166
Involvement of a DNA Polymerase III subunit in the Bacterial Response to Quinolones by Zakiya Whatley Department of Genetics & Genomics Duke University Date:_______________________ Approved: ___________________________ Ken Kreuzer, Supervisor ___________________________ Meta Kuehn ___________________________ Sue JinksRobertson ___________________________ Soman Abraham Dissertation submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy in the University Program in Genetics and Genomics in the Graduate School of Duke University 2014

ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

Page 1: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

 

Involvement  of  a  DNA  Polymerase  III  subunit  in  the  Bacterial  Response  to  Quinolones  

by  

Zakiya  Whatley  

Department  of  Genetics  &  Genomics  Duke  University  

 

Date:_______________________  Approved:  

 ___________________________  

 Ken  Kreuzer,  Supervisor    

___________________________  Meta  Kuehn  

 ___________________________  

Sue  Jinks-­‐‑Robertson    

___________________________  Soman  Abraham  

   

 

Dissertation  submitted  in  partial  fulfillment  of  the  requirements  for  the  degree    

of  Doctor  of  Philosophy  in  the  University  Program  in    Genetics  and  Genomics  in  the  Graduate  School  

of  Duke  University    

2014    

 

Page 2: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

ABSTRACT  

Involvement  of  a  DNA  Polymerase  III  subunit  in  the  Bacterial  Response  to  Quinolones  

by  

Zakiya  Nicole  Whatley  

Department  of  Genetics  &  Genomics  Duke  University  

 

Date:_______________________  Approved:  

 ___________________________  

Ken  Kreuzer,  Supervisor    

___________________________  Meta  Kuehn  

 ___________________________  

Soman  Abraham    

___________________________  Sue  Jinks-­‐‑Robertson  

   

 

An  abstract  of  a  dissertation  submitted  in  partial  fulfillment  of  the  requirements  for  the  degree  

of  Doctor  of  Philosophy  in  the  University  Program  in  Genetics  &  Genomics  in  the  Graduate  School  of  

Duke  University    

2014      

Page 3: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

                                                                       

Copyright  by  Zakiya  Whatley  

2014    

Page 4: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

iv  

Abstract Quinolone  treatment  induces  stabilized  cleavage  complexes  (SCCs),  consisting  of  

a  covalent  gyrase-­‐‑DNA  complex,  and  processing  of  these  complexes  is  thought  to  cause  

double-­‐‑strand  breaks  and  chromosome  fragmentation.  SCCs  are  required  but  not  

sufficient  for  cytotoxicity;  the  mechanism  that  converts  SCCs  to  double-­‐‑strand  breaks  is  

not  clearly  understood.  Evidence  of  chromosome  fragmentation  due  to  quinolones  

comes  from  indirect  measures  such  as  sedimentation  analysis  of  nucleoids  and  

measurements  of  lysis  viscosity.  This  work  outlines  a  method  that  combines  agarose  

plugs,  conditional  lysis  and  field  inversion  gel  electrophoresis  to  allow  direct  

visualization  of  chromosomal  fragmentation  resulting  from  quinolone  treatment.  We  are  

able  to  distinguish  between  latent  breaks  within  the  stabilized  cleavage  complex  and  

irreversible  breaks  that  result  from  downstream  processing.  

When  seeking  to  understand  the  genetic  requirements  for  quinolone-­‐‑induced  

SOS  response,  we  found  that  a  dnaQ  mutant  has  a  specific  defect  in  SOS  induction  

following  nalidixic  acid.  The  product  of  dnaQ  is  the  ε  subunit  of  DNA  polymerase  III,  

which  provides  3’  !  5’  exonuclease  activity.  In  addition  to  the  nalidixic  acid-­‐‑specific  

SOS  defect,  ΔdnaQ  has  multiple  phenotypes:  slow  growth,  high  mutation  frequency,  and  

constitutive  SOS.  We  propose  that  ε  has  a  role  in  the  quinolone  response  beyond  the  

normal  proofreading  function  of  the  subunit  in  the  polymerase  III  core.  Using  a  unique  

Page 5: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

v  

transposon  mutagenesis  system,  we  created  a  library  of  dnaQ  mutants  with  15  base  pair  

insertions  that  were  scored  phenotypically.  We  identified  mutants  that  separated  the  

various  phenotypes,  arguing  strongly  that  ε  has  multiple  functions.  The  isolation  of  a  

stable  dnaQ  mutant  with  SOS  phenotypes  allows  the  study  of  this  function  without  

confounding  results  from  spurious  mutations  throughout  the  chromosome.  We  also  

isolated  a  novel  class  of  SOS  “hyper-­‐‑inducible”  mutants.  Additionally,  my  findings  with  

weak  and  strong  β-­‐‑clamp  binding  mutants  provides  the  first  in  vivo  characterization  of  

these  ε  mutants  and  gives  insight  into  the  SOS  response  following  nalidixic  acid  

treatment.  

Page 6: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

vi  

Dedication

I  dedicate  this  dissertation  to  my  parents,  Mr.  and  Mrs.  Curtis  and  Vickie  

Whatley,  who  have  supported  me  at  every  step  in  my  academic  career.    

 

Page 7: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

vii  

Contents

Abstract  ..........................................................................................................................................  iv  

List  of  Tables  .................................................................................................................................  xi  

List  of  Figures  ..............................................................................................................................  xii  

Acknowledgements  ....................................................................................................................  xv  

List  of  Abbreviations  ..................................................................................................................  xvi  

1.  Background  ................................................................................................................................  1  

1.1  The  Bacterial  Replisome  ...................................................................................................  1  

1.2  Maintaining  Replication  Fidelity  ....................................................................................  4  

1.3  Repair  by  Recombination  Proteins  .................................................................................  5  

1.4  SOS  Response  ....................................................................................................................  7  

1.5  Translesion  Synthesis  .....................................................................................................  10  

1.6  Topoisomerases  ...............................................................................................................  11  

1.7  Quinolones  target  topoisomerases  ...............................................................................  14  

1.8  Requirements  for  Nalidixic  Acid-­‐‑Induced  SOS  .........................................................  16  

2.  DnaQ  Separation  of  Function  Plasmid-­‐‑Based  Assay  .........................................................  22  

2.1  Introduction  .....................................................................................................................  22  

2.1.1  Rationale  for  Linker  Scanning  Mutagenesis  ..........................................................  29  

2.2  Experimental  Procedures  ..............................................................................................  31  

2.2.1  Bacterial  Strains,  Plasmids  and  Materials  ..............................................................  31  

2.2.2  Linker  Scanning  Mutagenesis  ..................................................................................  32  

Page 8: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

viii  

2.2.2  Construction  of  Insertion  Library  ...........................................................................  32  

2.2.3  Plasmid-­‐‑Based  Screening  ..........................................................................................  35  

2.2.3.1  Qualitative  Measurement  of  Constitutive  SOS  ..............................................  35  

2.2.3.2  Preliminary  Mutator  Phenotype  Assessment  ................................................  36  

2.2.4  Gene  Replacement  .....................................................................................................  36  

2.2.5  Mutation  Rate  .............................................................................................................  37  

2.2.6  Quantitative  Western  Blots  ......................................................................................  37  

2.3  Results  ..............................................................................................................................  38  

2.3.1  Screening  for  Separation  of  Function  at  the  Plasmid  Level  .................................  38  

2.3.2  Separation  of  Function  after  Chromosomal  Gene  Replacement  ........................  44  

2.4  Discussion  ........................................................................................................................  53  

3.  Chromosome  Based  Separation  of  Function  Study  ............................................................  58  

3.1  Introduction  .....................................................................................................................  58  

3.2  Experimental  Procedures  ..............................................................................................  59  

3.2.1  Bacterial  Strains  and  Plasmids  .................................................................................  61  

3.2.2  Linker  Scanning  Mutagenesis  ..................................................................................  61  

3.2.3  Generating  dnaQ  mutant  library  .............................................................................  61  

3.2.4  Site-­‐‑Directed  Mutagenesis  ........................................................................................  63  

3.2.5  Gene  replacement  ......................................................................................................  63  

3.2.6  Quantitative  Beta-­‐‑galactosidase  Assay  ...................................................................  66  

3.2.7  Mutation  Rate  .............................................................................................................  67  

3.2.7  Site-­‐‑Directed  Mutagenesis  ........................................................................................  67  

Page 9: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

ix  

3.2.8  DnaE  Sequencing  .......................................................................................................  67  

3.3  Results  and  Conclusions  ................................................................................................  68  

3.3.1  Selection  and  Gene  Replacement  of  ε  Mutants  .....................................................  68  

3.3.2  Phenotypic  Scoring  ....................................................................................................  69  

3.3.3.  Overview  ....................................................................................................................  74  

3.3.4  Mutants  that  behave  like  Wildtype  .........................................................................  74  

3.3.5  Mutants  that  behave  like  ΔDnaQ  ............................................................................  75  

3.3.6  Proofreading  can  be  Uncoupled  from  SOS  Phenotypes  ......................................  80  

3.3.7  SOS  Phenotypes  are  Separable  ................................................................................  92  

3.3.8  β-­‐‑Clamp  Binding  Mutants  ........................................................................................  92  

3.3.9  Testing  for  dnaE  Suppressor  Mutations  .................................................................  94  

3.4  Discussion  ........................................................................................................................  95  

4.  Distinguishing  Between  Latent  SCC  Breaks  and  Irreversible,  Downstream  Breaks  ..  104  

4.1  Introduction  ...................................................................................................................  104  

4.1.1  Models  of  Quinolone  Cytotoxicity  ..........................................................................  105  

4.1.2  Importance  of  DNA  Extraction  Conditions  ...........................................................  111  

4.2  Experimental  Procedure  ..............................................................................................  114  

4.2.1  Materials  and  Bacterial  Strains  ..............................................................................  114  

4.2.2  Agarose  Plugs  ...........................................................................................................  115  

4.2.3  DNA  Preparations  ...................................................................................................  115  

4.2.5  Field  Inversion  Gel  Electrophoresis  ......................................................................  116  

4.3  Results  ............................................................................................................................  116  

Page 10: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

x  

4.3.1  Norfloxacin-­‐‑Treatment  Causes  Irreversible  Breaks  ............................................  117  

4.3.2  Nalidixic  Acid-­‐‑Treatment  Reveals  Fewer  Irreversible  Breaks  ..........................  119  

4.3.3.  ΔdnaQ,  ΔrecB,  and  gyrAR  Cleavage  &  Resealing  Conditions  ...........................  121  

4.4  Discussion  ......................................................................................................................  126  

5.  Conclusions  ............................................................................................................................  129  

5.1  Interpretations  &  Future  Directions  ...........................................................................  131  

5.2  Closing  Comments  .......................................................................................................  136  

Biography  ...................................................................................................................................  150  

Page 11: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

xi  

List of Tables Table  1:  Plasmids  used  for  Library  Construction  ...................................................................  33  

Table  2.  Dark  Blue  Plasmid  Mutants  with  Low  Mutator  Phenotypes  .................................  42  

Table  3.  Mutation  Rate  of  Chromosomal  Insertion  Mutants  ................................................  46  

Table  4.  Plasmid  Constructs  ......................................................................................................  62  

Table  5.  DnaQ  Insertion  Mutants  ..............................................................................................  71  

Table  6.  Mutation  Rate  and  SOS  Profiles  of  Insertion  Mutants  ............................................  78  

Table  7.  Summary  of  Separation  of  Function  Mutants  ..........................................................  98  

 

 

Page 12: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

xii  

List of Figures Figure  1.  Replisome  Structure  .....................................................................................................  3  

Figure  2.  Recombination  Pathways.  ...........................................................................................  6  

Figure  3.  Schematic  of  SOS  Induction  Pathways  ......................................................................  9  

Figure  4.  Gyrase  structure  ..........................................................................................................  13  

Figure  5.  dnaQ  Mutants  Fail  to  Induce  SOS  Following  Nalidixic  Acid  Treatment  ............  18  

Figure  6.  Quinolone  structures.  .................................................................................................  21  

Figure  7.  Replication  forks  can  proceed  with  the  α  subunit  alone  ......................................  23  

Figure  8.  Secondary  Structure  of  ε  ............................................................................................  25  

Figure  9.  Crystal  structure  of  ε186.  ...........................................................................................  26  

Figure  10.  Linker  Scanning  Mutagenesis.  ................................................................................  30  

Figure  11.  Schematic  of  dnaQ  Insertion  Mutant  Library  Construction  ...............................  34  

Figure  12.  Qualitative  measurement  of  SOS  Induction  .........................................................  39  

Figure  13.  Distribution  of  Chromosomal  Mutants  .................................................................  43  

Figure  14.  Quantitative  Western  Blots  of  RecA  in  Chromosomal  Mutants  ........................  49  

Figure  15.  Quantitative  Western  Blot  Confirms  ΔDnaQ  Partially  Constitutive  SOS  Phenotype  .....................................................................................................................................  50  

Figure  16.  Chromosomal  Insertion  Mutants  are  not  SOS  Constitutive  ...............................  51  

Figure  17.  Chromosomal  Mutants  are  not  SOS  defective  following  nalidixic  acid  treatment.  ......................................................................................................................................  52  

Figure  18.  Differential  expression  of  LexA-­‐‑regulated  genes  .................................................  56  

Figure  19.  Schematic  of  Insertion  Library  Construction  ........................................................  60  

Page 13: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

xiii  

Figure  20:  Mutant  Library  and  pMAK705  Cloning.  ...............................................................  65  

Figure  21.  Secondary  Structure  of  DnaQ  .................................................................................  70  

Figure  22.  Identification  of  Chromosomal  Gene  Replacement  Products  via  Gel  Electrophoresis  ............................................................................................................................  72  

Figure  23.  Kinetic  B-­‐‑Galactosidase  Assay  ................................................................................  73  

Figure  24.  SOS  &  Mutator  Profiles  of  Chromosomal  Insertion  Mutants  ............................  77  

Figure  25.  Mutant  115KLFKH  Insertion  Location  .....................................................................  79  

Figure  26.  Mutant  30IICLN  Insertion  Location  ........................................................................  81  

Figure  27.  Mutant  132V*  Insertion  Location  ..............................................................................  82  

Figure  28.  Crystal  structure  of  (εCTS35  -­‐‑  α270)  ........................................................................  83  

Figure  29.  Separation  of  Proofreading  and  Untreated  SOS  Levels  ......................................  84  

Figure  30.  Mutant  76CLNNK    Insertion  Location  ....................................................................  85  

Figure  31.  Separation  of  Proofreading  and  Nalidixic  Acid-­‐‑Induced  SOS  Activity  ...........  88  

Figure  32.  Relationship  between  the  Proofreading  and  Nalidixic  Acid-­‐‑Induced  SOS  .....  89  

Figure  33.  Mutant  143NSCLN  Insertion  Location  ....................................................................  90  

Figure  34.  Mutant  162LFKQL  Insertion  Location  .....................................................................  91  

Figure  35.  Separation  of  Constitutive  SOS  and  Nalidixic  Acid-­‐‑Induced  SOS  ....................  93  

Figure  36.  Model  for  Polymerase  Switching  .........................................................................  100  

Figure  37.  Replication-­‐‑independent  models  of  quinolone  cytotoxicity  ............................  109  

Figure  38.  Replication-­‐‑dependent  models  of  quinolone  cytotoxicity.  ...............................  110  

Figure  39.  Cleavage  conditions  reveal  the  latent  break  within  the  stabilized  cleavage  complex  .......................................................................................................................................  113  

Figure  40.  Norfloxacin-­‐‑treated  Wildtype  DNA  ....................................................................  118  

Page 14: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

xiv  

Figure  41.  Effects  of  Increasing  Nalidixic  Acid  on  Break  Formation  .................................  120  

Figure  42.  Break  Formation  in  Wildtype  and  ΔDnaQ  Treated  with  Nalidixic  Acid  &  Norfloxacin  .................................................................................................................................  122  

Figure  43.  Nalidixic  Acid  &  Norfloxacin-­‐‑treated  ΔRecB  and  gyrAR  under  cleavage  and  resealing  conditions.  .................................................................................................................  123  

 

 

Page 15: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

xv  

Acknowledgements I  would  like  to  acknowledge  the  supportive  group  of  family  and  friends  that  

have  been  the  wind  in  my  sails  throughout  my  graduate  school  career.  I  am  fortunate  to  

have  a  great  group  of  friends  that  I  can  call,  no  matter  the  hour.  I  would  like  to  

acknowledge  Dr.  Dawn  Scott,  Titilayo  Shodiya,  Dr.  Kia  Walcott,  Brittany  Curry,  Dr.  

Keisha  Findley,  and  Dr.  Fallon  Ukpe.  Additionally,  I  received  tremendous  support  

related  to  my  academic  growth  and  personal  well  being  from  Lana  Ben  David  and  the  

resources  available  through  the  Office  of  Graduate  Student  Affairs.  

I  was  fortunate  to  find  a  great  group  of  mentors  that  guided  my  trajectory  and  

supporters  that  pushed  me  to  try  harder  among  the  faculty  and  staff  at  Duke.  I  would  

like  to  acknowledge  my  thesis  committee  (Meta  Kuehn,  Sue  Jinks-­‐‑Robertson,  Shyam  

Unniraman,  Soman  Abraham  and  Ken  Kreuzer).  Thank  you  for  providing  valuable  

feedback  and  keeping  me  on  course  with  lots  of  help  and  suggestions.    

Thank  you  to  Dr.  Ken  Kreuzer  for  embracing  me  in  the  Kreuzer  Lab  Family.  We  

have  a  dynamic  lab  group  that  helped  make  my  graduate  school  experience  a  positive  

one.    

 

Page 16: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

xvi  

List of Abbreviations bp  –  basepairs  

DSBs  –  double-­‐‑strand  breaks  

dsDNA  –  double-­‐‑stranded  DNA  

CTS  –  C-­‐‑Terminal  Segment  

εCTS  –  Epsilon  C-­‐‑Terminal  Segment  

kb  –  kilobases  

NTD  –  N-­‐‑terminal  domain  

Pol  III  –  DNA  Polymerase  III    

SCCs  –  stabilized  cleavage  complexes  

ssDNA  –  single-­‐‑stranded  DNA  

TLS  –  translesion  synthesis  

 

Page 17: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

1  

1. Background Escherichia  coli  is  one  of  the  most  commonly  used  and  studied  prokaryotic  

organisms.  It  is  a  Gram-­‐‑negative  bacterium  that  has  been  used  in  basic  research  since  the  

early  20th  century.  Some  of  the  most  significant  strides  in  understanding  physiological  

processes,  molecular  genetics  and  evolution  have  been  made  using  E.  coli.  The  E.  coli  

chromosome  is  a  large  supercoiled  circle  of  ~4.6  million  basepairs  replicated  bi-­‐‑

directionally  from  the  origin.  Under  favorable  conditions,  the  bacterium  can  reproduce  

in  20  minutes,  replicating  the  chromosomal  DNA  and  separating  it  before  the  cell  

divides.  

1.1 The Bacterial Replisome

The  replication  machinery  in  E.  coli  consists  of  multiple  proteins  that  

simultaneously  unwind  double  stranded  DNA  (dsDNA)  and  synthesize  new  DNA.    The  

components  of  the  machinery  must  work  in  concert  to  replicate  the  template  DNA  both  

accurately  and  efficiently.  Generally,  replication  begins  at  the  origin,  oriC,  and  proceeds  

bidirectionally  by  two  separate  replisomes.  DnaA  unwinds  the  DNA  at  oriC  and  along  

with  DnaC,  assists  the  loading  of  the  replicative  helicase,  DnaB  at  the  origin  [1,  2].    

The  dsDNA  at  the  replication  fork  is  unwound  in  opposite  directions  by  each  

DnaB  helicase,  a  six-­‐‑subunit  structure  organized  into  a  ring  that  pulls  the  lagging  strand  

of  the  template  through  [3].  DnaB  maintains  contact  with  DNA  primase  (DnaG),  

Page 18: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

2  

stimulating  the  addition  of  RNA  primer  synthesis  along  the  lagging  strand  [4].  

Following  the  unwinding  of  DNA,  the  clamp  loader  binds  to  SSB  (single-­‐‑strand  binding  

protein)-­‐‑coated  ssDNA.  At  each  fork,  the  clamp  loader  complex  (τ3δδ’χψ)  loads  two  

dnaN  encoded  β  subunits,  which  dimerize  to  form  the  β-­‐‑clamps,  per  polymerase  

holoenzyme.  Interaction  of  β-­‐‑clamp  with  the  clamp  loader  complex  links  the  two  

polymerases  as  they  synthesize  nascent  strands  in  opposite  directions.  

The  replication  polymerase,  DNA  Polymerase  III  (pol  III),  executes  high  fidelity,  

processive  DNA  synthesis.  Single  molecule  studies  have  determined  that  pol  III  

synthesizes  DNA  at  653  ±  9  nucleotides  per  second  and  in  vitro  studies  of  wildtype  DNA  

polymerase  III  captured  fork  speeds  at  900  nucleotides  per  second  [5,  6].  Although  there  

are  only  two  nascent  strands  of  DNA,  recent  experiments  have  shown  there  are  actually  

three  pol  III  cores  that  associate  with  the  replication  fork  [7,  8]  (Figure  1).  The  pol  III  core  

consists  of  three  subunits,  α  ε  and  θ.  The  α  subunit  contains  the  catalytic  activity  of  the  

polymerase,  while  εis  responsible  for  proofreading  in  3’  !  5’  direction.  There  is  no  

clearly  defined  function  of  the  θ  subunit.  Rnase  H,  Pol  I  and  DNA  ligase  complete  

maturation  of  Okazaki  fragments  as  they  remove  the  RNA  primers  and  reseal  the  nicks  

left  behind  on  the  nascent  strand  synthesized  from  the  lagging  strand.  

Page 19: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

3  

 

Figure  1.  Replisome  Structure  

 Recent  single-­‐‑molecule  studies  have  shown  that  the  replisome  often  has  three  polymerases  and  beta  clamps  rather  than  two.  The  model  proposes  that  two  polymerases  are  allocated  for  synthesis  on  the  lagging  strand  enhancing  Okazaki  fragment  synthesis.  Figure  from  [9]  

     

Page 20: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

4  

1.2 Maintaining Replication Fidelity

Replication  must  be  reliable  and  relatively  error-­‐‑free  to  preserve  the  genetic  

information  transferred  from  the  mother  to  daughter  cell.  The  cell  dedicates  multiple  

pathways  towards  this  end.  This  includes  multiple  repair  pathways  that  assist  in  the  

preservation  of  DNA  before  it  is  replicated  and  measures  to  ensure  accurate  replication.  

In  vivo  studies  have  determined  the  rate  of  spontaneous  errors  per  round  of  replication  

is  5  per  1010  basepairs  [10,  11].  This  begins  with  the  selective  α  subunit,  which  houses  the  

catalytic  activity  of  the  polymerase.  As  an  additional  safeguard,  the  ε  subunit,  encoded  

by  dnaQ,  provides  the  3’  !  5’  exonuclease  activity,  editing  bases  that  were  incorrectly  

inserted  by  α.  Additional  incorrectly  paired  bases  are  removed  by  the  mismatch  repair  

system,  managed  by  proteins  MutH,  MutL,  and  MutS  [12,  13].  

When  replication  forks  progress,  they  are  constantly  threatened  and  the  cell  has  

many  pathways  dedicated  to  protecting  the  forks.  Endogenous  damage  can  block  

replication  fork  progression  and  lead  to  blockage,  collapse  or  reversal  of  the  fork.  

Recombination  proteins  facilitate  the  resolution  or  repair  of  forks  as  they  encounter  

damage.  As  replication  proceeds,  nicks  in  the  parental  DNA  can  lead  to  free  double  

strand  ends  that  must  be  protected  and  repaired.  Also  lesions  on  the  leading  or  lagging  

strand  can  lead  to  gaps  that  are  repaired  later.  

Page 21: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

5  

1.3 Repair by Recombination Proteins

There  are  two  major  pathways  of  recombinational  repair  mediated  by  RecBCD  or  

RecFOR.  Both  pathways  rely  on  homologous  recombination.  Also,  they  both  require  the  

protein  RecA,  which  loads  onto  single-­‐‑strand  DNA  (ssDNA)  and  promotes  strand  

invasion  of  a  homologous  duplex  of  DNA.  

 When  double  strand  ends  are  created,  they  are  recognized  and  processed  by  

RecBCD.  RecBC  is  responsible  for  the  exonuclease  activity  of  the  complex  while  RecD  

recognizes  the  Chi  sites.  Upon  Chi  site  recognition,  the  enzyme  catalyzes  RecA  loading  

onto  ssDNA.  The  RecA  filament  invades  homologous  DNA,  creating  a  D-­‐‑loop  (Figure  

2).    

When  the  replication  fork  encounters  a  nick  on  the  leading  strand,  it  can  create  a  

broken  fork  with  a  one-­‐‑ended  DSB.  RecBCD  and  RecA-­‐‑mediated  strand  invasion  forms  

a  D-­‐‑loop.  The  D-­‐‑loop  allows  the  loading  of  PriA,  which  then  recruits  the  other  

components  of  the  full  primosome  (PriB,  PriC  DnaT,  DnaC,  and  DnaG)  and  replication  

restarts  [14].  As  the  replication  fork  migrates,  this  creates  a  Holliday  junction  that  is  

resolved  by  RuvABC  [15].  

Page 22: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

6  

 

 Figure  2.  Recombination  Pathways.  

A)  Single  strand  gaps  are  repaired  via  RecJ  and  RecFOR.  RecFOR  replaces  SSB  with  RecA,  facilitating  strand  invasion  and  creating  Holliday  junctions.  Holliday  junctions  are  resolved  by  RuvABC.  In  RuvABC  mutants,  RecG  resolves  Holliday  junctions.    B)  Double  strand  ends  are  degraded  by  the  RecBCD  until  reaching  a  Chi  site.  RecA  is  loaded  onto  ssDNA,  facilitating  strand  invasion.  The  Holliday  junction  is  resolved  by  RuvABC.    Figure  from  [16]  

Page 23: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

7  

RecFOR  mediates  recombination  at  single-­‐‑strand  gapped  DNA  (Figure  2).  This  

pathway  is  useful  in  response  to  daughter  strand  gaps  that  may  arise  from  replication  

fork  blockage  by  a  non-­‐‑coding  lesion  and  subsequent  replication  reinitiation  

downstream.  To  begin  repair,  the  RecJ  exonuclease  enlarges  the  gap  with  the  help  of  SSB  

and  the  RecQ  helicase.  SSB  on  the  ssDNA  is  replaced  with  RecA  by  the  RecFOR  

complex.  RecFR  binds  at  the  ssDNA/dsDNA  junction  and  RecO  removes  SSB.  

RecA:ssDNA  filaments  invade  homologous  dsDNA  and  create  double  Holliday  

junctions  that  are  resolved  by  RuvABC  [16].  

1.4 SOS Response

RecA  has  a  role  in  stimulating  strand  invasion  following  recombinational  repair  

of  forks  but  it  also  has  a  role  in  the  induction  of  the  SOS  response.  The  SOS  response  was  

first  described  by  Miroslav  Radman  as  the  cell’s  reaction  to  a  DNA  damage  distress  

signal  [17].  The  SOS  regulon  contains  over  40  genes  involved  in  a  variety  of  processes  

including  DNA  damage  tolerance  and  repair  [18].  Previous  studies  have  used  SOS  to  

monitor  cells’  response  to  various  types  of  damage  including  irradiation  and  drug  

treatment.  Not  all  SOS  genes  are  involved  in  repair  and  recombination  and  delay  of  the  

cell  cycle;  some  have  unknown  function.  SOS  genes  are  generally  expressed  to  help  cell  

survival  in  the  presence  of  DNA  damage.    

Page 24: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

8  

The  SOS  pathway  is  under  the  control  of  LexA,  the  repressor,  and  RecA,  the  

activator.  When  damaged  DNA  induces  SOS,  it  is  funneled  through  one  of  two  

pathways,  RecBCD  or  RecFOR  (Figure  3).  DNA  damage  that  causes  single-­‐‑strand  gaps  is  

funneled  through  the  RecFOR  pathway  to  SOS,  while  damage  that  creates  double-­‐‑strand  

ends  is  processed  via  RecBCD.  UV  radiation  and  the  drug  mitomycin  C  induce  SOS  in  a  

RecFOR-­‐‑dependent  manner[19].  Nalidixic  acid  and  other  quinolones  induce  SOS  in  a  

RecBCD-­‐‑dependent  manner.  Both  pathways  lead  to  RecA:ssDNA  filaments  assisting  the  

proteolysis  of  LexA,  leading  to  the  transcription  of  genes  in  the  SOS  regulon.  

Earlier  studies  of  the  SOS  response  in  cell  populations  presented  a  simple  model  

of  SOS  induction  whereby  DNA  damage  caused  a  global  increase  in  LexA-­‐‑regulated  

gene  expression  [20,  21].  However,  single  cell  studies  revealed  a  temporal  pattern  of  SOS  

induction.  This  means  as  damage  increases,  the  cells  have  additional  cycles  of  SOS  

activation[22].  Additionally,  with  each  additional  wave  the  magnitude  of  the  response  

remains  the  same,  rather  than  a  greater  amount  of  expression  after  multiple  waves[22].  

The  expression  of  SOS  genes  directly  relates  to  how  tightly  LexA  binds  to  the  

promoter  .  SOS  induction  begins  with  increased  expression  of  genes  involved  in  

nucleotide  excision  repair  (uvrB,  uvrD)  then  slowly  recA  and  genes  involved  in  

homologous  recombination  are  expressed.  Cell  division  is  inhibited  by  expression  of  

sfiA.  If  SOS  persists  the  final  response  is  induction  of  the  error-­‐‑prone  polymerase,  Pol  V  

Page 25: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

9  

 

Figure  3.  Schematic  of  SOS  Induction  Pathways  

There  are  two  pathways  of  SOS  induction  –  RecBC  and  RecFOR.  Damaging  agents  like  quinolones,  which  create  double  strand  breaks,  induce  via  RecBC.  Damage  that  creates  single  strand  gaps,  such  as  UV  irradiation  and  mitomycin  C,  use  the  RecFOR  pathway.

