j8u mo. hill/67531/metadc... · was characterized and its interaction with other pyrimidine enzymes...

199
3~T\ /J8U Mo. Hill PYRIMIDINE BIOSYNTHESIS IN THE GENUS Streptomyces. CHARACTERIZATION OF ASPARTATE TRANSCARBAMOYLASE AND ITS INTERACTION WITH OTHER PYRIMIDINE ENZYMES DISSERTATION Presented to the Graduate Council of the University of North Texas in Partial Fulfillment of the Requirements for the Degree of DOCTOR OF PHILOSOPHY By Lee E. Hughes, B.A., M.S, Denton, Texas May, 1998

Upload: others

Post on 23-Oct-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

  • 3~T\

    /J8U

    Mo. Hill

    PYRIMIDINE BIOSYNTHESIS IN THE GENUS Streptomyces.

    CHARACTERIZATION OF ASPARTATE TRANSCARBAMOYLASE

    AND ITS INTERACTION WITH OTHER

    PYRIMIDINE ENZYMES

    DISSERTATION

    Presented to the Graduate Council of the

    University of North Texas in Partial

    Fulfillment of the Requirements

    for the Degree of

    DOCTOR OF PHILOSOPHY

    By

    Lee E. Hughes, B.A., M.S,

    Denton, Texas

    May, 1998

  • Hughes, Lee E., Pyrimidine biosynthesis in the crenus

    Streptomyces: Characterization of aspartate

    transcarbamoylase and its interaction with other pyrimidine

    enzymes. Doctor of Philosophy (Microbiology), May, 1998, 188

    pp., 16 tables, 43 illustrations, references, 181 titles.

    Aspartate transcarbamoylase (ATCase) of Streptomyces

    was characterized and its interaction with other pyrimidine

    enzymes explored. ATCase and dihydroorotase (DHOase) of S.

    griseus were assayed and purified using conventional methods

    and HPLC. The two activities were found to co-purify,

    suggesting both are found in a complex. The S. griseus

    holoenzyme has an estimated molecular mass of 480 kDa.

    Examination by SDS-PAGE revealed that the holoenzyme is

    composed of two types of subunits. Western blot analysis

    using antibody to E. coli PyrB (ATCase catalytic subunit)

    showed a cross reaction with the 38 kDa subunit from the S.

    griseus ATCase/DHOase complex. Only two other bacterial

    ATCases, from Deinococcus and Thermus, have been reported to

    contain both ATCase and DHOase catalytically-active

    subunits. The similarity of these ATCases to that of S.

    griseus is surprising considering the wide divergence

    between these organisms and streptomycetes.

    The S. griseus ATCase showed typical Michaelis-Menten

    kinetics for velocity versus substrate plots for both

    aspartate and carbamoylphosphate. These kinetics and the

    observed molecular mass place Streptomyces ATCase in the

  • Class A bacterial ATCases. Like other Class A enzymes, S.

    grriseus ATCase was inhibited by ATP, CTP, and UTP.

    Interactions between the biosynthetic and salvage

    pathways were also found. A 5-fluorouracil resistant mutant

    of S. griseus was selected which lacked uracil

    phosphoribosyltransferase (Upp) activity. This strain was

    found to be derepressed 8-10 fold for ATCase and DHOase.

    Furthermore, ATCase in this strain was more sensitive to

    effector molecules than was the wild-type enzyme.

    Considering the positive selection method used, there is

    likely to be a mutation in only one gene, upp. Therefore,

    the observed loss of Upp activity, derepression of the

    pathway, and increased effector sensitivity are linked.

    This would be possible only if the same polypeptide were

    involved in each.

    Genetic studies using PCR showed that the Streptomyces

    pyrimidine operon is organized similarly to that of

    Mycoba cterium.

  • 3~T\

    /J8U

    Mo. Hill

    PYRIMIDINE BIOSYNTHESIS IN THE GENUS Streptomyces.

    CHARACTERIZATION OF ASPARTATE TRANSCARBAMOYLASE

    AND ITS INTERACTION WITH OTHER

    PYRIMIDINE ENZYMES

    DISSERTATION

    Presented to the Graduate Council of the

    University of North Texas in Partial

    Fulfillment of the Requirements

    for the Degree of

    DOCTOR OF PHILOSOPHY

    By

    Lee E. Hughes, B.A., M.S,

    Denton, Texas

    May, 1998

  • ACKNOWLEDGMENT

    I would like to gratefully acknowledge that this work

    was made possible in part by the support of a Grant-in-Aid

    of Research from Sigma Xi, The Scientific Research Society.

    111

  • TABLE OF CONTENTS

    Page

    LIST OF TABLES vi

    LIST OF ILLUSTRATIONS vii

    INTRODUCTION 1

    Pyrimidine biosynthetic pathway 4 Genetic organization 6 Aspartate transcarbamoylase 13 Dihydroorotase 25 Streptomyces 35

    METHODS 44

    Bacterial strains 44 Media and growth conditions 44 Streptomyces spore suspension 46 Selection of 5-fluorouracil resistant strain 49 Harvesting of bacterial cultures 50 Preparation of cell extracts 51 Aspartate transcarbamoylase assay 52 Dihydroorotase assay 55 Growth curve 56 Protein concentration determination 57 Purification strategies 58 Streptomycin sulfate precipitation .59 Ammonium sulfate fractionation 59 Sephadex G-200 column chromatography 61 Sephacryl S-400 column chromatography 62 Concentration of size-exclusion fractions 62 High performance liquid chromatography ion-exchange .63 Nondenaturing polyacrylamide activity gels 63 Sodium dodecyl sulfate denaturing polyacrylamide gel electrophoresis 66

    Western blot 68 Isolation of chromosomal DNA 70 Design of oligonucleotides for polymerase chain

    reaction 72 Polymerase chain reaction 73

    xv

  • TABLE OF CONTENTS (continued)

    Page

    RESULTS 80

    Carbamoylaspartate standard curve 80 Determination of optimum pH 80 Linearity of ATCase activity . .82 Growth curve 82 Stability of Streptomyces enzymes 88 Purification of ATCase/DHOase 92 Kinetics of ATCase 112 Effector response of ATCase 121 DHOase 121 Transcriptional control 138 Polymerase chain reaction products 138

    DISCUSSION 145

    Pyrimidine enzyme levels during growth phases . . . 145 Enzyme complex 147 Genetic organization 149 Stability of Streptomyces enzymes 153 Kinetics and effector response of ATCase 155 Conclusions 159

    REFERENCES 161

    v

  • LIST OF TABLES

    Table Page

    1. Linkage map locations of pyrimidine genes in E. coli

    and S. typhimurium 6

    2. Overview of properties of bacterial ATCase classes . .17

    3. Properties of known bacterial Class A ATCases 26

    4. Properties of known bacterial Class B ATCases 28

    5. Properties of known bacterial Class C ATCases 29

    6. Properties of known bacterial ATCases - class not yet described 30

    7. Properties of known eukaryotic ATCases - plant and other monofunctional 31

    8. Properties of known eukaryotic ATCases -

    multifunctional 33

    9. Bacterial strains 45

    10. Oligonucleotide primers for PCR 74

    11. Results of various strategies for purification of S. griseus ATCase 100

    12. Results of various strategies for purification of S. griseus DHOase 102

    13. Representative purification of S. griseus ATCase and DHOase for SDS-PAGE 104

    14. Determination of molecular mass (kDa) of ATCase enzyme in Streptomyces Ill

    15. Comparison of percentage of inhibition of ATCase by effector molecules 126

    16. Comparison of concentration of effectors necessary for 50% inhibition of ATCase specific activity . . 135

    vx

  • LIST OF ILLUSTRATIONS

    Figure Page

    1. Pyrimidine biosynthetic pathway 2

    2. Schematic diagram of bacterial pyrimidine operons... 8

    3. Schematic diagram of the S. cerevisiae ura2 locus. . .11

    4. Schematic diagram of the Dictyostelium, Drosophila,

    and hamster CAD locus 12

    5. The reaction catalyzed by ATCase 14

    6. Classes of bacterial ATCase 15

    7. Model of the domain structure of CAD 22

    8. Genetic map of S. coelicolor A3(2) showing relative positions of markers conferring uracil requirement .43

    9. Apparatus used for preparation of spore suspensions

    10. Standard curve of carbamoylaspartate concentration

    11. Effect of pH on ATCase activity of S. griseus 10137

    12. Effect of pH on ATCase activity of S. venezuelae .

    13. Effect of pH on ATCase activity of S. lividans . .

    .47

    .81

    .83

    .84

    .85

    14. Change in absorbance over time for S. griseus 10137 ATCase activity 86

    15. S. griseus 10137 extract concentration and ATCase activity 87

    16. Dry weight determination, ATCase specific activity, and DHOase specific activity during growth 89

    17. Activity of ATCase of S. griseus 10137 after storage at various conditions 90

    18. Activity of DHOase of S. griseus 10137 after storage at various conditions 93

    VIX

  • LIST OF ILLUSTRATIONS (continued)

    Figure Page

    19. Activity of ATCase and DHOase after fractionation on a Sephacryl S-400 column 95

    20. Activity of ATCase and DHOase after fractionation on an HPLC ion-exchange column 97

    21. (a) SDS-PAGE of partially-purified S. griseus ATCase/DHOase. (b) Western blot of SDS-PAGE using antibody to E. coli PyrB 105

    22. Activity of ATCase and DHOase of S. griseus 10137 extract after fractionation on a Sephadex G-200 column 106

    23. Activity of ATCase and DHOase of S. griseus 10137 extract, which had been stored for six weeks, after fractionation on a Sephadex G-200 column 108

    24. ATCase activity on 4-20% gradient polyacrylamide gel 110

    25. Activity of ATCase and DHOase of S. venezuelae extract after fractionation on a Sephadex G-200 column 113

    26. Activity of ATCase and DHOase of S. lividans extract after fractionation on a Sephadex G-200 column

    27. Velocity versus substrate plot for S. griseus 10137 ATCase as a function of aspartate concentration. . 117

    28. Lineweaver-Burk plot for S. griseus 10137 ATCase as a function of aspartate concentration 118

    29. Velocity versus substrate plot for S. griseus 10137 ATCase as a function of carbamoylphosphate concentration

    30. Lineweaver-Burk plot for S. griseus 10137 ATCase as a function of carbamoylphosphate concentration . . 120

    vixi

  • LIST OF ILLUSTRATIONS (continued)

    Figure Page

    31. Velocity versus substrate plot for S. griseus 10137 ATCase with changing aspartate concentration in the presence and absence of effector molecules . . . . 122

    32. Lineweaver-Burk plot for S. griseus 10137 ATCase with changing aspartate concentration in the presence and absence of effector molecules 123

    33. Velocity versus substrate plot for S. griseus 10137 ATCase with changing carbamoylphosphate concentration in the presence and absence of effector molecules. 124

    34. Lineweaver-Burk plot for S. griseus 10137 ATCase with changing carbamoylphosphate concentration in the presence and absence of effector molecules . . . . 125

    35. Effector response of S. griseus ATCases in the presence of changing concentrations of ATP 127

    36. Effector response of S. griseus ATCases in the presence of changing concentrations of CTP 129

    37. Effector response of S. griseus ATCases in the presence of changing concentrations of GTP 131

    38. Effector response of S. griseus ATCases in the presence of changing concentrations of UTP 133

    39. Velocity versus substrate plot for S. griseus 10137 DHOase as a function of dihydroorotate concentration 136

    40. Lineweaver-Burk plot for S. griseus 10137 DHOase as a

    function of dihydroorotate concentration 137

    41. Products from PCR of S. coelicolor M145 DNA 139

    42. M. tuberculosis operon with PCR primer locations and orientations 140

    43. Probable Streptomyces operon with PCR primer locations and orientations 143

    xx

  • INTRODUCTION

    The role of pyrimidines as building blocks in the

    informational macromolecules ribonucleic acid (RNA) and

    deoxyribonucleic acid (DNA) makes the study of pyrimidine

    synthesis and regulation important. Along with purines, the

    pyrimidines are essential for cellular growth and for the

    passing of genetic information to subsequent generations.

