11.05.2013 Views

B12 METABOLISM IN HUMANS By NICOLE AURORA LEAL A ...

B12 METABOLISM IN HUMANS By NICOLE AURORA LEAL A ...

B12 METABOLISM IN HUMANS By NICOLE AURORA LEAL A ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>B12</strong> <strong>METABOLISM</strong> <strong>IN</strong> <strong>HUMANS</strong><br />

<strong>By</strong><br />

<strong>NICOLE</strong> <strong>AURORA</strong> <strong>LEAL</strong><br />

A DISSERTATION PRESENTED TO THE GRADUATE SCHOOL<br />

OF THE UNIVERSITY OF FLORIDA <strong>IN</strong> PARTIAL FULFILLMENT<br />

OF THE REQUIREMENTS FOR THE DEGREE OF<br />

DOCTOR OF PHILOSOPHY<br />

UNIVERSITY OF FLORIDA<br />

2004


Copyright 2004<br />

by<br />

Nicole Aurora Leal


The work in this dissertation is dedicated to my family, especially my parents, for<br />

their endless love and support. I would also like to dedicate this dissertation to my<br />

fiancé, Tearlach Bigger, whose love, patience, and encouragement guided me through<br />

this enlightening experience.


ACKNOWLEDGMENTS<br />

I would sincerely like to thank my mentor, Dr. Thomas A. Bobik. I find myself<br />

continually amazed by his keen intellectual insight and unceasing dedication to research<br />

and the furthering of science. I would like to thank the rest of my committee,<br />

Dr. Terence R. Flotte, Dr. Nemat O. Keyhani, Dr. Julie A. Maupin-Furlow, and<br />

Dr. Keelnatham T. Shanmugam. Their help and guidance along the way were greatly<br />

appreciated.<br />

I owe a debt of gratitude to Dr. Ruma Banerjee and Horatiu Olteanu, who provided<br />

purified human methionine synthase reductase, which was instrumental in the MSR-ATR<br />

studies. Special thanks go to Dr. Peter E. Kima for his constant optimism and<br />

much-needed expertise in the area of cell biology. Additional thanks go to the entire<br />

faculty, staff, and graduate students, especially Stephanie Havemann, at the Microbiology<br />

and Cell Science Department.<br />

I would also like to extend my appreciation to Dr. Gregory Havemann, Celeste<br />

Johnson, Dorothy Park, Edith Sampson, and all the other members of the Bobik lab, for<br />

making it a friendly and supportive place to work.<br />

The text of Chapter 2 in this dissertation, in part or in full, is a reprint of the<br />

material as it appears in the Journal of Biological Chemistry (volume 278, pp. 9227-9234,<br />

2003). The text of Chapter 4 in this dissertation, in part or in full, is a reprint of the<br />

material as it appears in Archives of Microbiology (volume 180, pp. 353-361, 2003).<br />

iv


TABLE OF CONTENTS<br />

Page<br />

ACKNOWLEDGMENTS ................................................................................................. iv<br />

LIST OF TABLES............................................................................................................. ix<br />

LIST OF FIGURES .............................................................................................................x<br />

LIST OF ABBREVIATIONS........................................................................................... xii<br />

ABSTRACT.......................................................................................................................xv<br />

CHAPTER<br />

1 <strong>IN</strong>TRODUCTION ........................................................................................................1<br />

History of Cobalamin ...................................................................................................1<br />

Structure of Cobalamin.................................................................................................2<br />

Biosynthesis of Cobalamin...........................................................................................3<br />

Cobalamin-Dependent Enzymes ..................................................................................4<br />

Cobalamin in Humans ..................................................................................................6<br />

Cobalamin-Dependent Enzymes in Humans.........................................................6<br />

Methylmalonyl-CoA mutase..........................................................................7<br />

Methionine synthase.......................................................................................9<br />

Cobalamin Absorption and Transport .................................................................11<br />

Intracellular Cobalamin Metabolism...................................................................12<br />

β-ligand transferase ......................................................................................12<br />

Cob(III)alamin reductase..............................................................................13<br />

Cob(II)alamin reductase...............................................................................14<br />

Adenosyltransferase .....................................................................................15<br />

Formation of methylcobalamin ....................................................................17<br />

Inherited Disorders of Cobalamin Metabolism ..........................................................18<br />

Combined Methylmalonic Aciduria and Homocystinuria ..................................19<br />

Methylmalonic Aciduria without Homocystinuria..............................................21<br />

Homocystinuria without Methylmalonic Aciduria..............................................23<br />

Salmonella as a Model Organism for the Study of Cobalamin Metabolism ..............25<br />

Research Overview.....................................................................................................26<br />

v


2 IDENTIFICATION OF THE HUMAN AND BOV<strong>IN</strong>E ATP:COB(I)ALAM<strong>IN</strong><br />

ADENOSYLTRANSFERASE CDNAS BASED ON COMPLEMENTATION OF<br />

A BACTERIAL MUTANT........................................................................................36<br />

Introduction.................................................................................................................36<br />

Materials and Methods ...............................................................................................38<br />

Chemicals and Reagents......................................................................................38<br />

Bacterial Strains and Growth Media ...................................................................39<br />

General Protein and Molecular Methods and ATR Assays.................................39<br />

P22 Transductions ...............................................................................................39<br />

Screening the Bovine and Human Liver cDNA Libraries...................................40<br />

Cloning the Bovine Adenosyltransferase Coding Sequence<br />

for High-Level Expression...............................................................................40<br />

Cloning the Human Adenosyltransferase Coding Sequence<br />

for High-Level Expression...............................................................................41<br />

Cloning the Human Adenosyltransferase Coding Sequence<br />

for Complementation Studies ..........................................................................42<br />

Growth of Adenosyltransferase Expression Strains and Preparation<br />

of Cell Extracts ................................................................................................43<br />

Growth Curves.....................................................................................................43<br />

Western Blots ......................................................................................................44<br />

DNA Sequencing and Analysis...........................................................................45<br />

Results.........................................................................................................................45<br />

Screening of the Bovine and Human Liver cDNA Libraries for Clones that<br />

Express Adenosyltransferase Activity .............................................................45<br />

Identification of a Human Gene Related to the Bovine Adenosyltransferase<br />

cDNA ...............................................................................................................47<br />

Putative Subcellular Localization of Adenosyltransferase Enzymes ..................48<br />

High-Level Expression of the Bovine and Human Adenosyltransferases<br />

in E. coli...........................................................................................................48<br />

Assay of the Bovine and Human Adenosyltransferase Fusion Proteins<br />

for ATR Activity..............................................................................................49<br />

Complementation of an S. enterica Mutant Deficient in ATR Activity by a<br />

Human Adenosyltransferase cDNA Clone ......................................................50<br />

Level of Human Adenosyltransferase Enzyme in Normal and cblB Mutant<br />

Human Skin Fibroblasts...................................................................................51<br />

Conserved Amino Acids and Distribution of Adenosyltransferase Enzymes.....52<br />

Discussion...................................................................................................................53<br />

3 PURIFICATION AND <strong>IN</strong>ITIAL CHARACTERIZATION OF THE HUMAN<br />

ATP:COB(I)ALAM<strong>IN</strong> ADENOSYLTRANSFERASE AND ITS <strong>IN</strong>TERACTION<br />

WITH METHION<strong>IN</strong>E SYNTHASE REDUCTASE .................................................62<br />

Introduction.................................................................................................................62<br />

Materials and Methods ...............................................................................................64<br />

Chemicals and Reagents......................................................................................64<br />

Bacterial Strains and Growth Media ...................................................................65<br />

vi


General Molecular and Protein Methods.............................................................65<br />

ATP:cob(I)alamin Adenosyltransferase Assays..................................................65<br />

Construction of ATR Expression Strains ............................................................66<br />

Purification of the Human ATR Variants............................................................66<br />

MSR-ATR Assay.................................................................................................68<br />

Measurement of Cob(I)alamin with Iodoacetate.................................................68<br />

Separation and Quantification of Cobalamins by HPLC ....................................68<br />

Effect of Ionic Strength on the MSR-ATR System.............................................69<br />

DNA Sequencing.................................................................................................69<br />

Results.........................................................................................................................70<br />

Purification of the Human ATR Variants............................................................70<br />

Linearity of the ATR reaction .............................................................................71<br />

ATR Reaction Requirements...............................................................................71<br />

Alternative Nucleotide Donors............................................................................71<br />

Km and Vmax Values for ATR 239K and 239M ...................................................71<br />

MSR Reduces Cob(II)alamin to Cob(I)alamin for AdoCbl Synthesis................72<br />

Cob(I)alamin is Sequestered by the MSR-ATR System .....................................73<br />

MSR Produces little Cob(I)alamin in the Absence of the ATR ..........................75<br />

Stoichiometry of the MSR-ATR system .............................................................75<br />

Ionic Strength Dependence of Cobalamin Reduction and Adenosylation<br />

by the MSR-ATR System ................................................................................75<br />

Discussion...................................................................................................................76<br />

4 PDUP IS A COENZYME-A-ACYLAT<strong>IN</strong>G PROPIONALDEHYDE<br />

DEHYDROGENASE ASSOCIATED WITH THE POLYHEDRAL BODIES<br />

<strong>IN</strong>VOLVED <strong>IN</strong> <strong>B12</strong>-DEPENDENT 1,2-PROPANEDIOL DEGRADATION BY<br />

SALMONELLA ENTERICA SEROVAR TYPHIMURIUM LT2...............................92<br />

Introduction.................................................................................................................92<br />

Materials and Methods ...............................................................................................95<br />

Bacterial Strains, Media, and Growth Conditions...............................................95<br />

General Molecular Methods ................................................................................95<br />

General Protein Methods.....................................................................................95<br />

P22 Transduction.................................................................................................96<br />

Cloning pduP into pLAC22.................................................................................96<br />

Construction of His8-PduP ..................................................................................97<br />

Purification of His8-PduP under Denaturing and Nondenaturing Conditions.....97<br />

Propionaldehyde Dehydrogenase Assays............................................................98<br />

Identification of Propionyl-CoA a Product of the Propionaldehyde<br />

Dehydrogenase Reaction .................................................................................99<br />

Construction of a Nonpolar pduP Deletion.......................................................100<br />

Antibody Preparation.........................................................................................101<br />

Western Blots ....................................................................................................101<br />

Electron Microscopy .........................................................................................102<br />

DNA Sequencing and Analysis.........................................................................102<br />

Chemicals and Reagents....................................................................................102<br />

Results.......................................................................................................................103<br />

vii


Effect of a Precise pduP Deletion on the Growth of S. enterica on Minimal<br />

1,2-Propanediol Medium ...............................................................................103<br />

Propionaldehyde Dehydrogenase Activity in Wild-type S. enterica<br />

and a ∆pduP Mutant.......................................................................................105<br />

High-level Production of the PduP Protein .......................................................105<br />

Purification of Recombinant His8-PduP Protein ...............................................106<br />

Propionyl-CoA is a Product of the PduP Reaction............................................107<br />

Preparation and Specificity of the Anti-PduP Antiserum..................................108<br />

Localization of PduP Immunoelectron Microscopy..........................................109<br />

Propionaldehyde Dehydrogenase Activity is associated<br />

with Purified pdu Bodies ...............................................................................110<br />

Discussion.................................................................................................................110<br />

5 CONCLUSIONS ......................................................................................................119<br />

Identification of the Bovine and Human Adenosyltransferase.................................119<br />

Biochemical Characterization of the Human Adenosyltransferase..........................120<br />

Methionine Synthase Reductase is a Cob(II)alamin Reductase<br />

for the Human Adenosyltransferase.....................................................................121<br />

Future Experimentation ............................................................................................123<br />

LIST OF REFERENCES.................................................................................................125<br />

BIOGRAPHICAL SKETCH ...........................................................................................139<br />

viii


LIST OF TABLES<br />

Table page<br />

1-1. Adenosylcobalamin- and methylcobalamin-dependent reactions .............................30<br />

2-1. Bacterial strains .........................................................................................................56<br />

2-2. Specific activities of bovine and human ATP:cob(I)alamin ATRs...........................59<br />

3-1. ATP:cob(I)alamin adenosyltransferase activity during ATR 239K and 239M<br />

purification ...............................................................................................................82<br />

3-2. Use of alternative nucleotide donors by the human ATR .........................................84<br />

3-3. The MSR-ATR system sequesters cob(I)alamin.......................................................85<br />

4-1. Bacterial strains .......................................................................................................113<br />

4-2. Propionaldehyde dehydrogenase activity in extracts from wild-type S. enterica<br />

and strain BE191 (∆pduP)......................................................................................114<br />

4-3. Propionaldehyde dehydrogenase activity during polyhedral body purification......115<br />

ix


LIST OF FIGURES<br />

Figure page<br />

1-1. Structure of cobalamins.............................................................................................29<br />

1-2. Different classes of cobalamin-dependent enzymes..................................................31<br />

1-3. Pathways of propionyl-CoA catabolism and cobalamin metabolism........................32<br />

1-4. Pathways involving methionine synthase and the reductive activation of<br />

methionine synthase by methionine synthase reductase ..........................................33<br />

1-5. Cobalamin metabolism in mammalian cells, and disorders associated with<br />

deficiencies in this pathway .....................................................................................34<br />

1-6. Proposed pathway of AdoCbl-dependent 1,2-propanediol degradation<br />

in S. enterica.............................................................................................................35<br />

2-1. Multiple sequence alignment of adenosyltransferase enzymes.................................57<br />

2-2. SDS-PAGE analysis of cell extracts from bovine and human ATR<br />

expression strains .....................................................................................................58<br />

2-3. Complementation of an ATR-deficient bacterial mutant for AdoCbl-dependent<br />

growth on 1,2-propanediol by a plasmid that expresses the human ATR................60<br />

2-4. Western blot analysis of ATR expression by normal and cblB mutant cell lines .....61<br />

3-1. Propionyl-CoA metabolism, methionine synthesis, and intracellular cobalamin<br />

metabolism ...............................................................................................................81<br />

3-2. Purification of the human ATR .................................................................................83<br />

3-3. Determination of kinetic constants for ATR 239K and ATR 239M .........................86<br />

3-4. Absorbance spectra of an MSR-ATR assay ..............................................................87<br />

3-5. Two schemes depicting the conversion of cob(II)alamin to AdoCbl by the<br />

MSR-ATR system....................................................................................................88<br />

3-6. MSR produces little cob(I)alamin in the absence of the ATR enzyme.....................89<br />

x


3-7. Stoichiometry of the MSR-ATR system ...................................................................90<br />

3-8. Effect of ionic strength on the MSR-ATR system ....................................................91<br />

4-1. SDS-PAGE analysis of His8-PduP ..........................................................................116<br />

4-2. Western blot using the anti-PduP antiserum............................................................117<br />

4-3. Immunogold localization of the PduP enzyme........................................................118<br />

xi


A adenine<br />

LIST OF ABBREVIATIONS<br />

ADP adenosine diphosphate<br />

AdoCbl adenosylcobalamin<br />

AIM aldehyde indicator medium<br />

Amp ampicillin<br />

ATP adenosine triphosphate<br />

ATR adenosyltransferase<br />

BSA bovine serum albumin<br />

C cytosine<br />

°C degree centigrade<br />

Cam chloramphenicol<br />

cbl cobalamin<br />

cDNA complementary deoxyribonucleic acid<br />

CH3Cbl methylcobalamin<br />

CH3THF methyl tetrahydrofolate<br />

CNCbl cyanocobalamin, Vitamin <strong>B12</strong><br />

CMCbl carboxymethylcobalamin<br />

ddH20 distilled deionized water<br />

DMB dimethylbenzimidazole<br />

DNA deoxyribonucleic acid<br />

xii


DTT dithiothreitol<br />

E. coli Escherichia coli<br />

FAD flavin adenine dinucleotide<br />

FldA flavodoxin<br />

FMN flavin mononucleotide<br />

Fre flavin oxidoreductase<br />

Fpr flavodoxin reductase<br />

G guanine<br />

GSCbl glutathionylcobalamin<br />

HOCbl hydroxycobalamin<br />

IF intrinsic factor<br />

IFCR intrinsic factor cobalamin receptor<br />

IPTG isopropyl-β-D-thiogalactopyranoside<br />

Kan kanamycin<br />

kDa kiloDaltons<br />

Km Michaelis Constant for enzyme activity<br />

LB Luria Bertani medium<br />

MCM methylmalonyl-CoA mutase<br />

MTS mitochondrial targeting sequence<br />

min minutes<br />

mg milligram<br />

mM millimolar<br />

MS methionine synthase<br />

xiii


MSR methionine synthase reductase<br />

NADPH nicotinamide adenine dinucleotide phosphate – reduced form<br />

NCE no carbon E medium<br />

OD optical density<br />

PAGE polyacrylamide gel electrophoresis<br />

PCR polymerase chain reaction<br />

SAM S-adenosylmethionine<br />

S. enterica Salmonella enterica<br />

SDS sodium dodecyl sulfate<br />

T thymine<br />

TCII transcobalamin II<br />

TCII-R transcobalamin II receptor<br />

TCA tricarboxylic acid<br />

THF tetrahydrofolate<br />

Tris tris hydroxymethyl aminomethane<br />

tRNA transfer ribonucleic acid<br />

U uracil<br />

V volt<br />

Vmax maximal rate of enzyme activity<br />

XCbl inactive cobalamin derivatives<br />

X-Gal 5-bromo-4-chloro-3-indolyl-β-D-galactopyranoside<br />

xiv


Abstract of Dissertation Presented to the Graduate School<br />

of the University of Florida in Partial Fulfillment of the<br />

Requirements for the Degree of Doctor of Philosophy<br />

<strong>B12</strong> <strong>METABOLISM</strong> <strong>IN</strong> <strong>HUMANS</strong><br />

<strong>By</strong><br />

Nicole Aurora Leal<br />

August 2004<br />

Chair: Thomas A. Bobik<br />

Major Department: Microbiology and Cell Science<br />

In humans, the <strong>B12</strong> coenzymes, adenosylcobalamin and methylcobalamin, are<br />

required cofactors for propionate metabolism and methionine biosynthesis, respectively.<br />

Humans are incapable of de novo synthesis of the <strong>B12</strong> coenzymes and require complex<br />

precursors such as vitamin <strong>B12</strong> in their diet. The metabolism of vitamin <strong>B12</strong> to<br />

adenosylcobalamin requires two successive reductions and an adenosylation reaction;<br />

however, many aspects of the genetics and biochemistry of this process have not been<br />

elucidated. This dissertation focuses on the isolation and characterization of the human<br />

genes and enzymes involved in the formation of adenosylcobalamin from vitamin <strong>B12</strong>.<br />

Deficiencies in this process result in methylmalonic aciduria, a rare disorder that is often<br />

fatal in newborns.<br />

Because of the sophisticated genetic methods that are available, Salmonella<br />

enterica was used as a model system. <strong>By</strong> complementation of an S. enterica mutant, a<br />

bovine adenosyltransferase cDNAs was isolated and subsequent sequence similarity<br />

xv


searches identified a homologous human cDNA. The bovine and human cDNAs were<br />

independently cloned, expressed in Escherichia coli, and the encoded proteins were<br />

shown to have adenosyltransferase activities similar to previously studied bacterial<br />

adenosyltransferases. Additional studies, using Western blots, showed that<br />

adenosyltransferase expression was altered in cell lines derived from patients with cblB<br />

methylmalonic aciduria when compared to cell lines from normal individuals.<br />

Two common human adenosyltransferase polymorphic variants were expressed in<br />

E. coli, purified, and biochemically characterized. Both had high specificity for ATP and<br />

comparable Km and Vmax values. An adenosyltransferase-linked cob(II)alamin reductase<br />

assay was developed and used to show that the human methionine synthase reductase<br />

functions as a cob(II)alamin reductase for adenosylcobalamin formation. A linked assay<br />

was also performed in the presence of a cob(I)alamin trapping agent, and results from<br />

these experiments suggested that human adenosyltransferase and methionine synthase<br />

reductase interact. Importantly, results from this work provide information that will<br />

allow improvements in diagnosis and treatment of methylmalonic aciduria, a devastating<br />

childhood disorder.<br />

xvi


CHAPTER 1<br />

<strong>IN</strong>TRODUCTION<br />

History of Cobalamin<br />

In 1925, Whipple and colleagues discovered that a diet of liver aided in the<br />

formation of blood cells in iron-deficient dogs. The next year, Minot and Murphy (Minot<br />

and Murphy 1926) were the first to document that patients with severe pernicious anemia<br />

(a disease that affects the formation of red blood cells) were successfully treated with a<br />

special diet consisting mainly of liver. Eight years later (in 1934) Whipple, Minot, and<br />

Murphy were awarded the Nobel Prize in Medicine and Physiology for their work on<br />

liver therapy against anemias.<br />

Following these studies, Castle (Kass 1978) concluded that the treatment of<br />

pernicious anemia by this specialized diet was dependent on the absorption of an<br />

extrinsic anti-pernicious anemia factor by an intrinsic protein in gastric juice. It is now<br />

known that Castle’s extrinsic anti-pernicious anemia factor is vitamin <strong>B12</strong> and the<br />

intrinsic factor is a specific cobalamin transporter (synthesized by the stomach) that was<br />

shown to enhance the curing effect by vitamin <strong>B12</strong> (Kass 1978). In 1948, two research<br />

groups led by Folkers and Smith independently isolated the red crystalline anti-pernicious<br />

anemia factor from bovine liver, and proposed the name vitamin <strong>B12</strong> (Rickes et al. 1948,<br />

Smith and Parker 1948). The isolation of vitamin <strong>B12</strong> inspired collaborative efforts<br />

between two laboratories (led by Alexander Todd and Dorothy Hodgkin) to elucidate its<br />

chemical structure. After 6 years, the crystal structure was resolved and finalized in the<br />

1


2<br />

laboratory of Dorothy Hodgkin (Hodgkin et al. 1955), a feat for which she was awarded<br />

the 1964 Nobel Prize in Chemistry.<br />

Today we know that two active cobalamin derivatives are required for enzyme<br />

catalysis. One is the adenosyl derivative (adenosylcobalamin) which was identified in<br />

1958, when Barker and coworkers reported the isolation of a light-sensitive cobalamin<br />

compound that was a required cofactor for glutamate mutase (Barker et al. 1958).<br />

Shortly after the discovery of this compound, its crystal structure was resolved by<br />

Lenhert and Hodgkin (Lenhert and Hodgkin 1961). The other biologically active<br />

cobalamin derivative was discovered later as methylcobalamin, a required cofactor for<br />

methionine synthase (Guest et al. 1962a, Guest et al. 1962b). The structure of<br />

methylcobalamin was resolved decades later by Jenny Gluskers laboratory (Rossi et al.<br />

1985).<br />

Structure of Cobalamin<br />

Cobalamin is the largest and most complex cofactor known to man. It is a<br />

water-soluble molecule composed of three general structural units: a central ring, an<br />

upper ligand, and a lower ligand (Figure 1-1). At the core of cobalamin is a ring structure<br />

designated as corrin, which is synthesized from uroporphyrinogen III, the last common<br />

intermediate of heme, siroheme, chlorophyll, and cobalamin synthesis. The corrin ring<br />

consists of four reduced pyrrole rings (A, B, C, and D). Unlike heme and chlorophyll (in<br />

which the pyrrole rings are linked by four methylene bridges), cobalamin has a direct<br />

bond between rings A and D. In cobalamin, the four nitrogen atoms of the pyrrole rings<br />

form coordinate bonds with a central cobalt atom; whereas the central atom of the<br />

tetrapyrrole ring is iron and magnesium, in heme and chlorophyll, respectively. Below<br />

the plane of the corrin ring is the α-ligand of the cobalt atom, dimethylbenzimidazole


3<br />

(DMB). The DMB moiety is covalently attached to the corrin ring by a ribose phosphate<br />

bonded to the aminopropanol side chain on ring D and is also coordinated to the central<br />

cobalt atom. The exact function of the lower ligand is not completely understood, but has<br />

been suggested that it may have a role in coenzyme binding and catalysis (Banerjee and<br />

Ragsdale 2003). Above the plane of the corrin ring is the β-ligand of the cobalt atom.<br />

The β-ligand varies in the different forms of cobalamins and includes cyano, hydroxy,<br />

glutathionyl, methyl, and adenosyl moieties. Cyanocobalamin (vitamin <strong>B12</strong>, CNCbl) is<br />

an artifact obtained during isolation of cobalamins from natural sources and is the most<br />

common pharmaceutical form (Rickes et al. 1948). CNCbl together with<br />

hydroxycobalamin (HOCbl), and glutathionylcobalamin (GSCbl) are precursors for<br />

synthesis of the biologically active coenzymes, methylcobalamin (CH3Cbl) and<br />

adenosylcobalamin (AdoCbl).<br />

Biosynthesis of Cobalamin<br />

Cobalamin or closely related corrinoid compounds are required in all kingdoms of<br />

life with perhaps the exception of plants and fungi. De novo synthesis of <strong>B12</strong> is restricted<br />

to some Bacteria and Archaea (Roth et al. 1996), requiring all other life forms that use<br />

<strong>B12</strong> to obtain complex precursors from their diet. Characterized as the most complex<br />

nonpolymeric natural substance known, cobalamin is synthesized de novo by a multi-step<br />

enzymatic process consisting of uroporphyrinogen III synthesis, side chain modification,<br />

cobalt insertion, aminopropanol addition, DMB biosynthesis, and addition of methyl or<br />

adenosyl moieties (Battersby 1994). The microbial biosynthesis of cobalamin can<br />

proceed along two different pathways: one aerobic and the other anaerobic (Scott 2003).<br />

These pathways differ both in their requirement for molecular oxygen and the timing of


cobalt insertion. Pseudomonas denitrificans and Bacillus megaterium use the aerobic<br />

pathway of cobalamin synthesis whereby molecular oxygen is used to assist with ring<br />

contraction and cobalt is inserted late in the synthesis. In contrast, cobalamin synthesis<br />

by the anaerobic pathway (used by Salmonella enterica and Propionibacterium<br />

4<br />

freundenreichii) is prevented by the presence of molecular oxygen, and cobalt is inserted<br />

early during synthesis. Unlike prokaryotes, humans are unable to synthesize cobalamin<br />

de novo and are dependent on a dietary source of a complex precursor such as vitamin<br />

<strong>B12</strong> from which AdoCbl and CH3Cbl are synthesized (Baker and Mathan 1981).<br />

Cobalamin-Dependent Enzymes<br />

AdoCbl and CH3Cbl share structural and chemical similarities; however, they<br />

catalyze very different biochemical reactions. The weak covalent carbon-cobalt bond of<br />

cobalamin, which is highly reactive and crucial for enzyme catalysis, is a shared trait of<br />

both cofactors (Frey and Reed 2000). In AdoCbl-dependent reactions, carbon-based free<br />

radicals are generated by the homolytic cleavage of the carbon-cobalt bond (Halpern<br />

1985); while in methyltransferase reactions, the carbon-cobalt bond of CH3Cbl is<br />

heterolytically cleaved, releasing a methyl group and cob(I)alamin (Banerjee 1997).<br />

AdoCbl-dependent enzymes can be classified into four types: eliminases (including<br />

dehydratases and deaminases), amino mutases, carbon skeleton mutases, and<br />

ribonucleotide reductases. The first three types are isomerases that catalyze<br />

1,2-rearrangement reactions (Table 1-1). Their main use is in the fermentation of small<br />

molecules and they use the same basic catalytic mechanism (Banerjee 1999). The<br />

reaction is triggered by substrate binding, causing homolytic cleavage of the<br />

carbon-cobalt bond, thereby generating an adenosyl radical and cob(II)alamin. The<br />

adenosyl radical abstracts a hydrogen atom from the substrate, forming a substrate radical


5<br />

and deoxyadenosine. After hydrogen abstraction, rearrangement of the substrate radical<br />

results in product radical formation. The product radical then abstracts a hydrogen atom<br />

from deoxyadenosine, forming both product and an adenosyl radical. Finally, the<br />

adenosyl radical recombines with cob(II)alamin, reforming the coenzyme (which can be<br />

used in additional catalytic cycles). The last type of AdoCbl-dependent enzymes is<br />

ribonucleotide reductases, which generate deoxyribonucleotides for DNA synthesis.<br />

There are three classes of ribonucleotide reductases. Class I ribonucleotide reductases<br />

are strict aerobic enzymes and are divided into two subclasses (Class Ia and Ib). Class Ia<br />

ribonucleotide reductases are found in eukaryotes, plants, viruses, and some prokaryotes.<br />

These enzymes contain three conserved cysteines, a tyrosyl radical, and a nonheme diiron<br />

center which are essential for catalysis (Aberg et al. 1989, Nordlund and Eklund 1993).<br />

Class Ib ribonucleotide reductases are found in prokaryotes and vary from class Ia in their<br />

lack of allosteric regulation and the use of different electron carriers for ribonucleotide<br />

reduction (Jordan et al. 1994). The Class II ribonucleotide reductases require AdoCbl,<br />

are found in bacteria and archaea and are active both aerobically and anaerobically.<br />

Their reaction mechanism is similar to that of the isomerase reaction, differing where the<br />

adenosyl radical does not interact directly with the substrate, but with a cysteinal radical<br />

in the active site of the enzyme (Booker et al. 1994). Class III ribonucleotide reductases<br />

are oxygen sensitive and are found in some prokaryotes and methanogens (Reichard<br />

1993). Although this class has not been well characterized it is thought that a glycyl<br />

radical, an iron sulfur center and S-adenosylmethionine (SAM) are required for catalysis<br />

(Harder et al. 1992).


