defect, we wanted to have a mitochondrially impaired cell line as a comparison for nutritional studies involving serine and glycine. We used pO cells that lack mitochondrial DNA (King and Attardi 1989) and thus cannot produce the proteins of the respiratory chain but retain a small mitochondrial membrane potential (Suchet and Godinot 1998). The mitochondria of these cells stain with

Mitotracker Red CMXRos (Minamikawa et al. 1999), which we also observed in this study (not shown).

98 FIGURE 2. Embryonic fibroblasts stained with Mitotracker Red CMXRos to detect mitochondria with an intact membrane potential.

99

Nutritianal Requirements far NMOMC Nu" Mutant Fibroblasts-A series af growth experiments were perfarmed on the NMDMC null mutant and wild type immortalized fibroblasts to evaluate their requirement for one-carbon precursors and products of folate-dependent metabolism. These fibroblasts, adapted for growth in dialyzed serum and defined medium, showed methionine dependence in that cell growth was much reduced by the replacement of methionine by homocysteine (not shown). This is not unexpected since many tumour and immortalized mammalian cells show a reduced or absent ability to replace methionine with homocysteine in the growth medium (Hoffman 1982, Tang et al.

2000). The nmdmc(-I-) cells are able to grow weil in the absence of serine (not shown) but are absolute glycine auxotrophs (Fig. 3, A and 8). Growth in glycine­ free medium was not stimulated by the addition of o-aminolevulinic acid, which might partially spare the requirement for glycine used in heme synthesis (Fig. 38).

Ali nmdmc(-I-) cell lines obtained from mice derived from two independently targeted ES cells required glycine for growth (Table 1), which could not be reversed with added hypoxanthine or formate (Fig. 38). However, in complete medium containing redialyzed serum, formate and hypoxanthine stimulate the growth of nmdmc(-I-) but not wild type cells, although neither is absolutely essential

(Fig. 3, C and 0). Addition of thymidine had no effect on cell growth. Control experiments that were not shown demonstrate that these cells readily incorporate radiolabeled thymidine into DNA, presumably by the thymidine kinase pathway.

100 FIGURE 3. Growth of fibroblasts in defined medium containing 10 % re­ dialyzed fetal bovine serum.

Effect of glycine on the growth of wild type (A) and null mutant cells (8). Symbols indicate: 0, complete medium; D, minus glycine; Il, minus glycine plus 30 !-lM hypoxanthine; 0, minus glycine plus 100 !-lM o-aminolevulinic acid. Effect of formate on cell growth of wild type (C) and null mutant cells (D). Symbols indicate: _, complete medium; complete medium containing: ., 100 !-lM sodium formate; Â, 30 !-lM hypoxanthine; or +, 30 !-lM thymidine.

101 50, 160 A lB 140 ~ 40 ~ z z 120 ~ ~ o 100 8 30 ü --l --l --l --l 80 W W ü 20 ü 60 >< >< 1" 1" 0 40 ~ 10 ...... 20 0 L....I 0 24 48 72 96 120 144 24 48 72 96 120 144 HOURS HOURS

80 C 300 1 D 70 ~ 250 ~ z z60~ ~ 0200 850 ü --l --l --l40w uj150 ü ü >< 30 1" '<1"><100 ~20 b 10 50 0 - 0 24 48 72 96 120 24 48 72 96 120 HOURS HOURS TABLE 1

Summary of the properties of established fibrobfast cel/fines

ES cell Primary Spontaneously Glycine Formatea/ line fibroblast Genotype.Immo rt a rIze d Transformed auxo t rop h Serine

05 PF22 +/+ SF22 no 0.3 PF74 -/- SF74 yes 2.5 PF134 +/+ SF134 no 0.4 PF139 +/+ IF139 no 0.3 PF121 -/- IF121 yes 1.1 1-B7 PF130 -/- IF130 yes 4.6 PF132 -/- SF132 yes 3.2 PF132 -/- IF132 yes 2.6 aRepresents the incorporation of radiolabel from formate relative to serine into total ONA of cell lines.

102 We asked whether cells that have another mitochondrial defect are also glycine auxotraphs and examined the pO cell line. This cell line, as weil as its parent cell line, are methionine-dependent as expected, and the replacement of methionine with homocysteine did not support the grawth of either. However, despite the fact that pO cells lack mitochondrial DNA and thus contain non­ respiring mitochondria, they graw equally weil on complete and glycine-minus media (not shown).

One-carbon Donor Incorporation into the DNA of Fibroblasts-As a preliminary approach to understanding the folate metabolism of these cells, we performed a series of radiolabeling experiments that measured the incorporation of C4C]-labeled glycine, serine and formate into the total DNA of exponentially growing nmdmc(-I-) cells, using wild type cells under identical conditions for comparison. Thymidine incorporation was measured in parallel and was equivalent in null and wild type cells. The incorporation of [1_ 14C]glycine and [2-

14C]glycine into DNA was identical for both wild type and null mutant cells (Fig. 4).

A major difference is seen in the incorporation of C4C]formate when compared with [3- 14C]serine into total DNA of wild type and mutant cell lines. Under the

conditions used in these experiments, more radiolabel is incorporated fram serine than fram formate in the wild type cells, and the mutant cell line incorporates

much higher amounts of label from formate than fram serine (Fig. 4). This

difference, shown as a ratio of radioactivity (formate/serine) in Table l, is

consistent with ail the wild type and mutant celllines examined.

103 FIGURE 4. Incorporation of radiolabeled one-carbon donors into total DNA of wild type and mutant fibroblasts.

Vertical bars, from the le ft, represent cpm .±. 8.0. incorporated into ONA of wild type and nul! mutant fibrablasts fram [1_14C]glycine, [2_ 14C]glycine, [3-14C]serine and C4C]formate.

104 ...... c ...... CCl :::J E As shown in Table Il, the incorporation of label from serine into adenine

and guanine bases in a similar experiment is much higher for the wild type cell than for the mutant, whereas the incorporation of formate is significantly less in

the wild type than in the mutant cells (Table Il). The radiolabeling of pO cells

under the same conditions does not show this difference, suggesting that its folate-dependent pathways are not affected.

105 TABLE Il

Incorporation of radiolabeled precursors into the bases of DNA

Following a 24-h incubation of cells with the one-carbon donors [3_14C] serine and [14C] formate, radioactivity was measured in thymidine, adenine and guanine of isolated DNA. The values are expressed in counts per minute (cpm) and represent an average of duplicate samples.

[3- ~4C] serine [~4C] formate Cell type and base Wild type Mutant Wild type Mutant Fibroblasts T 11550 6112 2834 829 A 70691 26759 8387 23821 G 86587 29142 9754 26036 Wild type pU Wild type pU 143B T 8816 4157 1794 1172 A 76512 17999 9457 5469 G 99133 25840 12638 8294

106 Supplementation of Pregnant Dams with Glycine and Formate-Since nmdmc(-I-) fibroblasts are glycine auxotrophs, we attempted to rescue the lethality of null mutant embryos by supplementing the standard mouse diet of pregnant dams with 6% glycine. Table III shows that out of a total of 73 embryos, dietary glycine supplementation does not reduce the numbers of resorbed fetuses or increase the viability of (-/-) embryos at E16.5 or even at E14.5. As control, we analyzed 74 embryos from pregnant females that received the standard rodent diet (Table III). Fig. 5 shows that supplementation with glycine does not improve the phenotype of E14.5 null embryos. Injection of the pregnant dams with sodium formate in the embryonic period beginning prior to liver development also did not enhance the phenotype of null mutants at E13.5 (Table III and Fig. 5).

107 TABLE III

Effects of supplements on embryonic phenotype

Embryos were isolated from dams receiving a standard diet, a glycine enriched diet, or a standard diet with daily injections of sodium formate.

No. of No. of No. Embryonic genotype Embryonic Age litters embryos resorbed +/+ +/- -/- No supplement E14.5 5 44 7 10 28 6 (5)a E16.5 4 30 5 10 15 5 (4) Total 9 74 12 20 43 11 (9) Plus glycine E14.5 5 35 15 9 23 3 (1) E16.5 5 38 9 11 27 0 Total 10 73 24 20 50 3 (1) No supplement E13.5 4 30 3 8 15 7 (1) Plus formate E13.5 3 31 3 5 18 8 (4) a Numbers in parentheses represent the number of dead (-/-) embryos.

108 FIGURE 5. Phenotypic comparison of embryos from pregnant dams supplemented with glycine or formate.

A, wild type embryo E14.5. B, nul! embryo. C, nul! embryo from dam supplemented with glycine. D, wild type embryo E13.5. E, nul! embryo. F, nul! embryo from dam supplemented with formate.

109 u..

w

o DISCUSSION

The expression pattern of NMDMC suggested a role in embryogenesis

(Mejia and MacKenzie 1985, Smith et al. 1990) that was confirmed by our recent knockout of the gene demonstrating that death begins at E 13.5 due to an inability of the liver to develop and take over hematopoiesis from the yolk sac (Di Pietro et al. 2002). We were unable to demonstrate obvious differences in the mitochondrial network of nmdmc(-I-) cells and in mitochondrial protein synthesis (Di

Pietro et al. 2002), and this is reinforced by the observation in this study that the mitochondria retain a membrane potential. The in vivo lethality seems therefore not to be due to a loss of mitochondrial integrity per se but due to a metabolic role that mitochondria perform for cellular metabolism that becomes significant around

E12 of embryonic development. Despite their lack of mitochondrial respiratory capability, in parallel studies, the pO cells exhibited properties very much like the wild type cell lines. These cells evidently retain a small mitochondrial membrane potential (Suchet and Godinot 1998) and import at least some of the enzymes normally located in the matrix (Herzberg et al. 1993), and in this study, we show that they are able to fulfill normal mitochondrial folate-mediated metabolism.

Metabolic studies with deuterated substrates indicate that much of the serine that is used as a one-carbon donor passes through the mitochondria and is released to the cytosol as formate (Sarlowe and Appling 1988, Fu et al. 2001,

Herbig et al. 2002). Fig. 6 illustrates that the NMDMC null cells can use mitochondrial serine hydroxymethyltransferase to convert the available

110 mitochondrial THF to methyleneTHF while producing glycine from serine.

However, the inability to oxidize the methyleneTHF to formylTHF due to lack of the 0* and C* activities prevents regeneration of the THF required to maintain the

production of glycine from serine and explains the glycine auxotrophy of these

cells. Both methyleneTHF from the conversion of serine to glycine, and that

produced from the glycine cleavage pathway would act as an intramitochondrial

"methylene trap" since folates do not exit to the cytoplasm to a significant extent

(Cybulski and Fisher 1976, 1981, Horne et al. 1989, Trent et al. 1991). Although

the cell lines used in this study incorporate [1_14C]glycine and [2-14C]glycine

equally into ONA, indicating that neither cell line carries out significant glycine

c1eavage activity, this defect could be very significant in tissues that require the

pathway in vivo. Since the nmdmc(-I-) cell Iines cannot provide mitochondrially

derived one-carbon units, they are entirely dependent on the cytoplasmic folate

pathways. Unlike wild type cells, growth of the nmdmc(-I-) cell Iines is stimulated

by added formate that can be substituted by supplementation with hypoxanthine

but not by thymidine. These results demonstrate that the nmdmd-I-) cells cannot

provide sufficient one-carbon units via cytoplasmic pathways to allow for optimal

purine synthesis. This conclusion is also supported by studies with radiolabeled

precursors that show a 3 to 10-fold enhancement of formate over serine

incorporation by the nmdmc(-I-) cells relative to the wild type cells. Since the

mitochondria of nmdmc(-I-) cells cannot produce intracellular formate, then the

radiolabeled formate added to the medium is preferentially incorporated into

nucleotides.