    Pyrimidines are found in all organisms and are six-membered,

    aromatic heterocyclic ring compounds. Pyrimidine

    nucleosides consist of a pyrimidine base plus a pentose

    sugar, generally either ribose or 2'-deoxyribose, while

    nucleotides are nucleosides plus one or more phosphate

    groups. The pyrimidine ring is a component of coenzymes

    such as nicotinamide adenine dinucleotide (NAD) and acetyl

    coenzyme A (acetylCoA), and nucleotides are used in the

    formation of activated intermediates in carbohydrate and

    lipid metabolism.

    There are six enzymatic steps in the biosynthesis of

    uridine-5'-monophosphate (UMP; Fig. 1), which itself serves

    as a precursor for all the pyrimidine nucleotides. The

    pathway appears to be universal and follows the same

    sequence in all organisms thus far studied (0'Donovan &

    Neuhard, 1970; Grogan & Gunsalus, 1993). This pathway has

    been studied extensively in bacteria, fungi, plants, and

  • Figure 1. Pyrimidine biosynthetic pathway in Escherichia

    coli and Salmonella typhimurium. Gene symbols and the

    enzymes they encode are: ndk - nucleoside diphosphate

    kinase; pyrA (carAB) - carbamoylphosphate synthetase; pyrBI

    - aspartate transcarbamoylase; pyrC - dihydroorotase; pyrD -

    dihydroorotate dehydrogenase; pyrE - orotate

    phosphoribosyltransferase; pyrF - OMP decarboxylase; pyrG -

    CTP synthetase; pyrH - UMP kinase. Adapted from Neuhard &

    Nygaard (1987) .

  • CD C "c

    as

    X +

    X Q <

    X

    X O O O =o

    CL Q_

    DC O-s

    Q_ CL X

    M Q <

    0=0 Z X I 0=0

    1 - Q

    x Q -

    X o O O X -O,

    0 = 0

    - Q

    ~o \ . * 0 \ o co Z X -o O

    / £1 T> O

    X y -

    / X O o o X -e>

    0 = 0 zx

    X Z °x 8

  • animals. Although the sequence is virtually the same in

    most organisms, regulation of the pathway and organization

    of the enzymes vary in different organisms.

    Pyrimidine biosynthetic pathway.

    The first step in the synthesis of pyrimidines is

    catalyzed by the enzyme carbamoylphosphate synthetase

    (CPSase, EC 6.3.5.5). The reaction utilizes bicarbonate,

    ammonium ions or glutamine, and two molecules of adenosine-

    s' -triphosphate (ATP) in the formation of one molecule of

    carbamoylphosphate and adenosine-5'-diphosphate (ADP)

    (Anderson & Meister, 1965; Kalman et al., 1966). The

    glutamine amidotransferase activity is analogous to that

    found in a later reaction of the pathway, CTP synthetase.

    Carbamoylphosphate is required for both arginine and

    pyrimidine synthesis (Abdelal et al., 1969). While bacteria

    other than Bacillus possess only one CPSase, which satisfies

    both pathways, eukaryotes possess two CPSase enzymes, one

    for each pathway.

    The formation of carbamoylaspartate (CAA) by aspartate

    transcarbamoylase (ATCase, EC 2.1.3.2) is the first

    committed step in pyrimidine biosynthesis. Aspartate is

    carbamoylated at the amino group, producing CAA and

    releasing inorganic phosphate. The structure and regulation

    of ATCase is discussed in detail elsewhere in this paper.

  • The next reaction involves the cyclization of CAA, with

    the release of a molecule of water, to produce

    dihydroorotate (DHO). This step is catalyzed by the enzyme

    dihydroorotase (DHOase, EC 3.5.2.3). This enzyme is

    discussed in more detail later in this paper. In the

    following step, DHO is oxidized to orotate (OA) in a

    reaction catalyzed by dihydroorotate dehydrogenase

    (DHOdehase, EC 1.3.3.1). The first pyrimidine nucleotide is

    then produced by the transfer of ribose-5'-phosphate from

    5'-phosphoribosyl-1'-pyrophosphate (PRPP) to OA to form

    orotidine-5'-monophosphate (OMP), a reaction catalyzed by

    orotate phosphoribosyltransferase (OPRTase, EC 2.2.4.10).

    OMP is decarboxylated by the enzyme OMP decarboxylase

    (OMPdecase, EC 4.1.1.23) in the final step in the production

    of UMP.

    The pyrimidine nucleoside triphosphates, uridine-5'-

    triphosphate (UTP) and cytidine-5'-triphosphate (CTP), are

    ultimately produced from UMP. In sequential steps, UMP is

    first phosphorylated to uridine-5'-diphosphate (UDP) by the

    highly specific UMP kinase (EC 2.7.4.4). UDP is further

    phosphorylated by a non-specific enzyme, nucleoside

    diphosphate kinase (NDK, EC 2.7.4.6), to form UTP. UTP is

    converted to CTP by the enzyme CTP synthetase (EC 6.3.4.2),

    which transfers an amino group from glutamine in a reaction

    analogous to that of the first step in the pathway, CPSase.

  • Table 1. Linkage map locations of pyrimidine genes in E. coli and S. typhimurium (Bachmann, 1987; Sanderson & Hurley, 1987; Neuhard & Kelln, 1996).

    Gene (enzyme)

    Map location (minutes)

    Gene (enzyme) E. coli S. typhimurium

    carAB/pyrA (CPSase) 0.6, 0.7 1

    pyrBI (ATCase) 96.3 98

    pyrC (DHOase) 24.2 23

    pyrD (DHOdehase) 21.6 20

    pyrE (OPRTase) 82 .1 79

    pyrF (OMPdecase) 28.8 33

    Genetic organization.

    In many bacterial systems, the six enzymatic steps of

    pyrimidine biosynthesis are encoded by unlinked genes. In

    Escherichia coli K-12 and Salmonella typhimurium, the

    pyrimidine genes are scattered throughout the genome (Table

    1) .

    Likewise, the genes in Pseudomonas are found at

    different chromosomal locations (Holloway et al., 1990).

    However, in both P. putida and P. aeruginosa, the

    scaffolding subunit of ATCase is a catalytically-inactive,

    DHOase-like protein which is encoded by a gene immediately

    downstream of the ATCase catalytic subunit gene, pyrB (Fig.

    2). In fact, there is a four base pair overlap of these two

    open reading frames (Schurr, 1992; Vickrey, 1993). The

    catalytically active DHOase is encoded by a separate pyrC

  • located in another region of the chromosome along with argG

    (D. Brichta, personal communication).

    In Bacillus, the genes for pyrimidine biosynthesis are

    organized into a single operon (Quinn et al., 1991). The

    segment of the B. subtilis chromosome containing this gene

    cluster has been sequenced and found to contain eight

    overlapping cistrons encoding the six enzymes of pyrimidine

    biosynthesis, as well as genes for uracil permease (pyrP)

    and a regulatory protein (pyrR) (Turner et al., 1994) (Fig.

    2) . A similar genetic arrangement has been found in B.

    caldolyticus (Ghim et al., 1994; Ghim & Neuhard, 1994).

    Other bacterial pyrimidine operons have also been described

    (Fig. 2).

    The mechanism for control of gene expression has been

    reported in two of these bacterial systems. The pyrBI

    operon of E. coli is regulated by a rho-independent

    attenuator sequence (Roof et al., 1982). Limitation of

    pyrimidine triphosphates would cause RNA polymerase to pause

    in the pyrimidine-rich tracts in that area. This pause

    permits the ribosome to reach this region, coupling

    transcription and translation. The ribosome disrupts the

    formation of the terminator hairpin and allows read-through

    by the RNA polymerase to the pyrBI structural genes.

    Likewise, the B. subtilis pyr gene cluster is preceded by

    three sets of mutually exclusive antiterminator/terminator

  • Figure 2. Schematic diagram of bacterial pyrimidine operons

    (Not drawn to scale). Overlapping genes are shown on

    separate lines. Gene symbols and the enzymes they encode:

    bbc, orf2, orfl, x - unknown; pyrAA - glutaminase; pyrAB -

    carbamoylphosphate synthetase; pyrB - aspartate

    transcarbamoylase; pyrC - dihydroorotase; pyrC' - inactive

    dihydroorotase-like; pyrD, pyrDa, pyrDb - dihydroorotate

    dehydrogenase; pyrE - orotate phosphoribosyltransferase;

    pyrF - OMP decarboxylase; pyrl - ATCase regulatory subunit;

    pyrP - uracil permease; pyrR - regulatory protein; upp -

    uracil phosphoribosyltransferase. Adapted from: Bacillus

    subtilis (Turner et al., 1994); B. caldolyticus (Ghim et

    al., 1994); Lactobacillus plantarum (Elagoz et al., 1996);

    L. leichmannii (Schenk-Groninger et al., 1995); Clostridium

    acetobutylicum (Genome Therapeutics Corp., available on the

    Internet at http://pandora.cric.com/htdocs/sequences/

    clostridium/clospage.html); Pseudomonas aeruginosa (Vickrey,

    1993); P. putida (Schurr et al., 1995); Thermus aquaticus

    (Van de Casteele et al., 1994); Mycobacterium leprae; and M.

    tuberculosis (PhiHipp et al., 1996).

    http://pandora.cric.com/htdocs/sequences/

  • B. subtilis

    pyrR pyrP pyrB pyrC pyrAA pyrAB or£2 pyrD pyrF pyrE

    B. caldolyticus

    pyrR pyrP pyrB pyrC pyrAA pyrAB orf2 pyrD pyrF pyrE

    L. pi ant arum

    pyrR pyrB pyrC pyrAA pyrAB pyrD pyrF pyrE

    L. leichmannii

    pyrB pyrC

    C. acetobutylicum

    pyrB pyrl pyrDb pyrDa pyrF

    P. aeruginosa and P. putida

    pyrR pyrB pyrC'

    T. aquaticus

    upp pyrB bbc pyrC

    M. tuberculosis and M. leprae

    pyrR pyrB pyrC orfl pyrAA pyrAB pyrF

  • 10

    structures which may be regulated by a protein-mediated

    equilibrium (Turner et al., 1994).