DNA  Damage

Double  Strand  Breaks Single  Strand  Gaps

RecBC RecFOR

RecA

SOS

Page 26: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

10  

(described  below).  SOS  is  a  finely  tuned  response  to  damage  in  the  cell;  it  begins  with  

error-­‐‑free  methods  of  repair  then  builds  the  response  with  the  final  effort  to  replicate  the  

DNA  using  the  error-­‐‑prone  polymerase.  This  strategy  gives  the  cell  another  cycle  to  deal  

with  the  damage  [23].  

1.5 Translesion Synthesis

Prolonged  SOS  induction  leads  to  the  expression  of  translesion  polymerases.  

Translesion  synthesis  allows  replication  through  damaged  sites  or  lesions.  Unlike  the  

highly  selective  replicative  polymerase,  translesion  polymerases  are  low  fidelity  and  

have  larger  active  sites  that  allow  them  to  replicate  across  from  bulky  lesions.  The  idea  

behind  these  error-­‐‑prone  polymerases  is  that  they  allow  replication  to  continue  and  the  

initial  lesion  can  be  repaired  in  the  next  cycle  of  replication.  SOS  induction  leads  to  the  

expression  of  the  three  translesion  polymerases,  Pol  II,  Pol  IV  and  V[24].  PolII  is  encoded  

by  polB/dinA,  and  Pol  IV  is  encoded  by  dinB/dinP  and  is  conserved  throughout  all  

domains  [25].  Pol  V  is  encoded  by  umuDC  and  comprised  of  UmuD’2C.    

Pol  II  plays  a  role  in  replication  restart  following  UV  exposure  and  SOS  

induction  increases  its  expression  ~7  fold  [26,  27].  It  has  proofreading  activity  and  has  

been  shown  to  be  involved  in  replicating  past  abasic  and  crosslinking  lesions  [28,  29].  

DNA  polymerases  IV  and  V  are  Y-­‐‑family  polymerases,  and  their  specificity  for  different  

templates  has  been  attributed  to  their  structural  differences[30].  Pol  IV  has  a  high  

Page 27: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

11  

affinity  for  N2-­‐‑guanine  adducts  while  Pol  V  has  high  fidelity  for  cyclobutane  pyrimidine  

dimers.  Pol  IV  is  not  processive  and  does  not  have  exonuclease  activity;  under  normal  

conditions  it  does  not  cause  spontaneous  mutations  but  overexpression  leads  to  mutator  

phenotype[31-­‐‑33].  Pol  V  is  both  transcriptionally  and  translationally  regulated.  In  

addition  to  expression  of  umuC  and  umuD  following  SOS  induction,  the  UmuD  protein  

must  be  cleaved  into  UmuD’,  then  form  the  homodimer  UmuD’2  and  associate  with  

UmuC  [34].  While  all  the  TLS  polymerases  interact  with  the  β-­‐‑clamp,  Pol  V  also  requires  

RecA  and  SSB  [35].  

1.6 Topoisomerases

Topoisomerases  are  ubiquitous  enzymes  that  create  transient  breaks  in  the  DNA  

backbone  to  regulate  the  topology  of  DNA.  In  bacteria  there  are  four  topoisomerases  

responsible  for  the  chromosome:  topoisomerases  I,  II,  III  and  IV.  These  dynamic  

enzymes  manipulate  the  double  helix,  making  it  accessible  for  essential  activities  such  as  

replication,  transcription,  and  recombination  [36].    

Topoisomerases  introduce  positive  and  negative  supercoils,  compact  the  DNA  

and  change  the  linking  number  by  creating  well-­‐‑protected  transient  breaks,  thus  altering  

the  topology  of  the  DNA  [37].  There  are  two  types  of  topoisomerase,  type  I  and  type  II.    

Type  I  topoisomerases,  topo  I  and  III,  cleave  single  strands  of  DNA,  one  at  a  time.  Type  

II  topoisomerases,  gyrase  (Figure  4)  and  topoisomerase  IV,  simultaneously  cleave  both  

Page 28: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

12  

strands  of  the  helix.  In  both  groups,  the  reaction  occurs  via  transesterification  reaction  

between  the  catalytic  tyrosine  of  the  enzyme  and  a  phosphate  group  on  the  backbone  of  

the  DNA.  

Page 29: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

13  

 

Figure  4.  Gyrase  structure  

Gyrase  is  a  tetramer  comprised  of  two  subunits  of  GyrA  (blue)  and  two  subunits  of  GyrB  (yellow).  The  DNA  interacts  with  the  C-­‐‑terminal  domain  of  GyrA,  wrapping  

around  the  enzyme.  The  G-­‐‑segment  represents  the  gate  segment  of  DNA  that  is  broken.  The  T-­‐‑segment  (red)  represents  the  transfer  segment  that  passes  through  the  gate  

segment  of  the  DNA.  Image  from  [38]  

Page 30: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

14  

Topoisomerases  are  drug  targets  in  both  eukaryotic  and  prokaryotic  cells.  Many  

antitumor  drugs,  such  as  etoposide  and  camptothecin,  target  eukaryotic  topoisomerase  

enzymes.  The  quinolone  class  of  drugs  target  type  II  topoisomerases  in  bacteria.  In  

Gram-­‐‑positive  bacteria,  the  primary  target  is  topoisomerase  IV,  while  the  target  is  gyrase  

in  Gram-­‐‑negative  bacteria  [39,  40].  Topoisomerase  IV  is  a  four  subunit  protein,  

comprised  of  two  subunits  each  of  gene  products  from  parC  and  parE  [41].  Gyrase  has  

two  subunits:  gyrA  and  gyrB.  The  A2B2  tetramer  allows  dsDNA  to  pass  through  its  gate-­‐‑

like  structure  (Figure  4).  Gyrase  cleaves  DNA  while  remaining  bound  to  the  5’  end  of  the  

DNA;  this  transient  intermediate  is  known  as  a  cleavage  complex.  Once  the  transfer  

segment  of  DNA  passes  through  the  gate,  gyrase  reseals  the  latent  break  and  dissociates  

[42].  

1.7 Quinolones target topoisomerases

Quinolones  are  a  widely  prescribed  class  of  anti-­‐‑bacterial  drugs  created  during  

the  last  45  years.  The  parent  drug  of  the  class,  nalidixic  acid,  was  created  in  1962.  In  

addition  to  novel  drugs  in  this  class,  a  subclass  was  created  by  the  addition  of  fluorine  at  

the  C-­‐‑6  position.  The  subclass,  fluoroquinolones,  showed  increased  bactericidal  activity  

against  Gram-­‐‑negative  bacteria.  Modifications  to  fluoroquinolones  and  novel  non-­‐‑

fluorinated  quinolones  have  exhibited  even  greater  activity,  proving  effective  against  

Gram-­‐‑negative  and  Gram-­‐‑positive  organisms  [43].  Common  agents  used  to  understand  

Page 31: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

15  

the  quinolone  mechanism  of  action  include  nalidixic  acid,  oxolinic  acid,  norfloxacin,  and  

ciprofloxacin.  The  broad  effectiveness  of  these  drugs  has  led  to  their  prevalent  use  in  the  

medical  and  veterinary  fields[43].    

Overall,  the  quinolone  mechanism  of  cytotoxicity  is  inhibition  of  DNA  synthesis.  

There  have  been  proposals  for  multiple  pathways  of  quinolone  cytotoxicity  based  on  the  

particular  drug.  Generally,  the  targeted  topoisomerases  are  gyrase  in  Gram-­‐‑negative  

organisms  and  topoisomerase  IV  in  Gram-­‐‑positive  organisms  [44].  Quinolones  that  

target  gyrase,  like  nalidixic  acid,  rapidly  inhibit  DNA  synthesis,  while  those  targeting  

topoisomerase  IV  slowly  inhibit  replication.  This  reflects  the  location  of  the  target  

enzymes  in  respect  to  the  replication  fork  machinery[45].  There  are  quinolones  that  

target  gyrase  in  Gram-­‐‑positive  organisms.  Also,  of  the  newer  quinolones,  there  are  some  

that  target  both  gyrase  and  topoisomerase  IV  equally[43].  

Quinolone  resistance  is  acquired  by  modifying  the  targeted  topoisomerases  or  by  

minimizing  the  amount  of  drug  in  the  cell.  Target  modification  includes  mutations  in  

either  of  the  subunits  of  gyrase  (gyrA  or  gyrB)  and/or  topoisomerase  IV  (parC  or  parE)  

[43,  46].  Resistance  to  nalidixic  acid  is  achieved  by  acquiring  mutations  in  gyrase;  

however,  while  one  mutation  in  gyrase  is  sufficient  for  nalidixic  acid  resistance,  multiple  

mutations  can  be  required  for  resistance  to  more  potent  fluoroquinolones  [43].  

Page 32: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

16  

Modifications  that  limit  the  absorption  of  quinolones  affect  membrane  permeability  and  

the  efflux  pump  systems  [43].    

1.8 Requirements for Nalidixic Acid-Induced SOS

Previous  studies  in  our  lab  aimed  to  understand  the  molecular  mechanism  of  

the  bacterial  response  to  quinolones.  Using  a  strain  that  has  the  lacZ  gene  fused  to  dinD,  

a  gene  in  the  SOS  regulon,  we  completed  a  series  of  genetic  screens  to  understand  gene  

products  involved  in  the  bacterial  response  to  nalidixic  acid.  Our  initial  screen  of  over  

19,000  transposon  mutants  selected  for  those  that  failed  to  induce  SOS  following  

nalidixic  acid  treatment.  Of  the  18  mutants  that  have  a  specific  defect  in  nalidixic  acid-­‐‑

induced  SOS,  all  were  insertions  mapped  to  recombination  genes  recB  and  recC  [47].  

This  result  was  not  completely  unexpected  given  that  previous  studies  have  shown  that  

the  RecBC  complex  is  necessary  for  nalidixic  acid  treated  cells  to  induce  SOS  [48].  

However,  our  screen  was  not  comprehensive  and  only  included  mutants  without  SOS  

expression  in  the  absence  of  drug.  Our  approach  excluded  mutants  that  were  SOS  

constitutive  and  had  detectable  levels  SOS  without  an  inducer.    

A  screen  for  constitutive  SOS  transposon  mutants  determined  that  dnaQ  mutants  

were  partially  constitutive  in  the  absence  of  drug  [49].  Additionally,  dnaQ  mutants  are  

unable  to  induce  SOS  following  nalidixic  acid  treatment.  DnaQ  encodes  ε,  the  

proofreading  subunit  of  DNA  polymerase  III.  The  dnaQ  mutant  did  not  have  a  general  

Page 33: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

17  

defect  in  SOS  induction,  as  it  was  able  to  induce  SOS  in  a  RecFOR-­‐‑dependent  manner  

following  mitomycin  C  treatment  (Figures  2-­‐‑5)[50].  The  nalidixic  acid-­‐‑specific  defect  in  

SOS  piqued  our  interest  in  dnaQ  and  led  us  to  hypothesize  that  epsilon  has  a  role  in  the  

E.  coli  quinolone  response.

Page 34: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

18  

 

 

Figure  5.  dnaQ  Mutants  Fail  to  Induce  SOS  Following  Nalidixic  Acid  Treatment  

Here,  the  partially  constitutive  SOS  phenotype  of  dnaQ-­‐‑  is  apparent.  Also,  it  is  clear  that  dnaQ-­‐‑  can  induce  SOS  in  response  to  mitomycin  C  (MMC)  but  not  in  response  to  nalidixic  acid.  Numbers  above  bars  indicate  replicates.  Modified  from  [50].  

Page 35: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

19  

Overview  

This  work  is  largely  centered  on  the  function  of  ε,  which  provides  proofreading  

and  structural  fidelity  to  the  pol  III  core.  There  are  multiple  phenotypes  of  ε  knockouts,  

and  our  proposal  that  it  has  a  specific  role  in  the  bacterial  response  to  quinolones  

prompted  our  search  for  separation  of  function  mutants.  Uncoupling  of  the  phenotypes  

proves  the  ability  of  ε  to  function  in  one  role  without  the  other.  I  will  outline  a  strategy  

of  generating  mutations  that  are  more  likely  to  generate  separation  of  function  ε  mutants  

and  the  phenotypic  scoring  of  these  alleles.  

This  work  also  focuses  on  the  well-­‐‑studied  parent  drug,  nalidixic  acid,  and  the  

first  generation  fluoroquinolone,  norfloxacin  (Figure  6).  Nalidixic  acid  stabilizes  the  

transient  intermediate  step  when  gyrase  is  covalently  bound  to  DNA.  These  stabilized  

cleavage  complexes  block  replication  [51].  Stabilized  cleavage  complexes  are  reversible,  

and  latent  breaks  at  gyrase  are  resealed  when  the  drug  is  washed  out  or  treated  with  

mild  heat  [39].  So  while  stabilized  cleavage  complexes  are  necessary,  they  are  not  

sufficient  for  quinolone  cytotoxicity.  Processing  of  stabilized  cleavage  complexes  leads  

to  strand  breaks;  however  the  details  of  this  processing  are  not  clearly  understood  [46,  

52].  

Multiple  models  for  the  processing  of  stabilized  cleavage  complexes  into  double-­‐‑

strand  breaks  have  been  proposed.  Some  studies  suggest  replication  is  necessary  for  

double-­‐‑strand  break  generation,  while  others  present  evidence  that  inhibiting  replication  

Page 36: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

20  

or  protein  synthesis  does  not  block  cytotoxicity  in  potent  fluoroquinolones  like  

ciprofloxacin  (Figure  6).  The  various  models  will  be  discussed  in  more  detail  in  later  

chapters.  In  the  following  chapters,  I  will  discuss  the  possible  role  of  the  replisome,  

specifically  the  ε  subunit  of  DNA  polymerase  III,  in  the  bacterial  response  to  quinolones.  

I  also  outline  a  method  of  distinguishing  between  latent  breaks  of  the  cleavage  complex  

and  downstream,  irreversible  breaks  following  quinolone  treatment.  

Page 37: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

21  

 

 

 

Figure  6.  Quinolone  structures.  

The  parent  drug  of  the  quinolone  class  is  nalidixic  acid.  The  fluoroquinolone  class,  which  includes  norfloxacin  and  norfloxacin,  has  fluorine  at  the  C-­‐‑6  position.    

Nalidixic  acid  

Norfloxacin  

Ciprofloxacin  

Page 38: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

22  

2. DnaQ Separation of Function Plasmid-Based Assay

2.1 Introduction

The  product  of  dnaQ  is  the  ε  subunit  of  the  replication  polymerase,  DNA  

polymerase  III,  and  provides  the  3’  !  5’  exonuclease  activity  of  the  core  [53].  The  

protein  physically  interacts  with  the  α  and  θ  subunits,  gene  products  of  dnaE  and  holE  

respectively;  together,  these  3  subunits  comprise  the  polymerase  core  [54,  55].  The  

polymerase  activity  is  in  α,  while  the  proofreading  function  resides  in  ε.  Beyond  binding  

to  ε,  the  function  of  θ  has  not  been  clearly  defined  [56,  57].  ε  bridges  the  α  and  θ  

subunits  as  in  vitro  studies  found  no  detectable  interaction  between  α  and  θ  [56].  In  

addition  to  fidelity,  ε  adds  structural  integrity.  There  is  conflicting  evidence  on  the  

necessity  of  ε  for  replication  fork  progression.  Marians  et  al.  showed  that  only  α  is  

required  for  replication  fork  processivity  on  rolling  circle  templates  (Figure  7)  [58].  Their  

in  vitro  analysis  of  replication  fork  speed  is  the  same,  even  in  the  absence  of  ε,  but  the  

presence  of  ε  makes  the  fork  much  more  efficient  (Figure  7).  Other  studies  found  that  the  

rate  of  processivity  on  primed  single-­‐‑stranded  templates  is  increased  upon  α-­‐‑ε  binding  

[56,  59].  Binding  of  α  stimulates  ε  exonuclease  activity  on  both  double-­‐‑strand  and  single-­‐‑

strand  templates  [60,  61].  In  return,  ε  is  also  able  to  stimulate  polymerase  activity  of  

α.[59]

Page 39: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

23  

 

 

Figure  7.  Replication  forks  can  proceed  with  the  α  subunit  alone  

In  vitro  reactions  demonstrate  replication  fork  speed  within  1  minute  on  rolling  circle  templates.  It  is  clear  that  the  fork  is  much  less  efficient  in  the  absence  of  ε.  However,  α  completes  synthesis  of  at  a  rate  similar  to  the  pol  III  core  and  αε.  All  fork  rates  are  within  650-­‐‑660  nucleotides  per  second.  Figure  modified  from  [58].  

Page 40: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

24  

Epsilon  is  a  243  amino  acid  protein  comprised  of  the  N-­‐‑terminal  proofreading  

domain  and  the  α-­‐‑subunit  binding  C-­‐‑terminal  segment  (εCTS)  (Figure  8).  The  N  

terminal  domain  (1-­‐‑186)  contains  three  conserved  exo  motifs,  the  exonuclease  active  site  

and  residues  responsible  for  binding  to  θ  [62,  63].  The  three  exo  motifs  were  originally  

proposed  to  be  exoI  (8-­‐‑21),  exoII  (95-­‐‑108)  and  exoIII  (209-­‐‑223)  as  determined  by  sequence  

analysis  of  33  DNA  polymerases  and  site-­‐‑directed  mutagenesis  of  T7,  ϕ29,  and  PolI  

polymerases  [64,  65].  However,  analysis  of  the  Bacillus  subtilis  3’!5’  exonuclease  

identified  a  novel  exo  motif,  exoIIIε,  which  was  required  for  exonuclease  activity  [66].  

The  ExoIIIε  motif  was  identified  in  ε  and  deemed  the  relevant  motif  rather  than  exoIII.  

This  was  later  confirmed  by  additional  sequencing  analysis  of  148  3’  exonucleases  [67].  

This  is  also  supported  by  mutational  analysis  indicating  that  residues  within  ExoIIIε  

(128-­‐‑192)  affect  mutation  frequency  [68].  Additionally,  in  vitro  exonuclease  reactions  

with  truncated  ε  (ε186)  show  the  catalytic  function  resides  in  the  first  186  residues  of  the  

protein  (Figure  9).  While  6  amino  acids  of  ExoIIIε  are  not  included  in  the  ε186  protein,  

none  of  exoIII  is  included  in  this  construct,  showing  that  residues  209-­‐‑223  (exoIII)  are  not  

necessary  for  exonuclease  activity  [60].  

 The  εCTS  is  flexible  with  a  proposed  Q-­‐‑linker  region  (residues  190-­‐‑212)  that  is  

known  to  join  regions  with  different  functions  within  proteins  [69,  70].  Within  the  Q-­‐‑

linker  there  is  a  stretch  of  four  glutamines  (194-­‐‑197)  that  remains  flexible  even  when  in  

complex  with  α,  as  identified  by  NMR  spectroscopy  [71].  The  α-­‐‑binding  function  of  

Page 41: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

25  

 

 Figure  8.  Secondary  Structure  of  ε  

A.  N186  contains  exo  motifs  I,  II  and  III.  The  Q-­‐‑linker  is  in  εCTS.  The  β-­‐‑clamp  binding  motif  is  at  the  junction  of  N186  and  εCTS.  B.  This  composite  structure  (based  on  PDB  ID:  2xy9  and  4gx9)  includes  residues  2  -­‐‑186  and  200-­‐‑243.  There  is  no  structural  information  for  the  flexible  region  187-­‐‑  210.  Yellow  arrows  represent  β-­‐‑sheets  and  red  waves  represent  α-­‐‑helices.  Purple  humps  are  turns.  Figure  from  [72,  73]  

| 210

| 220

| 230

Page 42: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

26  

 

 

 

 

 

Figure  9.  Crystal  structure  of  ε186.  

The  crystal  structure  of  the  N-­‐‑terminal  truncation  of  the  epsilon  subunit,  ε186  from  [74].  In  this  model,  the  truncated  protein  is  coupled  with  thymidine-­‐‑5’-­‐‑monophosphate  (blue  molecule),  two  Mn(II)  ions  at  the  active  site,  and  DNA  (orange).  

Page 43: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

27  

εCTS  has  been  shown  biochemically  and  genetically  [60,  75].  Full  length  ε  co-­‐‑fractionates  

with  α,  but  a  truncation  of  ε  missing  the  57  C-­‐‑terminal  amino  acids  does  not  [60].    

There  have  been  contradicting  models  in  the  literature  about  the  specific  residues  

necessary  for  binding  to  α.  Mutational  analysis  of  ε  revealed  dnaQ932  (W241ochre)  has  

mutation  frequencies  15  fold  higher  than  wildtype  even  though  it  has  a  fully  functional  

N-­‐‑terminal  exonuclease  domain.  This  mutant  lacks  the  last  three  C-­‐‑terminal  residues,  

WRA,  and  is  recessive  in  complementation  studies.  These  results  suggest  that  the  three  

terminal  residues  are  necessary  for  binding  to  α  [68].  However,  a  study  of  ε  proteolysis  

reported  that  residues  187-­‐‑213  are  responsible  for  binding  to  α  [76].  Recent  studies  also  

identified  a  β-­‐‑clamp  binding  motif  (QTSMAF  at  amino  acids  182-­‐‑187)  and  a  possible  site  

of  interaction  with  chaperone  DnaK  within  the  εCTS  that  has  not  been  well  defined  [76,  

77].  

Studies  of  mutations  that  lead  to  global  increases  in  mutation  frequency  

identified  multiple  alleles  that  later  mapped  to  dnaQ.  Three  of  the  strongest  alleles  were  

mutD5  (T15I),  dnaQ926  (D12A  &  Q14A)  and  dnaQ49  (V96G)  [78-­‐‑81].  These  proteins  were  

studied  to  determine  exactly  how  different  mutations  affected  the  proofreading  activity  

of  ε.  MutD5  is  an  extreme  mutator,  with  mutation  frequencies  as  high  as  3,000  fold  

higher  than  wildtype  levels  using  rifampicin  resistance  as  a  measure  [82,  83].  Studies  

using  a  yeast  two-­‐‑hybrid  system  showed  that  mutD5  and  dnaQ926  do  not  disrupt  the  

interaction  with  α  or  θ,  and  oddly,  DnaQ49  had  a  weaker  interaction  with  α  than  

Page 44: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

28  

wildtype  DnaQ  [84].  These  results  agreed  with  sequencing  data  that  was  used  to  

determine  the  functional  exo  motifs,  as  those  mutations  localized  to  motifs  exo  I  and  exo  

II.  DnaQ49  is  a  temperature  sensitive  mutant  with  negligible  mutator  activity  at  28-­‐‑30°C  

and  a  more  extreme  mutator  activity  at  37°C  [85].  Additional  alleles  that  were  studied  

include  moderate  mutators  such  as  dnaQ928  [68].    

Loss  of  ε  causes  multiple  phenotypes:  slow  growth,  high  mutation  frequency,  

constitutive  SOS,  a  SOS  defect  following  nalidixic  acid,  and  increased  deletion  of  direct  

repeats  [49,  50,  86].  Identification  of  dnaQ  in  our  screen  for  SOS  constitutive  mutants  that  

are  unable  to  induce  at  a  higher  level  following  nalidixic  acid  treatment  (see  

Introduction)  led  us  to  believe  that  ε  may  have  a  role  in  the  bacterial  response  to  

nalidixic  acid  (Figure  8).  Furthermore,  we  propose  that  this  role  of  ε  is  separate  from  the  

proofreading  role  in  the  pol  III  holoenzyme.  However  ε  binds  tightly  with  α  and  θ,  and  

this  implicates  the  replisome  and  replication  fork  in  the  quinolone  response.    

In  this  chapter,  I  present  a  genetic  screen  used  to  isolate  dnaQ  separation  of  

function  mutants.  I  created  a  plasmid  library  of  dnaQ  mutants  with  15-­‐‑basepair  

insertions  located  randomly  throughout  the  gene.  I  screened  the  library  of  mutants,  

searching  for  separation  of  the  proofreading  function  (assessed  by  mutation  frequency)  

and  the  constitutive  SOS  phenotype.  The  following  sections  detail  the  methods  and  

results  of  this  study.  

 

Page 45: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

29  

2.1.1 Rationale for Linker Scanning Mutagenesis

Various  dnaQ  alleles  have  been  discovered  and  created,  generating  many  of  the  

extremely  high  mutators  such  as  mutD5,  dnaQ49  and  dnaQ926  [78,  81].  These  studies  

have  confirmed  the  importance  of  the  exo  motifs  in  the  N-­‐‑terminal  region  to  the  

proofreading  function  of  ε.  Biochemical  and  genetic  assays  have  given  general  insight  

into  the  necessity  of  the  C-­‐‑terminal  segment  for  ε  association  with  α  within  the  

polymerase  III  core.  With  these  features  in  mind,  we  sought  a  mutagenesis  method  that  

would  allow  us  to  make  small  changes  to  the  amino  acid  sequence  rather  than  

producing  a  completely  defective  protein.  

We  opted  to  use  the  New  England  Biolabs  Genome  Priming  System-­‐‑  Linker  

Scanning  mutagenesis  kit  to  create  our  mutant  library  (Figure  10).  This  method  has  

proven  effective  in  the  separation  of  function  for  multi-­‐‑functional  proteins  XerD  and  

McrA  [87,  88].  Linker  scanning  mutagenesis  generates  both  insertions  and  truncations,  

and  even  the  latter  could  potentially  cause  separation  of  function,  especially  within  the  

less  well-­‐‑defined  C-­‐‑terminal  segment  (Figure  6).  It  also  allows  us  to  avoid  large  

insertions  that  would  render  the  protein  catalytically  defective.  

The  system  is  based  on  an  in  vitro  reaction  that  inserts  a  transprimer  donor  with  a  

drug  resistance  marker  using  a  modified  transposase  enzyme  (TnsABC*)  into  the  target  

DNA  on  a  plasmid.  At  this  point  the  insertion  is  a  large  region  of  DNA  over  1.3  

kilobases  in  length.  However,  after  isolation  of  this  collection  of  plasmids,  subsequent  

Page 46: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

30  

 

 

Figure  10.  Linker  Scanning  Mutagenesis.  

The  transprimer  is  inserted  throughout  the  target  DNA  (Column  2)  via  the  modified  transposase  TnsABC*.  The  large  transprimer  is  digested  and  ligated  to  leave  behind  a  15  basepair  insertion  (Column  3).  Image  from  NEB  GPS-­‐‑LS  Manual.  

Page 47: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

31  

digestion  with  the  rare  cutter  PmeI  excises  the  majority  of  the  transprimer.  Ligation  to  

close  the  plasmid  leaves  behind  15  basepairs  of  new  DNA:  5  basepairs  of  DNA  

duplicated  from  the  target  site  (varies  depending  on  the  target)  and  10  basepairs  that  

include  a  PmeI  restriction  site  (TGTTTAAACA).  

2.2 Experimental Procedures

2.2.1 Bacterial Strains, Plasmids and Materials

The  E.  coli  strains  used  in  this  study  were  AB1157  [F-­‐‑  thr-­‐‑1,  ara-­‐‑14,  leuB6,  Δ(gpt-­‐‑

proA)62,  lacY1,  tsx-­‐‑33,  supE44,  galK2,  rac-­‐‑,  hisG4(Oc),  rfbD1,  mgl-­‐‑51,  rpsL31,  kdgK51,  xyl-­‐‑5,  

mtl-­‐‑1,  argE3  (Oc),  thi-­‐‑1,  qsr,-­‐‑],  JH39  [F-­‐‑  sfiA11,  thr-­‐‑1,  leu-­‐‑6,  hisG4,  argE3,  ilv(Ts),  galK2,  srl(?),  

rpsL31,  lacΔU169,  dinD1::MudI(Apr  lac)],  and  DH5α  [F-­‐‑  endA1  glnV44  thi-­‐‑1  recA1  relA1  

gyrA96  deoR  nupG  Φ80dlacZΔM15  Δ(lacZYA-­‐‑argF)U169,  hsdR17(rK-­‐‑  mK+),  λ–]  derivatives.  

P1  transductions  were  performed  using  P1vir.  Plasmid  transformation  of  pMM5Z  and  

derived  constructs  were  performed  using  BioRad  Gene  Pulser  according  to  the  protocol  

provided  by  the  manufacturer.  Antibiotics  from  Sigma  Aldrich  were  added  as  follows:  

chloramphenicol  34  µμg/ml,  ampicillin  50  µμg/ml,  kanamycin  50  µμg/ml,  rifampicin  100  

µμg/ml.  Bluo-­‐‑gal  (5-­‐‑Bromo-­‐‑3-­‐‑indolyl-­‐‑β-­‐‑D-­‐‑galactopyranoside)  was  purchased  from  Gold  

Biotechnology.  Solid  media  contained  1.5%  agar.  

The  background  strain  for  screening  the  insertion  mutant  library  is  a  derivative  

of  JH39.  This  strain  contains  a  dinD::lacZ  fusion  which  allows  convenient  color  reporting  

of  SOS  induction  when  plated  on  Bluo-­‐‑gal  [89].  Using  P1  transduction,  a  single-­‐‑gene  

Page 48: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

32  

deletion  of  dnaQ  from  the  Keio  collection  was  transduced  into  the  JH39  background  

strain  [90].  P1vir  lysate  was  prepared  by  adding  50  µμl  phage  to  5  ml  of  donor  strain.  

