26.04.2013 Views

David William Haxton Girdwood PhD Thesis - University of St Andrews

David William Haxton Girdwood PhD Thesis - University of St Andrews

David William Haxton Girdwood PhD Thesis - University of St Andrews

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

DB0>C2B8@D8?>0< B46E =4380D43 D7B?E67 D74<br />

2?>9E60D8?> 0>3 342?>9E60D8?> ?5 D74 C=0$3 CE=?$*<br />

3GZOJ FOQQOGR 7G\XTS 6OVJ[TTJ<br />

0 DNKWOW CYHROXXKJ LTV XNK 3KMVKK TL @N3<br />

GX XNK<br />

ESOZKVWOX] TL CX% 0SJVK[W<br />

)''+<br />

5YQQ RKXGJGXG LTV XNOW OXKR OW GZGOQGHQK OS<br />

BKWKGVIN/CX0SJVK[W.5YQQDK\X<br />

GX.<br />

NXXU.&&VKWKGVIN$VKUTWOXTV]%WX$GSJVK[W%GI%YP&<br />

@QKGWK YWK XNOW OJKSXOLOKV XT IOXK TV QOSP XT XNOW OXKR.<br />

NXXU.&&NJQ%NGSJQK%SKX&('')*&),)-<br />

DNOW OXKR OW UVTXKIXKJ H] TVOMOSGQ ITU]VOMNX


Transcriptional Regulation Mediated through the Conjugation<br />

and Deconjugation <strong>of</strong> the Small Ubiquitin-Like Modifiers<br />

SUMO-1, SUMO-2, and SUMO-3<br />

<strong>David</strong> <strong>William</strong> <strong>Haxton</strong> <strong>Girdwood</strong><br />

A <strong>Thesis</strong> Presented for the Degree <strong>of</strong><br />

Doctor <strong>of</strong> Philosophy<br />

School <strong>of</strong> Biological and Medical Sciences<br />

<strong>University</strong> <strong>of</strong> <strong>St</strong>. <strong>Andrews</strong><br />

June 2004<br />

UNii/FRs'<br />

ý, ýEGR<br />

o,<br />

ýý'ýýREANý


DECLARATION<br />

I, <strong>David</strong> <strong>William</strong> <strong>Haxton</strong> <strong>Girdwood</strong>, hereby certify that this thesis, which is<br />

approximately 40,000 words in length, has been written by me, that it is a<br />

record <strong>of</strong> work carried out by me and that it has not been submitted in any<br />

previous application for a higher degree.<br />

Date<br />

\\ý<br />

Signature <strong>of</strong> candidat<br />

I was admitted as a research student in September 2000 and as a<br />

candidate for the degree <strong>of</strong> Doctor <strong>of</strong> Philosophy in September 2000: the<br />

higher study for which this record was carried out in the faculty <strong>of</strong> Sciences <strong>of</strong><br />

low<br />

the <strong>University</strong> <strong>of</strong> <strong>St</strong>. <strong>Andrews</strong> between 2000 and 2003.<br />

Date Signature <strong>of</strong> candidate<br />

I hereby certify that the candidate has fulfilled the conditions <strong>of</strong> the<br />

Resolution and Regulations appropriate for the degree <strong>of</strong> Doctor <strong>of</strong> Philosophy<br />

in the <strong>University</strong> <strong>of</strong> <strong>St</strong>. <strong>Andrews</strong> and that the candidate is qualified to submit<br />

this thesis in application for that degree.<br />

}..... ý'<br />

Date<br />

.<br />

Signature <strong>of</strong> supervisor .................................<br />

2


In submitting this thesis to the <strong>University</strong> <strong>of</strong> <strong>St</strong>. <strong>Andrews</strong> I understand<br />

that I am giving permission for it to be made available for use in accordance<br />

with the regulations <strong>of</strong> the <strong>University</strong> Library for the time being in force,<br />

subject to any copyright vested in the work not being affected thereby. I also<br />

understand that the title and abstract will be published, and that a copy <strong>of</strong> the<br />

work may be made and supplied to any bona fide library or research worker.<br />

Date Signature <strong>of</strong> candidate<br />

3


DECLARATION<br />

CONTENTS<br />

CONTENTS<br />

....................................................................................................................... 2<br />

............................................................................................................................... 4<br />

List <strong>of</strong> Figures ............................................................................................................................. 8<br />

List <strong>of</strong> Tables ............................................................................................................................... 12<br />

List <strong>of</strong> Abbreviations ................................................................................................................... 13<br />

Abbreviations for Amino Acids and Side-chains ........................................................................ 16<br />

Genetic Code ............................................................................................................................... 17<br />

ACKNOWLEDGEMENTS<br />

18<br />

.....................................................................................................<br />

ABSTRACT 19<br />

...............................................................................................................................<br />

1. INTRODUCTION 20<br />

................................................................................................................<br />

1.1. GENE EXPRESSION<br />

21<br />

.......................................................................................................<br />

1.1.1. Post-translational modification ................................................................................. 21<br />

1.2. UBIQUITIN-LIKE PROTEIN MODIFIERS .................................................................... 24<br />

1.2.1 Ubiquitin .................................................................................................................... 30<br />

. 1.2.1.1. The Proteasome ..................................................................................................... 36<br />

1.2.2. NEDD8 ......................................................................................................................... 38<br />

1.2.3. ISG15 39<br />

.................................................................................................................<br />

1.2.4. FAT10 39<br />

................................................................................................................<br />

1.2.5. Atg12 & Atg8 40<br />

.....................................................................................................<br />

1.2.6. URM1 ............................................................................................................................ 41<br />

1.2.7. HUB1 ............................................................................................................................. 42<br />

1.2.8. UFM1 ............................................................................................................................ 42<br />

1.3 SUMO .................................................................................................................................. 42<br />

1.3.1. Localisation & Nuclear domains ................................................................................. 49<br />

1.3.2. PML bodies ................................................................................................................... 49<br />

1.3.3. SUMO and genome integrity ........................................................................................ 56<br />

1.3.4. SUMO and transcriptional regulation ......................................................................... 58<br />

1.3.5. SUMO and neurodegenerative diseases ...................................................................... 59<br />

4


1.4. UBIQUITIN AND UBIQUITIN-LIKE PROTEASES: - CONTROL OVER<br />

CONJUGATION ......................................................................................................................... 60<br />

1.4.1. SUMO specific proteases and PML bodies ................................................................. 63<br />

1.5. COMPETION FOR LYSINE RESIDUES ........................................................................ 65<br />

1.6. THE UBL CONJUGATION PATHWAY ......................................................................... 67<br />

1.6.1. The El superfamily 66<br />

.......................................................................................................<br />

1.6.2. The E2 superfamily 74<br />

.......................................................................................................<br />

1.6.3. E3 ligases 77<br />

.......................................................................................................................<br />

1.6.3.1. Ubiquitin E3s 77<br />

...........................................................................................................<br />

1.6.3.2. SUMO E3s 79<br />

................................................................................................................<br />

1.7. p300/CB P 84<br />

............................................................................................................................<br />

1.7.1. p300/CBP and growth control and signal transduction .............................................. 85<br />

1.7.2. p300? CBP and disease ................................................................................................. 89<br />

1.7.3. p300/CBP and chromatin remodelling ........................................................................ 90<br />

1.7.4. An alternate role for p300 as an E3 ligase ................................................................... 91<br />

1.7.5 p300/CBP and transcriptional repression .................................................................... 95<br />

1.8. AIMS OF THE PROJECT<br />

95<br />

..................................................................................................<br />

2. MATERIALS & METHODS 97<br />

.............................................................................................<br />

2.1. MATERIALS 98<br />

......................................................................................................................<br />

2.2. ANTIBODIES 98<br />

.......................................................................................................:.............<br />

2.3. BACTERIAL STRAINS<br />

............................................................................. I......................<br />

98<br />

2.4. DNA PREPARATION 99<br />

.......................................................................................................<br />

2.4.1. cDNA cloning ................................................................................................................ 99<br />

2.4.1.1. Preparation <strong>of</strong> thermocompetent bacteria .............................................................. 99<br />

2.4.1.2. Transformation <strong>of</strong> thermocompetent bacteria ........................................................ 100<br />

2.4.1.3. Generated plasmid constructs & siRNA oligonucleotides ..................................... 100<br />

2.4.1.3.1. Cullin-2 .............................................................................................................. 101<br />

2.4.1.3.2. Synthetic siRNA oligonucleotides targeting SUMO-1/-2/-3 ............................ 101<br />

2.4.1.3.3. pSUPER retro siRNA expression plasmids ....................................................... 101<br />

2.4.2. DNA sequencing<br />

102<br />

............................................................................................................<br />

2.5. EXPRESSION & PURIFICATION OF UNLABELED RECOMBINANT PROTEINS<br />

102<br />

.....................................................................................................................................................<br />

2.6. QUANTITATION OF PROTEIN 104<br />

......................................................................................<br />

5


2.7. SDS/PAGE & WESTERN BLOT ANALYSIS<br />

................................................................ 104<br />

2.8. STRIPPING OF WESTERN BLOTS 105<br />

.............................................................................<br />

2.9. IN VITRO EXPRESSION OF PROTEINS ..................................................................... 105<br />

2.10. IN VITRO SUMO CONJUGATION ASSAY 105<br />

.............................................................<br />

2.11 IN VITRO SUMO DECONJUGATION ASSAY 106<br />

..........................................................<br />

2.12. PROTEIN INTERACTION ASSAY 106<br />

............................................................................<br />

2.13. MAMMALIAN CELL CULTURE TRANSIENT TRANSFECTIONS ..................... 108<br />

2.13.1. Calcium phosphate transient transfections & reporter gene assays ..................... 108<br />

2.14. PURIFICATION OF HIS6-TAGGED PROTEINS 109<br />

.....................................................<br />

2.15. INDIRECT IMMUNOFLUORESENCE ANALYSIS OF CELLS 109<br />

.............................<br />

2.16. LUCIFERASE ASSAY 110<br />

.................................................................................................<br />

3. RESULTS 111<br />

...........................................................................................................................<br />

3.1. BIOINFORMATICS: AN APPROACH TO IDENTIFY NOVEL MEMBERS OF THE<br />

UBIQUITIN-LIKE PROTEIN SUPERFAMILY 112<br />

...................................................................<br />

3.1.1. Identification <strong>of</strong> three potential novel members <strong>of</strong> the SUMO family ...................... 112<br />

3.1.2. SUMO-1 b/4/5 are conjugated both in vivo and in vitro ........................................... 113<br />

3.1.3. Deconjugation <strong>of</strong> SUMO proteins via a SUMO specific protease ........................... 117<br />

3.1.4. SUMO-1 b/4/5 may be psuedogenes<br />

120<br />

...........................................................................<br />

3.1.5. Discussion ................................................................................................................... 122<br />

3.2. SUMO MODIFICATION OF THE TRANSCRIPTION REGULATOR p300............ 124<br />

3.2.1. p300 is modified by SUMO in vivo 124<br />

............................................................................<br />

3.2.2. SUMO conjugation to the N-terminus <strong>of</strong> p300 in vitro ............................................. 126<br />

3.2.3. In vitro SUMO modification <strong>of</strong> the N-terminus <strong>of</strong> p300 does not require an E3..... 126<br />

3.2.4. Identification <strong>of</strong> the lysine residues, within the CRDI domain <strong>of</strong> p300, that are<br />

conjugated to by SUMO 131<br />

........................................................................................................<br />

3.2.5. SUMO modification is required for CRD1 activity in vivo ...................................... 137<br />

3.2.6. SUMO modified CRDI recruits a histone deacetylase<br />

140<br />

.............................................<br />

3.2.7. Conclusion<br />

145<br />

..................................................................................................................<br />

3.3. TRANSCRIPTIONAL REPRESSION IS PREFERENTIALLY MEDIATED<br />

THROUGH SUMO-2 AND SUMO-3 148<br />

....................................................................................<br />

3.3.1. Ablation <strong>of</strong> SUMO-1 & SUMO-2/3 expression following treatment with synthetic<br />

siRNA oligonucleotides ........................................................................................................ . 149<br />

6


3.3.2. Transcriptional repression is relieved preferentially-through siRNA targeting <strong>of</strong><br />

SUMO-2/3 .............................................................................................................................. 149<br />

3.3.3. The nuclear localisation <strong>of</strong> p80 coilin is differentially regulated by knockdown <strong>of</strong><br />

SUMO-1 compared to that <strong>of</strong> SUMO-2/3 ............................................................................. 157<br />

3.3.4. Conclusion .................................................................................................................. 161<br />

4. Discussion<br />

5. APPENDIX - VECTOR MAPS<br />

6. BIBLIOGRAPHY 193<br />

.............................................................................................................<br />

7. PUBLICATIONS 248<br />

..............................................................................................................<br />

164<br />

172<br />

7


List <strong>of</strong> Figures<br />

FIGURE 1. SCHEMATIC REPRESENTATION OF THE HISTONE CODE MODE ......... 23<br />

FIGURE 2. PHYLOGENIC TREE SHOWING THE EVOLUTIONARY RELATIONSHIP<br />

BETWEEN THE Ub-LIKE FAMILY OF PROTEINS 26<br />

.............................................................<br />

FIGURE 3. SCHEMATIC REPRESENTATION OF THE SUMO-1/2/3 AND UBIQUITIN<br />

CONJUGATION PATHWAYS 27<br />

.................................................................................................<br />

FIGURE 4. A RIBBON REPRESENTATION OF THE 3D STRUCTURE OF<br />

UBIQUITIN, A COMMON CHARACTERISTIC OF THE Ubls, AND A COMPARISON<br />

OF THE PREDICTED SECONDARY STRUCTURES OF SUMO-1, UBIQUITIN,<br />

NEDD8, AND UFM1 29<br />

.................................................................................................................<br />

FIGURE 5. SCHEMATIC REPRESENTATION OF UBIQUITIN, WITH THE<br />

MODIFIABLE LYSINE RESIDUES AND THE ROLES OF THE DIFFERENTLY<br />

LINKED POLYMERIC CHAINS 32<br />

.............................................................................................<br />

FIGURE 6. SCHEMATIC REPRESENTATION OF THE NF-KB PATHWAY, AND THE<br />

ROLES PLAYED BY Ub, THROUGH ITS ABILITY TO FORM POLYMERIC CHAINS<br />

WITH DIFFERENT INTERNAL LYSINE RESIDUES 34<br />

..........................................................<br />

FIGURE 7. ALIGNMENTS OF THE SEQUENCES FOR A NUMBER OF Ub-LIKE<br />

PROTEINS, AND THE SPACEFILL 3D MODELS OF SUMO-1, UBIQUITIN, NEDD8,<br />

AND REPRESENTATIONS OF SUMO-2 AND SUMO-3' BASED ON THE STRUCTURE<br />

OF SUMO-1 45<br />

................................................................................................................................<br />

FIGURE 8. COMPOSITION AND REULATION OF PML BODIES 50<br />

...................................<br />

8


FIGURE 9. SCHEMATIC REPRESENTATION OF PROTEIN ASSOCIATION WITHIN<br />

PML BODIES, WITH THE POTENTIAL MECHANISMS FOR TRANSCRIPTIONAL<br />

REGULATION<br />

........................................................................................................................... 54<br />

FIGURE 10. SCHEMATIC REPRESENTATION OF THE ACTIVE SITE OF A<br />

CYSTEINE PROTEASE, WITH A SCHEMATIC ILLUSTRATION OF A NUMBER OF<br />

SUMO-SPECIFIC PROTEAES INDICATING THE CONSERVED CORE DOMAIN AND<br />

RESIDUES. A SEQUENCE ALIGNMENT OF A NUMBER OF SUMO-SPECIFIC AND<br />

SUMO-SPECIFIC LIKE PROTEASES, INDICATING THE RESIDUES INVOLVED IN<br />

THE FORMATION OF THE PROTEASE ACTIVE SITE ...................................................... 62<br />

FIGURE 11. SCHEMATIC REPRESENTATION OF A GENERAL UBIQUITIN-LIKE<br />

PROTEIN CONJUGATION PATHWAY 69<br />

.................................................................................<br />

FIGURE 12. SCHEMATIC DIAGRAM SHOWING THE FOUR CONSERVERD<br />

REGIONS IN Ub AND Ub-LIKE E1 ACTIVATING ENZYMES, AND A PHYLOGENIC<br />

TREE DEPICTING THEIR EVOLUTIONARY RELATIONSHIPS 70<br />

......................................<br />

FIGURE 13. DIAGRAMMATIC REPRESENTATION OF THE SCF AND VHL E3<br />

UBIQUITIN LIGASE COMPLEXES 80<br />

.......................................................................................<br />

FIGURE 14. DIAGRAMMATIC REPRESENTATION OF p300 AND THE SIGNALING<br />

PATHWAYS WHICH ITS INVOLVED IN, AND THE KNOWN PROTEIN<br />

INTERACTORS WITH p300 86<br />

....................................................................................................<br />

FIGURE 15. SCHEMATIC REPRESENTATION OF THE p53 FUNCTIONAL CIRCUIT,<br />

INDICATING THE MECHANISMS INVOLVED IN THE CONTROL OVER ARF,<br />

HDM2, E2F-1, AND p53 91<br />

............................................................................................................<br />

FIGURE 16. DIAGRAMMATIC REPRESENTATION OF THE CHROMOSOMAL<br />

LOCATIONS OF THE SUMO GENES, A SUMMARY OF THEIR GENE MAKEUP,<br />

AND AN ALIGNMENT OF THE SEQUENCES OF SUMO-1/2/3 WITH SUMO-lb/4/5<br />

108<br />

..................................................................................................................................................<br />

9


FIGURE 17. RT-PCR OF SUMO-lb/4/5, AND TRANSFECTIONS OF HA-SUMOs INTO<br />

COST CELLS, SHOWING THEIR FUNCTIONAL STATUS 110<br />

.............................................<br />

FIGURE 18. IN VITRO CONJUGATION OF SUMO-lb/4/5 112<br />

..............................................<br />

FIGURE 19. SSP3 DECONJUGATES SUMO-lb AND SUMO-5 113<br />

.....................................<br />

FIGURE 20. RT-PCR OF THE NOVELS SUMO GENES WITH TAQ POLYMERASE<br />

115<br />

................................................................................................................................................<br />

FIGURE 21. p300 AND CBP CONTAIN SUMO CONSENSUS SITES IN A<br />

REPRESSION DOAMAIN, AND p300 IS MODIFIED, WITHIN THIS DOMAIN IN VIVO<br />

119<br />

..................................................................................................................................................<br />

FIGURE 22. p300 IS MODIFIED BY SUMO-1/2/3 IN VITRO<br />

121<br />

...........................................<br />

FIGURE 23. TITRATION OF UBC9 INTO THE p300 CONJUGATION ASSAY ........... 123<br />

FIGURE 24. ADDITION OF KNOWN SUMO E3s INTO p300 CONJUGATION ASSAY..<br />

124<br />

..................................................................................................................................................<br />

FIGURE 25. IDENTIFICATION OF THE RESIDUES REQUIRED FOR SUMO<br />

CONJUGATION 126<br />

......................................................................................................................<br />

FIGURE 26. SUMOs CONJUGATION CORRELATES WITH TRANSCRIPTIONAL<br />

REPRESSION<br />

130<br />

..........................................................................................................................<br />

FIGURE 27. TRANSCRIPTIONAL REPRESSION IS DEPENDENT ON SUMO<br />

CONJUGATION 132<br />

......................................................................................................................<br />

FIGURE 28. CONTROL OVER SUMO MODIFICATION OF p300 134<br />

.................................<br />

FIGURE 29. SUMO MODIFICATION OF CRDI RECRUITS A HISTONE<br />

DEACETYLASE 137<br />

.....................................................................................................................<br />

10


FIGURE 30. KNOCKDOWN OF ENDOGENOUS SUMO BY siRNA, AS SHOWN BY<br />

WESTERN BLOT ANALYSIS 143<br />

..............................................................................................<br />

FIGURE 31. KNOCKDOWN OF ENDOGENOUS SUMO BY siRNA, AS SHOWN BY<br />

IMMUNOFLUORESCENCE 144<br />

.................................................................................................<br />

FIGURE 32. AFFECT OF SUMO KNOCKDOWN ON p300 ACTIVITY 146<br />

........................<br />

FIGURE 33. AFFECT OF SUMO KNOCKDOWN ON ELK-I ACTIVITY 147<br />

.....................<br />

FIGURE 34. AFFECT OF SUMO KNOCKDOWN ON SP3 ACTIVITY 148<br />

..........................<br />

FIGURE 35. AFFECT OF SUMO KNOCKDOWN ON GAL4 ACTIVITY<br />

FIGURE 36. SUMO-1 AND SUMO-2/3 PLAY DIFFERENT ROLES IN THE<br />

...................... 149<br />

LOCALISATION OF p80 COILIN 152<br />

........................................................................................<br />

FIGURE 37. SCHEMATIC REPRESENTATION OF A NUMBER OF TRANSCRIPTION<br />

CO/FACTORS INDICATING DOMAINS AND SUMO CONJUGATION SITES ........... 158<br />

FIGURE 38. MODEL FOR TRANSCRIPTIONAL REPRESSION BY SUMO ................ 161<br />

11


LIST OF TABLES<br />

TABLE 1. THE DIFFERENT UBIQUITIN-LIKE PROTEINS IN SACCHAROMYCES<br />

CEREVISIAE AND HOMO SAPIENS ...................................................................................... 25<br />

TABLE 2. SUMMARY OF THE KNOWN SUBSTRATES FOR SUMO MODIFICATION<br />

46<br />

.....................................................................................................................................................<br />

TABLE 3. SUMMARY OF THE RELATIONSHIP BETWEEN SUMO CONJUGATION<br />

AND TRANSCRIPTIONAL REPRESSION 130<br />

.........................................................................<br />

TABLE 4. LYSINE RESIDUES WITHIN A SUMO CONSENSUS MOTIF THAT ARE<br />

MODIFIED BY UBIQUITN IN YEAST 163<br />

...............................................................................<br />

12


List <strong>of</strong> Abbreviations<br />

Ac Acetylation<br />

ADP Adenosine diphosphate<br />

AMP Adenosine monophosphate<br />

APL Acute promyelocytic leukaemia<br />

APP-BPI Amyloid precursor protein-binding protein 1<br />

Atg Autophagy<br />

ATP Adenosine triphosphate<br />

BLAST Basic local alignment search tool<br />

bp Base pairs<br />

BPV Bovine papilloma virus<br />

BSA Bovine serum albumen<br />

CDK Cyclin depedent Kinase<br />

CBP CREB-binding protein<br />

cDNA Complementary DNA<br />

CMV Cytomegalo virus<br />

DeUb De-ubiquitination<br />

D-MEM Dulbecco's modified Eagle's medium<br />

DNA Deoxyribonucleic acid<br />

DTT Dithiothrietol<br />

DUB Deubiquitination enzymes<br />

E. coli<br />

Eschericia coli<br />

El Activating enzyme<br />

E2 Conjugating enzyme<br />

E3 Ligase enzyme<br />

E6-AP E6 activating enzyme<br />

ECL Enhanced chemiluminesence<br />

EDTA Ethylenediaminetetracetic acid<br />

ERK Extracellular signal-regulated kinase<br />

FAT10 Diubiquitin<br />

FCS Foetal calf serum<br />

GST Glutathione S-transferase<br />

13


h Hours<br />

.<br />

HA Haemagglutinin<br />

HAT Histone acetyl-transferase<br />

HCl Hydrochloric acid<br />

HDAC Histone deacetylase<br />

HDM2 Human double minute 2<br />

HECT Homology to E6-AP C-terminus<br />

HIPK Homeodomain-interacting protein kinase<br />

HLH Helix-loop-helix<br />

HPV Human papilloma virus<br />

HUB I Homologous to ubiquitin 1<br />

Ig Immunglobulin<br />

IKB Inhibitor kappa B<br />

IPTG Isopropyl-ß-D-thiogalactopyranose<br />

ISG15 Interferon induced gene <strong>of</strong> 15 kDa<br />

IVTT In vitro transcription and translation<br />

KCl Potassium chloride<br />

kDa Kilo Dalton<br />

LB I Luria-Bertani broth<br />

MDM2 Mouse double minute 2<br />

Me Methylation<br />

min minute<br />

MonoUb Mono-ubiquitination<br />

MW Molecular weight<br />

NBs PML nuclear bodies<br />

NEDD8 Neuronal precursor cell expressed developmentally<br />

Down regulated protein 8<br />

NES Nuclear export signal<br />

NF-KB Nuclear Factor-Kappa B<br />

NLS Nuclear localisation signal<br />

NP-40 Nonidet P-40<br />

NPC Nuclear pore complex<br />

PAGE Polyacrylamide gel electrophoresis<br />

PBS Phosphate buffered saline<br />

PCR Polymerase chain reaction<br />

Pi Phosphate<br />

PIAS Protein inhibitor <strong>of</strong> activated STAT<br />

14


PML Promyelocytic leukaemia protein<br />

PODS PML oncogenic domains<br />

PolyUb Poly-ubiquitination<br />

PVDF Polyvinylidene difluoride<br />

RanGAP Ran GTPase activating protein<br />

RanBP Ran binding protein<br />

RAR Retinoic acid receptor<br />

RNA Ribonucleic acid<br />

RT-PCR Reverse transcriptase PCR<br />

Rubl Related to ubiquitin<br />

SAE1 SUMO activating enzyme subunit 1<br />

SAE2 SUMO activating enzyme subunit 2<br />

S. cerevisiae Saccharomyces cerevisiae<br />

SCF Skpl, cullin, F-box<br />

SENP Sentrin-specific protease<br />

SSP SUMO-specific protease<br />

STAT Signal transducer and activator <strong>of</strong> transcription<br />

SUMO Small ubiquitin-like modifier<br />

SUSP SUMO-specific protease<br />

SV40. Simian virus type 40<br />

3D Three-dimensional<br />

Tris 2-amino-2-(hydroxymethyl)propane-1,3-diol<br />

Ub Ubiquitin<br />

UBA1 Ubiquitin activating enzyme<br />

Ubc Ubiquitin conjugating enzyme<br />

UBL Ubiquitin-like modifiers<br />

UBP Ubiquitin-specific processing proteases<br />

UCH Ubiquitin C-terminal hydrolase<br />

UD Ubiquitin domain<br />

UDP Ubiquitin-domain protein<br />

UFM 1 Ubiquitin fold modifier 1<br />

ULP Ubiquitin-like protein<br />

Ulp Ubiquitin-like protease<br />

UV Ultra violet<br />

URM 1 Ubiquitin related modifier 1<br />

VCB VHL-Elongin C/ Elongin B<br />

VHL Von Hippel-Lindau<br />

Wrn Werner's disease protein<br />

WT Wild-type<br />

15


Abbreviations for Amino Acids and Side-chains<br />

Alanine ala A<br />

Arginine arg R<br />

Asparagine asn N<br />

Aspartic acid asp D<br />

Cysteine cys C<br />

Glutamine gln Q<br />

Glutaminc acid glu<br />

Glycine gly G<br />

Histidine his H<br />

Isoleucine He I<br />

Leucine leu L<br />

Lysine lys K<br />

E<br />

-CH3<br />

-(CH2)3-NH-CNH-NH2<br />

-CH2-CONH2<br />

-CH2-000H<br />

-CH2-SH<br />

-CH2-CH-CONH2<br />

-CH2-CH2-000H<br />

-H<br />

-CH2-imidazole<br />

-CH(CH3)-CH2-CH3<br />

-CH2-CH(CH3)2<br />

-(CH2)4-NH2<br />

Methionine met M -CH2-CH2-S-CH3<br />

Phenylalanine phe F<br />

-CH2-phi<br />

Proline pro P -[N]-(CH2)3-[CH]<br />

Serine ser S -CH2-OH<br />

Threonine thr T -CH-(CH3)-OH<br />

Tryptophan try w -CHZ-indole<br />

Tyrosine tyr Y -CH2-phi-OH<br />

Valine val v -CH-(CH3)2<br />

16


Genetic Code<br />

TTT'phe F TCT ser S TAT tyr Y TGT cys C<br />

TTC phe F TCC ser S TAC tyr Y TGC cys C<br />

TTA leu L TCA ser S TAA OCH Z TGA OPA Z<br />

TTG leu L TCG ser S TAG AMB Z TGG try W<br />

CTT leu L CCT pro P CAT his H CGT arg R<br />

CTC leu L CCC pro P CAC his H CGC arg R<br />

CTA leu L CCA pro P CAA gln Q CGA arg R<br />

CTG leu L CCA pro P CAG gln Q CGG arg R<br />

ATT ile I ACT thr T AAT asn N AGT ser S<br />

ATC ile I ACC thr T AAC asn N AGC ser S<br />

ATA ile I ACA thr T AAA lys K AGA arg R<br />

ATG met M ACG thr T AAG lys K AGG arg R<br />

GTT val V GCT ala A GAT asp D GGT gly G<br />

GTC val V GCC ala A GAC asp D GGC gly G<br />

GTA val V GCA ala A GGA glu E GGA gly G<br />

GTG val V GCG ala A GAG glu E GGG gly G<br />

17


ACKNOWLEDGEMENTS<br />

I would like to thank my supervisor, Ron Hay, for sharing with me his<br />

enthusiasm for the SUMO field, and providing me with opportunities,<br />

encouragement, and support throughout my <strong>PhD</strong>, both with my work and in<br />

developing my own ideas and interests. The work' within the thesis benefited<br />

greatly from the excellent help and assistance freely given by members <strong>of</strong> the<br />

lab and the BMS building as a whole, both past and present, and I gratefully<br />

acknowledge their help and assistance throughout my time here. I would<br />

additionally like both Graham Kemp and Rick Randall for their time and help<br />

in the lab both before and during my <strong>PhD</strong>.<br />

To my family and friends, special thanks, for your support and advice<br />

throughout, it made a real difference.<br />

Finally I would like to thank <strong>St</strong>ephen and Charlotte, their cats Haggis<br />

and Pepper, for both letting me stay in their home, and for making me feel so<br />

welcome, especially as the "just for a short while" turned into months.........<br />

This work was supported by the Biotechnology and Biological Sciences<br />

Research Council (BBSRC).<br />

18


ABSTRACT<br />

SUMO-1/2/3 are members <strong>of</strong> the ubiquitin-like family <strong>of</strong> protein<br />

modifiers. These proteins are covalently attached to numerous proteins in a<br />

directed and controlled manner. SUMO conjugation primarily occurs to<br />

proteins containing an exposed SUMO conjugation motif, (I, V, L, F)KxE,<br />

where x represents any amino acid. SUMO conjugation is controlled by key<br />

enzymes, a SUMO activating enzyme, SAE1/2 and a SUMO conjugating<br />

enzyme, Ubc9, which is responsible for substrate recognition, and the<br />

efficiency <strong>of</strong> this pathway can be increased by one <strong>of</strong> many SUMO ligase<br />

enzymes. This modification alters the substrate's characteristics and results in<br />

a change <strong>of</strong> state, such as stability, localisation, or activity.<br />

p300, a transcriptional co-activator, contains an evolutionary<br />

conserved tandem SUMO modification motif, located in a transcriptional<br />

repression domain. p300 was efficiently conjugated, both in vitro and in vivo,<br />

by SUMO-1/2/3, within this repression domain to both SUMO conjugation<br />

motifs. The SUMO conjugation to p300 correlated with p300 ability to repress<br />

transcription, requiring both SUMO conjugation motifs for full transcription<br />

repression activity. This repression activity was mediated through SUMO<br />

recruitment <strong>of</strong> histone deacetylase 6. Repression could be alleviated through<br />

co-expression <strong>of</strong> a SUMO-specific protease thereby suggesting a potential<br />

regulatory mechanism for transcription control <strong>of</strong> SUMO modified substrates.<br />

Despite utilising the same conjugation machinery, there remained the<br />

potential for distinct roles for the SUMO is<strong>of</strong>orms. SUMO -2/3, which form a<br />

distinct group from SUMO-1, were shown to preferentially mediate the<br />

transcription repression abilities <strong>of</strong> a number <strong>of</strong> known SUMO modifiable<br />

substrates: p300, Elk-1, and SP3. Further differences were observed in the<br />

ability <strong>of</strong> SUMO-1 and SUMO-2/3 to influence the nuclear organisation <strong>of</strong><br />

p80 coilin.<br />

19


1. INTRODUCTION<br />

20


1.1 Gene Expression<br />

1.1.1 Post-translational modification<br />

The specialization <strong>of</strong> an organism's cells depends upon the ability to<br />

control the expression <strong>of</strong> particular genes. Gene expression is controlled at six<br />

different levels: transcription; RNA processing; RNA transport; translation;<br />

mRNA turnover; and protein turnover. At the first level <strong>of</strong> control, that <strong>of</strong> gene<br />

transcription, determines how frequently a particular gene is transcribed and as<br />

such dictates the level <strong>of</strong> primary RNA transcript produced. The last level <strong>of</strong><br />

control, achieved through the modification <strong>of</strong> a protein, can determine the<br />

consequences <strong>of</strong> gene expression. The gene product can be selectively activated,<br />

inactivated, degraded, or compartmentalised following its translation. This<br />

regulation will affect a proteins capacity to interact with other proteins or DNA,<br />

and as such the proteins function. A protein, for instance, that is rapidly<br />

degraded may have little cellular impact although its gene may be expressed.<br />

This form <strong>of</strong> control, known as post-translational modification, can be achieved<br />

in numerous ways, phosphorylation, hydroxylation, glycosylation, and<br />

acetylation amongst others. Addition <strong>of</strong> these chemical groups to a target protein,<br />

is achieved through the activity <strong>of</strong> specific enzymes while removal <strong>of</strong> the<br />

chemical groups is mediated by other distinct enzymes. This represents a<br />

common method <strong>of</strong> controlling a proteins function. A more advanced system <strong>of</strong><br />

post-translational modification utilises the attachment <strong>of</strong> one protein to another,<br />

via an enzymatic cascade. Post-translational modification via the attachment <strong>of</strong><br />

other proteins provides unique mechanisms <strong>of</strong> controlling protein function.<br />

Ünlike chemical modification such as phosphorylation, adenylation, or<br />

glycosylation, the size <strong>of</strong> the protein attached can exert a direct influence on the<br />

proteins immediate environment, ubiquitin being approximately one hundred<br />

times bigger than a phosphate group. An example is histone ubiquitinylation<br />

(Histone code hypothesis) and this size also creates the addition <strong>of</strong> a more<br />

chemically complex surface, creating the possibility <strong>of</strong> further interactions with<br />

additional proteins. Histones are the fundamental constituents <strong>of</strong> the<br />

nucleosome. Nucleosomes consist <strong>of</strong> genomic DNA wrapped around an'octamer<br />

<strong>of</strong> the conserved histone proteins H2A, H2B, H3, and H4. H1 functions as linker<br />

histones, located between nucleosomes, which serve to promote the overall<br />

21


I<br />

folding <strong>of</strong> the chromatin fibre. Histones consist <strong>of</strong> a globular domain, a short C-<br />

terminal extension, and an extended N-terminal-tail (Figure l. i. ). Post-<br />

translational modifications <strong>of</strong> the tail domains <strong>of</strong> histones modulate chromatin<br />

structure and gene expression (Naar et al., 2001; Berger et al., -2002). This is<br />

achieved through the control <strong>of</strong> the accessibility ' <strong>of</strong> the gene expression<br />

machinery to the DNA. DNA is compacted within the nucleus <strong>of</strong> a eukaryotic<br />

cell, due to spatial limitations and as a means <strong>of</strong> protecting DNA (reviewed in<br />

Felsenfeld and<br />

Groudine, 2003).<br />

Histone acetylation is the best characterised post-translational<br />

modification <strong>of</strong> the histone tails. Histone acetyltransferases (HAT) acetylate<br />

specific lysine residues present in the tail regions, associated with transcription<br />

activation, conversely histone deacetylases (HDAC) reverse the acetylation <strong>of</strong><br />

lysine residues, and are associated with transcription repression. These distinct<br />

histone modifications can act sequentially or in combination- the histone code<br />

(reviewed in <strong>St</strong>rahl and Allis, 2000).<br />

The histone-code hypothesis proposes that post-translational<br />

modifications in histone tails are `read' by other histones and/or proteins, and are<br />

translated into silencing or activation <strong>of</strong> gene transcription (Jenuwein et al.,<br />

2001). For example, the phosphorylation <strong>of</strong> SerlO (S10) <strong>of</strong> histone H3<br />

antagonizes methylation <strong>of</strong> K9 and/or K14 (Figure l. ii. ). The modifications <strong>of</strong><br />

the histone tails will, under the hypothesis, regulate the interactions <strong>of</strong> proteins,<br />

such as Chromo and Bromo domain containing proteins, that specifically bind to<br />

methylated or acetylated lysine residues, respectively, in histones (Sun et al.,<br />

2002).<br />

Research into histone ubiquitination has provided much evidence in<br />

support <strong>of</strong> the histone-code hypothesis. Ubiquitination <strong>of</strong> K123 <strong>of</strong> H2B has been<br />

demonstrated to be a prerequisite for H3 methylation <strong>of</strong> K4 in yeast. H3<br />

methylation <strong>of</strong> K4 is a key event in the regulation <strong>of</strong> gene expression in yeast<br />

(Dover et al., 2002). The mutation <strong>of</strong> K123 <strong>of</strong> H2B results in the disruption <strong>of</strong><br />

gene silencing, indicating the links between the ubiquitination <strong>of</strong> K123, mediated<br />

by Rad6, K4 methylation <strong>of</strong> H3 by Setl, and transcriptional repression (Sun et<br />

al., 2002; Dover et al., 2002). Rad6 mediated ubiquitination <strong>of</strong> K123 on H2B<br />

has further been demonstrated to be a requirement for K79 methylation <strong>of</strong> H3,<br />

22


I.<br />

Histone tail<br />

-K9S1O-K14-R17-K18-K23-K27-K3<br />

Radio<br />

- ------------------<br />

Brei<br />

H2B B<br />

Globular<br />

domain<br />

Mb<br />

H2B K 123 ý`ý<br />

Figure 1. The histone-code model. In this model the histone tails extend to the left <strong>of</strong> the<br />

globular domain, ubiquitin (ub) is indicated in red. (i) Mono-ubiquitination <strong>of</strong> histone H2B is<br />

mediated by the E2 ubiquitin conjugating enzyme Rad6 and the ubiquitin protein ligase Brel. (ii)<br />

Cis- and trans-regulation <strong>of</strong> histone-tail modifications as part <strong>of</strong> the histone code on histories H2B<br />

and H3. Modified residues are indicated. Histone modifications include ubiquitination,<br />

methylation, phosphorylation, and acetylation. Cis regulation: phosphorylation <strong>of</strong> S10 on<br />

histone H3 negatively affects (indicated by red double blocked line) methylation <strong>of</strong> K9, whereas<br />

acetylation <strong>of</strong> K9 and/or K14 positively regulated (indicated by double pointed arrows) by S10<br />

phosphorylation (and vice versa). Both events take place on the same histone tail. Trans-<br />

regulation: methylation <strong>of</strong> K4 and K79 on histone H3 are dependent on the mono-ubiquitination<br />

<strong>of</strong> K123 on histone H2B (broken arrows).<br />

Abbreviations: Ac, acetylation; K, lysine; Me, methylation; P, phosphorylation; S, serine; R,<br />

arginine. (Adapted from Bach and Ostendorff, 2003).<br />

H3


providing a further example <strong>of</strong> trans-histone regulation (Briggs et al., 2002).<br />

Although the methylation <strong>of</strong> K4 <strong>of</strong> H3 had been implicated in gene silencing,<br />

there is also evidence for K4 methylation <strong>of</strong> H3 in gene activation. K4 <strong>of</strong> histone<br />

H3 exists in a di- or tri-methylated state in transcriptionally active chromosomal<br />

(euchromatin) regions, but not in inactive (heterochromatin) regions (Santos-<br />

Rosa et al., 2002). The role <strong>of</strong> H2B ubiquitination in this context remains<br />

ambiguous, although a general role <strong>of</strong> the ubiquitination <strong>of</strong> histone residues<br />

serving as key regulators <strong>of</strong> subsequent histone modifications (Figure 1. ) is<br />

strongly supported. Thus, the modification <strong>of</strong> histone tails will influence gene<br />

expression by controlling the accessibility <strong>of</strong> both RNA polymerases and<br />

transcription factors to DNA.<br />

1.2 Ubiquitin-like protein modifiers<br />

Ubiquitin was originally isolated by Goldstein and co-workers from the<br />

thymus and was thought to be a thymic hormone (Goldstein et al., 1975).<br />

Ubiquitin was subsequently independently identified by two further groups, as an<br />

'ATP-dependent proteolysis factor 1' (Ciehanover et al., 1978) and as attached to<br />

histone H2A (Goldknopf and Busch, 1975, Goldknopf and Busch, 1977).<br />

Subsequent work showed that ubiquitin was found in all tissues and all<br />

eukaryotic organisms, and this ubiquitous distribution gave the protein its name.<br />

Indeed ubiquitin has been shown to be one <strong>of</strong> the most phylogenetically well-<br />

conserved <strong>of</strong> all proteins in eukaryotes, although there are no known homologues<br />

in either archea or eubacterial organisms. Ubiquitins' identification was the first<br />

example <strong>of</strong> a polypeptide modifier, and as such lends its name to the family <strong>of</strong><br />

polypeptide protein modifiers, the ubiquitin-like modifiers. Since the discovery<br />

<strong>of</strong> ubiquitin, further ubiquitin-like protein modifiers<br />

have been identified:<br />

SUMO-1/-2/-3 (Boddy et al., 1996, Mannen et al., 1996, Matunis et al., 1996,<br />

Okura et al., 1996, Shen et al., 1996, Kamitani et al., 1997, Lapenta et al., 1997,<br />

Mahajan et al., 1997, Tsytsykova et al., 1998, Kamitani et al., 1998), NEDD8<br />

(Kamitani et al., 1997), ISG15 (Haas et al., 1987), FAT10 (Liu et al., 1999b),<br />

HUB 1 (Dittmar et al., 2002), An l a/An lb (Linnen et al., 1993), APG12<br />

(Mizushima et al., 1998), FUBI (Kas et al., 1992), URM1 (Furukawa et al.,<br />

2000), and UFM1 (Komatsu et al., 2004). Higher eukaryotic genomes contain<br />

24


an increased number <strong>of</strong> ubiquitin-like proteins, compared to the lower eukaryotes<br />

(Table. 1. ) and the evolution <strong>of</strong> new members <strong>of</strong> the ubiquitin-like family seem<br />

likely to have occurred from the ancestral members (Figure 2. ).<br />

Ubiquitin-like proteins fall into two separate classes (reviewed in Jentsch<br />

and Pyrowolakis, 2000). The first class, those mentioned above, function as<br />

modifiers in a similar manner to that <strong>of</strong> ubiquitin. They exist either in a free<br />

form or attached *covalently to other proteins via their matured C-termini<br />

(Figure. 3. ). These proteins are termed "ubiquitin-like modifiers", or type 1 Ubls.<br />

The conjugation <strong>of</strong> the ubl is achieved via an enzymatic cascade involving an El<br />

(activating enzyme), E2 (conjugating enzyme), and an E3 (ligase) enzyme. In all<br />

cases to date, the conjugation occurs at a lysine residue within the substrate,<br />

either within or without a "consensus motif", although recently p21<br />

has been<br />

shown to be ubiquitinated at its N-terminal methionine (Bloom et al., 2003). The<br />

second class <strong>of</strong> proteins contain ubiquitin related domains but also domains with<br />

no homology between. one another. Members <strong>of</strong> this second group are not<br />

conjugated to other proteins. Collectively these proteins are termed type-2 UBLs<br />

(also called UDPs, ubiquitin domain proteins), which contain ubiquitin-like<br />

structure embedded in a variety <strong>of</strong> different classes <strong>of</strong> large proteins with<br />

apparently distinct functions, such as Rad23, Elongin B, Scythe, Parkin, and<br />

HOIL-1 (Tanaka et al., 1998; Jentsch and Pyrowolakis, 2000; Yeh et al., 2000;<br />

Schwartz and Hochstrasser, 2003).<br />

The similarities between the different ubiquitin-like proteins and their<br />

component enzymes are as striking as their differences. Despite the lack <strong>of</strong><br />

sequence homology between the various ubiquitin-like modifiers, to date all<br />

share a common tertiary structure, the ubiquitin fold (Figure 4. i. ). The body <strong>of</strong><br />

each adopts this, ßßaßßaß fold (Figure 4. ii. ), resulting in a stable globular<br />

structure that is highly resistant to a range <strong>of</strong> denaturants (Vierstra and Callis,<br />

1999, Vijay-kumar et al., 1987, Melchior, 2000). Despite the similar structure<br />

there are significant differences in the electrostatic potentials <strong>of</strong> the various<br />

ubiquitin and ubiquitin-like proteins and unlike most ubiquitin-like modifiers, the<br />

members <strong>of</strong> the SUMO family possess a flexible N-terminal extension.<br />

25


Table 1. Identifed Ubls in Saccharomyces cerevisiae and Homo Sapiens. ;<br />

Saccharomyces cerevisiae Homo Sapiens<br />

Ubiquitin Ubiquitin<br />

ISG15<br />

FAT 10<br />

FUB1<br />

RUB! NEDD8<br />

SMT3 SUMO-1<br />

SUMO-2<br />

SUMO-3<br />

Atg8 GATE-16<br />

LC3<br />

GABARAP<br />

Atg l2 Atg l2<br />

Urm 1<br />

Ufml<br />

GenBank accession numbers: ScUbiquitin P04838; hUbiquitin AAA36789; ISG15<br />

P0161; FAT10 NP_001988; FUB1 NP001988; RUBI NP_009209; NEDD8<br />

NP_006147; SMT3 Q12306; SUMO-1 ACAA67898; SUMO-2 CAA67896; SUMO-3<br />

P55855; Atg8 NP_009475; GATE-16 P60520; LC3 NP_852610; GABARAP<br />

NP_009209; ScAtg12 NP_009776; hAtgl2 NP_004698; Ufml B0005193; Urml<br />

NP_012258<br />

AtgS<br />

GATE-16<br />

GARARAP<br />

MAP-113<br />

UFM1<br />

Atg12p<br />

Atg12<br />

MOCO1-A<br />

MoaD. E. coG<br />

Urml<br />

ThiS. E. coli<br />

FUBI<br />

Anla<br />

Anlb<br />

" Ubiquitin<br />

YeastUb<br />

- NED08<br />

Rubl<br />

tSG15<br />

FAT10<br />

- 50140-1<br />

- SUMO-2<br />

-S0140-3<br />

Sm13<br />

Figure 2. Phylogram <strong>of</strong> all known Ub-lke proteins and Ubl-like bacterial proteins. Phylogram<br />

was constructed via the ClustalW alignment program (www. ebi. ac. uk/clustalw) and was<br />

generated from full length sequences.


