...:3...: 2. . ... . Ix... .5 .2.
_.....i.“. . .35...
“Emu.“ £31123. :3. z: s»
21...}... ; I ..
mm... M.
was“?
El? b Ii.
THESIS
Z
2600
Willi”!!!"lull!!!ll!MlUllllllllllllllll
3925
This is to certify that the
dissertation entitled
DEVELOPMENTAL REGULATION OF STREPTOMYCES COELICOLOR
ANTIBIOTIC SYNTHESIS BY THE AbsA TWO—COMPONENT SYSTEM
presented by
Todd B. Anderson
has been accepted towards fulfillment
of the requirements for
mm ; degree in 11121.01:inng
Mééwm
ajor professor /
Date244?L 4.}an 0
MS U is an Affirmative Action/Equal Opportunity Institution 0-1277 1
LIBRARY
Michigan State
University
PLACE IN RETURN BOX to remove this checkout from your record.
TO AVOID FINES return on or before date due.
MAY BE RECALLED with earlier due date if requested.
DATE DUE
DATE DUE ' '
DATE DUE
11/00 c/CIRCJDatoDuepfis-plu
DEVELOP)
SY}
DEVELOPNIENTAL REGULATON OF Streptomyces coelicolor ANTIBIOTIC
SYNTHESIS BY THE AbsA TWO-COMPONENT SYSTEM
By
Todd B. Anderson
A DISSERTATION
Submitted to
Michigan State University
In partial fulfillment of the requirements
For the degree of
DOCTOR OF PHILOSOPHY
Department of Microbiology
2000
DEVELOP.
Si
Strepton
temporally com
coelz'color, the c
globally regular
absA produce d
igmlhiotic gm!
kimse gene. Ir
mutations in [h
to lock AbsAl
Site suppressor
Abs' mUlants v
ABSTRACT
DEVELOPMENTAL REGULATON OF Streptomyces coelicolor ANT IBIOTIC
SYNTHESIS BY THE AbsA TWO-COMPONENT SYSTEM
By
Todd B. Anderson
Streptomycetes synthesize antibiotics in a growth-phase dependent manner
temporally commensurate with, but spatially independent of sporulation. In Streptomyces
coelicolor, the absA locus encodes a two-component signal transduction system which
globally regulates antibiotic production independent of morphogenesis. Mutations in
absA produce drastically opposing phenotypes. Mutations responsible for an Abs'
(antibiotic synthesis deficient) phenotype were previously localized to the absAI histidine
kinase gene. In this study Abs' mutants C542 and C577 were shown to contain point
mutations in the region of absAI encoding the transmitter domain; these were proposed
to lock AbsAl into a kinase dominant, phosphatase deficient enzymatic state. Second-
site suppressor mutations, sab (suppressor of ng), that restored antibiotic synthesis to the
Abs‘ mutants were previously mapped very near to the absA locus. Sequence analysis
identified several sab mutations in the absAI or absAZ genes; the latter encodes a two-
component response regulator. A genetic analysis of the absA locus was performed to
examine the mechanism of AbsA regulation. A disruption in absAZ caused the
precocious hyperproduction of pigmented antibiotics (Pha phenotype), demonstrating that
AbsA2 negatively regulates antibiotic synthesis. Gene replacements in absA, that
disrupted phosphoryl- group transfer, also caused the Pha phenotype, demonstrating that
the phosphorylated form of AbsA2 is the active negative regulator of antibiotic synthesis.
These Pha strai:
sinthesis, in ad
ln-ritro phospi‘.
meated AbsA
biochemical ex"
phosphoryl-gro
ope back groun
also possesses .
demonstrated 1:
With-phase r
Pigmented anti
contrast to its a
activate transcr
F652, and Cda}
Emporally I61;
phosphorl'latic
high levels COi
Tmscfipllm] i
Pa‘allel with ll
.ibsA_ The p0
SWis in th:
9 s . .
OILEnaUOnO
These Pha strains demonstrated temporal acceleration of calcium-dependent antibiotic
synthesis, in addition to previously characterized acceleration of pigmented antibiotics.
In-vitro phosphorylation experiments utilizing a maltose-binding protein fusion to
truncated AbsAl ('AbsAl) and a His-tag fusion to AbsA2 provided preliminary
biochemical evidence supporting AbsAl autokinase activity and AbsAl-AbsAZ
phosphoryl-group signal transduction. High-copy expression of absAI alleles in a wild
type background resulted in a Pha phenotype, providing genetic evidence that AbsAl
also possesses AbsA2-P phosphatase activity. 81 nuclease mapping of the absA locus
demonstrated leaderless cotranscription of absAI and absAZ. Transcription of absA was
growth-phase regulated, experiencing a dramatic increase prior to the appearance of
pigmented antibiotics. Transcription of the absA locus was also autoregulated. In
contrast to its activity as a negative regulator of antibiotics, phospho-AbsA2 appeared to
activate transcription of abSA. AbsA regulation of pathway-specific regulators redD,
redZ, and cdaR was examined by $1 nuclease protection assays. Expression of redD was
temporally retarded with respect to absA and was clearly regulated by AbsA2 in a
phosphorylation-dependent manner. Conversely, redZ transcription was expressed at
high levels coincident with absA expression and showed no dependence on AbsA.
Transcription of the cdaR pathway-specific regulator was also temporally regulated in
parallel with the absA time course. Transcription of cdaR does not appear to depend on
AbsA. The possible role and mechanism of AbsA-mediated regulation of antibiotic
synthesis in the S. coelicolor life cycle is discussed. Preliminary evidence supporting the
conservation of absA in other species of Streptomyces is presented.
ldedicate this
healthy [3813]}:
also dedicatec
achieve this g
Anderson, in
ldedieate this
motivated me
I dedicate this work to my beautiful boys, Todd and David, who helped me keep a
healthy perspective and balance in my life over the course of this project. This work is
also dedicated to my wife, Marywbska, whose support and sacrifice allowed me to
achieve this goal; and to my mother, Alice Anderson, and the memory of my father, Ellef
Anderson, in appreciation of the values and work ethic they helped instill in me. Finally,
I dedicate this work to my sister, Pamela, whose commitment to good science has always
motivated me.
iv
lwm
patience and
Renqiu Kori;
support. I ti
also grateful
knowledge :
Dr. David .1
lhiern for i
Plfimids U:
ACKNOWLEDGEMENTS
I would like to thank my mentor, Dr. Wendy Champness, for her guidance,
patience and support during this period of my professional development. Thanks to '
Renqiu Kong, Marywbska Calderon, and Mark Kazmierczak for valuable technical
support. I thank Paul Brian for the lab training that helped me get up and nmning. I am
also grateful to N. Jamie Ryding, Gary Brown, and David Aceti for their sharing of
knowledge and ideas. My appreciation goes out to the members of my thesis committee,
Dr. David Amosti, Dr. Michael Bagdasarian, Dr. Michael Thomashow, and Dr. Suzanne
'I‘hiem for helpful guidance and discussion. Thank you to Mervyn Bibb for providing
plasmids used in this study.
LIST OF TABLE
llST 0F F lGUK
KEY TO AB BRI
NlRODL'CTlO
Regulatio
CHWIER 2
SEQUENCE M
MLTANT S ......
CHER 3
A GENETIC A}
OF AbsA-MED]
lNTROD
MATERJ
G
A
P.
D
C
S
R
S
RESLU
N
a
G
ll
P1
1
Pi
E
H.
S
G
DISCL'Sc1
TABLE OF CONTENTS
LIST OF TABLES ..................................................................... ix
LIST OF FIGURES .................................................................... x
KEY TO ABBREVIATIONS ........................................................ xii
lNTRODUCTION ..................................................................... 1
Regulation of Cellular Difi'erentiation and Antibiotic Production ...... 4
Factors triggering difl‘erentiation ................................... 5
Uncoupling of differentiation: regulation of antibiotics. . . .. 16
Two-Component Signal Transduction Systems ........................... 24
The AbsA two-component system of S. coelicolor .............. 32
CHAPTER 2
SEQUENCE ANALYSIS OF absA ALLELES OF Abs' AND sab
MUTANT S .............................................................................. 36
CHAPTER 3
A GENETIC AND TRANSCRIPT ANALYSIS INTO THE MECHANISM
OF AbsA-MEDIATED ANT IBIOTIC REGULATION ........................... 40
INTRODUCTION ............................................................. 41
MATERIALS AND METHODS ............................................. 45
Growth Conditions .................................................... 45
Antibiotic Assays ...................................................... 46
Plasmid and DNA Manipulations .................................... 46
Disruption of 0193/12 in C500 ......................................... 50
Construction of an In-Frame Deletion in absAIin C530 .......... 51
Site-Directed Mutagenesis ............................................ 54
RNA Isolation ........................................................... 58
$1 Nuclease Protection Assays ....................................... 59
RESULTS ........................................................................ 61
Negative Regulation of Antibiotics by the AbsA2 Regulator
and AbsAl Histidine Kinase ......................................... 61
Genetic Evaluation of the Role of Phosphorylation in AbsA2-
Mediated Regulation ................................................... 63
Precocious Hyperproduction of Calcium-Dependent Antibiotic,
Undecylprodigiosin, and Actinorhodin in absA Mutants. . . .. 64
Precocious Hyperproduction of Antibiotics Resulting from AbsA
Domain Overexpression ............................................... 66
High Resolution 81 Nuclease Mapping of the absA Transcription
Start Site ................................................................. 69
Growth-Phase Dependent Expression and Autoregulation of absA. 71
DISCUSSION ..................................................................... 73
vi
ClLlPIER 4
TEMPORAL EXP
REGULATORS A
LN'TRODL'
ll-L-iTERlA
RESULTS
De ,
on
De
DISC L'SSl
CHIPIER 5
0\EREXPRESSI
OEAbsAl AND .L‘
INTRODL
MATERLA
Bat
CHAPTER 4 -
TENTPORAL EXPRESSION OF red AND cda PATHWAY-SPECIFIC
REGULATORS AND THEIR DEPENDENCE ON AbsA ......................... 80
INTRODUCTION ............................................................... 81
MATERIALS AND METHODS ............................................... 83
RESULTS ......................................................................... 84
Dependence of Pathway-Specific Regulators redZ and redD
on AbsA ................................................................. 84
Dependence of cdaR Expression on AbsA .......................... 87
DISCUSSION ................................................................... 90
CHAPTER 5
OVEREXPRESSION, PURIFICATION, AND PHOSPHORYLATION
OF AbsAl AND AbsA2 PROTEINS .................................................. 95
INTRODUCTION ............................................................... 96
MATERIALS AND METHODS .............................................. 98
Bacterial Strains ......................................................... 98
Construction of AbsA2-His”) and AbsA2-Hi35 Expression
Plasmids ................................................................. 99
Construction of MBP-‘AbsAl Expression Plasmids ............... 101
Culture Media and Growth Conditions .............................. 102
Purification of MBP-‘AbsAl Proteins ................................ 103
Purification of AbsA2-His Proteins ................................... 104
'AbsAl and AbsA2 Phosphorylation Assays ........................ 105
RESULTS ......................................................................... 106
Overexpression and Purification of AbsA2-His“) from E. coli. .. 106
In- Vitro Phosphorylation of AbsA2-His“) with Acetyl Phosphate. 108
Overexpression and Purification of AbsA2-His; from S. lividans.. 109
In- Vitro and In- Vivo Analysis of AbsA2—His6 ........................ 111
Overexpression and Purification of 'AbsAl from E. coli ............ l 13
Autophosphorylation of 'AbsAl and Phosphorylation of
AbsA2-His6 ................................................................ 115
DISCUSSION ....................................................................... 1 17
CHAPTER 6
AbsA2 HOMOLOGUES IN OTHER STRAINS OF STREPTOMYCES ........... 120
INTRODUCTION .................................................................. 1 2 1
MATERIALS AND METHODS ................................................. 123
Bacterial Strains and Growth Conditions ............................... 123
PCR Amplification of Putative absA2 Homologs ..................... 124
absAZ Homolog Identification and Sequencing ........................ 125
vii
CHAPTER 7
CONCLUSION 1
MOlCCUlIil
for the At
The Role .
Model Su:
AbsA2 Ta
The AbsA
REFERENCES. ..
RESULTS ............................................................................ 127
PCR Amplification of Putative absAZ Homologs from
Heterologous DNA ........................................................ 127
Identification of Putative absAZ Homologs ............................. 128
Sequence Analysis of Putative Homologs .............................. 130
DISCUSSION ....................................................................... 1 34
CHAPTER 7
CONCLUSION AND FUTURE RESEARCH ........................................... «138
Molecular Genetic Characterization of absAI Mutations Responsible
for the Abs' Phenotype and Certain sab Suppressors of Abs' .................. 139
The Role of Phosphorylation in the AbsA Regulatory Mechanism ........... 141
Model Summary ..................................................................... 143
AbsA2 Targets ....................................................................... 146
The AbsA Signal and Signal-Sensing Mechanism .............................. 150
REFERENCES ................................................................................ 153
viii
TABLE 1. Oligo
LIST OF TABLES
TABLE 1. Oligonucleotide primers used in this study ................................. 47
ix
Figure I. ll
Figure 2. T
Figure3. P
figure 4. T
Figure 5.Pt
Figure 6. I
Figure 7. C
Figure 8. I
Figure 9. "i
Fig; 10.
Figure 11.
Figure 12.
Elm: 13,
Figure 14,
Flgurels
Figure 16
Fl
’1'?
re 17
FlSure 1 8
Flg‘dre 19
Figure 20
LIST OF FIGURES
Figure l. Hierarchy of antibiotic regulatory loci ...................................... 15
Figure 2. Two-component signal-transduction proteins .............................. 26
Figure 3. Paradigm two-component signal-transduction system activity ........... 30
Figure 4. The AbsA two-component system ....... ................................... 33
Figure 5. Position of absA within the cda gene cluster ................................. 44
Figure 6. The absA locus and plasmid inserts based on absA ........................ 49
Figure 7. Creation of the absAZ disruption in strain C500 ............................ 51
Figure 8. The absA IA530 in-frame deletion ............................................. 52
Figure 9. The effect of an absA gene disruption and gene replacements on
antibiotic production ............................................................ 62
Figure 10. Calcium-dependent antibiotic assays in Pha mutants ....................... 65
Figure 11. The effects of high-copy expression of absA alleles on antibiotic
production ...................................................................... 68
Figure 12. SI nuclease protection mapping of the (16571 locus ........................ 70
Figure 13. High-resolution Sl nuclease protection analysis of the absA transcript. 72
Figure 14. Position of absA within the cda gene cluster ................................ 82
Figure 15. High-resolution S1 nuclease protection analysis of the redD
and redZ transcripts ............................................................. 86
Figure 16. High resolution 81 nuclease protection analysis of the cdaR transcript. 89
Figure 17. Plasmids for 'absAI and absAZ overexpression ............................ 100
Figure 18. Purification of AbsA2 proteins overexpressed in E. coli ................. 108
Figure 19. Purification of AbsA2-Hi36 overexpressed in S. lividans ................. 110
Figure 20. In-vivo analysis of pAB8270 .................................................. 112
Figure Zl.
Figure 33.
Figure 33.
Figure 24.
Figure 25.
Figure 26. ‘
Figure 21.
Figure 22.
Figure 23.
Figure 24.
Figure 25.
Figure 26.
Purification of 'AbsAl expressed in E. coli ................................. l 14
In-vitro phosphorylation of AbsA2-Hi56 by 'AbsAl-phosphate .......... 1 15
PCR amplification of putative absAZ homologs ..................... . ...... 128
Southern blots of putative abSAZ homologs ................................. 130
Amino acid sequence alignment of S. coelicolor AbsA2 with putative
homologs from S. amobofaciens, S. griseus, S. peucetius, and
S. clavuligerus .................................................................. 132
Model of AbsA-mediated antibiotic regulation in S. coelicolor .......... 144
xi
Abs‘
Act
CDA
HI H
Mary
0er
Pha
PSR
Red
sub
SARP
TM
LIST OF ABREVIATIONS
antibiotic synthesis deficient
actinorhodin
calcium-dependent antibiotic
histidine kinase
helix-turn-helix
methylenomycin
Oxoid nutrient agar
precocious hyperproduction of antibiotics
pathway-specific regulator
undecylprodigiosin
response regulator
suppressor of abs
Streptomyces antibiotic regulatory proteins
transmembrane
xii
CHAPTER 1
INTRODUCTION
The ‘
order Act/no
may ofbiol
oi the thous.
half of the b
metabolites
antiparasitic
herbicides,
I"SEIher wit
other actino
if e also Und
characterize
the lncidem
“31111111131 81
Sire’pmm'ycl
DIESem We
the fabUlOu
lino“ acti
(filling the}
Sire
The genus Streptomyces is composed of high-G+C, Gram-positive bacteria of the
order Actinonrycetales. Streptomyces are most noted for their ability to produce a diverse
array of biologically active secondary metabolites. This genus accounts for over one-half
of the thousands of naturally produced antibiotics discovered and is the source of over
half of the bio-active compounds currently in clinical use (165). Streptomyces
metabolites important to health and agriculture include antibacterial, antifungal,
antiparasitic, and antitumor drugs, immunosuppressive agents, insecticides, and even
herbicides. The abundance of naturally occurring compounds produced by this genera,
together with the relative ease of culturing streptomycetes (in comparison to fungi or
other actinomycetes), makes them a prime target for bioexploration. Numerous efforts
are also underway to produce novel metabolites through genetic engineering of well
characterized Streptomyces biosynthetic pathways (82). Given the alarming increase in
the incidence of antibiotic resistance in common human pathogens, coupled with the
continual emergence of new human, animal and plant pathogens, it is critical that
Streptomyces continue to be exploited as a source of new drugs. Unfortunately, at
present we are able to culture only a handful of Streptomyces. Therefore, to fully harness
the fabulous diversity of chemical compounds synthesized by this genus, and other little
known actinomycetes, we must gain a better understanding of the biological processes
driving their growth and development.
Streptomyces are notable for their morphological similarity to filamentous fungi.
Spore germination and germ tube extension is followed by vegetative growth through
hyphal extension and branching to form a dense weave of substrate mycelium.
Vegetative growth frequently ends in response to some nutrient limitation (90a) at which
time the cell enle"
the Production of?
biomass EICCUmUla
mycelium. the CO“
differentiation Whl
physiological ditIC
emerge from the Si
34; 37). Meanvvhil
metabolically distii
secondary metabol
Streptonr_t‘c
(reviewed by 150;
consistencies in the
evolutionary con se
0rgatiizationally, l:
toEither on the chr
actinOfliodin genes
identified (cataIOgr
meme blOSYntheri
slIlllIESlS is grOBTl‘;
expreSSlOn of m0
5
L’e-nsi
I
I.
genus
(30)- Likeu
time the cell enters the second phase of biphasic biomass accumulation coincident with
the production of aerial mycelium. DNA replication continues throughout the period of
biomass accumulation (90a; 112). Prior to the appearance of antibiotics and aerial
mycelium, the colony undergoes the initial stages of a complex program of cellular
differentiation which has been divided into two distinct processes - morphological and
physiological differentiation (31). During morphological differentiation, aerial hyphae
emerge from the substrate mycelium, terminally septate, and develop spores (reviewed by
34; 37). Meanwhile, in the substrate mycelium, the temporally parallel but spatially and
metabolically distinct process of physiological differentiation results in the synthesis of
secondary metabolites.
Streptomyces produce several chemically distinct categories of antibiotics
(reviewed by 150; 173; 163). In spite of this tremendous diversity, there are certain
consistencies in their genetic organization and regulation that suggest substantial
evolutionary conservation in the mechanism controlling their synthesis.
Organizationally, biosynthetic genes encoding a particular antibiotic are clustered
together on the chromosome. Since first demonstrating the clustered organization of
actinorhodin genes (109), a great many Streptomyces antibiotic gene clusters have been
identified (cataloged by 119; 72). Clusters occupy approximately 15 to 100 kb and
include biosynthetic genes, resistance determinants, and regulatory genes. Antibiotic
synthesis is growth-phase dependent, occun'ing during the stationary phase. Temporal
expression of most regulatory genes for antibiotic production are upregulated during
transition phase growth. Cluster-specific regulatory features are common throughout the
genus (30). Likewise, preliminary evidence using PCR and Southern hybridization
techniques suggest
present in numeror
bidet, (99); absA 3.
introduced into het
suggesting some le
together, these I’CSL
may be conserved a
review advances i
0rf-fartism S. coy/rev
AbSA m'O‘COmpon
antibiotic productit‘
Regulation of Cell
Regulation
tiers. At the earlies
jointly regulated I
V ‘ .
egetative and/or tr
unthesis from me
r
Sps‘tlli
C r eSlllators
r1 l .
C.
i.
rereases greatly d
Show -
llll .
(A
techniques suggests that homologs of many antibiotic pleiotropic regulatory genes are
present in numerous streptomycetes (absB, (136); cfs‘Q, (84); afiR, (115) ; cutRS, (3 2);
bIdA, (99); absA2, see Chapter 6). Moreover, cloned antibiotic gene clusters that were
introduced into heterologous host streptomycetes showed proper temporal expression,
suggesting some level of regulatory conservation (109; reviewed in 119). Taken
together, these results imply that regulatory mechanisms governing antibiotic production
may be conserved at various stages of development. The remainder of this chapter will
review advances in the field of Streptomyces antibiotic regulation with a focus on model
organism S. coelicolor. Special attention will also be given to the topic of this study, the
AbsA two-component signal transduction system, and its contribution to regulation of
antibiotic production in S. coelicolor.
Regulation of Cellular Differentiation and Antibiotic Production
Regulation of antibiotic production in streptomycetes can be divided into three
tiers. At the earliest stages of development, antibiotic synthesis and sporulation are
jointly regulated. The precise timing of the uncoupling of joint regulation during the
vegetative and/or transition phases has not been widely studied. Uncoupling of antibiotic
synthesis from morphogenesis is characterized by two levels of regulation; pathway-
specific regulators control the synthesis of a single antibiotic, where as pleiotropic
regulators influence the production of two or more. Expression of these regulators
increases greatly during the transition phase and continues well into stationary phase
growth, which is characterized by the onset of antibiotic accumulation. In order to
understand the role of any individual regulator on differentiation, it is valuable to have an
appreciation of its
Therefore, Ivvill e»
the onset of dift ere
only in antibiotic p
Factors triggering
It has been
carbon, nitrogen, p
metabolism (revie\
mechanism ofiniti
required in second
repressed by grow
free assays with at
presence ofglucog
did not appear to 1
and references the
antibiotic synthes:
fiddie, ammOHIUn
used to generate F
Conversely, addit
min.
appreciation of its temporal and physiological significance in the global process.
Therefore, Iwill examine in greater detail, various physiological processes involved in
the onset of differentiation prior to focusing attention on specific regulators implicated
only in antibiotic production.
Factors triggering differentiation
It has been well established that exhaustion of one or more nutrients such as
carbon, nitrogen, phosphorous, or trace elements leads to the onset of secondary
metabolism (reviewed in 152). Easing of catabolite repression was demonstrated as one
mechanism of initiating antibiotic synthesis through derepression of specific enzymes
required in secondary metabolism. Production of numerous Streptomyces antibiotics was
repressed by growth on glucose, and certain other carbon sources (reviewed by 46). Cell-
free assays with antibiotic synthases showed that a number of these were repressed in the
presence of glucose. In contrast to the inducible enzymes of enteric bacteria, cyclic-AMP
did not appear to play a role in relieving carbon source repression in Streptomyces (45
and references therein). Readily assimilable ammonium salts also negatively affected
antibiotic synthesis in numerous strains (reviewed by 151; 46). In tylosin producing S.
fradie, ammonium repressed several enzymes associated with catabolism of amino acids
used to generate propionate and butyrate precursors of the macrolide ring (103).
Conversely, addition of branched chain amino acids promoted tylosin production in this
strain. In general, greater production of antibiotics was favored by growth on complex,
slowly assimilable nitrogen sources. The biosynthesis of a large number of antibiotics
was subject to regulation by phosphate (reviewed by 111). Those chemical classes of
antibiotics Cf
macrolideS a
much l€5S Set
biosynthetic 5
classes of anti
function at the
early in Colon}.
may be due In l
genes. or repres
Given ti
stationary phase
secondary metal:
examined whetht
of nutrient depm
venezue/ae ( l 72).
both induced at lo
granaticin was prc
absence of nutrien
Cases where some
mduce antibiotic s
Stn'ngent r
response 10 low 2'
IFSPOHSC IS 3550C?
antibiotics especially sensitive to phosphate included aminoglycosides, tetracyclines,
macrolides and polyenes, while antibiotics directly assembled from amino acids were
much less sensitive. Phosphate was shown to repress phosphatases required in
biosynthetic steps of aminoglycosides, and several other biosynthetic enzymes from other
classes of antibiotics. However, in the synthesis of macrolides, phosphate seemed to
function at the level of primary metabolism by inhibiting precursor formation. Thus,
early in colony development, growth-phase dependent regulation of antibiotic synthesis
may be due in part to catabolite repression of antibiotic biosynthetic and/or regulatory
genes, or repression of pathways involved in precursor formation.
Given that antibiotic production is associated with the low growth rates of
stationary phase metabolism, it was thought that nutrient deprivation may also induce
secondary metabolism through a growth-rate sensitive response. Several studies
examined whether antibiotics could be induced by low growth rate alone in the absence
of nutrient deprivation. In continuous culture experiments with S. cattleya (106) and S.
venezuelae (172), synthesis of cephamycin C and cloramphenicol, respectively, were
both induced at low grth rates independent of nutrient limitations. Likewise,
granaticin was produced by S. thermocyolaceus in response to low growth rate in the
absence of nutrient deficiencies (85). In contrast, Demain and Fang (1995) cite various
cases where some form of nutritional stress was required along with low growth rate to
induce antibiotic synthesis.
Stringent response is a common mechanism involved in bacterial adaptation in
response to low growth-rate or nutrient deprivation. In Enterobacteriaceae, the stringent
response is associated with the accumulation of the intracellular effector molecules
(:3)me in
regulation 01
some cases r
griserrs ( l 29,
subunit that 1
synthesis of l
coelicolor rel
{Red} synthe:
was unable tc
were also imp
to reduced lev
Were made in
PPGPP Synthe
Simulation ar
CDA Synthesi
to make Act 0
CODdIIIOna] d6
87 156715, Si n Ce ‘
(1996) witnegS
acturnulatiOn C
was no obligau
SIC/mm? ’rus
shift—down (1‘?)
(p)ppGpp (reviewed by 24). In Streptomyces, the involvement of ppGpp-dependent
regulation of antibiotic synthesis and sporulation appeared to be species specific and in
some cases metabolite specific. Mutations in the relC gene of S. coelicolor (128) and S.
griseus (129) produced "relaxed" mutants with altered L11 protein in the SOS ribosomal
subunit that were impaired in their ability to bind RelA - an enzyme that catalyzes the
synthesis of pppGpp from GTP (or GDP) and ATP during the stringent response . The S.
coelicolor relaxed mutants were deficient in actinorhodin (Act) and undecylprodigiosin
(Red) synthesis and unable to form aerial hyphae. Similarly, the S. griseus reIC mutant
was unable to produce streptomycin or form aerial hyphae. Unfortunately, these mutants
were also impaired in growth so that it was uncertain whether their phenotypes were due
to reduced levels of (p)ppGpp or general growth deficiencies. More recently, deletions
were made in the reIA gene of S. coelicolor (113; 114; 27; 26) that completely eliminated
ppGpp synthesis. Martinez-Costa, et al. (1996) found that the relA deletion reduced
sporulation and severely affected Act production, but had little or no effect on Red and
CDA synthesis. Chakraburtty and Bibb (1997) reported that the reIA null mutant failed
to make Act or Red; however, only under nitrogen limiting conditions. A similar
conditional dependence on ppGpp may also extend to streptomycin production in S.
griseus, since contrary to the finding of Ochi (1990b) mentioned above, Neumann et al.
(1996) witnessed substantial streptomycin production without any significant
accumulation of ppGpp in S. grisezrs when grown on minimal medium. Likewise, there
was no obligatory relationship between antibiotic production and ppGpp accumulation in
S. clavuligerus when grown in defined or complex medium and subjected to nutritional
shift-down (12). Therefore, (p)ppGpp may be required to trigger sporulation and/or
synthesis Of‘
role appears 1
products of 5:
Own
mechanism fc
effector mole
levels and elic
normally acyl
cytoplasmic n
sensing is four
hanerr‘. This
accumulate in
the phosphOr}.
reaching thresl
their indepmdl
5611501' kinasesg
ei‘PFeSSlon 0ft
QUOI‘un
reSlllallng d€\'e
are rel’le‘aed bv
synthesis of certain antibiotics under a particular set of culture conditions; however, its
role appears to be species dependent, nutritionally conditional, and limited to certain
products of secondary metabolism.
Quorum sensing, cell-density-dependent gene expression, is another common
mechanism for the initiation of developmental responses. In quorum-sensing systems,
effector molecules accumulate in the medium until such time that they reach threshold
levels and elicit a response. In Gram-negative bacteria the effector molecules are
normally acyl-homoserine lactone autoinducers that are able to diffuse across the
cytoplasmic membrane (58). A well characterized example of Gram-negative quorum
sensing is found in autoinducer regulated gene expression of bioluminescence in Vibrio
harveyi. This organism produces two signaling molecules, AL] and AI-2, that
accumulate in the medium during vegetative growth (13). In the absence of autoinducer,
the phosphorylated response regulator protein, LuxO, represses bioilluminescence. Upon
reaching threshold concentrations, AL] and AI-2 activate the phosphatase activity of
their independent cognate sensor molecules LuxN and LuxQ (hybrid two-component
sensor kinases) that are able to integrate their signal through LuxO to derepress
expression of bioilluminescence genes(56).
Quorum sensing is also a common mechanism among Gram-positive bacteria for
regulating developmental gene expression. Examples of the best characterized processes
are reviewed by Kleerebezem, et al. (1997) and include genetic competence in Bacillus
subtilis and Staphylococcus pneumoniae, virulence response in Staphylococcus aureus,
and the production of peptide antibiotics by various species of Gram-positive bacteria. In
contrast to their Gram-negative counterparts, these cell-density dependent systems
incorporale a p‘
atrrusible N'aci
dedicated ATP-
s}.'stem5 Which i
mommponenl
regulélte gene ex
the processed PC
the peptide pher'
cell-density deli“
mechanism has I
hierarchical regu
does appear to in
below). Likewis
(http://vmw. sang
ABC transporters
Interesting
involving peptide
lactone analogs, y
signaling molecul
species (eg, S. W
Q’aneofuscams (6
differentiation is f
Induction of aeria'
incorporate a post-translationally processed peptide as a signaling molecule. Unlike the
diffusible N-acyl homoserine lactones, these peptide pheromones are secreted by
dedicated ATP-binding-cassette exporters. Another common characteristic of these
systems which is not as ubiquitous in the Gram-negative mechanisms is the integration of
two-component signal transduction systems as sensor and response mechanisms to
regulate gene expression. The histidine kinase (HK) of these systems is thought to sense
the processed peptide pheromone and activate response regulator-mediated expression of
the peptide pheromone, the two-component system, and the genes responsible for the
cell-density dependent phenotype (96). A peptide pheromone quorum-sensing
mechanism has yet to be conclusively demonstrated in Streptomyces; however, a
hierarchical regulatory cascade controlling aerial mycelium formation in S. coelicolor
does appear to incorporate small peptide signaling molecules (185; 126; discussed
below). Likewise, the S. coelicolor genome sequencing project
(http://www.sanger.ac.uk/Projects/S_coelicolorl) has uncovered numerous homologs to
ABC transporters with as yet unassigned functions.
Interestingly, a variant of the Gram-positive quorum sensing systems not
involving peptide pheromones is found in the genus Streptomyces. N-acyl homoserine
lactone analogs, y-butyrolactones, have been implicated as cell-density-dependent
Signaling molecules in morphological and/or physiological differentiation in several
species (e. g., S. virginiae (189; 188); S. viridochromo (61); S. bikiniensis and S.
cyaneofuscatus (60)). The best known example of butyrolactone regulation of
differentiation is found in S. griseus where A-factor autoregulator is essential for the
induction of aerial mycelium formation and streptomycin production (reviewed by 75;
I30). A-factOI
the Ciiol’lasmi
spurulation in z
grisetrs (92) “in
regulatory funC
in the cyIOPlaS’l
independent lOC
Subsequent binc
promoter region
specific activato
(177; 130). A-fa
vegetative grout
genetically progr
accumulates to m
where it acts on A
of S. virginiae, vvl
mycelium) and is
It is uncert
SWP’OHU’CLJS di in
lllSllIlCi frOm A‘f"
t ‘ I
Wing exPerime:
L,
MFG/0r (14) 1-
130). A-factor is described as a microbial hormone that is able to diffuse freely across
the cyt0plasmic membrane. It was originally discovered by its ability to cause
sporulation in a bid mutant (deficient in sporulation and antibiotic production) of S.
griseus (92) when wild type and mutant were streaked side-by-side. A—factor exerts its
regulatory function through an interaction with A-factor binding protein (ArpA), present
in the cytoplasm of the cell. In the absence of A-factor, ArpA acts as a repressor of
independent loci required for activation of aerial mycelia and streptomycin genes.