111 The glycine auxotrophy demonstrates that the immortalized or transformed cells do not express a second mitochondrial OC or DCS activity. If a second enzyme were expressed to regenerate THF, we would not observe the glycine auxotrophy. Two types of evidence have shed some light on the origin of the

NMDMC protein. First, kinetic properties demonstrated a residual Mg2+_ dependent ability to use NADP and suggested that NMDMC evolved from an

NADP+ -dependent OC by using inorganic phosphate (Pi) and Mg2+ to bind the

NAD+ cofactor (Yang and MacKenzie 1993, Pawelek and MacKenzie 1996).

Second, our recent demonstration that the long 3'-untranslated region of the

NMDMC cDNA contains five sequences averaging more than 100 nucleotides each with significant homologies to the synthetase domain in both the mouse and human DCS enzymes and two su ch sequences in the Drosophila (Patel et al.

2002) support the proposai for the evolution from a trifunctional precursor through loss of the S domain. Yeast contains two nuclear-encoded NADP+ -dependent trifunctional DCS proteins, one located in mitochondria and the other in the cytoplasm (Shannon and Rabinowitz 1986). It now seems likely that the NMDMC protein evolved from a mammalian mitochondrial DCS through a change of cofactor requirement and the loss of the synthetase domain. The change of cofactor from NADP+ to NAD+ has been estimated to shift the equilibrium between methyleneTHF and formylTHF in mitochondria 60 to 200-fold toward the latter

(Pelletier and MacKenzie 1995) and thus would strongly benefit the ultimate production of formate from serine.

If NMDMC is the only OC expressed in the mitochondria of these celllines,

112 then there is no synthetase domain available to remove the formate from formylTHF as is proposed in the model of the yeast system (Pasternack et al.

1994, West et al. 1996, Kastanos et al. 1997, Piper et al. 2000). It is possible that there is a separate mitochondrial synthetase enzyme in mammalian cells, but it is not clear as to how this would be beneficial over a synthetase domain aUached to the OC domain. Moreover, having the synthetase operate in the "reverse direction" in the presence of significant concentrations of ATP is probably a less efficient system to release formate than, for example, the expression of a putative formylTHF hydrolase. Since mitochondrial protein synthesis does not seem to require formylTHF, such a hydrolase activity could be localized safely within the mitochondria. Transport of formylTHF out of the mitochondria for use in the cytoplasm would achieve the same goal and provide formylTHF directly to cytoplasmic enzymes, but this is believed to be too slow a process to support metabolic activity (Cybulski and Fisher 1976, 1981, Horne et al. 1989, Trent et al.

1991). Several metabolic studies have used immortalized mammalian cell lines to investigate fluxes of folate-related intermediates, and ail support the concept of formate release by mitochondria. If ail immortalized mammalian cell lines express only this single mitochondrial OC, then the mechanism of this formate production requires further elucidation.

113 FIGURE 6. Folate-dependent activities in the cytoplasm and mitochondria of mammalian cells.

Abbreviations are: SHMT, serine hydroxymethyltransferase; D, methyleneTHF dehydrogenase; C, methenylTHF cyclohydrolase; S, 10-formylTHF synthetase; GCS, glycine cleavage system. * indicates missing enzyme in null mutant cells. Adapted from Di Pietro et al. (2002).

114 Serine E ) Serine

Methionine THF~ 1--- ) THF SHMT SHMT l ----. Glycine E ) Glycine" ~ 5-methylTHF ~ ==Methylene TH F GCS'-... MethyleneTHF Thymidylate ... NADP~ NAD !i2+9 D' Pi NADPH.-1 D NADH

MethenylTHF MethenylTH F

lc 1C' Purines ~ ~ FormylTHF FormylTHF

ADP+Pi~S ATP~THF ~/\

Formate E ) Formate Met-tRNNmet

Cytoplasm Mitochondria Studies with the nmdmc(-I-) cell lines indicate that a metabolic role for

NMDMC is to provide one-carbon units for purine synthesis. Despite recognition of the metabolic consequences of the gene deletion at a cellular level, we cannot explain the embryonic lethality. The ability of heterozygous dams to synthesize glycine, the abundance of glycine in the diet and the failure of glycine to improve the phenotype of (-/-) embryos make it unlikely that embryonic lethality is due to the lack of glycine. Although formate stimulates the growth of null mutant fibroblasts, injection of sodium formate into pregnant females also did not improve the phenotype of embryos at E 13.5. How the nmdmc(-I-) block in mitochondrial folate metabolism causes the apparently rather specifie inability to establish hematopoiesis in the developing liver is still not clear.

115 Acknowledgments-We thank Dr. Eric Shoubridge (McGili University) for providing the pO cell lines and useful advice and Narciso Mejia from this laboratory for technical assistance.

116 CHAPTER FOUR

THE ROLE OF MITOCHONDRIAL METHYLENETETRAHYDROFOLATE

DEHYDROGENASE-CYCLOHYDROLASE ACTIVITIES IN

FORMATE PRODUCTION IN MAMMALIAN FIBROBLASTS

117 PREFACE

Since we had established that the NMDMC plays an important role in supporting purine biosynthesis in the cytoplasm of mammalian cells, in this chapter 1 attempted to substitute the NMDMC protein with different methylenetetrahydrofolate dehydrogenases in the mitochondria of the NMDMC null mutant fibroblasts.

The following chapter is a manuscript in preparation entitled "The Role Of

Mitochondrial Methylenetetrahydrofolate Dehydrogenase-Cyclohydrolase

Activities ln Formate Production ln Mammalian Fibroblasts" by Harshila Patel,

Erminia Di Pietro and Robert E. MacKenzie.

Ali of the results presented in this chapter were obtained by the candidate.

Erminia Di Pietro collaborated in the construction of the pL(NMDMC)SH expression vector and provided the NMDMC null cell lines. Narciso Mejia performed the polyacrylamide gel and Western analysis on the samples provided by the candidate.

118 SUMMARY

Mitochondrial NAD-dependent methylenetetrahydrofolate dehydrogenase­ cyclohydrolase (NMDMC) null mutant fibroblasts were infected with retroviral expression constructs of dehydrogenase/cyclohydrolase enzymes. Cellular fractionation confirmed that the expressed proteins were properly targeted to the mitochondria of these cells. Expression of the NAD-dependent methylene­ tetrahydrofolate dehydrogenase-cyclohydrolase cDNA corrected the glycine auxotrophy of the null mutant cells. A construct in which the cyclohydrolase activity of NMDMC was inactivated by point mutation also rescued the glycine auxotrophy, although poorly. This suggests that while chemical hydrolysis of methenyltetrahydrofolate can occur in mitochondria, the cyclohydrolase activity is required to ensure optimal production of 10-formyltetrahydrofolate. A cDNA construct expressing the NADP-dependent dehydrogenase-cyclohydrolase­ synthetase in the mitochondria also rescued the glycine auxotrophy of the null cells. Thus, the NAD cofactor specificity of the mitochondrial methylene­ tetrahydrofolate dehydrogenase is not absolutely essential to maintain the flux of one-carbon metabolites. The ability of ail rescued cell lines to grow without glycine correlated with a decrease in the ratio of incorporation of formate to serine in standardized radiolabeling studies from approximately 2.5 for nmdmc(-I-) cells toward 0.3 for the wild type cells. This ratio is a good qualitative indicator of the ability of cells to generate intracellular formate.

119 INTRODUCTION

ln eukaryotic cells, methylenetetrahydrofolate dehydrogenases are involved in the interconversions of tetrahydrofolate (THF) derivatives required for methionine regeneration as weil as thymidylate and purine biosynthesis

(MacKenzie 1984). There is an NADP-dependent dehydrogenase, which constitutes a trifunctional protein also containing methenyltetrahydrofolate cyclohydrolase and formyltetrahydrofolate synthetase activities in the cytoplasm. ln yeast, there is also an NADP-dependent DCS in the mitochondria (Shannon and Rabinowitz 1986). However, the mammalian mitochondrial counterpart is a bifunctional NAD-dependent methyleneTHF dehydrogenase-methenylTHF cyclohydrolase (NMDMC) (Mejia and MacKenzie 1988).

NMDMC activity was first detected in extracts of Ehrlich ascites tumor cells

(Scrimgeour and Huennekens 1960). It has since been detected in embryonic, undifferentiated and transformed cells but only in the developmental cells of normal adult tissues (Mejia and MacKenzie 1985). Unlike the yeast mitochondrial

NADP-dependent DCS, which was determined to be completely dispensable

(Shannon and Rabinowitz 1988), the NMDMC plays an essential role in mammals because the inactivation of this gene in mice was embryonic lethal (Di Pietro et al.

2002).

From kinetic evidence (Yang and MacKenzie 1993) and sequence comparisons (Patel et al. 2002) we have shown that the mitochondrial NAD­ dependent DC likely evolved from a pre-existing NADP-dependent DCS

120 precursor. It has been proposed that NMDMC changed cofactor specificity from

NADP to NAD to drive the mitochondrial folate pathway towards the production of

10-formylTHF (Yang and MacKenzie 1993, Pelletier and MacKenzie 1995)

because the ratio of NADH/NAD is 1 in the mitochondria of liver cells, whereas

the ratio of NADPH/NADP is 15 (Sies 1982). This shift in the equilibrium position

of THF derivatives towards 10-formylTHF in the mitochondria would allow for the

generation of formate, which could exit the mitochondria and provide one-carbon

units for purine synthesis in the cytoplasm. Metabolic studies on the NMDMC null

mutant cell lines (Patel et al. 2003) demonstrated that formate or hypoxanthine

stimulated their growth. The block in the mitochondrial pathway created by the

inactivation of the NMDMC gene does not allow for the generation of formate from

serine in this compartment. The question still remains as to how this

mitochondrial formate is being generated from the 10-formyITHF.

The NMDMC null cell lines provide a background in which we can ask

about the importance of the NAD specificity and the role of the cyclohydrolase

activity by using retrovirally expressed dehydrogenase-cyclohydrolase enzymes

targeted to the mitochondria. Full rescue would be expected to reverse the

glycine auxotrophy and reduce the dependence on exogenously added formate.

121 MATERIALS AND METHODS

Construction of Expression Vectors-The pCMVSport6-NMDMC vector was purchased from Open Biosystems, Inc. PCR primers were synthesized by

Integrated DNA Technologies, Inc. and PCR amplification was performed using a

Progene (Tech ne) thermocycler. Ali restriction enzymes were purchased from

New England Biolabs. The sequence of each cDNA expression construct was verified by automated sequencing at the Sheldon Biotechnology Center, McGill

University.