Recipient  strains  were  infected  and  transductants  were  isolated  on  LB  agar  +  kanamycin  

plates.  Single  transductants  were  purified  by  streaking  on  LB  agar  +  kanamycin  +  4  mM  

sodium  citrate.  PCR  products  were  generated  using  primers  dnaQseq1  (5’-­‐‑

ACCGCTCCGCGTTGTGTTCC-­‐‑3’)  and  dnaQseq2  (5’-­‐‑CGGTTGTTGGTGGTGCGGGT-­‐‑3’)  

and  subjected  to  sequencing  analysis  to  confirm  the  correct  transduction.  

2.2.2 Linker Scanning Mutagenesis

Mutagenesis  was  performed  with  the  New  England  Biolabs,  Inc.  GPS-­‐‑LS  linker  

scanning  system  according  to  provided  instructions.  The  transposon  reaction  of  plasmid  

pMM5Z  (see  Table  1)  was  electroporated  with  BioRad  Gene  Pulser  according  to  

manufacturer  instructions  into  DH5α  and  selected  with  chloramphenicol  and  ampicillin.  

Following  overnight  incubation  at  37°C,  1017  colonies  were  collected.  

2.2.2 Construction of Insertion Library

Plasmid  DNA  from  the  primary  mutant  library  (see  Table  1)  was  digested  with  

PstI  and  DraIII  and  run  on  a  0.8%  agarose  gel  (Figure  11A).  DnaQ  fragments  with  the  

large  transposon  DNA  (camR)  were  identified  via  gel  electrophoresis  and  extracted.  The  

extracted  dnaQ  +  transposon  fragments  were  ligated  to  fresh  pMM5Z  vector  that  was  

previously  cleaved  with  PstI  and  DraIII,  then  electroporated  into  DH5α  with  

chloramphenicol  and  ampicillin  selection  (Figure  11B).  A  pool  of  >5,000  CFU  was  

Page 49: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

33  

 

Table  1:  Plasmids  used  for  Library  Construction  

Plasmid   Relevant  Genotype  [derivation]   Source/Reference  pMM5   E.  coli  dnaQ+  and  rnhA+  inserted  into  EcoRI  

site  of  pBR322  [91]  

pMM5Z   pMM5;  AatII  restriction  site  replaced  with  BglII  using  primers  BglII#1  (5'ʹ-­‐‑TGAGATC  TCGCAACGT-­‐‑3'ʹ)  and  BglII#2  (5'ʹ-­‐‑TGCGAGA  TCTCAACGT-­‐‑3'ʹ)  

This  work  

Primary  Library   GPS-­‐LS  Mutagenesis  of  pMM5Z   This  work  Secondary  Library   Primary  library  subcloned  into  fresh  

pMM5Z  backbone,  removing  clones  with  insertions  outside  the  region  of  interest  (PstI-­‐DraIII)  

This  work  

Final  Library   Secondary  library  with  majority  of  transposon  excised,  leaving  behind  a  15-­‐  basepair  insertion  

This  work  

pACYC184   P15A  origin  plasmid  with  tetR  and  camR   [92]  pACYC-­‐based  Mutant  Library  

Fragments  of  final  library  containing  dnaQ  insertion  mutants  were  excised  via  NcoI-­‐HindIII  and  cloned  into  pACYC184  

This  work  

pMAK705   Rep(ts),  camR,  M13mp13  polylinker  insert   [93]    

Page 50: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

34  

 

Figure  11.  Schematic  of  dnaQ  Insertion  Mutant  Library  Construction  

 A.  Following  GPS-­‐‑LS  mutagenesis,  the  large  transprimer  region  (E)  is  randomly  distributed  throughout  the  entire  plasmid.  Each  plasmid  receives  only  one  insertion.  Here,  multiple  clones  are  shown.  B.  The  primary  mutant  library  from  (A)  is  digested  with  PstI  and  DraIII  and  run  on  a  gel  to  eliminate  fragments  that  do  not  have  an  insertion  within  the  desired  region.  Gel  extracted  PstI-­‐‑DraIII  fragment  is  ligated  to  fresh  PstI  and  DraIII-­‐‑cleaved  vector  that  has  not  been  mutagenized.  This  enriches  mutants  with  insertions  within  the  PstI-­‐‑DraIII  dnaQ  region.    C.  PmeI  digestion  of  the  secondary  library  removes  the  majority  of  the  transprimer,  resulting  in  the  construct  seen  in  (F).    D.  The  mutant  library  was  digested  with  BglII  and  HindIII  then  moved  onto  the  pMAK705  backbone  by  ligating  to  BclI  and  HindIII.    

 

Page 51: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

35  

prepped  to  isolate  the  secondary  library.  The  secondary  library  was  digested  with  PmeI  

for  1  hour  at  37°C.  Correct  products  were  identified  via  gel  electrophoresis,  extracted,  

ligated  and  transformed  into  DH5α.  Over  12,300  CFU  were  harvested  for  the  final  dnaQ  

insertion  mutant  library.  The  final  library  was  excised  with  BglII  and  HindIII  (Figure  

11C),  and  then  ligated  to  pACYC184  that  had  been  pre-­‐‑cleaved  with  BclI  and  HindIII.  

BglII  and  BclI  have  compatible  cohesive  ends  (Figure  11D).  The  library  was  moved  from  

the  pBR322  backbone,  which  is  higher  copy  number,  to  the  backbone  of  pACYC184,  a  

lower  copy  number  plasmid,  to  minimize  effects  from  overexpression.  

2.2.3 Plasmid-Based Screening

Plasmid  based  screening  was  conducted  in  JH39  ΔdnaQ  background  strain.  

Using  the  Keio  knockout  collection  as  a  donor,  ΔdnaQ  was  P1  transduced  into  JH39.  The  

transduction  was  confirmed  by  amplifying  with  primers  dnaQ5F  (5’-­‐‑  TGCCCC  

AAAACGAAGGCAGT  -­‐‑3’)  and  dnaQI3R2  (5’-­‐‑  TGCAGCCACAAAACGCAGTG  -­‐‑3’)  and  

analyzing  PCR  products  by  gel  electrophoresis.  Tranductants  were  confirmed  by  

sequencing  with  primers  dnaQseq1  (5’-­‐‑  ACCGCTCCGCGTTGTGTTCC  -­‐‑3’)  and  

dnaQseq2  (5’-­‐‑  CGGTTGTTGGTGGTGCGGGT-­‐‑3’).  

2.2.3.1  Qualitative  Measurement  of  Constitutive  SOS  

Once  the  pMAK-­‐‑based  library  was  electroporated  into  electrocompetent  JH39  

ΔdnaQ,  the  outgrowth  was  plated  onto  LB  +  chloramphenicol  +  Bluo-­‐‑Gal.  Colonies  were  

scored  based  on  intensity  of  blue  color  development.  

Page 52: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

36  

2.2.3.2  Preliminary  Mutator  Phenotype  Assessment  

To  determine  the  mutator  activity,  the  frequency  of  rifampicin  resistance  mutants  

were  determined  for  each  strain  by  plating  100  µμl  of  saturated  overnight  culture  onto  LB  

agar  +  rifampicin  (100  µμg/ml)  and  incubating  overnight  at  37°C.    

2.2.4 Gene Replacement

Following  the  method  of  Hamilton  et  al,  insertion  and  site-­‐‑specific  mutants  were  

incorporated  into  the  chromosome  of  AB1157  [93].  Mutants  were  excised  from  the  

pACYC184  backbone  with  NcoI  and  HindIII  and  ligated  into  temperature  sensitive  

replication  plasmid,  pMAK705,  using  the  same  restriction  sites.  Strains  carrying  

pMAK705  plasmid  hosting  dnaQ  mutants  were  grown  at  30°C  then  shifted  to  the  non-­‐‑

permissive  temperature  (44°C)  with  chloramphenicol.  This  selects  for  incorporation  of  

the  entire  plasmid  into  the  chromosome  as  the  temperature-­‐‑sensitive  origin  cannot  

replicate  at  44°C.  Shifting  back  to  30°C  selects  for  plasmid  resolution;  the  plasmid  can  

resolve  with  dnaQ+  or  with  mutant  dnaQ.  Following  growth  at  the  permissive  

temperature  (30°C),  colonies  were  screened  for  the  correct  gene  product  via  PCR  

amplification  using  chromosome-­‐‑specific  primers  dnaQ5F  (5’-­‐‑  TGCCCCA  

AAACGAAGGCAGT  -­‐‑3’)  and  dnaQI3R2  (5’-­‐‑  TGCAGCCACAAAACGCAGTG  -­‐‑3’).  PCR  

products  were  digested  with  PmeI  and  sequenced  using  primers  dnaQseq1  (5’-­‐‑ACCG  

CTCCGCGTTGTGTTCC-­‐‑3’)  or  dnaQseq2  (5’-­‐‑CGGTTGTTGGTGGTGCGGGT-­‐‑3’)  to  

confirm  gene  replacement.  

Page 53: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

37  

2.2.5 Mutation Rate

The  mutation  frequency  of  each  strain  was  determined  based  on  rifampicin  

resistant  CFU  and  total  CFU  of  at  least  8  independent  cultures  of  each  strain.  The  

mutation  rate  was  determined  with  the  Fluctuation  of  Analysis  Calculator  (FALCOR)  

using  the  Lea  and  Coulson  Method  of  the  Median  [94].  

2.2.6 Quantitative Western Blots

Overnight  cultures  of  strains  were  diluted  1:100  and  grown  until  OD600=0.5.  

Cultures  were  split  into  5  ml  aliquots  and  treated  with  nalidixic  acid  (100  µμg/ml)  or  no  

drug.  Samples  of  1  ml  were  collected  at  0  and  30  minutes  and  pelleted.  Pellets  were  

resuspended  in  SDS  Loading  Buffer.  Samples  were  loaded  onto  BioRad  4-­‐‑20%  TBE  

Precast  Protein  Gel  and  run  at  200  volts  for  35  minutes.  The  protein  gel  was  transferred  

to  nitrocellulose  membrane  using  the  LifeTechnologies  iBlot  Gel  Transfer  Device.  

Membranes  were  blocked  with  5%  non-­‐‑fat  milk  buffer  at  room  temperature  for  an  hour,  

then  supplemented  with  0.1%  Tween  and  primary  antibody  to  RecA  or  LexA  and  

incubated  overnight  at  4°C.  Membranes  were  washed  with  TBS  +  Tween  and  incubated  

with  secondary  antibody  in  5%  non-­‐‑fat  milk  buffer  +  SDS  at  room  temperature  for  1  

hour.  Membranes  were  rinsed  with  TBS  +  Tween  and  allowed  to  air  dry.  Reactive  

species  were  visualized  using  the  Odyssey  Li-­‐‑Cor  Imaging  System.  

Page 54: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

38  

2.3 Results

We  sought  to  isolate  separation  of  function  mutants  of  dnaQ  that  retained  

exonuclease  activity,  yet  remained  SOS  constitutive  in  the  absence  of  an  inducer.  The  

tests  for  these  phenotypes  were  the  most  convenient  to  use  as  we  screened  through  

hundreds  of  candidates.  

2.3.1 Screening for Separation of Function at the Plasmid Level

Preliminary  screening  of  the  insertion  mutant  library  was  done  in  JH39  ΔdnaQ.  The  

pMAK-­‐‑based  mutant  library  was  transformed  into  this  background  and  screened  for  

dark  blue  color  development  on  plate  with  Bluo-­‐‑gal.  Approximately  30-­‐‑35%  of  the  

clones  were  dark  blue,  indicating  that  we  were  isolating  constitutive  SOS  mutants.  

Colonies  with  strong  color  intensity  or  “dark  blue  mutants”  were  selected,  and  the  

constitutive  SOS  phenotype  confirmed  by  restreaking  on  Bluo-­‐‑gal  plates  (Figure  12).  

SOS  constitutive  mutants  were  then  selected  for  overnight  growth,  and  100  µμl  of  

saturated  culture  was  plated  on  rifampicin  agar  (100  µμg/mL)  to  assess  mutator  

phenotypes.  While  this  provided  only  a  crude  estimate  of  mutation  frequency,  it  was  

sufficient  to  isolate  mutants  with  proofreading  activity  similar  to  wildtype.  Wildtype  

strains  usually  returned  fewer  than  10  rifampicin  resistant  colonies  per  100  µμl  of  

saturated  culture;  whereas  ΔdnaQ  has  anywhere  from  350  –  1,500  rifampicin  resistant  

colonies  per  100  µμl  of  overnight  culture.

Page 55: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

39  

 

 

Figure  12.  Qualitative  measurement  of  SOS  Induction  

SOS  induction  of  JH39  strains  on  LB  agar  with  Bluo-­‐‑Gal,  a  chromogenic  substrate  that  allows  visual  analysis  of  SOS  induction.  A  constitutive  strain  is  streaked  on  the  right  side  of  the  plate,  displaying  the  dark  blue  phenotype.  Figure  from  Morgan  Henderson,  Kreuzer  Lab,  Duke  University.  

Light  Blue SOS  pathway  off  

Dark  Blue   SOS  pathway  on

   

Page 56: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

40  

Our  initial  tests  were  very  stringent  and  focused  on  the  darkest  blue  mutants,  

avoiding  those  with  intermediate  color  intensity.  Additionally,  we  used  a  threshold  for  

RifR  CFU/100  µμl  of  overnight  culture  at  less  than  10  CFU.  This  eliminated  any  moderate  

mutators  that  were  above  this  range,  yet  below  the  extreme  ΔdnaQ  range.  We  identified  

multiple  candidate  insertion  mutants  with  normal  proofreading  using  these  preliminary  

parameters  and  a  striking  constitutive  level  of  SOS  as  determined  by  visual  analysis.    

Several  insertion  mutants  that  were  both  SOS  constitutive  and  exhibited  low  

mutator  activity  were  selected  for  further  study.  First,  we  confirmed  the  phenotypes  

after  moving  the  plasmid  into  a  fresh  background.  The  plasmids  were  isolated  and  

transformed  into  fresh  JH39  ΔdnaQ  electrocompetent  cells.  Following  streaks  on  Bluo-­‐‑

Gal  agar  and  another  round  of  plating  on  rifampicin  agar  as  described  above,  dark  blue  

color  and  fewer  than  10  CFU  confirmed  the  phenotypes  of  these  insertion  mutants.    

Mutants  148,  149,  153,  200  and  201  were  SOS  constitutive  and  retained  

proofreading  at  levels  similar  to  wildtype.  We  also  found  that  Mutant  102  no  longer  

exhibited  the  low  mutator  phenotype  and  it  was  excluded  from  further  testing.  Plasmid  

preps  of  mutants  were  sequenced  to  determine  the  composition  and  location  of  the  15  

base  pair  insertions.  Sequencing  revealed  two  duplicates  among  our  mutants:  Mutant  

149  was  the  same  as  Mutant  148,  and  Mutant  153  was  the  same  as  Mutant  200.  We  

excluded  Mutants  149  and  153  from  further  testing.  Of  our  remaining  candidates,  one  of  

the  insertions  is  within  dnaQ  (Mutant  201),  the  other  two  insertion  mutants  were  

Page 57: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

41  

downstream  (Mutant  148  and  200)  leaving  a  wildtype  copy  of  the  gene  (Table  2,  Figure  

13).  We  were  surprised  to  isolate  insertion  mutants  in  the  region  downstream  of  dnaQ.  

These  mutants  are  actually  dnaQ+,  yet  they  showed  separation  of  function.  While  this  

was  an  unexpected  result,  we  thought  it  would  be  interesting  to  pursue,  and  these  

mutants  will  be  discussed  in  detail  below.  

Page 58: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

42  

 

Table  2.  Dark  Blue  Plasmid  Mutants  with  Low  Mutator  Phenotypes  

These  mutants  were  dark  blue  at  initial  selection  and  appeared  to  retain  proofreading  function  following  preliminary  tests.  Mutants  148  and  149  are  clones,  as  are  mutants  153  and  200.          

Mutant  #   RifR  CFU/100  μl   Insertion  

102   15   232  LFKQV  148   2   Downstream  80bp  149   4   Downstream  80bp  153   1   Downstream  81bp  200   0   Downstream  81bp  201   0   124  LFKHV  

Page 59: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

43  

 

 

Figure  13.  Distribution  of  Chromosomal  Mutants  

 Schematic  of  insertion  mutant  locations.  Mutant  148  and  200  are  near  the  aspV  tRNA

Page 60: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

44  

2.3.2 Separation of Function after Chromosomal Gene Replacement

Although  we  discovered  interesting  mutants  using  the  plasmid  library  system,  to  

be  confident  that  these  phenotypes  are  directly  from  the  mutations  introduced  and  not  

influenced  by  copy  number  or  dosage  effects,  we  moved  the  mutations  into  the  

chromosome.  While  the  JH39  strain  has  the  advantage  of  visual  SOS  reporting,  it  has  the  

dinD::lacZ  fusion,  which  abolished  dinD  expression.  A  recent  study  showed  that  DinD  

protein  can  interfere  with  RecA  filament  loading  onto  double  strand  DNA,  and  

disruption  of  DinD  with  the  lacZ  fusion  may  therefore  alter  RecA  function  [95].  To  

evaluate  the  mutations  in  a  wildtype  background,  we  used  the  gene  replacement  

method  employing  a  repTS  plasmid,  pMAK705,  to  introduce  the  mutation  into  the  

chromosome  of  AB1157  [93].    

Gene  replacement  relies  on  homologous  recombination  to  fully  incorporate  the  

entire  plasmid  into  the  chromosome  under  selection  at  the  non-­‐‑permissive  growth  

temperature  (44°C).  Shifting  to  the  permissive  growth  temperature  (30°C)  promotes  

plasmid  resolution,  and  depending  on  the  location  of  recombination,  the  dnaQ  allele  left  

behind  on  the  chromosome  can  be  either  wildtype  or  the  mutant  copy  of  the  gene.  Gene  

replacement  candidates  were  identified  by  gel  electrophoresis  following  PmeI  digest  of  

dnaQ  region  that  was  amplified  with  primers  dnaQ5F  and  dnaQ132.  Mutants  were  

confirmed  in  the  AB1157  background  via  sequencing  these  PCR  products  with  primers  

dnaQseq1  and  dnaQseq2.  

Page 61: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

45  

To  confirm  proofreading  function  following  chromosomal  gene  replacement,  we  

used  the  widely  accepted  rifampicin  resistance  assay  to  measure  mutation  rate  [94].  This  

method  is  much  more  strenuous  than  our  preliminary  testing.  Initial  and  secondary  

screening  of  plasmid  mutants  tested  only  1-­‐‑3  cultures  and  did  not  take  into  account  the  

total  CFU  per  culture.  Mutation  rate  calculations  for  the  chromosomal  dnaQ  mutant  

strains  included  at  least  8  independent  cultures  of  each  strain  and  measured  total  CFU  to  

determine  the  number  of  mutations  per  generation  of  growth.  While  there  is  some  

variation  in  the  mutation  rates,  all  the  mutants  have  rates  that  are  closer  to  wildtype  

levels  than  to  that  of  ΔdnaQ  (Table  3).  Interestingly,  mutant  201  exhibited  a  mutation  

rate  that  was  10  fold  higher  than  wildtype  levels.  Based  on  earlier  classifications  of  dnaQ  

mutator  activity,  this  insertion  is  considered  a  moderate  mutator.  

Because  AB1157  no  longer  allowed  colorimetric  monitoring  of  SOS,  we  used  

quantitative  Western  blots  to  measure  RecA  levels  within  the  cells.  We  first  wanted  to  

ensure  that  ΔdnaQ  exhibited  the  partially  SOS  constitutive  phenotype  in  AB1157  and  

that  it  was  not  an  artifact  of  the  JH39  screening  strain  which  was  used  in  earlier  

screening  steps.  We  observed  that  AB1157  ΔdnaQ  exhibited  SOS  levels  that  were  

statistically  significantly  higher  than  wildtype  levels  (Figures  14  and  15).  

To  determine  if  these  mutants  were  SOS  constitutive  or  possessed  the  SOS  defect  

following  nalidixic  acid  induction,  we  treated  cells  with  drug  and  allowed  them  to  grow  

shaking  at  37°C  for  an  additional  30  minutes.  We  then  analyzed  the  levels  of  RecA  

Page 62: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

46  

 

Table  3.  Mutation  Rate  of  Chromosomal  Insertion  Mutants  

Following  chromosomal  gene  replacement  in  strain  AB1157,  mutants  were  tested  to  determine  the  mutation  rate  according  to  FALCOR  analysis  using  the  Lea  and  Coulson  Method  [94].  

 

Strain  #  Cultures  

(n)  

Median  Mutation  Rate  (per  107)  

Upper  Difference  (95%  CI)  

Lower  Difference  (95%  CI)  

WT   16   0.015   0.0085   0.0019  ΔdnaQ   15   1.565   0.8305   0.7077  #148   12   0.089   0.0211   0.0245  #200   12   0.017   0.0263   0.0056  #201   8   0.102   0.0882   0.0242  

 

 

Page 63: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

47  

protein,  which  is  induced  by  the  SOS  response.  Lysates  were  normalized  according  to  

optical  density  and  loaded  onto  a  protein  gel  with  dilutions  of  purified  RecA  protein.  

The  gels  were  transferred  to  nitrocellulose  membrane  and  probed  with  appropriate  

primary  anti-­‐‑RecA  and  secondary  antibodies.  In  addition,  purified  RecA  was  loaded  

within  a  range  of  2-­‐‑20  ng  on  each  gel.  If  a  sample  did  not  fall  within  this  range,  the  

sample  concentration  was  adjusted  to  fall  within  this  range  of  measurement  in  a  

subsequent  gel.  Including  the  RecA  standard  curve  on  each  gel  allowed  comparisons  of  

samples  between  gels  (Figure  14).  

We  used  quantitative  Western  blots  to  measure  RecA  levels  of  the  chromosomal  

mutants  grown  to  log  phase  without  any  inducer.  We  first  validated  this  assay  by  

measuring  basal  levels  of  SOS  of  wildtype  and  a  dnaQ  knockout.  The  dnaQ  mutant  basal  

SOS  level  was  statistically  significantly  higher  than  wildtype  (Figure  15).  Surprisingly,  

the  insertion  mutants  were  not  SOS  constitutive  by  the  RecA  measure  even  though  they  

appeared  constitutive  with  the  colorimetric  test  in  the  JH39  background.  Values  were  

determined  based  on  an  average  of  at  least  three  replicates  and  none  of  the  mutants  

were  statistically  significantly  higher  than  wildtype  (Figure  16).    

Although  we  did  not  test  the  SOS  defect  following  nalidixic  acid  treatment  at  the  

plasmid  level,  we  measured  the  RecA  levels  following  growth  for  30  minutes  in  the  

presence  of  nalidixic  acid  at  100  µμg/mL.  The  insertion  mutant  RecA  levels  following  

nalidixic  acid  treatment  were  as  high  as,  and  in  some  cases  higher  than,  wildtype  levels  

Page 64: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

48  

(Figure  17).  We  determined  that  all  the  insertion  mutants  had  SOS  profiles  that  were  

similar  to  wildtype.  Therefore,  they  were  not  SOS  constitutive,  contrary  to  previous  

observations  in  the  plasmid-­‐‑based  assays.  

 

Page 65: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

49  

 

 

Figure  14.  Quantitative  Western  Blots  of  RecA  in  Chromosomal  Mutants  

A.  Plots  of  measured  intensity  vs  known  concentration  of  purified  RecA  protein  were  used  to  generate  a  standard  curve.  B.  Western  blot  with  samples  before  and  after  nalidixic  acid  treatment  (100  µμg/ml).  Following  the  sample  name,  0’  denotes  no  drug  treatment  and  30’  denotes  nalidixic  acid  treatment  for  30  minutes.  Each  blot  includes  a  RecA  standard  curve  (2-­‐‑20ng).  

 

R²  =  0.9959  

0  

10  

20  

0   5   10   15   20  

Intensity  

RecA  (ng)  

A  

B  

Page 66: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

50  

 

Figure  15.  Quantitative  Western  Blot  Confirms  ΔDnaQ  Partially  Constitutive  SOS  Phenotype  

Bars  are  RecA  averages  from  10  independent  cultures/lysates  based  on  western  blots  in  Figure  14.  Bars  indicate  standard  error  and  asterisk  indicates  a  p  value  <  0.05.  

0  0.5  1  

1.5  2  

2.5  3  

3.5  

WT   ΔDnaQ  

RecA  (ng)  

*

Page 67: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

51  

     

Figure  16.  Chromosomal  Insertion  Mutants  are  not  SOS  Constitutive  

Bar  graphs  indicate  the  mean  value  of  the  number  of  cultures  indicated  below  each  strain.  Error  bars  indicate  the  standard  error  of  all  values.  *  =  p<0.05  compared  to  wildtype.  Only  ΔDnaQ  is  statistically  significantly  different  from  wildtype  JH39.  Although  we  saw  the  partially  constitutive  SOS  phenotype  at  the  plasmid  level,  the  chromosomal  mutants  are  not  SOS  constitutive.

0  

0.5  

1  

1.5  

2  

2.5  

3  

3.5  

148   200   201   WT   ΔDnaQ  

RecA(ng)  

*  

    n=   3            7          9                        10                        10  

Page 68: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

52  

 

 

Figure  17.  Chromosomal  Mutants  are  not  SOS  defective  following  nalidixic  acid  treatment.  

Chromosomal  insertion  mutants  (numbered  148,  200  and  201)  induce  SOS  following  nalidixic  acid  at  levels  equal  to  or  greater  than  wildtype  levels.  Samples  were  treated  with  10  µμg/ml  nalidixic  acid  at  37°C  for  30  minutes  before  lysate  collection.  

 

0  

5  

10  

15  

20  

25  

30  

35  

148   200   201   WT   ΔDnaQ  

RecA  (ng)  

No  Drug  Nalidixic  Acid  

Page 69: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

53  

2.4 Discussion

The  goal  of  the  work  in  this  chapter  was  to  identify  separation  of  function  

mutants  using  a  plasmid  system  of  insertion  mutants.  Our  initial  efforts  focused  on  

separating  the  constitutive  SOS  phenotype  from  the  high  mutator  activity  due  to  loss  of  

proofreading.  Initial  screening  of  mutant  plasmids  electroporated  into  JH39  ΔdnaQ  

returned  three  interesting  candidates  that  were  extremely  dark  blue  (SOS  constitutive)  

and  appeared  to  retain  exonuclease  activity  with  low  levels  of  rifampicin  resistance.  

There  were  multiple  considerations  in  the  cloning  set-­‐‑up  and  screening  of  these  

mutants.  We  realized  that  there  could  have  been  a  dosage  effect  by  having  the  mutations  

on  a  plasmid  that  perhaps  contributed  to  the  constitutive  SOS  phenotype  we  observed  

by  qualitative  measurement.  In  our  cloning  strategy,  we  tried  to  avoid  this  by  moving  

the  insertion  mutant  library  from  the  pMM5Z  backbone,  which  is  pBR322  based,  to  the  

lower  copy  number  origin  p15A  on  the  backbone  of  pACYC184.  On  average,  the  pBR322  

origin  has  15-­‐‑20  copies  per  cell,  while  the  P15A  origin  has  10-­‐‑12  copies  per  cell  [96].    

As  a  final  safeguard,  we  moved  the  insertion  mutants  from  the  plasmid,  into  the  

chromosome  and  under  the  natural  promoter.  Following  chromosomal  gene  

replacement,  the  mutants  were  tested  with  more  quantitative  measures.  This  consisted  

of  mutation  rate  determination  using  fluctuation  analysis  and  monitoring  of  SOS  levels  

of  induced  and  uninduced  cells  via  quantitative  Western  blots  of  RecA  levels.  

Page 70: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

54  

In  our  plasmid-­‐‑based  assays,  we  initially  isolated  six  separation  of  function  

mutants.  Additional  screening  narrowed  our  pool  down  to  three  candidates,  Mutants  

148,  153,  200.  These  mutants  displayed  the  constitutive  SOS  phenotype  while  retaining  

exonuclease  activity.  The  mutants  were  then  cloned  into  the  chromosome  of  AB1157.  

Chromosomal  mutants  were  tested  for  exonuclease  activity  and  SOS  induction  with  and  

without  nalidixic  acid  treatment.  Overall,  the  mutation  rates  of  the  chromosomal  

mutants  are  similar  to  the  preliminary  measures  of  rifampicin  resistance  with  plasmid-­‐‑

based  mutants.  However,  none  of  the  chromosomal  mutants  were  SOS  constitutive,  

which  contradicted  the  plasmid-­‐‑based  qualitative  tests.  We  also  did  not  detect  an  SOS  

defect  following  nalidixic  acid  treatment.  Once  cloned  into  the  AB1157  chromosome,  all  

of  the  mutants  phenotypically  presented  similar  wildtype.  

There  are  clear  differences  between  the  strains  JH39  and  AB1157.  The  dinD::lacZ  

fusion  abolishes  DinD  expression  in  JH39,  and  this  could  have  affected  the  constitutive  

SOS  phenotype.  We  P1  transduced  a  knockout  of  dnaQ  from  the  Keio  collection  into  

JH39.  The  Keio  knockouts  are  single,  in-­‐‑frame  deletions  that  replace  the  gene  with  a  

kanamycin  resistance  cassette  [90].  However,  dnaQ  and  rnhA  are  adjacent  genes  encoded  

from  divergent,  overlapping  promoters  [91,  97].  The  complete  deletion  of  dnaQ  does  not  

remove  the  rnhA  promoter  sequence,  but  if  it  affects  gene  expression  at  all,  this  could  

confound  results.  RNase  HI  is  encoded  by  rnhA  and  is  used  to  remove  RNA/DNA  

hybrids,  particularly  the  RNA  primers  on  DNA  synthesized  from  the  lagging  strand  

Page 71: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

55  

template  [98].  Mutations  that  create  catalytically  defective  RNase  HI  cause  the  

accumulation  of  R-­‐‑loops,  which  can  lead  to  genomic  instability  and  the  activation  of  the  

SOS  response  [99].  Selecting  for  the  extremely  dark  blue  clones  could  have  also  selected  

for  SOS  that  is  activated  in  response  to  disruption  of  rnhA.  