Subsequent binding of A-factor to ArpA causes dissociation of the repressor from the
promoter region of adpA. AdpA then activates expression of the streptomycin pathway-
specific activator StrR and is also implicated in activating aerial mycelium formation
(177; 130). A-factor is produced in low concentrations during the “decision phase” of
vegetative growth, described by Neumann, et al. (1996), as an integral part of a
genetically programmed pathway required for the onset of differentiation. Later, it
accumulates to much higher concentrations during the transition and stationary phases
where it acts on ArpA. This is in contrast to VB, the y-butyrolactone signaling molecule
of S. virginiae, which induces production of the antibiotic virginiamycin (but not aerial
mycelium) and is synthesized just prior to the antibiotic itself (94).
It is uncertain how widespread y—butyrolactone signaling is in the regulation of
Streptomyces differentiation. S. coelicolor produces six such molecules (8) that are
distinct from A-factor, but are able to complement an A-factor deficient mutant in cross-
feeding experiments (66). One of these, expressed only during transition and stationary
phase growth, was purified and shown to stimulate both Act and Red production in S.
coelicolor (l4). Horinouchi and Beppu (1992) predicted y-butyrolactone hormonal
lO
regulators I0 be
regulatory “50
60% of streptor
say that as mOI’t
will be found IC
particular regul;
Wherea:
induction of dif
quomm-sensing
through a progr
cell ages. EV’lClt
convincingly de
for programmec
The idea
developmental p
became commit
nUtrient limitatit
IIutrient to (term,
Who POstulated t
pIOdUCfiOn Was l
exponential Qrou
t2 .-
auxtc growth (L
regulators to be general constituents of morphological and/or secondary metabolite
regulatory cascades. More conservatively, Yamada, et al. (1997) estimated that about
60% of streptomycetes produce butyrolactone signaling molecules. It is probably safe to
say that as more systems are studied, this type of quorum-sensing signaling mechanism
will be found to act both as global inducers of differentiation and as signals for a
particular regulatory cascade of a single antibiotic.
Whereas stringent response and easing of catabolite repression represent
induction of differentiation by way of a particular stress, synthesis of A-factor, as a
quorum-sensing signal or microbial hormone, seems to induce cellular differentiation
through a programmed developmental cycle; a regulatory cascade that progresses as the
cell ages. Evidence for this type of programmed developmental cycle has been
convincingly demonstrated in S. griseus. Inconclusive, but growing, evidence also exists
for programmed development in S. coelicolor.
The idea that S. griseus differentiation was under the control of a timed
developmental program was first proposed by Ensign (1988). He demonstrated that cells
became committed to sporulation at 10-12 hours after spore germination in the absence of
nutrient limitations and that this timing was not varied by 10X additions of any individual
nutrient to defined medium. That study was further elaborated by Neumann, et al. (1996)
who postulated that A-factor-induced aerial mycelium formation and streptomycin
production was initiated during the first stage (the "decision phase") of diauxic
exponential growth in S. griseus. The authors demonstrated, using an A-factor deficient
mutant, that low concentrations of A-factor were required during the first stage of
diauxic growth (up to 10 hours in the growth medium tested) for cellular differentiation
11
to take place, ever
entered stationary
decision phase, tli
by the concentrati
lag (IO-24hr), dtit
not induce differe I
production, L-val
prior to the diauxi
et al. (1996) note;
perceivable nutrit
doubling the conc
proposed that A-f
programmed difr‘t
mOlilhological an
exllOIIential grow
A second 1
Pattern of comppe
Sm’P’Omvces b] 1
v C
formation; most a
collaborators (1 84'
hierarchical casc
glotem .
rEqu 1 red f
on the ability ofc
to take place, even though appreciable quantities were not produced until the culture
entered stationary phase at 24 hours. The earlier that A-factor was supplied during the
decision phase, the higher the final yield of streptomycin, although yield was not affected
by the concentration of A-factor during this phase. Addition of A-factor after the diauxic
lag (IO-24hr), during the second stage of exponential growth (the "execution phase"), did
not induce differentiation. In the same manner, known inhibitors of streptomycin
production, L-valine and staurosporin (a kinase inhibitor), were only effective if added
prior to the diauxic lag (i.e., during the "decision phase"). Like Ensign (1988), Neumann,
et al. (1996) noted that the onset of the diauxic lag occurred in the absence of any
perceivable nutritional or physical perturbation, and its timing could not be altered by
doubling the concentration of any individual nutrient or all nutrients. The authors
proposed that A-factor is an inducer of cellular differentiation that forms part of a
programmed differentiation cycle in which the decision by the cell to switch on
morphological and physiological differentiation is made during the first phase of diauxic
exponential growth.
A second example of programmed development may come from the hierarchical
pattern of complementation witnessed in the bid (bald) mutants of S. coelicolor (185).
Streptomyces bld mutants are so called because they are deficient in aerial mycelium
formation; most are also blocked in antibiotics (reviewed by 31; 35). Losick and
collaborators (184; 185; 127; 126) believe that many of the bld mutants are involved in a
hierarchical cascade of extracellular signaling that controls SapB synthesis - a small
protein required for the initiation of aerial mycelium formation. This hypothesis is based
on the ability of certain bld mutants to complement the aerial mycelium deficiency of
12
other bid strains T
medium(184i 18'
cultivation of am‘
ordered by mutati
(and possibly cm
the first signal in
spanning transpor
associated with h,
tRVA that transla
be determined xvi:
identified in antib
bIdG mutations a:
antisigma genes (7.
the hierarchy, b/u'
molecule that is h
formation to all t}.
SapB SYnthesis (1
ability to Suppress
antibiotic defects
Several Ch
other bld strains through "cross feeding" when grown next to one another on solid
medium (184; 185), or by preconditioning agar medium with one b/d mutant prior to
cultivation of another (1 85; 126). The hierarchy of steps in the regulatory cascade are
ordered by mutations in bld261, -K, -A, -H, -G, -C, and -D. Briefly, bid 26] is required
(and possibly encodes) for the synthesis of an extracellular factor that is believed to be
the first signal in the cascade (126). The bldK gene encodes an ABC membrane-
spanning transporter (127) that is possibly involved in importing the extracellular signal
associated with b1d261. The most extensively studied bld gene, bldA, encodes the only
tRNA that translates the rare UUA leucine codon of Streptomyces (101; 104). It is yet to
be determined what function BldA has in this cascade, but UUA codons have been
identified in antibiotic pathway-specific regulators redZ (182) and actII-ORF 4 (52). The
bldG mutations are complemented by a locus that shows similarity to antisigma/anti-
antisigma genes of Bacillus subtilis (28; 30 and references therein). The final gene of
the hierarchy, bldD, is thought to code for or regulate synthesis of an extracellular
molecule that is heat and protease resistant and capable of restoring aerial mycelium
formation to all the other bld mutants of the cascade, presumably by its ability to restore
SapB synthesis (185). Interestingly, these complementation studies demonstrated the
ability to suppress morphological differentiation without pleiotrOpic suppression of
antibiotic defects.
Several characteristics of the bid mutants suggest their involvement in a
programmed developmental cycle. First, bld mutations tend to be pleiotropic for
morphological and physiological differentiation, and therefore represent early
components of the developmental process where these pathways are still coupled
l3
(Figure ll T
that are UUCOL
Spomiarion) a
134), Likewis
can Suppress b
differentiation
Very few trans:
above, b/dD tra
although this Wt
began to appear
predominantly u
second leg of the
Therefore, given
regulatory genes.
signaling cascade
that is active duri:
Obviously, there
Part of a developr
st: '
ess. Likewise
redilation of ant?‘
(Figure 1). There are many other regulators of either sporulation or antibiotic production
that are uncoupled from each other (discussed below). Whi mutants (impaired in
sporulation) are able to extracellularly compliment bld mutations (e. g., whiF C99, 33;
184). Likewise, overexpression of antibiotic pleiotropic or pathway specific regulators
can suppress bid mutations (29; 65), demonstrating that bld genes control cellular
differentiation upstream of uncoupled sporulation or antibiotic regulatory pathways.
Very few transcript studies have been performed on the bld genes. Of those mentioned
above, bldD transcription was most prominent at the earliest time point tested, 15 hr,
although this was predicted to have already been in transition phase since aerial mycelia
began to appear by 18 hr and antibiotics by 24 hr (48). Antibiotic regulatory genes are
predominantly up-regulated during the transition phase, which likely corresponds to the
second leg of the diauxic curve ("execution phase") described by Neumann, et al. (1996).
Therefore, given that the bid genes act upstream of both whi genes and antibiotic
regulatory genes, and the latter are expressed during transition and stationary phase, the
signaling cascade proposed by Willey, et al. (1993) may form part of a genetic program
that is active during the decision phase to commit the culture to differentiation.
Obviously, there is still much work to be done to define whether the proposed cascade is
part of a developmentally programmed pathway or is induced by nutritional or physical
stress. Likewise, it will be exciting to see how this pathway is integrated into the global
regulation of antibiotics.
l4
PleiOtrO
regulal’
Figure 1. Hi
ofantibiotic prod
.lutations in ma:
leiotropic regulz
pathway-specific
antibiotic gene cl
erar
blrl genes
antibiotics sporulation
Pleiotropic abaA
regulators “55A
absB
absC/mia
ast/K/R2
afsQ
l
redZ Pathway-specific
““114 cdaR ? mmyR regulators
whi genes
rerID
Red Act CDA Mmy
Figure l. Hierarchy of antibiotic regulatory loci. Genetic loci involved in the regulation
of antibiotic production in S. coelicolor can be divided into a three-tier hierarchy.
Mutations in many bld genes block both antibiotic production and sporulation.
Pleiotropic regulators influence the production of more than one antibiotic. Genes for the
pathway-specific regulators are linked to the biosynthetic genes of their respective
antibiotic gene clusters and regulate only the antibiotic that they are associated with.
15
Uncoupling or d
In additio
differentiation dc
also subject to ur‘
detail in model 0'
sporulation, will i
Losick (1997). I
production center
lividtms, but will
differences that 0
process to other 5
Champness (199‘
Streplonrt
undecylpfodigios
,oi.
(hit
Uncoupling of differentiation: regulation of antibiotics
In addition to the joint regulation of morphological and physiological
differentiation demonstrated by bld mutants, sporulation and antibiotic production are
also subject to uncoupled regulatory mechanisms. These have been studied in greatest
detail in model organism S. coelicolor. Numerous genes uniquely involved in
sporulation, whi genes, have recently been reviewed by Chater (1998) and Chater and
Losick ( 1997). This section will concentrate principally on the regulation of antibiotic
production centered around mechanisms discovered in S. coelicolor and closely related S.
lividans, but will also incorporate examples from other species to highlight similarities or
differences that will aid in conferring the current state of understanding of this complex
process to other species within the genus. Reviews on this topic can be found in
Champness (1999); Bibb (1996); Chater and Bibb (1997); and Hopwood, et al. (1995).
Streptomyces coelicolor synthesizes four distinct antibiotics, actinorhodin (Act),
undecylprodigiosin (Red), calcium-dependent antibiotic (CDA), and methylenomycin
(Mmy). The gene clusters for Act and Red have been cloned (109; 110) and the
biochemical pathway for Act synthesis has been extensively studied (reviewed by 72).
An advantageous characteristic of Act and Red is that they are pigmented (blue and red,
respectively, at alkaline pH) which facilitates phenotypic screens for factors affecting
their synthesis. The sequence of cda genes has recently been revealed through the
ongoing Streptomyces sequencing project
(http://www.sanger.ac.uk/Projects/S_coelicolor/). Previous work had partially cloned
and characterized this locus (39), but its regulatory and biochemical characteristics are
still largely unknown. Methylenomycin carries the distinction of being the only
16
Seaplane?“ 3"“
plasmid (95) pre
201th region and
A commr
more pathway-s;
that antibiotic. (
species. Most as
of S. per/celius h.
andrea’Z. Most l
more ofthe oper.
of the methyleno
apparently only c
while the others i
requred for expp
dnrl transcriptior
Dnrl REIdZ and
response requ lat c
(65; 59). A num"
pleiotropic reout ~.
c
Stggest they Con:
my -
and dttr biosynth.
Streptomyces antibiotic whose sequence is carried on a plasmid, SCPl, a 350 kb linear
plasmid (95) present at a copy number of four. The mmy gene cluster was localized to a
20 kb region and found to contain a regulatory gene, mmyR (3 6).
A common characteristic of antibiotic gene clusters is that they encode for one or
more pathway-specific regulators (PSR) that solely influence the expression of genes for
that antibiotic. Champness (1999) lists PSRs for antibiotics produced in numerous
species. Most antibiotic clusters encode a single PSR, although, the daunorubicin cluster
of S. perrcetius has three, dnrl, dnrN, and dnrO, and the Red cluster contains two, redD
and redZ. Most PSRs contain a DNA-binding motif that transcriptionally activates one or
more of the operons carrying antibiotic biosynthetic genes; although at least one, MmyIL
of the methylenomycin pathway is a repressor (36). Where multiple PSRs exist,
apparently only one interacts with biosynthetic gene promoters (e. g., RedD and Dan)
while the others regulate expression of the first. For example, Red Z and DnrN are
required for expression of red and dnr biosynthetic genes because they activate redD and
dnrl transcription, respectively (65; 59). They do not bypass the need for RedD and
Dan. RedZ and DnrN show extensive full-length sequence similarity to two-component
response regulators (RR) of the FixJ subfamily, but have lost the ability to phosphorylate
(65; 59). A number of other PSRs, RedD, ActII-ORF4, Dan, SnoA, and CcaIL and one
pleiotropic regulator, Ast, have recently been classified as a special group of
Streptomyces antibiotic regulatory proteins (SARPs) based on amino acid similarities that
suggest they contain a conserved C-terminal DNA-binding motif like the one found in the
two-component response regulator OmpR (183). Similar heptameric direct repeats in act
and dnr biosynthetic promoters have been predicted as binding sites for ActII-ORF4 and
17
Dnrl, resPeC‘lv'
another (I66)’ 5
with. One final
copy number TC
suggest that thei
Acting b
differentiation, 3
number ofgenet
affecting gromh
pleiotropic regul,
lirt'dans (which I
perform visual sc
loci which have
istKRz (1 15;
putative homolor:
“42» each om.
HERSduction path
The min a
cloned on a mu“;
Slii'idans and S
responsible fOr A
Fig!
Dan, respectively, and although these two SARPs can apparently substitute for one
another (166), SARPs generally only regulate genes of the antibiotic they are associated
with. One final observation about PSRs is that their overexpression at relatively low
copy number results in dramatic increases in antibiotic production (162; 62) which may
suggest that their expression is tightly controlled by the cell.
Acting between the bid genes that regulated morphological and physiological
differentiation, and the PSRs which influence the synthesis of individual antibiotics, are a
number of genetic loci that pleiotropically regulate more than one antibiotic without
affecting grth or sporulation. Remarkably, the only knowledge of antibiotic
pleiotropic regulators in the genus Streptomyces comes from work in S. coelicolor and S.
Iividcms (which like S. coelicolor produces Act and Red) due largely to the ability to
perform visual screens for the pigmented antibiotics Act and Red in these strains. The
loci which have so far been implicated in pleiotropic regulation are abaA (53), absA (2),
qst/K/RZ (115; 116; 174), mia (28), and micX (143). With the exception of absB (a
putative homolog of E. coli RNaseIII (136)) and micX (a possible antisense fragment
(142)) each of the loci have at least a suspected involvement in phosphorylation signal-
transduction pathways.
The mia and abaA loci were isolated because they produced a phenotype when
cloned on a multicopy plasmid (28; 53). The 2.7 kb abaA clone isolated from S.
coelicolor contained five open reading frames and caused overexpression of Act in both
S. Iividans and S. coelicolor when carried on a multicopy plasmid. The region
responsible for Act overexpression was isolated to cotranscribed orfA and orfB (53).
Fragments containing orfA and orfB, or oer alone were able to overexpress Act.
18
Furthermor e
severely redl
locus shows
OffB did not
there is signli
coelicolor, m
u'hiJ show sir
putative OrfA
In cont
four antibiotic.
been narrowed
intergenic regi:
possesses sequ.
region has rece
to contain the ti
shows sequence
however, the C.
re'r‘lacement in r
Promoter region
Caused by the titr
Containing 055C
QmP‘OHEnije k
Furthermore, single cross-over disruption of orfB abolished production of Act and
severely reduced that of Red and CDA while having no effect on Mmy. OrfA of this
locus shows similarity to the transmitter region of two-component sensor kinases (146).
OrfB did not show similarity to any proteins deduced from gene databases. Interestingly,
there is significant structural organization between abaA and the whiJ locus of S.
coelicolor, mutants of which are impaired in sporulation. Three open reading frames of
Mull show similarity to those in abaA, including Orfl which is homologous to the
putative OrfA kinase (146).
In contrast to abaA, the high copy clone of the mia locus abolished synthesis of all
four antibiotics in S. coelicolor (28). The region responsible for the niia phenotype has
been narrowed down to a 90 nt sequence of unknown fianction (30) which lies in an
intergenic region, contains an open reading frame for a 20 amino acid peptide, and
possesses sequence consistent with the formation of a large stem loop (29). The mia
region has recently been included in the newly described absC locus (29) and is believed
to contain the transcriptional start site for 01f! of absC (146). The N-terminus of ORF]
shows sequence similarity to two-component histidine kinases just as in abaA ORF A;
however, the C-terminal domain of ORF 1 is similar to protein phosphatases (146). Gene
replacement in orfl produced an increase in Act. Since the mia fragment lies in the
promoter region of absC orf], it is possible that the high-copy number mia phenotype is
caused by the titration of a regulator of orfl . Therefore, both abaA and the mia-
containing absC loci encode proteins that are not histidine kinases but possess two-
component-like kinase domains and appear to be closely associated with the pleiotropic
19
phenotypes ob:
phosphorylatio
The rer
associated witl
threonine phos
AC1 and Red or
54). In vitro, 1
product of aft}
l5 33% identic;
predicted to CC
can-ling mum]
consistent Wit}
rm” e Sl'nthes
not blpass the
“it“ on antib
devoid of Act
rich medium .5
also found to r
effect On Milly
Through 3 (“EC
Au and Red or
phenotypes obtained from strains overexpressing these regions. The role of these loci in
phosphorylation signal-transduction pathways is currently under investigation.
The remaining pleiotropic regulatory loci all encode for proteins commonly
associated with phosphorylation signal-transduction mechanisms. Ast is a serine—
threonine phosphoprotein that also was isolated by its ability to cause overexpression of
Act and Red when expressed on multicopy plasmids in S. lividans and S. coelicolor (78;
54). In vitro, Ast is phosphorylated by the downstream, convergently transcribed gene
product of asz, a serine-threonine protein kinase (115). The N-terminal domain of Ast
is 33% identical to the PSRs ActII-ORF4 and RedD (115), which have recently been
predicted to contain an OmpR-like DNA-binding motif (1 83). The strain of S. coelicolor
carrying multiple copies of ast produced increased levels of act/In! and redD transcripts
consistent with the effect of Ast on these antibiotics. Similarly, Ast was not able to
restore synthesis of Act or Red in AactII-4 or AredD mutants, demonstrating that it could
not bypass the PSRs (54). A chromosomal in-frame deletion of ast had a conditional
effect on antibiotic synthesis (54). On minimal medium, the ast mutant was essentially
devoid of Act or Red, while on complex medium Act was only slightly delayed, and on
rich medium Act and Red production were comparable to the wild-type strain. Ast was
also found to reduce CDA synthesis in a similar medium-dependent nature, but had no
effect on Mmy. Thus, the Ast Ser-Thr phosphoprotein appeared to regulate antibiotics
through a direct or indirect interaction with the PSRs; however, it was only essential for
Act and Red under certain nutritional conditions.
The final set of pleiotrOpic regulatory loci, absA, afsQ, and cutRS, all encode
independent two-component signal transduction systems. CutRS is a negative regulator
20
ofAct Si’mheSiS
or cmR resPO”S
which could be
reported 8 Sim”
however, recent
S, coelicolor (15
production.
The afSQ
production of pig
plasmid into S. 11
Surprisingly, no
the S. coelicolor
was able to suppi
synthesis). Givei
number, the abse
mean that, like A
conditions.
The final
regdlaior in S_ CU
hum mutants that
ambimics (2). L
is , .
We the AbsA
of Act synthesis in S. Iividans (32). Gene disruptions in either cutS histidine kinase (HK)
or cutR response regulator (RR) resulted in accelerated and increased production of Act,
which could be suppressed by introducing the cloned culR gene. Chang, et al. (1996)
reported a similar phenotype upon creating a gene replacement in cutS of S. coelicolor;
however, recent evidence suggests that this locus does not regulate antibiotic synthesis in
S. coelicolor (15). No information has been given as to the effect of CutRS on Red
production.
The afsQ locus of S. coelicolor was discovered by virtue of its ability to stimulate
production of pigmented antibiotics and A-factor when introduced on a multicopy
plasmid into S. Iividans (84). This locus encoded the AfsQl RR and AfsQ2 HK.
Surprisingly, no phenotype was obtained when either of these genes were disrupted on
the S. coelicolor chromosome. In contrast, a low copy-number plasmid carrying afsQI
was able to suppress an S. coelicolor absA mutation (globally deficient in antibiotic
synthesis). Given the ability of this gene to suppress the absA mutation in low copy
number, the absence of a phenotype in the afsQ chromosomal disruptions may simply
mean that, like Ast, this system is only essential or active under certain nutritional
conditions.
The final two-component system thus far identified as an antibiotic pleiotropic
regulator in S. coelicolor is encoded by the absA locus. The absA locus was isolated
from mutants that grew and sporulated normally but were deficient in pigmented
antibiotics (2). Like the absB and mia strains, absA mutants were shown to be globally
blocked in synthesis of all four S. coelicolor antibiotics. The absA locus was shown to
encode the AbsAl HK and the AbsA2 RR. It was later discovered that mutations in the
21
AbsA] HK W61
(19). Gene rep
expression of d
overexpression
antibiotics) (19
was a negative
phenotypes, th
uctII-ORF I V s
produced (I).
or indirect trar
It is ap
PhOSphorylatir
onset of antibi
sporulation 1.
cascade or are
repressing the
independent 52
in other Gram.
regulamrs, Sci:
Streptomdp,Ces_
The b 8
including Spor
i
50“! baCtti‘riurn
AbsAl HK were responsible for the Abs‘ phenotype exhibited by the original mutants
(19). Gene replacement or disruption of absA I, believed to have a polar effect on the
expression of downstream absA 2, resulted in the opposite phenotype, early onset and
overexpression of pigmented antibiotics (Pha phenotype; precocious hyperproduction of
antibiotics) (19). Therefore, it was hypothesized that the AbsA two-component system
was a negative regulator of antibiotic synthesis. Consistent with their antibiotic
phenotypes, the Abs‘ and Pha absA mutants affected expression of the PSRs redD and
actII-ORFIV such that their levels varied in agreement with the amount of antibiotic
produced (1). Therefore, AbsA seemed to regulate antibiotic synthesis through the direct
or indirect transcriptional control of the PSRs.
It is apparent fi'om a review of antibiotic pleiotropic regulators that
phosphorylation signal transduction plays an extremely important role in regulating the
onset of antibiotic synthesis after this process has been uncoupled from that of
sporulation. It is yet to be determined if these regulators comprise a linear regulatory
cascade or are part of an integrated network of independent cascades either activating or
repressing the production of multiple or individual antibiotics in response to numerous
independent signals. Examples of two-component regulation of developmental processes
in other Gram-positive organisms, in addition to data from S. coelicolor pleiotropic
regulators, seems to support an integrated network of antibiotic regulation in the genus
Streptomyces.
The best example of regulation of growth-phase dependent differentiation
including sporulation and secondary metabolite formation comes for the Gram-positive
soil bacterium B. subtilis (reviewed by 192). This organism demonstrates stationary-
22
phase productiO
Surfactin synthd
negative control I
signal transduct:
various compon
molecules (e. g.
cell tip the whol
regulatory netwt
pathway are alsc
competence, spc
component and t
(6g. surfactin s_~.
signals rather th.
Transcri p
anAlysis of the e;
phase production of a small enzymatically synthesized peptide antibiotic surfactin.
Surfactin synthesis is regulated by a complex integrated network of multiple positive and
negative controls involving various two-component systems and other phosphorylation
signal transduction molecules (e. g., aspartyl-phosphate phosphatases). In addition,
various components of this regulatory web are under the control of independent effector
molecules (e.g., CodY, ComX, Cfx), whose relative concentrations inside or outside the
cell tip the whole mechanism toward overall positive or negative control. The surfactin
regulatory network is not physiologically isolated; instead, the molecules active in this
pathway are also involved in regulation of other stationary-phase processes such as
competence, sporulation, and degradative enzyme synthesis. Thus, where numerous two-
component and other phosphoprotein signal transduction regulators mediate a process
(e.g., surfactin synthesis), there is precedent for integration of multiple pathways and
signals rather than a single linear regulatory cascade.
Transcript studies suggest that pleiotropic regulators absA, afsQ, ast, and cutRS
are transcribed simultaneously during transition and stationary phase growth; however, a
detailed examination of their simultaneous temporal expression is lacking, as is a detailed
analysis of the epistatic relationship between them. AbsA and Ast have been shown to
affect the transcription of PSRs. In addition, multiple copies of ast or afsQI suppressed
the absA mutation in an Abs’ strain to restore antibiotic synthesis. This would suggest
that Ast and AfsQ act either downstream of AbsA in a linear cascade or in a separate
cascade that integrates into a global antibiotic regulatory pathway. In support of the
latter, chromosomal disruptions of afsQI and ast did not affect antibiotics under
conditions in which 0123/! mutations do show a phenotype (84; 54; 2; 19). The fact that
23
45R and afsQ
demonstrated 1
pathways respt
deal of data is 1
suggests that tl
synthesis in S. .
In orde:
coelicolor, it is
involved. Onc
each independi
allow one to in
identified base
biochemical m
tWO‘COmPOIler
the Signal that
“Eduction, 5
WEEKS) OfIhe
Sl’stem FCgulai
pathway“) m
TmeOmPOi
Two“
ast and afsQ are not essential for antibiotic production under conditions in which AbsA
demonstrated global control suggests that they may be part of independent integrated
pathways responding to different environmental signals. Therefore, even though a great
deal of data is missing on the interactions of pleiotropic regulators, that which exists
suggests that they do not form a single linear regulatory cascade controlling antibiotic
synthesis in S. coelicolor.
In order to begin constructing a unified model for antibiotic regulation in S.
coelicolor, it is not sufficient to simply identify the regulatory molecules that are
involved. Once identified, it is necessary to determine the biochemical mechanism of
each independent system and establish the interactions that exist between systems to
allow one to integrate these into the overall pathway. Of the multiple loci that have been
identified based on their ability to produce a phenotype, very little is known about the
biochemical mechanisms which account for their activity. With respect to the various
two-component systems AbsA, AfsQ, and CutRS, questions of interest include: what is
the signal that modifies HK activity; is phosphorelay the mechanism of signal
transduction, and if so, what is its role in altering the activity of the RR; what is the
target(s) of the RR and is its activity positive or negative; and, how is expression of this
system regulated during the course of growth and in comparison to other genes of the
pathway(s) they regulate?
Two-Component Signal Transduction Systems
Two-component signal transduction systems are ubiquitous throughout eubacteria
as a means of regulating varied aspects of cell physiology in response to environmental
24
conditions, inc
The paradigm '
kinase and rest
component sys
component prc
Stock et al. (I
HK prc
domain and a I
two to eight tr
domain is hi gl
specificity of a
region r88pm]:
understood in
been best Chap
transmitter do
aSSOciated Wit
{minimal hor
mo‘compone
in these Prote
mimics; aUl
catFried Out in
axial-late phO
conditions, including physical, behavioral, nutritional, and developmental responses.
The paradigm two-component system (Figure 2) is composed of two proteins, a histidine
kinase and response regulator. The mechanism of signal transduction in all two-
component systems studied to date is phosphorelay. Extensive reviews of two-
component protein structure and function are presented in Parkinson and Kafoid (1992),
Stock, et al. (1995), and Volz (1995).
HK proteins are generally composed of two-domains, an N-terminal sensor
domain and a C-terminal transmitter domain. Sensor domains frequently contain from
two to eight transmembrane helices which anchor the HK to the cell membrane. This
domain is highly divergent between systems, which may in part be due to the high signal
specificity of each sensor. As the name implies the sensor domain is thought to contain a
region responsible for signal molecule recognition, although this mechanism is poorly
understood in most two-component systems. Among those HKs where signal sensing has
been best characterized are NarX (25; 102), VirA (154), and PhoR (179). The HK
transmitter domain is more highly conserved and ispresponsible for the enzymatic activity
associated with this molecule. Several subdomains of the transmitter share sequence and
functional homology and are conserved in their organization. The H-box, present in all
two-component HKs, contains a conserved His residue that is the site of phosphorylation
in these proteins. The transmitter domain is associated with two independent enzyme
activities; autophosphorylation at the conserved His residue which is believed to be
carried out in trans between the monomers of HK homodimers (161) and phospho-
aspartate phosphatase activity directed at the phosphorylated form of the R. The
25
Figure 2. Twc
N’O-componer
features conser
consists of two
transmembrane
domain has als
for the high SE .
terminal transn
“lth the amino
Which is the co
mains have i
activity. Th e U
the histidine ki
rEileiver domai
CC reSldLles
orm the acidic
e ECtor domai;
regulalOrs POSS
Elix‘m‘helin
t{afield-pt1.01121]
A. Histidine Kinase
I E U |H| lNlhrl IFIIGI l
Sensor Transmitter
Domain Domain
B. Response Regulator
[lDl lDl lKlHlHTHl l
Receiver Effector
Domain Domain
Figure 2. Two-component signal-transduction proteins. Primary structure diagrams of
two-component histidine kinase and response regulator proteins demonstrate common
features conserved among members of this family. (A.) The histidine kinase frequently
consists of two domains. The N-terminal sensor domain contains between two and eight
transmembrane helices that anchor the histidine kinase to the cell membrane. The sensor
domain has also been implicated in signal sensing in some systems, which may account
for the high sequence divergence of this domain among all histidine kinases. The C-
terminal transmitter domain has several highly conserved subdomains indicated by boxes
with the amino acid residue for which they are named. The H-box contains the histidine,
which is the conserved site of phosphorylation in all histidine kinases. Other sub-
domains have been implicated in nucleotide and Mg2+ binding required for enzymatic
activity. The transmitter domain possesses the auto kinase and phosphatase activity of
the histidine kinase. (B.) The response regulator normally consists of two domains. The
receiver domain is highly conserved among all members of the two-component family.
Three residues conserved in all response regulators are the Asp, Asp, and Lys, which
form the acidic pocket in which the central Asp is phosphorylated. The C-terminal
effector domain is responsible for the regulatory action of the protein. Most response
regulators possess a DNA binding region in the effector domain, indicated here by a
helix-tum-helix motif, which is consistent with the function of these proteins as
transcriptional regulators.
26
phosphorylat
HK to a cons
description, 1
additional dc
Lord and Li
C-terminusc
additional re
accepting his
accepting dc
The '
terminal rec.
With three if
central Asp
mm Ofien r
trartscriptio1
Se‘ernCe Sir
Wiants 0f ]
SW0}: 0f 8
the receiver
Ema-media:
83%“ dor
abseme or.
phosphorylated HK also acts as a donor for phosphoryl group transfer from the His of the
HK to a conserved Asp of its cognate RR. While the majority of HKs fit the above
description, there are also numerous hybrid HKs, most of which have one or more
additional domains C-terminal to the transmitter domain. For example, hybrid HKs
LuxN and LuxQ of Vibrio harveyi have response regulator receiver domains fused to the
C-terminus of their transmitter domains (13). Likewise, BarA and Ach have an
additional receiver domain, but also contain another C-terminal domain with a phospho-
accepting histidine independent of the transmitter His (83). These extra phospho-
accepting domains undoubtedly complicate signal transduction between the HK and R.