The K56Q mutation (used to create a monofunctional dehydrogenase) was introduced into the NMDMC cDNA using the in vitro overlap-extension method of

PCR mutagenesis (Ausubel 1997). The PCR reaction mix contained template

DNA (54 ng), 50 pmol of each primer, 125 I-IM of each dNTP, 1 mM MgS04 , 0.1 mg/ml BSA, 1x Thermopol buffer and 1 unit of Vent polymerase (New England

Biolabs) in a total volume of 50 1-11. The PCR conditions were as follows: 1 cycle at 94 oC for 5 min; 35 cycles at 94 Oc for 30 s,55 Oc for 30 s, and 72 Oc for 30 s; and 1 cycle at 72 Oc for 10 min. The products of the first two PCR reactions

(using sense mutagenic primer with antisense 3' flanking oligo and antisense mutagenic primer with 5' flanking oligo) were identified by gel electrophoresis and purified by the UltraClean DNA purification kit (Mo Bio Laboratories Inc.). The two

PCR products (27 ng) were used as template in a final PCR reaction with the 5' and 3' flanking oligo primers. This final PCR product was restricted with SaIl and

122 Msel and ligated into the wild-type pCMVSport6-NMDMC.

The mitochondrial leader sequence was amplified from the NMDMC cDNA by four rounds of PCR with a 5' flanking external primer and a series of overlapping primers using the mouse DCS cDNA (5.4 ng) as the template for the first PCR followed by the PCR product from the previous PCR as the template for the subsequent PCR. The PCR reaction mix contained template DNA (5.4 ng),

50 pmol of each primer, 125 !JM of each dNTP, 1 mM MgS04, 0.1 mg/ml BSA, 1x

Thermopol buffer and 1 unit of Vent polymerase (New England Biolabs) in a total volume of 50 !JI. The PCR conditions were as follows: 1 cycle at 94 oC for 5 min;

35 cycles at 94 oC for 30 s, 61°C for 30 s, and 72 oC for 45 s; and 1 cycle at 72

Oc for 10 min. The complete mitochondrial leader sequence was purified with the

Qiagen PCR purification kit, digested with Neol and Sbfl and ligated into the full­ length mouse DCS cDNA. The entire mouse DCS cDNA with the mitochondrial leader sequence was excised with Neol and Xhol and blunt-end ligated into the

Sail - Xhol sites of the pCMVSport6 vector.

Alternatively, the DC domain with the aUached mitochondrial leader sequence was amplified out of the full-Iength mouse DCS cDNA using the primer consisting of the 5' most sequence of the mitochondrial leader and a 3' primer designed to insert a stop codon after the DC domain. The PCR reaction mix contained template DNA (27 ng), 50 pmol of each primer, 125 !JM of each dNTP,

1 mM MgS04, 0.1 mg/ml BSA, 1x Thermopol buffer and 1 unit of Vent polymerase

(New England Biolabs) in a total volume of 50 !JI. The PCR conditions were as

123 follows: 1 cycle at 94 oC for 5 min; 35 cycles at 94 oC for 30 s,55 oC for 30 s, and

72 Oc for 45 s; and 1 cycle at 72 oC for 10 min. The PCR product was purified with the Qiagen PCR purification kit, digested with Ncol and Xhol and blunt-end ligated into the Sa/l-Xhol sites of the pCMVSport6 vector.

Each of the cDNA expression constructs in the pCMVSport6 vector were transferred into the pL(rfa)SH expression vector (Antonicka et al. 2003) via the successive LR and BP recombination steps (Gateway Cloning System) to yield the retroviral expression vectors. The polyA signal of the NMDMC cDNA, located between the two Hindlll sites, was excised out of the pL(NMDMC)SH and pL(K56Q)SH expression vectors.

GPE86 Transfection-GPE86 cells (obtained from Dr. E. Shoubridge) were plated at a density of 1 x 106 cells in two T75 flasks. The next day, the medium was changed 3 h prior to the transfection. For the transfection, 5 \Jg of expression construct DNA, 445 \JI of sterile distilled water and 50 \JI of 2.5 M calcium chloride were pipetted into a sterile 1.2-ml vial. Subsequently, 500 \JI of 2x HEPES were pipetted into another 1.2-ml vial. Then, the DNA/calcium chloride solution was added dropwise to the bubbling HEPES solution. The precipitate was allowed to sit for 20 min at room temperature before it was added dropwise to the GPE86 cells in one of the T75 flasks. The medium was pipetted gently up and down once. The next day, the transfected GPE86 cells were washed twice with phosphate-buffered saline pH 7.2 and fresh medium was added. The next day, the cells were subjected to 0.1 mg/ml hygromycin selection for 12 days with the

124 medium being changed daily. The hygromycin-resistant clones were trypsinized and plated in a new T75 flask.

PA317 Infection-PA317 cells (obtained from Dr. E. Shoubridge) were

plated at a density of 5 x 105 cells in two T75 flasks. The next day, virus­

containing supernatant was collected from a T75 flask containing confluent

transfected GPE86 cells grown in selection-free medium. The supernatant was filtered with a 0045 IJm membrane into a 15-ml conical tube. In another tube, 1.5

ml of the virus stock and 3 ml of medium containing 4 IJg/ml polybrene were

pipetted. The medium in the T75 flask containing PA317 cells was discarded and

replaced with the virion-containing medium. The flask was incubated at 37 Oc for

2 h. Another 5 ml of the medium containing 4 IJg/ml polybrene was added to the flask and the incubation was continued overnight at 37 oC. The next day, the

infected PA317 cells were washed once with 5 ml of medium and 10 ml of fresh

medium was added to the flask. The next day, the cells were subjected to 004

mg/ml hygromycin selection for five days with the medium being changed daily.

The hygromycin-resistant clones were trypsinized and plated in a new T75 flask.

Infection of NMOMC Null Mutant Immortalized Fibroblasts-IF74(-/-) cells

(Di Pietro et al. 2002) were plated at a density of 3 x 105 cells in two 10-cm plates.

The next day, virus-containing supernatant was collected from a T75 flask

containing confluent infected PA317 cells grown in selection-free medium. The

supernatant was filtered with a 0045 IJm membrane into a 15-ml conical tube. In

another tube, 1.5 ml of the virus stock and 3 ml of medium containing 4 IJg/ml

125 polybrene were pipetted. The medium in the 10-cm plate containing IF74(-/-) cells was discarded and replaced with the virion-containing medium. The plate was incubated at 37 oC for 2 h. Another 5 ml of the medium containing 4 I-lg/ml polybrene was added to the plate and the incubation was continued overnight at

37 oC. The next day, in order to ensure conditions to allow for the isolation of discrete colonies, the infected IF74(-/-) cells were split 1:10 and 9:10 into two new

10-cm plates. The next day, the cells were subjected to 0.2 mg/ml hygromycin selection for seven days with the medium being changed daily.

Selection, Expansion and Freezing of Clones-The hygromycin-resistant clones were individually trypsinized using cloning discs and transferred to 24-well plates. A few days later, the medium was changed in each weil and the cloning disc was removed. The clones were propagated in 24-well plates and stocks were frozen.

Western analysis-Cells from 24-well plates were pelleted and resuspended in 50 I-li of standard sample butter (Laemmli 1970). The cells were

Iysed in a boiling water bath for 5 min, centrifuged at 15 000 rpm for 5 min. The supernatant was collected from each sample, mixed with 20 I-li of sam pie butter and loaded onto a SDS-polyacrylamide slab gel consisting of a 9% separating gel and a 3% stacking gel. The proteins were separated by electrophoresis at a current of 25-35 mA for 3 h as described by Laemmli (1970) and were transferred electrophoretically onto a Protran nitrocellulose membrane (Mandel) for 3 h as described by Towbin et al. (1979). After the transfer, the membrane was rinsed in

126 washing solution #1 (TBS, 0.5% Tween 20). The membrane was incubated in

100 ml of blocking solution (TBS, 5% skim milk) for 2 h and washed two times for

30 min each in washing solution #1 at room temperature. The membrane was incubated in 50 ml of either the primary rabbit anti-mouse NMDMC or DCS antibody diluted 1000-fold in washing solution #2 (TBS, 1% skim milk) overnight at 4 oC. The following morning, the membrane was washed two times for 30 min each in 50 ml of washing solution #1 and four times for 30 min each in 50 ml of washing solution #2 at room temperature. Then, the membrane was incubated in

50 ml of peroxidase-conjugated affinipure donkey anti-rabbit IgG (H + L) secondary antibody (Jackson Immuno Research) diluted 10 OOO-foid in washing solution #2 for 1 h at room temperature. Once again, the membrane was washed two times for 30 min each in 50 ml of washing solution #1 and fourtimes for 30 min each in 50 ml of washing solution #2 at room temperature. The protein bands were visualized by Western Lightning Chemiluminescence Reagent (Perkin Elmer

Life Sciences).

Cellular Fractionation-Cells were grown to 100% confluency on two 10- cm TC plates. The cells were washed twice with 5 ml of cold phosphate-buffered saline, pH 7.2 and then incubated in 2 ml of PBS-citrate EDTA (4 g/I citrate, 0.6 mM EDTA) for 10 min at 37 oC until they started to lift, according to Ruffolo et al.

(2000). The cells were centrifuged at 1000 rpm for 5 min at 4 oC, washed once with 2 ml of PBS, pH 7.2. and recentrifuged at 1000 rpm for 5 min at 4 oC. Each cell pellet was resuspended in 100 \JI of HlM buffer (200 mM mannitol, 70 mM

127 sucrose, 10 mM HEPE8, 1 mM EGTA, adjusted to pH 7.5 with potassium hydroxide). The cells were homogenized in a 2-ml Teflon homogenizer with 30 up/down strokes at 2000 rpm. A 10-1..11 aliquot of the whole cell Iysate was retained. The remaining cell Iysate was centrifuged at 1000 rpm for 5 min at 4 oC.

The supernatant was transferred to another eppendorf tube (P10 fraction). Each cell pellet was resuspended in 100 1..11 of HlM buffer and the homogenization was repeated. The cells were centrifuged at 1000 rpm for 5 min at 4 oC. The supernatant was pooled with the first P10 fraction. The total supernatant was centrifuged at 9000 rpm for 5 min at 4 oC. The final supernatant was transferred to another eppendorf tube. This represented the 810 fraction, which contained the cytoplasm and endoplasmic reticulum. The resulting pellet was resuspended in an equal volume of HlM buffer. This represented the mitochondrial fraction.

The total proteins from the mitochondrial and cytoplasmic fractions were quantified and predetermined amounts were analyzed by Western analysis.

Cell Growth Studies-The clones with the highest levels of rescued protein expression used in these experiments were grown in glycine-free medium over the course of a five-day period as described by Patel et al. (2003).

Precursor Incorporation into DNA-The incorporation of [3-14C]serine and

C4C]formate into total DNA of each of the rescued cell lines was measured as described by Patel et al. (2003).