Another  factor  that  could  have  contributed  to  the  differences  seen  between  the  

plasmid  based  screen  and  the  quantitative  assessment  of  chromosomal  mutants  is  the  

actual  measure  of  SOS.  The  reporter  construct  in  JH39  measures  SOS  induction  as  it  

relates  to  dinD  promoter  activity,  while  the  quantitative  Western  blots  measure  RecA  

protein  levels  in  the  cells  to  determine  SOS  levels.  Previous  work  in  our  lab  showed  that  

loss  of  dnaQ  leads  to  differential  expression  of  SOS  genes  [50].  Figure  18  shows  the  recA  

and  dinD  mRNA  transcript  levels  before  and  after  nalidixic  acid  treatment  in  both  

wildtype  and  ΔdnaQ  JH39  strains.  In  this  study,  nalidixic  acid  treatment  of  JH39  ΔdnaQ  

leads  to  an  increase  in  recA  levels  while  dinD  levels  remain  the  same  [50].  Preliminary  

RecA  Western  blots  of  AB1157  ΔdnaQ  showed  a  clear  defect  in  SOS  induction  following  

nalidixic  acid  treatment  (Figures  15  &  17).  This  led  us  to  believe  that  RecA  levels  were  a  

good  indicator  of  SOS  levels  in  our  chromosomal  mutants.    

The  downstream  insertions  of  Mutant  148  and  200  were  in  the  promoter  region  

of  the  gene  for  tRNAaspV.  It  has  been  shown  that  many  tRNAs  are  homologous  to  the  5’  

end  of  the  dnaQ  transcript,  specifically  the  region  between  promoter  1  and  2  of  dnaQ  

(Figure  13).  This  provides  a  model  for  tRNAs  as  silencing  mRNA  [100].  There  are  three    

Page 72: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

56  

.    

 

 

Figure  18.  Differential  expression  of  LexA-­‐‑regulated  genes  

JH39  wildtype  and  dnaQ  mutant  strains  were  treated  with  or  without  nalidixic  acid  (+/-­‐‑)  as  indicated  above.  The  graphs  of  mRNA  transcript  levels  show  a  distinctly  different  pattern  of  levels  between  dinD  and  recA.  Modified  from  [50].  

 

Page 73: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

57  

aspartate  tRNA  genes:  aspT,  aspU  and  aspV.  The  aspU  tRNA  shows  homology  to  the  5’  

end  of  dnaQ  mRNA,  but  this  has  not  been  tested  with  aspV  which  is  immediately  

downstream  of  dnaQ.  If  aspV  does  bind  to  dnaQ  transcript,  the  expression  of  tRNAaspV  

from  the  plasmid  could  create  a  tRNA/dnaQ  mRNA  duplex  preventing  ribosome  

binding  and  decreasing  wildtype  epsilon  levels  [100,  101].  This  could  affect  the  

phenotypes  tested.  

After  observing  contradictory  results  with  our  insertion  mutants  at  the  plasmid  

level  vs.  chromosome  level  and  in  different  strains,  we  decided  to  restructure  our  

approach  to  the  separation  of  function.  Chapter  3  details  a  less-­‐‑directed  approach  to  

determine  if  we  can  isolate  separation  of  function  ε  mutants  in  a  strain  that  is  essentially  

wildtype  and  with  the  dnaQ  mutations  at  the  natural  location  in  the  chromosome.  

 

Page 74: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

58  

3. Chromosome Based Separation of Function Study

3.1 Introduction

In  Chapter  2,  we  screened  a  library  of  dnaQ  mutants,  searching  for  separation  of  

function.  Our  previous  approach  of  screening  for  separation  of  function,  isolating  

plasmid-­‐‑based  dnaQ  insertions,  then  moving  into  the  chromosome  for  additional  

screening  proved  ineffective.  The  phenotypes  seen  with  the  plasmid  level  were  not  seen  

at  the  chromosomal  level.  Earlier  iterations  of  this  study  rested  on  the  identification  of  a  

constitutive  SOS  mutant  that  retained  proofreading  activity.  However,  if  these  two  

phenotypes  are  linked,  we  would  never  isolate  that  type  of  separation  mutant.  While  we  

could  have  pursued  other  types  of  separation  of  function,  we  opted  to  restructure  our  

approach  after  separation  identified  at  the  plasmid  level  was  not  reproducible  when  the  

mutations  replaced  the  wildtype  chromosomal  allele  of  dnaQ.    

To  determine  if  ε  phenotypes  can  be  separated,  we  constructed  library  of  

plasmids  with  15  base  pair  insertions  randomly  inserted  throughout  the  dnaQ  gene.  We  

used  TransposaseABC*  to  insert  a  1.3  kilobase  transprimer  randomly  throughout  dnaQ.  

The  majority  of  the  transprimer,  including  a  chloramphenicol  marker,  was  removed  by  

PmeI  digest.  Subsequent  ligation  left  behind  a  15bp  scar:  10bp  of  transprimer  including  

PmeI  plus  an  additional  5bp  duplication  of  the  target  DNA.  

In  our  new  approach,  the  plasmid  library  of  dnaQ  insertion  mutants  is  moved  to  

the  backbone  of  pMAK705,  a  repTS  plasmid  that  facilitates  gene  replacement  via  

Page 75: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

59  

homologous  recombination  using  selective  temperature-­‐‑shifted  growth.  This  library  was  

transformed  into  wildtype  JH39,  which  has  a  convenient  dinD::lacZ  SOS  reporter  

construct.  Following  confirmation  that  the  majority  of  the  colonies  have  an  insertion,  we  

sequenced  over  100  plasmid  insertions  and  selected  approximately  25  interesting  

candidates  for  chromosomal  integration  using  the  method  described  above.  Those  

candidates,  once  in  the  JH39  chromosome,  were  tested  for  SOS  phenotypes,  both  

constitutive  and  defective  induction  after  nalidixic  acid  treatment,  and  mutation  rate.  

The  SOS  phenotypes  were  determined  using  a  kinetic  β-­‐‑galactosidase  assay  and  the  

mutation  rate  was  based  on  rifampicin  resistant  colony  forming  units.    

3.2 Experimental Procedures

To  determine  if  the  multiple  phenotypes  of  a  dnaQ  mutant  are  separable,  we  

used  the  GPS-­‐‑LS  Mutagenesis  kit  (NEB)  to  create  a  plasmid  library  of  dnaQ  insertion  

mutants.  Unlike  traditional  transposon  mutagenesis,  this  method  inserts  small,  15  

basepair  segments  randomly  throughout  the  target  sequence  (Figures  10  &  19).  Initially,  

a  1376-­‐‑basepair  transposon  is  introduced,  then  digested  with  PmeI  ,  which  excises  the  

majority  of  the  transposon,  leaving  behind  5-­‐‑amino  acid  insertions  and  in  a  subset  of  

cases,  a  stop  codon.  

Page 76: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

60  

 

Figure  19.  Schematic  of  Insertion  Library  Construction  

A) Following  GPS-­‐‑LS  mutagenesis,  the  large  transprimer  region  (E)  is  randomly  distributed  throughout  the  entire  plasmid.  Each  plasmid  receives  only  one  insertion.  Here,  multiple  clones  are  shown.  

B)  B)  The  primary  mutant  library  from  (A)  is  digested  with  PstI  and  DraIII  and  run  on  a  gel  to  eliminate  fragments  that  do  not  have  an  insertion  within  the  desired  region.  Gel  extracted  PstI-­‐‑DraIII  fragment  is  ligated  to  fresh  PstI  and  DraIII  cleaved  vector  that  has  not  been  mutagenized.  This  enriches  mutants  with  insertions  within  the  PstI-­‐‑DraIII  dnaQ  region.    

C) PmeI  digestion  of  the  secondary  library  removes  the  majority  of  the  transprimer,  resulting  in  the  construct  seen  in  (F).    

D) The  mutant  library  was  digested  with  ClaI  and  PstI  then  moved  onto  the  pMAK705  backbone  by  ligating  to  the  same  sites.  These  products  were  sequenced  and  used  to  select  mutants  that  would  be  evaluated  at  the  chromosomal  level.  

Page 77: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

61  

3.2.1 Bacterial Strains and Plasmids

The  E.  coli  strain  JH39  [F-­‐‑  sfiA11,  thr-­‐‑1,  leu-­‐‑6,  hisG4,  argE3,  ilv(Ts),  galK2,  srl(?),  

rpsL31,  lacΔU169,  dinD1::MudI(Apr  lac)]  was  the  background  strain  used  to  study  

mutants  at  the  chromosomal  level.  All  plasmid  constructions  were  done  in  the  DH5α  [F-­‐‑  

endA1  glnV44  thi-­‐‑1  recA1  relA1  gyrA96  deoR  nupG  Φ80dlacZΔM15  Δ(lacZYA-­‐‑argF)U169,  

hsdR17(rK-­‐‑  mK+),  λ–]  wildtype  strain.    

3.2.2 Linker Scanning Mutagenesis

Mutagenesis  was  performed  with  the  New  England  Biolabs,  Inc.  GPS-­‐‑LS  system  

according  to  provided  instructions.  The  transposon  reaction  of  plasmid  pMM5Z  (see  

Table  1)  was  electroporated  into  DH5α  with  BioRad  Gene  Pulser  according  to  

manufacturer  instructions  selecting  for  insertion  mutants  with  chloramphenicol  and  

ampicillin.  Following  overnight  incubation  at  37°C,  1017  colony  were  collected  and  

pooled  into  the  primary  mutant  library.  

3.2.3 Generating dnaQ mutant library

Plasmid  DNA  from  the  initial  transposon  reaction  (see  Table  4)  was  digested  

with  PstI  and  DraIII.  Fragments  with  the  large  transposon  insertion  were  identified  via  

gel  electrophoresis  and  extracted.  The  extracted  fragments  were  ligated  to  fresh  pMM5Z  

vector  that  was  not  subject  to  mutagenesis  electroporated  into  DH5α  with  

chloramphenicol  and  ampicillin  selection.  The  colonies  were  pooled  and  prepped  to  

isolate  the  secondary  plasmid  library.  The  secondary  library  was  digested  with  PmeI  to  

Page 78: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

62  

Table  4.  Plasmid  Constructs  

Plasmid   Relevant  Genotype  [derivation]   Source/Reference  pMM5   E.  coli  dnaQ+  and  rnhA+  inserted  into  EcoRI  

site  of  pBR322  [91]  

pMM5Z   pMM5;  AatII  restriction  site  replaced  with  BglII  using  primers  BglII#1    (5'ʹ-­‐‑TGAGATCTCGCAA  CGT-­‐‑3'ʹ)  &    BglII#2  (5'ʹ-­‐‑TGCGAGATCTCAACGT-­‐‑3'ʹ)  

This  work  

Primary  Library   GPS-­‐‑LS  Mutagenesis  of  pMM5Z   This  work  Secondary  Library   Primary  library  subcloned  into  fresh  pMM5Z  

backbone,  removing  clones  with  insertions  outside  the  region  of  interest  (PstI-­‐‑DraIII)  

This  work  

Final  Library   Secondary  library  with  majority  of  transposon  excised,  leaving  behind  a  15-­‐‑basepair  insertion  

This  work  

pMAK705   Rep(ts),  camR,  M13mp13  polylinker  insert   [93]  

Page 79: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

63  

remove  majority  of  the  transposon.  Desired  products  were  identified  via  gel  

electrophoresis,  extracted,  ligated  and  transformed  into  DH5α.  The  plasmids  were  

harvested  for  the  final  dnaQ  insertion  mutant  library  and  subcloned  into  the  polylinker  

of  pMAK705  via  ClaI  and  PstI  (Figures  19  &  20).  

3.2.4 Site-Directed Mutagenesis

The  Stratagene  QuikChange  Mutagenesis  kit  was  used  according  to  

manufacturer’s  instructions  to  generate  moderate  mutator  dnaQ928,  which  is  Gly17Ser  

[68].  We  created  strong  (QLSLPL)  and  weak  (ATSMAF)  β-­‐‑clamp  binding  mutants  using  

primers  QLSLPL  (5'ʹ-­‐‑CTGGCGATGACCGGTGGTCAACTATCGTTGCCTTTAGCGAT  

GGAAGGAGAGAC-­‐‑3'ʹ),  QLSLPL_anti  (5'ʹ-­‐‑GTCTCTCCTTCCATCGCTAAAGGCAACGA  

TAGTTGACCACCGGTCATCGCCAG-­‐‑3'ʹ),  Q182A  (5'ʹ-­‐‑GGCGATGACCGGTGGTGCAA  

CGTCGATGGCTTTT-­‐‑3'ʹ)  and  Q182A_anti  (5'ʹ-­‐‑AAAAGCCATCGACGTTGCACCAC  

CGGTCATCGCC-­‐‑3'ʹ).

3.2.5 Gene replacement

Following  the  method  of  Hamilton  et  al,  insertion  and  site-­‐‑specific  mutants  were  

incorporated  into  the  chromosome  of  JH39.  This  method  uses  a  temperature  sensitive  

replication  plasmid  pMAK705  to  facilitate  incorporation  into  the  chromosome  via  

selection  at  the  non-­‐‑permissible  temperature  (44°C)  with  chloramphenicol.  Following  

growth  at  the  permissive  temperature  (30°C),  colonies  were  screened  for  the  correct  gene  

product  via  PCR  amplification  using  chromosome-­‐‑specific  primers  dnaQ5F  (5’-­‐‑  

Page 80: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

64  

TGCCCCAAAACGAAGGCAGT  -­‐‑3’)  and  dnaQI3R2  (5’-­‐‑  TGCAGCCACAAAACGC  

AGTG  -­‐‑3’).  PCR  products  were  digested  with  PmeI  and  sequenced  using  primers  

dnaQseq1  (5’-­‐‑ACCGCTCCGCGTTGTGTTCC-­‐‑3’)  or  dnaQseq2  (5’-­‐‑CGGTTGTTGGTGG  

TGCGGGT-­‐‑3’)  to  confirm  gene  replacement.  

 

 

Page 81: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

65  

 

 

Figure  20:  Mutant  Library  and  pMAK705  Cloning.  

 Lane  1:  Marker,  Lanes  2&3:  Mutant  Library  Digested  with  ClaI  and  PstI,  Lane  4:  Empty,  Lane  5:  pMAK705  digested  with  ClaI  and  PstI.  The  panel  on  the  right  shows  gel  extraction  of  the  desired  bands  (white  arrows).  

 1                  2                3                  4                5    1                  2                3                  4                5  

Page 82: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

66  

3.2.6 Quantitative Beta-galactosidase Assay

The  convenient  dinD::lacZ  reporter  allows  monitoring  of  SOS  in  the  JH39  strain.  

SOS  levels  were  determined  using  a  beta-­‐‑galactosidase  protocol  modified  for  96-­‐‑well  

microtiter  plates  [102].  Overnight  cultures  were  diluted  1:100  in  LB  and  grown  for  2  

hours.  They  were  treated  with  10  µμg/mL  of  nalidixic  acid  or  no  drug  and  grown  2  

additional  hours.  Beta-­‐‑galactosidase  activity  was  measured  as  outlined  previously.  OD630  

was  measured  and  100  µμl  of  each  culture  was  lysed  with  Novagen  PopCulture  reagent  

at  room  temperature.  A  15  µμl  aliquot  of  lysed  cells  was  incubated  with  135  µμl  Z-­‐‑Buffer  

(60  mM  Na2HPO4,  40  mM  NaH2PO4,  10  mM  KCl,  1  mM  MgSO4),  0.5  µμl  TCEP  (tris(2-­‐‑

carboxyethyl)phosphine)  and  30  µμl  of  4  mg/mL  2-­‐‑Nitrophenyl  β-­‐‑D-­‐‑galactopyranoside,  

while  measuring  the  accumulation  2-­‐‑nitrophenol  at  A405  for  60  minutes.  Miller  units  

were  calculated  using  the  modified  equation  (V*1000*CF1*CF2)/(A595*CF3*relative  

volume  of  cell  lysate  used).  CF1  adjusts  for  the  difference  between  A420  in  the  plate  

reader  compared  to  A415  in  the  formula.  CF2  adjusts  for  the  omission  of  Na2CO,  which  

was  included  in  the  original  Miller’s  equation.  CF3  converts  the  optical  density  

measured  in  the  plate  reader  into  A595,  as  included  in  the  equation.  In  comparison  to  

traditional  Miller’s  assays,  this  method  assesses  the  rate  of  β-­‐‑galactosidase  accumulation  

over  time.  Because  it  does  not  rely  on  visual  color  detection  to  determine  the  reaction  

endpoint  and  manual  recording  of  color  development  over  time,  there  is  less  room  for  

Page 83: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

67  

error.  The  factors  in  this  formula  account  for  variability  between  machines  and  were  

calibrated  to  the  devices  used.  

3.2.7 Mutation Rate

The  mutation  rate  of  each  strain  was  determined  based  on  rifampicin  resistant  

CFU  and  total  CFU  of  at  least  8  independent  overnight  cultures  using  the  Lea  and  

Coulson  Method  of  the  Median  Fluctuation  Analysis  Calculator  (FALCOR)  program  

[94].  

3.2.7 Site-Directed Mutagenesis

The  Stratagene  QuikChange  Mutagenesis  kit  was  used  according  to  

manufacturer’s  instructions  to  generate  moderate  mutator  dnaQ928,  which  is  Gly17Ser  

[68].  We  created  strong  (QLSLPL)  and  weak  (ATSMAF)  β-­‐‑clamp  binding  mutants  using  

primers  QLSLPL  (5'ʹ-­‐‑CTGGCGATGACCGGTGGTCAACTATCGTTGCCTTTAGCGATG  

GAAGGAGAGAC-­‐‑3'ʹ),  QLSLPL_anti  (5'ʹ-­‐‑GTCTCTCCTTCCATCGCTAAAGGCAACGA  

TAGTTGACCACCGGTCATCGCCAG-­‐‑3'ʹ),  Q182A  (5'ʹ-­‐‑GGCGATGACCGGTGGTGCAA  

CGTCGATGGCTTTT-­‐‑3'ʹ)  and  Q182A_anti  (5'ʹ-­‐‑AAAAGCCATCGACGTTGCACCACCGG  

TCATCGCC-­‐‑3'ʹ).

3.2.8 DnaE Sequencing

The  DnaE  region  was  amplified  with  primers  dnaE_1  (5’-­‐‑GACTTGCGTCCT  

GATTCTTG-­‐‑3’)  and  dnaE_8  (5’-­‐‑ TAACCGCAGTCAGAGAATCG-­‐‑3’),  which  bind  to  the  

5’  and  3’  flanking  regions  of  the  gene.  Sequencing  of  PCR  products  was  performed  with  

Page 84: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

68  

dnaE_4  (5’-­‐‑ CTGGAGATGATCAACAAGCG-­‐‑3’),  dnaE_5  (5’-­‐‑ TGATCAGGTCCTTC  

ATGCC-­‐‑3’),  dnaE_1  and  dnaE_8.    

3.3 Results and Conclusions

3.3.1 Selection and Gene Replacement of ε Mutants

As  outlined  above,  our  goal  was  to  isolate  separation  of  function  mutants.  When  

selecting  a  subset  of  mutants  to  test,  we  relied  on  available  structural,  biochemical  and  

genetic  data  in  the  literature.  We  included  insertions  within  both  the  N186  domain  and  

the  C-­‐‑terminal  segment.  Insertions  within  the  N-­‐‑terminal  were  selected  based  on  

structural  data  mapped  using  Protein  Database  ID:  2XY8  and  2GUI  (Figure  21)  [103].  

While  selecting  insertions  within  N186,  we  chose  alleles  we  predicted  would  not  greatly  

disrupt  proofreading  and  exonuclease  function  based  on  homology.  This  included  

avoiding  highly  conserved  residues  and  early  truncations  that  would  likely  mimic  

ΔdnaQ.  Of  particular  interest  were  targets  on  the  outer  surfaces  of  the  protein  rather  

than  buried  in  the  core.    

In  addition  to  insertion  mutants,  we  generated  three  site-­‐‑specific  mutants  as  well.  

As  mentioned  earlier,  the  proofreading  defect  of  a  dnaQ  mutant  causes  mutation  

frequency  to  range  from  six  to  thousands  fold  higher  than  wildtype  levels.  As  an  

internal  control  we  recreated  allele  dnaQ928  to  identify  moderate  mutator  activity.  This  

allele  has  a  dominant  mutation  in  the  exoI  motif  (G17S)  indicating  it  retains  the  ability  to  

associate  with  α  [68].  Recent  evidence  of  a  β-­‐‑clamp  binding  motif  within  ε,  QTSMAF,  led  

Page 85: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

69  

us  to  generate  strong  (QLSLPL)  and  weak  (ATSMAF)  alleles  as  used  in  [104].  These  

studies  showed  a  stronger  β-­‐‑clamp  interaction  with  ε  led  to  a  more  stable  and  processive  

Pol  III  core.  Including  these  alleles  gives  greater  insight  into  replisome  fidelity,  

especially  the  effect  of  altering  the  ε-­‐‑β  interface  as  it  relates  to  proofreading  and  SOS  

phenotypes.  

Once  the  desired  subset  of  insertion  and  site-­‐‑specific  mutants  was  determined  

(Table  5),  plasmid  mutations  were  moved  onto  the  chromosome,  replacing  dnaQ+.  

Chromosomal  gene  replacement  was  facilitated  by  pMAK705,  a  temperature  sensitive  

plasmid  with  chloramphenicol  resistance.  The  mutant  fragment  was  excised  by  ClaI  and  

PstI  digestions  and  cloned  into  the  polylinker  of  pMAK705  (Figures  19  &  22).  The  

insertion  mutant  allele  replaced  dnaQ  as  described  in  [93].  Chromosomal  mutants  were  

confirmed  with  gel  electrophoresis  and  sequencing  primers  dnaQseq1  and  dnaQseq2  

(Figure  21).  

3.3.2 Phenotypic Scoring

Proofreading  activity  of  dnaQ  was  assessed  using  the  well-­‐‑established  rifampicin  

resistance  as  a  measure  of  mutation  rate.  At  least  8  independent  cultures  of  each  strain  

were  grown  overnight  and  appropriate  aliquots  were  plated  on  rifampicin  or  LB  agar  to  

determine  resistant  and  total  CFU/mL.  The  SOS  profile  was  determined  using  a  high-­‐‑

throughput  β-­‐‑galactosidase  assay  in  microtiter  plates  (Figure  23).  

Page 86: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

70  

 

Figure  21.  Secondary  Structure  of  DnaQ  

This  image  is  based  on  structural  data  from  PDB  ID:  2xy8  and  4gx9.  Yellow  arrows  indicate  β-­‐‑sheets,  red  waves  indicate  α-­‐‑helices,  and  purple  bumps  represent  turns  [72,  73].  

Page 87: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

71  

Table  5.  DnaQ  Insertion  Mutants  

The  final  amino  acid  position  that  shares  homology  with  wildtype  ε  is  listed.  Amino  acids  of  insertion  compositions  are  shown  (those  with  asterisks  indicate  stop  codons).  The  transposon  inserts  a  unique  10  base  pair  sequence,  while  duplicating  a  5  base  pair  target  for  a  final  insertion  of  15  base  pairs.  The  targeted  base  pairs  are  listed.  

 Amino  Acid  Position  

Insertion  Composition  

5-­‐‑bp  target  sequence  

30   IICLN   TCATT  57   LVFKQ   GCTGG  76   CLNNK   ATAAG  79   CLNTF   CGTTT  97   ILFKQ   GATCC  115   KLFKH   TAAGC  125   LFKHF   TTTCT  132   V*   CTTGC  143   NSCLN   GTTGA  162   LFKQL   GCTGC  196   LFKQQ   GCAAC  208   VFKHQ   CCTGA  224   AHV*   GCTCA  228   CLNKA   AAGCC  232   LFKQV   GGTGC  

Page 88: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

72  

 

 

Figure  22.  Identification  of  Chromosomal  Gene  Replacement  Products  via  Gel  Electrophoresis  

Gene  replacement  candidates  were  PCR  amplified  using  chromosome  binding  primers  dnaQ5F  and  dnaQI3R2.  The  PCR  products  were  digested  with  PmeI  to  identify  candidates  with  the  insertion  mutant  allele  on  the  chromosome  (white  arrow).  When  the  PCR  product  contains  the  insertion  mutant,  the  wildtype  fragment  (~3kb)  is  digested  into  two  fragments.  The  size  of  the  digested  fragments  depends  on  the  location  of  the  insertion  withi  PCR  product.  Following  identification  of  candidates  via  gel  electrophoresis,  the  samples  were  sent  to  sequencing  with  primers  dnaQseq1  and  dnaQseq2  to  confirm  gene  replacement.  

Page 89: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

73  

 

 

 

 

 

 

 

 

 

Figure  23.  Kinetic  B-­‐‑Galactosidase  Assay  

 Following  lysis  of  cells,  an  aliquot  of  the  lysed  cells  is  added  to  Z-­‐‑Buffer,  which  contains  ONPG.  Hydrolysis  of  ONPG  is  directly  correlated  to  the  amount  of  β-­‐‑galactosidase  present  in  the  cells.  Accumulation  of  O-­‐‑nitrophenol  is  measured  by  absorbance  at  405  over  time,  creating  a  kinetic  measure  of  β-­‐‑galactosidase  activity  that  can  be  directly  analysed  based  on  output  from  the  microtiter  plate  reader.  

 

 

 

 

                               

ONPG    (colorless)

β-­‐galactosidase  

H2O

Galactose  (colorless)

O-­‐nitrophenol  (yellow)

Yellow  color  directly  correlates  to    O-­‐‑nitrophenol  accumulation  is     Rate  of  O-­‐‑nitrophenol  

absorbance  over  time  

Page 90: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

74  

3.3.3. Overview

A  compilation  of  all  three  phenotypes  showed  a  striking  grouping  of  mutants  

based  on  mutation  rate  (Figure  24).  From  left  to  right  in  Figure  24,  the  insertion  mutants  

were  grouped  into  three  categories:  high  mutator  (like  ΔdnaQ),  moderate  mutator  (like  

DnaQ928),  and  non-­‐‑mutator  (like  WT).  It  was  clear  where  most  mutants  grouped  except  

the  two  β-­‐‑clamp  binding  mutants  (Q182A  and  QLSLPL)  and  162LFKQL.  These  mutation  

rates  of  these  three  were  2-­‐‑5  fold  higher  than  wildtype  but  not  high  enough  to  be  

confidently  considered  moderate  mutators.    The  lowest  mutation  rate  of  the  moderate  

mutator  group  is  11  fold  times  higher  than  wildtype  (228CLNKA).  

3.3.4 Mutants that behave like Wildtype

Wildtype  strains  that  are  DnaQ+  are  non-­‐‑mutators,  are  not  SOS  constitutive  and  respond  

to  nalidixic  acid  treatment  with  high  levels  of  SOS  induction.  Insertions  115KLFKH,  

125LFKHF  and  196LFKQQ  behave  like  wildtype  (Figure  24,  Table  6).  The  insertion  at  115  is  

at  the  end  of  alpha  helix  4  and  between  motifs  exoII  and  exoIIIε.  In  addition  to  this  it  is  

on  the  outside  edge  of  the  protein,  so  this  mutation  may  have  minimal  effects  on  protein  

folding  and  interactions  with  other  protein  partners  (Figure  25).  The  insertion  at  

125LFKHF  has  a  mutation  rate  that  is  not  significantly  higher  than  wildtype;  however  we  

collected  fluctuating  nalidixic  acid-­‐‑induced  SOS  results  with  this  mutant.  This  insertion  

is  also  between  exoII  and  exoIIIε  and  is  on  the  outside  edge  of  the  protein.  NMR  studies  

of  the  C-­‐‑terminal  segment  of  ε  determined  that  the  Q-­‐‑linker  region  and  especially  the  

Page 91: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

75  

stretch  of  four  glutamines  is  highly  flexible.  Insertion  196LFKQQ  is  within  the  four  

glutamines,  and  it  is  possible  that  the  flexible  nature  of  this  region  may  make  it  more  

tolerable  to  the  insertion  or  other  types  of  manipulation.    

3.3.5 Mutants that behave like ΔDnaQ

Loss  of  ε  leads  to  high  mutation  rate,  partially  constitutive  SOS  and  the  failure  to  

induce  SOS  following  nalidixic  acid  treatment.  Mutants  like  ΔDnaQ  are  30IICLN,  

79CLNTF,  97ILFKQ,  232LFKQV,  and  132V*  (Figures  25-­‐‑28).  Although  the  insertion  at  30  is  

not  within  the  conserved  exoI  or  exoII  motifs,  it  is  at  the  core  of  the  protein  on  β-­‐‑strand  2  

(Figure  26).  This,  more  than  likely,  disrupts  protein  folding.  The  truncation  at  132  is  

expected  to  behave  like  ΔDnaQ  as  it  eliminates  much  of  the  exoIIIε  motif  and  the  entire  

C-­‐‑terminal  segment  (Figure  27).  Insertion  97ILFKQ  maps  very  closely  to  the  conditionally  

lethal  mutator  dnaQ49  (V96G).  Both  mutations  are  within  the  exo  II  motif  of  ε.  DnaQ49  is  

recessive  and  cannot  bind  to  α,  but  we  do  not  know  if  the  insertion  at  97  is  simply  

catalytically  defective  or  deficient  in  α-­‐‑binding  as  well  [79,  84].  It  is  possible  that  the  

insertion  mutant  79CLNTF  disrupts  protein  stability  as  well.  