6<br />

CH3Cbl-dependent enzymes are involved in methyl transfer reactions (Table 1-1).<br />

Only the reaction mechanisms of methionine synthase and the acetyl-CoA synthase have<br />

been studied extensively (Ljungdahl and Wood 1982, Jarrett et al. 1998); however, all of<br />

the methyltransferases are thought to follow the same general mechanism. The reaction<br />

involves three steps, and begins with the methylation of enzyme-bound cob(I)alamin by a<br />

methyl-group donor, forming enzyme-bound CH3Cbl. The carbon-cobalt bond of<br />

CH3Cbl is then heterolytically cleaved, resulting in enzyme-bound cob(I)alamin and a<br />

methyl cation, which is transferred to the substrate forming the methylated product.<br />

Methyltransferases are essential for methionine synthesis and are also involved in acetate<br />

and methane synthesis (Schneider and Stroinski 1987). Figure 1-2 shows the different<br />

types of <strong>B12</strong>-dependent enzymes and the reactions they catalyze.<br />

Cobalamin in Humans<br />

Humans are incapable of de novo synthesis of cobalamin, and must take up<br />

complex cobalamin precursors from the diet. Cobalamin is a water-soluble vitamin that<br />

cannot freely diffuse across the intestinal wall because of its large size and requires<br />

specific carriers and transporters for uptake and delivery to cells. Once in the cells,<br />

cobalamin is then metabolized to CH3Cbl in the cytosol and AdoCbl in the mitochondria,<br />

by a series of enzyme-mediated reactions.<br />

Cobalamin-Dependent Enzymes in Humans<br />

In humans, AdoCbl and CH3Cbl are required for two enzymes that are essential in<br />

metabolism (Kolhouse and Allen 1977). One of these enzymes is methylmalonyl-CoA<br />

mutase (MCM), an AdoCbl-dependent enzyme used to metabolize propionyl-CoA<br />

resulting from the breakdown of odd-chain fatty acids, cholesterol, and some amino<br />

acids. The other is methionine synthase (MS), a CH3Cbl-dependent enzyme used for


7<br />

homocysteine recycling and methionine synthesis. Deficiencies in MCM or MS can lead<br />

to methylmalonic aciduria or homocystinuria (rare but often lethal childhood disorders).<br />

Methylmalonyl-CoA mutase<br />

MCM is the only AdoCbl-dependent enzyme that is present in both bacterial and<br />

mammalian systems. In humans, it resides in the mitochondrial matrix, and is used for<br />

the metabolism of propionyl-CoA to succinyl-CoA. In this metabolic pathway,<br />

propionyl-CoA is carboxylated to (2S)-methylmalonyl-CoA, isomerized to<br />

(2R)-methylmalonyl-CoA, and lastly rearranged to succinyl-CoA by MCM (Figure 1-3).<br />

In contrast, in some bacteria including Propionibacterium shermanii, MCM is used in the<br />

opposite metabolic flow to produce propionate from succinate (Allen et al. 1964).<br />

The gene encoding MCM has been sequenced and cloned from many organisms<br />

including human, P. shermanii, mouse, Streptomyces cinnamonensis, Porphyromonas<br />

gingivalis, and Sinorhizobium meliloti (Banerjee and Chowdhury 1999). In humans, the<br />

location of the MCM gene was mapped to chromosome 6p12-21.2, and the cDNA was<br />

shown to have a mitochondrial targeting sequence (MTS) (Ledley et al. 1988).<br />

In native conformation, the human and bacterial MCMs are 150 kiloDaltons (kDa),<br />

and are dimeric proteins. The crystal structure of P. shermanii MCM has been<br />

determined, and is an αβ heterodimer that binds one mole of AdoCbl (Marsh et al. 1989).<br />

The human MCM is a homodimer of two α subunits that binds two moles of AdoCbl and<br />

has 60% sequence identity to the α chain of the bacterial MCM (Mancia et al. 1996).<br />

UV, visible, and electron paramagnetic resonance spectroscopic studies have<br />

shown that AdoCbl binding by MCM occurs via “Base-off” (class I) interaction<br />

(Padmakumar and Banerjee 1995). “Base-off” binding occurs by the displacement of the


DMB ligand by a histidine residue on the α-subunit of MCM. DMB is still attached to<br />

the corrin ring through a nucleotide loop; however, in the class I interaction, the<br />

coordinate bond of DMB to cobalt is displaced and substituted by His610 of the<br />

α-subunit. The exact role of the His610 residue is not fully understood, but current<br />

8<br />

studies suggest that its major role is in DMB displacement and organization of the active<br />

site for AdoCbl binding and enzyme catalysis (Vlasie et al. 2002).<br />

The isomerization of methylmalonyl-CoA to succinyl-CoA by MCM requires<br />

homolytic cleavage of the carbon-cobalt bond of enzyme-bound AdoCbl (Padmakumar<br />

and Banerjee 1997). Homolysis of the organometallic bond occurs once<br />

methylmalonyl-CoA binds MCM. This results in the formation of an adenosyl radical<br />

and cob(II)alamin. The carbon-cobalt bond of free AdoCbl is weak and in solution has a<br />

rate of homolysis of 4x10 -10 sec -1 . With the addition of MCM, the rate of homolysis is<br />

increased to higher than 600 sec -1 , suggesting that MCM promotes the homolysis of<br />

AdoCbl for this rearrangement reaction (Padmakumar and Banerjee 1997). The adenosyl<br />

radical generated by homolysis abstracts hydrogen from the methyl-group of<br />

methylmalonyl-CoA, forming a primary-substrate radical and 5’-deoxyadenosine. The<br />

primary-substrate radical undergoes an intramolecular 1,2-rearrangement resulting in the<br />

formation of a secondary product radical that abstracts hydrogen from deoxyadenosine.<br />

Lastly, there is recombination of the adenosyl radical and cob(II)alamin, regenerating<br />

AdoCbl and releasing succinyl-CoA that can flow into the tricarboxylic acid (TCA)<br />

cycle.<br />

In humans, defects in the enzyme MCM cause a block in propionyl-CoA<br />

metabolism, resulting in a buildup of methylmalonyl-CoA that is hydrolyzed to


methylmalonic acid. Methylmalonic acid is excreted into the urine, resulting in<br />

methylmalonic aciduria, an inherited disorder that is often fatal in newborns.<br />

Methionine synthase<br />

MS catalyzes the conversion of CH3THF and homocysteine to THF and<br />

methionine. This enzyme is found in both bacteria and humans. In E. coli, two MS<br />

enzymes occur and they differ in their requirement for CH3Cbl (Drummond and<br />

Matthews 1993). In humans, only CH3Cbl-dependent MS is present. The E. coli and<br />

human genes encoding CH3Cbl-dependent MS have been cloned and encode large<br />

9<br />

monomeric proteins of 136 and 141 kDa, respectively (Banerjee et al. 1989, Leclerc et al.<br />

1996). The human gene was mapped to chromosome 1 at position q43 (Leclerc et al.<br />

1996). The bacterial and the human enzymes share 55% identity in amino acid sequence,<br />

and have established similarities in the mechanism of catalysis (Hall et al. 2000).<br />

However, only the E. coli CH3Cbl-dependent MS has been studied extensively.<br />

Methionine synthase from E. coli is a modular enzyme consisting of four domains,<br />

all of which are crucial for enzyme catalysis (Goulding et al. 1997). The N-terminal<br />

domain (residues 2-353) has been cloned, expressed, and shown to be involved in<br />

homocysteine binding, and methyl group transfer from CH3Cbl to homocysteine<br />

(Goulding et al. 1997). It contains three cysteine residues used to coordinate a zinc ion<br />

essential for homocysteine binding, which are conserved in other MS enzymes including<br />

Homo sapiens, Caenorhabditis elegans, and Synechocystis species (Goulding and<br />

Matthews 1997). The second domain (residues 354-649) was shown to catalyze methyl<br />

transfer from CH3THF to cob(I)alamin (Goulding et al. 1997). The third domain<br />

(residues 650-896) was lacking enzymatic activity; however, it was shown to be essential<br />

for CH3Cbl binding (Banerjee et al. 1989). This fragment was crystallized and shown to


ind CH3Cbl in the class I “Base-off” mode (Drennan et al. 1994). The last domain<br />

10<br />

(residues 897-1227), termed the activation domain, was shown to bind SAM required for<br />

the reductive activation of MS (Dixon et al. 1996, Hall et al. 2000).<br />

The mechanism of MS catalysis and reactivation is shown in Figure 1-4. The<br />

methyltransferase reaction requires heterolytic cleavage of the carbon-cobalt bond of<br />

CH3Cbl resulting in cob(I)alamin and a methyl group. After heterolysis, MS transfers its<br />

methyl group to homocysteine, forming methionine and MS-bound cob(I)alamin. During<br />

catalysis, additional methyl groups are provided by methyltetrahydrofolate (CH3THF),<br />

which regenerates CH3Cbl and releases tetrahydrofolate (THF) (Banerjee and Matthews<br />

1990). Occasionally, the MS-bound cob(I)alamin is oxidized to cob(II)alamin during<br />

catalytic turnover, and reductive activation is required to restore the methylation cycle<br />

(Fujii et al. 1977, Banerjee et al. 1990, Drummond and Matthews 1993). This involves<br />

reduction of cob(II)alamin and methylation by SAM to form CH3Cbl. The reduction of<br />

cob(II)alamin for MS activation differs in E. coli and humans. Reduced flavodoxin<br />

(FldA) provides the needed electron in E. coli (Fujii and Huennekens 1974). However,<br />

humans lack flavodoxin and instead use the dual flavoprotein methionine synthase<br />

reductase (MSR) (Olteanu and Banerjee 2001).<br />

In humans, a block in the methylation cycle caused by defects in MS results in<br />

homocystinuria, a severe and sometimes fatal childhood disease (Fenton and Rosenberg<br />

2000). In addition, partial defects in MS can lead to elevated homocysteine, a major<br />

cardiovascular disease risk factor (Refsum et al. 1998). A deficiency in MS also results<br />

in the trapping of cellular folate as CH3THF, making it unavailable for other


folate-dependent reactions including purine and pyrimidine biosynthesis (Wilson et al.<br />

1999).<br />

Cobalamin Absorption and Transport<br />

11<br />

As mentioned above, humans are incapable of de novo cobalamin synthesis, and<br />

require complex precursors in their diet. Suitable precursors (such as HOCbl or CNCbl)<br />

can be obtained by the consumption of beef, liver, poultry, fish, eggs, dairy products, and<br />

vitamin supplements (Stabler 1999). The cobalt of these cobalamin molecules is in the<br />

+3 oxidation state (cob(III)alamin), the form that is recognized for cobalamin absorption<br />

and transport. Once ingested, cobalamin molecules that are bound to food proteins are<br />

released by the combined action of proteases and acid in the stomach (Del Corral and<br />

Carmel 1990). Haptocorrin (a cobalamin carrier protein) binds released cobalamin and<br />

transports it from the stomach to the small intestines. Haptocorrin has a high specificity<br />

for cobalamin and when bound to cobalamin, protects it from damage by acids in the<br />

stomach, until it reaches the small intestines where cobalamin is liberated from<br />

haptocorrin via digestion by pancreatic enzymes. After being released from haptocorrin,<br />

cobalamin is bound by intrinsic factor (IF), a glycoprotein produced by the parietal cells<br />

(gastric glands lining the stomach). The IF-cobalamin complex is resistant to further<br />

digestion in the small intestines because of the carbohydrates on the IF. The<br />

IF-cobalamin complex recognizes the IF-cobalamin receptor (IFCR) on the epithelial<br />

cells of the distal third small intestines (ileum), and is transported into these cells via<br />

receptor-mediated endocytosis. In the epithelial cell, IF is degraded by the acidic<br />

environment of the lysosome, and cobalamin is released. Transcobalamin II (TCII), a<br />

serum-transport protein in the epithelial cells, binds released cobalamin, and transports it<br />

out of the cell and into the bloodstream until it is taken up by other cells. The


12<br />

TCII-cobalamin complex binds the TCII receptor on the target cell, and gains entry via<br />

endocytosis where it can be used for intracellular cobalamin metabolism (Rosenblatt and<br />

Fenton 1999, Seetharam 1999, Watkins and Rosenblatt 2001).<br />

Intracellular Cobalamin Metabolism<br />

Once the TCII-cobalamin complex is internalized by the target cell, the transport<br />

protein is degraded in the lysosomal vesicle, and cobalamin is released from this<br />

complex. A cobalamin transporter then shuttles free cobalamin from the lysosome to the<br />

cytosol of the cell. Intracellular cobalamin metabolism (Figure 1-5) is thought to be<br />

similar in prokaryotes and eukaryotes (Qureshi et al. 1994, Watkins and Rosenblatt<br />

2001). In the cytosol, cob(III)alamin, in the form of HOCbl or CNCbl, is converted to<br />

GSCbl and then reduced to cob(II)alamin. These steps are proposed to be catalyzed by a<br />

β-ligand transferase and a cob(III)alamin reductase, respectively. After cobalamin<br />

reduction in the cytosol, cob(II)alamin is localized to the mitochondrial matrix and is<br />

reduced to cob(I)alamin by a mitochondrial cob(II)alamin reductase. The final step in<br />

AdoCbl formation is the adenosylation of cob(I)alamin to form AdoCbl by an<br />

ATP:cob(I)alamin adenosyltransferase. The pathways of AdoCbl and CH3Cbl synthesis<br />

are shared, and thought to branch apart after the reduction of cob(III)alamin. For CH3Cbl<br />

metabolism, cytosolic cob(II)alamin associates with MS, forming<br />

MS-bound-cob(II)alamin which is reduced to cob(I)alamin by MSR and methylated by<br />

SAM to form MS-CH3Cbl.<br />

β-ligand transferase<br />

The first step in cobalamin metabolism is removal of the β-ligand. Cobalamin<br />

β-ligand transferase activity has been detected in crude cell extracts of Clostridium


tetanomorphum, P. shermanii, Euglena gracilis, human leukocytes, human skin<br />

fibroblasts, rat liver, and rabbit spleen (Weissbach et al. 1961, Brady et al. 1962,<br />

Watanabe et al. 1987, Pezacka et al. 1990, Pezacka 1993). The enzymes involved and<br />

13<br />

their encoding genes have not been identified. However, it was determined that for the<br />

enzymatic conversion of CNCbl to AdoCbl, cell extracts required supplementation with<br />

adenosine triphosphate (ATP), reduced flavin, and reduced glutathione (Weissbach et al.<br />

1962). Studies in mammalian cells suggest that GSCbl is a product of the β-ligand<br />

transferase reaction (Pezacka et al. 1990). In bacteria, this GSCbl intermediate has not<br />

been found.<br />

Cob(III)alamin reductase<br />

The second enzymatic step in cobalamin metabolism is cob(III)alamin reduction.<br />

Reductase activity has been detected in crude cell extracts of C. tetanomorphum,<br />

P. shermanii, E. gracilis, human, and rat (Weissbach et al. 1961, Brady et al. 1962,<br />

Watanabe et al. 1987, Pezacka et al. 1990, Pezacka 1993). Extracts of C. tetanomorphum<br />

required NADH and either flavin mononucleotide (FMN) or flavin adenine dinucleotide<br />

(FAD) to mediate reductase activity (Walker et al. 1969). The flavin oxidoreductase<br />

(Fre) of S. enterica was purified and shown to mediate the nonenzymatic reduction of<br />

cobalamin in vitro (Fonseca and Escalante-Semerena 2000). Fre reduces flavin<br />

nucleotides, which in turn chemically reduces cob(III)alamin. In E. gracilis,<br />

cob(III)alamin reductase activity was purified from the mitochondrial fraction, and was<br />

ultimately identified as a flavoprotein with an absolute requirement for NADPH but not<br />

FAD or FMN (Watanabe et al. 1987). Further analysis of this protein revealed that it<br />

contained one molecule of either FAD or FMN that remained bound during the


14<br />

purification, explaining why cob(III)alamin reductase activity was retained without the<br />

addition of these flavins. Cell extracts from human fibroblasts and rat liver have<br />

cob(III)alamin reductase activity associated with both microsomes and the mitochondrial<br />

membrane, that was derived from oxidation-reduction reactions by cytochromes and<br />

flavoproteins (Watanabe et al. 1989, Watanabe et al. 1996). There has been no genetic<br />

evidence that the flavoproteins studied in these systems are the physiological<br />

cob(III)alamin reductases. The studies by Watanabe and colleagues also indicate that<br />

both prokaryotic and eukaryotic systems produce redundant enzymes with the ability to<br />

reduce cobalamin; however, this reduction could be a primarily nonenzymatic process.<br />

Cob(II)alamin reductase<br />

Cob(I)alamin is one of the strongest nucleophiles known to exist in aqueous<br />

solution, and is a powerful reductant (E°’ = -0.61 V) (Banerjee et al. 1990). Because of<br />

the nucleophilic nature of cob(I)alamin, cob(II)alamin reduction is an energetically<br />

unfavorable reaction. Thus, it is no surprise that cob(II)alamin reductase activity can<br />

only be studied by coupling the reaction to an alkylation event such as adenosylation or<br />

methylation. Cob(II)alamin reductase activity has been detected in cell-free extracts of<br />

C. tetanomorphum and P. shermanii; however, these enzymes have not been identified,<br />

and the encoding genes are unknown (Brady et al. 1962, Weissbach et al. 1962, Walker et<br />

al. 1969). In humans, cob(II)alamin reductase activity has been detected in mitochondrial<br />

fractions, but the enzyme was never identified, and the gene is unknown (Fenton and<br />

Rosenberg 1978b).<br />

Purification of the cob(II)alamin reductase from cell extracts of C. tetanomorphum<br />

did not produce a single enzyme, but produced a complex containing the adenosylating


enzyme. This suggests that cob(II)alamin reductase and the adenosyltransferase may<br />

15<br />

exist as a structural complex under certain conditions (Walker et al. 1969). Advantages<br />

of this complex would decrease the probability of unfavorable cob(I)alamin side<br />

reactions and instead allow for adenosylation because of the close proximity of<br />

cob(I)alamin to the adenosyltransferase.<br />

The E. coli FldA protein has been shown to function as a cob(II)alamin reductase<br />

for the formation of AdoCbl (Fonseca and Escalante-Semerena 2001). In vitro,<br />

flavodoxin reductase (Fpr), flavodoxin (FldA), and purified CobA (an<br />

adenosyltransferase) catalyze AdoCbl synthesis from cob(III)alamin (Fonseca and<br />

Escalante-Semerena 2001). In this coupled assay, it is proposed that Fpr uses NADPH to<br />

reduce FldA, which then reduces cob(III)alamin to cob(I)alamin while bound to CobA<br />

(Fonseca and Escalante-Semerena 2001). To date, there are no reports showing an<br />

interaction between FldA and CobA. Although this reducing system is functional in<br />

vitro, there is no genetic evidence that FldA is the physiological cob(II)alamin reductase<br />

(Fonseca and Escalante-Semerena 2001).<br />

Adenosyltransferase<br />

Adenosylation is the terminal step in AdoCbl metabolism. ATP:cob(I)alamin<br />

adenosyltransferase (ATR) activity was detected in crude cell extracts of human<br />

fibroblast cells, but the enzyme involved and the encoding gene were never identified<br />

(Fenton and Rosenberg 1981). ATRs from P shermanii, C. tetanomorphum,<br />

P. denitrificans, Thermoplasma acidophilum, and S. enterica have been purified and<br />

partially characterized (Brady et al. 1962, Vitols et al. 1966, Debussche et al. 1991, Suh<br />

and Escalante-Semerena 1995, Johnson et al. 2001, Saridakis et al. 2004). The genes<br />

encoding these enzymes have been identified from P. denitrificans, T. acidophilum, and


16<br />

S. enterica (Crouzet et al. 1991, Suh and Escalante-Semerena 1993, Johnson et al. 2001,<br />

Saridakis et al. 2004).<br />

To date, three families of adenosyltransferases have been identified: EutT-type,<br />

PduO-type, and CobA-type. These families are classified based on amino acid sequence<br />

similarity and examples of each are found in S. enterica. EutT is thought to be involved<br />

in the conversion of CNCbl to AdoCbl for ethanolamine utilization (Kofoid et al. 1999).<br />

PduO adenosylates cob(I)alamin for AdoCbl synthesis which is required for<br />

<strong>B12</strong>-dependent 1,2-propanediol metabolism (Johnson et al. 2001). CobA can substitute<br />

for PduO in cob(I)alamin adenosylation for 1,2-propanediol metabolism, but its primary<br />

role is the adenosylation of cobalamin intermediates for de novo <strong>B12</strong> biosynthesis (Suh<br />

and Escalante-Semerena 1995, Johnson et al. 2001). The CobA-type and PduO-type<br />

ATRs have been extensively studied and well characterized biochemically.<br />

CobA ATR from S. enterica has been overexpressed and purified (Suh and<br />

Escalante-Semerena 1995). This enzyme was shown to have a broad specificity for<br />

reduced corrinoid substrates, which could be due to its role in de novo cobalamin<br />

synthesis (Suh and Escalante-Semerena 1995). The X-ray crystal structure of CobA in its<br />

free state, complexed with ATP and HOCbl has been resolved (Bauer et al. 2001). It is a<br />

homodimer, which is consistent with the subunit composition of CobO ATR from<br />

P. denitrificans (Debussche et al. 1991, Bauer et al. 2001). The N-terminal P-loop<br />

(GNGKGKT) of CobA coordinates the α-, β-, and γ-phosphates of ATP, however the<br />

crystal structure shows that this interaction occurs in the opposite orientation when<br />

compared to other P-loop dependent nucleotide hydrolases (Bauer et al. 2001). This<br />

unique mode of ATP binding positions the 2’-OH group to allow for nucleophilic attack


at the 5’-C of ATP by cob(I)alamin resulting in adenosylation (Fonseca et al. 2002).<br />

Additionally, triphosphate was identified as a reaction byproduct of adenosylation and<br />

was also found to be a strong inhibitor of the reaction (Fonseca et al. 2002).<br />

17<br />

The archaeal T. acidophilum PduO-type ATR has been purified, crystallized, and<br />

partially characterized (Saridakis et al. 2004). T. acidophilum ATR has 27% amino acid<br />

identity to PduO. This ATR is specific for ATP and deoxy-ATP. Additionally, the<br />

crystal structure revealed a trimeric configuration. Each of the subunits is composed of<br />

five alpha helical domains that aid in subunit assembly due to hydrophobic, ionic, and<br />

polar interactions depending on the domain. In contrast to CobA, there is no<br />

recognizable P-loop motif suggesting a different binding method for ATP.<br />

Formation of methylcobalamin<br />

The synthesis CH3Cbl is coupled to cob(II)alamin reduction (Huennekens et al.<br />

1982). Once cob(II)alamin is reduced, methylation is dependent on SAM as the methyl<br />

donor for the formation of CH3Cbl. The E. coli cob(II)alamin reductase requires FldA<br />

for the reductive activation of MS (Fujii and Huennekens 1974). Humans lack FldA,<br />

requiring an alternate reductase for CH3Cbl formation.<br />

Because of the role of FldA in bacteria, it was proposed that the human reductase<br />

would contain binding sites for FMN and FAD (Leclerc et al. 1998). Using homology<br />

based PCR, a cDNA with consensus sequences to predicted binding sites for FMN, FAD,<br />

and NADPH was cloned and proposed to encode methionine synthase reductase (MSR),<br />

the cob(II)alamin reductase required for MS activation (Leclerc et al. 1998). The MSR<br />

gene (MTRR) was mapped to chromosome 5 at position p15.2 (Leclerc et al. 1999).<br />

Analysis of the gene revealed that the N-terminal sequence is consistent with cytosolic<br />

targeting, but alternative splicing at the 5’ end of MTRR mRNA generates a


mitochondrial targeting sequence (Leclerc et al. 1999). There have been no studies<br />

18<br />

documenting the role of MSR in the mitochondria. Banerjee and Olteanu (2001) showed<br />

that MSR is a monomeric protein of 77 kDa, that when purified contains stoichiometric<br />

amounts of FAD and FMN. In an NADPH-dependent reaction, purified MSR reduced<br />

cob(II)alamin to cob(I)alamin for activation of MS in vitro (Olteanu and Banerjee 2001).<br />

In these same studies, it was found that the reduction of cobalamin bound to MS by MSR<br />

is dependent on electrostatic interactions.<br />

Inherited Disorders of Cobalamin Metabolism<br />

In humans, inherited disorders in the metabolic pathways of both AdoCbl and<br />

CH3Cbl synthesis result in methylmalonic aciduria and homocystinuria. Methylmalonic<br />

aciduria is caused by a block in the pathway of propionyl-CoA metabolism which causes<br />

an accumulation methylmalonyl-CoA that is hydrolyzed to methylmalonic acid, and<br />

excreted into the blood and urine (Rosenblatt and Cooper 1990). Clinically,<br />

methylmalonic aciduria presents itself by lethargy, recurrent vomiting, dehydration,<br />

respiratory distress, feeding difficulties, failure to thrive, hypotonia, and death (Ciani et<br />

al. 2000). Homocystinuria is the result of a block in the methylation cycle for methionine<br />

synthesis, and results in an accumulation of homocysteine, which is excreted into the<br />

blood and urine (Rosenblatt and Cooper 1990). Homocystinuria presents itself by<br />

megaloblastic anemia, delayed development, neurological disorders, cardiovascular<br />

disease, and death (Kapadia 1995). All of the disorders involved in cobalamin<br />

metabolism are inherited as autosomal recessive traits (Rosenblatt and Fenton 1999).<br />

There have been no reported heterozygotes with the presentation of these diseases.<br />

Inborn errors of cobalamin metabolism have been classified into eight distinct<br />

complementation groups, cblA to cblH (Figure 1-5). This classification is based on


iochemical studies and complementation analysis of human cultured fibroblast cells<br />

(Fenton and Rosenberg 1978a). Complementation analysis examines the uptake and<br />

19<br />

conversion of labeled precursors (propionate, CH3THF) by fused and unfused fibroblast<br />

cell lines derived from two patients with cobalamin disorders. In these studies, fused<br />

cells that stimulated the incorporation of labeled precursors complemented each other and<br />

were assigned to different complementation groups. If the fused cells were unable to<br />

stimulate incorporation of the precursors they were assigned to the same<br />

complementation group (Gravel et al. 1975). An assay that has helped to determine if<br />

defects are due to decreased synthesis of one or both cobalamin coenzymes involves<br />

incubating the cultured fibroblasts with labeled HOCbl and measuring the cofactors<br />

formed against control cell lines.<br />

Combined Methylmalonic Aciduria and Homocystinuria<br />

Defects in cellular cobalamin metabolism resulting in both methylmalonic aciduria<br />

and homocystinuria are defined by the complementation groups cblF, cblC, and cblD.<br />

Cultured fibroblasts from patients in these groups were analyzed and shown to have a<br />

deficiency in the conversion of CNCbl or HOCbl to AdoCbl or CH3Cbl cofactors<br />

(Qureshi et al. 1994). The absence of these cofactors renders both MCM and MS<br />

inactive. There are over 100 patients with combined methylmalonic aciduria and<br />

homocystinuria. The cblC complementation group has more than 100 patients, the cblD<br />

group has only two patients who are siblings, and the cblF group has six patients<br />

(Watkins and Rosenblatt 2001).<br />

Individuals with the cblF disorder have impaired MCM and MS activities. The<br />

total intracellular pools of AdoCbl and CH3Cbl are reduced when compared to control<br />

cells. Cultured cells incubated with [ 57 Co]-labeled CNCbl showed an accumulation of


20<br />

unmetabolized, non-protein bound CNCbl contained within the lysosomes (Qureshi et al.<br />

1994). There was no synthesis of either AdoCbl or CH3Cbl in these cells. This disorder<br />

is defined by a deficiency in the release of cobalamin from the lysosomes after<br />

endocytosis making cobalamin unavailable to intracellular enzymes (Kapadia 1995). The<br />

gene encoding cobalamin lysosomal transport has not been identified. Other clinical<br />

presentations of this disorder include hypotonia, stomatitis, facial abnormalities,<br />

developmental retardation, and failure to thrive (Fenton and Rosenberg 2000).<br />

Individuals diagnosed with the cblC and cblD disorders have very similar<br />

biochemical characteristics. Both of these groups have decreased total cobalamin content<br />

in the liver, kidney, and fibroblasts (Kapadia 1995). Cultured cells from these patients<br />

incubated with [ 57 Co]-labeled CNCbl or HOCbl have reduced conversion to labeled<br />

AdoCbl or CH3Cbl when compared to controls (Mahoney et al. 1971). Because of the<br />

reduced levels of these cofactors, MCM and MS activities are deficient in cultured cells<br />

even though the enzymes are normal (Mellman et al. 1978). In light of these results, it is<br />

proposed that cblC and cblD complementation groups result from a defect in cytosolic<br />

cob(III)alamin reductase activity (Watkins and Rosenblatt 2001). Based on<br />

complementation studies, the cblC and cblD groups are classified as mutations occurring<br />

at unique chromosomal locations; however, the specific role of these two groups has not<br />

been determined and their corresponding genes remain unidentified.<br />

The clinical presentation of the cblC disorder is more severe than cblD. In addition<br />

to methylmalonic aciduria and homocystinuria, the cblC group usually presents itself in<br />

the neonatal stage (early-onset) resulting in megaloblastic anemia, congenital<br />

malformations, lethargy, failure to thrive, poor feeding, and death (Rosenblatt et al.


1997). Both cases of the cblD disorders did not develop problems until later on in life<br />

21<br />

(late-onset) resulting in megaloblastic marrow, neuromuscular problems, methylmalonic<br />

aciduria, homocystinuria, and mental retardation (Rosenblatt and Cooper 1990). Since<br />

the cblD group is biochemically similar to cblC but presents itself as less severe, it is<br />

suggested that cblD is a leaky form of cblC, or that it may be affecting an unidentified<br />

step before or after the reduction of cobalamin (Qureshi et al. 1994). The cblC, cblD, and<br />

cblF groups respond to pharmacological doses of HOCbl which reduces methylmalonic<br />

aciduria and homocystinuria but does not eliminate them (Fenton and Rosenberg 2000).<br />

Methylmalonic Aciduria without Homocystinuria<br />

Defects in the cellular metabolism of cobalamin resulting in methylmalonic<br />

aciduria without homocystinuria are defined by cblA, cblB, and cblH complementation<br />

groups. Complementation analysis of fibroblast cells from these patients has shown that<br />

the cblA, cblB, and cblH defects occur at three distinct chromosomal locations (Qureshi<br />

et al. 1994, Watkins et al. 2000). Patients with these disorders have a deficiency in<br />

intracellular AdoCbl levels and have normal CH3Cbl levels (Watkins and Rosenblatt<br />

2001). There are at least 45 patients diagnosed with cblA, 36 individuals with cblB, and<br />

only one reported individual with cblH disorder (Watkins and Rosenblatt 2001).<br />

The cblA and cblH complementation groups are deficient in mitochondrial AdoCbl<br />

synthesis. Cultured cells from these patients were unable to convert [ 57 Co]-labeled<br />

HOCbl to labeled AdoCbl but had no abnormality in CH3Cbl synthesis (Fenton and<br />

Rosenberg 2000). Cell extracts from cblA patients required a reducing system for the<br />

synthesis of AdoCbl; however, intact mitochondria from these patients were unable to<br />

synthesize AdoCbl even in the presence of a reducing system. On the other hand, control<br />

intact mitochondria can synthesize AdoCbl from HOCbl without the addition of reducing


22<br />

agents (Fenton and Rosenberg 1978b), suggesting that the cblA disorder could be due to a<br />

deficiency in cobalamin reduction or abnormalities of mitochondrial cobalamin binding<br />

and transport (Mahoney et al. 1975).<br />

The gene encoding the cblA disorder was identified (MMAA) and shown to localize<br />

to chromosome 4q31.1-2 (Dobson et al. 2002). It appears that this gene product has<br />

mitochondrial functionality due to the presence of an N-terminal MTS. Dobson and<br />

colleagues suggested that the MMAA protein does not encode a mitochondrial cobalamin<br />

reductase because it lacks NADPH, and flavin binding motifs that would be essential for<br />

reductase activity (Watanabe et al. 1987), (Dobson et al. 2002). The MMAA protein has<br />

61% similarity to ArgK in E. coli, an accessory protein common to lysine (L), arginine<br />

(A), ornithine (O), and LAO transport systems (Dobson et al. 2002). It is thought that<br />

ArgK could have some function in <strong>B12</strong> transport in E. coli and has been proposed that<br />

MMAA could encode a mitochondrial cobalamin transporter (Dobson et al. 2002).<br />

Recently, MeaB, a homolog of the human MMAA, from Methylobacterium<br />

extorquens AM1 was identified (Korotkova et al. 2002). Homology searches (using<br />

MeaB and MCM) identified a fusion protein from Burkholderia fungorum with an<br />

N-terminal MeaB domain and a C-terminal MCM domain (Korotkova and Lidstrom<br />

2004). The MCM and MeaB homologue fusion protein in B. fungorum suggests that<br />

these two proteins would interact in systems where the genes are expressed separately<br />

(Korotkova and Lidstrom 2004). It has been shown that the M. extorquens MeaB exists<br />

as a complex with MCM (Korotkova and Lidstrom 2004). It is now thought that the cblA<br />

group (MMAA) encodes a protein that would interact with MCM and protect it from<br />

inactivation (Korotkova and Lidstrom 2004).


23<br />

The cblB complementation group is characterized by a deficiency in the conversion<br />

of cobalamin to AdoCbl. Cell extracts derived from patients with cblB disorder were<br />

unable to synthesize AdoCbl even when provided with a reducing system, labeled<br />

HOCbl, and ATP (Fenton and Rosenberg 1981). Patients with this disorder are defective<br />

in ATP:cob(I)alamin ATR activity. At the start of this study, the gene encoding the cblB<br />

defect was unknown and the protein had not been isolated.<br />

The clinical presentation of cblA, cblB, and cblH disorders are very similar and<br />

include methylmalonic aciduria, lethargy, failure to thrive, vomiting, dehydration,<br />

respiratory distress, and anemia (Fenton and Rosenberg 2000). These disorders have<br />

been treated by a diet restricted in amino acid precursors of methylmalonate and<br />

cobalamin supplementation (Kapadia 1995).<br />

Homocystinuria without Methylmalonic Aciduria<br />

Inborn errors of cellular cobalamin metabolism resulting in homocystinuria without<br />

methylmalonic aciduria are defined by cblE and cblG complementation groups. There<br />

has been only one patient in the cblE group that had combined homocystinuria and mild<br />

methylmalonic aciduria (Wilson et al. 1999). Cultured fibroblasts from these patients<br />

show a deficiency in CH3Cbl synthesis and decreased methionine synthase activity<br />

(Watkins and Rosenblatt 1989). Complementation analysis has shown that cblE and<br />

cblG defects occur at distinct chromosomal locations (Watkins and Rosenblatt 1988).<br />

Fibroblasts from patients with cblE and cblG disorders showed normal incorporation of<br />

labeled propionate into macromolecules; however, the incorporation of labeled CH3THF<br />

was reduced when compared to controls suggesting an abnormality in the methionine<br />

synthase pathway (Watkins and Rosenblatt 1989). To date there are 12 patients


24<br />

diagnosed with the cblE disorder and 28 individuals with the cblG disorder (Fenton and<br />

Rosenberg 2000).<br />

Members of the cblE complementation group are deficient in methionine synthase<br />

activity due to a block in CH3Cbl formation. Fibroblast extracts from patients with this<br />

disorder had normal methionine synthase activity when high concentrations of reducing<br />

agent were added to the assay, but activity was reduced or absent when the reducing<br />

agent was used at lower concentrations (Watkins and Rosenblatt 1988). From these<br />

studies it was shown that the cblE group is defective in cobalamin reduction. The gene<br />

corresponding to this group was identified (MTRR) and localized to chromosome<br />

5p15.2-p15.3 (Leclerc et al. 1998). The MTRR protein or MSR is required for the<br />

reductive activation of methionine synthase and was reviewed above (Wilson et al. 1999).<br />

The cblG complementation group is also deficient in methionine synthase activity<br />

but has normal levels of CH3Cbl. Fibroblast extracts from these patients were assayed<br />

and showed no MS activity under optimal conditions (Watkins and Rosenblatt 1989).<br />

This group is defined by defects in MS. Human MS is a modular enzyme with domains<br />

containing binding sites for homocysteine, CH3THF, SAM, and cob(II)alamin (Banerjee<br />

1997). The cDNAs from cblG mutants have been analyzed and shown to contain<br />

missense and nonsense mutations (Leclerc et al. 1996). It appears that a majority of<br />

deficiencies in the cblG group are due to the lack of MS protein (caused by a premature<br />

stop codon due to nonsense mutations) and not in cobalamin binding. The clinical<br />

presentations of the cblE and cblG groups include homocystinuria without methylmalonic<br />

aciduria, developmental delay, megaloblastic anemia, and neurological disorders


(Watkins et al. 2002). Some patients respond to treatment with HOCbl (Watkins and<br />

Rosenblatt 2001).<br />

25<br />

Salmonella as a Model Organism for the Study of Cobalamin Metabolism<br />

Salmonella enterica is an important model organism for studies of <strong>B12</strong>-dependent<br />

processes (Schneider and Stroinski 1987, Roth et al. 1996). More is known about<br />

cobalamin metabolism and physiology in S. enterica than in any other single organism.<br />

Cobalamin transport, cobalamin biosynthesis, intramolecular rearrangements, and methyl<br />

transfer reactions have been investigated in S. enterica (Roth et al. 1996). In addition,<br />

S. enterica has a well-defined genetic system that includes methods for transformation,<br />

transduction, gene expression, directed, chemical, and transposon mutagenesis and a<br />

known genome sequence.<br />

One of the best-studied AdoCbl dependent enzymes in S. enterica is diol<br />

dehydratase. This enzyme is crucial for the metabolism of 1,2-propanediol. The genes<br />

required for growth of S. enterica on 1,2-propanediol are organized at the propanediol<br />

utilization (pdu) locus. Determination of the DNA sequence indicated that this locus has<br />

23 genes (Bobik et al. 1997, Bobik et al. 1999): six pdu genes are thought to encode<br />

enzymes needed for the 1,2-propanediol degradative pathway (pduCDEPQW) (Bobik et<br />

al. 1997); two are involved in transport and regulation (pduF and pocR) (Bobik et al.<br />

1992, Chen et al. 1994); two are probably involved in diol dehydratase reactivation<br />

(pduGH) (Bobik et al. 1999); one is needed for the conversion of cobalamin to AdoCbl<br />

(pduO) (Johnson et al. 2001); five are of unknown function (pduLMSVX); and seven<br />

share similarities to genes involved in the formation of carboxysomes (pduABJKNTU),<br />

polyhedral bodies found in certain cyanobacteria and chemoautotrophs (Shively and<br />

English 1991, Shively et al. 1998). The large number of genes required for the


26<br />

degradation of propanediol suggests that this compound serves an important purpose in<br />

the lifestyle of Salmonella (Roth et al. 1996).<br />

Salmonella uses propanediol as a carbon and energy source in a <strong>B12</strong>-dependent<br />

manner. The proposed pathway of propanediol degradation was determined based on<br />

biochemical studies (Figure 1-6). The pathway begins with the conversion of<br />

1,2-propanediol to propionaldehyde, a reaction catalyzed by the AdoCbl-dependent diol<br />

dehydratase (Toraya et al. 1979, Obradors et al. 1988). Propionaldehyde is then<br />

disproportionated to propanol and propionic acid by a reaction series thought to involve<br />

propanol dehydrogenase, phosphotransacylase, and propionate kinase (Toraya et al. 1979,<br />

Obradors et al. 1988). Aerobic growth of S. enterica on 1,2-propanediol requires<br />

addition of CNCbl (or other complex cobalamin precursors) to growth media.<br />

Research Overview<br />

Coenzyme <strong>B12</strong>-dependent processes are essential for human health.<br />

Methylmalonyl-CoA mutase and methionine synthase, the only two reported<br />

<strong>B12</strong>-dependent enzymes in humans, are crucial for propionate metabolism and methionine<br />

synthesis. Inborn errors of cobalamin metabolism can block these pathways resulting in<br />

methylmalonic aciduria and homocystinuria, serious diseases that are often fatal in<br />

newborns. Cobalamin metabolism has been studied in both bacterial and eukaryotic<br />

systems, and has been found to be quite similar. Progress in identifying human genes<br />

involved in <strong>B12</strong> metabolism has been slow due to difficulties in purifying the<br />

corresponding enzymes and the lack of facile genetic techniques. To circumvent these<br />

problems, S. enterica was used as a model system for studies of <strong>B12</strong> metabolism in<br />

humans.