128 RESULTS

Rescue of NMDMC Nu" Mutant Ce" Lines-Retroviral expression constructs were prepared that contained the NMDMC cDNA or the K56Q mutant cDNA. After infection of each construct into NMDMC null mutant fibroblasts, hygromycin-positive clones were screened by Western analysis using an antibody to NMDMC to detect either NMDMC or NAD-dependent monofunctional dehydrogenase protein expression. The cell lines showed varying levels of protein expression in infected cells as shown in Fig. 1. This is expected because of different sites of integration of each construct in the genome of each ceilline.

129 FIGURE 1. Western blot analysis of rescued NMDMC nu" mutant fibroblasts.

Equal amounts of crude total cellular prote in extracts (1 OO~g) were separated by SDS-PAGE and either NMDMC or NAD-dependent monofunctional dehydrogenase (K56Q) protein levels were determined in the rescued cell lines by Western blot analysis. IF22(+/+) is a wild type cell li ne and IF74(-/-) is a null mutant cell line. Clones 2-1 D and 3-4D are NMDMC-rescued cell lines. Clones 1-58 and 1-6C are K56Q-rescued cell lines.

130 (-/-)17 L.:II Cellular Fractionation of NMDMC-Rescued Cells-Cellular fractionation was performed to verify that the NMOMC protein was properly targeted to the mitochondria of the rescued cell lines. The mitochondria were isolated from one

I of the NMOMC-rescued cell lines, clone 2-10, and a nmdmc- - cell line, IF74(-/-) and the presence of NMOMC was determined in each cellular fraction by Western analysis (Fig. 2). We could not detect NMOMC protein in any of the cellular fractions of the null mutant cell line, which served as a negative control. The

NMOMC protein was found in the whole cell Iysate of the rescued cell line, 2-10 and importantly, the protein is enriched in the mitochondrial fraction despite the higher amount of total protein loaded onto the SOS-PAGE gel from the cytoplasmic fraction.

131 FIGURE 2. Subcellular localization of NMDMC protein.

The mitochondria were isolated from NMDMC null mutant cell line, IF74(-I-) and one of the NMDMC-rescued cell lines, 2-10 by differential centrifugation. The expression of the NMDMC protein in each cellular fraction was determined by Western analysis. The cellular fractions are H, homogenate; M, mitochondrial and C, cytoplasmic. For the IF74(-I-) ceilline, 59 I-Ig of the M and 100 I-Ig of the C fractions were used. For the 2-10 cell line, 76 I-Ig of the M and 91 I-Ig of the C fractions were used. For both ceillines, 10% of the total H was used.

132 o

H

H Rescue of Nu" Mutant Ce" Unes with the NA DP-Dependent DCS

Construct-Hygromycin-resistant clones of NMDMC null mutant cells obtained following infections with retrovirus containing the cDNA expressing the

mitochondrially targeted NADP-dependent DCS were screened for the presence

of elevated amounts of the protein using an antibody against the DCS enzyme.

However, in this case the null mutant cells contain large amounts of the naturally­

occurring cytoplasmic DCS, and we had to determine if this rescued protein was

appropriately expressed in the mitochondria. To do this, the hygromycin-resistant

clones had to be subjected to subcellular fractionation to show the presence of

DCS in the mitochondria. Figure 3 shows the subcellular distribution of DCS in the null mutant ceilline, IF74(-/-) and a rescued ceilline, clone 2-5A. Comparing

equal amounts of total cytoplasmic and mitochondrial protein (Fig. 3A), it is

apparent that the DCS is enriched in the mitochondria of the rescued cell line. A

beUer approach is to compare 10% of the total protein in each of the cytoplasmic

and mitochondrial fractions of the null mutant and rescued cell lines (Fig. 38). In this case, no DCS can be detected in the mitochondria of the null mutant cell, but

good expression is seen in the mitochondria of the rescued cell. As shown in

Figure 4C, the presence of the DCS in the mitochondria relieves the strict glycine

auxotrophy.

133 FIGURE 3. Subcellular localization of NADP-dependent Des protein.

The mitochondria were isolated from NMDMC null mutant cell line, IF74(-I-) and one of DCS-rescued celilines, 2-5A by differential centrifugation. The expression of the DCS protein in each cellular fraction was determined by Western analysis. The cellular fractions are H, homogenate; M, mitochondrial and C, cytoplasmic. ln each case, 10% of the total homogenate was used. In A, 10 j.Jg of total protein from each cellular fraction was used. In B, 10% of total protein from each cellular fraction was used.

134 IF7 4( -1-) 2-5A

1 1

I u I u A

B Glycine Requirements for Rescued Ce" Lines-We had shown earlier

(Patel et al. 2003) that the NMDMC null mutant cell lines are glycine auxotrophs. ln order to determine whether the expression of the NMDMC, K56Q or NADP­ dependent DCS constructs in NMDMC null mutant cell lines rescued this phenotype, the glycine requirements of NMDMC-, K56Q- and DCS-rescued cell lines were determined by growth experiments. The wild type NMDMC protein rescued the glycine auxotrophy of the null mutant cells as demonstrated by the growth of several infected cell lines in glycine-free medium (Fig. 4A). A second construct in which the cyclohydrolase activity of NMDMC was inactivated by the

K56Q mutation, resulting in an NAD-dependent monofunctional dehydrogenase construct (Schmidt et al. 2000), was also weil expressed in these null mutant cells and reduced their dependency on glycine, although not as weil as did the wild type enzyme (Fig. 48). The expression of the NADP-dependent DCS protein in the mitochondria also reversed the glycine auxotrophy of the null mutant cells

(Fig. 4C). In fact, a couple of the DCS-rescued cell lines demonstrate better growth in glycine-free medium than the K56Q-rescued cell lines.

135 FIGURE 4. Growth of rescued cel! Iines in glycine-free medium containing 10% redialyzed fetal bovine serum.

The glycine requirements of the NMOMC-rescued cell lines (A), K56Q-rescued celllines (8) and OCS-rescued celllines (C). Symbols indicate: _, wild type cells and., null mutant cells. In the graph showing the growth of NMOMC-rescued celllines (A), symbols indicate: D, 1-18; 0, 1-3A; 0, 2-10 and Il, 3-40 cells. In the graph showing the growth of K56Q-rescued celllines (8), symbols indicate: D, 1-4C; 0, 1-58; 0, 1-6C and Il, 2-3C cells. In the graph showing the growth of OCS-rescued celllines (C), symbols indicate: D, 2-2C; 0, 2-20; 0, 2-5A and Il, 2- 50 cells.

136 140 A 1- Z 120 ::> 0 100 Ü 80 ....J ....J W 60 Ü >< 40 "'f 20 ...-0 0 24 48 72 96 120 HOURS

200 B 1- Z 160 ::> 0 ü 120 ....J ....J W 80 Ü >< "'f 40 ...-0 0 24 48 72 96 120 HOURS

140 C 1- 120 Z ::> 0 100 Ü 80 ....J ....J W 60 Ü >< 40 "'f 20 ...-0 0 24 48 72 96 120 HOURS Determination of the Exogenous Formate Versus Serine Incorporation into

the DNA of the Rescued Cell Unes-The NMDMC null mutants were also shown

previously to prefer formate as a one-carbon donor according to a series of

radiolabeling experiments that measured the incorporation of C4C]-labeled formate and [3-14C]-labeled serine into the total DNA of exponentially growing wild

type and null mutant cells. The ratio of formate/serine was determined to be

approximately 0.3 for the wild type cells and approximately 2.5 for the null mutant

cells. This ratio serves as a sensitive indicator of how efficiently the mitochondria

are generating formate from serine and contributing to the cytoplasmic folate

pathways. In the case of the NMDMC-rescued ceillines, the ratio varies from 0.7,

which is almost the wild type condition, to 2.5, which is closer to the null mutant

condition as shown in Table 1. Interestingly, of these four cell lines, those that

have the lower formate to serine ratio are the ones that grow faster in glycine-free

medium (Fig. 3A). In the case of the K56Q-rescued cell lines, the ratio varies

between 1.0 and 1.2 whereas in the case of the DCS-rescued cell lines, the ratio

varies between 0.9 and 1.7 as shown in Table 1. These ceillines show the same

relationship as seen with the NMDMC-rescued cell lines whereby the lower formate to serine ratio corresponds to better growth of the cell li ne in glycine-free

medium (Fig. 3, 8 and C).

137 TABLE 1

Incorporation of radiolabeled precursors into total DNA

Following a 24-h incubation of cells with the one-carbon donors [3-14C]serine or C4C]formate, radioactivity was measured in total ONA. Results are presented as the ratio of radiolabel incorporated fram formate relative to serine.

Glycine Cell type Ceilline Formate/Serine Auxotraph Wild type IF22 (+/+) no 0.3 NMOMC null mutant IF74 (-/-) yes 2.5 1-1 B no 0.7 1-3A no 1.7 NMOMC-rescued 2-10 no 0.8 3-40 no 2.5 1-4C no 1.0 1-5B no 1.2 K56Q-rescued 1-6C no 1.1 2-3C no 1.2 2-2C no 0.9 2-20 no 1.7 OCS-rescued 2-5A no 1.3 2-50 no 1.2

138 DISCUSSION

NMDMC null mutant fibroblasts were previously shown to be glycine auxotrophs (Patel et al. 2003). Furthermore, the block in the mitochondrial folate pathway as a result of the lack of NMDMC activity in these cells demonstrated a reduced capability of de novo purine synthesis provided by the remaining cytoplasmic pathways. Rescue of the null mutant fibroblasts with constructs expressing NMDMC partially reverses the metabolic defects, confirming that the defects are due to only the absence of this gene product.

The major questions on the role of this enzyme remain. The first is to demonstrate the advantage of the enzyme in evolving to use NAD rather than

NADP. Our earlier proposai (Yang and MacKenzie 1993, Pelletier and

MacKenzie 1995) was that this shifts the equilibrium of methyleneTHF and formylTHF much more in the direction of the latter. However, while the NMDMC construct rescued the glycine auxotrophy, so did the construct expressing the cytoplasmic NADP-dependent DCS activities in mitochondria of the null cells, which indicates that the NAD cofactor is not absolutely required. We were surprised to find that these DCS-rescued cells appear to grow almost as weil as the NMDMC-rescued cells. In fact, the DCS-rescued cell lines are now comparable to the yeast system (Pasternack et al. 1994, West et al. 1996,

Kastanos et al. 1997, Piper et al. 2000). This brings into question as to why the mammalian system evolved to substitute an NADP-dependent methyleneTHF dehydrogenase with one that is NAD-dependent. However, it is important to

139 remember that the a rtifi ci a 1 conditions of these cell lines most likely do not accurately reflect what is occurring during embryogenesis in the organism itself. It is not clear whether the extra synthetase activity expressed by the protein changes the overall conversion of formylTHF to formate in these cells. We tried to express only the NADP-dependent DC301 construct in these cells to better mirror the NMDMC, but the protein was degraded completely to yield a smaller product than the expected 34-kDA protein. It is possible that the cytoplasmic

NADP-dependent OC domain is not as stable as the mitochondrial NMDMC because the latter protein has an extension at the carboxyl terminus (Belanger and MacKenzie 1989).