The two-component R is usually a two-domain cytoplasmic protein. The N-
terrninal receiver domain is highly conserved among virtually all members of this family,
with three invariable residues, Asp, Asp, Lys, which form an acidic pocket in which the
central Asp is the site of phosphorylation (Figure 2). The C-terminal effector domain
most ofien contains a DNA-binding motif, such that most response regulators act in
transcriptional regulatiOn. Response regulators are grouped into subfamilies based on
sequence similarity within their effector domains (reviewed by 161). The most common
variants of Rs are proteins which possess only the receiver or effector domain, such as
SpoOF of B. subtilis (191) and GerE of E. coli (40). Those proteins which contain only
the receiver domain make up part of signal-transduction cascades in which they act as
intermediate signal-transduction molecules between an orthodox HK and RR. Where the
effector domain exists on its own, it is able to mediate transcriptional regulation in the
absence of signal. An offshoot of the signal-independent GerE protein are pseudo
27
response my
domain, full-l
Phosp
systems (revi
region of the
membrane tc
domain. Sui
This causes :
permits reco
transcription
suggests tha
that pl’Opoge
Cu (139), ,
bl’ Other res.
activity f0”
l50late the S
response regulators such as RedZ and DnrN mentioned earlier (65; 59), which have dual-
domain, full-length homology to RRs, but have lost the requirement for phosphorylation.
Phosphorylation is the mechanism of signal transduction in two-component
systems (reviewed by 161). Under the two-component paradigm (Figure 3), the sensor
region of the HK senses a signal which causes a conformational change across the cell
membrane to stimulate autophosphorylation at the conserved His of the transmitter
domain. Subsequently, the phosphoryl group is transferred to the Asp of its cognate RR.
This causes a conformational change between the receiver and effector domains that
permits recognition and binding of the effector domain to its target promoter and
transcriptional activation of the target gene. A review of two-component systems
suggests that most HKs are phosphatase dominant in the absence of signal. Exceptions
that propose kinase default HKs in the absence of signal include the EnvZ/OmpR (135),
Cpx (139), and LuxN/Q/O (56) systems; however, the status of the first two are debated
by other researchers (55; 132). The difficulty in unequivocally determining the default
activity for the great majority of two-component systems is the inability to define and
isolate the signal. Nevertheless, with the exception of Lux , in all other systems where
good signal identification data exists, the signal stimulates HK kinase activity.
While two-component phosphorylation-mediated signal transduction may appear
at first glance to confer all or nothing regulation, a closer look at the mechanism reveals
its ability to provide subtle, finely tuned responses. In part, this is due to the dual
28
Tune 3. Paradi
pen-1e of signal
11; Under the:
cgiator interacts
aresponse regul.
alts of INA
ante of signa
sline kinase ca
train stimulates
mime. The pl
calator. Again
fill region to re
ipfled here as 1
{non in which ‘
trigalconcentra
mithe histidin
:i‘phanse activ
fiOSphoryIated
Dpnrtion of p
‘ffél‘onding our
nations.
Figure 3. Paradigm two-component signal-transduction system activity. (A.) In the
absence of signal most histidine kinases are thought to be in a phosphatase dominant
state. Under these condition, the unphosphorylated receiver domain of the response
regulator interacts with the effector domain to inhibit binding to the target promoter. If
the response regulator is an activator, then transcription is blocked; indicated here by the
inability of RNA polymerase holoenzyme to recognize the promoter. (B.) In the
presence of signal, ligand binding to a recognition site in the sensor domain of the
histidine kinase causes a conformational change across the membrane. The transmitter
domain stimulates autokinase activity and phosphorylation occurs at the conserved
histidine. The phosphoryl group is transferred to the Asp of its cognate response
regulator. Again phosphorylation causes a conformational change, now allowing the
HTH region to recognize its target promoter, which effects transcriptional regulation
(depicted here as activation). (C.) Module C of this diagram depicts the hypothetical
situation in which the concentration of signal decreases or is lower than that in module B.
A signal-concentration dependent kinase/phosphatase equilibrium is established in which
part of the histidine kinase population is in the kinase state, while the remainder possess
. phosphatase activity. Some phosphorylated response regulator molecules may be
dephosphorylated such that the regulatory output at the target promoter reflects the
proportion of phosphorylated to unphosphorylated response regulator. The
corresponding output would be intermediate to signal saturation or signal deficient
conditions.
29
12.3% macs— oaaEvn uo~a>=o< 22.3%
.U .m
“53.555 on5ai~ao5~
Enummxo connects
.ft
noun—=9“ coo—60¢
Gama .9.qu
.0
55953... Enotmtomnauh
cum—EM m8~>uu< .anwE
.m
noun—awe“ oz
2358.5 832323
iuwmm a: 92522
.<
30
kinase-it’ll" 51
signal Slim“I
saturating le‘
eerClSe phOf
signal-concer
phosphatase a
output interm
159).
In addi
characteristics
more than one :
HR and, therefr
NarX autophosi
signal associate
target promoter
variable not onl
For example, th
Entensivel y Char
Sequence has be
diversity With re
OmpR rec0gniz
Similar ity betwe-
Ql require any,
kinase/phosphatase activities of the HK (161). In the two-component paradigm, the
signal stimulates HK kinase activity. However, unless the signal is present at sensor-
saturating levels, only part of the HK population is in the kinase mode, the rest still
exercise phosphatase activity toward the phosphorylated response regulator. Therefore, a
signal-concentration dependent equilibrium is established between the kinase and
phosphatase activities of the HK that combine to produce a corresponding regulatory
output, intermediate to that of the signal saturated or signal deficient extremes (145;
159)
In addition to the signal-concentration-dependent response, several other
characteristics of two-component systems add to their complexity. Several systems have
more than one signal that may vary in their ability to trigger autophosphorylation of the
HK and, therefore, produce different degrees of response as was recently exemplified by
NarX autophosphorylation in response to nitrate versus nitrite (102). In addition to the
signal associated properties of the I—IK, the response of two-component regulators at their
target promoters can also vary greatly. Promoter recognition sites can be extremely
variable not only between different response regulators, but even for the same regulator.
For example, the promoter region for at least eight operons regulated by NarL have been
extensively characterized (reviewed by 159). While a weak heptameric consensus
sequence has been proposed for NarL binding, heptameric sequences exhibit great
diversity with respect to number, location, orientation, and spacing. Similarly, while
OmpR recognizes multiple decameric sites in the ompF and ompC promoters, there is no
similarity between C-box and F-box sequences (135). In addition, regulatory responses
can require anywhere from a single to multiple copies of the regulator in only the
31
phosphoryle
phosphoryla
while some
response in
multiple twi
mechanism:
competence
appear to pl
outside of t.
signal to a 1
The AbsA
The
that negatis
HK is Pfed
highly Con:
ClOmain (a;
deduced pr
fideofag
the 5’0me
dgmatn (1s
F‘gmentatil
phosphorylated state (e. g., SpoOA, reviewed by 157), or a combination of the
phosphorylated and unphosphorylated states (e. g., OmpR (159) and UhpA (41)). Finally,
while some two-component systems seem to be the only regulator of a particular
response in simple pathways (e.g., Uhp transport system (89)), other pathways integrate
multiple two-component systems together with other transcription factors and feedback
mechanisms to fine tune a response through a combination of interacting regulators (e. g.,
competence gene expression in B. subtilis (121)). Therefore, while phosphorelay does
appear to play a central role in the signal transduction of all two-component systems,
outside of this mechanism enormous variability exists in the process of converting a
signal to a response.
The AbsA two-component system of S. colelicolor
The absA locus encodes a two-component signal transduction system (Figure 4)
that negatively regulates production of all four S. coelicolor antibiotics (19). The AbsA]
HK is predicted to be membrane bound and undergo phosphorylation at Hi5202 of the
highly conserved transmitter domain. AbsAl also has a relatively large C-terminal
domain (approximately 160 amino acids) that does not show sequence similarity to
deduced proteins from gene databases and is of yet unknown function. A vector-bome
allele of absA] used to complement the Abs' C542 and C577 mutants was truncated at
the BamHI site of absA 1, which removed the last 69 amino acids of the C-terminal
domain (19). Nevertheless, this truncated form of AbsAl was able to restore wild type
pigmentation to the Abs‘ mutants, calling into question the requirement
32
AA P X
m
w
X
Z
2
U3
1K
SCESJ 7c I I absAI m sass. 20c I
B.
IHI INIIDIlGI Ll El El lHTHl
Sensor Transmitter C-terminal Receiver Effector
Domain Domain Domain Domain Domain
Histidine Kinase Response Regulator
Figure 4. The AbsA two-component system. (A.) A physical map of the absA locus
and surrounding genome; (B.) Primary sequence diagrams of the AbsAl histidine kinase
and AbsA2 response regulator. AbsAl possesses three domains. The N-terminal sensor
domain is predicted to contain four transmembratne helices. The transmitter domain
contains sub-domains conserved in other histidine kinases including the H-box with the
putative site of phosphorylation at His202. The unique C-terminal domain of AbsAl is
of unknown function and shows no similarity to other proteins deduced from gene
databases. The AbsA2 response regulator consists of two domains. The receiver domain
contains the highly conserved residues of other response regulators including a putative
site of phosphorylation at Asp54. The C-terminal effector domain contains a helix-tum-
helix DNA-binding motif, which allows AbsA2 to effect transcriptional regulation at a
target promoter(s). Lettered boxes represent highly conserved sub-domains of functional
importance named after a highly conserved residues or motifs found therein. Dark boxes
represent transmembrane helices. Restriction sites are A, ApaI; B, BamHI; N, NaeI; P,
PstI; S, SacI; and X, )0101.
33
of the C-terminal domain in AbsAl signal sensing and transduction. The AbsA2 R is a
member of the FixJ subfamily of two-component regulators. It has two domains, an N-
terminal receiver domain with the putative site of phosphorylation at Asp54, and a C-
terminal effector domain containing a helix-tum-helix DNA-binding motif, which is
predicted to regulate target promoter transcription.
The absA locus was originally identified in Abs' mutants that were globally
deficient in antibiotic synthesis (2). The mutations responsible for the Abs' phenotype
were localized to a 1.45 kb region of absA 1 (19) in two independent isolates, C542 and
G577. Interestingly, sab (suppressors of Abs) mutants arose spontaneously at a
relatively high frequency (0.1%) in C542 or C577 protoplasts (19). Also, a gene
replacement and gene disruption of the absA 1 gene resulted in precocious
hyperproduction of antibiotics (Pha), characterized by early onset and overproduction of
antibiotics. The absAI gene knockouts were believed to have a polar effect on
downstream absAZ due to the close proximity of the two genes. Given these results, it
was hypothesized that absAZ encoded a negative regulator of antibiotics since its
elimination resulted in the Pha phenotype (19). The Abs' and Pha mutants were found to
affect the level of expression of redD and actII-ORF4 in a manner which corresponded
to their phenotypes (1). In relation to parental stain 11501, PSR expression was lower in
the Abs‘ mutant and higher in the Pha strain. Therefore, AbsA appeared to mediate
production of antibiotics through direct or indirect transcriptional regulation of the
antibiotic pathway-specific regulator genes.
34
fart]
13m:
is C0
In this study I continued to examine basic characteristics of alleles and gene
products of the absA locus in order to define in greater detail the mechanism of AbsA-
mediated antibiotic regulation. Given previous demarcation of the mutations causing the
Abs‘ phenotype to a 1.45 kb region of absA l, the C542 and C577 absA] mutant alleles
were sequenced in order to characterize, at a molecular level, changes responsible for
these phenotypes. Characterization of numerous sab mutants was also discussed,
including the identification of various second-site suppressor mutations localized within
the absA locus. The AbsAl HK and the AbsA2 RR possess conserved residues
consistent with the formation of active sites involved in phosphorelay signal transduction.
Thus, a genetic and biochemical approach was taken to determine whether phosphorelay
was active in this system, and to define the role of phosphorylation in mediating AbsA2
activity. Because antibiotic production is growth-phase dependent, a transcript analysis
of absA was performed. Identification of the absA transcription start site permitted
analysis of its promoter region. In addition, temporal expression of absA was
determined, which allowed for qualitative characterization of its timing with respect to
that of growth and antibiotic production. It was known from previous work that AbsA
influenced the expression of redD and actII-ORF 4 PSRs. In this study I examined the
temporal expression of the PSRs redD, redZ, and cdaR in comparison to that of absA and
further examined the influence of AbsA on expression of these regulators as putative
targets of AbsA2. Finally, preliminary results were presented which suggest that AbsA2
is conserved in other species of Streptomyces.
35
CHAPTER 2
SEQUENCE ANALYSIS OF absA ALLELES OF Abs' AND sub MUTANTS
36
The absA locus of S. coelicolor was isolated from mutants that demonstrated an
uncoupling of the temporally parallel processes of sporulation and antibiotic synthesis
(2). Two independent isolates, C542 and C577, that mapped to the same region of the
chromosome, were globally deficient in the synthesis of all four S. coelicolor antibiotics
(Abs', antibiotic synthesis deficient) while unaffected in sporulation. The absA locus was
later shown to encode a two-component signal transduction system comprised of the
AbsAl histidine kinase and AbsA2 response regulator (19). Marker exchange and
marker rescue experiments in the C542 and C577 Abs‘ mutants localized the mutations
responsible for this phenotype to a 1.45kb region of absAI (19). This phenotype was
dramatically opposed to the early onset and overproduction of antibiotics (Pha
phenotype) obtained from an absAI gene disruption and gene replacement (19). The Pha
phenotype resulting from absAI disruptions was hypothesized to result from polar effects
on downstream absAz. Therefore, it was proposed that AbsA2 encoded a negative
regulator of antibiotics and that the C542 and C577 Abs' strains mutationally locked the
AbsA system into a negative regulatory state.
Another characteristic of the C542 and C577 Abs' mutants was that they
underwent apparent spontaneous reversion. Pseudorevertants of the Abs‘ phenotype, sab
(auppressor of alas) mutants, which restored synthesis of all four antibiotics,
spontaneously arose in the C542 and C577 absA mutant protoplasts at a frequency of
0.1% (19). The sub mutants were of considerable interest because identification of
second site suppressors is a useful tool for finding additional members of a regulatory
pathway. Alternatively, if localized to the absA locus, these pseudorevertants might
provide insights into the mechanism of AbsAl/AbsA2 interactions.
37
In this study further characterization of the C542 and C577 Abs' mutants was
undertaken. It was of considerable interest to sequence the aim“ alleles of the Abs'
mutants C542 and C5 77 to confirm that mutations responsible for the Abs' phenotype
were indeed present. Additionally, it was hoped that these mutations would provide
evidence, when analyzed together with sab mutations (below) and site-directed mutations
(Chapter 3) into the biochemical mechanism that locks AbsAl into a negatively acting
state. This chapter was also concerned with the characterization of sab mutants,
beginning with colony purification and phenotypic analysis on through to genetic
mapping and sequence identification of various sab mutations within the absA locus.
This work is presented in " Genetic suppression analysis of non-antibiotic-producing
mutants of Streptomyces coelicolor absA locus" (6) . My contribution to this publication
includes designing the sequencing strategy, amplification and preparation of DNA for
sequencing, and analysis of raw sequencing data for the absAI and sub mutant alleles.
Conclusions drawn from this work as a whole are reprinted from the text (6):
(i) Non-antibiotic-producing (Abs-) mutants of the absA locus, which seem to
lock the AbsA regulatory system into a negatively regulating mode, contain point
mutations in conserved domains of the AbsA] histidine kinase sensor-transmitter
protein.
(ii) The absAI mutants spontaneously acquire suppressive mutations that restore
antibiotic synthesis.
(iii) Plasmid-mediated and protoplast fusion mapping techniques were useful for
genetic analysis of suppressive (sab) mutations, locating some close to absA.
38
(iv) Actinophage ¢C31-derived vectors were useful for marker rescue and marker
exchange experiments that verified the existence and location of sab mutations
and allowed transduction of sab mutations from strain to strain.
(v) Sequence analysis defined sab mutant residues in the absA two-component
system. Some sab alleles (Type I) restore a wild-type phenotype to Abs" mutants,
whereas some (Type 11) cause antibiotic overproduction.
(vi) Antibiotic overproduction in sab strains can result from deletion of absA,
consistent with absA's proposed role as a negative regulator, but the most strongly
pigmented sab strain contains a point mutation in the AbsA2 response regulator,
suggesting a complex role for the absA locus in production of antibiotics.
39
CHAPTER 3
GENETIC AND TRANSCRIPTIONAL ANALYSIS OF absA, AN ANTIBIOTIC
GENE CLUSTER-LINKED TWO-COMPONENT SYSTEM THAT REGULATES
MULTIPLE ANTIBIOTICS IN S. coelicolor
40
31
De
INTRODUCTION
Streptomycetes are notable among prokaryotes for their fiangal-like
developmental cycles. Early in the growth of a colony, multinucleoidal vegetative
hyphae extend through the growth medium, branching extensively to form a mycelia]
mat. Later, in response to poorly understood signals, the vegetative hyphae initiate a
program of multicellular differentiation. Morphological differentiation produces
sporulating aerial hyphae on the colony surface (reviewed by 37; 34) while the
temporally parallel but spatially distinct process of secondary metabolite (“antibiotic”)
production occurs in the substrate mycelium (reviewed by 31; 35).
Streptomycete antibiotic biosynthetic pathways involve multiple enzymes that are
encoded in large clusters of genes. Each species typically contains several antibiotic gene
clusters and these are subject to a complex network of regulation. Much of what is
known about the regulation of antibiotic genes has come from genetic studies in
Streptomyces coelicolor. One level of regulation that was discovered in S. coelicolor,
and is now known to be common to most if not all streptomycetes, involves so-called
“pathway-specific regulation,” a mechanism in which a cluster-linked transcriptional
regulator — usually an activator -— regulates expression of numerous polycistronic
transcripts in an antibiotic gene cluster. In the cases of the S. coelicolor antibiotics
actinorhodin and undecylprodigiosin, which are especially well characterized, the
pathway-specific activators are ActII-ORF4 and RedD, respectively (122; 167; 62; 57).
Both are OmpR-like DNA binding proteins and are founding members of the SARP (for
streptomycete antibiotic regulatory protein) family of regulators, which also includes
many of the known cluster-linked regulators for other streptomycete antibiotics (183).
41
Studies of S. coelicolor antibiotics have been facilitated by the ease of assaying
the antibiotics. Two are pigments: actinorhodin (Act) and undecylprodigiosin (Red) are
blue and yellow, respectively, at alkaline pH; both are red at acidic pH. The other two S.
coelicolor antibiotics, calcium-dependent antibiotic (CDA) and methylenomycin (Mmy),
can be assayed in simple plate culture-inhibition assays. Production of the S. coelicolor
antibiotic pigments can easily be seen to be growth-phase regulated in both plate and
liquid cultures. It has been demonstrated that this temporal regulation results fi'om
growth-phase regulated expression of the pathway-specific activators (167; 62; 182). It is
less well understood what regulates the pathway-specific activators. However, one such
control involves the absA two-component system, which was discovered in a genetic
analysis of global, or coordinate, antibiotic regulation. Mutants of absA were first
identified because of their actinorhodin/undecylprodigiosin-minus, sporulation-plus
phenotype; subsequently, they were shown to be calcium-dependent antibiotic-minus and
methylenomycin-minus, as well (2). This phenotype was named Abs’ and classical
genetic mapping showed that the Abs' phenotype was attributable to mutations in the
absA locus (2; 19) Further work showed a deficiency of actII-ORF 4 and redD
transcription in absA mutants (1), explaining the Abs' phenotype, at least with respect to
actinorhodin and undecylprodigiosin.
The genetic map location of absA was far from the act and red gene clusters, but
was close to the only existing cda mutant. Recent genomic sequencing of S. coelicolor
has revealed that absA is associated with the cda gene cluster
(http://www.sanger.ac.uk/Project/S_coelicolorl). Previous to the genomic sequencing,
only a peptide synthetase-encoding segment of the cda cluster had been defined (39).
42
Now, it is apparent that absA lies in a 12 kb region between the peptide synthetase genes
and a putative SARP-like regulator for the cda cluster, cdaR (Figure 5). The function of
absA as a regulator of multiple antibiotic clusters, while being genetically associated with
one cluster, makes absA highly unusual among antibiotic regulators.
In typical two-component systems, a dimeric histidine kinase uses ATP to
autophosphorylate, with one subunit transphosphorylating the other on a specific
conserved histidine residue (reviewed by 161). The phosphoryl group is then transferred
to an aspartate residue on a cognate response regulator, modulating its activity as a
transcriptional regulator. The absA-encoded two-component system is highly
“orthodox,” including the features common to many of the better-studied two-component
systems. The absA] gene is predicted to encode a histidine kinase and the adjacent,
downstream gene, absA2, is predicted to encode a response regulator with a C-terrninal
helix-turn-helix DNA binding domain. Following the two component paradigm, sequence
conservation predicts that the AbsAl protein would autOphosphorylate at His 202 and the
phosphoryl group would transfer to Asp54 of AbsA2. AbsA2 is highly homologous to
NarL of E. coli and the transmitter domain of AbsAl is similar to the cognate kinases,
NarX (63). Closely related two-component systems from Bacillus subtilis include
DegS/DegU and ComP/ComA (reviewed by 121).
43
17.5 kl:
Figure 5. Position of absA with respect to the cda gene cluster. This 58.3 kb region of
the cda cluster was reconstructed from sequence data made available by the Streptomyces
coelicolor Sequencing Project (The Sanger Centre). Genes shown in white have been
named and given putative functions based on genetic or functional analysis. cdaR is
homologous to pathway-specific activators. Biosynthetic genes cdaPSI, cdaPSII and
cdaPSIII encode peptide synthases which catalyze steps in the enzymatic synthesis of the
lipopeptide antibiotic CDA. Shaded genes have been assigned putative functions based
on sequence similarity to other proteins (annotated in
http://www.sanger.ac.uk/Projects/S_coelicolor/).
Marker rescue experiments (19) and subsequent sequence amlysis (6) of absA
mutants located the mutations that were responsible for the Abs" phenotype to the
transmitter domain of AbsA]. Below, we refer to these alleles as absA 1'. Additional
genetic experiments revealed that absA could also mutate to a phenotype essentially
opposite to Abs'; this phenotype was characterized by an early onset and increased level
of antibiotics (l9). Antibiotic gene transcription was correspondingly increased in the
44
overproducing mutants (1). Two absA disruption mutations caused the overproduction
phenotype, suggesting that the role of absA in antibiotic regulation was primarily
negative (19).
We have undertaken a genetic dissection of the absA locus, which we describe
here. This work evaluates the role of phosphorylation in absA-mediated regulation of
actinorhodin, undecylprodigiosin, and CDA and establishes the genetic basis for the two
opposing phenotypes observed in absA mutants. We also describe a transcriptional
analysis of the absA genes which reveals autoregulatory behavior of AbsA2. Together,
the results of these experiments have implications for the mechanism by which AbsA
signal transduction regulates antibiotics during the Streptomyces coelicolor life cycle.
MATERIALS AND METHODS
Growth Conditions
Streptomyces strains were cultured in YEME broth (73) for use in plasmid and
protoplast preparations. Cultures used for chromosomal DNA extraction were grown for
two days either in YEME broth or on SpMR (91) plates overlaid with cellophane disks.
Cultures used for RNA extraction were grown in 50 ml of SpMR broth in 300 ml baffled
flasks, inoculated with 108 spores, and incubated at 30°C, 250 rpm for 18, 30 or 54 hours.
Thiostrepton was added to obtain a final concentration of 10 ug/ml in liquid culture or
200 ug/ml in agar. Hygromycin (Hyg) was added to agar plates to a final concentration
of 200 pig/ml. Escherichia coli was grown in L broth or L agar (147). Ampicillin was
added to obtain a final concentration of 50 ug/ml in both agar and broth.
45
Antibiotics Assays
Assay conditions for the calcium-dependent antibiotic were as previously
described (2). Strains were grown on Oxoid nutrient agar (ONA), or R5 (73), and were
placed onto plates with or without added calcium (as Ca(N03)2 to 12 mM). Soft ONA or
ONA plus Ca was seeded with CDA-sensitive Staphylococcus aureus and was overlaid
around the plugs. Plates were incubated overnight at 37°C. Actinorhodin and
undecylprodigiosin determinations were as previously described (2).
Plasmid and DNA Manipulations
All Oligonucleotide primers used in this study (Table l) were prepared by the
Macromolecular Structure Facility at Michigan State University (East Lansing, MI,
USA). Streptomyces plasmid preparations and transformations were performed as
described by Hopwood, et al. (1985). Streptomyces chromosomal DNA was isolated
using the method of Pospiech and Neumann (1995). Escherichia coli plasmid
preparations were done by alkaline lysis (147) or using QIAprep spin columns
(QIAGEN). All replicative plasmids shown in Figure 6 were constructed by first cloning
the S. coelicolor absA region of interest into pBluscriptII SK+ (Stratagene) by standard
cloning techniques (147). Inserts flanked by BamHI sites were then subcloned directly
into the Bng site of pl] 702 (73) as in the case of pCB220 and pTBA155. Inserts used to
construct pCB520, pCBS30, pCB540, pTBA156, and pTBA175 were first subcloned into
pU2925 (86) and then excised as BglII fragments from the pIJZ925 polylinker for ligation
into the BgIII site of pl] 702. Replicative ligations were transformed into S. lividans
46
cocoa moan 5:235“ 05 mo 0.28205: “Em ofi an Eamon watonfiac
:88 gm :ouflmqab 05 mo 02820:: “we 05 “a mfimon matching
@5ch cum—Q88 u .3 659. wave
Epsom—mam Ho: n .ad
MQW
o H .m.o
N
.—.5.O.U.
?
.3 3 8 3. ad OUOUEm?
.3. v: 9 N: ohm/Rpm OUUOH?
.3 mam 8 SN Boa Hm N “and UEOO?
.md owl 8 Nat om ~?
.3 NB 9. wwm cm Fara UOOOHOODHO»
.3 oz: 8 3m dd UU‘
33 i E H202L 3f
.9. pTBA156 //
79
a: B D54E P
B B
pCB220 l l
B H202L B
pTBA155 l.___.!___l
49
1326; plasmids recovered from these transformants were then transformed into S.
coelicolor J 1 501 .
Disruption of absAZ in C500
The integrative plasmid pTBASOO (Figure 6) was constructed using an absAZ
fragment amplified by PCR from primers WC8 and WC9, both of which contained BgIII
restriction sites at their ends. The truncated region of AbsA2 encoded by the WC8/W C9
PCR product is illustrated in Figure 7. PCR amplification was carried out in a 100 pl
reaction volume with 100 ng J l 501 chromosomal DNA template, under buffer and
thermal-cycler conditions for absAZ amplification described in Anderson, et al. (1999).
The absAZ amplification product was purified on Wizard PCR preparatory columns
(Promega) prior to and following digestion with BgIII. The resulting 5’- and 3’-truncated
absAZ fragment was cloned as a BglII fragment into the BamHI site of pIJ963 (86) to
produce the integrative plasmid pTBASOO.
Strain C500 possessing a chromosomal disruption in absA2 was created by single
cross-over integration of pTBASOO. pTBASOO was passed through dam', dcm' E. coli
ET12567 prior to transformation into S. coelicolor J 1501. Hygr resistance was used to
select for single-crossover recombinants; these displayed the Pha phenotype. Plasmid
integration was analyzed using Southern hybridization with absA- and hyg-specific
probes.
SO
linker { helix ymmtfheux W
VLLADDETIW-uI/«D-n-l/m-Kn ----NAEIAQRLHLVEGTIKT -----
WC8 WC9
Receiver Domain Effector Domain
Figure 7. Creation of the absAZ disruption in strain C500. An internal region of absAZ
was generated by PCR from primers WC8 and WC9. Primer WC8 recognized the region
of absAZ around the highly conserved Asp13 codon. Primer WC9 annealed to the region
of absA2 encoding the first helix of the helix-tum-helix DNA-binding motif.
Construction of an In-Frame Deletion in absA] in C530
The integrative plasmid pTBA533 (Figure 6) was constructed by first digesting
pCB4OO (pIJ2925 with a 2 kb absA l BamHI fragment) with NaeI to remove the 0.8 kb
fragment internal to abaA] (Figure 8) which created pCB420. The resulting 1.2 kb
BamHI absA] fragment was ligated into the BamHI site of pCB3OO (pSK+ with a 1.8 kb
BamHI/X1101 absAZ-containing region) to produce pCBSOO. The 2.4 kb XhoI/PstI region
of pCBSOO carrying the whole absA locus with the absA 1 in-frame deletion was
subcloned from the pCBSOO polylinker as a SacI/Pstl fragment into pIJ2925 to create
pCBSOl. The entire pCBSOl insert was removed from the polylinker as a BgIII fragment
and ligated into BamHI-digested pIJ963 to produce the integrative plasmid pTBA533.
51
absA1A530
11.321 * IHI INHDIIGD-i 1
Sensor Transmitter C-terminal
Domain Domain Domain
Figure 8. The absA 1 A530 in-frame deletion (diagonal hatch) was created by the removal
of a 0.8kb NaeI region internal to absA] . Horizontally hatched boxes in the sensor
domain represent four transmembrane helices predicted for AbsAl. Lettered boxes of the
transmitter domain symbolize highly conserved sub-domains of two-component histidine
kinase transmitters. The H-box contains Hi3202 which is the putative site of
phosphorylation in AbsAl.
Initial attempts at gene replacement used a strain with a deletion/ermE
replacement in absA (C430), but this strain transformed extremely poorly (19).
Therefore, gene replacements were created in strain J 1501, as follows. Strain C530
possessing a chromosomal absA] in-frame deletion was created through double cross-
over gene replacement with integrative plasmid pTBA533. pTBA533 was passed
through dam', dcm' E. coli DM-l (GIBCO BRL) prior to transformation into S. coelicolor
J 1501. Hygr resistance was used to select for single-crossover recombinants. Plasmid
integration was analyzed using Southern hybridization with the same absA- and hyg-
specific probes described above (see Disruption of absAZ). Single-crossover
52
recombinants were subjected to multiple rounds of propagation and spore isolation on
solid and liquid media without Hyg to allow double-crossover curing of the plasmid.
Single colonies were chosen for making spore preparations and chromosomal extractions.
Initial screening for double crossovers was performed by PCR amplification from
primers WC12/WC13, which were both internal to the in-frame deletion, and
WC16/WC26, which produced a 1 kb product for absA] with the in—frame deletion,
versus a 1.8 kb product for wild type absA 1. If no PCR products corresponding to wild
type absA] were amplified, then Southern hybridization was performed on an XhoI digest
of chromosomal DNAs. Colonies which had successfiilly undergone double cross-over
integration of the absA] allele carrying the in-frame deletion showed no signal for the
hyg probe and a single signal of 2.4 kb for the absA probe. F inal confirmation of the
integrity and fidelity of the C530 absA locus was obtained by sequencing the entirety of
absA] and absAZ. Procedures for the amplification and sequencing of absA] and abs/12
are described elsewhere (6).
In each of the gene replacements described in this study, double-crossover
integration of the mutant allele required propagation of single-crossover transformants for
numerous generations under nonselective conditions. It should be noted that although
single-crossover transformants demonstrated Hyg sensitivity afier only a few generations
of growth in the absence of antibiotic, many of these still possessed the hyg marker as
determined by Southern analysis. Similarly, PCR screening that indicated complete
resolution of the double crossover was frequently contradicted by Southern analysis.
Final confirmation of successful gene replacement required careful analysis by Southern
hybridization and sequencing.