Figure 3. Schematic representation <strong>of</strong> the SUMO-1/-2J-3 (Su) and ubiquitin(Ub)<br />

conjugation pathways. Both pathways can be divided into five key stages. The first<br />

stage in the conjugation process is the maturation <strong>of</strong> the SUMO/ubiquitin translation<br />

products. Prior to conjugation a C-terminal inhibitory tag is required to be removed,<br />

to expose the di-glycine residues that allow conjugation. This maturation is<br />

catalysed by specific proteases; ubiquitin C-terminal hydrolayse (UCH) and SUMO<br />

specific protease(SSP), for ubiquitin and SUMO respectively. The processed<br />

modifier can then be attached to substrates via a three step enzymatic cascade. In the<br />

first step specific activating enzymes (Els) form ATP-dependent thioester bonds<br />

between internal cysteine residues and the conserved C-terminal glycine <strong>of</strong> the<br />

modifiers. The El for SUMO is a heterodimer, SAE1/SAE2, whilst the El for<br />

ubiquitin is monomeric, UBAI. In the second step specific conjugating enzymes<br />

(E2s) accept the activated modifier via a transesterification reaction to an internal<br />

cysteine residue. Many E2s are known to exist for ubiquitin (Ubcl-8, UbclO-11, and<br />

Ubcl3), whilst only a single E2 has been isolated for SUMO (Ubc9). The final step<br />

involves the formation <strong>of</strong> an isopeptide bond between the E-amino group <strong>of</strong> the<br />

substrate target lysine and the terminal glycine <strong>of</strong> the modifier. For the ubiquitin<br />

pathway the E3 activity is an absolute requirement for substrate modification, whilst<br />

in the SUMO pathway, E3 activity is not normally required for conjugation, but may<br />

serve to enhance the efficiency <strong>of</strong> the process. The modifiers can subsequently be<br />

removed from substrates via the actions <strong>of</strong> the SUMO-specific protease (SSP) or an<br />

ubiquitin isopeptidase. Following their deconjugation the protein modifiers are<br />

recycled and are available for conjugation, existing in a free "pool" <strong>of</strong> modifier.<br />

I


Ubiquitination<br />

E3 cycle<br />

5. Deconjugatio? 0 -<br />

4. Ligation<br />

LEL1<br />

Ubi<br />

-, C<br />

Maturation<br />

`Activation<br />

J3.<br />

ransestenfication


(ii)<br />

SUMO-1<br />

(1<br />

1)<br />

MSDQEAKYSTEDLGDKKEGEYIKL MGQDSSEIHFKVRMfTHLKKLKESYCQRQGVPMNSLRFLFEGQRIADNHTPKELQbIEEEDVIEVYQEQTGGHSTV<br />

Ubiquitin<br />

NEDD8<br />

Ufm l<br />

Key:<br />

I strand<br />

MQIrVKTLTGKTITLXVEPSDTIiM%IAXIQDKZOIDPDQQRLIFAG QLIDGRTLBDYNIQXZSTLHLVLRLRGG<br />

NQ, IRNETLTG EIEIDIZPTDEVERIEERVEEEEGIPPQQQRLITSGIIQMNDERTARDSEILGGSVLHLVLRLRGGGGM. RQ<br />

! (BEVBFRITLTBDPRLPTRVLBVPEBTPFTAVLEFAAEEFEVPAIºTSM IT$DGIOINPAQTAGNVFLRRGBELRI IPRDRVOBC<br />

J; )<br />

helix coil<br />

Figure 4. i. The "ubiquitin fold". A representation <strong>of</strong> the tertiary structure <strong>of</strong> ubiquitin.<br />

The adoption <strong>of</strong> this fold results in a stable globular structure, and is a common<br />

characteristic <strong>of</strong> all Uhl proteins. ii. Comparison <strong>of</strong> predicted secondary structures <strong>of</strong><br />

SUMO-1, Ubiquitin, NEDD8, and Ufml. All four <strong>of</strong> these proteins, despite having<br />

dissimilar amino acid sequences have all been reported to posses the "ubiquitin fold".<br />

Predicted secondary structures were generated through the secondary structure prediction<br />

s<strong>of</strong>tware Jpred (www. compbio. dundee. ac. uk/-wwwjpred)


The majority <strong>of</strong> Ubls are expressed as inactive precursors, which are<br />

made initially as fusions with C-terminal extensions, which prevent conjugation.<br />

These inhibitory tails, which can range from a single amino acid in length to a<br />

polypeptide, are specifically cleaved via the activity <strong>of</strong> proteases, resulting in the<br />

release <strong>of</strong> the active Ubl and its tail.<br />

1.2.1 Ubiquitin<br />

Despite the similarities in conjugation pathways <strong>of</strong> the ubl proteins, once<br />

conjugated their effects are distinctly unique. Ubiquitin was the first <strong>of</strong> the<br />

superfamily <strong>of</strong> protein tags to be identified, and for many years was thought to be<br />

unique in its ability to be attached to other proteins. The effects <strong>of</strong> ubiquitin<br />

conjugation on a substrate are varied and depend primarily on how many<br />

ubiquitin moieties are bound. The attachment <strong>of</strong> ubiquitin, in most cases, is not<br />

to a specific lysine residue that serves as an acceptor nor are they part <strong>of</strong> a<br />

recognition motif. There are exceptions to this rule such as in the case <strong>of</strong> IKBa,<br />

where lysine residues 21 and 22 are unique sites <strong>of</strong> ubiquitin ligation (Scherer et<br />

al., 1995; Baldi et al., 1996; Rodriguez et al., 1996). The conjugation <strong>of</strong> a single<br />

ubiquitin protein, monoubiquitination, is quite distinct from having multiple<br />

ubiquitin molecules conjugated, which is termed polyubiquitination. Ubiquitin<br />

chains can be formed in distinct linkages, due to the presence <strong>of</strong> lysine residues<br />

in ubiquitin molecules. The chain types and their roles will be discussed in detail<br />

later. Ubiquitin was first isolated bound to histone H2A, indeed in vertebrates up<br />

to 15% <strong>of</strong> histone H2A and around 5% <strong>of</strong> H2B, as well as the linker histone H1,<br />

are ubiquitinated, predominantly in the monoubiquitinated form (Rechsteiner ed,<br />

1988). Monoubiquitination has been shown to be involved in at least three<br />

distinct biological functions, these being histone regulation, endocytosis, and the<br />

budding <strong>of</strong> retroviruses from the plasma membrane (reviewed in Hicke, 2001).<br />

The function <strong>of</strong> histone modification by ubiquitin has been demonstrated to be<br />

important in gene expression, as previously mentioned in the histone code. The<br />

importance <strong>of</strong> histone ubiquitination can be illustrated in yeast cells, where H2B<br />

mutants, lacking the ubiquitinylation site do, not sporulate indicating a role in<br />

meiosis (Robzyk et al., 2000). It is well known that post-translational<br />

30


modifications <strong>of</strong> histone tail domains modulate chromatin structure and gene<br />

expression, although the role <strong>of</strong> ubiquitin in this process is poorly understood.<br />

Deubiquitination <strong>of</strong> monoubiquitinated histones is closely associated with<br />

mitotic chromatin condensation, and consequently transcriptional repression.<br />

The ubiquitin status <strong>of</strong> the histones is in a dynamic equilibrium, where ubiquitin<br />

is freely exchanged with intact nucleosomes (Seale, 1981), although it has been<br />

shown that apoptotic stimuli lead to a caspase-dependent deubiquitination <strong>of</strong><br />

monoubiquitinated histone H2A (Mimnaugh et al., 2001). Ubiquitination <strong>of</strong><br />

histone H2B in yeast (Saccharomyces cerevisiae) is required for the methylation<br />

<strong>of</strong> histone H3 (Sun and Allis, 2002) and as such shows the involvement <strong>of</strong><br />

ubiquitin as a determinant <strong>of</strong> the histone code hypothesis (reviewed in <strong>St</strong>rahl and<br />

Allis, 2000). Monoubiquitination is required as an internalisation signal on<br />

endocytic cargo at the plasma membrane and may play a role in the control <strong>of</strong> the<br />

activity <strong>of</strong> the endocytic machinery (reviewed in Hicke, 2001). Similarly the<br />

budding <strong>of</strong> enveloped viruses, such as Ebola, from the plasma membrane <strong>of</strong><br />

infected cells, requires monoubiquitination, although this process occurs in the<br />

opposite direction to that <strong>of</strong> vesicles in the process <strong>of</strong> endocytosis (Gar<strong>of</strong>f et al.,<br />

1998).<br />

Ubiquitin is also capable <strong>of</strong> forming polyubiquitin chains, through the<br />

attachment to internal lysine residues at positions 6,11,29,48, and 63 (Figure<br />

5). The functions <strong>of</strong> Lys ll and Lys29-linked chains are unknown. The<br />

functions <strong>of</strong> Lys63-linked chains have a range <strong>of</strong> fates: DNA repair, translation,<br />

IKB kinase activation, and endocytosis. The most abundant ubiquitin substrate in<br />

Saccharomyces cerevisiae is L28, a component <strong>of</strong> the large ribosomal subunit,<br />

which forms polymeric chains through Lys63 (Spence et al., 2000). The best<br />

documented role for polyubiquitination, is targeting <strong>of</strong> proteins for proteosomal<br />

degradation. In this case chains <strong>of</strong> four or more in length are linked via Lys48.<br />

Once bound to the 26S proteasome the protein to be degraded is unfolded and the<br />

ubiquitin chain deconjugates and is recycled.<br />

Ubiquitin is encoded for in the human genome both as a monoubiquitin<br />

form and as a polyubiquitin (Özkaynäk et al., 1984, Wiborg et al., 1985). In<br />

yeast, UB14 encodes a polyubiquitin gene, consisting <strong>of</strong> five consecutive<br />

ubiquitin-coding repeats in a spacerless head-to-tail arrangement. UB14 has been<br />

31


Ubiquitin<br />

Lys 6 11 9 3<br />

/<br />

/ I<br />

10. pIl<br />

E2 ? ? Many Ub Ub c 13 Ubc l, Ubc4, UbcS<br />

EZs<br />

E3 B Many Ra RAF6 Rsp5<br />

Eis Rad 1 8<br />

Substrates BRCA1/ Many Su bstrates? L28 ? Membrane<br />

BARD I substrates proteins<br />

Process DNA repair Proteasome DNA repair Trans lation IKB kinase Endocytosis<br />

/replication mediated activation and transport<br />

degradation<br />

Figure 5. Different functions for different ubiquitin linkages. Schematic representation <strong>of</strong><br />

ubiquitin with the modifiable lysines and roles <strong>of</strong> different linkages. Ubiquitin contains<br />

several lysine residues and, in vivo, can form multi-ubiquitin chains linked through positions<br />

6,11,29,48, and 63. The functions <strong>of</strong> Lysl I and Lys29-linked chains are unknown. Lys6-<br />

linked chains are formed by the autoubiquitination <strong>of</strong> BRCAI/BARD] during DNA repair<br />

and replication, Lys48-linked chains target substrates to the proteasome, but might have other<br />

functions, and Lys63-linked chains have a range <strong>of</strong> fates. Figure adapted in part from<br />

Weissman, 2001.<br />

C


shown to be essential in providing ubiquitin to cells under stress, and that<br />

ubiquitin is in itself is an essential component <strong>of</strong> the stress response system<br />

(Finley et al., 1987). Transcription <strong>of</strong> the polyubiquitin gene would allow cells<br />

to maintain the levels <strong>of</strong> free ubiquitin against the increased rate at which free<br />

ubiquitin is depleted through the formation <strong>of</strong> ubiquitin-protein conjugates.<br />

Following the attachment <strong>of</strong> a polyubiquitin chain, a substrate is most commonly<br />

degraded by the 26S proteasome, although ubiquitinated cell surface exposed<br />

membrane proteins, are targeted to the lysosome following ligand binding<br />

(reviewed in Hicke, 1997). Ubiquitin conjugation can also play important non-<br />

destructive functions, such as those involved in the NF-KB signalling pathway.<br />

Ubiquitin plays multiple roles in the NF-KB signalling pathway that ultimately<br />

leads to NF-KB translocation to the nucleus and gene expression (Figure 6). The<br />

family <strong>of</strong> NF-KB transcription factors are a group <strong>of</strong> structurally related and<br />

evolutionarily conserved proteins, which form either homodimers or<br />

heterodimers (Ghosh and Karin, 2002). In vertebrates, they comprise five<br />

subunits, p65 (ReIA), Re1B, c-Rel, and two which are transcribed and translated<br />

as precursors p100 and p105. Proteolytic processing <strong>of</strong> p100 and p105 generate<br />

p52 and p50 respectively. Both precursors possesss large C-terminal regions<br />

containing ankyrin repeat domains (ARD), resulting in their cytoplasmic<br />

retention. The ubiquitination <strong>of</strong> p105 leads to the removal <strong>of</strong> the C-terminus,<br />

exposure <strong>of</strong> its nuclear localisation sequence, and subsequent nuclear localisation<br />

(Palombella et al., 1994). Both p52 and p50 are only active as heterodimers,<br />

with p60, RelB, or c-Rel. The most abundant form <strong>of</strong> NF-KB dimer is the<br />

p5O/p65 heterodimer, which is found in most cell types, bound to the inhibitory<br />

protein, IKB in the cytoplasm. The second point at which ubiquitin is involved is<br />

in the proteolysis <strong>of</strong> IKB. IKBct is ubiquitinated on lysines 21 and 22, and this<br />

polyubiquitination, following its phosphorylation, results in its degradation and<br />

release <strong>of</strong> NF-KB, which subsequently translocates to the nucleus and activates<br />

transcription (Figure 6). Interestingly SUMO, an ubiquitin-like modifier<br />

competes with ubiquitin for conjugation <strong>of</strong> IKB, creating a privileged pool <strong>of</strong> NF-<br />

KB-IKB, held in the cytoplasm (Figure 6) (Desterro et al., 1998). The trigger for<br />

ubiquitination <strong>of</strong> IKB is its phosphorylation by the IKB kinase complex, and this<br />

phosphorylation<br />

33


u au -d on<br />

c4 tA<br />

0<br />

-: Ei<br />

14<br />

"ýe Oo<br />

o ýu oö<br />

äää<br />

, I: 1 ä0N<br />

°n U)<br />

"-1D<br />

Q '(U a .°v 2- a cý aý ö0<br />

"ö<br />

ä -, 4<br />

tu w Nö 4ý<br />

EI ,; 0'>G0b<br />

ca 0a<br />

ý `"+ý ýýCý<br />

ca<br />

IU :bOO<br />

ce<br />

IJ >,<br />

Ö<br />

U<br />

.ä<br />

0L<br />

4) be pp 2bÖ<br />

.c<br />

tom. iý Ri<br />

c0<br />

p+ G1 pb "v +`ý+ Ü<br />

c3 Qý cý vii a ýp U) yO<br />

"E ÄC 5G . cOC '" p.<br />

.0y:<br />

ýr<br />

)<br />

:jý,<br />

A'


ÜE<br />

a<br />

C<br />

ß<br />

ý.<br />

i-.<br />

..<br />

dýQ<br />

ý<br />

JV<br />

i<br />

ý<br />

x<br />

L<br />

oa<br />

ý' r<br />

Q'<br />

a<br />

c<br />

ýý ,n<br />

a<br />

U-<br />

I-<br />

F-<br />

v<br />

r.<br />

d Ký<br />

'`ý y<br />

l(i<br />

a<br />

ä ýý J<br />

1ý0<br />

Qýýý<br />

ý: ýýýý<br />

vý<br />

aý .<br />

öý<br />

O<br />

yC<br />

vý `U<br />

N "O<br />

ö<br />

pip<br />

.ý<br />

0<br />

u<br />

0<br />

rý<br />

W<br />

a<br />

U<br />

z


is also an ubiquitin-mediated event. The binding <strong>of</strong> interleukin-1 to the<br />

interleukin-1 receptor results in the association <strong>of</strong> adaptor proteins with the<br />

intracellular region <strong>of</strong> the receptor, and subsequent recruitment <strong>of</strong> TRAF6. The<br />

TRAF6 molecules are multiply ubiquitinated, forming chains via the Lys63<br />

residue in ubiquitin. By mechanisms as yet unclear this results in the activation<br />

<strong>of</strong> the TAK1 kinase, which phosphorylates IKK (Figure 6) (Wang et al., 2001,<br />

reviewed in Finley, 2001). Evidence has begun to emerge suggesting links<br />

between ubiquitin-dependent proteolysis and transcriptional activation. The<br />

most straightforward link between degradation and transcription is by controlling<br />

the levels <strong>of</strong> transcription factors, which is true for p53 (Carr, 2000), ß-catenin<br />

(Hart et al., 1999), and Rpn4 (Xie and Varshavsky, 2001). A slightly more<br />

sophisticated relationship between the two is shown by the observation that<br />

sequences within activators that specify proteolysis commonly overlap with the<br />

transcriptional activation domains (Salghetti et al., 2000), and that components<br />

<strong>of</strong> the transcription machinery recruited by activation domains can lead to the<br />

degradation <strong>of</strong> activators (Chi et al., 2001, Nelson et al., 2003). Recently it has<br />

been shown that ubiquitin-mediated degradation stimulates transcription<br />

promoted by the oestrogen receptor (Reid et al., 2003) and that the function <strong>of</strong><br />

the MYC proto-oncoprotein is also modulated via the ubiquitin pathway (von der<br />

Lehr et al., 2003, Kim et al., 2003), and there is compelling evidence that the<br />

proteasome itself plays a direct role in transcriptional regulation.<br />

1.2.1.1 The Proteasome<br />

The 26S Proteasome is the predominant, 2.5 MDa enzyme, that degrades<br />

proteins conjugated to ubiquitin, and is composed <strong>of</strong> two major subcomplexes,<br />

the 670 kDa proteolytic core particle (CP; also known as 20S proteasome) and<br />

the 900 kDa regulatory particle (RP; also known as PA700 and the 19S<br />

complex). The RP particle, consisting <strong>of</strong> six ATPases (Rptl-Rpt6) and at least<br />

twelve other non-ATPase subunits (Rpnl-12) is further subdivided into the base<br />

and lid, with all ATPases being located in the base (Glickman et al., 1998). The<br />

CP subunits are arranged into four seven-membered rings, the outer rings are<br />

formed by the a subunits, whilst the inner rings are formed by the proteolytically<br />

active ß subunits (Groll et al., 1997). These subunits can possess chymotrypsin-<br />

36


like activity, which cleaves after hydrophobic residues, a trypsin-like activity,<br />

which cleaves after basic residues, and a peptidylglutamyl peptidase that cleaves<br />

after acidic residues and are responsible for the disassembly <strong>of</strong> the substrate<br />

protein. (Baumeister et al., 1998). Peptidase activity is not limited to the CP <strong>of</strong><br />

the proteasome however. The Rnp11 subunit <strong>of</strong> RP lid has been shown to<br />

contain a metalloprotease motif (Verma et al., 2002, Yao et al., 2002). The<br />

proposed role for this subunit is in the removal <strong>of</strong> the multiubiquitin chain, which<br />

is essential for the efficient translocation <strong>of</strong> target protein into the CP <strong>of</strong> the<br />

proteasome (reviewed in Wilkinson, 2002, Hochstrasser, 2002). The RP has<br />

been demonstrated to have multiple functions, such as regulating the entry to the<br />

sealed internal chamber <strong>of</strong> the CP. The axial channel <strong>of</strong> the CP is gated by the<br />

Rpt2 ATPase and has been shown to control both substrate entry and product<br />

release (Köhler et al., 2001). The base <strong>of</strong> the lid has also been shown, through<br />

the Rpn1 subunit, to bind ubiquitin-like protein domains (Elsasser and Gali et al.,<br />

2002) and can recruit multi-ubiquitin chains through their association with<br />

ubiquitin-associated (UBA) domains with proteins also possesssing ubiquitin-<br />

like domains (Wilkinson et al., 2001). Interestingly, Gonzalez et al. (2002), have<br />

shown that the ATPases <strong>of</strong> the RP, and not the proteases <strong>of</strong> the CP, become<br />

associated with genes that are being transcribed. The ATPases not only did not<br />

interact with the CP when bound to DNA, they did not interact with the lid <strong>of</strong> the<br />

RP, the exact role <strong>of</strong> this association has yet to be elucidated, and it has yet to be<br />

determined whether the RP base plays a direct role, enzymatically, in<br />

transcriptional activation or is limited to transcription factor degradation<br />

(reviewed in Ottosen et al., 2002).<br />

Multiple associated proteins regulate proteasome structure and function.<br />

The most notable include Ubp6, a deubiquitinating enzyme, Hu15, an ubiquitin-<br />

ligase, Ecm29, which tethers the CP to the RP (Leggett et al., 2002), and the 11S<br />

complex. The i IS complex is composed <strong>of</strong> alpha, beta, and gamma subunits and<br />

is capable <strong>of</strong> associating with the 20S proteasome, and replaces the 19S<br />

proteasome on at least one side. 11 S is interferon inducible, as are three subunits<br />

from the 20S proteasome, ß1i, ß5i, and ß2i, and unlike the 19S complex, 11S<br />

play roles in antigen processing (Groettrup et al., 1995). The 11S complex is<br />

found in mature PML nuclear bodies, nuclear bodies which are dependent upon<br />

37


SUMO for their maturation, and are further recruited following exposure to<br />

arsenic trioxide, and interferon gamma (Fabunmi et al., 2000), which has<br />

implications in promyelocytic leukemia (Lallemand-Breitenbach et al., 2001).<br />

Thus the proteasome plays important roles, not only in protein<br />

degradation, antigen presentation, the recycling <strong>of</strong> ubiquitin, but also in<br />

transcriptional regulation. Despite these pivotal roles some proteasome<br />

inhibitors are already in clinical trials, showing particular promise in treating<br />

cancers <strong>of</strong> blood cells such as multiple myeloma (Adams, 2002).<br />

1.2.2 NEDD8<br />

NEDD8 (Neural precursor cell-Expressed Developmentally down-<br />

regulated) was first reported following the identification <strong>of</strong> a set <strong>of</strong> genes with<br />

developmentally down-regulated expression in the mouse brain (Kumar et al.,<br />

1992). Subsequently NEDD8 was shown to function in an analogous manner to<br />

that <strong>of</strong> ubiquitin and SUMO, and covalently conjugate to substrate proteins<br />

(Kamitani et al., 1997). NEDD8 substrates were subsequently identified as<br />

belonging to the Cullin family (Cullin-1, -2, -3, -4a, -4b, -5), a family <strong>of</strong> proteins,<br />

which bear significant sequence similarity to the yeast Cdc53 protein (Osaka et<br />

al., 1998, Wada et al., 1999, Hori et al., 1999). In yeast, Cdc53 forms part <strong>of</strong> a<br />

protein complex, with SKIP1 and Cdc4, that functions as an ubiquitin E3 ligase<br />

to target the yeast CDK inhibitor, p40sIC1, for ubiquitin-dependent degradation<br />

during the G1/S transition. Thus it was likely, and subsequently shown, that<br />

NEDD8 would be involved in the degradation pathway. This hypothesis was<br />

substantiated by the finding that NEDD8 modification <strong>of</strong> Cullin-1 activates the<br />

ubiquitin-ligase, SCF (beta-TrCP), and was required for IKBa degradation via<br />

the ubiquitin/proteasome pathway (Read et al., 2000). NEDD8 modification <strong>of</strong><br />

Cullin-1 was shown to enhance the recruitment <strong>of</strong> Ubc4, an ubiquitin E2, to the<br />

SCF complex (Kawakami et al., 2001). Recently NEDD8 was demonstrated to<br />

conjugate to pVHL, where it was required for fibronectin matrix assembly and<br />

suppression <strong>of</strong> tumour development (<strong>St</strong>ickle et al., 2004). Surprisingly, NEDD8<br />

modification <strong>of</strong> pVHL played no role in the ubiquitination <strong>of</strong><br />

hypoxia inducible<br />

factor-1 (HIF-1) a known substrate for the VCB E3 ligase complex (<strong>St</strong>ickle et<br />

38


al., 2004). Two reports show that another protein<br />

CAND1 (Cullin-associated<br />

NEDD8-deassociated 1) negatively modulates, the interaction <strong>of</strong> the E3 subunits,<br />

and selectively binds cullin-1, which has not been NEDD8 modified. The<br />

conjugation <strong>of</strong> NEDD8 to cullin leads to the dissociation <strong>of</strong> CAND1; and<br />

subsequent formation <strong>of</strong> the E3 complex (Liu et al., 2002; Zheng et al., 2002).<br />

NEDD8 conjugation is apparently modulated in a number <strong>of</strong> ways: the binding<br />

<strong>of</strong> Butl and But2 (proteins that bind to Uba three) to Uba3 (Yashiruda et al.,<br />

2003); and by NUB I. The association <strong>of</strong> Butl and But2 to the NEDD8 El<br />

component, Uba3 restricts NEDD8 conjugation by inhibiting the El. NUB1<br />

(NEDD8 Ultimate Buster-1) overexpression results in a severe reduction <strong>of</strong><br />

NEDD8 levels (Kito et al., 2001). NUB1 interacts with S5a subunit <strong>of</strong> the RP <strong>of</strong><br />

the proteasome, and proteasome inhibitors were shown to completely block<br />

NUB 1-mediated down-regulation <strong>of</strong> NEDD8 expression, which suggests that<br />

NUB I recruits NEDD8 and its conjugates to the proteasome where they are<br />

subsequently degraded (Kamitani et al., 2001).<br />

The COPS signalosome is an evolutionarily conserved multiprotein<br />

complex that possess similarities with the lid <strong>of</strong> the proteasome, RP (reviewed in<br />

Schwechheimer and Deng, 2001). Interestingly the COPS signalosome was<br />

demonstrated to remove NEDD8 from cullins, thus modulating the activities <strong>of</strong><br />

E3 ligases (Lyapina et al., 2001). The removal <strong>of</strong> NEDD8 is achieved via a<br />

metalloenzyme in the Jabl/Csn5 subunit <strong>of</strong> the complex. The Jabl/Csn5 appears<br />

to remove NEDD8 from cullins (Cope et al., 2002) in a similar fashion to the<br />

deubiquitinating metalloprotease <strong>of</strong> the RP. To date only one NEDD8 specific<br />

protease has been reported. NEDP1/DEN1 is a cysteine protease, which is<br />

homologous to the SUMO specific proteases (Mendoza et al., 2003; Gan-Erdene<br />

et al., 2003).<br />

1.2.3 ISG15<br />

ISG15 (interferon induced gene <strong>of</strong> 15 kDa), also known as UCRP<br />

(ubiquitin cross-reactive protein) was originally identified as a member <strong>of</strong> the<br />

ubiquitin-like proteins, which was induced upon exposure <strong>of</strong> cells to type I<br />

interferon (IFN), viral infection (Haas et al., 1987), and lipopolysaccharide<br />

(LPS) (Malakhov et al., 2003b). ISG15 is a 17 kDa protein comprised <strong>of</strong> two<br />

39


domains, each <strong>of</strong> which bears a regular pattern <strong>of</strong> homology to ubiquitin. ISG15<br />

shares a similar conjugation pathway but has its own unique<br />

conjugation/deconjugation enzymes that would appear to have evolved from the<br />

ubiquitin pathway enzymes (Narasimhan et al., 1996, Potter et al., 1999, Yuan<br />

and Krug, 2001). ISG15 has been shown to conjugate to Serpin 2a (Hamerman<br />

et al., 2002) as well as key regulators <strong>of</strong> signal transduction: PLCy1, Jakl,<br />

STAT1, and ERK1 (Malakhov et al., 2003a, Malakhov et al., 2003b). Loss <strong>of</strong><br />

UBP43, an ISG15 specific protease, in mice results in a decreased life span,<br />

brain cell injury, and hypersensitivity to interferon stimulation (Malakhov et al.,<br />

2002; reviewed in Kim and Zhang, 2003). Recently ISG15 has been shown to<br />

have functions in neutrophil-mediated immune mechanisms, as it has been<br />

shown to be a novel neutrophil chemotactic factor (Owhashi et al., 2003).<br />

1.2.4 FAT10<br />

FAT10 is similar to, but distinct from ISG15, and contains two ubiquitin-<br />

like domains seperated by a short linker region (Fan et al., 1996, Bates et al.,<br />

1997, Raasi et al., 1999). FAT10, formerly known as diubiquitin, has an<br />

exposed diglycine motif at its C-terminus, and is capable <strong>of</strong> being conjugated to<br />

as yet unidentified target protein(s). It induces apoptosis (Raasi et al., 2001), and<br />

is highly upregulated in hepatocellular carcinoma and other gastrointestinal and<br />

gynaecological cancers (Lee et al., 2003).<br />

1.2.5 Atgl2 and Atg8<br />

In yeast, Atg12 and Atg8 are ubiquitin-like proteins involved in the other<br />

pathway <strong>of</strong> protein breakdown, autophagy. The ubiquitin/proteasome is involved<br />

with, amongst other things, the breakdown <strong>of</strong> short-lived proteins. Its is reported<br />

that the half-lives <strong>of</strong> the majority <strong>of</strong> cellular proteins range from a few minutes to<br />

more than ten days, with the definition <strong>of</strong> a long-lived protein being one with a<br />

half-life <strong>of</strong> more than five hours. By this definition more than 99% <strong>of</strong> cellular<br />

proteins have relatively long lives (Ohsumi, 2001). The route for degradation <strong>of</strong><br />

these proteins lies in their targeting to the lysosome, which is equivalent to the<br />

vacuole in yeast. The interior <strong>of</strong> this organelle is acidic and contains multiple<br />

enzymes to breakdown cellular compounds. This process degrades aged proteins<br />

40


from the cytoplasm and can remove entire organelles under stress conditions,<br />

such as starvation (Ohsumi, 2001). Both Atgl2 and Atg8 are required for this<br />

process.<br />

Atg12 is a 186-amino acid hydrophilic protein, which possess a C-<br />

terminal glycine that, through a conjugation pathway analogous to that <strong>of</strong><br />

ubiquitin, is attached to a lysine residue in Atg5 (Mizushima et al., 1998) in<br />

yeast. The El enzyme in this pathway is Atg7 (Tanida et al., 1999), and the E2<br />

is AtglO (Shintani et al., 1999), although AtglO bears no homology with any<br />

known E2 protein. The sole substrate for this pathway seems to be Atg5, to<br />

which Atg12 is conjugated immediately after its translation. Apparently little or<br />

no free Atg12 exists in cells, with the majority conjugated to Atg5. The Atgl2-<br />

Atg5 complex then recruits Atgl6 to form a 350 kDa multimeric complex that is<br />

essential for autophagy (Mizushima et al., 1999, reviewed in Jentsch and Ulrich,<br />

1998, Ohsumi, 2001).<br />

Atg8 was the second ubiquitin-like protein shown to be involved in<br />

autophagy in yeast. Atg8 is processed by Atg4, a cysteine protease, removing a<br />

single amino acid to reveal a C-terminal glycine residue. Atg8 and Atg12 are<br />

activated by the same El enzyme, Atg7 but once activated are passed on to<br />

distinct E2 enzymes, Atg3, in the case <strong>of</strong> Atg8 (Ichimura et al., 2000). Uniquely<br />

Atg8 has been shown to conjugate to the amino group <strong>of</strong><br />

phosphatidylethanolamine (PE), and as such mediates protein lipidation<br />

(Ichimura et al., 2000, reviewed in Ohsumi, 2001). In mammals there are three<br />

ubls with homology to Atg8; GATE-16, GABARAP, and MAP-LC3.<br />

1.2.6 Urml<br />

Urml, ubiquitin-related modifier, is a ninety-nine amino acid protein that<br />

terminates in a diglycine motif, that is absent from higher eukaryoyic genomes.<br />

Urml is conjugated to substrate proteins, though these have yet to be identified,<br />

through the activation <strong>of</strong> Uba4, an El-like enzyme. Urml and Uba4 show<br />

sequence homology with the prokaryotic proteins essential for molybdopterin<br />

and thiamin biosynthesis (Furukawa et al., 2000), providing insights into the<br />

evolution <strong>of</strong> the ubiquitin-like proteins and their pathways. In Saccharomyces<br />

cerevisiae loss <strong>of</strong> Urml has been reported to enhance post-translational<br />

41


modification/proteolysis <strong>of</strong> Elongator subunit Totlp (Elplp) and abrogates its<br />

TOT (toxin target) function. This loss <strong>of</strong> function prevents Totip mediating cell<br />

cycle arrest in G1 (Fichtner et al., 2003).<br />

1.2.7 HUB1<br />

HUB1 (homologous to ubiquitin 1) is unique amongst ubiquitin-like<br />

modifiers, possesssing a dityrosine carboxyl-terminus, which is exposed<br />

following the processing <strong>of</strong> a single inhibitory amino acid residue. The<br />

dityrosine motif has been shown to conjugate to cell polarity factors Sphl and<br />

Hbtl in Saccharomyces cerevisiae (Dittmar et al., 2002). A subsequent<br />

investigation into HUB 1, suggests that although HUB I can indeed form SDS-<br />

resistant complexes with cellular proteins, these interactions are not through<br />

covalent conjugation with the dityrosine motif (Luders et al., 2003). As such<br />

HUB 1 may very well not be a true ubl, but may be a member <strong>of</strong> the type II<br />

category <strong>of</strong> Ubiquitin-like proteins.<br />

1.2.8 UFM1<br />

The newest member <strong>of</strong> the ubl family is the ubiquitin-fold modifier 1 (UFM1).<br />

Komatsu et al., whilst trying to identify GATE-16 interacting proteins, isolated a<br />

protein Uba5, which was similar to many <strong>of</strong> the previously identified El<br />

enzymes. Further investigation showed that Uba5 was indeed an El-like<br />

enzyme, and was the El enzyme for UFM1 (Komatsu et al., 2004). UFM1 was<br />

shown to be a true ubl as high molecular weight species were detected, indicating<br />

that UFM1 was capable <strong>of</strong> conjugating to target substrates. Ufml has no<br />

sequence homology with ubiquitin, yet still possess the characteristic ubiquitin<br />

fold as determined by computer modelling ((MOE program (2003.02; Chemical<br />

Computing Group Inc., Montreal, Quebec, Canada)) (Komatsu et al., 2004).<br />

1.3 SUMO<br />

The Schizosaccharomyces<br />

pombe and Saccharomyces cerevisiae yeasts<br />

contain a single gene encoding a SUMO homologue, Pmt3 and Smt3<br />

42


espectively, whilst in higher eukaryotes there are three distinct is<strong>of</strong>orms;<br />

SUMO-1, SUMO-2, and SUMO-3.<br />

SMT3 was originally isolated in a screen for high copy suppressors <strong>of</strong><br />

mutations in the centromere binding protein MIF2 (see GenBank Accession No.<br />

U33057). Mif2 protein is a yeast centromere protein with homology to the<br />

mammalian centromere protein, CENP-C (Brown et al., 1995; Meluh and<br />

Koshland, 1995). <strong>St</strong>udies using temperature-sensitive mutants showed that the<br />

loss <strong>of</strong> yeast Mif2p function results in chromosome missegregation, mitotic<br />

delay, and aberrant microtubule morphologies (Brown et al.,<br />

1993). Pmt3-null<br />

cells show various phenotypes, including aberrant mitosis, sensitivity to various<br />

DNA damaging reagents, and high-frequency loss <strong>of</strong> minichromosomes.<br />

Furthermore it was shown that loss <strong>of</strong> Pmt3 function resulted in a striking<br />

increase in telomere length (Tanaka et al., 1999). In contrast, deletion <strong>of</strong> SMT3<br />

in S. cerevisiae is lethal and yeast with mutations in UBC9 accumulate at the G2-<br />

M boundary <strong>of</strong> the cell cycle. Mutations in the Pmt3 UBC9 equivalent, hus5, or<br />

an E1 subunit, rad31, result in mitotic defects and impaired survival after ultra-<br />

violet radiation (Ho and Watts, 2003; reviewed in Hay, 2001). Genetic studies in<br />

mice have yet to be reported although RNA interference studies in<br />

Caenorhabditis elegans indicate that the knockdown <strong>of</strong> SUMO resulted in a high<br />

incidence <strong>of</strong> embryonic lethality, with the surviving progeny having high<br />

incidences <strong>of</strong> a protruding vulva phenotype (Fraser et al., 2000).<br />

Although Smt3/Pmt3 is attached to many yeast proteins, only six have<br />

been characterised to date. In Saccharomyces cerevisiae, where Smt3<br />

conjugation is reported to be essential for viability, three members <strong>of</strong> the septin<br />

family <strong>of</strong> proteins are Smt3 conjugated (Johnson et al., 1999). Septins are<br />

components <strong>of</strong> a belt <strong>of</strong> 10-nm filaments that encircle the yeast bud neck; Smt3<br />

is attached to the septins Cdc3, Cdcll, and Shsl/Sep7, specifically during<br />

mitosis. The same is not true, however, for Pmt3 in Schizosaccharomyces<br />

pombe. Whereas the staining pattern <strong>of</strong> Smt3 at the mother/bud neck is observed<br />

at mitosis in Saccharomyces cerevisiae, no similar pattern is observed at the<br />

mother/bud neck suggesting that the septins may not be major Pmt3-modified<br />

targets in Schizosaccharomyces pombe (Tanaka et al., 1999). Pds5 is another<br />

43


Smt3 substrate that is maximally conjugated during mitosis, where Pds5 Smt3-<br />

conjugation promotes the dissolution <strong>of</strong> cohesion (<strong>St</strong>ead et al., 2003).<br />

Topoisomerase II is similarly involved in controlling centromeric cohesion,<br />

regulating either a component <strong>of</strong> chromatin structure or topology required for<br />

centromeric cohesion, and like Pds5, this process can be regulated by Ulp2<br />

(Bachant et al., 2002).<br />

In vertebrates SUMO are a family <strong>of</strong> ubiquitin-like modifiers, comprising<br />

three members. SUMO-1 is approximately 45% identical to either SUMO-2 or<br />

SUMO-3, following processing, whilst SUMO-2 and SUMO-3 are highly<br />

homologous, being 96% identical to each other following processing (Figure 7).<br />

The SUMO proteins have been characterised as covalently modifying numerous<br />

cellular and viral proteins (Table 2) with substrate dependent consequences.<br />

Functional differences also exist between the various is<strong>of</strong>orms, with SUMO-2<br />

and SUMO-3 being capable <strong>of</strong> forming polymeric chains (Tatham et al., 2001).<br />

The specific roles <strong>of</strong> SUMO-1, SUMO-2, and SUMO-3 have yet to be fully<br />

elucidated, partially due to research focussing on SUMO-1. Recent work on<br />

topoisomerase II during mitosis highlighted the need for care. Azuma et al.,<br />

showed that although exogenous SUMO-1 was capable <strong>of</strong> conjugating to<br />

topoisomerase II, none was detected when endogenous material was analysed. It<br />

was further shown that topoisomerase II was specifically modified by SUMO-2/-<br />

3 during mitosis, and that SUMO-2/-3 conjugation was dramatically reduced<br />

following the addition <strong>of</strong> exogenous SUMO-1. Given all three is<strong>of</strong>orms share<br />

the same conjugation enzymes it is therefore quite conceivable that exogenous<br />

SUMO-1 competes with SUMO-2/-3 for the enzymes and as a result inhibits<br />

their conjugation (Azuma et al., 2003). This is unlikely to represent the only<br />

case where exogenous expression does not represent the endogenous situation,<br />

and it is probable that some <strong>of</strong> the reported SUMO-1 substrates may in fact turn<br />

out to be specifically modified by SUMO-2, SUMO-3, or both, although how<br />

this is achieved as yet remains unsolved.<br />

The three SUMO is<strong>of</strong>orms all conjugate to the same conserved motif,<br />

'KxE, where IF represents a large hydrophobic amino acid and x represents any<br />

amino acid. The sequence requirements for the SUMO consensus binding site<br />

were investigated, both in vitro and in vivo. W could<br />

be leucine, isoleucine,<br />

44


hSUMO-2<br />

hSUMO-3<br />

ScSmt3<br />

hSUMO-1<br />

hUbiquitin<br />

ScUbiquitln<br />

ScRubl<br />

WK x<br />

Figure 7. (i) Primary sequence alignment <strong>of</strong> the human (h) and S. cerevisiae (Sc)<br />

homologues <strong>of</strong> Ubiquitin, SUMO, and NEDD8. Similar residues are shown in blue.<br />

Position <strong>of</strong> the SUMO modification motifs present in SUMO-2/3 is indicated by pKxE.<br />

Scissors denote site <strong>of</strong> protease activity. (ii) Space-fill 3D models <strong>of</strong> SUMO-1,<br />

Ubiquitin, NEDD8, and representations <strong>of</strong> SUMO-2 and SUMO-3. Models for SUMO-2<br />

and SUMO-3 were predicted by submission <strong>of</strong> primary sequences with the PDB co-<br />

ordinates for the 3D structure <strong>of</strong> SUMO-1 (Bayer et al., 1998) to the ExPASy proteomics<br />

server <strong>of</strong> the Swiss Institute <strong>of</strong> Bioinformatics `Swiss Model' (Guex and Peitsch, 1997;<br />

Peitsch et al., 1996). Electrostatic calculations are based upon a dielectric constants <strong>of</strong><br />

80.0 (solvent) and 40.0 (protein) at physiological pH. Blue indicates negative surface<br />

potentials whilst red indicates positive surface potentials. Images generated using Swiss<br />

PDB viewer v3.6b (Guex et al., 1999)


Table 2. List <strong>of</strong> SUMO modified substrates.<br />

Transcription co-/factor Effect on transcription Reference<br />

Tcf-4 Positive Yamamoto et at, 2003<br />

HSF-1 Positive Hong et at, 2001<br />

HSF-2 Positive Goodson et at, 2001<br />

CREB Positive Comerford et at, 2003<br />

p53 Positive(? ) Rodriguez et at, 1999; Gostissa et at, 1999<br />

p73a Not characterised Minty et at, 2000<br />

Dnmt3a Positive Ling et at, 2004<br />

Dnmt3b Not characterised Kang et at, 2001<br />

Pdxl Positive Kishi et at, 2003<br />

APA-1 <strong>St</strong>abilisation Benanti et al, 2002<br />

C/EBP Negative Kim et at, 2002<br />

ELK-1 Negative Yang et at, 2003<br />

SREBPs Negative Hirano et at, 2003<br />

Sp3 Negative<br />

-<br />

Sapetsching et at, 2002; Ross et at, 2002<br />

IRF-1 Negative Nakagawa et at, 2002<br />

ARNT Negative Tojo et at, 2002<br />

AP-2 Negative Eloranta et at, 2002<br />

Jun Negative Muller et at, 2000<br />

c-Myb Negative Bies et at, 2002<br />

Lef1 Negative Sachdev et at, 2001<br />

SRF Negative Matasuzaki et at, 2003<br />

AR Negative (? ) Poukka et at, 2000<br />

PR Negative (? ) Takimoto et at, 2003<br />

GR Negative (? ) Tian et at, 2002<br />

GATA-2 Negative (? ) Chun et at, 2003<br />

GRIP1 Negative Kotaja et at, 2002<br />

HDAC 1 Enhanced repression <strong>David</strong> et at, 2002<br />

HDAC4 Enhanced repression Kirsh et at, 2002<br />

Histone H4 Negative Shiio and Eisenman, 2003<br />

p300 Negative <strong>Girdwood</strong> et at, 2003<br />

CtBPI Negative Kagey et at, 2003; Lin et at, 2003<br />

PL"ZF Negative Kang et al, 2003<br />

SOP-2 Negative Zhang et at, 2004


Nuclear body proteins<br />

Property Reference<br />

PML PML body formation Boddy et al, 1996<br />

Sp100 Uncertain <strong>St</strong>erndorf et al, 1997<br />

HIPK2 Localisation Kim et al, 1999<br />

TEL, TEL-AMLI Nuclear export Chakrabarti et al, 2000<br />

Daxx Ryu et al, 2000<br />

Genome integrity proteins<br />

Rad52/Rad22, Rad51/Rhpl<br />

TDG<br />

Topo I<br />

Topo II<br />

WRN<br />

BLM<br />

PCNA<br />

Signal transduction proteins<br />

IKBa<br />

CamkIl<br />

Smad4<br />

Axin<br />

Mekl<br />

Ttk69<br />

Dorsal<br />

FAK<br />

MDM2<br />

Double-strand break repair<br />

Increase in enzymatic turnover<br />

DNA repair<br />

DNA repair<br />

Not characterised<br />

Localisation<br />

DNA repair<br />

<strong>St</strong>abilisation<br />

Regulates neuronal differentiation<br />

Localisation/stabilisation<br />

JNK activation<br />

Nuclear export<br />

Regulates neuronal differentiation<br />

Immune response<br />

Activation <strong>of</strong> autophosphorylation<br />

<strong>St</strong>abilisation<br />

Viral proteins Property Reference<br />

Ho et al, 2001; Shen et al, 1996<br />

Hardeland et al, 2002<br />

Mao et al, 2000<br />

Mao et al, 2000<br />

Kawabe et al, 2000<br />

Suruki et al, 2001<br />

Hoege et al, 2002<br />

Desterro et al, 1998<br />

Long et al, 2000<br />

Lin et al, 2003<br />

Rui et al, 2002<br />

Sobko et al, 2002<br />

Lehembre et al, 2000<br />

Bhaskar et al, 2002<br />

Kadare et al, 2003<br />

Buschmann et al, 2001<br />

EBV-BZLF1 PML body disruption Adamson et al, 2001<br />

BPV-E1 Localisation Rangasamy et al, 2001<br />

CMV-IE1/IE2 Disrupts SUMO localisation Ahn et al, 2001; Spengler et al, 2002<br />

AdV-E1B 55K Localisation Endter et al, 2001<br />

Neuronal proteins Property Associated disorder Reference<br />

APP Abeta production Alzheimers Li et al, 2003<br />

Httex 1p <strong>St</strong>ability Huntingtons <strong>St</strong>effan et al, 2004<br />

Neuronal tracks NIID Pountney et al, 2004<br />

Atrophin mutant (? ) Polyglutamine tracts Terashima et al, 2002; Ueda et<br />

Cytoplasmic<br />

al, 2002<br />

Yeast septins Not characterised Takahashi et al, 1999 Johnson et al, 1999<br />

GLUT, GLUT4 Activity Giorgino et al, 2000<br />

Nuclear pore complex<br />

RanGAPI Localisation Matunis et al, 1996<br />

RanBP2 Unclear, SUMO E3 Pichler et al, 1998


valine, phenyalanine, and methionine with no detrimental affects on SUMO<br />

conjugation, whilst alanine and proline permitted weak conjugation and<br />

tryptophan prevented conjugation. The requirement <strong>of</strong> the glutamic acid was<br />

also shown although weak SUMO conjugation did occur when this was<br />

substituted to an aspartic acid (Rodriguez et al., 2001). These requirements were<br />

tested on transferable sequences containing the'PKxE motif, allowing SUMO<br />

modification. Although the WKxE motif was demonstrated to be sufficient for<br />

modification in vitro, modification in vivo required the additional presence <strong>of</strong> a<br />

nuclear localization signal. The SUMO conjugation motif is an. absolute<br />

requirement for'the vast majority <strong>of</strong> SUMO substrates but there are a number <strong>of</strong><br />

exceptions to this rule. APA-1, CREB, IRF-1, Tcf-4, Daxx, Axin, Hdm2,<br />

Histones, Huntingtin, and SENP1 have all been demonstrated to be SUMO<br />

substrates, that do not contain the consensus SUMO modification motif, with the<br />

exceptions <strong>of</strong> SENP1, which although it contains a motif, this is not the site <strong>of</strong><br />

SUMO modification and Tcf-4 which is modified at a consensus site but also<br />

possessses a second site <strong>of</strong> modification that lies out-with the consensus (Benanti<br />

et al., 2002, Comerford et al., 2003, Nakagawa et al., 2002, Yamamoto et al.,<br />

2003, Ryu et, 2000, Rui et al., 2002, Buschmann et al., 2001, Shiio and<br />

Eisenman, 2003, Bailey and O'Hare, 2003). CtBP2 is also modified in vitro by<br />

SUMO, but only in the presence <strong>of</strong> Pc2 although it lacks a consensus motif<br />

(Kagey et al., 2003), which raises the possibility that other substrates, that lack<br />

consensus motifs, are SUMO modified in the presence <strong>of</strong> appropriate SUMO E3<br />

enzymes.<br />

The effects <strong>of</strong> SUMO conjugation onto a substrate can broadly be<br />

characterised into one <strong>of</strong> three groups: altering substrate stability, changing sub-<br />

cellular localisation, and altering protein-protein interactions. As a large number<br />

<strong>of</strong> SUMO substrates are either transcription factors or co-factors, SUMO<br />

conjugation has been demonstrated to play an important role in transcriptional<br />

regulation, exerting both positive and negative influences in a substrate<br />

dependent manner.<br />

The classic example <strong>of</strong> SUMO modification increasing substrate stability<br />

is that <strong>of</strong> IiBa. SUMO conjugation to IKBa directly competes with ubiquitin,<br />

thereby preventing IKBa degradation. This antagonism <strong>of</strong> ubiquitin is achieved<br />

48


due to the requirements <strong>of</strong> lysine residues 21 and 22, which are blocked and<br />

masked by SUMO attachment. Ubiquitination does not usually require specific<br />

lysine residues for conjugation and as such SUMO conjugation to a substrate<br />

would not prevent degradation, as ubiquitin would bind at any available lysine<br />

residue. In the case <strong>of</strong> CREB, c-Myb, Smad4, and APA-1 the increase in<br />

stability could be explained by a reduction in ubiquitin moieties attached<br />

(Comerford et al., 2003, Bies et al., 2002, Benanti et al., 2002, Lin et al., 2003).<br />

1.3.1 Localisation and Nuclear domains<br />

The nucleus, like the cytoplasm, is compartmentalised, containing distinct<br />

nuclear domains. Distinct nuclear domains thus far identified include, nucleoli,<br />

Cajal bodies (CBs), gems, splicing speckles, clastosomes, paraspeckles, and<br />

Promyelocytic leukaemia (PML) bodies (reviewed in Lamond and Sleeman,<br />

2003). These nuclear bodies may function by accumulating factors in distinct<br />

localisations, thereby creating concentrations <strong>of</strong> components, promoting efficient<br />

interactions. The compartmentalisation <strong>of</strong> these components could create<br />

domains <strong>of</strong> active/inactive factors, and as such control <strong>of</strong> their<br />

assembly/disassembly may regulate many cellular functions (reviewed in Isogai<br />

and Tjian, 2003).<br />

1.3.2 PML bodies<br />

PML bodies are known under a variety <strong>of</strong> names; PML bodies, ND 10,<br />

kreb bodies, and PODs (PML oncogenic domains). PML bodies are in essence<br />

an association <strong>of</strong> numerous proteins, co-localising within the nucleus. The<br />

protein components <strong>of</strong> the PML bodies are <strong>of</strong>ten very transient and many <strong>of</strong> the<br />

proteins identified as components are conditional on their expression levels<br />

(Figure 8). Mature PML bodies, where PML forms the outer shell (Lallemand-<br />

Breitenbach et al., 2001, LaMorte et al., 1998) are intimately linked with the<br />

SUMO conjugation system. Much work has focused on the Promyelocytic<br />

leukaemia (PML) gene since its integral role in Acute Promyelocytic leukaemia<br />

was established, brought on by the t(15; 17) translocation resulting in a<br />

PML/retinoic acid receptor (RAR) a fusion protein. In healthy cells PML<br />

49


Figure 8. Composition and regulation <strong>of</strong> PML bodies. Table consisting <strong>of</strong> the known<br />

components <strong>of</strong> PML bodies, indicating the conditions used to show their localisation, and<br />

a brief summary <strong>of</strong> their function. 'Ishov et al, 1999; 2Mu et al, 1994; 3Alcalay et at,<br />

1998; °Guo et al, 2000; 5Seeler et al, 1998; 6Lehming et a1,1998; 7Yankiwski et al, 2000;<br />

SBoddy et al, 1996; 9<strong>St</strong>ernsdorf et al, 1997; 1°Kamitani et al, 1998; UAscoli and Maul,<br />

1991; 12Maul et al, 1995; '3Gong et al, 1997; "Gong et at, 2000; "Everett et al, 1999;<br />

16Seeler<br />

et al, 2001; "Doucas et al, 1999; '8Boisvert et al, 2001; '! Everett et al, 1999;<br />

'Maul et al, 1998; 21Fabunmi<br />

et al, 2001; 'Anton et al, 1999; 'Cao et al, 1998; 'Carlile<br />

et al, 1998; 'Lai and Borden, 2000; 'Ruthardt<br />

et al, 1998; 'Goodson et al, 2001;<br />

2'Doucas and Evans, 1999; 'Jiang et al, 1996; 30Gongora et al, 1997; 31Skinner et al,<br />

1997; 'Klement et al, 1998; 33Trost et at, 2000; 34Desbois et al, 1996; 35 Lehembre et at,<br />

2001; 'Vallian et al, 1998; 37Rochat-<strong>St</strong>einer<br />

et al, 2000; 'Maul, 1998; 'Yeager et al,<br />

1999; 42Johnson et al, 2001; `Alcalay et al, 1998; 'Bloch et al, 1999. And a schematic<br />

representing how the components <strong>of</strong> the PML bodies can be both recruited and released<br />

from PML bodies following different stimuli. (Adapted in part from Negorev and Maul,<br />

2001).


