88
Isolation and characterization of SOS constitutive mutations in Escherichia coli. Item Type text; Dissertation-Reproduction (electronic) Authors Ossanna, Nina. Publisher The University of Arizona. Rights Copyright © is held by the author. Digital access to this material is made possible by the University Libraries, University of Arizona. Further transmission, reproduction or presentation (such as public display or performance) of protected items is prohibited except with permission of the author. Download date 07/07/2018 02:38:28 Link to Item http://hdl.handle.net/10150/184398

,U·M·I - Open Repositoryarizona.openrepository.com/arizona/bitstream/10150/184398/6/azu_td... · Dissertation Director Date . 3 ... This dissertation has been submitted in partial

Embed Size (px)

Citation preview

Isolation and characterization of SOSconstitutive mutations in Escherichia coli.

Item Type text; Dissertation-Reproduction (electronic)

Authors Ossanna, Nina.

Publisher The University of Arizona.

Rights Copyright © is held by the author. Digital access to this materialis made possible by the University Libraries, University of Arizona.Further transmission, reproduction or presentation (such aspublic display or performance) of protected items is prohibitedexcept with permission of the author.

Download date 07/07/2018 02:38:28

Link to Item http://hdl.handle.net/10150/184398

INFORMATION TO USERS

The most advanced technology has been used to photo­graph and reproduce this manuscript from the microfilm master. UMI films the original text directly from the copy submitted. Thus, some dissertation copies are in typewriter face, while others may be from a computer printer.

In the unlikely event that the author did not send UMI a complete manuscript and there are missing pages, these will be noted. Also, if unauthorized copyrighted material had to be removed, a note will indicate the deletion.

Oversize materials (e.g., maps, drawings, charts) are re­produced by sectioning the original, beginning at the upper left-hand corner and continuing from left to right in equal sections with small overlaps. Each oversize page is available as one exposure on a standard 35 mm slide or as a 17" x 23" black and white photographic print for an additional charge.

Photographs included in the original manuscript have been reproduced xerographically in this copy. 35 mm slides or 6" x 9" black and white photographic prints are available for any photographs or illustrations appearing in this copy for an additional charge. Contact UMI directly to order.

I' ,U·M·I Accessing the World's Information since 1938

300 North Zeeb Road, Ann Arbor, M148106-1346 USA

Order Number 8814266

Isolation and characterization of SOS constitutive mutations in Escherichia coli

Ossanna, Nina, Ph.D.

The University of Arizona, 1988

U·M·I 300 N. Zeeb Rd. Ann Arbor, MI 48106

ISOLATION AND CHARACTERIZATION OF SOS CONSTITUTIVE

MUTATIONS IN ESCHERICHIA COLI

by

Nina Ossanna

A Dissertation Submitted to the Faculty of the

DEPARTMENT OF MOLECULAR AND CELLULAR BIOLOGY

In Partial Fulfillment of the Requirements For the Degree of

DOCTOR OF PHILOSOPHY WITH A MAJOR IN MOLECULAR BIOLOGY

In the Graduate College

THE 'UNIVERSITY OF ARIZONA

1 988

1

THE UNIVERSITY OF ARIZONA GRADUATE COLLEGE

2

As members of the Final Examination Committee, we certify that we have read

the dissertation prepared by ---------------------------------------------Nina Ossanna

entitled Isolation and Characterization of SOS Constitutive -------------------------------------------------------------Mutations in Escherichia coli

and recommend that it be accepted as fulfilling the dissertation requirement

for the Degree of __ ~Ph~.D~. __ ~(~M_o~l~e_c_u_la_r __ B_i_o_l_o~g~y~) ________________________ _

Date

))f. John Little Date I I 17, /If??

Date

Date Dr. Marty

O~ Date Dr. Danny Brower

Final approval and acceptance of this dissertation is contingent upon the candidate's submission of the final copy of the dissertation to the Graduate College.

I hereby certify that I have read this dissertation prepared under my direction and r~commend that it be accepted as fulfilling the dissertation requirement.

Dissertation Director Date

3

STATEMENT BY AUTHOR

This dissertation has been submitted in partial fulfillment of requirements for an advanced degree at The University of Arizona and is deposited in the University Library to be made available to borrowers under rules of the Library.

Brief quotations from this dissertation are allowable without special permission, provided that accurate acknowledgment of source is made. Requests for permission for extended quotation from or reproduction of this manuscript in whole or in part may be granted by the head of the major department or the Dean of the Graduate College when in his or her judgment the proposed use of the material is in the interests of scholarship. In all other instances, however, permission must be obtained from the author.

4

AKNOWLEDGEMENTS

I would like to thank my committee members for their help during

this project and particularly for their assistance in writing this

dissertation. My committee included Dr. Danny Brower, Dr. Marty

Hewlett, Dr. June Ito and Dr. John Little.

I would like to thank my dissertation director, Dr. David Mount,

for the opportunity to work in his lab and his guidance to help me

become (in his words) a "true geneticist".

I have been fortunate during my time in David's lab to work with

fine people. The work contained in this dissertation would not have

been possible without their moral and scientific support, and I am

grateful. Particularly I wish to thank Mark Dubnick, Belva Fisher,

Nancy Istock, Ken Peterson, Andy Thliveris and Ken Wertman.

I would like to thank Cindy Parrill (fellow tech-nerd) for the

computer-generated graphics in this dissertation. My typist gets no

thanks--she could have done a better job.

Friends and family have helped enormously and I am grateful.

I am going to miss Tuesday lunches at McDonald's with Val. Particularly

I want to thank Anne Ryan for her support and encouragement throughout

this entire Ph.D. project. Also, Marcy, Phoebe and Willa for their

constant companionship even when there wasn't a milk bone in it for

them.

LIST OF ILLUSTRATIONS

LIST OF TABLES

ABSTRACT

1. INTRODUCTION

TABLE OF CONTENTS

..

Escherichia coli SOS response

The SOS inducing signal

Role of RecA protein

Inducing signals for RecA activation

Objective and rationale of this study

2. MATERIALS AND METHODS

Media ..

Materials

Plasmids, bacteriophage and strains

Strain used in mutant isolation

MNNG mutagenesis

STS test

Mapping mutant alleles

B-galactosidase assays

Spontaneous mutagenesis

UV survival assays

Recombination proficiency

5

page

8

9

11

12

12

13

14

15

19

21

21

21

22

29

29

30

31

32

32

33

33

6

TABLE OF CONTENTS--continued

3. RESULTS . . . . . . . . . . . . . . . . . . . . . . . . . . 34

4.

Results of mutagenesis . . . . . . . . . . . . . lOa mutant (lig252) . . . . . . . . . . . . . . . . . .

Mapping · . . . . . . . . . . . . . . . . . . . . SOS expression . . . . . . . . . . . . . . . . . . 1ig mutant phenotypes . . . . . . . . . . . . . .

9a mutant • • · . . . . . . . . . . . . . · . . . . . . 3e. 41 and 8a mutants (uvrD242. 243. 244) · . . . . . .

Mapping mutations in 3e, 41 and 8a · . . . . . . UV sensitivity of uvrD mutants . . . . . . . . . SOS expression in uvrD mutants . . . . . . . . . Spontaneous mutagenesis in uvrD mutants . . . . . uvrD spontaneous mutagenesis when SOS

induction is blocked •••••• . . . . uvrD mutagenesis in mismatch repair deficient

strains •• •• • • • • • • • • •

· . · .

Viability of uvrD244 rep5 double mutants . . . . . recA mutants · . . . . . . . . . . . . . . . . . . . .

SOS expression . . . . . . . . . . . . . . . . .

34

35

36

37

37

39

41

41

43

45

48

51

53

55

56

56

recA mutant recombination proficiency ••• • •• 56

Spontaneous mutagenesis in ~mutants •• 56

Viable and inviable mutations with recA81 and recAB 5 • • • • • • • • • • • • • • • • • • · . 60

DISCUSSION • • . . . . . . . . . . . . . . . . . . . . . . . 62

Isolation of SOS constitutive mutations . . . . . . . . 62

lig252 mutant • • • • • • • • • • • • • • • • • • • •• 66

7 .' r

TABLE OF CONTENTS--continued

dam17 mutant • • • • 67

uvrD mutants • • • • • • 69

~ mutants • • • • • • • 73

7. APPENDIX A Definitions of genetic terms • • • 76

6. REFERENCES . • • • · 77

LIST OF ILLUSTRATIONS

Figure

1. B-ga1actosidase expression from su1A::MuQ fusion in lig and dam strains . . . . . .

2. UV sensitivity of ~ mutants

3. B-galactosid~se expression from sulA::MuQ fusion in uvrD mutants . . . . . . . . . . .

4. Mitomycin C induced B-galactosidase expression from

8

page

38

44

46

sulA::MuQ fusion in uvrD mutants 49

5. Spontaneous mutagenesis in uvrD strains 50

6. B-galactosidase levels from sulA::MuQ fusion in mutant recA strains .... . . . . . 57

7. Spontaneous mutagenesis levels in recA mutants 59

LIST OF TABLES

Table

1. Bacteria, phage and plasmids used in this study

2. PI transductional mapping of mutant 10a(lig252)

3. PI transductional mapping of mutant 9a(daml7) .

4. Results of three factor crosses to map uvrD alleles

5. B-galactosidase expression from sulA::MuQ fusion in previously isolated mutants . . . . . . . . . . . .

6. Spontaneous mutagenesis in uvrD244 strain when SOS response blocked . . . . . . . . . . . . . .

7. Spontaneous mutagensis in uvrD strains when mismatch repair blocked . . . . . . . . . . . . . . . . .

8. Recombination proficiency in mutant recA strains

9

page

23

36

40

42

47

52

54

58

10

ABSTRACT

Early events occuring during induction of the SOS response in

Escherichia coli are poorly understood. In order to understand the

early steps in SOS induction more fully, we have isolated several

mutations which constitutively express the SOS regulon.

Using a Mug(ApR,lac) fusion to the SOS regulated gene sulA, we

isolated Lac+ colonies as mutants in which RecA protein is

constitutively activated for repressor cleavage. The mutations map to

four loci: darn, lig, uvrD and recA. The extent of constitutive SOS

induction in these mutants varied greatly, indicating different levels

or types of signal in the cell.

The mutations isolated demonstrate two early steps in SOS

induction. The first step in SOS induction is signal generation and

includes mutations found in darn, lig and uvrD genes. The. mutant gene

products presumably alter DNA metabolism to produce an inducing signal.

These non-lethal mutations lead to sub-induction and probably generate

very specific signals, such as abnormally unwound DNA in the case of DNA

helicase II mutants or unsealed DNA nicks that result from deficient

ligation in l!g mutants. Greater induction may require quantitatively

more signal or different types of signal generated by severe defects

leading to cell death. These mutations also show that signal is a

variable quantity, allowing the cell to fine tune the levels of SOS

repair activity according to the amount or type of signal (damage)

perceived. In some cases (such as darn mutations), blocking the SOS

response by lexA(Ind-) alleles leads to cell death. In this type of

constitutively activated strain, the increased level of repair from SOS

11

induction is required to allow the cell to tolerate potentially lethal

DNA structures generated by the mutant gene product. The second step in

induction is the interaction of signal with RecA protein and is shown by

isolating 8 recA mutants. Mutant recA alleles caused the strongest SOS

induction of any mutants obtained, similar to the level found in strains

lacking repressor (lexA(Def) mutants). This full induction in the

absence of lethal DNA damage underscores the pivotal role of RecA

protein in regulating the SOS response.

12

CHAPTER 1

INTRODUCTION

Escherichia coli SOS response

The SOS response is normally induced when Escherichia coli

suffers DNA damage from physical agents such as ultra-violet (UV) light

or mutagenic chemicals such as Mitomycin C (Witkin, 1976). Induction of

SOS functions results in a diverse set of responses such as an increased

capacity for DNA repair, enhanced mutagenesis (SOS mutagenesis),

inhibition of cell division, cessation of respiration, alleviation of

host-controlled restriction, induction of stable DNA replication and

prophage induction (Little and Mount, 1982).

Many of the approximately twenty genes induced during the SOS

response that participate in repair phenomena have been identified. The

major repair pathways induced are excision repair and RecF

recombinational repair (also known as daughter-strand gap repair,

Peterson et al., 1987). Excision repair requires the SOS-regulated

uvrA, uvrB, uvrC (whether uvrC is SOS regulated is unclear) and uvrD

gene products. Inducible RecF recombinational repair requires the

products of the SOS regulated recA, recN, recO, uvrD and ~ genes, as

well as recF (not SOS regulated), and recO and recJ whose regulation is

unknown. Inhibition of cell division, or filamentation, requires the

SOS regulated sulA gene. SOS mutagenesis requires the products of the

recA and umuC.D genes. Induction of prophage requires the recA gene

13

product. Induction of stable DNA replication (replication in the

absence of protein synthesis) requires a functional RecA protein (Witkin

and Kogoma, 1984; Kogoma et a1., 1985). Other genes induced by the SOS

response, called din (for damage inducible) genes, have been identified

but their functions are unknown (Kenyon and Walker, 1980). For

additional reviews concerning the SOS response see Walker (1984) and

Ossanna et a1. (1986).

At the molecular level, the SOS response is regulated by the

gene products of the recA and lexA genes. When DNA damage occurs, an

inducing signal is produced that alters RecA protein to an activated

form. This activated RecA promotes cleavage of LexA protein, a

repressor of the various genes of the SOS regu1on. Cleaved LexA protein

is unable to act as a repressor and as a result, the genes of the SOS

regu10n are expressed, enhancing the ability of the cell to recover from

DNA damage.