Chapter 2 of this dissertation reports the identification of the human<br />

27<br />

ATP:cob(I)alamin adenosyltransferase involved in AdoCbl metabolism. An S. enterica<br />

mutant deficient in ATR activity allowed for the isolation of a bovine ATR cDNA.<br />

Subsequent sequence similarity searching was used to identify a homologous human<br />

cDNA. Both the bovine and human cDNAs were independently cloned and<br />

overexpressed in E. coli. Enzyme assays showed that both the bovine and human<br />

enzymes had ATR activity in vitro. Subsequent studies showed that the human cDNA<br />

clone complemented an ATR-deficient S. enterica strain for AdoCbl-dependent growth<br />

on 1,2-propanediol in vivo. In addition, Western blots were used to show that ATR<br />

expression is altered in cell lines derived from cblB methylmalonic aciduria patients<br />

compared with cell lines from normal individuals.<br />

Chapter 3 of this dissertation reports the biochemical properties of the human ATR<br />

and its interaction with methionine synthase reductase. Two common polymorphic<br />

variants of the ATR, which are found in normal individuals, were expressed in E. coli and<br />

purified to apparent homogeneity. Purified ATR variants were used for kinetic studies<br />

and the catalytic properties of both ATR variants including the Vmax and the Km for ATP<br />

and cob(I)alamin were determined. Investigations also showed that the purified<br />

methionine synthase reductase in combination with purified ATR can convert<br />

cob(II)alamin to AdoCbl in vitro. In this system, MSR reduced cob(II)alamin to<br />

cob(I)alamin which was adenosylated to AdoCbl by the ATR enzyme and results<br />

indicated that MSR and ATR interact in such a way that the highly reactive intermediate<br />

(cob(I)alamin) was sequestered. The finding that MSR can reduce cob(II)alamin to<br />

cob(I)alamin for AdoCbl synthesis (in conjunction with the prior finding that MSR


educed cob(II)alamin for the activation of methionine synthase) indicates a dual<br />

physiological role for the MSR enzyme.<br />

28<br />

Chapter 4 of this dissertation is not directly related to <strong>B12</strong> metabolism in humans,<br />

but was the research performed during my first two years in Dr. Bobik’s Laboratory.<br />

This work focused on the identification of PduP, a coenzyme-A-acylating<br />

propionaldehyde dehydrogenase associated with polyhedral bodies involved in<br />

AdoCbl-dependent 1,2-propanediol degradation by S. enterica. A PCR based method<br />

was used to construct a precise nonpolar chromosomal deletion of the gene pduP. The<br />

resulting pduP deletion strain grew poorly on 1,2-propanediol minimal media and<br />

expressed 105-fold less propionaldehyde dehydrogenase activity than did wild-type<br />

S. enterica grown under similar conditions. An E. coli strain was constructed for<br />

high-level production of His8-PduP, which was purified by nickel-affinity<br />

chromatography and was shown to have propionaldehyde dehydrogenase activity.<br />

Reverse-phase High Performance Liquid Chromatography (HPLC) and mass<br />

spectrometry were used to verify the product of the PduP reaction. Immunogold electron<br />

microscopy was used to show that PduP is associated with a polyhedral organelle<br />

involved in AdoCbl-dependent 1,2-propanediol degradation.


Corrin<br />

Ring<br />

Dimethyl<br />

benzimidazole<br />

H H2NC 2NC<br />

O H O 3C H3C O<br />

H H2NC 2NC<br />

Aminopropanol<br />

H H3C 3C<br />

H H3C 3C<br />

H H3C 3C<br />

C<br />

C<br />

H H3C 3C<br />

H H2NC 2NC<br />

CH<br />

C<br />

C N N<br />

Co 3+<br />

Co 3+<br />

C<br />

C N N<br />

O<br />

HN<br />

A<br />

O<br />

C<br />

D C<br />

C<br />

C C<br />

C<br />

C<br />

C<br />

H H3C 3C<br />

H<br />

C<br />

CH 3<br />

R<br />

C<br />

CH 3<br />

N<br />

N<br />

OH<br />

29<br />

R= upper β-ligand<br />

H H3C 3C CNH 2<br />

C<br />

C<br />

CNH 2<br />

O<br />

O<br />

O<br />

B<br />

O<br />

O<br />

C<br />

C<br />

P O<br />

O<br />

CH<br />

CH CH2OH 2OH<br />

CH 3<br />

CH 3<br />

O<br />

CNH 2<br />

R= Active<br />

Adenosyl<br />

AdoCbl<br />

Methyl<br />

CH CH3Cbl 3Cbl<br />

R= Inactive<br />

Cyano<br />

CNCbl,<br />

vitamin B <strong>B12</strong> 12<br />

Hydroxy<br />

HOCbl<br />

Glutathionyl<br />

GSCbl<br />

N<br />

NH 2<br />

N<br />

OH<br />

CH 2<br />

R<br />

R<br />

O<br />

N<br />

N<br />

OH<br />

R CH 3<br />

C N<br />

R OH<br />

R GS<br />

Figure 1-1. Structure of cobalamins. The major parts of cobalamin are labeled: upper<br />

β-ligand, corrin ring, aminopropanol, and dimethylbenzimidazole. The four<br />

pyrrole rings of the corrin are labeled with capital letters. The upper ligand<br />

can be any of the groups listed in the legend. Active and inactive forms of<br />

cobalamin are dependent on the upper ligand.


Table 1-1. Adenosylcobalamin- and methylcobalamin-dependent reactions<br />

AdoCbl-dependent Isomerases MeCbl-dependent Methyltransferases<br />

Diol dehydratase:<br />

Methionine synthase:<br />

1,2-propanediol propionaldehyde + homocysteine CH3THF methionine +<br />

H20<br />

THF<br />

Glycerol dehydratase:<br />

glycerol β-hydroxypropionaldehyde<br />

+ H20<br />

Ethanolamine ammonia lyase:<br />

ethanolamine acetaldehyde + H20<br />

L-β-Lysine aminomutase:<br />

L-β-lysine <br />

L-erythro-3,5-diaminohexanoic acid<br />

D-α-Lysine aminomutase:<br />

D-α-lysine 2,5-diaminohexanoic acid<br />

D-Ornithine aminomutase:<br />

D-ornithine 2,4-diaminovaleric acid<br />

Methylmalonyl-CoA mutase:<br />

methylmalonyl-CoA succinyl-CoA<br />

Glutamate mutase:<br />

glutamate β-methylaspartate<br />

Isobutyryl-CoA mutase:<br />

isobutyryl-CoA butyryl-CoA<br />

Methyleneglutarate mutase:<br />

2-methylene glutarate <br />

3-methylitaconate<br />

Ribonucleotide reductase:<br />

NTP dNTP<br />

30<br />

Acetyl -CoA synthesis via Corrinoid/Fe<br />

protein:<br />

(a) CH3THF + corrinoid/FeS protein <br />

THF + CH3-corrinoid Fe/S protein<br />

(b) CH3-corrinoid Fe/S protein + carbon<br />

monoxide corrinoid Fe/S protein +<br />

acetyl-CoA<br />

Methyltetrahydromethanopterin:coenzyme<br />

M methyltransferase:<br />

CH3-H4MPT + coenzyme M <br />

H4MPT + CH3-coenzyme M<br />

Methanol:2-mercaptoethanesulfonic acid<br />

methyltransferase:<br />

CH3OH + coenzyme M <br />

CH3-coenzyme M + H20


Classes<br />

Eliminases<br />

Dehydratases<br />

OH<br />

Aminomutases<br />

OH<br />

1,2-propanediol<br />

H 2N COH<br />

NH 2<br />

Carbon skeleton mutases<br />

HOC<br />

Ribonucleotide reductases<br />

(P)PPO<br />

ethanolamine<br />

methylmalonyl-CoA<br />

H<br />

H<br />

OH<br />

O<br />

NDP, NTP<br />

Methyltransferases<br />

O<br />

Deaminases<br />

H 2N<br />

O<br />

H<br />

C<br />

CH 3<br />

NH 2<br />

H<br />

OH<br />

H<br />

OH<br />

O<br />

BASE<br />

31<br />

AdoCbl<br />

diol dehydratase<br />

ethanolamine ammonia lyase<br />

L-β-lysine 5,6 aminomutase<br />

O<br />

propionaldehyde<br />

H 3C<br />

CO SCoA<br />

AdoCbl<br />

methylmalonyl CoA-mutase<br />

HOC C<br />

cob(I)alamin<br />

Reactions<br />

AdoCbl<br />

AdoCbl<br />

CH 3Cbl<br />

O<br />

CH<br />

acetaldehyde<br />

β-lysine L-erythro 3,5 diaminohexanoic acid<br />

AdoCbl<br />

ribonucleotide reductase<br />

H 2<br />

C<br />

H 2<br />

+<br />

+<br />

succinyl-CoA<br />

HOC CH CH2 CH2 SH<br />

CH3THF THF<br />

HOC CH CH2 CH2 homocysteine<br />

methionine synthase<br />

methionine<br />

O<br />

H 3C<br />

O<br />

(P)PPO<br />

NH 2<br />

NH 2<br />

H<br />

H<br />

OH<br />

NH 2<br />

O<br />

dNDP, dNTP<br />

H 2O<br />

NH 3<br />

O<br />

COH<br />

CO SCoA<br />

H<br />

H<br />

H<br />

BASE<br />

S CH 3<br />

Figure 1-2. Different classes of cobalamin-dependent enzymes. The enzymes involved<br />

and the reactions they catalyze are highlighted.


H3C C CO SCoA<br />

H 2<br />

propionyl-CoA<br />

carboxylase<br />

HOOC<br />

propionyl-CoA<br />

CH 3<br />

CH CO<br />

Biotin, HCO 3 2- ,<br />

ATP, Mg 2+<br />

SCoA<br />

(2S)-methylmalonyl-CoA<br />

methylmalonyl-CoA<br />

epimerase<br />

HOOC<br />

(2R)-methylmalonyl-CoA<br />

methylmalonyl-CoA<br />

mutase<br />

HOOC C C<br />

H 2<br />

H<br />

C CO SCoA<br />

CH 3<br />

H 2<br />

succinyl-CoA<br />

CENTRAL<br />

<strong>METABOLISM</strong><br />

adenosylcobalamin<br />

CO SCoA<br />

32<br />

Valine<br />

Isoleucine<br />

Methionine<br />

Threonine<br />

Odd chain fatty acids<br />

Thymine<br />

Cholesterol<br />

HOOC<br />

adenosyltransferase<br />

CH 3<br />

CH COOH<br />

methylmalonic acid<br />

cob(I)alamin<br />

cob(II)alamin<br />

cob(II)alamin<br />

reductase<br />

cob(III)alamin<br />

reductase<br />

GSCbl<br />

β-ligand<br />

transferase<br />

HOCbl<br />

Figure 1-3. Pathways of propionyl-CoA catabolism and cobalamin metabolism.<br />

Propionyl-CoA carboxylase catalyzes the formation of<br />

(2S)-methylmalonyl-CoA. Methylmalonyl-CoA epimerase catalyzes the<br />

conversion of (2S)-methylmalonyl-CoA to the (2R) isomer.<br />

AdoCbl-dependent methylmalonyl-CoA mutase catalyzes the conversion of<br />

(2R)-methylmalonyl-CoA to succinyl-CoA. In humans, a deficiency in<br />

MCM or in any of the metabolic steps needed for AdoCbl synthesis leads to<br />

methylmalonic aciduria.


A)<br />

B)<br />

CH 3THF<br />

33<br />

MS-cob(I)alamin<br />

Methionine<br />

MS-CH3Cbl Synthase<br />

MS-cob(II)alamin<br />

methionine<br />

Methionine<br />

Synthase<br />

Reductase<br />

THF<br />

homocysteine<br />

MS-CH 3Cbl<br />

SAM SAH<br />

Figure 1-4. Pathways involving methionine synthase and the reductive activation of<br />

methionine synthase by methionine synthase reductase. A) Proposed<br />

pathway of methionine synthesis. In this pathway, CH3Cbl-dependent MS<br />

transfers a methyl from CH3THF to homocysteine forming methionine. B)<br />

Reductive activation of MS-bound cob(II)alamin. MSR reduces cobalamin,<br />

and SAM provides a methyl for the reactivation of MS which reenters the<br />

methylation cycle.


TCII<br />

HOCbl<br />

β-ligand<br />

transferase<br />

Lysosome<br />

TCII<br />

HOCbl<br />

TCII<br />

HOCbl<br />

HOCbl<br />

GSCbl<br />

cblF<br />

Cytoplasm<br />

34<br />

Methylmalonyl-CoA<br />

Methylmalonyl-CoA<br />

Mutase mut<br />

Adenosylcobalamin<br />

Succinyl-CoA<br />

cblC, cblD<br />

Mitochondrion<br />

Cob(III)alamin Reductase<br />

cblA<br />

Adenosyltransferase<br />

Cob(II)alamin<br />

Reductase<br />

MS Reductase<br />

SAM cblE<br />

MS-CH MS-CH3Cbl 3Cbl<br />

cblB<br />

Cob(I)alamin<br />

cblH<br />

Cob(II)alamin<br />

Cob(II)alamin<br />

MS-Cob(II)alamin<br />

CH CH3THF 3THF<br />

MS-Cob(I)alamin<br />

Homocysteine<br />

cblG<br />

Methionine Synthase<br />

Methionine<br />

Figure 1-5. Cobalamin metabolism in mammalian cells, and disorders associated with<br />

deficiencies in this pathway. The pathways by which inactive cobalamin<br />

precursors are taken up and metabolized to active cofactors are shown. The<br />

cytoplasmic and mitochondrial compartments are labeled. The specific steps<br />

affected by inborn errors of cobalamin metabolism (mut, cblA-cblH) are<br />

labeled.


propanol<br />

dehydrogenase<br />

(PduQ)<br />

propanol<br />

OH<br />

OH<br />

diol dehydratase<br />

(PduCDE)<br />

NAD +<br />

35<br />

1,2-propanediol<br />

OH<br />

O<br />

propionaldehyde<br />

NADH<br />

+ H +<br />

adenosylcobalamin<br />

NAD +<br />

NADH<br />

+ H +<br />

adenosyltransferase<br />

(PduO)<br />

CoA-dependent<br />

propionaldehyde dehydrogenase<br />

(PduP)<br />

O<br />

SCoA<br />

propionyl-CoA<br />

O<br />

PO 3 2-<br />

propionyl-phosphate<br />

ADP<br />

propionate<br />

Figure 1-6. Proposed pathway of AdoCbl-dependent 1,2-propanediol degradation in<br />

S. enterica. AdoCbl-dependent diol dehydratase catalyzes the conversion of<br />

1,2-propanediol to propionaldehyde. Propionaldehyde is then<br />

disproportionated into propanol and propionate presumably by aldehyde<br />

dehydrogenase (PduP), phosphotransacylase, propionate kinase (PduW) and<br />

propanol dehydrogenase (PduQ). When grown aerobically, S. enterica is<br />

unable to synthesize cobalamin de novo and requires cobalamin precursors<br />

(XCbl) for the metabolism to AdoCbl.<br />

ATP<br />

phosphotransacylase<br />

propionate kinase<br />

(PduW)<br />

O<br />

O<br />

XCbl


CHAPTER 2<br />

IDENTIFICATION OF THE HUMAN AND BOV<strong>IN</strong>E ATP:COB(I)ALAM<strong>IN</strong><br />

ADENOSYLTRANSFERASE CDNAS BASED ON COMPLEMENTATION OF A<br />

BACTERIAL MUTANT<br />

Introduction<br />

Enzymes dependent on the vitamin <strong>B12</strong> coenzymes, AdoCbl and CH3Cbl have a<br />

broad but uneven distribution among the three domains of life (Dolphin 1982, Schneider<br />

and Stroinski 1987, Banerjee 1999). In higher animals, there are two known<br />

cobalamin-dependent enzymes. CH3Cbl dependent MS is needed for the methylation of<br />

homocysteine to methionine (Cauthen et al. 1966, Drummond and Matthews 1993), and<br />

AdoCbl-dependent MCM plays an essential role in the conversion of propionyl-CoA to<br />

the TCA cycle intermediate, succinyl-CoA (Banerjee and Chowdhury 1999, Matthews<br />

1999). This latter process occurs in three steps: propionyl-CoA is carboxylated to<br />

(2S)-methylmalonyl-CoA, isomerized to (2R)-methylmalonyl-CoA, and finally<br />

rearranged to succinyl-CoA in a reaction catalyzed by AdoCbl-dependent MCM<br />

(Figure 1-3). In higher animals, propionyl-CoA is produced from the breakdown of the<br />

amino acids valine, isoleucine, methionine, and threonine, as well as thymine, cholesterol<br />

and odd-chain fatty acids; hence, MCM is essential for the complete catabolism of each<br />

of these compounds (Banerjee and Chowdhury 1999).<br />

In humans, inherited defects that impair the activity of AdoCbl-dependent MCM<br />

lead to methylmalonic aciduria, a rare but severe disease that is often fatal in the first year<br />

of life (Rosenblatt and Fenton 1999, Fenton et al. 2000, Olteanu and Banerjee 2001).<br />

Such inherited deficiencies can result from mutations in the MCM structural gene (mut)<br />

36


or from mutations that impair the acquisition of the required cofactor, AdoCbl<br />

(Rosenblatt and Fenton 1999, Watkins and Rosenblatt 2001). Higher animals are<br />

incapable of de novo synthesis and, hence, must obtain AdoCbl by synthesis from<br />

37<br />

complex precursors taken up from their diet (Banerjee 1999). Suitable precursors include<br />

vitamin <strong>B12</strong> (CNCbl) and other cobalamins with various β-ligands (XCbls) (Banerjee<br />

1999). The pathway by which XCbls are converted to AdoCbl has been studied in<br />

several organisms and is thought to be similar in both prokaryotes and eukaryotes<br />

(Figure 1-3): XCbl is converted to glutathionyl-cobalamin (GSCbl), reduced to<br />

cob(II)alamin, further reduced to cob(I)alamin and finally adenosylated to AdoCbl<br />

(Friedmann 1975, Huennekens et al. 1982, Pezacka et al. 1990, Pezacka 1993, Watanabe<br />

et al. 1996). In humans, four complementation groups (cblABCD), associated with<br />

methylmalonic aciduria, are thought to correspond to genes involved in the conversion of<br />

XCbls to AdoCbl (Figure 1-5). Enzymatic assays of fibroblast extracts have indicated<br />

that the cblC and cblD complementation groups encode cytoplasmic enzyme(s) needed<br />

for the conversion of XCbls to cob(II)alamin (Mellman et al. 1979). Similar studies have<br />

indicated that the cblA and cblB complementation groups correspond to a mitochondrial<br />

cob(II)alamin reductase and an ATP:cob(I)alamin ATR enzymes, respectively (Mahoney<br />

et al. 1975, Fenton and Rosenberg 1981). To date, the human genes that correspond to<br />

the cblABCD complementation groups have not been identified. Progress in this area has<br />

been slow due to the difficulties in purifying the relevant proteins (Rosenblatt and Cooper<br />

1990).<br />

We recently identified the PduO ATR of Salmonella, and showed that this enzyme<br />

has partial functional redundancy with the CobA enzyme (Johnson et al. 2001). In these


studies, it was also shown that a Salmonella pduO cobA double mutant lacked<br />

38<br />

AdoCbl-dependent diol dehydratase activity due to the ATR deficiency. Furthermore, it<br />

was found that a plasmid-encoded source of ATR enzyme restored diol dehydratase<br />

activity to a cobA pduO double mutant (Johnson et al. 2001). Because this enzymatic<br />

activity can be readily detected on aldehyde indicator medium (AIM), we expected that<br />

the ATR-deficient Salmonella strain could be used to screen expression libraries for<br />

cDNAs that encode ATR enzymes. Such an approach would circumvent the difficulties<br />

associated with ATR purification.<br />

Here, we report the isolation of an ATR cDNA, from a bovine liver library, by<br />

complementation of an ATR-deficient Salmonella mutant for color formation on<br />

aldehyde indicator medium. Subsequently, sequence similarity searching identified the<br />

homologous human cDNA and its corresponding gene. Both the human and bovine<br />

cDNAs are shown to express ATR activity and complement an ATR-deficient bacterial<br />

mutant. In addition, Western blots were used to show that expression of the human ATR<br />

was altered in cell lines derived from three cblB patients compared with a cell line<br />

derived from a normal individual. We propose that the human gene identified here<br />

coincides with the cblB complementation group, defects in which lead to methylmalonic<br />

aciduria. The identification of genes involved in methylmalonic aciduria is important for<br />

the development of improved methods for the diagnoses and treatment of this devastating<br />

disorder.<br />

Chemicals and Reagents<br />

Materials and Methods<br />

CNCbl and HOCbl were purchased from Sigma Chemical Company (St. Louis,<br />

MO). Titanium (III) citrate was prepared as previously described (Bobik and Wolfe


1989). 5-bromo-4-chloro-3-indolyl-β-D-galactopyranoside (X-Gal) and<br />

isopropyl-β-D-thiogalactopyranoside (IPTG) were from Diagnostic Chemicals Ltd.<br />

39<br />

(Charlottetown, Prince Edward Island, Canada). Restriction enzymes and T4 DNA ligase<br />

were from New England Biolabs (Beverly, MA). Other chemicals were from Fisher<br />

Scientific (Norcross, GA).<br />

Bacterial Strains and Growth Media<br />

Bacterial strains used in this study are listed in Table 2-1. The minimal medium<br />

used was NCE (Vogel and Bonner 1956, Berkowitz et al. 1968) supplemented with 0.4%<br />

1,2-propanediol, 200 ng/ml of CNCbl, 1 mM MgSO4, 0.3 mM each of valine, isoleucine,<br />

leucine, and threonine. LB medium was the rich medium used (Difco Laboratories,<br />

Detroit, MI) (Miller 1972). MacConkey and aldehyde indicator media were<br />

supplemented with 1,2-propanediol and CNCbl and prepared as previously described<br />

(Johnson et al. 2001).<br />

General Protein and Molecular Methods and ATR Assays<br />

Bacterial transformation, polymerase chain reaction (PCR), restriction enzyme<br />

digests, and other standard molecular and protein methods were performed as previously<br />

described (Maniatis et al. 1982, Johnson et al. 2001). ATR assays were also performed<br />

as previously reported (Johnson et al. 2001).<br />

P22 Transductions<br />

Transductional crosses were performed as described previously (Davis et al. 1980).<br />

In preparing lysates of galE mutant strains, overnight cultures were grown on LB<br />

medium supplemented with 0.2% galactose and 0.2% glucose with the addition of<br />

appropriate antibiotics. A P22 HT105/1 int-201 phage was used for transductional<br />

crosses at a concentration of 2x10 8 phage/ml (Schmieger 1971).


40<br />

Screening the Bovine and Human Liver cDNA Libraries<br />

Uni-ZAP XR bovine and human liver cDNA libraries were from Stratagene (La<br />

Jolla, CA). Titering, amplification, and mass excision of the cDNA carried on<br />

pBlueScript SK(+/-) were done according to the Manufacturer’s protocol except that the<br />

titering procedure used F-top agar with 0.2 mM thymine, but no NaCl (Miller 1972)<br />

instead of NZY top agar. Following mass excision, the cDNA expression plasmids were<br />

purified using a QIAprep spin mini prep kit (Qiagen, Chatsworth, CA). A portion of the<br />

resulting expression library was used to transform S. enterica strain BE253 by<br />

electroporation. Strain BE253 is ATR-deficient due to pduO and cobA mutations, and<br />

carries pSJS1240 which provides rare tRNAs that enhance expression of heterologous<br />

genes in S. enterica. Following electroporation, transformation mixtures were suspended<br />

in molten aldehyde indicator medium that had been supplemented with 1,2-propanediol,<br />

HOCbl and 100 µg/ml ampicillin (Amp), and cooled to 45 ºC. The molten medium was<br />

poured into 100 x 15 mm sterile Petri dishes, allowed to solidify, and incubated at 37 ºC<br />

in the dark for ~12 h. Resultant colonies were screened for red/brown color formation.<br />

This procedure allowed screening of about 5,000 transformants per plate.<br />

Cloning the Bovine Adenosyltransferase Coding Sequence for High-Level<br />

Expression<br />

PCR was used to amplify the bovine ATR coding sequence. Plasmid pNL121,<br />

isolated in this study from a bovine cDNA library, provided the template DNA. The<br />

primers used for amplification of the full-length coding sequence were<br />

5’-GCCGCCGGTACCGATGACGACGACAAGTTCGGCACGAGCCCGGGAGGT-3’<br />

(forward) and 5’-GCCGCCAAGCTTGCTTGGTTCCTCGATGAAGCA-3’ (reverse).<br />

To eliminate the predicted mitochondrial targeting sequence (MTS), forward primer


41<br />

5’-GCCGCCGGTACCGATGACGACGACAAGCCCCAGGGCGTGGAAGACGGG-3’<br />

was used in conjunction with the reverse primer described above. These primers<br />

introduced KpnI and HindIII restriction sites into the PCR products and were designed<br />

such that following cloning, the bovine ATR coding sequence would be fused to both<br />

N-terminal glutathione S-transferase and His6 tags. PCR products obtained using the<br />

primers described above were restricted with KpnI and HindIII, and ligated to the<br />

pET-41a expression vector (Novagen, Cambridge, MA) that had been similarly digested<br />

(Maniatis et al. 1982). Ligation mixtures were used to transform E. coli DH5α by<br />

electroporation and transformants were selected by plating on LB-kanamycin (Kan)<br />

medium. Pure cultures were prepared from selected transformants and plasmid DNA<br />

isolated from these strains was analyzed by restriction digestion and DNA sequencing.<br />

Clones having the expected DNA sequences were transformed into E. coli strain BL21<br />

(DE3) RIL (Stratagene) for high-level expression.<br />

Cloning the Human Adenosyltransferase Coding Sequence for High-Level<br />

Expression<br />

The human ATR coding sequence was cloned via PCR using a strategy similar to<br />

that described above for the bovine enzyme. IMAGE cDNA clone 2822202 provided the<br />

template DNA (Incyte Genomics, Palo Alto, CA). For amplification of the full-length<br />

coding sequence, the following primers were used: forward<br />

5’-GCCGCCAGATCTGGATGACGACGACAAGATGGCTGTGTGCGGCCTGG-3’<br />

and reverse 5’-GCCGCCAAGCTTTCAGAGTCCCTCAGACTCGGCCG-3’. To<br />

eliminate the putative MTS, primer 5’-GCCGCCAGATCTGGATGACGACGACAAGC<br />

CTCAGGGCGTGGAAGACGGG-3’ was used as the forward primer. The primers<br />

described above, introduced BglII and HindIII restriction sites that were used for cloning


42<br />

into the pET-41a expression vector. The BglII site was positioned such that the resulting<br />

clones would express the human ATRs as fusion proteins with N-terminal GST and His6<br />

tags. Ligation, transformation and analysis of clones were performed as described above<br />

for the bovine expression clones. Clones with the expected DNA sequence were<br />

transformed into E. coli strain BL21 (DE3) RIL for high-level expression.<br />

Cloning the Human Adenosyltransferase Coding Sequence for Complementation<br />

Studies<br />

For complementation studies, the human ATR coding sequence was amplified by<br />

PCR and cloned into plasmid pLAC22 which allows IPTG-inducible expression in<br />

S. enterica (Warren et al. 2000). The primers used for amplification were forward,<br />

5’-GCCGCCAGATCTTATGCCTCAGGGCGTGGAAGACGGG-3’ and reverse,<br />

5’-GCCGCCAGCTTTCAGAGTCCCTCAGACTCGGCCG-3’. These primers eliminate<br />

the putative MTS and provide the needed ATG start triplet such that the first five amino<br />

acids of the expressed protein would be MPQGV. The PCR product was restricted with<br />

BamHI and HindIII and ligated into pLAC22 that had been digested with BglII and<br />

HindIII. Ligation mixtures were used to transform S. enterica strain BE253<br />

(ATR-deficient) by electroporation. Transformants were selected on AIM supplemented<br />

with 1,2-propanediol and Amp. Use of this medium allowed the identification of clones<br />

expressing ATR activity. Pure cultures were prepared from selected transformants and<br />

plasmid DNA isolated from these strains was analyzed by restriction digestion and DNA<br />

sequencing. Clones having the expected DNA sequences were moved into strain BE266<br />

(pduO cobA) via P22 transduction. Pure cultures prepared from the resultant colonies<br />

were determined to be phage-free by cross-streaking against P22 H5 and used for<br />

complementation studies.