The second question on NMDMC function relates to the role of the cyclohydrolase activity. MethenylTHF hydrolyzes nonenzymatically to formylTHF at a significant rate (Kay et al. 1960) and at neutral pH, the concentration of formylTHF is 10-fold that of methenylTHF. The cytoplasmic NADP-dependent OC domain is optimized to catalyze the conversion of formylTHF to methyleneTHF.

This conclusion is supported by kinetic evidence that demonstrates 100% channeling of the formylTHF to methyleneTHF (Pawelek and MacKenzie 1998) but only 50% channeling of methyleneTHF to formylTHF in the forward direction

(Rios-Orlandi and MacKenzie 1988). Moreover, the dehydrogenase activity is rate limiting in the forward direction (Green et al. Biochemistry 1988, Pawelek and

MacKenzie 1998) whereas the cyclohydrolase activity is severely rate limiting in the reverse direction (Pawelek and MacKenzie 1998). If the one-carbon flow in

140 mammalian cells is the same as proposed for yeast, th en formylTHF to methyleneTHF conversion is important for cytoplasmic function and we would

predict that the cyclohydrolase activity is essential.

However, if the flow in mitochondria is metabolically important only in the

direction of methyleneTHF to formylTHF, it is possible that chemical

transformation could at least partially contribute to this role. Site directed

mutagenesis studies on the human cytoplasmic DC301 protein showed that the

substitution of the at residue 56 of this protein with a glutamine results in a

monofunctional dehydrogenase enzyme (Schmidt et al. 2000). This K56Q mutant

protein has no detectable cyclohydrolase activity but retains approximately 50% of

the wild type dehydrogenase activity. The results from the growth studies of the

K56Q-rescued cell lines demonstrate a reversai of the glycine auxotrophy of the

NMDMC nUII mutant fibroblasts but these cell lines do not grow as weil as the

NMDMC-rescued celllines in glycine-free medium. Therefore, it appears that the

cyclohydrolase activity is required to ensure optimal production of formylTHF in

the mitochondria. However, the K56Q mutant has only 50% of the wild type NAD­

dependent dehydrogenase activity (Schmidt et al. 2000) so that the less effective

rescue of the glycine auxotrophy could actually over-estimate the apparent

requirement for the cyclohydrolase reaction.

NMDMC null mutant fibroblasts were also previously shown to prefer

formate as a one-carbon donor as demonstrated by their increased incorporation

of radiolabeled formate as compared to radiolabeled serine into their total DNA

141 (Patel et al. 2003), which is a result of the block in the mitochondrial folate pathway in these cells. Consequently, exogenously supplied formate is required to meet the one-carbon demands for purine biosynthesis in the cytoplasm. In the case of each rescued cell line, although the formate/serine incorporation ratio does shift away from the null mutant condition of 2.5, none actually attain the wild type condition, which is represented by 0.3. Furthermore, there is a correlation for each rescued cell line between the low formate/serine ratio and its ability to grow on glycine-free medium. Since this radiolabeling assay can demonstrate up to a

10-fold change between the wild type and null mutant condition, it represents a more sensitive assay than the stimulation of the growth of the null mutant cells with either formate or hypoxanthine, which only demonstrates a 2-fold change

(Patel et al. 2003). Despite the limitation of this assay being that it only permits comparisons within the sa me cell line, this ratio still serves as a good qualitative measure of how efficiently the mitochondrial folate pathway is producing formate and contributing to total cellular function.

ln summary, we were able to rescue NMDMC null mutant fibroblasts with

NMDMC, NAD-dependent dehydrogenase and NADP-dependent DCS constructs.

As predicted, the restoration of only the NAD-dependent dehydrogenase activity demonstrates that the cyclohydrolase activity is not essential in mammalian mitochondria but it is not clear why NADP-dependent DCS appears to rescue almost as weil as the NMDMC.

142 Acknowledgments-We thank Dr. Eric Shoubridge for providing the pL(rfa)SH vector, GPE86 and PA317 cells as weil as useful advice.

143 CHAPTER FIVE

GENERAL DISCUSSION

144 Folate metabolism is essential for cellular division and proliferation as it is required for purine and thymidylate synthesis as weil as for the regeneration of methionine from homocysteine that is necessary to support cellular methylation reactions. In eukaryotic cells, the folate pathways are compartmentalized between the mitochondria and the cytoplasm and a major effort in the field in recent years has been to understand how mitochondrial folate metabolism contributes to the function of the cell. One of the key enzymes in these pathways is the trifunctional NADP-dependent DeS that is found as different gene products in both compartments in yeast. It has been proposed that the mitochondrial DeS generates formate, which exits into the cytoplasm and is interconverted to the various folate intermediates by the cytoplasmic DeS (Pasternack et al. 1994,

West et al. 1996, Kastanos et al. 1997, Piper et al. 2000). It was certainly reasonable that the yeast model might weil function in mammalian systems. In the mammalian system, the NADP-dependent DeS is present in the cytoplasm.

However, there has been much controversy as to whether mammalian mitochondria also contain a trifunctional DeS and this has been pursued by severallaboratories. One group (Barlowe and Appling 1988) claimed that through subcellular fractionation of rat liver, they could demonstrate the three enzyme activities of the DeS protein in the isolated mitochondria. However, it was possible that the apparent mitochondrial DeS activities could actually be accounted for by the contamination of these activities by the cytoplasmic DeS protein. This conclusion could not be supported or disproved by results obtained

145 in this laboratory because the low amounts made it difficult to rule out such contamination.

However, the yeast model was weil established and several groups used this as the model for interpreting mammalian studies (Barlowe and Appling 1988,

Garcfa-Martfnez and Appling 1993, Fu et al. 2001). Clearly, it was important to establish the nature of any methylenetetrahydrofolate dehydrogenases in mitochondria. Our laboratory pursued an interesting observation by Scrimgeour and Huennekens (1960) that led to the isolation and purification of an NAD­ dependent methylenetetrahydrofolate dehydrogenase-cyclohydrolase (Mejia et al.

1986) that was later shown to be a mitochondrial enzyme (Mejia and MacKenzie

1988). It seemed pointless for mammalian mitochondria to express both NAD­ and NADP-dependent dehydrogenases since their differing cofactor specificities would lead to the creation of a futile loop of the flow of folate intermediates in the mitochondria. If there are both OC and DCS enzymes in mitochondria, it would be reasonable to expect that they not be expressed under the same conditions. ln the absence of a purified enzyme from mitochondria, the cloning of a putative mitochondrial DCS protein became necessary to help resolve this issue.

There are questions that remain regarding the mammalian folate system, including whether the NMDMC is the only methyleneTHF dehydrogenase expressed in the mitochondria. Furthermore, is there a synthetase activity present in the mitochondria, which would allow for the production of formate and subsequent contribution to total cellular function? If there is no mitochondrial

146 synthetase activity, but mammalian mitochondria do produce formate, how is this done? ln order to address these questions, this thesis has dealt with the interplay between the mitochondrial and cytoplasmic folate pathways in mammalian cells, which has involved determining the role of NMDMC in contributing to the cytoplasmic folate pathway.

EVOLUTION OF NMDMC

It has been proposed that bacterial and mitochondrial OC enzymes

diverged from a common ancestral OC enzyme due to the observation that

bacterial sequences share a higher degree of similarity with the eukaryotic NAD­ dependent mitochondrial enzymes than with the OC domains from eukaryotic

NADP-dependent trifunctional enzymes (Pawelek and MacKenzie 1996).

However, recent sequence analyses have revealed regions with 50% identity

between sequences in the 3' untranslated region of the NMDMC cDNA and the synthetase region of the cytoplasmic NADP-dependent DCS cDNA and gene of the mouse, human and Drosophila (Chapter 2). It is even more significant that these regions even include intron sequences in the synthetase region of DCS in

both the mouse and human. This strongly suggests that the mitochondrial NAD­

dependent OC evolved trom a pre-existing NADP-dependent DCS precursor with

altered cofactor substitution from NADP to NAD and a requirement for

magnesium and inorganic phosphate (Chapter 2). This is supported by an earlier

147 observation that the human NMDMC has a 44% amine acid sequence identity with the OC domain of the yeast mitochondrial trifunctional DCS encoded by the

MIS1 gene (Yang and MacKenzie 1993). This observation made it even more

compelling to conclude that there is no trifunctional DCS in mammalian

mitochondria. 1 was able to pursue this further using mouse cell lines lacking

NMDMC activity.

MAMMALIAN MITOCHONDRIAL DEHYDROGENASES

It has been demonstrated in yeast that the folate compartments are

metabolically linked. It has been one of the objectives of this thesis to investigate

the contribution of mitochondria to cytoplasmic folate pathways. The initial

characterization of mou se fibroblasts lacking NMDMC activity revealed that these

cells do not have glycine cleavage in the mitochondria, which indicates that

glycine is not a major one-carbon donor. A series of growth experiments on

NMDMC null mutant fibroblasts established from NMDMC knockout mice have

demonstrated that these cells are glycine auxotrophs (Chapter 3). This had been

seen in a Chinese Hamster Ovary cell line, glyA, that lacks mitochondrial SHMT

activity (Chasin et al. 1974), which is the activity required to produce the

methyleneTHF substrate acted upon by the dehydrogenase. The glycine

auxotrophy of the NMDMC null mutant cells cannot be rescued with

hypoxanthine, which indicates that the mitochondrially-produced glycine is not

148 providing one-carbon units for purine biosynthesis. Also, Q-aminolevulinic acid does not even partially reverse the glycine auxotrophy, which indicates that the growth is not limited by heme synthesis (Chapter 3).

We have hypothesized that the glycine auxotrophy is due to an inability of the methyleneTHF to be metabolized in the mitochondria as a result of the inactivation of the NMOMC gene (Chapter 3). Consequently, the methyleneTHF accumulates and the lack of THF reduces the activity of the mitochondrial SHMT preventing glycine synthesis resulting in the observed glycine auxotrophy.

The growth studies also demonstrate a 2-fold stimulatory effect on the growth of the NMDMC null mutant cells with the addition of formate, which is not seen in the wild type cells (Chapter 3). The addition of hypoxanthine also produces this stimulatory effect, which suggests that formate is supplying one­ carbon units for purine biosynthesis. The addition of thymidine did not have any effect on these null mutant cells, which indicates that formate is not providing one­ carbon units for thymidine synthesis. These results were strengthened by a series of radiolabeling experiments that demonstrated that there is a 3 t010-fold enhancement in the incorporation of C4C]formate as compared to [3-14C]serine into total DNA of NMDMC null mutant cells (Chapter 3). In addition, the incorporation of [14C]formate is more significant into the adenine and guanine bases rather than thymidine in a single experiment. Therefore, it appears that in the NMDMC null mutant cells, the block in the mitochondrial folate pathway prevents the conversion of serine to formate, which would normally exit the

149 mitochondria and contribute to purine biosynthesis in the cytoplasm.