Studies  have  shown  that  the  εCTS  end  of  the  protein  is  responsible  for  

association  with  the  α  subunit,  and  the  two  proteins  can  be  cross-­‐‑linked  via  residue  235  

in  εCTS  [60,  71,  75,  105].  Residues  218-­‐‑237  form  an  alpha  helix  and  insertion  mutant  

232LFKQV  maps  within  this  region  (Figure  28)[77].  Because  this  mutant  has  no  mutations  

Page 92: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

76  

in  the  exonuclease  domain,  it  is  assumed  the  insertion  disrupts  the  ability  of  ε  to  bind  to  

α  or  protein  stability.    

Page 93: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

77  

 

Figure  24.  SOS  &  Mutator  Profiles  of  Chromosomal  Insertion  Mutants  

 Mutation  rate  is  indicated  by  the  small  black  rectangles  and  corresponds  to  the  axis  on  the  right.  Light  gray  bars  indicate  β-­‐‑galactosidase  activity  in  the  absence  of  drug;  dark  gray  bars  indicate  β-­‐‑galactosidase  activity  following  nalidixic  acid  treatment.  Dotted  lines  group  mutation  rate  into  high,  moderate  and  low  from  left  to  right.  

Page 94: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

78  

Table  6.  Mutation  Rate  and  SOS  Profiles  of  Insertion  Mutants  

P  values  are  compared  to  WT  for  each  phenotype,  and  *  indicates  p  <  0.05;  **  indicates  p  <  0.005.  

 

 

 

 

Page 95: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

79  

 

Figure  25.  Mutant  115KLFKH  Insertion  Location  

The  structure  of  ε186  from  PDB  ID:  2xy8  is  shown.  The  red  residue  indicates  the  location  of  the  insertion  site  for  amino  acids  KLFKH  at  site  115  [106]

Page 96: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

80  

3.3.6 Proofreading can be Uncoupled from SOS Phenotypes

To  examine  the  relationship  between  proofreading  and  SOS  phenotypes,  we  

analyzed  mutation  rates  of  various  insertion  mutants  in  comparison  to  SOS  activity.  

Most  of  the  mutants  grouped  with  either  wildtype  or  the  ε  knockout.  We  did  not  find  

complete  separation  of  function  mutants,  particularly  high  mutator  with  low  basal  SOS  

or  low  mutator  with  partially  constitutive  SOS.  In  Figure  29,  mutants  exhibiting  

moderate  mutator  activity  yet  wildtype  basal  levels  of  SOS  are  grouped  in  the  lower  left  

quadrant.  Mutants  displaying  this  type  of  functional  separation  are  DnaQ928  and  

insertions  76CLNNK  (Figure  30),  143NSCLN,  and  57LVFKQ.  While  DnaQ928  and  

143NSCLN  are  in  motifs  exoI  and  exoIIIε  respectively,  the  other  insertions  are  not  within  

the  conserved  residues  related  to  proofreading.  Insertion  57LVFKQ  lies  between  motifs  

exo  I  and  exo  II,  and  this  agrees  with  published  results  that  implicate  this  region  for  full  

proofreading  function  [68].  

The  mutants  with  mutator  activity  yet  low  basal  SOS  could  be  explained  by  two  

models.  Either  SOS  is  induced  by  only  extremely  high  levels  of  misincorporation  or  the  

two  SOS  phenotypes  are  caused  by  two  different  functions  of  ε.  Further  experiements  

are  necessary  to  distinguish  between  these  models.

Page 97: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

81  

 

Figure  26.  Mutant  30IICLN  Insertion  Location  

Mutant  30IICLN  behaves  like  ΔDnaQ.  The  red  residue  indicates  the  location  of  the  insertion.  It  is  on  one  of  the  central  β  sheets  at  the  core  that  structural  analysis  has  proven  essential  for  protein  stability.

Page 98: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

82  

 

 

 

 

 

Figure  27.  Mutant  132V*  Insertion  Location  

A.  The  insertion  at  residue  132  creates  a  truncation.  The  red  residue  in  the  ε186  structure  represents  the  end  of  the  protein  in  this  insertion  mutant,  eliminating  the  remainder  of  the  yellow-­‐‑colored  helix  through  the  turn  and  salmon-­‐‑colored  helix  [106].  B.  The  secondary  structure  shows  the  amino  acids  that  are  eliminated  in  this  insertion  mutant  (right  of  the  arrow),  which  is  a  truncation.    

B  

A  

Page 99: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

83  

 

Figure  28.  Crystal  structure  of  (εCTS35  -­‐‑  α270)  

This  is  a  fusion  protein  of  the  α  subunit  (cyan  blue)  and  35  C-­‐‑terminal  residues  of  ε  (green).  Image  from  [71]  

Page 100: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

84  

 

Figure  29.  Separation  of  Proofreading  and  Untreated  SOS  Levels  

Dot  plot  of  mutation  rate  based  on  FALCOR  Lea  and  Coulson  Method  of  the  Median  compared  to  untreated  SOS  levels.  Mutants  are  color-­‐‑coded:  separation  of  function  mutants  (orange),  wildtype  (green),  ΔDnaQ  (red).

Page 101: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

85  

   

 

Figure  30.  Mutant  76CLNNK    Insertion  Location  

The  structure  of  ε186  from  PDB  ID:  2xy8  is  shown.  The  red  residue  indicates  the  location  of  the  insertion  site  for  amino  acids  CLNNK.  The  panel  on  the  right  shows  the  surface  model  of  ε186  (orange)  in  complex  with  homolog  of  θ  (grey).  The  insertion  mutant  76CLNNK  may  disrupt  the  interaction  with  homolog  of  θ.  Figure  created  with  [106,  107].

Page 102: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

86  

To  evaluate  the  relationship  between  proofreading  and  the  ability  to  induce  SOS  

following  nalidixic  acid  treatment,  we  plotted  mutation  rate  in  comparison  to  β-­‐‑

galactosidase  activity  (Figure  31).  Many  of  the  mutants  grouped  with  wildtype  or  the  ε  

knockout  (Figure  32),  but  we  found  striking  separation  of  function  mutants  -­‐‑  208VFKHQ  

and  143NSCLN  (Figure  33).    

Insertion  mutant  208VFKHQ  has  functional  proofreading  yet  is  unable  to  induce  

SOS  after  nalidixic  acid  exposure.  This  is  within  the  εCTS,  outside  the  proofreading  

domain.  While  this  mutant  has  no  changes  within  the  exonuclease  domain,  the  C-­‐‑

terminal  segment  is  implicated  in  the  ability  of  ε  to  associate  with  the  α  subunit  (Figure  

28).  This  indicates  that  an  insertion  at  this  site  does  not  disrupt  that  ability.  

Insertion  mutant  143NSCLN  presents  a  different  type  of  separation  of  function.  

With  diminished  proofreading  function  that  leads  to  a  22-­‐‑fold  increase  in  mutation  rate  

but  retains  the  ability  to  induce  SOS  following  nalidixic  acid.  Mutant  143NSCLN  is  within  

ExoIIIε,  consistent  with  the  observed  mutator  phenotype  (Figure  33,  Table  6).  The  ability  

of  this  mutant  to  induce  SOS  after  nalidixic  acid  suggests  the  insertion  does  not  disrupt  

whatever  interaction  is  necessary  for  SOS  induction.  

Interestingly,  insertion  mutant  162LFKQL  has  a  mutation  rate  similar  to  wildtype  and  

induces  SOS  after  nalidixic  acid  at  levels  significantly  higher  (74%,  p=0.0064)  than  

wildtype  (Figure  24,  Figure  34,  Table  6).  This  insertion  is  within  the  exoIIIε  motif  and  

immediately  follows  the  conserved  residue  162H,  which  caused  a  1,000-­‐‑fold  increase  in  

Page 103: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

87  

rifampicinR-­‐‑measured  mutation  frequency  when  it  was  mutated  [108].  It  is  likely  that  

this  mutant  is  a  non-­‐‑mutator  because  the  insertion  does  not  disrupt  the  location  of  the  

conserved  exonuclease  residues  (DEDDH)  that  group  around  the  active  site  in  ε.  Mutant  

162LFKQL  may  represent  a  novel  class  of  dnaQ  mutants  that  are  hyper-­‐‑responsive  to  

nalidixic  acid  –induced  cleavage  complexes.    

 

Page 104: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

88  

 

Figure  31.  Separation  of  Proofreading  and  Nalidixic  Acid-­‐‑Induced  SOS  Activity  

Dot  plot  of  mutation  rate  based  on  FALCOR  Lea-­‐‑Coulson  Method  of  the  Medians  compared  to  SOS  induction  following  nalidixic  acid  treatment  for  2  hours.  ΔNal  is  the  difference  between  SOS  levels  when  treated  with  nalidixic  acid  and  when  uninduced  (Nal  SOS  –  uninduced  SOS).  Mutants  are  color  coded:  separation  of  function  mutants  (orange),  wildtype  (green),  ΔDnaQ  (red).    

Page 105: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

89  

 

Figure  32.  Relationship  between  the  Proofreading  and  Nalidixic  Acid-­‐‑Induced  SOS  

Possible  linear  relationship  between  the  nalidixic  acid-­‐‑induced  SOS  and  the  mutation  rate  of  the  insertion  mutants.  Bars  correspond  to  left  axis  (nalidixic-­‐‑acid  induced  SOS)  and  diamonds  correspond  to  right  axis  (mutation  rate).  Extremely  high  mutators  (near  ΔDnaQ,  red)  have  SOS  defects  while  mutants  like  wildtype  (green)  have  high  levels  of  SOS  induction  following  nalidixic  acid.  

Page 106: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

90  

 Figure  33.  Mutant  143NSCLN  Insertion  Location  

The  structure  of  ε186  from  PDB  ID:  2xy8  is  shown.  The  red  residue  indicates  the  location  of  the  insertion  site  for  amino  acids  NSCLN.  The  panel  on  the  right  shows  the  surface  model  of  ε186  in  complex  with  θ.  Figure  created  with  [106].  

Page 107: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

91  

 

   

   

Figure  34.  Mutant  162LFKQL  Insertion  Location  

The  structure  of  ε186  from  PDB  ID:  2xy8  is  shown.  The  red  residue  indicates  the  location  of  the  insertion  site  for  amino  acids  LFKQL.  The  panel  on  the  right  shows  the  surface  model  of  ε186  (orange)  in  complex  with  θ  (grey).  Figure  created  with  [106,  107].  

 

Page 108: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

92  

3.3.7 SOS Phenotypes are Separable

To  assess  if  the  SOS  phenotypes  were  linked,  we  compared  the  basal  SOS  levels  

with  the  increase  in  SOS  following  nalidixic  acid  treatment,  and  the  mutants  scattered  

widely  (Figure  35).  Of  the  mutants  with  defective  SOS  induction  following  nalidixic  acid  

treatment,  phenotypes  ranged  from  basal  SOS  similar  to  wild  type  (Q182A,  76CLNNK  and  

208VFKHQ;  separation  of  function  for  SOS)  to  significantly  higher  constitutive  levels  than  

the  ε  knockout  (97ILFKQ,  30IICLN  and  232LFKQV).  We  also  identified  partially  

constitutive  mutants  that  are  able  to  induce  SOS  following  nalidixic  acid  treatment  

(162LFKQL  and  QLSLPL;  discussed  above  and  below,  respectively).  

3.3.8 β-Clamp Binding Mutants

The  β-­‐‑clamp  binding  mutants  present  as  interesting  separation  of  function  alleles.  

QLSLPL,  the  strong  β-­‐‑clamp  binding  mutant,  retains  catalytic  function  as  expected.  

Surprisingly,  this  mutant  has  a  basal  SOS  level  that  is  statistically  significantly  higher  

than  wildtype  (p=0.0000134)  and  closer  to  the  ε  knockout.  This  mutant  also  

demonstrates  separation  of  SOS  phenotypes.  Following  nalidixic  acid  treatment,  

QLSLPL  is  able  to  fully  induce  SOS  at  levels  higher  than  wildtype.  This  mutant  was  

shown  to  have  a  stronger  ε-­‐‑β  interaction  than  wildtype  [77,  104].  If  the  replisome  needs  

to  be  rapidly/easily  dissociated  when  encountering  endogenous  damage,  this  could  be  a  

barrier  to  that  process.  This  could  lead  to  the  partially  constitutive  phenotype.  

Page 109: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

93  

 

 

Figure  35.  Separation  of  Constitutive  SOS  and  Nalidixic  Acid-­‐‑Induced  SOS  

Dot  plot  of  chromosomal  insertion  mutants  comparing  no  drug  SOS  levels  to  nalidixic  acid-­‐‑induced  SOS  activity.  

Page 110: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

94  

The  weak  β-­‐‑clamp  binding  mutant,  Q182A,  reveals  an  interesting  collection  of  

phenotypes.  It  demonstrated  rather  weak  mutator  activity  (only  5  fold  higher  than  

wildtype),  basal  SOS  that  was  significantly  lower  than  wildtype  (p=0.0042)  and  greatly  

reduced  SOS  induction  following  nalidixic  acid  treatment.  These  results  implicate  the  ε-­‐‑

β  clamp  interaction  in  the  SOS  phenotypes  caused  by  the  absence  of  ε.  

3.3.9 Testing for dnaE Suppressor Mutations

DnaQ  mutants  rapidly  acquire  suppressor  mutations  in  dnaE  to  relieve  the  effects  

of  losing  ε  function.  We  analyzed  the  dnaE  sequence  for  the  spq-­‐‑2  allele  (V832G)  known  to  

complement  defective  growth  of  dnaQ  mutants  [109,  110].  We  also  searched  for  two  of  

the  stronger  antimutators  in  rich  media,  dnaE911  (P257L)  and  dnaE925  (F388L).  DnaE911  

reduced  the  rifampicinR-­‐‑determined  mutation  frequency  of  mutD5  strain  by  10  fold  and  

dnaE925  reduced  the  mutation  frequency  by  100  fold  [111,  112].  None  of  our  mutant  

strains  had  any  of  the  suppressor  mutations  described  above  nor  additional  mutations  in  

sequence  reads  of  good  quality.  However,  we  did  find  a  mutation  in  ΔdnaQ  that  was  

previously  uncharacterized,  I846V.  This  mutation  is  very  close  to  spq-­‐‑2  and  could  assist  in  

growth  of  the  ΔdnaQ  strain.  

Page 111: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

95  

3.4 Discussion

The  multiple  phenotypes  of  dnaQ  knockout  mutants  are  consistent  with  the  

possibility  that  ε  has  multiple  functions  (see  Introduction).  This  possibility  was  raised  

decades  ago,  supported  by  the  isolation  of  suppressor  mutations  within  dnaE  that  

restore  the  growth  deficiency  but  not  the  proofreading  defect  of  a  dnaQ  mutant  [110,  

113].  The  isolation  of  separation  of  function  mutants  in  this  chapter  strongly  supports  

the  proposal  of  multiple  ε  functions  (Table  7).  Certain  separation  of  function  mutants  

isolated  here  also  provide  a  very  useful  reagent,  namely  ε variants  with  defects  in  

functions  relating  to  SOS  induction  but  wild-­‐‑type  mutation  rates,  i.e.  mutationally-­‐‑stable  

strains  that  do  not  produce  confounding  results  due  to  many  spurious  global  mutations.  

Our  data  also  identify  sites  in  ε  that  are  tolerant  of  5  amino-­‐‑acid  insertions  with  no  

phenotypic  consequence  (115,  125,  196).  

Many  of  the  mutants  with  a  high  mutator  phenotype  also  failed  to  induce  SOS  after  

nalidixic  acid  treatment.  While  this  seems  to  suggest  that  proofreading  and  SOS  

induction  might  be  linked,  this  is  not  necessarily  the  case.  Any  mutant  that  is  completely  

inactive  (e.g.  encoding  unstable  proteins)  is  equivalent  to  an  ε  knockout,  which  cannot  

cause  separation  of  function.  Similarly,  strains  with  normal  mutational  rates  and  the  

ability  to  induce  SOS  could  simply  encode  fully  functional  ε.  These  results  would  appear  

as  a  correlation  when  sorted,  but  the  correlation  is,  in  a  sense,  trivial.  The  more  rare  

separation  of  function  mutant,  such  as  208VFKHQ,  is  much  more  informative,  

Page 112: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

96  

demonstrating  that  SOS  induction  can  be  compromised  without  loss  of  proofreading,  

and  providing  the  strongest  evidence  for  a  separable  function  relating  to  damage  

response.  

Regarding  the  relationship  between  proofreading  function  and  the  SOS  

constitutive  phenotypes,  we  did  not  isolate  mutants  with  complete  separation  of  

function  (high  mutator/low  basal  SOS  or  low  mutator/partially  constitutive  SOS).  

However,  we  did  identify  a  number  of  mutants  with  moderate  mutator  activity  

(reduced  proofreading)  yet  wild-­‐‑type  levels  of  basal  SOS  expression  (dnaQ928,  

76CLNNK,  143NSCLN,  and  57LVFKQ).  These  results  are  consistent  with  the  possibility  

that  SOS  is  related  to  proofreading,  but  is  induced  only  by  extreme  levels  of  

misincorporation.  They  are  also  consistent  with  the  possibility  that  the  two  phenotypes  

are  caused  by  two  different  functions  of  ε.  

Results  with  several  of  the  mutants  analyzed  here  can  be  directly  compared  to  

prior  biochemical  studies  to  allow  molecular  inferences.  For  example,  ε  Q208L,  also  

called  dnaQ516,  was  able  to  associate  with  the  α  subunit  (as  tested  via  gel  filtration  

chromatography)  and  had  slightly  diminished  proofreading  activity  [114,  115].  Residue  

208  is  immediately  before  Exo  III  and  our  insertion  mutant  at  the  same  site  (208VFKHQ)  

exhibited  functional  proofreading.  In  contrast,  our  insertion  143NSCLN  is  within  the  Exo  

IIIε  motif  and  exhibits  moderate  mutator  activity.  Our  results  thus  echo  past  studies  

Page 113: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

97  

demonstrating  the  importance  of  Exo  IIIε,  rather  than  the  earlier  proposed  Exo  III  motif,  

to  catalytic  activity  of  ε  [66,  68].    

 

Page 114: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

98  

 

Table  7.  Summary  of  Separation  of  Function  Mutants  

Yes  or  no  indicates  the  presence  or  absence,  respectively,  of  the  listed  phenotype.  Mutator  activity  marked  with  M  represents  the  moderate  mutator  class.  

 

 

Page 115: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

99  

The  insertion  57LVFKQ  is  one  amino  acid  away  from  a  previously  characterized  

allele,  dnaQ920  (R56W),  located  between  motifs  ExoI  and  ExoII  [68].  Both  are  moderate  

mutators,  and  R56W  has  also  been  characterized  as  a  dominant  allele.    

Our  results  can  be  viewed  in  the  context  of  recent  studies  on  the  clamp  binding  

motif  of  ε and  the  interactions  between  the  clamp,  ε  and  α  [77,  104,  105].  Toste  Rêgo  et  al  

(2013)  analyzed  the  interactions  of  the  three  proteins,  demonstrating  that  ε  enhances  the  

stability  of  the  clamp-­‐‑α  complex  and  identifying  sites  where  each  pair  of  proteins  can  be  

crosslinked  to  each  other.  The  increase  in  stability  of  the  complex  due  to  ε binding  is  

presumably  related  to  the  stimulation  of  polymerase  processivity  by  ε  [59,  116],  and  

clamp  binding  by  ε  in  turn  improves  proofreading  ability  [105].  Based  on  these  and  

other  results,  Toste  Rêgo  et  al  (2013)  developed  a  model  in  which  α  and  ε  

simultaneously  bind  to  the  canonical  binding  pocket  of  the  two  protomers  of  the  β  

clamp  dimer.  Importantly,  they  also  identified  crosslinks  between  α  and  both  the  

proofreading  NTD  of  ε  and  the  extreme  εCTS  (residue  235).  These  results  led  to  an  

elegant  model  for  how  the  polymerase  complex  responds  to  both  template  damage  and  

misincorporation  (Figure  36).  

Page 116: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

100  

 

Figure  36.  Model  for  Polymerase  Switching  

This  model  was  proposed  by  Toste  Rêgo  et  al.  (2013)  and  suggests  ε  is  removed  from  the  β  clamp  protomer  while  remaining  bound  to  α.  This  gives  translesion  polymerases  access  to  the  β  clamp.  By  remaining  bound  to  α,  ε  can  easily  regain  access  to  the  DNA  following  synthesis  past  the  lesion.

Page 117: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

101  

As  mentioned  above,  misincorporated  bases  are  more  efficiently  proofread  by  ε  

in  the  context  of  the  tripartite  complex.  Template  damage,  on  the  other  hand,  cannot  be  

corrected  by  ε,  but  rather  is  proposed  to  cause  weakening  of  the  clamp-­‐‑ε  interaction,  

which  can  allow  binding  of  the  translesion  polymerase  Pol  IV.  One  important  feature  of  

the  model  is  that  ε  remains  associated  with  α  via  the  flexible  εCTS  even  when  the  ε-­‐‑

clamp  association  is  lost.  Several  aspects  of  this  model  could  relate  to  the  phenotypes  of  

ε  mutants  analyzed  here,  and  we  propose  that  the  dynamic  behavior  of  the  tripartite  

complex  described  above  could  directly  relate  to  the  response  of  the  replisome  to  

endogenous  lesions  and  to  blocking  gyrase  cleavage  complexes.  

As  the  tripartite  complex  encounters  endogenous  damage,  the  weak  clamp  

binder  (Q182A)  would  be  expected  to  more  readily  dissociate  from  the  clamp,  allowing  

translesion  polymerases  to  access  the  free  site  on  the  β  clamp  (Figure  36,  Translesion  

Orientation).  Q182A  has  weak  mutator  activity  (5-­‐‑fold  higher  than  wildtype)  and  we  

propose  two  mechanisms  that  could  lead  to  this.  One  scenario  is  that  Q182A  has  weaker  

exonuclease  activity  and  this  leads  to  less  efficient  proofreading.  Diminished  

proofreading,  rather  than  a  complete  loss  of  proofreading,  could  explain  the  weak  

mutator  activity  observed.  The  other  scenario  is  that  the  weak  binder  allows  Pol  IV  to  

access  the  β  clamp  more  often,  with  or  without  a  lesion  present.  Pol  IV  is  a  low-­‐‑fidelity  

polymerase  and  this  occasional  access  could  be  the  cause  the  weak  mutator  activity.  If  

the  weak  clamp  binder  allows  the  translesion  polymerase  to  associate  more  easily,  it  

Page 118: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

102  

could  synthesize  past  the  lesion  and  dissociate,  then  Q182A  (still  bound  to  α)  would  

regain  access  to  the  clamp  to  continue  high-­‐‑fidelity  replication  without  inducing  SOS.  

More  efficient  TLS  across  the  endogenous  template  damage  is  proposed  to  reduce  SOS.  

This  model  aligns  with  the  (lower  than  wildtype)  basal  SOS  levels  reported  for  Q182A.    

Under  these  same  assumptions,  the  strong  clamp  binder,  QLSLPL,  would  remain  

bound  to  the  β  clamp,  preventing  access  by  the  translesion  polymerase  (Figure  36,  

Replication  Orientation).  This  is  consistent  with  the  similar-­‐‑to-­‐‑wildtype  mutation  rate  

observed  with  this  mutant.  When  facing  endogenous  damage,  the  replication  fork  with  ε  

tightly  bound  to  α  would  stall,  as  the  translesion  polymerase  does  not  have  access  to  the  

β  clamp  in  QLSLPL  and  the  damage  would  remain  not  repaired.  This  prolonged  

polymerase  stalling  could  result  in  increased  SOS  induction.  This  would  explain  the  

constitutive  SOS  phenotype  of  QLSLPL.    

It  is  unclear  how  the  tripartite  model  fits  with  the  observed  SOS  levels  following  

nalidixic  acid  treatment.  The  weak-­‐‑binding  mutant  Q182A  remains  unable  to  induce  SOS  

after  nalidixic  acid,  while  QLSLPL  induces  SOS  at  levels  significantly  higher  than  

wildtype.  This  suggests  that  SOS  induction  is  prevented  by  the  formation  of  the  

“translesion”  orientation  of  the  complex  in  Figure  36,  and  favored  by  the  (stalled)  

version  of  the  tighter  complex,  in  which  α,  β,  and  ε  are  fully  interacting.  The  only  known  

requirements  for  SOS  induction  after  nalidixic  acid  are  RecBC  and  ε.  This  indicates  that  

Page 119: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

103  

double  strand  break  formation  is  critical  for  nalidixic  acid-­‐‑induced  SOS.  We  will  discuss  

possible  models  for  these  observations  in  the  context  of  DSB  formation  in  Chapter  5.    

There  are  multiple  models  for  quinolone-­‐‑induced  cytotoxicity  and  this  reflects  

the  multiple  groups  of  drugs  within  the  overarching  quinolone  class.  There  is  evidence  

that  some  quinolones  require  protein  synthesis  for  cytotoxicity,  while  others  do  not.  

There  are  also  replication-­‐‑dependent  models,  which  require  active  replication  forks  for  

quinolone  treatment  to  cause  cell  death.  Future  studies  with  these  ε  mutants  to  

determine  how  they  affect  quinolone  cytotoxicity  could  provide  useful  information  to  

further  define  the  role  of  ε  and  possibly  the  interaction  with  the  β  clamp  in  the  bacterial  

quinolone  response.  

 

Page 120: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

104  

4. Distinguishing Between Latent SCC Breaks and Irreversible, Downstream Breaks

4.1 Introduction

Quinolones  bind  topoisomerase-­‐‑DNA  complexes  to  create  stabilized  cleavage  

complexes  [39,  40].  Stabilized  cleavage  complexes  are  necessary  for  quinolone  

cytotoxicity,  but  they  are  not  sufficient.  Cleavage  complexes  accumulate  in  vivo  

following  exposure  to  topoisomerase  poisons,  even  under  conditions  that  do  not  cause  

lethality  [52].  Furthermore,  stabilized  cleavage  complexes  are  reversible  under  various  

physiological  conditions.  Washing  the  drug  out  or  treating  with  heat  (60  –  65°C)  prior  to  

protein  denaturation  allows  resealing  of  the  latent  break  within  the  cleavage  complex  

[39,  46,  117,  118].    

The  literature  about  quinolones  suggests  that  cytotoxicity  arises  from  conversion  

of  SCCs  into  irreversible  double  strand  breaks  [46].  RecBC  mutants  are  hypersensitive  to  

quinolone  treatment;  this  indicates  that  RecBC,  which  is  necessary  to  repair  double  

strand  breaks,  is  required  to  repair  quinolone-­‐‑induced  damage.  Additionally,  quinolone  

treatment  causes  RecBC-­‐‑dependent  SOS  induction  [47,  48,  119].  

The  processing  of  quinolone-­‐‑induced  SCCs  and  subsequent  double  strand  break  

generation  causes  chromosome  fragmentation.  Chromosome  fragmentation  was  proven  

by  sedimentation  analysis  [52,  117].  More  recent  work  using  lysate  viscosity  confirmed  

earlier  reports  that  nalidixic  acid  causes  chromosome  fragmentation  [52,  120].  

Page 121: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

105  

4.1.1 Models of Quinolone Cytotoxicity

The  mechanism  between  accumulation  of  SCCs  and  double  strand  break  

generation  is  not  clearly  understood.  Multiple  models  have  been  proposed  to  explain  the  

mechanisms  that  lead  to  cytotoxicity;  some  are  specific  to  particular  subgroups  of  

quinolones  and  they  are  not  necessarily  mutually  exclusive.  The  most  widely  accepted  

models  can  be  separated  into  two  groups:  replication  independent  and  replication  

dependent.  Replication-­‐‑independent  models  include  gyrase  subunit  dissociation  and  

endonucleolytic  cleavage  processing  of  SCCs,  while  the  replication-­‐‑dependent  models  

include  replication  run-­‐‑off  and  the  collateral  damage  model.    

The  gyrase  subunit  dissociation  model  suggests  a  replication-­‐‑independent  mode  

of  quinolone  cytotoxicity,  whereby  the  gyrase  tetramer  dissociates,  releasing  two  DNA  

ends  that  are  covalently  attached  to  the  free  subunits  (Figure  37A)  [52].  Protein  synthesis  

inhibition  blocks  the  cytotoxicity  of  nalidixic  acid  and  oxolinic  acid,  but  it  does  not  block  

killing  by  ciprofloxacin  [52,  117,  121,  122].  Ciprofloxacin,  a  more  potent  fluoroquinolone,  

caused  DNA  breaks,  even  when  treated  with  chloramphenicol.  This  led  to  the  proposal  

that  the  protein  synthesis-­‐‑independent  mode  of  killing  is  due  to  release  of  DNA  ends  in  

the  quinolone-­‐‑DNA-­‐‑gyrase  complex  [52].  

An  in  vivo  study  reported  that  illegitimate  recombination  was  stimulated  by  

oxolinic  acid  and  could  be  blocked  by  a  gyrA  mutation  that  conferred  resistance  rather  

Page 122: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

106  

than  a  recA  mutation  [123].  In  Streptomyces  ambofaciens,  fluoroquinolone  treatment  

(ciprofloxacin,  enoxacin  and  norfloxacin)  also  stimulated  genome  instability  and  

rearrangements  [124].  These  findings  supported  a  model  of  gyrase  subunit  dissociation  

and  exchange.  As  described  above,  oxolinic  acid  cytoxicity  is  blocked  by  

chloramphenicol  treatment,  while  ciprofloxacin  is  not.  Based  on  these  findings,  it  was  

suggested  that  recombination  assays  are  a  more  sensitive  measure  of  gyrase  subunit  

dissociation  than  cell  death  and  that  fluoroquinolones  are  more  effective  at  stimulating  

illegitimate  recombination  than  first  generation  quinolones  [46].  