43<br />

Growth of Adenosyltransferase Expression Strains and Preparation of Cell Extracts<br />

The E. coli strains used for expression of the bovine and human ATRs were grown<br />

on LB supplemented with 25 µg/ml Kan at 37°C with shaking at 275 rpm in a New<br />

Brunswick Scientific shaker incubator. Cells were grown to an absorbance of 0.6-0.8 at<br />

600 nm and protein expression was induced by the addition of 1 mM IPTG. Cells were<br />

incubated at 37°C with shaking at 275 rpm for an additional 3 hours. Cells were removed<br />

from the incubator, held on ice for 10 minutes, and then harvested by centrifugation at<br />

6,690 x gmax for 10 minutes using a Beckman JLA-10.500 rotor. The cells were<br />

resuspended in 3 ml of 50 mM sodium phosphate, 300 mM NaCl, pH 7, and broken using<br />

French Pressure Cell (SLM Aminco, Urbana, IL) at 20,000 psi. Phenylmethylsulfonyl<br />

fluoride was added to the cell extract to a concentration of 100 µg/ml to inhibit proteases.<br />

The crude cell extract was centrifuged at 16,000 rpm (31,000 x gmax) for 30 minutes using<br />

a Beckman JA20 rotor to separate the soluble and insoluble fractions. The supernatant<br />

was decanted (soluble fraction), and the pellet was resuspended in 1 ml of 50 mM sodium<br />

phosphate, 300 mM NaCl, pH 7 (inclusion body fraction). Both fractions were used for<br />

the ATR assays.<br />

Growth Curves<br />

Cells to be used as inocula for growth curves were grown overnight at 37°C in 2 ml<br />

LB medium or LB medium supplemented with Amp at 100 µg/ml for strains carrying<br />

pLAC22. A portion of the LB culture (1.5 ml) was pelleted by centrifugation, and<br />

resuspended in 1 ml of minimal medium. Then resuspended cells (125 µl) were used as<br />

the inoculum. Cells were grown in 16 x 150-mm culture tubes containing 5 ml of<br />

minimal medium and incubated at 37°C in a New Brunswick model C-24 water bath with


44<br />

a shaking speed of 250 rpm. Cell growth was determined by measuring the absorbance at<br />

600 nm using a Milton Roy model Spectronic 20D + spectrophotometer.<br />

Western Blots<br />

Human skin fibroblasts from one normal individual (MCH45) and three patients<br />

diagnosed with cobalamin b (cblB) disorder (WG1680, WG1879, and WG2127) were<br />

obtained from the Repository for Mutant Human Cell Strains at Montreal Children’s<br />

Hospital. These cell lines were grown in Eagle’s minimum essential medium<br />

supplemented with Earle’s salts, L-glutamine, 1% nonessential amino acids, 10%<br />

heat-inactivated fetal bovine serum, 100 IU/ml penicillin, 100 µg/ml streptomycin, and<br />

0.2 µg/ml HOCbl and incubated at 37°C with 5% CO2, 95% air mixture. For lysate<br />

preparation, each of the cell lines were cultured in 3, T150 cm 2 tissue culture flasks and<br />

grown to confluence. Harvested fibroblasts were washed twice with phosphate-buffered<br />

saline and lysed in 50 mM Tris, pH 7.3, 100 mM KCl, 5 mM MgCl2, 1.6% Triton-X-100<br />

and a protease inhibitor mixture (Roche Diagnostics, Manheim, Germany). Insoluble<br />

cellular matter was removed by centrifugation at 12,000 x g at 4 °C for 15 minutes.<br />

Protein concentration of the soluble extract was determined using a Micro BCA assay<br />

(Pierce, Rockford, IL). Cell Extracts (220 µg protein) were fractionated by 10% sodium<br />

dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE) and transferred onto a<br />

Hybond-P, polyvinylidene difluoride membrane (Amersham Biosciences). The<br />

membrane was incubated at room temperature for one hour with rocking in 10 mM<br />

Tris-HCl, pH 8, 150 mM NaCl, 0.05% Tween 20 (TBST) containing 3% bovine serum<br />

albumin and 3% mouse serum. The primary antibody used was affinity-purified rabbit<br />

IgG anti-human-ATR diluted 1:1000. The secondary antibody was monoclonal


anti-rabbit IgG (γ-chain specific) biotin conjugate (Sigma) and was used at a 1:25,000<br />

45<br />

dilution. All membrane incubations were carried out at room temperature for one hour<br />

with rocking in TBST. The membrane was prepared for detection using streptavidin<br />

horseradish peroxidase conjugate (Caltag Laboratories, Burlingame, CA) diluted<br />

1:10,000, and Immun-Star HRP Chemiluminescent Kit (Bio-Rad, Hercules, CA)<br />

following the Manufacturer’s protocol. Membranes were exposed to X-ray film from<br />

Research Products International (Mt. Prospect, IL).<br />

DNA Sequencing and Analysis<br />

DNA sequencing was carried out by University of Florida, Interdisciplinary Center<br />

for Biotechnology Research, DNA Sequencing Core facility, and University of Florida,<br />

Department of Microbiology and Cell Science, DNA Sequencing Facility as described<br />

previously (Bobik et al. 1999). Sequence similarity searches were carried out using the<br />

Blast family of sequence analysis software (Altschul et al. 1990, Altschul et al. 1997).<br />

Multiple sequence alignments were done with ClustalX using default parameters except<br />

as noted below (Thompson et al. 1997). The slow and accurate method was used for<br />

pairwise alignments with a gap opening and extension penalties of 35 and 0.75,<br />

respectively. Multiple sequence alignment parameters were gap opening and extension<br />

penalties of 15 and 0.3, respectively, and delay divergent sequences were set at 25%.<br />

Results<br />

Screening of the Bovine and Human Liver cDNA Libraries for Clones that Express<br />

Adenosyltransferase Activity<br />

Our previous studies showed that S. enterica strain BE253 forms red/brown<br />

colonies on AIM supplemented with 1,2-propanediol and CNCbl when it is transformed<br />

with an ATR expression plasmid (Johnson et al. 2001). The basis of the color formation


46<br />

is that ATR activity restores the ability of strain BE253 to convert CNCbl to AdoCbl, the<br />

required cofactor for diol dehydratase. Given AdoCbl, diol dehydratase converts<br />

1,2-propanediol to propionaldehyde which reacts with components in the aldehyde<br />

indicator medium to form a red/brown precipitate (Johnson et al. 2001). Hence, we<br />

expected that BE253 would be useful for screening expression libraries for ATR activity.<br />

Accordingly, we attempted to identify a human ATR cDNA by transforming strain<br />

BE253 with a human liver cDNA expression library and screening the resultant colonies<br />

for red/brown color formation on AIM. Approximately 250,000 human cDNAs were<br />

screened, but no red/brown colonies were found. Subsequently, a bovine liver cDNA<br />

library was similarly screened and four positive clones were isolated. The DNA<br />

sequences of the four clones were determined and found to be identical. The clone was<br />

1058 bp in length, and included a poly(A) tail indicating that its 3' end was complete.<br />

However, analyses indicated that the 5' end of the clone was incomplete. The clone<br />

included a 228-amino acid open reading frame that would be expressed as an N-terminal<br />

β-galactosidase fusion protein (as was expected given the cloning method employed), but<br />

it lacked the expected ATG triplet. Hence, the bovine clone apparently lacked a portion<br />

of its 5' end.<br />

After the DNA sequence of the bovine cDNA clones was determined, Blastp<br />

searches of the NCBI nonredundant database were conducted, and these searches showed<br />

that the 228-amino acid protein encoded by the bovine cDNA clone was 29% identical<br />

(Expect = 8 x 10 -15 ) to the N-terminal domain (165 amino acids) of the PduO ATR of<br />

S. enterica (Johnson et al. 2001). The findings that the bovine cDNA isolated both<br />

complemented an ATR-deficient mutant of S. enterica and encoded a protein with


sequence similarity to a known ATR enzyme suggested that the bovine clone also<br />

encoded an ATR enzyme.<br />

47<br />

Identification of a Human Gene Related to the Bovine Adenosyltransferase cDNA<br />

Blastp searches showed that the putative bovine ATR enzyme was 88% identical to<br />

a human protein encoded by IMAGE cDNA clone 2822202 (accession number<br />

BC005054). The human cDNA clone was 1128 bp in length and included an appropriate<br />

ATG triplet and a poly(A) tail indicating that it was a full-length clone. Further sequence<br />

analyses showed that the predicted protein encoded by the human cDNA was 250 amino<br />

acids in length and was homologous over its entire length to the putative bovine ATR<br />

except that it included 19 additional N-terminal amino acids (Figure 2-1). This suggested<br />

that the bovine clone was nearly full-length, lacking only a portion of its N-terminus,<br />

probably a small region of its MTS (see below). Perhaps more importantly, however, the<br />

high identity between the human and bovine sequences indicated that both encoded<br />

proteins with similar functions. Furthermore, sequence similarity searches showed<br />

human cDNA BC005054 encoded a protein with 26% amino acid identity the PduO ATR<br />

of S. enterica. Thus, these results tentatively identified sequence BC005054 as the<br />

human ATR cDNA.<br />

Subsequent sequence analyses showed human cDNA BC005054 was 99% identical<br />

to nine regions of human chromosome XII. The IMAGE clone BC005054 encoded a<br />

protein with 2 amino acid substitutions compared with the putative protein encoded by<br />

chromosome XII (Arg-19 to Gln and Met-239 M to Lys). These changes may represent<br />

natural polymorphisms or may have resulted from mutations introduced during cDNA<br />

preparation. The nine regions on chromosome XII were the only regions of the human<br />

genome that showed significant homology to human cDNA BC005054. Hence, it seems


48<br />

likely that these regions correspond to nine exons that encode the human ATR enzyme.<br />

The exon structure of the human ATR is indicated in Figure 2-1.<br />

Putative Subcellular Localization of Adenosyltransferase Enzymes<br />

The amino acid sequences of the human and bovine ATR enzymes were aligned<br />

with 20 related sequences obtained from GenBank (Figure 2-1). This alignment<br />

showed that the human, bovine, and mouse ATR sequences included about 50-60<br />

additional N-terminal amino acids compared with their prokaryotic homologues. In<br />

higher organisms, ATRs localize to mitochondria (Banerjee and Chowdhury 1999);<br />

hence, the additional N-terminal amino acids of the eukaryotic sequences might represent<br />

MTSs. Prediction of the subcellular localization of the human and mouse proteins using<br />

Predotar (Small et al. 2004) and MitoProt II (Claros and Vincens 1996) software<br />

indicated mitochondrial localization: human, MitoProt and Predotar scores both = 0.89;<br />

and mouse MitoProt and Predotar scores = 0.97 and 0.99, respectively. Because the<br />

bovine clones were apparently partial sequences, reliable prediction of subcellular<br />

localization of the presumptive bovine ATR enzyme was not possible (Claros and<br />

Vincens 1996, Small et al. 2004).<br />

High-Level Expression of the Bovine and Human Adenosyltransferases in E. coli<br />

E. coli strains were constructed for high-level expression of the presumptive bovine<br />

and human ATR enzymes. Clones were constructed to allow expression of these<br />

enzymes with and without sequences presumed to be involved in mitochondrial targeting.<br />

Our concern was that the MTS sequences might interfere with enzyme activity as such<br />

sequences are normally removed during mitochondrial localization (Paschen and Neupert<br />

2001). The arrows in Figure 2-1 indicate the regions of the bovine and human ATRs that<br />

were expressed by clones that eliminated their MTSs. Furthermore, in all cases, the


ATRs were produced as fusion proteins with N-terminal glutathione S-transferase and<br />

His6 tags that could be removed by enterokinase cleavage if desired.<br />

49<br />

Production of the presumptive bovine and human ATR fusion proteins was<br />

monitored by SDS-PAGE (Figure 2-2). Both the soluble (lanes 2 thru3) and insoluble<br />

fractions (lanes 5-7) were analyzed. Strains BE255 and BE256 produced large amounts<br />

of protein with molecular masses of 55 and 52 kDa, respectively (Figure 2-2 A, lanes 3,<br />

4, 6, and 7). These values are near the predicted masses (58 and 54 kDa) for the bovine<br />

ATR fusion proteins with and without the presumptive MTS and including the expression<br />

tags. In contrast, a strain carrying the expression plasmid without insert (BE237)<br />

produced little protein near these molecular masses (Figure 2-2 A, lanes 2 and 5)<br />

indicating that BE255 and BE256 were indeed producing large amounts of bovine ATR<br />

fusion proteins with and without the presumptive MTS.<br />

Expression of the human ATR fusion proteins was also monitored by SDS-PAGE<br />

(Figure 2-2 B). The predicted molecular masses of the human ATR fusion proteins with<br />

the expression tags, with and without the predicted MTS are 58 and 55 kDa, respectively.<br />

As shown in Figure 2-2 B, large amounts of proteins with calculated molecular masses of<br />

56 and 52 kDa were produced by expression strains BE257 (ATR + MTS) and BE258<br />

(ATR without MTS), (Figure 2-2 B, lanes 3, 4, 6, and 7) but not by the control strain<br />

BE237 (lanes 2 and 5) which carried that expression plasmid without insert.<br />

Assay of the Bovine and Human Adenosyltransferase Fusion Proteins for ATR<br />

Activity<br />

The cell extracts analyzed by SDS-PAGE (Figure 2-2) were tested for<br />

ATP:cob(I)alamin ATR activity (Table 2-2). Substantial activity was found in both<br />

soluble and inclusion body extracts from strains BE255 (bovine ATR + MTS), BE256


(bovine ATR), BE257 (human ATR + MTS), and BE258 (human ATR). Of the four<br />

50<br />

bovine ATR extracts tested, the highest specific activity measured was 85.7 nmol min -1<br />

mg -1 found in the inclusion body fraction from strain BE255 (ATR + MTS). The highest<br />

specific activity found in the human ATR extracts tested was 98 nmol min -1 mg -1 in the<br />

inclusion body fraction containing human ATR without its presumptive MTS. For each<br />

ATR fusion protein, the majority of the activity was found in the soluble fraction<br />

(Table 2-2, column 4). For both the human and bovine cell extracts, ATR activity was<br />

linear with protein concentration (data not show). Furthermore, when the substrates,<br />

ATP and/or cob(I)alamin were excluded from the assay mixture, no activity was detected.<br />

Also, as expected, cell extracts from strain BE237 (T7 expression vector without insert)<br />

expressed no measurable ATR activity.<br />

Complementation of an S. enterica Mutant Deficient in ATR Activity by a Human<br />

Adenosyltransferase cDNA Clone<br />

We previously showed that an S. enterica strain deficient for both the CobA and<br />

PduO ATRs (cobA pduO) was unable to grow on 1,2-propanediol due to an ATR<br />

deficiency (Johnson et al. 2001). To test whether the human ATR cDNA could<br />

complement this defect, we compared the growth of the wild-type strain, BE265 (pduO<br />

cobA /pLAC22-human ATR) and BE263 (pduO cobA /pLAC22-no insert) on<br />

1,2-propanediol minimal medium (Figure 2-3). The doubling times of the wild-type and<br />

strain BE265 (pduO cobA /pLAC22-human ATR cDNA) were found to be comparable;<br />

they were 8.4 and 8.9 h, respectively. Control experiments showed that strain BE263,<br />

which carried the pLAC22 vector without insert, grew minimally on 1,2-propanediol, and<br />

that growth of strain BE265 (pduO cobA/pNL135-human ATR) required the addition of<br />

IPTG as was expected since expression of the human ATR in this strain is under control


51<br />

of the lacI repressor (not shown). Thus, clearly, the human ATR cDNA complemented<br />

the ATR-deficient bacterial mutant for AdoCbl-dependent growth on 1,2-propanediol.<br />

These results provided further evidence that human cDNA BC005054 encodes an ATR<br />

enzyme, and also showed that this enzyme can function as an ATR under physiological<br />

conditions albeit in a heterologous host.<br />

Level of Human Adenosyltransferase Enzyme in Normal and cblB Mutant Human<br />

Skin Fibroblasts<br />

Above, we presented biochemical and genetic evidence that human cDNA<br />

BC005054 encodes an ATR enzyme. If defects in this enzyme underlie cblB<br />

methylmalonic aciduria, its expression or stability might be altered in cblB mutant cell<br />

lines. To examine the expression of this enzyme in normal and cblB mutant human skin<br />

fibroblasts, Western blots were performed on lysates from these cells using antibody<br />

against recombinant human ATR described above. Figure 2-4 shows a representative<br />

experiment. Lane 5 contained recombinant human ATR lacking both the mitochondrial<br />

targeting sequence and the expression tags. In this lane, a single protein band with an<br />

approximate molecular mass of 25 kDa was recognized by the anti-ATR antibody as<br />

expected. In lane 1, which contained cell extracts from normal skin fibroblasts, the<br />

anti-human ATR antibodies recognized two bands. One of these bands was similar in<br />

molecular mass to the recombinant human ATR (approximately 25 kDa) and the second<br />

band was of lower mass (approximately 23 kDa). Neither the 25 kDa nor the 23 kDa<br />

band was observed in blots performed with pre-immune serum (data not shown)<br />

indicating that these bands most likely represent two processed forms of the human ATR.<br />

Interestingly, ATR-reactive polypeptides of these molecular masses were differentially<br />

expressed in the three cblB cell lines tested (lanes 2-3). The cblB mutants WG1680 and


52<br />

WG2127 did not express detectable levels of the 25 kDa form of the ATR and expressed<br />

reduced levels of the 23 kDa form (lanes 2 and 4). For cblB mutant WG1879, the 25 kDa<br />

form was reduced, and the 23 kDa form was not detectable. These results indicated that<br />

expression of the ATR was significantly altered in cblB mutant human skin fibroblasts.<br />

These results provide direct evidence that defects in the ATR identified in this study<br />

underlie cblB methylmalonic aciduria.<br />

Conserved Amino Acids and Distribution of Adenosyltransferase Enzymes<br />

The human and bovine ATR enzymes were aligned with 20 related sequences<br />

obtained from GenBank (Figure 2-1). Of the sequences shown, the function of the<br />

H. sapiens, B. taurus, and S. enterica ATRs is supported by biochemical evidence<br />

presented here and previously (Johnson et al. 2001). To obtain additional information<br />

about the function of the remaining prokaryotic ATRs, we examined their genomic<br />

context. Genes encoding ATR homologues from S. enterica, L. innocua,<br />

L. monocytogenes, B. halodurans, C. perfringes, C. pasteurianum, and K. pneumonia<br />

were found proximal to either AdoCbl-dependent enzymes or AdoCbl biosynthetic genes.<br />

Since bacteria frequently cluster genes of related function, these finding suggest that the<br />

genes from these organisms do indeed encode ATR enzymes. It is also notable that there<br />

are several regions of high conservation in all the ATRs aligned suggesting that many are<br />

indeed ATR enzymes (Figure 2-1). Thus, PduO-type ATR enzymes appear to have broad<br />

phylogenetic distribution, being found in mammals, gram positive and gram negative<br />

bacteria, and the archaea.<br />

The highly conserved regions found in the ATRs also indicate amino acids likely to<br />

be essential for enzymatic activity, and may represent sites directly involved in catalysis<br />

or in binding of substrates, ATP and cob(I)alamin. There are at least two different ways


in which proteins bind <strong>B12</strong>, the "base-on" and "base-off" modes. Base-off binding<br />

53<br />

involves displacement of the lower axial ligand of <strong>B12</strong> (dimethylbenzimidazole) with an<br />

imidazole side-chain of a histidine residue (Drennan et al. 1994, Shibata et al. 1999,<br />

Marsh and Drennan 2001). The sequence containing the coordinating histidine<br />

(D-X-H-X-X-G), which is conserved among proteins that bind <strong>B12</strong> in the “base-off”<br />

mode, is absent from PduO-type ATRs examined in this study. This suggests that the<br />

ATR enzymes described in this study bind <strong>B12</strong> in the “base-on” mode.<br />

Discussion<br />

In this report we identified human and bovine cDNAs that encode ATR enzymes.<br />

The highest specific activities measured in cell extracts from expression strains were 85.7<br />

nmol min -1 mg -1 protein for the bovine ATR and 98 nmol min -1 mg -1 protein for the<br />

human ATR. These values are comparable to those previously reported for purified<br />

CobA ATR, and partially purified PduO ATR which were 53 and 312 nmol min -1 mg -1<br />

protein, respectively (Suh and Escalante-Semerena 1993, Suh and Escalante-Semerena<br />

1995). The bovine ATR cDNA was isolated based on its ability to complement an<br />

ATR-deficient S. enterica mutant for color formation on indicator medium, and the<br />

human ATR cDNA was shown to restore the ability of an ATR-deficient mutant to grow<br />

on 1,2-propanediol in an AdoCbl-dependent fashion. This demonstrated that the human<br />

and bovine cDNAs described here encode ATR enzymes that are active under<br />

physiological conditions, although in a heterologous host. Moreover, the bovine and<br />

human ATR cDNAs had 29% and 26% identity to the PduO ATR of S. enterica (Johnson<br />

et al. 2001) and the human enzyme included a presumptive MTS as expected (Fenton et<br />

al. 2000). Thus, genetic, biochemical, and bioinformatic evidence indicate that the<br />

bovine and human cDNAs analyzed here encode ATR enzymes.


54<br />

Prior studies have indicated that ATR defects in humans lead to methylmalonic<br />

aciduria (Fenton and Rosenberg 1981, Banerjee and Chowdhury 1999). MCM is known<br />

to require AdoCbl for activity (Banerjee and Chowdhury 1999) and fibroblast extracts<br />

from methylmalonic aciduria patients with defects in the cblB complementation group<br />

have been shown to lack ATR activity (Fenton and Rosenberg 1981). The ATR cDNA<br />

identified in this study corresponded to a single human gene comprised of nine exons<br />

found on chromosome XII. Furthermore, Western blots indicated that expression of this<br />

ATR was altered in cblB mutant human cell lines compared to normal cell lines. Thus,<br />

we propose that the human gene identified here corresponds to the cblB complementation<br />

group that is involved in methylmalonic aciduria.<br />

The approach used in this study to identify the human ATR cDNA involved<br />

screening mammalian cDNA expression libraries for clones that complemented an<br />

ATR-deficient bacterial mutant. This general strategy may have application to the<br />

identification of additional cDNAs involved in cobalamin metabolism. Human cDNAs<br />

encoding the β-ligand transferase and the cob(III)alamin and cob(II)alamin reductases<br />

(the cblACD complementation groups) have not yet been identified (Rosenblatt and<br />

Fenton 1999, Watkins and Rosenblatt 2001). In this regard, it is notable that in this study<br />

the human ATR cDNA was not directly isolated, but rather the bovine ATR cDNA was<br />

isolated and sequence similarity searching was used to identify the homologous human<br />

gene. The fact that our screen allowed direct isolation of a bovine ATR cDNA, but not a<br />

human ATR cDNA may reflect the relative abundance of ATR mRNA in the human<br />

versus the bovine liver. Ruminant livers have among the highest levels of <strong>B12</strong>-dependent<br />

enzymes measured in any tissue (Schneider and Stroinski 1987). Thus, ruminant liver


libraries may be particularly useful for isolating mammalian cDNAs involved in<br />

cobalamin metabolism.<br />

55<br />

Importantly, the identification of human genes involved in methylmalonic aciduria<br />

should help with the development of improved methods for diagnosis and treatment of<br />

this rare but devastating disease. Knowledge of the relevant genes will allow DNA-based<br />

methods of diagnosis following amniocentesis or chorionic villi sampling. Such<br />

techniques will allow the identification of the specific genetic lesions involved, and this<br />

may help guide treatment. Some cases of methylmalonic aciduria respond to high-dose<br />

<strong>B12</strong> therapy and it seems likely that such cases will be associated with specific mutations<br />

(Ampola et al. 1975, Matsui et al. 1983, Zass et al. 1995). Moreover, as the needed<br />

methodologies become available, knowledge of the genes involved in cobalamin<br />

metabolism may also help with the development of gene or enzyme therapies as<br />

treatments for methylmalonic aciduria.


Table 2-1. Bacterial strains<br />

Species<br />

E. coli<br />

Strain Genotype<br />

BL21 (DE3) RIL<br />

(E. coli B) F - ompT hsdS (rB - mB - ) dcm + Tet r gal<br />

SOLR TM<br />

(Stratagene)<br />

XL1-Blue MRF’<br />

(Stratagene)<br />

BE237<br />

56<br />

λ (DE3) endA Hte (argU ileY leuW Cam r )<br />

e14 - (McrA - ) ∆(mcrCB-hsdSMR-mrr)171 sbcC<br />

recB recJ uvrC umuC::Tn5 (Kan r ) lac gyrA96<br />

relA1 thi-1 endA1 λ R [F’ proAB lacI q Z∆M15] c<br />

Su - (nonsuppressing)<br />

D(mcrA)183 ∆(mcrCB-hsdSMR-mrr)173 endA1<br />

supE44 thi-1 recA1 gyrA96 relA1 lac [F’ proAB<br />

lacI q Z∆M15 Tn10 (Tet r )] c<br />

BL21 (DE3) RIL/pET-41a (T7 expression vector<br />

without insert)<br />

BE255 BL21 (DE3) RIL/pNL128 (bovine ATR+MTS)<br />

BE256<br />

BL21 (DE3) RIL/pNL129 (bovine ATR without<br />

MTS)<br />

BE257 BL21 (DE3) RIL/pNL132 (human ATR+MTS)<br />

BE258<br />

BL21 (DE3) RIL/pNL133 (human ATR without<br />

MTS)<br />

S. enterica serovar Typhimurium LT2<br />

TR6579<br />

metA22 metE551 trpD2 ilv-452 hsdLt6 hsdSA29<br />

HsdB - strA120 GalE - Leu - Pro -<br />

BE253<br />

TR6579 cobA::dcam ∆pduO651<br />

yeex::kan r /pSJS1240 (ileX, argU)<br />

BE254 BE253/pNL121 (pBlueScript with bovine ATR)<br />

BE266<br />

∆pduO651 cobA366::Tn10dCam/pSJS1240<br />

BE263<br />

BE265<br />

(ileX, argU)<br />

∆pduO651<br />

cobA366::Tn10dCam/pLAC22/pSJS1240 (ileX,<br />

argU)<br />

∆pduO651 cobA366::Tn10dCam/pNL135<br />

(human ATR without MTS under plac control)


57<br />

Figure 2-1. Multiple sequence alignment of adenosyltransferase enzymes. ClustalX was<br />

used to align the prokaryotic, eukaryotic, and archaeal homologues of ATR<br />

enzymes. Identical residues are shaded black, and highly conserved residues<br />

are shaded gray. The predicted mitochondrial targeting sequence determined<br />

by MitoProtII software is underlined in the human sequence. Triangles<br />

separate the human ATR into nine regions that correspond to exons on<br />

chromosome XII. Arrows represent the PCR primers used for amplification<br />

of the human ATR without its presumptive MTS.


A.<br />

58<br />

kDa 1 2 3 4 5 6 7 kDa 1 2 3 4 5 6 7<br />

116.25<br />

97.4<br />

97.4<br />

66.2<br />

66.2<br />

45<br />

31<br />

21.5<br />

14.4<br />

45<br />

31<br />

21.5<br />

Figure 2-2. SDS-PAGE analysis of cell extracts from bovine (A) and human (B) ATR<br />

expression strains. Panels A and B lane 1, molecular mass markers. Panel<br />

A, lanes 2 thru 4 are soluble extracts of BE237 (T7 expression vector without<br />

insert), BE255 (bovine ATR + MTS), and BE256 (bovine ATR without<br />

MTS). Panel A, lanes 5 thru 7 are inclusion body fractions from strains<br />

BE237, BE255, and BE256. Panel B, lanes 2 thru 4 are soluble extracts from<br />

BE237 (T7 expression vector without insert), BE257 (human ATR + MTS),<br />

and BE258 (human ATR without MTS), respectively. Lanes 5 thru 7 are<br />

inclusion body fractions from strains BE237, BE257, and BE258.<br />

B.


Table 2-2. Specific activities of bovine and human ATP:cob(I)alamin ATRs<br />

Strain S/I *<br />

59<br />

Total activity<br />

nmol min -1<br />

Total<br />

activity<br />

%<br />

Specific activity<br />

nmol min -1 mg -1<br />

protein<br />

BE255<br />

S 2021 95 47<br />

(bovine ATR+MTS) I 107 5 85.7<br />

BE256<br />

S 3364 84 44.1<br />

(bovine ATR without MTS) I 632 16 30<br />

BE257<br />

S 2445 52 37.5<br />

(human ATR+MTS) I 2296 48 60.8<br />

BE258<br />

S 5481 57 63<br />

(human ATR without MTS) I 4144 43 98<br />

* S=soluble fraction, I=inclusion body fraction


OD 600<br />

1.0<br />

0.5<br />

0.1<br />

60<br />

0 10 20 30 40<br />

Time (hours)<br />

Figure 2-3. Complementation of an ATR-deficient bacterial mutant for<br />

AdoCbl-dependent growth on 1,2-propanediol by a plasmid that expresses<br />

the human ATR. Cells were grown in minimal 1,2-propanediol medium<br />

supplemented with HOCbl and 0.1 mM IPTG. Growth was determined by<br />

measuring the absorbance of cultures at 600 nm (OD600), and data are shown<br />

as a semi-log plot. , Wild-type (S. enterica serovar Typhimurium LT2);<br />

, BE265, ATR-deficient/pNL135(human ATR under plac control); ,<br />

BE263, ATR-deficient /pLAC22 (vector without insert).


61<br />

1 2 3 4 5<br />

25 kDa<br />

Figure 2-4. Western blot analysis of ATR expression by normal and cblB mutant cell<br />

lines. Extracts were prepared from cultured fibroblast lines obtained from a<br />

normal individual and three cblB patients. Lane 1: cell extract from the<br />

MCH45 control cell line. Lanes 2 thru 4: cell extracts from cblB mutant cell<br />

lines WG1680, WG1879 and WG2127, respectively. 220 µg of protein was<br />

loaded into each of lanes 1 thru 4 (It was necessary to load relatively high<br />

amounts of protein because of the low abundance of the ATR enzyme in skin<br />

fibroblasts). Lane 5 contained 100 pg recombinant human ATR lacking the<br />

mitochondrial targeting sequence and affinity tags. The arrow shows the<br />

location of the recombinant ATR enzyme (approximately 25 kDa). The<br />

experiment was performed twice with similar results.


CHAPTER 3<br />

PURIFICATION AND <strong>IN</strong>ITIAL CHARACTERIZATION OF THE HUMAN<br />

ATP:COB(I)ALAM<strong>IN</strong> ADENOSYLTRANSFERASE AND ITS <strong>IN</strong>TERACTION WITH<br />

METHION<strong>IN</strong>E SYNTHASE REDUCTASE<br />

Introduction<br />

The vitamin <strong>B12</strong> coenzymes, adenosylcobalamin (AdoCbl) and methylcobalamin<br />

(CH3Cbl) are required cofactors for at least fifteen different enzymes that have a broad<br />

but uneven distribution among living forms (Schneider and Stroinski 1987, Banerjee<br />

1999). In humans, two <strong>B12</strong>-dependent enzymes are known (Kolhouse and Allen 1977,<br />

Mellman et al. 1977). AdoCbl -dependent methylmalonyl-CoA mutase (MCM) catalyzes<br />

the reversible rearrangement of methylmalonyl-CoA to succinyl-CoA (Banerjee and<br />

Chowdhury 1999). MCM is found in the mitochondrial matrix and it is required for the<br />

complete catabolism of compounds degraded via propionyl-CoA including<br />

branched-chain amino acids, odd-chain fatty acids, and cholesterol (Figure 3-1) (Banerjee<br />

and Chowdhury 1999). The second <strong>B12</strong>-dependent enzyme known in humans is<br />

CH3Cbl-dependent methionine synthase (MS). This enzyme is found in the cytoplasm<br />

where it catalyzes the conversion of methyltetrahydrofolate and homocysteine to<br />

tetrahydrofolate and methionine (Matthews 1999). In mammals, methionine is an<br />

essential amino acid, and MS plays a role in recycling the S-adenosylhomocysteine<br />

formed from S-adenosylmethionine-dependent methylation reactions (Weissbach et al.<br />

1963).<br />

Humans are incapable of synthesizing AdoCbl and CH3Cbl de novo and are<br />

dependent on dietary sources of complex cobalamin precursors, such as vitamin <strong>B12</strong><br />

62


63<br />

(cyanocobalamin, CNCbl) and hydroxycobalamin (HOCbl) (Stabler 1999). The pathway<br />

by which these precursors are metabolized to the coenzyme forms is thought to be similar<br />

in both prokaryotes and eukaryotes (Figure 3-1). For AdoCbl synthesis, CNCbl is<br />

proposed to be converted to glutathionylcobalamin (GSCbl), reduced successively to<br />

cob(II)alamin and cob(I)alamin respectively and finally, adenosylated to AdoCbl<br />

(Figure 3-1) (Friedmann 1975, Huennekens et al. 1982, Pezacka et al. 1990, Pezacka<br />

1993, Watanabe et al. 1996). For the synthesis of CH3Cbl, cob(II)alamin associated with<br />

MS, is reductively methylated to form CH3Cbl (Figure 3-1) (Drummond and Matthews<br />

1994, Matthews 1999). For this reaction (which is also used for the reductive activation<br />

of MS following adventitious oxidation of the cobalamin cofactor),<br />

S-adenosylmethionine is the methyl-group donor and methionine synthase reductase<br />

(MSR) reduces cob(II)alamin to cob(I)alamin (Leclerc et al. 1998, Olteanu and Banerjee<br />

2001).<br />

In humans, inherited defects in the MCM or MS structural genes or in the genes<br />

needed for the synthesis of <strong>B12</strong> coenzymes result in methylmalonic aciduria,<br />

homocystinuria, or combined disease. These rare disorders, which are often fatal in the<br />

first year of life, result from recessive autosomal mutations that fall into nine<br />

complementation groups (mut, cblABCDEFGH) (Qureshi et al. 1994, Kapadia 1995,<br />

Rosenblatt and Fenton 1999, Watkins and Rosenblatt 2001). The cblF, cblC and cblD<br />

defects lead to combined disease. Prior studies indicated that cblF mutations affect<br />

cobalamin transport from the lysosome to the cytoplasm, while cblC and cblD mutations<br />

impair the conversion of complex precursors to cob(II)alamin (Mellman et al. 1979,<br />

Watkins and Rosenblatt 2001). The cblE and cblG complementation groups correspond


64<br />

to the genes for MSR (MTRR) and MS (MTR), respectively, and defects in either gene<br />

results in homocystinuria (Cooper and Rosenblatt 1987, Watkins and Rosenblatt 2001).<br />

The mut group corresponds to the gene that encodes MCM (MUT). The cblA group<br />

(MMAA gene) encodes a protein proposed to protect MCM from inactivation (Korotkova<br />

and Lidstrom 2004) and recent studies have shown that the cblB complementation group<br />

(MMAB gene) encodes a mitochondrial ATP cob(I)alamin adenosyltransferase (ATR)<br />

(Dobson et al. 2002, Leal et al. 2003). Mutations in the cblH complementation group<br />

also result in methylmalonic aciduria, but the nature of the underlying defect and the<br />

corresponding gene are unknown (Watkins and Rosenblatt 2001).<br />

The human enzyme that catalyzes the reduction of cob(II)alamin to cob(I)alamin<br />

for AdoCbl synthesis has not been identified. Recent studies showed that MSR mediates<br />

cob(II)alamin reduction for MS activation (Leclerc et al. 1999, Olteanu and Banerjee<br />

2001), raising the possibility that this enzyme might play a role in AdoCbl synthesis.<br />

Here we describe the purification and initial biochemical characterization of the human<br />

ATR from recombinant E. coli and demonstrate that in vitro human MSR can reduce<br />

cob(II)alamin to cob(I)alamin for AdoCbl synthesis by the ATR enzyme.<br />

Chemicals and Reagents<br />

Materials and Methods<br />

Restriction enzymes and T4 DNA ligase were from New England Biolabs<br />

(Beverly, MA). Titanium III citrate was prepared as previously described (Bobik and<br />

Wolfe 1989). AdoCbl, HOCbl, β-nicotinamide adenine dinucleotide phosphate reduced<br />

form (NADPH), iodoacetic acid (IA), flavin adenine dinucleotide (FAD), and bovine<br />

serum albumin (BSA) were purchased from Sigma Chemical Company (St. Louis, MO).<br />

Isopropyl-ß-D-thiogalactopyranoside (IPTG), ultra-pure ammonium sulfate, adenosine


65<br />

5’- triphosphate (ATP), and dithiothreitol (DTT) were purchased from ICN Biomedicals,<br />

Inc. (Aurora, Ohio). Pefabloc SC PLUS was from Roche Diagnostics (Penzberg,<br />

Germany). All other chemicals and reagents were from Fisher Scientific (Norcross, GA).<br />

Bacterial Strains and Growth Media<br />

The bacterial strains used were E. coli DH5α and BL21 (DE3) RIL (Stratagene, La<br />

Jolla, CA). LB was the rich medium used (Difco, Detroit, MI) (Miller 1972). LB was<br />

supplemented with 25 µg/ml kanamycin and 20 µg/ml chloramphenicol or as indicated.<br />

General Molecular and Protein Methods<br />

Agarose gel electrophoresis and restriction enzyme digests were performed using<br />

standard protocols (Maniatis et al. 1982, Johnson et al. 2001). PCR products and plasmid<br />

DNA were gel purified using QIAquick ® gel extraction kit (QIAGEN, Inc., Valencia,<br />

CA). DNA ligation was accomplished using T4 DNA ligase according to manufacturer’s<br />

instructions. Bacterial transformation and SDS-PAGE were performed as previously<br />

described (Johnson et al. 2001).<br />

ATP:cob(I)alamin Adenosyltransferase Assays<br />

ATR assays were performed as previously reported with some modifications<br />

(Johnson et al. 2001). Reaction mixtures contained: 200 mM Tris-HCl (pH 8.0), 2.8 mM<br />

MgCl2, 10 mM KCl, 0.05 mM HOCbl, 0.4 mM ATP, and 1 mM titanium(III) citrate in a<br />

total volume of 2 ml. Reaction components (except for the ATR) were dispensed into<br />

cuvettes inside an anaerobic chamber (Coy Laboratories, Ann Arbor, MI). The cuvettes<br />

were sealed, removed from the chamber, and incubated at 37°C for 2 minutes. Reactions<br />

were initiated by addition of purified recombinant ATR, and AdoCbl formation was<br />

measured by following the decrease in absorbance at 388 nm (∆ε388 = 24.9 cm -1 mM -1 ).


