• Keine Ergebnisse gefunden

Snapshots of DNA polymerase processing aberrant substrates : Structural insights into abasic site bypass and polymerization of 5-alkynylated nucleotide analogs

N/A
N/A
Protected

Academic year: 2022

Aktie "Snapshots of DNA polymerase processing aberrant substrates : Structural insights into abasic site bypass and polymerization of 5-alkynylated nucleotide analogs"

Copied!
120
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

       

Snapshots  of  DNA  polymerase  processing    aberrant  substrates:    

Structural  insights  into  abasic  site  bypass  

&  

 polymerization  of  5-­‐alkynylated  nucleotide  analogs  

         

Dissertation    

zur  Erlangung  des  akademischen  Grades  des   Doktors  der  Naturwissenschaften  

 (Dr.  rer.  nat.)    

 

an  der  Universität  Konstanz  

Naturwissenschaftliche  Sektion  Fachbereich  Chemie    

         

vorgelegt  von    

Samra  Obeid    

      2011  

 

   

(2)

   

                                                             

Referent  1:       Prof.  Dr.  Andreas  Marx   Referent  2:       Prof.  Dr.  Valentin  Wittmann   Referent  3  und  Prüfungsvorsitz:   Prof.  Dr.  Wolfram  Welte   Tag  der  mündlichen  Prüfung:                04.05.2012  

 

(3)

Teile  dieser  Arbeit  sind  veröffentlicht  in:  

 

 

Proc.  Natl.  Acad.  Sci.  U.  S.  A.,  2010,  

Obeid  S,  Baccaro  A,  Welte  W,  Diederichs  K,  &  Marx  A;  

107:21327-­‐21331.   „Structural  basis  for  the  synthesis  of  nucleobase  modified  

DNA  by  Thermus  aquaticus  DNA  polymerase.“  

 

EMBO  J.,  2010,  29:1738-­‐1747.  

Obeid   S,   Blatter   N,   Kranaster   R,   Schnur   A,   Diederichs   K,   Welte   W,   &   Marx   A;   „Replication   through   an   abasic   DNA   lesion:  structural  basis  for  adenine  selectivity.“  

 

J.  Biol.  Chem.,  2012,  287:14099-­‐  

Obeid  S,  Welte  W,  Diederichs  K,  &  Marx  A;  „  Amino  acid  

14108   templating   mechanisms   in   selection   of   nucleotides  

opposite  abasic  sites  by  a  family  a  DNA  polymerase.”  

 

Chem.  Commun.,  2012,  

Obeid  S,  Bußkamp  H,  Welte  W,  Diederichs  K,  &  Marx  A   DOI:  10.1039/c2cc34181f.   „Interactions   of   non-­‐polar   and   “Click-­‐able”   nucleotides   in  

the  confines  of  a  DNA  polymerase.“  

 

Weitere  Publikationen:  

   

ChemBioChem,  2011,  12:1574-­‐1580.   Obeid   S,   Schnur   A,   Gloeckner   C,   Blatter   N,   Welte   W,  

Diederichs   K   ,   &   Marx   A;   „   Learning   from   Directed   Evolution:   Thermus   aquaticus   DNA   Polymerase   Mutants   with  Translesion  Synthesis  Activity.“  

 

Angew.  Chem.,  Int.  Ed.  Engl.,  2008,  

Obeid  S,  Yulikov  M,  Jeschke  G,  &  Marx  A;  

47:6782-­‐6785.   „Enzymatic  synthesis  of  multiple  spin-­‐labeled  DNA.“  

 

Nucleic  Acids  Sym  Ser  (Oxf).,  2008,  

Obeid  S,  Yulikov  M,  Jeschke  G,  &  Marx  A;  

52:373-­‐374.   „Enzymatic  synthesis  of  multi  spin-­‐labeled  DNA.“  

(4)

   

(5)

 

Danksagung    

Die  vorliegende  Arbeit  entstand  in  der  Zeit  von  Nov  2007  bis  Dez  2011  in  der  Arbeitsgruppe  von  Prof.  Dr.  

Andreas  Marx  am  Lehrstuhl  für  Organische  und  Zelluläre  Chemie  im  Fachbereich  Chemie  der  Universität   Konstanz.  

 

In  erste  Linie  möchte  ich  mich  ganz  herzlich  bei  Herrn  Prof.  Dr.  Andreas  Marx  für  die  Vergabe  eines  sehr   vielseitigen   und   interessanten   Promotionsthemas   bedanken.   Insbesondere   möchte   ich   mich   für   die   Betreuung  und  Unterstützung,  die  fordernd  aber  auch  sehr  motivierend  waren,  bedanken.  Hierbei  möchte   ich   vor   allem   das   in   mich   gesetzte   Vertrauen   und   die   Freiheit   zu   selbstständigen   Bearbeitung   und   Gestaltung  des  Themas  erwähnen.  

Ich  möchte  mich  an  dieser  Stelle  auch  für  die  gute  Zusammenarbeit  mit  der  Arbeitsgruppe  von  Prof.  Dr.  

Welte   und   Prof.   Dr.   Diederichs   bedanken.   Ich   wurde   in   Ihrem   Arbeitskreis   herzlich   aufgenommen   und   tatkräftig  unterstützt.  

 

Ich   danke   der   ganzen   Arbeitsgruppe   Marx   für   die   super   Arbeitsatmosphäre.   Mit   jedem   konnte   ich   wissenschaftliche  Fragen  diskutieren  und  alle  waren  dabei  sehr  offen  und  immer  hilfsbereit.  Danke  dafür!  

 

Ein  offenes  Ohr  hatten  vor  allem  meine  Laborkollegen  Sascha  Keller,  Anna  Baccaro  und  Holger  Bußkamp   für   mich.   Hierbei   war   der   Schwierigkeitsgrad   der   Frage,   ob   wissenschaftlicher   Natur   oder   nicht,   immer   zweitrangig.  Ich  bin  euch  dafür  so  unendlich  dankbar.  

 

Ein  besonderer  Dank  gilt  auch  Bastian  Holzberger,  der  mich  schon  während  dem  Studium  begleitet  und   unterstützt  hat.  Ich  werde  nie  das  AC-­‐III  Lernen  bei  Riccardo  Behr  vergessen  J!  Ich  danke  dir  auch  für  das   zahlreiche  Korrekturlesen  von  Paperversionen.  

 

Ich  möchte  mich  auch  bei  Nina  Blatter,  Ramon  Kranaster,  Andreas  Schnur,  Christian  Glöckner  und  Anna   Baccaro  für  die  produktive  Zusammenarbeit,  die  sogar  mit  Publikationen  veredelt  wurde,  bedanken.  

 

Allen  Freunden,  die  ich  hier  leider  nicht  alle  namentlich  erwähnen  kann,  danke  ich  für  die  Unterstützung,   sowie  die  schöne  und  lustige  Zeit,  die  einem  den  Laborstress  auch  mal  vergessen  ließen  J  -­‐  bin  mir  sicher,   dass  sich  die  richtigen  Personen  angesprochen  fühlen!  

 

Meinem  Freund  Christoph  und  meiner  Familie  möchte  ich  für  die  mentale  Stütze  und  die  Ablenkungen,   sowohl   während   meines   Studiums,   wie   auch   während   der   Promotion,   danken.   Ohne   euch   wäre   das   niemals  möglich  gewesen!  