Although my studies on the NMDMC null mutant ce Il lines have demonstrated that the mitochondria of mammalian cells likely generate formate to support purine biosynthesis in the cytoplasm, a possible candidate enzyme to generate mitochondrial formate has remained elusive. We have attempted to search for a hydrolase activity in the mitochondria, which would hydrolyse 10- formylTHF to formate but we have not been successful. Recently, a human mitochondrial DCS cDNA was cloned (Prasannan et al. 2003), which the authors propose generates formate. However, their expression of this cDNA in yeast failed to demonstrate any dehydrogenase activity and they did not assay for cyclohydrolase activity (Prasannan et al. 2003). Despite their claims that this novel cDNA encodes for mitochondrial trifunctional DCS protein, it is likely a monofunctional synthetase protein. This is also supported by another study

(Sugiura et al. 2004), which was unsuccessful in their attempt to express the OC domain of this putative mitochondrial DCS protein in Escherichia coli and baculovirus expression systems. In addition, there is a missing stretch of 12 amine acids, which the first group (Prasannan et al. 2003) states is located near the linker region connecting the OC and S domains. However, according to the crystal structure of the DC301 protein (Allaire et al. 1998), this deletion is actually located within the OC domain where it encodes a portion of an important

Rossman fold of the nucleotide binding site.

Furthermore, even if the protein was a trifunctional DCS, we would still

150 question its expression in our NMDMC null mutant cell lines because the glycine auxotrophy would not manifest if there was another methyleneTHF dehydrogenase present in the mitochondria to metabolize the trapped methyleneTHF. Therefore, we decided to test our hypothesis that this was actuallya monofunctional synthetase. 1 have established mouse fibroblasts from embryonic stem cells in which the cytoplasmic DCS was inactivated by homologous recombination (Christensen, Patel et al. unpublished results). We assayed these cell lines for any remaining dehydrogenase and synthetase activities, which would be attributed to the proposed mitochondrial DCS protein.

Although we did not detect any dehydrogenase activity, we did detect synthetase activity (Christensen, Patel et al. unpublished results).

We propose that this newly isolated cDNA encodes a monofunctional synthetase protein. In addition, we propose that this protein may fit in our mammalian model for folate metabolism in that it is able to generate formate in the mitochondria. If so, the mammalian folate system may have evolved to substitute an NADP-dependent DCS protein in the mitochondria with two separate proteins: an NAD-dependent OC and a monofunctional synthetase. This would be to ensure that the mitochondrial folate pathway would be poised to proceed in the forward direction as a result of the use of the NAD cofactor. This would also allow for the production of formate in the mitochondria, which could contribute to purine synthesis in the cytoplasm during periods of rapid growth. Perhaps there is an instability associated with an NAD-dependent DCS protein, as no su ch

151 protein is known to exist in nature. However, the existence of a bifunctional OC protein separate from a monofunctional synthetase protein is documented for example in spinach leaves (Nour and Rabionowitz 1991).

It is more likely that the mammalian folate system evolved to encode the

NAD-dependent OC and synthetase activities by two separate proteins so as to

be able to allow differential expression patterns of these two proteins. This is

supported by a recent study do ne on the putative human mitochondrial DCS

(Sugiura et al. 2004), which demonstrates that there is ubiquitous expression of

this protein in normal tissues in addition to its upregulated expression in human

colon adenocarcinoma. This is in contrast to the expression profile of the

NMDMC protein, which is restricted to mammalian embryonic and undifferentiated

tissues as weil as transformed cell lines (Mejia and MacKenzie 1985, Mejia et al.

1986) whereas the mRNA is barely detectable and no activity is found in normal

adult differentiated cells and tissues.

METABOLIC ROLE OF NMDMC

The NMDMC protein is required to optimize purine biosynthesis in the

cytoplasm. This is particularly true during periods of rapid growth and

development such as embryogenesis and tumorigenesis, when accelerated

purine synthesis is supported by the generation of formate in the mitochondria,

which exits into the cytoplasm and serves as a one-carbon donor. This is

152 consistent with its expression in embryonic, undifferentiated and transformed cells as weil as in only the developmental cells of normal adult tissues (Mejia and

MacKenzie 1985). This implies that normal cells and rapidly proliferating cells possess alternative pathways to produce one-carbon units for processes such as purine biosynthesis.

This raises the question as to how normal adult tissues or even undifferentiated ceillines generate one-carbon units for purine synthesis. To date ail the studies of folate metabolism on mammalian cells have been performed on transformed ceillines, such as the MCF-7 ceillines (Fu et al. 2001, Oppenheim et al. 2001, Herbig et al. 2002) and in our case, transformed mouse fibroblasts

(Chapter 3). It is difficult to perform metabolic studies on primary fibroblasts, for instance, because they do not survive for many passages in culture. It is possible that in adult tissues, purine synthesis may use one-carbon units from the cytoplasmic DCS, which would be required to proceed in the forward direction.

ANTIFOLATES

Antifolates are a category of drugs that compete with the binding of natural folate cofactors to various important biosynthetic enzymes that are required for nucleotide synthesis for DNA and RNA, such as thymidylate synthase, glycinamide ribonucleotide formyltransferase, aminoimidazole carboxamide ribonucleotide formyltransferase and dihydrofolate reductase (Curtin et al. 2001).

153 The inhibition of these enzymes will affect the growth of rapidly proliferating cells, including cancer cells as they need a constant supply of nucleotides. There are many antifolates that have been developed, including aminopterin and methotrexate, which are OH FR inhibitors as weil as 5-fluorouracil and raltitrexed, which are T8 inhibitors.

A recently developed drug, 5,1 O-dideazatetrahydrofolate (OOATHF), which is clinically known as sodium lometrexol is a reversible competitive inhibitor of

GAR transformylase in the purine synthetic pathway (8anghani and Moran 1997).

According to our metabolic studies of the NMOMC null mutant ceillines, NMOMC appears to be generating one-carbon units for purine biosynthesis in the cytoplasm. We hypothesized that the inactivation of NMOMC should increase the efficacy of sodium lometrexol as an anti-cancer agent by reducing the mitochondrial pool of 10-formylTHF that can contribute to purine biosynthesis.

However, when 1 performed a toxicity curve, we were quite surprised to observe that there was no difference in the ability of sodium lometrexol to kill NMOMC null mutant and wild type cells. Perhaps these NMOMC null fibroblasts can compensate for the block in the mitochondrial generation of formate by obtaining one-carbon units from the cytoplasmic pool of 10-formylTHF to support purine biosynthesis. It is also possible that the mitochondrial pool of 10-formylTHF is not significantly reduced in the NMOMC null fibroblasts in order to enhance the toxicity of sodium lometrexol.

154 REPLACEMENT OF DEHYDROGENASE-CYCLOHYDROLASE ACTIVITIES IN MITOCHONDRIA OF MAMMALIAN CELLS

The rescue of the NMDMC null mutant cell lines with the NMDMC cDNA proves that the glycine auxotrophy is a result of the inactivation of the Nmdmc gene and not some other adjacent gene (Chapter 4). A reversai of the glycine auxotrophy was also seen with the rescue of the null cells with the monofunctional

NAD-dependent dehydrogenase although their growth was poorer as compared to the NMDMC-rescued cells (Chapter 4). This indicates that although the cyclohydrolase activity is not essential in the mitochondria, it is required to ensure optimal production of 10-formylTHF in this compartment. Furthermore, the rescue of the null cells with an NADP-dependent DCS protein targeted to the mitochondria also alleviated the glycine auxotrophy (Chapter 4). This indicates that the NAD cofactor specificity is not absolutely required to maintain the flow of one-carbon metabolites towards 10-formylTHF in the mitochondria, which would allow the generation of formate in this compartment.

The attempt to substitute the NMDMC with an NADP-dependent DC301 protein was not successful because the protein was degraded as determined by

Western analysis (Chapter 4). The NMDMC protein may be stable in mammalian mitochondria because it has an extension at the carboxyl terminus that is not present in the cytoplasmic NADP-dependent DC domain.

155 PREFERENCE FOR FORMATE AS ONE-CARBON DON OR

Another method of characterization of the various rescued NMDMC null mutant fibroblasts involved the determination of the ratio of the incorporation of

C4C]formate as compared to [3-14C]serine into total DNA. It was established that this ratio is 0.3 for the wild type cells (Chapter 3). However, in the null mutant cells, this ratio is approximately 2.5 because these cells preferentially incorporate exogenously supplied formate (Chapter 3). Consequently, this assay represents a qualitative indicator of how efficiently the mitochondria are producing formate and contributing to cellular function. Furthermore, since it can demonstrate up to a 10-fold change between the wild type and null mutant cells, it represents a more sensitive assay than the stimulation of the growth of the null mutant cells with either formate or hypoxanthine, which only demonstrates a 2-fold change. The only limitation of this radiolabeling assay is that it only allows comparisons within the same cell line. It was seen in ail of the rescued cell lines that this ratio varied between the wild type condition and the null condition but it never attained the wild type condition (Chapter 4). Furthermore, a decrease of this ratio among the rescued cell lines correlated with their ability to grow in glycine-free medium.

FUTURE DIRECTIONS

ln this thesis, 1 have studied the role of mitochondrial folate metabolism.

To further elucidate the nature of mammalian folate metabolism, it will be

156 necessary to knockout the cytoplasmic DCS gene and rescue with one or more of the three activities of its trifunctional protein. As 1 mentioned previously, 1 have already established mouse fibroblasts from embryonic stem cells in which DCS has been inactivated in both alleles. These mutant cells were determined to be purine auxotrophs (Christensen, Patel et al. unpublished results) as is expected and has already been seen with ADE3 mutant cells in yeast (Barlowe and Appling

Mol. Cell Biol. 1990). A key experiment that will confirm the contribution of the mitochondrial folate pathway to purine synthesis consists of rescuing the DCS null mutant cells with only the synthetase activity and demonstrating that [3-14C]serine is incorporated into purines. The lack of the cytoplasmic NADP-dependent dehydrogenase and cyclohydrolase activities would only leave the mitochondrial pathway to contribute one-carbon units for purine synthesis. Furthermore, the rescue of the DCS null mutant cells with both the dehydrogenase and synthetase activities will better define the role of the cyclohydrolase activity. This activity is required for the 100% channeling of formylTHF to methyleneTHF in the cytoplasm

(Pawelek and MacKenzie 1998) and we would predict that in its absence, no one­ carbon units from formate could flow to thymidylate or methionine. Such studies would close the circle on the proposai that mitochondria generate one-carbon units for purine synthesis and help to understand if these units also contribute to the methylation part of the folate pathway.