    There are also several known schemes for pyrimidine

    gene organization in eukaryotes. In Saccharomyces

    cerevisiae, genetic analysis has shown that five independent

    genes are involved in the biosynthesis of UMP and expression

    of these genes are sequentially induced by intermediates of

    the pathway (Lacroute, 1968) . In this yeast, the activities

    of the first two enzymes, CPSase and ATCase, are coded by

    the same genetic region (ura2) and form a single enzymatic

    complex. The four enzymes that follow later in the pathway

    are encoded by separate genetic loci and are induced in a

    sequential way by the intermediary products. This same

    genetic organization has been found for Neurospora crassa

    (Caroline, 1969). Sequencing of the ura2 locus in S.

    cerevisiae showed that this gene includes a DHOase-like

    domain between the CPSase and ATCase domains (Souciet et

    al., 1989) (Fig. 3). The same has been found for

    Schizosaccharomyces (Lollier et al., 1995).

    In the slime mold Dictyostelium discoideum, in

    Drosophila melanogaster, and in mammals, a similar genetic

    arrangement to that of yeast is found for the first three

    enzymatic activities of the pyrimidine pathway (Rawls &

    Fristrom, 1975; Freund & Jarry, 1987; Faure et al., 1989)

    (Fig. 4). However, in these cases, the CPSase, DHOase and

    ATCase domains are all catalytically active (Coleman et al..

  • 11

    Figure 3. Schematic diagram of the S. cerevisiae ura2

    locus. Numbers indicate the number of base pairs from the

    start of the gene. White spaces are interdomain linker

    regions. Adapted from Souciet et al. (1989).

    400 440

    ii IlllpllMll 1

    1480 1490 1819 1907 2212

    GATase domain DHOase-like domain

    CPSase domain ATCase domain

  • 12

    Figure 4. Schematic diagram of the Dictyostelium,

    Drosophila, and hamster CAD locus. White spaces are

    interdomain linker regions. Adapted from Davidson et al.

    (1993).

    GATase domain DHOase domain

    CPSase domain ATCase domain

  • 13

    1977; Simmer et al., 1989, 1990a, 1990i>) . This

    multifunctional protein is known as CAD or multienzyme

    pyrl-3 (Jones, 1980) . A similar situation has also been

    shown for the final two steps of UMP biosynthesis in these

    organisms. A single polypeptide, UMP synthase or

    multienzyme pyr5, 6, contains both OPRTase and OMPdecase

    activities (Jones, 1980). Jones (1980) suggests that a

    protein similar to multienzyme pyrl-3 may exist in all

    animals higher than Diptera.

    In other eukaryotes, though, these multienzymic,

    pyrimidine proteins are not found. In Chlorella, a green

    alga, ATCase and DHOase activities reside on different

    proteins (Dunn et al., 1977), while the activities for

    CPSase, ATCase and DHOase have also been shown to be on

    separate proteins in higher plants (Yon, 1972; Achar et al.,

    1974; Doremus, 1986; Bartlett et al., 1994). For parasitic

    protozoans, biosynthesis of pyrimidines is catalyzed by six

    discrete enzyme activities (Krungkrai et al., 1990).

    Aspartate transcarbamoylase.

    ATCase catalyzes the transfer of the carbamoyl group of

    carbamoylphosphate to the a-amino group of L-aspartate to

    form CAA and phosphate (Fig. 5). As the enzyme catalyzing

    the first step unique to pyrimidine biosynthesis, the study

    of the structure and regulation of ATCase is important in

  • 14

    COO" COO"

    I I CH2 0 CH2 0 I II I II CH2-NH, + C-NH2 CH2-NH2-C-NH2 + P04"

    3

    I 1 , 1 COO" O-PO3"2 COO"

    Aspartate Carbamoyl- Carbamoyl- Phosphate phosphate aspartate

    Figure 5. The reaction catalyzed by aspartate transcarbamoylase.

    understanding the catalytic mechanism and evolutionary

    history of this enzyme in different organisms.

    The catalytic activity of ATCase resides in a trimer of

    identical subunits. This activity is dependent on the

    formation of an active site from half-sites on separate

    subunits (Honzatko et al., 1982; Rosenbusch & Weber, 1971;

    Stevens et al., 1991). Each trimeric unit forms three

    active sites. Regulation of enzyme activity is accomplished

    in a different manner in each of the types of ATCase

    discussed below.

    Three classes of bacterial ATCase have been described

    (Fig. 6), with varying molecular weights of the holoenzyme,

    quaternary structure, and enzyme kinetics (Bethell & Jones,

    1969; Wild et al., 1980) (Table 2).

  • 15

    Figure 6. Classes of bacterial ATCase. Adapted from Bergh

    & Evans (1993).

  • Class A ATCases

    Catalytic Trimer

    16

    45 kDa

    34 kDa

    Class B ATCases

    Catalytic Trimer

    Regulatory Dimer

    17 kDa

    34 kDa

    Class C ATCases

    Catalytic Trimer • 34 kDa

  • 17

    Table 2. Overview of properties of bacterial ATCase classes.

    Class Properties

    A e.g. Pseudomonas aeruginosa

    ~500 kDa molecular mass Dodecamer, only holoenzyme active Michaelis-Menten kinetics Inhibition by ATP, CTP, UTP

    B e.g. Escherichia coli

    ~300 kDa molecular mass Dodecamer, catalytic trimer also active

    Sigmoidal kinetics for dodecamer, Michaelis-Menten for trimer

    Allosteric Activation by ATP Inhibition by CTP, UTP

    C e.g. Bacillus subtilis

    ~100 kDa molecular mass Trimer Michaelis-Menten kinetics No effector response

    The largest ATCases are found in Class A. One example,

    fluorescens ATCase, has a complex with a molecular mass

    of 464 kDa. This complex appears to consist of a 1:1 ratio

    of 34 kDa and 45 kDa polypeptides (Bergh & Evans, 1993) .

    The presence of two polypeptides is consistent with recent

    sequencing studies in P. aeruginosa and P. putida which show

    two open reading frames encoding polypeptides of

    approximately similar sizes (Schurr, 1992; Vickrey, 1993;

    Schurr et al., 1995). Based on the mass of the constituent

    subunits and the stoichiometry of the complex, the protein

    must be a dodecamer (Bergh & Evans, 1993). Similar findings

    have also been reported for P. fluorescens and P. syringae

    (Shepherdson & McPhail, 1993). The ATCase of P. putida has

  • 18

    been shown to have a molecular mass of 482 kDa, with the

    smaller catalytic polypeptide having a molecular mass of

    36.4 kDa. The larger polypeptide, 44.2 kDa, is encoded by

    the DHOase-like gene but does not have DHOase activity. The

    DHOase-like subunit is required for activity in the

    holoenzyme, as no activity has been found solely for the

    catalytic trimers (Schurr et al., 1995). Characteristics

    for Class A enzymes include hyperbolic substrate saturation

    curves and inhibition by ATP, UTP, and CTP which is

    competitive with carbamoylphosphate and noncompetitive with

    aspartate (Neumann & Jones, 1964; Bethell & Jones, 1969;

    Linscott, 1996). The nucleotide effector binding site has

    been localized to the catalytic polypeptide, not to the 45

    kDa polypeptide (Bergh & Evans, 1993; Schurr et al., 1995).

    Class A ATCases have been found to be widespread (Table 3).

    The Class B ATCases are smaller and are distinguished

    by sigmoidal substrate saturation curves as typified by the

    enzyme from E. coli. The enzyme is a dodecamer. Six

    catalytic polypeptides, each with a molecular mass of 34

    kDa, are grouped into two trimers. The trimers are

    connected by three regulatory dimers. Each subunit of the

    regulatory dimer has a molecular mass of 17 kDa (Kantrowitz

    & Lipscomb, 1988). The holoenzyme molecular mass is 306 kDa

    (Fig. 6). Activity of the trimers is retained when they are

    separated from the regulatory subunits by chemical means,

    such as p-chloromercuribenzoate (Blackburn & Schachman,

  • 19

    1977; Subramani & Schachman, 1981). The enzyme is under

    allosteric control, whereby the activity is influenced

    through the binding of effector molecules at a site other

    than the active site. The nucleotide binding site is

    located between pairs of regulatory subunits. E. coli

    ATCase is inhibited by physiological levels of CTP (Yates &

    Pardee, 1956; Gerhart & Pardee, 1962, 1964) and activated by

    ATP. Subunit interactions are responsible for sigmoidal

    concentration curves for both aspartate (Gerhart & Pardee,

    1962) and carbamoylphosphate (Bethell et al., 1968). Class

    B ATCases have been isolated from several representatives of

    the Enterobacteriaceae (Wild et al., 1980), as well as other

    Gram negative bacteria and even an archaebacterium,

    Pyrococcus abyssi (Purcarea et al., 1994) (Table 4).

    The Class C ATCases are characterized by their small

    size, insensitivity to pyrimidine nucleotide effectors, and

    typical Michaelis-Menten kinetics in the carbamoylphosphate

    and the aspartate saturation curves. The B. subtilis enzyme

    represents a typical Class C ATCase. The native enzyme is a

    trimeric protein with a molecular mass of 102 kDa,

    consisting of three 33.5 kDa polypeptides (Brabson &

    Switzer, 1975) (Fig. 6). Class C ATCases have been found in

    other Gram-positive organisms, Streptococcus faecalis (Chang

    et al., 1974) and Staphylococcus epidermidis (Kenny et al.,

    1996), as well as in several Gram-negative organisms,

  • 20

    Stenotrophomonas maltophilia, Xanthomonas campestris, and

    Lysobacter enzymogenes (Kenny et al., 1996).

    Recently, several bacterial ATCases have been described

    which do not at present appear to fit within any of the

    three classes. Ishihara et al. (1992) cloned the ATCase

    gene of Treponema denticola in their studies of antigenic

    proteins from this organism. This sequence produces a 55

    kDa protein with 33.8% homology to the ATCase of E. coli.

    No pyrT-like sequence was found downstream of the gene, and

    the enzyme was not inhibited by CTP. Unfortunately, no

    holoenzyme size is reported, so it not possible to determine

    if the enzyme is a homotrimer or if it contains additional

    subunits.