Construction of ATR Expression Strains<br />

66<br />

Plasmid pNL166 (Leal et al. 2003) was restricted with BglII and HindIII releasing a<br />

645 bp fragment that encodes the ATR enzyme with a lysine at position 239 (ATR<br />

239K). This fragment was gel purified and ligated into a modified pET-41a expression<br />

vector, pTA925 (Johnson et al. 2001). Ligation mixtures were used to transform E. coli<br />

DH5α by electroporation and transformants were selected by plating on LB kanamycin<br />

medium. Pure cultures were obtained from selected transformants and plasmid DNA<br />

isolated from these strains was analyzed by restriction digestion and DNA sequencing.<br />

Clones with the expected DNA sequence were transformed into E. coli strain BL21<br />

(DE3) RIL (Stratagene) for high-level protein production.<br />

PCR was used to construct the ATR polymorphic variant with a methionine at<br />

position 239 (ATR 239M). Plasmid pNL166 (Leal et al. 2003) provided the template<br />

DNA. The primers used for amplification were<br />

5’-GCCGCCAGATCTTATGCCTCAGGGCGTGGAAGACGGG-3’ (forward) and<br />

5’-GCCGCCAAGCTTTCAGAGTCCCTCAGACTCGGCCGATGGGTCATTTTTCAT<br />

GTATATTTTCTCTTGATTCCC-3’ (reverse). These primers introduced BglII and<br />

HindIII restriction sites into the PCR product and were designed to change the lysine at<br />

position 239 to a methionine. Ligation, transformation, and analysis of plasmid DNA<br />

were performed as described above. Clones with the expected DNA sequence were<br />

transformed into E. coli BL21 (DE3) RIL for protein production.<br />

Purification of the Human ATR Variants<br />

Soluble cell extracts of the E. coli expression strains were used for the purification<br />

of both ATR variants. The growth of the cells and the preparation of cell extracts were<br />

carried out as described (Johnson et al. 2001). For ammonium sulfate precipitation, cell


extracts of ATR 239K (160 mg protein) and 239M (180 mg protein) were diluted to 1<br />

mg/ml in 50 mM sodium phosphate (pH 7.0), 300 mM NaCl. Then, ultra-pure<br />

67<br />

ammonium sulfate was added as a fine powder with stirring. Precipitated proteins were<br />

centrifuged for 15 min at 12,000 x gmax using a Beckman JLA-10.500 rotor. The protein<br />

pellets were resuspended in 4 ml of 5 mM potassium phosphate (pH 6.8), and passed<br />

through a 0.45 µm filter. ATR 239K (60 mg) and 239M (35 mg) were applied to a 60 ml<br />

ceramic hydroxyapatite column (Bio-Rad) equilibrated with 5 mM potassium phosphate<br />

(pH 6.8). The column was washed with 60 ml equilibration buffer and eluted with a 600<br />

ml linear gradient of 5 mM to 200 mM potassium phosphate (pH 6.8) at a flow rate of 5<br />

ml/min. Fractions containing ATR of the highest purity were pooled and exchanged into<br />

10 mM potassium phosphate (pH 7.0), 50 mM KCl using a Vivaspin centrifugal<br />

concentrator with a molecular weight cutoff of 10,000 Daltons (Viva Science, Binbrook,<br />

UK). Samples were filtered through a 0.45 µm pore-size membrane and further purified<br />

using a Mono Q HR 10/10 anion exchange column (Amersham Pharmacia Biotech,<br />

Piscataway, NJ). ATR 239K (10 mg) and 239M (9 mg) were applied to the column<br />

which had been equilibrated with 10 mM potassium phosphate (pH 7.0), 50 mM KCl.<br />

Then, the column was washed with 8 ml equilibration buffer and eluted with a 160 ml<br />

linear gradient from 50 to 1000 mM KCl in 10 mM potassium phosphate (pH 7.0) at a<br />

flow rate of 4 ml/min. Protein elution was monitored by following the absorbance of the<br />

column effluent at 280 nm and 8 ml fractions were collected. Purified ATR was<br />

concentrated and stored in 10 mM phosphate buffer (pH 7.0), 130 mM KCl, 50%<br />

glycerol at -20°C.


MSR-ATR Assay<br />

68<br />

The basis of this assay is that MSR converts cob(II)alamin to cob(I)alamin which is<br />

in turn converted to AdoCbl by the ATR enzyme. The conversion of cob(I)alamin to<br />

AdoCbl proceeds with an increase in absorbance at 525 nm that allows quantitation<br />

(∆ε525 = 4.8 cm -1 mM -1 ). Assays contained 200 mM Tris (pH 8.0), 1.6 mM potassium<br />

phosphate, 2.8 mM MgCl2, 100 mM KCl, 0.1 mM cob(II)alamin, 0.4 mM ATP, 1 mM<br />

DTT, 1 mM NADPH, purified MSR and ATR as indicated in the text. The total volume<br />

was 2 ml and anaerobic procedures were used as described for the ATR assay.<br />

Cob(II)alamin was prepared by exposing an anoxic solution of 1 mM AdoCbl to a 150<br />

watt incandescent light at a distance of 15 cm for 30 minutes (Johnson et al. 2001). ATR<br />

239K was purified as described above and MSR was purified as described previously<br />

(Olteanu and Banerjee 2001). Reaction mixtures were incubated at 37°C for 2 min prior<br />

to initiating reactions by the addition of an anoxic solution of NADPH.<br />

Measurement of Cob(I)alamin with Iodoacetate<br />

Iodoacetate reacts rapidly and quantitatively with cob(I)alamin to form<br />

carboxymethylcobalamin (CMCbl). This reaction proceeds with an increase in<br />

absorbance at 525 nm (∆ε525 = 5.1 cm -1 mM -1 ) and hence provides a facile method for<br />

quantitation of cob(I)alamin (Debussche et al. 1991). Anoxic stock solutions of<br />

iodoacetate (40 mM) were prepared and used the same day. The stock solution was<br />

shielded from light using aluminum foil and was added to assay mixtures just prior to the<br />

initiation of reactions.<br />

Separation and Quantification of Cobalamins by HPLC<br />

HOCbl, AdoCbl, and CMCbl were separated and quantified by HPLC using a<br />

NovaPak C18 column (3.9×150 mm) equipped with a C18 Sentry guard column (Waters,


Milford, MA). Samples (200 µl) were loaded onto the column and eluted with 30 ml<br />

69<br />

linear gradient of 10 to 90% methanol in 50 mM sodium acetate (pH 4.6) at a flow rate of<br />

1 ml/min. The absorbance of column effluent was monitored at 365 nm, and analytes<br />

were quantified by comparison of peak areas to a standard curve. The CMCbl standard<br />

was prepared by incubating 2 ml of 200 mM Tris (pH 8.0), 1.6 mM potassium phosphate,<br />

2.8 mM MgCl2, 100 mM KCl, 0.1 mM cob(II)alamin, 1 mM DTT, 0.1 mM FAD, and 0.4<br />

mM iodoacetate at 37°C for 1 hour which resulted in quantitative conversion of<br />

cob(II)alamin to CMCbl (data not shown). Standard curves were prepared using 5, 10,<br />

20, 30, and 40 nmoles of CMCbl, AdoCbl, or HOCbl. Peak areas were integrated using<br />

Waters Breeze software. The standard curves for CMCbl, AdoCbl, and HOCbl, had<br />

regression coefficients (R 2 ) of 0.9985, 0.9993, and 0.9956, respectively. Solutions that<br />

contained CMCbl or AdoCbl were shielded from light to prevent photolysis.<br />

Effect of Ionic Strength on the MSR-ATR System<br />

The ionic strength of MSR-ATR assays was varied by the addition of KCl. Ionic<br />

strength was calculated using the formula I = 1/2Σ(ciz) 2 where I is the ionic strength, ci is<br />

the molar concentration of each type of ion, and z is the charge of each ion (Segel 1976).<br />

For lower values of ionic strength, it was necessary to reduce the Tris-HCl<br />

concentrations.<br />

DNA Sequencing<br />

DNA sequencing was carried out by University of Florida, Interdisciplinary Center<br />

for Biotechnology Research, DNA Sequencing Core facility and University of Florida,<br />

Department of Microbiology and Cell Science, DNA Sequencing facility as described<br />

previously (Bobik et al. 1999).


Purification of the Human ATR Variants<br />

70<br />

Results<br />

The human ATR catalyzes the transfer of a 5'-deoxyadenosyl group from ATP to<br />

the central cobalt atom of cob(I)alamin to form AdoCbl (Chapter 2, Leal et al. 2003).<br />

Two common polymorphic variants of the human ATR having either lysine or<br />

methionine as amino acid 239 (ATR 239M and 239K) were recently shown to be present<br />

in 54% and 46% of 72 control cell lines, respectively (Dobson et al. 2002). We<br />

previously produced ATR 239K as a GST-fusion protein in E. coli (Leal et al. 2003).<br />

Here, subcloning and site-directed mutagenesis were used to construct E. coli expression<br />

strains that produce high levels of ATR 239K and 239M without fusion tags and lacking<br />

their predicted mitochondrial targeting sequences. Both ATR variants were produced at<br />

high levels and purified to apparent homogeneity from soluble cell extracts of<br />

recombinant E. coli. The purification consisted of three steps including ammonium<br />

sulfate precipitation and ceramic-hydroxyapatite and Mono Q chromatography (Table<br />

3-1). Both ATR variants precipitated between 40 and 50% ammonium sulfate saturation<br />

and eluted from the hydroxyapatite and Mono Q columns at 110-135 mM potassium<br />

phosphate and 114-130 mM KCl. ATR 239K and 239M were purified 4.5 and 5.4-fold<br />

with specific activities of 220 and 190 nmol min -1 mg -1 proteins, respectively. The yields<br />

were 14 and 9%. During the purification, inclusion of KCl in chromatography and<br />

storage buffers increased enzyme stability.<br />

SDS-PAGE was used to follow the purification of both ATR variants. Since<br />

purification of both enzymes proceeded similarly, only the gel used to assess the purity of<br />

ATR 239K is depicted (Figure 3-2). Following the final Mono Q chromatography step,<br />

both variants were homogenous following staining with Coomassie Brilliant Blue.


Linearity of the ATR reaction<br />

71<br />

The effect of ATR concentration on activity was determined. Results showed that<br />

the rate of adenosylation was proportional to ATR concentration ranging from 5 µg to<br />

115 µg for both variants and linear regression yielded R 2 values of 0.9995 and 0.9993 for<br />

ATR 239K and 239M, respectively.<br />

ATR Reaction Requirements<br />

To determine the ATR reaction requirements, key assay components were<br />

individually omitted. For both variants, there was no detectable activity in the absence of<br />

ATR, ATP, HOCbl, or titanium (III) citrate (the reducing agent used to produce<br />

cob(I)alamin). In the complete reaction mixture, there was 215 and 194 nmol min -1 mg -1<br />

activity for ATR 239K and 239M, respectively. In the absence of MgCl2, activity<br />

decreased by 20% for both variants. There was no significant difference in activity in the<br />

absence of KCl.<br />

Alternative Nucleotide Donors<br />

The specificity of the ATR variants for ATP, CTP, GTP, UTP, ADP, and AMP was<br />

examined. ATP was found to be the best substrate giving a specific activity of 207 and<br />

197 nmol min -1 mg -1 for ATR 239K and 239M, respectively. These values were set as<br />

100% activity. When CTP, GTP, and UTP were tested as substrates for ATR 239K there<br />

was 9, 16, or 8% activity, respectively (Table 3-2). Similar results were found for ATR<br />

239M (6, 14, or 6% activity with CTP, GTP, and UTP, respectively). For both ATR<br />

variants there was no detectable activity when AMP or ADP was substituted for ATP.<br />

Km and Vmax Values for ATR 239K and 239M<br />

Lineweaver-Burk double-reciprocal plots were used to calculate Km and Vmax values<br />

(Segel 1976). For ATR 239K and ATR 239M, the Km values for ATP were 6.3 µM and


72<br />

6.9 µM, and the Km for cob(I)alamin were 1.2 µM and 1.6 µM, respectively (Figure 3-3).<br />

The Vmax values for ATR 239K and 239M were 250 and 200 nmol min -1 mg -1 ,<br />

respectively. When the Km values for cob(I)alamin were determined, saturating levels of<br />

ATP (500 µM) were added to assay mixtures while varying the concentration of<br />

cob(I)alamin. Similarly, saturating levels of cob(I)alamin (50 µM) were added to assays<br />

when the Km values for ATP were determined. Each value shown in Figure 3-3 is the<br />

average of three measurements of the initial reaction rate at the indicated substrate<br />

concentrations.<br />

MSR Reduces Cob(II)alamin to Cob(I)alamin for AdoCbl Synthesis<br />

The human enzyme that catalyzes the third step of CNCbl assimilation, the<br />

reduction of cob(II)alamin to cob(I)alamin for AdoCbl synthesis by the ATR enzyme, is<br />

unknown. Recently, MSR was shown to catalyze cob(II)alamin reduction for MS<br />

activation (Olteanu and Banerjee 2001). To test whether MSR can also reduce<br />

cob(II)alamin to cob(I)alamin for AdoCbl synthesis, an in vitro assay was used<br />

(MSR-ATR assay). In assays that contained standard components, cob(II)alamin, ATP,<br />

DTT, NADPH, 100 µg purified MSR, and 10 µg purified ATR, AdoCbl was formed at a<br />

rate of 0.96 nmol min -1 . Controls showed that no measurable AdoCbl was formed in the<br />

absence of ATR, MSR, ATP, or cob(II)alamin. When NADPH was excluded from the<br />

reaction mixture, 65% of the total activity remained indicating that DTT can partially<br />

replace NADPH as the electron donor for this reaction. When DTT was omitted, 100%<br />

activity remained with NADPH as the sole electron donor, showing that DTT is not<br />

required for the reaction to occur. In addition, several further tests were used to establish<br />

that AdoCbl was the product of the reaction. The UV-Visible spectrum of completed


73<br />

reactions was characteristic of AdoCbl, and the reaction product was photolyzed by a 30<br />

min exposure to incandescent light with the formation of cob(II)alamin (Figure 3-4).<br />

Moreover, the product of the MSR-ATR reaction co-migrated with authentic AdoCbl by<br />

reverse-phase HPLC following co-injection (not shown). These results show that the<br />

combination of the MSR and ATR enzymes converted cob(II)alamin to AdoCbl. Since<br />

prior studies have shown that the human ATR requires cob(I)alamin for AdoCbl<br />

synthesis and is inactive with cob(II)alamin, these results indicate that MSR reduced<br />

cob(II)alamin to cob(I)alamin for AdoCbl synthesis by ATR.<br />

Cob(I)alamin is Sequestered by the MSR-ATR System<br />

Above, we presented evidence that MSR can reduce cob(II)alamin to cob(I)alamin<br />

for adenosylation by the ATR. In principle, there are two ways in which this might occur<br />

(Figure 3-5). MSR could reduce cob(II)alamin to cob(I)alamin directly, which is<br />

subsequently released into solution and diffuses to ATR. Alternatively, MSR could<br />

reduce cob(II)alamin bound to ATR and the interaction between the two proteins could<br />

sequester cob(I)alamin.<br />

To test whether cob(I)alamin was sequestered or released into solution during<br />

AdoCbl synthesis by the MSR-ATR system, a chemical trap for cob(I)alamin was used.<br />

Iodoacetate (which reacts rapidly and quantitatively with cob(I)alamin to form CMCbl)<br />

was added to an MSR-ATR assay at a 1000-fold excess compared to the ATR (400 µM<br />

compared to 0.4 µM). The assay was allowed to proceed to completion. Then, CMCbl<br />

and AdoCbl were resolved and quantitated by reverse-phase HPLC. Results showed that<br />

35% of the cob(I)alamin formed was converted to CMCbl by reaction with iodoacetate<br />

and that 65% was converted to AdoCbl via the ATR enzyme (Table 3-3, line 1). If "free"


cob(I)alamin had been produced under the conditions used, it is expected that the<br />

majority should have been converted to CMCbl by reaction with the large excess of<br />

iodoacetate prior to diffusion to the ATR for conversion to AdoCbl. Hence, the above<br />

74<br />

results indicated that a significant portion of the cob(I)alamin was sequestered (protected<br />

from reaction with iodoacetate) during conversion of cob(II)alamin to AdoCbl by MSR<br />

and ATR.<br />

As a control, assays were performed that were similar to those described above<br />

except that MSR was replaced with the combination of 1 mM DTT plus 50 µM FAD to<br />

chemically generate cob(I)alamin in situ. Chemical reduction of cob(II)alamin<br />

necessarily produces “free” cob(I)alamin which must then diffuse to the ATR before it<br />

can be converted to AdoCbl. Under these conditions, only 10% of the cob(I)alamin<br />

formed was converted to AdoCbl while 90% was converted to CMCbl (Table 3-3, line 2).<br />

This was in contrast to results obtained when MSR was used to generate cob(I)alamin for<br />

the ATR enzyme where 65% of the cob(I)alamin formed was converted to AdoCbl<br />

(Table 3-3, line1). Hence these results indicate that the MSR-ATR system sequesters<br />

cob(I)alamin during the conversion of cob(II)alamin to AdoCbl.<br />

Interestingly, further studies showed that iodoacetate had relatively little effect on<br />

the total amount of AdoCbl formed by the MSR-ATR system. In the absence or presence<br />

of iodoacetate, 94 or 85 nmol of AdoCbl was produced, respectively (Table 3-3). This<br />

indicated that the majority of the cob(I)alamin produced by the MSR-ATR system was<br />

sequestered.


75<br />

MSR Produces little Cob(I)alamin in the Absence of the ATR<br />

Iodoacetate was also used to test whether MSR can produce “free” cob(I)alamin in<br />

the absence of ATR. Assays were prepared that were similar in composition to the<br />

MSR-ATR assay except that ATR was replaced by iodoacetate. During the first 10 min<br />

of the reaction, cob(I)alamin was formed at a very low rate (0.02 nmol min -1 )<br />

(Figure 3-6). Then, at the 10 min time point, 10 µg of purified ATR was added to the<br />

assay and the rate of cob(I)alamin formation increased approximately 46-fold. These<br />

findings showed that the presence of the ATR greatly enhanced the production of<br />

cob(I)alamin by MSR suggesting a physical interaction between the MSR and ATR.<br />

Control experiments showed that BSA had no effect on the rate of cob(I)alamin<br />

formation when substituted for ATR in the second phase of the reaction shown in<br />

Figure 3-6. Moreover, the controls showed that iodoacetate was working effectively<br />

under our assay conditions. In similar reaction mixtures containing DTT and FAD, the<br />

cob(I)alamin formed was efficiently trapped by iodoacetate (data not shown).<br />

Stoichiometry of the MSR-ATR system<br />

The dependence of ATR activity on the molar ratio of MSR/ATR was determined<br />

at ratios from 0 to 40 (Figure 3-7). Maximal activity occurred at a ratio of approximately<br />

4 moles MSR (which is a monomer) per mol ATR monomer. This ratio is similar to the<br />

optimal stoichiometry for activation of MS by MSR and is a reasonable value for a<br />

physiological process (Olteanu and Banerjee 2001).<br />

Ionic Strength Dependence of Cobalamin Reduction and Adenosylation by the<br />

MSR-ATR System<br />

Variation of the ionic strength from 15 to 720 mM had relatively little effect on the<br />

synthesis of AdoCbl by the MSR-ATR system (Figure 3-8). Within this range, greater


76<br />

than 65% activity was retained. This is in contrast to the relatively narrow ionic strength<br />

dependence of MS activation by MSR (Olteanu and Banerjee 2001).<br />

Discussion<br />

Prior studies identified the human ATR and showed that defects in its encoding<br />

gene (MMAB) underlie cblB methylmalonic aciduria (Dobson et al. 2002, Leal et al.<br />

2003). Investigations also identified two common polymorphic variants of the ATR that<br />

are found in normal individuals (Dobson et al. 2002). Here, both ATR variants were<br />

expressed in E. coli, purified, and found to have similar kinetic properties. The specific<br />

activities of variants 239K and 239M were 250 and 200 nmol min -1 mg -1 , and the Km<br />

values were 6.3 and 6.9 µM for ATP and 1.2 and 1.6 µM for cob(I)alamin, respectively.<br />

These values are roughly similar to those previously reported for bacterial ATR enzymes<br />

where specific activities range from 53 to 619 nmol min -1 mg -1 , and Km values from 2.8 to<br />

110 µM for ATP and from 3 to 5.2 µM for cob(I)alamin (Suh and Escalante-Semerena<br />

1995, Johnson et al. 2001, Saridakis et al. 2004). The Km values for the human ATR are<br />

appropriate to its physiological role since cellular levels of cobalamin are typically 1 µM<br />

or less in higher organisms and ATP concentrations are in the low mM range (Schneider<br />

and Stroinski 1987). Thus, the results reported here show that both common ATR<br />

variants are essentially wild-type and have kinetic properties sufficient to meet cellular<br />

needs for AdoCbl synthesis.<br />

Three classes of ATRs that are unrelated in amino acid sequence have been<br />

identified, the PduO-type, CobA-type, and EutT-type (Johnson et al. 2001). Members of<br />

the CobA-type and the PduO-type have been purified, partially characterized, and their<br />

crystal structures determined (Suh and Escalante-Semerena 1995, Bauer et al. 2001,


Johnson et al. 2001, Saridakis et al. 2004). The CobA enzyme has a relatively broad<br />

specificity for nucleotide donors. It is active with ATP, CTP, GTP, UTP, and ITP and<br />

contains a modified P-loop for nucleotide binding (Bauer et al. 2001). In contrast, the<br />

77<br />

Thermoplasma acidophilum ATR which is a PduO-type enzyme and lacks a recognizable<br />

P-loop, uses only ATP and deoxy-ATP as nucleotide donors (Saridakis et al. 2004).<br />

Thus, different modes of ATP binding are thought to account for the observed differences<br />

in nucleotide specificity by CobA- and PduO-type enzymes. The human ATR is a<br />

PduO-type enzyme, but results presented here show that its nucleotide specificity is<br />

intermediate to that of CobA and T. acidophilum enzymes. In addition, the Km of the<br />

human and T. acidophilum enzymes for ATP is 7 versus 110 µM suggesting some<br />

differences between the two enzymes and the mode of ATP binding which might be<br />

expected since these enzymes are only 32% identical in amino acid sequence.<br />

Results reported here also showed that an in vitro system containing the human<br />

MSR and ATR converted cob(II)alamin to AdoCbl. In this system, MSR reduced<br />

cob(II)alamin to cob(I)alamin which in turn was converted to AdoCbl by ATR. When<br />

0.2 µM purified ATR and 0.6 µM purified MSR was used, the rate of AdoCbl formation<br />

was 0.96 nmol min -1 . It is probable that this level of activity is sufficient to meet<br />

physiological needs. Cells require only very small quantities of AdoCbl. Typical<br />

intracellular levels are about 1 µM whereas other well-known coenzymes are present at<br />

100 to 1000-fold higher concentrations or more (Schneider and Stroinski 1987).<br />

Furthermore, the MSR-ATR system is about 10-fold more active than the analogous<br />

bacterial system which consists of the Fpr, FldA, and CobA proteins (Fonseca and<br />

Escalante-Semerena 2001). Thus, the MSR-ATR system mediates the conversion of


78<br />

cob(II)alamin to AdoCbl at a physiologically relevant rate. To our knowledge, this is the<br />

first reported evidence that MSR can catalyze the reduction of cob(II)alamin to<br />

cob(I)alamin for AdoCbl synthesis.<br />

In addition to its role in AdoCbl synthesis, prior studies showed that MSR also<br />

reduces MS-bound cob(II)alamin to cob(I)alamin for the reductive activation of MS<br />

(Leclerc et al. 1998, Olteanu and Banerjee 2001). It is known that MS activation occurs<br />

in the cytoplasm but that AdoCbl synthesis occurs in the mitochondrion. Therefore, to<br />

carry this dual function MSR must reside in both compartments. Interestingly, recent<br />

studies have indicated that alternative transcript splicing leads to the production of MSR<br />

enzymes with and without mitochondrial targeting sequences providing further evidence<br />

of a dual role of this enzyme (Leclerc et al. 1999).<br />

Findings reported here which indicate dual functionality for MSR bring to light an<br />

interesting parallel between cobalamin reduction in bacteria and in humans. Biochemical<br />

evidence indicated that the E. coli flavodoxin (FldA) catalyzes the reduction of<br />

cob(II)alamin to cob(I)alamin for both the reductive activation of MS (MetH) and the<br />

synthesis of AdoCbl by the bacterial ATR (CobA) (Fujii and Huennekens 1974, Fujii et<br />

al. 1977, Fonseca et al. 2002). Similarly, human MSR reduces cob(II)alamin for both<br />

MS activation (Leclerc et al. 1998, Olteanu and Banerjee 2001) and AdoCbl synthesis by<br />

the human ATR (this study). Thus, although the human MSR and ATR lack significant<br />

sequence similarity to their bacterial counterparts (FldA and CobA), evolution appears to<br />

have driven analogous dual physiological roles for both FldA and MSR enzymes.<br />

In this report, we also conducted studies to determine whether cob(I)alamin was<br />

sequestered or released “free” in solution during the conversion of cob(II)alamin to


AdoCbl by the MSR-ATR system. Experiments in which iodoacetate was used as a<br />

chemical trap indicated that cob(I)alamin was sequestered. This is likely to be<br />

79<br />

physiologically important. Cob(I)alamin is one of the strongest nucleophiles that exists<br />

in aqueous solution and an extremely strong reductant (E°' = -0.61 V) (Schrauzer et al.<br />

1968, Schrauzer and Deutsch 1969, Banerjee et al. 1990). It oxidizes instantaneously in<br />

air and rapidly reduces protons to H2 gas at pH 7 (Schneider 1987). Hence, within the<br />

cells, sequestration of cob(I)alamin would be important by preventing nonspecific or<br />

deleterious side reactions.<br />

The finding that cob(I)alamin was sequestered during the conversion of<br />

cob(II)alamin to AdoCbl also indicates a physical interaction between the MSR and<br />

ATR. This raises the question of whether MSR interacts with both MS and ATR by a<br />

conserved mechanism. Prior studies of the MSR-MS reaction indicated that the optimal<br />

stoichiometry is ~4 MSR/MS and that activity is maximal over a narrow range of ionic<br />

strength indicating that electrostatic interactions are important to the MSR-MS interaction<br />

(Olteanu and Banerjee 2001). Here we found that the optimal ratio for the MSR-ATR<br />

system is also ~4 MSR/ATR, but in contrast to the MSR-MS system, there was no strict<br />

dependence on ionic strength. Hence, although MSR has a conserved function in both<br />

systems (cob(II)alamin reductase), the details of the protein-protein interactions between<br />

MSR and either MS or ATR are apparently somewhat different.<br />

The investigations reported here support prior studies that indicate redundant<br />

systems for mitochondrial cobalamin reduction in mammals. Several lines of evidence<br />

indicated a role for MSR in the reduction of cob(II)alamin to cob(I)alamin for AdoCbl<br />

synthesis: the MSR and ATR enzymes converted cob(II)alamin to AdoCbl at a significant


80<br />

rate; the stoichiometry needed for maximal activity was reasonable (4 MSR/ATR); and<br />

cob(I)alamin was sequestered as is expected for the physiologically relevant system. On<br />

the other hand, prior investigations showed that patients with inherited defects in MSR<br />

(cblE disorder) generally have homocystinuria with the exception of one reported case<br />

resulting in combined homocystinuria and mild methylmalonic aciduria (Wilson et al.<br />

1999). This indicates that in vivo cob(II)alamin reduction for AdoCbl synthesis can occur<br />

independently of MSR. Of the known complementation groups that result in<br />

methylmalonic aciduria, only the cblH and cblA groups have phenotypes consistent with<br />

a role in mitochondrial cobalamin reduction. Recent studies have indicated that the cblA<br />

group functions to protect MCM from inactivation (Korotkova and Lidstrom 2004) and<br />

there has only been one reported case of cblH methylmalonic aciduria (Watkins et al.<br />

2000, Dobson et al. 2002). Thus at this time, there are no strong indications as to the<br />

identity of the alternative reductase(s). Redundant systems for cobalamin metabolism are<br />

known in bacterial systems (Fonseca and Escalante-Semerena 2000, Johnson et al. 2001),<br />

and unpublished results). There, inducible systems provide extra capacity during times of<br />

high demand. Similarly, in humans multiple cobalamin reductases might be required for<br />

an efficient response to physiological stresses of diets that result in higher rates of<br />

propionyl-CoA formation. For example, MSR could be targeted to the mitochondria<br />

under circumstances where the diet is rich in amino acids metabolized via<br />

propionyl-CoA.


propionyl-CoA<br />

PCC<br />

(2S)-methylmalonyl-CoA<br />

MCEE<br />

(2R)-methylmalonyl-CoA<br />

MCM<br />

succinyl-CoA<br />

CENTRAL <strong>METABOLISM</strong><br />

breakdown breakdown (AdoCbl (AdoCbl is is unstable unstable in in vivo) vivo)<br />

β-ligand<br />

transferase<br />

81<br />

CNCbl<br />

GSCbl<br />

cob(III)alamin<br />

reductase<br />

cob(II)alamin<br />

MSR<br />

cob(I)alamin<br />

ATR<br />

AdoCbl<br />

Methionine<br />

synthase<br />

[cob(II)alamin]<br />

MSR<br />

CH CH3THF 3THF +<br />

homocysteine<br />

Methionine<br />

synthase<br />

[CH [CH3Cbl] 3Cbl]<br />

inactivation inactivation<br />

THF +<br />

methionine<br />

Figure 3-1. Propionyl-CoA metabolism, methionine synthesis, and intracellular<br />

cobalamin metabolism. In humans, the pathways shown are needed for the<br />

complete catabolism of compounds degraded via propionyl-CoA and for<br />

recycling homocysteine. Abbreviations: PPC, propionyl-CoA carboxylase;<br />

MCEE, methylmalonyl-CoA epimerase; MCM, AdoCbl-dependent<br />

methylmalonyl-CoA mutase; CNCbl, vitamin <strong>B12</strong>; GSCbl,<br />

glutathionylcobalamin, MSR, methionine synthase reductase; ATR,<br />

ATP:cob(I)alamin adenosyltransferase; THF, tetrahydrofolate.


Table 3-1. ATP:cob(I)alamin adenosyltransferase activity during ATR 239K and 239M purification<br />

Purification step<br />

Total protein<br />

(mg)<br />

Specific activity<br />

(nmol min -1 mg -1 )<br />

Total activity<br />

(nmol min -1 )<br />

Yield<br />

(%)<br />

Purification<br />

(fold)<br />

239K 239M 239K 239M 239K 239M 239K 239M 239K 239M<br />

Cell free extract 160 180 49 35 7840 6300 100 100 1 1<br />

Ammonium<br />

sulfate precipitate<br />

60 35 110 70 6600 2450 84 39 2.2 2<br />

Hydroxyapatite 10 9 118 97 1180 873 15 14 2.4 2.8<br />

Mono Q 5 3 220 190 1100 570 14 9 4.5 5.4<br />

82


kDa<br />

97<br />

66<br />

45<br />

31<br />

21<br />

14<br />

7<br />

83<br />

1 2 3 4<br />

Figure 3-2. Purification of the human ATR. SDS-PAGE was used to assess the<br />

purification of ATR 239K. Lane 1: molecular mass markers, lane 2: 10 µg<br />

of soluble cell free extract from the E. coli expression strain, lane 3: 5 µg of<br />

ammonium sulfate precipitate, lane 4: 4 µg hydroxyapatite eluate, lane 5: 2<br />

µg of Mono Q fraction containing purified ATR. The purification of ATR<br />

239M proceeded similarly (not shown).<br />

5


84<br />

Table 3-2. Use of alternative nucleotide donors by the human ATR<br />

Nucleotide Donor<br />

Activity %<br />

a<br />

ATR 239K ATR 239M<br />

ATP (adenosine-5’-triphosphate) 100 100<br />

ADP (adenosine-5’-diphosphate) ND b ND<br />

AMP (adenosine-5’-monophosphate) ND ND<br />

CTP (cytidine-5’-triphosphate) 9 6<br />

GTP (guanosine-5’-triphosphate) 16 4<br />

UTP (uridine-5’-triphosphate) 8 6<br />

a<br />

ATP is the physiological substrate for the ATR and activity with this nucleotide was set<br />

to 100%.<br />

b ND, none detected


85<br />

Table 3-3. The MSR-ATR system sequesters cob(I)alamin<br />

Variable assay components a<br />

Amount detected (nmol)<br />

AdoCbl CMCbl HOCbl<br />

NADPH, MSR, ATR, and iodoacetate 85 46 45<br />

DTT, FAD, ATR, and iodoacetate 18 165 6<br />

NADPH, MSR, and ATR 94 ND b 81<br />

a Reaction mixtures contained 200 mM Tris (pH 8.0) 1.6 mM KPi, 2,8 mM MgCl2,<br />

100 mM KCl, 0.1 mM (200 nmol) cob(II)alamin, 0.4 mM ATP, and the components<br />

indicated in the table in the following amounts: 10 µg purified ATR, 150 µg purified<br />

MSR, 1 mM NADPH, 1 mM DTT, 50 µM FAD.<br />

Cob(I)alamin was generated enzymatically using NADPH and purified MSR<br />

(lines 1 and 3) or chemically by the combination of DTT and FAD which necessarily<br />

generated “free” cob(I)alamin (line 2).<br />

HOCbl, CMCbl, and AdoCbl were resolved and quantified by HPLC. In the<br />

reverse-phase system used, their retention times were 11.9, 13.8, and 15.3 min,<br />

respectively.<br />

b ND, none detected.