The  endonucleolytic  cleavage  model  suggests  that  SCCs  are  detected  by  an  

endonuclease  that  cleaves  the  DNA  flanking  the  SCCs  and  simultaneously  creates  free  

double  strand  ends  (Figure  37B).  If  this  endonuclease  is  already  present,  then  cell  death  

may  be  protein  synthesis-­‐‑independent.  This  model  mimics  the  meiosis-­‐‑specific  removal  

of  Spo11,  a  topoisomerase-­‐‑like  protein  in  Saccharomyces  cerevisiae.  Spo11  shares  sequence  

homology  with  the  type  II  topoisomerase,  topo  VI,  and  it  creates  double  strand  breaks  

while  remaining  covalently  bound  to  the  5’  termini  of  the  DNA  via  the  catalytic  tyrosine  

residue  [125,  126].  In  budding  yeast,  Spo11  is  removed  from  the  DNA  via  the  

endonuclease  activity  of  the  MRX  complex  (Mre11-­‐‑Rad50-­‐‑Xrs2)  and  Sae2  [127,  128].  

There  is  also  evidence  that  Top2-­‐‑DNA  complexes  are  subject  to  endonucleolytic  

processing  as  well  [127].  

Page 123: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

107  

The  replication  run-­‐‑off  model  suggests  that  the  replication  machinery  physically  

encounters  SCCs  causing  gyrase  to  release  the  strand  of  DNA  that  is  not  covalently  

bound  to  the  enzyme  [129].  This  would  result  in  free  double  strand  ends  from  the  latent  

break  within  the  cleavage  complex,  potentially  causing  cell  death  (Figure  38A).  This  

model  has  supporting  evidence  with  eukaryotic  type  I  topoisomerase  inhibitors,  but  in  

vitro  studies  in  E.  coli  show  that  replication  machinery  approaching  SCCs  does  not  lead  

to  DNA  breaks,  but  rather,  is  physically  blocked  [130,  131].  

The  collateral  damage  model  posits  that  the  replication  fork  is  stalled  as  SCCs  

present  physical  barriers  to  its  progress.  The  stalled  replication  fork  leaves  exposed  

DNA,  which  is  then  converted  into  double  strand  breaks  by  recombination  nucleases  

(Figure  38B).  A  recent  study  in  our  lab  examining  replication  fork  dynamics  of  plasmid  

pBR322  found  norfloxacin  treatment  caused  stalled  replication  forks  and  breaks  [118].  In  

bacteriophage  T4,  a  similar  mechanism  has  been  documented  as  Y-­‐‑form  DNA  

accumulates  when  active  replication  forks  encounter  drug-­‐‑induced  cleavage  complexes.  

Inactivating  the  T4  recombination  endonuclease,  endonuclease  VII,  causes  further  

accumulation  of  stalled  replication  forks  indicating  that  endo  VII  is  responsible  for  

clearing  the  forks  [132].  This  further  supports  the  collateral  damage  model.    

There  can  be  multiple  modes  of  quinolone  cytotoxicity.  The  mechanism  of  

cytotoxicity  varies  depending  on  cell  conditions  or  may  be  specific  to  individual  drugs.  

Page 124: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

108  

These  models  do  not  have  to  be  mutually  exclusive,  and  it  is  possible  that  different  

mechanisms  are  activated  depending  on  cellular  conditions.  

Page 125: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

109  

 Figure  37.  Replication-­‐‑independent  models  of  quinolone  cytotoxicity  

 A.  The  gyrase  dissociation  model  proposes  the  subunits  simply  come  apart  and  the  double  strand  break  arises  from  the  latent  break  within  the  SCC.  B.  The  endonucleolytic  cleavage  model  proposes  an  endonuclease  detect  the  SCC  and  cleaves  the  flanking  regions  of  DNA  (purple  arrows).

A  

B  

Page 126: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

110  

 

Figure  38.  Replication-­‐‑dependent  models  of  quinolone  cytotoxicity.  

 A.  The  replication  run-­‐‑off  model  proposes  the  replication  fork  physically  encounters  SCCs  (red  bubble)  and  the  double  strand  break  comes  from  within  the  SCC.  B.  The  collateral  damage  model  proposes  the  replication  fork  stalls  when  it  encounters  SCCs  and  the  double  strand  break  generates  from  replication  fork  cleavage  by  a  recombination  endonuclease  (purple  arrow).

A  

B  

Page 127: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

111  

4.1.2 Importance of DNA Extraction Conditions

As  mentioned  above,  the  latent  break  within  the  gyrase-­‐‑DNA  complex  can  be  

resealed  under  certain  conditions.  Many  of  the  studies  used  to  work  out  the  conditions  

of  DSB  formation  and  quinolone  cytotoxicity  extracted  DNA  in  the  presence  of  the  

protein  denaturant,  SDS.  This  complicates  results,  as  it  reveals  not  only  breaks  that  are  

due  to  SCC  formation,  but  also  breaks  within  the  stabilized  cleavage  complex.    

Recent  studies  in  our  lab  have  used  distinct  DNA  extraction  conditions  to  

distinguish  between  reversible  breaks  within  the  stabilized  cleavage  complex  and  those  

that  are  irreversible  and  likely  due  to  downstream  events  following  cleavage  complex  

formation.  Cleavage  conditions  treat  with  SDS,  which  denatures  gyrase  and  abolishes  its  

ability  to  reseal  the  covalently  attached  DNA  ends.  Resealing  conditions  include  heat  or  

EDTA  treatment  prior  to  SDS,  allowing  the  ends  to  reseal  before  gyrase  is  denatured.    

Following  norfloxacin  treatment,  Pohlhaus  et  al.  extracted  plasmid  DNA  under  

the  cleavage  and  resealing  conditions  to  reveal  the  accumulation  of  SCCs  and  their  effect  

on  replication  fork  progression  [118].  This  intracellular  system  monitored  cleavage  

complex  formation  on  plasmid  pBR322  that  was  modified  to  include  a  strong  gyrase  

binding  site.  Southern  blot  analysis  of  a  wildtype  strain  in  the  absence  of  norfloxacin  

showed  no  plasmid  cleavage.  Norfloxacin  treatment  and  DNA  extraction  under  

cleavage  conditions  revealed  the  latent  breaks  within  cleavage  complexes.  However,  

resealing  conditions  showed  no  plasmid  cleavage,  confirming  that  the  breaks  observed  

Page 128: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

112  

were  from  the  latent  break  within  the  cleavage  complex  (Figure  398,  Lanes  1-­‐‑3).  The  

same  conditions  were  used  to  isolate  DNA  from  strains  parCR  and  gyrAR.  Norfloxacin  

targets  gyrase  and  the  gyrAR  mutation  showed  no  plasmid  cleavage,  while  cleavage  of  

the  plasmid  was  evident  in  the  parCR  strain  (Figure  39,  Lanes  4-­‐‑6  &  7-­‐‑9).  This  study  also  

found  that  norfloxacin  treatment  led  to  SCCs  and  the  blockage  of  replication  forks  by  

analyzing  plasmid  DNA  extracted  under  cleavage  and  resealing  conditions  using  two-­‐‑

dimensional  gel  electrophoresis  [118].    

Using  this  system  as  a  foundation,  particularly  the  well-­‐‑defined  resealing  and  

cleavage  conditions,  I  sought  to  provide  direct  evidence  of  chromosomal  fragmentation  

due  to  quinolone  treatment.  The  results  presented  in  this  chapter  are  preliminary  and  

should  be  treated  as  such.  However,  they  demonstrate  the  potential  of  this  system  to  

distinguish  between  the  different  types  of  double  strand  ends  resulting  from  quinolone  

treatment,  those  within  the  latent  break  and  those  generated  as  a  result  of  irreversible  

processing  mechanisms.  

 

Page 129: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

113  

 

 Figure  39.  Cleavage  conditions  reveal  the  latent  break  within  the  stabilized  cleavage  

complex  

Southern  blot  analysis  reveals  that  norfloxacin-­‐‑induced  cleavage  complex  breaks  are  visible  under  cleavage  conditions  (C).  However,  resealing  conditions  (R)  allow  the  latent  break  within  the  complex  to  be  resealed.  Lanes  1-­‐‑3  are  JH39  WT,  Lanes  4-­‐‑6  are  parCR,  Lanes  7-­‐‑9  are  gyrAR  cells.  Norfloxacin  concentration  was  1  µμg/mL.  Figure  from  [118]

Page 130: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

114  

4.2 Experimental Procedure

This  method  uses  field  inversion  gel  electrophoresis  (FIGE)  to  visualize  genomic  

DNA.  Field  inversion  gel  electrophoresis  is  a  method  of  electrophoresis  that  periodically  

switches  the  current  in  opposing  directions  (180°),  creating  alternating  electric  fields  

between  the  two  electrodes.  This  forces  the  DNA  molecules  to  reorient  with  each  switch,  

and  eventually  separates  small  and  large  fragments  of  DNA.  Unlike  conventional  

electrophoresis,  FIGE  can  resolve  much  larger  fragments  ranging  from  100  basepair  to  

250  kilobases  [133].  Chromosomal  DNA  was  isolated  in  agarose  plugs.  Lysing  cells  

within  the  plugs  prevents  mechanical  shearing  of  the  DNA  and  minimizes  breaks  due  to  

manipulation.  Cell  lysis  under  different  conditions,  cleavage  or  resealing,  allows  us  to  

distinguish  between  reversible,  latent  breaks  and  irreversible  double  strand  breaks.  

4.2.1 Materials and Bacterial Strains

Nalidixic  acid  and  norfloxacin  were  purchased  from  Sigma-­‐‑Aldrich.  Clean  Cut  

agarose  and  agarose  plug  molds  were  from  the  BioRad  CHEF  Genomic  DNA  Plug  Kit.  

Markers  used  for  field  inversion  gel  electrophoresis  experiments  include  Lambda  

Ladder  Pulsed-­‐‑Field  Gel  Marker,  Low  Range  Pulsed-­‐‑Field  Gel  Marker,  Mid  Range  I  

Pulsed-­‐‑Field  Gel  Marker  and  Yeast  Chromosome  Pulsed-­‐‑Field  Gel  Marker.  All  markers  

were  purchased  from  New  England  Biolabs.  The  strains  used  in  this  assay  are  

derivatives  of  JH39  [F-­‐‑  sfiA11,  thr-­‐‑1,  leu-­‐‑6,  hisG4,  argE3,  ilv(Ts),  galK2,  srl(?),  rpsL31,  

lacΔU169,  dinD1::MudI(Apr  lac)].    

Page 131: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

115  

4.2.2 Agarose Plugs

Overnight  cultures  of  appropriate  strains  were  diluted  1:100  and  grown  shaking  

at  37°C  until  OD595  reached  0.5  –  0.7  as  measured  by  the  biophotometer.  Once  cells  

reached  the  appropriate  optical  density,  aliquots  of  each  culture  were  incubated  with  the  

selected  drug  at  indicated  concentrations  and  lengths  of  time.  Two-­‐‑milliliter  aliquots  of  

cultures  were  pelleted  in  1.5-­‐‑ml  tubes  using  the  tabletop  microcentrifuge  and  suspended  

with  spent  LB  and  2  mM  EDTA  to  a  final  volume  of  200  µμL.    Suspended  cells  were  

mixed  with  200  µμl  of  molten  2%  Clean  Cut  Agarose  and  immediately  pipetted  into  plug  

molds.  Approximately  500  µμL  of  culture  was  used  for  each  well.  Plugs  set  at  4°C  for  10  

minutes  to  solidify  agar.    

4.2.3 DNA Preparations

To  lyse  cells  and  extract  DNA,  plugs  were  incubated  in  either  Resealing  Lysis  

Buffer  (50  mM  Tris  HCl  pH  7.8,  10  mM  EDTA,  100  mM  NaCl,  1%  Triton,  1.8  µμg/mL  

Lysozyme)  or  Cleavage  Lysis  Buffer  (100  mM  NaCL,  10  mM  EDTA,  50  mM  Tris  HCl  pH  

7.8  and  0.2%  SDS).  Plugs  in  Resealing  Lysis  Buffer  were  incubated  for  two  hours  at  65°C,  

while  plugs  in  Cleavage  Lysis  Buffer  were  incubated  at  37°C.  Following  incubation,  

Proteinase  K  was  added  to  all  plugs  at  1  mg/mL  and  0.2%  SDS  was  added  to  resealing  

plugs.  Plugs  were  incubated  overnight  in  buffers  at  37°C,  then  washed  4  times  for  1  hour  

at  room  temperature  in  1X  Wash  Buffer  (200  mM  Tris,  pH  8.0  and  50  mM  EDTA)  from  

the  BioRad  CHEF  Genomic  Kit  (1  mL/plug).  During  the  second  wash,  PMSF  was  added  

Page 132: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

116  

at  1  mM  to  the  Wash  Buffer.  Following  washes,  the  plugs  were  equilibrated  on  ice  for  

15-­‐‑30  minutes  in  0.5X  TBE.  

4.2.5 Field Inversion Gel Electrophoresis

Plugs  were  loaded  into  the  wells  of  a  1%  agarose  gel  and  sealed  with  molten  1%  

agarose  to  prevent  dislodging  from  wells  during  the  runs.  The  gel  was  run  at  120  volts  

for  24  hours.  The  MJ  Research  programmable  power  inverter,  PPI-­‐‑200,  was  used  to  

switch  the  electrical  field  during  electrophoresis  using  the  pre-­‐‑loaded  setting,  Program  

5,  with  recirculating  TBE  buffer  according  to  manufacturer’s  instructions.  Program  5  is  

optimized  for  50  –  600kb  resolution  of  DNA  with  the  following  parameters:  A,  initial  

reverse  time  –  0.1  minutes;  B,  reverse  increment  –  0.01  minutes;  C,  initial  forward  time  –  

0.3  minutes;  D,  forward  increment  –  0.03  minutes;  E,  steps  –  45;  F,  reverse  increment  –  

0.01  minutes,  G,  forward  increment  –  0.03  minutes.  This  allows  resolution  of  DNA  

between  50-­‐‑600kb.  

4.3 Results

This  method  was  created  with  the  goal  of  providing  physical  analysis  of  

quinolone-­‐‑induced  chromosome  fragmentation.  FIGE  allows  large  fragments  of  DNA  to  

migrate  into  the  gel  matrix.  Combining  this  method  with  varying  DNA  extraction  

conditions  allows  visual  analysis  of  chromosomal  breaks.  Cleavage  conditions  use  

protein  denaturants  and  reveal  both  reversible  and  irreversible  breaks.  The  program  

selected  for  the  gels  resolves  DNA  ranging  from  50-­‐‑600  kb.    

Page 133: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

117  

An  often  overlooked  feature  of  FIGE  is  the  trapping  of  large,  circular  and/or  

branched  DNA  [134].  This  means  branched  or  circular  DNA,  even  if  it  is  the  same  size  as  

linear  DNA  in  the  gel,  may  not  enter  the  gel  matrix.  Another  caveat  is  band  inversion,  

which  makes  it  difficult  to  distinguish  between  large  DNA  fragments  at  the  region  of  

limiting  mobility  (see  Discussion)  [135,  136].  Large  fragments  of  DNA  enter  the  gel  but  

their  migration  is  limited.  This  makes  it  possible  for  linear  fragments  as  large  as  1  

megabase  to  migrate  with  those  that  are  in  the  600kb  range.  Assuming  the  agarose  plugs  

yield  completely  intact  chromosomes  and  the  double  strand  breaks  are  at  equal  

distances,  chromosomal  DNA  could  have  only  2  breaks  and  still  migrate  into  the  gel  

under  the  inversion  conditions  used.  

4.3.1 Norfloxacin-Treatment Causes Irreversible Breaks

 In  our  test  of  untreated  wildtype  strains,  most  of  the  DNA  remains  in  the  wells  

under  both  cleavage  and  resealing  conditions,  indicating  that  most  of  the  DNA  is  intact  

(Figure  40).  There  is  a  small  level  of  smearing  in  the  cleavage  conditions  sample  and  the  

breaks  in  the  untreated  resealing  lanes  most  likely  are  due  to  endogenous  levels  of  

chromosomal  damage  as  reported  in  other  studies  [137].    

Norfloxacin  treatment  of  wildtype  cells  with  DNA  extracted  under  cleavage  

conditions  reveals  fragments  that  migrate  close  to  48  kb  near  the  midpoint  of  the  gel  

(Figure  40).  This  would  require  around  95  –  100  breaks  for  DNA  fragments  of  that  size.  

Cleavage  conditions  reveal  both  latent  breaks  and  any  overt  breaks  within  the  sample.  

Page 134: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

118  

 

 

 

Figure  40.  Norfloxacin-­‐‑treated  Wildtype  DNA  

Genomic  DNA  from  WT  JH39  isolated  under  cleavage  conditions  reveals  breaks  in  both  untreated  and  norfloxacin-­‐‑treated  cells.  Under  resealing  conditions  with  drug  treatment,  the  DNA  smear  migrates  higher,  indicating  that  breaks  are  resealed.  Even  under  resealing  conditions  without  drug  treatment,  there  is  endogenous  damage  that  migrates  into  the  gel.  Norfloxacin  is  at  1  µμg/mL.  

Norfloxacin                      -­‐‑              -­‐‑            +          +                                    -­‐‑          +          +     Resealing     Cleavage  

-­‐‑  48.5  kb

-­‐‑  97.0  kb  

-­‐‑  145.5  kb

-­‐‑  194.0  kb

-­‐‑  242.5  kb

-­‐‑  291.0  kb

Page 135: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

119  

DNA  extracted  under  resealing  conditions  from  norfloxacin-­‐‑treated  cells  migrates  at  a  

range  over  291  kb,  which  corresponds  to  roughly  15  breaks.  The  norfloxacin-­‐‑treated  

resealing  samples  show  far  more  breaks  than  the  untreated  samples  on  the  same  gel.  The  

FIGE  of  norfloxacin-­‐‑treated  DNA  indicates  that  norfloxacin  leads  to  irreversible  breaks.  

4.3.2 Nalidixic Acid-Treatment Reveals Fewer Irreversible Breaks

Norfloxacin  is  one  of  the  more  potent  fluoroquinolones  and  we  wanted  to  test  

this  method  using  a  first  generation  quinolone,  nalidixic  acid.  Figure  41  shows  the  

chromosomal  fragmentation  of  samples  treated  over  a  range  of  nalidixic  acid  

concentrations  (30  -­‐‑  90  µμg/ml)  and  a  norfloxacin-­‐‑treated  sample  for  comparison.  DNA  

smears  are  visible  ranging  from  48.5  to  600  kb  in  the  nalidixic  acid-­‐‑treated  cleavage  

condition  samples.  However,  under  resealing  conditions,  DNA  in  the  nalidixic  acid  

samples  migrates  much  higher  and  looks  similar  to  the  untreated  control.  The  

norfloxacin-­‐‑treated  resealing  sample  shows  chromosome  fragmentation  in  comparison  

to  the  nalidixic  acid  samples.  The  norfloxacin-­‐‑treated  wildtype  samples  are  more  intense  

than  the  nalidixic  acid  samples  and  migrate  at  a  smaller  size  (Figure  41).  It  is  possible  

that  nalidixic  acid  breaks  arise  directly  from  the  latent  break  within  the  cleavage  

complex.  It  seems  apparent  that  nalidixic  acid  causes  fewer  irreversible  breaks  than  

norfloxacin,  which  results  in  larger  chromosomal  fragments.  

Page 136: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

120  

 

Figure  41.  Effects  of  Increasing  Nalidixic  Acid  on  Break  Formation  

JH39  wildtype  cells  treated  with  increasing  concentrations  of  nalidixic  acid  from  0  –  90  µμg/mL  or  norfloxacin  (N)  at  1  µμg/mL.  The  ladders  are  λ  monomer  @  48.5kb  (M)  or  λ  concatemers  which  are  successively  larger  (C).  Field  inversion  gel  electrophoresis  reveals  chromosomal  breaks  at  various  concentrations  of  nalidixic  acid  treated  samples  under  cleavage  conditions  and  the  majority  of  breaks  are  from  cleavage  complexes.  

Nalidixic  Acid              0        30    60    90      N    0      30    60      90    N  

Resealing   |   Cleavage  

C        M   M      C  

48.5kb-­‐‑  

97.0  kb-­‐‑  

145.5  kb-­‐‑  

194.0  kb-­‐‑  

242.5  kb-­‐‑  

291.0  kb-­‐‑  

339.5  kb-­‐‑  

Page 137: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

121  

4.3.3. ΔdnaQ, ΔrecB, and gyrAR Cleavage & Resealing Conditions

Our  findings  that  dnaQ  mutants  fail  to  induce  SOS  after  nalidixic  acid  treatment,  

led  us  to  test  a  ΔdnaQ  strain  with  this  method.  Samples  were  treated  with  nalidixic  acid,  

norfloxacin  or  no  drug.  The  untreated  ΔdnaQ  samples  look  similar  to  wildtype,  but  the  

extent  of  chromosome  fragmentation  in  the  nalidixic  acid-­‐‑treated  samples  of  ΔdnaQ  is  

difficult  to  interpret  (Figure  42).  In  the  ΔdnaQ  norfloxacin-­‐‑treated  cleavage  sample,  the  

fragments  seem  to  migrate  close  to  48.5  kb  and  under.  In  comparison,  the  wildtype  

norfloxacin-­‐‑treated  cleavage  sample  migrates  between  48.5  –  194  kb.  Also  the  ΔdnaQ  

DNA  extracted  under  resealing  conditions  migrates  over  48.5  kb  to  the  upper  resolution  

limit  (~600  kb).  In  the  wildtype  DNA  sample  under  the  same  conditions,  the  majority  of  

the  DNA  migrates  within  the  region  above  375  kb  to  the  higher,  600  kb  resolution  limit  

(Figure  42).  The  nalidixic  acid  ΔdnaQ  samples  should  be  repeated,  but  it  appears  that  

norfloxacin  treatment  may  cause  more  breaks  in  ΔdnaQ  than  in  the  wildtype  strain.  It  is  

interesting  that  even  the  apparent  number  of  cleavage  complexes  (cleavage  conditions)  

seems  higher  in  the  dnaQ  mutant  than  in  the  wildtype  sample,  suggesting  the  existence  

of  more  gyrase  cleavage  complexes.    

Page 138: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

122  

 

Figure  42.  Break  Formation  in  Wildtype  and  ΔDnaQ  Treated  with  Nalidixic  Acid  &  Norfloxacin    

A  dnaQ  mutant  appears  to  be  more  sensitive  to  norfloxacin  treatment  than  the  wildtype  strain.  Y  indicates  yeast  chromosome  marker,  which  ranges  from  225-­‐‑1900kb.  λC  indicates  lambda  concatemer  marker,  which  starts  at  48.5kb.  Nalidixic  acid:  90  µμg/mL;  norfloxacin:  1  µμg/mL.    

C      R        C        R        C        R                    C        R        C        R        C      R  No  Drug  |  Nal  Acid  |    Norf   No  Drug  |  Nal  Acid  |    Norf  

Wildtype       ΔDnaQ  

Y   λC  

48.5  kb-­‐‑  

242.5  kb-­‐‑  

97.0  kb-­‐‑  

145.5  kb-­‐‑  

194.0  kb-­‐‑  

375  kb-­‐‑  

291.0  kb-­‐‑  

Page 139: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

123  

 

Figure  43.  Nalidixic  Acid  &  Norfloxacin-­‐‑treated  ΔRecB  and  gyrAR  under  cleavage  and  resealing  conditions.  

The  ΔRecB  mutant  shows  migration  into  the  gel  even  without  drug  treatment,  which  is  consistent  with  the  role  of  RecBCD  in  the  processing  of  double  strand  ends.  The  nalidixic  acid-­‐‑treated  gyrAR  strain  shows  some  protection  against  chromosome  fragmentation  under  cleavage  conditions  in  comparison  to  wildtype  (Figure  41).  Y  indicates  yeast  chromosome  marker,  which  ranges  from  225-­‐‑1900kb.  λC  indicates  lambda  concatemer  marker,  which  starts  at  48.5kb.

C      R        C        R        C        R                    C          R      C      R        C        R  No  Drug  |  Nal  Acid  |    Norf   No  Drug  |  Nal  Acid  |    Norf  

  ΔRecB         gyrAR  

Y   λC  

48.5  kb-­‐‑  

242.5  kb-­‐‑  

97.0  kb-­‐‑  

145.5  kb-­‐‑  

194.0  kb-­‐‑  

375  kb-­‐‑  

291.0  kb-­‐‑  

Page 140: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

124  

RecBC  mutants  are  unable  to  repair  double  strand  breaks,  and  our  results  with  

ΔrecB  aligns  with  its  function.  In  Figure  43,  the  untreated  ΔrecB  samples  show  

chromosome  fragmentation  that  migrates  around  375  –  600  kb.  This  is  striking  result  in  

comparison  to  the  wildtype  strain,  which  does  not  reveal  DNA  fragmentation  when  

untreated  (Figure  42).  Furthermore,  the  fragmentation  appears  in  both  cleavage  and  

resealing  conditions,  indicating  these  breaks  are  not  related  to  the  latent  break  within  the  

gyrase-­‐‑DNA  complex.  They  are  most  likely  from  DSBs  caused  by  other  processes.  The  

majority  of  the  ΔrecB  nalidixic  acid-­‐‑treated  DNA  extracted  under  cleavage  conditions  

migrates  closer  to  48.5  kb.  Wildtype  DNA  under  the  same  condition  migrates  from  48.5  -­‐‑  

600  kb  with  a  more  intense  smear  ranging  from  375  –  600  kb  (Figure  41).  Chromosome  

fragmentation  in  norfloxacin-­‐‑treated  ΔrecB  samples  looks  similar  to  wildtype  except  

there  is  more  DNA  migrating  closer  to  the  lower  ranges  (48.5  –  145  kb).  

GyrAR  confers  nalidixic  acid  resistance  and  our  results  show  a  higher  migrating,  

less  intense  DNA  smear  from  the  mutated  gyrase  sample  under  cleavage  conditions  

than  in  wildtype.  Norfloxacin-­‐‑treated  gyrAR  DNA  extracted  under  cleavage  conditions  

reveals  DNA  ranging  from  48.5  –  145.5  kb  and  291  –  600  kb  (Figure  42).  In  the  wildtype  

strain,  norfloxacin-­‐‑treated  DNA  extracted  under  cleavage  conditions  all  migrates  over  

the  range  of  48.5  –  242.5  kb.  Neither  gyrAR    or  the  wildtype  strain  show  chromosome  

fragmentation  in  nalidixic  acid-­‐‑treated  resealing  samples.  Under  norfloxacin-­‐‑treatment  

and  resealing  conditions,  both  wildtype  and  gyrAR  DNA  migrates  at  375  –  600  kb,  but  the  

Page 141: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

125  

amount  of  DNA  (intensity)  is  much  greater  in  wildtype.  These  results  indicate  that  gyrAR  

protects  against  SCCs  induced  by  nalidixic  acid  and  norfloxacin  as  evidence  by  fewer  

breaks  in  the  cleavage  condition  samples.  The  degree  of  gyrAR  protection  from  

norfloxacin-­‐‑induced  irreversible  breaks  (resealing  conditions)  is  not  immediately  clear  

and  should  be  followed  up  with  additional  experiments.  

Page 142: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

126  

4.4 Discussion

Our  results  confirm  multiple  earlier  findings  in  the  literature,  validating  our  

system  of  physically  analyzing  breaks.  Norfloxacin,  a  fluoroquinolone,  showed  a  

migration  pattern  indicative  of  much  smaller  DNA  fragments,  indicating  a  greater  

degree  of  breaks  than  nalidixic  acid.  While  both  norfloxacin  and  nalidixic  acid-­‐‑treated  

samples  revealed  breaks  in  cleavage  conditions,  the  smear  in  nalidixic  acid  resealing  

samples  was  undetectable,  unlike  the  smear  of  norfloxacin-­‐‑treated  samples.  This  is  

consistent  with  evidence  that  the  fluoroquinolones  are  more  potent  than  the  first-­‐‑

generation  quinolones  like  nalidixic  acid  [43,  138].    

The  results  of  nalidixic  acid-­‐‑treated  ΔdnaQ  were  inconclusive  and  will  require  

more  tests.  Norfloxacin-­‐‑treated  ΔdnaQ  appeared  to  have  smaller  fragments  than  

wildtype  under  cleavage  and  resealing.  This  was  an  interesting  finding  and  warrants  

further  study.  The  recB  mutant  results  further  validated  the  system  and  were  in  

agreement  with  the  known  role  of  RecB  in  double  strand  end  processing.  Additionally,  

results  of  the  nalidixic  acid  and  norfloxacin-­‐‑treated  gyrAR  strain  aligned  with  findings  

that  this  allele  offers  more  resistance  against  nalidixic  acid  than  it  does  against  

norfloxacin  and  other  fluoroquinolones,  which  require  both  gyrAR  and  additional  

mutations  in  parC  for  high  levels  of  resistance  [139,  140].    

While  this  method  is  still  in  its  preliminary  stages,  it  shows  a  lot  of  promise  for  

understanding  the  underlying  mechanisms  of  quinolone  treatment  as  it  relates  to  

Page 143: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

127  

chromosome  fragmentation.  This  system  provides  a  useful  tool  to  analyze  latent  breaks  

within  the  cleavage  complex  in  comparison  to  irreversible,  double  strand  breaks  that  

arise  from  downstream  processing.  While  killing  and  survival  assays  measure  the  final  

endpoint  of  cell  death,  this  assay  allows  direct  analysis  of  breaks.  Beyond  visualization  

of  the  broken  DNA,  it  is  possible  to  see  the  extent  of  breakage  in  comparison  to  

untreated  samples.  Coupling  this  assay  of  chromosomal  fragmentation  with  a  measure  

of  survival  would  provide  a  powerful  method  to  assess  various  quinolone  methods  of  

cytotoxicity.  