53
Site-Directed Mutagenesis
The absAZ D54E allele was generated using PCR amplification with mutagenic
primers. Separate upstream and downstream absAZ fragments, with an overlapping
region centered at the site of the D54E-encoded mutation, were amplified from pCB46O
(pSK+ carrying a 3.9 kb BamHI/Xhol fragment with the entire absA locus) using primer
pairs WC24/W C29 and WC15/W C28 (Table 1). The resulting GAC to GAG change
also introduced an XhoI site into the PCR products. Thus, the D54E-containing
fi'agments were digested with XhoI and an additional restriction enzyme, the site for
which was present in the upstream or downstream region surrounding absA2: BamHI for
the (WC24/W C29) product upstream of absA 2, -and PM for the downstream
(WCl 5/W C28) product. A three-way ligation between these fi’agments and pSK+
BamHI/Pstl produced pTBA16O containing a 1.2 kb insert with the entire absAZ D54E
allele. Confirmation of the site-directed change was obtained by sequence analysis of the
entire absA2 D54E allele from pTBAl60. Subsequently, a 2 kb BamHI absA] region
was ligated into pTBAl6O BamHI to create pTBA162. The 3.2 kb BamHI/Pstl insert of
pTBA162 containing the entire absA locus was removed as a XbaI/Kpnl fragment from
the pTBA162 polylinker for ligation into pIJZ925 to produce pTBA166. The same 3.2 kb
insert was excised from the pTBA166 polylinker as a Bng fragment and ligated into
pU963 BamHI to generate the integrative plasmid pTBA57O (Figure 6).
Strain C570 possessing a chromosomal absAZ D54E mutation was created
through double-crossover gene replacement with integrative plasmid pTBA5 70.
pTBA570 was demethylated as described above prior to transformation into S. coelicolor
54
J 1501. Hygr resistance was used to select for single-crossover recombinants. Plasmid
integration was analyzed using Southern hybridization with the same absA- and hyg-
specific probes described above (see Disruption of absA2). Single-crossover
recombinants were subjected to multiple rounds of propagation on solid and liquid media
without Hyg to allow double-crossover curing of the plasmid. Single colonies were
chosen for making spore preparations and chromosomal extractions. Initial screening for
plasmid curing was performed by PCR amplification from primers WC30/WC28, which
amplified a 1.2 kb product containing the entire absA2 allele. Only the absA2 D54E
allele was susceptible to digestion with X7101. Therefore, if no PCR products
corresponding to wild type absA2 were amplified, then Southern hybridization was
performed on XhoI digests of chromosomal DNAs. Colonies which had successfully
undergone double cross-over integration of the absA2 D54E allele showed no signal for
the hyg probe and signals of 1.7 and 1.5 kb for the absA probe. Final confirmation of the
integrity and fidelity of the C570 absA locus was obtained by sequencing the entirety of
absA] and absA2 as described (6).
The D54A and D54N alleles of absA2 were created by PCR-mediated
introduction of these mutations using the QuickChange Site-Directed Mutagenesis Kit
(Stratagene). The D54A mutation was introduced into absA2 through a single nucleotide
change (GAC to GCC) with complimentary primers WC65 and WC66. Likewise,
complimentary primers WC67 and WC68 produced a single mismatch (GAC to AAC) in
absA2 to generate the D54N mutation. The mutagenesis reactions were carried out on 50
ng of pTBA400 (pSK+ carrying a 1.8 kb BamHI/X1101 absA 2-containing fragment) under
the manufacturer-prescribed buffer conditions with the addition of 5% glycerol and 2.5%
55
dimethyl sulfoxide (DMSO). The thermal cycler conditions were 95°C for 5 min,
followed by 12 cycles of 95°C for l min, 65°C for 45 sec, and 72°C for 12 minutes. The
PCR products were digested with DpnI and transformed into E. coli DHSa (GIBCO
BRL) to create pTBA410 (containing absA2 [D54A]) and pTBA430 (containing absA2
[D54N]). Both the D54A and D54N mutations removed a T an restriction site at the
mutagenized codon. Therefore, to screen for successful incorporation of site-directed
mutations, plasmid preparations from transformant colonies were used as template in
PCR reactions with primers WC35 and P11, which amplified a 360 nt region internal to
absA2. Amplification products were digested with Tan and analyzed on 1.2% agarose
gel. In each case, over 90% of the transformants tested screened positive for the
mutation. Confirmation of the desired mutations was obtained by sequencing the absA]
and absA 2 portions of the mutagenized plasmids, pTBA410 and pTBA430. Integrative
plasmid pTBA532 (Figure 6) encoding the absA2 D54N mutation was made by
subcloning the 1.8 kb BamHI/Kpnl fragment from the polylinker of pTBA430 into
pU963. In order to construct integrative plasmid pTBA516 (Figure 6), the pTBA410
insert was increased in size to 3.8 kb by cloning in a 2 kb BamHI fragment containing the
upstream portion of absA] to produce pTBA516. The 3.2 kb XhoI fragment of pTBA516
(containing absA2 [D54A]) was then cloned into the SaII site of pIJ2927 (86) to create
pTBA41‘4. This same 3.2 kb insert was then removed as a 3.2 kb BgIII fragment from the
polylinker of pTBA414 and cloned into pIJ963 BamHI to create the integrative plasmid
pTBA5 16.
Strains C516 and C532, possessing chromosomal mutations absA2 [D54A] and
absAZ [D54N], respectively, were created through double-crossover gene replacements
56
with integrative plasmids pTBA516 and pTBA532, respectively. Single- and double-
crossover integration procedures were the same as described for C570 gene replacement.
Initial screening for plasmid curing was performed by PCR amplification of T an
restriction digest analysis of the WC3 5/P11 absA2 products. Chromosomal DNA from
strains that did not amplify wild type absA2 alleles were digested with XhoI and analyzed
by Southern hybridization with hyg and absA probes as described above. Colonies which
had successfirlly undergone gene replacement with the absA2 (D54A) and (D54N) alleles
showed no signal for the hyg probe and signals of 3.2 kb for the absA probe. Final
confirmation of the integrity and fidelity of the C516 and C532 absA loci was obtained
by sequencing the entirety of absA] and absA 2 using methods described previously (6).
The absA] H202L allele was generated using PCR overlap extension (171).
Separate upstream and downstream absA 1 fragments, with an overlapping region
centered at the site of the H202L mutation, were amplified from pCB460 (pSK+ carrying
a 3.9 kb BamHI/Xhol fragment with the entire absA locus) using primer pairs
WCl6/WC30 and WC14/WC26 (Table 1). Then, the full-length absA] H202L allele was
amplified by combining 5 to 10 ng of the upstream and downstream PCR products as
template together with primers WC 16 and WC26. PCR amplifications were carried out
in 50 pl reactions using high-fidelity Pfu DNA polymerase (Stratagene). Bufi‘er
conditions and thermal cycler settings were as previously described (6) except that the
extension time was increased from 1 to 4 minutes. The expected size product was
agarose-gel purified and digested with XhoI and BamHI to produce a 1.45 kb fragment
that was ligated into pSK+ to create pTBA140. Confirmation of the site-directed change
was obtained by sequence analysis of the absA I XhoI/BamHI fragment from pTBA140.
57
Subsequently, a 5 kb BamHI fragment (containing the 3' region of absA] and all of
absA2) was ligated into pTBA14O BamHI to produce pTBAl42. The 3.2 kb XhoI region
(containing most of the absA] H202L allele and all of absA 2) was excised from
pTBAl42 and ligated into XhoI-digested pCB3 60 (pIJZ925 containing a 1.6 kb SacI/Xhol
insert with the 5‘ region of absA] and upstream SCE8. 17c) to create pTBA144. The
entire 5.4 kb insert was removed from pTBA144 as a BgIII fragment for ligation into
pIJ963 BamHI to create the integrative plasmid pTBAISO (Figure 6).
Strain C550 possessing a chromosomal absA 1 H202L mutation was created by
gene replacement with pTBAISO. Single- and double-crossover integration procedures
were the same as described for the C5 70 gene replacement. Initial screening for plasmid
curing took advantage of a T an restriction site introduced by the H202L mutation. An
internal region of absA] was amplified from chromosomal DNA of putative C550 strains
using primers WC12 and WC13, followed by T an digest analysis. Chromosomal DNA
from strains that did not amplify wild type absA] were digested with X7101 and analyzed
by Southern hybridization with hyg and absA probes as described above. Colonies which
had successfiilly undergone gene replacement with the absA] H202L showed no signal
for the hyg probe and a signal of 3.2 kb for the absA probe. Final confirmation of the
integrity and fidelity of the C550 absA locus was obtained by sequencing all of absA]
and absA2 using methods described previously (6).
RNA Isolation
Streptomyces RNA isolation was carried out as described by Hopwood, et al.
(1985) using the preparation method for dot-blotting and northern blotting . Two
58
independent isolations at 18, 30 and 54 hours of growth were performed for S. coelicolor
strains J 1501 (hisAI uraAI strAl SCPl' SCP2" Pgl'), C542 (absA1-542 (6)), and C570
(absA2 [D54E] hisAI uraAI strAI SCPI" SCPZ-Pgl'). Four 50 ml cultures were pooled
for each 18 hour RNA preparation, whereas two 50 ml cultures were pooled for 30 and
54 hr samples. The concentration, purity and integrity of the RNA samples was
evaluated by spectrophotometry and agarose gel electrophoresis. Isolation of E. coli
RNA, for use as a negative control, was performed with an RN Aeasy RNA-purification
column (QIAGEN).
Sl Nuclease Protection Assays
All experiments were performed using 50 pg of RNA and 60,000-100,000 cpm of
32P-end-labeled double-stranded DNA probe. The absA transcript time-course analysis
incorporated an absA] probe together with a glk (glucose kinase (7)) probe - which
served as an internal standard for normalizing the quantity of RNA in each assay; (1)
The 455 bp absA] probe was generated by PCR using the WC64 forward primer and the
5’-32P-end-labeled WC20 reverse primer. The template for absA! probe synthesis was
pCB400, containing the 2 kb BamHI region of the absA locus cloned into pU2925. A
309 bp 32P-end-labeled glk probe was also generated by PCR from the primers and
template described by Aceti and Champness (1998). Primers (50 pmoles) were end-
labeled by the T4 polynucleotide kinase (Promega) forward reaction as described by the
manufacturer with minor modifications. Prior to initiating the end-labeling reaction, each
primer was incubated with spermidine (10 mM final concentration) at 70°C for 10
minutes. Likewise, ethanol precipitation of the labeled oligos was facilitated by the
59
addition of 2 ug of glycogen. The labeled oligo was divided between duplicate 50 ul
PCR reactions. The reaction mix contained 20 pmoles of each primer, 100 ng template,
0.2 mM dNTPs, 1.5 mM MgC12, 5% glycerol, 2.5% DMSO, 1% formamide and 1.25 U
T aq polymerase (Perkin Elmer). Thennal-cycler conditions were 95°C for 5 min,
followed by 30 cycles of 95°C for 1 min, 65°C for 45 sec, and 72°C for 1 min, and a final
extension at 72°C for 10 minutes. The S] nuclease protection assay was performed as
previously described (1). Replicate RNA isolates were tested in independent 81
experiments. Time-course experiments included E. coli RNA as a negative control. In
addition, the presence of excess probe was verified by treating an RNA sample with 2-
fold concentrations of each probe and comparing their signals to the same sample treated
with normal levels of probe. Results were analyzed by electrophoresis on 6%
polyacrylamide sequencing gels (147) and autoradiography. Transcript sizes were
estimated by running 32P-end-labeled ¢X174lHian molecular weight markers (Promega)
on the same gel. To map the absA] transcription start site, a sequencing ladder was
generated from primer WC20 using the fmoiD DNA Sequencing System (Promega) and
compared to Sl-treated 18 hr C542 RNA hybridized to the absA] probe.
The region upstream of the absA 2 translation start site was examined for promoter
activity using a 504 nt probe. The absA2 probe protected the region from 330 nt
upstream of the absA2 translational start to 174 nt downstream into the coding region.
The absA2 probe was generated by PCR using forward primer WC24 and 5'-32P-labeled
WC29 reverse primer. The PCR template was pCB46O (pSK+ carrying a 3.9 kb
BamHI/Xhol insert with the entire absA locus). Primer end-labeling, PCR reaction
60
conditions, and probe purification were performed as described for the absA] probe. The
absA2 probe was used in SI nuclease protection assays with 18 and 30 hr C542 RNA.
RESULTS
Negative Regulation of Antibiotics by the absA2-Encoded Response Regulator and
absA] Histidine Kinase
In previous genetic studies of the absA locus, certain mutations that disrupted the
locus suggested that the absA-encoded two-component system fimctioned as a negative
regulator of antibiotics (19). These mutations caused a visible phenotype of early,
enhanced production of the actinorhodin and undecylprodigiosin antibiotics; we refer to
this phenotype as Pha, for precocious hyperproduction of antibiotics. The Pha mutant
alleles were created by insertions into the absA! gene, which is upstream in a putative
absA I-absAZ operon. The phenotype in these mutants may have resulted from
disruption of absA] or from polar effects on expression of absA 2. To distinguish
between these possibilities, we directly tested the fiinction of absA2 by specifically
disrupting the absA2 gene. A fragment internal to absA2 (Figure 6) was cloned into the
nonreplicating plasmid pIJ963 to create pTBASOO, which was then integrated into the
absA locus of strain J 1501. The resulting strain, C500, was absA 1+ absA2::pTBA500,
with absA2 truncated upstream of the predicted helix-tum-helix domain (Figure 7).
Disruption of absA2 in C500 caused a Pha phenotype (Figure 9), thereby demonstrating
the involvement of absA2 in negative regulation of antibiotic production. Both repressor
and activator functions are well-documented for two-component response regulators;
61
functional regions of the transmitter domain of histidine kinases, including the conserved
H, N, D, and G boxes. The phenotype of the absA 121530, strain C530, was Pha (Figure
9) and, moreover, was identical to that of C500. This result implicated the protein kinase
activity of AbsA] in the negative regulation effected by AbsA2.
J1501
C530
Figure 9. The effect of an absA gene disruption and gene replacements on antibiotic
production. Strains were grown for 4 days on SpMR agar. Strains are S. coelicolor
J 1501 (wild type), C550 (absAI [H202L]), C570 (absA2 [D54E]), C500
(absA2zszBA500), C532 (absA2 [D54N]), and C530 (absA I A530). Actinorhodin and
undecylprodigiosin pigments were assessed as described in Materials and Methods.
Images in this dissertation are presented in color.
62
Genetic Evaluation of the Role of Phosphorylation in AbsA2-Mediated Regulation
For most response regulators, phosphorylation of a conserved aspartate residue is
essential for the regulatory functions of the proteins in viva. Following this precedent,
the AbsA2 regulatory activity would likely require that AbsA2 be phosphorylated;
AbsAl would likely be responsible for AbsA2 phosphorylation. The Pha phenotypes of
C500 and C530 would be consistent with this. scenario, but it was important to consider
the additional factor that many of the characterized two-component system histidine
kinases are bifunctional enzymes that possess both kinase and phosphatase activities; the
phosphatase activity dephosphorylates the phosphorylated response regulator. In the case
of AbsAl, the in-frame deletion in C530 would remove AbsA2-specific phosphatase
activity, as well as the kinase activity associated with the transmitter domain. Thus, in
strain C530, phospho-AbsA2 may be present if AbsA2 can be phosphorylated by an
alternative kinase or low molecular weight phosphate donors, and the C530 phenotype
might be caused by a lack of the AbsA] phosphatase and a resulting overabundance of
phospho-AbsAZ. In this case the negatively regulating form of AbsA2 would be the
unphosphorylated form.
In order to distinguish whether phospho-AbsA2 or unphosphorylated AbsA2
functions as the negative regulator, we constructed several mutants with site-directed
changes to the chromosomal absA2 gene (Figure 6), altering the AbsA2 aspartate residue
(D54) that is analogous to the conserved phosphorylated aspartate of response regulators.
Separate gene replacements created strains C570, C516, and C532 with AbsA2 amino
acid replacements, D54E, D54A, and D54N respectively. These aspartate substitutions
have been shown to prevent phosphorylation of numerous response regulators (88; 42;
63
22). All three mutant strains exhibited the Pha phenotype (Figure 9; C516 not shown).
Thus, these results supported the hypothesis that phospho-AbsAZ functions as the
negative regulator.
The histidine residue in AbsAl that corresponds to the site of phosphorylation in
well-characterized members of the histidine kinase family is His 202 (Figure 6). A site-
directed mutation, H202L, was made in the chromosomal absA! gene of 11501, creating
strain C550. Strain C550 exhibited a Pha phenotype, a result consistent with a
requirement for histidine kinase activity in negative regulation. However, the phenotype
differed fi'om that of C530 (AabsA 1) in several respects. First, C550 visibly produced
undecylprodigiosin earlier than actinorhodin, whereas C530 produced both antibiotics ‘
precociously. Second, hyperproduction of antibiotics, relative to strain 11501, never
reached the levels seen for C530 (Figure 9). The reason for the weaker Pha phenotype of
C550 is not clear at this time. We considered the possibility that AbsAl (HZOZL)
contained a second site of phosphorylation; however a fiision protein containing the
AbsAl (HZOZL) transmitter domain fused to maltose-binding protein did not demonstrate
autokinase activity in vitro, whereas an MBP:: AbsAl+ fusion did (Chapter 5).
Precocious Hyperproduction of Calcium-Dependent Antibiotic, Undecylprodigiosin
and Actinorhodin in absA Mutants
We sought to determine whether Pha mutations affected synthesis of calcium
dependent antibiotic, in addition to actinorhodin and undecylprodigiosin. To assess CDA
activity, plugs from plate grown cultures were tested for anti-Staphylococcus aureus
activity. In the presence of added calcium, the lipopeptide CDA is active, damaging cell
64
de:
act
membranes (100). For CDA assays, culture plugs were tested on plates with and without
added calcium. In a 2-day time course, shown in Figure 10, CDA activity was detected
in Pha mutants at least 7 hrs earlier than in J 1501. Pha mutants C550 and C530 are
shown. Similar results were obtained with C550, C530 and C570 on R5 media (data not
shown). These results showed that AbsA negatively regulates CDA as well as
actinorhodin and undecylprodigiosin.
41hr f5
Figure 10. Calcium-dependent antibiotic assays in Pha mutants. Growth conditions are
described in Methods. Plugs were taken from ONA plates at the times indicated. CDA
activity is detected in the presence of calcium, right.
65
Over the course of cultivating Pha mutants we have observed variability in how
much earlier a given Pha mutant produces antibiotics compared to J1501. The amount of
acceleration has ranged from at least 7 hours to several days on different media, e.g. R5,
SpMR and ONA In addition, in quantitative assays of actinorhodin and
undecylprodigiosin, the Pha-related overproduction has varied from 5-fold to more than
60-fold (data not shown). An exploration of this phenomenon will be reported in more
detail elsewhere. Besides the effect on antibiotics, the Pha phenotype includes a defect in
morphology. Pha mutants produce aerial hyphae relatively sparsely and their colony
surfaces are notably crenulated.
Precocious Hyperproduction of Antibiotics Resulting from AbsA Domain
Overexpression
In some cases, overexpression of an unphosphorylated response regulator can
mimic the regulation of target promoters that normally is effected by a phosphorylated
response regulator (e. g., 180). To evaluate whether overexpression of unphosphorylated
AbsA2 could regulate antibiotics, we introduced a high copy clone of the absA2 (D54E)
mutant allele (pTBA175; Figure 6) into J 1501 and C577SZS, a strain deleted for absA2
and most of absA] (6). The pTBAl75 plasmid included absA1+ and the absA promoter
region. If unphosphorylated AbsA2 could negatively regulate antibiotics we might have
observed a delay of antibiotics in the Pha C577825 strain. However, we observed no
change in the Pha phenotype (data not shown) suggesting that phosphorylation of AbsA2
is required even at high protein abundance, for negative regulation.
66
When pTBA175 was introduced into J1501, the resulting phenotype was Pha,
indicating an interference with normal AbsA-mediated regulation (Figure 11B). To
further examine the phenomenon, we evaluated the effects of overexpressing selected
domains of the AbsAl and AbsA2 proteins (Figure 6). First, we excluded an effect of the
absA promoter region by introducing plasmid pCB540; this plasmid did not alter the
J 1501 phenotype (Figure 11A). Second, we observed that multiple copies of the entire
absA locus, in pCB520, produced no change in the Abs+ phenotype (Figure 11A). Next,
we evaluated a set of high-copy plasmids that expressed the wild-type AbsA2 but carried
phosphorylation-minus absA] alleles; these included pCBS30, carrying the in-frame
deletion of absA 121530, and pTBA156, carrying the absA! [HZOZL] allele. These
produced no change in the Abs+ phenotype (Figure 11). In contrast, a Pha phenotype
resulted from plasmids that lacked absA2+, but contained absA] sequences. Two such
plasmids were pCB220 and pTBA155. A pattern that emerged from these results was
that an increase in gene dosage of absA 2", with or without an increase in abs/11+, did not
alter antibiotic regulation. However, an increase in absA 1 sequences without an increase
in absAZ+ deregulated antibiotics. One interpretation of these results is that a high absA]
gene dosage causes a shift in the ratio of AbsA] kinase to phosphatase activity to favor
the phosphatase activity, and relatively low expression of AbsA2+ may not allow
sufficient AbsA2-P accumulation to down-regulate antibiotics.
What would cause the ratio of AbsAl phosphatase to kinase activity to be higher
than normal in these strains? In the cases of the pCB220 and pTBA155 plasmids, the C-
terminal 69 aa of absA] are truncated. In previous complementation analyses, the
67
115011
pABSl7S
11501/ 11501/
pCBS30 pABSlS6
Figure 11. The effects of high-copy expression of absA alleles on antibiotic production.
All plasmids were derivatives of pIJ702 expressed in a S. coelicolor J 1501 background.
Plasmid inserts are shown in Figure 6. Strains were grown for 3 or 4 days on SpMR agar.
Early onset of both Red and Act are characteristics of the Pha phenotype. Frame A
demonstrates the early production of Red in the Pha phenotype, whereas frame B shows
the early onset of Act synthesis.
pCB220 absA] allele was capable of restoring a wild-type phenotype to absAI', but it is
possible that this allele and the H202L version in pTBA155 have higher than normal
phosphatase activities. However, absAI is wild-type on pTBA175. One speculation
about the observed effects might be that the AbsAl-kinase activity is normally activated
by the binding of a low-abundance signal molecule and the AbsAl polypeptides
produced by the high-copy constructs are present in quantities sufficient to titrate the
ligand; thus the population of AbsAl molecules may be predominantly in the
phosphatase mode.
68
High Resolution 81 Nuclease Mapping of the absA Transcription Start Site
The absA] and absA2 ORFs are separated by only 17 at and are likely
cotranscribed. To define the transcription start site for absA, high resolution 81 nuclease
mapping was performed. First, a PCR-generated double-stranded DNA probe specific to
the predicted promoter region for absA] (Figure 12B) was used to map the transcription
start site of absA] by analyzing the 81 product alongside a sequencing ladder generated
from the same 32P-labeled primer used to synthesize the probe. The absA] probe
protected a single product of 291 nt (Figure 12A), identifying the transcription start site
for absA] as the adenosine nucleotide that is also predicted to be the putative translation
start site. To evaluate cotranscription of absA 1 and absA 2, the region upstream of absA2
was probed with a 504 bp double-stranded DNA probe (Materials and Methods). The 81
product showed no indication of independent promoter activity for absA2 (data not
shown). These results indicated that absA] and absA2 are transcribed as a single,
leaderless transcript. Leaderless transcripts are not uncommon in actinomycetes, as
documented by Strohl (1992), who reported that 11 of 139 promoters analyzed produced
leaderless transcripts.
Inspection of the sequence upstream of absA! revealed a —10 region with the
sequence TAGCGT (Figure 12); this is similar to the consensus sequence proposed by
Strohl (1992) for transcription from Streptomyces RNA polymerase that contains an E.
coli-like Eo7° sigma factor, e.g. HrdB or HrdD (23). There was, however, no
recognizable consensus sequence in the —35 region.
69
A T
- T
G
T*
Air
G*
c* -10region
G*
T*
G
C
T
G
G
A
A‘M
T
G
C H
A
C
C R
G
A
B.
455m DNA probe
<_136%6291nt
«ac
BamHI XhoI BamHI
Figure 12. S] nuclease protection mapping of the absA locus. (A.) High resolution S1
nuclease protection mapping on total RNA isolated from an 18 hour culture of S.
coelicolor C542 grown in SpMR liquid medium. The AGCT sequencing ladder was
generated from 5'-labeled Oligonucleotide WC20 (see Materials and Methods section).
The transcription start site (a) and the hexameric -10 promoter region (*) are shown. (B.)
The absA probe was a 455 bp PCR product amplified fi'om primer WC64 (Table 1) and
primer WC20 - uniquely labeled with 32P at the 5' end. The shaded areas represent
coding regions of absA] and SCE8. 1 7c contained on pCB400. The absA probe extends
291 nt downstream of the putative translation start site and 136 nt upstream.
7O
Growth-Phase Dependent Expression and Autoregulation of absA
To evaluate the temporal profile of absA expression, RNA was isolated over a 54
hr time course fiom cultures grown in liquid media. The media used, SpMR (see
Materials and Methods), supported production of the actinorhodin and
undecylprodigiosin antibiotics by strain J1501. Streptomyces coelicolor does not
sporulate when grown in liquid cultures, so temporal comparisons of antibiotic
production and sporulation could not be made in this experiment. However, as is
generally observed, the antibiotics showed growth-phase dependent production kinetics,
appearing only after a period of biomass accumulation.
Sl nuclease protection assays were performed on the absA promoter region from
RNA isolated from J 1501 cultures grown for 18 hours, 30 hours, and 54 hours.
Antibiotics were not produced in the 18 hour culture but were visible in the 30 hr J] 501
culture. Figure 13 shows that the absA transcript was present in the 18 hr culture, and it
increased significantly in abundance from 18 hours to 30 hours. The transcript then
remained at a constant level through 54 hours. Comparisons of transcript abundance in
different cultures were aided by the addition of a probe for the glk (glucose kinase) gene
to each S1 assay. The absA signal increased about five-fold relative to the glk signal over
the course of culture growth.
Figure 13 also includes 81 nuclease protection assays of RNAs isolated from two
strains that are mutant for the absA locus. One strain was C542, an Abs“ strain mutant in
absA] (i.e., an absA1* strain), as described above. The second was C570, the Pha strain
carrying the D54E mutation in absA2 that was described in Figure 6. The profile of absA
expression was altered in both absA mutants. In C570, the absA transcript abundance
71
11501 (Abs‘)
C542 (Abs')
0170 (Pha)
11501 (Absl)
€170 (Pha)
11501 (Abs+)
c542 (Abs')
c170 (Pha)
MW Marker (nt)
;
a‘
‘3
E
'3
:3
N
v
V)
O
30hr
«absA
Figure 13. High resolution 81 nuclease protection analysis of the absA transcript, using
RNA isolated fi'om 18, 30 and 54 hour S. coelicolor cultures in SpMR liquid medium. S.
coelicolor strains are 11501 (absA +), C542 (absA 1-542, (6)), and C570 (absA2 [D54E]).
The absA probe was the 455 bp probe in Figure 12. Glucose kinase (glk) was measured
to normalize the amount of RNA assayed at each time point (1).
72
was very low at all time points. In contrast, in C542, the absA transcript was several-fold
more abundant than in J1501, at all time points.
The 81 protection assays revealed several aspects of absA regulation. First, the
effects of the absA mutations indicated that absA expression is autoregulated. Second,
the mutant effects on the absA transcript were opposite to the previously-observed effects
on antibiotic transcripts: whereas the Abs' and Pha phenotypes were found to correlate
with decreased or increased antibiotic gene transcription, respectively, the absA transcript
was decreased in the Pha strain but increased in the Abs‘ strain. These results suggest
that autoregulation by absA is positive, in contrast to absA negative regulation of
antibiotics. Third, the low level of absA transcript in C570, the absA2 (D54E) mutant,
suggests that phospho-AbsAZ is the autoregulatory form of AbsA2, which is consistent
with data from the genetic analysis that implicates phospho-AbsA2 as the antibiotic-
regulatory form. Finally, the absence of any growth phase-related increase of absA
transcript in C570 suggests that phospho-AbsAZ was responsible for the growth-phase
regulation observed in J 1501 and C542. Thus, the growth phase regulation of absA
appears to result from phosphorylation-dependent AbsA2-mediated autoregulation. We
have not determined at this time whether the absA autoregulation is direct or indirect.
DISCUSSION
In this paper we have described a genetic and transcriptional analysis of the absA
locus which firrther characterizes aspects of the mechanism of AbsA-mediated regulation
of antibiotic production. Disruptions in the absA] and absA2 genes demonstrated that the
AbsA two-component system is a negative regulator of the multiple antibiotics produced
73
by S. coelicolor, including calcium-dependent antibiotic, actinorhodin and
undecylprodigiosin. In addition, gene replacements in the absA locus altered the putative
sites of phosphorylation of AbsAl or AbsA2. As predicted from sequence conservation
with other two-component systems, both the His at position 202 of AbsAl and the Asp
residue at position 54 of AbsA2 were required for normal regulation of antibiotic
synthesis: each of the gene replacement strains tested attained an antibiotic
overproducing phenotype (Pha) consistent with a mechanism in which the
phosphorylated form of AbsA2 is the active negative regulator of antibiotic synthesis.
Our results did not, however, distinguish whether or not AbsA2~P is a direct repressor of
the antibiotic genes or whether it is the activator of a repressor.
Without absA regulation, the timing of antibiotic production is advanced but, even
in Pha cultures, a period of approximately 2 days passes before antibiotics appear. One
interpretation of this observation is that the appearance of antibiotics in Pha cultures
indicates the time at which the culture enters an antibiotic production-competent state,
but the AbsA system normally imposes a delay on production. The heterogeneities in a
growing mycelia] biomass complicate distinctions of growth phases, but for the purposes
of firrther discussion, we refer to the postulated “AbsA-repressed” period as the
“transition stage.” We can envision several models for how AbsA, as a signal
transduction system, could modulate production of antibiotics during culture growth.
One model, which accommodates both the genetic and transcriptional data,
supposes the following. Early in growth, a culture is not competent for antibiotic
synthesis; also, the absA genes are expressed at a low level. Following a period of
growth, the culture enters the “transition stage.” During this time, the signal that
74
regulates AbsA may be present at significant levels. If AbsAl is like many sensor
kinase/phosphatases, it will require signal binding to activate the kinase fiinction, and
exist in a phosphatase-dominant mode if signal is absent (131; 69; 186). Once the signal
is present, and AbsAl is shifted to a kinase-dominant form, AbsA2-P will accumulate
and negatively regulate antibiotics and also positively autoregulate, accounting for the
AbsA2~P-dependent, growth-phase-related increase of absA transcript seen in 11501.
Easing of AbsA-repression may require that the signal be depleted or degraded, allowing
AbsAl to switch to the phosphatase form and dephosphorylate AbsA2, allowing
antibiotic gene expression. At present, we have no information regarding the nature of
the signal hypothesized to regulate AbsAl.
Ifthe normal fimction of the AbsA system is in negative regulation of antibiotics,
what explains the Abs” phenotype in the mutants that first defined the absA locus? We
hypothesize that these absA 1‘ alleles lock the AbsA system into the negatively-regulating
mode, i.e., in which AbsA2 is phosphorylated. In support of this notion, the Abs"
phenotype requires absA 2+ (6). The mutant AbsAl' proteins might be constitutively
kinase-dominant forms, either lacking phosphatase capability or functioning as signal-
independent kinases. The latter possibility would be most consistent with the increased
level of absA transcript observed in C542, e.g. AbsA2-P would be present even in young
cultures lacking signal and would autoregulate. Another observation that could be
explained by signal-independent AbsAl kinase activity is the persistence of the Abs'
phenotype over the life of mutant cultures: even colonies that grow for several weeks
remain unpigmented.
75
An alternative model for signal regulation in the AbsA system is that AbsAl is a
kinase in the absence of signal and a phosphatase in the presence of signal, as a few
sensor kinase-phosphatases are proposed to fiinction (135; 139; 138; 56). In this case, the
transition stage culture lacks the signal regulating AbsAl and AbsAl -kinase activity
would generate AbsA2-P. Later a signal would switch AbsA] to the AbsAl-phosphatase
mode so it could dephosphorylate AbsA2-P, allowing antibiotic synthesis. We consider
this model to be less compelling than the first because the AbsA2-P-dependent
transcription profiles are more simply explained if the AbsAl -kinase activity is activated
by a transition stage signal.
What purpose does AbsA regulation of antibiotics serve in the S. coelicolor life
cycle? One relevant observation is the substantial perturbation of morphogenesis
observed in most Pha mutants: these mutants usually produce only sparse aerial hyphae.