157 REFERENCES

Allaire, M., Li, Y., MacKenzie, RE. and Cygler, M. (1998) Structure 6,173-182

Anderson, RG.W., Kamen, B.A., Rothberg, K.G. and Lacey, S.W. (1992) Science

255, 410-411

Anguera, M.C., Suh, J.R, Gahndour, H., Nasrallah, LM., Selhub, J. and Stover,

P.J. (2003) J. Biol. Chem. 278, 29856-29862

Antonicka, H., MaUman, A., Carlson, C.G., Glerum, D.M., Hoffbuhr, K.C., Leary,

S.C., Kennaway, N.G. and Shoubridge, E.A. (2003) Am. J. Hum. Genet. 72, 101-

114

Antony, A.C. (1996) Annu. Rev. Nutr. 16,501-521

Appling, D.R (1991) FASEB J. 5, 2645-2651

Ashfield-Watt, P.A.L., Moat, S.J., Doshi, S.N. and McDowell, LF.W. (2001)

Biomed. Pharmacother. 55,425-433

158 Atkinson, 1., Garrow, T., Brenner, A. and Shane, B. (1997) Methods Enzymol.

281, 134-140

Ausubel, F.M. (1997) in Current Protocols in Molecular Biology Vol. 1, pp. 8.01-

8.5.10, John Wiley and Sons, New York

Banerjee RV. and MaUhews RG. (1990) FASEB J. 4,1450-1459

Barber, RC., Bennett, G.O., Greer, K.A. and Finnell, RH. (1999) Mol. Genet. And

Metabolism 66,31-39

Barlowe, C.K. and Appling, O.R (1988) Biofactors 1, 171-176

Barlowe, C.K. and Appling, O.R (1990) Biochemistry 29,7089-7094

Barlowe, C.K. and Appling, O.R (1990) Mol. Ce" Biol. 10,5679-5687

Belanger, C. and MacKenzie, RE. (1989) J. Biol. Chem. 264,4837-4843

Belanger, C., Peri, K.G. and MacKenzie, RE. (1991) Nucleic Acids Research 19,

4341-4345

159 Bertrand, R, Beauchemin, M., Dayan, A., Ouimet, M. and Jolivet, J. (1995)

Bioehim. Biophys. Acta 1266, 245-249

Bertrand, R and Jolivet, J. (1989) J. Biol. Chem. 264, 8843-8846

Birn, H., Selhub, J. and Christensen, E.1. (1993) Am. J. Physiol. 264, C302-C310

Botto, L.O. and Yang, Q. (2000) Am. J. Epidemiol.151, 862-877

Brigle, K.E., Spinella, M.J., Sierra, E.E. and Goldman, 1.0. (1995) J. Biol. Chem.

270, 22974-22979

Brigle, K.E., Westin, E.H., Houghton, M.T. and Goldman, 1.0. (1991) J. Biol.

Chem. 266, 17243-17249

Buchet, K. and Godinot, C. (1998) J. Biol. Chem. 273,22983-22989

Champion, K.M., Cook, RJ., Tollaksen, S.L. and Giometti, C.S. (1994) Proe. Natl.

Aead. Sei. USA 91, 11338-11342

Chasin, L.A., Feldman, A., Konstam, M. and Urlaub, G. (1974) Proe. Nat!. Aead.

Sei. USA 71,718-722

160 Chen, L., Qi, H., Korenberg, J., Garrow, T.A., Choi, Y.-J. and Shane, B. (1996) J.

Biol. Chem. 271, 13077-13087

Chen, Z., Karaplis, A.C., Ackerman, S.L., Pogribny, LP., Melnyk, S., Lussier­

Cacan, S., Chen, M.F., Pai, A., John, S.W.M., Smith, RS., Bottiglieri, T., Bagley,

P., Selhub, J., Rudnicki, M.A., James, S.J. and Rozen, R (2001) Hum. Mol.

Genet. 10, 433-443

Chiao, J.H., Roy, K., Tolner, B., Yang, C.-H. and Sirotnak, F.M. (1997) J. Biol.

Chem. 272,11165-11170

Choi, S.-W. and Mason, J.B. (2000) J. Nutr. 130, 129-132

Choi, S.-W. and Mason, J.B. (2002) J. Nutr. 132, 2413S-2418S

Cook, RJ. (2001) Arch. Biochem. Biophys. 392,226-232

Cook, J.O., Cichowicz, D.J., George, S., Lawler, A. and Shane, B. (1987)

Biochemistry 26, 530-539

Cybulski, RL. and Fisher, RR (1976) Biochemistry 15,3183-3187

161 Cybulski, R.L. and Fisher, R.R. (1981) Biochim. Biophys. Acta. 646, 329-333

Czeizel, A.E. and Dudas, 1. (1992) N. Engl. J. Med. 327,1832-1835

De Bree, A., Verschuren, W.M.M., Kromhout, D., Kluijtmans, L.A.J. and Blom,

H.J. (2002) Pharmacol. Rev. 54,599-618

Di Pietro, E., Sirois, J., Tremblay, M.L. and MacKenzie, R.E. (2002) Mol. Cel/.

Biol. 22, 4158-4166

Dixon, K.H., Lampher, B.C., Chiu, J., Kelley, K. and Cowan, K.H. (1994) J. Biol.

Chem. 269, 17-20

Dolk, H., Dewals, P., Gillerot, Y., Lechat, M.F., Ayme, S., Cornel, M., Cuschieri,

A., Garne, E., Goujard, J., Laurence, K.M., Ullis, D. and Lys, F. (1991) Teratology

44,547-559

Drury, E.J., Bazar, L.S. and MacKenzie, R.E. (1975) Arch. Biochem. Biophys.

169,662-668

Elwood, P.C. (1989) J. Biol. Chem. 264, 14893-14901

162 Finkelstein, J.o. and Martin, J.J. (1984) J. Biol. Chem. 259,9508-9513

Fishman, S.M., Christian, P. and West, K.P. (2000) Public Health Nutr. 3,125-150

Freemantle, S.J., Taylor, S.M., Krystal, G. and Moran, RG. (1995) J. Biol. Chem.

270, 9579-9584

Fu, T.-F., Rite, J.P. and Schirch, V. (2001) Arch. Bioehem. Biophys. 393,42-50

Garcfa-Martfnez, L.F. and Appling, D.R (1993) Bioehemistry 32,4671-4676

Garrow, T.A., Admon, A. and Shane, B. (1992) Proe. Nat!. Aead. Sei. USA 89,

9151-9155

Garrow, T.A., Brenner, A.A., Whitehead, V.M., Chen, X.-N., Duncan, RG.,

Korenberg, J.R and Shane, B. (1993) J. Biol. Chem. 268,11910-11916

Giometti, C.S., Tollaksen, S.L. and Grahn, D. (1994) Mutat. Res. 320,75-85

Girgis, S., Suh, J.R, Jolivet, J. and Stover, P.J. (1997) J. Biol. Chem. 272,4729-

4734

163 Girgis, S., Nasrallah, I.M., Suh, J.R., Oppenheim, E., Zanetti, K.A., Mastri, M.G. and Stover, P.J. (1998) Gene 210,315-324

Goldman, 1 (1971) Biochim. Biophys. Acta. 233, 624-634

Green, J.M., MacKenzie, R.E. and Matthews, R.G. (1988) Biochemistry 27,8014-

8022

Green, J.M., Ballou, D.P. and Matthews, R.G. (1988) FASEB J. 2,42-47

Gregory III, J.F., Cuskelly, G.J., Shane, B., Toth, J.P., Baumgartner, T.G. and

Stacpoole, P.W. (2000) Am. J. Clin. Nutr. 72, 1535-1541

Henderson, G.B. (1990) Annu. Nutr. Rev. 10,319-335

Herbig, K., Chiang, E.-P., Lee, L.-R., Hills, J., Shane, B. and Stover, P.J. (2002) J.

Biol. Chem. 277, 38381-38389

Herzberg, N.H., Middelkoop, E., Adorf, M., Dekker, H.L., Van Galen, M.J., Van den Berg, M., Bolhuis, P.A. and Van den Bogert, C. (1993) Eur. J. Cell. Biol. 61,

400-408

164 Hoffman, RM. (1982) ln Vitro 18, 421-428

Holmes, W.B. and Appling, D.R (2002) J. Biol. Chem. 277,20205-20213

Horne, D.W., Patterson, D. and Cook, RJ. (1989) Arch. Biochem. Biophys. 270,

729-733

Hum, D.W., Bell, A.W., Rozen, R and MacKenzie, RE. (1988) J. Biol. Chem.

263, 15946-15950

Hum, D.W. and MacKenzie, RE. (1991) Prot. Eng. 4,493-500

Italiano, C., John, S.W.M., Hum, D.W., MacKenzie, RE. and Rozen, R (1991)

Genomics, 10, 1073-1074

James, S.J., Pogribna, M., Pogribny, I,P., Melnyk, S., Hine, RJ., Gibson, J.B., Yi,

P., Tafoya, D.L., Swenson, D.H., Wilson, V.L. and Gaylor, D.W. (1999) Am. J.

Clin. Nutr. 70, 495-501

Kastanos, E.K., Woldman, Y.Y. and Appling, D.R (1997) Biochemistry 36,14956-

14964

165 Kay, L.O., Osborn, M.J., Hatefi, Y. and Huennekens, F.M. (1960) J. Biol. Chem.

235, 195-201

Kim, J.S. and Shane B. (1994) J. Biol. Chem. 269,9714-9720

Kim, Y.1. (1999) J. Nutr. Biochem. 10,66-88

King, M.P. and Attardi, G. (1989) Science 246,500-503

Laemmli, U.K. (1970) Nature 227, 680-685

Li, Y., Holmes, W.B., Appling, O.R and RajBhandary, U.L. (2000) J. Bacteriol.

182, 2886-2892

Lin, B.-F., Huang, R-F.S. and Shane, B. (1993) J. Biol. Chem. 268,21674-21679

Lin. B.-F. and Shane, B. (1994) J. Biol. Chem. 269,9705-9713

Lowe, K.E., Osborne, C.B., Lin, B.-F., Kim, J.-S., Hsu, J.-C. and Shane, B. (1993)

J. Biol. Chem. 268, 21665-21673

Lu, Y.-Z., Aiello, P.O. and Matthews, RG. (1984) Biochemistry 23,6870-6876

166 Lucock, M. (2000) Mol. Genetics and Metabolism 71, 121-138

Luka, Z., Cerone, R, Phillips III, J.A., Mudd, S.H. and Wagner, C. (2002) Hum

Genet. 110,68-74

MacKenzie, RE. (1984) in Folates and Pterins. Chemistry and Biochemistry of

Folates (Blakley, RL. and Benkovic, S.J., eds) Vol. 1, pp. 255-306, Wiley

Interscience, New York

Mascisch, A. and Rozen, R (1991) Soma tic Cell Mol. Genet. 17,391-398

MaUhews RG. and Baugh, C.M. (1980) Biochemistry 19,2040-2045

MaUhews, RG., Ross, J., Baugh, C.M., Cook, J.o. and Davis, L. (1982)

Biochemistry 21, 1230-1238

MaUson, M.P. and Shea, T.B. (2003) Trends in Neurosciences 26,137-146

Mayor, S., Rothberg, K.G. and Maxfield, F.R (1994) Science 264,1948-1951

McBurney, M.W. and Whitmore, G.F. (1974) Ce1l2, 173-182

167 McCully, K.S. (1969) Am. J. Pathol. 56, 111-128

McKinley, M.C. (2000) Proe. Nutr. Soc. 59,221-237

McNeil, J.B., Mclntosh, E.M., Taylor, B.V., Zhang, F.-R, Tang, S. and Bognar,

AL. (1994) J. Biol. Chem. 269,9155-9165

McNeil, J.B., Bognar, AL. and Pearlman, RE. (1996) Geneties 142,371-381

Mejia, N.R and MacKenzie, RE. (1985) J. Biol. Chem. 260, 14616-14620

Mejia, N.R, Rios-Orlandi, E.M. and MacKenzie, RE. (1986) J. Biol. Chem. 261,

9509-9513

Mejia, N.R and MacKenzie, RE. (1988) Bioehem. Biophys. Res. Commun. 155,

1-6

Miller, J.W. (2002) Nutr. Rev. 60,410-413

Minamikawa, T., Sriratana, A, Williams, D.A., Bowser, D.N., Hills, J.S. and

Nagley, P. (1999) J. Cell. Sei. 112,2419-2430

168 Narkewicz, M.R., Sauls, S.o., Tjoa, S.S., Teng, C. and Fennessey, P.v. (1996)

Bioehem. J. 313, 991-996

Newton, D.T., Creuzenet, C. and Mangroo, D. (1999) J. Biol. Chem. 274,22143-

22146

Neymeyer, V.R. and Tephly, T.R. (1994) Lite Sei. 54,395-399

Nour, J.M. and Rabinowitz, J.C. (1991) J. Biol. Chem. 266, 18363-18369

Oppenheim, E.W., Adelman, C., Liu, X. and Stover, P.J. (2001) J. Biol. Chem.