This  system  would  be  particularly  useful  in  addressing  the  various  proposed  

models  of  quinolone  double  strand  break  generation  that  were  outlined  earlier  in  the  

chapter.  We  could  test  whether  protein  synthesis  is  required  for  downstream,  

irreversible  breaks  by  treating  with  chloramphenicol.  We  could  test  whether  or  not  

replication  affects  the  migration  of  breaks  under  resealing  conditions  using  DnaB  

mutants  or  treating  with  hydroxyurea  or  azidothymidine  (AZT)  [141,  142].  Studies  of  a  

DnaB-­‐‑temperature  sensitive  mutant  revealed  that  the  active  replication  is  not  necessary  

for  nalidixic  acid  or  ciprofloxacin-­‐‑mediated  cytotoxicity  [120].  This  study  was  based  on  

survival  of  cells  following  treatment,  and  it  is  possible  that  there  are  multiple  methods  of  

quinolone-­‐‑cytotoxicity,  even  with  only  one  drug.  It  is  possible  that  the  extent  of  

chromosome  fragmentation  is  much  higher  when  there  is  ongoing  replication,  but  this  

Page 144: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

128  

level  of  damage  is  not  sufficient  to  cause  cell  death.  Our  assay  would  allow  direct  testing  

of  this  hypothesis.  

As  stated  in  the  Introduction,  this  method  is  still  in  the  very  early  stages  and  the  

results  should  be  treated  as  such.  The  priority  of  future  directions  for  this  project  is  to  

optimize  the  resolution  and  reproducibility.  In  many  of  the  samples,  there  is  a  large  

region  of  larger  DNA  that  almost  looks  like  two  distinct  bands.  This  is  due  to  band  

inversion  at  the  limiting  mobility  of  the  gel  [136].  To  examine  broader  ranges  of  DNA  

with  FIGE,  you  sacrifice  resolution.  In  these  gels,  there  is  poor  resolution  of  products  

over  ~330kb  and  they  all  migrate  together  in  one  band.  A  clear  example  of  this  is  with  

the  yeast  chromosome  marker,  which  ranges  from  225  –  1900  kb  (Figures  42  &  43).  

However,  everything  over  375  kb  runs  together  in  a  smear;  the  same  is  true  with  many  

of  the  samples.  This  can  be  addressed  by  changing  the  switching  parameters  to  resolve  a  

narrow  range  of  larger  DNA  sizes.    

Even  though  the  voltage  and  run  time  remain  constant,  the  extent  of  migration  

differs  between  gels.  This  is  most  likely  due  to  temperature  differences  with  each  run.  

The  gels  shown  were  run  at  room  temperature  with  a  peristaltic  pump  in  0.5X  TBE.  

Using  TBE  +  glycine  and  running  the  gels  in  the  cold  can  help  prevent  overheating  and  

give  us  better  reproducibility  [133].  Once  the  technical  issues  are  resolved  and  we  have  

an  efficient  temperature-­‐‑regulated  set  up,  the  possibilities  of  testing  with  various  

mutants  (xseA,  uvrD,  sbcCD)  and  physiological  conditions  are  vast.  

Page 145: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

129  

5. Conclusions The  work  presented  above  aimed  to  understand  the  bacterial  response  to  

quinolone.  More  specifically,  I  proposed  that  the  ε  subunit  of  DNA  polymerase  III  has  

an  additional  role  beyond  proofreading.  To  test  this  hypothesis  I  created  a  library  of  

dnaQ  mutants  to  screen  for  separation  of  function.  Chapter  2  outlined  initial  plasmid-­‐‑

based  screen  of  the  dnaQ  insertion  mutant  library  targeting  mutants  that  uncoupled  the  

mutator  phenotype  from  the  constitutive  SOS  phenotype.  I  saw  conflicting  results  

between  mutants  on  the  plasmid  compared  to  chromosomal  mutants,  arguing  that  

inappropriate  expression  from  the  plasmid  was  influencing  the  phenotypes.  I  decided  to  

abandon  this  method  of  isolation  and  restructure  my  approach.    

Chapter  3  detailed  our  second  attempt  to  isolate  separation  of  function  mutants,  

which  was  less  biased  and  allowed  me  to  capture  a  collection  of  mutants  based  solely  on  

the  location  of  the  insertions.  Based  on  the  results  of  Chapter  2,  I  was  able  to  optimize  

the  cloning  and  screening  strategy,  eliminate  multiple  strain  switches  and  eliminate  

phenotypic  scoring  of  the  mutants  at  the  plasmid  level.  The  less  biased  approach  was  

fruitful  and  I  isolated  multiple  types  of  separation  of  function  mutants.  These  results  

showed  that  proofreading  can  be  uncoupled  from  the  SOS  phenotypes  (constitutive  and  

defective  induction  following  nalidixic  acid).  I  also  uncoupled  the  constitutive  SOS  

phenotype  from  the  defective  SOS  following  nalidixic  acid  phenotype,  indicating  that  

these  two  can  be  unlinked  as  well.    

Page 146: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

130  

My  results  support  the  hypothesis  that  ε  has  a  role  outside  of  proofreading.  I  

identified  insertion  mutants  that  uncouple  the  various  phenotypes  of  a  dnaQ  mutant.  I  

found  that  most  of  the  high  mutators  led  to  constitutive  SOS,  and  found  many  moderate  

mutators  with  low  basal  SOS  levels.  Based  on  this,  I  proposed  a  model  whereby  at  least  

some  of  the  constitutive  SOS  expression  is  induced  by  extreme  mutator  activity  and  

incorporation  of  wrong  bases.  Additionally,  proofreading  partially  correlated  with  

nalidixic  acid-­‐‑induced  SOS.    

In  Chapter  4,  I  presented  preliminary  results  of  a  method  that  allows  direct  

analysis  of  chromosomal  breaks,  distinguishing  between  reversible  breaks  within  the  

cleavage  complex  and  irreversible  double  strand  breaks.  While  the  system  is  in  its  early  

stages,  the  results  obtained  so  far  agree  with  previously  published  reports.  This  further  

validates  the  usefulness  of  the  system,  and  I  think  it  could  serve  as  a  useful  tool  to  

directly  test  proposed  models  of  quinolone-­‐‑induced  double  strand  break  generation.  

The  latent  break  within  the  topoisomerase-­‐‑DNA-­‐‑quinolone  complex  complicates  

the  study  of  quinolone-­‐‑induced  DNA  damage  at  the  chromosomal  level.  These  breaks  

are  reversible  with  EDTA,  high  heat,  or  removal  of  the  drug;  cell  death  is  not  reversible.  

The  method  I  proposed  in  Chapter  4  allows  direct  visualization  of  chromosomal  breaks  

based  on  migration  from  the  wells  into  the  body  of  the  agarose  gel.  This  distinguishes  

between  the  latent  break  within  the  cleavage  complex,  which  is  reversible,  and  those  

breaks  that  may  arise  from  processing  of  stabilized  cleavage  complexes,  which  are  not  

Page 147: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

131  

reversible  and  most  likely  the  cause  of  quinolone  cytotoxicity.  Preliminary  results  

suggested  that  norfloxacin-­‐‑treated  ΔdnaQ  has  more  cleavage  complexes  than  wildtype  

DNA  extracted  under  cleavage  conditions.  This  was  an  interesting  finding  and  should  

be  pursued  to  determine  what  role  additional  cleavage  complexes  play  in  a  ΔdnaQ  

mutant.  

5.1 Interpretations & Future Directions

The  various  separation  of  function  mutants  contribute  a  set  of  mutationally  

stable  proteins  to  study  the  role  of  ε.  I  also  identified  a  “hyper-­‐‑inducible”  mutant  that  

shows  overwhelming  SOS  induction  in  response  to  nalidixic  acid.  We  also  identified  

proofreading  phenotypes  arising  from  loss  of  binding  to  α,  weak  binding  to  β  and  

mutations  in  the  N-­‐‑terminal  domain  that  affect  exonuclease  activity.    

A  recent  characterization  of  the  α-­‐‑β-­‐‑ε  interaction  proposed  that  ε  binding  to  the  

clamp  enhances  α  binding  and  prevents  clamp  access  to  translesion  polymerases  (see  

Chapter  3  Discussion)  [105].  Putting  the  results  of  my  clamp  mutants  into  the  tripartite  

model  proposed  by  Toste  Rêgo  et  al.  (2013),  I  provided  a  model  for  basal  SOS  induction  

in  which  the  weak  clamp  binder  allows  Pol  IV  access  to  the  clamp  and  synthesis  past  

lesions;  this  results  in  low  basal  SOS  levels.  On  the  other  hand,  the  strong  binder  

prevents  Pol  IV  access  to  the  clamp  and  when  encountering  endogenous  damage,  it  

cannot  synthesize  past  it  and  leads  to  prolonged  stalled  forks.  This  accounts  for  the  

constitutive  SOS  levels  observed  in  the  strong  clamp-­‐‑binding  mutant.  

Page 148: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

132  

It  is  unclear  how  the  tripartite  model  relates  to  SOS  induction  following  nalidixic  

acid.  Our  results  indicate  that  the  ε-­‐‑β  clamp  interaction  is  required  for  nalidixic  acid-­‐‑

induced  SOS.  Conversely,  the  weak  clamp  binder  leads  to  defective  SOS  induction  

following  nalidixic  acid.  We  know  that  nalidixic  acid  leads  to  SCCs  and  that  RecBC  is  

required  for  the  subsequent  SOS  induction,  so  the  generation  of  DSBs  is  important  for  

this  mechanism.  Based  on  this  knowledge  and  our  results,  we  propose  two  possible  

models  for  the  nalidixic  acid  SOS  phenotypes  in  the  context  of  the  tripartite  model.  

One  model  is  based  on  replication  run  off  as  described  in  Chapter  4  (Figure  38A).  

We  know  that  SCCs  cause  replication  fork  stalling,  and  stalling  leads  to  DnaB  helicase  

unloading.  We  propose  that  the  tripartite  complex  usually  cannot  proceed  but  

occasionally  may  move  forward,  encountering  the  SCC.  The  broken  arm  of  the  

replication  fork  and  destabilized  gyrase-­‐‑cleavage  complex  creates  a  double  stranded  end  

for  RecBC  to  create  ssDNA  and  load  RecA,  which  then  generates  the  signal  for  SOS  

induction.  This  is  the  replication  run-­‐‑off  model.  

As  outlined  above,  the  weak  clamp  mutant  easily  allows  translesion  polymerases  

to  associate  with  the  tripartite  complex.  We  propose  that  this  weaker,  less-­‐‑processive  

tripartite  complex  would  not  have  the  ability  to  push  the  fork  forward  and  destabilize  

the  gyrase-­‐‑cleavage  complex.  This  means  the  weak  clamp  binder  does  not  create  a  break  

or  double  strand  end  for  RecBC,  and  accounts  for  the  nalidixic  acid-­‐‑induced  SOS  defect  

observed  in  Q182A.  

Page 149: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

133  

With  the  strong  clamp-­‐‑binding  mutant,  the  tripartite  complex  is  tightly  

associated.  According  to  this  model,  this  mutant  pushes  forward,  and  more  often  than  in  

wildtype,  it  destabilizes  the  SCCs  and  generates  double-­‐‑strand  ends  for  RecBC  

processing.  Biochemical  characterization  shows  QLSLPL  is  more  processive  than  

wildtype  and  this  would  explain  the  higher-­‐‑than-­‐‑wildtype  SOS  levels  in  response  to  

nalidixic  acid  [104].    

It  may  be  possible  to  propose  another  model,  the  collateral  damage  model,  with  

our  data  (Figure  38B).  In  this  model,  the  tightly  associated  tripartite  complex  causes  

stalled  forks  that  are  cleaved  by  an  unidentified  endonuclease.  The  cleavage  generates  a  

free  end  for  RecBC  processing  and  subsequent  SOS  induction.  However,  this  model  

would  require  many  assumptions,  including  that  only  the  strong  clamp  binder  would  

lead  to  double  strand  breaks.  This  model  is  not  as  clear  and  does  not  align  with  our  data  

as  well  as  the  replication  run-­‐‑off  model.    

Additionally,  we  propose  that  ε  must  be  present  at  the  complex  for  SOS  

induction  after  nalidixic  acid.  Using  these  models,  we  are  able  to  address  not  only  the  

clamp-­‐‑binding  mutants,  but  the  insertion  mutants  as  well  –  those  that  present  separation  

of  function  and  those  that  do  not.  

In  our  model,  an  ε  knockout  always  presents  with  a  nalidixic  acid-­‐‑induced  SOS  

defect.  In  several  of  the  SOS  defective  insertion  mutants,  the  high  mutators  are  

essentially  knockouts  due  to  protein  instability,  misfolding,  or  inability  to  associate  with  

Page 150: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

134  

α  at  the  tripartite  complex.  This  explains  the  coupled  constitutive  SOS  phenotype  (due  

to  overwhelming  mutator  activity)  and  the  nalidixic  acid  SOS  defect  (unable  to  create  a  

processive  tripartite  complex  that  causes  SCC  disassembly  or  SOS  induction).  This  

would  explain  the  phenotypes  of  mutants  such  as  30IICLN,  79CLNTF,  132V*,  224AHV*,  and  

232LFKQV.    

Additionally,  these  models  can  address  some  of  the  striking  separation  of  

function  mutants.  For  example,  162LFKQL  behaves  just  like  the  strong  clamp  binder;  it  is  

a  low  mutator  (consistent  with  preventing  Pol  IV  access  to  the  clamp),  SOS  constitutive,  

and  able  to  induce  SOS  after  nalidixic  acid  at  levels  higher  than  wildtype.  Mutant  

76CLNNK  has  moderate  mutator  activity  that  is  likely  due  to  the  insertion  in  the  NTD.  

This  mutant  fits  the  model  of  a  weak  β  binder  as  it  has  basal  SOS  levels  lower  than  

wildtype  and  nalidixic  acid-­‐‑induced  SOS  levels  lower  than  wildtype  as  well.  

Biochemical  characterization  of  these  mutants  could  tell  us  if  they  tightly  associate  with  

β  and  prevents  TLS  polymerase  access  to  the  clamp.    

When  thinking  about  the  tripartite  complex,  the  isolation  of  a  dnaE  I846V  

mutation  in  JH39  ΔdnaQ  raises  even  more  questions.  Does  this  mutation  confer  any  

protection  to  this  strain  (similar  to  spq-­‐‑2)?  How  do  the  strong,  weak  and  wildtype  β-­‐‑

clamp  mutants  behave  in  this  background?  Does  this  mutation  limit  or  enhance  ε  

binding  to  β?  How  does  this  affect  the  observed  SOS  phenotypes?  It  would  be  

interesting  to  see  how  various  dnaE  alleles,  such  as  spq-­‐‑2,  affect  the  SOS  profiles  of  

Page 151: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

135  

various  ε  mutants.  This  would  give  more  insight  into  whether  the  SOS  defect  following  

nalidixic  acid  is  due  to  core  stability  or  another  interacting  partner.    

Testing  these  models  with  the  assay  described  in  Chapter  4  could  also  provide  

valuable  information.  It  would  be  interesting  to  see  if  a  strong  tripartite  complex  and  

SOS  induction  directly  correlates  with  DSBs  both  in  our  β  clamp  mutants  and  separation  

of  function  mutants.  Single  or  double  endonuclease  mutants  may  show  significant  

changes  in  DSBs  following  nalidixic  acid  treatment  if  the  collateral  damage  model  is  

true.    

The  proposed  models  could  be  tested  with  additional  biochemical  studies  that  

show  whether  the  insertion  mutants  associate  with  α  and  β.  If  they  do,  what  is  the  extent  

of  the  association?  Moreover,  does  the  degree  of  association  correlate  with  the  SOS  

profiles?  Assays  that  measure  the  ability  of  the  core  to  persist  and  processivity  of  the  

separation  of  function  mutants  would  give  us  a  more  complete  picture  of  what  is  

happening.  These  results  would  give  us  more  information  about  what  is  happening  at  

the  replication  fork,  and  at  the  same  time  they  prompt  another  set  of  questions  about  the  

interaction  of  fork  stability  as  it  relates  to  SOS  induction  and  survival  in  the  presence  of  

quinolones.  Quinolone  treatment  coupled  with  strengthening  the  tripartite  complex  may  

make  the  antibiotics  much  more  potent.  Given  the  prevalence  of  antibiotic  resistance,  

this  is  a  line  of  investigation  that  is  worth  pursuing.  

Page 152: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

136  

5.2 Closing Comments

Overall,  this  work  has  provided  a  small  glimpse  of  what  could  be  happening  at  

the  replisome  in  E.  coli.  My  findings  implicate  the  replisome  in  SOS  induction  and  in  the  

bacterial  response  to  nalidixic  acid.  It  further  confirms  the  importance  of  replisome  

fidelity  and  provides  valuable  insight  into  the  recently  identified  β-­‐‑clamp  motif  within  ε.    

There  are  still  many  avenues  to  be  explored  based  on  the  results  presented.  

Many  questions  can  be  addressed  with  the  insertion  and  β-­‐‑clamp-­‐‑binding  mutants.  Does  

the  ability  or  inability  to  induce  SOS  facilitate  nalidixic  acid  survival?  Is  there  an  

advantage  or  disadvantage  of  rapid  replisome  disassembly?  Does  this  change  

depending  on  the  type  (endogenous  or  exogenous)  or  level  of  damage?  Does  it  differ  

depending  on  the  type  of  quinolone  used?    

The  FIGE  method  combined  with  distinguishing  between  latent,  reversible  and  

irreversible  breaks  offers  a  valuable  method  of  analysis.  We  could  address  the  multiple  

questions  surrounding  the  various  quinolone  cytotoxicity  models.  How  does  the  

mechanism  of  quinolone  cytotoxicity  change  based  on  subclass  and  structural  

modifications  unique  to  each  drug?  With  models  that  are  proposed  under  high  

concentrations  of  drug,  such  as  gyrase  subunit  dissociation,  what  is  the  threshold  for  

shifting  to  that  mode  of  cytotoxicity?  Can  we  detect  that  shift?    

Combining  the  methods  of  physical  analysis  of  chromosome  fragmentation  

presented  in  Chapter  4  with  the  separation  of  function  mutants  isolated  in  Chapter  3  

Page 153: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

137  

could  push  the  field  with  a  better  understanding  of  the  additional  function  of  ε  and  how  

it  relates  to  quinolone  cytotoxicity.    

Page 154: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

138  

References  

1.   Marszalek,  J.  and  J.M.  Kaguni,  DnaA  protein  directs  the  binding  of  DnaB  protein  in  initiation  of  DNA  replication  in  Escherichia  coli.  J  Biol  Chem,  1994.  269(7):  p.  4883-­‐‑90.  

2.   Wahle,  E.,  R.S.  Lasken,  and  A.  Kornberg,  The  dnaB-­‐‑dnaC  replication  protein  complex  of  Escherichia  coli.  II.  Role  of  the  complex  in  mobilizing  dnaB  functions.  J  Biol  Chem,  1989.  264(5):  p.  2469-­‐‑75.  

3.   Baker,  T.A.,  B.E.  Funnell,  and  A.  Kornberg,  Helicase  action  of  dnaB  protein  during  replication  from  the  Escherichia  coli  chromosomal  origin  in  vitro.  J  Biol  Chem,  1987.  262(14):  p.  6877-­‐‑85.  

4.   Johnson,  S.K.,  S.  Bhattacharyya,  and  M.A.  Griep,  DnaB  helicase  stimulates  primer  synthesis  activity  on  short  oligonucleotide  templates.  Biochemistry,  2000.  39(4):  p.  736-­‐‑44.  

5.   Sugaya,  Y.,  et  al.,  Hyper-­‐‑processive  and  slower  DNA  chain  elongation  catalysed  by  DNA  polymerase  III  holoenzyme  purified  from  the  dnaE173  mutator  mutant  of  Escherichia  coli.  Genes  Cells,  2002.  7(4):  p.  385-­‐‑99.  

6.   Pham,  T.M.,  et  al.,  A  single-­‐‑molecule  approach  to  DNA  replication  in  Escherichia  coli  cells  demonstrated  that  DNA  polymerase  III  is  a  major  determinant  of  fork  speed.  Mol  Microbiol,  2013.  90(3):  p.  584-­‐‑96.  

7.   McInerney,  P.,  et  al.,  Characterization  of  a  triple  DNA  polymerase  replisome.  Mol  Cell,  2007.  27(4):  p.  527-­‐‑38.  

8.   Georgescu,  R.E.,  I.  Kurth,  and  M.E.  O'ʹDonnell,  Single-­‐‑molecule  studies  reveal  the  function  of  a  third  polymerase  in  the  replisome.  Nat  Struct  Mol  Biol,  2012.  19(1):  p.  113-­‐‑6.  

9.   Kurth,  I.  and  M.  O'ʹDonnell,  New  insights  into  replisome  fluidity  during  chromosome  replication.  Trends  Biochem  Sci,  2013.  38(4):  p.  195-­‐‑203.  

10.   Drake,  J.W.,  A  constant  rate  of  spontaneous  mutation  in  DNA-­‐‑based  microbes.  Proc  Natl  Acad  Sci  U  S  A,  1991.  88(16):  p.  7160-­‐‑4.  

11.   Drake,  J.W.,  et  al.,  Rates  of  spontaneous  mutation.  Genetics,  1998.  148(4):  p.  1667-­‐‑86.  

12.   Iyer,  R.R.,  et  al.,  DNA  mismatch  repair:  functions  and  mechanisms.  Chem  Rev,  2006.  106(2):  p.  302-­‐‑23.  

Page 155: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

139  

13.   Fijalkowska,  I.J.,  R.M.  Schaaper,  and  P.  Jonczyk,  DNA  replication  fidelity  in  Escherichia  coli:  a  multi-­‐‑DNA  polymerase  affair.  FEMS  Microbiol  Rev,  2012.  36(6):  p.  1105-­‐‑21.  

14.   Ng,  J.Y.  and  K.J.  Marians,  The  ordered  assembly  of  the  phiX174-­‐‑type  primosome.  I.  Isolation  and  identification  of  intermediate  protein-­‐‑DNA  complexes.  J  Biol  Chem,  1996.  271(26):  p.  15642-­‐‑8.  

15.   Kuzminov,  A.,  Recombinational  repair  of  DNA  damage  in  Escherichia  coli  and  bacteriophage  lambda.  Microbiol  Mol  Biol  Rev,  1999.  63(4):  p.  751-­‐‑813,  table  of  contents.  

16.   Michel,  B.,  et  al.,  Recombination  proteins  and  rescue  of  arrested  replication  forks.  DNA  Repair  (Amst),  2007.  6(7):  p.  967-­‐‑80.  

17.   Radman,  M.,  SOS  repair  hypothesis:  phenomenology  of  an  inducible  DNA  repair  which  is  accompanied  by  mutagenesis.  Basic  life  sciences,  1975.  5A:  p.  355-­‐‑367.  

18.   Butala,  M.,  D.  Zgur-­‐‑Bertok,  and  S.J.  Busby,  The  bacterial  LexA  transcriptional  repressor.  Cell  Mol  Life  Sci,  2009.  66(1):  p.  82-­‐‑93.  

19.   Whitby,  M.C.L.,  Robert  G.,  Altered  SOS  induction  associated  with  mutation  in  recF,  recO  and  recR.  Mol  Gen  Genet,  1995.  246:  p.  174-­‐‑179.  

20.   Courcelle,  J.,  et  al.,  Comparative  gene  expression  profiles  following  UV  exposure  in  wild-­‐‑type  and  SOS-­‐‑deficient  Escherichia  coli.  Genetics,  2001.  158(1):  p.  41-­‐‑64.  

21.   Ronen,  M.,  et  al.,  Assigning  numbers  to  the  arrows:  parameterizing  a  gene  regulation  network  by  using  accurate  expression  kinetics.  Proc  Natl  Acad  Sci  U  S  A,  2002.  99(16):  p.  10555-­‐‑60.  

22.   Friedman,  N.,  et  al.,  Precise  temporal  modulation  in  the  response  of  the  SOS  DNA  repair  network  in  individual  bacteria.  PLoS  Biol,  2005.  3(7):  p.  e238.  

23.   Michel,  B.,  After  30  years  of  study,  the  bacterial  SOS  response  still  surprises  us.  PLoS  Biol,  2005.  3(7):  p.  e255.  

24.   Iwasaki,  H.,  et  al.,  The  Escherichia  coli  polB  gene,  which  encodes  DNA  polymerase  II,  is  regulated  by  the  SOS  system.  J  Bacteriol,  1990.  172(11):  p.  6268-­‐‑73.  

25.   Ohmori,  H.,  et  al.,  The  Y-­‐‑family  of  DNA  polymerases.  Mol  Cell,  2001.  8(1):  p.  7-­‐‑8.  

Page 156: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

140  

26.   Oh,  T.J.,  C.W.  Lee,  and  I.G.  Kim,  The  damage-­‐‑inducible  (din)  genes  of  Escherichia  coli  are  induced  by  various  genotoxins  in  a  different  way.  Microbiol  Res,  1999.  154(2):  p.  179-­‐‑83.  

27.   Qiu,  Z.  and  M.F.  Goodman,  The  Escherichia  coli  polB  locus  is  identical  to  dinA,  the  structural  gene  for  DNA  polymerase  II.  Characterization  of  Pol  II  purified  from  a  polB  mutant.  J  Biol  Chem,  1997.  272(13):  p.  8611-­‐‑7.  

28.   Tessman,  I.  and  M.A.  Kennedy,  DNA  polymerase  II  of  Escherichia  coli  in  the  bypass  of  abasic  sites  in  vivo.  Genetics,  1994.  136(2):  p.  439-­‐‑48.  

29.   Kanuri,  M.,  et  al.,  Evidence  for  Escherichia  coli  polymerase  II  mutagenic  bypass  of  intrastrand  DNA  crosslinks.  DNA  Repair  (Amst),  2005.  4(12):  p.  1374-­‐‑80.  

30.   Chandani,  S.  and  E.L.  Loechler,  Y-­‐‑Family  DNA  polymerases  may  use  two  different  dNTP  shapes  for  insertion:  a  hypothesis  and  its  implications.  J  Mol  Graph  Model,  2009.  27(7):  p.  759-­‐‑69.  

31.   Kim,  S.R.,  et  al.,  Roles  of  chromosomal  and  episomal  dinB  genes  encoding  DNA  pol  IV  in  targeted  and  untargeted  mutagenesis  in  Escherichia  coli.  Mol  Genet  Genomics,  2001.  266(2):  p.  207-­‐‑15.  

32.   Wagner,  J.  and  T.  Nohmi,  Escherichia  coli  DNA  polymerase  IV  mutator  activity:  genetic  requirements  and  mutational  specificity.  J  Bacteriol,  2000.  182(16):  p.  4587-­‐‑95.  

33.   Galhardo,  R.S.,  et  al.,  DinB  upregulation  is  the  sole  role  of  the  SOS  response  in  stress-­‐‑induced  mutagenesis  in  Escherichia  coli.  Genetics,  2009.  182(1):  p.  55-­‐‑68.  

34.   Burckhardt,  S.E.,  et  al.,  UmuD  mutagenesis  protein  of  Escherichia  coli:  overproduction,  purification,  and  cleavage  by  RecA.  Proc  Natl  Acad  Sci  U  S  A,  1988.  85(6):  p.  1811-­‐‑5.  

35.   Tang,  M.,  et  al.,  UmuD'ʹ(2)C  is  an  error-­‐‑prone  DNA  polymerase,  Escherichia  coli  pol  V.  Proc  Natl  Acad  Sci  U  S  A,  1999.  96(16):  p.  8919-­‐‑24.  

36.   Champoux,  J.J.,  DNA  topoisomerases:  structure,  function,  and  mechanism.  Annu  Rev  Biochem,  2001.  70:  p.  369-­‐‑413.  

37.   Wang,  J.C.,  Cellular  roles  of  DNA  topoisomerases:  a  molecular  perspective.  Nat  Rev  Mol  Cell  Biol,  2002.  3(6):  p.  430-­‐‑40.  

Page 157: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

141  

38.   Nollmann,  M.,  N.J.  Crisona,  and  P.B.  Arimondo,  Thirty  years  of  Escherichia  coli  DNA  gyrase:  from  in  vivo  function  to  single-­‐‑molecule  mechanism.  Biochimie,  2007.  89(4):  p.  490-­‐‑9.  

39.   Sugino,  A.,  et  al.,  Mechanism  of  action  of  nalidixic  acid:  purification  of  Escherichia  coli  nalA  gene  product  and  its  relationship  to  DNA  gyrase  and  a  novel  nicking-­‐‑closing  enzyme.  Proc  Natl  Acad  Sci  U  S  A,  1977.  74(11):  p.  4767-­‐‑71.  

40.   Gellert,  M.,  et  al.,  Nalidixic  acid  resistance:  a  second  genetic  character  involved  in  DNA  gyrase  activity.  Proc  Natl  Acad  Sci  U  S  A,  1977.  74(11):  p.  4772-­‐‑6.  

41.   Kato,  J.,  et  al.,  New  topoisomerase  essential  for  chromosome  segregation  in  E.  coli.  Cell,  1990.  63(2):  p.  393-­‐‑404.  

42.   Wang,  J.C.,  DNA  topoisomerases.  Annu  Rev  Biochem,  1996.  65:  p.  635-­‐‑92.  

43.   Ruiz,  J.,  Mechanisms  of  resistance  to  quinolones:  target  alterations,  decreased  accumulation  and  DNA  gyrase  protection.  J  Antimicrob  Chemother,  2003.  51(5):  p.  1109-­‐‑17.  