The SOS induction process is reversible. As DNA damage is

repaired the level of inducing signal drops and RecA protein is no

longer activated. LexA repressor can again accumulate and repression of

the SOS genes is reestablished. The reversibility of SOS induction is

crucial to cell survival since one SOS protein, SulA, causes the cell to

lethally filament if continually expressed (George et a1., 1975).

The SOS inducing signal

The nature of signal(s) and cellular alterations that generate

inducing signal(s) to activate RecA protein after DNA damage is unclear.

Many suggestions have been made, and all involve aspects of DNA

14

metabolism or structure. Biochemical studies (discussed below) suggest

that a possible inducing signal is single-stranded regions of DNA.

These regions may arise from blocks in DNA synthesis due to damage, or

gaps in double-stranded DNA opposite damaged regions (Rupp and Howard-'

Flanders, 1968; Roberts and Devoret, 1982). DNA degradation products,

such as oligonucleotides, resulting from DNase action on damaged DNA

have also been suggested as inducing signals (Smith and Oishi, 1978).

Alternatively, induction may result from alterations in the nucleotide

pools after DNA damage occurs (Das and Loeb, 1984), changes in

superhe1icity of DNA (Little and Mount, 1982), or a combination of any

of the above.

Early events in the cell that generate signal may not give rise

to the same types of signal. Genetic evidence supports different types

of signal generation that may be dependent upon the nature of the

cellular damage. Cells with mutations in recB or recC (Exonuclease V)

are not inducible with nalidixic acid, but are inducible by UV

irradiation. Strains deficient in recF are poorly inducible by UV light

and show normal induction when treated with nalidixic acid (McPartland

et a1., 1980; Little and Hanawalt, 1977).

Role of RecA protein

Extensive genetic evidence exists to support a central role for

RecA protein in SOS regulation as the target for the inducing signal.

The recA430 mutation drastically reduces the ability of RecA protein to

participate in LexA repressor cleavage without affecting its

recombinationa1 properties (Morand et a1., 1977). Another recA allele,

15

recA441 (formerly tif-1), makes a protein which is constitutively

activated when grown at 41°C in the absence of inducing treatment. The

presence of adenine further stimulates activation of RecA441 protein,

while a mixture of cytidine and guanosine inhibit its activity (George

et a1., 1975). Many RecA mutations, termed PrtC , have now been isolated

which show constitutive SOS induction at all temperatures of growth

(Witkin et a1., 1982; Wang and Tessman, 1986). Additionally, deletions

or null alleles of recA prevent induction of any SOS functions. These

observations indicate a central role for RecA protein as a positive

regulator of the SOS response.

As well as being the target for inducing signal(s), it is clear

that RecA protein must be altered to an active form in order to induce

the SOS response. Mutations that result in the overproduction of RecA

protein (recAoc , operator-constitutive mutations) do not lead to SOS

induction (Quillardet et al., 1982). Conversely, cleavage of LexA

repressor can take place if RecA protein is activated but amplification

does not occur (Little, 1983; Baluch et a1., 1980).

Further support for the regulatory role of RecA protein in SOS

induction comes from biochemical studies. The earliest measured event

after UV irradiation is the appearance of LexA protein cleavage

fragments. These fragments appear 1-2 minutes after irradiation and are

dependent upon RecA protein function (Little, 1983). It is believed

that signal interacts directly with RecA protein to effect activation,

leading to LexA cleavage.

16

Inducing signals for RecA activation

Biochemical studies have provided insight into the molecules

that interact with RecA protein and may be involved in activation

during SOS induction. The proteolytic function of RecA protein was

demonstrated in vitro by Craig and Roberts (1980). RecA protein was

shown to cleave lambda £1 repressor in a reaction requiring ATP and a

polynucleotide. Further studies showed that ATP1S (a weakly-hydrolyzed

ATP analog) satisfied the ATP requirement, indicating that hydrolysis of

the nucleotide was not necessary for cleavage activity, and that a

ternary complex of RecA protein, DNA and ATP was active for cleavage of

£1 repressor (Craig and Roberts, 1981). A comparison of RecA protein

and the more easily activated RecA441 protein demonstrated that the

RecA441 protein interacts with DNA and ATP more efficiently than wi1d­

type RecA protein (Phizicky and Roberts, 1981; McEntee and Weinstock,

1981). RecA protein was also shown to cleave purified LexA protein in

vitro, with requirements similar to £1 repressor of ATP and DNA for

activation (Little et a1., 1980; Horii et a1., 1981). These studies

suggest that a nucleotide cofactor and DNA may be important for

activation of RecA protein in inducing the SOS response.

An approach utilized to study factors effecting RecA protein

activation in vivo was to supply an inducing signal exogenously to

permeabi1ized cells (Irbe et a1., 1981). In this system, prophage ~80

induction was followed in the presence of different oligonucleotides

that mimick products of DNA degradation. The study was designed to test

the idea that recBC Dnase action produces oligonucleotides and is

responsible for generating the SOS inducing signal (Smith and Oishi,

17

1978). Induction occured in the presence of d(A-G), d(G-G), d(A-G-G)

and d(pG)8' while other similar molecules did not lead to ~80 induction.

While these studies indicated that some oligonucleotides may be signals

for SOS induction, the significance of the oligonucleotide specificity

is unclear.

Indirect induction results when UV irradiated replicons are

introduced into unirradiated cells. These types of experiments have

shed light on the relationship between replication of damaged DNA and

SOS induction. Introducing UV irradiated replicons such as F or F'

factors into cells induced SOS functions (George et al., 1974), as did

infection with irradiated Pl, lambda or M13 phages (D'Ari and Huisman,

1982). Irradiated lambda or Pl phage unable to replicate were

introduced into ~ coli. The ability of these non-replicating phage to

induce SOS functions was reduced, but not abolished, leading the authors

to suggest that two pathways or two signals exist to induce SOS

functions (D'Ari and Huisman, 1982). A possible signal for activation

in non-replicating phage may be the damaged DNA template. Lu et al.

(1986) found that RecA protein was better activated for LexA cleavage in

vitro by UV irradiated double-stranded DNA than unirradiated double­

stranded DNA.

Certain mutations are also capable of generating an SOS inducing

signal leading to prophage induction. These mutations lie in genes

whose products are involved in replication of DNA and provide insight

into what perturbations of DNA synthesis induce SOS functions. In

strains with a temperature sensitive DNA ligase activity, prophage

induction occurred when cells were grown at the non-permissive

18

temperature (Gottesman et al., 1973). Strains with temperature

sensitive alleles of two DNA polymerase genes, dnaE (e subunit of

polymerase III) and polA (polymerase I), induced prophage when grown at

non-permissive temperatures (Schuster et al., 1973; Blanco, and Pomes,

1977).

Casaregola et al. (1982) surveyed the effect of several dna

mutants on the induction of a recA::lacZ fusion and found that not all

mutations affecting DNA replication are capable of inducing the SOS

response. They showed that strains with a temperature sensitive allele

of dnaB (involved in initiation and elongation during DNA synthesis)

induced SOS functions when grown at the non-permissive temperature

(dnaB(Ts) strains also induced prophage, Schuster et al., 1973).

Strains with a dnaG(Ts) allele, which is the RNA primase for lagging

strand DNA synthesis, did not induce the ~::lac fusion when grown at

the non-permissive temperature (but induced prophage, see Schuster et

al., 1973). This mutation did increase the basal level of recA

expression at all temperatures, however. Strains with dnaA and dnaC

alleles, which are required for initiation of new rounds of DNA

replication, did not induce recA::lacZ when grown at elevated

temperature. Mitomycin C was capable of inducing these strains during

the residual replication occuring after a temperature shift, but when

this replication ceased the dnaC strain was no longer inducible.

Another gene involved in replication, ssb (single-stranded binding

protein), does not induce prophage when mutants are UV irradiated at the

non-permissive temperature (Baluch et al., 1980a).

Several features about the induction of the SOS response emerge

19

from the studies presented. The SOS system is induced in response to

DNA damage or perturbations of DNA replication. During induction, a

signal (or signals) is generated, possibly by different pathways. The

signal(s) interacts with RecA protein, altering it to an activated form.

This activated form of RecA protein is required for the in vivo cleavage

of LexA repressor, resulting in SOS induction. What actually

constitutes the signal(s) that interacts with RecA protein in vivo, how

it is generated and its relationship to DNA replication is not clear.

Objective and rationale of this study

The work undertaken for this dissertation was designed to

provide additional information about the early events surrounding

induction of the SOS response in ~ coli. I have taken a genetic

approach to this problem by isolating mutants in which RecA protein is

constitutively activated for the SOS response. These mutants are

altered in functions that are associated with generating signal during

induction of the SOS response.

Using a screen described below, I isolated and characterized

mutants in which an SOS-regulated gene is derepressed without ~ priori

knowledge of specific genes that may be involved in SOS induction.

These mutants constitutively expressed SOS functions in the absence of

an inducing treatment. Unlike previously identified SOS inducing

mutants (except for those in the known SOS regulatory genes recA and

lexA) , these were viable at all temperatures, causing induction under

non-lethal conditions. Because these mutants induced the SOS response

under non-lethal conditions, I anticipated that the mutations would have

20

more specific defects leading to SOS induction with fewer secondary or

unforseeable effects (such as would be expected in conditionally lethal

mutants that stall DNA replication).

In order to identify SOS constitutive mutants, I utilized a MUQ

fusion to the SOS regulated sulA gene. The MUQ fusion placed

promoterless Lac structural genes (lacZYA) under control of the sulA

promoter. This fusion was useful for a number of reasons. First, since

the sulA gene product is responsible for lethal filamentation, I needed

a sulA(Null) mutation so that SOS inducing mutants isolated would not

be lethal. Second, the sulA gene has one of the highest induction

ratios of the SOS regulated genes. This provided me with a large window

of sulA expression levels. Third, this operon fusion made it possible

to easily follow SOS induction by scoring Lac phenotype on MacConkey

lactose plates. Strains induced for the SOS response had a Lac+

phenotype (red colonies), while uninduced strains were Lac- (white

colonies). An additional advantage of this fusion was the ability to

quantitate expression of the sulA fusion colorimetrically by B­

galactosidase assays.

21

CHAPTER 2

MATERIALS AND METHODS

A glossary of genetic terms and definitions is in the Appendix.

Media

Strains were routinely grown in L broth containing 10 g Bacto­

tryptone (Difco), 5 g yeast extract (Difco) and 10 g NaCl per liter.

Solid media also contained 15 g Difco agar. Cells to be infected with

lambda were grown in Lambda broth, which contained 10 g tryptone (BBL) ,

5 g NaCl and' 2 g maltose per liter. Lambda plates also contained 10 g

Difco agar per liter. lXA minimal medium for B-galactosidase assays was

prepared according to Miller (1972) and supplemented with 0.5% casamino

acids (Difco) for strains with auxotrophic markers. LC media for PI

transductions was L media supplemented with 25 mM CaC12 , 0.005%

thymidine and 0.2% glucose. MacConkey agar (Difco) was prepared

according to the manufacturer's directions. Antibiotics, when required,

were used in the following concentrations: 20 ug/ml chloramphenicol, 25

ug/ml tetracycline, 80 ug/ml kanamycin, 100 ug/ml streptomycin, 300

ug/ml 2-aminopurine, 100 ug/ml ampicillin for plasmid selection or 25

ug/ml ampicillin for chromosomal markers and 100 ug/ml spectinomycin.

Unless otherwise specified Mitomycin C was used at 0.25 ug/ml.

Materials

N-methyl-N'-nitro-N-nitrosoguanidine (MNNG) and o-nitrophenyl-B­

D-galactopyranoside (ONPG) were purchased from Sigma Chemical.

22

Plasmids, Bacteriophage and Strains

Table 1 lists plasmids,oscteriophage and ~ coli strains used

in this work. All bacterial genetic nomenclature and map positions for

alleles are in accordance with the current version of Bachmann's ~ coli

genetic map (Bachmann, 1983). Strains were constructed by standard

Plvir transductions (Miller, 1972). Transductants were selected for on

the appropriate antibiotic media, which also contained 5 glliter sodium

citrate, or minimal media (plus citrate) when selecting an amino acid

marker. Selection for TnlO excision was by the Bochner method (Davis et

al., 1980), while precise excisions (TnlO and Tn2) were selected for on

the appropriate minimal medium or screened for on MacConkey indicator

media. Strains bearing MUQ(Amp,Lac) gene fusions were temperature

stabilized to the XCam derivative (Baker et al., 1983). The MuQXCam

phage contained a Tn2 insertion in the Mu h gene which prevents

replication (and therefore transposition). This stabilized the

MUQ(Amp,Lac) fusion by allowing growth at elevated temperatures. The

XCam conversion also conferred chloramphenicol resistance on the strain.

Backcrosses, as noted in the strain list (Table 1), were done to

move transposon-linked mutant alleles into unmutagenized backgrounds. A

transposon (TnlO with a known map location) was crossed into the mutant

strains by PI transduction. If the transposon was closely linked to the

mutant allele, the allele would be crossed out by the wild-type gene and

the phenotype change to Lac- according to the co-transduction frequency.

An isolate that was still Lac+ was saved and used as a PI donor to cross

the transposon-linked mutant allele into a wild-type parent, either N056

or N0120, again saving a Lac+ transductant.