   

   

(6)

   

(7)

Table  of  Contents  

1   Introduction   1  

1.1   History  of  DNA  ...  1  

1.2   DNA  structure  and  characteristics  ...  1  

1.3   DNA  function  ...  2  

1.4   History  of  DNA  polymerases  ...  2  

1.5   DNA  polymerases  features  and  function  ...  3  

1.5.1   Biological  role  of  DNA  polymerases  ...  3  

1.5.2   DNA  polymerase  catalysis  –  the  two  metal  ion  mechanism  ...  5  

1.5.3   DNA  polymerase  fidelity  and  selectivity:  Watson  Crick  base  pairing  vs.  active  site   tightness  ...  6  

1.5.4   DNA  polymerase  as  tool  for  molecular  biology,  biotechnology  and  diagnostics  ...  8  

1.5.5   Model  system  for  sequence  family  A  DNA  polymerases:  KlenTaq  DNA  polymerase  ....  8  

1.6   DNA  lesions  ...  9  

1.6.1   Overview  ...  9  

1.6.2   Abasic  site  ...  10  

1.7   Template  independent  incorporation  at  a  blunt-­‐ended  DNA  duplex  ...  12  

1.8   Functionalized  DNA  ...  12  

1.8.1   Solid  support  synthesis  of  modified  DNA  ...  13  

1.8.2   Enzymatic  Synthesis  of  modified  nucleotides  ...  13  

1.9   Crystallography  ...  14  

1.9.1   Synchrotrons  as  X-­‐ray  source  ...  14  

1.9.2   Protein  crystals  ...  14  

1.9.3   From  data  collection  to  refined  model  ...  14  

1.10   Concepts  and  Objectives  ...  16  

1.10.1   Elucidation  of  lesion  bypass  mechanism  ...  16  

1.10.2   Elucidation  of  the  mechanism  by  which  blunt-­‐ended  DNA  is  elongated  ...  16  

1.10.3   Elucidation  of  process  of  functionalized  nucleotides  by  DNA  polymerases  ...  16  

2   Results  and  Discussion   19   2.1   Abasic  site  bypass  ...  19  

2.1.1  

KlenTaq  follows  the  ‘A-­‐rule’  ...  19

 

2.1.2   Crystal  structure:  KlenTaq  bound  to  ddATP  opposite  an  abasic  site  analog  F   (KlenTaq

F-­‐ddATP

)  ...  19  

2.1.2.1

 

Overall  KlenTaqF-­‐ddATP  structure  ...  21

 

2.1.2.2

 

Active  site  arrangement  of  KlenTaqF-­‐ddATP  ...  22

 

2.1.2.3

 

Tyrosine  671  mimics  the  absent  nucleobase  in  the  template  strand  ...  23

 

(8)

2.1.3   Purine  selection:  preference  of  adenosine  over  guanosine  ...  25  

2.1.4   Crystal  structure:  KlenTaq  bound  to  ddGTP  opposite  an  abasic  site  analog  F   (KlenTaq

F-­‐ddGTP

)  ...  25  

2.1.5   The  unfavored  cases:  Pyrimidine  incorporation  opposite  an  abasic  site  ...  28  

2.1.6   Crystal  structure:  KlenTaq  bound  to  ddTTP  opposite  an  abasic  site  analog  F   (KlenTaq

F-­‐ddTTP

)  ...  28  

2.1.7   Crystal  structure:  KlenTaq  in  presence  of  ddCTP  and  an  abasic  site  analog  F   (KlenTaq

F-­‐ddCTP-­‐binary

)  ...  30  

2.1.8   Crystal  structure:  binary  complex  of  KlenTaq  bound  to  primer/template  construct   containing  an  abasic  site  ...  31  

2.1.9   Stacking  probes:  enhance  the  geometric  fit  to  the  active  site  ...  32  

2.1.10   Crystal  structure:  KlenTaq  bound  to  dNITP  opposite  an  abasic  site  analog  F   (KlenTaq

F-­‐dNITP

)  ...  33  

2.1.11   Discussion  ...  35  

2.1.11.1

 

Abasic  site  bypass  by  different  DNA  polymerases  ...  35

 

2.1.11.2

 

Role  of  Tyr671  in  abasic  site  bypass  by  KlenTaq  DNA  polymerase  and  transfer  to  other  A-­‐ family  DNA  polymerases  ...  36

 

2.1.11.3

 

Features  of  KlenTaqF-­‐ddATP  ...  36

 

2.1.11.4

 

Possible  impact  of  the  ‘A-­‐rule’  ...  37

 

2.1.11.5

 

Relevance  of  the  obtained  KlenTaq  structures  in  presence  of  an  abasic  site  lesion  ...  37

 

2.1.11.6

 

Selectivity  of  adenosine  over  guanosine  opposite  abasic  site  lesions  ...  38

 

2.1.11.7

 

Geometric  constrains  and  distinct  hydrogen  binding  patterns  account  for  the  decrease  in   incorporation  efficiency  of  pyrimidines  opposite  an  abasic  site  lesion  ...  40

 

2.1.11.8

 

Nucleobase  analog  dNITP  lacking  hydrogen  bonding  capability  opposite  an  abasic  site   lesion  ...  40

 

2.1.11.9

 

Relevance  of  the  binary  KlenTaq  structure  in  presence  of  an  abasic  site  ...  41

 

2.2   Template  independent  incorporation  at  a  blunt-­‐end  DNA  duplex  ...  43  

2.2.1   Nucleotide  selectivity  at  a  blunt-­‐end  DNA  duplex  ...  43  

2.2.2   Crystal  structure:  KlenTaq  bound  to  ddATP  at  a  bunt  end  DNA  duplex  ...  44  

2.2.3   Discussion  ...  46  

2.3   Functionalized  nucleotides  enabling  numerous  biomolecular  applications  ...  47  

2.4   Incorporation  of  modified  nucleotide  analogs  ...  47  

2.4.1   Single  nucleotide  incorporation  of  C5  modified  dNTPs  ...  47  

2.4.2   Structure  of  KlenTaq  in  Complex  with  DNA  and  C5  Modified  dNTP.  ...  49  

2.4.3   Structure  of  KlenTaq  in  Complex  with  DNA  and  dT

spin

TP.  ...  51  

2.4.4   Structure  of  KlenTaq  in  Complex  with  DNA  and  dT

dend

TP.  ...  52  

2.4.5   Discussion  ...  55  

2.5   Elongation  of  modified  nucleotide  analogs  ...  57  

(9)

2.5.1   Acceptance  of  dT

alkyne

TP  or  dC

alkyne

TP  by  KlenTaq  DNA  polymerase  ...  58  

2.5.2   Structure  of  KlenTaq  in  complex  with  DNA  and  alkyne  modified  substrates  ...  59  

2.5.2.1

 

Structure  of  KlenTaq  in  complex  with  DNA  and  dTalkyneTP  ...  61

 

2.5.2.2

 

Structure  of  KlenTaq  in  complex  with  DNA  and  dCalkyneTP  ...  62

 

2.5.2.3

 

Structure  of  KlenTaq  in  complex  with  DNA  and  ddCalkyneTP  ...  63

 

2.5.3   Increase  in  incorporation  efficiency  of  the  alkyne  modified  substrates  by  using  a   mutated  KlenTaq  variant  ...  64  

2.5.4   Discussion  ...  65  

3   Conclusive  summary   69   4   Zusammenfassung   73   5   Materials  and  methods   79   5.1   General  ...  79  

5.1.1   Chemicals  and  solvents  ...  79  

5.1.2   Chromatography  ...  79  

5.1.3   Instrumental  and  chemical  analysis  ...  80  

5.1.4   Chemical  DNA  synthesis  ...  80  

5.1.5   Ethanol  precipitation  ...  80  

5.1.6   Determination  of  the  extinction  coefficient  ε  of  modified  nucleotide  analogs  ...  81  

5.2   Chemical  synthesis  ...  81  

5.2.1   5-­‐Nitro-­‐1-­‐indolyl-­‐2′-­‐deoxyriboside-­‐5′-­‐triphosphate  (dNITP)  ...  81  