(i)<br />

Proteins present at PML bodies at endogenous levels <strong>of</strong> expression<br />

PML Fssential in PML, body formation; in SUMO-modified form recruits Daxx; interacts with CBP, pRb, p. 53; repressor;<br />

considered a co-<br />

''.<br />

SplOO<br />

Interacts with HPI; PML bodies are limited in their capacity to recruit Spl005.1,.<br />

Daxx Recruited by PMLSUMO; binds to an increasing number <strong>of</strong> proteins.<br />

BLM<br />

Low relative amount <strong>of</strong> BLM is stationed in PML bodies".<br />

SUMO<br />

NDP55<br />

Modifies PML and SpIOO covalently changing their capacity to bind other proteins"').<br />

Uncharacterised protein(s); potentially a protein modification recognised by Mab 138; increases after heat shock; present in<br />

all vertebrate species tested"'2.<br />

Proteins conditionally present in or at PML bodies<br />

TRF1l2<br />

NBS 1<br />

P95/nibrin<br />

Mrel 1<br />

Present in a subset <strong>of</strong> PMI, bodies containing telomere repeats in cells depending on an alternative telomere maintenance<br />

RAD-51 mechanism. These proteins colocalise with PML in RAD52<br />

Repl. Factor A<br />

WRN<br />

telomerase-negative immortalised cells during the late SIG,<br />

-phase.<br />

SENPI<br />

HPI<br />

SpIOOC<br />

HMG2<br />

CBP/p30p<br />

pRb<br />

USP7/HAUSP<br />

pA28<br />

proteosome<br />

p$p<br />

GAPDH<br />

eIF-4<br />

PLZF<br />

HSF2<br />

Tax<br />

p53<br />

ISO-20<br />

Mutant ataxin<br />

PKM<br />

Int6<br />

PAX3<br />

Spl<br />

FISTIHIPK3<br />

hGCNP<br />

Sp140<br />

BRCA 1<br />

(ii)<br />

Removes <strong>St</strong> IMO from PMI. ".<br />

Interacts with Sp10 but is present in PML bodies in variable amounts depending on the cell cycle Phase 5,6.1.5<br />

.<br />

Present in some PML bodies".<br />

Interacts with SplOOB6.<br />

Interacts with PMI. "; found in some cell types but not other'.<br />

Phosporylated form found by some groups but not others; possibly cell cycle related colocalisation'.<br />

At or beside some PML bodies; probably cell cycle dependent"-'<br />

Proteosomal activator; increased recruitment to PML bodies by interferon y exposure".<br />

Only after application <strong>of</strong> proteosome inhibitors=' or after interferon y as immunoproteosomes'".<br />

Interacts with PML and associates with a subset <strong>of</strong> PML bodies'.<br />

Interaction with PML is RNA dependent.<br />

Interaction with PML23.<br />

No interaction with PML but side by side localisation in nucleus".<br />

Heat shock transcription factor 2 after overexpression in low number <strong>of</strong> cells, mostly cytoplasmic2.<br />

Recruited to PML bodies when PML is overexpressee.<br />

In specific cells, after overexpression or proteosome inhibition° or beside PML bodies with T-ag down regulations.<br />

After overexpression <strong>of</strong> a fragment".<br />

PML but not the other PML body proteins are located in ataxin- l aggregates after polyglutamine tract extension3`32<br />

Mx interaction kinase after transient expression°.<br />

In some cells possibly by binding to overexpressed Daxxx.<br />

Interacts with PML'.<br />

After overexpression interacts with Daxe.<br />

Beside PML bodies upon transient overexpressio& .<br />

After overexpression with some PML bodies4'.<br />

Breast cancer protein I beside PML bodies upon transient overexpression.<br />

Recruitment<br />

IFN UP motion Recycling<br />

PML<br />

SUMO Ubc9<br />

Davy<br />

Release<br />

Viral Proteins<br />

E4 ORF3<br />

ICPO<br />

IEI<br />

External Insults<br />

Heat Shock<br />

I<br />

MMS<br />

Cd'<br />

Uv<br />

40"AO-Daxx SENP-1 SITMO<br />

PML body PML<br />

pl W'mo-HP1<br />

nax, i


functions as the organiser <strong>of</strong> the PML bodies, and is essential for the recruitment<br />

<strong>of</strong> other PML body associated proteins, a function lost by the PML/retinoic acid<br />

receptor (RAR) a fusion protein. PML is a common form <strong>of</strong> acute myeloid<br />

leukaemia (AML), and it is common for patients to possess the PML/RARa<br />

fusion protein, although four other different fusion partners have been reported to<br />

date, these being the promyelocytic leukemia zinc finger gene (PLZF), the<br />

nucleophosmin gene (NPM), the nuclear mitotic apparatus gene (NuMA), and<br />

the signal transducer and activator <strong>of</strong> transcription 5b gene (STAT5b) (Lin et al.,<br />

2001).<br />

Cells from APL patients show a disruptedmicropunctate pattern in<br />

place <strong>of</strong> the usual 10-20 PML bodies. SUMO modification was first suspected<br />

<strong>of</strong> being a crucial component <strong>of</strong> PML bodies as both PML and SplOO, another<br />

principal component <strong>of</strong> PML bodies, are SUMO substrates, as are many PML<br />

body associated proteins. Furthermore it was demonstrated that upon exposure<br />

to arsenic trioxide (As203) SUMO modification <strong>of</strong> PML was increased and<br />

nucleocytoplasmic PML is transferred to the nuclear matrix, resulting in an<br />

increase in size <strong>of</strong> the PML bodies (Muller et al., 1998). Treatment <strong>of</strong> APL cells<br />

with retinoic acid or As203 restores normal NB organisation. As203 treatment<br />

results in the proteasome-mediated degradation <strong>of</strong> SUMO modified PML and<br />

PML/RARa, mediated through Lys160 (Lallemand-Breitenbach et al., 2001) and<br />

subsequent cell apoptosis. PML/RARa strongly represses key regulatory genes<br />

and slows down differentiation resulting in the arrest <strong>of</strong> the cells at the<br />

promyelocyte stage (Du et al., 1999). Following PML degradation, healthy cells<br />

will synthesise additional PML, whereas PML/RARa cells being arrested at the<br />

promyelocyte stage will fail to. This results in the restoration <strong>of</strong> NBs localisation<br />

and remission <strong>of</strong> PML (reviewed in Zhu et al., 2002). Although SUMO<br />

modification is not required for either matrix targeting or formation <strong>of</strong> primary<br />

PML bodies, it is required for secondary shell-like PML body formation<br />

(Lallemand-Breitenbach et al., 2001). The ability <strong>of</strong> PML to function as a<br />

scaffold protein for NB recruitment is shown by PML -/- cells which had nuclear<br />

diffuse or mislocalised patterns <strong>of</strong> normally PML body-associated proteins<br />

(Ishov et al., 1999, Zhong et al., 2000). PML mutants that lack SUMO<br />

conjugation sites are defective in recruiting proteins to PML bodies, and this has<br />

52


lead to the hypothesis that PML bodies are sites <strong>of</strong> nuclear storage <strong>of</strong> SUMO<br />

modified substrates. The storage <strong>of</strong> these substrates can be affected by a number<br />

<strong>of</strong> early gene products from several DNA viruses, which act to disrupt the PML<br />

body. DNA viruses such as herpes simplex virus (HSV) and cytomegalovirus<br />

virus (CMV), both code for proteins which disrupt the SUMO-1 modification <strong>of</strong><br />

PML and SplOO (Muller and Dejean, 1999; Everett, 2001) and the nuclear<br />

accumulation <strong>of</strong> the CMV protein IE1 that mediates PML NB disruption, is<br />

SUMO-1 modification-dependent (Muller and Dejean, 1999). These<br />

observations suggest that DNA viruses may target PML bodies as a means <strong>of</strong><br />

relieving the transcriptionally repressive environment created by SUMO<br />

modification. Additionally PML body composition is affected by external<br />

stresses, such as heat shock and DNA damage, suggesting a role for PML bodies<br />

in many response pathways<br />

(Figure 8).<br />

PML is able to regulate transcription via association with numerous<br />

transcription factors/c<strong>of</strong>actors (Figure 9), and shows intrinsic repression activity<br />

as a Ga14 fusion (Ahn et al., 1998; Vallian et al., 1997). Furthermore over<br />

expression <strong>of</strong> PML leads to the repression <strong>of</strong> various promoters (Mu et al.,<br />

1994). Other examples <strong>of</strong> SUMO modified substrates, which are associated with<br />

PML bodies and exhibit repressive . capabilities include HIPK2, Daxx, and<br />

Sp100. HIPK2 is a member <strong>of</strong> the homeodomain interacting protein kinases<br />

(HIPK). SUMO modification <strong>of</strong> HIPK2 is required for its localisation to PML<br />

bodies (Kim et al., 1999) where it subsequently forms a stable corepressor<br />

complex with Groucho corepressor and HDAC1 (Choi et al., 1999). Daxx and<br />

Sp100, like PML, have been demonstrated to have repressive capabilities when<br />

fused to Ga14 DBD (Szostecki et al., 1990). Daxx accumulation in PML bodies<br />

is dependent upon SUMO modification <strong>of</strong> PML (Duprez et al., 1999). Daxx<br />

may potentially act as an adaptor protein, recruiting other PML body-associated<br />

proteins that do not interact directly with PML (reviewed in Maul et al., 2000).<br />

Daxx has the ability to recruit HDACs (Li et al., 2000), and thus likely represses<br />

transcription by remodelling chromatin. Interestingly the repressive ability <strong>of</strong><br />

Daxx is lost by PML coexpression, suggesting a mechanism by which PML<br />

sequesters Daxx, possibly by its retention to PML bodies (Lehembre et al., 2001;<br />

53<br />


Figure 9. PML associated proteins and transcriptional regulation. (i) Summary <strong>of</strong><br />

direct protein-protein interactions among PML associated proteins which are involved<br />

in transcriptional activation (green) or repression (orange). (ii) Models for<br />

transcriptional regulation by PML bodies: Left, PML bodies regulate transcription by<br />

sequestering and thereby inactivating activators or repressors from the active pool in the<br />

nucleoplasm. Middle, PML bodies serve as sites for assembly or activation <strong>of</strong><br />

transcription modulatory complexes by providing the post-translational modifying<br />

enzymes and a high local concentration <strong>of</strong> the factors involved. Right, PML bodies<br />

create a microenvironment that facilitates repression or activation in their direct<br />

surroundings. Figure adapted in part from Lin et al, 2001.


4<br />

HP1<br />

PML<br />

Y<br />

vpl<br />

A<br />

(-'Oact, vatRrs ('orcpressors<br />

fra cription Factors<br />

Scqucs(ation <strong>of</strong><br />

't,<br />

y<br />

ql<br />

% qw<br />

/ riýr1<br />

0<br />

R<br />

0<br />

ýd r<br />

ý, rRr<br />

ý`ý<br />

Z_<br />

, doch,<br />

EOML<br />

PMI. f><br />

Iy<br />

PML<br />

ý--<br />

-<br />

X53<br />

ý<br />

(3 1<br />

ýJ<br />

GA<br />

0<br />

HDACS<br />

corepressor/activator protci ns repressor/activator complexes<br />

mlcroenvironment<br />

Y<br />

Modification and/or assembly <strong>of</strong><br />

®SUMO<br />

"Phosphorylation =Interaction,<br />

Ac Acetylation<br />

('realign <strong>of</strong> activating/repressing<br />

Interaction, Repression<br />

Activation


Li et al., 2000). SplOO has been shown to possesss a heterochromatin 1(HP1)<br />

binding site (Lehming et al., 1998; Seeler et al., 1998). SUMO conjugation to<br />

SplOO enhances its interaction with HPIcc, at least in vitro (Seeler et al., 2001),<br />

providing an insight into how SUMO modification might regulate both PML<br />

bodies and chromatin. SplOO is currently thought to be recruited to PML bodies<br />

following the association <strong>of</strong> SplOO with an as yet unidentified adaptor protein.<br />

This adaptor proteins recruits Sp l00 either by binding to the self-assembly<br />

domain or by a conformational change in PML. (Negorev et al., 2001, reviewed<br />

in Negorev and Maul, 2001). Sp l00, like Daxx, in general appears to be<br />

recruited to PML bodies in order to reduce its availability in the nucleus and not<br />

as a necessity <strong>of</strong> PML body structural formation.<br />

In addition to the number <strong>of</strong> transcription factors and transcription co-<br />

factors found in PML bodies, a number <strong>of</strong> DNA damage associated protein are<br />

also found within this nuclear domain.<br />

1.3.3 SUMO and genome integrity<br />

The DNA damage response involves the coordination <strong>of</strong> numerous<br />

cellular pathways, including DNA repair events and checkpoint arrest<br />

mechanisms. Abnormalities in these processes are identified by their increased<br />

sensitivities to DNA damaging agents, such as UV and ionising radiation, or to<br />

the DNA synthesis inhibitor hydroxyurea (HU).<br />

SUMO was co-discovered as an interactor with the DNA repair enzymes<br />

Rad5l and Rad52 (Shen et al., 1996), although the authors termed their novel<br />

gene UBL1 (ubiquitin-like 1), this was subsequently shown to be the same gene<br />

as SUMO-1. Ubc9 has also been reported to associate with Rad51 (Kovalenko et<br />

al., 1996). In yeast Rad5l/Rad52-dependent DNA repair pathways are involved<br />

in DNA recombination and DNA double-strand break repair in yeast. In fission<br />

yeast, Schizosaccharomyces pombe, Pmt3 (SUMO) has been shown to be<br />

conjugated to both Rad22 and Rhp5 1, the fission yeast homologues <strong>of</strong> Rad51 and<br />

Rad52 (Ho et al., 2001). Similarly, mammalian Rad5l proteins have been<br />

demonstrated to play essential roles in DNA homologous recombination, DNA<br />

repair, and cell proliferation. Rad5l and Rad52 have been reported to form a<br />

complex with non-conjugated SUMO-1 (UBL1) in human cells (Li et al., 2000).<br />

56


The function <strong>of</strong> this complex seems likely to regulate homologous<br />

recombination, based on the observations that overexpression <strong>of</strong> SUMO-1 down<br />

regulated DNA double-strand break-induced homologous recombination and<br />

reduced resistance to ionizing radiation. Interestingly Rad51 foci colocalise with<br />

a subset <strong>of</strong> PML NBs and do not form in cells that express a dominant-negative<br />

PML form, suggesting that functional PML is necessary for their assembly<br />

(Bisch<strong>of</strong> et al., 2001). The role played by SUMO modification in recruitment <strong>of</strong><br />

Rad51 foci, either to PML or another as yet unidentified component,<br />

is further<br />

supported by the findings that other components <strong>of</strong> the Rad51 foci are indeed<br />

SUMO modified, such as WRN (Kawabe et al., 2000). WRN localises to the<br />

nucleolus under normal growth conditions, but following exposure to DNA-<br />

damaging agents, relocates into Rad51 foci (reviewed in Muller et al., 2004).<br />

Topoisomerase-mediated DNA damage represents an unique type <strong>of</strong><br />

DNA damage, and is <strong>of</strong> particular interest due to the effect many antibiotics,<br />

anticancer drugs, toxins, carcinogens, and physiological stresses have on the<br />

catalytic cycles <strong>of</strong> topoisomerases, and as such result in the formation <strong>of</strong><br />

topoisomerase-mediated DNA damage. Both topoisomerase I and topoisomerase<br />

II are substrates for SUMO modification (Mao et al., 2000a, Mao et al., 2000b).<br />

SUMO conjugation to topoisomerase is induced by the presence <strong>of</strong> camptothecin<br />

(CPT), an anti-tumour drug. CPT was shown not to be the only drug to induce<br />

SUMO modification <strong>of</strong> topoisomerase, ICRF-193, a drug which does not induce<br />

topoisomerase II-mediated DNA damage, but traps topoisomerase II into a<br />

circular clamp conformation, also resulted in SUMO conjugation. The<br />

consequence <strong>of</strong> SUMO modification on topoisomerase I, is to enhance the<br />

cleavable complex formation (Horie et al., 2002). Indeed sumoylation deficient<br />

mutants <strong>of</strong> topoisomerase I, reduced the CPT-induced cleavable complexes but<br />

had no effect on the in vitro catalytic activity.<br />

DNA damage has also been shown to activate the NF-KB pathway via the<br />

activation <strong>of</strong> NEMO (NF-KB essential modulator) / IKB kinase Y (IKKy) by<br />

ubiquitin-like proteins. NEMO is held in the cytoplasm as a member <strong>of</strong> the IKB<br />

kinase complex, consisting <strong>of</strong> NEMO and IKKI/a and IKK2/ß. Following DNA<br />

damage NEMO is transported to the nucleus and SUMO modified (Huang et al.,<br />

2003, reviewed in Hay, 2004). It is not yet clear whether NEMO shuttles<br />

57


etween the nucleus and cytoplasm, in a manner analogous to that <strong>of</strong> both NF-<br />

KB and IKBa or whether the translocation is DNA damage induced. Following<br />

NEMO sumoylation and nuclear localisation, NEMO is desumoylated, and<br />

phosphorylated by the ATM kinase. The ATM/ATR kinase family are <strong>of</strong>ten<br />

associated with phosphorylation <strong>of</strong> substrates following DNA damage, and<br />

signal the DNA damage event to the appropriate response enzymes.<br />

Interestingly, following NEMO phosphorylation, NEMO is then modified by<br />

ubiquitin, an event that seemingly occurs at the same lysine residues required for<br />

SUMO conjugation. Following NEMOs ubiquitination, NEMO exits the<br />

nucleus, associates with the IKK complex, and activates NF-KB.<br />

Smt3 (SUMO) modification <strong>of</strong> the proliferating cell nuclear antigen<br />

(PCNA), a DNA-polymerase sliding clamp involved in DNA synthesis and<br />

repair, in Saccharomyces cerevisiae plays a crucial role in DNA repair. Hoege et<br />

al., demonstrated that Smt3 conjugation occurred and competed with<br />

ubiquitination <strong>of</strong> the same lysine residue (Hoege et al., 2002, reviewed in<br />

Pickart, 2002, <strong>St</strong>elter and Ulrich, 2003). PCNA is monoubiquitinated through<br />

Rad6 and Radl8, and subsequently modified by lysine-63-linked multi-<br />

ubiquitination, through Mms2, Ubc13, and Rad5. The ubiquitination <strong>of</strong> PCNA is<br />

induced following DNA damage, and the resulting multi-K63 linked ubiquitin<br />

chain leads to error-free DNA repair. Smt3 modification <strong>of</strong> PCNA conversely<br />

leads to an inhibition <strong>of</strong> error-free DNA repair, via the competition for the<br />

acceptor lysine with ubiquitin (Hoege et al., 2002). It is attractive to speculate<br />

that the ubiquitin/SUMO switch present in Saccharomyces<br />

cerevisiae for PCNA,<br />

might be a prevalent regulatory mechanism, and there are a growing list <strong>of</strong><br />

documented cases where ubiquitin and SUMO directly compete for lysine<br />

residues (Desterro et al., 1998; Sobko et al., 2002; Hoege et al., 2002; Huang et<br />

al., 2003; <strong>St</strong>effan et al., 2004). However human PCNA has yet to be shown to<br />

be SUMO modified. Indeed the authors whilst, acknowledging the absence <strong>of</strong><br />

SUMO modified PCNA, also reported the detection, indeed prevalence <strong>of</strong> mono-<br />

ubiquitinated PCNA in HeLa cells, and that <strong>of</strong> multi-ubiquitinated forms at lower<br />

levels under normal conditions (Hoege et al., 2002).<br />

58


1.3.4 SUMO and transcriptional regulation<br />

The addition <strong>of</strong> a SUMO moiety to components <strong>of</strong> the transcriptional<br />

apparatus does not have a common consequence as it can both activate and<br />

repress transcription. Currently, a greater number <strong>of</strong> transcription factors, around<br />

half <strong>of</strong> all known substrates, are repressed following SUMO conjugation.<br />

A number <strong>of</strong> SUMO acceptor sites in many <strong>of</strong> the substrate transcription<br />

factors, have been mapped to previously characterised repression domains, and<br />

mutation <strong>of</strong> target lysine(s) leads to an increase in transcriptional activity.,<br />

SUMO modification <strong>of</strong> transcription factors may also result in an increase in<br />

transcriptional activity. Modification <strong>of</strong> the heat-shock transcription factors<br />

HSF1 and HSF2 with SUMO-1 results in increased DNA-binding activity and<br />

mutation <strong>of</strong> the acceptor lysine reduces the transcriptional activity <strong>of</strong> HSF1<br />

(Hong et al., 2001; Goodson et al., 2001). Also the UV light stimulated SUMO<br />

modification <strong>of</strong> p53, has been shown to result in a mild enhancement <strong>of</strong><br />

transcriptional and apoptotic responses (Rodriguez et al., 1999; Gostissa et al.,<br />

1999), although, as yet, the molecular mechanisms have not been elucidated.<br />

1.3.5 SUMO and neurodegenerative diseases<br />

The majority <strong>of</strong> the studies exploring the links between SUMO and<br />

neurodegeneration have focused on conditions in which there are apparent<br />

intranuclear aggregates. Evidence linking SUMO-modified proteins, and their<br />

accumulation in aggregates is shown in the case <strong>of</strong> neuronal intranuclear<br />

inclusion disease (NIID). NIID is a rare multisystem neurodegenerative disease<br />

characterised by large intranuclear inclusions in neurons <strong>of</strong> the central and<br />

peripheral nervous systems. Intranuclear aggregates are the pathologic hallmark<br />

<strong>of</strong> this disease, and are also a common feature <strong>of</strong> many other neurodegenerative<br />

diseases including many CAG/polyglutamine disorders. The role <strong>of</strong> ubiquitin in<br />

these diseases has been shown by the presence <strong>of</strong> ubiquitinated proteins<br />

in long<br />

glutamine tracts, representing aggregated proteins, a hallmark <strong>of</strong> many<br />

neurodegenerative diseases (Lieberman et al., 1998). Recently these<br />

intraneuronal aggregates, in NIID, were shown to stain with antibody<br />

recognising SUMO-1 (Pountney et al., 2004). The role <strong>of</strong> SUMO in other<br />

neuronal degenerative diseases is implicated by the observations that there is<br />

59


increased SUMO modification in brain tissue samples from patients with<br />

spinocerebellar ataxia type 3 (SCA3), dentatorubropallidoluysian atrophy<br />

(DRPLA), and in a mouse model <strong>of</strong> spinocerebellar ataxia type 1 (Terashima et<br />

al., 2002; Ueda et al., 2002).<br />

The role for SUMO in neuronal degenerative disorders has yet to be<br />

understood, although through the use <strong>of</strong> Drosophila experimental models <strong>of</strong><br />

polyglutamine disease, Chan et al., 2002, have demonstrated that SUMO is a<br />

modifier <strong>of</strong> toxicity. Further clues to SUMOs role comes from investigations<br />

showing SUMO-3 conjugation to amyloid precursor protein, and a possible role<br />

played in its proteolytic processing (Li et al., 2003).<br />

Recently SUMO has also been implicated in Huntington's disease (HD),<br />

a condition characterised by the accumulation <strong>of</strong> a pathogenic protein Htt, which<br />

contains an abnormal polyglutamine expansion. Marsh and colleges reported<br />

that a pathogenic fragment <strong>of</strong> Htt (Httexlp) can either be modified by SUMO-1<br />

or Ubiquitin. In a Drosophila model <strong>of</strong> HD, it was demonstrated that SUMO<br />

modification <strong>of</strong> Httexlp exacerbates neurodegeneration, whilst ubiquitination <strong>of</strong><br />

Httexlp abrogates neurodegeneration (<strong>St</strong>effan et al., 2004).<br />

1.4 Ubiquitin and ubiquitin-like proteases: - control over<br />

conjugation<br />

Ubiquitin is expressed as a protein fusion or as polyubiquitin, and free<br />

ubiquitin is generated by the actions <strong>of</strong> a family <strong>of</strong> cysteine proteases termed C-<br />

terminal hydrolases (UCHs) revealing the Gly-Gly motif required for its<br />

subsequent conjugation. Ubiquitin can subsequently be removed following its<br />

covalent attachment to substrate proteins, by actions <strong>of</strong> a second group <strong>of</strong><br />

proteases known as Ub-specific proteases (UBPs/USPs), and recently two further<br />

classes <strong>of</strong> ubiquitin proteases have been identified, a Ub-specific JAMM<br />

(JAB 1/MPN/Mov34) motif containing an metalloprotease and cysteine proteases<br />

containing an OTU (ovarian tumour) domain (Verma et al., 2002, Balakirev,<br />

2003, Evans et al., 2003). Similarly two distinct proteases are required for the<br />

control <strong>of</strong> ISG15. As yet no proteases have been proven to process ISG15, but<br />

ISG15 is removed from substrates via the protease activity <strong>of</strong> UBP43. Unlike<br />

60


oth ubiquitin and ISG15, the three SUMO ubls and NEDD8 are processed and<br />

removed following conjugation by the same proteases.<br />

The SUMO specific proteases are members <strong>of</strong> the cysteine protease<br />

family, falling into the C48 class <strong>of</strong> proteases (Rawlings and Barrett, 2000), and<br />

as such are distinct from the ubiquitin C-terminal hydrolyases and DUBs which<br />

are in the classes C12 and C19 respectively. Cysteine proteases contain at their<br />

active sites the amino acids Cys, His, and Asp, and this catalytic triad is used to<br />

define the cysteine protease family (Figure 10). Members <strong>of</strong> the SUMO family<br />

<strong>of</strong> SUMO specific proteases include proteases with no activity towards SUMO,<br />

rather. they contain a conserved catalytic domain which is found in all members<br />

<strong>of</strong> this group (Figure 10. ii and iii. ); furthermore the catalytic core shares a similar<br />

tertiary structure. Initially eight SUMO specific proteases were identified due to<br />

the presence <strong>of</strong> this core domain, termed "ULP domain" (Ubiquitin like protease)<br />

after the yeast Smt3 protease first identified, these being SENP1,2,3,5,6,7,<br />

and 8 (Figure 10. iii. ) (Yeh et al., 2000). The Schizosaccharomyces pombe and<br />

Saccharomyces cerevisiae genome each contain two genes encoding distinct<br />

Smt3/Pmt3 proteases, Ulpl and U1p2 (SMT4). Ulpl is localised to distinct<br />

regions <strong>of</strong> S. pombe in a cell cycle dependent manner (Taylor et al., 2002).<br />

During S and G2 phases <strong>of</strong> the cell cycle the Ulpl protein is localised to the<br />

nuclear periphery. Although not essential for cell viability, cells lacking the Ulpi<br />

gene display severe cell and nuclear abnormalities. Ulpl-null (Ulpl. d) cells are<br />

sensitive to ultraviolet radiation and have similar, although less pronounced<br />

phenotypes, to rad31. d and hus5.62. These cells contain mutations in one subunit<br />

<strong>of</strong> the activator and the conjugator for Smt3/Pmt3 respectively (Ho and Watts,<br />

2003). The second protease, U1p2, is localised throughout the nucleus, and the<br />

distinct localisations <strong>of</strong> the proteases accounts, in part, for its distinct in vivo<br />

substrate selectivity (Li et al., 2000). Ulp2-null cells, like U1p1-null cells,<br />

display multiple defects, including temperature-sensitive growth, abnormal cell<br />

morphology, decreased plasmid and chromosome stability, and severe<br />

sporulation defects. U1p2-null cells are hypersensitive to DNA-damaging agents,<br />

hydroxyurea, and benomyl. DNA damage results in an impaired ability to<br />

recover from cell cycle arrest. The studies on the proteases in yeast also revealed<br />

61


Figure 10. Comparison <strong>of</strong> SUMO-specific and SUMO-like cysteine proteases. (i)<br />

Representation <strong>of</strong> the active site <strong>of</strong> the cysteine protease, and schematic <strong>of</strong> its<br />

activation. The cysteine residue, activated by a histidine residue, plays a role in the<br />

nucleophile that attacks the peptide carbonyl bond. (ii) Schematic illustration <strong>of</strong> two<br />

<strong>of</strong> the six human SUMO-specific proteases (SENP1 and SENP2) and the two yeast<br />

homologues (ScUlpI and ScUlp2). The conserved core domain is in blue, whilst the<br />

arrows indicate the positions <strong>of</strong> conserved residues <strong>of</strong> the catalytic triad (histidine,<br />

asparate, and cysteine) an additional glutamine residue predicted to aid in the<br />

formation <strong>of</strong> the active site. (iii) Sequence alignment <strong>of</strong> the conserved catalytic core<br />

domain <strong>of</strong> the SUMO specific-like protease family. Vertical blue bars indicate<br />

highly conserved amino acids. Full triangles mark the amino acid residues <strong>of</strong> the<br />

catalytic triad, (His, Asn/Asp/Glu, and Cys) and the Gln residue involved in the<br />

formation <strong>of</strong> the oxyanion hole in the active site. Sequences: Human sentrin-specific<br />

protease 1 (SENP1) (AF149770); Human SENP2 (Q9HC62); Human SENP3<br />

(Q9H4L4); Human SENP5 (Q96HI0); Human SENP7 (Q9BQF6); Ulpl,<br />

Saccharomyces cerevisiae Ulpl protease; Human NEDD8 protease<br />

(NEDP1/DEN1/SENP8); African swine fever virus (ASFV) protease (Q00946);<br />

Human adenovirus 2 protease (ADE2) (P03252).


(i)<br />

(ii)<br />

(iii)<br />

SENP1<br />

SENP2<br />

SENP3<br />

SENP7<br />

Ulpl<br />

NEDP1<br />

ASFV<br />

PDE2<br />

SENP1<br />

SENP3<br />

SENP5<br />

SENP7<br />

Ulpi<br />

NEDP1<br />

ASFV<br />

ADE2<br />

ýZ<br />

ISO&<br />

ý' R<br />

ScUlp 1 (621 aa)<br />

SENPI (643 aa)<br />

SENP2 (1112 aa)<br />

ScUlp2 (1034 aa)


the presence <strong>of</strong> a possible feedback mechanism. Bylebyl et al (2003), showed<br />

that overproduction <strong>of</strong> a catalytically inactive version <strong>of</strong> Ulpl can substantially<br />

overcome the U1p2-null defects, a result that was shown to be reciprocal. The<br />

double protease mutant accumulates Smt3-protein conjugates to a much lesser<br />

extent than either single mutant, suggesting a mechanism that limits Smt3-<br />

protein ligation when Smt3 deconjugation by both Ulpl and Ulp2 are<br />

compromised.<br />

Ulpl is the major Smt3 precursor-processing enzyme in yeast, and<br />

whenpresent at endogenous levels is capable <strong>of</strong> cleaving Smt3 conjugates that<br />

accumulate in Ulp2-null cells. In contrast Ulp2, possessses a more limited ability<br />

to cleave Ulpl cell-specific Smt3 conjugates (Li et al., 2000). Ulp2 has been<br />

demonstrated to cleave the Smt3-Smt3 chains, formed via covalent attachment to<br />

the lysine residue <strong>of</strong> the conjugation motif in the N-terminus <strong>of</strong> Smt3 (Bylebyl et<br />

al., 2003), although despite relieving several <strong>of</strong> the Ulp2-null phenotypes, the<br />

polymeric chains were shown not to be essential in yeast. Wild type strains<br />

lacking any <strong>of</strong> the nine lysines in SUMO were viable, had no obvious growth<br />

defects or stress sensitivities, and showed similar patterns <strong>of</strong> SUMO conjugation<br />

as the wild type strain (Bylebyl et al., 2003). The specificity <strong>of</strong> the proteases<br />

towards substrates is controlled in part by the non-essential amino-terminal<br />

region <strong>of</strong> Ulpl. The Ulpl, Ulp domain (UD), a 200-residue segment, conserved<br />

among Ulps, which includes the core cysteine protease domain, is capable <strong>of</strong><br />

supporting wild-type growth and is capable <strong>of</strong> cleaving Smt3 from substrates in<br />

vitro (Li et al., 2003). Interestingly, the yeast strain containing the amino-<br />

terminally deleted Ulpl protease was capable <strong>of</strong> suppressing defects associated<br />

with cells lacking the Ulp2 protease, whilst wild-type Ulpl proteases are not.<br />

Additionally it was observed that many Smt3 conjugates accumulated to high<br />

levels in cells lacking the Ulp2 protease. These results showed that, although the<br />

catalytic domain was capable <strong>of</strong> processing Ulp2 protease substrates, the amino-<br />

terminal domain <strong>of</strong> Ulpl restricted protease activity towards certain substrates<br />

whilst enabling the cleavage <strong>of</strong> other substrates.<br />

In vertebrates, only three <strong>of</strong> the seven potential SUMO proteases, SENP1,<br />

SENP2, and SENP6 have been shown, to have SUMO processing and<br />

deconjugating activity in vitro (Gong et al., 2000; Hang and Dasso, 2002; Kim et<br />

al., 2000; Zhang et al., 2002). SENP8, as mentioned earlier, is a NEDD8<br />

64


specific protease, and it is possible that other members <strong>of</strong> this protease group will<br />

turn out to have specificity towards other Ubl proteins. The gene products from<br />

several animal viruses, including the adenovirus L3 protease, the I7 products <strong>of</strong><br />

fowlpox and vaccinia virus, and openreading frame (ORF) S273R <strong>of</strong> African<br />

swine fever virus (ASFV) (Li and Hochstrasser, 1999) also possesss this Ulp<br />

domain, the "core Ulp domain" <strong>of</strong> -90 amino acids, containing the conserved<br />

catalytic cysteine and histidine residues. Andres and colleagues subsequently<br />

investigated this observation and demonstrated that the ASFV protease was<br />

indeed capable <strong>of</strong> processing at Gly-Gly-Xaa sites, and was involved in the<br />

processing <strong>of</strong> the viral polyproteins pp62 and pp220 (Andres et al., 2000). The<br />

adenovirus protease has been well characterized, and is similarly capable <strong>of</strong><br />

processing at Gly-Gly-Xaa (Webster et al., 1993; Mangel et al., 1996; Ding et<br />

al., 1996).<br />

1.4.1 SUMO specific proteases and PML bodies<br />

Given the fact that SUMO modification <strong>of</strong> PML is required for both NB<br />

formation and PML recruitment <strong>of</strong> proteins such as, Daxx, CBP it was likely that<br />

the role <strong>of</strong> the family <strong>of</strong> SUMO specific proteases would prove important in both<br />

the regulation <strong>of</strong> PML body formation and regulation <strong>of</strong> transcriptional activity.<br />

Repression mediated through SUMO conjugation to Sp3 has been shown to be<br />

relieved via the expression <strong>of</strong> SUMO specific proteases (Ross et al., 2002),<br />

thereby providing a mechanism for controlling transcription. Work on a murine<br />

SENP2 is<strong>of</strong>orm, SuPr-1, provided further detail into the workings <strong>of</strong> the<br />

proteases. Best and colleagues demonstrated that PML is required for SuPr-1<br />

activity and enhances SuPr-1 action, presumably by having SUMO substrates in<br />

the PML body, which can subsequently be removed by the protease. A<br />

catalytically inactive form <strong>of</strong> the protease was capable <strong>of</strong> binding to SUMO<br />

modified PML, whereas the wild type was not, and it was further noticed that the<br />

inactive mutant was capable <strong>of</strong> activating c-Jun dependent transcription just as<br />

well as the wild type (Best et al., 2002). Recently it has been demonstrated that<br />

SENP1, is capable <strong>of</strong> being a target for SUMO modification in itself, and this<br />

modification is enhanced in catalytically inactive forms <strong>of</strong> the protease (Bailey<br />

and<br />

O'Hare, 2003).<br />

65


1.5 Competition for lysine residues<br />

SUMO is one member <strong>of</strong> the protein tags, which is capable <strong>of</strong> modifying<br />

substrates on their lysine residues, yet a further number <strong>of</strong> post-translational<br />

modifiers also attach via a substrates lysine residue, most notably acetylation.<br />

To date the number <strong>of</strong> identified substrates, where ubiquitin and SUMO compete<br />

for the same lysine residues is small. The first such substrate identified was<br />

IKBa (Desterro et al., 1999). As previously mentioned, Hay and colleagues<br />

demonstrated that the conjugation <strong>of</strong> SUMO to IKBa directly inhibited the<br />

conjugation <strong>of</strong> ubiquitin, thereby creating a "privileged pool" <strong>of</strong> IKBa resistant<br />

to degradation. NEMO is also modified by SUMO and ubiquitin on identical<br />

lysine residues. SUMO modification <strong>of</strong> NEMO results following DNA damage<br />

and nuclear localisation. Subsequently following SUMO deconjugation, NEMO<br />

is ubiquitinated and exported to the cytoplasm (Huang et al., 2003, reviewed in<br />

Hay, 2004). Furthermore it was demonstrated that PCNA, in yeast, was also<br />

modified by SUMO and ubiquitin, and again there was competition for the same<br />

lysine residue and controlled PCNAs DNA repair process (Hoege et al., 2002).<br />

The ubiquitination <strong>of</strong> PCNA is induced following DNA damage, and the<br />

resulting multi-K63 linked ubiquitin chain leads to error-free DNA repair.<br />

SUMO modification <strong>of</strong> PCNA conversely leads to an inhibition <strong>of</strong> error-free<br />

DNA repair, via the competition for the acceptor lysine with ubiquitin. A further<br />

example <strong>of</strong> competition is that involving MEK1, in Dictyostelium. SUMO<br />

conjugation <strong>of</strong> MEK1 occurs in response to elevated levels <strong>of</strong> cAMP that are a<br />

consequence <strong>of</strong> nutrient starvation. Under these conditions nuclear MEK1 is<br />

SUMO modified and transported to the cytoplasmic cortex. Interestingly the<br />

lysine residue for SUMO modification is also the target for ubiquitination, and<br />

following desumoylation, MEK1 translocates back to the nucleus where it is<br />

ubiquitinated and degraded (Sobko et al., 2002). Competition by ubiquitin and<br />

SUMO for the same lysine residue also occurs on the Huntingtin (Htt) protein, as<br />

previously mentioned. Interestingly Htt does not contain a consensus SUMO<br />

modification motif, and raises the likelihood <strong>of</strong> further examples <strong>of</strong> substrates<br />

that are modified by both SUMO and ubiquitin on identical lysine residues.<br />

66


Since modification by all members <strong>of</strong> the ubiquitin-like family <strong>of</strong> protein<br />

modifiers, and acetylation can all occur on lysine residues, transcription factors<br />

can potentially undergo a cascade <strong>of</strong> modifications that modulate their functions.<br />

An example <strong>of</strong> this is shown in the case <strong>of</strong> Sp3, where the major site <strong>of</strong> SUMO<br />

modification is identical to the major, site <strong>of</strong> acetylation (Braun et al., 2001).<br />

SUMO modification <strong>of</strong> Sp3 results in transcriptional repression (Sapetschnig et<br />

al., 2002; Ross et al., 2002), whilst the role played by Sp3 acetylation has yet to<br />

be determined (Braun et al., 2001). In addition to SUMO preventing<br />

ubiquitination <strong>of</strong> IKBa, the acetylation <strong>of</strong> Smad7 has been shown to protect<br />

Smad7 from ubiquitin-mediated degradation, suggesting competition between<br />

acetylation and Ubiquitin at critical lysine residues (Gronroos et al., 2002). Thus<br />

the complex relationship between the various post-translational modifiers can<br />

potentially enhance or repress the functional potential <strong>of</strong> various substrates. This<br />

modulation <strong>of</strong> protein function will provide an important mechanism to help<br />

control the diverse arrays <strong>of</strong> gene expression programs that are required for the<br />

lives <strong>of</strong> a complex organism (reviewed in Freiman and Tjian, 2003).<br />

1.6 The Ubl Conjugation Pathway<br />

The Ubl modifier is first activated by its cognate El enzyme, that uses ATP to<br />

adenylate the exposed C-terminal glycine residue <strong>of</strong> the Ubl (2) (Figure 11,<br />

numbers in brackets refer to position in schematic), following which a thioester<br />

bond is formed between the C-terminus <strong>of</strong> the Ubl and a cysteine residue in the<br />

El, releasing AMP (3). The Ubl is transferred from the El to the E2, in a<br />

transesterification reaction (4). The thioester linked E2-Ubl, either with or<br />

without the help <strong>of</strong> an E3 ligase activity, forms an isopeptide linkage between the<br />

C-terminus <strong>of</strong> the Ubl and an E-amino group <strong>of</strong> a lysine or N-terminal residue <strong>of</strong><br />

a substrate protein (5).<br />

1.6.1 The El super family<br />

The common feature <strong>of</strong> all the known ubiquitin-like protein modification<br />

systems is in the activation <strong>of</strong> the enzymatic cascade. This is achieved via a<br />

unique El, activating enzyme, for each <strong>of</strong> the conjugation pathways. Currently<br />

El enzymes have been found for ubiquitin, ISG15, NEDD8, APG12, SUMO,<br />

67<br />


URM1, and UFM1 conjugation pathways (Figure 12). The ubiquitin-activating<br />

enzyme exists as two is<strong>of</strong>orms, Ela (117 kDa) and Elb (110 kDa). Ela is<br />

phosphorylated and localised to the nucleus in a cell cycle-dependent manner<br />

(<strong>St</strong>ephen et al., 1996), due to the presence <strong>of</strong> a nuclear localisation sequence<br />

(SPLSKKRR), which is essential for both Ela phosphorylation and nuclear<br />

localisation (<strong>St</strong>ephen et al., 1997). In contrast Elb is predominantly cytoplasmic<br />

and is not phosphorylated (<strong>St</strong>ephen et al., 1996). Whilst the El's for ubiquitin,<br />

ISG15, APG12, and URM1 are monomeric, the El enzymes for NEDD8 and<br />

SUMO are heterodimeric, being composed <strong>of</strong> APPBPI/UBA3 (Gong et al.,<br />

1999), and SAE1/SAE2 (Desterro et al., 1999) respectively (Figure 12). UBA3<br />

and SAE2 are homologous to the C-terminal <strong>of</strong> the ubiquitin El, and contain<br />

both the nucleotide-binding motif involved in adenylation and the catalytic<br />

cysteine involved in the thioester intermediate. SAE1 and APPBP1 conversely,<br />

share homology with the N-terminal segment <strong>of</strong> the ubiquitin-activating enzyme.<br />

Although no ubiquitin-like modifiers exist in bacteria, there are members <strong>of</strong> the<br />

El super family present. There are distinct parallels between ubiquitin-like<br />

activation and the biosynthesis <strong>of</strong> several sulphur-containing enzyme c<strong>of</strong>actors<br />

(Furukawa et al., 2000). The microbial synthesis <strong>of</strong> thiamine and molybdopterin<br />

each requires a specific sulphurtransferase, ThiF or MoeB, respectively. All the<br />

El proteins identified to date share, along with ThiF and MoeB, conserved<br />

motifs and the same structural folds. In both cases the sulphur chemistry <strong>of</strong> both<br />

involves the addition <strong>of</strong> a sulphur atom to the carboxyl-terminal carboxyl group<br />

<strong>of</strong> short polypeptides, ThiS or MoaD, for ThiF and MoeB respectively, each <strong>of</strong><br />

which terminates in a glycine-glycine dipeptide. Similar to El enzymes, MoeB<br />

activates the C terminus <strong>of</strong> MoaD to form an acyl-adenylate. Subsequently, a<br />

sulphurtransferase converts the MoaD acyl-adenylate to a thiocarboxylate that<br />

acts as the sulphur<br />

donor during Molybdenum (Moco) biosynthesis (Leimkuhler<br />

et al., 2001 and Pitterle et al., 1993). The structures <strong>of</strong> these bacterial members<br />

<strong>of</strong> the El super family have been resolved (Wang et al., 2001 and Rudolph et al.,<br />

2001) including the MoeB-MoaD complex (Lake et al., 2001). The similarities<br />

in both structure and function <strong>of</strong> ubiquitin and the ubls, together with their<br />

activating enzymes, and those <strong>of</strong> the ThiS-ThiF and MoeB-MoaD pathways<br />

68<br />

4'


O<br />

EI-Sll EI-SII E2-Sll ýý<br />

Ubl/JýbFi ATP Ub MP Ubl/\S-E1 -Ubl/\S-E2 t~ tc Ub NH-Protein<br />

(2) (3) (4) (5)<br />

Figure 11. Schematic representation <strong>of</strong> a Ubl protein conjugation. The Ubl modifier is<br />

first activated by its cognate E1 enzyme, in an ATP dependent reaction, to adenylate the<br />

exposed C-terminal glycine residue <strong>of</strong> the Ubl (2), following which a thioester bond is<br />

formed between the C-terminus <strong>of</strong> the Ubl and a cysteine residue in the El, releasing<br />

AMP (3). The Ubl is transferred from the El to the E2, in a transesterification reaction<br />

(4). The thioester lonked E2-Ubl, either with or without the help <strong>of</strong> an E3 ligase, forms<br />

an isopeptide linkage between the C-terminus <strong>of</strong> the Ubl and an amino group <strong>of</strong> a lysine<br />

residue <strong>of</strong> a substratre protein (5).