Conversely, antibiotic production is not altered in Abs' strains. One possibility is that
precocious antibiotic synthesis per se is deleterious to normal sporulation. Calcium-
dependent antibiotic may be especially inhibitory as suggested by strain C577825 (6;
Champness, unpublished). This strain demonstrates a strong Pha phenotype for
pigmented antibiotic production, but is blocked in synthesis of CDA due to a deletion in
this gene cluster. In contrast to other Pha mutants, C5 77825 is wildtype for sporulation.
Thus, it may be that S. coelicolor acquires competence for antibiotic production before
the sporulation process has proceeded adequately and the function of the AbsA system is
to delay antibiotic production to allow optimal sporulation.
What factors establish the state that we have referred to as "antibiotic-production
competent"? Likely candidates include the genes that have been identified on the basis of
76
mutant defects in antibiotic synthesis. Among these is a second gene found in Abs’
mutant hunts, absB, which encodes the S. coelicolor homolog of RNase H1 (136).
Another large group of genes is known to regulate both antibiotic synthesis and the onset
of sporulation. Some genes in this group are the bid genes, several of which encode
regulators of gene expression (reviewed by 30). Another is reIA, which encodes pppGpp
synthetase (27; 26; 113; 114). Also important are the components of y-butyrolactone
signaling pathways (188; 130).
Additional antibiotic regulatory genes have been isolated on the basis of multi-
copy stimulation of antibiotic production. The best characterized of these are the
AfsQl/Q2 two-component system (84) and the Ast/K serine-threonine
phosphoprotein/kinase pair (76; 77; 71; 115; 116; 54). Mutations in the afsQl/QZ genes
cause no phenotype, but disruptions to the cy'sR/K locus conditionally reduce antibiotic
synthesis, especially on high phosphate media (115; 54). The multicopy effect of qst/K
has been shown to correlate with increased antibiotic pathway-specific activator
transcription (54). Multi-copy clones of the ast/K locus can restore antibiotic synthesis
to Abs' absA]. mutants (28), and overexpression of the AfsQ response regulator does the
same (84). These observations imply that these genes can compete against the postulated
persistent negative regulation imposed by absA 1‘ alleles.
It is widely observed that phosphorylation of response regulators modifies their
activities, likely causing conformational changes that affect promoter recognition or
cooperative binding at the target promoter (161). However, the extent to which
phosphorylation is required for DNA binding and transcriptional regulation in viva varies
for different response regulators. In the Nar system of E. coli (reviewed by 159), which
77
regulates nitrate/nitrite-responsive anaerobic respiratory pathways, phosphorylation of
NarL is absolutely required for DNA binding and regulatory activity (11). Conversely, in
E. coli UhpA-mediated regulation of sugar-phosphate uptake, high-copy expression of
the unphosphorylated UhpA D54N protein allowed phosphorylation-independent
activation of the uhpT promoter (180; 181). If unphosphorylated AbsA2 was fiinctional
in viva, high-copy expression of the absA2 D54E allele, on plasmid pTBA175 (Figure 6),
might have repressed antibiotic synthesis. Since it did not (Figure 11B), it appears that
AbsA2 regulatory activity is strongly dependent on phosphorylation.
It is noteworthy to contrast AbsA2 with several other recently discovered
antibiotic cluster-linked regulators that are closely related in sequence. One such protein
is RedZ, a red-cluster-linked activator of redD transcription (182). The amino acid
sequence of RedZ has end-to-end similarity to AbsA2, including the putative helix-tum-
helix region, with 27% identical residues overall. However RedZ lacks the conserved
aspartate residue that is normally the site of phosphorylation in response regulators (65).
A homolog of RedZ, DnrN, is found in the daunorubicin gene cluster of S. peucetius,
where it regulates dan, which encodes a SARP pathway-specific regulator of the dnr
cluster. Although the DnrN protein sequence has retained the conserved aspartate, other
residues of the phosphorylation pocket are not conserved and phosphorylation appears to
not be involved in DnrN function in viva (59). Thus DnrN and RedZ appear to serve as
regulators in the unphosphorylated state. It is not known if a modification other than
phosphorylation regulates the activity of RedZ or DnrN.
Our results have shown that the AbsA two-component system is a negative
regulator of the calcium-dependent antibiotic, actinorhodin, and undecylprodigiosin. For
78
the later two antibiotics, absA negatively regulates the SARP pathway-specific activator
genes. It will be important to determine whether AbsA2 regulates the SARP genes
directly. For the cda cluster, interesting questions are whether AbsA2 regulates the
SARP homolog, cdaR, or whether it regulates another, as yet unidentified cda regulator,
or directly represses cda biosynthetic genes.
79
CHAPTER 4
TEMPORAL EXPRESSION OF red AND cda PATHWAY-SPECIFIC
REGULATORS AND THEIR DEPENDENCE ON AbsA
8O
INTRODUCTION
Global negative regulation of antibiotics by AbsA has been demonstrated to act
through transcriptional control of redD and actII-4 pathway-specific regulators (PSRs)
(1). While these genes constitute possible targets of AbsA2, the red gene cluster contains
a second PSR redZ, which is required for expression of redD, but can not activate red
biosynthetic gene expression in the absence of redD (182). The effect of AbsA on redZ
expression is unknown. Another plausible target of AbsA2 is the putative PSR of the cda
gene cluster, cdaR, the sequence of which was recently made available by the S.
coelicolor genome sequencing project (http://www.sanger.ac.uk/Projects/S_coelicolor/).
Very little is known about the temporal expression of the cda gene cluster of S.
coelicolor. The deduced cdaR gene product shows sequence similarity to the N—terminal
region of Streptomyces antibiotic regulatory proteins (146), including redD and actII-4,
that contain an OmpR-like DNA-binding motif (183). A chromosomal disruption of
cdaR was found to eliminate CDA production, suggesting it had a positive regulatory
function (156). Sequencing of the cda cluster also revealed that the absA locus lies
within it (Figure 14). The significance of absA association with the cda genes is not yet
known. The only other Streptomyces antibiotic gene cluster known to encode a two-
component signal transduction system is the rapamycin gene cluster of S. hygroscopicus,
which contains genes encoding a putative two-component system of as yet unknown
function (149). However, various Gram-positive bacteria which produce class I and class
II antimicrobial peptide (AMP) antibiotics do contain two-component genes as part of
the antibiotic gene cluster
81
17.5 kb 40 51 58.3
Figure 14. Position of absA with respect to the cda gene cluster. This 58.3 kb region of
the cda cluster was reconstructed from sequence data made available by the Streptomyces
coelicolor Sequencing Project (The Sanger Centre). Genes shown in white have been
named and given putative fimctions based on genetic or functional analysis. cdaR is
homologous to pathway-specific activators. Biosynthetic genes cdaPSI, cdaPSII and
cdaPSIII encode peptide synthases which catalyze steps in the enzymatic synthesis of the
lipopeptide antibiotic CDA. Shaded genes have been assigned putative functions based
on sequence similarity to other proteins (annotated in
http://www.sanger.ac.uk/Projects/S_coelicolor/).
82
(reviewed by 123; 96). AMPS have a conserved regulatory gene organization in which an
open-reading frame encoding an autoinducing peptide pheromone precedes the genes of a
two-component system. While demonstration of response regulator binding to operons
within AMP clusters is still pending, disruption of either the autoinducer or two-
component genes abolishes AMP production (47; 10; ‘98). It is hypothesized that two-
component systems regulate most or all of the regulatory, biasynthetic and
immunity/transport operons of AMP clusters (123; 96). Although the precise target(s) of
AMP two-component systems remains to be elucidated, unlike AbsA, these systems have
not been implicated in regulation of genes outside of the cluster in which they are found.
Given the recent revelation of the association of absA with the cda cluster, and
dependence of other PSRs on AbsA, it was of interest to determine whether inhibition of
CDA synthesis in the Abs' mutant was correlated with transcriptional regulation of cdaR.
Similarly, I examined the effect of AbsA mutants on the expression of refl and the
possibility that AbsA-mediated regulation of redD was a consequence of its effect on
red. Finally, growth-phase dependent expression of PSRs was compared to that ofabsA
to establish temporal relationships between AbsA activity and the appearance of PSRs of
the red and cda clusters.
MATERIALS AND METHODS
Time-course analyses of pathway specific regulators redD, refl, and cdaR
transcripts were performed using high-resolution Sl nuclease protection assays. The
same RNA samples used for Si analyses of the absA locus were used here, therefore,
83
growth conditions and RNA isolation procedures were as described in the Materials and
Methods of Chapter 3. Double-stranded DNA probes were also labeled and synthesized
under the same conditions described above (Chapter 3, Materials and Methods). A redD
probe of 497 bp, with a predicted 330 nt 81 product (167; l), was generated from primers
and template described in Aceti and Champness (1998). The expression of redZ was
evaluated with a 405 bp probe synthesized from template pIJ4132 (White and Bibb,
unpublished) with forward primer WC43 (5' AGATCTTGGAGCGGGAACTCTC
CCTGC) and ”P-labeled reverse primer WC96 (5' GTCGCAGCACACACCAGGA
CACG). Previous studies of this gene predict this probe would produce an 81 product of
141 nt (65). A cdaR probe of 584 bp was amplified from S. coelicolor 11501
chromosomal DNA with forward primer WC106 (5' GGCGCACTGACGAAA
GCAAGGGC) and 32P-labeled reverse primer WC94 (5' CCGCCCACCGTAAGACC
TCGGCC). The transcription start site for this locus had not previously been determined.
81 nuclease assay conditions were identical to those used for absA. Likewise, the same
glk probe was used to normalize RNA loading, and S1 product sizes were estimated
alongside 32P-labeled ¢XI74lHinfl molecular weight markers (Promega).
RESULTS
Dependence of Pathway-Specific Regulators rail and redD on AbsA.
Red antibiotic synthesis is under the control of two pathway-specific regulators,
RedD and RedZ (reviewed by 35; 31). The RedD activator is required for Red synthesis
(144; 50) and overexpression of redD causes an increase in red biosynthetic gene
expression and Red production (122; 167). Expression of redZ is required for both Red
84
synthesis and redD expression, suggesting that RedZ is an activator of redD transcription
(182). Negative regulation of Red synthesis by AbsA was previously demonstrated to act
through transcriptional regulation of redD (1). Therefore, it was of interest to see if the
dependence of redD on AbsA2 was mediated through redZ, a possible target of AbsA2.
Similarly, while redD (167), red (182) and absA (Chapter 3) have all been shown to be
temporally regulated, I sought to gain a better understanding of the temporal relationship
between the global regulator AbsA2 and expression of pathway-specific regulators which
are possibly under its control. As such, S1 nuclease protection assays were performed to
evaluate redD and redZ transcription with the same RNA samples used to analyze absA.
These were extracted from 18, 30, and 54 hr cultures representing time points prior to and
during antibiotic synthesis. AbsA-mediated regulation of red PSRs was tested in wild
type (J 1501) versus absA mutant strains C542 (Abs’) and C570 (Pha).
The redD transcript gave a single product of 330m (Figure 15) with a temporal
pattern and AbsA2 dependence in agreement with that observed previously (1). The 18
hr cultures although turbid, still had rather sparse growth and no pigmentation. As
expected, no redD transcript was discemable. redD was not present until 30 hours when
it was strongest in J1501, coinciding with the Red pigment observed in this culture at that
time. The level of redD in 30 hour C570 was surprisingly low since this strain
reproducibly demonstrated early onset of Red on SpMR agar plates.
85
bp,. f‘r‘,‘ (‘6 ‘5
3171 m3 chm" v
$88 £55. .388 12
F" v— o —-a
88$. 9.8:; 328$ E
:00 :00 :00 E
Fishrll 30hr1154hrl
Figure 15. High-resolution Sl nuclease protection analysis of redD and redZ transcripts.
Temporal regulation of redD and redZ was followed by analyzing RNA samples isolated
from S. coelicolor strains grown for 18,30 and 54 hours 1n liquid SpMR. The
dependence of redD and redZ expression on the AbsA two-component system was tested
by examining their transcription in strains 11501 (wild type), C542 (Abs'), and C570
(Pha) at each time point. Glucose kinase (gIk) transcript was measured to normalize the
quantities of RNA loaded at each time point.
86
Nevertheless, the low level of signal did coincided with the lack of pigmentation
observed in this culture. By 54 hours, J1501 cultures had attained a strong red hue, while
the C570 cultures were maroon, having overtaken J1501 in Red as well as visibly
producing Act. At no point was there any pigmentation in the Abs' C542 cultures. As
expected, based on previous observation (1), the level of redD transcript in the C542 Abs'
strain was extremely low at all times tested.
Contrary to the marked temporal regulation seen for redD, redZ (141 nt signal)
was strongly expressed at all stages of growth examined in this culture medium. There
was, however, an increase in redZ from 18 to 30 hours, indicating that this gene is
temporally regulated. The fact that it was expressed so strongly hours before redD is in
contrast to the almost simultaneous upregulation of redD and redZ reported by White and
Bibb (1997). Equally apparent from the transcript data was the lack of dependence that
refl had on AbsA2, as witnessed by the similarity in signal intensities for each of the
strains tested.
Dependence of cdaR Expression on AbsA
No previous transcript data was available for cdaR, but its expression was
hypothesized to be growth-phase dependent like other antibiotic PSRs. The intergenic
region between the translation start codon of the deduced open reading frame for cdaR
and the stop codon of upstream gene SCE8.09 is 744 bp. 81 nuclease protection assays
of cdaR transcript employed a 584 bp double-stranded DNA probe that protected a region
439 nt upstream of the predicted translation start codon. Expression of cdaR was
evaluated at 18, 30, and 54 hours in strains J1501 (wild type), C542 (Abs-) and C570
87
(Pha). S1 nuclease protection results presented in Figure 16 revealed a major product of
approximately 500 nt, corresponding to a transcription start site about 380 nt upstream of
the translation start codon. Although undigested probe was present in some samples, its
inconsistency lead me to believe that it did not represent an additional transcription start
site beyond that of 380 nt. There were some minor signals of smaller size that could
represent additional promoters of this gene; however, these were not firrther analyzed in
this study. Inspection of sequence upstream from the region encompassing the
transcription start point identified hexamers around -10 and -35 with strong similarities to
consensus sequences proposed by Strohl (1992) to be recognized by Streptomyces RNA
polymerase containing and E. coli 67°
-like sigma factor (data not shown). Comparison of
cdaR expression in J1501 between 18 and 30 hours suggested that this gene was under
growth-phase dependent regulation. It also appeared as though cdaR was expressed
considerably earlier than redD under these growth conditions, but not earlier than refl.
It was not clear fiom the results presented in Figure 16 whether AbsA regulated
the expression of cdaR. A comparison of 18 hour J 1501 and C542 transcripts indicated
that there was a strong effect. Given that C542 is an Absi strain that does not produce
antibiotics and cdaR expression was dramatically increased in the Abs’ strain, the
transcript results suggested that CdaR was a negative regulator. In contrast, a
chromosomal disruption of cdaR reportedly blocked synthesis of CDA (156) implying
that CdaR was an activator of CDA. An additional point of confirsion was that the 18
hour C570 signal was also stronger than that of J1 501. C570 is an overproducing Pha
strain, but like the Abs' strain showed increased expression of cdaR in comparison to
88
:3 FAA CG... QGA
5 B 36’ 2 —° 3 a D B a
3 332-, $216: 53%.
‘23 § 8 E § 8 E 5’» 9. E
3 L: DU :4 U U 'v—i U U
2
I 18hr ll 30hr l I 54hr l
713»
553) "
’ ” ‘AbSAZ'HiSw
‘* era-u—
18.4P
Figure 18. Purification of AbsA2 proteins overexpressed in E. coli.
(A.) SDS-PAGE (12%) analysis oinS1o-tagged AbsA2 D54E protein purified fi'om E.
coli BL21(DE3) containing pTBA240. Lanes: (1) Molecular weight markers; (2)
Whole-cell lysate of BL21(DE3)/pET16b, 3 hours post-induction; (3) Whole-cell lysate
of BL21(DE3)/pTBA240, 3 hours post-induction; (4) Souble phase of
BL21(DE3)/pTBA240 whole-cell lysate; (5) 6M urea-solubilized cell lysate pellet fi'om
BL21(DE3)/pTBA240; (6) Ni2+-column flow through of 6M urea-solubilized pellet; (7)
Niz+-column wash with binding buffer; (8) Ni2+-column wash with wash buffer (60mM
imidazole); (9) Dilute eluate of 6M urea-solubilized pellet from BL21(DE3)/pTBA240;
(10) Molecular weight markers. (B.) SDS-PAGE (12%) analysis of Ni2+-column
purified Hisro-tagged AbsA2 and AbsA2 D54E. Lanes: (1) Molecular weight markers;
(2) Eluate of 6M urea-solubilized AbsA2-His“) fiom BL21(DE3)/pTBA235; (3) Eluate
of 6M urea-solubilized AbsA2(D54E)-Hisro fi'om BL21(DE3)/pTBA240.
In- Vitra Phosphorylation of AbsA2-His". with Acetyl Phosphate
Response regulators can be readily phosphorylated by small molecular weight
phosphate donors such as acetyl phosphate (107; 117). Moreover, many cases exist
where His-tagged RRs retained both in-vitra phosphorylation and in-viva physiological
activity without the necessity of cleaving the fusion tag (137; 132; 70; 105; 108). Thus, I
108
sought to examine the requirement of Asp54 for the phosphorylation of AbsA2 by in-
vitra phosphorylation assays with AbsA2-His“, and AbsA2 (D54E)-Hisro using
enzymatically synthesized 32P-acetyl phosphate as a phosphate donor. Prior to use in in-
vitra phosphorylation, purified 6M urea-denatured AbsA2-His") proteins were refolded
by the gradual removal of urea through dialysis. Refolding was also attempted while
AbsA2-His“) was still bound to the nickel column by applying a 20 ml linear gradient of
6 M to O M urea. Quon, et al. (1996) used E. coli acetate kinase to synthesize 32P-acetyl
phosphate for subsequent phosphorylation of RR CtrA-Hiss. Phosphorylation of
renatured AbsA2-H1810 (after dialysis or gradient renaturation) was unsuccessful by this
method. However, from the controls that were included in the assay it did appear as
though the enzymatic synthesis of 32P-acetyl phosphate was proceeding to completion.
Thus, it was suspected that my attempts to refold the denatured AbsA2-His“) was not
producing an active conformer of AbsA2.
Overexpression and Purification of AbsA2-His; from S. lividans
Unsuccessful phosphorylation of heterologously produced-renatured AbsA2-
Hism led me to overproduce AbsA2 in a Streptomyces expression system. It was
hypothesized that weaker expression from the tipA promoter, together with production in
its natural environment, might favor accumulation of AbsA2-His6 in a soluble and active
form. The same allele of absA2 used to construct pTBA235 was ligated into pIJ4123
(168) to create pTBA270, which encoded an N-terminal Hiss fusion to AbsA2. Fusion
protein expression was under the control of the thiostrepton-inducible tipA promoter.
Given the slow growth of Streptomyces relative to E. coli, S. Iividans 1326/pTBA270 was
109
grown for 16 hours prior to induction and an additional 12 hours post induction.
Comparison of the whole-cell lysates from strains 1326 and 1326/pTBA270 did not
clearly indicate a band due to overproduced AbsA2-His; fusion protein (Figure 19, lanes
1 and 2). Nevertheless, there was a significant product with an Mr of ~29kDa in the
whole cell lysate and the soluble and particulate fi'actions of the lysate which could be the
fusion protein (Figure 19, lanes 2, 3 and 4). When the soluble phase of the lysate was
purified on a nickel column, a 29 kDa protein was eluted at better that 90% purity (Figure
19, lanes 8 and 9), which was in close agreement with the predicted size of 26.2 kDa for
AbsA2-Hig.
12345678910 rd),
.1. “Manama“: w! .‘ a
A
A
N
4 18.4
‘ 14.3
‘ 6.2
Figure 19. Purification of AbsA2-Hig overexpressed in S. Iividans. SDS-PAGE (12%)
analysis of Hisa-tagged AbsA2 purified from S. Iividans 1326 containing pTBA270.
Lanes: (1) Uninduced 16-hour S. Iividans 1326; (2) Whole-cell lysate of 12-hour post
induction S. Iividans l326/pTBA270; (3) Soluble phase of 1326/pTBA270; (4) 6 M urea-
solubilized pellet fi'om 1326/pTBA270; (5) Molecular weight mkers; (6) Ni2+-column
flow through from soluble phase; (7) Ni2+-column wash with 60 mM imidazole; (8) Ni”-
column soluble-phase post-wash eluate (Fraction 1); (9) Soluble-phase eluate (Fraction
2); (10) Molecular weight markers.
110
It was noticed after repeated purification of 1326/pTBA27O cultures that very low
yields of AbsA2-Hig were obtained in the final eluted fraction. Examination of a low
imidazole wash fiom the nickel column revealed that most of the 29 kDa AbsA2-His;
was eluted in this fraction, suggesting that at least part of the reduction in yield was due
to deterioration of the nickel-column affinity matrix from repeated use. Therefore, the
wash sample was dialyzed and analyzed by SDS-PAGE, revealing that AbsA2-Hig
represented the major single band and approximately 30% of the total protein in this
sample.
In- Vitra and In-viva Analysis of AbsA2-Hi“
Purified AbsA2-Hig produced in Streptomyces was tested by in-vitra
phosphorylation experiments as described above. Once again I was unable to
demonstrate phosphorylation of AbsA2-Hig. Therefore, it was of concern that the His;
and Him fusion peptides might be rendering AbsA2 inactive. To test the activity of
AbsA2-Hig in viva, pTBA270 was transformed into C570 (see Chapter 3), which carried
a chromosomal allele that encoded AbsA2 with a D54E mutation at the proposed site of
phosphorylation. Strain C570 caused an antibiotic overproducing phenotype (Pha) which
has been associated with an inactive form of AbsA2. When pTBA270 was introduced
into C570, antibiotic production was regulated at the wild type level (Figure 20).
Conversely, when pIJ6021 (tipA-containing overexpression vector without insert (168)
was introduced into C570, there was no effect on overproduction of antibiotics (Figure
20). This suggests that the AbsA2-His; fiision protein encoded by pTBA270 was
111
C570
C570/
pIJ6021
Figure 20. In-viva analysis of pTBA270. The biological function of AbsA2-His;
encoded by pTBA270 was tested by its ability to complement the Pha phenotype of strain
C570, which carries a chromosomal absA2 (D54E) mutation. Plasmid pTBA270 was
transformed into strain C570. Antibiotic production of C570/pTBA270 was compared to
that of C570 and wild type J1501. As a control, pIJ6021 (Hisr; overexpression plasmid
identical to pTBA270 without the absA2 insert) was also transformed into C570 to test its
ability to complement the Pha phenotype. Strains were analyzed after 4 days of growth
on SpMR agar.
functional when expressed in viva. Therefore, it was hypothesized that the inability to
phosphorylate AbsA2-H156 in vitro may have been due to an inherent instability of this
protein under the purification or assay conditions, or a faulty phosphorylation reaction.
An alternative to the enzymatic generation of 32P-acetyl phosphate used in this study is
the chemical synthesis of 32P-labeled small molecular weight phosphate donors. The
latter method has been more widely used for in-vitra phosphorylation of Rs (107; 117)
and will be implemented in firture trials with AbsA2.
112
Overexpression and Purification of 'AbsAl from E. coli
Although small molecular weight phosphate donors have proven useful for in-
vitra phosphorylation of various response regulators, there is at least one report in which
it did not serve as a suitable donor (107). A preferred approach with several two-
component systems has been the purification of cognate HR for use as an in-vitra
phosphate donor (148; 178; 153; 42). This approach allows for more specific and
efficient phosphorylation of Rs (161) as well as the opportunity to study properties of
the HK and HK-RR interactions. Given difficulties phosphorylating AbsA2-His proteins
with acetyl phosphate, I decided to overproduce and purify AbsAl to use as a phosphate
donor in vitro.
Since AbsAl is membrane bound, a truncated form, 'AbsAl, which lacks the N-
terrninal transmembrane domain but possesses the entire transmitter and C-terminal
domains, was expressed to favor accumulation in the cytoplasm. The truncated alleles of
wild type 'absAI and 'absAI (HZOZL), which encodes an H202L change at the putative
site of phosphorylation, were ligated into pMAL-C2 to produce N-terminal maltose-
binding protein fusions (MBP-‘AbsA1) (Figure 17). Protein expression from pTBA350
(MBP-‘AbsAl) and pTBA360 (MBP-‘AbsAl- [1-1202L]) was IPTG inducible from a tac
promoter. The MBP-‘AbsAl fiisions have a predicted Mr of 90 kDa. The whole cell
lysates from these cultures demonstrated a band of minor intensity at about 90 kDa in
comparison to the same strain carrying the pMal-c2 vector with no insert (data not
shown). The soluble phase of the JM109/TBA350 cell lysate (Figure 21, lane 1) was
purified on an amylose resin column as described in Materials and Methods. Purified
MBP-‘AbsAl fusion protein with a predicted Mr of 90 kDa produced a diffuse band of
113
products ranging from about 75 to 90 kDa (Figure 21, lane 2) suggesting possible
proteolysis in the host strain E. coli JM109. pMAL-c2-generated fusion proteins contain
a Factor Xa recognition sequence situated between MBP and 'AbsAl. Cleavage of
purified MBP-'AbsAl with Factor Xa (Figure 21, lane 4) produced a major band in the
range of ~42 to 47 kDa, which was probably composed principally of the liberated MBP
domain (~42 kDa) and some 'AbsAl (predicted M,of 47 kDa). Another major band
represented a product of ~38 kDa and there was also one minor band at ~30 kDa, again
suggesting either proteolysis of the fusion protein in the host strain or secondary cleavage
sites within 'AbsAl that were recognized by Factor Xa
175
83
62
47.5
32.5
25
16.5
Figure 21. Purification of 'AbsAl expressed in E. coli. SDS-PAGE (10%) analysis of
maltose binding protein fusions to truncated absA] ('absAI) purified fi'om E. coli JM109
containing pTBA350. Lanes: (1) Soluble-phase whole-cell lysate fi'om 3-hour post-
induction JM109/pTBA350; (2) Post-wash eluate of soluble-phase JM109/pTBA350
purified on an amylase resin column; (3) Concentrated JM109/pTBA350 eluate; (4)
Concentrated JM109/pTBA350 eluate treated with Factor Xa; (5) Molecular weight
markers.
114
Autophosphorylation of 'AbsAl and Phosphorylation of AbsA2-His;
Autokinase activity of MBP-'AbsAl and the 'AbsAl Factor Xa cleavage products
was tested by addition of [y-3ZP] ATP. The major phosphorylation product in the
untreated MBP-'AbsAl sample (Figure 22, lane 1) corresponded to the same diffuse band
from ~75 to 90 kDa observed by Coomassie blue staining (Figure 21, lane 2). Afier
proteolysis with Factor Xa, the major phosphorylation products corresponded to the 38
and 30 kDa proteins (Figure 22, lane 4) previously observed by Coomassie blue staining
(Figure 21 , lane 4). An equivalent concentration of MBP-’AbsAl H202L was treated
MBP-’AbsAl D
Figure 22. In-vitro phosphorylation of AbsA2-Hig by 'AbsAl-phosphate.
Radioactively labeled MBP—'AbsAl-phosphate and 'AbsAl-phosphate were prepared by
incubating purified MBP-'AbsAl and 'AbsAl (cleaved fiom MBP by Factor Xa) with [y-
32P] ATP in the presence of Mg“. After incubation, semi-purified soluble-phase AbsA2-
Hisa from S. lividans was added to each sample. Samples were analyzed by SDS-PAGE
(10%) and autoradiography. Lanes: (1) MBP-'AbsAl + [y-3ZP] ATP; (2) MBP-'AbsAl-
32P + AbsA2-Hig; (3) Molecular weight marker; (4) Factor Xa-treated MBP-'AbsAl +
[y-3’P] ATP; (5) 'AbsAl-32P + AbsA2-His6.
115
with Factor Xa and subjected to the same phosphorylation conditions, demonstrating no
phosphorylation of the 'AbSAl H202L mutant (data not shown). These results indicate
that phosphorylation in MBP-'AbsAl is occurring in the 'AbsAl portion of the firsion
protein, and that His202 of AbsAl is required for phosphorylation.
MBP-'AbsAl and the 'AbsAl cleavage product were phosphorylated with [y-32P]
ATP and used as donors for the in-vitra phosphorylation of AbsA2-His; produced in
1326/pTBA270. Semi-purified AbsA2-Hig was added to each of MBP-'AbsAl-
phosphate (Figure 22, lane 2) and 'AbsAl -phosphate (Figure 22, lane 5). Addition of
AbsA2-Hig to 'AbsAl-phosphate produced a strong signal at ~29 kDa (Figure 22, lane
5), which coincided with the size of AbsA2-H156 in the semi-purified sample. The fact
that this product did not appear when incubated with MBP-'AbsAl-phosphate suggested
that its phosphorylation was dependent upon the liberated 'AbsAl domain. In contrast,
heterologously produced AbsA2-His“), renatured by dialysis after 6 M urea solubilization
and purification, was not phosphorylated under similar conditions. Addition of AbsA2-
Hiss to MBP-'AbsAl-phosphate produced a faint signal at ~36 kDa (Figure 22, lane 2).
This same signal appeared when [y-nP] ATP was added to soluble phase 1326/pTBA270
whole-cell lysate in the absence of MBP-'AbsAl or 'AbsAl (data not shown). Therefore,
this signal appeared to represent a contaminating protein whose phosphorylation did not
depend upon the presence of 'AbsAl. Preliminary evidence from these experiments
suggests that AbsA2-Hig could be phosphorylated from 'AbsAl-phosphate after
cleavage of the MBP fitsion domain.
116
DISCUSSION
Purification of AbsA2 and AbsA2 D54E was pursued for use in in-vitra
phosphorylation experiments to support genetic evidence for the role of Asp54 in AbsA2
phosphorylation and activity. Highly purified AbsA2-His“) proteins renatured from the
insoluble fraction of E. coli cultures, and soluble phase AbsM-Hi36 from S. lividans were
not phosphorylated in vitra by acetyl phosphate. Subsequent overproduction of MBP-
‘AbsAl fusions was performed to provide an alternative phosphate donor for AbsA2-His
fusions. When 'AbsAl-phosphate was reacted with semi-purified AbsA2-His; from
C270, preliminary results suggested cognate HK-RR phosphoryl-group transfer to form
AbsA2-Hiss-phosphate. Formation of an active phosphorylated conformer by the
AbsA2-Hiss fusion was fisrther supported by its ability to complement mutant AbsA2
D54E in viva. In addition, comparison of autokinase activity by 'AbsAl and 'AbsAl
H202L provided evidence that His202 is the site of phosphorylation in AbsA] due to the
inability of 'AbsAl H202L to become phosphorylated under identical conditions in which
wild type 'AbsAl was readily phosphorylated.
Numerous two-component regulators overexpressed in E. coli and renatured after
6 M urea solubilization have been successfully phosphorylated in vitra by low molecular
weight phosphate donors (137; 132; 108). Initial attempts to phosphorylate AbsA2-His"),
produced in E. coli and renatured after 6 M urea solubilization, were unsuccessfirl
utilizing an enzymatic preparation of 32P- acetyl phosphate. Similarly, AbsA2-His;
purified from the soluble phase of S. lividans was not phosphorylated in this reaction. It
is uncertain at this time if these experiments were unsuccessful because the AbsA2-His
fission conformers were inactive as a result of purification or reaction conditions, if there
117
was a problem with the enzymatic synthesis of acetyl phosphate, or if AbsA2 can not
utilize acetyl phosphate as a phosphate donor. The affinities of response regulators to
small molecular weight phosphate donors are low (161) and differences in the reactivities
of Rs to acetyl phosphate have been reported (118). At least one RR, CheB, has been
reported that can not use acetyl phosphate as a phosphate donor (107). An alternative
method to that presented in the Materials and Methods for in-vitra phosphorylation of
RRs utilizes chemically synthesized 32P--labe1ed small molecular weight phosphate
donors (107; 117). Given that certain studies of two-component protein behavior are
facilitated by phosphorylation with species such as acetyl phosphate (e. g., AbsA2-P
DNA-binding, AbsAl phosphatase activity), we will continue to investigate expression,
purification, and reaction conditions necessary to achieve small molecular weight donor
in-vitra phosphorylation of AbsA2.