    Unusual ATCases have also been reported for Thermatoga

    maritima and Thermus aquaticus (Van de Casteele et al.,

    1994) . The Thermatoga clone in the study produced a

    truncated gene product of approximately 33 kDa, however the

    authors suggest that the actual polypeptide is closer in

    size to that of T. denticola based on homology of the two

    sequences. The gene of T. aquaticus produces a polypeptide

    of about the same size as the E. coli or Bacillus catalytic

    subunit. The holoenzyme, which is inhibited by UTP, has a

    molecular mass of 480 kDa and contains both ATCase and

    DHOase activities. This enzyme adds a curious twist to the

    evolutionary puzzle when compared to the Class A bacterial

  • 21

    ATCases with their inactive, DHOase-like subunit and similar

    holoenzyme sizes.

    In the multienzymic proteins of higher organisms,

    ATCase is not under allosteric control. Instead, CPSase is

    the site of regulation, with UTP and CTP acting as

    inhibitors and ATP as an activator (Jones, 1980) . The yeast

    ura2 gene product of Saccharomyces has a predicted molecular

    mass of 245 kDa (Souciet et al., 1989), while the CAD

    polypeptide has a molecular mass of about 240 kDa (Kelly et

    al., 1986). The CAD polypeptide associates as trimers and

    hexamers (Coleman et al., 1977). Trimers form by an

    association around the ATCase domain (Fig. 7), much like

    that seen for bacterial ATCases, while the hexamers are due

    to dimeric interactions between DHOase domains of adjacent

    trimers (Carrey, 1993; Davidson et al., 1993). It has been

    suggested that channeling of substrates is an advantage

    conferred by multienzymic proteins (Lue & Kaplan, 1970).

    This may occur in effect through a process of facilitated

    diffusion between active sites that are close together in

    space (Carrey, 1993). Evidence has also been found for

    channeling in the yeast enzyme (Penverne & Herve, 1983;

    Belkaid et al., 1987) .

    In plants, ATCase was found to be a regulatory enzyme.

    The ATCase of plants is noteworthy since it combines

    regulatory behavior with a comparatively small molecular

    mass, 104 kDa in wheat germ (Yon et al., 1982) and 128 kDa

  • 22

    Figure 7. Model of the domain structure of CAD. ATC =

    aspartate transcarbamoylase domain; CPS = carbamoylphosphate

    synthetase domains; DHO = dihydroorotase domain; and GLN =

    glutaminase domain. Adapted from Carrey (1993) .

  • 24

    in mung bean (Achar et al., 1974). Yon et al. (1982) showed

    that the wheat germ ATCase is also a trimer. Inhibition was

    seen with ATP, CTP, and UTP, with the greatest inhibition of

    activity by the addition of UMP (Achar et al., 1974). The

    CPSase and ATCase enzymes of radish (Raphanus sativus),

    spinach (Spinacia oleracea), soybean (Glycine max), and corn

    (Zea mays) have been localized in the chloroplasts (Shibata

    et al., 1986) .

    In the protozoan Crithidia fasciculata, neither

    inhibition nor activation of ATCase was seen in the presence

    of pyrimidine ribonucleotides (Kidder et al., 1976).

    Kurelec (1974) reports that the ATCase of parasitic

    platyhelminths is not affected by CTP, but is activated by

    ATP. In some invertebrates, evidence suggests that CPSase

    and ATCase are separate polypeptides. The single CPSase of

    the land snail Strophocheilus oblongus is localized in the

    mitochondrion, while ATCase is found in the soluble fraction

    of the cell (Tramell & Campbell, 1970). A similar finding

    exists for other snails (Otala lactea and Helix aspersa),

    earthworm (Lumbricus terrestris), and land planarian

    (Bipalium kewense) (Tramell & Campbell, 1971). In some

    parasitic platyhelminths, CPSase is apparently absent,

    although ATCase is still present (Kurelec, 1972, 1973,

    1974) . These organisms require arginine and pyrimidines and

    probably derive their carbamoylphosphate from the arginine

  • 25

    deaminase pathway as is done in Lactobacillus (Hutson &

    Downing, 1968).

    Properties of reported ATCases are summarized in Tables

    3-8. The following organisms have been found to lack

    ATCase: Chlamydia psittaci (McClarty & Qin, 1993);

    Mycoplasma mycoides (Mitchell & Finch, 1977); and

    Trichomonas vaginalis (Heyworth et al., 1984).

    Dihydroorotase.

    DHOase catalyzes the reversible cyclization of

    carbamoylaspartate to dihydroorotate. The enzyme from

    Clostridium oroticum has been purified to homogeneity and

    shown to be a zinc-metalloenzyme (Taylor et al., 1976;

    Pettigrew et al., 1985a). It has also been suggested that

    the mammalian dihydroorotase is a zinc-metalloenzyme

    (Christopherson & Jones, 1980, Simmer et al., 1990£>) .

    The DHOase enzyme in prokaryotes has been found to be a

    homodimer, with subunit molecular masses in the range of 40-

    50 kDa (Pettigrew et al., 1985b; Ogawa and Shimizu, 1995;

    Schenk-Groninger et al., 1995). While all described

    prokaryotic DHOases are active in the dimeric form,

    Krungkrai et al. (1990) report that the protozoan enzyme,

    which is also approximately 40 kDa, is active as a monomer.

    The DHOase of a eukaryotic green alga, Chlorella, has an

    apparent molecular mass of 88 kDa (Dunn et al. 1977), which

    would be consistent with the prokaryotic enzymes if it is

  • 26

    Table 3. Properties of known bacterial Class A ATCases.

    Vmax reported in nmol CAA min"1 (mg protein). KmAsp and KmCP

    given in mM of the substrate. References denoted as: 1)

    Barron, 1994; 2) Bergh & Evans, 1993; 3) Bethell & Jones,

    1969; 4) Burns et al., 1997; 5) Hooshdaran, personal

    communication; 6) Kenny et al., 1996; 7) Linscott, 1996; 8)

    Masood & Venkitasubramanian, 1988; 9) Mendz et al., 1994;

    10) O'Donovan, personal communication; 11) Schurr et al.,

    1995; 12) Shepherdson & McPhail, 1993; 13) Shepherdson et

    al., 1997; 14) Van de Casteele et al., 1994; 15) Van de

    Casteele et al., 1997; 16) Vickrey, 1993; and 17) West,

    1994.

  • 27

    Organism kDa V vraax KmABp Kmcp Ref.

    Bacterial Class A ATCases

    Acinetobacter calcoaceticus 480 6

    Azomonas agilis -450 6

    Azotobacter vinelandii -450 3,6

    Brevundimonas diminuta -480 139 1.0 1.0 7

    Comamonas acidovorans -500 0.6 1.0 7

    C. testosteroni -500 1.0 0.7 7

    Deinococcus radiophilus 500 6

    Helicobacter pylori 23 11.6 0.6 4,9

    Leucothrix mucor -450 6

    Micrococcus luteus -480 1

    Mycobacterium smegmatis 465 34 5,8

    Paracoccus denitrificans -450 6

    Pseudomonas aeruginosa 484 120 1.5 0.13 3,16

    Ps. aureofaciens -480 1.3 1.0 7

    Ps. fluorescens 464 40 1.1 0.08 2,3

    Ps. mendocina -480 1.0 1.0 7

    Ps. pseudoalcaligenes 73 2.7 0.29 17

    Ps. putida 482 8 2.2 0.60 11

    Ps. syringae 490 31 1.3 0.9 12, 7

    Ps. stutzeri -480 1.0 1.0 7

    Rhizobium meliloti -480 10

    Synechocystis sp. -490 13

    Thermus aquaticus 480 20 14,15

  • 28

    Table 4. Properties of known bacterial Class B ATCases. Vmax reported in nmol CAA min"

    1 (mg protein)"1. [S]os for both substrates given in triM. References denoted as: 1) Bethell & Jones, 1969; 2) Jyssum, 1992; 3) Kenny et al., 1996; 4) Purcarea et al., 1994; 5) Purcarea et al., 1997; and 6) Wild et al., 1980.

    Organism kDa V vmax Asp [S] o s

    CP [S] o.s Ref.

    Bacterial Class B ATCases

    Aeromonas hydrophila 285 18 17 . 0 6

    Citrobacter diversus 280 46 5.5 6

    C. freundii 290 23 7.5 1,6

    Enterobacter aerogenes 295 196 4.8 6

    En. liquefaciens 295 23 17 .5 6

    Erwinia carnegiana 275 38 2.6 6

    Er. herbicola 315 37 2.9 6

    Escherichia coli 310 53 5.0 1,6

    Neisseria canis 6.7 3.4 2

    N. caviae 6.2 1.8 2

    N. elongata 5.7 1.8 2

    N. gonorrhoeae 9.2 5.4 2

    N. meningitidis Ml 295 30.1 8.1 2

    Proteus vulgaris 310 9 33 .0 1,6

    Pyrococcus abyssi 310 3.0 0.06 4,5

    Rhodopseudomonas spheroides 1

    Salmonel1 a typhimuriurn 300 51 7.0 6

    Serratia marcescens 275 20 19 .5 1,6

    Shigella flexneri 285 454 5 . 6 6

    Vibrio natriegens 280 3

    Yersinia enterocolitica 315 10 4.2 6

  • 29

    Table 5. Properties of known bacterial Class C ATCases. Vmax reported in nmol CAA min"

    1 (mg protein)_1. Km^ and Km^ given in mM of the substrate. References denoted as: 1) Barron, 1994; 2) Bethell & Jones, 1969; 3) Brabson & Switzer, 1975; 4) Chang & Jones, 1974; 5) Ghim et al., 1994; 6) Kenny et al., 1996; 7) Linscott, 1996; and 8) O'Donovan & Shanley, 1995.

    Organism kDa V v max ^^•Asp Ktricp Ref.

    Bacterial Class C ATCases

    Bacillus caldolyticus 10 5

    B. subtilis 102 380 7.0 0 .11 2,3

    Lactobacillus fermentum -100 1

    Lysobacter enzymogenes 120 6

    Shewanella putrefaciens -100 2.0 0.5 7,8

    Sporosarcina ureae -100 2.5 1

    Staphylococcus epidermidis -100 6

    Stenotrophomonas maltophilia 112 0.7 0.7 6,7,8

    Streptococcus faecalis 128 4

    Xanthomonas campestris 126 6

  • 30

    Table 6. Properties of known bacterial ATCases - other, class not yet described. Vmax reported in nmol CAA min"

    1 (mg protein) Km, LAsp and Kmcp given in mM of the substrate. References denoted as: 1) Ahonkhai et al., 1989; 2) Currier & Wolk, 1978; 3) Elagoz et al., 1996; 4) Grogan & Gunsalus, 1993; 5) Hutson & Downing, 1968; 6) Ishihara et al., 1992; 7) Li & West, 1995; 8) Linscott, 1996; 9) Makoff & Radford, 1978; 10) Norberg et al., 1973; 11) 0'Donovan & Shanley, 1995; 12) Van de Casteele et al., 1994; 13) Wheeler, 1989; and 14) Wheeler, 1990.