Figure 12. Schematic diagram showing the four conserved regions found in Ub and Ub-<br />

like protein, El activating proteins, and -a phylogram showing the evolutionary<br />

relationship <strong>of</strong> the El proteins. (i) The, monomeric El enzymes for the activation <strong>of</strong> Ub<br />

in both humans- and S. cerevisiae (hUBA1 and scUBA! ), and the heterodimeric El<br />

enzymes for NEDD8, its yeast homologue, Rubl; SUMO and its yeast homologue Smt3<br />

are shown. Four conserved regions are found in the majority <strong>of</strong> all Ub and Ub-like<br />

protein El activating enzymes, indicated I to IV, as defined by (Johnson et al., 1997).<br />

Recently an El described for the Ufml Ubl has been characterised,<br />

that is significaantly<br />

smaller than the other El proteins and contains only one <strong>of</strong> the conserved domains.<br />

Adapted in part from (Desterro et al., 1999). (ii) Phylogram <strong>of</strong> various El proteins and<br />

bacterial El-like proteins. Phylogram was constructed via the ClustaiW alignment<br />

program (www. ebi. ac. uk/clustalw) and was generated with full length sequences.


CFs 632<br />

hUbiquitin hUBAI IV<br />

. 4-197 301-438 4-50406 608-664 926-913<br />

0 6(X)<br />

ScUbiquitin ScUBA I IV<br />

(ii)<br />

Rub]<br />

NEDD8<br />

S mB p<br />

Ula lp<br />

18-162 322-40.; 416-574 576-632 793-995 1024<br />

(168<br />

148<br />

UBA3p . 391-402 IV<br />

APP-BPI<br />

hUBA3<br />

14-160 451 5341<br />

1-138<br />

Aos lp cys 177<br />

UBA2p<br />

SUMO-1<br />

SAE]<br />

SUMO-2<br />

15-161 26(-347<br />

2(X)-276 299<br />

c ý. s 213<br />

IV<br />

49-1£ 3 245-321 441<br />

in<br />

I IV<br />

3-151 153-209 303-395 636,<br />

(y' 1 - 1-3<br />

SUMO-3 SAE2 18-)(4 263-_346; IV<br />

Uba5<br />

Move<br />

THF<br />

9-147 149-3(9 ? li-*X (AO<br />

Moee-like<br />

('ý s 250<br />

72-229 404<br />

IJba4<br />

SAE2<br />

- USE I<br />

UBE1L<br />

- UbaS<br />

SAE1<br />

X97<br />

1058<br />

APPBP1<br />

1 -. 9


make it likely that these eukaryotic protein post-translational modification<br />

pathways evolved from a similar protein-based sulphide donor system in<br />

bacteria.<br />

The initial catalytic step in ubl activation involves the formation <strong>of</strong> a<br />

high-energy thiol-ester linkage between the El and the carboxyl-terminal glycine<br />

<strong>of</strong> the ubiquitin-like protein. This process is ATP dependent and results in the<br />

production <strong>of</strong> AMP and PPi. The mechanism for this energy dependent<br />

activation was described some twenty-one years ago (Haas et al., 1982). The<br />

crystallization <strong>of</strong> the El heterodimer for NEDD8, APPBPI-UBA3, provides<br />

insight into functionality <strong>of</strong> the El (Walden et al., 2003). The El displays a<br />

large active site groove that is divided into two distinct clefts, with ATP being<br />

bound in one cleft whilst NEDD8 is bound in the other. In the NEDD8 El, there<br />

is a distance <strong>of</strong> some 30A between the adenylation site and the catalytic cysteine,<br />

Cys216. In order for NEDD8 (Ubl) to transfer from the adenylate, a-significant<br />

conformational change is required, as the catalytic cysteine is required to move<br />

by at least 10A, to allow the formation <strong>of</strong> the thioester bond (reviewed in<br />

VanDemark et al., 2003). The crystal structure also revealed that UBA3, which<br />

shares sequence similarity with the C-terminal half <strong>of</strong> the ubiquitin El, adopts an<br />

ubiquitin-like fold. The ubiquitin-like fold <strong>of</strong> UBA3 sits primarily in cleft one,<br />

and its deletion reduces the binding <strong>of</strong> the E2 enzyme (Walden et al., 2003).<br />

Thus El functions in a two-step process.<br />

Biochemical analysis <strong>of</strong> point mutants <strong>of</strong> ubiquitin had previously<br />

identified the important roles <strong>of</strong> arginine 42 and 72 in the interaction with<br />

ubiquitin-activating enzyme (Burch et al., 1994). The corresponding residues in<br />

SUMO family members is either glutamate or glutamine and alanine in NEDD8.<br />

The crystal structure <strong>of</strong> the NEDD8 El enzyme, indicates that Ala 72 <strong>of</strong> NEDD8<br />

interacts with the hydrophobic Leu 206 and Tyr 207 residues <strong>of</strong> UBA3. The<br />

orientation <strong>of</strong> the residues corresponding to Leu 206 and Tyr 207 in other El<br />

enzymes correlates with the nature <strong>of</strong> the side chain at position 72 in the other<br />

Ubls (Walden et al., 2003). Mutation <strong>of</strong> these two key residues <strong>of</strong> UBA3 to<br />

aspartate, reduces NEDD8 adenylation, but was shown not to be sufficient for<br />

72


APPBPI-UBA3 to promote adenylation <strong>of</strong> ubiquitin, showing that specificity is<br />

regulated by other conformation restraints (Walden et al., 2003).<br />

The El enzyme for the ISG15 ubl, UBE1L, was originally identified as<br />

being reduced in expression in lung cancers (Kok et al., 1993 and McLaughlin et<br />

al., 2000), but it was not shown to be the El enzyme for ISG15 until later (Yuan<br />

et al., 2001). Although pharmacological targeting <strong>of</strong> a Ubls El enzyme may be<br />

unlikely to be <strong>of</strong> therapeutic interest due to lack <strong>of</strong> specificity, viral proteins do<br />

target El function. Influenza B virus strongly induces the ISG15 protein during<br />

infection via the interferon response but the virus has evolved ways to counteract<br />

this response. Influenza B virus NS1 protein, is capable <strong>of</strong> preventing ISG15<br />

from being activated by UBE1L (Yuan et al., 2001), thereby preventing ISG15<br />

conjugation to substrates. El activity can also be regulated by cellular stress,<br />

such as oxidative stress. The activity <strong>of</strong> the ubiquitin-activating enzyme<br />

increases 2-6.7 fold, as determined by thiol ester formation, during recovery<br />

from oxidation (Shang et al., 1997).<br />

Several chemicals have also been reported which act to inhibit the ubiquitin-<br />

activating enzyme. Panepophenanthrin, from the mushroom strain Panus rudis,<br />

was isolated during screening trials for proteasome inhibitors (Sekizawa et al.,<br />

2002). Following the discovery <strong>of</strong> this natural product, Lei and co-workers,<br />

reported on the total synthesis <strong>of</strong> (+)-panepophenanthrin (Lei et al., 2003).<br />

However although both report the inhibition <strong>of</strong> the El-ubiquitin complex there is<br />

no evidence as to the specificity <strong>of</strong> this mechanism, as to whether it functions as<br />

a general E1 inhibitor or an ubiquitin specific inhibitor..<br />

Interestingly there is an example <strong>of</strong> a protein with both E1/E2 activities,<br />

TAF, 1250 in Drosophila (Pharr et al., 2000). TAFI, 250 is the central subunit<br />

within the general transcription factor TFIID, which was shown by Pham and<br />

colleagues to mediate monoubiquitination <strong>of</strong> histone 1.<br />

73


1.6.2 The E2 super family<br />

The second enzyme in the cascade, known as E2, forms a thioester with<br />

its cognate Ubl, following their maturation and subsequent activation by the El.<br />

Ubcs (ubiquitin conjugating enzymes) can be placed into one <strong>of</strong><br />

four broad<br />

groups, by comparison <strong>of</strong> their amino acid sequences (Jentsch, 1992). Class I<br />

enzymes possess a conserved catalytic core domain <strong>of</strong> about 150 residues with a<br />

minimum <strong>of</strong> 25% sequence identity. Class II enzymes have an extra C-terminal<br />

extension attached to the core domain, whilst Class III enzymes have an N-<br />

terminal extension to the core domain. Class IV enzymes possess both the N-<br />

and C-terminal extensions. Specificity for both Ubl and substrate recognition is<br />

in part determined by these extensions to the core domain, whilst amino acid<br />

insertions and other sequence and structural heterogeneity will undoubtedly play<br />

important roles.<br />

As it was not initially recognised that Ubls other than ubiquitin had<br />

distinct E2's, and as they are highly homlogous, the E2's are known as Ubcs.<br />

Indeed the E2 for SUMO, Ubc9, when initially discovered was thought to act in<br />

the ubiquitin system (Yasugi and Howley, 1996), whilst the E2 for the NEDD8<br />

pathway is named Ubcl2 (Gong et al., 1999). Subsequently Ubc9 was shown<br />

not be involved in the ubiquitin pathway, but the SUMO conjugation pathway<br />

(Desterro et al., 1997 and Schwartz et al., 1998).<br />

The Ubcs from the SUMO and NEDD8 pathways are incapable <strong>of</strong><br />

forming thioesters with either ubiquitin or any other Ubl (Desterro et al., 1997,<br />

Schwartz et al., 1998, and Gong et al., 1999), demonstrating the specificity for<br />

its own unique conjugation pathway. The ubiquitin pathway is reported to<br />

consist <strong>of</strong> eleven Ubc is<strong>of</strong>orms, and these is<strong>of</strong>orms have different substrate<br />

specificities. Ubcl, Ubc4, and Ubc5 would appear to be involved in general<br />

ubiquitination pathways which lead to the degradation <strong>of</strong> abnormal proteins<br />

(Seufert et al., 1990a, Seufert et al., 1990b). Ubc2 (Rad6) is involved in DNA<br />

damage repair and sporulation (Finley and Chau, 1991). Ubc3 (cdc34) is linked<br />

to the control <strong>of</strong> cyclin stability (Finley and Chau, 1991). Ubc6 has been<br />

demonstrated to be essential for protein translocation through the membrane<br />

(Sommer and Jentsch, 1993). UbclO is involved in the biogenesis <strong>of</strong><br />

74


peroxisomes (Wiebel and Kunau, 1992). Ubc13 forms a heterodimeric complex<br />

with mms2, to promote Lys63 ubiquitin chain formation. This heteromeric<br />

complex is largely cytosolic, but relocates to the nucleus following DNA<br />

damage, where they are subsequently recruited to chromatin (Ulrich et al., 2000).<br />

Thus despite belonging to the same class <strong>of</strong> enzyme and possesssing homologous<br />

amino acid sequences, there is distinct, and specific, functional versatility, such<br />

as the promotion <strong>of</strong> different ubiquitin chain linkages. Modification <strong>of</strong> the<br />

ubiquitin-conjugating enzyme, to create chimeric proteins containing appropriate<br />

protein-binding peptides fused to their C-termini, redirects the specificity <strong>of</strong><br />

ubiquitination (Sullivan and Vierstra, 1991, Gosink and Vierstra, 1995).<br />

The crystal structures for both Ubc9 and Ubc9 bound to a substrate<br />

RanGAPI have been resolved (Tong et al., 1997; Giraud et al., 1998; Bernier-<br />

Villamor et al., 2002) and along with mutagenesis and biochemical studies has<br />

provided a working hypothesis for the role <strong>of</strong> Ubc9 in SUMO transfer. A<br />

catalytic cysteine residue (C93 in Ubc9) is required to accept the transfer <strong>of</strong> the<br />

Ubl from the El to the E2. The Ubcs also possess a conserved asparagine<br />

residue, Asp85 in Ubc9, which whilst being dispensable for both E2 folding and<br />

E2-Ub thiol ester formation is required for E2-catalysed Ubl conjugation (Wu et<br />

al., 2003). Mutational analysis <strong>of</strong> residues in Ubc9, which are not conserved<br />

between Ubcs, AsplOO and LyslOl, which are located close to the active site<br />

cysteine in the tertiary structure <strong>of</strong> Ubc9, inhibit both the transesterification<br />

reaction from SAE1/SAE2 and substrate recognition (Tatham et al., 2003).<br />

NMR experiments had previously shown that SUMO interacted with Ubc9<br />

through its ubiquitin domain and that Ubc9 interacted with SUMO through a<br />

structurally conserved region (Liu et al., 1999). The elucidation <strong>of</strong> the crystal<br />

structure for Ubc9-RanGAPI revealed extensive RanGAPI interactions with the<br />

surface <strong>of</strong> Ubc9 and provided the mechanism for interaction between the SUMO<br />

modification motif WKXE/D and Ubc9 recognition. Conserved residues on the<br />

surface <strong>of</strong> Ubc9 create shallow grooves, complementary electrostatic, and<br />

hydrogen bonding interactions form a platform into which the IFKXE/D motif<br />

can specifically dock, resulting in the proper orientation and coordination <strong>of</strong> the<br />

acceptor lysine (Bernier-Villamor et al., 2002). This was consistent with earlier<br />

work, which had previously identified a substrate recognition site on Ubc9 by<br />

75


NMR (Lin et al., 2002). It is important to remember though, that not all <strong>of</strong> the<br />

substrates for SUMO modification possess the motif. The N-terminus <strong>of</strong> Ubc9,<br />

and by analogy other E2 enzymes due to its conservation, also plays important<br />

regulatory roles. R 13A/K 14A and R 17A/K 18A mutations in Ubc9 were shown<br />

to disrupt the interaction with SUMO, whilst retaining some degree <strong>of</strong> interaction<br />

with the El enzyme (Tatham et al., 2003). Furthermore, although the ability <strong>of</strong><br />

the mutant to transfer SUMO-1 from El to E2 was significantly disrupted, the<br />

transfer <strong>of</strong> SUMO-1 to substrate was unaffected. Thus, although distant from the<br />

active site these conserved residues within the N-terminus, and hence the N-<br />

terminus still plays a modulatory role.<br />

Work on Ubc9p, the Ubc9 from Saccharomyces cerevisiae, identified<br />

regions for Smt3p-Smt3p chain formation and thioester bond formation.<br />

Bencsath and co-workers showed that the surfaces involved in thioester bond<br />

formation map to the surface <strong>of</strong> the El binding site, and that the same surface<br />

binds Smt3p. Furthermore, it was demonstrated that a mutation in Ubc9p that<br />

disrupts El binding, but not Smt3p binding, impairs thioester bond formation,<br />

which suggests that it is the El interaction at this site that is crucial (Bencsath et<br />

al., 2002). The authors raise the point that as other E2s utilise this surface for<br />

binding to ubls, E3s, and other proteins and as such creating a multipurpose<br />

protein scafold, this could suggest that the entire El-E2-E3 conjugation pathway<br />

has coevolved for a given ubiquitin-like protein.<br />

As in the case <strong>of</strong> the El enzyme, certain viruses have evolved with these<br />

pathways. The African swine fever virus genome encodes its own viral<br />

ubiquitin-conjugating enzyme, UBCvI (Hingamp et al., 1992, Rodriguez et al.,<br />

1992). In vitro recombinant UBCvI can self-ubiquitinate, as well as ubiquitinate<br />

histones, and the ASFV virion protein, PIG 1 (Hingamp et al., 1992, Hingamp et<br />

al., 1995). In vivo evidence indicates that a DNA binding protein, SMCp is a<br />

substrate for the virally subverted ubiquitin pathway (Bulimo et al., 2000).<br />

76


1.6.3 E3 ligases<br />

1.6.3.1 Ubiquitin E3s<br />

The high degree <strong>of</strong> sequence and functional similarity <strong>of</strong> the various E1<br />

and E2 enzymes <strong>of</strong> the Ubl pathways is not found to the same extent for the final<br />

enzymes, the ligases, or E3 protein. The most studied E3s belong to the<br />

ubiquitin pathway although, due to the large number <strong>of</strong> E3s <strong>of</strong> this pathway (they<br />

number into the hundreds) there still remains a considerable amount <strong>of</strong> effort<br />

before they are fully understood. In general the E3 reaction involves at least two<br />

distinct steps, beginning with the binding to the substrate via the ubiquitination<br />

signal, followed by the covalent ligation <strong>of</strong> one or more ubiquitin proteins to the<br />

substrate, and as such ubiquitin ligases are defined as "enzymes that bind,<br />

directly or indirectly, specific protein substrates and promote the transfer <strong>of</strong><br />

ubiquitin, directly or indirectly, from a thioester intermediate to amide linkages<br />

with proteins or polyubiquitin chains". Despite the large number <strong>of</strong> ubiquitin<br />

ligases, all known E3s feature one <strong>of</strong> two structural elements, with the exception<br />

<strong>of</strong> p300, which will be discussed later. The first <strong>of</strong> these ligases are known as<br />

HECT domain E3s. Work on the human papiloma virus (HPV) showed that the<br />

ubiquitin dependent degradation <strong>of</strong> p53 depended on the HPV E6 gene product<br />

and a -100 kDa cellular protein named E6-AP (E6-associated protein). E6 and<br />

E6-AP form a complex that functions as a p53-specific E3. The final third <strong>of</strong> the<br />

E6-AP sequence is 35-45% identical to distinct domains in a large number <strong>of</strong><br />

identified proteins. These domains are collectively known as the HECT domain<br />

(homologous to E6-AP carboxyl terminus). The HECT domain contains a<br />

strictly conserved cysteine residue positioned -35 residues upstream <strong>of</strong> the C-<br />

terminus. The conserved cysteine <strong>of</strong> E6-AP is required for p53 ubiquitination, as<br />

it acts as a site <strong>of</strong> thiol ester formation with ubiquitin. There are numerous lines<br />

<strong>of</strong> evidence to indicate that all HECT domain E3s use a similar mechanism <strong>of</strong><br />

covalent catalysis, <strong>of</strong>ten with UbcH7 or UbcH8 as their E2. The construction <strong>of</strong><br />

HECT E3s is modular, the unique N-terminus <strong>of</strong> each family member interacts<br />

with specific substrate(s), whilst the HECT domain mediates E2 binding. The<br />

second ubiquitin ligase types, are known as RING E3s. ' Many eukaryotic<br />

proteins display a series <strong>of</strong> histidine and cysteine residues with a characteristic<br />

spacing that allows for the coordination <strong>of</strong> two zinc ions in a cross-brace<br />

77


structure called the Really Interesting New Gene (RING) finger. RING finger<br />

proteins number in the hundreds and have been implicated in a wide range <strong>of</strong><br />

cellular functions. The conservation in the RING fingers is not in the sequence,<br />

but rather the spacing <strong>of</strong> the zinc ligands. As the zinc ions and their ligands are<br />

catalytically inert, it suggests that RING fingers do not function as chemical<br />

catalysts, but rather as molecular scaffolds that function to bring other proteins<br />

together. There are two varieties <strong>of</strong> ubiquitin RING E3s, those that function as<br />

single-subunit and those that form part <strong>of</strong> a multi-subunit complex (reviewed in<br />

Deshaies, 1999; Borden, 2000; Joazeiro and Weissman, 2000). That the single-<br />

subunit E3s function alone has yet to be proven categorically, as many can form<br />

complexes with proteins besides their substrate, but so far only the E2 has been<br />

shown as a requirement for substrate ubiquitination. Following the binding <strong>of</strong><br />

their E2, single-subunit RING E3s have the ability to catalyse their own<br />

ubiquitination in vitro. This autoubiquitination, in some circumstances, is due to<br />

ubiquitin attachment to the RING protein, or fusion protein, and as such<br />

represents. an in vitro artefact. However, in other situations the RING E3<br />

autoubiquitination has been demonstrated in vivo, and as such has the potential to<br />

represent a physiological mechanism <strong>of</strong> E3 down-regulation, as is the case for<br />

both c-Cbl and Mdm2. c-Cbl is involved in the ubiquitin-dependent down-<br />

regulation <strong>of</strong> activated receptor protein tyrosine kinases (RTKs) (Joazeiro et al.,<br />

1999; Waterman et al., 1999; Joazeiro and Weissman, 2000). Mdm2<br />

ubiquitinates itself and the tumour suppressor protein, p53, and this relationship<br />

will be discussed in detail later. The second type <strong>of</strong> RING E3 requires that it is<br />

incorporated into a multiprotein complex to display ligase activity. There are<br />

three types <strong>of</strong> known multisubunit E3s, in which a small RING finger protein is<br />

an essential component. The three E3s are the SCF E3s (Skp1-Cullin-F-box<br />

protein), the APC (Anaphase-Promoting Complex), and the VCB (VHL-Elongin<br />

C/ Elongin B) E3s (reviewed in Tyers and Willems, 1999). These multisubunit<br />

E3s possesss a -100-residue RING finger protein, Rbxl (Hrtl/Rocl) that<br />

functions as an organiser, suportive <strong>of</strong> the scaffold hypothesis. Rbxl interacts<br />

strongly with the cullin protein family; cullinl in SCF and culling in VCB<br />

complexes. The assembly <strong>of</strong> the appropriate cullin, Rbxl, and specific E2, at<br />

least in vitro, results in an active ligase that will either autoubiquitinate the E2 or<br />

78


produce free polyubiquitin chains. The APC E3 contains Apc 11, a RING finger<br />

protein, amongst its many subunits.<br />

Single-subunit RING E3s directly recognise the ubiquitination signals <strong>of</strong><br />

their substrates, through domains that are distinct from their RING finger. In<br />

multisubunit RING Eis, the substrate recognition is mediated through a separate<br />

subunit. The Skpl protein <strong>of</strong> the SCF complex, for instance, recognises the<br />

substrate-specific F-box proteins through their F-box motif (Figure 13), whilst<br />

the recognition <strong>of</strong> the VCB E3 ligase is mediated through the pVHL, the product<br />

<strong>of</strong> the Von Hippel-Lindau tumour suppressor gene. The pVHL protein is<br />

recruited to the complex through interactions with the heterodimeric adaptor<br />

proteins Elongin B and Elongin C. Mutations <strong>of</strong> pVHL are <strong>of</strong>ten found in a<br />

subset <strong>of</strong> cancers, and these mutations are predicted to destabilise the interaction<br />

<strong>of</strong> pVHL with Elongin B/C (Figure 13).<br />

1.6.3.2 SUMO E3s<br />

Since the initial discovery <strong>of</strong> the SUMO conjugation system, an<br />

intriguing question was whether or not an E3 ligase activity was required for<br />

SUMO modification. <strong>St</strong>udies had shown that there was no necessity for an<br />

SUMO E3-like enzyme for modification in vitro, and unlike the ubiquitin<br />

pathway, the SUMO E2, Ubc9 directly bound to a substrate through the WKXE<br />

motif. This however, did not rule out the possibility <strong>of</strong> an E3-like activity in<br />

vivo, being required for the conjugation <strong>of</strong> those substrates that lack the WKXE<br />

motif. This question remained unanswered for some five years until the<br />

discovery <strong>of</strong> Sizl a yeast member <strong>of</strong> the mammalian protein inhibitor <strong>of</strong><br />

activated STAT (PIAS) family, as a potential SUMO E3-like protein (Takahashi<br />

et al., 2001, Sachdev et al., 2001, Johnson and Gupta, 2001).<br />

PIAS proteins, as their name would suggest, are involved in the STAT<br />

signalling transduction pathway. The STAT signal transduction pathway is<br />

activated following the binding <strong>of</strong> cytokines to the cell surface receptors, and<br />

induces a variety <strong>of</strong> cellular responses. Janus (JAK) kinases are constitutively<br />

79


ED,<br />

0<br />

SCF (Skipl-Cullin-F-box protein)<br />

F-Box > Substrate<br />

-0 -ý<br />

1i2 c'dc ýn<br />

P 13<br />

ß-TrCP 1KB, ß--catcnin<br />

VHL (VHL-Elongin C/Elongin B)<br />

w<br />

ffli<br />

F-Box Substrate<br />

pVHL<br />

HIFIu<br />

Figure 13. E, 3 ubiquitin ligase complexes. Similar overall architecture between the SCF<br />

and VHL complexes which participate in the ubiquitin-dependent degradation <strong>of</strong> many<br />

cellular proteins. Each complex is composed <strong>of</strong> a set <strong>of</strong> adapter proteins that recruit<br />

different binding partners through specific protein-protein interactions. Complex subunits<br />

that share sequence similarity are shown in the same colour, and their respective substrates<br />

are indicated below. Adapted in part from Desterro et al, 2000.


associated with the cytoplasmic domain <strong>of</strong> cytokine receptors. Binding <strong>of</strong> the<br />

ligand to its receptor induces dimerisation <strong>of</strong> the receptor chains, bringing<br />

together two JAK kinases that are activated by transphosphorylation. Activated<br />

JAKs phosphorylate various substrates in the cell, including the receptor itself<br />

and STAT transcription factors. Phosphorylated STATs then dimerise and<br />

migrate to the nucleus where they activate the transcription <strong>of</strong> genes, which<br />

mediate the cytokine-induced biological response.<br />

PIAS family <strong>of</strong> proteins were identified as STAT interacting proteins<br />

using yeast two-hybrid screens. There are four PIAS proteins in vertebrates;<br />

PIAS1, PIAS3, PIASx, and PIASy (Liu etal., 1998). PIAS1 was identified as a<br />

specific interaction partner for STAT1, and PIAS3 was subsequently cloned on<br />

the basis <strong>of</strong> homology to PIAS 1. PIAS 1 and PIAS3 inhibit DNA binding <strong>of</strong> the<br />

STAT1 and STAT3 factors, thereby inhibiting signal transduction (Chung et al.,<br />

1997, Liu et al., 1998), whilst PIASy represses STAT1 and the activity <strong>of</strong> the<br />

androgen receptor without affecting DNA binding (Gross et al., 2001). The role<br />

<strong>of</strong> PIAS/Siz as SUMO E3s further increases the important roles played by these<br />

proteins.<br />

Interestingly Sizl/PIAS members contain a conserved domain with a<br />

similarity to a zinc-binding RING-domain, which is <strong>of</strong>ten found in ubiquitin<br />

ligases. Ubiquitin RING domain E3 ligases do not possess intrinsic enzymatic<br />

activity, rather they function as adapters that bring together the E2 and the<br />

substrate, and this was shown to be true for the PIAS family <strong>of</strong> SUMO E3s.<br />

Both Siz1 and PIASy were shown to physically bind Ubc9 and substrate, which<br />

were septins (Takahashi et al., 2001, Johnson and Gupta, 2001) and LEF1, in the<br />

initial examples (Sachdev et al., 2001). Sizl was shown to be essential for<br />

septin-Smt3 conjugation in vivo. While Sizl was not required for conjugation in<br />

vitro, the presence <strong>of</strong> an E3 greatly enhanced the efficiency <strong>of</strong> the reaction.<br />

In yeast there are only two known SUMO ligases, Sizl and Siz2.<br />

However sizl-siz2 double mutants do not show mitotic arrest suggesting that<br />

other E3 ligases remain to be discovered. In higher eukaryotes two other<br />

families <strong>of</strong> SUMO E3 ligases, RanBP2 (Nup358) (Pichler et al., 2002) and Pc2<br />

(Kagey et al., 2003), have been identified.<br />

81


RanBP2(Nup358) is a giant nucleoporin, comprising an amino-terminal<br />

700-residue leucine-rich region, four RanBP1-homologous domains, eight zinc-<br />

finger motifs, and a carboxyl terminus with high homology to cyclophilin<br />

(Yokoyama et al., 1995). RanBP2 is one <strong>of</strong> the key constituents <strong>of</strong> the<br />

RanGTPase system, a system that is integral for nuclear transport, serving as a<br />

docking factor for import complexes on their way to the nucleus. The other<br />

constituents <strong>of</strong> this system are the guanine nucleotide exchange factor RCC1<br />

(Ohtsubo et al., 1987); the RanGTPase-activating protein RanGAP 1 (Bisch<strong>of</strong>f et<br />

al., 1994); the Ran-binding protein RanBP1 (Coutavas et al., 1993); NTF2<br />

(Moore and Blobel, 1994); and the Impß-related transport receptors. The<br />

evolution <strong>of</strong> the distinct nuclear and cytoplasmic compartments that characterise<br />

eukaryotic cells, requires a means <strong>of</strong> separating the two compartments whilst<br />

allowing exchange <strong>of</strong> factors between the two. This is achieved by the nuclear<br />

envelope (NE), that is penetrated by nucleopore complexes (NPCs), thus<br />

allowing the exchange <strong>of</strong> macromolecules between the compartments. RanBP2<br />

has recently been shown to be involved in nuclear mRNA export (Forler et al.,<br />

2004). The compartmentalisation <strong>of</strong> eukaryotic cells, although considerably<br />

energy intensive, allows the regulation <strong>of</strong> key cellular events, such as the control<br />

<strong>of</strong> access <strong>of</strong> transcriptional regulators to chromatin. The nuclear envelope is<br />

spanned in places by the nuclear pore complexes (NPCs). The number <strong>of</strong> NPCs<br />

in a cell will depend on the demand for nuclear transport-and will vary depending<br />

on cell size, synthetic, and proliferative activity, with about 3000-5000 NPCs in a<br />

proliferating human cell (Gorlich and Kutay, 1999). Metazoan nuclear<br />

envelopes contain an additional structural element, called lamina, which is<br />

associated with the nuclear face <strong>of</strong> the inner nuclear membrane.<br />

RanBP2(Nup358) is a component <strong>of</strong> the short (100nm) filaments that extend<br />

from the cytoplasmic face <strong>of</strong> the NPC during interphase, which relocates to both<br />

spindle microtubules and kinetochores (Joseph et al., 2002). The relocation <strong>of</strong><br />

RanBP2 occurs in association with RanGAP I. RanBP2 has been demonstrated<br />

to play an important role in nuclear envelope breakdown. Higher eukaryotes<br />

have an open mitosis, requiring that once per. cell cycle the nuclear envelope<br />

breaks, and the nuclear and cytoplasmic compartments mix. This involves the<br />

disassembly <strong>of</strong> all major structural components <strong>of</strong> the nuclear envelope,<br />

including the nuclear pore complex. RanBP2 is essential for kinetochore<br />

82


function, shown by the defects <strong>of</strong> chromosome congression and segregration in<br />

the absence <strong>of</strong> RanBP2, and aberrant kinetochore formation, due to the inhibition<br />

<strong>of</strong> the assembly <strong>of</strong> kinetochore components (Salina et al., 2003). The<br />

implication <strong>of</strong> this is that RanBP2 plays an essential role in integrating nuclear<br />

envelope breakdown. with kinetochore maturation and function (Salina et al.,<br />

2003).<br />

RanBP2 also possessses SUMO E3 ligase activity. It was shown that<br />

RanBP2 interacted directly with Ubc9 and promoted the SUMO-1 modification<br />

<strong>of</strong> Sp100, by stimulating transfer <strong>of</strong> SUMO-1 from Ubc9 to Sp100 (Pichler et al.,<br />

2002). Although possesssing a RING finger domain, it was shown not to be<br />

required for SUMO E3 ligase ability, and both RanBP2-DRING and point<br />

mutations <strong>of</strong> the cysteine residues still possesssed E3 ability. Furthermore the<br />

treatment <strong>of</strong> recombinant protein with alkylating agents did not prevent RanBP2<br />

E3 activity towards Sp100, suggesting that thioester bond formation is not<br />

essential for this activity. RanBP2, as with the Siz/PIAS family <strong>of</strong> SUMO Eis,<br />

was capable <strong>of</strong> stimulating the formation <strong>of</strong> SUMO chains on itself, although the<br />

physiological relevance <strong>of</strong> this still remains to be elucidated.<br />

The third known SUMO E3 ligase identified to date, is a member <strong>of</strong> the<br />

Polycomb group <strong>of</strong> proteins. Polycomb group (PcG) proteins form large<br />

multimeric complexes (PcG bodies) and are associated with the stable repression<br />

<strong>of</strong> gene expression. PcG proteins were originally identified in Drosophila as<br />

repressors <strong>of</strong> homeotic gene expression (Kennison et al., 1995; Simon et al.,<br />

1995). Flies containing mutations in PcG genes showed aberrant expression <strong>of</strong><br />

homeotic genes, outside their normal segment boundaries, leading to the<br />

alteration <strong>of</strong> body part identity (Simon et al., 1995; Schumacher and Magnuson,<br />

1997). Many homologs <strong>of</strong> the Drosophila PcG proteins have been identified in<br />

mammals (Brock and van Lohuizen, 2001), and multiple core PcG complexes are<br />

found in a variety <strong>of</strong> cell types (Seawalt et al., 1998; Francis et al., 2001; Gunster<br />

et al., 2001). Although PcG complexes are associated with DNA, individual<br />

PcG proteins do not seem capable <strong>of</strong> directly binding DNA. (Francis et al.,<br />

2001). It is thought that the recruitment <strong>of</strong> PcG complexes to DNA, is mediated<br />

by sequence specific DNA binding proteins, such as, YY1, and E217 (Dahiya et<br />

al., 2001; Satijin et al., 2001; Ogawa et at., 2002). PcG complexes form<br />

83


discrete foci within the nucleus, termed PcG bodies (Gerasimova and Corces,<br />

1998; Saurin et al., 1998), and these bodies are <strong>of</strong>ten found near centromeres.<br />

The SUMO E3 ligase ability <strong>of</strong> a Polycomb protein, Pc2, was found, in the<br />

context <strong>of</strong> the SUMO modification <strong>of</strong> the transcriptional corepressor carboxyl-<br />

terminus binding proteins (CtBP), CtBP1 and CtBP2 (Kagey et al., 2003). CtBP<br />

was originally identified by its interaction with the adenovirus E1A protein,<br />

which was required for E1A transcriptional repression (Boyd et al., 1993;<br />

Schaeper et al., 1999). CtBP interacts with Pc2, leading to its recruitment in PcG<br />

bodies (Sewalt et al., 1999). Pc2 has recently been shown to be able to recruit<br />

Ubc9 to PcG bodies (Kagey et al., 2003), and could provide the first indication<br />

that PcG bodies, similar to PML bodies, may function as centres <strong>of</strong> SUMO<br />

modification within the nucleus. It was also noted that Pc2 could stimulate<br />

SUMO modification <strong>of</strong> CtBP2, both in vivo and in vitro, despite the absence <strong>of</strong> a<br />

consensus SUMO conjugation motif. This suggests that other potential SUMO<br />

substrates, which do not possessses a SUMO conjugation motif, could be<br />

modified in the presence <strong>of</strong> the correct E3.<br />

1.7 p300/CBP<br />

p300 and its paralogue CREB-binding protein (CBP) are large multi-domain<br />

proteins that play key roles in transcriptional regulation, signal transduction<br />

pathways, and protein stability. p300/CBP contain distinct, conserved, domains:<br />

a bromodomain, three cysteine-histidine (CH)-rich domains (CHI, CH2, and<br />

CH3), a KIK domain. The CBP binds to CREB via the KIK domain as well as<br />

binding to transactivation domains <strong>of</strong> other nuclear factors including Myb and<br />

Jun. Additionally p300/CBP possess an ADA2-homology domain; which shows<br />

homology to Ada2p, a yeast transcriptional coactivator (Figure 14) (Chan and La<br />

Thangue, 2001). Both p300 and CBP interact with numerous proteins and are<br />

components <strong>of</strong> many signalling pathways (Figure 14). p300 and CBP were both<br />

initially identified in protein interaction assays, p300 through its interaction with<br />

the adenoviral-transforming E1A protein (<strong>St</strong>ein et al., 1990; Eckner et al., 1994)<br />

and CBP through its association with the transcription factor CREB (Chrivia et<br />

al., 1993). CBP can specifically interact with the protein protein kinase A-<br />

phosphorylated, activated form <strong>of</strong> CREB, and in transfection experiments<br />

84


enhances the CREB"mediated activation <strong>of</strong> reporter constructs in vivo in a<br />

manner that depends on the integrity and function <strong>of</strong> the protein kinase A<br />

phosphorylation site (Kwok et al, 1994). E1A is a transforming viral protein that<br />

forces quiescent cells into S phase. E! A binds to the CH3 domain through the<br />

EIA amino-terminus and CR1 (conserved region 1) domains, and represses p300<br />

transcriptional activity. EIA also targets retinoblastoma protein (Rb), which is<br />

essential for the regulation <strong>of</strong> GI/S transition by controlling the transcription<br />

factor E2F. E2F activates a subset <strong>of</strong> genes whose products are involved in<br />

DNA synthesis or in cell cycle control. p300/CBP possessses intrinsic histone<br />

acetyltransferasc (tiAT) activity and this activity is regulated in a cell cycle<br />

dependent manner and shows a peak at the G1/S transition (Ait-Si-Ali et a!.,<br />

1998; Ait-Si-Ali et at., 2000). The intrinsic HAT activity <strong>of</strong> p300/CBP provides<br />

a means <strong>of</strong> influencing chromatin activity by modifying nucleosomal histones.<br />

p300/CBP is required for cell proliferation, performing an important function<br />

during the G 1/S transition, by influencing the activities <strong>of</strong> the genes required for<br />

cell cycle progression. Additionally p300 has a cell cycle inhibitory function,<br />

and is likely to regulate the promoters for genes required for arrest in GOIG 1.<br />

7.1.1 p300/CUP and growth control and signal transduction<br />

p300/CBP do not solely associate with Ela and E2F, rather there is a<br />

diverse and constantly increasing number <strong>of</strong> transcription factors, that have been<br />

shown to form stable physical complexes with, and respond to, the coactiviating<br />

properties <strong>of</strong> p300/CBP. Many insights into the functioning <strong>of</strong> p300/CBP come<br />

from investigating the role Ela plays in compromising the functions <strong>of</strong><br />

p300/CBP. Proto-oncogenes, such as c-Afyb, utilise CBP as a co-activator,<br />

unless compromised by Ela. Much thought has gone into understanding why a<br />

viral oncoprotcin would compromise mitogenic transcription factors. Theories<br />

put forward include the possibility p300/CBP may allow for a stronger<br />

coactivator to play a role in transcriptional activation. Alternatively, p300/CBP<br />

may be involved with regulating transcription <strong>of</strong> a subgroup <strong>of</strong> AP-1 or c-Myb-<br />

responsive genes that antagonise proliferation. It has been speculated, therefore,<br />

that some transcription factors use p300/CBP preferentially for regulating a<br />

category <strong>of</strong> genes or, alternatively, for responding to a particular transcriptional<br />

85


stimulus that, if inactivated, would be advantageous to the ultimate aims <strong>of</strong> a<br />

viral oncoprotein (Shikama et al., 1997).<br />

Ela can also cause an increase in the overall transcriptional activity <strong>of</strong><br />

some transcription factors, such as YY1. YY1 has transcriptional repression<br />

properties in the absence <strong>of</strong> Ela (Shi et al., 1991). Although YY1-binding sites<br />

are frequently present in transcriptional control regions, some sites function to<br />

repress transcription, and, in these cases, p300 is believed to be responsible for<br />

the repressive activity (Lee et al., 1995). This repression can be relieved by Ela<br />

binding, an effect dependent upon the N-terminal p300/CBP-binding domain in<br />

Eta. YY1 and Eta associate in vivo with p300 through distinct C-terminal<br />

domains, and it is thought that both proteins can associate with p300 at the same<br />

time (Shikama et al., 1997).<br />

p300/CBP are also connected to a further class <strong>of</strong> sequence-specific<br />

transcription factors, the ligand-dependent nuclear receptors. The ligand-binding<br />

domains <strong>of</strong> multiple nuclear receptors, including the retinoic acid (RAR and<br />

RXR), oestrogen (ER), progesterone (PR), thyroid hormone (TR), and<br />

glucocorticoid (GR) receptors, interact with the N-terminal region <strong>of</strong> p300/CBP<br />

(Figure 14). Several other coactivators are implicated in the hormone-dependent<br />

activation <strong>of</strong> nuclear receptors which can associate with receptor bound to<br />

p300/CBP.<br />

Signal transduction mediated by the JAK-STAT pathway also involves<br />

p300/CBP (Bhattacharya et al., 1996). Interferon a (IFN-a) induces the<br />

transcription <strong>of</strong> a variety <strong>of</strong> genes required to produce antiviral effects, many <strong>of</strong><br />

which require the ISGF3 transcription factor (Darnell et al., 1994). ISGF3 is a<br />

multicomponent activity composed <strong>of</strong> three subunits, STAT1 and STAT2,<br />

together with a weak DNA-binding component known as p48. Adenovirus Ela<br />

impedes transcriptional activation mediated by IFN-a by inactivating p300/CBP,<br />

which binds through its N-terminal region to STAT2. The compromisation <strong>of</strong><br />

the IFN signalling pathway is likely to have beneficial effects for viral<br />

replication.<br />

86<br />

ý!<br />

ff<br />

S<br />

ý1ý


Ei °ö<br />

ä<br />

Ei<br />

rA<br />

ö<br />

`ý ao<br />

0-0 .<br />

= 4. w<br />

"ggE<br />

LE öo<br />

2 "v<br />

BUö<br />

Nö<br />

cý UoE<br />

l. i «<br />

U<br />

aý y<br />

ävö<br />

3ý_ö<br />

öb<br />

00<br />

><br />

2<br />

.0<br />

,ý<br />

to o eu<br />

a ý. 'd Q<br />

°<br />

Ü<br />

2 e: .<br />

txo<br />

UmAb<br />

g<br />

..<br />

C<br />

U<br />

ice, JZ "r'<br />

12 -v<br />

U<br />

=0<br />

eK<br />

.G<br />

~ >'b N<br />

"<br />

Nw Ei<br />

0


ýy ÜO<br />

0<br />

0<br />

ÜUU 4)<br />

ÜÜ z<br />

u<br />

A. -1 x<br />

O<br />

N<br />

X CD<br />

Co rrý<br />

wýM<br />

o°,<br />

äk<br />

Ct<br />

Gq<br />

U<br />

U<br />

a1<br />

M<br />

U<br />

N<br />

U<br />

oWOJ g<br />

UND<br />

U<br />

ä<br />

U<br />

U<br />

s.,<br />

C/1<br />

bA<br />

cri ý<br />

>w E<br />

Cl1 x ^-ý<br />

M<br />

ý Wä ßi ý'n.<br />

O<br />

N<br />

3 Ö<br />

C it<br />

ti H<br />

a1 cat<br />

Uri<br />

W<br />

N<br />

E-<br />

In<br />

110<br />

M4<br />

z<br />

N<br />

H<br />

C/ý<br />

a<br />

H<br />

. -ý<br />

ýC<br />

z<br />

a<br />

pa<br />

h<br />

>.<br />

ýU<br />

-S


7.1.2 p300/CBP and disease<br />

The essential role <strong>of</strong> p300/CBP is further shown by genetic studies.<br />