276, 19855-19861

Park, E.1. and Garrow, T.A. (1999) J. Biol. Chem. 274, 7816-7824

Pasternack, L.B., Laude, D.A., Jr. and Appling, D.R. (1992) Biochemistry 31,

8713-8719

Pasternack, L.B., Laude, D.A., Jr. and Appling, D.R. (1994) Bioehemistry 33,74-

82

169 Patel, H., Christensen, K.E., Mejia, N.R. and MacKenzie, R.E. (2002) Arch.

Biochem. Biophys. 403, 145-148

Patel, H., Di Pietro, E. and MacKenzie, R.E. (2003) J. Biol. Chem. 278, 19436-

19441

Pawelek, P.O. and MacKenzie, RE. (1996) Biochim. Biophys. Acta. 1296,47-54

Pawelek, P.o. and MacKenzie, RE. (1998) Biochemistry 37, 1109-1115

Pawelek, p.o., Allaire, M., Cygler, M. and MacKenzie, R.E. (2000) Biochim.

Biophys. Acta. 1479, 59-68

Pelletier, J.N. and MacKenzie, RE. (1994) Biochemistry 33,1900-1906

Pelletier, J.N. and MacKenzie, R.E. (1995) Biochemistry 34,12673-12680

Peri, K.G. and MacKenzie, RE. (1991) FEBS Lett. 294,113-115

Pfendner, W. and Pizer, L.1. (1980) Arch. Biochem. Biophys. 200,503-512

Piedrahita, J.A., Oetama, B., Bennett, GD., Van Waes, J., Kamen, B.A.,

170 Richardson, J., Lacey, S.W., Anderson, RG.W. and Finnell, RH. (1999) Nat.

Genet. 23, 228-232

Piper, M.D., Hong, S.-P., Bali, G.E. and Dawes, I.W. (2000) J. Biol. Chem. 275,

30987 -30995

Prasannan, P., Pike, S., Peng, K., Shane, B. and Appling, D.R (2003) J. Biol.

Chem. 278,43178-43187

Quinlivan, E.P. and Gregory, J.F. (2003) Am. J. Clin. Nutr. 77,221-225

Ratnam, M., Marquardt, H., Duhring, J.L. and Freisheim, J.H. (1989) Biochemistry

28, 8249-8254

Refsum, H. (2001) British. J. Nutr. 85, S109-S113

Rios-Orlandi, E.M. and MacKenzie, RE. (1988) J. Biol. Chem. 263,4662-4667

Rosenquist, T.H. and Finnell, RH. (2001) Proc. Nat!. Soc. 60, 53-61

Ross, J., Green, J., Baugh, C.M., MacKenzie, RE. and Matthews, RG. (1984)

Biochemistry 23, 1796-1801

171 Ross, J.F., Chaudhuri, P.K. and Ratnam, M. (1994) Cancer Res. 73,2432-2443

Rothberg, K.G., Ying, Y., Kolhouse, J.F., Kamen, B.A. and Anderson, RG.W.

(1990) J. Ce". Biol. 110,637-649

Rozen, R, Barton, O., Du, J., Hum, D.W., MacKenzie, RE. and Francke, U.

(1989) Am.J. Hum. Genet. 44, 781-786

Ruffolo, S.C., Breckenridge, D.G., Nguyen, M., Goping, I.S., Gross, A.,

Korsmeyer, S.J., Li, H., Yuan, J. and Shore, G.C. (2000) Ce" Death Diff. 7, 1101-

1108

Sanghani, S.P. and Moran, RG. (1997) Biochemistry 36,10506-10516

Schalinske, K.L. and Steele, RD. (1996) Arch. Biochem. Biophys. 328, 93-100

Schirch, V. (1997) Methods Enzymol. 281,77-81

Schirch, V. and Strong, W.B. (1989) Arch. Biochem. Biophys. 269, 371-380

Schmidt, A., Wu, H. MacKenzie, RE., Chen, V.J., Bewly, J.R, Ray, J.E., Toth,

J.E. and Cygler, M. (2000) Biochemistry 39, 6325-6335

172 Scrimgeour, K.G. and Huennekens, F.M. (1960) Biochem. Biophys. Res.

Commun. 2, 230-233

Selhub, J. (1999) Annu. Rev. Nutr. 19,217-246

Selhub, J. and Miller, J.W. (1992) Am. J. Clin. Nutr. 55, 131-138

Shane, B. (1989) Folypolyglutamate synthesis and role in the regulation of one­ carbon metabolism. Vitamins and Hormones 45,263-335

Shane, B. and Stokstad, E.L. (1985) Annu. Rev. Nutr. 5, 115-141

Shannon, K.W. and Rabinowitz, J.C. (1986) J. Biol. Chem. 261, 12266-12271

Shannon, K.W. and Rabinowitz, J.C. (1988) J. Biol. Chem. 263, 7717-7725

Shen, F., Wu, M., Ross, J.F., Miller, D. and Ratnam, M (1995) Biochemistry 34,

5660-5665

Sierra, E.E. and Goldman, I.D. (1998) Biochem. Pharmacol. 55,1505-1512

173 Sies, H. (1982) in Metabolic Comparlmentation (Sies, H., Ed.) pp. 205-231,

Academie Press, London

Sirotnak, F.M. and Tolner, B. (1999) Annu. Rev. Nutr. 19,91-122

Smith, O.O.S. and MacKenzie, R.E. (1985) Biochem.Biophys. Res. Commun.

128,148-154

Smith, G.K., Banks, S.O., Monaco, T.J., Rigual, R, Ouch, O.S., Mullin, RJ. and

Huber, B.E. (1990) Arch. Biochem. Biophys. 283,367-371

Song, J., Kim, H., Hay, K. et al. (1999) Proc. Am. Assoc. Cancer Res. 40, 3490

Song, J.M. and Rabinowitz, J.C. (1993) Proc. Nat!. Acad. Sei. USA 90,2636-2640

Spiegelstein, O., Eudy, J.O. and Finnell, RH. (2000) Gene 258,117-125

Spinella, M.J., Brigle, K.E. Sierra, E.E. and Goldman, I.D. (1995) J. Biol. Chem.

270, 7842-7849

Steinberg, S.E. (1984) Am. J. Physiol. 246, G319-G324

174 Stover, P. and Schirch, V. (1990) J. Biol. Chem. 265,14227-14233

Stover, P.J., Chen, L.H., Suh, J.R, Stover, D.M., Keyomarsi, K. and Shane, B.

(1997) J. Biol. Chem. 272, 1842-1848

Sugiura, T., Nagano, Y., Inoue, T. and Hirotani, K. (2004) Biochem. Biophys. Res.

Comm. 315,204-211

Sundararajan, S. and MacKenzie, RE. (2002) J. Biol. Chem. 277,18703-18709

Sunden, S.L.F., Renduchintala, M.S., Park, E.I., Miklasz, S.D. and Garrow, T.A.

(1997) Arch. Biochem. Biophys. 345, 171-174

Swanson, D.A., Liu, M.-L., Baker, P.J., Garrett, L., Stitzel, M., Wu, J., Harris, M.,

Banerjee, R, Shane, B. and Brody, L.C. (2001) Mol. Ce". Biol. 21, 1058-1065

Tan, L.U.L. and MacKenzie, RE. (1977) Biochem. Biophys. Acta. 485, 52-59

Tan, L.U.L. and MacKenzie, RE. (1979) Cano J. Biochem. 57,806-812

Tang, B., Li, Y.N. and Kruger, W.D. (2000) Cancer Res. 60,5543-5547

175 Taylor, RT. and Hanna, M.L. (1982) Arch. Biochem. Biophys. 217,609-623

Thigpen, A.E., West, M.G. and Appling, D.R (1990) J. Biol. Chem. 265, 7907-

7913

Titus, S.A. and Moran, RG. (2000) J. Biol.Chem. 275,36811-36817

Towbin, H., Staehelin, T. and Gordon, J. (1979) Proc. Natl. Acad. Sci. USA 76,

4350-4354

Tremblay, G.B., Sohi, S.S., Retnakaran, A. and MacKenzie, RE. (1995) FEBS

368,177-182

Trent, D.F., Seither, RL. and Goldman, I.D. (1991) Biochem. Pharmacol. 42,

1015-1019; Correction (1981) Biochem. Pharmacol. 42, 2405

Trippett, T.M., Garcia, S., Manova, K., Mody, R, Cohen-Gould, L., Flintoff, W. and

Bertino, J.R (2001) Cancer Res. 61, 1941-1947

Van der Put, N.M.J., Van Straaten, H.W.M., Trijbels, F.J.M. and Blom, H.J. (2001)

Exp. Biol. Med. 226, 243-270

176 Verhaar, M.C., Stroes, E. and Rabelink, T.J. (2002) Arlerioseler. Thromb. Vase.

Biol. 22, 6-13

Watanabe, M., Osada, J., Aratani, Y., Kluckman, K., Reddick, R, Malinow, M.R and Maeda, N. (1995) Proe. Natl. Aead. Sei. USA 92,1585-1589

West, M.G., Barlowe, C.K. and Appling, O.R (1993) J. Biol. Chem. 268,153-160

West, M.G., Horne, O.W. and Appling, O.R (1996) Bioehemistry 35,3122-3132

Wittwer, A.J. and Wagner, C (1981) J. Biol. Chem. 256, 4102-4108

Yang, C.-H., Sirotnak, F.M. and Oembo, M. (1984) J. Membr. Biol. 79,285-292

Yang, X.-M. and MacKenzie, RE. (1993) Bio ch emistry 32, 11118-11123

Yeo, E.-J., Briggs, W.T. and Wagner, C. (1999) J. Biol. Chem. 274,37559-37564

Yoshida, T. and Kikuchi, G. (1973) J. Bioehem. 73, 1013-1022

Zhao, R, Russell, RG., Wang, Y., Liu, L., Gao, F., Kneitz, B., Edelmann, W. and

Goldman, 1.0. (2001) J. Biol. Chem. 276, 10224-10228

177