5.2.2   5’-­‐Tert-­‐butyldiphenylsilyl-­‐2’-­‐deoxyuridine  ...  81  

5.2.3   2’,3’-­‐dideoxyuridine  ...  82  

5.2.4   5-­‐Iodo-­‐2’,3’-­‐dideoxyuridine  ...  82  

5.2.5   5-­‐(2-­‐(4-­‐ethynylphenyl)ethynyl)-­‐2’,3’-­‐dideoxyuridine  (ddT

alkyne

)  ...  83  

5.2.6   5-­‐(2-­‐(4-­‐ethynylphenyl)ethynyl)-­‐2’,3’-­‐dideoxycytidine  (ddC

alkyne

)  ...  83  

5.2.7   5-­‐(2-­‐(4-­‐ethynylphenyl)ethynyl)-­‐2’,3’-­‐dideoxyuridine-­‐5’-­‐triphophate  (ddT

alkyne

TP)  .  84   5.2.8   5-­‐(2-­‐(4-­‐ethynylphenyl)ethynyl)-­‐2’,3’-­‐dideoxycytidine-­‐5’-­‐triphosphate  (ddC

alkyne

TP)   84   5.2.9   3’-­‐O-­‐acetyl-­‐5-­‐(2-­‐(4-­‐ethynylphenyl)ethynyl)-­‐2’-­‐deoxyuridine  (dT

alkyne

)  ...  84  

5.2.10   3’-­‐O-­‐acetyl-­‐5-­‐(2-­‐(4-­‐ethynylphenyl)ethynyl)-­‐2’-­‐deoxycytidine  (dC

alkyne

)  ...  85  

5.2.11   5-­‐(2-­‐(4-­‐ethynylphenyl)ethynyl)-­‐2’-­‐deoxyuridine-­‐5’-­‐triphosphate  (dT

alkyne

TP)  ...  85  

5.2.12   5-­‐(2-­‐(4-­‐ethynylphenyl)ethynyl)-­‐2’-­‐deoxycytidine-­‐5’-­‐triphosphate  (dC

alkyne

TP)  ...  86  

5.3   Molecular  biological  Experiments  ...  86  

5.3.1   General  procedures  ...  86  

5.3.1.1

 

Buffers  and  solutions  ...  86

 

5.3.1.2

 

Gel  electrophoresis  ...  87

 

(10)

5.3.1.3

 

Protein  concentration  determination  by  Bradford  ...  87

 

5.3.1.4

 

Protein  concentration  determination  using  SDS-­‐PAGE  analysis  ...  88

 

5.3.1.5

 

Protein  expression  and  purification  ...  88

 

5.3.1.6

 

Radioactive-­‐labeling  of  primers  ...  88

 

5.3.2   Incorporation  opposite  an  abasic  site  ...  88  

5.3.2.1

 

Primer  extension  assay  ...  88

 

5.3.2.2

 

Pre-­‐steady  state  kinetics  for  incorporation  of  dNITP  opposite  an  abasic  site  ...  89

 

5.3.3   Incorporation  of  functionalized  nucleotides  ...  89  

5.3.3.1

 

Primer  extension  assay  ...  89

 

5.3.3.2

 

Pre-­‐steady  state  kinetics  for  incorporation  of  functionalized  dTRMPs  ...  89

 

5.4   Crystallization  Experiments  ...  90  

5.4.1   General  procedures  ...  90  

5.4.1.1

 

Buffers  and  solutions  ...  90

 

5.4.1.2

 

Gene  construct  of  KlenTaq  DNA  polymerase  ...  90

 

5.4.1.3

 

Site-­‐directed  mutagenesis  ...  90

 

5.4.1.4

 

Protein  expression  and  purification  ...  91

 

5.4.1.5

 

Protein  crystallization  ...  92

 

5.4.1.6

 

Data  collection  ...  93

 

5.4.2   Crystallization  trials  in  the  presence  of  an  abasic  site  ...  93  

5.4.3   Crystallization  trials  at  the  blunt-­‐ended  DNA  ...  95  

5.4.4   Crystallization  trials  in  the  presence  of  a  functionalized  nucleotide  ...  95   5.4.5   Crystallization  trials  with  the  enthynylphenylethynyl  modified  pyrimidine  analogs  96  

6   Appendix   97  

7   Abbreviation   99  

 

(11)

 

1 Introduction  

1.1 History  of  DNA  

In   1944,   Avery,   MacLeod   and   McCarty   were   able   to   isolate   DNA   and   support   “the   belief   that   a   nucleic   acid   of   the   desoxyribose   type   is   the   fundamental   unit   of   the   transforming  principle  of  Pneumocuccus  Type  III.”  and  thereby  established  DNA  as  the   genetic   material   (1).   Almost   one   decade   later   (1953)   Watson   and   Crick   postulated   the  three-­‐dimensional  structure  of  DNA  (2,  3)  (Figure  1)  considering  the  X-­‐ray  data   of  Wilkins  et  al  (4)  and  Franklin  et  al  (5)  as  well  as  the  Chargaff’s  observation  that  

“the  ratio  of  Adenine  to  Thymine  and  Guanine  to  Cytosine  were  nearly  1.0  in  all  species   studied”   (6).   Based   on   chemical   and   stereo-­‐chemical   arguments   Watson   and   Crick   disproved   the   previously   three-­‐chain   models   proposed   by   Pauling   and   Corey   (7).  

Whereas  the  X-­‐ray  data  of  Wilkins  and  Franklin  only  verified  the  helical  structure  and   a   repeat   of   the   polynucleotide   composition,   nearly   three   decades   later   Wing   et   al.  

(1980)   were   able   to   crystallize   and   solve   a   structure   of   a   self-­‐complementary   dodecamer  (PDB-­‐ID:  1BNA)  (8,  9).  These  revolutionary  findings  added  significantly   to  the  understanding  of  DNA  and  to  the  possible  copying  mechanism  of  the  genetic   material.  

1.2 DNA  structure  and  characteristics  

DNA   (deoxyribonucleic   acid)   is   a   polymer   consisting   of   four   monomeric   units   (nucleotides).   The   nucleotides   are   composed   of   a   phosphate,   sugar   and   a   base   moiety,   which   is   N-­‐glycosidic   bound   to   the   sugar   part.   Furthermore,   they   can   be   classified   into   pyrimidines   (thymidine   (T)   and   cytidine   (C))   and   purines  (adenosine  (A)  and  guanosine  (G))  adapted  from  the  respective  nucleobase  (Figure  2A).  Based  on   hydrogen  bonds  the  nucleotides  form  specific  base  pairs.  Thereby,  A  pairs  with  T  and  G  with  C  (Watson-­‐

Crick   base   pairing)   resulting   in   two   base   pairs   with   nearly   the   same   size   (Figure   2B).   Thus   the   three-­‐

dimensional   structure   of   DNA   is   a   double   helix   with   an   alternating   phosphate-­‐sugar   backbone   and   the   base  pairs  in  the  core  perpendicular  to  the  common  helix  axis.  Watson  and  Crick  correctly  concluded  “if   only  specific  pairs  of  the  bases  can  be  formed,  it  follows  that  if  the  sequence  of  bases  on  one  chain  is  given,   than  the  sequence  on  the  other  chain  is  automatically  determined.”  (2).  Not  only  the  ability  to  predict  the   sequence   of   the   complementary   strand   by   a   given   sequence   is   characteristic   for   DNA,   it   also   forms   a   geometrically   well-­‐defined   duplex   structure   with   major   and   minor   groove   (Figure   1).   These   advantageous   characteristics   as   self-­‐assembly,   hybridization   specificity   and   the   formation   of   a   geometrically  well-­‐defined  helical  structure  make  DNA  an  interesting  tool  for  various  applications.    