23

Table 1. Bacteria, phage and p1asmids used in this study

Plasmid

pNS-1

pVMK35

Phage

A38

A103

A120

Alnd-

AClear

Strain

594

DE309

Markers

Markers

bl453 nin5

i 21 cI- biol Q73 S7

i 434 cI- bio10 029

wild type phage

cI(Ind-)

cI(Inds )

P1asmids

Phage

Source

Samuel Sch1agman U of Rochester

Sidney Kushner U of Georgia, Athens

Source

This work

Don Ennis

Don Ennis

Don Ennis

D. Mount lab

D. Mount lab

D. Mount lab

D. Mount lab

Escherichia coli K-l2 strains

Genetic markers

ssb-l, ma1E::TnlO

Source or reference

D. Mount lab

Don Ennis

Strain

DE310

DM1901

DM7012

JIA07

KP281

KP315

KP419

KP479

KW88

N14-4

N2668

N056

N057

N060

N061

N062

N063

N064

N073

N074

N075

Table 1., continued

Genetic markers

ssb-113, rna1E::TnlO

HfrH lacU169 re1A1 thi-l spr55 malB: :Tn.2. sulA: : MUQ(Ap ,lac) (Mu)

lexA3 zja505::TnI0 malB45

srl- 300: : TnlO

as N0120 but 1exA71::Tn~

dam13 : : Tn.2. cysG: : Tn~

recB268: :TnlO

mutL: :TnlO

thil pyrD34 his68 trp45 mtl2 xyl7 ma1Al galk35 strAl18 recB21 argA81::TnlO

uvrD3

ligts-7

HfrH(?) lacU169 re1Al thi-l cps3 malF59::Tn~ su1A::MuQX(Cam) (Mu)

as N056 but de1(sr1R-recA)306::Tn10

as N056 but srl-300::TnI0 recA80(2a)

as N056 but srl-300::TnI0 recA81(2b)

as N056 but srl-300::TnlO recA730

as N056 but srl-300::TnlO recA441

as N056 but Xi434cI(Ind-)lexA+

as N056 but srl-300::TnI0 recA82(lf)

as N056 but srl-300::TnI0 recA83(8g)

as N056 but srl-300::TnI0 recA84(6i)

Source or reference

Don Ennis

(Huisman and D'Ari, 1981)

D. Mount lab

John Little

24

Kenneth Peterson

Kenneth Peterson

Kenneth Peterson

Kenneth Peterson

Kenneth Wertman

D. Mount lab

Martin Gellert

This work

This work

This work1

This workl

This work

This work

This work

This workl

This workl

This workl

25

Table 1., continued

Strain Genetic markers Source or reference

N076 as NOs6 but srl-300::TnlO recA8s(7f) This workl

N077 as NOs6 but srl-300::TnlO recA86(6j) This workl

N078 as NOs6 but srl-300::TnlO recA87(lh) This workl

NOll2 as N2668 but z£b::TnlO (ligts7) This workS

NOl20 as NOs6 but mal+ (Kans ) This work

NOl2l as N076 but TetS (recA8s) This work2

NOl22 as N06l but TetS (recA8l) This work2

NOl23 as NOl20 but ssb-l, malE::TnlO This work3

NOl24 as NOl20 but ssbll3, malE:: TnlO This work4

NOl34 as NOl23 but TetS (ssb-l) This work2

NOl3s as NOl24 but TetS (ssbll3) This work2

NOl37 as NOl20 but lig2s2(lOa) z£b: : TnlO This workS

NOl38 as NOl20 but z£b::TnlO This workS

NOl39 as NOl20 but uvrD244(8a) metE: :TnlO This work6

NOl40 as NOl20 but metE::TnlO This work6

NOls7 as NOl20 but lig2s2(lOa) nupC: : TnlO This work7

NOls8 as NOl20 but nupC::TnlO This work7

NOl62 as NOl20 but ligts-7 nupC::TnlO This work8

NOl73 as NOl20 but uvrD244(8a) fad7s1::TnlO This work9

NOl84. as NOl39 but met+ TetS (uvrD244) This work1O

NOl90 as NOl20 but dam13:: Tn.2, cysG:: Tn,2. This workll

NOl97 as NOl20 but uvrD242(3e) metE::TnlO This work6

26

Table 1., continued

Strain Genetic markers Source or reference

N0198 as N0120 but uvrD243(41) metE: :TnlO This work6

N0199 as N0120 but metE- Tets This work12

N0217 as N0120 but daml7(9a) malT::TnlO This work13

N0219 as N0120 but daml3: : Tn2 malT: : TnlO This work

N0220 as N0120 but malT: :TnlO This work13

N0221 as N0140 but uvrD3 met+ This work14

N0222 as N0140 but met+ This work14

N0223 N0139/pVMK35 (uvrD244/uvrD+) Transformation

N0224 N0140/pVMK35 (uvro+;uvrD+) Transformation

N0225 N0197/pVMK35 (uvrD242/uvrD+) Transformation

N0226 N0198/pVMK35 (uvrD243/uvrD+) Transformation

N0230 as N0137 but Tets (lig252) This work2

N0248 as N0184 but lexA3 zjaSOS::TnlO malB45 This work15

(uvrD244)

N0249 as N0184 but del(srlR-recA)306: :TnlO This work16

(uvrD244)

N0250 as N0197 but met+ Tets (uvrD242) This work10

N0251 as N0198 but met+ Tets (uvrD243) This work10

N0266 as N0120 but mutL: :TnlO This work17

N0267 as N0184 but mutL: :TnlO (uvrD244) This work17

N0268 as N0250 but mutL: :TnlO (uvrD242) This work17

N0269 as N0251 but mutL: :TnlO (uvrD243) This work17

N0276 as N0120 but repS metE: : TnlO This work18

Table 1., continued

Strain Genetic markers

N0278 as N0276 but rnet+ (repS)

N0280

Source or reference

This work19

This work20

27

N0282

as N0276 but uvrD244 (repS)

N0232/pNS-1 (dam17/dam+) Transformation

PMS

RK4349

RS3087

S~1024

YMC

repS re1Al spoT1 thil

metE163::Tn10 pro3 entA403 his2l8 metB1 xy1 rpsL109 de1(lac)6 supE44

fad-7Sl::TnlO relAl spoTl thil HfrH

HfrH relAl rnetC69 spoTl thi-l nupCS10: : TnlO

supFS8

Barbara Bachmann

Barbara Bachmann

Barbara Bachmann

Barbara Bachmann

D. Mount lab

1Resu1t of backcross (described in text), srl-300::TnlO source was JL407.

2TnlO excision by Bochner selection.

3ssb- l source DE309.

4ssbl13 sourc~ DE3l0.

SResult of backcross, zfb::TnlO generated in this study.

6Resu1t of backcross, metE: : TnlO source was RK4349.

7Resu1t of backcross, nupC: : TnlO source was S~1024.

8ligts-7 source was N2668, nupC::TnlO source was S~1024.

9Resu1t of backcross, fad7Sl::tnlO source was RS3087.

10Se1ection for Met+ on minimal media.

11Donor was KP31S.

12Bochner selection of N0140.

l3Result of backcross, malT::TnlO source was TST3.

l4uvrD3 met+ source was N14-4.

l5Donor was DM70l2.

l6Donor was N057.

l7Donor was KP479.

l8rep5 donor was PM5, metE::TnlO donor was RK4349.

19Selection for Met+ on minimal media.

20Donor was N0184.

Escherichia coli Hfr strains

Point of Chromosome Selectable Hfr Origin movement marker

HfrH 98' clockwise leu: :TnlO

KL99 22' clockwise trp: : TnlO

KL14 68' clockwise rpsL (Str)

P4X 6' counter-clockwise leu: : TnlO

KL98 51' counter-clockwise cps3: :TnlO

KL209 90' counter-clockwise zih5l0: :TnlO

28

Map position

2'

28'

73'

2'

45'

89'

29

Strain used in mutant isolation

The su1A::Mug fusion was constructed by Huisman and D'Ari (1981)

using the Mug(Lac,Amp) phage of Casadaban and Cohen (1979). This

strain, DM1901, was made 1exA+ via P1 transduction. The mucoid

phenotype of the original strain made it very difficult to lysogenize or

plate lambda phage; therefore I introduced a cps3::Tn10 allele which

relieved the mucoid phenotype. The Tn10 was removed by Bochner

selection (described above) to render the strain TetS and allow further

use of Tn10 in mutation mapping. The Mug fusion was temperature

stabilized as described above using MugxCam conversion, resulting in the

strain N056. Since the inactivation of 1exA was a likely mutation to

arise and result in derepression of su1A, a 1exA diploid of N056 (N064)

was constructed by lysogenizing N056 with ~i434cI(Ind-)lexA+. Using

the heteroimmune i 434 phage still allowed the use of ~ in further

screenings.

MNNG Mutagenesis

Mutagenesis by MNNG treatment was similar to previously

described methods (Adelberg et a1., 1965). N064 was grown to mid-log

phase and resuspended in 0.1 M sodium citrate (pH 5.5). MNNG, dissolved

in 0.1 M citrate, was added to the cells at a final concentration of 100

ug/m1. The culture was incubated for 20 minutes at 37oC. The

mutagenized culture was diluted 1 to 100 in L broth, dispensed into 10

culture tubes (to isolate independent mutants) and grown overnight at

32oC. Cells were spread onto MacConkey lactose plates to screen for

Lac+ colonies.

30

The Lac+ mutants to be studied further were ~creened for two

types of interfering mutations. Although the strain was a lexA diploid,

I checked for the unlikely possibility that a lexA (Null) mutation was

responsible for Lac expression. Lysogens of the mutants were made with

~lexA+ phage, an allele dominant to lexA (Null) mutations. If the

isolate became white on MacConkey lactose plates (Lac-), the mutation

may have been in lexA and not interesting to this study. The other type

of mutants to be screened out were sulA operator mutants or Mug(Ap,Lac)

phage transpositions. To detect this type of mutation, mutant strains

were lysogenized with ~lexA3 phage. This phage provided a non-cleavable

repressor to check that the sulA::Mug fusion was still repressable by

LexA. In this screen, white (Lac-) lysogens on MacConkey lactose agar

indicated an intact sulA::Mug fusion and were kept for further study.

STS Test

The STS test was a rapid screen to determine which mutants were

capable of constitutively inactivating £1 repressor (Mount, 1979).

Since £1 repressor is more resistant to cleavage than LexA, this enabled

me to determine which mutants were most strongly induced for the SOS

response. The mutant colonies were patched onto maltose-containing L

plates and incubated for four hours (until patch growth was barely

perceptible). These patch plates were replica plated sequentially onto

maltose-containing Lambda plates spread with ~cI(Ind-), ~cI(Ind+),

~cI(Inds) and finally ~cI-(Clear). This order of replica plating was

important so that phage with repressors less susceptible to cleavage

would not be carried over to a plate seeded with phage more likely to

31

grow 1ytica11y and interfere with the test results. Patch growth on any

of the phage seeded plates indicated lysogenization of the phage, not

constitutive cleavage of repressor. Lack of patch growth indicated

lytic growth of the phage (resulting in cell death) and constitutive

repressor cleavage. The Inds mutant produces a repressor that is more

easily cleaved than wild-type (Cohen et a1., 1981; Crowl et a1., 1981).

Ind- and clear plaque phage mutants were used as positive and negative

controls.

Mapping mutant alleles

Hfr matings for mapping mutations were performed according to

standard procedures (Miller, 1972). Hfr strains constructed for mapping

are shown at the end of Table 1. Since the recipient strain (mutant

N056) was an Hfr, these cultures were grown to saturation overnight at

320 C to discourage pilus formation. Donors were grown to mid-log phase

at 370 C and added to recipient cultures at one-tenth volume. Matings

were disrupted after 20 minutes by vortexing and an equal volume of L

broth containing 40 ug/m1 chloramphenicol was added and the cultures

allowed to grow out for one hour before plating. Recombinants were

selected on tetracycline chloramphenicol MacConkey lactose plates, since

donor strains mobilized an early Tn10 (Tet marker) and recipient strains

contained Tn2 (Cam marker). MacConkey plates allowed me to distinguish

between Lac+ and Lac- colonies and determine if the given Hfr crossed

out the mutant allele. This gave me a rough estimate of the map

location for a mutant allele.

Map location to the minute was determined by PI transductions.

P1 mapping is described with backcrosses in the section entitled

P1asmids, Bacteriophage and Strains.

B-ga1actosidase Assays

32

B-ga1actosidase assays were performed according to the method of

Miller (1972). Overnight cultures grown in the presence of chloramphen­

icol were diluted back 1/200 in 1XA media and grown to early log phase

(Klett 15). The cultures were split and Mitomycin C was added (or not

added); cultures were allowed to grow for an additional two hours (to

mid-log), and the assays were performed. For each assay two samples

were taken and each experiment was done two or more times. Unless

specifically noted the cultures were grown at 37oC. The B-ga1actosidase

results are expressed as Miller units.

Spontaneous Mutagenesis

Single colonies were used to start overnight cultures in L

broth. The overnights were diluted 1/40 in L broth and grown to 1ate­

log phase. The cells were concentrated by centrifugation, resuspended

in 0.85% saline and plated on streptomycin-containing L plates. Total

cell numbers were determined by plating appropriate dilutions on non­

antibiotic L plates. All incubations were done at 37oC. Mutation

frequency was measured in at least three independent cultures. In two

cases jackpot cultures were found and the values not used.

UV Survival Assays

Cultures were grown to mid-log phase (3 X 108 ce11s/m1) in L

broth, diluted 1/100 in 0.85% saline and irradiated. Appropriate

33

dilutions were plated on L plates and incubated overnight at 370 C to

determine survival.