An alternative method that offers specific and efficient in-vitra phosphorylation of
Rs has been routinely demonstrated using purified truncations of cognate 1-IKs(148;
178; 153). Preliminary data presented here suggests that AbsA2-His; was successfully
phosphorylated in vitra by 'AbsAl -phosphate. Nevertheless, fisrther work needs to be
done to optimize overexpression and purification of 'AbsAl and AbsA2-Hig. The
conditions for consistent recovery of highly-purified AbsA2-His; from S. lividans need to
be reestablished. Similarly, multiple phosphorylation products due to proteolysis of
MBP-'AbsAl tend to confound the results of the in-vitra phosphorylation reactions.
Proteolysis of MBP-'AbsAl might be avoided by expression of pTBA350 in an
alternative host, such as BL21 which has protease gene knockouts, or possibly by
118
expressing only the transmitter domain of AbsAl instead of the transmitter and C-
terrninal domains.
Finally, while the scope of the experiments presented in this chapter were
originally oriented at analyzing the role of Asp54 in phosphorylation and activity of
AbsA2, the utility of obtaining purified AbsAl and AbsA2, and establishing the
conditions for in-vitra phosphorylation, project into various fisture pursuits of this project.
I have already mentioned determining the phosphatase activity of AbsAl. In addition,
AbsA2 in the phosphorylated and unphosphorylated states will be used in gel-shift and
footprint assays to determine targets and a possible consensus binding sequence(s) of
AbsA2, as well as the requirement for phosphorylation on DNA binding. Another
projected use for purified AbsA2 is in the preparation of antibodies for
immunoprecipitation and in-viva phosphorylation studies aimed at differentiating
between the kinase and phosphatase default models of the AbsA mechanism discussed in
Chapter 3.
119
CHAPTER 6
AbsA2 HOMOLOGUES IN OTHER STRAINS OF ST KEPT OMYCES
120
INTRODUCTION
Streptomyces are presently the single greatest natural source of chemotherapeutic
agents for health and agriculture (165). Nevertheless, a. paucity of knowledge on the
genetic regulation of antibiotic synthesis in the genus as a whole has traditionally led
industry to depended on expensive and laborious random mutagenesis and screening
programs to achieve improvements in product yield. Currently, numerous groups are
working in the model organism S. coelicolor to piece together the complex array of
pathways involved in antibiotic regulation (reviewed in Chapter 1; 30; 74; 35). It is
hoped that many of the regulatory processes uncovered in S. coelicolor will be conserved
in other streptomycetes such that a general model for the regulation of antibiotic
synthesis can be constructed. New data presented here, along with existing evidence
from Streptomyces and other genera, lend support to generalized conservation of
developmentally regulated processes.
Precedent for the evolutionary conservation of genetic regulation of
developmental processes such as sporulation and antibiotic synthesis exists in other
Gram-positive bacteria. Producers of class II antimicrobial peptides (AMP) from the
genera Camabacierium and Lactabacillus show remarkable similarities in the
organization and function of genes within clusters encoding their function (reviewed by
96). Regulatory features conserved in these clusters include genes encoding a peptide
pheromone precursor, a two-component system, and an ATP-binding cassette (ABC)
exporter. Similarly, gene clusters for the regulation and synthesis of lanbiotics by species
of Lactabacillus, Bacillus, and Staphylococcus show pronounced conservation of gene
organization and fisnction (reviewed by 96). Once again, conserved regulatory elements
121
include signal precursors, two-component regulators, and transport systems. While these
examples demonstrate conserved regulation of the products of a single gene cluster
within which both regulatory and biosynthetic genes are found, more complex regulatory
schemes may also be conserved.
Conservation of a two-component regulator that functions pleiotropically in
differentiation is exemplified by SpoOA. The phOSphorylated state of SpoOA triggers the
initiation of sporulation in B. subtilis by activating and/or repressing transcription of
several other regulators encoded on independent operons (reviewed by 68). Youngman
and collaborators (21) presented compelling evidence for evolutionary conservation of
SpoOA in phylogenetically diverse species of Bacillus and Clastridium. Based on the
similarity of partial or complete DNA sequences, spOOA was proposed to be conserved in _
each of 8 Bacillus and 6 Clastridium species analyzed. Furthermore, spaOA homolog
gene disruptions performed in B. anthracis and C. acetabutylicum demonstrated
sporulation deficient phenotypes similar to spaOA mutants of B. subtilis. Thus, both
structural and functional homology were shown to be conserved.
In addition to the examples sited in other Gram-positive bacteria, several
observations suggest that numerous aspects of antibiotic regulation may be conserved
among streptomycetes. First, throughout the genus Streptomyces, antibiotic synthesis is
growth-phase dependent, which logically suggests an evolutionarily conserved genetic
basis for their temporal regulation. A possible link to temporal regulation is the cell-
density dependent accumulation of structurally similar y-butyrolactones, which has been
implicated in signaling the onset of morphological and physiological differentiation in
several species of Streptomyces (reviewed in Chapter 1). As seen in the production of
122
peptide antibiotics in other Gram-positive bacteria, a routinely conserved feature of
Streptomyces antibiotic gene clusters is the presence of one or more pathway specific
regulators (reviewed by 30). On a broader scale, preliminary evidence for the
conservation of many S. coelicolor pleiotropic antibiotic regulatory genes in other
streptomycetes has been implied fi'om PCR and Southern hybridization data (absB (136);
afsQ (84); (1st (115); thS (32); bIdA (101)).
In this chapter, I examine the conservation of the global antibiotic regulator
AbsA2 in the genus Streptomyces. An attractive feature of AbsA2 is that it acts as a
global negative regulator of antibiotics, so that by knocking it out, antibiotic production
begins earlier and reaches greater concentrations. If this mechanism were conserved, it
could have important economic implications in industrial relevant strains. Directed
mutagenesis of AbsA2 homologs would offer an alternative to traditional random
mutagenesis and screening protocols for increasing antibiotic yields. Here, I present
preliminary data that suggests the existence of AbsA2 homologs in each of ten industrial
strains of Streptomyces examined.
MATERIALS AND METHODS
Bacterial Strains and Growth Conditions
Streptomyces strains utilized in this study were S. coelicolor M600, S. albus, S.
ambafaciens 2035, S. clavuligerus, S. halstedii JM8, S. halsredii 2581, S. lincalnensis,
and S. peucetius. All studies were done with laboratory stocks of chromosomal DNA or
samples received from other labs.
123
Escherichia coli K12 strain DHSor was used for proliferation of plasmids that
carried inserts to be sequenced. E. coli was grown in culture tubes containing 10 ml of L
broth supplemented with 100 pg/ml ampicillin. Cultures were incubated for 12 to 16
hours at 37°C and 250 rpm prior to plasmid extraction and purification on QIAprep spin
columns (Qiagen).
PCR Amplification of Putative absA2 Homologs
Amplification of absA2 homologs was based on the strategy for spaOA homolog
amplification (21). Members of the response regulator superfamily share extensive
structural (primary, secondary and tertiary) similarity in the N-terminal receiver domain
(176; 11). Sequence similarities are even stronger within subfamilies, such that a primer
designed to a canserved region of the receiver domain would be expected to prime PCR
synthesis from most response regulators of this subfamily. The C-terminal effector
domain of Rs is more highly divergent such that primers specific to important
fisnctional regions of this domain (e. g., the helix-turn-helix [HTH] DNA-binding motif)
would be expected to prime more specifically for homologs of absA2. In addition,
streptomycetes have a codon preference favoring a high GC content, especially at
position three which has a G or C over 90% of the time (16). Consequently, where a
codon allowed either G or C at position three, degeneracy was designed into the absA2
homolog primers to provide this option. Implementing these parameters, 5' forward
primers were designed to conserved blocks around residues Asp9 (HO, 5' GCS GAC
GAC GAG ACS ATC ATC CGS GCS; where S = G or C) and Asp54 (P9, 5' GCS CTS
CTS GAC ATC CGS ATG CCS G) of the receiver domain (Figure 25). To impart
124
specificity toward absA2 homologs, a 3‘ reverse primer (P8, G TGS AGS CGC TGS
GCG ATC TCS GCG) was designed around the HTH motif of the effector domain.
Primer P8 is specific to the first helix of the HTH motif. While this is not the so-called
'recognition helix', it is projected to possess residues that are important in DNA
recognition and binding (67; 11). Five of the eight codons recognized by P8 encode
amino acids that are moderately to highly divergent with respect to a HTH motif
alignment of RRs of the same subfamily (11). Of these, the most highly divergent
codons lie at the 3' and 5' ends of P8.
PCR amplification was carried out in a 100 pl reaction volume containing 100 ng
chromosomal DNA template, 1X PCR buffer (with 1.5mM MgClz), 5% glycerol, 0.2 mM
of each dNTP, 40 pmole of each primer, and 2.5U Taq polymerase. Thermal cycler
conditions were: denaturation at 95°C for 5 min, followed by 30 cycles of 1 min at 94°C,
1.5 min at 60°C, and 1 min at 72°C, and a final extension at 72°C for 10 min. PCR
amplification products were sized by separation on a 1.5% agarose gel with comparison
to pBR322 HaeIII DNA molecular size marker (Boehringer Mannheim).
absA2 Homolog Identification and Sequencing
Southern hybridization of PCR amplification products was utilized to identify
putative absA2 homologs. PCR amplification products were separated on a 1.5% agarose
gel. DNA was transferred from the gel to a positively charged nylon membrane
(Hybond-N+, Amersham) by capillary transfer (147). DNA was fixed to the membrane
by UV. crosslinking. Hybridization and colorometric detection of an absA 2-dioxigenin
labeled probe was performed as recommended (The Genius System User's Guide for
125
Filter Hybridization, Boehringer Mannheim) with the exception that all washes were
performed at room temperature. Probes were prepared by purifying S. coelicolor absA2
PCR products with Wizard PCR preparatory columns (Promega) and using random
primed DNA labeling of 100 to 300 ng of absA2 DNA overnight as recommended (The
Genius System). In addition to probes prepared from the S. coelicolor P8/P9 and P8/P10
PCR products, another primer, P11 (5' SSWSAGGC ASSWSCCSCCSSWSGCSAC; S =
G or C; W = A, C, G, or T), was used in combination with P9 to amplify a region of the
AbsA2 receiver domain internal to the P8/P10 product to use as a probe against P8/P10
products. '
Sequencing of putative absA2 homologs was performed by cutting the P8/P10
PCR product of interest from 1% low-melting-point agarose gel and purifying with a
QIAquick Gel Extraction Kit (Qiagen). The purified product (10 ng) was used as
template for PCR amplification with primers WC8 (5' TTT TAG ACT TGA CGA CGA
GAC SAT CAT CCG SGC SGG G) and WC9 (5' TTT TAG ATC TGT GSA GSC GCT
GSG CGA TCT CSG CG), which are identical to primers P10 and P8, respectively,
except that they contain BlgII restriction sites on their ends. The WC8/W C9 PCR
products were digested with BgIII and purified on Wizard PCR preparatory columns
(Promega). The absA2 BgIII homologs were ligated into BamHI-digested pBluescript II
SK+ (Stratagene) sequencing vectors and sequenced as described by Anderson, et al.
(1999). The resulting sequence was compiled, analyzed and compared to that of S.
coelicolor absA2 using the Wisconsin GCG software package.
126
RESULTS
PCR Amplification of Putative absA2 Homologs from Heterologous DNA
Primers specific to highly conserved regions of the receiver domain and the HTH
motif of the effector domain were used to amplify putative absA2 homologs from
chromosomal DNA of industrially important strains of Streptomyces. Receiver domains
of the same subfamily of RRs show extensive sequence similarity. Therefore, primers
P10 and P9 were expected to prime numerous RRs of the AbsA2-containing subfamily
since they were specific to highly conserved regions around codons for Asp9 and Asp54
of AbsA2. Conversely, primer P8, which was specific to the more highly divergent HTH
encoding region of the effector domain, was predicted to permit more specific priming of
absA2 homologs. Furthermore, G—C base degeneracy was designed into the third position
of codons where either base encoded the same residue to allow for silent mutations in
evolutionary divergence, while maintaining the high G-C codon preference of
streptomycetes.
The PCR products obtained using primer combinations P8/P9 and P8/P10 on
chromosomal templates from S. coelicolor and eight industrial stains are shown in Figure
23. S. coelicolor DNA generated a single product of expected molecular size from each
primer combination (PS/P9, 360 nt, Figure 23A lane 2; P8/P10, 522 nt, Figure 23B lane
2). Conversely, an assortment of products of various sizes and intensities were amplified
from the other strains. Every strain except S. halstedii JM8 and S. albus generated a
product of the same size as S. coelicolor M600 fi'om primers P8/P9, some of which were
very concentrated (Figure 23A). Primer combination P8/PlO only generated a product
equivalent in size to M600 from S. ambafaciens, although both S. griseus and S. albus
127
generated products only slightly smaller or larger (Figure 23B). In addition, each of these
P8/P10 products was of very modest abundance in comparison to many of those
amplified with P8/P9.
A. B.
1234567891011 1234567891011
-
-
a:
..--=
- .-
~
Figure 23. PCR amplification of putative absA2 homologs. Putative homologs of absA2
were amplified with primer pairs P8/P9 (A.) and P8/PlO (B.). Reactions were carried out
in 100 pl volumes containing 100 ng chromosomal DNA template, 1X PCR buffer (with
1.5 mM MgClz), 5% glycerol, 0.2 mM of each dNTP, 40 pmole of each primer, and 2.5U
Taq polymerase. Thermal cycler conditions were: denaturation at 95°C for 5 min,
followed by 30 cycles of 1 min at 94°C, 1.5 min at 60°C, and 1 min at 72°C, and a final
extension at 72°C for 10 min. PCR amplification products were sized by separation on a
1.5% agarose gel. Lanes; (1) DNA molecular weight marker pBR322 HaeIII; (2) S.
coelicolor M600; (3) S. Iincalnensis; (4) S. halstedii JM8; (5) S. griseus; (6) S. albus;
(7) S. avermitilis; (8) S. cinnemanium; (9) S. halstedii 2581; (10) S. ambafaciens; (l 1)
DNA molecular weight marker pBR322 HaeIII.
Identification of Putative abs/42 Homologs
Although major amplification products from P8/P9 corresponded in size to absA 2,
there were numerous other products of significant abundance generated in almost every
128
strain. Moreover, the major products amplified from P8/P10 were not of expected size.
Therefore, in order to determine if any of these products represented possible absA2
homologs, they were transferred to nylon membranes and hybridized against S. coelicolor
P8/P9 or P8/PlO absA2 probes.
The Southern blot of P8/P9-generated products provided strong evidence that
genes of similar size and sequence to that of absA2 were amplified from each of the
strains tested (Figure 24A). Not only was hybridization specific to fragments of the
expected size on these blots, but the absA2 probe revealed products of similar size in S.
halstedii JM8 and S. albus that were not visible on the gel alone. The lack of
hybridization to fi'agments of other sizes together with the appearance of signals in S.
halstedii JM8 and S. albus, suggested high annealing specificity for full-length products
of similar sequence to that of absA 2.
In marked contrast, the Southern blot of P8/P10-generated products (Figure 24B)
demonstrated a hybridization pattern virtually identical to the pattern of amplification
products seen on the gel. There was no apparent specificity for products of equivalent
size to P8/P10 absA2. It was suspected that hybridization conditions for this blot were
not stringent enough to exclude annealing based solely on primer recognition. Therefore,
a fragment internal to P8/PlO was generated fiom M600 and hybridized to the P8/P10
products. The new P9/Pll probe recognized a 210 bp region of the receiver domain of
absA2. Remarkably, when this probe was hybridized to P8/P10 PCR products,
essentially all signals corresponded in size to absA2 (Figure 24C). As observed in the
P8/P9 blot, hybridization signals were revealed for S. Iincalnensis, S. halstedii JM8, and
S. cinnemanium that were not visible on the gel. In their initial rounds of PCR
129
amplification of spoOA homologs using a similar approach, Brown et a1. (1996) obtained
products of similar size to those of SpoOA from four out of eight Bacillus species. Here,
Southern hybridization of PCR products suggested that homologs of similar size to absA2
existed in each of the strains tested. Indeed, sequence analysis of the amplified region of
four putative homolog genes confirmed that these were of similar size to the same region
of absA2 (see below).
A. B. C.
123456789 123456789123456789
Figure 24. Southem blots of putative absA2 homologs. Putative absA2 homologs
generated by PCR from primer pairs P8/P9 and P8/P10 were hybridized to absA2 probes
generated from S. coelicolor M600. (A.) P8/P9 PCR products hybridized to a P8/P9
absA2 probe; (B.) P8/P10 PCR products hybridized to a P8/P10 absA2 probe; (C.)
P8/P10 PCR products hybridized to a P9/Pll absA2 probe. Lanes: (1) S. coelicolor
M600; (2) S. IincaInensis; (3) S. halstedii JM8; (4) S. griseus; (5) S. albus; (6) S.
avermitilis; (7) S. cinnemanium; (8) S. halstedii 2581; (9) S. ambofaciens.
Sequence Analysis of Putative Homologs
Given the promising results obtained from Southern hybridizations, several PCR
products were sequenced to examine their similarity to absA2. Although products fiom
P8/P9 were more abundant, I chose to sequence P8/P10 products because they were 180
nt larger, encoding a region from the extreme N-terminus through to the HTH motif of
AbsA2. PCR products fiom S. griseus and S. ambafaciens were gel purified and used as
130
templates for amplification from primers WC8 and WC9, which were identical to P8 and
P l 0 but had BgIII restriction sites designed into their ends to facilitate cloning of the
amplified genes. In addition, it was found that P8/P10 PCR products of similar size to
that of absA2 could be generated from S. peucetius and S. clavuligerus (data not shown).
The sequence of these products was also determined.
An alignment of the translated sequences from S. ambafaciens, S. griseus, S.
peucetius, and S. clawligerus with that of AbsA2 are shown in Figure 25. Inspection of
the consensus sequence from this alignment revealed substantial amino acid similarity
over the length of the partial gene products with the exception of the region fi'om residue
1 3 8 to 154 of AbsA2. This region of the RR encompasses a solvent accessible loop
which acts as a flexible tether joining the receiver and effector domains. It is devoid of
secondary structure (prediction data not shown) and is highly divergent among response
regulators (131; 11). A second apparently divergent region was observed between
residues 67 and 81 of AbsA2; however, closer inspection revealed that only a four amino
acid stretch fi'om position 70 to 73 was highly variable. Comparison with crystolgraphic
data fi'om NarL RR predicted this small region to encompass a loop between helix 01-3
and strand B-4 of the receiver domain (11; AbsA2 modeling data not shown).
Although the alignment shows significant sequence similarity over most of the
amplified region, receiver domains from different RRs of the same sub-family tend to be
131
P10
1 I 50
AbsA2 mirvllaDDE TIIRAGvrsI LttepgiEVV.AEAsdGreAv eLarkHRPDV
Ambofaciens DDE TIIRAGaraI LsadpeiEVV.AEAstGreAv eLvrrHRPDV
Griseus DDE TIIRAGvraI LardphvEVV AEAgdGheAi‘aLtraHRPDV
Peucetius DDE TIIRAGvraI LsadtgiEVV AEAddGrqu eLaerRPDV
Clavaligerus DDE TIIRAGvcaI LaaepgiEVV.AEAadGheAv eLterRPDV
Consensus ------- ans: Tunas-"I r. ------ svv mus-41- -r.---rramv
*
99
100
AbsA2 aLlDirMPem DGLtAAgemr ttnpdtavvv lTTFgeDrYi eRAquGvaG
Ambofaciens aLleqMqu DGLdAAseil ksdagtavii fTTanDin aRALgeGasG
Griseus vLmDirMPgl DGLtAAarlh resatvglim lTTngDeYv tRALeeGadG
Peucetius aLlDirMPrl DGLaAAeelr raaprtavvm 1TTFseDeYv eRALgansG
Clavaligerus leDvrMPrf DGLrAAeeiq rvapdtavvm lTTFseDeYi aRALdsGasG
Consensus -L-D--MP-- DGL-AA -------------- -TTr--D-!- -RAL--G--G
*
P11
101 ‘__— 150
AbsA2 FLLKasDPRd LisGVnAvas GgscLSPlvA rletelrra pspRsevsge
Ambofaciens FLLngDPRd LiaGVhAvad GaayLSPeaA thirglpta rmaRgsaare
Griseus FLLKadDPRe LlnGV2Avga GgayLSPrvA ngiagm... rahRaahphr
Peucetius FLLngDPRe LiaGVnAavq GaacLSPeiA erldrlggg rmsRaaeara
Clavaligerus FLLKagDPRe LiaGVrAvad GaacLSPevA eriarlgdg rlsRawaarr
Consensus ILLR--DPR- L--GVeA--- G---LSP--A -R ----------- R ------
*
P8
151 4— 200
AbsA2 rttlLthEq eVlgmlgaGL SNAEIAQRLH lvegtiktyv saiftqlevr
Ambofaciens rversaREr eVltllgeGL SNAEIAQRLH
Griseus slarLteREr eVlaglgaGL SNAEIAQRLH
Peucetius alethgREr eVValvaaGL SNAEIAQRLH
Clavaligerus tlethrREr dVvalvadGL SNAEIAQRLH
Consensus ----L--RE- -V------GL SNAEIAQRLH --------------------
\_J L_l \ J
helix turn helix
201 223
AbsA2 nrvqaaiiay eaglvkdadl nr*
Ambofaciens
Griseus
Peucetius
Clavaligerus
Consensus -----------------------
Figure 25. Amino acid sequence alignment of S. coelicolor AbsA2 with putative
homologs from S. amabafaciens, S. griseus, S. peucetius, and S. clavuligerus. A
consensus sequence of residues conserved in all five strains are in bold type. Primers P8
and P10 were used to amplify these genes. The orientation and location of amino acid
condons recognized by primers P8, P9, P10, and P11 are indicated by arrows. Highly
conserved residues within the acidic pocket are indicated C“). The predicted location of
the helix-tum-helix DNA-binding region of AbsA2 is also identified.
132
the most highly conserved region of the RR. Therefore, regions of similarity were sought
that might set the AbsA2 homolog apart from other RRs. Various residues were
conserved among the putative AbsA2 homologs that were highly variable in Rs of the
same sub-family as AbsA2. For example, amino acids H46 and R47 of AbsA2 were
highly divergent not only among RRs from different genera, but also from an alignment
of eleven randomly chosen S. coelicolor RRs of the same subfamily as AbsA2 (data not
shown). These residues were predicted by homology modeling of AbsA2 to the NarL
crystalography structureto represent the last amino acid of helix 01-2 and the first amino
acid of the 01-2 to B-3 loop (1 l), which is one of the most highly solvent exposed regions
of the receiver domain (176). It is tempting to speculate that these and possibly other
uniquely conserved residues are important in cognate HK-RR recognition, or some other
system-specific fisnction.
A comparison of each homolog sequence with that of AbsA2 produced the
following amino acid similarities and identities: S. griseus, 78% similarity, 60% identity;
. S. amabafaciens, 78% similarity, 61% identity; S. clavuligerus, 67% similarity, 63%
identity; and S. peucetius, 69% similarity, 63% identity. Excluding the highly divergent
linker region, which constituted about 10% of the sequence, amino acid identities would
have been around 70%. Brown, et al. (1996) did not report amino acid identities between
B. subtilis SpoOA and its homologs retrieved by PCR. They did, however, mention that
DNA sequence identity for spaOA homologs from different strains of C. acetaburylicum
was as low as 66%. DNA sequence identity between putative absA2 homologs from
different species tested here ranged fi'om 70 to 72%. Sequence identity between
homologous proteins of Streptomyces has been shown to vary greatly. The sigma factor,
133
OF, required for sporulation in streptomycetes was found to be 87% identical between S.
coelicolor and S. aureafaciens (134). In another study, five polyketide synthase (PKS)
genes encoding the type II polyketide antibiotic frenolicin in S. raseafulvus, demonstrated
approximately 40 to 70% amino acid identity with similar PKS proteins from four other
species of Streptomyces (17). Finally, a BLAST search of the nearly complete
Streptomyces genome sequence (httpz/lwww.sanger.ac.uk/Projects/S_coelicolor/)
recognized 52 RS (in addition to absA 2) with full-length similarity to AbsA2. Amino
acid identities for all but one RR (excluding AbsA2) ranged from 26 to 49%, well below
the 60 to 63% identity obtained here from putative absA2 homologs fi'om other Species of
Streptomyces. However, one RR found on cosmid St8D11 was 55% identical to AbsA2
and contained residues highly conserved in the putative homolog proteins. This locus
will be discussed below.
DISCUSSION
Evidence exists for the evolutionary conservation of developmental regulation in
Gram-positive bacteria. The well characterized SpoOA response regulator, which triggers
sporulation in B. subtilis, was shown to be conserved in all species of highly diverse
Bacillus and Clastridium tested (21). Using a strategy similar to that employed for
SpoOA, degenerate PCR primers with varying specificities for absA2 successfully
amplified nucleotide sequences of the same size as absA2 from ten industrial strains of
Streptomyces. Southern hybridization analysis demonstrated that only those products that
were the same size as absA2 annealed with absA2 probes. Yet, it was quite possible that
R genes with full-length sequence similarity to absA 2, but different biological function,
134
were being primed and amplified. Therefore, the next step in confirming whether or not
these were actual homologs of absA2 was to inspect their sequence.
Sequence analysis of putative homologs fi'om S. griseus, S. ambafaciens, S.
clawligerus, and S. peucerius revealed amino acid identities with absA2 of
approximately 60%. By comparison, S. coelicolor and S. aureafaciens homologs of the
highly conserved o'F sigma factor, required for transcription of late sporulation genes,
shared 87% sequence identity (134). In contrast, homologous PKS genes, that synthesize
different polyketide antibiotics through similar enzymatic reactions, have amino acid
similarities that commonly range from just 40 to 60% (17). Moreover, only one S.
, coelicolor gene encoding a R with fisll-length similarity to AbsA2 has yet demonstrated
greater than 49% amino acid identity. Therefore, the fact that partial gene sequences
from heterologous strains demonstrate at least 60% identity is suggestive evidence that
these genes encode AbsA2 homologs.
A blast search conducted on the nearly complete S. coelicolor genome identified a
RR from cosmid St8D11 that is 55% identical to AbsA2. The St8Dll RR possesses the
highly conserved residues of putative AbsA2 homologs that are missing from other RRs
of the same sub-family. Also, the HTH region of St8Dll RR shows 54% identity and
83% similarity to that of AbsA2. Not only is the St8Dll RR highly similar to AbsA2,
but the St8D11 I-IK is a remarkable 44% identical to full-length AbsAl. These HKs
share significant similarity in the C-terminus of the sensor domain, which is normally
highly divergent among HKs. Moreover, the St8D11 HK has a C-terminal domain that is
approximately 40%.identical to that of AbsAl, making it the only protein from any
database to demonstrate similarity to this region of AbsAl.
135
Streptomyces have linear chromosomes, the ends of which are genetically
unstable, undergoing deletion and duplication events at a rather high frequency (reviewed
by 175). As a likely consequence of this instability, no S. coelicolor housekeeping genes
are found within at least several hundred kilobases of the chromosomal ends. Cosmid
St8Dll lies very close to the end of the chromosome on restriction fragment AseI-A
(140). According to Kieser et al. (1992), this region of the J1501 chromosome seems to
have undergone several deletions and possibly duplication events in comparison to strain
M145, which is the progenitor of J1501 and the strain currently being sequenced. Thus,
it is uncertain whether the St8D11 two-component system is present in S. coelicolor
11501, the strain from which absA was isolated and the strain used in this study. Given
its remarkable similarity to AbsA, it is possible that the St8D11 two-component system
functions in antibiotic regulation, but is missing from J 1501. Studies are currently
underway to determine whether the St8D11 two-component system is present in J1501
and whether it is a regulator of antibiotic synthesis or sporulation.
Future amplification of absA homologs might take advantage of the unique C-
terrninal domain of AbsAl. This domain is about 160 amino acids long, is of unknown
function, and shows homology to only one other S. coelicolor HK. Assuming that this
domain is conserved, it could serve as a criterion for amplifying absA homologs. One
strategy would be to prime for regions of the RR receiver domain and the HK transmitter
domain that are highly conserved and of specialized function in two-component systems.
This should reduce the amplification of spurious products of varying size. At the same
time, putative homologs with a C-terminal absA] domain would generate products
136
approximately 500 bp larger than those of two-component systems with 'orthodox'
primary structure.
The recently revealed association of absA within the cda gene cluster raises obvious
questions about the possible conservation of absA throughout the genus Streptomyces.
Can it be simply coincidental that a global regulator of antibiotics is located in an
antibiotic gene cluster? If AbsA was originally associated only with cda regulation, how
did it expand its range to regulate all S. coelicolor antibiotics? If AbsA was originally
specific to cda regulation, in a manner similar to pathway-specific regulators, why don't
more antibiotic gene clusters possess two-component systems? It is still not clear
whether the AbsA two-component system is an idiosyncrasy of S. coelicolor or a highly
conserved mechanism of antibiotic regulation throughout streptomycetes. Further
analysis into the conservation of absA may help establish whether specific mechanisms of
pleiotropic antibiotic regulation tied to multicellular development are highly conserved in
Streptomyces or predominantly unique due to random evolutionary pressures of the
organisms environment.
137
CHAPTER 7
CONCLUSIONS AND FUTURE RESEARCH
138
Prior to this study the AbsA two-component signal transduction system had been
shown to globally regulate antibiotics of S. coelicolor. Interestingly, absA mutants were
associated with dramatically opposing phenotypes. Mutations mapped to the absA!
histidine kinase gene caused global inhibition of all four S. coelicolor antibiotics (Abs'
phenotype), while gene disruptions of absA caused the early anset and overproduction of
Act and Red (Pha phenotype) (19). Given the probable cotranscription of absA] and
absA2, it was hypothesized that AbsA2 was a negative regulator of antibiotic production
since gene disruptions in absA caused the Pha phenotype. This study sought to
understand the basis for the dramatically opposed phenotypes obtained from absA
mutants through further molecular genetic characterization of the absA locus and an
examination of basic aspects of the AbsA two-component mechanism. Definition of
AbsA2 as a positive or negative regulator was essential to more fully understanding any
possible interactions with other antibiotic pathway-specific or pleiotropic regulators.
Likewise, knowledge of the biochemical mechanism of signal transduction was
considered prerequisite to defining more complex characteristics of the AbsA mechanism
such as target binding and signal sensing.
Molecular Genetic Characterization of absA] Mutations Responsible for the Ahs'
Phenotype and Certain sab Suppressors of Abs'
Mutations responsible for the Abs' phenotype had previously been localized to the
absA] histidine kinase gene (2; 19). Sequence analysis of absA] from two independently
isolate Abs' mutants, C542 and C577, identified point mutations that caused amino acid
substitutions to the transmitter domain of AbsAl. The mutations to C542 (1360L and
139
R365Q) were contained within the G-box, which is proposed to play a critical role in
phosphotransfer (158). The C577 mutation (L253R) was found in a region of moderate
conservation among histidine kinases termed the X-box (79). X-box mutations in EnvZ
(79), NtrB (9) and DegS (169) lock each of these HKs into a kinase dominant,
phosphatase deficient state. Sequence characteristics of the Abs' mutations, taken
together with results obtained from the genetic and transcript analyses of the absA locus
(Chapter 3), lead to the hypothesis that mutations identified in C542 and C57 7 lock
AbsAl into a kinase dominant, phosphatase deficient state which causes constitutive
negative regulation of antibiotics. It is uncertain at this point whether the Abs’ mutants
are signal-independent; however, the early accumulation of absA transcript together with
the stability of the Abs’ phenotype over days and weeks suggests signal-independent
behavior.