ATR Activity (nmol min-1mg-1 ATR Activity (nmol min ) -1mg-1 )<br />

ATR Activity (nmol min-1mg-1 ATR Activity (nmol min ) -1mg-1 )<br />

A.<br />

200<br />

150<br />

100<br />

1 / Rate<br />

0.012<br />

0.009<br />

0.006<br />

0.003<br />

50<br />

0<br />

0 100<br />

0<br />

0 0.05 0.1 0.15<br />

1 / [ATP]<br />

200 300 400<br />

[ATP], (µM)<br />

0.2<br />

500<br />

150<br />

1 / Rate<br />

0.0120<br />

0.0090<br />

86<br />

250<br />

200<br />

150<br />

100<br />

B.<br />

C. D.<br />

200<br />

200<br />

ATR Activity (nmol min-1mg-1 ATR Activity (nmol min ) -1mg-1 )<br />

ATR Activity (nmol min-1mg-1 ATR Activity (nmol min ) -1mg-1 )<br />

1 / Rate<br />

0.007<br />

0.00525<br />

0.0035<br />

0.00175<br />

0<br />

50<br />

0 0.1 0.2 0.3 0.4 0.5<br />

0<br />

0 10 20<br />

1 / [cob(I)alamin]<br />

30 40 50<br />

[cob(I)alamin], (µM)<br />

100<br />

0.0060<br />

0.0030<br />

120<br />

80<br />

0.006<br />

0.003<br />

50<br />

0<br />

0 100<br />

0<br />

0 0.03 0.05 0.08 0.10 0.13<br />

1 / [ATP]<br />

200 300 400 500<br />

40<br />

0<br />

0 10<br />

0<br />

0<br />

20<br />

0.1 0.2 0.3 0.4<br />

1 / [cob(I)alamin]<br />

30 40<br />

0.5<br />

50<br />

[ATP], (µM) [cob(I)alamin], (µM)<br />

Figure 3-3. Determination of kinetic constants for ATR 239K and ATR 239M. Panels A<br />

and B are for ATR 239K and panels C and D are for ATR 239M. ATR<br />

assays were performed as described in the materials and methods except that<br />

the ATP and cob(I)alamin concentrations were varied as indicated. The<br />

insets are Lineweaver-Burk plots of the data shown in the larger graph.<br />

160<br />

1 / Rate<br />

0.009


Absorbance<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

400<br />

87<br />

440 480 520 560 600<br />

Wavelength, (nm)<br />

Figure 3-4. Absorbance spectra of an MSR-ATR assay. Prior to the addition of NADPH<br />

the spectrum is that of cob(I)alamin (············); 1 hour after the addition of<br />

NADPH and incubation at 37°C, the spectrum is that of AdoCbl (——); after<br />

photolysis, the spectrum is characteristic of cob(II)alamin (⎯ ⎯ ⎯).


cob(II)alamin<br />

NADPH + H +<br />

MSR<br />

cob(II)alamin<br />

NADPH + H +<br />

MSR<br />

cob(II)alamin<br />

NADPH + H +<br />

MSR<br />

NADP + NADP + NADP +<br />

88<br />

Scheme 1: Cob(I)alamin is released in solution and diffuses to the ATR<br />

cob(I)alamin<br />

ATP<br />

ATR<br />

Scheme 2: MSR reduces ATR-bound cob(II)alamin<br />

ATR + cob(II)alamin<br />

ATR—cob(II)alamin<br />

ATR<br />

MSR<br />

tight tight complex<br />

complex<br />

NADPH NADPH NADPH + H +<br />

+<br />

ATP<br />

AdoCbl<br />

+ PPPi<br />

NADP + NADP + NADP +<br />

AdoCbl<br />

+ PPPi<br />

+ MSR<br />

+ ATR<br />

Figure 3-5. Two schemes depicting the conversion of cob(II)alamin to AdoCbl by the<br />

MSR-ATR system. In scheme 1, MSR reduces cob(II)alamin to<br />

cob(I)alamin and releases it free in solution for diffusion to the ATR. This<br />

scheme seems unlikely due to the high nucleophilic nature of cob(I)alamin.<br />

In scheme 2, MSR reduces cob(II)alamin while bound to the ATR thereby<br />

sequestering the highly reactive cob(I)alamin molecule and protecting it from<br />

oxidation and then adenosylation can occur. MSR, a dual flavoprotein,<br />

mediates the one electron transfer to cob(II)alamin from the two-electron<br />

donor, NADPH.


Wavelength, (525 nm)<br />

0.24<br />

0.23<br />

0.22<br />

0.21<br />

89<br />

ATR addition<br />

0.20<br />

0 5 10 15 20<br />

Time, (min)<br />

Figure 3-6. MSR produces little cob(I)alamin in the absence of the ATR enzyme.<br />

During the first phase of the reaction little cob(II)alamin was reduced to<br />

cob(I)alamin. During this phase, iodoacetate was used to detect<br />

cob(I)alamin. Iodoacetate reacts chemically and quantitatively with<br />

cob(I)alamin to form CMCbl and this reaction proceeds with an increase in<br />

absorbance at 525 nm. At the 10 min time point, 10 µg of ATR was added to<br />

the assay mixture and the rate of cob(I)alamin production increased<br />

approximately 46-fold. This large increase in the rate of cob(I)alamin<br />

formation suggested a physical interaction between the MSR and ATR<br />

enzymes. The assay conditions used were similar to those for the MSR-ATR<br />

assay and controls showed that iodoacetate worked effectively under these<br />

conditions.


Activity, (nmol min-1 Activity, (nmol min ) -1 Activity, (nmol min ) -1 )<br />

2.0<br />

1.6<br />

1.2<br />

0.8<br />

0.4<br />

90<br />

0.0<br />

0 10 20 30 40<br />

MSR/ATR,(mole/mole)<br />

Figure 3-7. Stoichiometry of the MSR-ATR system. The MSR-ATR assay was used for<br />

these studies. 10 µg (0.4 nmol) purified ATR was used with the molar ratio<br />

of MSR indicated. Maximal activity required ~4 MSR/ATR.


Activity, (nmol min-1 Activity, (nmol min ) -1 )<br />

4.0<br />

3.0<br />

2.0<br />

1.0<br />

91<br />

0.0<br />

0 120 240 360 480 600 720<br />

Ionic strength, (mM)<br />

Figure 3-8. Effect of ionic strength on the MSR-ATR system. The MSR-ATR assay<br />

containing 40 µg purified ATR and 28 µg purified MSR was used for these<br />

studies. The ionic strength was varied by the addition of KCl.


CHAPTER 4<br />

PDUP IS A COENZYME-A-ACYLAT<strong>IN</strong>G PROPIONALDEHYDE<br />

DEHYDROGENASE ASSOCIATED WITH THE POLYHEDRAL BODIES<br />

<strong>IN</strong>VOLVED <strong>IN</strong> <strong>B12</strong>-DEPENDENT 1,2-PROPANEDIOL DEGRADATION BY<br />

SALMONELLA ENTERICA SEROVAR TYPHIMURIUM LT2<br />

Introduction<br />

Virtually all Salmonella degrade 1,2-propanediol in a coenzyme-<strong>B12</strong>-dependent<br />

manner and this ability (which is absent in Escherichia coli) is thought to be an important<br />

aspect of the Salmonella-specific lifestyle (Roth et al. 1996, Price-Carter et al. 2001).<br />

The degradation of 1,2-propanediol may play a role in the interaction of Salmonella with<br />

its host organisms. In Salmonella enterica, in vivo expression technology (IVET) has<br />

indicated that 1,2-propanediol utilization (pdu) genes may be important for growth in<br />

host tissues, and competitive index studies with mice have shown that pdu mutations<br />

confer a virulence defect (Conner et al. 1998, Heithoff et al. 1999). 1,2-Propanediol is<br />

likely to be abundant in anoxic environments, such as the large intestine, since it is<br />

produced by the anaerobic degradation of the common plant sugars rhamnose and fucose.<br />

In addition, 1,2-propanediol degradation by S. enterica provides an important model<br />

system for understanding coenzyme-<strong>B12</strong>-dependent processes some of which are<br />

important in human physiology, industry, and the environment (Roth et al. 1996).<br />

On the basis of biochemical studies, a pathway for coenzyme-<strong>B12</strong>-dependent<br />

1,2-propanediol degradation by S. enterica was proposed (Figure 1-6). This pathway<br />

begins with the conversion of 1,2-propanediol to propionaldehyde, a reaction that is<br />

catalyzed by coenzyme-<strong>B12</strong>-dependent diol dehydratase (Toraya et al. 1979, Obradors et<br />

92


al. 1988). The aldehyde is then disproportionated to propanol and propionic acid by a<br />

reaction series thought to involve propanol dehydrogenase, coenzyme-A<br />

93<br />

(CoA)-dependent propionaldehyde dehydrogenase, phosphotransacylase, and propionate<br />

kinase (Toraya et al. 1979, Obradors et al. 1988, Palacios et al. 2003). This pathway<br />

generates one ATP, an electron sink, and a 3-carbon intermediate (propionyl-CoA),<br />

which can feed into central metabolism via the methyl-citrate pathway (Horswill and<br />

Escalante-Semerena 1997). In S. enterica, the degradation of 1,2-propanediol occurs<br />

aerobically, or anaerobically when tetrathionate is supplied as an electron acceptor (Price-<br />

Carter et al. 2001).<br />

The genes specifically required for growth of S. enterica on 1,2-propanediol are<br />

organized as a single contiguous cluster named the propanediol utilization (pdu) locus.<br />

Determination of the DNA sequence of this locus showed that 1,2-propanediol<br />

degradation was much more complex than prior biochemical studies had suggested.<br />

DNA sequence analyses indicated the pdu locus included 23 genes (Bobik et al. 1997,<br />

Bobik et al. 1999): six pdu genes are thought to encode enzymes needed for the<br />

1,2-propanediol degradative pathway (Bobik et al. 1997); two are involved in transport<br />

and regulation (Bobik et al. 1992, Chen et al. 1994); two are probably involved in diol<br />

dehydratase reactivation (Bobik et al. 1999); one is needed for the conversion of vitamin<br />

<strong>B12</strong> (CN-<strong>B12</strong>) to coenzyme <strong>B12</strong> (Johnson et al. 2001); five are of unknown function; and<br />

seven share similarity to genes involved in the formation of carboxysomes, polyhedral<br />

bodies found in certain cyanobacteria and some chemoautotrophs (Shively and English<br />

1991, Shively et al. 1998).


94<br />

The finding that the pdu locus included a number of genes similar to those involved<br />

in carboxysome formation suggested that a related polyhedral body might be involved in<br />

<strong>B12</strong>-dependent 1,2-propanediol degradation. Indeed, S. enterica was recently shown to<br />

form polyhedral bodies (also called polyhedral organelles) during growth on<br />

1,2-propanediol (Bobik et al. 1999, Havemann et al. 2002). In general appearance, the<br />

pdu polyhedra are similar to carboxysomes. They are 100-150 nm in cross-section and<br />

are composed of a proteinaceous interior covered by a 3- to 4-nm protein shell.<br />

However, carboxysomes and the pdu polyhedra differ in several ways. Carboxysomes,<br />

which contain most of the cell’s ribulose 1,5-bisphosphate carboxylase/oxygenase<br />

(RuBisCO), function to improve autotrophic growth at low CO2 concentrations and are<br />

thought to do so by concentrating CO2 (Shively et al. 2001, Price et al. 2002). <strong>By</strong><br />

contrast, the pdu polyhedra of S. enterica do not contain RuBisCO. They consist of a<br />

protein shell composed in part of the PduA protein (Havemann et al. 2002) and are<br />

associated with <strong>B12</strong>-dependent diol dehydratase and other unidentified proteins (Bobik et<br />

al. 1999). The function of the S. enterica polyhedral bodies is uncertain. Work published<br />

to date suggests that they may serve to minimize aldehyde toxicity by sequestration and<br />

channeling of propionaldehyde and/or by moderating the rate of aldehyde formation<br />

through control of diol dehydratase activity (Chen et al. 1994, Rondon et al. 1995,<br />

Stojiljkovic et al. 1995, Bobik et al. 1999, Havemann et al. 2002, Havemann and Bobik<br />

2003).<br />

If the pdu polyhedra do indeed function to minimize aldehyde toxicity, it might be<br />

expected that aldehyde-degrading enzymes are also associated with these structures. In<br />

this report, we show that the pduP gene encodes a polyhedral-body-associated


CoA-acylating propionaldehyde dehydrogenase important for 1,2-propanediol<br />

degradation.<br />

95<br />

Materials and Methods<br />

Bacterial Strains, Media, and Growth Conditions<br />

The bacterial strains used in this study are listed in Table 4-1. The rich medium<br />

used was LB medium (Difco, Detroit, MI) (Miller 1972). The minimal medium used was<br />

no-carbon-E (NCE) medium (Vogel and Bonner 1956, Berkowitz et al. 1968)<br />

supplemented with 53 mM propanediol, 1 mM MgSO4, 148 nM CNCbl, and 0.3 mM<br />

each valine, isoleucine, leucine, and threonine. In addition, for strains carrying pLAC22<br />

media were also supplemented with 1 mM IPTG, and 100 µg Amp/ml. Cultures were<br />

incubated at 37°C with shaking at 250 rpm as previously described (Leal et al. 2003), and<br />

cell growth was determined by measuring the optical density of cultures at 600 nm.<br />

General Molecular Methods<br />

Agarose gel electrophoresis was done as described previously (Sambrook et al.<br />

1989). Plasmid DNA was purified by the alkaline lysis procedure (Sambrook et al. 1989)<br />

or by using Qiagen products (Qiagen, Chatsworth, CA) according to the manufacturer's<br />

instructions. Following restriction or PCR amplification, DNA was purified using<br />

Qiagen PCR purification or gel extraction kits. Restriction enzymes were used according<br />

to standard protocols (Sambrook et al. 1989). DNA fragments were ligated using T4<br />

DNA ligase according to the manufacturer's directions. Electroporation was carried out<br />

as previously described (Bobik et al. 1999).<br />

General Protein Methods<br />

SDS-PAGE was performed using Bio-Rad Redigels and Bio-Rad Mini-Protean II<br />

electrophoresis units according to the Manufacturer's instructions. Following gel


96<br />

electrophoresis, proteins were stained with Coomassie Brilliant Blue R-250. The protein<br />

concentration of solutions was determined using Bio-Rad protein assay reagent (Bio-Rad)<br />

(Bradford 1976).<br />

P22 Transduction<br />

Transductional crosses were performed as described using P22 HT105/1 int-210<br />

(Davis et al. 1980), a mutant phage that has high transducing ability (Schmieger 1971).<br />

For the preparation of P22 transducing lysates from strains having galE mutations,<br />

overnight cultures were grown on LB-medium supplemented with 11 mM glucose and 11<br />

mM galactose. Transductants were tested for phage contamination and sensitivity by<br />

streaking on green plates against P22 H5.<br />

Cloning pduP into pLAC22<br />

PCR was used to amplify pduP from the template pMGS2 (Havemann et al. 2002).<br />

The primers used for amplification were 5’-GGAATTCGGATCCTATGAATACTTCTG<br />

AACTCGAAAC-3’ (forward) and 5’-GGAATTCAAGCTTCAGTTAGCGAATAGAAA<br />

AGCC-3’ (reverse). These primers introduced BamHI and HindIII restriction sites that<br />

were used for cloning into pLAC22 cut with BglII and HindIII (Warren et al. 2000).<br />

Cloning was carried out by ligation of the complementary cohesive ends formed by<br />

cutting with BamHI and BglII since pduP contains an internal BglII site. Following<br />

ligation, clones were introduced into S. enterica strain TR6579 by electroporation and<br />

transformants were selected by plating on LB-agar supplemented with ampicillin (Amp)<br />

at 100 µg/ml. Pure cultures were prepared from selected transformants, and plasmid<br />

DNA isolated from these strains was cut with NruI, PstI, and HindIII. The DNA<br />

sequence of one clone that released products of the expected sizes following restriction<br />

(2,635, 3,322 and 943 base pairs) was determined and found to be identical to the


97<br />

previously published DNA sequence of pduP (Bobik et al. 1999). This clone (pNL9) as<br />

well as pLAC22 without insert were moved into strain BE191 by P22 transduction and<br />

the resulting strains (BE270 and BE269) were used for complementation studies.<br />

Construction of His8-PduP<br />

PCR was used to fuse an N-terminal His-tag to the PduP protein. Template<br />

pMGS2 was used with the following primers:<br />

5’-GGGGATCCATG(CAT)8ATGAATACTTCTGAACTCGAAAC-3’ (forward); and<br />

5’-GGAATTCAAGCTTCAGTTAGCGA ATAGAAAAGCC-3’ (reverse). These<br />

primers introduced BamHI and HindIII restriction sites that were used for cloning into<br />

pTA925 cut with BglII and HindIII. The ligation mixture was used to transform E. coli<br />

strain DH5α and transformants were selected by plating on LB agar supplemented with<br />

25 µg kanamycin/ml. Plasmid DNA isolated from selected transformants was analyzed<br />

by restriction mapping and DNA sequencing. One plasmid encoding His8-PduP (pNL69)<br />

was introduced into E. coli expression strain BL21 (DE3) RIL by electroporation to form<br />

expression strain BE273.<br />

Purification of His8-PduP under Denaturing and Nondenaturing Conditions<br />

For purification of PduP under denaturing conditions, E. coli strain BE273 was<br />

grown in 500 ml LB Kan (25 µg/ml) broth incubated at 37°C with shaking at 275 rpm in<br />

a 1-liter baffled Erlenmeyer flask. Cells were grown to an optical density of 0.6-0.8 at<br />

600 nm and PduP production was induced by addition of IPTG to 1 mM. Cells were<br />

incubated for an additional 2 hours, and harvested by centrifugation at 6,690 x g for 10<br />

minutes using a Beckman JLA-10.500 rotor. Two grams of cells (wet weight) were<br />

resuspended in 3 ml of 50 mM sodium phosphate, 300 mM NaCl, pH 7 and broken using


98<br />

a French pressure cell at 20,000 psi (SLM Aminco, Urbana, IL). The protease inhibitor<br />

phenylmethylsulfonylflouride was added to the cell extract to a concentration of 100<br />

µg/ml. To separate soluble and insoluble fractions, cell extract was centrifuged at 31,000<br />

x g for 30 minutes using a Beckman JA-20 rotor. Inclusion bodies were purified from the<br />

insoluble fraction using BPER-II according to the manufacturer’s instructions. Purified<br />

inclusion bodies were solubilized overnight in binding buffer (10 mM imidazole, 500<br />

mM NaCl, 20 mM Tris-HCl pH 7.9, and 6 M urea) and filtered using a 0.45-µm filter.<br />

Affinity purification was carried out using 1-ml Amersham Pharmacia Hi-trap chelating<br />

(Ni 2+ ) column according the manufacturer’s instructions with the following<br />

modifications: 6 M urea was added to all buffers, and two wash steps were used, the first<br />

with 20 mM imidazole and the second with 60 mM imidazole. The PduP protein was<br />

then eluted using 150 mM imidazole.<br />

For purification of PduP under nondenaturing conditions, inclusion bodies (isolated<br />

as described above) were resuspended in 50 mM potassium phosphate, pH 7, 25 mM<br />

NaCl and stored at 4 °C for 24 h. Insoluble proteins were pelleted by centrifugation and<br />

PduP that remained in the supernatant was purified by nickel-affinity chromatography<br />

using Ni-NTA resin (Qiagen) according to the manufacturer's instructions. Briefly, the<br />

column was washed with 5 bed volumes of binding buffer (50 mM sodium phosphate,<br />

300 mM NaCl, pH 7) and PduP was eluted with binding buffer supplemented with 40<br />

mM imidazole.<br />

Propionaldehyde Dehydrogenase Assays<br />

Two assays for propionaldehyde dehydrogenase activity were used. Assay one was<br />

carried out as previously described (Walter et al. 1997). Reaction mixtures contained 50


mM CHES (2-[N-cyclohexylamino]ethanesulfonic acid) pH 9.5, 10 mM<br />

propionaldehyde, 1 mM dithiothreitol (DTT), 75 µM NAD + , 100 µM HS-CoA, and an<br />

99<br />

appropriate amount of cell extract. Assays were incubated at 37°C. Enzyme activity was<br />

measured by following the conversion of NAD + to NADH by monitoring the absorbance<br />

of reaction mixtures at 340 nm, and the amount of NADH formed was determined using<br />

∆ε340 = 6.22 mM -1 cm -1 . Assay two was done as described above except that DTT was<br />

omitted and the absorbance of reaction mixtures was followed at 232 nm. An increase in<br />

absorbance at 232 nm occurs when thioester bonds are formed and ∆ε232 ≅ 4.5 mM -1 cm -1<br />

(Dawson et al. 1969).<br />

Identification of Propionyl-CoA a Product of the Propionaldehyde Dehydrogenase<br />

Reaction<br />

Propionaldehyde dehydrogenase reactions were performed as described above<br />

except that CHES buffer was replaced by 50 mM potassium phosphate, pH 7.<br />

Immediately after the reactions reached completion, they were analyzed by reverse-phase<br />

HPLC using conditions previously described (Bobik and Rasche 2003). For the<br />

purification of reaction products, HPLC solvents were buffered with 10 mM ammonium<br />

formate, pH 6.4. Selected reaction products were collected, lyophilized, resuspended in<br />

distilled water at a concentration of 4 µM. For the MALDI-TOF MS (Matrix Assisted<br />

Laser Desorption Ionization-Time of Flight Mass Spectrometry) analyses, samples were<br />

mixed 1:1 with 10 mg/ml α-cyano-4-hydroxycinnamic acid matrix in 0.1% trifluoroacetic<br />

acid in 30% acetonitrile. The sample (1 µl) was spotted on a MALDI plate and analyzed<br />

using an Applied Biosystems Voyager DE-Pro MALDI-TOF MS operated in reflector<br />

mode with a delay of 100 nsec, acc voltage of 20 KV, and a grid voltage of 71.5%. The


100<br />

sample was irradiated with a nitrogen laser (337 nm) and 100 individual laser shots were<br />

collected for each sample.<br />

Construction of a Nonpolar pduP Deletion<br />

Bases 28 to 1,386 of the pduP coding sequence were deleted via a PCR-based<br />

method (Miller and Mekalanos 1988). The deletion was designed to leave the predicted<br />

translational start and stop signals of all pdu genes intact. The following primers were<br />

used for PCR amplification of the flanking regions of the pduP gene: primer 1,<br />

5’-GCTCTAGACCAGGCCAACATCATCCGTGAAGTTAG-3’; primer 2,<br />

5’-TCATCGCGACCTCAGCAGGGTTTCGAGTTCAGAAGTAATCATTG-3’; primer<br />

3, 5’-CGCTAACTGAGGTCGCGATGAATACC-3’; and primer 4,<br />

5’-GAAGAGCTCAATTCTGCGGCGGTACGCTGACCACC-3’. Primers 1 and 2 were<br />

used to amplify 496 DNA bases upstream of the pduP gene, and primers 3 and 4 were<br />

used to amplify 508 DNA bases downstream of pduP gene. The upstream and<br />

downstream amplification products were purified and then fused by a PCR reaction that<br />

included 1 ng/µl of each product and primers 1 and 4. The fused product was cut with<br />

XbaI and SacI (these sites were designed into primers 1 and 4, respectively), and ligated<br />

to suicide vector pCVD442 (Miller and Mekalanos 1988) that had been similarly cut.<br />

The ligation mixture was used to transform E. coli S17.1 by electroporation and<br />

transformants were selected on LB medium supplemented with Amp (100 µg/ml).<br />

Plasmids were extracted from four of the isolated transformants, screened by restriction<br />

analysis, and all released an insert of the expected size (1004 bp). One of these<br />

transformants was used to introduce the pduP deletion into the S. enterica chromosome<br />

using the procedure of Miller and Mekelanos (1988) with the following modification.


101<br />

For the conjugation step, strain BE47 was used as the recipient and exconjugants were<br />

selected by plating on LB agar supplemented with ampicillin (100 µg/ml) and<br />

chloramphenicol (20 µg/ml). Deletion of the pduP coding sequence was verified by PCR<br />

using chromosomal DNA as a template. Lastly, the thr-480 dCAM insertion used for<br />

selection of exconjugants was crossed-off by P22 transduction using a phage lysate<br />

prepared with the wild-type strain and by selecting for prototrophy on NCE glucose<br />

minimal medium.<br />

Antibody Preparation<br />

His8-PduP purified using Ni 2+ -affinity chromatography was resolved on a 12%<br />

Tris-HCl SDS-PAGE gel. The protein band corresponding to His8-PduP was excised and<br />

used as a source of antigen for the preparation of polyclonal antibody in a New Zealand<br />

white rabbit by Cocalico Biologicals (Reamstown, PA). To eliminate antibodies<br />

cross-reacting with E. coli proteins, the antiserum was preadsorbed using acetone powder<br />

prepared from E. coli BL21DE3 RIL containing pTA925 (the expression strain lacking<br />

pduP) as described previously (Harlow and Lane 1988).<br />

Western Blots<br />

Cultures were grown and prepared for SDS-PAGE and electroblotting as<br />

previously described (Havemann et al. 2002). Membranes were probed using adsorbed<br />

anti-PduP antiserum (diluted 1:3,500 in blocking buffer) as the source of the primary<br />

antibody, and goat anti-rabbit conjugated to alkaline phosphatase (diluted 1:3,000 in<br />

blocking buffer) as the secondary antibody. Chromogenic developing agents were used<br />

following the manufacturer’s instructions (Bio-Rad).


Electron Microscopy<br />

102<br />

For electron microscopy, cells were grown in minimal medium supplemented with<br />

37 mM succinate and 26 mM propanediol. Cultures (10 ml) were incubated in 125-ml<br />

Erlenmeyer flasks at 37º C, with shaking at 275 rpm in a New Brunswick C24 Incubator<br />

Shaker.<br />

For immunogold localization of the PduP protein, cells were prepared as previously<br />

described (Bobik et al. 1999). The source of the primary antibody was rabbit polyclonal<br />

anti-PduP antiserum or preimmune serum diluted in 1:100 in PBS. The secondary<br />

antibody used was goat anti-rabbit IgG conjugated to12-nm colloidal gold (Jackson<br />

ImmunoResearch Laboratories, Inc. West Grove, PA) diluted 1:30 in PBS.<br />

DNA Sequencing and Analysis<br />

DNA sequencing was carried out at the University of Florida Interdisciplinary<br />

Center for Biotechnology Research DNA Sequencing Core Facility using Applied<br />

Biosystems automated sequencing equipment (Perkin Elmer, Norwalk, CT) or at the<br />

University of Florida, Department of Microbiology and Cell Science DNA Sequencing<br />

Facility using a LI-COR model 4000L DNA sequencer, automated sequencing<br />

equipment, and Base ImagIR Analysis Software version 04.1h (LI-COR, Lincoln, NE).<br />

The template for DNA sequencing was plasmid DNA purified using Qiagen 100 tips or<br />

Qiagen mini-prep kits. BLAST software was used for sequence similarity searches<br />

(Altschul et al. 1997).<br />

Chemicals and Reagents<br />

Formaldehyde, (R, S)-1,2-propanediol, and antibiotics were from Sigma Chemical<br />

Company (St. Louis, MO). Isopropyl-β-D-thiogalactopyranoside (IPTG) was from<br />

Diagnostic Chemicals Limited (Charlotteville PEI, Canada). Restriction enzymes were


103<br />

from New England Biolabs (Beverly, MA) or Promega (Madison, WI). T4 DNA ligase<br />

was from New England Biolabs. Glutaraldehyde was from Tousimis (Rockville, MD),<br />

and uranyl acetate was from EM Sciences (Washington, PA). LR White resin was from<br />

Ted Pella Inc. (Redding, CA). Acrylamide, agarose, ammonium persulfate, Coomassie<br />

Brilliant Blue R-250, EDTA, ethidium bromide, 2-mercaptoethanol, N,<br />

N-bis-methylene-acrylamide, powdered milk, SDS, and TEMED were from Bio-Rad<br />

(Hercules, CA). Bacterial Protein Extraction Reagent II (BPER-II) was from Pierce<br />

(Rockford, IL). Other chemicals were from Fisher Scientific (Pittsburgh, PA).<br />

Results<br />

Effect of a Precise pduP Deletion on the Growth of S. enterica on Minimal<br />

1,2-Propanediol Medium<br />

Prior biochemical studies indicated that a CoA-acylating propionaldehyde<br />

dehydrogenase was involved in the degradation of 1,2–propanediol by S. enterica<br />

(Toraya et al. 1979, Obradors et al. 1988). Recent DNA sequence analyses of the pdu<br />

operon (a cluster of genes required for 1,2-propanediol degradation by S. enterica)<br />

showed that the PduP protein has sequence similarity to a number of CoA-acylating<br />

aldehyde dehydrogenases and is 32% identical in amino acid sequence to the E. coli<br />

aldehyde/alcohol dehydrogenase, AdhE (Kessler et al. 1991, Bobik et al. 1999). This<br />

suggested that PduP is a CoA-acylating propionaldehyde dehydrogenase needed for the<br />

degradation of 1,2-propanediol by S. enterica.<br />

To examine the role of PduP in 1,2-propanediol degradation, a precise deletion of<br />

the pduP gene was constructed using a PCR-based method, and the effects of this<br />

mutation on the growth of S. enterica on 1,2-propanediol/CN-<strong>B12</strong> minimal medium were<br />

examined. The wild-type strain and strain BE191 (∆pduP) grew with generation times of


104<br />

about 7.4 and 21.7 hours, respectively, and reached maximum optical densities at 600 nm<br />

of about 1.38 and 0.49, respectively. This showed the ∆pduP mutant was significantly<br />

impaired for growth on 1,2-propanediol/CNCbl minimal medium.<br />

As a control, we tested whether the growth defect of strain BE191 (∆pduP) could<br />

be corrected by the expression of pduP in trans. The generation times of strains BE270<br />

(∆pduP/pLAC22-pduP) and BE269 (∆pduP/pLAC22-no insert) on<br />

1,2-propanediol/CNCbl minimal medium were 3.2 and 25.4 hours, respectively, and the<br />

maximum optical densities reached by these strains were 1.7 for BE270 and 0.45 for<br />

BE269. Hence, expression of pduP in trans corrected the growth defect of ∆pduP<br />

mutation on 1,2-propanediol/CNCbl minimal medium. This result indicated that the<br />

observed growth defect of BE191 was due to deletion of pduP, but not to polarity or a<br />

mutation inadvertently introduced during strain construction. This was the expected<br />

result since BE191 was constructed by a PCR-based method designed to produce a<br />

precise nonpolar deletion. Strain BE270 was found to grow about 2.2-fold faster than did<br />

the wild-type strain. It appeared that pLAC22 enhanced growth of this strain on<br />

1,2-propanediol minimal medium for unknown reasons, since the wild-type strain also<br />

grew faster when it carried this vector without insert (not shown). Growth tests were<br />

repeated twice with similar results. The fact that BE191 (∆pduP) was impaired for<br />

growth on 1,2-propanediol/CNCbl minimal medium, in conjunction with the control<br />

experiments described above indicates that pduP plays an important role in<br />

1,2-propanediol degradation in vivo.