Recombination Proficiency

Recombination proficiency was assayed using lambda by lambda

. crosses. The phage used, Xl03 and X120, were deficient in recombination

due to red deletions. Also, they were unable to replicate in sup+

strains due to amber mutations. The test strains were co-infected with

both phage at an moi (multiplicity of infection) of 5 and incubated 1.5

hours to allow phage recombination and growth. The cells were killed

with chloroform and reSUlting phage titers determined on supF and sup+

strains. Only recombinant phage in which both amber markers were

crossed out plated on the sup+ strain. The percentage of recombinants

was determined by:

titer on sup+ strain titer on supF strain

34

CHAPTER 3

RESULTS

Results of mutagenesis

Approximately 40,000 colonies were screened for Lac+ phenotype

following MNNG mutagenesis of strain N064. I initially picked 116

strong Lac+ colonies as possible SOS constitutive mutants. Upon

restreaking these colonies, 84 had retained the Lac+ phenotype and the

remainder were discarded. The 84 Lac+ colonies were tested for the

ability to cleave h£I repressor, an in vivo indication of activated RecA

protein (STS test). Since the £1 repressor is more resistant to

cleavage than LexA, I had hoped that in conjunction with Lac+ phenotype

I could identify those mutants in which the SOS response was more

strongly induced. The STS screen identified 10 mutants with differing

abilities to cleave lambda repressor, while the remainder showed no

evidence of hcI cleavage by this test.

By P1 transduction analysis, eight of the ten mutants able to

cleave lambda repressor were found to reside in the recA gene (described

below). Strikingly, the two mutants that were not in recA had only a

marginal ability to cleave h repressor constitutively based upon plaque

morphology of hInd+ and the more easily cleaved hIndS. One of these

mutants mapped to the dam locus and is described in a later section. I

was not able to map the second mutation, and believe that there was more

than one mutation responsible for the Lac+ phenotype. Thus, the STS

test did unveil the most strongly induced mutants, and these were

exclusively recA mutations. Perhaps other mutations that altered

cellular metabolism to the extent that activated RecA could cleave

lambda repressor were lethal.

35

Several other mutants (of the 76/84 that did not cleave cI

repressor) were chosen for further study. These isolates had strong

Lac+ phenotypes and appeared stable. The mutants were screened for lexA

or suIA::MuQ fusion mutations as described in Materials and Methods.

Those mutants that checked out were roughly mapped on the ~ coli

chromosome using Hfrs (described in Materials and Methods). When a

given Hfr crossed out a mutant allele the strain phenotype changed from

Lac+ to Lac-. Once the mutations were roughly mapped, I mapped more

precisely using PI transduction of various known TnlO markers in that

area. Again, if the mutation was near the TnlO, a percentage

(transduction frequency) of transductants became phenotypically Lac- .

As well as providing map location to the minute, this provided a

selectable marker close to the mutation. Thus I could backcross the

mutation into a clean background (i.e., unmutagenized N056 or N0120) in

order to characterize the effect of the mutation. The mutants are each

discussed in detail below.

lOa mutant (lig252)

Mapping

Mapping data for the lOa mutant is shown in Table 2. This

mutation mapped to 52' on the ~ coli chromosome. When ligts7 was

crossed into lOa, no wild-type recombinants were recovered, indicating

36

Table 2. P1 transductiona1 mapping of mutant lOa (lig252). Donor strains were N0137 (zfb::Tn10, 1ig252), NOl12 (zfb::Tn10, 1igts7), N0157 (nupC::Tn10, 1ig252), N0162 (nupC::Tn10, ligts7). The recipient strain was N0120 for the first four transductions and N0230 (lig252, Tets ) for the last.

selected unse1ected total # of marker marker transductants

zfb: :Tn10a lig252°(10a) 31 zfb: : Tn10 1igts7b 32

nupC: : Tn10c lig252(10a) 168 nupC: :Tn10 1igts7 48 nupC: : Tn10d llgts7 59

a Tn10 generated in this study. b llgts7 source was strain N2668.

# of unse1ected co-transd. marker frequency(%)

19 61 18 56

120 71 37 77 41 69

c nupC::Tn10 source was strain S~1024. d The recipient in this cross was a 1ig252 mutant in an attempt to

recover wild-type recombinants. None were found.

37

that these two mutations lie very close together. This observation,

along with the linkage frequencies, led me to conclude that the mutation

in lOa resided in the 1ig gene, which codes for DNA ligase. The allele

was designated 1ig252.

SOS expression

The extent of SOS induction in mutant lOa is shown in Figure 1.

This mutation resulted in a seven-fold increase of su1A expression,

indicating that this !ig mutation increases the basal level of SOS

expression. This increase is perhaps not surprising, since condition­

ally lethal 1ig mutants induce prophage at non-permissive temperatures

(Gottesman et a1., 1973). This result does, however, show that 1ig

mutants can increase SOS expression under non-lethal conditions of cell

growth.

lig mutant phenotypes

Other observations indicate that mutant lOa is a lig mutation.

If lOa were a 1ig mutation, I expected Spi- lambda phage to grow very

poorly on the mutant host (Smith, 1983). Therefore, I did a burst size

experiment with mutant lOa. The phage us~a, ~38, has a b1453 deletion

which removes the red and gam genes and confers the Spi- phenotype.

After two hours of growth in isogenic lig+ and lig252 strains the burst

sizes were 64 and 4 phage, respectively, per input phage. This small

burst size indicates the growth of ~Spi- was much poorer in the strain

bearing the lig252 allele. characteristic of lig mutants.

Since recB mutations inhibit SOS induction by nalidixic acid

(Gudas and Pardee, 1975), I wanted to test whether recB would inhibit

CD 1/1

1750

1500

.g 1250 -;; o -u .g 1000 01 I

m !.! 750 C :::J

"-CD 500

SE

250

wt 110252 daml7 ~ daml7/ Mutant allele dam+

LEGEND

~-mmc D+mmc

38

Figure 1. B-ga1actosidase expression from sulA::MuQ fusion in lig and dam strains. Assays were performed as described in Materials and Methods. Mitomycin C was added to O.25ug/m1. Strains used were as follows: wt(N0137 or N0220), 1ig252(N0138), dam17(N0217), dam13(N0219) and dam17/dam+(N0282).

39

the high basal level of induction seen in the lOa mutant. I was unable

to construct the li&252 double mutant using recB21 or recB::TnlO as the

mutant donors. In the case of recB2l, ar&::TnlO (KW88 was the donor and

N0230 the li&252 recipient) 0/40 Tetr transductants screened were RecB- .

The normal co-transduction frequency for these markers is 65%. In

attempting to make the lig252 recB::TnlO double mutant (KP4l9 was the

recB::TnlO donor), 0 Tetr transductants were obtained in the N0230

recipient, while 300 Tetr transductants were obtained in a control

transduction into the wild-type N0120 strain. The 1i&252 recB

inviability may be characteristic of !1g mutants as an earlier study

showed that li&4 recB21 double mutants were barely viable at 420 C

(Gottesman et al., 1973).

9a mutant (dam17)

This mutant was one of those shown in the early screening to

have a slight ability to cleave £1 repressor, in addition to a strong

Lac+ phenotype. Subsequently, this mutation was mapped at 74' on the

~ coli linkage map (see Table 3) and had a similar linkage to

malA: :Tn10 as dam13: : Tn,2,.

Mutant 9a demonstrated several characteristics similar to those

of strains bearing dam mutations. First, when screening for su1A::MuQ

fusion mutations, I was unable to lysogenize 9a with ~lexA3, indicating

that the mutation was inviable with 1exA3. dam strains are a1'so

inviable with lexA3 mutations (Marinus and Morris, 1975). Second, the

9a strain was sensitive to 2-aminopurine, as are dam mutants. Third,

the basal level of SOS expression in this mutant was quite high and

40

Table 3. PI transductional mapping of mutant 9a(dam17). Donor strains were N02l7(malA::TnlO, dam17) or N02l9(malA::TnlO, dam13). The recipient strain was N0120.

selected marker

unselected marker

total # of transductants

# of unselected co-transd.

malA: :TnlOa

malA: : TnlO 120

60

a malT45::TnlO (in MalA operon) source was TST3. b dam13::Tn2 source was KP3l5.

marker frequency(%)

8 7 5 8

41

approximately l3-fold greater than wild-type (see Fi~ure 1). For

comparison, sulA expression in an isogenic dam13::Tn1 strain was

measured and shown to be nearly identical to 9a. Previous work had also

shown that da~ mutations increased the basal level of SOS expression

(Craig et al., 1983; Peterson et al., 1985). Finally, construction of a

partial diploid strain by introducing dam+ (on plasmid pNS-l) into 9a

restored sulA expression to near wild-type levels (see Figure 1) and

alleviated 2-aminopurine sensitivity. This confirmed that the 9a

mutation was a darn allele. The 9a mutation was designated dam17.

3e, 41. and 8a mutants (uvrD242, 243, 244)

Mapping mutations in 3e, 41 and 8a

These mutants were distinguished from the others by their

sensitivity to UV light. The mutations all mapped to the same region,

at 85' on the linkage map. Results of a three factor cross are shown in

Table 4. A three factor cross was also included with a known uvrD

mutation, uvrD3, to compare linkages and map order. The order of the

markers on the ~ coli chromosome was shown to be fad met uvrD, in

agreement with the published genetic map (Bachmann, 1983). The UV

sensitivity of mutants 3e, 41 and 8a and their map position suggested

they were alleles of uvrD, the gene which encodes DNA helicase II.

Consequently the mutations were designated as uvrD242 for 3e, uvrD243

for 41 and uvrD244 for 8a.

Previously isolated mutants of uvrD have a variety of

phenotypes, differing from one allele to another, which were useful in

designing further experiments. The uvrD3 allele was isolated by Ogawa

42

Table 4. Results of three factor crosses to map uvrD alleles. The donor strains were fad::TnlO and uvrD 8a(N0173), 3e(N0227), 41(228) or D3(N0229), with selection for fad::TnlO. The recipient strain was met -(N0199).

Linkages(%)

Number of % of fad/ met/ fad/ Possible Markers Transductants Total met uvrD uvrD

fad: :TnlO, uvrD3, met+ 12 22 80 27 22 fad: :TnlO, uvrD3, met 0 0 fad: :TnlO, uvrD+, met+ 32 56 fad: :TnlO, uvrD+, met 11 20

fad: :TnlO, uvrD8a, met+ 33 28 84 51 28 fad: :TnlO, uvrD8a, met - 1 0.9 fad: :TnlO, uvrD+ met+ 64 55 --' --fad: :TnlO, uvrD+, met 18 16

fad: : TnlO , uvrD3e, met+ 16 30 78 39 21 fad: :TnlO, uvrD3e, met - 1 2 fad: :Tn10, uvrD+ met+ 25 56 --, --fad: :TnlO, uvrD+, met 10 19

fad: :TnlO, uvrD4l, met+ 30 31 64 48 31 fad: :TnlO, uvrD41, met - 0 0 fad: :TnlO, uvrD+ met+ 32 33 --, --fad: :TnlO, uvrD+, met- 35 36

43

(1968) as a mutant sensitive to UV light and methyl methanesulfonate. A

spontaneous mutator and UV sensitive mutation was isolated by Siegel

(1973) and designated mutU4. A similar mutant was isolated

independently as uvrE (Smirnov and Skavronskaya, 1971). While looking

for recombination deficient mutations in recBC sbcB strains Horii and

Clark (1973) isolated the recL152 mutant, which is not a spontaneous

mutator. Subsequently, these mutations have all been found to reside in

a single locus, uvrD, and share the common phenotype of UV sensitivity.

The product of the uvrD gene plays roles in several pathways of DNA

metabolism that include the SOS regulated repair functions excision

repair and RecF recombination repair, as well as methyl-directed

mismatch repair. In addition, uvrD is regulated by the SOS response

(Siegel, 1983). Identification of DNA helicase II as the product of the

uvrD gene uncovered its role in DNA replication as one of the known ~

coli DNA helicases that unwind duplex DNA (Maples and Kushner, 1982;

Oeda et al., 1982). The mutants were screened for some of these known

uvrD phenotypes in an attempt to understand which altered properties

were associated with SOS induction.

UV sensitivity of uvrD mutants

The UV sensitivity of strains bearing uvrD242, 243 and 244 is

shown in Figure 2. In order to show definitively that these mutations

did lie in uvrD, I constructed merodiploids of each mutant with a low­

copy number uvrD+ plasmid, pVMK35. A low copy number plasmid was used

for these complementation studies because overproduction of the wild­

type UvrD protein causes UV sensitivity (Map1es and Kushner, 1982).

44

100 uvrD allele

+

+/+ P e 242/+

r 243/+

C 244/+ e n 10 t

s u r v i v a 1 1 242

244

243

0.1--------------------------~-------o 4.5 9.0 18 27

UV dose ~Joules/square meter)

Figure 2. UV sensitivity of ~ mutants. UV survival assays were performed as described in Materials and Methods. Strains used were as follows: uvrD+(N0140), uvrD244(N0139), uvrD242(N0197), uvrD243(N0198), uvrD+/uvrD+(N0224), uvrD244/uvrD+(N0223), uvrD242/uvrD+(N0225)and uvrD243/uvrD+(N0226).

45

As can be seen in Figure 2, uvrD+ largely complemented the UV

sensitivity in the mutant strains. This result indicates that the

mutations do lie in the uvrD gene and that the repair defect in these

strains is recessive and can be complemented.