Abs“ stains C577 and C542 undergo apparent pseudoreversion to attain wild type
(Type I) or Pha (Type H) levels of antibiotic production. Genetic mapping of five sab
(suppressor of glgs) mutants placed the mutations responsible for this phenotype very
close to the absA locus (Appendix A). Marker exchange experiments demonstrated that
restoration of antibiotic production was due to second-site suppressors of the Abs'
mutations. The mapping data suggested that the sab mutations were located downstream
of the absA! mutations. Therefore absA2 was sequenced in the five mapped sab strains
plus six unmapped sab mutants. Two sab mutations were identified in absA2 while a
third contained a deletion that included most of absA 1 and all of downstream absA2. The
absA! gene was then sequenced in 2 of the 4 remaining mapped sab strains. Both of
these were found to contain a sab mutation. Six remaining sab mutants that are wild type
140
for absA2, have not been sequenced for absA]. Second-site suppressors are a useful tool
for finding additional members of a regulatory pathway. Identifying whether the
remaining sab mutations lie in absA] would provide valuable insight into whether
another locus is involved in the AbsA regulatory pathway. If all sab mutations were
localized to absA, the likelyhood that AbsA functions through an intermediate regulator
(e.g., that AbsA2 is an activator of a repressor) would be diminished. If a mutation was
found to lie outside absA, then evidence would exist for the involvement of a second gene
product. Four sab mutants that do not possess mutations in absA2 are Type II (Pha);
three of these are unmapped. An initial experiment to test whether these mutants possess
a sab mutation in absA] would be to transform them with pCBZOO (19) containing a
truncated, fisnctional form of absA 1, or with pCBSZO (Chapter 3) carrying the entire absA
locus. Ifa sab mutation causing the Pha phenotype lies outside of absA 1, the absA clones
would not be expected to complement the mutation. Sab mutants not complemented by
the absA clones could first be sequenced for absA] to assure that there is not a dominant
Pha-producing mutant, and then studied through mapping and complementation to isolate
the suppressor locus.
The Role of Phosphorylation in the AbsA Regulatory Mechanism
Disruptions in the absA] and absA2 genes demonstrated that the AbsA two-
component system is a negative regulator of the multiple antibiotics produced by S.
coelicolor, including calcium-dependent antibiotic, actinorhodin and undecylprodigiosin.
While synthesis of Act and Red had previously been shown to be accelerated in absA
disruptions (19), this is the first demonstration that absA mutants also cause early onset
141
and overproduction of CDA. The absA locus was recently found to lie within the cda
gene cluster. Thus, AbsA is also the first antibiotic-gene-cluster-associated regulator that
demonstrates regulatory activity outside of its cluster.
Much of the attention of this study was focused on defining the role of
phosphorylation in AbsA2-mediated negative regulation. Gene replacements in the absA
locus altered the putative sites of phosphorylation of AbsAl or AbsA2. As predicted
from sequence conservation with other two-component systems, both the His at position
202 of AbsAl and the Asp residue at position 54 of AbsA2 were required for normal
regulation of antibiotic synthesis. Furthermore, each of the gene replacement strains
tested attained an antibiotic overproducing phenotype (Pha) consistent with a mechanism
in which AbsA2-P is the active negative regulator of antibiotic synthesis. In addition,
high-copy expression of AbsA2 D54E was not able to complement the Pha phenotype of
an absAA strain; moreover, it caused overexpression of antibiotics in a absA wild type
background. Taken together, these results suggest that AbsA2 activity is strongly
dependent upon phosphorylation.
Transcription of the absA locus also demonstrated phosphorylation-dependent
autoregulation of absA expression. Transcript studies of absA suggest either positive
autoregulation by AbsA2-P or negative autoregulation by unphosphorylated AbsA2.
Data fi'om this study coupled with precedent from other two-component systems favors a
model for phospho-AbsA2-mediated positive autoregulation. First, the genetic data
presented in Chapter 3 showed AbsA2-P to be the active regulator of antibiotics with a
strong dependence on phosphorylation for activity. Likewise, the absence of growth-
phase-related change in the absA transcript in C570 Pha strain (Chapter 3) suggests that
142
AbsA2-P was responsible for the growth-phase increase observed in J 1501 and C542.
Finally, the majority of response regulators from other two-component systems are
activated by phosphorylation (161). Therefore, I predict that AbsA2-P is not only the
active negative regulator of antibiotic synthesis, but that it also positively autoregulates
its own expression. An experiment that could resolve whether absA autoregulation is
positive or negative would examine absA expression in absAA strain C577S25 (Appendix
A). C577S25 possesses a deletion in all of absA2 and most of absA 1, excluding the
promoter region and part of the 5' region of absA] encoding the sensor domain. IfabsA
positively autoregulates its own expression, this strain would be expected to contain low
levels of transcript over the course of growth, similar to C570. Conversely, if absA
expression is negatively autoregulated, C577S25 absAA should produce elevated levels
of transcript similar to that of C542.
Model Summary
A working model which accommodates both the genetic and transcriptional data
summarized above is illustrated in Figure 26 and supposes the following. Early in
growth, a'culture is not competent for antibiotic synthesis and the absA genes are
expressed at a basal level. In the absence of signal, AbsAl is in a phosphatase dominant
form. Following a period of growth, the culture enters the “transition stage.” During this
time, the signal that regulates AbsA may be present at significant levels. Once the signal
is present, AbsAl is shifted to a kinase-dominant form and AbsA2-P will accumulate.
AbsA2-P negatively regulates antibiotics and also positively autoregulates absA
expression, accounting for the AbsA2-P-dependent, growth-phase-related increase of
143
Signal Activates
Absence of Signal .
AbsAl Phosphatase AbSAl Kinase
Dominant
. . . .
A1) A2
Inaztive 3’ n
V”
9
9
Less Signal
(stationary phase) 7
Relaxation of negative regulation EdaR _ abs cdar
A
ltedZ - -?redD red r '
Focal-4 act r
Transcriptional regulation
Figure 26. Model of AbsA-mediated regulation of antibiotic production in S. coelicolor.
144
absA transcript seen in 11501. A signal-concentration-dependent equilibrium is
established between the kinase and phosphatase forms of AbsAl, such that the level of
signal determines the extent to which antibiotics are repressed. Easing of AbsA-
repression may require that the signal be depleted or degraded. As signal decreases, the
AbsAl kinase/phosphatase equilibrium would shift toward the phosphatase form, causing
dephosphorylation of AbsA2 and a decrease in the negative regulation of antibiotic gene
expression.
This model proposes the establishment of an AbsAl signal-dependent
kinase/phosphatase equilibrium. Genetic evidence for AbsAl autokinase activity was
demonstrated by loss of function phenotypes in gene replacements that targeted the
transmitter domain or Hi5202 of AbsAl. Furthermore, purified MBP-'AbsAl (HZO2L)
was not phosphorylated in vitro under the same conditions in which wild type MBP-
‘AbsAl was readily phosphorylated, suggesting that AbsAl is phosphorylated at this
residue. Similar direct evidence for AbsAl phosphatase activity toward AbsA2-P has yet
to be demonstrated. However, genetic evidence from two experiments can be interpreted
to support phosphatase activity by AbsAl. First, overexpression of wild type or an
H202L mutant of AbsA] in a wild type absA background caused a Pha phenotype
(Chapter 3). These results are consistent with a decrease in the proportion of AbsA2-P,
and thus an increase in the number of AbsAl molecules exercising phosphatase activity.
Second, in a recent experiment, pTBA155 carrying absA1* (HZOZL) (Chapter 3) was
transformed into C542. If AbsAl * H202L possesses phosphatase activity, I would
expect some level of restoration of antibiotic production to strain C542, which is
hypothesized to be locked in a kinase dominant, phosphatase deficient state. The
145
C542/pTBA155 strain attained wild type levels of antibiotic regulation, which suggests
AbsAl H202L phosphatase activity. The assumption that AbsAl H202L can possesses
phosphatase activity while being kinase deficient is consistent with phosphatase activity
demonstrated by other HKs carrying mutations at the conserved histidine residue (80). A
more direct analysis of AbsAl phosphatase activity would test purified 'AbsAl (Chapter
5) in-vitro phosphatase activity against purified AbSAZ‘HISG‘P.
The principle focus for the immediate future of this project will be to refine and
extend the model of AbsA-mediated regulation to include the identification of targets of
AbsA2 regulation and the signal-sensing mechanism of AbsAl. A discussion of
additional data generated in this study will be presented in the context of future work in
these two areas.
AbsA2 Targets
A previous study demonstrated that mutations responsible for Abs' and Pha
phenotypes correspondingly affected transcription of redD and actII-ORF 4 (1), thus
establishing these pathway-specific activators as part of the AbsA regulatory pathway
and possible targets of AbsA2. RedZ is a second pathway-specific regulator of the red
biosynthetic pathway which appears to activate expression of redD (182). 1 found that
mutations in absA had no affect on the expression of redZ, suggesting that it is not a
target of AbsA2 regulation. Interestingly, RedZ is a "pseudo response regulator” which
possesses full-length sequence similarity to AbsA2, but has lost the requirement for
phosphorylation (65). RedZ possesses a HTH motif in its effector domain, making it
possible that AbsA2 and RedZ compete for binding at the redD promoter, although
146
probably recognizing different sequences. Exclusion ofredZ from the AbsA regulatory
pathway makes the redD promoter an attractive potential target of AbsA2-mediated
regulation of Red synthesis.
The effect of AbsA mutants was also tested on the expression of cdaR, a putative
pathway-specific activator of the cda gene cluster (156). While redD and actII-4 were
consistently repressed in Abs' mutants over the course of growth (1; Chapter 4), the
expression of cdaR was equivalent in Abs', Abs+ and Pha cultures by the time these
reached the stationary phase (Chapter 4). Only at the early 18 hr time point was there a
difference in the levels of expression of cdaR; however, both the Abs‘ and Pha cultures
had more abundant levels of cdaR transcript than J1501. The Abs' strain does not
produce CDA, so if cdaR is an activator of CDA biosynthesis and a target of AbsA2, I
would expect it to be repressed in the Abs‘ mutant. New evidence has demonstrated that
Abs' mutants dramatically reduce expression of various cda biosynthetic operons over the
course of grth (Ryding, unpublished), explaining the CDA‘ phenotype of Abs’
mutants. Inspection of the S. coelicolor genome sequence reveals potential additional
regulatory genes adjacent to the cda cluster (146). Thus, AbsA may be acting at a target
other than cdaR in the cda gene cluster. Nonetheless, CDA production is evident several
hours before the appearance of Red, and cdaR expression was evident at the earliest time
point in which absA was witnessed. Therefore, I can not yet exclude that AbsA may be
regulating the early expression of cdaR. As such, cdaR expression will be examined at
closely grouped time points between mid-exponential and stationary phase grth of
Abs', Abs+, and Pha strains. Growth curves will be generated to assure that similar time
points from each culture represent the same stage of development. Likewise, assays will
147
be conducted to determine the timing and abundance of CDA in the Abs+ and Pha
cultures. At the same time, 81 analyses will be performed on other putative regulators of
the cda gene cluster to determine their dependence on AbsA and possible involvement in
the AbsA regulatory pathway.
It is uncertain at this time whether the autoregulation of absA expression is direct
or indirect. Nevertheless, an obvious target to begin testing for AbsA2 binding is the
absA promoter region. The absA promoter has an imperfect heptameric inverted repeat, a
motif frequently recognized by response regulators. The sequence and arrangement of
the heptameric repeat sequence is loosely conserved in other possible targets of AbsA2
(e.g., redD and actII-4). Initial targets to be used in gel-shift AbsA2 DNA-binding assays
will include the promoter regions of absA, redD, and actII-4. Successful target
identification with gel-shift assays could be followed by DNaseI or hydroxy-radical
footprint analyses to identify a possible AbsA2 consensus binding sequence, the
identification of which could be valuable in the search for other targets of AbsA2, for
example in the cda cluster.
Gel-shifi assays will utilize purified AbsA2. Purification of AbsA2 and AbsA2
D54E was pursued for use in in-vitra phosphorylation experiments to support genetic
evidence for the role of Asp54 in AbsA2 phosphorylation and activity. AbsA2-IEs-tag
proteins purified from E. coli and S. lividans were not phosphorylated in vitro with
enzymatically synthesized 32P-acetyl phosphate under the conditions tested. Genetic data
strongly suggest that AbsA2-P is required for target recognition. Therefore, it may prove
necessary to resolve the problems with overexpression and in-vitra phosphorylation of
AbsA2 in order to perform in-vitra DNA-binding studies. The following strategy is
148
being pursued. AbsA2-His; overexpression will concentrate on the Streptomyces
overexpression system since, i.) AbsA2-His; was shown to be functional in viva when
expressed in S. coelicolor C570; ii.) a good yield of soluble phase AbsA2-His6 was
produced in the soluble phase, thus avoiding denaturation and renaturation (i.e., excess
handling); and iii.) preliminary evidence exists for the in-vitra phosphorylation of
AbsA2-His; by purified phospho-‘AbsAl. A steady decrease in product yield was seen
over time in the Streptomyces/Ni2+ expression/purification system. Suspected causes
include plasmid instability in the S. lividans host and purification column degradation. I
am currently examining the stability of the pTBA27O AbsA2 expression plasmid in strain
C570 and establishing overexpression and purification conditions that will lead to
consistent recovery of soluble phase AbsA2-Hig.
Both AbsA2 DNA-binding studies and determination of AbsAl phosphatase
activity would be facilitated by AbsA2-Hist; in-vitra phosphorylation by small molecular
weight phosphate donors. Once reproducible recovery of AbsA2-His; has been achieved,
in-vitra phosphorylation trials will be conducted using chemical synthesis of 32P-acetyl
phosphate. This method (118) has been employed more frequently and for a greater
number of response regulators than the enzymatic preparation of 32P-acetyl phosphate
discussed in Chapter 5. Phosphorylation of AbsA2-Hi56 utilizing purified ‘AbsAl as a
phosphate donor will be pursued simultaneously. Preliminary experiments with MBP-
'AbsA1 provided evidence that the Factor Xa-liberated phospho-‘AbsAl could serve as a
phosphate donor for AbsA2-His; from S. lividans. Nonetheless, the MBP-'AbsAl fusion
protein appears to undergo proteolysis in the E. coli JM109 host, leading to purification
products of multiple sizes. Two alternatives are currently being pursued to eliminate
149
proteolysis of MBP-'AbsAl. First, the fusion protein will be expressed in E. coli BL21, a
strain containing ampT and Ian protease knockouts. A second alternative is to
overexpress only the transmitter domain of 'AbsAl versus the current truncation which
includes both the transmitter domain and the C-terminal domain of unknown function.
The size of the MBP-'AbsAl fusion products suggests that proteolysis is occurring in the
C-terminal domain of AbsAl. Removal of the C-terminal domain should not affect the
function of the transmitter domain since a 69 aa tmncation of the AbsAl C-terminus has
previously been shown to complement Abs' mutants (19).
The AbsA Signal and Signal-Sensing Mechanism
Identification of the signal and signal-sensing mechanism has proven elusive for
most two-component systems. Understanding the signal-sensing mechanism of AbsAl
would help define whether AbsAl is kinase or phosphatase default in the absence of
signal (i.e., whether the signal activates or relieves negative regulation of antibiotics).
Likewise, identifying the signal recognized by AbsAl may provide insight into what
external environmental or cell-generated factors are involved in the onset of antibiotic
synthesis and how this is coordinated with other deve10pmentally regulated processes
such as sporulation.
Transcriptional analysis of absA expression revealed that it was growth-phase
regulated, experiencing dramatic upregulation prior to the appearance of antibiotics in a
wild type culture. The temporal profile of absA expression suggests that the signal
recognized by AbsAl increases during transition phase growth. At this time, the signal
that AbsAl senses is not known. Thus, a more immediate focus on elucidating the
150
AbsAl signal-sensing mechanism could serve to fisrther refine the AbsA mechanism and
possibly provide clues into the type of signal recognized by AbsAl.
Precedent from other two-component systems establishes two general trends for
HK signal sensing mechanisms. The first class of HK proteins, exemplified by EnvZ
(reviewed by 55) and NarX (25), have a large periplasmic loop with a proposed ligand-
binding box lying between two transmembrane (TM) helices in the N-terrninal sensor
domain. Ligand (signal) binding is proposed to cause movement of one transmembrane
helix with respect to the other, resulting in modulation of enzymatic activity in the
transmitter domain (38). Alternatively, HKs such as FixL sense the signal on a
cytoplasmic domain of the HK situated between the transmembrane and transmitter
domains (reviewed by (4)). Deletion of the membrane-bound sensor domain of NarX
locks the HK into a signal-independent kinase dominant state (25). Conversely, deletion
of the transmembrane domain of FixL does not alter its signal sensitivity (43). An initial
approach to identifying the region of AbsAl involved in signal sensing could implement
a series of in-frame deletions and amino acid substitutions. Unlike EnvZ and NarX,
AbsAl is predicted to contain four TM helices with relatively small external loops.
Deletion mutations might include the entire TM domain or the two central TM helices. If
the signal sensing region is contained within the deletions, I would expect an Abs'
phenotype consistent with constitutive kinase activity demonstrated in viva by truncations
of HKs with the signal-sensing region contained within the transmembrane domain of the
HK (170; 25). A recently identified HK from the S. coelicolor genome sequence of
cosmid St8D11 is 40% identical to absA! and shows unusual similarity in the C-terminus
of the normally highly divergent sensor domain. Striking within the alignment of this
151
region are 5 consecutive residues conserved in a predicted external solvent exposed loop
between TM helix 3 and TM helix 4, which is also the largest of the helical loops. It is
tempting to speculate that this highly conserved region represents a ligand-binding box
within the sensor domain. Site-directed amino acid substitutions could provide useful in
examining this region’s functional significance.
This study has broadened our knowledge of the AbsA two-component system and
allowed us to establish a working model which sets clear goals for future research.
Whether the AbsA system is part of a genetically-programmed developmental cycle or an
isolated component of an integrated network subject to external environmental stimuli
will become more clear as the targets and signal-sensing mechanism of AbsA are
elucidated. Finally, as approximately eighty novel two-component systems emerge from
the Streptomyces Sequencing Project, knowledge of the structural features and
mechanism of AbsA will surely provide insight into the function of other Streptomyces
two-component systems.
152
REFRENCES
153
10.
11.
Aceti, D. J., and W. C. Champness. 1998. Transcriptional regulation of
Streptomyces coelicolor pathway-specific antibiotic regulators by the absA and
absB loci. J Bacteriol. 180(12):3100-6.
Adamidis, T., P. Riggle, and W. Champness. 1990. Mutations in a new
Streptomyces coelicolor locus which globally black antibiotic biosynthesis but
not sporulation. J Bacteriol. 172(6):2962-9.
Agron, P. G., G. S. Ditta, and D. R. Helinski. 1993. Oxygen regulation of nifA
transcription in vitro. Proc Natl Acad Sci U S A. 90(8):3506-10.
Agron, P. G., and D. R. Helinski. 1995. Syrnbiotic expression of Rhizobium
melilati nitrogen fixation genes is regulated by oxygen, p. 275-287. In J. A. Hoch
and T. J. Silhavy (ed.), Two-Component Signal Transduction. ASM Press,
Washington, DC.
Aiba, H., F. Nakasai, S. Mizushima, and T. Mizuno. 1989. Evidence for the
physiological importance of the phosphotransfer between the two regulatory
components, EnvZ and OmpR, in osmoregulation in Escherichia coli. J Biol
Chem. 264(24): 14090-4.
Anderson, T., P. Brian, P. Riggle, R. Kong, and W. Champness. 1999. Genetic
suppression analysis of non-antibiotic-producing mutants of the Streptomyces
coelicolor absA locus [In Process Citation]. Microbiology. 145(Pt 9):2343-5 3.
Angel], 8., E. Schwarz, and M. J. Bibb. 1992. The glucose kinase gene of
Streptomyces coelicolor A3(2): its nucleotide sequence, transcriptional analysis
and role in glucose repression. Mol Microbiol. 6(19):2833-44.
Anisova, L. N ., I. N. Blinova, O. V. Efremenkova, P. Koz'min Iu, and V. V.
Onoprienko. 1984. [Regulators of development in Streptomyces coelicolor
A3(2)]. Izv Akad Nauk SSSR [Biol](1):98-108.
Atkinson, M. R., and A. J. Ninfa. 1993. Mutational analysis of the bacterial
signal-transducing protein kinase/phosphatase nitrogen regulator 11 (NRH or
NtrB). J Bacteriol. l75(21):7016-23.
Axelsson, L., and A. Holck. 1995. The genes involved in production of and
immunity to sakacin A, a bacteriocin from Lactobacillus sake Lb706. J Bacteriol.
l77(8):2125-37.
Baikalov, 1., 1. Schroder, M. Kaczor-Grzeskowiak, 1c Grzeskowiak, R. P.
Gunsalus, and R. E. Dickerson. 1996. Structure of the Escherichia coli response
regulator NarL. Biochemistry. 35(34): 1 1053-61.
154
12.
l3.
14.
15.
l6.
17.
18.
19.
20.
21.
22.
23.
Bascaran, V., L. Sanchez, C. Hardisson, and A. F. Brana. 1991. Stringent
response and initiation of secondary metabolism in Streptomyces clavuligerus. J
Gen Microbial. 137(Pt 7):]625-34.
Bassler, B. L., and M. R. Silverman. 1995. Intercellular communication in
marine Vibria species: density-dependent regulation of the expression of
bioluminescence, p. 431-445. In J. A. Hoch and T. J. Silhavy (ed.), Two-
Component Signal Transduction. ASM Press, Washington, DC.
Bibb, M. 1996. 1995 Colworth Prize Lecture. The regulation of antibiotic
production in Streptomyces coelicolor A3(2). Microbiology. 142(Pt 6):1335-44.
Bibb, M. J. personal communication
Bibb, M. J., P. R. Findlay, and M. W. Johnson. 1984. The relationship between
base composition and codon usage in bacterial genes and its use for the simple
and reliable identification of protein- coding sequences. Gene. 30(1-3):157-66.
Bibb, M. J., D. H. Sherman, S. Omura, and D. A. Hopwood. 1994. Cloning,
sequencing and deduced functions of a cluster of Streptomyces genes probably
encoding biosynthesis of the polyketide antibiotic frenolicin. Gene. 142(1):31-9.
Bourret, R. B., J. F. Hess, and M. 1. Simon. 1990. Conserved aspartate residues
and phosphorylation in signal transduction by the chemotaxis protein CheY. Proc
Natl Acad Sci U S A. 87(1):41-5.
Brian, P., P. J. Riggle, R. A. Santos, and W. C. Champness. 1996. Global
negative regulation of Streptomyces coelicolor antibiotic synthesis mediated by
an absA-encoded putative signal transduction system. J Bacteriol. 178(1 1):3221-
31.
Brissette, R. E., K. L. Tsung, and M. Inouye. 1991. Suppression of a mutation
in OmpR at the putative phosphorylation center by a mutant EnvZ protein in
Escherichia coli. J Bacteriol. l73(2):601-8.
Brown, D. P., L. Ganova-Raeva, B. D. Green, S. R. Wilkinson, M. Young,
and P. Youngman. 1994. Characterization of spoOA homologues in diverse
Bacillus and Clostridium species identifies a probable DNA-binding domain. Mal
IVIicrobiol. l4(3):411-26.
Burbulys, D., K. A. Trach, and J. A. Hoch. 1991. Initiation of sporulation in B.
subtilis is controlled by a multicomponent phosphorelay. Cell. 64(3):545-52.
Buttner, M. J., K. F. Chater, and M. J. Bibb. 1990. Cloning, disruption, and
transcriptional analysis of three RNA polymerase sigma factor genes of
Streptomyces coelicolor A3(2). J Bacteriol. 172(6):3367-78.
155
24.
25.
26.
27.
28.
29.
30.
31.
32.
33.
34.
Cashel, M., D. R. Gentry, V. J. Hernandez, and D. Vinella. 1996. The
Stringent Response, p. 1458-1496. In F. C. Neidhardt (ed.), Escherichia coli and
Salmonella Cellular and Molecular Biology, Second ed, vol. I. ASM Press,
Wahington, D. C.
Cavicchioli, R., R. C. Chiang, L. V. Kalman, and R. P. Gunsalus. 1996. Role
of the periplasmic domain of the Escherichia coli NarX sensar- transmitter protein
in nitrate-dependent signal transduction and gene regulation. Mol Microbiol.
21(5):901-11.
Chakraburtty, R., and M. Bibb. 1997. The ppGpp synthetase gene (relA) of
Streptomyces coelicolor A3(2) plays a conditional role in antibiotic production
and morphological differentiation. J Bacteriol. l79(18):5854-61.
Chakraburtty, R., J. White, E. Takano, and M. Bibb. 1996. Cloning,
characterization and disruption of a (p)ppGpp synthetase gene (relA) of
Streptomyces coelicolor A3(2). Mol Microbiol. l9(2):357—368.
Champness, W., P. Riggle, T. Adamidis, and P. Vandervere. 1992.
Identification of Streptomyces coelicolor genes involved in regulation of
antibiotic synthesis. Gene. 115(1-2):55-60.
Champness, W. C. personal communication.
Champness, W. C. 1999. Cloning and analysis of regulatory genes involved in
streptomycete secondary metabolite biosynthesis, p. 725-739. In A. L. Demain
and J. E. Davies (ed.), Industrial Microbiology and Biotechnology, Second ed.
ASM Press, Washington, D. C.
Champness, W. C., and K. F. Chater. 1994. Regulation of antibiotic production
and morphological differentiation in Streptomyces spp., p. 61-94. In P. Piggot, C.
P. Moran, and P. Youngman (ed.), Regulation of Bacterial Differentiation.
American Society for Microbiology, Washington, DC.
Chang, H. M., M. Y. Chen, Y. T. Shieh, M. J. Bibb, and C. W. Chen. 1996.
The cutRS signal transduction system of Streptomyces lividans represses the
biosynthesis of the polyketide antibiotic actinorhodin. Mol Microbiol.
21(5): 1075-85. '
Chater, K. F. 1972. A morphological and genetic mapping study of white colony
mutants of Streptomyces coelicolor. J Gen Microbial. 72(1):9-28.
Chater, K. F. 1998. Taking a genetic scalpel to the Streptomyces colony.
Microbiology. 144: 1465-1478.
156
35.
36.
37.
38.
39.
40.
41.
42.
43.
44.
45.
Chater, K. F., and M. J. Bibb. 1997. Regulation of Bacterial Antibiotic
Production, p. 57—105. In H. Kleinkauf and H. van Dohren (ed.), Products of
Secondary Metabolism, vol. 6. VCH, Weinheirn, Germany.
Chater, K. F., and C. J. Bruton. 1985. Resistance, regulatory and production
genes for the antibiotic methylenomycin are clustered. Embo J. 4(7): 1 893-7.
Chater, K. F., and R. Losick. 1997. The mycelial life-style of Streptomyces
coelicolor A3(2) and its relatives, p. 149-182. In J. H. Shapiro and M. Dworkin
(ed.), Bacteria as Multicellular Organisms. Oxford University Press, New York.
Chervitz, S. A., and J. J. Falke. 1996. Molecular mechanism of transmembrane
signaling by the aspartate receptor: a model. Proc Natl Acad Sci U S A.
93(6):2545-50.
Chang, P. P., S. M. Podmore, H. M. Kieser, M. Redenbach, K. Turgay, M.
Marahiel, D. A. Hopwood, and C. P. Smith. 1998. Physical identification of a
chromosomal locus encoding biosynthetic genes for the lipopeptide calcium-
dependent antibiotic (CDA) of Streptomyces coelicolor A3(2). Microbiology.
144(Pt 1):193-9. '
Cutting, S., and J. Mandelstam. 1986. The nucleotide sequence and the
transcription during sporulation of the gerE gene of Bacillus subtilis. J Gen
Microbiol. 132(Pt 11):3013-24.
Dahl, J. L., B. Y. Wei, and R. J. Kadner. 1997. Protein phosphorylation afi‘ects
binding of the Escherichia coli transcription activator UhpA to the-uhpT
promoter. J Biol Chem. 272(3):1910-9.
Dahl, M. K., T. Msadek, F. Kunst, and G. Rapoport. 1991. Mutational analysis
of the Bacillus subtilis DegU regulator and its phosphorylation by the DegS
protein kinase. J Bacteriol. l73(8):2539-47.
de Philip, P., E. Soupene, J. Batut, and P. Boistard. 1992. Modular structure of
the F ixL protein of Rhizobium melilati. Mol. Gen. Genet. 235:49-54.
Delgado, J ., S. Forst, S. Harlocker, and M. Inouye. 1993. Identification of a
phosphorylation site and fimctional analysis of conserved aspartic acid residues of
OmpR, a transcriptional activator for ompF and ompC in Escherichia coli. Mol
‘ Microbial. 10(5):1037-47.
Demain, A. L. 1989. Carbon Source Regulation of Idiolite Biosynthesis in
Actinomycetes, p. 127-134. In S. Shapiro (ed.), Regulation of Secondary
Metabolites in Actinomycetes. CRC Press, Boca Raton.
157
46.
47.
48.
49.
50.
51.
52.
53.
54.
55.
56.
57.
Demain, A. L., and A. Fang. 1995. Emerging concepts of secondary metabolism
in actinomycetes. Actinomycetologica. 9:98-117.
Diep, D. B., L. S. Havarstein, and I. F. Nes. 1996. Characterization of the locus
responsible for the bacteriocin production in Lactobacillus plantarum C11. J
Bacteriol. 178:4472-4483.
Elliot, M., F. Damji, R. Passantino, K. Chater, and B. Leskiw. 1998. The bldD
gene of Streptomyces coelicolor A3(2): a regulatory gene involved in
morphogenesis and antibiotic production. J Bacteriol. 180(6): 1549-55.
Ensign, J. C. 1988. Physiological Regulation of Sporulation of Streptomyces
griseus, p. 309-315. In Y. Okami, T. Beppu, and H. Ogawara (ed.), Biology of
Actinomycetes '88. Japan Scientific Societies Press, Tokyo.
Feitelson, J. S., F. Malpartida, and D. A. Hopwood. 1985. Genetic and
biochemical characterization of the red gene cluster of Streptomyces coelicolor
A3(2). J. Gen. Microbiol. 131:2431-2441.
Feng, J., M. R. Atkinson, W. McCleary, J. B. Stock, B. L. Wanner, and A. J.
N infa. 1992. Role of phosphorylated metabolic intermediates in the regulation of
glutamine synthetase synthesis in Escherichia coli. J Bacterial. 174(19):6061-70.
Fernandez-Moreno, M. A., J. L. Caballero, D. A. Hopwood, and F.
Malpartida. 1991. The act cluster contains regulatory and antibiotic export
genes, direct targets for translational control by the bldA tRNA gene of
Streptomyces. Cell. 66(4):769-80.
Fernandez-Moreno, M. A., A. J. Martin-Triana, E. Martinez, J. Niemi, H. M.
Kieser, D. A. Hopwood, and F. Malpartida. 1992. abaA, a new pleiotropic
regulatory locus for antibiotic production in Streptomyces coelicolor. J Bacteriol.
174(9):295 8-67.
Floriano, B., and M. Bibb. 1996. ast is a pleiotropic but conditionally required
regulatory gene for antibiotic production in Streptomyces coelicolor A3(2). Mol
Microbiol. 21(2):385-96.
Forst, S. A., and D. L. Roberts. 1994. Signal transduction by the EnvZ-OmpR
phosphotransfer system in bacteria. Res Microbial. 145(5-6):363-73.
Freeman, J. A., and B. L. Bassler. 1999. A genetic analysis of the function of
LuxO, a two-component response regulator involved in quorum sensing in Vibria
harveyi. Mol Microbiol. 31(2):665-77.
Fujii, T., H. C. Gramajo, E. Takano, and M. J. Bibb. 1996. redD and actII-
ORF4, pathway-specific regulatory genes for antibiotic production in
158
58.
59.
60.
61.
62.
63.
64.
65.
66.
67.
68.
Streptomyces coelicolor A3(2), are transcribed in vitro by an RNA polymerase
holoenzyme containing sigma hrdD. J Bacteriol. 178(11):3402-5.
Fuqua, W. C., S. C. Winans, and E. P. Greenberg. 1994. Quorum sensing in
bacteria: the LuxR-LuxI family of cell density- responsive transcriptional
regulators. J Bacteriol. l76(2):269-75.
Furuya, K., and C. R. Hutchinson. 1996. The DnrN protein of Streptomyces
peucetius, a pseudo-response regulator, is a DNA-binding protein involved in the
regulation of daunorubicin biosynthesis. J Bacteriol. l78(21):6310-8.
Grafe, U., G. Reinhardt, W. Schade, I. Erritt, W. F. Fleck, and L. Radics.
1983. Interspecific inducers of cytodifferentiation and anthracycline biosynthesis
from Streptomyces bikiniensis and S. cyaneafuscatus. Biotechnol Lett. 5:591-596.
Grafe, U., W. Schade, I. Eritt, W. F. Fleck, and L. Radics. 1982. A new
inducer of anthracycline biosynthesis from Streptomyces viridochromogenes. J
Antibiot (Tokyo). 35(12):l722-3. '
Gramajo, H. C., E. Takano, and M. J. Bibb. 1993. Stationary-phase production
of the antibiotic actinorhodin in Streptomyces coelicolor A3(2) is transcriptionally
regulated. Mol Microbiol. 7(6):837-45.