105<br />

Propionaldehyde Dehydrogenase Activity in Wild-type S. enterica and a ∆pduP<br />

Mutant<br />

To test whether the pduP gene encodes a propionaldehyde dehydrogenase, enzyme<br />

assays were performed on cell extracts of wild-type S. enterica and strain BE191<br />

(∆pduP). The ∆pduP mutant used was the same strain used for the growth studies<br />

described above, and assays were conducted on both the supernatants and the pellets that<br />

resulted from centrifugation of cell extracts at 31,000 x g (Table 4-2). For the wild-type<br />

strain, the majority of the propionaldehyde dehydrogenase activity was found in the<br />

31,000 x g pellet which had a specific activity 1.15 µmol min -1 mg -1 protein. Similar<br />

extracts from the pduP mutant had only 1% of the specific activity of the wild-type strain<br />

(0.011 µmol min -1 mg -1 ). When the substrates, propionaldehyde, NAD + or HS-CoA were<br />

omitted from the assay mixture, no activity was detected. To our knowledge, this was the<br />

first experimental evidence that the pduP gene encodes a CoA-acylating propionaldehyde<br />

dehydrogenase.<br />

It was of interest that the majority of propionaldehyde dehydrogenase activity was<br />

found in the 31,000 x g pellet from the wild-type strain (99%). Prior studies showed that<br />

polyhedral bodies are involved in AdoCbl-dependent 1,2-propanediol degradation by<br />

S. enterica (Bobik et al. 1999, Havemann et al. 2002). Hence, the finding that the PduP<br />

aldehyde dehydrogenase was associated with the 31,000 x g pellet suggested that it might<br />

be associated with the polyhedral bodies which pellet under similar centrifugation<br />

conditions (unpublished results).<br />

High-level Production of the PduP Protein<br />

E. coli strain BE273 (pET41a-His8-pduP) was constructed to produce high levels of<br />

recombinant His8-PduP protein. Protein production by this strain as well as by the


106<br />

control strain BE237 (which is isogenic to BE273 except that it contains the expression<br />

plasmid without insert) was analyzed by SDS-PAGE (Figure 4-1). Both the soluble and<br />

inclusion-body fractions of cell extracts were examined. High amounts of a protein with<br />

a molecular mass near 50 kDa were found in the inclusion-body fraction of cell extracts<br />

from expression strain BE273, and moderate amounts of a 50-kDa protein were also<br />

found in the soluble fraction of cell extracts from this strain (Figure 4-1, lanes 3 and 5).<br />

This is near the predicted molecular mass of His8-PduP (50 kDa). In contrast, the soluble<br />

and inclusion-body fractions from control strain BE237 (expression vector without insert)<br />

contained relatively little protein of 50 kDa molecular mass (Figure 4-1, lanes 2 and 4)<br />

indicating that the observed 50-kDa protein produced by BE273 was His8-PduP.<br />

The cell extracts analyzed by SDS-PAGE (Figure 4-1) were tested for<br />

propionaldehyde dehydrogenase activity. Substantial activity (2.4 µmol min -1 mg -1<br />

protein) was found in inclusion bodies isolated from strain BE273, and in the soluble<br />

fraction of this strain (0.19 µmol min -1 mg -1 protein). However, little activity was found<br />

in either the soluble or inclusion-body fractions of control strain BE237 (0.02 µmol min -1<br />

mg -1 protein). For the inclusion-body fraction from the expression strain,<br />

propionaldehyde dehydrogenase activity was linear with protein concentration (data not<br />

shown). Furthermore, when the substrates, propionaldehyde, NAD + or HS-CoA were<br />

omitted from the assay mixture, no activity was detected. Thus, these data provided<br />

additional evidence that PduP is a CoA-acylating propionaldehyde dehydrogenase.<br />

Purification of Recombinant His8-PduP Protein<br />

It was observed that following storage of purified inclusion bodies at 4 °C for 24<br />

hours a significant amount of active His8-PduP protein was eluted into the soluble


107<br />

fraction. Following centrifugation of this preparation at 31,000 x g for 30 min.,<br />

approximately 0.5 mg of soluble protein remained in the supernatant. Enzyme assays<br />

showed that this supernatant contained 9.22 µmol min -1 mg -1 propionaldehyde<br />

dehydrogenase activity. This sample was further purified by nickel-affinity<br />

chromatography. The purity of the resulting preparation was analyzed by SDS-PAGE<br />

followed by staining with Coomassie. Results indicated that the His8-PduP obtained was<br />

highly purified (Figure 4-1, lane 6) and enzyme assays showed that the preparation was<br />

enriched in propionaldehyde dehydrogenase activity, with a specific activity of 15.2 µmol<br />

min -1 mg -1 protien. The reaction requirements for purified PduP were the same as those<br />

described above for the partially purified enzyme. These results demonstrated that the<br />

His8-PduP protein had propionaldehyde dehydrogenase activity.<br />

Propionyl-CoA is a Product of the PduP Reaction<br />

The fact that CoA-SH was required for the propionaldehyde dehydrogenase activity<br />

of PduP suggested that propionyl-CoA was a reaction product. Several additional<br />

experiments were conducted to test this possibility. Propionaldehyde dehydrogenase<br />

reactions containing purified PduP protein were allowed to proceed to completion. Then,<br />

reverse-phase HPLC was used to identify CoA-SH and CoA-derivatives present in<br />

reaction mixtures. A single CoA compound with a retention time of 10.2 min was<br />

detected. Further analyses showed that this compound co-eluted with authentic<br />

propionyl-CoA following co-injection and resolution via C18 reverse-phase HPLC. This<br />

indicated that propionyl-CoA was a product of the PduP aldehyde dehydrogenase<br />

reaction. Next, the reaction product identified as propionyl-CoA by reverse-phase HPLC<br />

was analyzed by MALDI-TOF MS. Three major peaks at m/z 824.8, 848.1, and 878.1


108<br />

were observed in the mass spectrum. These peaks correspond to the [M+H] + ,<br />

[M+H+Na] + , and [M+3NH4] + of propionyl-CoA, further supporting the assignment of<br />

propionyl-CoA as a PduP reaction product. In addition, the propionaldehyde<br />

dehydrogenase reaction catalyzed by the PduP enzyme was followed<br />

spectrophotometrically at 232 nm, the wavelength at which thioester bonds<br />

characteristically absorb (Dawson et al. 1969). Within experimental error, the specific<br />

activity of purified PduP was the same when reactions were followed at 340 nm (NAD +<br />

reduction) or 232 nm (thioester bond formation). Thus, based on the evidence described<br />

above, we conclude that propionyl-CoA is a product of the PduP propionaldehyde<br />

dehydrogenase reaction.<br />

Preparation and Specificity of the Anti-PduP Antiserum<br />

To obtain the antigen needed for preparation of antisera, His8-PduP was purified<br />

from inclusion bodies isolated from expression strain BE273 by Ni 2+ -affinity<br />

chromatography under denaturing conditions. SDS-PAGE indicated that the PduP<br />

protein obtained was highly purified (data not shown). Denaturing conditions were used<br />

to obtain antigen because a nondenaturing protocol was unavailable at the time.<br />

Following purification, His8-PduP was used to prepare polyclonal anti-PduP antiserum in<br />

rabbit. The antiserum obtained was subjected to a preadsorption procedure to remove<br />

cross-reacting proteins (Warren et al. 2000). To determine the specificity of the adsorbed<br />

anti-PduP antiserum, Western blot analysis was done on boiled cell lysates. The<br />

anti-PduP antibody preparation recognized one major protein band in extracts from<br />

wild-type S. enterica at approximately 50 kDa (Figure 4-2, lane 2); however, this band<br />

was not detected in strain BE191, which contained a nonpolar pduP deletion, but was<br />

otherwise isogenic to the wild-type strain (Figure 4-2, lane 2). This indicated that the


109<br />

antiserum obtained was specific for the PduP protein. Furthermore, the anti-PduP<br />

antibody preparation detected a 50-kDa band in boiled cell extracts from PduP expression<br />

strain BE270 (pLAC22-PduP) but not in extracts from strain BE269 which carried the<br />

pLAC22 expression plasmid without the insert but is otherwise isogenic to strain BE270<br />

(Figure 4-2, lanes 4 and 5). This provided additional evidence that the antibody<br />

preparation was specific for the PduP protein. The minor band observed in Figure 4-2,<br />

lane 5, was most likely a degradation product of PduP that resulted from overproduction<br />

since this band was not detected in extracts from the wild-type strain (Figure 4-2, lane 2).<br />

Localization of PduP Immunoelectron Microscopy<br />

Wild-type S. enterica and BE191 (∆pduP) were grown on<br />

succinate/1,2-propanediol minimal medium which induces formation of the polyhedral<br />

bodies involved in 1,2-propanediol degradation (Bobik et al. 1999, Havemann et al.<br />

2002). Immunogold labeling of wild-type S. enterica and BE191 (∆pduP) was then<br />

carried out using anti-PduP antiserum described above. In the micrograph (Figure 4-3),<br />

the antibody-conjugated gold particles (solid black circles) indicate the location of the<br />

PduP protein. Note that the gold particles localized to the polyhedral bodies (which<br />

appear as uniformly stained regions within the cytoplasm) in the wild-type strain, but no<br />

labeling was observed in the ∆pduP mutant, although polyhedra were present. These<br />

results indicated the antibody preparation reacted specifically with the PduP protein<br />

under the labeling conditions used and that the PduP protein localizes to the polyhedral<br />

bodies. Additional electron microscopy studies of standard thin sections showed that the<br />

∆pduP mutant formed normal-appearing polyhedra (standard thin sections result in<br />

higher contrast than does fixation for immunolabeling and the fine structure of the


110<br />

polyhedra is more easily discerned). This result indicated that PduP is nonessential for<br />

polyhedral body formation.<br />

Propionaldehyde Dehydrogenase Activity is associated with Purified pdu Bodies<br />

To further investigate the subcellular localization of the PduP enzyme, the<br />

polyhedral bodies involved in 1,2-propanediol degradation were purified and<br />

propionaldehyde dehydrogenase activity was followed during the course of their<br />

purification (Table 4-3). The polyhedra were purified by a combination of detergent<br />

treatment and differential and density-gradient centrifugation as described in Havemann<br />

and Bobik (2003). Electron microscopy and SDS-PAGE indicated that the polyhedra<br />

obtained were highly purified (not shown). During the purification the specific activity<br />

of the PduP propionaldehyde dehydrogenase increased 13.5-fold from 0.3 - 4.0 µmol<br />

min -1 mg -1 protein (Table 4-3). This suggested that the polyhedral bodies comprised<br />

about 7% of the total cell protein, which is consistent with previous electron microscopy<br />

(Havemann et al. 2002). Furthermore, Western blots verified that the PduP enzyme was<br />

associated with the purified polyhedra (not shown). Thus, enzyme assays and Western<br />

blots with purified polyhedra supported the immunolabeling studies described above and<br />

indicated that the PduP propionaldehyde dehydrogenase is associated with the polyhedral<br />

bodies involved in 1,2-propanediol degradation.<br />

Discussion<br />

Prior biochemical studies indicated that a CoA-acylating propionaldehyde<br />

dehydrogenase was required for AdoCbl-dependent 1,2-propanediol degradation (Toraya<br />

et al. 1979, Obradors et al. 1988). Subsequent DNA sequence analyses of the pdu operon<br />

identified a presumptive enzyme (PduP) that was 32% identical to the E. coli<br />

aldehyde/alcohol dehydrogenase, AdhE (Bobik et al. 1999). In this report, genetic and


111<br />

biochemical evidence was presented that established PduP as a CoA-acylating<br />

propionaldehyde dehydrogenase involved in AdoCbl-dependent 1,2-propanediol<br />

degradation by S. enterica.<br />

Previously, Jorge Escalante's group proposed that two propionaldehyde<br />

dehydrogenases might be involved in 1,2-propanediol degradation by S. enterica: one<br />

that produces propionyl-CoA and a second that oxidizes propionaldehyde directly to<br />

propionate (Palacios et al. 2003). The PduP enzyme studied here produced<br />

propionyl-CoA as a reaction product and was inactive in the absence of HS-CoA showing<br />

that it is incapable of oxidizing propionaldehyde directly to propionate under the assay<br />

conditions used. It was also shown that an S. enterica strain with a nonpolar pduP null<br />

mutation grew at one-third the rate of the wild-type strain. The residual growth of a<br />

∆pduP mutant could have resulted from the activity of an alternative propionaldehyde<br />

dehydrogenase, such as the one proposed to convert propionaldehyde directly to<br />

propionate (Palacios et al. 2003), since the propionate produced by this enzyme could be<br />

used as a carbon and energy source via the methyl citrate pathway (Horswill and<br />

Escalante-Semerena 1997).<br />

Previously, we established that S. enterica formed polyhedral bodies during<br />

AdoCbl-dependent growth on 1,2-propanediol (Bobik et al. 1999). Immunolabeling<br />

studies demonstrated that these bodies consisted of a protein shell (partly composed of<br />

the PduA protein) AdoCbl-dependent diol dehydratase and additional unidentified<br />

proteins (Bobik et al. 1999, Havemann et al. 2002). In this report, we presented several<br />

lines of evidence that indicated PduP is polyhedral-body-associated. The finding that the<br />

pdu polyhedra include both diol dehydratase (aldehyde-producing) and propionaldehyde


112<br />

dehydrogenase (aldehyde-consuming) is consistent with the prior proposal that these<br />

structures function to minimize aldehyde toxicity (Chen et al. 1994, Stojiljkovic et al.<br />

1995, Havemann et al. 2002). Presumably, the shell of the polyhedral body could trap<br />

the propionaldehyde produced by the PduCDE diol dehydratase until it is consumed by<br />

the PduP propionaldehyde dehydrogenase. Sequestering propionaldehyde until it is<br />

converted to propionyl-CoA might protect sensitive cytoplasmic components.<br />

Alternatively, the inclusion of diol dehydratase and propionaldehyde dehydrogenase in<br />

the pdu polyhedra might simply serve to juxtapose these enzymes in order to facilitate<br />

propionaldehyde channeling, although it is unclear why such a complex structure would<br />

be needed for this sole purpose. Prior studies suggested that the pdu polyhedra minimize<br />

aldehyde toxicity by limiting aldehyde production through control of AdoCbl availability<br />

(Havemann et al. 2002). Such a mechanism would be improved by either aldehyde<br />

channeling or sequestration, and a combined mechanism that fine tunes propionaldehyde<br />

production and consumption could explain the selective advantage provided by these<br />

complex polyhedral bodies involved in 1,2-propanediol degradation.


113<br />

Table 4-1. Bacterial strains<br />

Species Strain Genotype<br />

E. coli<br />

BL21<br />

(DE3) RIL<br />

BE237<br />

BE273<br />

(E. coli B) F - ompT hsdS (rB - mB - ) dcm + Tet r gal λ (DE3) endA<br />

Hte (argU ileY leuW Cam r )<br />

BL21 (DE3) RIL/pET-41a (T7 expression vector without<br />

insert, Kan r )<br />

BL21 (DE3) RIL/pNL69 (T7expression vector with insert<br />

encoding His8-PduP)<br />

S17.1λpir recA(RP4-2-Tc::Mu) λpir<br />

BE47 thr-480::Tn10dCam<br />

S. enterica serovar Typhimurium LT2<br />

TR6579<br />

metA22 metE551 trpD2 ilv-452 hsdLt6 hsdSA29 HsdB -<br />

strA120 GalE - Leu - Pro -<br />

BE191 ∆pduP659<br />

BE268 LT2/ pNL9 (pLAC22-pduP, Ap r )<br />

BE269 ∆pduP659/ pLAC22 (vector without insert, Ap r )<br />

BE270 ∆pduP659/pNL9 (pLAC22-pduP, Ap r )


114<br />

Table 4-2. Propionaldehyde dehydrogenase activity in extracts from wild-type<br />

S. enterica and strain BE191 (∆pduP).<br />

Extract S/P a<br />

Wild-type<br />

BE191<br />

(∆pduP)<br />

Total<br />

protein<br />

(mg)<br />

Propionaldehyde dehydrogenase activity<br />

Specific activity<br />

(µmol min -1 mg -1<br />

protein)<br />

Total activity<br />

(µmol min -1 )<br />

Total activity<br />

(%)<br />

S 45 0.004 0.18 1<br />

P 25 1.15 29.23 99<br />

S 51 ND b ND 0<br />

P 30 0.011 0.33 100<br />

a Assays were performed on the supernatant (S) and pellet (P) fractions following<br />

centrifugation at 31,000 x g.<br />

b ND indicates, none detected.


115<br />

Table 4-3. Propionaldehyde dehydrogenase activity during polyhedral body purification<br />

Sample<br />

Crude extract<br />

Protein<br />

(mg)<br />

Total activity<br />

(µmol min -1 )<br />

Specific activity a<br />

(µmol min -1 mg -1<br />

protein)<br />

Yield<br />

(%)<br />

Fold -<br />

purification<br />

515 152 0.30 100 1.0<br />

Detergent/salts<br />

treatment<br />

338 205 0.61 135 2.1<br />

12,000 x g<br />

super<br />

338 169 0.50 111 1.7<br />

48,000 x g<br />

pellet<br />

5.0 9.6 1.92 6.3 6.5<br />

12,000 x g<br />

super<br />

Sucrose<br />

4.3 8.6 2.00 5.7 6.8<br />

density<br />

gradient<br />

0.3 1.2 4.00 0.8 13.6<br />

a<br />

Activities were calculated from the initial rate


97<br />

66<br />

45<br />

31<br />

22<br />

14<br />

7<br />

116<br />

kDa 1 2 3 4 5 6<br />

PduP<br />

Figure 4-1. SDS-PAGE analysis of His8-PduP. Lane 1: Molecular mass markers, lane 2:<br />

12 µg soluble extract from the control strain (BE237), lane 3: 12 µg soluble<br />

extract from the His8-PduP expression strain (BE273), lane 4: 12 µg<br />

inclusion-body extract from the control strain, lane 5: 5 µg inclusion-body<br />

extract from the His8-PduP expression strain, lane 6: 2 µg His8-PduP purified<br />

by nickel-affinity chromatography. The control and the expression strains<br />

were isogenic except that the control strain lacked the pduP coding sequence.


kDa<br />

209<br />

80<br />

49<br />

35<br />

29<br />

117<br />

1 2 3 4<br />

Figure 4-2. Western blot using the anti-PduP antiserum. Lane 1: Molecular mass<br />

standards, lane 2: wild-type (Salmonella enterica), lane 3: ∆pduP (strain<br />

BE191), lane 4: ∆pduP/pLAC22 (vector without insert), lane 5:<br />

∆pduP/pLAC22-pduP. Each lane was loaded with 20 µl boiled whole cells<br />

(OD600 = 1.0).<br />

5


118<br />

B<br />

Figure 4-3. Immunogold localization of the PduP enzyme. A, Wild-type S. enterica; B,<br />

strain BE191 (∆pduP). The 12 nm gold particles (small black circles)<br />

indicate the location of the PduP protein. The arrows point to the polyhedral<br />

bodies. Bars 100 nm


CHAPTER 5<br />

CONCLUSIONS<br />

In humans, cobalamin cofactors are required coenzymes for two enzymatic<br />

reactions that are vital to human health. These coenzymes are used in propionate<br />

metabolism and methionine biosynthesis, and in humans, deficiencies in cobalamin<br />

metabolism result in methylmalonic aciduria and homocystinuria. Nine complementation<br />

groups associated with such deficiencies have been identified. Historically, progress on<br />

understanding these diseases has been slow due to difficulties in purifying the enzymes<br />

involved as well as the lack of facile genetic methods. A bacterial model system was<br />

used to circumvent some of these problems and allowed progress on the identification<br />

and characterization of the genes and enzymes involved in these rare, but devastating<br />

disorders.<br />

Identification of the Bovine and Human Adenosyltransferase<br />

In Chapter 2, S. enterica was used as a model system for the identification of the<br />

bovine and human ATR cDNAs by complementation analysis. In this screening<br />

technique, the bovine cDNA was isolated by complementing an ATR deficient<br />

S. enterica strain. Subsequently, sequence similarity searches using the bovine ATR<br />

cDNA identified a homologous human gene. Both human and bovine expression<br />

libraries were screened; however, complementing clones were only found from the<br />

bovine library. This could have been due to the relative abundance of ATR mRNA in<br />

these systems. Ruminant livers have the highest reported concentration of <strong>B12</strong>-dependent<br />

119


120<br />

enzymes measured in any tissue. Similar methods could be applied for the identification<br />

of additional cDNAs involved in cobalamin metabolism.<br />

To determine whether the bovine and human cDNAs that we identified encode<br />

ATRs, the cDNAs were cloned and over expressed in E. coli. Cell extracts from the<br />

expression strains were then used for biochemical studies. Both the soluble and insoluble<br />

fractions of the crude cell extracts of bovine and human expression strains were found to<br />

have ATR activity. This suggests that the cDNAs we identified encode ATR enzymes.<br />

Additional studies were used to show that the human cDNA could function as an ATR<br />

in vivo. An S. enterica ATR mutant transformed with a plasmid-encoded source of the<br />

human ATR was used for growth studies on 1,2-propanediol supplemented with HOCbl.<br />

From these experiments, we showed that the human ATR restored wild-type phenotype<br />

to an ATR-deficient S. enterica strain proving that the human ATR can function<br />

physiologically in a heterologous host.<br />

Previous studies have shown that patients with cblB methylmalonic aciduria lack<br />

ATR activity (Fenton and Rosenberg 1981). To determine if mutations in the ATR<br />

underlie cblB methylmalonic aciduria, expression of this enzyme was monitored in<br />

normal and cblB mutant fibroblast cells. Western blot analysis using antibodies specific<br />

for the ATR showed that expression of this protein was altered in fibroblast cells from<br />

patients with cblB when compared to control cells. Additional studies conducted<br />

concurrently in another lab showed that the ATR gene (MMAB) corresponded to the cblB<br />

complementation group of methylmalonic aciduria (Dobson et al. 2002).<br />

Biochemical Characterization of the Human Adenosyltransferase<br />

In Chapter 3, we investigated the biochemical properties of the human ATR.<br />

Dobson et al. previously analyzed the MMAB gene and identified two amino acid


121<br />

substitutions that are common polymorphic variants in control cell lines (Dobson et al.<br />

2002). Both of these variants were independently expressed, purified, and used for<br />

biochemical analysis. In the studies conducted, the variants behaved similarly suggesting<br />

that both are normal forms of the enzyme. The kinetic properties including Vmax, and Km<br />

for ATP and cob(I)alamin were determined. The Vmax falls within the range of previously<br />

reported ATRs including CobA, PduO, and T. acidophilum ATR (Suh and Escalante-<br />

Semerena 1995, Johnson et al. 2001, Saridakis et al. 2004). The Km for cob(I)alamin<br />

(1 µM) is physiologically significant when compared to the intracellular concentration of<br />

<strong>B12</strong> in human tissue (Schneider and Stroinski 1987). The Km for ATP is low when<br />

compared to the physiological levels of this substrate but is considered relevant since<br />

ATP is required for many different reactions in human tissue.<br />

In this study, the nucleotide specificity of the human ATR variants was also<br />

analyzed. It was found that ATR variants are highly specific for ATP (the physiological<br />

substrate for adenosylation). This result is not surprising since catalysis with other<br />

nucleotide substrates (GTP, CTP, and UTP) resulting in GCbl, CCbl, and UCbl are<br />

generally inhibitory to AdoCbl-dependent enzymes and would be counter-productive<br />

(Toraya and Mori 1999). Establishing the biochemical properties of control ATRs<br />

provides a basis of comparison for the characterization of disease causing variants of this<br />

enzyme.<br />

Methionine Synthase Reductase is a Cob(II)alamin Reductase for the Human<br />

Adenosyltransferase<br />

In addition to analyzing the biochemical properties of the human ATR, we also<br />

investigated whether human MSR can act as a cob(II)alamin reductase for the synthesis<br />

of AdoCbl by the ATR (Chapter 3). Previous studies have shown that MSR functions as


122<br />

a cob(II)alamin reductase for reductive activation of MS (Olteanu and Banerjee 2001).<br />

Here, an MSR-ATR linked assay was developed and this provided biochemical evidence<br />

that MSR can also function as a cob(II)alamin reductase for AdoCbl synthesis. The rate<br />

of AdoCbl formation in these assays (0.96 nmol min -1 ) is physiologically significant<br />

since humans require only small amounts of cobalamin (Schneider 1987). This rate is<br />

also significant when compared to previous measurements of AdoCbl formed in human<br />

fibroblast cells (0.021 pmol min -1 mg -1 ) (Fenton and Rosenberg 1981).<br />

In mammalian systems, MSR is a cytosolic protein while ATR is localized to the<br />

mitochondria. In these studies we demonstrated that MSR is as a cob(II)alamin reductase<br />

for the ATR in vitro. The proposed requirement of MSR for AdoCbl synthesis<br />

necessitates the presence of MSR in both the cytoplasm and the mitochondria. Recent<br />

studies have shown that alternative transcript splicing of MSR produces protein with and<br />

without a mitochondrial targeting sequence (Leclerc et al. 1999). These findings<br />

strengthen the hypothesis that MSR has dual functionality in the cytoplasm and the<br />

mitochondria.<br />

Investigations also suggest an interaction between MSR and ATR. In MSR-ATR<br />

linked assays containing iodoacetate (an alkylating agent that forms CMCbl in the<br />

presence of cob(I)alamin), very little CMCbl was formed suggesting that MSR does not<br />

release cob(I)alamin free in solution. In addition, HPLC analyses showed that AdoCbl<br />

formation was favored even when a large excess of iodoacetate was added to the assays.<br />

These studies indicate that an interaction between MSR and ATR triggers cobalamin<br />

reduction.


123<br />

Studies also indicated that humans have redundant systems for mitochondrial<br />

cob(II)alamin reduction. Prior complementation analysis demonstrated that cblE (MSR)<br />

disorders usually result in homocystinuria without methylmalonic aciduria with only one<br />

reported case having both homocystinuria and mild methylmalonic aciduria (Wilson et al.<br />

1999). The fact that most cblE patients lack methylmalonic aciduria suggests that there<br />

are redundant mitochondrial cob(II)alamin reductase(s) involved in AdoCbl synthesis.<br />

There is precedence for redundant cobalamin reductases in bacteria. The role of MSR<br />

could be to provide extra capacity under certain conditions, such as a high protein diet,<br />

rich in amino acids that are metabolized via the propionyl-CoA pathway.<br />

Future Experimentation<br />

The development of the screening method employed in this study can be used in<br />

the future as a general technique for the isolation of other cDNAs involved in cobalamin<br />

metabolism. Salmonella is an excellent model system for these studies since<br />

AdoCbl-dependent diol dehydratase for 1,2-propanediol degradation has been well<br />

studied. Once other enzymes involved in cobalamin metabolism are identified in<br />

S. enterica, strains deficient in these activities can be used to screen mammalian<br />

expression libraries by complementation. Hence, this technique may allow for the<br />

identification of other genes crucial for cobalamin metabolism in eukaryotic systems.<br />

Now that a purification method has been developed for the ATR, efforts can be<br />

focused on crystallizing this protein. Determining the three dimensional structure of the<br />

ATR will allow us to understand how mutations in this enzyme lead to dysfunction by<br />

comparing amino acid residues in normal and mutant ATRs and mapping the altered<br />

amino acid residues onto the structure. Potentially, this process would help lend insight<br />

in identifying the residues of the ATR that are crucial for enzymatic activity.


124<br />

Future studies should also be focused on determining the mode of interaction<br />

between MSR and ATR. Approaches to verifying the interaction between these enzymes<br />

include gel filtration, immuno-precipitation, and native PAGE analyses. All of these<br />

techniques would verify a physical interaction between MSR and ATR and give a better<br />

understanding of how this system functions in vivo.<br />

Most importantly, knowledge of the ATR gene will allow for DNA based methods<br />

of methylmalonic aciduria diagnosis following amniocentesis or chorionic villi sampling<br />

and could aid in determining the mode of treatment. Additionally, identification of the<br />

ATR gene makes possible the development of gene therapy as a treatment for cblB<br />

methylmalonic aciduria.


LIST OF REFERENCES<br />

Aberg, A., S. Hahne, M. Karlsson, A. Larsson, M. Ormo, A. Ahgren and B. M.<br />

Sjoberg. 1989. Evidence for two different classes of redox-active cysteines in<br />

ribonucleotide reductase of Escherichia coli. J. Biol. Chem. 264:12249-52.<br />

Allen, S. G. H., R. W. Kellermeyer, R. L. Stjernholm and H. G. Wood. 1964.<br />

Purification and properties of enzymes involved in the propionic acid fermentation.<br />

J. Bacteriol. 87:171-187.<br />

Altschul, S. F., W. Gish, W. Miller, E. W. Myers and D. J. Lipman. 1990. Basic local<br />

alignment search tool. J. Mol. Biol. 215:403-410.<br />

Altschul, S. F., T. L. Madden, A. A. Schäffer, J. Zhang, Z. Zhang, W. Miller and D.<br />

J. Lipman. 1997. Gapped BLAST and PSI-BLAST: a new generation of protein<br />

database search programs. Nucleic Acids Res. 25:3389-3402.<br />

Ampola, M. G., M. J. Mahoney, E. Nakamura and K. Tanaka. 1975. Prenatal therapy<br />

of a patient with vitamin <strong>B12</strong> responsive methylmalonic acidemia. N. Engl. J. Med.<br />

293:313-317.<br />

Baker, S. J. and V. I. Mathan. 1981. Evidence regarding the minimal daily requirement<br />

of dietary vitamin <strong>B12</strong>. Am. J. Clin. Nutr. 34:2423-33.<br />

Banerjee, R. 1997. The Yin-Yang of cobalamin biochemistry. Chem. Biol. 4:175-86.<br />

Banerjee, R. (eds.). 1999. Chemistry and Biochemistry of <strong>B12</strong>. John Wiley & Sons, Inc,<br />

New York.<br />

Banerjee, R. and S. Chowdhury. 1999. Methylmalonyl-CoA mutase, p. 707-730. In R.<br />

Banerjee (ed.), Chemistry and Biochemistry of <strong>B12</strong>. John Wiley & Sons, Inc., New<br />

York.<br />

Banerjee, R. and S. W. Ragsdale. 2003. The many faces of vitamin <strong>B12</strong>: catalysis by<br />

cobalamin-dependent enzymes. Annu. Rev. Biochem. 72:209-47.<br />

Banerjee, R. V., S. R. Harder, S. W. Ragsdale and R. G. Matthews. 1990. Mechanism<br />

of reductive activation of cobalamin-dependent methionine synthase: an electron<br />

paramagnetic resonance spectroelectrochemical study. Biochemistry 29:1129-35.<br />

125


126<br />

Banerjee, R. V., N. L. Johnston, J. K. Sobeski, P. Datta and R. G. Matthews. 1989.<br />

Cloning and sequence analysis of the Escherichia coli metH gene encoding<br />

cobalamin-dependent methionine synthase and isolation of a tryptic fragment<br />

containing the cobalamin-binding domain. J. Biol. Chem. 264:13888-95.<br />

Banerjee, R. V. and R. G. Matthews. 1990. Cobalamin-dependent methionine synthase.<br />

Faseb. J. 4:1450-9.<br />

Barker, H. A., H. Weisbach and R. D. Smyth. 1958. A Coenzyme containing<br />

pseudovitamin <strong>B12</strong>. Proc. Natl. Acad. Sci. U S A 44:1039-1097.<br />

Battersby, A. R. 1994. How nature builds the pigments of life: the conquest of vitamin<br />

<strong>B12</strong>. Science 264:1551-7.<br />

Bauer, C. B., M. V. Fonseca, H. M. Holden, J. B. Thoden, T. B. Thompson, J. C.<br />

Escalante-Semerena and I. Rayment. 2001. Three-dimensional structure of<br />

ATP:corrinoid adenosyltransferase from Salmonella typhimurium in its free state,<br />

complexed with MgATP, or complexed with hydroxycobalamin and MgATP.<br />

Biochemistry 40:361-74.<br />

Berkowitz, D., J. Hushon, H. Whitfield, Jr., J. Roth and B. Ames. 1968. Procedure for<br />

identifying nonsense mutations. J. Bacteriol. 96:215-220.<br />

Bobik, T. A., M. Ailion and J. R. Roth. 1992. A single regulatory gene integrates<br />

control of vitamin <strong>B12</strong> synthesis and propanediol degradation. J. Bacteriol.<br />

174:2253-66.<br />

Bobik, T. A., G. D. Havemann, R. J. Busch, D. S. Williams and H. C. Aldrich. 1999.<br />

The propanediol utilization (pdu) operon of Salmonella enterica serovar<br />

Typhimurium LT2 includes genes necessary for the formation of polyhedral<br />

organelles involved in coenzyme B 12-dependent 1,2-propanediol degradation. J.<br />

Bacteriol. 181:5967-5975.<br />

Bobik, T. A. and M. E. Rasche. 2003. HPLC assay for methylmalonyl-CoA epimerase.<br />

Anal. Bioanal. Chem. 375:344-9.<br />

Bobik, T. A. and R. S. Wolfe. 1989. Activation of formylmethanofuran synthesis in cell<br />

extracts of Methanobacterium thermoautotrophicum. J. Bacteriol. 171:1423-1427.<br />

Bobik, T. A., Y. Xu, R. M. Jeter, K. E. Otto and J. R. Roth. 1997. Propanediol<br />

utilization genes (pdu) of Salmonella typhimurium: three genes for the propanediol<br />

dehydratase. J. Bacteriol. 179:6633-9.<br />

Booker, S., S. Licht, J. Broderick and J. Stubbe. 1994. Coenzyme <strong>B12</strong>-dependent<br />

ribonucleotide reductase: evidence for the participation of five cysteine residues in<br />

ribonucleotide reduction. Biochemistry 33:12676-85.


127<br />

Bradford, M. 1976. A rapid and sensitive method for the quantitation of microgram<br />

quantities of protein utilizing the principle of protein-dye binding. Anal. Biochem.<br />

72:248-254.<br />

Brady, R. O., E. G. Castanera and H. A. Barker. 1962. The enzymatic synthesis of<br />

cobamide coenzymes. J. Biol. Chem. 237:2325-32.<br />

Cauthen, S. E., M. A. Foster and D. D. Woods. 1966. Methionine synthesis by extracts<br />

of Salmonella typhimurium. Biochem. J. 98:630-635.<br />

Chen, P., D. I. Andersson and J. R. Roth. 1994. The control region of the pdu/cob<br />

regulon in Salmonella typhimurium. J. Bacteriol. 176:5474-82.<br />

Ciani, F., M. A. Donati, G. Tulli, G. M. Poggi, E. Pasquini, D. S. Rosenblatt and E.<br />

Zammarchi. 2000. Lethal late onset cblB methylmalonic aciduria. Crit. Care Med.<br />

28:2119-21.<br />

Claros, M. G. and P. Vincens. 1996. Computational method to predict mitochondrially<br />

imported proteins and their targeting sequences. Eur. J. Biochem. 241:770-786.<br />

Conner, C. P., D. M. Heithoff, S. M. Julio, R. L. Sinsheimer and M. J. Mahan. 1998.<br />

Differential patterns of acquired virulence genes distinguish Salmonella strains.<br />

Proc. Natl. Acad. Sci. U S A 95:4641-5.<br />

Cooper, B. A. and D. S. Rosenblatt. 1987. Inherited defects of vitamin <strong>B12</strong> metabolism.<br />

Annu. Rev. Nutr. 7:291-320.<br />

Crouzet, J., S. Levy-Schil, B. Cameron, L. Cauchois, S. Rigault, M. C. Rouyez, F.<br />

Blanche, L. Debussche and D. Thibaut. 1991. Nucleotide sequence and genetic<br />

analysis of a 13.1-kilobase-pair Pseudomonas denitrificans DNA fragment<br />

containing five cob genes and identification of structural genes encoding<br />

Cob(I)alamin adenosyltransferase, cobyric acid synthase, and bifunctional<br />

cobinamide kinase-cobinamide phosphate guanylyltransferase. J. Bacteriol.<br />

173:6074-87.<br />

Davis, R. W., D. Botstein and J. R. Roth. 1980. Advanced bacterial genetics : a manual<br />

for genetic engineering. Cold Spring Harbor Laboratory, Cold Spring Harbor, New<br />

York.<br />

Dawson, R. M. C., D. C. Elliott, W. H. Elliott and K. M. Jones (eds.). 1969. Data for<br />

biochemical research, 2nd ed. Oxford University Press, Oxford.<br />

Debussche, L., M. Couder, D. Thibaut, B. Cameron, J. Crouzet and F. Blanche.<br />

1991. Purification and partial characterization of cob(I)alamin adenosyltransferase<br />

from Pseudomonas denitrificans. J. Bacteriol. 173:6300-6302.


128<br />

Del Corral, A. and R. Carmel. 1990. Transfer of cobalamin from the cobalamin-binding<br />

protein of egg yolk to R binder of human saliva and gastric juice.<br />

Gastroenterologist 98:1460-6.<br />

Dixon, M. M., S. Huang, R. G. Matthews and M. Ludwig. 1996. The structure of the<br />

C-terminal domain of methionine synthase: presenting S-adenosylmethionine for<br />

reductive methylation of <strong>B12</strong>. Structure 4:1263-75.<br />

Dobson, C. M., T. Wai, D. Leclerc, H. Kadir, M. Narang, J. P. Lerner-Ellis, T. J.<br />

Hudson, D. S. Rosenblatt and R. A. Gravel. 2002. Identification of the gene<br />

responsible for the cblB complementation group of vitamin <strong>B12</strong>-dependent<br />

methylmalonic aciduria. Hum. Mol. Genet. 11:3361-9.<br />

Dobson, C. M., T. Wai, D. Leclerc, A. Wilson, X. Wu, C. Dore, T. Hudson, D. S.<br />

Rosenblatt and R. A. Gravel. 2002. Identification of the gene responsible for the<br />

cblA complementation group of vitamin <strong>B12</strong>-responsive methylmalonic acidemia<br />

based on analysis of prokaryotic gene arrangements. Proc. Natl. Acad. Sci. U S A<br />

99:15554-9.<br />

Dolphin, D. 1982. <strong>B12</strong>. John Wiley & Sons, Inc., New York.<br />

Drennan, C. L., S. Huang, J. T. Drummond, R. G. Matthews and M. L. Ludwig.<br />

1994. How a protein binds <strong>B12</strong>: a 3.0 Å x-ray structure of <strong>B12</strong>-binding domains of<br />

methionine synthase. Science 266:1669-1674.<br />

Drummond, J. T. and R. G. Matthews. 1993. Cobalamin-dependent and cobalaminindependent<br />

methionine synthases in Escherichia coli: two solutions to the same<br />

chemical problem. Adv. Exp. Med. Biol. 338:687-692.<br />

Drummond, J. T. and R. G. Matthews. 1994. Nitrous oxide inactivation of cobalamindependent<br />

methionine synthase from Escherichia coli: characterization of the<br />

damage to the enzyme and prosthetic group. Biochemistry 33:3742-50.<br />

Fenton, W. A., R. A. Gravel and D. Rosenblatt. 2000. Disorders of propionate and<br />

methylmalonate metabolism. In C. R. Scriver, A. L. Beaudet, W. S. Sly, et al. (ed.),<br />

The Metabolic and molecular basis of inherited disease. McGraw-Hill, New York.<br />

Fenton, W. A. and L. E. Rosenberg. 1978a. Genetic and biochemical analysis of human<br />

cobalamin mutants in cell culture. Annu. Rev. Genet. 12:223-48.<br />

Fenton, W. A. and L. E. Rosenberg. 1978b. Mitochondrial metabolism of<br />

hydroxocobalamin: synthesis of adenosylcobalamin by intact rat liver<br />

mitochondria. Arch. Biochem. Biophys. 189:441-7.<br />

Fenton, W. A. and L. E. Rosenberg. 1981. The defect in the cblB class of human<br />

methylmalonic acidemia: deficiency of cob(I)alamin adenosyltransferase activity in<br />

extracts of cultured fibroblasts. Biochem. Biophys. Res. Commun. 98:283-289.