SOS expression in uvrD mutants

The basal level of SOS expression in the uvrD mutants is shown

in Figure 3. sulA expression is increased five to seven-fold in these

strains. The basal levels of two previously identified uvrD alleles,

uvrD3 and uvrDl56 (formerly mutU4) are shown in Table 5. The uvrD3

mutation did not lead to any increase in the basal level of SOS

expression, indicating that it is a different mutation from those

isolated in this study. This result also shows that a repair defect,

per se, does not increase sulA expression. In contrast, uvrDl56

increased sulA expression approximately ten-fold, indicating the new

uvrD mutations may be of a similar class.

Complementing the uvrD alleles isolated in this study with the

wild-type uvrD gene did not alleviate SOS induction, indicating there is

a dominant or incompletely dominant defect in the newly isolated

mutations. This dominance is interesting since the UV sensitivity of

the uvrD mutants was shown to be recessive (see previous section and

Figure 2). This result shows that the new uvrD alleles are not null

alleles and that the protein product still functions in some capacity.

The uvrD- strains containing the uvrD+ plasmids were unstable.

This was particularly true of uvrD243, strain N0226. Standard devia­

tions in B-galactosidase units were approximately 5% in the uvrD strains

Q) UJ 0

-0 • iii 0 -(,) 0 0 Ol I

CD

UJ

== c ::J

~ Q)

:::::!!

400

350

300

250

200

150

100

50

0 + 244 242 243 +/+ 244/+ 242/+ 243/+

uvrD allele

46

Figure 3. B-galactosidase expression from sulA::Mug fusion in uvrD mutants. Assays were performed as described in Materials and Methods. B-galactosidase expression in a fully induced (lexA7l) strain is 8000 Units. Strains used are as follows: uvrD+(NOl40), uvrD244(NOl39), uvrD242(NOl97), uvrD243(NOl98), uvrD+/uvrD+(N0224), uvrD244/uvrD+(N0223), uvrD242/uvrD+(N0225)and uvrD243/uvrD+(N0226).

47

Table 5. B-ga1actosidase expression from su1A::MuQ fusion in previously isolated mutants. Assays were performed as described in Materials and Methods.

Units B-ga1actosidase Strain Relevant genotype

-Mitomycin C +Mitomycin C

N0120 wild-type 59 1630

N0221 uvrD3 61 1160

N0265 uvrD156 620 2740

N0266 mutL: :Tn10 54 1870

N0276 repS 170 1140

48

(not the merodip1oids). The standard deviation in B-ga1actosidase units

for the diploids were much greater, probably due to plasmid instability.

For example, in the most stable strain, N0223 (uvrD244), the standard

deviation in B-ga1actosidase units was 7%, for N0224 (uvrD242) the

standard deviation was 12%, and in N0226 (uvrD243) the standard

deviation was 20%.

Since uvrD plays a major role in excision repair, I considered

the possibility that a lack of excision repair results in an increased

level of SOS expression. I constructed a sulA::Mug fusion strain

containing uvrA6, a mutation which inactivates UvrABC excinuc1ease and

excision repair. While the strain was sensitive to UV irradiation, the .

basal level of sulA expression was not enhanced (data not shown),

suggesting that a defect in excision repair does not necessarily sub-

induce the SOS regu1on.

The levels of Mitomycin C induced SOS expression are presented

in Figure 4. The uvrD mutants were further inducible by the addition of

Mitomycin C, showing that the strains can respond to the addition of

further SOS inducing signal. This Mitomycin C induced expression was

enhanced above wild-type levels in strains bearing the new uvrD alleles.

This enhanced response to Mitomycin C induction was not complemented by

the uvrD+ plasmid.

Spontaneous mutagenesis in uvrD mutants

Since some alleles of uvrD confer a mutator phenotype, I

measured levels of spontaneous mutagenesis in uvrD242, 243 and 244. As

shown in Figure 5, the new uvrD alleles gave spontaneous mutator

49

Q) 2500 en

c "'C ·m 0 -u c 2000 c C)

I CD

en 1500 := c: ::J

L-Q)

::E 1000

+ 244 242 243 +/+ 244/+ 242/+ 243/+

uvrD allele

Figure 4. Mitomycin C induced B-galactosidase expression from sulA::MuQ fusion in uvrD mutants. Assays were performed as described in Materials and Methods. Strains used are as follows: uvrD+(NOl40), uvrD244(NOl39), uvrD242(NOl97), uvrD243(NOl98), uvrD3(N022l), uvrD+/uvrD+(N0224), uvrD244/uvrD+(N0223), uvrD242/uvrD+(N0225)and uvrD243/uvrD+(N0226).

100

90

eo

70

UJ 60 -c: c -::l 50 E '--(/) 40

30

20

10

0 + 244 242 243 3 +/+ 244/+ 242/+ 243/+

uvrD allele

50

Figure 5. Spontaneous mutagenesis in uvrD strains. Streptomycin resistant mutation assays were performed as diBcribed in Materials and Methods and expressed as Strr mutants/ 10 cells. Strains used are as follows: uvrD+(N0140), uvrD244(N0139), uvrD242(N0197), uvrD243(N0198), uvrD3(N0221), uvrD+/uvrD+(N0224), uvrD244/uvrD+(N0223), uvrD242/uvrD+(N0225)and uvrD243/uvrn+(N0226).

51

activity, unlike uvrD3. The uvrD156 mutation, however, is known to

exhibit a mutator phenotype (Siegel, 1973), again indicating the new

alleles may be similar. Complementation studies with the newly isolated

uvrD mutants showed that increased spontaneous mutagenesis was recessive

to the wild-type allele, since mutagenesis was restored to near wi1d­

type levels. Prior work has shown that uvrD156 spontaneous mutability

and UV sensitivity can be complemented by the uvrD+ allele (Siegel,

1981).

uvrD spontaneous mutagenesis when SOS induction is blocked

Since the new uvrD mutations resulted in an increase in SOS

expression, I wanted to see if this increase was responsible for the

elevated spontaneous mutation rate. SOS mutagenesis requires the

products of LexA regulated UmuC,D proteins as well as RecA protein.

Using the uvrD244 strain as a representative, I blocked SOS induction by

introducing a non-cleavable repressor, LexA3(Ind-). In another

strain, I introduced a deletion of recA, which not only prevented

derepression of SOS functions, but abolished SOS mutagenesis. As can be

seen in Table 6, the spontaneous mutation rate in the SOS blocked

uvrD244 strains was not reduced.

Construction of the uvrD244 de1recA or uvrD244 lexA3 double

mutant shows that a uvrD244 strain is viable when the induction of SOS

functions is blocked. su1A derepression is also blocked in the double

mutants as judged by Lac phenotype on MacConkey lactose indicator

plates. All of the new uvrD mutant strains were found to be viable with

1exA3 when I was screening for sulA fusion mutations early in this

52

Table 6. Spontaneous mutagenesis in uvrD244 strain when SOS response blocked. Streptomycin resistant mutation assays were performed as described in Materials and Methods.

Strain Relevant Genotype Strepr 'Butants per 101 cells

!!Y.!:Q recA lexA

N0139 244 + + 41

N0249 244 del + 57

N0248 244 + .1 97

N0120 + + + 3

53

study. This result stands in contrast to darn mutant$ which require SOS

derepression for viability.

uvrD mutagenesis in mismatch repair deficient strains

Mutations in mutH, mutL or mutS abolish mismatch repair in ~

coli, and strains bearing these mutations have a high level of

spontaneous mutagenesis (Radman and Wagner, 1986). Since uvrD

participates in mismatch repair, I wanted to see if a lack of mismatch

repair is responsible for the increased level of mutagenesis seen in the

mutants I isolated. I reasoned that, if the lack of mismatch repair

caused the mutagenesis seen in these strains, double mutants with

mutL::TnlO would have a level of spontaneous mutagenesis similar to that

in a single (mutL or uvrD) mutant strain. If another deficiency of the

uvrD strains was responsible for mutagenesis, then spontaneous

mutagenesis in the double mutants would be greater than in strains

bearing either of the single mutant alleles.

The spontaneous mutability of mutL strains is shown in Table 7.

The levels of spontaneous mutagenesis in mutL uvrD strains is not

increased above that seen in mutL single mutation strains. This result

agrees with previous work (Glickman and Radman, 1980) which suggests that

the spontaneous mutagenesis seen in uvrD strains is due to a lack of

mismatch repair.

I also measured sulA expression in the mutL strain to check

whether a lack of mismatch repair by itself would elevate SOS

expression. The basal level of ~ulA expression in the mutL strain was

that of wild-type strains (shown in Table 5). These results indicate

54

Table 7. Spontaneous mutagenesis in uvrD strains when mismatch repair blocked. Streptomycin resistant mutation assays were performed as described in Materials and Methods.

Strain Relevant genotype

N0266

N0267 mutL uvrD242

N0268 mutL uvrD243

N0269 mutL uvrD244

Strepr mutants/ 1010 cells

134

132

83

140

55

that the lack of mismatch repair in the uvrD strains is unlikely to be

responsible for the increased expression of su1A in the uvrD mutants.

Viability of uvrD244 repS double mutants

The dominance of the new uvrQ alleles with respect to SOS

induction observed in the mutant strains indicates that they are not

null alleles. Previous studies showed that uvrD rep double mutants are

inviable or very unstable, which suggests that there is an interaction

between these two gene products during replication or that they perform

a similar essential function (Taucher-Scholz et a1., 1983; Siegel,

1973). Taucher-Scholz et a1. (1983) propose that UvrD and Rep proteins

act jointly during DNA replication to unwind duplex DNA, with each

protein acting on opposite arms of the replication fork. In the case of

Rep or UvrD mutations, they suggest that these proteins can substitute

for each other during replication. In order to determine if the uvrD244

strain was defective in replication functions, I attempted to construct

a repS uvrD244 double mutant. I was able to make this construction with

normal cotransduction frequencies and conclude that UvrD244 protein has

at least some function as a DNA helicase. It may be that an altered

helicase function in this mutant is responsible for the increased

expression of SOS functions.

Additionally, I measured B-galactosidase expression in the repS

strain. sulA expression was elevated 2-3 fold over wild-type levels,

demonstrating some sub-induction of sulA in this mutant (Table 5).

56

recA mutants

SOS expression

Expression levels from the sulA::MuQ fusion in the new recA

alleles is shown in Figure 6. For comparison, I included data for a

lexA7l(Def) mutant, the recA44l (formerly tif-l) allele and recA730

(Prtc ) which was isolated by Witkin et al. (1982). The expression of

sulA in these new recA mutants was very high, near that of a cell in

which LexA is non-functional. The high expression of SOS functions was

also seen in recA730, as would be expected in mutants of a similar

class. Standard deviations for the B-galactosidase values in the recA

mutant strains were approximately 5%. The addition of Mitomycin Chad

little effect on these strains, probably because they were nearly fully

induced.

recA mutant recombination proficieny

Since RecA protein has an additional function in promoting

homologous recombination, I assayed levels of recombination in the recA

mutant strains. Recombination efficiency is shown in Table 8. All

mutants appear to be proficient in homologous recombination of phage

lambda.

Spontaneous mutagenesis in recA mutants

Spontaneous mutagenesis in the new recA mutants is shown in

Figure 7. All strains have increased spontaneous mutator activity

compared to recA wild-type strains. recA44l strains had previously been

shown to have increased spontaneous mutations, but under conditions of

Q) VI C

'"C ·iii o -(,) c c en I

en

..... Q)

57

10000 LEGEND

9000 -mmc

+mmc 8000

7000

6000

5000

4000

3000

2000

~ O.L.-.......f..-n ~~~ ~~~~-'---I

1000

+ + 441 730 82 87 80 81 86 84 85 83 ----- - --lexA71 recA allele

Figure 6. B-galactosidase levels from sulA::Mug fusion in mutant recA strains. Assays were performed as described in Materials and Methods. Strains used are as follows: recA+(N056), lexA7l(KP28l), recA441(N063), recA730(N062), recA82(N073), recA87(N078), recA80(N060), recA81(N061), recA86(N077), recA84(N075), recA85(N076), recA83(N083).

58

Table 8. Recombination proficiency in mutant recA strains. Recombination frequency was determined using lambda by lambda crosses as described in Materials and Methods.

Relevant sus+ phage/ Strain genotype total phage

N056 recA+ . 4.4xlO- 2

KP28l recA+, lexA71 2.0xlO- 2

N063 recA441 2.9xlO- 2

N062 recA730 3.7xlO- 2

N073 recA82 3.4xlO- 2

N078 recA87 4.6xlO- 2

N060 recA80 4.4xlO- 2

N06l recA8l 4.lxlO- 2

N077 recA86 3.9xlO- 2

N075 recA84 4.7xlO- 2

N076 recA85 1. 9xlO- 2

N083 recA83 2.8xlO- 2

59

250

225

200

175

2 150 s:::: c -::l 125 E L. -Vl 100

75

50

25

a + + 441 730 82 87 80 81 8S 84 85 83

lexA71 recA allele

Figure 7. Spontaneous mutagenesis levels in recA mutants. Streptomycin resistance mutagenesis assays were performed as dIBcribed in Materials and Methods and are express~d as Strr mutants/ 10 cells. The strains used for these studies are as follows: recA+(N056). lexA7l(KP28l). recA441(N063). recA730(N062). recA82(N073). recA87(N078). recA80(N060). recA81(N06l). recA86(N077). recA84(N075). recA85(N076). recA83(N083).

60

growth different than used in this study. The growth conditions

conditions used in this study, 370 C versus 42oC, may account for the

relatively low level of mutagenesis for the recA44l strain shown in

Figure 7. Two mutants that exhibited extremes of mutagenesis and sulA

expression, recA8l and recA85 were chosen for further study.