 

Figure   1   DNA   double   helix   (PDB-­‐

ID:  1BNA)  

(12)

 

1.3 DNA  function  

With   understanding   the   structure   of   DNA   in   1953   simultaneously   the   function   became   obvious.   Since   then,   the   knowledge   that   every   cell   contains   all   the   genetic   information,   allowing   it’s   functioning,   was   established.  DNA  serves  as  carrier  of  genetic  information  and  provides  the  information  to  construct  RNA   molecules   and   proteins.   Thereby   the   sequence   of   these   four   nucleobases   along   a   sugar   phosphodiester   backbone  encodes  the  information  making  DNA  the  “life  molecule”.  

 

1.4 History  of  DNA  polymerases  

Proposing  the  DNA  double  helix  model  Watson  and  Crick  instantly  pointed   out:  “It  has  not  escape  our  notice  that  the  specific  pairing  we  have  postulated   immediately  suggests  a  possible  copying  mechanism  for  the  genetic  material.”  

(2).   The   discovery   of   DNA   structure   was   the   initial   point   for   new   field   in   biology,  which  is  addicted  to  elucidate  the  genetic  code.  In  this  connection   understanding  the  copy  mechanism  played  a  central  role.  Watson  and  Crick   mentioned   already   “Whether   a   special   enzyme   is   required   to   carry   out   the   polymerization,   or   whether   the   single   helical   chain   already   formed   acts   effectively  as  an  enzyme,  remains  to  be  seen.”  (3).  However,  Arthur  Kornberg   (1918-­‐2007),   maybe   the   most   famous   enzymologists   of   that   time,   was   convinced   that   an   enzyme   is   responsible   for   the   copy   mechanism   of   DNA   and  started  to  search  for  it,  which  was  then  named  DNA  polymerase  (DNA   pol).   After   he   identified   dNTPs   as   the   right   substrate   for   the   enzyme   Figure   3   DNA   pol   (PDB-­‐ID:  

1DPI);   shown   are   the   finger-­‐,   thumb-­‐  and  palm-­‐domain.  

Figure  2  (A)  Chemical  structure  of  DNA  buildings  blocks.  (B)  Watson-­‐Crick  base  pairs.  Dashed  lines  indicate   hydrogen  bonding  interaction.  

(13)

performing  the  DNA  synthesis,  he  was  able  to  isolate,  purify  and  characterize  this  enzyme  (10,  11).  Nearly   one   decade   later   Kornberg   succeeded   in   the   synthesis   of   a   viral   DNA   and   demonstrated   that   the   fully   synthetic   circular   DNA   is   still   infectious   (12).   Based   on   Kornberg’s   and   co-­‐workers   effort   another   ten   years  later  other  DNA  polymerases  were  found  and  it  became  obvious  that  the  DNA  replication  process   requires   several   DNA   polymerases.   1985,   Steitz   O.   and   co-­‐workers   presented   the   first   high-­‐resolution   structure   of   the   Klenow   fragment   (or   large   fragment)   of  E.   coli   DNA   pol   I   (13).   Based   on   this   crystal   structure  the  DNA  polymerase  structure  is  associated  with  a  cupped  right  hand  containing  a  finger,  thumb   and  palm  sub-­‐domain  (Figure  3).  The  deep  groove  formed  by  the  three  sub-­‐domains  already  suggests  the   DNA   bining   site.   Subsequent   crystallization   of   DNA   polymerases   in   a   ternary   complex   bound   to   a   primer/template   duplex   and   an   incoming   nucleotide   followed.   Together   with   functional   studies   the   crystal  structure  helped  to  elucidate  the  basic  mechanism  of  nucleotide  incorporation  by  DNA  pols.  

1.5 DNA  polymerases  features  and  function  

1.5.1 Biological  role  of  DNA  polymerases  

DNA  polymerases  catalyze  all  DNA  synthesis  occurring  in  nature,  which  can  be  categorized  in  three  main   processes   -­‐   DNA   replication,   repair   and   recombination   (14).   Replicative   polymerases   -­‐   members   of   the   sequence  family  A  and  B  -­‐  copy  a  template  strand  by  selectively  incorporating  nucleotide  monophosphates   to   the   3’-­‐primer   terminus   (15,   16).   As   mentioned   before,   the   specific   structure   of   complementary   DNA   double  strands  enables  the  copying  of  genetic  information  in  a  template  directed  manner  (Figure  4).  

The  replisome,  a  multimeric  protein  complex,  facilitates  the  replication  in  E.  coli  comprising  enzymes  and   proteins  with  different  functions  (17-­‐19).  Starting  from  the  oriC  (origin  of  replication)  the  genomic  DNA  is   unwound   by   the   ATP-­‐dependant   DnaB   helicase.   The   resulting   two   single   DNA   strands   are   stabilized   by   SSB   proteins   (single   strand   binding   proteins)   (20).   The   unwinding   of   the   helical   structure   triggers   the   release   of   topoisomerases   (21).   Next,   short   RNA   oligonucleotides   (primers)   of   roughly   12   nucleotides  

Figure  4  Scheme  of  enzyme  catalyzed  DNA  synthesis  in  5’→3’  direction  

(14)

complementary  to  the  template  strand  are  synthesized  by  a  specialized  RNA  polymerase  DnaG  and  DnaB   form   the   so-­‐called   primosome   (22).   Then   the   DNA   polymerase   III   (pol   III)   holoenzyme   complex   is   assembled   at   the   replication   fork   containing   ten   different   proteins.   These   proteins   can   be   grouped   into   three   major   groups:   (i)   the   catalytic   core   of   pol   III,   (ii)   the   sliding   β-­‐clamp,   and   (iii)   the   γ-­‐clamp   loader   (19).   Synthesis   of   a   new   DNA   strand   always   proceeds   in   5’→3’   direction,   thereby   the   5’-­‐end   of   the   incoming   nucleotides   are   coupled   to   the   free   3’-­‐hydroxyl   group   of   the   growing   primer   strand.   The   existence   of   two   core   units   in   the   holoenzyme   ensures   simultaneous   replication   of   both   single   stands.  

Upon   binding   of   the   β-­‐clamp   the   processivity   –   the   incorporation   number   of   nucleotides   before   dissociation   –   is   drastically   increased   allowing   the   polymerization   of   more   than   50   kb   before   the   next   dissociation  event  occurs  (19,  20).  The  β-­‐clamp  is  a  ring-­‐like  protein  dimer  encompassing  the  DNA  strand   tethering   the   core   enzyme   to   the   DNA.   Since   DNA   synthesis   can   only   take   place   in   5’→3’   direction,   the   simultaneous  copying  of  both  parental  template  strands  cannot  be  conducted  in  a  continuous  way  on  both   strands.  While  the  replisome  is  moving  along  the  DNA,  pushing  the  replication  fork  forward,  the  leading   strand   is   copied   continuously,   whereas   the   lagging   strand   is   copied   in   fragments   of   roughly   1000   nucleotides  (Okazaki  fragments)  (23,  24).  The  synthesis  of  lagging  strand  points  away  from  the  replication   fork.  After  completion  of  an  Okaziaki  fragment  the  pol  III  has  to  dissociate  from  its  template  and  bind  to  a   freshly  unwound  single-­‐stranded  template  section.  To  ensure  the  disassembly  and  reassembly  of  the  β-­‐

clamp  the  γ-­‐clamp  loader  is  essential  and  in  this  way  leading  pol  III  to  a  new  primer/template  complex   (19,   20).   Each   Okazaki   fragment   starts   with   an   RNA   primer,   which   is   removed   by   the   intrinsic   5’-­‐3’-­‐

endonuclease  activity  of  E.  coli  DNA  polymerase  I  (pol  I)  after  replication.  Upon  binding  of  pol  I  to  the  3’-­‐

end  of  an  Okazaki  fragment  the  RNA  primer  are  successive  hydrolyzed  and  causes  simultaneous  synthesis   of   DNA.   In   this   way   RNA   primers   are   translated   into   DNA   (nick   translation).   The   remaining   nicks   are   removed  by  ligases  catalyzing  the  formation  of  phosphodiester  bonds  between  the  fragments.  