    Organism kDa V vraax KmAsp Kiticp Ref.

    Bacterial ATCases - Other, Class Not Yet Described

    Anabaena variabilis 3 1.0 0.70 2

    Burkholderia cepacia -600 74 2.5 2.2 7,8,11

    B. pickettii -500 30 3.5 0.9 8

    Halobacterium cutirubrum 160 15.0 2.7 9,10

    Lactobacillus leichmannii 1.4 30.0 5,9

    L. plantarum 3

    Mycobacterium avium 0.8 14

    M. leprae 13

    M. microti 1.2 14

    Pseudomonas indigofera -400 9.0 0.7 8, 11

    Sulfolobus acidocaldarius .15 4

    Thermatoga maritima 400 12

    Treponema denticola 55a 6

    Vibrio costicola » ̂ i « , — _ i - i

    130 1

  • 31

    Table 7. Properties of known eukaryotic ATCases - plant and

    other monofunctional. Vmax reported in nmol CAA min"1 (mg

    protein)"1. Km^ and Km^ given in mM of the substrate.

    References denoted as: 1) Achar et al., 1974; 2) Asai et

    al., 1983; 3) Doremus, 1986; 4) Dunn, 1977; 5) Grayson &

    Yon, 1979; 6) Hammond & Gutteridge, 1980; 7) Holland et al.,

    1983; 8) Kidder et al., 1976; 9) Kurelec, 1974; 10)

    Landstein et al., 1996; 11) Lovatt et al., 1979; 12)

    Mukherjee et al., 1988; 13) Nasr et al., 1994; 14) Neumann &

    Jones, 1964; 15) Ong & Jackson, 1972; 16) Overduin et al.,

    1993; 17) Shibata et al., 1986; 18) Tampitag & 0'Sullivan,

    1986; 19) Tramell & Campbell, 1970; 20) Williamson & Slocum,

    1994; 21) Yon, 1972; and 22) Yon et al., 1982.

  • 32

    Organism kDa V "max Kir̂ Ref.

    Eukaryotic ATCases - Plant and other monofur ictional

    Arabidopsis thaliana 13

    Babesia rodhaini 8 1.9 0.01 7

    Chlorella Virus PBCV-1 200 10

    Chlorella sorokiniana 160 4

    Crithidia fasciculata 100 320 5.0 0.5 8

    Cr. luciliae 150 21 2.6 0.03 18

    Cucurbita pepo 11

    Fasciola hepatica 9

    Glycine max 17

    Lactuca sativa 14

    Leishmania donovani 135 35 7.6 0.3 12

    Lycopers icon esculentum 16

    Moniezia benedeni 9

    Paramphistomum cervi 9

    Phaseolus aureus 128 150 4.0 0.09 1,15

    Pi sum sativum 110 16 3,20

    Raphanus sativus 17

    Spinacia oleracea 17

    Strophocheilus oblongus 19

    Toxoplasma gondii 140 17 17.6 0.03 2

    Triticum vulgare 104 4 0.6 0.01 5,21,22

    Trypanosoma cruzi 22 6

    Vinca rosea 17

    Zea mays 17

  • 33

    Table 8. Properties of known eukaryotic ATCases -multifunctional. Vmax reported in nmol CAA min"

    1 (mg protein)"1. KmAsp and Km^ given in mM of the substrate. References denoted as: 1) Caroline, 1969; 2) Christopherson & Jones, 1980; 3) Faure et al., 1989; 4) Freund & Jarry, 1987; 5) Hirsch, 1968; 6) Hong et al., 1995; 7) Hoogenraad & Lee, 1974; 8) Jarry, 1976; 9) Kent et al., 1975; 10) Kim et al., 1992; 11) Koskimies et al., 1971; 12) LaCroute, 1968; 13) Lollier et al., 1995; 14) Mally et al., 1980; 15) Penverne & Herve, 1983; and 16) Soderholm et al., 1975.

    Organism kDa V vmax KmAsp KmCP Ref.

    Eukaryotic ATCases - Multifunctional

    Chicken 900 11

    Coprinus radiatus 800 5

    Dictyostelium discoideum 3

    Drosophila melanogaster 800 5.4 0.44 4,8,16

    Neurospora crassa 46 1

    Hamster 243a 240 44

    9.05 .004 .02

    2,10,14

    Mouse 900 0.7 2,11

    Rana catesbeiana 900 30 9

    Rat 900 10 7, 11

    Saccharomyces cerevisiae 600 16.6 1.18 12, 15

    Schizosaccharomyces pombe 13

    Squalus acanthias 6

  • 34

    also a dimer. The catalytically-active yeast DHOase is

    similar to those described previously for prokaryotes, while

    this activity is found within the CAD multienzyme for higher

    eukaryotes.

    Dendograms of DHOases show that they fall into two

    distinct groups, with the multifunctional hamster,

    Dictyostelium, Drosophila, and yeast inactive domain forming

    one group while the monofunctional proteins from yeast, E.

    coli, and Salmonella typhimurium constitute a second class

    (Simmer et al., 1990Jb) . Even so, a 44 kDa proteolytic

    fragment of the CAD protein, which forms a dimer, has been

    shown to catalyze only the dihydroorotase reaction (Simmer

    et al., 1990jb) . Comparison of this domain with well

    characterized prokaryotic dihydroorotases indicated that

    many features of the common ancestral protein had been

    conserved (Kelly et al., 1986).

    More recent analysis has shown that additional DHOase

    sequences confirm the division into two diverged types

    (Shepherdson et al., 1997). Active DHOases from T.

    aquaticus, Synechocystis species, and P. aeruginosa align

    with the E. coli group described above. This group is

    designated the family C type DHOases by Shepherdson et al.

    (1997) and generally has a molecular mass of 40 kDa. The

    other group, designated family C', includes the CAD and CAD-

    like DHOases as shown by Simmer et al. (1990jb) as well as

    those from two archaeas, Methanococcus jannaschii and

  • 35

    Sulfolobus solfataricus. The family C' also contains the

    DHOase of Lactobacillus leichmanni and the DHOase-like PyrC'

    polypeptides of P. putida and Synechocystis sp., hence the

    name of the family.

    The existence of divergent families of DHOase adds to

    the debate on the evolution of the CAD polypeptide. The

    discovery of bacterial complexes containing ATCase and

    either active or inactive DHOase subunits suggested that

    these complexes could represent a CAD precursor (0'Donovan &

    Shanley, 1995). This theory is further supported by the

    presence of both CAD and these bacterial proteins in the

    same DHOase family.

    Streptomyces.

    Pyrimidine metabolism has not been studied extensively

    in one important order, the Actinomycetes. This order

    contains a wide range of genera and species which are

    generally divided into eight broad families based on partial

    sequencing of 16S rRNA. Of these families, only one, the

    family Streptomycetaceae, does not contain diverse taxa

    (Goodfellow, 1989). The members of this family are

    morphologically complex and have the ability to form spores

    in or on the mycelium. The family contains a number of

    genera, one of the most widely known being Streptomyces.

    The genus Streptomyces was proposed in 1943 by Waksman

    & Henrici for aerobic, spore-forming actinomycetes which had

  • 36

    been previously termed Actinomyces. The streptomycetes are

    found within the high G+C Gram-positive subdivision based on

    rRNA homology (Woese, 1987). They are most notable for

    their large genome size, 10.5 x 106bp {Benigni et al.,

    1975), their DNA base composition of 69 to 78 mol % guanine

    (G) plus cytosine (C) (Korn-Wendisch & Kutzner, 1992), and

    the wide range of antibiotics, vitamins, enzymes and enzyme

    inhibitors they have been found to produce (Goodfellow &

    Cross, 1983). As a result of these useful secondary

    metabolites, thousands of strains have been isolated from

    soils and sediments around the world. In addition,

    Streptomyces also stands out as one of only a few

    prokaryotes in which a linear chromosome is present

    (Hinnebusch & Tilly, 1993; Lin et al., 1993; Lezhava et al.,

    1995) .

    Streptomyces and related genera occupy a unique

    evolutionary position, being filamentous, spore-forming

    prokaryotes. The spores are hyphal in origin, developing as

    a result of septation and fragmentation of pre-existing

    hyphal elements. These spores are, at maturity, relatively

    unspecialized compartments of hyphae. They are bounded by

    walls of true peptidoglycan which are normally about 30-50

    nm thick, compared to 10-20 nm for vegetative cells (Locci &

    Sharpies, 1983).

    Following germination of the spore, the first type of

    tissue to develop is the substrate mycelium. This fungus-

  • 37

    like growth is a branching network of multinucleate hyphae,

    which are occasionally interrupted by septa (Chater, 1993).

    Substrate mycelial growth continues until, presumably, the

    nutrient supply is exhausted. At such a time,

    differentiation takes place.

    Aerial mycelia, upon which the spores will develop,

    begin to grow on the colony. The aerial hyphae form a dense

    lawn and grow perpendicular to the substrate mycelium.

    These hyphae escape the aqueous environment of the colony

    surface in order to grow into the air. This gives a fuzzy

    white appearance to the surface of the colony. During this

    stage, lysis occurs in the substrate mycelium, releasing

    nutrients which can be utilized by the aerial mycelium for

    growth (Mendez et al., 1985). The aerial mycelium then

    undergoes septation into uninucleate units, which develop

    into chains of gray-pigmented, hydrophobic spores. The

    aerial mycelium is a device for the dispersal of the spores

    (Kaiser & Losick, 1993). The use of mutants and molecular

    genetics has begun to reveal how differentiation is brought

    about in the multicellular, mycelial Streptomyces spp.

    (Chater, 1993).

    Classification within the genus has been complicated by

    the proliferation of named species. The number of species

    names that have been used for a streptomycete in the

    scientific and patent literature rose from 41 in 1920 to 153

    in 1957 to now more than 3,000. This explosion in species

  • 38

    number is certainly due in part to the many isolates

    investigated in the course of screening for antibiotics and

    other useful industrial compounds. This resulted in the

    discovery of thousands of different secondary metabolites,

    each one suggesting the existence of a distinct producer or

    "special type", which were often considered as a "new

    species" (Korn-Wendisch & Kutzner, 1992).