Viable and inviable mutations with recA8l and recA85

I wanted to make double mutants of recA8l and recA85 with other

alleles whose mutant products affect SOS induction. These included

ssbl, ssbl13, recBC and recF. I successfully made double mutants of all

alleles in recA8l. The additional mutations did not affect constitutive

induction resulting from recA8l. This result indicates that the

mutations act at different stages in SOS induction. A similar result

was obtained by Lieberman and Witkin (1983) in finding that ssb

mutations did not affect SOS-constitutivity in recA730(prtc ) mutants.

In constructing the double mutants of ssb or recB with recA85 the

situation was very different. Pl transductions which attempted to move

ssbl or ssbl13 into a recA85 strain (N01.2l) yielded many colonies at the

expected transduct~on frequency; however, on patching or streaking,

those colonies which had the double mutant completely failed to grow.

In the case of recB, I was unable to construct recA85 recB2l or recA85

recB::TnlO double mutants. In attempting to construct recA85 recB2l

mutants (Pl KW88 --> NOl2l), 30 transductants failed to yield recB2l

mutations. In another Pl transduction to move recB::TnlO (donor KP4l9)

into recA85 strain NOl2l, no tetracycline resistant transductants were

obtained. A control transduction (recB::TnlO --> wild-type N0120

61

strain) run side-by-side yielded 300 transductants. recA85 recF142

double mutants were constructed without difficulty. The recA85 recF142

double mutant strain still expressed su1A constitutively as judged on

MacConkey plates. The difference in ability to construct double mutants

demonstrates that recA85 and recA81 are different" mutations that lead to

different effects within the cell.

62

CHAPTER 4

DISCUSSION

Isolation of SOS constitutive mutants

I have isolated, mapped and characterized several ~ coli

mutations which lead to constitutive induction of the SOS response.

These mutations include 8 recA (Prtc ) mutations, 3 uvrD mutations, and

single mutations in lig and dam. All of these mutations lead to

constitutive increased expression of the SOS response under conditions

of viable growth; that is, they are not conditionally lethal mutations.

The mutants isolated fall into two groups. The first group

consists of the uvrD, lig and dam mutations. These mutations subinduce

the SOS response, as shown by an increase in sulA expression (Figs. I

and 3). SOS induction in these mutants is most likely mediated through

RecA and LexA proteins, since LexA3(Ind-) blocks sulA expression in lig

and uvrD strains (dam mutants are inviable with lexA3 and require

expression of some SOS-regulated genes). This is in agreement with the

current model of SOS response regulation through RecA and LexA proteins

as discussed in the Introduction of this dissertation. The second type

of mutants, the recA mutations, arG those in which sulA is fully induced

to levels of expression similar to lexA(Null) mutants (Fig. 6). These

mutations are similar to others isolated previously and designated PrtC

for constitutive protease activity. SOS expression appears to be

mediated through LexA repressor destruction at least for RecA85, as the

63

LexA2(Ind-) hypocleavable repressor drastically reduces sulA expression

in a recA85 strain (Peterson et al., 1988a).

We interpret our results by proposing that the mutations

isolated represent two early steps in the induction of the SOS response.

The first step is generation of an SOS inducing signal mimicking (or

causing) DNA damage and is represented by the dam, lig and uvrD

mutations. The products of these genes all affect DNA metabolism and

the mutant proteins most likely alter DNA metabolism in such a way as to

produce an SOS inducing signal and activate RecA protein. The second

step of induction is the signal target and is represented by mutations

in recA, a key regulatory gene of the SOS response.

The alterations in DNA metabolism that induce the SOS response

are unclear. Finding mutations in !ig and uvrD which induce SOS is

consistent with other previously observed mutations causing induction,

such as !ig(Ts) , dnaB(Ts), dnaE(Ts) and polA(Ts). These mutations all

specifically affect synthesis of DNA indicating that the integrity of

DNA synthetic machinery plays a role in SOS induction. This appears to

be limited to the actual process of DNA synthesis, however, not

including re-initiation of new rounds of chromosomal replication.

Mutations that specifically halt initiation of new rounds of DNA

replication, such as dnaA(Ts) or dnaC(Ts), do not lead to spontaneous

SOS induction. Possible specific signals generated by each of the

mutations isolated in this study will be discussed later in this

chapter.

The uvrD, !ig and dam mutations lead to intermediate levels of

SOS induction, a condition sometimes refer.red to as sub-induction

64

(Bailone, et al., 1979; Little, 1983). These mutations all lead to

different basal levels of SOS expression, as shown in Figures land 3,

indicating that the level of inducing signal is a variable quantity.

The variable nature of signal is underscored by the observation that the

mutants isolated were capable of further induction by Mitomycin C

treatment. In control experiments, wild-type cells were induced with

Mitomycin C. The levels of SOS expression did not approach that of

strains which lacked LexA repressor. This difference indicates that

there was still intact LexA repressor remaining in the cell and may be

referred to as a sub-induced state as well. In the mutant strains, sub­

induction may result from low levels of SOS inducing signal being

generated or incomplete signal if there are multiple' paths or types of

signal needed. My studies do not distinguish between these

possibilities.

The variable nature of signal probably allows the cell to fine

tune levels of SOS-associated repair proteins according to the amount of

DNA damage perceived. Based on the current model of SOS regulation,

signal most likely mediates this fine tuning by causing low levels of

RecA protein activation leading to cleavage of a fraction of the LexA

repressor population, partially derepressing SOS-regulated operons. A

steady state would be established producing a certain level of repair

proteins. Another factor infuencing the level of activated RecA protein

and SOS repair proteins is negative regulation of recA and lexA by the

lexA product. In the case of the darn mutation, the increased basal

level of SOS expression is required to accommodate some type of damage

generated by the defective protein, as dam strains are inviable when

65

lexA3(Ind-). The other mutations, li& and uvrD, do not generate damage

requiring SOS repair since they are viable when the SOS response is

blocked.

One problem with this type of study is trying to determine how

the mutant phenotype relates to signal generation. In these sub­

induction mutants, one hopes that the mutant protein subtly perturbs a

specific and identifiable cellular process, allowing speculation as to

what constitutes an inducing signal. This direct type of perturbation

may be in DNA synthesis, causing a DNA structural alteration or an

immbalance in nucleotide pools for example. In reality, however, the

mutant protein may have wide-ranging effects that indirectly produce an

inducing signal by affecting other cellular processes. One possible

example would be of a mutation causing DNA damage that interferes with

DNA syt~thesis, subjecting DNA to exonuclease action or leading to DNA

repair intermediates that generate inducing signal. Mutations that

cause SOS induction under lethal conditions (such as Ts mutations) could

certainly cause such an effect. I will attempt to address this issue

when each of the specific mutations are discussed in detail.

The failure to find fully induced mutants (sulA levels analogous

to those found in lexA(Def) strains) in genes other than recA may be a

consequence of selecting non-lethal, SOS inducing mutations. The

mutants chosen for characterization in this study were those which

exhibited the strongest SOS induction, indicated by Lac+ phenotype.

Only mutants of recA, which represent the signal target, induced the SOS

system to the extetlt that lambda £1 repressor was cleaved constitutively

or that most (or all) LexA protein was cleaved. The ~ mutants

66

isolated were identified in the STS ~I repressor cleavage screen and

were shown to have levels of sulA expression similar to a lexA(Def)

strain. Mutations which caused this extent of SOS induction were not

found in any other genes--those that would be involved in the first step

of induction described above. It may be that severe defects in cellular

physiology (such as inhibition of DNA synthesis) generating an inducing

signal (or signals) strong enough to cleave most of the LexA in the cell

would be lethal. In fact, conditional mutations previously known to

induce prophage (cleave £1 repressor) do so at the non-permissive

temperature, as discussed in the Introduction. Since I was screening

for non-lethal SOS inducing mutants, this may explain why none of the

mutations I isolated (besides ~) lead to high levels of SOS

induction.

lig2S2 mutation

I have isolated a DNA ligase mutation which leads to sub­

induction of the SOS response. This mutation is not conditionally

lethal, but leads to increased SOS expression under normal conditions

for growth. Since ligase is an essential gene, the lig2S2 mutation is

not a null allele but probably ligates replicating DNA and naturally

occuring nicks at a reduced efficiency.

We propose that the 1!g mutation generates an inducing signal by

the disrupting of DNA metabolism. One possibility is that unligated

nicks occurring in DNA during replication may serve directly as inducing

signals, being recognized by RecA protein. Alternatively, unligated

nicks may serve as a substrate for the action of further cellular

67

processes that generate SOS signals, thus inducing the SOS response

indirectly. These structures may be single-stranded regions of DNA or

oligonucleotides resulting from DNA degradation.

One clue bearing on SOS induction in ligase mutations is that

the mutant is inviable with recB mutations. This suggests that RecBCD

enzyme may act at DNA nicks to repair potentially lethal DNA damage

resulting from an over-abundance of nicks in DNA. This action most

likely involves recombination, since RecBCD is a major participant in

homologous recombination. The possible damage in the DNA ligase mutant

does not, however, require SOS repair functions to be expressed since

this mutant is viable when SOS is blocked (in lexA3 lig252 double

mutants).

daml7 mutant

The isolation of a dam mutation in this study comfirms earlier

observations that dam mutations increase the basal level of SOS

expression (Craig et al., 1984; Peterson et al., 1985). The allele

isolated in this study appears to be a null allele of dam since the

level of sulA expression in the mutant is similar to that found in an

isogenic strain bearing a known dam null allele, dam13::Tn2 (Figure 1).

Similar to other dam mutants, daml7 is sensitive to 2-aminopurine and

inviable in combination with lexA3.

The inviability of dam mutations with lexA3 indicates that

partial SOS derepression is required in strains lacking DNA-adenine

methylase and suggests that the lack of methylation produces a cellular

abnormality that needs to be repaired. Strains deficient in dam do not

68

simply lack mismatch repair, but are subject to indiscriminant mismatch

repair. Since neither DNA strand is methylated, MutHLS repair enzymes

attempt to repair both strands of DNA when mismatches occur (Lu et al,

1983; Pukkila et al., 1983). The lack of mismatch repair itself does

not lead to increased SOS expression, demonstrated in mutL strains

(Table 5) and in mutS strains (Craig et al., 1984). The results of

Peterson et ale (1985) show that strains bearing dam mutations require

SOS regulated recombination repair genes for viability. The required

genes are recA, ~ and recJ. The recB and Q genes, which are

associated with recombination but not under LexA repressor control, are

also required in dam strains (Marinus and Morris, 1974). This indicates

that in dam strains some type of DNA damage occurs that requires repair

for cell viability. Glickman and Radman (1980) suggest that

indiscriminant mismatch repair results in the formation of double­

stranded breaks in DNA, which would require DNA repair processes to

correct these potentially lethal structures.

How dam mutations lead to SOS induction is not clear.

Certainly, DNA damage resulting from indiscriminant mismatch repair,

such as double- or single-strand DNA breaks, could produce an inducing

signal and lead to increased expression of sulA. Many of the phenotypes

associated with dam mutations are alleviated by second site mutations in

the mismatch repair enzymes (MutH, L or S). These phenotypes include 2-

aminopurine sensitivity and inviability with other allele combinations

such as recA(Def) and lexA(Ind-) (McGraw and Marinus, 1980). Other

evidence, however, shows that mutations in mutH, 1 or ~ do not suppress

SOS sub-induction in dam strains (Craig et al, 1984; K.R. Peterson,

69

personal communication). This result indicates that another de~ect,

which does not require SOS associated repair, sub-induces the SOS

response in dam mutants.

The extent of SOS induction in dam mutants as measured by su1A

expression may not be a true reflection of recA activation. The su1A

promoter contains a Dam methylation site: GATC (Cole, 1983). How the

lack of methylation affects promoter activity is unclear. In the case

of the trpR promoter, lack of methylation increases expression 2-3 times

(Marinus, 1985), however, lack of methylation in one of the dnaA

promoters decreases expression of the gene (Braun and Wright, 1986).

While the effect of methylation on su1A expression is not known, it is

still clear that dam mutations increase expression of other SOS

regulated genes (Craig et a1., 1984; Peterson et al., 1985).

uvrD mutants

The results of this study show that uvrD mutations can lead to

sub-induction of the SOS response (Figure 3 and Table 5). This includes

the new mutations isolated (uvrD242, 243 and 244), as well as a

previously isolated mutation, uvrD156. Not all uvrD mutations share

this property, however, since uvrD3 does not increase SOS expression.

All known uvrD mutations, including those isolated in this study, share

a UV sensitive phenotype (shown in Figure 2).

Several phenotypes are shared by uvrD242, 243, 244 and uvrD156,

suggesting that these may be of a similar class. These phenotypes

include SOS sub-induction, spontaneous mutagenesis and UV sensitivity.

Also, the spontanteous mutagenesis seen in all these uvrD alleles can be

70

complemented by uvrD+. Siegel (1983) found the spontaneous mutator

phenotype in a uvrD::Mu4 fusion strain, prompting him to suggest that

the mutator alleles of uvrD are the null phenotype. The uvrD mutants

isolated in this study, however, are not null alleles. This activity is

demonstrated by a lack nf uvrD+ complementation in SOS expression levels

of the mutant strains, shown in Figure 3. Also the viability of uvrD244

with the repS allele suggests at least some functioning of the mutant

protein. It is possible that the mutations isolated in this study may

not be as defective as uvrDls6 since the basal level of su1A expression

is higher in the uvrD1s6 strain. Alternatively, uvrD may be an

essential gene-- no known uvrD deletions have been constructed.