Eucaryotes   contain   at   least   15   different   DNA   pols   and   replication   is   facilitated   in   a   similar   manner.  

However,  two  different  DNA  polymerases  –  DNA  polymerase  δ  (pol  δ)  and  DNA  polymerase  ε  (pol  ε)  –  are   mainly  responsible  for  the  replication  (20).  It  remains  to  be  elucidated  which  polymerase  is  copying  the   leading   and   which   the   lagging   strand.   Several   studies   suggest   that   pol   ε   copies   the   leading   strand   in   Saccharomyces   cervisiae   (16,   25).   Eucaryotic   DNA   polymerase   α   (pol   α)   is   part   of   the   primosome   and   essential   for   initiation   of   replication,   analog   to   the   primase   DnaG   in   prokaryots.   Other   important   eukaryotic  DNA  polymerases  are  pol  β,  which  closes  gaps  resulting  from  the  repair  of  DNA  lesions  during   base   excision   repair   (BER)   (26),   and   pol   γ,   which   is   responsible   for   mitochondrial   replication   (27,   28).  

Further  eucaryotic  DNA  polymerases  play  a  role  in  DNA  repair  and  translesion  bypass  like  pol  ζ,  η,  θ,  ι,  κ,  λ,   μ,  σ  and  φ,  which  have  been  discovered  during  the  last  years  (26,  29,  30).  

             

(15)

1.5.2 DNA  polymerase  catalysis  –  the  two  metal  ion  mechanism  

 

DNA  polymerases  incorporate  nucleotides  following  a  simplified  five-­‐step-­‐

procedure.  The  first  step  is  the  binding  of  the  primer/template  duplex  to   the  DNA  polymerase  forming  the  binary  complex  (Figure  5  step  1).  In  a   second   step   the   2’-­‐deoxynucleoside   triphosphate   (dNTP)   binds   to   the   complex   resulting   in   the   ternary   complex   (step   2).   Upon   nucleotide   binding  discrimination  for  canonic  and  mispaired  nucleotides  take  place.  

Thereby  the  correctly  pairing  nucleotides  usually  display  a  higher  affinity   to   the   complex.   Furthermore,   nucleotide   binding   triggers   a   conformational  change  of  the  polymerase  binding  pocket  (step  3).  At  this   level   the   so-­‐called   ‘induced   fit’   mechanism   (see   chapter   1.5.3)   provides   further  discrimination  for  the  correct  Watson-­‐Crick  base  pair.  It  leads  to  a   tighter   enclosure   of   the   dNTP   in   a   pocket   shaped   to   fit   the   correct   nucleotide  excluding  water  from  the  active  site  (31).  This  conformational   change,  which  includes  the  closure  of  the  finger  domain,  was  discussed  to   be  the  rate-­‐limiting  step  in  DNA  synthesis.  However,  recent  studies  have   shown   that   local   reorganizations   in   the   active   site   are   the   rate-­‐limiting   factor  (32).  The  closed  conformation  of  the  DNA  pol  places  the  reactive  3’-­‐

hydroxyl  group  of  the  primer  in  an  ideal  position  to  enable  a  nucleophilic  

attack   to   the   5’-­‐α-­‐phosphate   group   of   the   incoming   dNTP.   Two   divalent   metal   ions   (usually   Mg2+)   are   octahedrally  coordinated  by  three  highly  conserved  side  chains  in  the  active  site  and  the  triphosphate  of   the   incoming   nucleotide.   Metal   ion   A   facilitates   the   deprotonation   of   the   3’-­‐OH   group,   promoting   the  

Figure   5   Kinetic   model   of   nucleotide   incorporation.   The   various   complexes   are   indicated  as  mentioned  in  the  text.  The  rate  constant  of  the  rate-­‐limiting  step:  k3  is   indicated.   E   =   DNA   polymerase.   Graphic   adapted   from   Rothwell   and   Waksman   2005  (32).  

Figure   6   The   two-­‐metal-­‐ion   mechanism   of   polynucleotide   polymerases   adapted   from   Steitz,   1998   (34).   Two   divalent   metal   ions,   A   and   B   are   coordinated  by  two  aspartic-­‐acid   residues   in   the   active   site   (here   D705   and   D882   for  E.   coli   DNA   polymerase   I).   Water   molecules   bound   to   the   metal   ion   A   are   shown  as  filled  black  circles.  

(16)

nucleophilic   attack   of   3’-­‐O-   on   the   α-­‐phosphate   (step   4)   (33).   Metal   ion   B   is   mainly   coordinated   by   the   triphosphate  moiety  as  well  as  the  aspartic  acid  residues  assisting  the  release  of  pyrophosphate  (Figure   6).  Both  metal  ions  stabilize  the  structure  and  charge  of  the  trigonal  bipyramidal  transition  state  of  the  α-­‐

phosphate   during   the   nucleophilic   substitution   (SN2   reaction)   (34-­‐37).   Therefore,   the   mechanism   is   known   as   the   ‘Two   metal   ion   mechanism’.   The   phosphoryl   transfer   is   completed   by   the   release   of   pyrophosphate,   which   is   subsequently   hydrolyzed   in   aqueous   solution   pushing   the   equilibrium   to   the   product   side.   Then   the   polymerase   switches   back   into   the   open   conformation   (step   5).   Afterwards,   translocation  of  the  polymerase  along  the  template  strand  and  incorporation  of  a  second  nucleotide  can   occur,  otherwise  dissociation  from  the  primer/template  complex  can  take  place  or  –  if  present  –  edit  the   just   incorporated   nucleotide   with   intrinsic   or   extrinsic   exonuclease   proof-­‐reading   activity.   Thereby   the   adjacent  steps  are  reversibly  connected.  

1.5.3 DNA   polymerase   fidelity   and   selectivity:   Watson   Crick   base   pairing   vs.  

active  site  tightness  

The   accuracy   of   DNA   synthesis   is   crucial   for   the   maintenance   of   the   genome   stability.   Therefore,   replicative  DNA  polymerases  are  high  fidelity  DNA  polymerases  with  low  error  rates.  Studies  found  that   replication  in  E.  coli  and  bacteriophages  displays  a  base  substitution  error  rate  of  10−7–10−8  per  nucleotide   in  vivo  in  the  absence  of  mismatch  repair  (38).  The  error  rate  can  be  improved  to  10−8–10−10  in  E.  coli  by   proof-­‐reading,   mismatch   repair   and   numerous   other   factors   (16,   39)   making   DNA   synthesis   a   highly   accurate  process.  In  general,  the  fidelity  and  selectivity  of  DNA  polymerases  is  linked  to  their  biological   function   and   the   organism   from   which   they   are   derived.   Translesion   synthesis   DNA   polymerases   show   relative  low  fidelity  e.g  an  error  rate  of  approximately  1/10,  meaning  a  remarkable  one  in  ten  error  rate,   for  pol  η  (eta)  (40).  Even  more  remarkable  is  pol  ι  (iota),  which  inserts  in  vitro  G  opposite  T  rather  than  A   opposite  T  (41).  The  difference  in  fidelity  begs  the  question:  What  are  the  determinants  of  fidelity  of  these   enzymes?    