129<br />

Fenton, W. A. and L. E. Rosenberg. 2000. Inherited Disorders of Cobalamin Transport<br />

and Metabolism. In C. R. Scriver, A. L. Beaudet, W. S. Sly, et al. (ed.), The<br />

Metabolic and molecular basis of inherited disease. McGraw-Hill, New York.<br />

Fonseca, M. V., N. R. Buan, A. R. Horswill, I. Rayment and J. C. Escalante-<br />

Semerena. 2002. The ATP : Co(I)rrinoid adenosyltransferase (CobA) enzyme of<br />

Salmonella enterica requires the 2 '-OH group of ATP for function and yields<br />

inorganic triphosphate as its reaction byproduct. J. Biol. Chem. 277:33127-33131.<br />

Fonseca, M. V. and J. C. Escalante-Semerena. 2000. Reduction of Cob(III)alamin to<br />

Cob(II)alamin in Salmonella enterica serovar Typhimurium LT2. J. Bacteriol.<br />

182:4304-9.<br />

Fonseca, M. V. and J. C. Escalante-Semerena. 2001. An in vitro reducing system for<br />

the enzymic conversion of cobalamin to adenosylcobalamin. J. Biol. Chem.<br />

276:32101-8.<br />

Frey, P. A. and G. H. Reed. 2000. Radical mechanisms in adenosylmethionine- and<br />

adenosylcobalamin-dependent enzymatic reactions. Arch. Biochem. Biophys.<br />

382:6-14.<br />

Friedmann, H. C. 1975. Biosynthesis of corrinoids, p. 75-103. In B. M. Babior (ed.),<br />

Cobalamin. John Wiley and Sons, New York.<br />

Fujii, K., J. H. Galivan and F. M. Huennekens. 1977. Activation of methionine<br />

synthase: further characterization of flavoprotein system. Arch. Biochem. Biophys.<br />

178:662-70.<br />

Fujii, K. and F. M. Huennekens. 1974. Activation of methionine synthetase by a<br />

reduced triphosphopyridine nucleotide-dependent flavoprotein system. J. Biol.<br />

Chem. 249:6745-53.<br />

Goulding, C. W. and R. G. Matthews. 1997. Cobalamin-dependent methionine<br />

synthase from Escherichia coli: involvement of zinc in homocysteine activation.<br />

Biochemistry 36:15749-57.<br />

Goulding, C. W., D. Postigo and R. G. Matthews. 1997. Cobalamin-dependent<br />

methionine synthase is a modular protein with distinct regions for binding<br />

homocysteine, methyltetrahydrofolate, cobalamin, and adenosylmethionine.<br />

Biochemistry 36:8082-91.<br />

Gravel, R. A., M. J. Mahoney, F. H. Ruddle and L. E. Rosenberg. 1975. Genetic<br />

complementation in heterokaryons of human fibroblasts defective in cobalamin<br />

metabolism. Proc. Natl. Acad. Sci. U S A 72:3181-5.<br />

Guest, J. R., M. A. Foster and S. Friedman. 1962a. Alternative pathways for<br />

methylation of homocysteine by Escherichia coli. Biochem. J. 84:93P-94P.


130<br />

Guest, J. R., E. L. Smith, D. D. Woods and S. Friedman. 1962b. A methyl analogue of<br />

cobamide coenzyme in relation to methionine synthesis by bacteria. Nature<br />

195:340-342.<br />

Hall, D. A., T. C. Jordan-Starck, R. O. Loo, M. L. Ludwig and R. G. Matthews.<br />

2000. Interaction of flavodoxin with cobalamin-dependent methionine synthase.<br />

Biochemistry 39:10711-9.<br />

Halpern, J. 1985. Mechanisms of coenzyme <strong>B12</strong>-dependent rearrangements. Science<br />

227:869-75.<br />

Harder, J., R. Eliasson, E. Pontis, M. D. Ballinger and P. Reichard. 1992. Activation<br />

of the anaerobic ribonucleotide reductase from Escherichia coli by Sadenosylmethionine.<br />

J. Biol. Chem. 267:25548-52.<br />

Harlow, E. and D. Lane. 1988. Antibodies: a laboratory manual. Cold Spring Harbor<br />

Laboratory, Cold Spring Harbor, New York.<br />

Havemann, G. D. and T. A. Bobik. 2003. Protein content of polyhedral organelles<br />

involved in coenzyme <strong>B12</strong>-dependent degradation of 1,2-propanediol in Salmonella<br />

enterica serovar Typhimurium LT2. J. Bacteriol. 185:5086-95.<br />

Havemann, G. D., E. M. Sampson and T. A. Bobik. 2002. PduA is a shell protein of<br />

polyhedral organelles involved in coenzyme <strong>B12</strong>-dependent degradation of 1,2propanediol<br />

in Salmonella enterica serovar Typhimurium LT2. J. Bacteriol.<br />

184:1253-61.<br />

Heithoff, D. M., C. P. Conner, U. Hentschel, F. Govantes, P. C. Hanna and M. J.<br />

Mahan. 1999. Coordinate intracellular expression of Salmonella genes induced<br />

during infection. J. Bacteriol. 181:799-807.<br />

Hodgkin, D. C., J. Pickworth, J. H. Robertson, K. N. Trueblood, R. J. Prosen and J.<br />

G. White. 1955. Crystal structure of the hexacarboxylic acid derived from <strong>B12</strong> and<br />

the molecular structure of the vitamin. Nature 176:325-328.<br />

Horswill, A. R. and J. C. Escalante-Semerena. 1997. Propionate catabolism in<br />

Salmonella typhimurium LT2: two divergently transcribed units comprise the prp<br />

locus at 8.5 centisomes, prpR encodes a member of the sigma-54 family of<br />

activators, and the prpBCDE genes constitute an operon. J. Bacteriol. 179:928-40.<br />

Huennekens, F. M., K. S. Vitols, K. Fujii and D. W. Jacobsen. 1982. Biosynthesis of<br />

the cobalamin coenzymes, p. 145-164. In D. Dolphin (ed.), <strong>B12</strong>, vol. volume 1.<br />

John Wiley & Sons, Inc., New York.<br />

Jarrett, J. T., D. M. Hoover, M. L. Ludwig and R. G. Matthews. 1998. The<br />

mechanism of adenosylmethionine-dependent activation of methionine synthase: a<br />

rapid kinetic analysis of intermediates in reductive methylation of cob(II)alamin<br />

enzyme. Biochemistry 37:12649-58.


131<br />

Johnson, C. L. V. J., E. Pechonick, S. D. Park, G. D. Havemann, N. A. Leal and T.<br />

A. Bobik. 2001. Functional genomic, biochemical, and genetic characterization of<br />

the Salmonella pduO gene, an ATP:cob(I)alamin adenosyltransferase gene. J.<br />

Bacteriol. 183:1577-1584.<br />

Jordan, A., E. Pontis, M. Atta, M. Krook, I. Gibert, J. Barbe and P. Reichard. 1994.<br />

A second class I ribonucleotide reductase in Enterobacteriaceae: characterization of<br />

the Salmonella typhimurium enzyme. Proc. Natl. Acad. Sci. U S A 91:12892-6.<br />

Kapadia, C. R. 1995. Vitamin <strong>B12</strong> in health and disease: part I--inherited disorders of<br />

function, absorption, and transport. Gastroenterologist 3:329-44.<br />

Kass, L. 1978. William B. Castle and intrinsic factor. Ann. Intern. Med. 89:983-91.<br />

Kessler, D., I. Leibrecht and J. Knappe. 1991. Pyruvate formate-lyase-deactivase and<br />

acetyl-CoA reductase activities of Escherichia coli reside on a polymeric protein<br />

particle encoded by adhE. FEBS Lett. 281:59-63.<br />

Kofoid, E., C. Rappleye, I. Stojiljkovic and J. R. Roth. 1999. The 17-gene<br />

ethanolamine (eut) operon of Salmonella typhimurium encodes five homologues of<br />

carboxysome shell proteins. J. Bacteriol. 181:5317-5329.<br />

Kolhouse, J. F. and R. H. Allen. 1977. Recognition of two intracellular cobalamin<br />

binding proteins and their identification as methylmalonyl-CoA mutase and<br />

methionine synthetase. Proc. Natl. Acad. Sci. U S A 74:921-5.<br />

Korotkova, N., L. Chistoserdova, V. Kuksa and M. E. Lidstrom. 2002. Glyoxylate<br />

regeneration pathway in the methylotroph Methylobacterium extorquens AM1. J.<br />

Bacteriol. 184:1750-8.<br />

Korotkova, N. and M. E. Lidstrom. 2004. MeaB is a component of the methylmalonyl-<br />

CoA mutase complex required for protection of the enzyme from inactivation. J.<br />

Biol. Chem. 279:13652-8.<br />

Leal, N. A., S. D. Park, P. E. Kima and T. A. Bobik. 2003. Identification of the human<br />

and bovine ATP:Cob(I)alamin adenosyltransferase cDNAs based on<br />

complementation of a bacterial mutant. J. Biol. Chem. 278:9227-34.<br />

Leclerc, D., E. Campeau, P. Goyette, C. E. Adjalla, B. Christensen, M. Ross, P.<br />

Eydoux, D. S. Rosenblatt, R. Rozen and R. A. Gravel. 1996. Human methionine<br />

synthase: cDNA cloning and identification of mutations in patients of the cblG<br />

complementation group of folate/cobalamin disorders. Hum. Mol. Genet. 5:1867-<br />

74.<br />

Leclerc, D., M. Odievre, Q. Wu, A. Wilson, J. J. Huizenga, R. Rozen, S. W. Scherer<br />

and R. A. Gravel. 1999. Molecular cloning, expression, and physical mapping of<br />

the human methionine synthase reductase gene. Gene 240:75-88.


132<br />

Leclerc, D., A. Wilson, R. Dumas, C. Gafuik, D. Song, D. Watkins, H. H. Heng, J.<br />

M. Rommens, S. W. Scherer, D. S. Rosenblatt and R. A. Gravel. 1998. Cloning<br />

and mapping of a cDNA for methionine synthase reductase, a flavoprotein<br />

defective in patients with homocystinuria. Proc. Natl. Acad. Sci. U S A 95:3059-<br />

64.<br />

Ledley, F. D., M. R. Lumetta, H. Y. Zoghbi, P. VanTuinen, S. A. Ledbetter and D.<br />

H. Ledbetter. 1988. Mapping of human methylmalonyl CoA mutase (MUT) locus<br />

on chromosome 6. Am. J. Hum. Genet. 42:839-46.<br />

Lenhert, P. G. and D. C. Hodgkin. 1961. Structure of 5,6dimethylbenzimidazolylcobamide<br />

coenzyme. Nature 192:937-938.<br />

Ljungdahl, L. and H. Wood. 1982. Acetate biosynthesis. In D. Dolphin (ed.), <strong>B12</strong>. John<br />

Wiley & Sons, New York.<br />

Mahoney, M. J., A. C. Hart, V. D. Steen and L. E. Rosenberg. 1975.<br />

Methylmalonicacidemia: biochemical heterogeneity in defects of 5'deoxyadenosylcobalamin<br />

synthesis. Proc. Natl. Acad. Sci. USA 72:2799-2803.<br />

Mahoney, M. J., L. E. Rosenberg, S. H. Mudd and B. W. Uhlendorf. 1971. Defective<br />

metabolism of vitamin <strong>B12</strong> in fibroblasts from children with methylmalonic<br />

aciduria. Biochem. Biophys. Res. Commun. 44:375-81.<br />

Mancia, F., N. H. Keep, A. Nakagawa, P. F. Leadlay, S. McSweeney, B. Rasmussen,<br />

P. Bosecke, O. Diat and P. R. Evans. 1996. How coenzyme <strong>B12</strong> radicals are<br />

generated: the crystal structure of methylmalonyl-coenzyme A mutase at 2 Å<br />

resolution. Structure 4:339-50.<br />

Maniatis, T., E. F. Fritsch and J. Sambrook. 1982. Molecular Cloning; A Laboratory<br />

Manual. Cold Spring Harbor Laboratory, Cold Spring Harbor, New York.<br />

Marsh, E. N. and C. L. Drennan. 2001. Adenosylcobalamin-dependent isomerases:<br />

new insights into structure and mechanism. Curr. Opin. Chem. Biol. 5:499-505.<br />

Marsh, E. N., N. McKie, N. K. Davis and P. F. Leadlay. 1989. Cloning and structural<br />

characterization of the genes coding for adenosylcobalamin-dependent<br />

methylmalonyl-CoA mutase from Propionibacterium shermanii. Biochem. J.<br />

260:345-52.<br />

Matsui, S. M., M. J. Mahoney and L. E. Rosenberg. 1983. The natural history of the<br />

inherited methylmalonic acidemias. N. Engl. J. Med. 293:857-861.<br />

Matthews, R. G. 1999. Cobalamin-dependent Methionine Synthase, p. 681-706. In R.<br />

Banerjee (ed.), Chemistry and biochemistry of <strong>B12</strong>. John Wiley & Sons, Inc., New<br />

York.


133<br />

Mellman, I., H. F. Willard and L. E. Rosenberg. 1978. Cobalamin binding and<br />

cobalamin-dependent enzyme activity in normal and mutant human fibroblasts. J.<br />

Clin. Invest. 62:952-60.<br />

Mellman, I., H. F. Willard, P. Youngdahl-Turner and L. E. Rosenberg. 1979.<br />

Cobalamin coenzyme synthesis in normal and mutant human fibroblasts. Evidence<br />

for a processing enzyme activity deficient in cblC cells. J. Biol. Chem. 254:11847-<br />

11853.<br />

Mellman, I. S., P. Youngdahl-Turner, H. F. Willard and L. E. Rosenberg. 1977.<br />

Intracellular binding of radioactive hydroxocobalamin to cobalamin-dependent<br />

apoenzymes in rat liver. Proc. Natl. Acad. Sci. U S A 74:916-20.<br />

Miller, J. H. 1972. Experiments in molecular genetics. Cold Spring Harbor Laboratory,<br />

Cold Spring Harbor, New York.<br />

Miller, V. L. and J. J. Mekalanos. 1988. A novel suicide vector and its use in<br />

construction of insertion mutations: osmoregulation of outer membrane proteins<br />

and virulence determinants in Vibrio cholerae requires toxR. J. Bacteriol.<br />

170:2575-83.<br />

Minot, G. R. and W. P. Murphy. 1926. Treatment of pernicious anemia by a special<br />

diet. J. Am. Med. Assoc. 87:470-476.<br />

Nordlund, P. and H. Eklund. 1993. Structure and function of the Escherichia coli<br />

ribonucleotide reductase protein R2. J. Mol. Biol. 232:123-64.<br />

Obradors, N., J. Badía, L. Baldomà and J. Aguilar. 1988. Anaerobic metabolism of<br />

the L-rhamnose fermentation product 1,2-propanediol in Salmonella typhimurium.<br />

J. Bacteriol. 170:2159-2162.<br />

Olteanu, H. and R. Banerjee. 2001. Human methionine synthase reductase, a soluble P-<br />

450 reductase-like dual flavoprotein, is sufficient for NADPH-dependent<br />

methionine synthase activation. J. Biol. Chem. 276:35558-63.<br />

Padmakumar, R. and R. Banerjee. 1995. Evidence from electron paramagnetic<br />

resonance spectroscopy of the participation of radical intermediates in the reaction<br />

catalyzed by methylmalonyl-coenzyme A mutase. J. Biol. Chem. 270:9295-300.<br />

Padmakumar, R. and R. Banerjee. 1997. Evidence that cobalt-carbon bond homolysis<br />

is coupled to hydrogen atom abstraction from substrate in methylmalonyl-CoA<br />

mutase. Biochemistry 36:3713-8.<br />

Palacios, S., V. J. Starai and J. C. Escalante-Semerena. 2003. Propionyl coenzyme A<br />

is a common intermediate in the 1,2-propanediol and propionate catabolic pathways<br />

needed for expression of the prpBCDE operon during growth of Salmonella<br />

enterica on 1,2-propanediol. J. Bacteriol. 185:2802-10.


134<br />

Paschen, S. A. and W. Neupert. 2001. Protein import into mitochondria. IUBMB Life<br />

52:101-12.<br />

Pezacka, E. 1993. Identification of and characterization of two enzymes involved in the<br />

intracellular metabolism of cobalamin. Cyanocobalamin β-ligand transferase and<br />

micosomal cob(III)alamin reductase. Biochem. et. Biophys. Acta. 1157:167-177.<br />

Pezacka, E., R. Green and D. W. Jacobsen. 1990. Glutathionylcobalamin as an<br />

intermediate in the formation of cobalamin coenzymes. Biochem. Biophys. Res.<br />

Commun. 169:443-450.<br />

Price, G. D., S. Maeda, T. Omata and M. R. Badger. 2002. Modes of active inorganic<br />

carbon uptake in the cyanobacterium, Synechococcus sp PCC7942. Functional<br />

Plant Biology 29:131-149.<br />

Price-Carter, M., J. Tingey, T. A. Bobik and J. R. Roth. 2001. The alternative electron<br />

acceptor tetrathionate supports <strong>B12</strong>-dependent anaerobic growth of Salmonella<br />

enterica serovar typhimurium on ethanolamine or 1,2-propanediol. J. Bacteriol.<br />

183:2463-75.<br />

Qureshi, A. A., D. S. Rosenblatt and B. A. Cooper. 1994. Inherited disorders of<br />

cobalamin metabolism. Crit. Rev. Oncol. Hematol. 17:133-51.<br />

Refsum, H., P. M. Ueland, O. Nygard and S. E. Vollset. 1998. Homocysteine and<br />

cardiovascular disease. Annu. Rev. Med. 49:31-62.<br />

Reichard, P. 1993. The anaerobic ribonucleotide reductase from Escherichia coli. J.<br />

Biol. Chem. 268:8383-6.<br />

Rickes, E. L., N. G. Brink, F. R. Koniuszy, T. R. Wood and K. Folkers. 1948.<br />

Crystalline vitamin <strong>B12</strong>. Science 107:396-397.<br />

Rondon, M. R., R. Kazmierczak and J. C. Escalante-Semerena. 1995. Glutathione is<br />

required for maximal transcription of the cobalamin biosynthetic and 1,2propanediol<br />

utilization (cob/pdu) regulon and for the catabolism of ethanolamine,<br />

1,2-propanediol, and propionate in Salmonella typhimurium LT2. J. Bacteriol.<br />

177:5434-9.<br />

Rosenblatt, D. S., A. L. Aspler, M. I. Shevell, B. A. Pletcher, W. A. Fenton and M. R.<br />

Seashore. 1997. Clinical heterogeneity and prognosis in combined methylmalonic<br />

aciduria and homocystinuria (cblC). J. Inherit. Metab. Dis. 20:528-38.<br />

Rosenblatt, D. S. and A. B. Cooper. 1990. Inherited disorders of vitamin <strong>B12</strong> utilization.<br />

Bioessays 12:331-334.<br />

Rosenblatt, D. S. and W. A. Fenton. 1999. Inborn errors of cobalamin metabolism, p.<br />

367-384. In R. Banerjee (ed.), Chemistry and Biochemistry of <strong>B12</strong>. John Wiley &<br />

Sons, Inc., New York.


135<br />

Rossi, M., J. P. Glusker, L. Randaccio, M. F. Summers, P. J. Toscano and L. G.<br />

Marzilli. 1985. The Structure of a <strong>B12</strong> coenzyme: Methylcobalamin Studies by Xray<br />

and NMR Methods. J. Am. Chem. Soc. 107:1729-1738.<br />

Roth, J. R., T. A. Bobik and J. G. Lawrence. 1996. Cobalamin (coenzyme <strong>B12</strong>):<br />

synthesis and biological significance. Annual Rev. Microbiol. 50:137-181.<br />

Sambrook, J., E. F. Fritsch and T. Maniatis. 1989. Molecular cloning : a laboratory<br />

manual, 2nd ed. Cold Spring Harbor Laboratory, Cold Spring Harbor, New York.<br />

Saridakis, V., A. Yakunin, X. Xu, P. Anandakumar, M. Pennycooke, J. Gu, F.<br />

Cheung, J. M. Lew, R. Sanishvili, A. Joachimiak, C. H. Arrowsmith, D.<br />

Christendat and A. M. Edwards. 2004. The structural basis for methylmalonic<br />

aciduria. The crystal structure of archaeal ATP:cobalamin adenosyltransferase. J.<br />

Biol. Chem. 279:23646-53.<br />

Schmieger, H. 1971. A method for detection of phage mutants with altered transducing<br />

ability. Mol. Gen. Genet. 110:378-381.<br />

Schneider, Z. 1987. The occurence and distribution of corrinoids, p. 157-223. In Z.<br />

Schneider and A. Stroinski (ed.), Comprehensive <strong>B12</strong>. Walter de Gruyter, Berlin.<br />

Schneider, Z. 1987. Purification and Estimation of Vitamin <strong>B12</strong>, p. 111-155. In Z.<br />

Schneider and A. Stroinski (ed.), Comprehensive <strong>B12</strong>. Walter de Gruyter, Berlin.<br />

Schneider, Z. and A. Stroinski (eds.). 1987. Comprehensive <strong>B12</strong>. Walter de Gruyter &<br />

Co., Berlin.<br />

Schrauzer, G. N. and E. Deutsch. 1969. Reactions of cobalt(I) supernucleophiles. The<br />

alkylation of vitamin <strong>B12</strong>s cobaloximes(I), and related compounds. J. Am. Chem.<br />

Soc. 91:3341-50.<br />

Schrauzer, G. N., E. Deutsch and R. J. Windgassen. 1968. The nucleophilicity of<br />

vitamin <strong>B12</strong>. J. Am. Chem. Soc. 90:2441-2.<br />

Scott, A. I. 2003. Discovering nature's diverse pathways to vitamin <strong>B12</strong>: a 35-year<br />

odyssey. J. Org. Chem. 68:2529-39.<br />

Seetharam, B. 1999. Receptor-mediated endocytosis of cobalamin (vitamin <strong>B12</strong>). Annu.<br />

Rev. Nutr. 19:173-95.<br />

Segel, I. H. 1976. Biochemical calculations : how to solve mathematical problems in<br />

general biochemistry, 2 nd ed. Wiley, New York<br />

Shibata, N., J. Masuda, T. Tobimatsu, T. Toraya, K. Suto, Y. Morimoto and N.<br />

Yasuoka. 1999. A new mode of <strong>B12</strong> binding and the direct participation of a<br />

potassium ion in enzyme catalysis: X-ray structure of diol dehydratase. Structure<br />

Fold Des. 7:997-1008.


136<br />

Shively, J. M. and R. S. English. 1991. The carboxysome, a prokaryotic organelle: a<br />

mini-review. Can. J. Bot. 69:957-962.<br />

Shively, J. M., R. S. English, S. H. Baker and G. C. Cannon. 2001. Carbon cycling:<br />

the prokaryotic contribution. Curr. Opin. Microbiol. 4:301-306.<br />

Shively, J. M., G. van Keulen and W. G. Meijer. 1998. Something from almost<br />

nothing: carbon dioxide fixation in chemoautotrophs. Annu. Rev. Microbiol.<br />

52:191-230.<br />

Small, I., N. Peeters, F. Legeai and C. Lurin. 2004. Predotar: A tool for rapidly<br />

screening proteomes for N-terminal targeting sequences. Proteomics 4:1581-90.<br />

Smith, E. L. and L. F. J. Parker. 1948. Purification of anti-pernicious anaemia factor.<br />

Biochem. J. 43:R8-R9.<br />

Stabler, S. P. 1999. <strong>B12</strong> and Nutrition, p. 343-365. In R. Banerjee (ed.), Chemistry and<br />

biochemistry of <strong>B12</strong>. John Wiley & Sons, Inc., New York.<br />

Stojiljkovic, I., A. J. Baumler and F. Heffron. 1995. Ethanolamine Utilization in<br />

Salmonella Typhimurium - Nucleotide-Sequence, Protein Expression, and<br />

Mutational Analysis of the Ccha Cchb Eute Eutj Eutg Euth Gene-Cluster. J.<br />

Bacteriol. 177:1357-1366.<br />

Suh, S. J. and J. C. Escalante-Semerena. 1993. Cloning, sequencing and<br />

overexpression of cobA, which encodes ATP:corrinoid adenosyltransferase in<br />

Salmonella typhimurium. Gene 129:93-97.<br />

Suh, S.-J. and J. C. Escalante-Semerena. 1995. Purification and initial characterization<br />

of the ATP:corrinoid adenosyltransferase encoded by the cobA gene of Salmonella<br />

typhimurium. J. Bacteriol. 177:921-925.<br />

Thompson, J. D., T. J. Gibson, F. Plewniak, F. Jeanmougin and D. G. Higgins. 1997.<br />

The ClustalX windows interface: flexible strategies for multiple sequence<br />

alignment aided by quality analysis tools. Nucleic Acids Res. 24:4876-4882.<br />

Toraya, T., S. Honda and S. Fukui. 1979. Fermentation of 1,2-propanediol and 1,2ethanediol<br />

by some genera of Enterobacteriaceae, involving coenzyme <strong>B12</strong>dependent<br />

diol dehydratase. J. Bacteriol. 139:39-47.<br />

Toraya, T. and K. Mori. 1999. A reactivating factor for coenzyme <strong>B12</strong>-dependent diol<br />

dehydratase. J. Biol. Chem. 274:3372-7.<br />

Vitols, E., G. A. Walker and R. M. Huennekens. 1966. Enzymatic conversion of<br />

vitamin B 12s to a cobamide coenzyme, α-(5,6dimethylbenzimidazolyl)deoxyadenosylcobamide<br />

(adenosyl-<strong>B12</strong>). J. Biol. Chem.<br />

241:1455-1461.


137<br />

Vlasie, M., S. Chowdhury and R. Banerjee. 2002. Importance of the histidine ligand to<br />

coenzyme <strong>B12</strong> in the reaction catalyzed by methylmalonyl-CoA mutase. J. Biol.<br />

Chem. 277:18523-7.<br />

Vogel, H. J. and D. M. Bonner. 1956. Acetylornithase of Escherichia coli: partial<br />

purification and some properties. J. Biol. Chem. 218:97-106.<br />

Walker, G. A., S. Murphy and F. M. Huennekens. 1969. Enzymatic conversion of<br />

vitamin <strong>B12</strong>a to adenosyl <strong>B12</strong>: evidence for the existence of two separate reducing<br />

systems. Arch. Biochem. Biophys. 134:95-102.<br />

Walter, D., M. Ailion and J. Roth. 1997. Genetic characterization of the pdu operon:<br />

use of 1,2-propanediol in Salmonella typhimurium. J. Bacteriol. 179:1013-22.<br />

Warren, J. W., J. R. Walker, J. R. Roth and E. Altman. 2000. Construction and<br />

characterization of a highly regulable expression vector, pLAC11, and its<br />

multipurpose derivatives, pLAC22 and pLAC33. Plasmid 44:138-51.<br />

Watanabe, F., Y. Nakano, S. Maruno, N. Tachikake, Y. Tamura and S. Kitaoka.<br />

1989. NADH- and NADPH-linked aquacobalamin reductases occur in both<br />

mitochondrial and microsomal membranes of rat liver. Biochem. Biophys. Res.<br />

Commun. 165:675-9.<br />

Watanabe, F., Y. Oki, Y. Nakano and S. Kitaoka. 1987. Purification and<br />

characterization of aquacobalamin reductase (NADPH) from Euglena gracilis. J.<br />

Biol. Chem. 262:11514-8.<br />

Watanabe, F., H. Saido, R. Yamaji, K. Miyatake, Y. Isegawa, A. Ito, T. Yubisui, D.<br />

S. Rosenblatt and Y. Nakano. 1996. Mitochondrial NADH- or NADPH-linked<br />

aquacobalamin reductase activity is low in human skin fibroblasts with defects in<br />

synthesis of cobalamin coenzymes. J. Nutr. 126:2947-2951.<br />

Watkins, D., N. Matiaszuk and D. S. Rosenblatt. 2000. Complementation studies in<br />

the cblA class of inborn error of cobalamin metabolism: evidence for interallelic<br />

complementation and for a new complementation class (cblH). J. Med. Genet.<br />

37:510-3.<br />

Watkins, D. and D. S. Rosenblatt. 1988. Genetic heterogeneity among patients with<br />

methylcobalamin deficiency. Definition of two complementation groups, cblE and<br />

cblG. J. Clin. Invest. 81:1690-4.<br />

Watkins, D. and D. S. Rosenblatt. 1989. Functional methionine synthase deficiency<br />

(cblE and cblG): clinical and biochemical heterogeneity. Am. J. Med. Genet.<br />

34:427-34.<br />

Watkins, D. and D. S. Rosenblatt. 2001. Cobalamin and inborn errors of cobalamin<br />

absorption and metabolism. The Endocrinologist 11:98-104.


138<br />

Watkins, D., M. Ru, H. Y. Hwang, C. D. Kim, A. Murray, N. S. Philip, W. Kim, H.<br />

Legakis, T. Wai, J. F. Hilton, B. Ge, C. Dore, A. Hosack, A. Wilson, R. A.<br />

Gravel, B. Shane, T. J. Hudson and D. S. Rosenblatt. 2002.<br />

Hyperhomocysteinemia due to methionine synthase deficiency, cblG: structure of<br />

the MTR gene, genotype diversity, and recognition of a common mutation, P1173L.<br />

Am. J. Hum. Genet. 71:143-53.<br />

Weissbach, H., A. Peterkofsky, B. G. Redfield and H. Dickerman. 1963. Studies on<br />

the Terminal Reaction in the Biosynthesis of Methionine. J. Biol. Chem. 238:3318-<br />

24.<br />

Weissbach, H., B. Redfield and A. Peterkofsky. 1961. Conversion of vitamin <strong>B12</strong> to<br />

coenzyme <strong>B12</strong> in cell-free extracts of Clostridium tetanomorphum. J. Biol. Chem.<br />

236:PC40-2.<br />

Weissbach, H., B. G. Redfield and A. Peterkofsky. 1962. Biosynthesis of the <strong>B12</strong><br />

coenzyme: requirements for release of cyanide and spectral changes. J. Biol. Chem.<br />

237:3217-22.<br />

Wilson, A., D. Leclerc, D. S. Rosenblatt and R. A. Gravel. 1999. Molecular basis for<br />

methionine synthase reductase deficiency in patients belonging to the cblE<br />

complementation group of disorders in folate/cobalamin metabolism. Hum. Mol.<br />

Genet. 8:2009-16.<br />

Zass, R., D. Leupold, M. A. Fernandez and U. Wendel. 1995. Evaluation of prenatal<br />

treatment in newborns with cobalamin-responsive methylmalonic acidaemia. J. Inh.<br />

Met. Dis. 18:100-101.


BIOGRAPHICAL SKETCH<br />

Nicole Aurora Leal was born to Raul and Adrianne Leal, on August 2, 1976, in<br />

Santo Domingo, Dominican Republic. Nicole and her family moved to New York City<br />

during the summer of 1979; and 2 years later, to South Florida, where she learned<br />

English while attending Pinecrest Elementary School. She attended Plantation Key<br />

Middle School, and Coral Shores High School, in the Florida Keys. It was during middle<br />

school that Nicole first starting taking an active interest in music and science. During her<br />

final years of middle school, the tragic loss of her brother Ryan to cancer awakened a<br />

desire to study the life sciences. In high school, she continued to pursue both music and<br />

science. During this time, Nicole was a member of the honors choir, competed in<br />

regional competitions, and was selected to represent her school at choral camp at the<br />

University of Miami. She was a key member of the Coral Shores High School Odyssey<br />

of the Mind team, where she competed at both the regional and state levels winning the<br />

Ranatra Fusca award for creativity in consecutive years. Throughout High School,<br />

Nicole volunteered extensively with Youthwish, a nonprofit organization helping<br />

children displaced by Hurricane Andrew; and Habitat for Humanity. She earned a degree<br />

in microbiology and cell science (with a minor in chemistry) at the University of Florida,<br />

in Gainesville, in the Spring of 1999. During this time, she conducted undergraduate<br />

research in the laboratory of Dr. Thomas Bobik. In the Summer of 1999, she continued<br />

working in Dr. Bobik’s laboratory, studying <strong>B12</strong>-dependent 1,2-propanediol degradation<br />

in Salmonella. She was accepted into Graduate School in the Fall of 1999, and continued<br />

139


140<br />

her work on <strong>B12</strong>-dependent propanediol degradation, with the successful identification<br />

and partial characterization one of the metabolic enzymes involved in this pathway. In<br />

the summer of 2001, she switched her focus of study to <strong>B12</strong> metabolism in humans, with<br />

specific interest in identifying the enzymes involved in AdoCbl formation, which is the<br />

basis of this dissertation. She plans to continue her academic study and intellectual<br />

development as a postdoctoral fellow.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!