That the mutator alleles of uvrD sub-induce the SOS response may

affect what is currently thought about the levels of uvrD gene

expression. The uvrD gene is controlled by LexA repressor and is part

of the SOS regu1on. This SOS-mediated regulation was identified by

constructing Mu4 fusions to the uvrD gene and the resultant mutations

are UV sensitive and spontaneous mutators (Siegel, 1983). The level of

uvrD expression is increased 2.s-fo1d when induced with lug/ml Mitomycin

C and s-fold in 1exA71(Def) strains. If, however, the basal level of

SOS expression is increased due to the uvrD fusion mutation, the

induction ratio may actually be greater than what was observed in the

fusion strains.

This study demonstrates that certain defects in the uvrD mutants

can be complemented by a wild-type allele. These defects are UV

sensitivity and spontaneous mutator activity, shown in Figures 2 and 5.

Complementation of the UV sensitivity indicates that DNA repair

71

deficiencies associated with the uvrD mutation are complemented.

Experiments showed that spontaneous mutagenesis was not a result of SOS­

mutagenesis (see Table 6), but may be due to a lack of mismatch repair

in the mutant strains (Table 7). This conclusion is in agreements with

previous studies that show the following: 1) mismatch repair deficient

strains (mutH, mutL, mutS or dam mutants) are spontaneous mutators

(Marinus, 1984), 2) evidence suggesting uvrE (now known to be uvrD) and

the mutHLS genes operate within the same pathway (Glickman and Radman,

1980), and 3) uvrD mutants and mutHLS and dam mutants all share the same

mutational spectrum (Cox, 1976; Glickman, 1979). I' conclude that one of

the defects in the newly isolated uvrD mutants is a repair defect which

can be complemented. This defect may be in the ability of UvrD protein

to recognize other repair proteins at DNA lesions. Pseudoreversion

analysis of these mutants by isolating UV resistant strains would

determine other proteins which interact with UvrD during DNA repair.

The mutant uvrD strains have another defect, possibly in DNA

replication, that leads to SOS induction and is not fully complemented

by uvrD+. Evidence presented in this study shows that a lack of DNA

repair functions is not sufficient to sub-induce the SOS response.

Excision repair defective uvrA6 strains and mismatch repair deficient

mutL strains were shown not to cause sub-induction of the sulA fusion.

Furthermore, uvrD3 strains which are deficient in DNA repair (UV

sensitive) do not cause increased expression of sulA.

The viability of uvrD244 in combination with repS suggests that

the mutant UvrD protein is proficient in DNA replication, since a uvrD

rep double mutant of other uvrD alleles cannot be made. I suggest that

72

an altered DNA helicase function during replication, rather than DNA

repair defects, is responsible for the SOS sub-induction seen in this

mutant. Furthermore, finding that the repS strain leads to a low level

of sulA expression suggests that improper helicase function during

replication can produce an inducing signal.

The specific role of DNA helicase II in DNA replication is not

clear. Taucher-Scholz et al. (1983) propose that during DNA replication

Rep and UvrD proteins act jointly on opposite strands to unwind duplex

DNA. This proposal was based on several observations. First, they were

unable to make double mutants of rep and uvrD (although the alleles used

in the attempted construction were not specified). Second, Rep helicase

unwinds DNA in the 3'--> 5' direction, the parental strand leading in

DNA synthesis (Yarranton and Gefter, 1979). Third, UvrD helicase was

thought at the time to unwind DNA on the parental strand for lagging DNA

synthesis, in the 5'--> 3' direction (Kuhn et al., 1979). In addition

to proposing joint action of Rep and UvrD helicases on opposite strands

at the replication fork, Taucher-Scholz et al. (1983) further suggested

that the two helicases could substitute for each other if necessary

(such as when one helicase is mutationally inactivated). More recent

studies, however, report that UvrD translocates on the 3'--> 5'strand,

the same direction as Rep (Matson, 1986). Perhaps UvrD and Rep can

substitute for one another on the 3'--> 5', leading strand during DNA

replication. Recently, DnaB protein has been shown to be a helicase,

unwinding DNA on the lagging strand template, or 5'--> 3' direction

(LeBowitz and McMacken, 1986). Rep, UvrD and DnaB may all function as

DNA helicases during replication.

73

DNA he1icases appear to play a role in signal metabolism and

the SOS response. rep and uvrD mutations have been shown to sub-induce

the SOS response and dnaB(Ts) mutations induce SOS functions at non­

permissive temperatures. Another DNA unwinding protein, RecBCD, has

been shown to affect SOS induction as well (see Chapter I). Chaudhury

and Smith (1985) showed that the exonuclease activity of this enzyme was

not needed for SOS induction and suggested that RecBCD DNA unwinding

activity instead was involved in SOS signal generation. He1icases may

be important in regulating the amounts of single-stranded DNA in the

cell, an imbalance of which may be an inducing signal for RecA protein.

Alternatively, they may affect the rate of DNA replication. If

unwinding is impeded, the slowed replication fork movement may produce

an inducing signal, such as a nucleotide cofactor for RecA activation.

Biochemical studies of the mutant UvrD proteins may help to unravel the

alterations in DNA unwinding activity that produce an SOS inducing

signal.

recA mutants

The recA mutations isolated exhibited the strongest SOS

induction in any of the mutants screened. These mutations are of the

PrtC class because they are ~o~ease £onstitutive for assisting in

repressor cleavage. Additionally, the mutants all have increased

spontaneous mutability and are recombination proficient. Finding that

the only fully SOS-induced mutations isolated in this work map to the

recA gene underscores the central role of this gene product in

regulating the SOS response.

74

PrtC mutations have been isolated previously. These include the

initial temperature-sensitive allele, recA441 (tif-1)(Go1dthwait and

Jacob, 1964), the non-Ts PrtC allele, recA730 (Witkin et a1., 1982) and

several isolated by Tessman (Tessman and Peterson, 1985). All of these

alleles share phenotypes similar to the recA mutations I have isolated

including spontaneous phage induction, high levels of SOS expression,

recombination proficiency and spontaneous mutagenesis. It should be

noted that Tessman also isolated some recA(Prtc ) class alleles that were

recombination deficient; however, they cleaved LexA protein to a lesser

extent and did not spontaneously induce prophage (Tessman and Peterson,

1985a).

One of the recA mutations isolated in this study, recA85, had

unusual behavior. This mutation appeared to be inviable in combination

with ssbl, ssbll3 or recB mutations. This is in contrast to another of

the recA mutations isolated, recASl, in which the same double mutant

combinations could be made. In another study, a recA730 mutation was

viable in combination with ssbl or sbll3 mutations (Lieberman and

Witkin, 1983). These results demonstrate that the recA85 defect is

different from the defect in recASl or recA730, although other

phenotypes are similar. It is possible that these different mutations

define different sites on RecA protein that interact with inducing

signal(s) or show different abilities to bind ssDNA.

The relationship between RecA protein and Ssb (single-stranded

DNA binding) protein has been the subject of much investigation. In

vitro studies implicate Ssb protein in assisting strand exchange

reactions of RecA protein in homologous recombination (reviewed by

75

Radding, 1982). Repressor cleavage reactions, however, are enhanced at

optimal Ssb concentrations and inhibited by higher concentrations of Ssb

protein (Resnick and Sussman, 1982). Induction of the SOS response in

vivo is inhibited in ssb1 and ssb113 strains, indicating that this

protein plays a role in signal generation (Ba1uch et a1., 1980a).

Recent work suggests that there is no specific protein-protein

interaction between RecA and Ssb proteins, but that competitive binding

occurs on single-stranded DNA (Kowa1czykowski et a1., 1987). Based on

this evidence, a plausible explanation for the inviability of recA85 and

ssb mutations is that RecA85 protein may bind single-stranded regions of

DNA more tightly than wild-type RecA protein. In mutants with defective

Ssb, RecA85 may outcompete the protein for single-stranded DNA sites,

excluding Ssb. This competition could be lethal since Ssb is an

essential protein and its only known function is binding single-stranded

DNA. Tighter binding of RecA85 to single-stranded DNA could also

explain the constitutive activation seen in this protein.

Biochemical studies would provide more insight into the nature

of the RecA85 defects. Protein DNA binding studies could determine

whether this protein does bind single-stranded DNA more tightly.

Current collaboration with Dr. Patricia Weber (DuPont Corp.) to

investigate the biochemistry of purified RecA85 may provide this

information.

2-aminopurine

Mu.Q. fus ion

mutHLS

recBCD

SOS mutagenesis

uvrABC

76

APPENDIX A

Definitions of genetic terms

DNA base analogue that incorporates as adenine and causes transition mutations.

Gene coding for DNA adenine methylase, a protein involved in mismatch repair.

Gene encoding the repressor of the SOS regu10n.

Gene coding for DNA ligase.

Mu phage which inserts randomly into the chromosome and contains promoterless Lac structural genes (lac ZYA). The Lac genes are expressed when this phage inserts downstream from a promoter in the chromosome.

Genes encoding the mismatch repair proteins.

A gene encoding a key regulatory protein of the SOS response, also involved in strand exchange reactions during homologous recombination.

Genes encoding the multi-subunit protein Exonuclease V. As well as exonuclease activity, this protein has DNA unwinding ability and participates in homologous recombination.

Gene encoding replicase, an ATP-dependent DNA unwinding protein.

An increase in mutagenesis resulting from expression of the SOS genes recA and umuCD.

Gene encoding a cell division inhibitor, also known as sfiA. Mutations in sulA prevent lethal filamentation when the SOS response induced.

Genes encoding a protein complex that perfoms excision repair.

Gene encoding DNA helicese II, an ATP-dependent DNA unwinding protein. This protein also participates in excision repair, RecF recombination and mismatch repair.

77

REFERENCES

Ade1berg, E.A., M.Mande11 and G.Chein Chin Hem. 1965 Optimal conditions for mutagenesis by N-methy1-N-nitro-N-nitrosoguanidine in Escherichia coli K-12. Biochem. Biophys. Res. Commun. 18:788-795.

Bachmann, B.J. 1983 Linkage map of Escherichia coli K-12, Edition 7. Microbio1. Rev. 47:180-230.

Bai1one, A., A.Levine and R.Devoret. 1979 Inactivation of prophage lambda repressor in vivo. J. Mol. Bio. 131:553-572.

Baker, T.A., M.M.Howe and C.A.Gross. 1983 MudX, a derivative of Mug1(lac,Ap ) which makes stable 1acZ fusions at high temperature. J. Bacterio1 156:970-974.

Ba1uch, J., R.Sussman and J.W.Resnick. 1980 Induction of prophage lambda without amplification of recA protein. Mol. Gen. Genet. 178:317-323.

Ba1uch, J., J.W.Chase and R.Sussman. 1980a Synthesis of RecA protein and induction of bacteriophage lambda in single-strand deoxyribonucleic acid-binding protein mutants of Escherichia coli. J. Bacteriol. 144:489-498.

Blanco, M., and L.Pomes. 1977 Prophage induction in Escherichia coli K12 cells deficient in DNA polymerase I. Mol. Gen. Genet. 154:287-292.

Braun, R.E., and A.Wright. 1986 DNA methylation differentially enhances the expression of one of the two E. coli dnaA promoters in vivo and in vitro. Mol. Gen. Genet. 202:246-250.

Casadaban, M.J., and S.N.Cohen. 1979 Lactose genes fused to exogenous promoters in one step using a Mu-1ac bacteriophage: in vivo probe for transcriptional control sequences. Proc. Natl. Acad. Sci. USA 76:4530-4533.

Casaregola, S., R.D'Ari and O.Huisman. 1982 Role of DNA replication in the induction and turn-off of the SOS response in Escherichia coli. Mol. Gen. Genet. 185:440-444.

Chaudhury, A.M., and G.R.Smith. 1985 Role of Escherichia coli RecBC enzyme in SOS induction. Mol. Gen. Genet. 201:525-528.

Cohen, S., B.J.Knoll, J.W.Little and D.W.Mount. 1981 Preferential cleavage of phage lambda repressor monomers by RecA protease. Nature 294:182-184.

78

Cole, S.T. 1983 Characterization of the promoter for the lexA regulated sulA gene of Escherichia coli. Mol. Gen. Genet. 189:400-404.

Cox, E.C. 1976 Bacterial mutator genes and the'control of spontaneous mutation. Ann. Rev. Genet. 10:135-156.

Craig, N.L., and J.W.Roberts. 1980 ~ coli RecA protein-directed cleavage of phage lambda repressor requires polynucleotide. Nature 283:26-30.

Craig, N.L., and J.W.Roberts. 1981 Function of nucleoside triphosphate and polynucleotide in Escherichia coli RecA protein-directed cleavage of phage lambda repressor. J. BioI. Chem. 256:8039-8044.

Craig, R.L., J.A. Arraj and M.G. Marinus. 1984 Induction of damage inducible (SOS) repair in dam mutants of Escherichia coli exposed to 2-aminopurine. Mol. Gen. Genet. 194:539-540.

Crowl, R.M., R.P. Boyce, and H. Echols. 1981 Repressor cleavage as a prophage induction mechanism: hypersensitivity of a mutant AcI protein to RecA-mediated proteolysis. J. Mol. Bio. 152:815-820.

D'Ari, R., and O.Huisman. 1982 DNA replication and indirect induction of the SOS response in Escherichia coli. Biochimie 64:623-627.

Das, S.K., and L.A.Loeb. 1984 UV irradiation alters deoxynucleoside triphosphate pools in Escherichia coli. Mut. Res. 131:97-100.