The   elucidated   structure   of   DNA   by   Watson   and   Crick   immediately   suggested   a   copy   mechanism,   since   only  specific  base  pairs  are  formed.  Thereby,  the  idea  was  established  that  nucleotide  selectivity  of  DNA   polymerases   is   manifested   in   their   hydrogen   bonding   capacity   and   the   formation   of   the   correct   base   pairing   according   to   Watson   and   Crick   (Figure   2).   Thus   the   Watson-­‐Crick   hydrogen   bonds   between   canonic  base  pairs  have  been  thought  to  account  for  the  high  accuracy  in  DNA  synthesis  (2,  3).  However,   the  small  free  energy  barriers  between  machted  and  mismachted  base  pairs  showed  that  the  selectivity  of   DNA   polymerases   depends   not   primarily   on   hydrogen   bonding   capability.   Moreover,   the   base   pair   geometry  contributes  to  the  DNA  polymerase  selectivity.  The  active  site  of  a  DNA  polymerase  is  designed   in  way  to  accept  Watson-­‐Crick  base  pairs  or  base  pairs  imitating  the  geometry  of  such  a  base  pair.  

 

(17)

   

For   instance,   extensive   studies   with   designed   isosteric   nucleotide   analogs   lacking   hydrogen   bonding   capability,   but   showing   enhanced  stacking  capacity  have  shown  that  efficient  and  selective   DNA   synthesis   is   possible   (37,   42-­‐46).   With   these   steric   probes   in   hand   one   could   show   that   Watson-­‐Crick   hydrogen   bonds   are   not   the   only   important   factor   assigning   selectivity.   In   addition   to   hydrogen   bond   interaction   between   the   minor   groove   and   the   protein,   base   stacking   and   solvation   effects,   but   especially   sterical   effects   are   taken   into   account   to   explain   the   highly   accurate   performance   of   DNA   polymerases.   Based   on   these   and   other   (47)   findings   Kool   postulated   the   model   of   ‘active   site   tightness   and   substrate   fit   in   DNA   replication’   (37,   42,   43,   48).   Therefore,   he   defined   first   the   active   site   binding   pocket.   The   analysis   of   crystal   structures   suggests   that   the   active   site   of   selective   polymerases   forms  a  tight  binding  pocket  whose  geometry  is  complementary  to   the  respective  canonical  dNTP  (49-­‐54)  depending  on  the  respective   template   base   (42,   55).   In   addition,   it   could   be   shown   that   the   canonical   base   pairs   only   slightly   differ   in   their   shape   and   size   (Figure   7and  Figure   8).   Their   geometric   constraints   show   minor   alterations  in  the  minor  and  major  groove  defining,  but  no  variation   in   the   over-­‐all   length   (48,   56)   (Figure   8A),   suggesting   a   size   exclusion   hypothesis   that   a   base   pair   must   fit   into   the   consensus  

base   pair   shape.   Finally,   the   incoming   nucleotide   placed   opposite   the   templating   nucleobase   have   to   fit   into  the  geometric  constrains  defined  by  the  minor  and  major  groove  sides,  as  illustrated  by  the  consensus   base  pair  shape  (Figure  8B).  In  the  case  of  non-­‐canonical  base  pairing  steric  clashes  can  occur.  Thereby  

Figure  7  Schematic  diagram  illustrating  the  space-­‐filling  shapes  of  the   four   base   pairs   in   isomorphous   orientation.   Graphic   is   adapted   from   Kool  2002  (42).  

Figure   8   (A)   Schematic   diagram   showing   the   overlay   of   the   four   base   pair  shapes.  The  variability  are  marked   by  arrows  at  sides  of  major  groove  and   in  center  of  minor  groove.  R  represents   deoxyribose   and   phosphodiester   backbone.   (B)   Overlay   showing   the   consensus   largest   dimensions   along   the   outer   surface.   The   graphic   is   adapted  from  Kool  2002  (42).  

(18)

either  the  incoming  nucleotide  can  not  enter  the  binding  pocket  or  if  it  does  partially  insert  itself  it  will  not   allow   the   triphosphate   moiety   to   be   aligned   correctly   for   efficient   phosphodiester   bond   formation.  

However,  in  this  scenario  one  should  not  forget  that  the  tightness  of  the  active  site  binding  pocket  is  only   defined  in  the  closed  conformation  of  the  enzyme.  The  closure  of  the  active  site  is  likely  only  for  correctly   shaped  pairs,  thus  the  steric  clashes  resulting  from  non-­‐canonical  base  pairing  might  prevent  the  closed   conformation,  rather  than  the  reverse,  where  the  formed  active  site  prevents  the  steric  clashes.  

However,   the   theory   of   active   site   tightness   does   not   exclude   the   contribution   of   further   non-­‐covalent   interactions,  such  as  base  stacking,  it  only  completes  the  picture,  how  DNA  polymerases  perform  accurate   DNA  synthesis.  

1.5.4 DNA   polymerase   as   tool   for   molecular   biology,   biotechnology   and   diagnostics  

Functional  studies  added  significantly  to  the  understanding  of  DNA  polymerase  reaction  and  elucidated  an   universal   basic   reaction   mechanism   (34).   For   efficient   catalysis   the   DNA   polymerase   needs   (i)   the   four   natural  nucleotides,  (ii)  a  DNA  template  directing  the  incorporation  events,  (iii)  a  DNA  primer  hybridizing   to  the  template  strand  and  harboring  a  3’-­‐OH  group,  (iv)  a  divalent  metal  ion  e.g.  magnesium  as  cofactor,   and  (v)  in  some  cases  DNA  pol  auxiliary  proteins  such  as  PCNA  (proliferating  cell  nuclear  antigen).  With   understanding   how   DNA   polymerases   function   their   applications   in   research   area   such   as   genetics   or   diagnostics  explored.  The  major  breakthrough  was  the  proposed  concept  of  the  polymerase  chain  reaction   (PCR)   by   Saiki   and   Mullis   in   1985   (57-­‐59).   After   several   rounds   of   optimization,   the   method   enables   in   vitro   amplification   of   target   DNA   sequences   by   using   thermostable   DNA   polymerases,   2’-­‐

deoxyribonucleotides   and   short   oligonucleotides   as   primers.   Thereby,   exponential   amplification   of   the   original  genetic  material  is  achieved  by  successive  temperature  cycles  causing  DNA  denaturation,  primer   annealing   and   DNA   polymerization   (57,   58).   Because   of   this   application   DNA   polymerases   reaped   the   following  title  ‘The  Molecule  of  the  Year’  in  1989  (60,  61).  

To   date,   numerous   PCR-­‐based   methods   cater   for   specialized   needs.   The   possibility   to   amplify,   modify,   analyze   or   tag   DNA   by   simple   experimental   set-­‐ups   has   been   of   great   benefit   in   the   fields   of   genetics,   medicine  and  diagnostics.  Further  progression  in  this  direction  elucidated  the  time-­‐resolved  amplification   of  genetic  material  in  real-­‐time  PCR,  also  known  as  quantitative  PCR  (62)  or  the  use  of  error  prone  PCR  to   create  mutant  protein  libraries  (63).  