    Initially, classification was based purely on

    morphological features. This led, however, to many new

    genera and families being created, as well as many species

    and genera being rearranged. Overall, a very complex

    situation was created. Chemical composition of the cell

    wall provided a tool for distinguishing streptomycetes from

    other actinomycetes. Lechevalier and colleagues (Becker et

    al., 1964, 1965; Lechevalier & Lechevalier, 1970)

    demonstrated that Streptomyces and other genera of the

    family Streptomycetaceae contained LL-diaminopimelic acid

    (LL-A2pm) in its peptidoglycan. Most other actinomycetes

    described thus far contain meso-A2pm. Streptomycete

    peptidoglycan is also characterized by an interpeptide

    bridge composed of a glycine residue (Schleifer & Kandler,

    1972) . A number of other biochemical criteria have been

    used to examine Streptomyces. The Streptomycetaceae form a

    homogeneous family in regard to these, including the pattern

    of sugars in whole-cell hydrolysates (Lechevalier &

    Lechevalier, 1970), phospholipids (Lechevalier et al.,

  • 39

    1977), fatty acids (Kroppenstedt, 1985), menaquinones

    (Alderson et al., 1985; Kroppenstedt, 1985), and acetylated

    muramic acid residues (Uchida & Aida, 1977).

    Williams et al. (1983) studied 475 strains, which

    included 394 type cultures of genus Streptomyces and

    representatives of 14 other actinomycete genera. Overall

    similarities of these strains for 139 unit characters were

    determined. The results of this study provided a basis for

    the reduction of the large number of Streptomyces species

    which have been described. It demonstrated that the

    previous use of a limited number of subjectively chosen

    characters to define species-groups or species had resulted

    in artificial classifications. Their data, together with

    those from previous diverse studies, indicated that the

    genera Actinopycnidium, Actinosporangium, Chainia,

    Elytrosporangium, Kitasatoa, and Microellobosporia should be

    reduced to synonyms of Streptomyces. The status of

    Streptomyces clusters defined by Williams et al. have been

    supported by chemical (Saddler et al., 1987; Manchester et

    al., 1990), serological (Ridell & Williams, 1983), and

    nucleic acid sequencing methods (Mordarski et al., 1986;

    Witt & Stackebrandt, 1990; Labeda & Lyons, 1991; Labeda,

    1992) .

    Species classification within the genus has been

    attempted using a number of characters. Streptomyces phage-

    typing has shown that the host range pattern of a set of

  • 40

    polyvalent phages does not cross genus-boundaries and can

    also be used in taxonomic studies for identification at the

    species level (Korn-Wendisch & Schneider, 1992).

    The application of polymerase chain reaction (PCR) to

    obtain rDNA sequences has dramatically extended the

    potential to use these sequences to elucidate natural

    relationships between organisms (Stackebrandt et al., 1992).

    In the genus Streptomyces, Stackebrandt and colleagues

    observed that the 16S rRNA of too many species contains too

    similar sequences that the analysis of the complete

    sequences seems to be "unpractical and unnecessary".

    Instead, restricting analysis to two regions of the

    sequence, one for species-specific signatures and another

    for intragenic classification, appears to be sufficient for

    the purposes of streptomycete classification (Stackebrandt

    et al., 1992; Kim et al., 1993). Stackebrandt et al. also

    note that, given the limitations of highly similar 16S rRNA

    sequences, analysis of this molecule cannot be hailed as the

    solution to all problems of streptomycete taxonomy.

    Analysis of rRNA is only one piece on the jigsaw puzzle that

    eventually may lead to a more complete picture about

    streptomycete evolution (Stackebrandt et al., 1992).

    Analysis of ATCases could be another.

    While much work has been done to identify the medically

    and industrially important secondary metabolites produced by

    the members of this genus, there are many basic questions of

  • 41

    primary metabolism which have yet to be answered. These

    studies are not only necessary to help further understand

    the events leading to and controlling secondary metabolite

    production, but may also be useful as additional taxonomic

    markers in the classification of species within the genus.

    As a basic and universal pathway, the pyrimidine

    biosynthetic pathway provides an opportunity for such study.

    Enzymes of the pathway in other organisms have been used as

    taxonomic and evolutionary markers with success (Neumann &

    Jones, 1964; Bethell & Jones, 1969; Wild et al., 1980;

    Foltermann et al., 1981; Beck et al., 1989; Major et al.,

    1989; Tricot et al., 1989; Wild & Wales, 1990; Jyssum, 1992;

    Kern et al., 1992; Kimsey & Kaiser, 1992; Nagy et al., 1992;

    Davidson et al., 1993; Schofield, 1993; van den Hoff et al.,

    1995; Kenny et al., 1996; Lawson et al., 1996; Linscott,

    1996; 0'Donovan & Shanley, 1996).

    A representative species, Streptomyces griseus, was

    chosen for study. S. griseus was one of the original

    antibiotic-producing organisms discovered by Waksman. The

    genetically well-characterized species S. coelicolor was

    also used in these experiments as several pyrimidine mutants

    were available.

    Genetic mapping studies of S. coelicolor A3(2) have

    identified four uracil-requiring strains which have been

    kindly supplied for this study by Dr. D. A. Hopwood. One of

    these loci, designated uraA, maps in a separate location

  • 42

    from the others. However, the remaining three, uraB, uraC,

    and uraD, are found in close proximity of each other and are

    likely placed in the operon discovered in this research

    (Hopwood & Kieser, 1990)(Fig. 8). The enzymes encoded by

    these loci have not been identified, though the uraD strain

    requires both uracil and arginine, suggesting that this is

    the gene for CPSase. Since these mutations are found in the

    same region of the chromosome, they could represent

    mutations in different genes of an operon or within

    different domains of a multienzymatic polypeptide.

    In this research, I sought to determine the nature of

    the enzyme ATCase, its structure and regulation, in

    Streptomyces. These properties as well as the relationship,

    both enzymatically and at the genetic level, of ATCase to

    other enzymes of the pyrimidine pathway are important in

    examining the classification of this enzyme relative to

    other known ATCases. This study may also serve as a basis

    for determining the type and distribution of ATCases within

    the genus and in relation to other genera in the high G+C

    Gram-positive subdivision.

  • 43

    Figure 8. Genetic map of S. coelicolor A3(2) showing

    relative positions of markers conferring uracil requirement.

    Adapted from Hopwood & Kieser (1990).

    uraB,C,D

    uraA

  • METHODS

    Bacterial strains.

    The wild-type strain Streptomyces griseus (ATCC 10137)

    and S. lividans (ATCC 19844) were obtained from the American

    Type Culture Collection (ATCC). S. griseus LU2 was selected

    from the wild-type for 5-fluorouracil (5FU) resistance. S.

    coelicolor M145 was kindly provided by Dr. D. A. Hopwood,

    John Innes Institute, Norwich, England, while S. venezuelae

    was obtained from Ward's Biological Supply. Table 9

    includes a list of the strains utilized and their

    genotype/phenotype.

    Media and growth conditions.

    Streptomyces spore suspensions were obtained from

    growth on sporulation agar plates (Hopwood et al., 1985) or

    from nutrient agar slants. Sporulation agar contained

    1 g l"1 yeast extract, 1 g l"1 beef extract, 2 g l"1 tryptose,

    trace FeS04, 10 g l"1 glucose, and 15 g l"1 agar in ddH20.

    Difco nutrient agar contained 3 g l"1 beef extract, 5 g l"1

    peptone, and 15 g l"1 agar in H20.

    Two different defined media were used to grow the

    Streptomyces strains. The solid medium used was

    Streptomyces minimal agar medium (Hopwood, 1967), modified

    for alternate carbon source. This medium contained 2 g l"1

    44

  • 45

    Table 9. Bacterial strains

    Organism Source Genotype and/or Phenotype

    S. griseus 10137 ATCC wild-type

    S. griseus LU2 This study 5FU resistant

    S. coelicolor M145 Hopwood wild-type

    S. lividans 19844 ATCC wild-type

    S. venezuelae Ward's Biological wild-type

    B. cepacia 25416 ATCC wild-type

    E. coli K12 Bachmann, B. wild-type

    E. coli TB2 Roof, W. D. pyrB argF requires uracil and arginine

    P. aeruginosa PAOl Holloway, B. W. wild-type

    (NH4)2S04, 0.5 g l'1 K2HP04, 0.2 g l'

    1 MgS04'7H20, O.Olgl"1

    FeS04'7H20, and 15 g l"1 agar in ddH20. Succinate, pH7.0, was

    added after autoclaving to a final concentration of 20 TOM.

    Growth in liquid medium was done in Streptomyces minimal

    liquid medium (Hopwood et al., 1985), which contained 2 g l"1

    (NH4)2S04, 5 g l"1 Difco Casamino acids, 0.6 g I"1 MgS04'7H20,

    and 50 g l"1 polyethyleneglycol (PEG) 8000 in 800 ml ddH20.

    To this, 1ml minor elements solution (per liter: lg

    ZnS047H20, lg FeS04'7H20, lg MnCl24H20, and lg CaCl2,

    anhydrous) was added. After autoclaving, added 150 ml final

    volume of 0.1M NaH2P04/K2HP04 buffer, pH6.8, and carbon

    source, succinate, pH 7.0, to 20 mM final concentration.

  • 46

    Required growth factors were added at the time of

    inoculation.

    For isolation of total DNA, Streptomyces cultures were

    grown in nutrient broth with 34% (w/v) sucrose. Difco

    nutrient broth contained 3 g l"1 beef extract and 5 g l"1

    peptone in water to which 340 g l"1 sucrose was added.

    Streptomyces cultures were grown at 30 °C. Liquid

    cultures were shaken vigorously on an orbital shaker, in a

    baffled flask when available.

    Streptomyces spore suspension.

    Spores were harvested from plate cultures of

    Streptomyces according to a modified version of the

    procedure of Hopwood et al. (1985). All solutions and

    apparati were sterilized by autoclaving prior to use.

    Water, typically 9 ml, was added to a well-sporulating plate

    (2-3 ml for slant cultures). The agar surface was scraped

    with a sterile loop to suspend the spores. The liquid was

    then removed with a pipette and transferred to a test tube.

    The suspension was mixed vigorously on a vortex mixer for 1-

    2 minutes (min), then filtered by vacuum through fiberglass

    wool to remove mycelial fragments (Fig. 8). The filtrate

    was centrifuged for 10 min at 1800 x g in a Sorvall H1000B

    rotor. The supernatant was immediately poured off. After

    resuspending the pellet in the remaining drop (approximately

    0.5ml), lml of water was added, mixed well, and the

  • 47

    Figure 9. Apparatus used for preparation of spore

    suspensions and removal of mycelial fragments. Modified

    from the procedure of Hopwood et al. (1985) .

  • 48

    Crude spore

    suspension

    Fiberglass wool plug

    Vacuum^

    Filtered spore

    suspension

    125 ml Filter flask

  • 49

    mixture transferred to a screw-top vial containing 0.5ml

    80% (w/v) glycerol to achieve a final concentration of 20%

    glycerol. The suspension was mixed and stored at -20°C.

    Selection of 5-fluorouracil resistant strain.