Davis, R.W., D.Botstein and J.R.Roth. 1980 Advanced Bacterial Genetics. New York: Cold Spring Harbor.

Geider, K., and H.Hoffmann-Berling. 1981 Proteins controlling the helical structure of DNA. Ann. Rev. Biochem. 50:233-260.

George, J., R.Devoret and M.Radman. 1974 Indirect ultraviolet­reactivation of phage lambda. Proc. Natl. Acad. Sci. USA 71:144-147.

George, J., M.Castellazzi and G.Buttin. 1975 Prophage induction and cell division in ~ coli. III. Mutations sfiA and sfiB restore division in tif and Ion strains and permit the expression of mutator properties of tif. Mol. Gen. Genet. 140:309-332.

Glickman, B.W., and M.Radman. 1980 Escherichia coli mutator mutants deficient in methylation-instructed DNA mismatch correction. Proc. Natl. Acad. Sci. USA 77:1063-1067.

Glickman, B.W. 1979 Spontaneous mutagenesis in ~ coli strains lacking 6-methyladenine residues in their DNA. An altered mutational spectrum in dam mutants .. Mutat. Res. 61:153-162.

79

Goldthwait, D., and F.Jacob. 1964 Sur Ie mechanisme de l'induction du developpment du prophage chez les bacteries lysogenes. C.R. Acad. Sci. 259:661-664.

Gottesman, M.M., M.L.Hicks and M.Gellert. 1973 Genetics and function of DNA ligase in Escherichia coli. J. Mol. Bio. 77:531-547.

Gudas, L.J., and D.W.Mount. 1977 Identification of the recA (tif) gene product of Escherichia coli. Proc. Natl. Acad. Sci. USA 74:5280-5284.

Horii, Z.I., and A.J.Clark. 1971 Genetic analysis of the RecF pathway to genetic recombination in Escherichia coli: Isolation and characterization of mutants. J. Mol. Bio. 80:327-344.

Horii, T., T.Ogawa, T.Nakatani, T.Hase. H.Matsubara and H.Ogawa. 1981 Regulation of SOS functions: Purification of h coli LexA protein and determination of its specific site cleaved by the RecA protein. Cell 27: 515-522.

Huisman, 0., and R.D'Ari. 1981 An inducible DNA replication-cell division coupling mechanism in ~ coli. Nature 290:797-799.

Irbe, R.M., L.M.Morin and M.Oishi. 1981 Prophage (080) induction in Escherichia coli K-12 by specific deoxyoligonucleotides. Proc. Natl. Acad. Sci. USA 78:138-142.

Kenyon, C.J., and G.C.Walker. 1980 DNA-damaging agents stimulate gene expression at specific loci in Escherichia coli. Proc. Natl. Acad. Sci. USA 77:2819-2823.

Kowalczykowski, S.C., R.Clow, R.Somani and A.Varghese. 1987 Effects of the Escherichia coli SSB protein on the binding of Escherichia coli RecA protein to single-stranded DNA. J. Mol. BioI. 193:81-95.

Kuhn. B., M.Abdel-Monem and J.Hoffmann-Berling. 1979 Evidence for two mechanisms for DNA unwinding catalyzed by DNA helicases. J. BioI. Chem. 254:1134311350-11350.

LeBowitz, J.H., and R.McMacken. 1986 The Escherichia coli DnaB replication protein is a DNA helicase. J. BioI. Chem. 261:4738-4748.

Lieberman, H.B., and E.M.Witkin. 1983 DNA degradation, UV sensitivity and SOS-mediated mutagenesis in strains of Escherichia coli deficient in single-strand DNA binding protein: Effects of mutations and treatments that alter levels of exonuclease V or RecA protein. Mol. Gen. Genet. 190:92-100.

Little, J.W., and P.C.Hanawalt. 1977 Induction of protein X in Escherichia £Q!i. Mol. Gen. Genet. 150:237-248.

Little, J.W., and D.W.Mount. 1982 The SOS regulatory system of Escherichia coli. Cell 29:11-22.

80

Little, J.W., S.H.Edmiston, L.Z.Pace1Ii and D.W.Mount." 1980 Cleavage of the Escherichia coli LexA protein by the RecA protease. Proc. Natl. Acad. Sci. USA 77:3225-3229.

Little, J.W. 1983 The SOS regulatory system: Control of its state by the level of RecA protease. J. Mol. BioI. 167:791-808.

Lu, A.L., S.Clark and P.Modrich. 1983 Methyl-directed repair of DNA base-pair mismatches in vitro. Proc. Natl. Acad. Sci. USA 80:4639-4643.

Lu, C., R. Scheuermann, and H. Echols. 1986 Capacity of RecA protein to bind preferentially to UV lesions and inhibit the editing (E)

subunit of DNA polymerase III: a possible mechanism for SOS-induced targeted mutagenesis. Proc. Nat1. Acad. Sci. USA 83:619-623.

Maples, V.F., and S.R.Kushner. 1982 DNA repair in Escherichia coli: Identification of the uvrD gene product. Proc. Natl. Acad. Sci. USA 79:5616-5620.

Marinus, M.G., and N.R.Morris. 1974 Biological function for 6-methyladenine residues in the DNA of Escherichia coli K-12. J. Mol. Bio. 85:309-322.

Marinus, M.G., and N.R.Morris. 1975 Pleiotropic effects of a DNA adenine" methylation mutation (dam-) in Escherichia coli K-12. Mut. Res. 28:15-26.

Marinus, M. 1984 Methylation of procaryotic DNA. In DNA Methylation. A.Razin, M.Cedar, A.Riggs, (eds.) Pp. 81-109. New York. Springer­Verlag.

Marinus, M.G. 1985 DNA methylation influences trpR promoter activity in Escherichia coli K-12. Mol. Gen. Genet. 200:185-186.

Matson, S.W. 1986 Escherichia coli he1icase II (uvrD gene product) trans locates unidirectiona1ly in a 3' to 5' direction. J. BioI. Chem. 261:10169-10175.

McEntee, K., and G.M.Weinstock. 1981 tif-1 mutation alters polynucleotide recognition by the RecA protein of Escherichia coli. Proc. Natl. Acad. Sci. USA 75:6061-6065.

McGraw. B.R., and M.G.Marinus. 1980 Isolation and characterization of dam+ revertants and suppressor mutations that modify secondary phenotypes of dam-3 strains of Escherichia coli K-12. Mol. Gen. Genet. 178:309-315.

McEntee, K., and G.M.Weinstock. 1981 tif-1 mutation alters polynucleotide recognition by the RecA protein of Escherichia coli. Proc. Nat1. Acad. Sci. USA 75:6061-6065.

81

McPartland, A., L.Green and H.Echo1s. 1980 Control of recA gene RNA in ~ coli: regulatory and signal genes. Cell 20:731-737.

Miller, J.H. 1972 Experiments in Molecular Genetics. New York: Cold Spring Harbor.

Morand./P., M.B1anco and R.Devoret. 1977 Characterization of 1exB mutations in Escherichia coli K-12. J. Bacterio1. 131:572-582.

Mount, D.W. 1979 Isolation and characterization of mutants of recA which synthesize a hyperactive RecA protein. Virology 98:484-488.

Oeda, K., T.Horiuchi and M.Sekiguchi. 1982 The uvrD gene of ~ coli encodes a DNA-dependent ATPase. Nature 298:98-100.

Ogawa, H., K.Shimada and J.Tomizawa. 1968 Studies on radiation­sensitive mutants of ~ coli. I. Mutants defective in the repair synthesis. Mol. Gen. Genet. 101:227-244.

Ossanna, N., K.R.Peterson and D.W.Mount. 1986 Genetics of DNA repair in bacteria. TIGS 2:55-58.

Ossanna, N., K.R.Peterson and D.W.Mount. 1987 UV-inducib1e SOS response in Escherichia coli. Photochem Photobio1 45:905-908.

Peterson, K.R., N. Ossanna, and D.W. Mount. 1988a K-12 1exA2 gene encodes a hypoc1eavab1e repressor. (in press).

The Escherichia coli J. Bacterio1. 170:

Peterson, K.R., N.Ossanna, A.T.Th1iveris, D.G.Ennis and D.W.Mount. 1988 Derepression of specific genes promotes DNA repair and mutagenesis in Escherichia ,coli. J. Bacterio1. 170:1-4.

Peterson, K.R., K.F.Wertman, D.W.Mount and M.G.Marinus. 1985 Viability of Escherichia coli K-12 DNA adenine methylase (dam) mutants requires increased expression of specific genes in the SOS regu1on. Mol. Gen. Genet. 201:14-19.

Phizicky, E.M., and J.W.Roberts. 1981 Induction of SOS functions: Regulation of proteolytic activity of ~ coli RecA protein by interaction with DNA and nucleoside triphosphate. Cel125:259-267.

Pukki1a, P.J., J.Peterson, G.Herman, P.Modrich and M.Mese1son. 1983 Effects of high levels of DNA adenine methylation on methyl-directed mismatch repair in Escherichia coli. Genetics 104:571-582.

82

Quillardet, P., P.L.Moreau, H.Ginsburg, D.W.Mount and R.Devoret. 1982 Cell survival, UV-reactivation and induction of prophage lambda in Escherichia coli K12 overproducing RecA protein. Mol. Gen. Genet. 188:37-43.

Radding, C. 1982 Homologous pairing and strand exchange in genetic recombination. Ann. Rev. Genet. 16:405-437.

Radman, M. and R. Wagner. 1986 Mismatch repair in Escherichia coli. Ann. Rev. Genet. 20:523-538.

Resnick, J., and R.Susssman. 1982 Escherichia coli single-strand DNA binding protein from wild-type and lexCl13 mutant affects in vitro proteolytic cleavage of phage lambda repressor. Proc. Natl. Acad. Sci. USA 79:2832-2835.

Roberts, J.W., and R.Devoret. 1983 Lysogenic induction. In Lambda II. R.W. Hendrix , J.W.Roberts, F.W.Stahl, R.A.Weisberg, (eds.) Pp. 123-144. Cold Spring Harbor. Cold Spring Harbor Laboratory.

Rupp, W.D., and P.Howard-Flanders. 1968 Discontinuities in the DNA synthesized in an excision-defective strain of Escherichia coli following ultraviolet irradiation. J. Mol. Bio. 31:291-304.

Schuster, H., D.Beyersmann, M.Mikolajczyk and M.Sch1icht. 1973 Prophage induction by high temperature in thermosensitive dna mutants lysogenic for bacteriophage lambda. J. Virol. 11:879-885.

Siegel, E.C. 1973 Ultraviolet-sensitive mutator strain of Escherichia coli K-12. J. Bacteriol. 113:145-160.

Siegel, E.C. 1981 Complementation studies with the repair-deficient uvrD3, uvrE156, and recL152 mutations in Escherichia coli. Mol. Gen. Genet. 184:526-530.

Siegel, E.C. 1983 The Escherichia coli uvrD gene is inducible by DNA damage. Mol. Gen. Genet. 191:397-400.

Smirnov, G.B., and A.G.Skavronskaya. 1971 Location of uvr502 mutation on the chromosome of Escherichia coli K-12. Mol. Gen. Genet. 113:217-221.

Smith, C.L., and M.Oishi. 1978' Early events and mechanisms in the induction of bacterial SOS functions: Analysis of the phage repressor inactivation process in vivo. Proc. Natl. Acad. Sci. USA 75:1657-1661.

Smith, G., 1983 General recombination. In Lambda II. R.W.Hendrix, J.W.Roberts, F.W.Stah1, R.A.Weisberg, (eds.) Pp. 175-209. New York. Cold Spring Harbor Laboratories.

83

Taucher-Scholz, G., M.Abde1-Monem and H.Hoffmann-Ber1ing. 1983 Functions of DNA he1icases in Escherichia coli. In Mechanisms of DNA replication and recombination. N.R.Cozzare11i, (ed.) Pp. 65-76. New York. Alan R. Liss, Inc •.

Tessman, E.S., and P.Peterson. 1985 Plaque color method for rapid isolation of novel recA mutants of Escherichia coli K-12: New classes of protease-constitutive recA mutants. J. Bacterio1. 163:677-687.

Tessman, E.S., and P.K.Peterson. 1985a Isolation of Protease­proficient, recombinase-deficient recA mutants of Escherichia coli K-12. J. Bacterio1. 163:688-695.

Walker, G.C. 1984 Mutagenesis and inducible responses to deoxyribonucleic acid damage in Escherichia coli. Microbio1. Revs. 48:60-93.

Wang, W.B., and E.S.Tessman. 1986 Location of functional regions of the Escherichia coli RecA protein by DNA sequence analysis of RecA protease-constitutive mutants. J. Bacterio1. 168:901-910.

Witkin, E.M., and T.Kogoma. 1984 Involvement of the activated form of RecA protein in SOS mutagenesis and stable DNA replication in Escherichia coli. Proc. Nat1. Acad. Sci. USA 81:7539-7543.

Witkin, E.M., J.O.McCa11, M.R.Vo1kert and I.E.Wermundsen. 1982 Constitutive expression of SOS functions and modulation of mutagenesis resulting from resolution of genetic instability at or near the recA locus of Escherichia coli. Mol. Gen. Genet. 185:43-50.

Witkin, E. 1976 Ultraviolet mutagenesis and inducb1e DNA repair in Escherichia coli. Bact. Revs. 40:869-907.

Yarranton, G.T., and M.L.Gefter. 1979 Enzyme-catalyzed DNA unwinding studies on Escherichia coli Rep protein. Proc. Nat1. Acad. Sci. USA 76:1658-16662.