1.5.5 Model   system   for   sequence   family   A   DNA   polymerases:  

KlenTaq   DNA  

polymerase  

The  large  fragment  of  Thermus  aquaticus  (Taq)  DNA  polymerase  (in  short  KlenTaq,  N-­‐terminally  truncated   form  of  Taq  polymerase  (aa  293-­‐832))  is  the  ortholog  of  DNA  polymerases  I  from  E.  coli  and  belongs  to  the   A-­‐family   of   DNA   polymerases.  KlenTaq   DNA   polymerase   shows   the   characteristic   right-­‐hand   shaped   structure  consisting  of  a  finger,  thumb  and  palm  sub-­‐domain.  DNA  polymerase  I  enzymes  are  involved  in   nucleotide   excision   repair   and   in   the   processing   of   Okazaki   fragments   in   procaryotes.   With   a   reaction   temperature  optimum  of  75-­‐80°C  KlenTaq  DNA  polymerase  is  applicable  in  various  experiments  such  as   PCR.   Since   this   enzyme   class   is   heavily   employed   and   well   characterized   on   a   functional   and   structural   level  (32,  52,  54,  64-­‐67),  it  is  used  in  this  study  as  a  model  system  for  DNA  polymerases  from  sequence  

(19)

family  A.  G.  Waksman  and  coworkers  were  able  to  crystallize  KlenTaq  DNA  polymerase  in  apo  form  as  well   as   in   complex   with   suitable   substrates   (52,   54,   65).   Based   on   these   structures   the   enzyme   substrates   interactions  in  the  different  reaction  states  during  nucleotidyl  transfer  are  well  known.  Starting  with  DNA   binding  the  thumb  domain  changes  forming  a  cylinder-­‐like  crevice,  which  almost  completely  surrounds   the   primer/template   complex   (binary   complex:   pdb   4KTQ).   This   fact   explains   the   very   low   DNA   dissociation  rate.  By  the  entrance  of  a  nucleotide,  the  finger  domain  switches  from  the  open  conformation   (open  ternary  complex:  pdb  2KTQ)  to  the  closed  from  (closed  ternary  complex:  pdb  3KTQ).  Thereby  the  O   helix  of  the  finger  domain  packs  against  the  nascent  base  pair  and  closes  the  active  site.  A  tight  binding   pocket  is  formed  aligning  all  components  required  for  catalysis.  After  the  nucleotidyl  transfer  reaction  the   KlenTaq  DNA  polymerase  changes  back  to  the  open  confromation  releasing  the  pyrophosphate,  followed   by  translocation  of  the  polymerase  along  the  DNA.  

 

1.6 DNA  lesions  

1.6.1 Overview  

Endo-­‐  and  exogenous  agents  constantly  damage  DNA.  For  instance,  exposure  to  UV  radiation,  alkylating   agents  and  oxidative  species  leads  to  the  formation  of  abasic  sites,  pyrimidine  dimers,  alkylated  adducts   and  oxidative  lesion  products.  To  maintain  the  genomic  integrity  and  reduce  the  mutagenic  potential  cells   allocate  with  multiple  repair  pathways  and  specialized  enzymes.  However,  several  health  statistics  could   show  that  DNA  lesions  can  be  highly  mutagenic  and  sometimes  carcinogenic  e.g.  in  Europe  in  2000  ∼35   000   new   cases   of   UV   radiation   damage-­‐induced   skin   cancer   were   diagnosed   (68).   Further,   the   tobacco-­‐

derived  nitrosamine  NNK  is  associated  with  lung  cancer  resulting  in  ∼334  800  deaths  in  Europe  in  2006   (69).   Therefore   the   biological   prevalence   of   the   DNA   lesions   and   their   chemical   structures   need   to   be   determined.   The   main   aspect   here   are   (i)   identification   and   quantification   of   DNA   lesions   in   model   systems  and  in  vivo,  (ii)  to  assess  influences  of  lesions  on  physical  properties  of  DNA  e.g.  thermal  stability,   and  (iii)  to  elucidate  the  impact  of  the  lesions  on  DNA  function  e.g.  enzyme-­‐mediated  processes  such  as   replication.    

Within   the   last   decade   years   specialized   DNA   polymerases,   responsible   for   translesion   DNA   synthesis   (TLS),   were   identified   and   characterized.   W.   Yang   and   R.   Woodgate   published   a   clear   summary   of   this   class  of  enzymes  emphasizing  the  relationship  of  the  bypass  properties  and  the  structural  features  (29).  In   brief,   many   of   the   TLS   enzymes   are   member   of   the   Y-­‐family   of   DNA   polymerases   exhibiting   universal   features   to   manage   bypass   of   a   variety   of   DNA   lesions.   In   a   simplified   model   TLS   polymerases   can   be   categorized  into  two  classes.  The  first  class  of  enzyme  is  highly  specialized  and  responsible  for  bypassing   a  certain  DNA  lesion  e.g.  the  human  pol  η  is  able  to  bypass  thymine-­‐thymine  cyclobutane  dimer  with  high   efficiency.   Interestingly,   patients   showing   mutations   or   defects   in   the   human   pol   η   gene   suffer   from   sunlight-­‐sensitive   and   cancer-­‐prone  Xeroderma   pigmentosum   variant   (XP-­‐V)   syndrome   (70,   71).   The   second  class  of  enzymes  is  the  all-­‐rounder  and  has  the  ability  to  accommodate  different  DNA  lesions  e.g.  

the  archaeal  Dpo4  DNA  polymerase  from  the  Y-­‐family.  A  series  of  structural  studies  show  this  low  fidelity   polymerase   bound   to   damaged   substrates   such   as   oxidative   damage   (72,   73),   UV   cross-­‐linking   (74),  

(20)

benzo-­‐[a]pyrene  diol  epoxide  adduct  (BPDE)  (75),  and  abasic  site  lesions  (76).  However,  efficient  catalysis   is  mainly  observed  in  case  of  an  abasic  site  lesion(76,  77).  

 

1.6.2 Abasic  site  

The   most   common   DNA   damage   under   physiological   conditions   are   abasic   sites   resulting   mainly   from   spontaneous  hydrolysis  of  the  N-­‐glycosidic  bond  between   the   sugar   moiety   and   the   nucleobase   in   DNA   (78).   Abasic   sites  also  occur  as  intermediates  during  excision  repair  of   damaged  nucleotides  (79)  or  can  be  manifested  in  several   chemical   structures   such   as   C4’-­‐oxidized   abasic   site   (C4-­‐

AP)  after  treatment  of  DNA  with  antitumor  antibiotics  like   bleomycin   (80,   81).   The   abasic   site   L   (2’-­‐

deoxyribonolacetone)  results  from  one-­‐electron  nucleotide   oxidation   (82,   83).   In   general,   it   has   been   estimated   that   10000  abasic  sites  are  formed  in  human  cell  per  day  (78,  

84,  85).  Guanine  and  adenine  nucleobases  are  cleaved  most  efficiently  resulting  in  the  abasic  sugar  moiety   (AP,  Figure  9A).  To  investigate  the  biochemical  impact  of  AP  a  stabilized  tetrahydrofuran  analog  is  used   as  a  model.  

Since   the   genetic   information   gets   lost   by   the   cleavage   of   the   nucleobase,   abasic   sites   bear   a   high   mutagenic  potential  (85-­‐87).  To  face  this  problem  nature  offers  a  whole  arsenal  of  enzymes  and  possible   pathways.  In  most  cases,  the  lesion  is  removed  by  DNA  repair  systems  using  the  sister  strand  to  guide  for   incorporation  of  the  right  nucleotide.  However,  undetected  lesions  or  those,  formed  during  S  phase,  pose  a   challenge   to   DNA   polymerases   and   block   replication   (26,   88).   Additionally,   it   was   found   that   the   mutagenic  potential  of  these  lesions  in  translesion  synthesis  is  more  pronounced  in  animal  compared  with   bacterial  cells  presumably  because  of  higher  translesion  synthesis  in  eukaryotes  (87,  89,  90).  