    The 5-fluoropyrimidine analogues are toxic only after

    being converted by the corresponding pyrimidine salvage

    enzymes to the nucleotide level (0'Donovan & Neuhard, 1970).

    Positive selection of spontaneous mutants utilizing these

    analogues, in a manner similar to that of Martinussen &

    Hammer (1995) , was used to obtain strains lacking specific

    salvage enzymes. Selection of a 5FU resistant strain of S.

    griseus was done by spreading a spore suspension of the

    wild-type strain onto a minimal plate. To the center of the

    plate, a crystal of 5FU was placed. Growth of the plate was

    monitored. A zone of clearing indicated that the compound

    was toxic. Colonies which arose within the zone were

    allowed to sporulate and then collected using a sterile

    cotton swab. These spores were transferred to a fresh plate

    of minimal medium and streaked for isolation. Several

    isolated colonies were then selected and transferred to

    separate plates. Spores from colonies on these plates were

    collected for spore suspensions. The spores were again

    exposed to the analogue. Strains which showed no zone of

    clearing were resistant to 5FU.

  • 50

    Harvesting of bacterial cultures.

    Cells for assays were typically prepared by inoculating

    a flask containing 100 ml of Streptomyces minimal liquid

    medium with 50-100 jul of a dense spore suspension of the

    required Streptomyces strain. After 3 days incubation, 5 or

    10 ml of this starter culture were transferred to one or two

    liter volumes of the same medium. After 40 to 48 hours (h),

    the cells were harvested by centrifugation in a Sorvall RC5C

    centrifuge in a GS-3 rotor at 10,000 xg for 20 min at 4 °C

    with the centrifuge brake turned off. Most of the

    supernatant was carefully poured off to avoid disturbing the

    loose pellet. The pellet was resuspended in the remaining

    amount (10-15 ml) . The cells from 500 ml of the culture

    were transferred to 50ml, disposable conical tubes. The

    centrifuge tubes were washed with 10 ml of distilled water

    to remove residual cells and the wash added to the cells in

    the conical tubes. The conical tubes were then centrifuged

    in a Sorvall H1000B rotor for 15 min at 1800 x g at 4 °C. The

    supernatant was poured off. The cell pellet was either

    frozen at -20 °C for storage or used immediately.

    Cells for total DNA isolation were harvested by vacuum

    filtration using a modified version of the method of Hopwood

    et al. (1985). Cells were first grown in nutrient broth

    with 34% (w/v) sucrose. The mycelium was then collected by

    filtration in a Buchner funnel containing two sheets of

    Whatman No. 50 filter paper. The mycelium was washed with

  • 51

    100 ml of 10% (v/v) glycerol per 500 ml of culture, and the

    paste then transferred to a sterile conical tube using a

    sterile spatula. The cell paste was stored at -20 °C or

    used immediately.

    Preparation of cell extracts.

    Cell extract was prepared by breaking the cells using

    sonication. Cell pellets from freshly harvested cells or

    frozen samples were resuspended in the appropriate breaking

    buffer for the assays to be performed. Typically, 1ml of

    ATCase breaking buffer (2 mM /3-mercaptoethanol, 20 ptM ZnS04,

    and 50 mM Tris-HCl, pH8.0) was added per 1 g of wet weight

    of pellet. The cell suspension was sonicated using a

    Branson Cell Disruptor 200 for 5 min intervals while the

    tube was in an ethanol-ice water slurry to control heating

    of the sample. The sample was mixed by inversion, and the

    sonication repeated for a total of 15 min sonication per

    sample. The sonicated suspension was transferred to an

    SA600 centrifuge tube and centrifuged at 4 °C for 1 h at

    33,000 xg. The resulting supernatant was transferred to a

    sterile 15 ml, disposable conical tube and stored at room

    temperature until the cell extract was used in an assay.

    Assays on crude extract were performed within 48 h of cell

    extract preparation.

  • 52

    Aspartate transcarbamoylase assay.

    Cell extract was prepared as described earlier using

    ATCase breaking buffer.

    ATCase activity was assayed by measuring the amount of

    carbamoylaspartate (CAA) produced in 20 min at 30 °C

    according to the method of Gerhart & Pardee (1962), with

    modifications using the color development procedure of

    Prescott & Jones (1969). The optimal pH for ATCase in S.

    griseus was first determined by assaying ATCase activity

    with pH varied from 6.5 to 10.0. All the subsequent assays

    were done at the optimal pH.

    It was also necessary to determine the best storage

    conditions for the S. griseus extracts. During initial

    experiments, it was found that the extracts rapidly lost

    activity during storage at 4°C. This was similar to the

    cold-inactivation of the wheat-germ ATCase (Grayson, 1979).

    A cell extract was prepared and then divided into 0.5ml

    aliquots. An initial assay was done on the sample to

    establish a zero-point for comparison. An aliquot was

    stored under each of the following conditions: 1) room

    temperature; 2) filtered through 0.2 fjM membrane and stored

    at room temperature; 3) heated at 65°C for 5 min.,

    centrifuged 2 min. at 10, 000 x g, and stored at room

    temperature; 4) stored at 4 °C; 5) stored at 37 °C; 6) with 1

    mM UTP and stored at room temperature; 7) in 20% glycerol

    and stored at 4 °C; and 8) in 20% glycerol and stored at

  • 53

    -20 °C. ATCase assays were performed on these samples after

    storage for 1 day, 2 days, 7 days and 14 days.

    Assays were performed to determine the Km for

    aspartate

  • 54

    ATCase assay tubes were prepared in advance, without

    addition of carbamoylphosphate, and preincubated at 30 C

    for 4 min. The reaction was initiated by the addition of

    100 pi of carbamoylphosphate (Smgml"1) for the aspartate

    curve and, for the carbamoylphosphate curve, by 100 /il of

    carbamoylphosphate of varying concentrations (0.081 mM - 5.2

    mM) . At 20 min, 1 ml of the color mix was pipetted into the

    reaction tubes. The color mix contained 2 parts of 5 mg ml"1

    antipyrine in 50% (v/v) sulfuric acid and 1 part of 8 mg ml"1

    2,3-butanedione monoxime in 5% (v/v) acetic acid. Color was

    developed by incubating the tubes in a 65 °C water bath with

    the tubes exposed to the light of the room and capped by

    marbles to limit evaporation. After 2 h of incubation, the

    absorbance was read in a Beckman DU-40 spectrophotometer at

    a wavelength of 466 nm. Controls were utilized to determine

    background readings in tubes containing all reactants except

    cell extract and in tubes with cell extract but lacking the

    substrate carbamoylphosphate. All readings were blanked to

    these controls.

    The nmol CAA was determined using a CAA standard curve.

    The standard curve was prepared using known concentrations

    of CAA ranging from 50 to 500 nmol in the standard assay

    reaction mix and under the same color development

    conditions.

    The kinetic curves were generated by plotting specific

    activity of the enzyme (nmol CAA min"1 (mg protein)"1) versus

  • 55

    the concentration of aspartate or carbamoylphosphate. The

    Km, and Kmcp were determined from the intercepts of the

    x-axis on Lineweaver-Burk plots (l/velocity versus

    1/substrate).

    When checking fractionation samples for ATCase

    activity, the volume of sample used was typically 100 /xl,

    with the water adjusted accordingly to achieve a final

    reaction volume of 1ml. Aspartate concentration was 10 mM

    for all such samples and the reaction time was 20 min. Stop

    tubes were typically incubated for 2 h at 65 °C.

    Repression of ATCase in cells grown in the presence of

    exogenous pyrimidine compounds (100 pig ml"1) was also studied.

    ATCase activity was measured as previously described using

    varied aspartate and the specific activity of ATCase in

    cells grown with pyrimidine compounds was compared to the

    specific activity of cells grown in minimal medium without

    any additions.

    Dihydroorotase assay.

    With an equilibrium constant of 1.0, the DHOase

    reaction can be assayed in either direction. Here, DHOase

    was assayed in the degradative direction. The DHOase assay

    reaction tubes contained in a 1 ml total volume: 1 mM EDTA;

    20 jtil fraction sample; 100 mM Tris, pH8.6; and ddH20 to

    which 100 pi of 20 mM dihydroorotate (DHO) in 0.1 M phosphate

    buffer, pH 7.5, were added to start the reaction. A

  • 56

    background control was prepared for each cell extract used

    which received cell extract, but no DHO. A blank was

    prepared for each assay which contained all reactants, but

    no enzyme. All readings were zeroed using these controls

    before calculations were performed. After a 20 min

    incubation at 30 °C, 1 ml of color mix (same as used for

    ATCase assay) was added directly to the assay tubes. These

    tubes were incubated at 65 °C for 2 h and then the absorbance

    read at 466 nm. Conditions for storage of the enzyme were

    determined in the same manner as was done for Streptomyces

    ATCase. Kinetic studies were performed using varying

    amounts of the substrate. A blank was prepared for each

    concentration of DHO used, and all readings were zeroed to

    the blanks prior to calculations.

    Growth Curve.

    A growth curve was prepared for S. griseus 10137. A

    flask containing 2 liters of Streptomyces minimal liquid

    medium was inoculated with 20 ml of a starter culture. At

    the time of inoculation, a zero point sample was taken. Two

    50 ml aliguots were removed and placed in pre-weighed, 50 ml

    conical tubes. One was used for dry weight determination

    and the other for ATCase and DHOase assays. Samples were

    taken at 8 h intervals throughout the growth of the culture

    up to 72 h. All samples were stored at 4 °C until collected

    by centrifugation at 1800 x g. Pellets were stored at

  • 57

    -20 °C. Samples for dry weight determination were dried

    together in a vacuum desiccator for 120 h. ATCase and

    DHOase assays were performed on cell extracts from each time

    point.

    Protein concentration determination.

    Protein was quantified by the method of Lowry et al.

    (1951) . Bovine serum albumin (BSA) at 0-100 fig was used to

    produce a standard curve. The BSA was divided into 11 tubes

    at 10 fig increments as follows: the 0 fig tube contained

    only 200 fil ddH20, while the 10 fig tube contained 190 fil ddH20

    + 10 fil 0.1% BSA (100 fil of 1% desiccated BSA in water

    diluted in 900 fi 1 ddH20 yields a solution of BSA at 1 fig

    ixl'1) . The rest of the tubes used in the standard curve

    were set up accordingly. Three different volumes (5 fil, 10

    fil, and 20 fj.1) of the unknown sample cell extracts,

    typically 1:10 dilutions in ddH20 of each sample, were added

    and brought up to 200 fil total volume with ddH20. To the

    standard and unknown tubes, 800 fil alkaline copper reagent

    (0.5ml 2% sodium-potassium tartrate and 0.5 ml 1% copper

    sulfate mixed together before adding 49 ml of 2% sodium

    carbonate in 0.1 N sodium hydroxide) were added. The tubes

    were allowed to stand at room temperature for 10 min. Folin

    reagent (100 fil, 2 N commercial preparation diluted 1:1 with

    ddH20) was added to all tubes while mixing on a vortex

    mixer. After incubating at room temperature for 30 min, the

  • 58

    absorbance was read at 660 nm. The BSA concentration was

    plotted against the A660 to establish a standard curve.