Inactivation <strong>of</strong> a single allele <strong>of</strong> cbp leads to bone malformations and<br />

monoallelic deletion <strong>of</strong> p300 results in a partially penetrant neural tube closure<br />

defect. In humans haploinsufficiency <strong>of</strong> cbp leads to Rubinstein-Taybi<br />

Syndrome (RTS) (Petri] et al., 1995; Kalkhoven et al., 2003), a congenital<br />

disease characterised by malformation <strong>of</strong> facial bones and digits as well as<br />

increased incidence <strong>of</strong> heart defects. <strong>St</strong>udies in mice show that CBP loss and<br />

histone deacetylation were observed in two separate pathological contexts, both<br />

in amyloid precursor protein-dependent signalling and amyotrophic lateral<br />

sclerosis, indicating that these modifications may well have roles in<br />

neurodegenerative diseases. It has been demonstrated that p300 and CBP play<br />

essential roles in neuronal apoptosis. p300/CBP also binds phospho-CREB, a<br />

transcription factor shown to be involved in neuroprotection (Lonze and Ginty,<br />

2002). p300/CBP degradation normally occurs through the proteasome pathway<br />

although p300/CBP can also be degraded by Caspase-6 (Rouaux et al., 2003).<br />

The controlled degradation <strong>of</strong> p300/CBP and resulting decrease in HAT activity<br />

is critical in neuronal apoptosis.<br />

A role for the gene encoding p300 as a tumour suppressor is suggested<br />

from studies where the gene is mutated in human tumour cells. Specifically,<br />

missense mutations occur in both colorectal and gastric carcinomas. The<br />

mutations identified alter residues 1399 and 1680, located within the C-terminal<br />

C/H-rich region. Further observations for p300/CBP roles as a tumour<br />

suppressor is provided from the studies in which the oncogenic capacity <strong>of</strong> Ela is<br />

compromised by coexpressing p300 (Smits et al., 1996). Further evidence<br />

suggesting a role for CBP in tumourigenesis, is provided by CBP translocation<br />

events that occur during myeloid leukaemia. In myeloid leukaemia, a large<br />

proportion <strong>of</strong> the MOZ gene is translocated to the gene encoding CBP, resulting<br />

in the fusion <strong>of</strong> the MOZ acetyltransferase domain to the largely intact CBP<br />

(Muraoka et al., 1996). The physiological significance <strong>of</strong> this translocation<br />

remains to be elucidated, although it is possible that the generation <strong>of</strong> this fusion<br />

protein will the acetyltransferase to be targeted constitutively to particular groups<br />

<strong>of</strong> p300/CBP-regulated genes.<br />

89<br />

F, ,


Despite the many functional similarities between p300 and CBP, they are<br />

not functionally interchangeable, as there are subtle differences in the expression<br />

<strong>of</strong> p300 and CBP during development, although mouse embryos nullizygous for<br />

either p300 or CBP die at midgestation, heterozygosity <strong>of</strong> cbp causes distinct<br />

haematological defects and an increased risk <strong>of</strong> developing cancer, an effect not<br />

seen' in mice with a single allele <strong>of</strong> p300 (Kung et al., 2000; reviewed in<br />

Goodman and Smolik, 2000). Further differences include the observations that<br />

fibroblasts derived from homozygous p300 knockouts are defective for retinoic<br />

acid receptor but not CREB signalling (Yao et al., 1998; Kawasaki et al., 1998).<br />

Full activation <strong>of</strong> CREB-dependent transcription by CBP depends on CBP<br />

phosphorylation at Ser301, an event that does not occur for p300 (Impey et al.,<br />

2002). p300 also plays a role in the apoptotic response to DNA damage<br />

following ionizing damage (IR) whilst CBP does not (Yuan et al., 1999).<br />

Additionally the Karposi sarcoma-associated virus protein vIRF - has been<br />

reported to be stimulated by CBP whilst repressed by p300 (Jayachandra et al.,<br />

1999).<br />

7.1.3 p300/CBP and chromatin remodeling<br />

Although p300/CBP possessses intrinsic HAT activity, in many<br />

circumstances, it appears to function as part <strong>of</strong> larger complexes. Indeed<br />

p300/CBP can bind to other acetyltransferases such as GCN5 and PCAF<br />

(p300/CBP-associated factor), resulting in the assembly <strong>of</strong> protein complexes<br />

consisting <strong>of</strong> two or more acetyltransferases (Yang et al., 1996). Acetylation <strong>of</strong><br />

multiple sites in the core histone tails is commonly associated with<br />

transcriptional activity, whilst hypo-acetylation is associated with transcriptional<br />

repression. The consequences <strong>of</strong> acetylation <strong>of</strong> lysine residues, within the N-<br />

terminal tails <strong>of</strong> histones can be varied, and forms part <strong>of</strong> the histone code, as<br />

described earlier. The acetylation <strong>of</strong> histones by p300/CBP, directed by the<br />

histone chaperone RbAp48 (Zhang et al., 2000), helps the transfer <strong>of</strong> H2B-H2B<br />

dimers from nucleosomes to NAP-1, a chaperone protein (Ito et al., 2000).<br />

Histones H2A and H2B are known to be absent from transcriptionally active<br />

chromatin and is readily exchanged out <strong>of</strong> transcribed chromatin in vivo.<br />

However increased histone acetylation alone is usually insufficient for<br />

90


transcriptional activation (Hebbes et al., 1994; Gregory et al.,<br />

1999). Efficient<br />

gene expression requires cooperation between acetyltransferases and other<br />

chromatin modifying activities such as ATP-dependent remodelling complexes<br />

(Hassan et al., 2001; Narlikar et al., 2002). Potentially as important may be the<br />

acetylation <strong>of</strong> transcription factors through factor acetyl-transferase activities<br />

(FAT). This activity was first determined in the context <strong>of</strong> the tumour<br />

suppressor p53 (Gu and Roeder, 1997), but there is growing evidence that<br />

acetylation <strong>of</strong> transcription factors, is a common feature in gene expression.<br />

p300/CBP form complexes with proteins other than acetyltransferases and can<br />

function as protein bridges, connecting different DNA sequence-specific<br />

transcription factors to the transcriptional apparatus (reviewed in Chan and La<br />

Thangue, 2001). The interactions <strong>of</strong> p300/CBP with components <strong>of</strong> the general<br />

transcriptional machinery, such as TFIID, TFIIB, and the RNA polymerase II<br />

holoenzyme (RNAPII) suggests that these interactions are likely to contribute to<br />

p300/CBP function (Kwok et al., 1994; Kee et al., 1996; Yuan et al., 1996;<br />

Mikecz et al., 2000). The simultaneous interaction <strong>of</strong> multiple transcription<br />

factors with p300/CBP could contribute to transcriptional synergy, where the<br />

interaction <strong>of</strong> two or more transcription factors has a combined effect greater<br />

than the sum <strong>of</strong> the individual transcription factors. Conversely, the competition<br />

for the binding to p300/CBP may mediate signal induced repression.<br />

Furthermore it has been rationalised that given the limited repertoire <strong>of</strong><br />

activators, co-activators, and c<strong>of</strong>actors that respond to diverse regulatory cues,<br />

cells may use cooperativity and transcriptional synergy so that a combination <strong>of</strong><br />

the few available ubiquitous, signal- and tissue-specific activators can create a<br />

potentially very large number <strong>of</strong> regulatory complexes.<br />

1.7.4 An alternate role for p300 as an E3 ligase<br />

The relationship between p300 and p53 is complex, with seemingly '<br />

contradictory reports <strong>of</strong> effects <strong>of</strong> p300 on the levels <strong>of</strong> p53 (Li et al., 1997;<br />

Kawai et al., 2001; Livengood et al., 2002). The p53 gene is a tumour<br />

suppressor gene that plays a critical role in safeguarding the integrity <strong>of</strong> the<br />

genome. p53 is mutated or part <strong>of</strong> its regulatory circuit is functionally inactivated<br />

in almost all cancers, which highlights its importance in preventing<br />

91


tumourigenesis. p53 is regulated by two interactive feedback loops. p53 and<br />

Hdm2 (Mdm2 in the mouse), form one <strong>of</strong> these feedback loops, in which p53<br />

positively regulates Hdm2 by activating HDM2 transcription, and Hdm2<br />

negatively regulates p53 by promoting p53 ubiquitination and subsequent<br />

degradation via the 26S proteasome, by functioning as an E3 ligase (Figure 15)<br />

(Sherr, 1998; Zhang and Xiong, 2001). E2F-1 and p14M (p19"' in the mouse)<br />

form an additional feedback loop, in which E2F-1 activates ARF transcription,<br />

and p14"RF promotes the ubiquitin dependent degradation <strong>of</strong> E2F-l. Furthermore<br />

p14 "RF interacts with Hdm2, inhibiting Hdm2 E3 ligase function, which will<br />

stabilise p53 (Bates et al., 1998; Zhang et al., 1998; <strong>St</strong>ott et al., 1998)(Fig. 15. ).<br />

Additionally p53 represses the transcription <strong>of</strong> the ARF gene (Fig. 15). SUMO<br />

modification adds an extra layer <strong>of</strong> complexity to this regulatory mechanism.<br />

Hdm2 and p53 are both modified by SUMO (Gostissa et al., 1999; Rodriguez et<br />

al., 1999; Muller et al., 2002; Xirodimas et -al., 2002). HDM2 and ARF<br />

expression has been shown to increase the level <strong>of</strong> SUMO modified p53 (Chen<br />

and Chen, 2003). This increase in SUMO modification was shown to be<br />

regulated by Hdm2 and Arf mediated nucleolar targeting (Chen and Chen, 2003).<br />

The acetylation, <strong>of</strong> three specific lysine residues in the C-terminus <strong>of</strong><br />

p53 increases the DNA binding capacity <strong>of</strong> p53, possibility due to a<br />

conformational change <strong>of</strong> an inhibitory regulatory domain. Ionizing radiation<br />

promotes the N-terminal phosphorylation <strong>of</strong> p53, and as a consequence increases<br />

the affinity between p53 and p300/CBP and results in the stimulation <strong>of</strong> p53<br />

DNA binding activity following its acetylation. However it has been reported<br />

that although acetylated p53 binds to short fragments <strong>of</strong> DNA with a greater<br />

affinity than that <strong>of</strong> non-acetylated, acetylated p53 does not appear to<br />

significantly affect p53 DNA binding activity on chromatin templates (Espinosa<br />

and Emerson, 2001). As such the growth suppression and apoptotic promoting<br />

functions <strong>of</strong> p300/CBP are mediated, in part, by their interaction with p53.<br />

CBP has been shown to act as a p53 coactivator that potentiates p53<br />

transcriptional activity. p53 interacts via its N-terminus activation domain with<br />

several domains within the C-terminus <strong>of</strong> p300/CBP, which can up-regulate<br />

expression <strong>of</strong> the Mdm2 gene. Mdm2 is the negative regulator <strong>of</strong> p53, acting as<br />

an ubiquitin ligase, and as such, increasing levels <strong>of</strong> Hdm2 lead to decreased<br />

92


(i) Transcriptional control.<br />

(ii) Post-translational control.<br />

do ý-.<br />

p53<br />

Proteosomal degradation<br />

\ýý<br />

Nucleolar targeting<br />

-<br />

p14ARF<br />

E2F-1<br />

Proteosomal degradation<br />

Figure 15. The p_53 functional circuit. (i)Transcriptional regulation <strong>of</strong> the core circuit.<br />

p53 positively regulates the expression <strong>of</strong> the HDM2 gene, whilst repressing expression<br />

<strong>of</strong> ARF. E2F-1 positively regulates ARF expression. (ii) Post-translational regulation<br />

<strong>of</strong> the core circuit. p14ARF promotes the SUMO modification <strong>of</strong> Hdm2 and p53, when<br />

in a complex with Hdm2. Hdm2 also undergoes autoubiquitination, thereby regulating<br />

its own stability. p14ARF further promotes the ubiquitination <strong>of</strong> E2F-1, whilst<br />

repressing Hdm2 ability to ubiquitinate p53. Effects on protein levels are indicated by (-<br />

), reduction, and (+), increase.


levels <strong>of</strong> p53. Decreasing levels <strong>of</strong> p53 will in turn lead to decreased expression<br />

<strong>of</strong> Hdm2, as previously mentioned. This auto-inhibitory feed back loop is<br />

subverted by viral protein such as E1A, leading to decreased levels <strong>of</strong> p53, and<br />

allows viral replication. p53 is the most commonly mutated tumour suppressor<br />

in the majority <strong>of</strong> cancers investigated to date. The Tax (human T-cell leukemia<br />

virus (HTLV-1)) protein is also capable <strong>of</strong> binding to p300/CBP, and directly<br />

competes for the binding <strong>of</strong> p53 at the CR2 (conserved region 2) domain <strong>of</strong><br />

p300/CBP. The mechanism by how E1A subverted the feedback loop, was only<br />

recently proposed. The N-terminal <strong>of</strong> p300 shows partial sequence homology to<br />

the E6-AP, an ubiquitin ligase and is found mutated in Angelman's syndrome.<br />

Angelman's syndrome is a neurological disorder, characterised by<br />

neurodevelopmental impairments. E6-AP when bound to E6 was also capable<br />

<strong>of</strong> targeting p53 for ubiquitination and this induces its degradation. Purified<br />

p300 showed ubiquitin ligase activity, in vitro, an activity that was abolished in<br />

the presence <strong>of</strong> E1A (Grossman et al., 2003). The ubiquitin ligase activity was<br />

further mapped to the N-terminal region <strong>of</strong> p300. In vitro experiments showed<br />

that p300 and Mdm2 were required for promoting polyubiquitination <strong>of</strong> p53<br />

(Grossman et al., 2003). In the absence <strong>of</strong> p300, only single ubiquitins were<br />

conjugated to p53. Expression <strong>of</strong> E1A caused a decrease in polyubiquitinated<br />

forms <strong>of</strong> p53, but not monoubiquitinated forms (Grossman et al., 2003). A<br />

caveat <strong>of</strong> this work was that p300 was only able to polyubiquitinate<br />

monoubiquitinated p53, and thus, may be functioning as an E4, promoting rather<br />

than controlling p53 polyubiquitination.<br />

However the intrinsic ubiquitin ligase<br />

ability <strong>of</strong> p300 could explain how it can control both p53 activation and its<br />

destruction. As p300 regulates a range <strong>of</strong> short-lived transcription factors, it is<br />

possible that it may act as an ubiquitin ligase in other cases, although this is yet<br />

to be proven. However there are conflicting reports on this as Li et al., have<br />

shown that p53 is monoubiquitinated and localised in the cytoplasm when Mdm2<br />

is expressed at low levels, whereas p53 is polyubiquitinated and subsequently<br />

degraded when Mdm2 is expressed at high levels (Li et al., 2003).<br />

94


1.7.5 p300 and transcriptional repression<br />

p300/CBP can also act as a transcriptional repressor via the modulation <strong>of</strong><br />

adaptor proteins, such as the proliferating cellular nuclear antigen. (PCNA). The<br />

PCNA protein can associate with p300 (Hasan et al., 2001), an association that<br />

potently inhibits both the acetyltransferase activity and transcriptional activation<br />

properties <strong>of</strong> p300 (Hong and Chakravarti, 2003). p300/CBP also contains a<br />

conserved transcriptional repression domain (Snowden et al., 2000). This<br />

repression domain was located in the N-terminal half <strong>of</strong> p300/CBP upstream<br />

from the bromodomain, termed cell cycle regulatory domain 1 (CRDI).<br />

Repression mediated by this domain could be relieved via coexpression <strong>of</strong><br />

p21"'af/`ipl a cyclin-dependent kinase inhibitor (Snowden et al., 2000). This<br />

stimulatory effect <strong>of</strong> p21waflcipl, in the context <strong>of</strong> p300 coactivation, was first<br />

reported for the p65 NF-xB subunit, resulting in the stimulation <strong>of</strong> KB-dependent<br />

gene expression (Perkins et al., 1997). Of special interest was the observation<br />

that the repression domain <strong>of</strong> p300 contained two contiguous "classic" SUMO<br />

consensus modification motifs.<br />

1.8. Aims <strong>of</strong> the project<br />

The aims <strong>of</strong> the work described herein were to investigate whether p300<br />

was a bona fide SUMO substrate, identify whether the tandem SUMO<br />

modification motifs were the sites <strong>of</strong> SUMO conjugation and elucidate the<br />

biological significance <strong>of</strong> the potential modification. SUMO modification <strong>of</strong><br />

p300 was to be investigated through use <strong>of</strong> a permanent cell line expressing<br />

6HIS-tag SUMO, allowing the pull-down <strong>of</strong> any associated proteins, under<br />

conditions that inactivate the SUMO-specific proteases, preventing substrate<br />

deconjugation. The identification <strong>of</strong> the SUMO conjugation site was to be<br />

investigated through the use <strong>of</strong> p300 mutants lacking the predicted conjugation<br />

domain in vivo, and subsequently in vitro using site-directed mutagenesis and in<br />

vitro SUMO conjugation assays.<br />

The occurrence <strong>of</strong> the tandem SUMO modification motifs in a previously<br />

identified transcriptional repression domain, CRD1, <strong>of</strong> p300 suggested a role for<br />

SUMO in the regulation <strong>of</strong> p300 transcriptional capacity. The relationship<br />

95


etween that <strong>of</strong> SUMO conjugation and transcriptional activity <strong>of</strong> p300 was<br />

investigated, through use <strong>of</strong> a Ga14 luciferase reporter system, with SUMO<br />

competent and deficient versions <strong>of</strong>'p300.<br />

This relationship could<br />

be further<br />

investigated through the control <strong>of</strong> the SUMO modification pathway, by either a<br />

dominant negative version <strong>of</strong> Ubc9 or over-expression <strong>of</strong> a SUMO-specific<br />

protease.<br />

Finally through use <strong>of</strong> emerging siRNA technology the potential <strong>of</strong><br />

functional differences between the three SUMO paralogues were to be<br />

researched. This work was to focus on the potential roles played in<br />

transcriptional regulation and localisation by the various SUMO is<strong>of</strong>orms.<br />

96


2. MATERIALS & METHODS<br />

97


2.1. Materials<br />

All materials not produced in house were purchased from Sigma unless<br />

otherwise stated.<br />

2.2. Antibodies<br />

HA-SUMO-1, HA-SUMO-2, and HA-SUMO-3 were detected in<br />

Western-blot experiments using mAb 12CA5 (at 1: 5000 dilution), which<br />

recognises YPYDVPDYA from influenza HA, obtained from BabCO.<br />

Ubc9 was detected using an affinity purified sheep pAB (in house) at a<br />

1: 1000 dilution.<br />

SUMO-1 was detected using the mouse anti-GMP1 mAb (1: 1000)<br />

(Zymed), and SUMO-2 and SUMO-3 were detected using a rabbit anti-Sentrin2<br />

pAb (1: 2500) (Zymed). The SUMO-1 and SUMNO-2/3 antibodies were both<br />

used at 1: 300 dilutions for visualisations in immun<strong>of</strong>luorescence.<br />

p300 and CBP were detected by anti-p300 mAb (Pharmingen) and anti-<br />

CBP pAb (Santa Cruz) respecfuly.<br />

Sheep anti-mouse (Amersham), donkey anti-sheep (Jackson<br />

ImmunoResearch) and donkey anti-rabbit (Amersham) horseradish-peroxidase<br />

conjugated IgGs were used to detect the primary antibodies at 1: 2500 dilution in<br />

Western-blot experiments. The secondary Immun<strong>of</strong>lurescence antibodies; goat<br />

anti-mouse FITC conjugate, and goat anti-rabbit Texas Red conjugate (Oxford<br />

Biotechnology Ltd. ) were used at a concentration <strong>of</strong><br />

2.3. Bacterial <strong>St</strong>rains<br />

1: 250.<br />

E. coli DH5a (genotype: ý80dlacZAM15, rec Al, end Al, gyr A96, thi-<br />

1, hsd R17 (rk-, mk+), sup E44, rel Al, deoR, D(lacZYA-argF) U169) was used<br />

for routine DNA preparation. E. coli B834 (F-, ompT, hsdSB, (rB-, m8-) dcm,<br />

gal) was used for all protein expression unless otherwise stated.<br />

98


2.4. DNA preparation<br />

The DNA preparations (maxi-preps, mini-preps, and gel extractions) used<br />

for cloning and transfections were prepared with Qiagen`"' kits, in accordance<br />

with the manufactures instructions. DNA restriction enzymes were obtained<br />

from both Promega ` "' and New England Biolabs' ' (NEB). The Vent polymerase<br />

used in polymerase chain reactions (PCR) was obtained from NEB. A<br />

Boeringher "TitanTm one tube RT-PCR System" was used for reverse<br />

transcription followed by PCR amplification (RT-PCR). The quality and<br />

quantity <strong>of</strong> DNA were analysed by spectrophotometric readings at 260 nm and<br />

280 nm and by electrophoresis in an agarose gel in the presence <strong>of</strong> ethidium<br />

bromide, followed by U. V. measurement.,<br />

2.4.1. cDNA cloning<br />

2.4.1.1. Preparation <strong>of</strong> thermocompetent bacteria<br />

A single colony <strong>of</strong> the appropriate bacterial strains were used to inoculate<br />

5 mis <strong>of</strong> L-broth medium overnight, which was subsequently used to inoculate 5<br />

L <strong>of</strong> L-broth medium. The bacteria cultures were allowed to grow at room<br />

temperature until the optical density at 600 nm were in the 0.5-0.8 region. The<br />

cultures were subsequently divided into pre-chilled 50 ml sterile centrifuge tubes<br />

and kept on ice for thirty minutes. The cultures were centrifuged at 3600 rpm,<br />

for five minutes at 4°C, the supernatants were removed and 20 mls <strong>of</strong> both pre-<br />

chilled 100 mM CaC12 and 40 mM MgSO4 were used to resuspend the bacterial<br />

pellet, and subsequently left to incubate on ice for a further thirty minutes.<br />

Following the incubation, the bacterial solution was re-centrifuged at 3600 rpm<br />

for five minutes at 4°C. The supernatant was subsequently removed and the<br />

pellet resuspended in 2.5 mis <strong>of</strong> CaCl2 and 2.5 mis <strong>of</strong> MgSO4. Glycerol was<br />

added to 10%, the cells were aliquoted as desired, snap-frozen and stored at<br />

-70°C.<br />

99


2.4.1.2. Transformation <strong>of</strong> thermocompetent bacteria<br />

Approximately 10 ng <strong>of</strong> DNA was added to 100 µl <strong>of</strong> bacteria, and left on ice for<br />

thirty minutes, following which the bacteria and DNA mixture was heat-shocked<br />

by being left at 42°C for between two and four minutes, before returned to ice for<br />

a further twenty minutes. Subsequently 1 ml <strong>of</strong> LB was added and the mixture<br />

was incubated at 37°C for between thirty and sixty minutes. Following this<br />

incubation the mixture was centrifuged for one minute at 13000 rpm, the pellet<br />

was subsequently resuspended in 100 µl <strong>of</strong> the supernatant and plated out onto<br />

L-agar plates, containing the appropriate selection antibiotic, and left at 37°C for<br />

twelve hours.<br />

2.4.1.3 Generated plasmid constructs and siRNA oligonucleotides<br />

siRNA<br />

Oigonucleotide sequences were designed against 21 nucleotide<br />

(nt) long<br />

sequences that terminated in double thymidine nucleotide (nt). Subsequently the<br />

21-nt was blasted to ensure the 21-nt was unique to the gene <strong>of</strong> interest.<br />

Synthetic oligonucleotides take advantage <strong>of</strong> a naturally occurring RNA<br />

interference pathway. The siRNA oligonucleotide mimics the naturally forming<br />

siRNA duplex created by the double stranded RNA cleavege mediated through<br />

the Rnase III family member, Dicer. The siRNA oligonucleotides are<br />

subsequently incorporated into the RNA-inducing silencing complex (RISC).<br />

Following' the unwinding <strong>of</strong> the <strong>of</strong> the siRNA duplex in an ATP dependent<br />

manner, the single-stranded antisense strand guides RISC to messenger RNA that<br />

has a complementary sequence which results in the endonucleolytic cleavage <strong>of</strong>f<br />

the target mRNA.<br />

An alternative to the synthetic oligonucleotides, is plasmid based siRNA<br />

vectors. These vectors are considerably cheaper than the synthetic<br />

oligonucleotides, and can be efficiently co-transfected along with other plasmids.<br />

Mammalian expression vectors have been generated that direct the synthesis <strong>of</strong><br />

siRNA-like transcripts (pSUPER: supression <strong>of</strong> endogenous RNA). PSUPER<br />

possess a polymerase-Ill H1-RNA gene promoter, and it produces a small RNA<br />

transcript lacking a polyadenosine tail and a well defined start <strong>of</strong> transcription<br />

100


and termination signal consisting <strong>of</strong> five thymidines in a row. Importantly, the<br />

cleavage <strong>of</strong> the transcript at the termination site is after the second uridine<br />

yielding a transcript resembling the end <strong>of</strong> the synthetic siRNAs. The pSUPER,<br />

and other shRNA vectors, contains a 19-nt sequence derived from the target<br />

transcript, separated by a short spacer from the reverse complement <strong>of</strong> the same<br />

19-nt sequence. The resulting RNA hairpin can be cleaved by Dicer into siRNA,<br />

and subsequently mediated mRNA destruction.<br />

2.4.1.3.1 Cullin-2<br />

The full length open reading frame <strong>of</strong> human Cullin-2 cDNA was obtained via<br />

RT-PCR from HeLa cell mRNA (kind gift from Dr. <strong>David</strong> C. Hay), with the<br />

primers Cul-2F (5'-CGAAGGATCCATGTCTTTGAAACCAAGAGTAGTAG-<br />

3') and Cul-2R (5'-TCAGGAATTCACGCGACGTAGCTGTATTC-3').<br />

2.4.1.3.2. Synthetic siRNA oligonucleotides targeting the SUMO is<strong>of</strong>orms<br />

Synthetic siRNA oligonucleotides were created against the following target<br />

sequences; SUMO-1, AACUGGGAAUGGAGGAAGAAG; SUMO-2,<br />

AAGAUCAAGAGGCACACGUCG; SUMO-3,<br />

AACAGACACACCUGCACAGUU.<br />

2.4.1.3.3. pSUPER retro siRNA expression plasmids<br />

The siRNA expression plasmids were designed as described previously<br />

(Brummelkamp et al, 2002), were inserted into the pSUPER retro plasmid<br />

(oligoengine). The siRNA oligonucleotides created were (upper strand only):<br />

SUMO-1,<br />

GATCCCCCTGGGAATGGAGGAAGAAGTTCAAGAGACTTCTTCCTCCA<br />

TTCCCAGTITFI GGAAA;<br />

SUMO-2,<br />

GATCCCCGATCAAGAGGCACACGTCGTTCAAGAGACGACGTGTGCCT<br />

CTTGATCTTTTTGGAAA;<br />

SUMO-3,<br />

GATCCCCCGACAGGGATTGTCAATGATTCAAGAGATCATTGACAATC<br />

CCTGTCGT =GGAAA.<br />

101


Expression plasmids for HDAC1, HDAC4, and HDAC6 were kind gifts<br />

from T. Kouzarides (Cambridge, UK). SSP3/SENP2, wild type and mutant<br />

DNA were generous gifts from Barbara Ink (GalaxoWellcome), Heidi Mendoza,<br />

and Owen A. Vaughan (<strong>University</strong> <strong>of</strong> <strong>St</strong>. <strong>Andrews</strong>). The dominat negative Ubc9<br />

mutant plasmid was kindly provided by Joanna M. P. Desterro. The expression<br />

plasmids for HA-SUMO-1, HA-SUMO-2, HA-SUMO-3, GST-SUMO-1 GG,<br />

GST-SUMO-2 GG, and GST-SUMO-3 GG were provided by Michael H.<br />

Tatham (<strong>University</strong> <strong>of</strong> <strong>St</strong>. <strong>Andrews</strong>). Purified SAE1/2, Ubc9, RanBP2, and all<br />

PIAS proteins were generously provided by Ellis Jaffray. The expression<br />

plasmids for His-p300 CRD+, His-p300 CRD-, Ga14 N-, Ga14 N+, and Ga14<br />

(Figure 16a) were provided as part <strong>of</strong> collaboration by Neil D. Perkins<br />

(<strong>University</strong> <strong>of</strong> Dundee). The Ga14 fusions both wild-type and sumoylation<br />

deficient for Sp3 and Elk-1 (Figure 16a) were provided by Grace Gill (Harvard<br />

<strong>University</strong>) and Andrew Sharrocks (Manchester <strong>University</strong>) respectively.<br />

2.4.2. DNA sequencing<br />

All constructs were verified by automated DNA sequencing on an ABI<br />

PRISMT' 377 DNA sequencer (<strong>St</strong>. <strong>Andrews</strong> <strong>University</strong> DNA sequencing unit).<br />

2.5. Expression and Purification <strong>of</strong> unlabeled Recombinant Proteins<br />

GST-SUMO-1-GG, GST-SUMO-2-GG, GST-SUMO-3-GG, GST-<br />

SAE1/2, and all p300 CRD1 domain mutants were expressed from E. coli strain<br />

B834 as described previously (Jaffray et al, 1995). All GST-SUMO proteins<br />

were cleaved by thrombin while bound to glutathione-agarose beads. For freeze-<br />

drying, proteins were dialysed against 20 mM ammonium bicarbonate, 1 mm<br />

dithiothrietol (DTT), before concentration calculations. Lyophilised protein<br />

samples were taken up in 50 mM Tris/HCL pH - 7.5,1 mM DTT to a<br />

concentration <strong>of</strong> 10 mg. ml-'.<br />

The CRD 1 domain proteins were purified on glutathione-agarose beads<br />

(GGA) equilibrated with PBS + 0.5M NaCl. The sonicated supernatant was<br />

filter sterilised and mixed with the GGA beads for 10-20 minutes under agitation<br />

102


p300 N+<br />

p300 N-<br />

107IKDE10 538IKEE-"<br />

Sp3 wt AD AD ID<br />

Sp3 mut<br />

Elk-1 wt<br />

Elk-1<br />

Ga14 AdMLP reporter<br />

229LKSE23'<br />

LRSE232<br />

Ga14 bs AdMLP Luciferase<br />

Figure 16a. Schematic representation <strong>of</strong> the Ga14 fusion proteins, as indicated, and Ga14<br />

AdMLP reporter construct. Plasmids were originally described in <strong>Girdwood</strong> et al., 2003<br />

(p300); Ross et al., 2002 (Sp3), and Yang et al., 2003 (Elk-1).<br />

1004-1044<br />

.. LKTEIKEE..


on a roller. The bead/supernatant mix was added to a plastic column and the<br />

flow-through was re-circulated three times. The column was washed with four<br />

column volumes <strong>of</strong> PBS + 0.5M NaCl, one volume at a time. The GST fusion<br />

protein was eluted by the addition <strong>of</strong> glutathione buffer, and 0.5 column volume<br />

fractions were collected. The fractions were analysed by running on 10% SDS-<br />

PAGE, and the fractions containing protein were pooled and dialysed overnight<br />

at 4°C to remove glutathione. The proteins were subsequently concentrated<br />

through use <strong>of</strong> a centrifugal device (MicroconTM)<br />

2.6. Quantitation <strong>of</strong> protein<br />

Protein concentrations were " determined using Bradford's method<br />

(Bradford, 1976). Protein samples were mixed with Bradford's (BioradTM) and<br />

the absorbance at 595 nm was measured on a spectrophotometer (Hitachi U-<br />

1100). Protein absorbencies were converted to mg/ml concentrations using a<br />

standard curve constructed by measuring the absorbencies <strong>of</strong> a range <strong>of</strong> bovine<br />

serum albumin (BSA) concentrations.<br />

2.7. SDS/PAGE and Western Blot analysis<br />

Protein samples were resuspended in disruption buffer (1X: 20mM<br />

Tris/HCL pH 6.8,2% SDS, 5% f -mercaptoethanol, glycerol and 2.5%<br />

bromophenol blue) and denatured at 100°C for five minutes, before loading on<br />

6%-15% SDS-polyacrylamide gel (acrylamide percentage appropriate for the<br />

size <strong>of</strong> proteins to be separated), or following lysis SDS sample buffer, samples<br />

were diluted 1: 3 in RIPA buffer (25mM Tris-Hcl pH8.2,50mM NaCl, 0.5% NP-<br />

40,0.5% Deoxycholate, 0.1% Azide) containing protease inhibitor tablets. Bio-<br />

Rad' mini-gel equipment was used in accordance with the manufactures<br />

instructions. New England Biolabs' prestained molecular weight markers were<br />

used as standards to establish the apparent molecular weights <strong>of</strong> proteins<br />

resolved on SDS-polyacrylamide gels. Separated polypeptides were either<br />

stained with Coomassie Brillant Blue (0.2% Coomassie brilliant blue R250; 50%<br />

methanol; 10% acetic acid) for thirty minutes and then destained (20% methanol;<br />

10% acetic acid) or transferred to a polyvinylidene diflouride mebrane (PVDF)<br />

(Sigma) using a wet blotter (Bio-Rad Systems). The membranes were blocked<br />

104


with PBS containing 5% skimmed milk powder (Tesco) and 0.1% Tween 20, and<br />

subsequently incubated with monoclonal or polyclonal antibodies diluted in<br />

blocking buffer. Horseradish peroxidase conjugated anti-mouse IgG, anti-rabbit<br />

IgG (Amersham) and anti-sheep IgG were used as' secondary antibodies.<br />

Western blotting was performed using ECL detection system.<br />

2.8. <strong>St</strong>ripping <strong>of</strong> P. V. D. F. membranes<br />

After ECL, the membranes were washed in PBS for 20 minutes before<br />

being transferred into stripping buffer (142 µl ß-mercaptoethanol, 2 ml 20%<br />

SDS, 1.25 ml Tris pH 6.7 and 16.6 ml dH2O) and incubated for 30 minutes in a<br />

hybridisation oven at 72°C. The membranes were subsequently washed twice in<br />

250 ml PBS containing 0.1% Tween 20. The membranes were blocked for 1<br />

hour in PBS containing 5% skimmed milk powder and 0.1% Tween 20 and then<br />

probed as in the Western blot procedure.<br />

2.9. In vitro Expression <strong>of</strong> Proteins<br />

In vitro transcription/translation <strong>of</strong> proteins was performed using 1 µg <strong>of</strong><br />

plasmid DNA and a wheat germ-coupled transcription/translation system<br />

according to the manufacturer's instructions (Promega). 35S-methionine<br />

(Amersham Pharmacia Biotech) was used in the reactions to generate<br />

radiolabelled protein.<br />

2.10. In vitro SUMO conjugation assay<br />

SUMO conjugation assays were performed in 10 µl volumes containing<br />

0.84 µ1 <strong>of</strong> 1251-labeled<br />

SUMO-1 or 1 µg <strong>of</strong> SUMO-1, SUMO-2, SUMO-3, GST-<br />

SUMO-1, GST-SUMO-2, and GST-SUMO-3 (as indicated), an ATP<br />

regenerating system and buffer (50 mM Tris pH 7.5,5 MM MgC12,2 mM ATP,<br />

10 mM creatine phosphate, 3.5 U. ml'' creatine kinase) and 0.6 U. ml"' IPP with<br />

either 1µ1 <strong>of</strong> 35S-methionine labelled substrate (5'-p300) or varying<br />

concentrations <strong>of</strong> unlabeled GST-p300 fragments (as indicated). Assays<br />

contained 120 ng (1.1 pmoles) purified recombinant SAE1/SAE2 and 650 ng<br />

(35.9 pmoles) Ubc9, unless otherwise shown. Reactions were incubated at 37°C<br />

for 4 h. Reactions were stopped following the addition <strong>of</strong> SDS sample buffer,<br />

105


eaction products were fractionated by electrophoresis in polyacrylamide gels (6-<br />

10%) containing SDS, stained, destained and dried, before analysis by<br />

phosphorimaging (Fujix BAS 1500, MacBAS s<strong>of</strong>tware).<br />

2.11. In vitro SUMO Deconjugation Assay<br />

Recombinant GST-PML substrate was conjugated with SUMO as<br />

described previously. Conjugation was terminated by the addition <strong>of</strong><br />

iodoacetamide to 10mM and incubated at 20°C for 30 minutes. lodoacetamide<br />

was quenched by the addition <strong>of</strong> ß-mercaptoethanol to 15mM and incubated at<br />

20°C for a further 15 minutes.<br />

Deconjugation <strong>of</strong> SUMO-GST-PML was performed in a total volume <strong>of</strong> 10µl;<br />

containing 2µg <strong>of</strong> GST-PML and the indicated amounts <strong>of</strong> SSP3 (provided by H.<br />

Mendoza)(0-243ng) in deconjugation buffer (50mM Tris pH7.5,2mM MgC12,<br />

and 5mM ß-mercaptoethanol. Reactions were incubated at 37°C for 3 hours,<br />

before being terminated with SDS sample buffer, reaction products were<br />

fractionated by electrophoresis in polyacrylamide gels (10%) containing SDS,<br />

stained, and destained.<br />

4F<br />

2.12. Protein Interaction Assays<br />

GST, unmodified GST-p300-CRD1 and GST-p30085210719<br />

or SUMO-<br />

modified GST-p300-CRD1 and GST-p300852_1071 (Figure 16b) were purified from<br />

reaction components on glutathione agarose. HDAC4 and HDAC6 were 15S_<br />

labeled by in vitro translation in a wheat germ extract and diluted 10-fold in 150<br />

mM NaCl, 50 mM Tris (pH 7.5), 1 mM DTT, and 0.1% NP40 before<br />

clarification by centrifugation. The soluble fraction (100 µl) was added to the<br />

loaded glutathione agarose beads (10 µl packed volume) and incubated for 2 hr<br />

at 37°C. Beads were collected by centrifugation, washed extensively with the<br />

above buffer and bound proteins fractionated by SDS PAGE, and analysed by<br />

phosphorimaging.<br />

106


(1)<br />

C/H1 CRDI Bromo C/H2 C/H3<br />

PMO wt III on IIIII<br />

p300 i-1255<br />

p300 ss2-ion<br />

p3001017-1029<br />

p300 CRD1<br />

p300 192-1044<br />

(ii)<br />

p300 Aloo4-1o44<br />

RanBP2 2532-2767<br />

1 1255<br />

0-91<br />

852 CRD 1 1071<br />

1017 029<br />

192<br />

1044<br />

2532 2633 2685 2710 2761 2767<br />

Figure 16b. Fusion constructs. (i) Schematic representation <strong>of</strong> p300 fusion constructs,<br />

as described in material and methods. (ii) Schematic representation <strong>of</strong> a fragment<br />

from the SUMO E3, RanBP2, that possess the E3 ligase activity.


2.13. Mammalian cell culture transient transfections<br />

HeLa His6-SUMO-1 stable cell line was grown in DMEM (Gibco)<br />

supplemented with 10% FCS and puromycin (2µg/ml). HeLa His6-SUMO-2 and<br />

HeLa His6-SUMO-3 stable cell lines were grown in DMEM supplemented with<br />

10% FCS and puromycin (0.5µg/ml). COST and HeLa cells were maintained in<br />

exponential growth in Dulbecco's modified Eagle's medium, containing 10%<br />

foetal bovine serum. For transfections, a total <strong>of</strong> 12µg <strong>of</strong> plasmid DNAs were<br />

transfected for 10 hours in subconfluent 75cm3 flasks using Lip<strong>of</strong>ectamineTM<br />

according to manufacturers instructions (Invitrogen). After 16 hours <strong>of</strong><br />

expression, cells were washed in PBS and cellular extracts were either added to<br />

1X disruption buffer or 6M guanidinium-HCL.<br />

2.13.1. Calcium phosphate transient transfections and reporter gene assays.<br />

Cells growing in log phase were plated into individual wells <strong>of</strong> a six well<br />

plate at 40% confluency, 1 to 2h prior to transfection to allow the cells to<br />

adhere. All transfection solutions were equilibrated to room temperature. siRNA<br />

encoding plasmids (1µg) were co-transfected into U2-OS cells along with Ga14-<br />

p300, Gal4-Sp3, and Gal4-Elk-1 expression plasmids (0.5 ng) and 2 µg <strong>of</strong> Ga14<br />

AdML luciferase reporter plasmids. Synthetic siRNA oligos were similarly<br />

transfected at 100 nM each. The reporter plasmid DNAs along with the<br />

pSUPER retro plasmid DNA or synthetic siRNA oligos was diluted into 438 µl<br />

<strong>of</strong> water containing 61 pl <strong>of</strong> 2M CaC12<br />

and added in drops with gentle agitation<br />

to 500 gl <strong>of</strong> 2X HEPES-buffered saline (0.274 M NaCl, 1.5 mM Na2HPO4,<br />

54.6 mM HEPES [pH 7.1]). The DNA, CaC12, or synthetic siRNA oligo, and<br />

HEPES-buffered saline were mixed by pipetting and, split three ways, then<br />

immediately sprinkled onto the cells, in three wells <strong>of</strong> a six well plate, ensuring<br />

that all the cells were covered. Following a 24-h incubation, the precipitate was<br />

removed and fresh complete medium was added to the cells, and the cells left for<br />

a further 24-h, before being subjected to SDS-page and western blotting or<br />

harvested and luciferase activity determined as appropriate.<br />

108


2.14. Purification <strong>of</strong> His6-tagged proteins<br />

After twenty-six hours the transfected cells were washed with PBS and<br />

resuspended in 10 ml <strong>of</strong> PBS. 1 ml was removed, pelleted and lysed in SDS<br />

sample buffer (5% SDS, 0.15 M Tris-HCI pH 6.7,30% glycerol, 0.72 M ß-<br />

mercaptoethanol) diluted 1: 3. The remaining 9 ml was pelleted and lysed in 5 ml<br />

6M guanidinium-HCI, 0.1 M Na2HPO4/NaH2PO4,0.01 M Tris-HC1 pH 8.0 plus<br />

5 mM imidazole and 10mM ß-mercaptoethanol. per 75 cm3 flask. Endogenous<br />

material were directly lysed in 5m1 6M guanidinium-HCI, 0.1 M<br />

Na2HPO4/NaHZPO4,0.01 M Tris-HC1 pH 8.0 plus 5 mM imidazole and 10mM ß-<br />

mercaptoethanol. per 75 cm3 flask. Both transfected and endogenous cell lysates<br />

were sonicated, to reduce viscosity, the lysates were mixed with 50µ1 <strong>of</strong><br />

Nie+NTA-agarose beads prewashed with lysis buffer and incubated for 3h at<br />

room temperature. The beads were subsequently washed twice with 6M<br />

guanidinium-HCI, 0.1 M Na2HPO4/NaHZPO4,0.01 M Tris-HC1 pH 8.0,10mM f-<br />

mercaptoethanol; three times with 8M urea, 0.1 M Na2HPO4/NaH2PO4,0.01 M<br />

Tris-HC1 pH 8,10 mM ß-mercaptoethanol; twice with 8M urea, 0.1<br />

Na2HPO4/NaHZPO4,0.01 M Tris-HC1 pH 6.3,10 mM ß-mercaptoethanol (buffer<br />

A); buffer A plus 0.2% Triton X-100; buffer A; buffer A plus 0.1% Triton X-<br />

100, and then three times with buffer A. The beads were finally eluted with 200<br />

mM imidazole in 5% SDS, 0.15 M Tris-HC1 pH 6.7,30% glycerol, 0.72 M ß-<br />

mercaptoethanol. The eluates were subjected to SDS-PAGE (6-10%) and the<br />

proteins transferred to a PVDF membrane (Sigma). Western blotting was<br />

performed with either pAb against p300 (Santa Cruz) or mAb specific to the HA-<br />

epitope (Babco).<br />

2.15. Indirect immunotluorescence analysis <strong>of</strong> cells<br />

Cells were cultured on 13 mm glass coverslips (Scientific Laboratory<br />

Supplies Ltd. ) in 6 well plates and subjected to transient transfections as<br />

necessary. On completion <strong>of</strong> transfection, cells were washed twice in PBS<br />

containing 1 mM MgCl2 and 1 mM CaCl2. Cells were subsequently fixed with<br />

warm 3% Paraformaldehyde/PBS for 10 minutes, washed in PBS three times,<br />

and fixation quenched by two 10 minute 0.1 M glycine/PBS incubations. After a<br />

5 minute PBS wash, cells were permeabilised for 10 minutes in 0.1% Triton-X-<br />

109<br />

4


100/PBS. Following three further 5 minute PBS washes, the cells were blocked<br />

in 0.2%/PBS BSA for 10 minutes. Immunostaining with the relevant<br />

concentration <strong>of</strong> primary antibodies was performed for 2 hours at room<br />

temperature, in 0.2% BSA/PBS before being subjected to three 5 minute 0.2%<br />

BSA/PBS washes. The cells were then incubated for a further 30 minutes with<br />

the corresponding fluorescently labelled secondary antibodies diluted in 0.2%<br />

BSA/PBS at 1: 250 dilution. The cells were given a final three 5 minute PBS<br />

washes before the coverslips were gently mounted onto slides using Movial<br />

(Calbiochem) and stored overnight at 4oC. Fluorescence microscopy was carried<br />

out with a DeltaVision microscope (Olympus IX70) with a 1.40 NA 100X<br />

objective and Photometric CH300 CCD camera. 10X 0.2 µm sections were<br />

taken. Images were deconvoluted using S<strong>of</strong>tWorx (Applied Precision) s<strong>of</strong>tware.<br />

2.16. Luciferase Assay<br />

Transfections to be assayed for luciferase activity were stopped by<br />

washing cells twice with PBS, and then subsequently lysed in 200 µl <strong>of</strong><br />

luciferase lysis buffer ( 25 mM Tris-phosphate pH 7.8,8 MM MgC12,1 mM<br />

DTT, 1% Triton X-100 and 15% glycerol; made up in dH2O). 50 µl <strong>of</strong> the lysate<br />

- was injected with luciferase assay buffer (0.25mM luciferin, 1mM ATP, and 1%<br />

BSA diluted in the lysis buffer) into a luminometer (Berthold 'Sirius<br />

Luminometer) and the luciferase activity was recorded. The activity was stated<br />

in relative light units (R. L. U. ).<br />

110


3. RESULTS & DISCUSSIONS<br />

111


3.1 Bioinformatics: an approach to identify novel members <strong>of</strong> the<br />

ubiquitin-like protein superfamily<br />

Previous bioinformatics within the group had focused primarily on<br />

database searching, <strong>of</strong> submitted proteins, to identify those proteins containing<br />

the reported SUMO modification motif. This approached allowed the generation<br />

<strong>of</strong> a list <strong>of</strong> putative SUMO modified substrates. The completion <strong>of</strong> many<br />

genomes potentially allowed the generation <strong>of</strong> a comprehensive list <strong>of</strong> potential<br />

SUMO substrates. The draft human genome sequence, as accessible through the<br />

Sanger Institute (httl2: //www. satiger. ac. uk/HGP/) enables us to search for similar<br />

nucleotide sequences or "blasting" (Basic Local Alignment Search Tool) <strong>of</strong><br />

proteins <strong>of</strong> interest against the draft human genome sequence. Thus to utilise the<br />

completion <strong>of</strong> the draft version <strong>of</strong> the human genome, the protein sequences <strong>of</strong><br />

all known ubiquitin-like proteins were "blasted" against the draft sequence in an<br />

attempt to identify novel members <strong>of</strong> the ubiquitin-like protein modifier<br />

superfamily.<br />

3.1.1 Identification <strong>of</strong> three possible new members <strong>of</strong> the SUMO<br />

family<br />

The database searching <strong>of</strong> the draft sequence <strong>of</strong> the human genome<br />

revealed the presence <strong>of</strong> an additional three members <strong>of</strong> the SUMO family,<br />

located on chromosomes six, seven, and twenty (Figure 17). The sequences for<br />

all three novel members had previously been submitted to GenBank<br />

(Accession numbers XP_066029, CAA20019, and XP_168354 respectively).<br />

These genes were annotated by the Ensembl automatic analysis pipeline using<br />

either a GeneWise model from a human/vertebrate protein, a set <strong>of</strong> aligned<br />

human cDNAs followed<br />

. by GenomeWise for open reading frame (ORF)<br />

prediction or from Genscan exons supported by protein, cDNA and EST<br />

evidence. GeneWise models are further combined with available aligned cDNAs<br />

to annotate UTRs. All three annotated genes possesssed ORFs and contained a<br />

112


di-glycine motif towards the end <strong>of</strong> their C-terminal region. Of these three<br />

potential new genes, one had high homology to SUMO-1, that on chromosome<br />

20, termed SUMO-lb, and two possesssed high homology to SUMO-2/-3, those<br />

on chromosomes six and seven, termed SUMO-4 and SUMO respectively.<br />

Interestingly, all three contained an internal SUMO modification motif (Figure<br />

16a), the new genes however, possesssed no introns unlike the previous three<br />

members.<br />

3.1.2 SUMO-lb/4/5 are conjugated both in vitro and in vivo<br />

To clone the genes, mRNA was extracted from HeLa cells (mRNA<br />

Isolation system, Promega), which included a DNase treatment step to digest any<br />

contaminating genomic DNA. Subsequently RT-PCR was preformed to amplify<br />

the genes from the mRNA (Figure. 18. ). The' DNA products were then cloned<br />

into HA-pcDNA3 and pGEX-4T vectors, and the correct sequences were verified<br />

by DNA sequencing.<br />

All members <strong>of</strong> the SUMO family are capable <strong>of</strong> being covalently bound<br />

to substrates via lysine residues, following their processing by the proteases, and<br />

conjugation through the enzymatic cascade involving SAE1/2, Ubc9, SUMO<br />

ligase. Due to the sequence similarities <strong>of</strong> the novel members with SUMO-1/-2/-<br />

3, it was likely that these novel members <strong>of</strong> the SUMO gene family would be<br />

functionally similar. To test this hypothesis HA-SUMOs la/lb/2/3/4/5 and<br />

empty pcDNA3 vectors were transfected into COST cells. Following<br />

transfection cells were lysed in RIPA buffer and subjected to Western blot<br />

analysis with anti-HA antibody (Figure. 18. ). All three potential new members <strong>of</strong><br />

the SUMO family were shown to be capable <strong>of</strong> covalently modifying a range <strong>of</strong><br />

cellular substrates, as indicated by the analogous pattern <strong>of</strong> detection. Western<br />

blot analysis indicated that the levels <strong>of</strong> SUMO-lb were greatly elevated,<br />

compared to the other SUMO is<strong>of</strong>orms, there being approximately 100 times<br />

more SUMO-lb conjugated than SUMO-la (Figure 18). The differences in<br />

levels <strong>of</strong> conjugation <strong>of</strong> SUMO-lb, suggest that SUMO-lb has greater stability<br />

than the other SUMO is<strong>of</strong>orms, or is more resistant to removal via the actions <strong>of</strong><br />

the SUMO specific proteases.<br />

113


Figure 17(i). Diagrammatic representation <strong>of</strong> the human chromosomes. The<br />

chromosomal locations <strong>of</strong> both the SUMO gene members, and potential new SUMO gene<br />

members are shown (indicated by red arrow). Locations were identified through the blast<br />

s<strong>of</strong>tware service option <strong>of</strong> the human genome sequencing project, through the sanger<br />

centre (www. sanger. ac. uk/HGP). (ii) Table providing GenBank accession numbers,<br />

chromosome location, number <strong>of</strong> exons, and number <strong>of</strong> pseudogenes for the SUMO gene<br />

family. (iii) Sequence alignment <strong>of</strong> SUMO-1/-2/-3 with the predicted sequences for the<br />

potential new members <strong>of</strong> the SUMO family.