44.   Drlica,  K.,  et  al.,  Quinolone-­‐‑mediated  bacterial  death.  Antimicrob  Agents  Chemother,  2008.  52(2):  p.  385-­‐‑92.  

45.   Khodursky,  A.B.  and  N.R.  Cozzarelli,  The  mechanism  of  inhibition  of  topoisomerase  IV  by  quinolone  antibacterials.  J  Biol  Chem,  1998.  273(42):  p.  27668-­‐‑77.  

46.   Drlica,  K.  and  X.  Zhao,  DNA  gyrase,  topoisomerase  IV,  and  the  4-­‐‑quinolones.  Microbiol  Mol  Biol  Rev,  1997.  61(3):  p.  377-­‐‑92.  

47.   Newmark,  K.G.,  et  al.,  Genetic  analysis  of  the  requirements  for  SOS  induction  by  nalidixic  acid  in  Escherichia  coli.  Gene,  2005.  356:  p.  69-­‐‑76.  

48.   McPartland,  A.,  L.  Green,  and  H.  Echols,  Control  of  recA  gene  RNA  in  E.  coli:  regulatory  and  signal  genes.  Cell,  1980.  20(3):  p.  731-­‐‑737.  

49.   O'ʹReilly,  E.K.  and  K.N.  Kreuzer,  Isolation  of  SOS  constitutive  mutants  of  Escherichia  coli.  J  Bacteriol,  2004.  186(21):  p.  7149-­‐‑60.  

50.   Pohlhaus,  J.R.,  et  al.,  The  epsilon  subunit  of  DNA  polymerase  III  Is  involved  in  the  nalidixic  acid-­‐‑induced  SOS  response  in  Escherichia  coli.  J  Bacteriol,  2008.  190(15):  p.  5239-­‐‑47.  

Page 158: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

142  

51.   Hiasa,  H.,  D.O.  Yousef,  and  K.J.  Marians,  DNA  strand  cleavage  is  required  for  replication  fork  arrest  by  a  frozen  topoisomerase-­‐‑quinolone-­‐‑DNA  ternary  complex.  J  Biol  Chem,  1996.  271(42):  p.  26424-­‐‑9.  

52.   Chen,  C.R.,  et  al.,  DNA  gyrase  and  topoisomerase  IV  on  the  bacterial  chromosome:  quinolone-­‐‑induced  DNA  cleavage.  J  Mol  Biol,  1996.  258(4):  p.  627-­‐‑37.  

53.   Scheuermann,  R.,  et  al.,  Identification  of  the  epsilon-­‐‑subunit  of  Escherichia  coli  DNA  polymerase  III  holoenzyme  as  the  dnaQ  gene  product:  a  fidelity  subunit  for  DNA  replication.  Proc  Natl  Acad  Sci  U  S  A,  1983.  80(23):  p.  7085-­‐‑9.  

54.   Welch,  M.M.  and  C.S.  McHenry,  Cloning  and  identification  of  the  product  of  the  dnaE  gene  of  Escherichia  coli.  J  Bacteriol,  1982.  152(1):  p.  351-­‐‑6.  

55.   McHenry,  C.S.  and  W.  Crow,  DNA  polymerase  III  of  Escherichia  coli.  Purification  and  identification  of  subunits.  J  Biol  Chem,  1979.  254(5):  p.  1748-­‐‑53.  

56.   Studwell-­‐‑Vaughan,  P.S.  and  M.  O'ʹDonnell,  DNA  polymerase  III  accessory  proteins.  V.  Theta  encoded  by  holE.  J  Biol  Chem,  1993.  268(16):  p.  11785-­‐‑91.  

57.   Slater,  S.C.,  et  al.,  holE,  the  gene  coding  for  the  theta  subunit  of  DNA  polymerase  III  of  Escherichia  coli:  characterization  of  a  holE  mutant  and  comparison  with  a  dnaQ  (epsilon-­‐‑subunit)  mutant.  J  Bacteriol,  1994.  176(3):  p.  815-­‐‑21.  

58.   Marians,  K.J.,  et  al.,  Role  of  the  core  DNA  polymerase  III  subunits  at  the  replication  fork.  Alpha  is  the  only  subunit  required  for  processive  replication.  J  Biol  Chem,  1998.  273(4):  p.  2452-­‐‑7.  

59.   Studwell,  P.S.  and  M.  O'ʹDonnell,  Processive  replication  is  contingent  on  the  exonuclease  subunit  of  DNA  polymerase  III  holoenzyme.  J  Biol  Chem,  1990.  265(2):  p.  1171-­‐‑8.  

60.   Perrino,  F.W.,  S.  Harvey,  and  S.M.  McNeill,  Two  functional  domains  of  the  epsilon  subunit  of  DNA  polymerase  III.  Biochemistry,  1999.  38(48):  p.  16001-­‐‑9.  

61.   Maki,  H.  and  A.  Kornberg,  Proofreading  by  DNA  polymerase  III  of  Escherichia  coli  depends  on  cooperative  interaction  of  the  polymerase  and  exonuclease  subunits.  Proc  Natl  Acad  Sci  U  S  A,  1987.  84(13):  p.  4389-­‐‑92.  

62.   Keniry,  M.A.,  et  al.,  Structure  of  the  theta  subunit  of  Escherichia  coli  DNA  polymerase  III  in  complex  with  the  epsilon  subunit.  J  Bacteriol,  2006.  188(12):  p.  4464-­‐‑73.  

Page 159: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

143  

63.   DeRose,  E.F.,  et  al.,  Elucidation  of  the  epsilon-­‐‑theta  subunit  interface  of  Escherichia  coli  DNA  polymerase  III  by  NMR  spectroscopy.  Biochemistry,  2003.  42(13):  p.  3635-­‐‑44.  

64.   Bernad,  A.,  et  al.,  A  conserved  3'ʹ-­‐‑-­‐‑-­‐‑-­‐‑5'ʹ  exonuclease  active  site  in  prokaryotic  and  eukaryotic  DNA  polymerases.  Cell,  1989.  59(1):  p.  219-­‐‑28.  

65.   Blanco,  L.,  A.  Bernad,  and  M.  Salas,  Evidence  favouring  the  hypothesis  of  a  conserved  3'ʹ-­‐‑5'ʹ  exonuclease  active  site  in  DNA-­‐‑dependent  DNA  polymerases.  Gene,  1992.  112(1):  p.  139-­‐‑44.  

66.   Barnes,  M.H.,  et  al.,  The  3'ʹ-­‐‑5'ʹ  exonuclease  site  of  DNA  polymerase  III  from  gram-­‐‑positive  bacteria:  definition  of  a  novel  motif  structure.  Gene,  1995.  165(1):  p.  45-­‐‑50.  

67.   Moser,  M.J.,  et  al.,  The  proofreading  domain  of  Escherichia  coli  DNA  polymerase  I  and  other  DNA  and/or  RNA  exonuclease  domains.  Nucleic  Acids  Res,  1997.  25(24):  p.  5110-­‐‑8.  

68.   Taft-­‐‑Benz,  S.A.  and  R.M.  Schaaper,  Mutational  analysis  of  the  3'ʹ-­‐‑-­‐‑>5'ʹ  proofreading  exonuclease  of  Escherichia  coli  DNA  polymerase  III.  Nucleic  Acids  Res,  1998.  26(17):  p.  4005-­‐‑11.  

69.   Wootton,  J.C.  and  M.H.  Drummond,  The  Q-­‐‑linker:  a  class  of  interdomain  sequences  found  in  bacterial  multidomain  regulatory  proteins.  Protein  Eng,  1989.  2(7):  p.  535-­‐‑43.  

70.   Sutrina,  S.L.,  et  al.,  The  glucose  permease  of  Bacillus  subtilis  is  a  single  polypeptide  chain  that  functions  to  energize  the  sucrose  permease.  J  Biol  Chem,  1990.  265(30):  p.  18581-­‐‑9.  

71.   Ozawa,  K.,  et  al.,  The  proofreading  exonuclease  subunit  epsilon  of  Escherichia  coli  DNA  polymerase  III  is  tethered  to  the  polymerase  subunit  alpha  via  a  flexible  linker.  Nucleic  Acids  Res,  2008.  36(15):  p.  5074-­‐‑82.  

72.   Deshpande,  N.,  et  al.,  The  RCSB  Protein  Data  Bank:  a  redesigned  query  system  and  relational  database  based  on  the  mmCIF  schema.  Nucleic  Acids  Res,  2005.  33(Database  issue):  p.  D233-­‐‑7.  

73.   Kabsch,  W.  and  C.  Sander,  Dictionary  of  protein  secondary  structure:  pattern  recognition  of  hydrogen-­‐‑bonded  and  geometrical  features.  Biopolymers,  1983.  22(12):  p.  2577-­‐‑637.  

74.   Hamdan,  S.,  et  al.,  Structural  basis  for  proofreading  during  replication  of  the  Escherichia  coli  chromosome.  Structure,  2002.  10(4):  p.  535-­‐‑46.  

Page 160: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

144  

75.   Taft-­‐‑Benz,  S.A.  and  R.M.  Schaaper,  The  C-­‐‑terminal  domain  of  dnaQ  contains  the  polymerase  binding  site.  J  Bacteriol,  1999.  181(9):  p.  2963-­‐‑5.  

76.   Bressanin,  D.,  et  al.,  Proteolysis  of  the  proofreading  subunit  controls  the  assembly  of  Escherichia  coli  DNA  polymerase  III  catalytic  core.  Biochim  Biophys  Acta,  2009.  1794(11):  p.  1606-­‐‑15.  

77.   Ozawa,  K.,  et  al.,  Proofreading  exonuclease  on  a  tether:  the  complex  between  the  E.  coli  DNA  polymerase  III  subunits  alpha,  epsilon,  theta  and  beta  reveals  a  highly  flexible  arrangement  of  the  proofreading  domain.  Nucleic  Acids  Res,  2013.  41(10):  p.  5354-­‐‑67.  

78.   Echols,  H.,  C.  Lu,  and  P.  Burgers,  Mutator  strains  of  Escherichia  coli,  mutD  and  dnaQ,  with  defective  exonucleolytic  editing  by  DNA  polymerase  III  holoenzyme.  Proc  Natl  Acad  Sci  U  S  A,  1983.  80(8):  p.  2189-­‐‑2192.  

79.   Takano,  K.,  et  al.,  Structure  and  function  of  dnaQ  and  mutD  mutators  of  Escherichia  coli.  Mol  Gen  Genet,  1986.  205(1):  p.  9-­‐‑13.  

80.   Maruyama,  M.,  et  al.,  A  dominant  (mutD5)  and  a  recessive  (dnaQ49)  mutator  of  Escherichia  coli.  J  Mol  Biol,  1983.  167(4):  p.  757-­‐‑71.  

81.   Fijalkowska,  I.  and  R.  Schaaper,  Mutants  in  the  Exo  I  motif  of  Escherichia  coli  dnaQ:  defective  proofreading  and  inviability  due  to  error  catastrophe.  Proc  Natl  Acad  Sci  U  S  A,  1996.  93(7):  p.  2856-­‐‑2861.  

82.   Degnen,  G.E.  and  E.C.  Cox,  Conditional  mutator  gene  in  Escherichia  coli:  isolation,  mapping,  and  effector  studies.  J  Bacteriol,  1974.  117(2):  p.  477-­‐‑87.  

83.   Schaaper,  R.M.,  Mechanisms  of  mutagenesis  in  the  Escherichia  coli  mutator  mutD5:  role  of  DNA  mismatch  repair.  Proc  Natl  Acad  Sci  U  S  A,  1988.  85(21):  p.  8126-­‐‑30.  

84.   Jonczyk,  P.,  et  al.,  In  vivo  protein  interactions  within  the  Escherichia  coli  DNA  polymerase  III  core.  J  Bacteriol,  1998.  180(6):  p.  1563-­‐‑6.  

85.   Nowosielska,  A.,  et  al.,  Mutator  activity  and  specificity  of  Escherichia  coli  dnaQ49  allele-­‐‑-­‐‑effect  of  umuDC  products.  Mutat  Res,  2005.  572(1-­‐‑2):  p.  113-­‐‑22.  

86.   Saveson,  C.J.  and  S.T.  Lovett,  Enhanced  deletion  formation  by  aberrant  DNA  replication  in  Escherichia  coli.  Genetics,  1997.  146(2):  p.  457-­‐‑70.  

Page 161: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

145  

87.   Cao,  Y.,  et  al.,  Structure-­‐‑function  correlations  in  the  XerD  site-­‐‑specific  recombinase  revealed  by  pentapeptide  scanning  mutagenesis.  Journal  of  molecular  biology,  1997.  274(1):  p.  39-­‐‑53.  

88.   Anton,  B.  and  E.  Raleigh,  Transposon-­‐‑mediated  linker  insertion  scanning  mutagenesis  of  the  Escherichia  coli  McrA  endonuclease.  J  Bacteriol,  2004.  186(17):  p.  5699-­‐‑5707.  

89.   Heitman,  J.  and  P.  Model,  SOS  induction  as  an  in  vivo  assay  of  enzyme-­‐‑DNA  interactions.  Gene,  1991.  103(1):  p.  1-­‐‑9.  

90.   Baba,  T.,  et  al.,  Construction  of  Escherichia  coli  K-­‐‑12  in-­‐‑frame,  single-­‐‑gene  knockout  mutants:  the  Keio  collection.  Mol  Syst  Biol,  2006.  2:  p.  2006  0008.  

91.   Maki,  H.,  T.  Horiuchi,  and  M.  Sekiguchi,  Structure  and  expression  of  the  dnaQ  mutator  and  the  RNase  H  genes  of  Escherichia  coli:  overlap  of  the  promoter  regions.  Proc  Natl  Acad  Sci  U  S  A,  1983.  80(23):  p.  7137-­‐‑7141.  

92.   Rose,  R.E.,  The  nucleotide  sequence  of  pACYC184.  Nucleic  Acids  Res,  1988.  16(1):  p.  355.  

93.   Hamilton,  C.M.,  et  al.,  New  method  for  generating  deletions  and  gene  replacements  in  Escherichia  coli.  J  Bacteriol,  1989.  171(9):  p.  4617-­‐‑22.  

94.   Foster,  P.L.,  Methods  for  determining  spontaneous  mutation  rates.  Methods  Enzymol,  2006.  409:  p.  195-­‐‑213.  

95.   Uranga,  L.A.,  et  al.,  The  Escherichia  coli  DinD  protein  modulates  RecA  activity  by  inhibiting  postsynaptic  RecA  filaments.  J  Biol  Chem,  2011.  286(34):  p.  29480-­‐‑91.  

96.   Sambrook,  J.  and  D.W.  Russell,  Molecular  cloning  :  a  laboratory  manual.  3rd  ed.  2001,  Cold  Spring  Harbor,  N.Y.:  Cold  Spring  Harbor  Laboratory  Press.  

97.   Nomura,  T.,  H.  Aiba,  and  A.  Ishihama,  Transcriptional  organization  of  the  convergent  overlapping  dnaQ-­‐‑rnh  genes  of  Escherichia  coli.  J  Biol  Chem,  1985.  260(11):  p.  7122-­‐‑5.  

98.   Ogawa,  T.  and  T.  Okazaki,  Function  of  RNase  H  in  DNA  replication  revealed  by  RNase  H  defective  mutants  of  Escherichia  coli.  Mol  Gen  Genet,  1984.  193(2):  p.  231-­‐‑7.  

99.   Kogoma,  T.,  et  al.,  Requirement  of  homologous  recombination  functions  for  viability  of  the  Escherichia  coli  cell  that  lacks  RNase  HI  and  exonuclease  V  activities.  Biochimie,  1993.  75(1-­‐‑2):  p.  89-­‐‑99.  

Page 162: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

146  

100.   Dorazi,  R.,  J.J.  Lingutla,  and  M.Z.  Humayun,  Expression  of  mutant  alanine  tRNAs  increases  spontaneous  mutagenesis  in  Escherichia  coli.  Mol  Microbiol,  2002.  44(1):  p.  131-­‐‑41.  

101.   Dorazi,  R.,  Can  tRNAs  act  as  antisense  RNA?  The  case  of  mutA  and  dnaQ.  J  Theor  Biol,  2003.  225(3):  p.  383-­‐‑8.  

102.   Thibodeau,  S.A.,  R.  Fang,  and  J.K.  Joung,  High-­‐‑throughput  beta-­‐‑galactosidase  assay  for  bacterial  cell-­‐‑based  reporter  systems.  Biotechniques,  2004.  36(3):  p.  410-­‐‑5.  

103.   Berman,  H.M.,  et  al.,  The  Protein  Data  Bank.  Nucleic  Acids  Res,  2000.  28(1):  p.  235-­‐‑42.  

104.   Jergic,  S.,  et  al.,  A  direct  proofreader-­‐‑clamp  interaction  stabilizes  the  Pol  III  replicase  in  the  polymerization  mode.  EMBO  J,  2013.  32(9):  p.  1322-­‐‑33.  

105.   Toste  Rego,  A.,  et  al.,  Architecture  of  the  Pol  III-­‐‑clamp-­‐‑exonuclease  complex  reveals  key  roles  of  the  exonuclease  subunit  in  processive  DNA  synthesis  and  repair.  EMBO  J,  2013.  32(9):  p.  1334-­‐‑43.  

106.   Moreland,  J.L.,  et  al.,  The  Molecular  Biology  Toolkit  (MBT):  a  modular  platform  for  developing  molecular  visualization  applications.  BMC  Bioinformatics,  2005.  6:  p.  21.  

107.   Xu,  D.  and  Y.  Zhang,  Generating  triangulated  macromolecular  surfaces  by  Euclidean  Distance  Transform.  PLoS  One,  2009.  4(12):  p.  e8140.  

108.   Slupska,  M.M.,  et  al.,  Examination  of  the  role  of  DNA  polymerase  proofreading  in  the  mutator  effect  of  miscoding  tRNAs.  J  Bacteriol,  1998.  180(21):  p.  5712-­‐‑7.  

109.   Lancy,  E.D.,  et  al.,  Nucleotide  sequences  of  dnaE,  the  gene  for  the  polymerase  subunit  of  DNA  polymerase  III  in  Salmonella  typhimurium,  and  a  variant  that  facilitates  growth  in  the  absence  of  another  polymerase  subunit.  J  Bacteriol,  1989.  171(10):  p.  5581-­‐‑6.  

110.   Lancy,  E.D.,  et  al.,  Isolation  and  characterization  of  mutants  with  deletions  in  dnaQ,  the  gene  for  the  editing  subunit  of  DNA  polymerase  III  in  Salmonella  typhimurium.  J  Bacteriol,  1989.  171(10):  p.  5572-­‐‑80.  

111.   Schaaper,  I.J.F.A.R.M.,  Effects  of  Escherichia  coli  dnaE  Antimutator  Alleles  in  a  Proofreading-­‐‑Deficient  mutD5  Strain.  Journal  of  Bacteriology,  1995.  177(20):  p.  5979-­‐‑5986.  

Page 163: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

147  

112.   Schaaper,  R.M.  and  R.  Cornacchio,  An  Escherichia  coli  dnaE  mutation  with  suppressor  activity  toward  mutator  mutD5.  J  Bacteriol,  1992.  174(6):  p.  1974-­‐‑82.  

113.   Lifsics,  M.R.,  E.D.  Lancy,  Jr.,  and  R.  Maurer,  DNA  replication  defect  in  Salmonella  typhimurium  mutants  lacking  the  editing  (epsilon)  subunit  of  DNA  polymerase  III.  J  Bacteriol,  1992.  174(21):  p.  6965-­‐‑73.  

114.   Perrino,  D.A.L.a.F.W.,  Dysfunctional  proofreading  in  the  Escherichia  coli  DNA  polymerase  III  core.  Journal  of  Biochemistry,  2004.  384:  p.  337-­‐‑348.  

115.   Lehtinen,  D.A.  and  F.W.  Perrino,  Dysfunctional  proofreading  in  the  Escherichia  coli  DNA  polymerase  III  core.  Biochem  J,  2004.  384(Pt  2):  p.  337-­‐‑48.  

116.   Kim,  D.R.  and  C.S.  McHenry,  In  vivo  assembly  of  overproduced  DNA  polymerase  III.  Overproduction,  purification,  and  characterization  of  the  alpha,  alpha-­‐‑epsilon,  and  alpha-­‐‑epsilon-­‐‑theta  subunits.  J  Biol  Chem,  1996.  271(34):  p.  20681-­‐‑9.  

117.   Snyder,  M.  and  K.  Drlica,  DNA  gyrase  on  the  bacterial  chromosome:  DNA  cleavage  induced  by  oxolinic  acid.  J  Mol  Biol,  1979.  131(2):  p.  287-­‐‑302.  

118.   Pohlhaus,  J.R.  and  K.N.  Kreuzer,  Norfloxacin-­‐‑induced  DNA  gyrase  cleavage  complexes  block  Escherichia  coli  replication  forks,  causing  double-­‐‑stranded  breaks  in  vivo.  Mol  Microbiol,  2005.  56(6):  p.  1416-­‐‑29.  

119.   McDaniel,  L.S.,  L.H.  Rogers,  and  W.E.  Hill,  Survival  of  recombination-­‐‑deficient  mutants  of  Escherichia  coli  during  incubation  with  nalidixic  acid.  J  Bacteriol,  1978.  134(3):  p.  1195-­‐‑8.  

120.   Zhao,  X.,  et  al.,  Lethal  action  of  quinolones  against  a  temperature-­‐‑sensitive  dnaB  replication  mutant  of  Escherichia  coli.  Antimicrob  Agents  Chemother,  2006.  50(1):  p.  362-­‐‑4.  

121.   Lewin,  C.S.,  B.M.  Howard,  and  J.T.  Smith,  Protein-­‐‑  and  RNA-­‐‑synthesis  independent  bactericidal  activity  of  ciprofloxacin  that  involves  the  A  subunit  of  DNA  gyrase.  J  Med  Microbiol,  1991.  34(1):  p.  19-­‐‑22.  

122.   Deitz,  W.H.,  T.M.  Cook,  and  W.A.  Goss,  Mechanism  of  action  of  nalidixic  acid  on  Escherichia  coli.  3.  Conditions  required  for  lethality.  J  Bacteriol,  1966.  91(2):  p.  768-­‐‑73.  

123.   Shimizu,  H.,  H.  Yamaguchi,  and  H.  Ikeda,  Molecular  analysis  of  lambda  bio  transducing  phage  produced  by  oxolinic  acid-­‐‑induced  illegitimate  recombination  in  vivo.  Genetics,  1995.  140(3):  p.  889-­‐‑96.  

Page 164: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

148  

124.   Volff,  J.N.,  D.  Vandewiele,  and  B.  Decaris,  Stimulation  of  genetic  instability  and  associated  large  genomic  rearrangements  in  Streptomyces  ambofaciens  by  three  fluoroquinolones.  Antimicrob  Agents  Chemother,  1994.  38(9):  p.  1984-­‐‑90.  

125.   Bergerat,  A.,  et  al.,  An  atypical  topoisomerase  II  from  Archaea  with  implications  for  meiotic  recombination.  Nature,  1997.  386(6623):  p.  414-­‐‑7.  

126.   Diaz,  R.L.,  et  al.,  Identification  of  residues  in  yeast  Spo11p  critical  for  meiotic  DNA  double-­‐‑strand  break  formation.  Mol  Cell  Biol,  2002.  22(4):  p.  1106-­‐‑15.  

127.   Neale,  M.J.,  J.  Pan,  and  S.  Keeney,  Endonucleolytic  processing  of  covalent  protein-­‐‑linked  DNA  double-­‐‑strand  breaks.  Nature,  2005.  436(7053):  p.  1053-­‐‑7.  

128.   Hartsuiker,  E.,  et  al.,  Ctp1CtIP  and  Rad32Mre11  nuclease  activity  are  required  for  Rec12Spo11  removal,  but  Rec12Spo11  removal  is  dispensable  for  other  MRN-­‐‑dependent  meiotic  functions.  Mol  Cell  Biol,  2009.  29(7):  p.  1671-­‐‑81.  

129.   Kuzminov,  A.,  Collapse  and  repair  of  replication  forks  in  Escherichia  coli.  Mol  Microbiol,  1995.  16(3):  p.  373-­‐‑84.  

130.   Hiasa,  H.  and  K.J.  Marians,  Two  distinct  modes  of  strand  unlinking  during  theta-­‐‑type  DNA  replication.  J  Biol  Chem,  1996.  271(35):  p.  21529-­‐‑35.  

131.   Strumberg,  D.,  et  al.,  Conversion  of  topoisomerase  I  cleavage  complexes  on  the  leading  strand  of  ribosomal  DNA  into  5'ʹ-­‐‑phosphorylated  DNA  double-­‐‑strand  breaks  by  replication  runoff.  Mol  Cell  Biol,  2000.  20(11):  p.  3977-­‐‑87.  

132.   Hong,  G.  and  K.N.  Kreuzer,  Endonuclease  cleavage  of  blocked  replication  forks:  An  indirect  pathway  of  DNA  damage  from  antitumor  drug-­‐‑topoisomerase  complexes.  Proc  Natl  Acad  Sci  U  S  A,  2003.  100(9):  p.  5046-­‐‑51.  

133.   Finney,  M.,  Pulsed-­‐‑field  gel  electrophoresis.  Curr  Protoc  Mol  Biol,  2001.  Chapter  2:  p.  Unit2  5B.  

134.   Akerman,  B.,  Effects  of  supercoiling  in  electrophoretic  trapping  of  circular  DNA  in  polyacrylamide  gels.  Biophys  J,  1998.  74(6):  p.  3140-­‐‑51.  

135.   Carle,  G.F.,  M.  Frank,  and  M.V.  Olson,  Electrophoretic  separations  of  large  DNA  molecules  by  periodic  inversion  of  the  electric  field.  Science,  1986.  232(4746):  p.  65-­‐‑8.  

136.   Birren,  B.  and  E.  Lai,  Pulsed  Field  Gel  Electrophoresis:  A  Practical  Guide.  1993:  Academic  Press.  272.  

Page 165: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

149  

137.   Miranda,  A.  and  A.  Kuzminov,  Chromosomal  lesion  suppression  and  removal  in  Escherichia  coli  via  linear  DNA  degradation.  Genetics,  2003.  163(4):  p.  1255-­‐‑71.  

138.   Newsom,  S.W.,  The  antimicrobial  spectrum  of  norfloxacin.  J  Antimicrob  Chemother,  1984.  13  Suppl  B:  p.  25-­‐‑31.  

139.   Vila,  J.,  et  al.,  Association  between  double  mutation  in  gyrA  gene  of  ciprofloxacin-­‐‑resistant  clinical  isolates  of  Escherichia  coli  and  MICs.  Antimicrob  Agents  Chemother,  1994.  38(10):  p.  2477-­‐‑9.  

140.   Ruiz,  J.,  et  al.,  High  prevalence  of  nalidixic  acid  resistant,  ciprofloxacin  susceptible  phenotype  among  clinical  isolates  of  Escherichia  coli  and  other  Enterobacteriaceae.  Diagn  Microbiol  Infect  Dis,  2002.  42(4):  p.  257-­‐‑61.  

141.   Elwell,  L.P.,  et  al.,  Antibacterial  activity  and  mechanism  of  action  of  3'ʹ-­‐‑azido-­‐‑3'ʹ-­‐‑deoxythymidine  (BW  A509U).  Antimicrob  Agents  Chemother,  1987.  31(2):  p.  274-­‐‑80.  

142.   Sinha,  N.K.  and  D.P.  Snustad,  Mechanism  of  inhibition  of  deoxyribonucleic  acid  synthesis  in  Escherichia  coli  by  hydroxyurea.  J  Bacteriol,  1972.  112(3):  p.  1321-­‐‑4.  

 

Page 166: ZNWThesis 042314 Final - COnnecting REpositories · iv! Abstract Quinolone!treatment!induces!stabilized!cleavage!complexes(SCCs),!consisting!of! acovalent!gyraseKDNAcomplex,!and!processing!of!these!complexes!is!thought

 

 

150  

Biography Zakiya  Nicole  Whatley  was  born  in  Albemarle,  North  Carolina  on  June  6,  1985  to  

Mr.  and  Mrs.  Curtis  and  Vickie  Whatley.  She  grew  up  in  Greensboro,  North  Carolina  

and  graduated  from  James  B.  Dudley  Senior  High  School.  She  then  completed  her  

Bachelor’s  of  Science  at  in  Biology,  concentrating  on  Cellular  and  Molecular  Biology,  at  

Hampton  University  in  Hampton,  VA  in  May  2007.  She  joined  the  Cellular  and  

Molecular  Biology  umbrella  program  at  Duke  University  in  August  2007.  At  Duke,  she  

received  the  James  B.  Duke  Endowment  Fellowship  and  the  Dean’s  Award  for  

Excellence  in  Mentoring.  Zakiya’s  campus  activities  included  serving  as  President  of  the  

Edward  Alexander  Bouchet  Society,  a  Graduate  Student  Liaison  for  the  Office  of  

Graduate  Student  Affairs,  a  graduate  student  mentor  for  the  Summer  Research  

Opportunities  Program  and  the  Institute  for  Genome  Science  and  Policy’s  Summer  

Research  Program.  Zakiya  attended  multiple  conferences  including  the  American  

Society  of  Microbiology’s  Annual  Biomedical  Conference  for  Minorities  in  the  Sciences,  

Leadership  Alliance  National  Symposium,  and  the  Bacteria,  Archaea  and  Phage  

Conference  at  Cold  Spring  Harbor,  New  York.  She  affiliated  with  the  laboratory  of  Dr.  

Kenneth  Kreuzer  of  the  Biochemistry  Department  and  graduated  with  her  PhD  from  the  

University  Program  in  Genetics  and  Genomics  at  Duke  University  in  May  2014.