    Purification strategies.

    Several purification strategies were utilized with

    Streptomyces aspartate transcarbamoylase (ATCase) and

    dihydroorotase (DHO). These strategies are outlined below,

    with the specific procedures discussed in following

    sections. All steps of each strategy were performed at room

    temperature.

    The first strategy was begun by streptomycin sulfate

    precipitation of cell extract. The supernatant was then

    fractionated using ammonium sulfate. The resuspended pellet

    of the fraction containing maximal ATCase activity was

    passed over a Sephacryl S-400 size-exclusion column. The

    column fractions which contained maximal ATCase activity

    were pooled and concentrated. The concentrate was then

    fractionated using high performance liquid chromatography

    (HPLC) ion-exchange.

    In an alternate strategy, the extract was first passed

    over a Sephadex G-200 size-exclusion column. The fractions

    containing peak activity were pooled and concentrated. The

    concentrate was separated through a Sephacryl S-400 size-

    exclusion column. The tubes containing maximal activity

    were pooled and then fractionated with ammonium sulfate.

  • 59

    Streptomycin sulfate precipitation.

    A streptomycin sulfate solution was made by mixing 10%

    (w/v) streptomycin sulfate in a 10 mM phosphate buffer, pH

    7.5 in the manner of Adair & Jones (1972). A one-half

    volume of this 10% streptomycin sulfate solution was slowly

    added to a known volume of cell extract (typically 30 ml)

    while stirring slowly. After stirring for 1 h, precipitates

    were removed by centrifugation for 30 min at 33,000 x g at

    4 °C. A 0.5 ml sample of the supernatant was removed for

    later assays and protein analysis. The remainder of the

    supernatant was transferred to sterile conical tubes where

    it was measured before proceeding to the next purification

    step. Although Adair & Jones (1972) dialyzed their samples

    before proceeding, excessive loss of activity was seen for

    dialyzed Streptomyces extracts. Upon noting this in pilot

    purifications of the samples, the dialysis step was omitted

    in all subsequent experiments.

    Ammonium sulfate fractionation.

    Ammonium sulfate fractionation first involved

    determination of the correct percent saturation for

    precipitation of the desired protein. Depending on the

    sample source, either streptomycin sulfate supernatant or

    column fractions, the saturation for precipitation varied

    due to differences in initial salt contents of these

    samples. To determine the correct percent saturation for

  • 60

    precipitation, a sample was fractionated in 5% (w/v)

    ammonium sulfate saturation increments beginning at 15%

    saturation at 25 °C. The pellet at each step was

    resuspended in ATCase buffer, typically 400 fil, while 500 (il

    of the supernatant was saved in a sterile 1.5 ml

    microcentrifuge tube. Activity for ATCase and DHOase was

    monitored in the pellet and supernatant. In this manner, it

    was determined that maximal ATCase activity for Streptomyces

    from streptomycin sulfate supernatants precipitated between

    40% and 55% ammonium sulfate saturation at 25 °C while

    ATCase activity from column fractions precipitated at 45%

    ammonium sulfate saturation at 25 °C.

    In the purification experiments using streptomycin

    sulfate precipitation, the initial ammonium sulfate cut was

    made by slowly adding ammonium sulfate to the streptomycin

    sulfate supernatant until reaching the desired saturation.

    The sample was stirred for 1 h. Precipitates were removed

    by centrifugation for 30 min at 33,000 xg at 4 °C. The

    supernatant was transferred to a sterile conical tube and

    the volume noted. A 0.5ml supernatant sample was removed

    and the pellet was resuspended in 0.5-1 ml ATCase buffer.

    Both were stored for ATCase and protein analysis.

    Additional ammonium sulfate was slowly added to the

    supernatant to bring to the second saturation percentage.

    The sample was stirred for 1 h before centrifugation for 30

    min at 33,000 xg to remove precipitates. A 0.5ml

  • 61

    supernatant sample was removed and stored for ATCase and

    protein analysis. The pellet was resuspended in 0.5-1 ml

    ATCase buffer and stored at room temperature.

    When using ammonium sulfate precipitation of S-400

    column fractions, only one cut was done. Ammonium sulfate

    was slowly added to the combined column fractions until

    reaching the desired saturation. The sample was stirred for

    1 h. Precipitates were removed by centrifugation for 30 min

    at 33,000 xg at 4 °C. The supernatant was transferred to a

    sterile conical tube. The pellet was resuspended in 0.25ml

    ATCase buffer and stored at room temperature.

    Sephadex 6-200 column chromatography.

    A 1 cm x 30 cm column was filled to a bed height of

    23.5 cm with Sephadex G-200 size-exclusion matrix. Size-

    exclusion buffer (1 mM EDTA, 20 jiM zinc acetate, 0.5 mM

    /3-mercaptoethanol, 150 mM potassium acetate and 10 mM Tris-

    HC1, pH 8.2, in ddH20) was degassed by vacuum and used as

    the elution buffer. The column was loaded with 1 ml of

    extract which was mixed with 50 fil of Blue Dextran (0.01 g

    ml"1) and 2 /xl of Phenol Red (0.01 g ml"1) . After the entire

    sample was in the gel matrix, the buffer reservoir was

    connected, and collection of 1 ml fractions began. For all

    determinations, the Blue Dextran peak tube was designated

    Tube #1 and all other tubes were labelled relative to this

    tube. Samples were stored at 25 °C.

  • 62

    Each fraction was assayed for both ATCase and DHOase

    activity. The assays were performed as described later in

    this paper.

    Sephacryl S-400 column chromatography.

    A 3 cm x 100 cm BioRad column was gravity-packed with a

    Sephacryl S-400 size-exclusion matrix. The same size-

    exclusion buffer was used as with the G-200 column. The

    column was first loaded with 0.5ml of blue dextran (0.01 g

    ml"1}. The loading apparatus was washed with 0.5ml of

    column buffer prior to loading 1-3 ml of sample. A gravity

    flow system was used, and flow rate was typically 0.8ml

    min"1. The void volume of typically 150 ml was removed and

    discarded, after which 80 tubes of 5 ml fractions were

    collected. In a few experiments, 2.5ml fractions were

    collected in the tubes where ATCase was expected to elute.

    Samples were stored at 25 °C.

    Each fraction was assayed for both ATCase and DHOase

    activity. These assays were performed as described later in

    this paper.

    Concentration of size-exclusion fractions.

    The fractions from the size-exclusion chromatography

    containing maximal ATCase activity were pooled and

    concentrated using a Centriprep-30 concentrator (Amicon,

    Inc., Beverly, MA). The concentrators were centrifuged at

    1500 x g for 15 min at 25°C. The filtrate was decanted and

  • 63

    centrifugation repeated according to the instructions until

    the retentate was concentrated to approximately 3 ml in

    preparation for use on an S-400 column and 0.7 ml in

    preparation for HPLC fractionation.

    High performance liquid chromatography ion-exchange.

    The concentrated sample was loaded on a Waters Protein-

    Pak Q 8HR AP Mini (5 mm x 50 mm) ion-exchange column. The

    initial buffer (Buffer A) was 20 mM Tris-HCl, pH8.2, which

    had been filtered and degassed. The flow rate was 1 ml

    min"1. A linear gradient was applied over 40 min to go from

    100% Buffer A to a 25%:75% mixture with Buffer B, which was

    20 mM Tris-HCl, pH8.2 and 1 M KC1. During the gradient

    change, 0.5ml fractions were collected. Samples were

    stored at 25 °C. Each fraction was assayed for both ATCase

    and DHOase activity. The ATCase and DHOase assays for

    fractionation samples were used as described later in this

    paper.

    Nondenaturing polyacrylamide activity gels.

    Cell extract was prepared as described previously for

    the ATCase activity assay. Samples of typically 24 [j.1 of

    cell extract were mixed with 6 jul of 5X loading buffer

    (312.5 mM Tris, pH 6.8, 50% v/v glycerol, and 0.05% w/v

    bromophenol blue in ddH20) . The total volume of 30 (xl was

    loaded onto a nondenaturing polyacrylamide gel with a 5%

    stacking gel and an 8% separating gel. The gel was

  • 64

    typically run in a Bio-Rad Mini-Protean II cell at 100V for

    2.5 h at room temperature. The gel was prepared by first

    pouring the separating gel, which contained 2.67ml of the

    stock solution of acrylamide (29% w/v acrylamide and 1% w/v

    bis-acrylamide in ddH20) , 2.5 ml of Buffer B (1.5 M

    Tris-HCl, pH 8.8), and 4.83 ml of ddH20. Ammonium persulfate

    (0.02 g) was added to the mixture to remove dissolved

    oxygen. Just prior to pouring the gel, 5 /z 1 of N,N,N',N'-

    tetramethylethylenediamine (TEMED) was added and mixed in by

    gentle inversion. The solution was poured into the gel

    plates leaving a 2 cm gap at the top. The separating gel

    was covered with a layer of N-butanol to prevent drying and

    allowed to stand at room temperature for 1 h to polymerize.

    The butanol layer was poured off, and the stacking gel

    poured. The stacking gel contained 0.67ml of the stock

    solution of 30% acrylamide, 1ml of Buffer C (0.5 M Tris, pH

    6.8), and 2.3 ml of ddH20. Ammonium persulfate (0.01 g) and

    TEMED (5 /xl) were added. Once the stacking gel was poured,

    a comb was inserted between the plates to form the wells.

    The stacking gel was allowed to polymerize for 30 min. The

    running buffer contained 25 mM Tris and 192 mM glycine in

    ddH20 at a pH of 8.8.

    For molecular mass determinations, samples were loaded

    on a 4-20% Bio-Rad polyacrylamide ready-gel. These gels

    were run at 150V for 3 h at room temperature using the same

    non-denaturing running buffer described above. Partially-

  • 65

    purified ATCase from Escherichia coli K12 (holoenzyme and

    catalytic trimer) and Pseudomonas aeruginosa PA01 with known

    molecular mass were used as standards, as were Sigma protein

    molecular weight markers urease and BSA.

    The gels were stained specifically for ATCase activity

    by a procedure developed by Bothwell (Bothwell, 1975) and

    further modified by K. Kedzie (1987), who used histidine

    instead of imidazole. In this study, two slight

    modifications to the procedure were made as indicated. When

    the electrophoresis was complete, the plates were separated,

    and the gel was placed in 250 ml 50 mM histidine, pH7.0, at

    room temperature (modified from ice-cold at this st