(I)<br />

(ii)<br />

123456789 10 ii 12<br />

13 14 15 16 17 18 19 20 21 22 XY<br />

Genes GenBank Accession number Chromosome Exons Pseudogenes<br />

SUMO-1 CAA67898 2 5 8<br />

SUMO-2 CAA67896 21 4 0<br />

SUMO-3 P55855 17 4 23<br />

SUMO-Ih XP_066029 20 1<br />

SUMO-4 CAA2(X)19 6 1<br />

SUMO-5 XP 168354 7 1<br />

(iii)<br />

Sabo-2<br />

SW®0-3<br />

SUND-5<br />

SUMO-4<br />

SUM-2<br />

SWF-3<br />

SUMO-5<br />

SUMO-4<br />

SU-1 s T$DLGDKPMGZY<br />

SUND-lb ruimZD sn lJAII'<br />

IB<br />

ýA'! ý<br />

SIDE)-1 sIýý<br />

sý------sasa-ý<br />

s xa ,4 VW z srv-------<br />

-<br />

-


(t An<br />

z<br />

(iii q rý<br />

x<br />

175<br />

83<br />

62<br />

m<br />

OO o o<br />

vý ýn<br />

ý' x xx x x<br />

a t t ý ..<br />

HA-SUMO<br />

conjugates<br />

Figure 18. (i) RT-PCR<br />

was performed on HeLa<br />

SUMO DNA poly(A)+ RNA.<br />

from RT-PCR Amplification<br />

reaction. (ii) Conjugation <strong>of</strong><br />

proteins in vivo. COS7 <strong>of</strong> transfected SUMO<br />

cells were transfected<br />

la/lb/2/3/4/5-GG.<br />

with either<br />

Total pcDNA3 or HA-SUMOcell<br />

lysates<br />

were analyzed by<br />

anti-HA Western<br />

mAb. Molecular blotting<br />

mass standards using<br />

are expressed in kilodaltons.<br />

6


To investigate the mechanisms <strong>of</strong> their conjugation, a purified GST-PML<br />

construct (Tatham et al., 2001) or 35S-Met-labelled Sp100, generated by in vitro<br />

transcription/translation, was incubated with recombinant Ubc9, SAE1/2, and<br />

" either SUMO-la, SUMO-lb, SUMO-2, SUMO-3, GST-SUMO-4, or SUMO-5<br />

(as described in material and methods). The incubation <strong>of</strong> the substrates, with<br />

the components <strong>of</strong> the conjugation pathway, resulted in the appearance <strong>of</strong> slower<br />

migrating species, with all <strong>of</strong> the SUMO forms tested (Fig. 18i. ). The specificity<br />

<strong>of</strong> the reactions was demonstrated by the absence <strong>of</strong> any other forms <strong>of</strong> GST-<br />

PML or 35S-SplOO when incubated in the absence <strong>of</strong> purified SUMO, Ubc9, or<br />

SAE1/2, furthermore substitution <strong>of</strong> SUMO for a tagged GST-SUMO further<br />

reduces the mobility <strong>of</strong> the substrate, demonstrating that the occurrence <strong>of</strong> the<br />

slower migrating forms observed are due to the conjugation <strong>of</strong> SUMO to the<br />

substrates. In the reactions lacking substrate, the formation <strong>of</strong> poly(SUMO)<br />

chains was observed, indicating that the all three potential new members,<br />

functioned in a manner analogous to that <strong>of</strong> SUMO-2/-3 are were capable <strong>of</strong><br />

forming the SUMO chains (Fig. 19. ii. ).<br />

3.1.3 Deconjugation <strong>of</strong> SUMO proteins via a SUMO specific<br />

protease<br />

To test whether the new SUMO family members were deconjugated in a<br />

similar manner to those already identified, deconjugation assays were set up.<br />

SUMO modification <strong>of</strong> substrates is controlled on many levels, including their<br />

deconjugation from their substratcWsing the GST-PML substrate that had<br />

previously been demonstrated to be modified by all SUMO species, the<br />

deconjugating ability <strong>of</strong> the SUMO specific protease, available within the group,<br />

SSP3/SENP2 was investigated. Recombinant GST-PML was independently<br />

modified by both SUMO-lb and SUMO-5 in the presence <strong>of</strong> recombinant<br />

SAE1/2 and Ubc9; and subsequently conjugation was terminated by the addition<br />

<strong>of</strong> iodoacetamide. After quenching <strong>of</strong> the iodoacetamide with ß-<br />

mercaptoethanol the reaction products were used as substrates for a range <strong>of</strong><br />

SSP3/SENP2 concentrations with a three hour incubation at 37°C (Figure 20)(as<br />

117


(i)<br />

(ii)<br />

SUMO-1aSUMO-2 SUMO-lb SUMO-4 SUMO-5<br />

SUMO - ++ - ++ +- ++++- ++++ -<br />

GST-SUMO - -- --- ----- + +- ++ ++ "<br />

- +<br />

SAE1l2 + ++ +++ - ++- +++ -++- ++ ++ +<br />

-++-<br />

Ubc9 +++ +++ -+++-++ - ++ + -+<br />

++++ +<br />

GST-PML ++- ++- -++++-+ -+++ +- -++++-<br />

175<br />

83<br />

63<br />

47<br />

32<br />

25 r `" ý. ý" W- go or<br />

16<br />

SUMO-1a<br />

SUMO-lb<br />

SUMO -+-+++-<br />

GST-SUMO -+<br />

SAE1! 2<br />

Ubc9<br />

++++-++<br />

+++++-+<br />

175-<br />

83-<br />

62-<br />

!! '<br />

....<br />

w.. * w.,.<br />

. ý.<br />

-<br />

wow<br />

SUMO-4<br />

SUMO-5<br />

+<br />

sumo<br />

mº conjugates<br />

ow<br />

- - - - + + + -<br />

- + + + - - - - +<br />

+ + - + + + - + +<br />

+ + + - + + + - +<br />

a<br />

06<br />

SUMO<br />

conjugates<br />

-44<br />

+ SUMO<br />

Figure 19. SUMO-1b/4/5 are capable <strong>of</strong> conjugation to target substrates<br />

and to a similar extent. (i) GST-PML or (ii) 35S Sp1OO. In the absence <strong>of</strong><br />

substrate a reduction in monomeric SUMO levels were observed, and high<br />

weight molecular species were observed, consistent with the formation <strong>of</strong><br />

poly(SUMO) chains. Due to the presence <strong>of</strong> an internal thrombin site<br />

SUMO-4 was left as a GST fusion, GST-SUMO-4-GG. Reaction products<br />

were fractionated by SDS-PAGE before the gels were either (i) stained<br />

with coomassie brilliant blue or (ii) dried and analysed by<br />

phosphorimaging. The positions <strong>of</strong> conjugated SUMO are indicated.


+I SUMO- I blo.<br />

GST-PML º,,<br />

SUMO-lb<br />

+1 SUMO-5<br />

0'<br />

GST-PML ºr<br />

SUMO-5 º<br />

SSP3/SENP2<br />

039 27 81 243 (ng)<br />

.,.,,<br />

ý<br />

q... , new __ ý<br />

SSP3/SENP2<br />

039 27 81 243 (ng)<br />

Figure 20. SUMO modification <strong>of</strong> GST-PML is removed following the<br />

addition <strong>of</strong> a SUMO specific protease. Following a3h incubation in<br />

deconjugation buffer, with varying amounts <strong>of</strong> SSP3/SENP2 present (as<br />

indicated) at 37°C, the SUMO modified substrate was not seen. With no<br />

SSP3/SENP2 added or at very low concentrations, SUMO modification was<br />

left intact. Reaction products were fractionated by SDS-PAGE before the gels<br />

werevisualised via staining with coomassie<br />

brilliant blue.


detailed in materials and methods). As indicated by Figure 20, assays that<br />

contained no or very low concentrations <strong>of</strong> SSP3/SENP2 did not affect the<br />

SUMOylation status <strong>of</strong> the substrate. The assays containing the higher<br />

concentrations <strong>of</strong> resulted in the removal <strong>of</strong> all SUMO modified GST-PML<br />

forms and increase in the available pool <strong>of</strong> unconjugated "free" SUMO.<br />

3.1.4 SUMO-lb, SUMO-4, and SUMO-5 may be psuedogenes<br />

Data obtained from the human genome is analysed against a range <strong>of</strong><br />

critera, to check for validity, before being submitted to the accessible databases.<br />

To confirm that the newly identify genes are expressed and that the original RT-<br />

PCR constructs generated were amplified directly from mRNA and not<br />

contaminating DNA, RT-PCR reactions were repeated in duplicate, one set<br />

containing the reverse transcriptase DNA polymerase and the second set had the<br />

reverse transcriptase enzyme substituted by Taq polymerase. The resulting PCR<br />

products demonstrated that there was possible DNA contamination (Figure 21).<br />

Thus there was the possibility that the newly identified "novel" SUMO genes<br />

may in fact represent pseudogenes. Pseudogenes are not true genes, as although<br />

their DNA sequence is found within the genome, they are not transcribed, and<br />

there are many pseudogenes within the human genome.<br />

120


+ Reverse transcriptase +++---<br />

+ Taq polymerase ---+++<br />

SUMO lb 45 lb 45<br />

Figure 21. Contamination <strong>of</strong> mRNA preparation with genomic DNA. The<br />

RT-PCR for the amplification <strong>of</strong> the potential new SUMO genes was<br />

repeated under standard conditions with the reverse transcriptase enzyme<br />

included. To exclude artefacts from DNA targets like residual genomic DNA<br />

contaminations from RNA preparations, a control reaction was set up where<br />

the RT was not preformed. This was achieved by the addition <strong>of</strong> Taq DNA<br />

polymerase to the RT-PCR mixture instead <strong>of</strong> the normal enzyme mix.<br />

4-


3.1.5 Conclusion<br />

That the RNA used for the RT-PCR was potentially contaminated with<br />

DNA did not directly rule out that these potentially novel genes were expressed.<br />

However the fact that none <strong>of</strong> these potential new genes contained any introns<br />

unlike SUMO-l/-2/-3, was suggestive that they were in fact pseudogenes.<br />

Further evidence was provided by bioinformatics analysis that indicated that<br />

there was no similarity between the mouse and human promoter regions and that<br />

represented they were not in EST databases. (C. W. Crumbley, Glaxo Smithkline,<br />

personal communication). With the combination <strong>of</strong> evidence suggesting that the<br />

genes were not expressed, the decision was taken not to follow up this line <strong>of</strong><br />

study, although one possibility was to clone the gene from a cDNA library.<br />

Recently, however Owerback and colleagues reported the expression <strong>of</strong> SUMO-<br />

4, in the liver (Bohren et al., 2004). SUMO-4 is polymorphic, the predominant<br />

form having a methionine at amino acid position 55, whilst a second form exists<br />

with a valine residue in place <strong>of</strong> the methionine. This finding was similar to the<br />

findings from the potential SUMO-4 gene identified through the human genome,<br />

which was submitted as having a valine at amino acid 55, but following cloning<br />

and sequencing from HeLa mRNA, all clones contained a methionine at this<br />

residue. Subsequently it was confirmed that the reported SUMO-4 gene was the<br />

same gene (NCBI accession number: CAA20019) as the potential SUMO-4 gene<br />

identified by our bioinformatics search. Owerback and colleagues demonstrated<br />

that SUMO-4 is expressed by performing control RT-PCR reactions. The reverse<br />

transcriptase step <strong>of</strong> the RT-PCR is omitted, and no PCR product is generated,<br />

thereby showing that the positive reaction was not due to DNA contamination.<br />

As this would seem to indicate that SUMO-4 is a bona fide gene, and novel<br />

member <strong>of</strong> the SUMO family, the same may yet be true for SUMO-lb and<br />

SUMO-5. Interestingly the SUMO-4 gene that we identified through the human<br />

genome database searching was preformed in 2002, and in the intervening two<br />

years, the draft version <strong>of</strong> the human genome has undergone many revisions, and<br />

currently the potential SUMO-4 gene has been reannotated as being a<br />

pseudogene. As such there remains the possibility that the SUMO-4 sequence<br />

identified independently both by our group and by Bohren et al., may yet turn<br />

122


out to be a pseudogene. Conversely there remains the possibility that SUMO-lb<br />

and SUMO-5 may be expressed and as such novel members <strong>of</strong> the SUMO<br />

family.<br />

123<br />

t


3.2 SUMO modification <strong>of</strong> the transcriptional regulator p300<br />

p300 and CREB binding protein can both activate and repress<br />

transcription. Contained within the previously identified cell cycle regulatory<br />

domain (CRD1), a transcriptional repression domain <strong>of</strong> p300 that lies between<br />

residues 1017 to 1029, we identified two copies <strong>of</strong> the SUMO modification<br />

sequence, iKxE (Figure 22. i. ). Mutations within the conjugation motif, which<br />

reduced SUMO modification, resulted in an increase in p300-mediated<br />

transcriptional activity. Expression <strong>of</strong> a SUMO-specific protease or a<br />

catalytically inactive Ubc9 relieved repression, demonstrating that the repressive<br />

effect <strong>of</strong> p300 CRD 1 domain was dependent on the conjugation <strong>of</strong> SUMO.<br />

SUMO-modified CRD1 domain bound HDAC6 in vitro, and the ability <strong>of</strong><br />

SUMO conjugation to cause transcriptional repression was relieved both by<br />

histone deacetylase inhibitors and siRNA targeting <strong>of</strong> HDAC6. Thus the ability<br />

<strong>of</strong> p300 to act as a transcriptional repressor is dependent on the conjugation <strong>of</strong><br />

SUMO to the CRD 1 repression domain, and suggests that SUMO-dependent<br />

repression is mediated by recruitment <strong>of</strong> HDAC6.<br />

3.2.1 p300 is modified by SUMO in vivo<br />

To demonstrate directly that p300 was modified by SUMO in vivo, a<br />

stable cell line expressing 6His SUMO-1 was lysed directly with guanidine<br />

hydrochloride and 6His proteins isolated on Ni-NTA agarose. Western blotting<br />

analysis <strong>of</strong> the eluted proteins revealed that endogenous p300 was modified by<br />

6His SUMO-1 in vivo (Figure 22. ii. ). The specificity <strong>of</strong> this procedure was<br />

demonstrated by the simultaneous analysis <strong>of</strong> the parental HeLa cells, which<br />

lacked 6His-SUMO-1 (Figure 22. ii. ).<br />

To establish that the tandem modification motif within the CRD1 domain<br />

was indeed required for SUMO modification and to determine if p300 could also<br />

be modified by SUMO-2 and SUMO-3, COS7 cells were cotransfected with<br />

6His p300, residues 192-1044 containing the CRD1 domain, or a version that<br />

124<br />

1, Iý


(i)<br />

p300<br />

WKxEWKxE<br />

y9ISESKVEDCKMESTETEERSTELKTEIKEEEDQPSTSATQSSPAPGQSPý<br />

1035<br />

CBP KEETDTTEQKSEPMEVEEKKPEVKVEAKEEEEENSSNDTASQSTSPSQP082<br />

CRD I domain<br />

HA-SUMO-1 HA-SUMO-2 HA-SUMO-3<br />

NaNN<br />

Cl<br />

C, a<br />

6His oö<br />

s°<br />

(ii) HeLa SUMO-1 (iii)<br />

Q z° Z ä° x<br />

p300 4050<br />

4175<br />

1175<br />

Ni-beads<br />

Anti-HA<br />

extract<br />

Anti-HA<br />

extract wrr"+<br />

Anti-p300 I<br />

Figure 22. SUMO modifies p300 and CBP. (i) p300 and CBP possess a tandem SUMO<br />

modification motif (W)KxE) within the p21-inducible transcriptional repression domain, CRD I.<br />

(ii) Endogenous p300 and CBP are modified by SUMO-1. The SUMO-1 modification <strong>of</strong> p300<br />

is occurs at the CRDI domain. HeLa His6-SUMO-1 stable cell line and parental HeLa cell line<br />

were lysed in a buffer containing 6M Gaunadine Hydrochloride and purified on Ni' heads as<br />

described in materials and methods. Eluates were subjected to Western blotting with either<br />

p300 or CBP specific antibodies. (iii) COST cells were transfected with His6 p300192-1044 or<br />

His6 p300l92-1(X)4 and Ha-SUMO. Following transfection, samples were taken and crude cell<br />

extracts were prepared for Western blotting and probed with antibodies against Ha and p300, to<br />

test for transfection efficiency. The reaming cellular extracts were lysed in 6M Guanadine<br />

Hydrochloride and purified on Ni' beads. Eluates were Western blotted and probed with anti-<br />

Ha antibody.<br />

4175<br />

/ 16.5<br />

4175


lacks the CRD1 domain (01004-1045), and HA versions <strong>of</strong> SUMO. CRD1<br />

containing p300 is modified by SUMO-1, SUMO-2, and SUMO-3, whereas<br />

mutant p300 that has the CRD1 domain removed is not modified (Figure 22. iii. ).<br />

To monitor expression <strong>of</strong> HA-SUMO and p300 a fraction <strong>of</strong> the lysate, prior to<br />

Ni-NTA chromatography, was analysed by Western blotting using the HA-<br />

specific mAb 12CA5 and p300 specific mAb. Thus the observed differences<br />

were not due to differences in the levels <strong>of</strong> expression <strong>of</strong> either HA-SUMO or<br />

p300 (Figure 22. iii. ). SUMO modification <strong>of</strong> endogenous CBP was also<br />

detected, albeit to a lesser extent (Figure 22 (ii), bottom panel).<br />

3.2.2 SUMO conjugation to the N-terminus <strong>of</strong> p300 in vitro<br />

modification,<br />

To confirm that the N terminus <strong>of</strong> p300 is an in vitro substrate for SUMO<br />

"S-Met-labelled<br />

p300 (1-1255) was generated by in vitro<br />

transcription/translation, and incubated with recombinant SUMO, Ubc9, and<br />

SAE1/2. Incubation <strong>of</strong> p300 (1-1255) with either SUMO-1, SUMO-2, or<br />

SUMO-3, and the purified components <strong>of</strong> the SUMO conjugation pathway,<br />

resulted in the appearance <strong>of</strong> a more slowly (labelled) migrating species that was<br />

consistent with SUMO modification (Figure 23). To demonstrate that the more<br />

slowly migrating species <strong>of</strong> p300 was indeed SUMO modified, SUMO, Ubc9, or<br />

SAE1/2 were omitted from the in vitro reactions. Under these conditions<br />

appearance <strong>of</strong> the more slowly migrating species was abolished. Substitution <strong>of</strong><br />

SUMO with a tagged GST-SUMO further reduced the mobility <strong>of</strong> p300 (1-1255)<br />

confirming that the occurrence <strong>of</strong> the slower migrating forms observed are due to<br />

the conjugation <strong>of</strong> SUMO to the N terminus <strong>of</strong> p300 (1-1255).<br />

3.2.3 In vitro SUMO modification <strong>of</strong> N-terminus <strong>of</strong> p300 does not<br />

require an E3<br />

A major difference between the ubiquitin and SUMO conjugation<br />

pathways, is that SUMO conjugation in vitro does not require the presence <strong>of</strong> an<br />

E3 protein. Rather, the addition <strong>of</strong> an E3-like protein can enhance the in vitro<br />

modification <strong>of</strong> target substrates. These E3 ligases include RanBP2, and<br />

126


Assay mix<br />

CD<br />

2<br />

O<br />

SUMO<br />

SUMO-1 -p300<br />

SUMO-2<br />

SUMO-3<br />

bpw ' mw s A! WOM **4r* 44- p3001_,,<br />

5;<br />

ow, 44<br />

Zim-m<br />

0-4 w-*<br />

tww4<br />

".,, (^r rw<br />

_<br />

1ý<br />

k<br />

A_<br />

w.<br />

SUMO<br />

-p300<br />

r* rº 4- p300 1-1255<br />

_, ., _ f-p300<br />

SUMO<br />

-p300<br />

i-1255<br />

Figure 23. p300 is modified by SUMO-l/-2/-3 in vitro. The N-terminal half <strong>of</strong> p300,<br />

as 1-1255, containing the CRD 1 domain, was 35S-labeled by in vitro translation. The<br />

35S_<br />

labeled p3001 1255 was incubated in the assay for SUMO conjugation or in assay<br />

mixtures missing elements necessary for SUMO conjugation, as<br />

indicated. Reaction<br />

products were fractionated by SDS-PAGE before the gels were dried and analysed by<br />

phosphorimaging. The positions <strong>of</strong> the free and conjugated forms <strong>of</strong> p3001_1255 are<br />

indicated.


members <strong>of</strong> the PIAS family; PIAS 1, PIASy, and PIASxa. Theses four SUMO<br />

E3 were available within the group, and were available for study. E3 activity can<br />

be tested in vitro by determining the lowest concentration <strong>of</strong> Ubc9 that results in<br />

substrate concentration and then testing a range <strong>of</strong> E3 concentrations. If there is<br />

an E3-like affect, the amount <strong>of</strong> substrate concentration will increase as the E3<br />

concentration increases. Initially a range <strong>of</strong> recombinant Ubc9 concentrations<br />

were titrated into the in vitro modification assays, whilst keeping all other<br />

variables as standard (Figure 24), using both recombinant SUMO-1 or the mutant<br />

SUMO-2 (KI IR) protein that is incapable <strong>of</strong> forming polymeric chains with<br />

itself under the in vitro conditions. This is due to the mutation <strong>of</strong> the lysine<br />

residue within the SUMO consensus motif located within SUMO-2/3. At Ubc9<br />

concentration <strong>of</strong> 25ng a very weak modification was observed, and the band<br />

intensity increased at concentrations <strong>of</strong> Ubc9 above this (Figure 24). In vitro<br />

SUMO modification assays were performed using 25ng <strong>of</strong> Ubc9 and titrating in<br />

SUMO E3 ligases over a range <strong>of</strong> concentrations. Recombinant RanBP22532-2767<br />

fragment (provided by Ellis Jaffray, Figure 16b), which contains the SUMO E3<br />

ligases catalytic domain or members <strong>of</strong> the PIAS family (provided by Ellis<br />

Jaffray) were titrated as indicated (Figure 25), with either SUMO-1 or SUMO-2<br />

(KIIR) recombinant proteins as the modifiers. Enhancement <strong>of</strong> SUMO<br />

modification <strong>of</strong> the N terminus <strong>of</strong> p300 was not observed with any <strong>of</strong> the<br />

concentrations <strong>of</strong> RanBP225322767<br />

or PIAS proteins, with either SUMO-1 or<br />

SUMO-2 (KIIR) (Figure 25). As a positive control, to show the assay was<br />

functional SplOO was used as a substrate to show the effects <strong>of</strong> an E3 ligase on<br />

substrate modification. SUMO modification <strong>of</strong> SplOO has been reported to be<br />

enhanced by the RanBP2 E3 (Pichler et al., 2002). Modification <strong>of</strong> SplOO by<br />

SUMO-2 (KIIR) was enhanced by RanBP22532<br />

2767, although at higher E3<br />

concentrations there was an inhibitory effect. Modification <strong>of</strong> Sp100 by SUMO-1<br />

was also observed, albeit to a lesser extent than SUMO-2 (KIIR), and the same<br />

inhibitory effect was noticed at high E3 concentrations (Figure 25). A caveat to<br />

these experiments was the possibility that full length p300 may be required for an<br />

interaction with an E3, and thus function. Although as yet no SUMO E3 ligases<br />

have been shown to bind directly to their substrates<br />

128


Ubc9(ng) 05 10 15 20 25 50 100 200 375<br />

SUMO- I<br />

SUMO-2 (KIIR)<br />

z WW son Awa W= olm sI<br />

VMw<br />

I<br />

-7<br />

]+ SUMOs<br />

4 p3001-1255<br />

J+ SUMOs<br />

4 p3001-i,; 5<br />

Figure 24. Titration <strong>of</strong> Ubc9 into SUMO conjugation assay. The 35S_<br />

labeled<br />

p300, _, 255 was incubated in the assay for SUMO conjugation with increasing<br />

amounts <strong>of</strong> Ubc9, to determine the minimal amount <strong>of</strong> Ubc9 required for<br />

SUMO conjugation. Assays were carried out in the presence <strong>of</strong> SUMO-1<br />

and SUMO-2 (K 11 R) as indicated. Reaction products were fractionated by<br />

SDS-PAGE before the gels were dried and analysed by phosphorimaging.<br />

The positions <strong>of</strong> the free and conjugated forms <strong>of</strong> p300, _, 255 are indicated.


(i)<br />

(iii)<br />

, ý.. - .ý<br />

SUMO-1<br />

RanBP2,,,,,<br />

, r, w+ . "A" -ý -'ý+ --~ wý+ý" tai<br />

PIAS I<br />

,, ý,, l. º<br />

- +1 SUMO<br />

' Sp l 00<br />

(ii)<br />

SUMO-2(KIIR)<br />

RanBP2,,,,,<br />

767<br />

p3001-1255 00 sw +.....,, .. º ývr ,sw. r. -... ,. ", "., r, 'ww<br />

(x)<br />

RanBP225I'1767<br />

mWI! --<br />

Figure 25. RanBP2 and the PIAS family; PIASI, y, and xa, have all been reported as<br />

SUMO E3's. Neither RanBP2 nor the PIAS family <strong>of</strong> SUMO E3 ligases have ligase<br />

activity towards p300 modification in vitro. SUMO modification assays were set up,<br />

with increasing amounts <strong>of</strong> E3 ligase present. (i-iv)RanBP2 and PIAS I were added at 0,<br />

1,2,4,8,16,25,50, and 100ng, whilst (v-viii) PIASy and PIASxa were added at 0,2,4,<br />

6,25,50,100ng. (ix, x) 35S-labeled SplOO was included as a positive control, RanBP2 is<br />

known to enhance its SUMO modification, showed the reported preference<br />

for SUMO-2<br />

modification in the presence <strong>of</strong> RanBP2. At high concentrations <strong>of</strong> RanBP2, SUMO<br />

modification <strong>of</strong> Sp l00 was inhibited. Reaction products were fractionated by SDS-<br />

PAGE before the gels were dried and analysed by phosphorimaging. The positions <strong>of</strong><br />

the free and conjugated forms <strong>of</strong> p300i_i255 and Sp100 are indicated.


3.2.4 Identification <strong>of</strong> the lysine residues, within the CRD1<br />

domain <strong>of</strong> p300, that are conjugated to by SUMO<br />

Modification by SUMO usually occurs at the lysine residue contained within the<br />

iKxE conserved motif, requiring both a large hydrophobic and glutamic acid for<br />

efficient conjugation (Rodriguez et al, 2001). To investigate the sequence<br />

requirement for SUMO modification <strong>of</strong> p300, p300 sequences 1017-1029,<br />

comprising the CRD1 repression domain, were fused to GST and produced as<br />

recombinant protein, and sequenced to check for correct sequences (Figure 16b).<br />

The purity <strong>of</strong> the GST-CRD1 proteins was determined by SDS-PAGE analysis<br />

following staining with Coomassie brilliant blue R250 and subsequent de-<br />

staining (Fig. 26. see Appendix). The purified GST-CRD1 was incubated with the<br />

required components <strong>of</strong> the SUMO conjugation assay, containing "I-labelled<br />

SUMO-1. Under these experimental conditions, two species <strong>of</strong> GST-p300<br />

mCRD1 were generated that corresponded to single and<br />

double SUMO-1<br />

modified forms. The conjugation was particularly efficient at low concentrations<br />

<strong>of</strong> GST-p300 CRD 1 substrate (Figure 25. iii). To determine the sites at which<br />

p300 is modified by SUMO, alanine scanning mutagenesis <strong>of</strong> the 1017-1029<br />

region was performed in the context <strong>of</strong> the GST-p300 CRD1 fusion protein.<br />

CRD1 mutants were created by the hybridisation <strong>of</strong> oligonucleotides containing<br />

the appropriate mutations to encode an alanine residue, as indicated (Fig. 26. i. ).<br />

Annealed oligonucleotides were subsequently cloned into a modified GST<br />

expression vector, and transformed into the E. coli strain B834. Proteins were<br />

expressed and purified by GST affinity chromatography. The purity <strong>of</strong> the<br />

expressed proteins was analysed on SDS-PAGE, following staining with<br />

Coomassie brilliant blue R250 and subsequent de-staining, which indicated that<br />

fusion proteins were not contaminated by any additional proteins (Fig. 26. ii).<br />

These purified proteins were subsequently used as substrates<br />

for SUMO<br />

modification in vitro (Figure 26. iii. ). Mutation <strong>of</strong> single residues in the iKxE<br />

SUMO modification motif, L1019A, K1020A, E1022A, 11023A, K1024A, and<br />

K1026A abolished the presence <strong>of</strong> the higher molecular weight band,<br />

representing that <strong>of</strong> the double SUMO modified form <strong>of</strong> the CRD 1 peptide. The<br />

131


IM<br />

moo`-<br />

t4<br />

bx<br />

CU<br />

N<br />

(A 0<br />

. .'a. C<br />

.<br />

CZ .Uy<br />

4"<br />

Z<br />

Oäö<br />

bA ..,<br />

C- U MU,<br />

b Cl<br />

"0 TJ cu<br />

mä<br />

z3<br />

0 r- C, m3>,<br />

~ fr'<br />

(U ýi4 . -<br />

."4., .CU<br />

CD 0ö<br />

42<br />

"C "0 F- Qx<br />

- tom.<br />

oA(74"_ý<br />

Wö_0.1<br />

y<br />

u<br />

N0C rn vi<br />

Ub Cý C" M`<br />

0<br />

Z oää<br />

Ü<br />

f


ww<br />

kW<br />

xw<br />

.x<br />

WH<br />

a a a a a a a c+ a a a a o<br />

C)<br />

w<br />

o<br />

w<br />

W<br />

Q<br />

w<br />

W<br />

o<br />

w<br />

W<br />

0<br />

w<br />

W<br />

o<br />

w<br />

W<br />

Q<br />

w<br />

W<br />

o<br />

w<br />

W<br />

o<br />

w<br />

W<br />

M M<br />

a<br />

W<br />

w<br />

x w<br />

x w<br />

x w<br />

x w<br />

x w<br />

x w<br />

a x w<br />

x w<br />

,ý w<br />

oG w<br />

H<br />

w<br />

1-1<br />

w<br />

h-1<br />

w<br />

H<br />

w<br />

H<br />

a<br />

04<br />

w<br />

H<br />

w<br />

H<br />

w<br />

F-I<br />

w<br />

F-I<br />

w<br />

Q<br />

W<br />

F--I<br />

w<br />

Q<br />

<<br />

W<br />

I-I<br />

w<br />

1<br />

x<br />

w x<br />

H H H H H H H H H H H H H<br />

H a H H H H H H H H H H H Q<br />

U<br />

E a < < < < <<br />

o<br />

N N 7<br />

O O 7)<br />

x x<br />

p4 co rn o r Cl Cl) v Ln C [- o o w<br />

w<br />

[-H<br />

I<br />

u<br />

1--I<br />

O<br />

r-I<br />

O<br />

(N<br />

C)<br />

(N<br />

O<br />

N<br />

O<br />

N<br />

C)<br />

w 1 x H w H x w w w x<br />

(N<br />

O<br />

N<br />

O<br />

N<br />

O<br />

N<br />

C)<br />

(N<br />

O<br />

(N<br />

C)<br />

-i<br />

x<br />

4<br />

m<br />

a<br />

oF\InS 8IT IM<br />

21tzo I Nf IOZO IN<br />

dbzo I N/Vozo lx<br />

VLZOI<br />

d9ZOI<br />

vSZOIH<br />

VI7ZO Ix<br />

V %O II<br />

bZZOI3<br />

VIZOIJ<br />

Bozo IN<br />

b'6[O["I<br />

V81019<br />

I Q2Iý M91-3<br />

ISO<br />

QN 00 `DI- N(<br />

It M^


En ö w+ ööO ýs<br />

0 "0 A Ts<br />

4 , 0<br />

LwA<br />

u. cCD ziýU UA CZ c)<br />

b<br />

4. o c<br />

CS' UO >ý O<br />

Cý<br />

N 'b 1,12 ==<br />

U it Q<br />

m<br />

3ýr<br />

pyNQO<br />

N'-<br />

aý+ ýr y3, O 4r<br />

2 :1 CD<br />

...<br />

öW-o<br />

C7 'd<br />

a0¢A c) CD E4<br />

o A-ti ^<br />

G"ý+ CD Uö<br />

(q -ci<br />

a> i3 r;<br />

,V "v<br />

O<br />

oý`. ci<br />

rA<br />

CD "ctiNAQ<br />

öWo...<br />

W vn<br />

.r<br />

ÜÜ C7<br />

ýr r- vý C<br />

"o ><br />

0<br />

4r UO<br />

cu CD -- CD<br />

a+ Oý+.<br />

a<br />

Oý'"ma.<br />

Z) vii Uý+ O ^.<br />

cu º-+ t<br />

C/] ö-ý /<br />

4<br />

cl<br />

Gn<br />

ý<br />

L3+ 'em" =Ö<br />

N<br />

u<br />

42<br />

aý 33 cl +<br />

- k4 CA -0ý.<br />

o<br />

C:<br />

U<br />

4 o a, Gn Oo o= Gn<br />

r. uA<br />

i: 4E U<br />

VJ ä C7 iäcä<br />

I


o, N<br />

w<br />

ý G°<br />

a<br />

--<br />

wr<br />

F" C<br />

c<br />

[r<br />

V<br />

r t<br />

iE+<br />

.<br />

O<br />

Il E1rll\Id<br />

1rZO<br />

I N'1OZO<br />

IN<br />

VIZOr)'bozo[)<br />

VLZOl9<br />

d9ZOl ýl<br />

vczori<br />

Vt7 OIN<br />

b'£ZO[I<br />

VZZ013<br />

VIZOI. L<br />

Bozolx<br />

V6101'I<br />

V81013<br />

IM<br />

ISO<br />

I<br />

ö ö<br />

N<br />

+ +<br />

A<br />

{<br />

Ö<br />

L<br />

T


C<br />

. 230<br />

N<br />

N<br />

C<br />

0<br />

U<br />

0<br />

.5<br />

«r<br />

U<br />

ca<br />

h-<br />


mutants were still capable <strong>of</strong> being modified by a single SUMO protein, as a<br />

higher molecular weight form was still observed (Figure 26. iv. ). A further set <strong>of</strong><br />

mutants, changing both the SUMO acceptor lysines to either alanine or arginine<br />

was generated. When used as substrate in the in vitro SUMO conjugations, the<br />

K1020A/K1024A and K1020R/K1024R mutants failed to show any<br />

modification, as indicated by the absence <strong>of</strong> any higher molecular weight species<br />

(Figure 26. iv. ). The fused SUMO modification sites from PML and adenovirus<br />

E1B55K were also modified in vitro, although the extent <strong>of</strong> modification to the<br />

tandem motif was less than that <strong>of</strong> the GST-p300 CRD1 (Figure 26. iv. ).<br />

As part <strong>of</strong> the collaborative effort with Neil D. Pekins (<strong>University</strong> <strong>of</strong><br />

Dundee) to investigate the SUMO modification <strong>of</strong> p300, the same alanine<br />

mutants- <strong>of</strong> the CRD 1 domain, within a Ga14 reporter construct, were used to<br />

investigate the relationship between SUMO conjugation and transcriptional<br />

activity (Fig. 27. ). The results, originally preformed by D. Bumpass and<br />

reproduced with kind permission, indicated that there was a strong correlation<br />

between the ability <strong>of</strong> SUMO to conjugate to the CRD 1 domain, and its ability to<br />

mediate transcriptional repression in vivo (Table. 3. ).<br />

3.2.5 SUMO modification is required for CRD1 activity in vivo<br />

The construction <strong>of</strong> the Lys to Ala/Arg mutants <strong>of</strong> the CRD1 domain<br />

showed that the SUMO modification motif was required for SUMO conjugation,<br />

and that mutations affecting SUMO conjugation also affected p300<br />

transcriptional repression capabilities. To provide direct evidence that SUMO is<br />

required for p300 repressive effect on transcription, and rule out the possibility <strong>of</strong><br />

modification by other posttranslational modifiers that act through the CRD 1<br />

domain, further experiments were undertaken. These experiments involved<br />

SUMO-specific proteases, which would remove any SUMO proteins bound to<br />

CRD1 and a dominant negative Ubc9 mutant <strong>of</strong> the SUMO conjugation pathway,<br />

which would inhibit SUMO conjugation. These were utilised to show the role <strong>of</strong><br />

SUMO in p300 mediated repression.<br />

To measure the repressive effect <strong>of</strong> SUMO modified CRD1, as indicated<br />

through luciferase activity, Gal4-p300 N+, containing the CRD 1 domain or<br />

Ga14-p300 N-, which lacks the CRD1 domain (Figure 16a), were cotransfected<br />

137


into U-2 OS cells along with expression plasmids for SUMO-specific protease<br />

(SSP3/SENP2), catalytically inactive form <strong>of</strong> the protease C548A SSP3/SENP2,<br />

and catalytically inactive SUMO E2, C93S Ubc9, that acts a dominant negative<br />

mutant, as indicated (Figure 28). The SSP3/SENP2 protease was utilised due to<br />

its availability within the lab, and its nuclear localisation (O. A. Vaughan,<br />

personal communication).<br />

Cotransfections with pcDNA3 empty vector and<br />

Ga14-p300 N-, increased expression by almost 300-fold when compared to Ga14<br />

expression plasmid (Figure 28). In comparison cotransfection with Gal4-p300<br />

N+, which contains CRD1, lead to a much smaller increase <strong>of</strong> expression by<br />

around 11-fold (Figure 28). Reporter activity is therefore repressed by 24-fold,<br />

when CRD 1 is expressed in the presence <strong>of</strong> pcDNA3 empty vector.<br />

Cotransfection <strong>of</strong> a plasmid expressing SSP3/SENP2 with<br />

Ga14-p300 N-, had<br />

little influence on reporter activity, however when SSP3/SENP2 was<br />

coexpressed with Gal4-p300 N+ plasmid, a dramatic increase <strong>of</strong> reporter activity<br />

was observed (Figure 28). Indeed under these conditions only residual<br />

repressive activity, 1.4-fold, was observed (Figure 28). The relief <strong>of</strong> repression<br />

was not observed when Ga14-p300 N+ was cotransfected with the plasmid<br />

encoding the catalytically inactive C548A SSP3/SENP2 (Figure 28). Titration <strong>of</strong><br />

SSP3 and C548A SSP3/SENP2 revealed that SSP3/SENP2 relieved CRD1-<br />

mediated repression even at very low levels <strong>of</strong> cotransfected DNA. At these<br />

levels <strong>of</strong> cotransfected DNA, C548A SSP3/SENP2 did not alter reporter activity,<br />

but higher levels <strong>of</strong> C548A SSP3/SENP2 do appear to relieve repression. This<br />

may be due to a transdominant effect on the processing <strong>of</strong> SUMO (Fig. 28. ii. ).<br />

The results clearly demonstrate that SUMO deconjugation by SSP3/SENP2<br />

protease activity relieves p300 CRD1-mediated transcriptional repression.<br />

Further evidence to support SUMO modified CRD1 role in transcriptional<br />

repression came via use <strong>of</strong> C93S Ubc9, a catalytically inactive form <strong>of</strong> the<br />

SUMO conjugating enzyme Ubc9, which functions as a dominant-negative<br />

mutant. The dominant negative Ubc9 mutant, lacking the catalytic cysteine<br />

residue (C93S), is not able to accept the transfer <strong>of</strong> SUMO from SAE1/2 to<br />

Ubc9. The Ubc9 mutant will form an oxy-ester with SUMO, preventing SUMO<br />

transfer and inhibiting SUMO conjugation.<br />

138


(Iý<br />

-}<br />

4001<br />

300<br />

2 00 1<br />

100<br />

(ii) 20<br />

c<br />

ca<br />

U<br />

fC<br />

a8<br />

ö<br />

16<br />

12<br />

4<br />

Ga14 p300 N+<br />

(1005-1044)<br />

Ga14 p300 N<br />

Ga14<br />

0 0 20 40 60 80<br />

cotransfected DNA (ng)<br />

SP3/SENP2<br />

Figure 2tß. SUMO modification <strong>of</strong> the CRD domain is required for repression.<br />

(i) SUMO-specific protease 3 (SSP3) relieves p300-mediated repression. Expression<br />

plasmids encoding the indicated Ga14 or Ga14-p30() fusion proteins (511o) were<br />

cotransfected into U-2 OS cells with Gal4 AdMLP luciferase reporter plasmid (2p, g) and<br />

either pcDNA3 or pcDNA expression constructs for SSP3 or C548A SSP3 (50ng).<br />

(ii) Dose response <strong>of</strong> SSP3 action. An expression plasmid encoding<br />

1044 fusion protein (5ng) was cotransfected into U-2 OS cells with a<br />

Gal4-p300 N+l005-<br />

Ga14 AdMLP<br />

luciferase reporter plasmid (2µg)and increasing amounts <strong>of</strong> pcDNA3 expression<br />

constructs for either SSP3 or C548A SSP3. Results are displayed as fold activation<br />

compared to transtection with Ga14 lueiferase and the pcDNA3 empty. Results shown in<br />

(i), and (ii) are the means <strong>of</strong> three separate experiments, and standard deviations are<br />

shown.