A  set  of  studies  concerning  the  behavior  of  DNA  polymerases,  belonging  to  different  families,  showed  that   there  are  multiple  mechanisms  to  overcome  an  abasic  site.  Most  translesion  DNA  polymerases  from  family   X   and   Y   follow   various   loop   out   mechanisms   (76,   77,   91-­‐94).   Thereby,   the   nucleotide   selection   is   influenced   by   the   following   upstream   templating   bases   resulting   in   deletions   and   complex   mutation   spectra.  Recently,  an  amino  acid  templating  mechanism  was  found  for  the  “error-­‐free”  bypass  of  an  abasic   site   by   the   yeast   Rev1   DNA   polymerase   belonging   to   the   family   Y   (95).   Since   guanine   is   cleaved   most   efficiently   (85),   the   preference   of   Rev1   for   dCMP   incorporation   opposite   an   abasic   site   represents   the  

“best-­‐guess”.    

In   contrast,   in   vitro   and   in   vivo   studies   of   the   replicative   DNA   polymerases   from   family   A   (including   human  DNA  polymerases  γ  and  θ)  and  B  (including  human  DNA  polymerases  α,  ε  and  δ)  in  the  presence  of   the   stabilized   tetrahydrofuran   abasic   site   analog   F   (Figure   9D)   have   shown   that   purines,   in   particular   adenosine,   and   to   a   lesser   extent   guanosine,   are   most   frequently   incorporated   opposite   the   lesion.   The   strong  preference  for  adenosine  incorporation  opposite  an  abasic  site  has  been  termed  ‘A-­‐rule’  (89,  91,   Figure  9   Structures   of   different  forms   of   abasic   DNA  lesions.  

(21)

96-­‐104).   The   apparent   selectivity   for   incorporation   of   purines   ultimately   results   in   transversion   mutations  commonly  found  in  human  cancers  (86).  

However,   the   determinants   of   the   ‘A-­‐rule’   are   still   controversially   discussed.   Structural   and   functional   studies  have  added  significantly  to  our  understanding  of  the  basic  mechanisms  of  translesion  synthesis  by   DNA  polymerases  (29,  105).  Since  in  the  canonical  case  several  points  account  for  the  high  selectivity  of   DNA   polymerases.   Besides   Watson-­‐Crick   base   paring,   stacking   interaction   and   solvation   properties   the   induced   fit   to   the   active   site   of   an   enzyme   plays   a   paramount   role.   Sine   the   Watson-­‐Crick   recognition   cannot  take  place  in  the  presence  of  an  abasic  site  it  seems  obvious  that  the  other  properties  account  for   the   selectivity   opposite   an   abasic   site.   Therefore,   superior   stacking   as   well   as   solvation   properties   of   adenine  have  been  discussed  to  be  the  driving  force  behind  the  adenine  selection  (47,  99,  106,  107).  Based   on  this  assumption  numerous  of  non-­‐natural  nucleotide  analogs  were  studied  regarding  their  behavior  in   the  presence  of  an  abasic  site.  If  the  induced  fit  model  is  taken  as  a  selection  criteria  opposite  abasic  sites,   a  non-­‐natural  nucleotide  analog  with  nearly  identical  size  to  the  Watson-­‐Crick  base  pair,  should  show  the   highest  incorporation  efficiency.  By  steric  examination  Matray  and  Kool  identified  the  pyrene  nucleoside   triphosphate  (dPTP)  as  a  perfect  match  in  the  absence  of  a  templating  base  (Figure  10)  (47).  Indeed,  they   could   show   that   the   pyrene   modified   nucleotide   is   incorporated   by   DNA   polymerase   I   from  E.   coli   with   higher  efficiency  than  any  other  natural  nucleotide,  demonstrating  that  a  simple  steric  model  is  sufficient   for   efficient   incorporation.   Further   the   fluorescent   nucleobase   analog   is   used   to   identify   and   sequence   abasic   site   lesions   in   DNA.   Studies   of   several   nucleotide  

analogs   identified   5-­‐nitro-­‐1-­‐indoyl-­‐nucleotide   (dNITP)   as   the  ‘specific  partner’  opposite  an  abasic  site,  since  dNIMP  is   incorporated   with   increased   efficiency   by   RB69   DNA   polymerase,   a   α-­‐like   DNA   polymerase,   compared   to   dPMP   (106).  The  structure  of  RB69  DNA  polymerase  capturing  an   artificial   5-­‐nitro-­‐1-­‐indoyl-­‐nucleotide   (dNITP)   opposite   an   abasic  site  in  the  active  site  of  the  enzyme  elucidated  that  a   dipole-­‐induced   dipole   stacking   interaction   between   the   5-­‐

nitro   group   and   base   3′   to   the   templating   lesion   might   explain   the   increase   in   incorporation   efficiency.   These   findings  might  suggest  that  base  stacking  is  likely  to  have  a   paramount   role   in   the   selective   incorporation   of   dAMP   opposite  abasic  sites  (106).  DNA  polymerases  from  family  A   and  B,  which  are  involved  in  the  majority  of  DNA  synthesis   in   DNA   replication   and   repair,   follow   the   A-­‐rule   when   bypassing   abasic   sites.   Up   to   now   only   a   structure   from   RB69  DNA  polymerase,  a  member  of  the  B  sequence  family,   was   reported   showing   an   incoming   guanosine   opposite   an   abasic   site   lesion   (97).   This   structure   is   a   hybrid   structure   between  the  enzyme  in  replicating  and  editing  or  apo  mode.  

The   approach   to   crystallize   RB69   DNA   polymerase  

Figure  10  (A)  Chemical  Structures  of  dPTP  and   the   abasic   site   analog   F.   (B)   Space-­‐filling   models   of   the   A-­‐T   (top)   and   the   P-­‐F   (bottom)   base  pairs  in  B-­‐form  geometry,  illustrating  the   steric  fit  of  pyrene  opposite  an  abasic  site.  The   graphic  is  adapted  from  Matray  and  Kool  1999   (47).  

Referenzen

ÄHNLICHE DOKUMENTE

In accordance with the hypothesis that the selectivity of nucleotide insertion during DNA replication is achieved by editing of nucleotide geometry within a tight binding

Lane P: primer only; lane 1: primer extension by KlenTaq in the presence of dATP, dGTP and dTTP; lane 2: all four natural dNTPs; lane 3: the same as lane 2 followed by incubation

A, aetive site assembly in the presenee of an abasie site analog F: The naseent base pair of Tyr-671 and ddGTP (KlenTaq F_G_I) is depieted_ The surfaee ofthe

When all four dNTPs are present in the primer extension experiment, KF- is able to bypass all three types of uridine modifications (1PydU, 2PydU, and BodU) but not the modified

In all single nucleotide incorporation experiments it became clear that the mutants incorporate a nucleotide opposite an abasic site more preferentially than the

the detailed molecular mechanism by which CS modifications are processed by a DNA polymerase is poorly understood, Here, we present the first crystal structures

Figure 1. Structures of KlenTaq ternary complexes. a) Overall structures of KlenTaq (ribbon representation) containing primer, template, dT Me TP (yellow), and dT Et TP

We were able to obtain several crystals and could solve the structure of a ternary complex of KlenTaq bound to an abasic site, containing a primer, template, and an incoming 2 0 ,3 0