Grebe, T. W., and J. B. Stock. 1999. The histidine protein kinase superfamily.
Advances in Microbial Physiology. 41:139-227.
Grimsley, J. K., R. B. Tjalkens, M. A. Strauch, T. H. Bird, G. B. Spiegelman,
Z. Hostomsky, J. M. Whiteley, and J. A. Hoch. 1994. Subunit composition and
domain structure of the SpoOA sporulation transcription factor of Bacillus subtilis.
J Biol Chem. 269(24):]6977-82.
Guthrie, E. P., C. S. Flaxman, J. White, D. A. Hodgson, M. J. Bibb, and K. F.
Chater. 1998. A response-regulator-like activator of antibiotic synthesis fiom
Streptomyces coelicolor A3(2) with an amino-terminal domain that lacks a
phosphorylation pocket [published erratum appears in Microbiology 1998
Jul;l44(Pt 7):2007]. Microbiology. 144(Pt 3):727-38.
Hara, 0., S. Horinouchi, T. Uozumi, and T. Beppu. 1983. Genetic analysis of
A-factor synthesis in Streptomyces coelicolor A3(2) and Streptomyces griseus. J
Gen Microbiol. 129(Pt 9):2939—44.
Harrison, S. C., and A. K. Aggarwal. 1990. DNA recognition by proteins with
the helix-turn-helix motif. Annu Rev Biochem. 59:933-69.
Hoch, J. A. 1995. Control of Cellular Development in Sporulating Bacteria by the
Phosphorelay Two-Component Signal Transduction System, p. 129-144. In J. A.
159
69.
70.
71.
72.
73.
74.
75.
76.
77.
78.
79.
Hoch and T. J. Silhavy (ed.), Two-Component Signal Transduction. ASM Press,
Washington, D. C.
Hoch, J. A., and T. J. Silhavy (ed.). 1995. Two-Component Signal
Transduction. ASM Press, Washington, DC.
Holman, T. R., Z. Wu, B. L. Wanner, and C. T. Walsh. 1994. Identification of
the DNA-binding site for the phosphorylated VanR protein required for
vancomycin resistance in Enterococcus faecium. Biochemistry. 33(15):4625-31.
Hang, S. K., M. Kito, T. Beppu, and S. Horinouchi. 1991. Phosphorylation of
the Ast product, a global regulatory protein for secondary-metabolite formation
in Streptomyces coelicolor A3(2). J Bacteriol. l73(7):23l 1-8.
Hopwood, D. A. 1997. Genetic Contributions to Understanding Polyketide
Synthases. Chem. Rev. 97 :2465-2497.
Hopwood, D. A., M. J. Bibb, K. F. Chater, T. Kieser, C. J. Bruton, H. M.
Kieser, D. J. Lydiate, C. P. Smith, J. M. Ward, and H. Schrempf. 1985.
Genetic Manipulation of Streptomyces; a Laboratory Manual. The John Innes
Foundation, Norwich, United Kingdom.
Hopwood, D. A., K. F. Chater, and M. J. Bibb. 1995. Genetics of antibiotic
production in Streptomyces coelicolor A3(2), a model streptomycete.
Biotechnology. 28:65-102.
Horinouchi, S. 1996. Streptomyces genes involved in aerial mycelium formation.
FEMS Microbiol Lett. 141:1-9.
Horinouchi, S., O. Hara, and T. Beppu. 1983. Cloning of a pleiotropic gene that
positively controls biosynthesis of A-factor, actinorhodin, and prodigiasin in
Streptomyces coelicolor A3(2) and Streptomyces lividans. J Bacteriol.
155(3):1238-48.
Horinouchi, S., M. Kito, M. Nishiyama, K. Furuya, S. K. Hang, K. Miyake,
and T. Beppu. 1990. Primary structure of Ast, a global regulatory protein for
secondary metabolite formation in Streptomyces coelicolor A3(2). Gene.
95(1):49-56.
Horinouchi, S., F. Malpartida, D. A. Hopwood, and T. Beppu. 1989. afsB
stimulates transcription of the actinorhodin biosynthetic pathway in Streptomyces
coelicolor A3(2) and Streptomyces lividans. Mol Gen Genet. 215(2):355-7.
Hsing, W., F. D. Russo, K. K. Bernd, and T. J. Silhavy. 1998. Mutations that
alter the kinase and phosphatase activities of the two-component sensor EnvZ. J
Bacteriol. 180(17):4538-46.
160
80.
81.
82.
83.
84.
85.
86.
87.
88.
89.
90.
Hsing, W., and T. J. Silhavy. 1997. Function of conserved histidine-243 in
phosphatase activity of EnvZ, the sensor for porin osmoregulation in Escherichia
coli. J Bacteriol. 179(1 1):3729—35.
Huang, K. J., C. Y. Lan, and M. M. Igo. 1997. Phosphorylation stimulates the
cooperative DNA-binding properties of the transcription factor OmpR. Proc Natl
Acad Sci U S A. 94(7):2828-32.
Hutchinson, C. R. 1997. Antibiotics from Genetically Engineered
Microorganisms, p. 683-702. In W. R. Strohl (ed.), Biotechnology of Antibiotics,
Second ed. Marcel Dekker, Inc., New York.
Ishige, K., S. N agasawa, S. Tokishita, and T. Mizuno. 1994. A novel device of
bacterial signal transducers. Embo J. 13(21):5195-202.
Ishizuka, H., S. Horinouchi, H. M. Kieser, D. A. Hopwood, and T. Beppu.
1992. A putative two-component regulatory system involved in secondary
metabolism in Streptomyces spp. J Bacteriol. 174(23):7585-94.
James, P. D. A., C. Edward, and M. Dawson. 1991. The effects of temperature,
pH, and growth rate on secondary metabolism in Streptomyces thermavialaceus
grown in a chemostat. J. Gen. Microbiol. 133:1715-1720.
Janssen, G. R., and M. J. Bibb. 1993. Derivatives of pUCl8 that have BglII
sites flanking a modified multiple cloning site and that retain the ability to
identify recombinant clones by visual screening of Escherichia coli colonies.
Gene. 124(1):l33-4.
Jin, S., T. Roitsch, R. G. Ankenbauer, M. P. Gordon, and E. W. N ester.
1990a. The VirA protein of Agrobacterium tumefaciens is autophosphorylated
and is essential for vir gene regulation. J Bacteriol. l72(2):525-30.
Jin, S. G., R. K. Prusti, T. Roitsch, R. G. Ankenbauer, and E. W. N ester.
1990b. Phosphorylation of the VirG protein of Agrobacterium tumefaciens by the
autophosphorylated VirA protein: essential role in biological activity of VirG. J
Bacteriol. l72(9):4945-50.
Kadner, R. J., M. D. Island, J. L. Dahl, and C. A. Webber. 1994. A
transmembrane signalling complex controls transcription of the Uhp sugar
phosphate transport system. Res Microbial. l45(5-6):381-7.
~ Kahn, D., and G. Ditta. 1991. Modular structure of FixJ: homology of the
transcriptional activator domain with the -35 binding domain of sigma factors.
Mol Microbiol. 5(4):987-97.
161
90a.
91.
92.
93.
94.
95.
96.
97.
98.
99.
100.
101.
Karandikar, A., Sharples, G. P., and G. Hobbs. 1997. Differentiation of
Streptomyces coelicolor A3(2) under nitrate-limited conditions. Microbial. 143:
3581-3590.
Kendrick, K. E., and J. C. Ensign. 1983. Sporulation of Streptomyces griseus in
submerged culture. J Bacteriol. 155(1):357-66.
Khokhlov, A. S., I. I. Tovarova, L. N. Borisova, S. A. Pliner, L. A.
Schevchenko, E. Y. Kornitskaya, N. S. Ivkina, and I. A. Rapoport. 1967. A-
factor responsible for the biosynthesis of streptomycin by a mutant strain of
Actinomyces streptomycini. Dokl. Akad. Nauk SSSR. 177:232-235.
Kieser, H. M., T. Kieser, and D. A. Hopwood. 1992. A combined genetic and
physical map of the Streptomyces coelicolor A3(2) chromosome. J Bacterial.
l74( 17):5496—507.
Kinoshita, H., H. Ipposhi,-S. Okamoto, H. Nakano, T. Nihira, and Y.
Yamada. 1997. Butyrolactone autoregulator receptor protein (BarA) as a
transcriptional regulator in Streptomyces virginiae. J Bacteriol. 179(22):6986-93.
Kirby, R., and D. A. Hopwood. 1977. Genetic determination of methylenomycin
synthesis by the SCPl plasmid Streptomyces coelicolor A3(2). J. Gen. Microbiol.
98:239-252.
Kleerebezem, M., L. E. Quadri, O. P. Kuipers, and W. M. de Vos. 1997.
Quorum sensing by peptide pheromones and two-component signal- transduction
systems in Gram-positive bacteria. Mol Microbiol. 24(5):895-904.
Kleman-Leyer, K., D. W. Armbruster, and C. J. Daniels. 1997. Properties of
H. volcanii tRNA intron endonuclease reveal a relationship between the archaeal
and eucaryal tRNA intron processing systems. Cell. 89(6):839-47.
Kuipers, O. P., M. M. Beerthuyzen, P. G. de Ruyter, E. J. Luesink, and W.
M. de V08. 1995. Autoregulation of nisin biosynthesis in Lactococcus lactis by
signal transduction. J Biol Chem. 270(45):27299-304.
Kwak, J., L. A. McCue, and K. E. Kendrick. 1996. Identification of bldA
mutants of Streptomyces griseus. Gene. 17 1(1):75-8.
Lakey, J. H., E. J. Lea, B. A. Rudd, H. M. Wright, and D. A. Hopwood. 1983.
A new channel-forming antibiotic from Streptomyces coelicolor A3(2) which
requires calcium for its activity. J Gen Microbiol. 129(Pt 12):3565-73.
Lawlor, E. J., H. A. Baylis, and K. F. Chater. 1987. Pleiotropic morphological
and antibiotic deficiencies result from mutations in a gene encoding a tRNA-like
product in Streptomyces coelicolor A3(2). Genes Dev. 1(10): 1305-10.
162
102.
103.
104.
105.
106.
107.
108.
109.
110.
111.
Lee, A. I., A. Delgado, and R. P. Gunsalus. 1999. Signal-dependent
phosphorylation of the membrane-bound NarX two- component sensor-
transmitter protein of Escherichia coli: nitrate elicits a superior anion ligand
response compared to nitrite. J Bacteriol. l8l(17):5309-16.
Lee, S. H., and K. J. Lee. 1993. Aspartate aminotransferase and tylosin
biosyntheis in Streptomycesfradiae. Appl. Environ. Microbiol. 59:822-827.
Leskiw, B. K., E. J. Lawlor, J. M. Fernandez-Abalos, and K. F. Chater. 1991.
TTA codons in some genes prevent their expression in a class of developmental,
antibiotic-negative, Streptomyces mutants. Proc Natl Acad Sci U S A.
88(6):2461-5.
Li, J., S. Kustu, and V. Stewart. 1994. In vitro interaction of nitrate-responsive
protein NarL with DNA target sequences in the fdnG, narG, narK and frdA
operon control regions of Escherichia coli K-12. J. Mol. Biol. 241: 150-165.
Lilley, G., A. E. Clark, and G. C. Lawrence. 1981. Control of the production of
cephamycin C and thienamycin by Streptomyces cattleya NRRL 8057. J. Chem.
Technol. Biotechnol. 31:127.
Lukat, G. S., W. R. McCleary, A. M. Stock, and J. B. Stock. 1992.
Phosphorylation of bacterial response regulator proteins by low molecular weight
phospho-donors. Proc Natl Acad Sci U S A. 89(2):718-22.
Lynch, A. S., and E. C. Lin. 1996. Transcriptional control mediated by the ArcA
two-component response regulator protein of Escherichia coli: characterization of
DNA binding at target promoters. J Bacteriol. l78(21):6238-49.
Malpartida, F., and D. A. Hopwood. 1984. Molecular cloning of the whole
biosynthetic pathway of a Streptomyces antibiotic and its expression in a
heterologous host. Nature. 309(5967):462-4.
Malpartida, F., J. Niemi, R. N avarrete, and D. A. Hopwood. 1990. Cloning
and expression in a heterologous host of the complete set of genes for
biosynthesis of the Streptomyces coelicolor antibiotic undecylprodigiosin. Gene.
93(1):91-9.
Martin, J. F. 1989. Molecular Mechanisms for the Control by Phosphate of the
Biosynthesis of Antibiotics and Other Secondary Metabolites, p. 213-237. In S.
Shapiro (ed.), Regulation of Secondary Metabolism in Actinomycetes. CRC
Press, Boca Raton.
163
112.
113.
114.
115.
116.
117..
118.
119.
120.
121.
122.
Martin, J. F., and L. E. McDaniel. 1975. Specific inhibition of candicidin
biosynthesis by the lipogenic inhibitor cerulenin. Biochirn BiOphys Acta.
411(2):186-94.
Martinez-Costa, O. H., P. Arias, N. M. Romero, V. Parro, R. P. Mellado, and
F. Malpartida. 1996. A relA/spoT homologous gene from Streptomyces
coelicolor A3(2) controls antibiotic biosynthetic genes. J Biol Chem.
271(18):]0627-34.
Martinez-Costa, O. H., M. A. Fernandez-Moreno, and F. Malpartida. 1998.
The relA/spoT-homologous gene in Streptomyces coelicolor encodes both
ribosome-dependent (p)ppGpp-synthesizing and -degrading activities. J Bacteriol.
l80(16):4123-32.
Matsumoto, A., S. K. Hang, H. Ishizuka, S. Horinouchi, and T. Beppu. 1994.
Phosphorylation of the Ast protein involved in secondary metabolism in
Streptomyces species by a eukaryotic-type protein kinase. Gene. 146(1):47-56.
Matsumoto, A., H. Ishizuka, T. Beppu, and S. Horinouchi. 1995. Involvement
of a small ORF downstream of the ast gene in the regulation of secondqry
metabolism in Streptomyces coelicolor A3(2). Actinomycetologica. 9:37-43.
McCleary, W. R., and J. B. Stock. 1994. Acetyl phosphate and the activation of
two-component response regulators. J Biol Chem. 269(50):31567-72.
McCleary, W. R., J. B. Stock, and A. J. Ninfa. 1993. Is acetyl phosphate a
global signal in Escherichia coli“? J. Bacteriol. 175:2793-2798.
Meurer, G., and C. R. Hutchinson. 1999. Genes for the Biosynthesis of
Microbial Secondary Metabolites, p. 740-758. In A. L. Demain and J. E. Davies
(ed.), Manual of Industrial Microbiology and Biotechnology, Second ed. ASM
Press, Washington, D. C. '
Moore, J. B., S. P. Shiau, and L. J. Reitzer. 1993. Alterations of highly
conserved residues in the regulatory domain of nitrogen regulator I (NtrC) of
Escherichia coli. J Bacteriol. l75(9):2692-701.
Msadek, T., F. Kunst, and G. Rapoport. 1995. A signal transduction network in
Bacillus subtilis includes the DegS/DegU and ComP/ComA two-component
systems, p. 447-471. In J. A. Hoch and T. J. Silhavy (ed.), Two-Component
Signal Transduction. ASM Press, Washington, DC.
N arva, K. E., and J. S. Feitelson. 1990. Nucleotide sequence and transcriptional
analysis of the redD locus of Streptomyces coelicolor A3(2). J Bacteriol.
l72(1):326-33.
164
123.
124.
125.
126.
127.
128.
129.
130.
131.
132.
133.
134.
N es, 1. F., D. B. Diep, L. S. Havarstein, M. B. Brurberg, V. Eijsink, and H.
Halo. 1996. Biosynthesis of bacteriacins in lactic acid bacteria. Antonie Van
Leeuwenhoek. 70(2-4):1 13-28.
N eumann, T., W. Piepersberg, and J. Distler. 1996. Decision phase regulation
of streptomycin production in Streptomyces griseus. Microbiology. 142:1953-
1963.
Ninfa, E. G., M. R. Atkinson, E. S. Kamberov, and A. J. Ninfa. 1993.
Mechanism of autophosphorylation of Escherichia coli nitrogen regulator H (NRH
or NtrB): trans-phosphorylation between subunits. J Bacteriol. l75(21):7024-32.
N odwell, J. R., and R. Losick. 1998. Purification of an extracellular signaling
molecule involved in production of aerial mycelium by Streptomyces coelicolor. J
Bacteriol. 180(5): 1334-7.
Nodwell, J. R., K. McGovern, and R. Losick. 1996. An oligopeptide permease
responsible for the import of an extracellular signal governing aerial mycelium
formation in Streptomyces coelicolor. Mol Microbiol. 22(5):881-93.
Ochi, K. 1990a. A relaxed (rel) mutant of Streptomyces coelicolor A3(2) with a
missing ribosomal protein lacks the ability to accumulate ppGpp, A-factor and
prodigiosin. J Gen Microbiol. 136(Pt 12):2405-12.
Ochi, K. 1990b. Streptomyces relC mutants with an altered ribosomal protein ST-
L11 and genetic analysis of a Streptomyces griseus relC mutant. J Bacteriol.
172(7):4008-16.
Ohnishi, Y., S. Kameyama, H. Onaka, and S. Horinouchi. 1999. The A-factor
regulatory cascade leading to streptomycin biosynthesis in Streptomyces griseus :
identification of a target gene of the A-factor receptor. Mol Microbiol. 34(1):]02-
11.
Parkinson, J. S., and E. C. Kafoid. 1992. Communication modules in bacterial
signaling proteins. Annu Rev Genet. 26:71-112.
Pogliano, J., A. S. Lynch, D. Belin, E. C. Lin, and J. Beckwith. 1997.
Regulation of Escherichia coli cell envelope proteins involved in protein folding
and degradation by the Cpx two-component system. Genes Dev. ll(9):1169-82.
Pospiech, A., and B. Neumann. 1995. A versatile quick-prep of genomic DNA
from gram-positive bacteria. Trends Genet. ll(6):217-8.
Potuckova, L., G. H. Kelemen, K. C. Findlay, M. A. Lonetto, M. J. Buttner,
and J. Kormanec. 1995. A new RNA polymerase sigma factor, sigma F, is
165
135.
136.
137.
138.
139.
140.
141.
142..
143.
144.
145.
required for the late stages of morphological differentiation in Streptomyces spp.
Mol Microbiol. l7(1):37-48.
Pratt, L. A., and T. J. Silhavy. 1995. Porin regulation of Escherichia coli, p.
105-127. In J. A. Hoch and T. J. Silhavy (ed.), Two-Component Signal
Transduction. ASM Press, Washington, DC.
Price, B., T. Adamidis, R. Kong, and W. Champness. 1999. A Streptomyces
coelicolor antibiotic regulatory gene, absB, encodes an RNase III homolog. J
Bacteriol. 181(19):6142-51.
Quon, K. C., G. T. Marczynski, and L. Shapiro. 1996. Cell cycle control by an
essential bacterial two-component signal transduction protein. Cell. 84(1):83-93.
Raivio, T. L., D. L. Popkin, and T. J. Silhavy. 1999. The Cpx envelope stress
response is controlled by amplification and feedback inhibition. J Bacteriol.
181(17):5263-72.
Raivio, T. L., and T. J. Silhavy. 1997. Transduction of envelope stress in
Escherichia coli by the Cpx two- component system. J Bacteriol. 179(24):7724-
33.
Redenbach, M., H. M. Kieser, D. Denapaite, A. Eichner, J. Cullum, H.
Kinashi, and D. A. Hopwood. 1996. A set of ordered casmids and a detailed
genetic and physical map for the 8 Mb Streptomyces coelicolor A3(2)
chromosome. Mol Microbiol. 21(1):77-96.
Reyrat, J. M., M. David, C. Blonski, P. Boistard, and J. Batut. 1993. Oxygen-
regulated in vitro transcription of Rhizobium meliloti nifA and fixK genes. J
Bacteriol. l75(21):6867-72.
Romero, N. M., and R. P. Mellado. 1995. Activation of the actinorhodin
biosynthetic pathway in Streptomyces lividans. FEMS Microbiol Lett. 127(1-
2):79-84.
Romero, N. M., V. Parro, F. Malpartida, and R. P. Mellado. 1992.
Heterologous activation of the actinorhodin biosynthetic pathway in Streptomyces
lividans. Nucleic Acids Res. 20(11):2767-72.
Rudd, B. A., and D. A. Hopwood. 1980. A pigmented mycelial antibiotic in
Streptomyces coelicolor: control by a chromosomal gene cluster. J Gen
Microbiol. 119(Pt 2):333-40.
Russo, F. D., J. M. Slauch, and T. J. Silhavy. 1993. Mutations that affect
separate frmctions of OmpR the phosphorylated regulator of porin transcription in
Escherichia coli. J Mol Biol. 231(2):261-73.
166
146.
147.
148.
149.
150.
151.
152.
153.
154.
155.
156.
157.
Ryding, N. J. personal communication.
Sambrook, J., E. F. Fritsch, and T. Maniatis. 1989. Molecular Cloning: a
Laboratory Manual, 2nd ed. Cold Springs Harbor Laboratory Press, Cold Springs
Harbor, N. Y. -
Schroder, I., C. D. Wolin, R. Cavicchioli, and R. P. Gunsalus. 1994.
Phosphorylation and dephosphorylation of the NarQ, NarX, and NarL proteins of
the nitrate-dependent two-component regulatory system of Escherichia coli. J
Bacteriol. l76(16):4985-92.
Schwecke, T., J. F. Aparicio, I. Molnar, A. Konig, L. E. Khaw, S. F.
Haydock, M. Oliynyk, P. Caffrey, J. Cortes, J. B. Lester, and et al. 1995. The
biosynthetic gene cluster for the polyketide irnmunasuppressant rapamycin. Proc
Natl Acad Sci U S A. 92(17):7839-43.
Seno, E. T., and R. H. Baltz. 1989. Structural Organization and Regulation of
Antibiotic Biosynthesis and Resistance Genes in Actinomycetes. In S. Shapiro
(ed.), Regulation of Secondary Metabolites in Actinomycetes. CRC Press, Boca
Raton.
Shapiro, S. 1989a. Nitrogen Assimilation in Actinomycetes and the Influence of
Nitrogen Nutrition on Actinomycete Secondary Metabolism, p. 135-212. In S.
Shapiro (ed.), Regulation of Secondary Metabolites in Actinomycetes. CRC
Press, Boca Raton.
Shapiro, S. (ed.). 1989b. Regulation of Secondary Metabolism in Actinomycetes.
CRC Press, Boca Raton.
Shi, L., and F. M. Hulett. 1999. The cytoplasmic kinase domain of PhoR is
sufficient for the law phosphate-inducible expression of pho regulon genes in
Bacillus subtilis. Mol Microbiol. 31(1):211-22.
Shimoda, N., A. Toyoda-Yamamoto, S. Aoki, and Y. Machida. 1993. Genetic
evidence for an interaction between the VirA sensor protein and the ChvE sugar-
binding protein of Agrobacterium. J Biol Chem. 268(35):26552-8.
Simms, S. A., M. G. Keane, and J. Stock. 1985. Multiple forms of the CheB
methylesterase in bacterial chemosensing. J Biol Chem. 260(18): 10161-8.
Smith, C. P. personal communication
Spiegelman, G. B., T. H. Bird, and V. Voon. 1995. Transcription regulation by
the Bacillus subtilis response regulator SpoOA, p. 159-179. In J. A. Hoch and T. J.
167
158.
159.
160.
161.
162.
163.
164.
165.
166.
167.
Silhavy (ed.), Two-Component Signal Transduction. ASM Press, Washington,
DC.
Stewart, R. C., R. VanBruggen, D. D. Ellefson, and A. J. Wolfe. 1998. TNP-
ATP and TNP-ADP as probes of the nucleotide binding site of CheA, the
histidine protein kinase in the chemotaxis signal transduction pathway of
Escherichia coli. Biochemistry. 37(35): 12269-79.
Stewart, V., and R. S. Rabin. 1995. Dual sensors and dual response regulators
interact to control nitrate- and nitrite-responsive gene expression in Escherichia
coli, p. 233-252. In J. A. Hoch and T. J. Silhavy (ed.), Two-Component Signal
Tansduction. ASM Press, Washington, DC.
Stock, J. B., A. J. Ninfa, and A. M. Stock. 1989. Protein phosphorylation and
regulation of adaptive responses in bacteria. Microbiol Rev. 53(4):450-90.
Stock, J. B., M. G. Surette, M. Levit, and P. Park. 1995. Two-Component
Signal Transduction Systems: Structure-Function Relationships and Mechanisms
of Catalysis, p. 25-52. In J. A. Hoch and T. J. Silhavy (ed.), Two-Component
Signal Transduction. ASM Press, Washington, D. C.
Strauch, E., E. Takano, H. A. Baylis, and M. J. Bibb. 1991. The stringent
response in Streptomyces coelicolor A3(2). Mol Microbiol. 5(2):289-98.
Strohl, W. R. (ed.). 1997. Biotechnology of Antibiotics, Second ed, vol. 82.
Marcel Dekker, New York.
Strohl, W. R. 1992. Compilation and analysis of DNA sequences associated with
apparent streptomycete promoters. Nucleic Acids Res. 20(5):961-74.
Strohl, W. R. 1997. Industrial Antibiotics: Today and the Future, p. 1-48. In W.
R. Strohl (ed.), Biotechnology of Antibiotics, Second ed, vol. 82. Marcel Dekker,
Inc., New York.
Stutzman-Engwall, K. J., S. L. Otten, and C. R. Hutchinson. 1992. Regulation
of secondary metabolism in Streptomyces spp. and overproduction of
daunorubicin in Streptomyces peucetius. J Bacteriol. 174(1): 144-54.
Takano, E., H. C. Gramajo, E. Strauch, N. Andres, J. White, and M. J. Bibb.
1992. Transcriptional regulation of the redD transcriptional activator gene
accounts for growth-phase-dependent production of the antibiotic
undecylprodigiosin in Streptomyces coelicolor A3(2). Mal Microbial.
6(19):2797-804.
168
168.
169.
170.
171.
172.
173.
174.
175.
176.
177.
178.
179.
Takano, E., J. White, C. J. Thompson, and M. J. Bibb. 1995. Construction of
thiostrepton-inducible, high-copy-number expression vectors for use in
Streptomyces spp. Gene. l66(1):133-7.
Tanaka, T., M. Kawata, and K. Mukai. 1991. Altered phosphorylation of
Bacillus subtilis DegU caused by single amino acid changes in DegS. J Bacteriol.
173(17): 5507-15. .
Tokishita, S., A. Kojima, H. Aiba, and T. Mizuno. 1991 . Transmembrane
signal transduction and osmoregulation in Escherichia coli. Functional importance
of the periplasmic domain of the membrane-located protein kinase, EnvZ. J Biol
Chem. 266(11):6780-5.
Vallejo, A. N., R. J. Pogulis, and L. R. Pease. 1995. Mutagenesis and synthesis
of novel recombinant genes using PCR, p. 603-612. In C. W. Dieffenbach and G.
S. Dveksler (ed.), PCR Primer; A Labaraboratory Manual. Cold Spring Harbor
Laboratory Press, Plainview, NY.
Vining, L. C., and S. Shapiro. 1984. Chloramphenicol production in carbon-
lirnited media: efi'ect of methyl alpha-glucoside. J Antibiot (Tokyo). 37(1):74-6.
Vining, L. C., and C. Stuttard (ed.). 1995. Genetics and Biochemistry of
Antibiotic Production, vol. 28. Butterworth-Heinemann, Boston.
Vogtli, M., P. C. Chang, and S. N. Cohen. 1994. ast2: a previously tmdetected
gene encoding a 63-amino-acid protein that stimulates antibiotic production in
Streptomyces lividans. Mol Microbiol. 14(4):643-53.
Volff, J. N., and J. Altenbuchner. 1998. Genetic instability of the Streptomyces
chromosome. Mol Microbiol. 27(2):239-46.
Volz, K. 1995. Structure and Functional Conservation in Response Regulators, p.
53-64. In J. A. Hoch and T. J. Silhavy (ed.), Two-Component Signal
Transduction. ASM Press, Washington, D. C.
Vujaklija, D., S. Horinouchi, and T. Beppu. 1993. Detection of an A-factor-
responsive protein that binds to the upstream activation sequence of strR, a
regulatory gene for streptomycin biosynthesis in Streptomyces griseus. J
Bacteriol. 175(9):2652-6l.
Walker, M. S., and J. A. DeMoss. 1993. Phosphorylation and dephosphorylation
catalyzed in vitro by purified components of the nitrate sensing system, NarX and
NarL. J. Biol. Chem. 268:8391-8393.
Wanner, B. L. 1995. Signal Transduction and Cross Regulation inthe
Escherichia coli Phosphate Regulon by PhoR, CreC, and Acetyl Phosphate, p.
169
180.
181.
182.
183.
184.
185.
186.
187.
188.
189.
203-222. In J. A. Hoch and T. J. Silhavy (ed.), Two-Component Signal
Transduction. ASM Press, Washington, D. C.
Webber, C. A., and R. J. Kadner. 1995. Action of receiver and activator
modules of UhpA in transcriptional control of the Escherichia coli sugar
phosphate transport system. Mol Microbiol. 15(5):883-93.
Webber, C. A., and R. J. Kadner. 1997. Involvement of the amino-terminal
phosphorylation module of UhpA in activation of uhpT transcription in
Escherichia coli. Mol Microbiol. 24(5): 1039-48.
White, J., and M. Bibb. 1997. bldA dependence of undecylprodigiosin
production in Streptomyces coelicolor A3 (2) involves a pathway-specific
regulatory cascade. J Bacteriol. l79(3):627-33.
Wietzorrek, A., and M. Bibb. 1997. A novel family of proteins that regulates
antibiotic production in streptomycetes appears to contain an OmpR-like DNA-
binding fold [letter]. Mol Microbiol. 25(6): 1 181-4.
Willey, J., R. Santamaria, J. Guijarro, M. Geistlich, and R. Losick. 1991.
Extracellular complementation of a developmental mutation implicates a small
sporulation protein in aerial mycelium formation by S. coelicolor. Cell.
65(4):641-50.
Willey, J., J. Schwedock, and R. Losick. 1993. Multiple extracellular signals
govern the production of a morpha genetic protein involved in aerial mycelium
formation by Streptomyces coelicolor. Genes Dev. 7(5):895-903.
Williams, S. B., and V. Stewart. 1997. Discrimination between structurally
related ligands nitrate and nitrite controls autokinase activity of the NarX
transmembrane signal transducer of Escherichia coli K-12. Mol Microbiol.
26(5):91 1-25.
Wolfe, A. J., and R. C. Stewart. 1993. The short form of the CheA protein
restores kinase activity and chemotactic ability to kinase-deficient mutants. Proc
Natl Acad Sci U S A. 90(4):]518-22.
Yamada, Y., T. Nihira, and S. Sakuda. 1997. Butyrolactone Autoregulators,
Inducers of Virginiamycin in Streptomyces virginiae: Their Sturctures,
Biosynthesis, Receptor Proteins, and Induction of Virginiamycin Biosynthesis, p.
63-80. In W. R. Strohl (ed.), Biotechnology of Antibiotics, Second ed, vol. 82.
Marcel Dekker, New York.
Yamada, Y., K. Sugamura, K. Kondo, M. Yanagimoto, and H. Okada. 1987.
The structure of inducing factors for virginiamycin production in Streptomyces
virginiae. J Antibiot (Tokyo). 40(4):496-504.
170
190.
191.
192.
Yang, Y., and M. Inouye. 1992. Intermolecular compalementation between two
defective mutant signal-transducing receptors. Proc. Natl. Acad. Sci. USA.
88:1 1057-11061.
Yoshikawa, H., J. Kazami, S. Yamashita, T. Chibazakura, H. Sane, F.
Kawamura, M. Oda, M. Isaka, Y. Kobayashi, and H. Saito. 1986. Revised
assignment for the Bacillus subtilis spoOF gene and its homology with spaOA and
with two Escherichia coli genes. Nucleic Acids Res. 14(2): 1063-72.
Zuber, P., and M. A. Marahiel. 1997. Structure, Function, and Regulation of
Genes Encoding Multidomain Peptide Synthetases, p. 187-216. In W. R Strohl
(ed.), Biotechnology of Antibiotics, Second ed, vol. 82. Marcel Dekker, New
York.
171
"11111111111111.1000“