The effects <strong>of</strong> SSP3/SENP2 and C93S Ubc9 on SUMO conjugation to<br />

p300 were directly demonstrated by expressing HA-SUMO-1 and 6His-p300<br />

with either SSP3/SENP2 or C93S Ubc9.6His-tagged p300 was isolated on Ni-<br />

NTA agarose and eluted proteins analysed by Western blotting with antibody<br />

specific to the HA-tag. Analysis <strong>of</strong> the Western blot indicated that there was no<br />

SUMO conjugation to p300 in the presence <strong>of</strong> SSP3/SENP2, whilst SUMO<br />

conjugation to p300 was severely reduced by expression <strong>of</strong><br />

C93S Ubc9 (Figure<br />

29. i). p21 has been demonstrated to relieve p300 transcriptional repressive<br />

capabilities (Snowden et al., 2000). Thus there was the possibility that p21<br />

expression may result in a decrease <strong>of</strong> SUMO conjugation to p300. Although<br />

p21 does not directly bind to p300 itself, and functions through an as yet<br />

unidentified adaptor protein(s). p21 expression was shown not influence the<br />

level <strong>of</strong> SUMO conjugation to the CRD1 domain (Figure 29. ii), showing that<br />

p21's repressional relieving capabilities is not mediated through the removal <strong>of</strong><br />

SUMO or direct competition. A control reaction consisted <strong>of</strong> cotransfecting<br />

alcohol dehydrogenase (ADH), which was cloned into the same expression<br />

vector as p21, RSV (Figure 29. ii).<br />

3.2.6 SUMO-Modified CRD1 recruits a Histone Deacetylase<br />

SUMO modification <strong>of</strong> the CRD 1 domain could lead to transcriptional<br />

repression by a variety <strong>of</strong> mechanisms. A common feature <strong>of</strong> transcriptional<br />

repression in higher eukaryotes is the recruitment <strong>of</strong> histone deacetylases<br />

(HDACs). To examine the possibility that HDAC activity was<br />

involved in<br />

SUMO-dependent, CRD1-mediated repression, we employed the HDAC<br />

inhibitor Trichostatin A. The addition <strong>of</strong> Trichostatin A had no effect on reporter<br />

activity mediated by Ga14 alone, it dramatically increased transcription mediated<br />

by a Ga14 fusion protein containing the CRD 1 domain (Figure 30. i. ) These<br />

results were originally preformed by D. Bumpass and reproduced with kind<br />

permission. Trichostatin A had only a marginal effect on transcription mediated<br />

by a Ga14-p300 fusion protein lacking the CRD1 domain (Figure 30. i. ). To<br />

identify the HDAC responsible for the interaction with SUMO modified CRD1<br />

domain, an in vitro binding assay was established. This consisted <strong>of</strong> GST fusion<br />

140


ýiý<br />

Ni-beads<br />

Anti-HA<br />

extract<br />

Anti-HA<br />

extract<br />

Anti-p300<br />

extract<br />

Anti-SSP3<br />

extract<br />

Anti-Ubc9<br />

Q<br />

Z<br />

da<br />

HA-SUMO-1<br />

6His p30O<br />

(ii)<br />

u N<br />

M<br />

c<br />

-fl<br />

Z.<br />

++<br />

a vD -d<br />

qw<br />

Ni-beads<br />

Anti-HA<br />

extract<br />

HA-SUMO- I<br />

6His p300<br />

Q<br />

Q<br />

4 175<br />

Anti-HA 4 175<br />

extract<br />

Anti-p300 4 175<br />

Figure 29. Control over SUMO conjugation to p300, SSP3 and do Ubc9 prevent<br />

SUMO modification <strong>of</strong> p300, whilst p2l has no direct control over conjugation. (i)<br />

COST cells were transfected with HA-SUMO-1,6His-p300, and either pcDNA3, SSP3,<br />

or the dominant negative Ubc9, as indicated. Following transfection crude samples<br />

were taken to test for transfection efficiency, whilst the remaining cell lysate were lysed<br />

in 6M Guanadine HCI and purified on Ni-beads. (ii) To investigate whether p21<br />

competed with SUMO for binding to the CRDI domain p21 and control vector ADH,<br />

were co-transfected with HA-SUMO-1 and 6His-p300, and analysed as previously<br />

described.


construct, possessing either the minimal CRD 1 domain, or in case, both SUMO<br />

conjugation sites and the surrounding p300 sequence was required for binding, a<br />

larger construct possessing amino acids 852-1071 were used (Figure 16b). As<br />

such recombinant GST fusion constructs, containing either an extended 220<br />

amino acid fragment <strong>of</strong> p300, that includes the CRD1 domain (residues 852-<br />

1071) or the CRD1 domain (residues 1018-1027) were used as substrates in<br />

SUMO conjugation assays (Figure 16b). Following incubation with individual<br />

HDACs, generated by in vitro transcription/ translation in the presence <strong>of</strong> 3SS-<br />

Met, GST fusion proteins were recovered on glutathione agarose and their<br />

binding capacity for individual HDACs was determined. Both SUMO modified<br />

p300 fragments (p3008521071<br />

and p3001005-1o) bound to HDAC6, whereas<br />

unmodified fragments and that <strong>of</strong> GST were unable to bind to HDAC6 (Figure<br />

29. ii. ). The interaction was specific to HDAC6 as no other HDAC was seen to<br />

interact with the SUMO modified fragments (Figure 30. ii. ).<br />

To establish the importance <strong>of</strong> HDAC6 in the regulation <strong>of</strong> p300 activity,<br />

endogenous HDAC6 expression was ablated using siRNA. Plasmids expressing<br />

HDAC6-specific siRNA or a scrambled siRNA were cotransfected with Gal4-<br />

p300 N- or Gal4-p300 N+ (containing CRD1) and a Ga14 luciferase reporter<br />

plasmid. Gal4-p300 N-, which cannot be SUMO modified, drives high levels <strong>of</strong><br />

reporter activity which are unaffected by plasmids expressing siRNA to HDAC6.<br />

In contrast, transcriptional repression mediated by Gal4-p300 N+CRD1 is<br />

relieved by siRNA to HDAC6 but not by the control scrambled siRNA (Figure<br />

30. iii. ) These experiments were originally preformed by D. Bumpass, and are<br />

reproduced with kind permision.<br />

142


Figure 30. SUMO-mediated p300 repression is mediated by a Histone Deacetylase<br />

(i)Trichostatin A alleviates CRDl-dependent p300 repression. HEK 293 cells were<br />

transfected with 5 µg Ga14 E1B CAT reporter together with 10 ng <strong>of</strong> Ga14,10 ng <strong>of</strong> Ga14-<br />

p300 (192-1044), or 0.2 ng <strong>of</strong> Ga14-p300 (192-1004) expression plasmids. Trichostatin A<br />

was added to the cells after 4 hr and processesed for CAT assay 20 hr later(ii) GST-p300-<br />

CRD1 or GST-p300852.1071<br />

(Figure 16b) were modified by SUMO-1 in vitro and the<br />

modified species recovered on glutathione agarose. 35S-labeled in vitro-translated HDAC4<br />

and HDAC6 were incubated with glutathione agarose bound GST, unmodified GST-p300<br />

proteins, or the SUMO modified GST-p300 proteins as indicated. After extensive washing,<br />

bound proteins were analysed by SDS PAGE and phosphorimaging. 10% <strong>of</strong> the 35S-labeled<br />

HDAC4 and HDAC6 input was also analysed. (iii) Ga14-p300 N- and Ga14-p300 N+mCRD1<br />

expression plasmids (0.5ng) were cotransfected into U-2 OS cells together with a Ga14-<br />

dependent luciferase reporter (1.5 µg) and 2 µg <strong>of</strong> either the scrambled siRNA plasmid or a<br />

pool <strong>of</strong> the three HDAC6 siRNA expression plasmids as indicated. Plotted results presented<br />

are the means <strong>of</strong> three separate experiments, and standard deviations are shown. The results<br />

shown in Figure. 29i and iii are taken from work carried out by D. Bumpass et al and are<br />

reproduced with kind permission.


ý11)<br />

41<br />

.S<br />

ý-<br />

I.<br />

(i)<br />

v<br />

C<br />

0<br />

L<br />

4)<br />

i C<br />

0<br />

U<br />

w<br />

.5<br />

U<br />

F-<br />

Q<br />

U<br />

Tv<br />

C9<br />

D O ö<br />

U<br />

Ö<br />

Ö<br />

n ä ä<br />

In<br />

(D VI) (D 0+ (D (D +<br />

V '0<br />

w w<br />

Ö p<br />

NN 40 w<br />

AA<br />

uiu 4HDAC6<br />

HDAC4<br />

cp<br />

_:<br />

siRNA scr.<br />

100 - siRNA HDAC6<br />

80<br />

60<br />

40<br />

20<br />

E<br />

p<br />

W<br />

Z<br />

1<br />

Ä<br />

W<br />

Ö<br />

Z<br />

O


3.2.7 Conclusion<br />

Through the generation <strong>of</strong> p300 CRD 1 mutants the minimal repression<br />

domain within p300 was mapped; residues 1017-1029. Occuring within this<br />

domain are two copies <strong>of</strong> the iKxE which represent the consensus modification<br />

site for the small ubiquitin-like protein SUMO. Here, we demonstrate that<br />

endogenous p300 is modified by SUMO in vivo and the isolated sequence is<br />

modified by SUMO in vitro in the presence <strong>of</strong> SUMO-activating enzyme and<br />

Ubc9. We also demonstrate that endogenous CBP is SUMO modified, although<br />

to a lesser extent than p300, although no further characterisation <strong>of</strong> SUMO<br />

modification <strong>of</strong> CBP was preformed. Although Best et al, do report a<br />

relocalisation <strong>of</strong> CBP following overexpression <strong>of</strong> a SUMO protease (Best et al.,<br />

2002). Mutagenesis reveals a strict correlation between SUMO modification and<br />

repression (Figure 25 and 26). Each SUMO modification motif appears to<br />

function independently, and both motifs are required for full repression activity<br />

(Figure, 26). The CRD 1 SUMO modification sequence is also present in the<br />

equivalent repression domain <strong>of</strong> CBP (Snowden et al., 2000), is highly<br />

conserved in human, mouse, and Xenopus versions <strong>of</strong> the proteins, and is also<br />

found in Drosophila CBP. Not only is the double SUMO modification motif<br />

conserved, but its location, just to the amino terminal side <strong>of</strong> the bromodomain,<br />

is also conserved.<br />

SUMO is conjugated to a numerous cellular substrates via an enzymatic<br />

pathway, analogous to, but distinct from that <strong>of</strong> ubiquitin conjugation (Hay,<br />

2001; Pichler et al., 2002; VanDemark and Hill, 2002) that results in the transfer<br />

<strong>of</strong> SUMO from Ubc9 to the target protein. Substrate recognition is mediated<br />

through acceptor lysine residues on target proteins within the consensus motif<br />

ipKxE, and structural analysis has indicated that this sequence is contacted by<br />

Ubc9 (Bernier-Villamor et al., 2002; Lin et al., 2002). Thus, the sequence<br />

LKTEIKEE in p300 is likely to be directly recognised by Ubc9. Recently, it has<br />

been shown that selection <strong>of</strong> target proteins for modification by SUMO is aided<br />

by a number <strong>of</strong> proteins that appear to function in a fashion that is similar to the<br />

specificity-determining E3s in the ubiqüitin system (Johnson and Gupta, 2001;<br />

Kahyo et al., 2001; Takahashi et al., 2001; Kagey et al., 2003). We could not<br />

145


show any E3-ligase activity mediated through members <strong>of</strong> the PIAS family or<br />

RanBP2 towards SUMO modification <strong>of</strong> the N-terminal half <strong>of</strong> p300. This does<br />

not rule out E3-ligase activity towards the SUMO modification <strong>of</strong><br />

full length<br />

p300. A more direct role for SUMO modification is evident in the case <strong>of</strong> c-Myb<br />

and SP3, in which SUMO modification <strong>of</strong> a negative regulatory domain mediates<br />

transcriptional repression (Bies et al., 2002; Ross et al. 2002; Sapetsching et al.,<br />

2002).<br />

p300 transcriptional repression is relieved by Trichostatin A in vivo , and<br />

the SUMO-dependent binding <strong>of</strong> HDAC6 to the CRDI domain suggests that<br />

p300 transcriptional repression is mediated by SUMO-dependent recruitment <strong>of</strong><br />

HDAC6 to the CRD 1 <strong>of</strong> p300. This is supported by the observation that CRD 1-<br />

dependent, p21-mediated relief <strong>of</strong> repression was blocked by coexpression <strong>of</strong><br />

HDAC6 and that siRNA-mediated ablation <strong>of</strong> endogenous HDAC6 expression<br />

leads to CRD1-dependent derepression (<strong>Girdwood</strong> et al., 2003). That HDAC6<br />

expression did not result in any further repression <strong>of</strong> the CRD1 domain when<br />

expressed in the absence <strong>of</strong> p21, suggests that this domain is maximally<br />

repressed. This is consistent with the very low basal levels <strong>of</strong> activity seen with<br />

the Gal4-p300 N+CRD 1 vectors which (Figure 16a) display similar levels <strong>of</strong><br />

activity to Ga14 alone. Thus, it is likely that SUMO-modified p300 recruits<br />

HDAC6 to a subset <strong>of</strong> promoters that are susceptible to both p300-mediated<br />

repression and p21 inducibility (Gregory et al., 2002). Deacetylation <strong>of</strong> core<br />

histones and other transcriptional regulators could then lead to the observed<br />

effects on transcription. Histone deacetylase 1,4, and 6 have all been reported to<br />

be substrates for SUMO modification (<strong>David</strong> et al., 2002; Kirsh et al., 2002;<br />

Tatham et al., 2001), and while non-modified forms <strong>of</strong> HDACI and HDAC4<br />

appear to be less active deacetylases, the role <strong>of</strong> SUMO modification in vivo has<br />

yet to be established. Members <strong>of</strong> the class II HDACs (-4, -5, -6, and -7) all<br />

appear to undergo rapid nucleo-cytoplasmic shuttling (Khochbin et al., 2001),<br />

and recent evidence suggests that HDAC6 binds ubiquitin through its conserved<br />

C-terminal zinc finger domain (Hook et al., 2002; Seigneurin-Berny et al.,<br />

2001). Thus, regulated nuclear entry <strong>of</strong> HDAC6 coupled with SUMO<br />

modification <strong>of</strong> the CRD 1 domain could serve to regulate the transcriptional<br />

output <strong>of</strong> p300.<br />

146


p300-mediated repression activity in vivo is relieved by SSP3 (Figure 27<br />

and 28), a SUMO-specific protease also known as SENP2, SMT3IP2, or Axam<br />

(Hang and Dasso, 2002; Kadoya et at., 2002; Nishida et al., 2001) which<br />

removes SUMO from modified substrates. While this directly demonstrates that<br />

p300 repression is mediated by SUMO and not some other ubiquitin-like protein,<br />

it also suggests a mechanism for control <strong>of</strong> p300 repression activity. Modulation<br />

<strong>of</strong> SSP3/SENP2 activity or subcellular localisation by signalling molecules could<br />

allow the protease to deconjugate SUMO modified p300 and thus relieve<br />

repression in response to various signals. It has recently been demonstrated that<br />

Axam/SMT3IP2/SENP2/SSP3 can activate transcription in a PML-dependent<br />

fashion (Best et al., 2002). In addition to transcriptional repression, SUMO<br />

modification also appears to have additional diverse effects. These include<br />

antagonism <strong>of</strong> ubiquitination, alteration <strong>of</strong> subcellular localisation, and<br />

transcriptional activation. It is likely that the diverse biological effects <strong>of</strong> SUMO<br />

modification, including the repression <strong>of</strong> p300-dependent transcription reported<br />

here, are mediated by the ability <strong>of</strong> SUMO-modified proteins to selectively<br />

recruit new protein partners that alter their properties.<br />

147


3.3 Transcriptional repression is preferentially mediated<br />

through SUMO-2 and SUMO-3<br />

siRNA technology provides a useful mechanism <strong>of</strong> decreasing if not<br />

completely ablating the levels <strong>of</strong> the targeted gene, through the endonuclolytic<br />

cleavage <strong>of</strong> the target mRNA (as detailed in material and methods). To<br />

investigate the potential for distinct roles <strong>of</strong> SUMO-1. and SUMO-2/3 as<br />

transcriptional regulators, siRNA has been used to ablate expression <strong>of</strong> the<br />

different SUMO is<strong>of</strong>orms individually. Using three previously identified<br />

substrates repressed by SUMO modification it was demonstrated that ablation <strong>of</strong><br />

SUMO-2/3 substantially derepressed transcriptional activity <strong>of</strong> p300, Sp3, and<br />

Elk-1. Ablation <strong>of</strong> SUMO-1 results in only a modest derepression <strong>of</strong><br />

transcription.<br />

The conjugation <strong>of</strong> all three SUMO species to p300 occurs with equal<br />

efficiency in the presence <strong>of</strong> El and E2 in vitro, and all three conjugate to the<br />

same iKxE SUMO conjugation motif (Figure 23). Despite the sub-groupings <strong>of</strong><br />

the SUMO is<strong>of</strong>orms, there have been few reported differences in the functional<br />

consequences regarding modification by SUMO-1 or SUMO-2/-3. There are<br />

large pools <strong>of</strong> free SUMO-2/-3 in cells and under certain stress situations the<br />

conjugation <strong>of</strong> SUMO-2/-3 (Saitoh and Hinchey, 2000) to substrates can be<br />

induced, although hypoxia has been demonstrated to increase the levels <strong>of</strong><br />

SUMO-1 mRNA, suggesting that stress responses may not be restricted to<br />

SUMO-2/-3 alone. Indeed SUMO-1 has been shown to modify NEMO<br />

following genotoxic stress and result in its translocation to the nucleus (Huang et<br />

al., 2003).<br />

148


3.3.1 Ablation <strong>of</strong> SUMO-1 and SUMO-2/-3 expression following<br />

treatment with synthetic siRNA oligonucleotides<br />

Initially synthetic oligonucleotides were the only siRNA technology<br />

available, although subsequently plasmid siRNA expression constructs were<br />

developed. As such the initial work was preformed with synthetic<br />

oligonucleotides, although later, due to their availability, plasmid based siRNA<br />

technology was utilized. To check the specificity and efficiency <strong>of</strong> the siRNA<br />

mediated silencing synthetic siRNA oligonucleotides (100nM) were transfected<br />

by calcium phosphate as described in material and methods, and SUMO levels<br />

were determined by both Western blotting and immun<strong>of</strong>luorescence. Western<br />

blot analysis demonstrated that SUMO-1 levels were reduced following<br />

transfection <strong>of</strong> SUMO-1 siRNA oligonucleotides, whilst the levels <strong>of</strong> SUMO-2/-<br />

3 were unaffected. Experiments with siRNA oligonucleotides against SUMO-2<br />

+3 gave the reciprocal results (Figures 31). As shown by immun<strong>of</strong>luorescence,<br />

U2-OS cells transfected with the random sequence (RSC) oligonucleotide, cells<br />

showed the predicted distribution with all three SUMO is<strong>of</strong>orms being<br />

predominatly nuclear in distribution, and could be observed in discrete nuclear<br />

foci (PML bodies). SUMO-1 levels were observed to be lower than those <strong>of</strong><br />

SUMO-2/-3 as determined by Western blot analysis. Following treatment <strong>of</strong> U2-<br />

OS with the SUMO-1 specific siRNA oligonucleotide, the reduction in the levels<br />

<strong>of</strong> SUMO-1 staining was observed, and similarly following treatment with<br />

SUMO-2 and SUMO-3 oligonucleotides, a reduction in the amounts <strong>of</strong> SUMO-<br />

20 was noticed (Figure 32).<br />

3.3.2 Transcriptional repression is relieved preferentially<br />

through shRNA targeting <strong>of</strong> SUMO-2/-3<br />

To identify whether there are any functional differences between<br />

conjugation by SUMO-1 and that <strong>of</strong> SUMO-2/-3, in their transcriptional<br />

repression roles, pSUPER retro constructs (Oligoengine) were constructed as per<br />

manufacturers instructions, for each SUMO is<strong>of</strong>orm (as described in material and<br />

149


1.<br />

11.<br />

Anti-SUMO-1 Anti-SUMO-2/-3<br />

RSC S1 S2+S3 siRNA RSC Sl S2+S3<br />

_<br />

1Conjugated<br />

SUMO<br />

f Free SUMO f `..... ý.. ý.<br />

Actin -º<br />

Amount <strong>of</strong> SUMO-1 Amount <strong>of</strong> SUMO-2/3<br />

too 100 1<br />

90 { 90<br />

io<br />

60<br />

so 50<br />

40<br />

m<br />

2<br />

SO<br />

RSC SUMO I SUMO-2/7<br />

FIRM oIIg. (o) tnn. /hcted<br />

i<br />

70<br />

60<br />

30<br />

20<br />

10<br />

0<br />

ý-<br />

0<br />

-- _____ý-..... .. -.. _. -,<br />

RSC SUMO-1 SUMO-2/3<br />

"IRMA eliye(ý) trendected<br />

Figure 31. i. U2-OS cells were transfected with siRNA targeting SUMO-1 or SUMO-2 and<br />

SUMO-3, or control siRNA respectively and left for 48 hours, before whole cell extracts<br />

were run on 10% SDS-PAGE gels and subjected to Western blotting with either antibodies<br />

against SUMO-1 or SUMO-3, which also recognizes SUMO-2. The knocking down <strong>of</strong><br />

SUMO-1 mRNA was demonstrated not to affect levels <strong>of</strong> SUMO-2 or SUMO-3, and this<br />

was demonstrated to be reciprocal. ii. The Western blots were subjected to quantification<br />

via MacBas s<strong>of</strong>tware and the relative amounts <strong>of</strong> the proteins, as a percentage <strong>of</strong> the control<br />

were determined.


Figure 32. Knockdown <strong>of</strong> endogenous SUMO levels through siRNA mediated<br />

silencing. Endogenous SUMO shows a predominantly nuclear localisation, and<br />

accumulates into punctate nuclear dot like structures. Transfection RSC<br />

oligonucleotides has no effect on the levels <strong>of</strong> either SUMO-1 or SUMO-2/3.<br />

Transfection with the oligonucleotide targeting SUMO-1 specifically knockdown the<br />

levels <strong>of</strong> SUMO-1 and does not affect the levels <strong>of</strong> SUMO-2/3. Conversly transfection<br />

<strong>of</strong> oligonucleotides targeting SUMO-2/3 reduced SUMO-2/3 levels and did not affect<br />

SUMO-1 levels. Synthetic siRNA oligonucleotides (100nm) were transfected by<br />

calcium phosphate as described in material and methods. Endogenous SUMO-1 was<br />

detected with anti-GMPI (Zymed) antibody with Texas Red as secondary, whilst<br />

SUMO-2/-3 was detected with an anti-Sentrin-2 antibody (Zymed) with FITC as the<br />

secondary, as indicated. White text denotes siRNA oligonucleotide transfected.<br />

Fluorescence microscopy was carried out with a DeltaVision microscope (Olympus<br />

IX70) with a 1.40 NA 100X objective and Photometric CH300 CCD camera. lOX 0.2<br />

[um sections were taken. Images were deconvoluted using S<strong>of</strong>tWorx (Applied<br />

Precision) s<strong>of</strong>tware.


methods, and Appendix). Co-transfection <strong>of</strong> the synthetic oligonucleotides with<br />

the appropriate reporter constructs had proved lethal to the U2-OS cells, so the<br />

pSUPER retro plasmids were utilised. To identify differential roles played by<br />

the two SUMO subgroups, these pSUPER retro SUMO-1 or pSUPER retro<br />

SUMO-2 with SUMO-3 vectors were to be co-transfected with a luciferase<br />

reporter construct and Ga14 fusions <strong>of</strong> known SUMO substrates, whose<br />

modification has been documented to result in transcriptional repression (Figure<br />

16a). The siRNA targeting <strong>of</strong> the appropriate SUMO or SUMOs RNA, would<br />

result in a reduction in the amount <strong>of</strong> SUMO available for conjugation and hence<br />

affect a substrates repressional status. Additional experiments were to be<br />

preformed co-transfecting SUMO conjugation deficient versions <strong>of</strong> the<br />

substrates, to act as a control, to shown that the changes in luciferase activity<br />

were directly related to the SUMO modification status <strong>of</strong> a substrate, as there<br />

should be no additional affect on luciferase activity. The three SUMO substrates<br />

chosen were; p300, Sp3 (Sapetschnig et al., 2002; Ross et al., 2002), and Elk-1<br />

(Yang et al., 2003) (Figure 16a), and the repression status <strong>of</strong> each following<br />

siRNA targeting <strong>of</strong> the different SUMO is<strong>of</strong>orms was investigated.<br />

, Knockdown <strong>of</strong> SUMO-2/3 resulted in a substantial relief <strong>of</strong> p300-SUMO<br />

mediated repression, whilst the knockdown <strong>of</strong> SUMO-1 resulted in only a<br />

modest increase in transcription (Figure 33. i and ii). The knockdown <strong>of</strong> SUMO-<br />

2/3 also resulted in an increase in the transcriptional activity <strong>of</strong> the SUMO<br />

conjugation deficient mutant <strong>of</strong> p300, N- (Figures 16 and 33. i and ii). The<br />

shRNA mediated targeting <strong>of</strong> SUMO-2/3 similarly preferentially relieved Elk-1<br />

mediated transcriptional repression, compared to the relief through SUMO-1<br />

knockdown (Figure 34. i and ii). The Elk-1 constructs that were SUMO<br />

conjugation deficient were shown not to be significantly effected by SUMO<br />

knockdown. Co-expression <strong>of</strong> Ga14-Sp3 with the pSUPER retro SUMO-1 vector<br />

resulted in a minimal relief <strong>of</strong> repression, whilst similarly to those results<br />

obtained from Elk-l, the co-expression with the pRETRO super plasmids for<br />

SUMO-2/3 resulted in a substantial relief <strong>of</strong> repression, but only with the SUMO<br />

conjugation competent Sp3 construct (Figure35. i and ii).<br />

To investigate how SUMO deficient constructs, such as p300 N-, were<br />

152


(i)<br />

(II)<br />

1200000<br />

1000000<br />

800000<br />

6<br />

. 600000<br />

oC<br />

30<br />

400000<br />

25<br />

c<br />

o 20<br />

A<br />

t 15<br />

a<br />

o 10<br />

U.<br />

200000<br />

s<br />

o.<br />

0<br />

o<br />

Effect <strong>of</strong> SUMO siRNA on p300 mediated repression.<br />

N+ N- & Si N- & S2+S3<br />

Figure 33. Relief <strong>of</strong> repression is preferentially mediated through the siRNA targeting <strong>of</strong><br />

SUMO-2 and SUMO-3.2 µg <strong>of</strong>'the luciferase reporter construct were translected along with<br />

the indicated plasmid 5no N+/N- (Figure l6a). Additionally transfections were carried out<br />

with either pSUPER retro, empty vector (2 ug), pSUPER retro targeting SUMO-1 (I1. tg) with<br />

empty vector (I «g), or pSUPER retro targeting SUMO-2 (11t,, ) and SUMO-3 (l u") as<br />

indicated. Plasmids were transfected by calcium phosphate into U2-OS cells. Results <strong>of</strong> the<br />

lucif'erase assays are plotted both as raw R. L. U. readings and as fold activation compared to the<br />

plasmid without shRNA coexpression, and are the means <strong>of</strong> three separate experiments, and<br />

standard deviations are shown where applicable.


(I)<br />

(ii)<br />

1400000<br />

1200000<br />

1000000<br />

800000<br />

J<br />

Ci 600000<br />

400000<br />

200000<br />

18<br />

16<br />

14<br />

0 12<br />

> 10<br />

Q8<br />

V<br />

p6<br />

U.<br />

4<br />

2<br />

0 elk<br />

Effect <strong>of</strong> SUMO siRNA on Elk-1 mediated repression.<br />

0 F']<br />

elk elk & S1 elk & 52+53 elk k-r<br />

DNA<br />

Effect <strong>of</strong> SUMO siRNA on ELk-1 Fold Activation<br />

Liii<br />

0 elk & Si elk & S2+S3<br />

DNA<br />

O00<br />

elk k-r k-r & Si k-r & 52+S3<br />

Figure 34. Relief <strong>of</strong> repression is preferentially mediated through the siRNA targeting <strong>of</strong><br />

SUMO-2 and SUMO-3.2 µg <strong>of</strong> the luciferase reporter construct were transtected along with<br />

the indicated plasmid Elk- I/Elk- I K-R (Figure 16a). Additionally transfections were carried<br />

out with either pSU PER retro, empty vector (2 µg), pSUPER retro targeting SUMO-1 (l p-)<br />

with empty vector (I dug), or pSUPER retro targeting SUMO-2 (l µg) and SUMO-3 (lug) as<br />

indicated. Plasmids were translected by calcium phosphate into U2-OS cells. Results <strong>of</strong> the<br />

luciferase assays are plotted both as raw R. L. U. readings and as mold activation, compared to<br />

the plasmid without shRNA coexpression, and are the means <strong>of</strong> three separate experiments, and<br />

standard deviations are shown where applicable.<br />

k-r & Si k-r & S2+S3


C<br />

O<br />

O<br />

LL<br />

25<br />

20<br />

10<br />

0<br />

O<br />

Effect <strong>of</strong> SUMO siRNA on Spa mediated repression.<br />

Sp3<br />

Effect <strong>of</strong> SUMO siRNA on Sp3 Fold Activation.<br />

EL<br />

___<br />

Sp3 & S1 Sp3 & S2+S3<br />

DNA<br />

oQ<br />

K-R Si K-R S2+S3<br />

Figure 35. Relief <strong>of</strong> repression is preferentially mediated through the siRNA targeting <strong>of</strong><br />

SUMO-2 and SUMO-3.2 «g <strong>of</strong> the luciferase reporter construct were transfected along with<br />

the indicated plasnild 200ng Sp3/Sp3 K-R (Figure 16a). Additionally transfect ions were<br />

carried out with either pSUPER retro, empty vector (2 µg), pSUPER retro targeting SUMO-1<br />

(l pg) with empty vector (I pg), orpSUPER retro targeting SUMO-2 (I LIg) and SUMO-3 (lug)<br />

as indicated. Plasmids were transfected by calcium phosphate into U2-OS cells. Results <strong>of</strong> the<br />

luciferase assays are plotted both as raw R. L. U. readings and as fold activation compared to the<br />

plasmid without shRNA coexpression, and are the means <strong>of</strong> three separate experiments, and<br />

standard deviations are shown where applicable.<br />

Sp3 K-R


(1) 1<br />

(11)<br />

8000<br />

7000<br />

6000<br />

5000<br />

J 4000<br />

oc<br />

c8<br />

0<br />

LL<br />

3000<br />

2000<br />

1000<br />

12<br />

10<br />

6<br />

4<br />

2<br />

0<br />

0<br />

Effect <strong>of</strong> SUMO siRNA on basal Ga14 activity<br />

Ga14 Ga14 & Si<br />

DNA<br />

Effect <strong>of</strong> SUMO siRNA on Ga14 fold activation<br />

Ga14 Ga14 & Si<br />

DNA<br />

Ga14 &S2/S3<br />

Ga14 &S2/S3<br />

Figure 36. Ga14 is activated by siRNA targeting <strong>of</strong> SUMO-2 and SUMO-3.2 it,, <strong>of</strong> the<br />

luciferase reporter construct were transfected with 5ng <strong>of</strong> the Ga14 plasmid. Additionally<br />

transfections were carried out with either pSUPER retro, empty vector (2 [tg ), pSUPER retro<br />

targeting SUMO-1 (l. t(T) with empty vector (lug), or pSUPER retro targeting SUMO-2 (I 1-)<br />

and SUMO-3 (11.<br />

(g) as indicated. Plasmids were transfected by calcium phosphate<br />

into U2-OS<br />

cells. Results <strong>of</strong> the luciferase assays are plotted both as raw R. L. U. readings and as fold<br />

activation compared to the plasmid without shRNA coexpression and are the means <strong>of</strong> three<br />

separate experiments, and standard deviations are shown where applicable.


eing activated following the siRNA knockdown <strong>of</strong> SUMO, Ga14 constructs<br />

were transfected with either empty vector as a control, SUMO-1, or SUMO-2<br />

and SUMO-3. pSUPER retro plasmids targeting SUMO-1 resulted in a modest<br />

increase in Ga14 activation whilst those targeting SUMO-2 and SUMO-3 lead to<br />

a more substantial increase in Ga14 activation (Figure 36. i and ii) as observed for<br />

Ga14 p300 N-.<br />

3.3.3 The nuclear localisation <strong>of</strong> p80 coilin is differentially<br />

regulated by knockdown <strong>of</strong> SUMO-1 compared to that <strong>of</strong> SUMO-<br />

2/-3<br />

SUMO modification controls the nuclear localization <strong>of</strong> numerous<br />

proteins, most notably recruiting many to PML bodies. To investigate the<br />

differential roles potentially played by SUMO-1 and SUMO-2/3 in nuclear<br />

localization <strong>of</strong> SUMO substrates siRNA targeting <strong>of</strong> the SUMO was preformed.<br />

The effect <strong>of</strong> substrate localization following ablation <strong>of</strong> SUMO-1 or SUMO-2/3<br />

could then be monitored through imun<strong>of</strong>louresence.<br />

The Cajal (coiled) body is a discrete nuclear organelle that was first<br />

described in mammalian neurons in 1903. Because the molecular composition,<br />

structure, and function <strong>of</strong> Cajal bodies were unknown, these enigmatic structures<br />

were largely ignored for most <strong>of</strong> the last century. This was changed by the<br />

discovery <strong>of</strong> p80-coilin, which can be used as a protein marker for cajal bodies.<br />

However despite current widespread use <strong>of</strong> coilin to identify Cajal bodies in<br />

various cell types, its structure and function are still little understood. Previous<br />

work had identified that p80 coilin is SUMO modified in vivo (J. M. P. Desterro,<br />

personal communication). p80 coilin could not be SUMO modified in vitro,<br />

possibly due to the lack <strong>of</strong> a specific E3-like ligase, and no functional test<br />

existing for p80 coilin, thereby excluding any further work into the biological<br />

significance <strong>of</strong> this modification at that time.<br />

To further investigate p80 coilin SUMO modification in vivo, U2-OS<br />

cells transfected with the RSC control oligonucleotide showed coilin localized<br />

into either cajal bodies or as nuclear aggregates. Treatment <strong>of</strong> cells with SUMO-<br />

157


1 siRNA completely removed the appearance <strong>of</strong> coilin in cajal bodies, whilst<br />

treatment with SUMO-2/-3 siRNA left the cajal body coilin intact but completely<br />

removed the appearance <strong>of</strong> the aggregates (Figure 37. ii. ). PML localisation in<br />

contrast was not affected by SUMO siRNA, presumably due to the effectiveness<br />

<strong>of</strong> the knockdown. As SUMO levels were not completely knockout, PML,<br />

which is efficiently modified by SUMO may out compete many <strong>of</strong> the numerous<br />

SUMO substrates for the available SUMO (Figure 37. ii. ).<br />

158


Figure 37. Endogenous PML and p80 coilin both localise to different nuclear dot<br />

formations within the nucleus. The knockdown <strong>of</strong> SUMO has no effect on PML<br />

localisation, but dramatically influences p80 coilin distribution. Synthetic siRNA<br />

oligonucleotides (100nm) were transfected by calcium phosphate as described in<br />

materials and methods. (i) Endogenous PML was detected with anti-PML antibody,<br />

(ii) endogenous p80 coilin was detected with an anti-coilin antibody with FITC as the<br />

secondary, as indicated by the coloured text. White text denotes siRNA<br />

oligonucleotide transfected. Fluorescence microscopy was carried out with a<br />

DeltaVision microscope (Olympus IX70) with a 1.40 NA 100X objective and<br />

Photometric CH300 CCD camera. IOX 0.2 µm sections were taken. Images were<br />

deconvoluted using S<strong>of</strong>tWorx (Applied Precision) s<strong>of</strong>tware.


BEST COPY<br />

AVAILABLE<br />

Variable print quality


(i)<br />

(ii)


3.3.4 Conclusion<br />

In vertebrates the number <strong>of</strong> SUMO genes increases to three: SUMO-1,<br />

SUMO-2, SUMO-3 and numerous pseudogenes, as compared to lower<br />

eukaryotes. At the protein level the SUMO-2 and SUMO-3 are highly<br />

homologous (Kamitani et al., 1998), being 95% identical to each other following<br />

processing and are regarded as being functionally homologous. SUMO-1<br />

homology is significantly reduced compared to that <strong>of</strong><br />

SUMO-2/-3, being<br />

approximately 50% identical following processing. Furthermore SUMO-2/-3,<br />

unlike SUMO-1, are capable <strong>of</strong> forming polymeric SUMO chains due to the<br />

presence <strong>of</strong> internal SUMO modification consensus motifs iKxE (where 1 is a<br />

large hydrophobic residue and x can be any amino acid) at their N-terminal<br />

extensions. The consequences <strong>of</strong> these polymeric chains in vivo have yet to be<br />

established although their existence has been determined. The conjugation <strong>of</strong> all<br />

three SUMO species occurs with equal efficiency in the presence <strong>of</strong> El and E2 in<br />

vitro, however in vivo there are examples <strong>of</strong> substrates that are predominantly<br />

modified by one SUMO paralogue. RanGAP1 is preferentially modified by<br />

SUMO-1 while Left (Sachdev et al., 2001) is preferentially modified<br />

by SUMO-<br />

2. As cells progress through the cell cycle, topoisomerase II, is preferentially<br />

modified by SUMO-2/3 during mitosis (Azuma et al., 2003). It is as yet unclear<br />

how this specificity/control <strong>of</strong> conjugation is controlled.<br />

Targeting <strong>of</strong> SUMO-1 or SUMO-2/-3 by synthetic siRNA<br />

oligonucleotides specifically reduces the expression <strong>of</strong> targeted SUMO species,<br />

and does not result in the decreased levels <strong>of</strong> the SUMO species not targeted<br />

(Figures 30 and 31). The extent <strong>of</strong> the reduction <strong>of</strong> SUMO-1 is greater than that<br />

for the two other SUMO species SUMO-2 and SUMO-3. This result is not<br />

unexpected in that it has been reported previously that, whilst<br />

SUMO-1 has<br />

greater mRNA levels compared to those <strong>of</strong> SUMO-2 and SUMO-3, there are<br />

greater amounts <strong>of</strong> SUMO-2 and SUMO-3 at the protein level, than SUMO-1.<br />

Due to the high degree <strong>of</strong> identity between SUMO-2 and SUMO-3,95%<br />

following processing, the available antibody, raised against SUMO-3 (Sentrin-2)<br />

also recognizes SUMO-2. The reduction in protein levels <strong>of</strong> SUMO-1 and<br />

SUMO-2/-3 was also characterised by immun<strong>of</strong>luorescence in U-2 OS cells.<br />

Targeting <strong>of</strong> SUMO-1 resulted in the disappearance <strong>of</strong> SUMO-1 punctate dots,<br />

161


and a general reduction in SUMO-1 background staining, in the majority <strong>of</strong> cells<br />

transfected (Figure 31). The removal <strong>of</strong> SUMO-1 punctate dots was shown not<br />

to inhibit the formation <strong>of</strong> SUMO-2/-3 punctate dots, suggesting that SUMO-1 is<br />

not a requirement <strong>of</strong> SUMO-2/-3 localisation or accumulation (Figure 31).<br />

Furthermore knockdown <strong>of</strong> SUMO-1 levels did not prevent the formation <strong>of</strong><br />

PML bodies (Figure 37), as observed via fluorescence, although it has previously<br />

been noted that SUMO is not required for primary PML body formation, only<br />

secondary PML body formation (Lallemand-Breitenbach et al., 2001). The<br />

knockdown <strong>of</strong> SUMO-2 and SUMO-3 was also observed in U-2 OS cells,<br />

although this was achieved with a lower degree <strong>of</strong> efficiency than that <strong>of</strong> the<br />

SUMO-1 knockdown. The reduction <strong>of</strong> SUMO-2 and SUMO-3 resulted in the<br />

reduction in SUMO-2/-3 punctate dots, although this was not universal and<br />

presumably depended upon the transfection efficiency. As observed with<br />

SUMO-1, knockdown with SUMO-2 and SUMO-3 did not affect PML<br />

localisation.<br />

Many substrates <strong>of</strong> SUMO are either transcription factors or c<strong>of</strong>actors,<br />

and whilst not exclusively, predominantly have been shown to mediate<br />

transcriptional repression, following conjugation by SUMO. Three such<br />

examples are the transcription co-activator p300 (<strong>Girdwood</strong> et al., 2003), the<br />

ETS domain transcription factor Elk-1 (Yang et al., 2003), and the ubiquitous<br />

transcription factor Sp3 (Sapetschnig et al., 2002; Ross et al., 2002). To further<br />

expedite the characterisation <strong>of</strong> the roles played by SUMO-1 and those <strong>of</strong><br />

SUMO-2/-3 in transcriptional repression, pSUPER retro vectors were<br />

constructed for each <strong>of</strong> the SUMO species, and were used to investigate their<br />

roles in the repression <strong>of</strong> theses substrates. As such Ga14 fusions bearing both<br />

SUMO modifiable p300, ELK-1, and Sp3 and mutants either lacking the required<br />

domain or bearing lysine to alanine or arginine point mutations (as previously<br />

described in <strong>Girdwood</strong> et al., 2003; Yang et al., 2003; Ross et al., 2002; Figure<br />

16a) were co-transfected in the presence <strong>of</strong> empty pSUPER retro plasmid,<br />

pSUPER retro SUMO-1, pSUPER retro SUMO-2, or pSUPER retro SUMO-3 as<br />

indicated (Figures 33,34, and 35). The co-transfection <strong>of</strong> SUMO modifiable<br />

substrate with the pSUPER retro SUMO-1 plasmid resulted in a modest increase<br />

in transcription when compared to those co-transfected with only empty vector.<br />

162


This result was repeated when the pSUPER retro SUMO-1 plasmid was co-<br />

transfected with the SUMO modification deficient version <strong>of</strong> the Ga14 fusion<br />

substrates. The co-transfection assays were repeated, with pSUPER retro<br />

plasmids targeting SUMO-2 and SUMO-3, a substantial relief <strong>of</strong> repression was<br />

observed (Figures 33,34, and 35), although it was also observed with the SUMO<br />

modification deficient version <strong>of</strong> the Ga14 fusion substrates (Figures 32,33, and<br />

34). As these substrates were not being modified by SUMO, the observed<br />

increase in luciferase readings is likely due to an activation event rather than a<br />

relief <strong>of</strong> repression. In support <strong>of</strong> this theory, when the pSUPER retro plasmids<br />

were co-transfected with Ga14 a similar pattern <strong>of</strong> activation was noticed,<br />

indicating that SUMO may be having a repressive effect at the promoter level <strong>of</strong><br />

transcription (Figure 36). Whilst this work was in progress Holmstrom et al.,<br />

reported that SUMO-2 and SUMO-3 Ga14 fusions had a greater capacity to<br />

repress Ga14 activity than SUMO-1, although this was only seen to a modest<br />

degree (Holmstrom et al., 2003). Interestingly p80 coilin's nuclear localisation<br />

was affected differently depending on whether SUMO-1 or<br />

SUMO-2/-3 levels<br />

were being reduced. SUMO-1 knockdown prevented the localisation <strong>of</strong> p80<br />

coilin to cajal bodies, and instead resulted in nuclear aggregation. The<br />

knockdown <strong>of</strong> SUMO-2/-3 had no effect on coilin cajal body localisation, and<br />

indeed reduced the levels <strong>of</strong> the nuclear aggregates, and further suggests that<br />

SUMO-1 and SUMO-2/-3 are involved in differential roles, outside that <strong>of</strong><br />

transcriptional repression.<br />

163


A<br />

4. DISCUSSION<br />

164


4.1 Discussion<br />

SUMO-1/-2/-3 are key members <strong>of</strong> the growing family <strong>of</strong> protein<br />

modifiers, that covalently modify selected cellular substrates. The SUMO<br />

conjugation pathway proceeds through an enzymatic cascade that is analogous,<br />

though distinct from that <strong>of</strong> ubiquitin. These differences between the<br />

conjugation pathways <strong>of</strong> ubiquitin and SUMO are not at the mechanistic level;<br />

rather each pathway possessses distinct enzymes for their pathways.<br />

While both<br />

ubiquitin and SUMO conjugation pathways possess a sole El enzyme, the<br />

ubiquitin pathway contains around ten E2s whereas there appears to be only a<br />

single SUMO conjugating enzyme. There are reportedly around seven hundred<br />

ubiquitin E3 enzymes, 'whilst only six individual E3-like enzymes have been<br />

reported for the SUMO conjugation pathway to date. Prior to conjugation<br />

SUMO-1/-2/-3 are processed via actions <strong>of</strong> cysteine proteases to remove an<br />

inhibitory C-terminal tag, and exposes the characteristic di-glycine motif, and<br />

again distinct from that <strong>of</strong> the ubiquitin the same class <strong>of</strong> protease is capable <strong>of</strong><br />

both the initial processing and the deconjugation <strong>of</strong> SUMO from the target<br />

substrates. The conjugation <strong>of</strong> SUMO to a substrate usually requires the<br />

presence <strong>of</strong> a SUMO consensus motif, although there are a growing number <strong>of</strong><br />

substrates that lack this consensus motif, and potentially, only through use <strong>of</strong><br />

first principal proteomic approaches will the comprehensive list <strong>of</strong> SUMO<br />

substrates be identified. A large number <strong>of</strong> previously characterised repression<br />

domains, contain SUMO modification motifs, and have subsequently been<br />

shown to be SUMO substrates (Figure 38).<br />

The ability <strong>of</strong> SUMO to recruit HDAC6 following its conjugation to p300<br />

is the first proven example <strong>of</strong> the long held theory in the SUMO field, that the<br />

conjugation <strong>of</strong> SUMO provides a novel interaction surface, one capable <strong>of</strong><br />

recruiting additional proteins. SUMO has been shown to enhance the<br />

recruitment <strong>of</strong> target proteins into repressive domains, such as PML bodies. This<br />

newly identified second mechanism for SUMO to mediate repression would<br />

appear to be due to the ability <strong>of</strong> SUMO to impart repressive properties upon<br />

target proteins. Following the identification <strong>of</strong> SUMO dependent recruitment <strong>of</strong><br />

HDAC6 to p300, a further example <strong>of</strong> SUMO dependent HDAC recruitment has<br />

been identified. SUMO mediated repression <strong>of</strong> the Elk-1 transcription factor had<br />

165


(i) Modification <strong>of</strong> histone deacetylases.<br />

Catalytic domain<br />

(ii) SUMO dependent repression.<br />

HLH<br />

Catalytic domain<br />

Gln Gln C2H2<br />

ADI ADII bZIP<br />

ETS BRDC<br />

C/HI CRD BROMO C/H2 C/H3<br />

DBD LZ<br />

hHDAC1<br />

hHDAC4<br />

h c-Myb<br />

hSp3<br />

hC/EBPa<br />

r//, 1111A h c-Jun<br />

hELK<br />

hp300<br />

KAKRVKTEDEKEK45'<br />

EAKGVKEEVKLA482<br />

AGVGVKGEPIESD566<br />

LLKKIKQEVESPT530<br />

AIDRIKEEEPDFE546<br />

RPLVIKQEPREED166<br />

RLQALKKEPQTVP233<br />

EAPNLKSEELNV136<br />

LPPEVKVEGPKE256<br />

TELKTEIKEEEDQ'°29<br />

Figure 38. SUMO conjugation and transcription repression. SUMO conjugation<br />

has been<br />

mapped to a number <strong>of</strong> previously identified repression domains, <strong>of</strong> transcription factors and<br />

co-factors.<br />

Solid vertical bars indicate SUMO conjugation motifs, horizontal lines indicate previously<br />

identified repression domains. HLH, helix-loop-helix; Gln, rich in glutamine; AD, activation<br />

domain; bZIP, basic zipper; LZ, leucine zipper C/H, cysteine-histide rich domains; Bromo,<br />

bromodomain; ETS, B, R, D, and C represent designated regions <strong>of</strong> ELK-l; C/EBP,<br />

CCAAT/enhancer-binding protein. h, human.


previously been attributed to SUMO conjugation to R domain <strong>of</strong> Elk-1, a domain<br />

that contains two SUMO conjugation sites, and subsequently the repression was<br />

shown to be by SUMO recruitment <strong>of</strong> HDAC2 (Yang and Sharrocks, 2004).<br />

Thus the recruitment <strong>of</strong> HDACs, by SUMO-modified proteins would lead to the<br />

alteration in the local balance between acetylases and deacetylases, increasing<br />

the preference for the deacetylation <strong>of</strong> chromatin, associated with states <strong>of</strong><br />

transcriptional repression. Sharrocks and colleagues further identified that<br />

HDAC3 also has a role in the repression <strong>of</strong> p300, suggesting that p300 may<br />

recruit both HDACs (Yang and Sharrocks, 2004). That p300 and Elk-1 are able<br />

to recruit different HDACs following their modification by SUMO demonstrates<br />

that the substrate <strong>of</strong> the SUMO modification also plays a key role in determining<br />

the specificity <strong>of</strong> HDAC recruited, presumably by the overall tertiary fold <strong>of</strong> the<br />

substrate. The importance <strong>of</strong> the substrate in HDAC recruitment was further<br />

demonstrated by the observation that GAL-SUMO fusion proteins activity are<br />

not affected by the siRNA targeting <strong>of</strong> the HDACs, demonstrating that<br />

unconjugated SUMO is unable to recruit HDACs to the promoters. Furthermore<br />

the ability <strong>of</strong> the protease to control SUMO mediated transcription repression<br />

also provides strong evidence that unconjugated SUMO is not able to mediate<br />

transcriptional repression. The exact mechanisms as to how SUMO directs these<br />

different repressive functions still remains to be elucidated. Further questions<br />

remaining to be resolved include the mechanisms by which SUMO-2/-3<br />

preferentially mediate transcriptional repression, although this could be<br />

explained by the relative abundance <strong>of</strong> the SUMO-2/-3 is<strong>of</strong>orms compared to<br />

that <strong>of</strong> SUMO-1, and that a significant proportion <strong>of</strong> SUMO-1 is conjugated to<br />

RanGAP1, thereby limiting its availability further.<br />

A common feature <strong>of</strong> all ubl-proteins that have been functionally<br />

characterised to date, with the exception <strong>of</strong> Atgl2, is that only a small proportion<br />

<strong>of</strong> a substrate is modified at any one time. Although only relatively low levels <strong>of</strong><br />

SUMO-modified substrates are observed, there can be dramatic consequences <strong>of</strong><br />

SUMO conjugation/deconjugation on the transcriptional activities <strong>of</strong> a number <strong>of</strong><br />

transcription factors, including those <strong>of</strong> p300, Elk-1, and Sp3. Thus far few<br />

theories have been put forward to try and explain this apparent paradox. One<br />

hypothesis on this matter, (<strong>Girdwood</strong> et al, 2004) draws on analogies from a<br />

167


wide range <strong>of</strong> biological systems, in which specific proteins are required for- the<br />

assembly, but not maintenance <strong>of</strong> multiprotein complexes. The model for<br />

SUMO proposes that modification <strong>of</strong> a substrate may result in the recruitment <strong>of</strong><br />

chaperonins that assemble the modified protein into a stable complex. Once<br />

formed, the chaperonins can dissociate and SUMO can be deconjugated leaving<br />

the once modified protein locked into the repressed state. A good example <strong>of</strong><br />

this, state A to state B transition comes from the field <strong>of</strong> DNA replication. In<br />

DNA replication, helicase loaders assemble monomeric helicase units into the<br />

ring shaped hexameric replicative helicases that encircle the naked<br />

DNA. The<br />

loading <strong>of</strong> the monomeric helicase units by the helicase loaders is ATP<br />

dependent, but once loaded they remain loaded following the hydrolysis <strong>of</strong> ATP.<br />

Applying this model to that <strong>of</strong> SUMO conjugation and transcriptional repression,<br />

Hay proposes that following the ATP dependent SUMO conjugation to a<br />

substrate there follows a rapid SUMO-dependent incorporation into a repression<br />

complex, a complex that will remain intact following the deconjugation <strong>of</strong><br />

SUMO, and that will remain active in transcriptional repression (<strong>Girdwood</strong> et al,<br />

2004). The eventual break up <strong>of</strong> the complex and hence loss <strong>of</strong> transcriptional<br />

repression will be a more gradual event, far slower than the incorporation, and<br />

thus explain why despite the low levels <strong>of</strong> SUMO conjugated substrates are<br />

observed, the consequences are biologically significant (Figure 39) (<strong>Girdwood</strong> et<br />

al, 2004). Following the transition from active to repressed state, there must be<br />

mechanisms to revert this state and relieve repression. In principle<br />

transcriptional activation could be accomplished by signal-induced modification<br />

<strong>of</strong> the repressed transcription factor, such as its phosphorylation or acetylation,<br />

which in turn mediates disassembly <strong>of</strong> the complex and the release <strong>of</strong> the active<br />

transcription factor. Variation <strong>of</strong> the activation pathway can be envisaged in<br />

which signal-induced modification <strong>of</strong> the transcription could lead to enhanced<br />

SUMO deconjugation and dissociation <strong>of</strong> the complex. Signal-induced relief <strong>of</strong><br />

repression is evident in the case <strong>of</strong> Elk-1 where signalling through the MAP<br />

kinase pathway results in phosphorylation <strong>of</strong> Elk-1 with concomitant loss <strong>of</strong><br />

SUMO modification and activation <strong>of</strong> Elk-l-dependent transcription (Yang et al,<br />

2003). In this scenario Elk-1 phosphorylation could have multiple consequences,<br />

and could act to recruit SUMO-specific proteases and induce conformational<br />

168


Active state<br />

INIX NF-X<br />

SUMOylatio<br />

FAST<br />

1NIF-X<br />

i<br />

(U-M0<br />

SUMO-dependent<br />

Incorporation into<br />

Repression comply<br />

(FAST)<br />

i Release from<br />

Repression comple<br />

ý<br />

ý (SLOW)<br />

i<br />

i<br />

Repressed state<br />

Figure 39. Model for transcriptional repression by SUMO. See text for details.<br />

Figure courtesy <strong>of</strong> Ron Hay.<br />

SUMO-independent'<br />

Retention in<br />

repression comple


changes in Elk-1 that allow release from the "repression complex".. In the case<br />

<strong>of</strong> p300, expression <strong>of</strong> the cyclin-dependent kinase inhibitor p21 relieves p300-<br />

dependent repression without affecting the level <strong>of</strong> SUMO-modified p300<br />

(Figure 28). This could be explained if p21 simply induced the disassembly <strong>of</strong><br />

the repression complex containing predominantly unmodified p300. Indeed this<br />

hypothetical model can be directly applied to other SUMO dependent events,<br />

such as NEMOs recruitment to the ATM scaffold, and subsequent release<br />

following ubiquitination (Huang et al, 2003; Hay, 2004).<br />

v<br />

Due to the many similarities shared between the various ubl-proteins,<br />

research into SUMO is likely to provide insights into the functioning <strong>of</strong> the other<br />

ubl-proteins, with the techniques developed to study SUMO being directly<br />

applied to the other members <strong>of</strong> the superfamily <strong>of</strong> polypeptide modifiers. This<br />

research into the field <strong>of</strong> SUMO will undoubtedly identify novel substrates for<br />

SUMO modification, and likely provide further examples <strong>of</strong> the importance <strong>of</strong><br />

SUMO modification in the regulation <strong>of</strong> transcriptional activity, protein stability,<br />

sub-nuclear domain organisation, and competition amongst the other post-<br />

translational modifiers for lysine residues. Indeed a proteomic approach to<br />

identify ubiquitinated substrate was recently undertaken in S. cerevisiae. Four <strong>of</strong><br />

the identified peptides contained an ubiquitin conjugating lysine residue, in a<br />

classic SUMO conjugation motif (Table 4) (Peng et al, 2003). Two peptides<br />

came from uncharacterised proteins, YGR268C and YOL109W. The remaining<br />

two were identified in an inositol transporter protein, ITR1, and interestingly the<br />

fourth belonged to UIP3, an Ulpl interacting protein. Although research into the<br />

SUMO field <strong>of</strong> post-translation modification has answered many <strong>of</strong> the initial<br />

questions regarding its function and similarity to ubiquitin, both in conjugation<br />

and regulation. Future research will undoubtedly answer many more questions<br />

and help determine the physiological role <strong>of</strong> SUMO and the role played by<br />

SUMO in the development <strong>of</strong> diverse processes associated with cancer,<br />

neurodegenerative conditions and viral infections.<br />

170


Protein Peptide sequence<br />

YGR268C DTHDDELPSYEDVIKEEER<br />

YOL109W EQAEASIDNLKNEATPEAEQVK<br />

ITR 1 VHELKYEPTQEIIDI<br />

UIP3 MQTPSENTDVKLDTLDEPSAH<br />

Table 4. Proteins with identified ubiquitination sites from S. cerevisiae. Table<br />

showing the identified ubiquitin conjugated lysine residues (in bold and larger size)<br />

and potential SUMO modification sites (underlined).


5. Appendix I Vector Maps<br />

172


Cloning primers:<br />

Pvul<br />

NEOr<br />

B<br />

AMPr P CMV<br />

pcDNA3 Cullin-2<br />

7683bp<br />

B. GH<br />

40 on<br />

amHl<br />

Cu12<br />

Tthl11I vpp- 1%<br />

EcoRl<br />

Smal<br />

Upstream (BamHl): 5'-C GAA GGA TCC ATG TCT TTG AAA CCA AGA GTA GTA G-3'<br />

Downstream (EcoRl): 5'-T CAG GAA TTC ACG CGA CGT AGC TGT ATT C-3'<br />

CUL2:<br />

Subunit <strong>of</strong> the pVHL ubiquitin E3 ligase complex<br />

Organism: Homo sapiens<br />

Coding region: 2283bp<br />

GenBank accession number: NP_003582<br />

Notes:<br />

Modified by the ubl-protein NEDD8.<br />

173


Bss<br />

Mlu 1<br />

Oligos annealed:<br />

Upper (BamHl): 5'-GA TCC TCT AGA CCC GGG TCG ACT AGA TCT G-3'<br />

Lower (EcoRl): 5-TT AAC AGA TCT AGT CGA CCC GGG TCT AGA G-3'<br />

Notes:<br />

Introduces Xbal and BgIII cloning sites.<br />

Ban-Hl<br />

Xbal<br />

Bglll<br />

Smal<br />

EcoRl<br />

thIII l<br />

w11<br />

174


Mlu l<br />

Bss<br />

Oligos annealed:<br />

BamH 1<br />

Xbal<br />

Bgll1<br />

Smal<br />

EcoRl<br />

l<br />

ýat11<br />

Upper (Xbal): 5'-CT AGA ACT GAG TTA AAA ACT GAA ATA AAA GAG GAG GAA<br />

GAC CAG TGA A-3'<br />

Lower (Bg1II): 5-GA TCT TCA CTG GTC TIC CTC CTC TIT TAT TTC AGT TTT TAA<br />

CTC AGT T-3'<br />

Amino acid Sequence:<br />

TELKTEIKEEEDQ<br />

SUMO conjugation status:<br />

Two sites, full modification<br />

175


Mlu l<br />

Bss<br />

Oligos annealed:<br />

BamHl<br />

Xbal<br />

Bglll<br />

Smal<br />

EcoRl<br />

t ml<br />

kat11<br />

Upper (Xbal): 5'-CT AGA ACT GCC TTA AAA ACT GAA ATA AAA GAG GAG GAA<br />

GAC CAG TGA A-3'<br />

Lower (Bg1II): 5-GA TCT TCA CTG GTC TTC CTC CTC TTT TAT TTC AGT TTT TAA<br />

GGC AGT T-3'<br />

Amino acid Sequence:<br />

TALKTEIKEEEDQ<br />

SUMO conjugation status:<br />

Two sites, full modification<br />

176


Mlu l<br />

Bss<br />

Oligos annealed:<br />

BamHl<br />

Xbal<br />

Bg111<br />

Smal<br />

EcoRl<br />

LhIHI<br />

Upper (Xbal): 5'-CT AGA ACT GAG GCC AAA ACT GAA ATA AAA GAG GAG GAA<br />

GAC CAG TGA A-3'<br />

Lower (Bg1II): 5-GA TCT TCA CTG GTC TTC CTC CTC TTT TAT TTC AGT TTT GGC<br />

CTC AGT T-3'<br />

Amino acid Sequence:<br />

TEAKTEIKEEEDQ<br />

SUMO conjugation status:<br />

One site, partial modification<br />

w11<br />

177


Mlul<br />

Bss<br />

Oligos annealed:<br />

BamHl<br />

Xbal<br />

Bglll<br />

Smal<br />

EcoR1<br />

Upper (Xbal): 5'-CT AGA ACT GAG TTA AAA ACT GAA ATA AAA GAG GAG GAA GAC<br />

CAG TGA A-3'<br />

Lower (Bg1II): 5-GA TCT TCA CTG GTC TTC CTC CTC TTT TAT TTC AGT TTT TAA<br />

CTC AGT T-3'<br />

Amino acid Sequence:<br />

TELKTEIKEEEDQ<br />

SUMO conjugation status:<br />

Two sites, full modification<br />

thIIII<br />

w11<br />

178


Bss<br />

Mlu 1<br />

Oligos annealed:<br />

BamHi<br />

Xbal<br />

Bgll1<br />

Smal<br />

EcoRl<br />

thIII I<br />

kat11<br />

Upper (Xbal): 5'-CT AGA ACT GAG TTA AAA GCC GAA ATA AAA GAG GAG GAA<br />

GAC CAG TGA A-3'<br />

Lower (Bg1II): 5-GA TCT TCA CTG GTC TTC CTC CTC TTT TAT TTC GGC TTT TAA<br />

CTC AGT T-3'<br />

Amino acid Sequence:<br />

TELKAEIKEEEDQ<br />

SUMO conjugation status:<br />

Two sites, full modification<br />

179


Mlul<br />

Bss<br />

Oligos annealed:<br />

BamHl<br />

Xbal<br />

BgIll<br />

Smal<br />

EcoRl<br />

thM 1<br />

kat11<br />

Upper (Xbal): 5'-CT AGA ACT GAG TTA AAA ACT GCC ATA AAA GAG GAG GAA<br />

GAC CAG TGA A-3'<br />

Lower (Bg1II): 5-GA TCT TCA CTG GTC TTC CTC CTC TTT TAT TTC GGC TTT TAA<br />

CTC AGT T-3'<br />

Amino acid Sequence:<br />

TELKTAIKEEEDQ<br />

SUMO conjugation status:<br />

One site, partial modification<br />

180


Mlul<br />

Bss<br />

Oligos annealed:<br />

BamHl<br />

Xbal<br />

BgIll<br />

Smal<br />

EcoRl<br />

thIII 1<br />

Upper (Xbal): 5'-CT AGA ACT GAG TTA AAA ACT GAA GCC AAA GAG GAG GAA<br />

GAC CAG TGA A-3'<br />

Lower (Bg1II): 5-GA TCT TCA CTG GTC TTC CTC CTC TTT GGC TTC AGT TTT TAA<br />

CTC AGT T-3'<br />

Amino acid Sequence:<br />

TELKTEAKEEEDQ<br />

SUMO conjugation status:<br />

One site, partial modification<br />

Ut11<br />

181


Mlu l<br />

Bss<br />

Oligos annealed:<br />

BamH1<br />

Xbal<br />

Bgll1<br />

Smal<br />

EcoRl<br />

Upper (Xbal): 5'-CT AGA ACT GAG TTA AAA ACT GAA ATA GCC GAG GAG GAA<br />

GAC CAG TGA A-3'<br />

Lower (Bg1II): 5-GA TCT TCA CTG GTC TTC CTC CTC GGC TAT TTC AGT TTT TAA<br />

CTC AGT T-3'<br />

Amino acid Sequence:<br />

TELKTEIAEEEDQ<br />

SUMO conjugation status:<br />

One site, partial modification<br />

ýl<br />

thil<br />

w11<br />

182


Mlul<br />

Bss<br />

Oligos annealed:<br />

BamHl<br />

Xbal<br />

BgIll<br />

Smal<br />

EcoRl<br />

Upper (Xbal): 5'-CT AGA ACT GAG TTA AAA ACT GAA ATA AAA GCC GAG GAA<br />

GAC CAG TGA A-3'<br />

Lower (Bg1II): 5-GA TCT TCA CTG GTC TTC CTC GGC TTT TAT TTC AGT TTT TAA<br />

CTC AGT T-3'<br />

Amino acid Sequence:<br />

TELKTEIKAEEDQ<br />

SUMO conjugation status:<br />

One site, partial modification<br />

sui<br />

w11<br />

183


Mlu l<br />

Bss<br />

Oligos annealed:<br />

BamHl<br />

Xbal<br />

BgIl1<br />

Smal<br />

EcoRl<br />

thIII l<br />

ýat11<br />

Upper (XbaI): 5'-CT AGA ACT GAG TTA AAA ACT GAA ATA AAA GAG GCC GAA<br />

GAC CAG TGA A-3'<br />

Lower (BgIII): 5-GA TCT TCA CTG GTC TTC GGC CTC TTT TAT TTC AGT TTT TAA<br />

CTC AGT T-3'<br />

Amino acid Sequence:<br />

TELKTEIKEAEDQ<br />

SUMO conjugation status:<br />

One site, partial modification<br />

184


Miul<br />

ßm<br />

Oligos annealed:<br />

BamH 1<br />

Xbal<br />

Bgl ll<br />

Smal<br />

EcoRl<br />

hm 1<br />

Upper (Xba 1): 5'-Q ACA ACT GAG TTA AAA ACT GAA ATA AAA GAG GAG GCC<br />

GACCAGTGAA-3'<br />

Lowcr (Bg1I<br />

I): 5-GA Jý, I' TCA CTG GTC GGC CTC CTC M TAT TTC AGT TTT TAA<br />

CTC AGT T-3'<br />

Amino acid Sequence:<br />

TELKTEIKEEQDQ<br />

SUMO conjugation status:<br />

Two sites, full modification<br />

, at11<br />

185<br />

ýý<br />

ý,


Mlul<br />

BssHl l<br />

Oligos annealed:<br />

Narl<br />

lacIq<br />

tacp<br />

pGEX2T TEV K1020A/K1024A<br />

ORI<br />

4985bp<br />

GST s<br />

Ampr<br />

BamHl<br />

Xbal<br />

- Bglll<br />

Smal<br />

EcoRl<br />

TthIIIl<br />

Aatl 1<br />

Upper (Xbal): 5'-CT AGA ACT GAG TTA GCC ACT GAA ATA GCC GAG GAG GAA<br />

GAC CAG TGA A-3'<br />

Lower (BgIII): 5-GA TCT TCA CTG GTC TTC CTC CTC GGC TAT TTC AGT GGC TAA<br />

CTC AGT T-3'<br />

Amino acid Sequence:<br />

TELATEIAEEEDQ<br />

SUMO conjugation status:<br />

No sites, no modification<br />

186


Mlu l<br />

BssHll<br />

Oligos annealed:<br />

Narl<br />

lacIq<br />

tacp<br />

GST<br />

pGEX2T TEV K1020R/K1024R<br />

oRI<br />

4985bp<br />

Ampr<br />

BanH l<br />

Xbal<br />

- Bglll<br />

Smal<br />

EcoR l<br />

TthIII 1<br />

Aatl l<br />

Upper (Xbal): 5'-CT AGA ACT GAG TTA AGG ACT GAA ATA AGG GAG GAG GAA<br />

GAC CAG TGA A-3'<br />

Lower (Bg1II): 5-GA TCT TCA CTG GTC TTC CTC CTC CCT TAT TTC AGT CCTTAA<br />

CTC AGT T-3'<br />

Amino acid Sequence:<br />

TELRTEIREEEDQ<br />

SUMO conjugation status:<br />

No sites, no modification<br />

187


Mlul<br />

Bss<br />

Oligos annealed:<br />

BamH1<br />

Xbal<br />

Bgll1<br />

Smal<br />

EcoR 1<br />

Upper (Xbal): 5'-CT AGA CCC AGG GTG ATC AAG ATG GAG GTG AAG AGG GAG-3'<br />

Lower (Bg1II): 5-GA TCT TCA CTC CCT CTT CAC CTC CAT CTT GAT CAC CTT CCT<br />

GGG T-3'<br />

Amino acid Sequence:<br />

PRLVIKMEVKRE<br />

SUMO conjugation status:<br />

Two sites, full modification<br />

aiml<br />

Ut11<br />

188


pUC on<br />

B3pw III<br />

Anip<br />

Empty Vector<br />

1-1 ow<br />

$gill sbtiCI<br />

ý1 7000<br />

pFUP Bi. rrtm<br />

fw/Ptrfier]<br />

7195 bp<br />

%RI PLW<br />

Fpulo<br />

Purchased from Oligoengine.<br />

Sequencing primers:<br />

1000<br />

ýaooý<br />

Prn<br />

Dra III 9ecII BsiWI<br />

SLANT<br />

DgG IAwl<br />

I-h Prams B. u<br />

PGK Age[<br />

s ,i<br />

spulü21<br />

5'-GGAAGCCTTGGCTTTTG-3' binding site: 1241-1257<br />

5'-GATGACGTCAGCGTTCG-3' binding site: 2645-2629<br />

I<br />

! %II<br />

Hindill<br />

Dam HI<br />

I LORI<br />

Not<br />

Ptnel<br />

Nool<br />

ßäi<br />

189


SUMO-1<br />

cl<br />

pUC «i<br />

Armp<br />

_, -'sow<br />

AI BM I<br />

70W<br />

I5<br />

pFUPBi. wtm<br />

/<br />

loon<br />

SUMO-1 2xo--<br />

4ao<br />

BrG ICI<br />

51buflor<br />

HIM ai IET BýII<br />

BIPW III 6aoRI<br />

, pl<br />

CIO<br />

Pur<br />

PGK<br />

Para<br />

BM)<br />

BSI<br />

Dra III CI I g5iwl<br />

siRNA oligonucleotides (upper strand only):<br />

Axel<br />

Bpulü2l<br />

ii<br />

Hinall<br />

Bam Hl<br />

E n'<br />

NotI<br />

Pmd<br />

GATCCCCCTGGGAATGGAGGAAGAAGTTCAAGAGACTTCTTCCTCCATTCCCAGTTT<br />

TTGGAAA<br />

Notes:<br />

Oligonucleotide replaces the stuffer sequence indicated in the plasmid map.<br />

Sequencing primers:<br />

5' -GGAAGCCTTGGCTTTTG-3'<br />

5'-GATGACGTCAGCGTTCG-3'<br />

binding site: 1241-1257<br />

binding site: 2645-2629<br />

rkl<br />

190


SUMO-2<br />

9acl<br />

pJC«i<br />

SgAl Mci<br />

II<br />

70M<br />

SUMO-2 anaaýýý<br />

oaa<br />

BxG ICI<br />

gel l<br />

H r<br />

Hl<br />

6ccRI<br />

1044 Not I<br />

MPrams<br />

Bepw ul 6MAI<br />

$bPI PLU<br />

PGK<br />

Agel<br />

I4<br />

BMnI<br />

Posr11 ßI<br />

Dra III ecl I Bti l<br />

siRNA oligonucleotides (upper strand only):<br />

BPUIID21<br />

\-_-PmeI<br />

GATCCCCGATCAAGAGGCACACGTCGTTCAAGAGACGACGTGTGCCTCTTGATCTTT<br />

TTGGAAA<br />

Notes:<br />

Oligonucleotide replaces the stuffer sequence indicated in the plasmid map.<br />

Sequencing primers:<br />

5'-GGAAGCCTTGGCTTTTG-3'<br />

5'-GATGACGTCAGCGTTCG-3'<br />

binding site: 1241-1257<br />

binding site: 2645-2629<br />

191


SUMO-3<br />

Sgßi BE,<br />

c i<br />

&xGl4acl<br />

siI sell<br />

Hindill<br />

Anip Bam Hl<br />

6caRl<br />

1o Nct I<br />

Pmcl<br />

pUC on<br />

OFUPER. ICtlO <strong>St</strong>ufk! T<br />

Naal<br />

SUMO-3 B<br />

aaaa<br />

3000<br />

Hl<br />

MIET Bcj11<br />

B6pw III 6aoRl<br />

%pI Pur<br />

PGK<br />

Age[<br />

Ppulo<br />

Posrl<br />

BMm I<br />

gel<br />

pra III %cl NMI<br />

siRNA oligonucleotides (upper strand only):<br />

Bpulü2l<br />

GATCCCCCGACAGGGATTGTCAATGATTCAAGAGATCATTGACAATCCCTGTCGTTT<br />

TTGGAAA.<br />

Notes:<br />

Oligonucleotide replaces the stuffer sequence indicated in the plasmid map.<br />

Sequencing primers:<br />

5' -GG AAGCCTTGGCTTTTG-3' binding site: 1241-1257<br />

5' -GATGACGTCAGCGTTCG-3'<br />

binding site: 2645-2629<br />

192


6. Bibliography<br />

193


Adams, J. (2002). Proteasome inhibitors as new anticancer drugs. Curr Opin Oncol<br />

14,628-634.<br />

Adamson, A. L., and Kenney, S. (2001). Epstein-barr virus immediate-early protein<br />

BZLF1 is SUMO-1 modified and disrupts promyelocytic leukemia bodies. J Virol<br />

75,2388-2399.<br />

Ahn, J. H., Xu, Y., Jang, W. J., Matunis, M. J., and Hayward, G. S. (2001).<br />

Evaluation <strong>of</strong> interactions <strong>of</strong> human cytomegalovirus immediate-early IE2<br />

regulatory protein with small ubiquitin-like modifiers and their conjugation enzyme<br />

Ubc9. J Virol 75,3859-3872.<br />

Ait-Si-Ali, S., Polesskaya, A., Filleur, S., Ferreira, R., Duquet, A., Robin, P.,<br />

Vervish, A., Trouche, D., Cabon, F., and Harel-Bellan, A. (2000). CBP/p300 histone<br />

acetyl-transferase activity is important for the G1/S transition. Oncogene 19,2430-<br />

2437.<br />

Ait-Si-Ali, S., Ramirez, S., Barre, F. X., Dkhissi, F., Magnaghi-Jaulin, L., Girault, J.<br />

A., Robin, P., Knibiehler, M., Pritchard, L. L., Ducommun, B., et al. (1998). Histone<br />

acetyltransferase activity <strong>of</strong> CBP is controlled by cycle-dependent kinases and<br />

oncoprotein E1A. Nature 396,184-186.<br />

Alcalay, M., Tomassoni, L., Colombo, E., <strong>St</strong>oldt, S., Grignani, F., Fagioli, M.,<br />

Szekely, L., Helin, K., and Pelicci, P. G. (1998). The promyelocytic leukemia gene<br />

product (PML) forms stable complexes with the retinoblastoma protein. Mol Cell<br />

Biol. 18,1084-1093.<br />

Andres, G., Alejo, A., Simon-Mateo, C., and Salas, M. L. (2001). African swine<br />

fever virus protease, a new viral member <strong>of</strong> the SUMO-1-specific protease family. J<br />

Biol Chem 276,780-787.<br />

194


Anton, L. C., Schubert, U., Bacik, I., Princiotta, M. F., Wearsch, P. A., Gibbs, J.,<br />

Day, P. M., Realini, C., Rechsteiner, M. C., Bennink, J. R., and Yewdell, J. W.<br />

(1999). Intracellular localization <strong>of</strong> proteasomal degradation <strong>of</strong> a viral antigen.<br />

Biol 146,113-124.<br />

Ascoli, C. A., and Maul, G. G. (1991). Identification <strong>of</strong> a novel nuclear<br />

Cell Biol 112,785-795.<br />

Azuma, Y., Arnaoutov, A., and Dasso, M. (2003). SUMO-2/3 regulates<br />

topoisomerase II in mitosis. J Cell Biol 163,477-487.<br />

domain. J<br />

Bach, I., and Ostendorff, H, P,. (2003). Orchestrating nuclear functions: ubiquitin<br />

sets the rhythm. Trends Biochem Sci 4,189-195.<br />

Bachant, J., Alcasabas, A., Blat, Y., Kleckner, N., and Elledge, S. J. (2002). The<br />

SUMO-1 isopeptidase Smt4 is linked to centromeric cohesion through SUMO-1<br />

modification <strong>of</strong> DNA topoisomerase II. Mol Cell 9,1169-1182.<br />

Bailey, D., and O'Hare, P.. (2004). Characterization <strong>of</strong> the' localization and<br />

J Cell<br />

proteolytic activity <strong>of</strong> the SUMO-specific protease, SENP1. J Biol Chem 279,692-<br />

703.<br />

Baldi, L., Brown, K., Franzoso, G., and Siebenlist, U. (1996). Critical role for<br />

lysines 21 and 22 in signal-induced, ubiquitin-mediated proteolysis <strong>of</strong> I kappa B-<br />

alpha. J Biol Chem 271,376-379.<br />

Bates, E. E., Ravel, 0., Dieu, M. C., Ho, S., Guret, C., Bridon, J. M., Ait-Yahia, S.,<br />

Briere, F., Caux, C., Banchereau, J., and Lebecque, S. (1997). Identification and<br />

analysis <strong>of</strong> a novel member <strong>of</strong> the ubiquitin family expressed in dendritic cells and<br />

mature B cells. Eur J Immunol 27,2471-2477.<br />

195


Bates, S., Phillips, A. C., Clark, P. A., <strong>St</strong>ott, F., Peters; G., Ludwig, R. L., and<br />

Vousden, K. H. (1998). p14ARF links the tumour suppressors RB and p53. Nature<br />

395,124-125.<br />

Baumeister, W., Walz, J., Zuhl, F., and Seemuller, E. (1998). The proteasome:<br />

paradigm <strong>of</strong> a self-compartmentalizing protease. Cell 92,367-380.<br />

Benanti, J. A., <strong>William</strong>s, D. K., Robinson, K. L., Ozer, H. L., and Galloway, D. A.<br />

(2002). Induction <strong>of</strong> extracellular matrix-remodeling genes by the senescence-<br />

associated protein APA-1. Mol Cell Biol 22,7385-7397.<br />

Bencsath, K. P., Podgorski, M. S., Pagala, V. R., Slaughter, C. A., and Schulman, B.<br />

A. (2002). Identification <strong>of</strong> a multifunctional binding site on Ubc9p required for<br />

Smt3p conjugation. J Biol Chem 277,47938-47945.<br />

Berger, S. L. (2002). Histone modifications in transcriptional regulation. Curr Opin<br />

Genet Dev 12,142-148.<br />

Bernier-Villamor, V., Sampson, D. A., Matunis, M. J., and Lima, C. D. (2002).<br />

<strong>St</strong>ructural basis for E2-mediated SUMO conjugation revealed by a complex between<br />

ubiquitin-conjugating enzyme Ubc9 and RanGAP1. Cell 108,345-356.<br />

Best, J. L., Ganiatsas, S., Agarwal, S., Changou, A., Salomoni, P., Shirihai, 0.,<br />

Meluh, P. B., Pandolfi, P. P., and Zon, L. I. (2002). SUMO-1 protease-1 regulates<br />

gene transcription through PML. Mol Cell 10,843-855.<br />

Bhattacharya, S., Eckner, R., Grossman, S., Oldread, E., Arany, Z., D'Andrea, A.,<br />

Livingston, D. M. (1996). Cooperation <strong>of</strong> STAT2 and p300/CBP in signalling<br />

induced by interferon-alpha. Nature 383,344-347.<br />

196


Bhaskar, V., Smith, M., and Courey, A. J. (2002). Conjugation <strong>of</strong> Smt3 to dorsal<br />

may potentiate the Drosophila immune response. Mol Cell Biol 22,492-504.<br />

Bies, J., Markus, J., and Wolff, L. (2002). Covalent attachment <strong>of</strong> the SUMO-1<br />

protein to the negative regulatory domain <strong>of</strong> the c-Myb transcription factor modifies<br />

its stability and transactivation capacity. J Biol Chem 277,8999-9009.<br />

Bisch<strong>of</strong>, 0., Kirsh, 0., Pearson, M., Itahana, K., Pelicci, P. G., and Dejean, A.<br />

(2002). Deconstructing PML-induced premature senescence. Embo J 21,3358-3369.<br />

Bisch<strong>of</strong>f, F. R., Klebe, C., Kretschmer, J., Wittingh<strong>of</strong>er, A., and Ponstingl, H.<br />

(1994). RanGAPI induces GTPase activity <strong>of</strong> nuclear Ras-related Ran. Proc Natl<br />

Acad Sci USA 91,2587-2591.<br />

Bloch, D. B., Chiche, J. D., Orth, D., de la Monte S. M., Rosenzweig, A., Bloch, K. D.<br />

(1999). <strong>St</strong>ructural and functional heterogeneity <strong>of</strong> nuclear bodies. Mol Cell Biol. 6,<br />

4423-4430.<br />

Bloom, J., Amador, V., Bartolini, F., DeMartino, G., and Pagano, M. (2003).<br />

Proteasome-mediated degradation <strong>of</strong> p21 via N-terminal ubiquitinylation. Cell 115,<br />

71-82.<br />

Boddy, M. N., Howe, K., Etkin, L. D., Solomon, E., and Freemont, P. S. (1996). PIC<br />

1, a novel ubiquitin-like protein which interacts with the PML component <strong>of</strong> a<br />

multiprotein complex that is disrupted in acute promyelocytic leukaemia. Oncogene<br />

13,971-982.<br />

Bohren, K. M., Nadkarni, V., Song, J. H., Gabbay, K. H., and Owerbach, D. (2004).<br />

A M55V polymorphism in a novel SUMO gene (SUMO-4) differentially activates<br />

heat shock transcriptional factors and is associated with susceptibility to type I<br />

diabetes mellitus. J Bio Chem 26,27233-27238.<br />

197


Boisvert, F. M., Kruhlak, M. J., Box, A. K., Hendzel, M. J., and Bazett-Jones, D. P.<br />

(2001). The transcription coactivator CBP is a dynamic component <strong>of</strong> the<br />

promyelocytic leukemia nuclear body. J Cell Biol 152,1099-1106.<br />

Borden, K. L. (2000). RING domains: master builders <strong>of</strong> molecular scaffolds? J Mol<br />

Biol 295,1103-1112.<br />

Boyd, J. M., Subramanian, T., Schaeper, U., La Regina, M., Bayley, S., and<br />

Chinnadurai, G. (1993). A region in the C-terminus <strong>of</strong> adenovirus 2/5 Ela protein is<br />

required for association with a cellular phosphoprotein and important for the<br />

negative modulation <strong>of</strong> T24-ras mediated transformation, tumorigenesis and<br />

metastasis. Embo J 12,469-478.<br />

Braun, H., Koop, R., Ertmer, A., Nacht, S., and Suske, G. (2001). Transcription<br />

factor Sp3 is regulated by acetylation. Nucleic Acids Res 29,4994-5000.<br />

Briggs, S. D., Xiao, T., Sun, Z. W., Caldwell, J. A., Shabanowitz, J., Hunt, D. F.,<br />

Allis, C. D., and <strong>St</strong>rahl, B. D. (2002). Gene silencing: trans-histone regulatory<br />

pathway in chromatin. Nature 418,498.<br />

Brock, H. W., and van Lohuizen, M. (2001). The Polycomb group--no longer an<br />

exclusive club? Curr Opin Genet Dev 11,175-181.<br />

Brown, M. T. (1995). Sequence similarities between the yeast chromosome<br />

segregation protein Mif2 and the mammalian centromere protein CENP-C. Gene<br />

160,111-116.<br />

Brown, M. T., Goetsch, L., and Hartwell, L. H. (1993). MIF2 is required for mitotic<br />

spindle integrity during anaphase spindle elongation in Saccharomyces<br />

cerevisiae. J<br />

Cell Bio1123,387-403.<br />

198


Bulimo, W. D., Miskin, J. E., and Dixon, L. K. (2000). An ARID family protein<br />

binds to the African swine fever virus encoded ubiquitin conjugating enzyme,<br />

UBCv1. FEBS Lett 471,17-22.<br />

Burch, T. J., and Haas, A. L. (1994). Site-directed mutagenesis <strong>of</strong> ubiquitin.<br />

Differential roles for arginine in the interaction with ubiquitin-activating enzyme.<br />

Biochemistry 33,7300-7308.<br />

Buschmann, T., Lerner, D., Lee, C. G., and Ronai, Z. (2001). The Mdm-2 amino<br />

terminus is required for Mdm2 binding and SUMO-1 conjugation by the E2 SUMO-<br />

1 conjugating enzyme Ubc9. J Biol Chem 276,40389-40395.<br />

Bylebyl, G. R., Belichenko, I., and Johnson, E. S. (2003). The SUMO isopeptidase<br />

U1p2 prevents accumulation <strong>of</strong> SUMO chains in yeast. J Biol Chem 278,44113-<br />

44120.<br />

Cao, T., Duprez, E., Borden, K. L., Freemont, P. S., and Etkin, L. D. (1998). Ret<br />

finger protein is a normal component <strong>of</strong> PML nuclear bodies and interacts directly<br />

with PML. J Cell Sci 111 (Pt 10), 1319-1329.<br />

Carlile, G. W., Tatton, W. G., and Borden, K. L. (1998). Demonstration <strong>of</strong> a RNA-<br />

dependent nuclear interaction between the promyelocytic leukaemia protein and<br />

glyceraldehyde-3-phosphate dehydrogenase. Biochem J 335 (Pt 3), 691-696.<br />

Carr, A. M. (2000). Cell cycle. Piecing together the p53 puzzle. Science 287,1765-<br />

1766.<br />

Chakrabarti, S. R., Sood, R., Nandi, S., and Nucifora, G. (2000). Posttranslational<br />

modification <strong>of</strong> TEL and TEL/AML1 by SUMO-1 and cell-cycle-dependent<br />

assembly into nuclear bodies. Proc Natl Acad Sci USA 97,13281-13285.<br />

199


Chan, H. M., and La Thangue, N. B. (2001). p300/CBP proteins: HATs for<br />

transcriptional bridges and scaffolds. J Cell Sci 114,2363-2373.<br />

Chen, L., and Chen, J. (2003). MDM2-ARF complex regulates p53 sumoylation.<br />

Oncogene 22,5348-5357.<br />

Chi, Y., Huddleston, M. J., Zhang, X., Young, R. A., Annan, R. S., Carr, S. A., and<br />

Deshaies, R. J. (2001). Negative regulation <strong>of</strong> Gcn4 and Msn2 transcription factors<br />

by SrblO cyclin-dependent kinase. Genes Dev 15,1078-1092.<br />

Choi, C. Y., Kim, Y. H., Kwon, H. J., and Kim, Y. (1999). The homeodomain<br />

protein NK-3 recruits Groucho and a histone deacetylase complex to repress<br />

transcription. J Biol Chem 274,33194-33197.<br />

Chrivia, J. C., Kwok, R. P., Lamb, N., Hagiwara, M., Montminy, M. R., and<br />

Goodman, R. H. (1993). Phosphorylated CREB binds specifically to the nuclear<br />

protein CBP. Nature 365,855-859.<br />

Chun, T. H., Itoh, H., Subramanian, L., Iniguez-Lluhi, J. A., and Nakao, K. (2003).<br />

Modification <strong>of</strong> GATA-2 transcriptional activity in endothelial cells by the SUMO<br />

E3 ligase PIASy. Circ Res 92,1201-1208.<br />

Chung, C. D., Liao, J., Liu, B., Rao, X., Jay, P., Berta, P., and Shuai, K. (1997).<br />

Specific inhibition <strong>of</strong> <strong>St</strong>at3 signal transduction by PIAS3. Science 278,1803-1805.<br />

Comerford, K. M., Leonard, M. 0., Karhausen, J., Carey, R., Colgan, S. P., and<br />

Taylor, C. T. (2003). Small ubiquitin-related modifier-1 modification mediates<br />

resolution <strong>of</strong> CREB-dependent responses to hypoxia. Proc Natl Acad Sci USA 100,<br />

986-991.<br />

200


Cope, G. A., Suh, G. S., Aravind, L., Schwarz, S. E., Zipursky, S. L., Koonin, E. V.,<br />

and Deshaies, R. J. (2002). Role <strong>of</strong> predicted metalloprotease motif <strong>of</strong> Jabl/Csn5 in<br />

cleavage <strong>of</strong> Nedd8 from Cull. Science 298,608-611.<br />

Coutavas, E., Ren, M., Oppenheim, J. D., D'Eustachio, P., and Rush, M. G. (1993).<br />

Characterization <strong>of</strong> proteins that interact with the cell-cycle regulatory protein<br />

Ran/TC4. Nature 366,585-587.<br />

Dahiya, A., Wong, S., Gonzalo, S., Gavin, M., and Dean, D. C. (2001). Linking the<br />

Rb and polycomb pathways. Mol Cell 8,557-569.<br />

Darnell, J. E. Jr., Kerr, I. M., <strong>St</strong>ark, G. R. (1994). Jak-STAT pathway and<br />

transcriptional activation in response to IFNs and other extracellular signaling<br />

proteins. Science 264,1415-1421.<br />

<strong>David</strong>, G., Neptune, M. A., and DePinho, R. A. (2002). SUMO-1 modification <strong>of</strong><br />

histone deacetylase 1 (HDAC1) modulates its biological activities. J Biol Chem 277,<br />

23658-23663.<br />

Deng, L., Wang, C., Spencer, E., Yang, L., Braun, A., You, J., Slaughter, C., Pickart,<br />

C., and Chen, Z. J. (2000). Activation <strong>of</strong> the IkappaB kinase complex by TRAF6<br />

requires a dimeric ubiquitin-conjugating enzyme complex and a unique<br />

polyubiquitin chain. Cell 103,351-361.<br />

Desbois, C., Rousset, R., Bantignies, F., and Jalinot, P. (1996). Exclusion <strong>of</strong> Int-6<br />

from PML nuclear bodies by binding to the HTLV-I Tax oncoprotein. Science 273,<br />

951-953.<br />

Deshaies, R. J. (1999). SCF and Cullin/Ring H2-based ubiquitin ligases. Annu Rev<br />

Cell Dev Biol 15,435-467.<br />

201


Desterro, J. M., Rodriguez, M. S., and Hay, R. T. (1998). SUMO-1 modification <strong>of</strong><br />

IkappaBalpha inhibits NF-kappaB activation. Mol Cell 2,233-239.<br />

Desterro, J. M., Rodriguez, M. S., Kemp, G. D., and Hay, R. T. (1999).<br />

Identification <strong>of</strong> the enzyme required for activation <strong>of</strong> the small ubiquitin-like<br />

protein<br />

SUMO-1. J Biol Chem 274,10618-10624.<br />

Desterro, J. M., Thomson, J., and Hay, R. T. (1997). Ubch9 conjugates SUMO but<br />

not ubiquitin. FEBS Lett 417,297-300.<br />

Ding, J., McGrath, W. J., Sweet, R. M., and Mangel, W. F. (1996). Crystal structure<br />

<strong>of</strong> the human adenovirus proteinase with its 11 amino acid c<strong>of</strong>actor. Embo J 15,<br />

1778-1783.<br />

Dittmar, G. A., Wilkinson, C. R., Jedrzejewski, P. T., and Finley, D. (2002). Role <strong>of</strong><br />

a ubiquitin-like modification in polarized morphogenesis. Science 295,2442-2446.<br />

Doucas, V., Tini, M., Egan, D. A., and Evans, R. M. (1999). Modulation <strong>of</strong> CREB<br />

binding protein function by the promyelocytic (PML) oncoprotein suggests a role for<br />

nuclear bodies in hormone signaling. Proc Natl Acad Sci USA 96,2627-2632.<br />

Dover, J., Schneider, J., Tawiah-Boateng, M. A., Wood, A., Dean, K., Johnston, M.,<br />

and Shilatifard, A. (2002). Methylation <strong>of</strong> histone H3 by COMPASS requires<br />

ubiquitination <strong>of</strong> histone H2B by Rad6. J Biol Chem 277,28368-28371.<br />

Duprez, E., Saurin, A. J., Desterro, J. M., Lallemand-Breitenbach, V., Howe, K.,<br />

Boddy, M. N., Solomon, E., de The, H., Hay, R. T., and Freemont, P. S. (1999).<br />

SUMO-1 modification <strong>of</strong> the acute promyelocytic leukaemia protein PML:<br />

implications for nuclear localisation. J Cell Sci 112 (Pt 3), 381-393.<br />

202


Eckner, R., Arany, Z., Ewen, M., Sellers, W., and Livingston, D. M. (1994). The<br />

adenovirus EIA-associated 300-kD protein exhibits properties <strong>of</strong> a transcriptional<br />

coactivator and belongs to an evolutionarily conserved family. Cold Spring Harb<br />

Symp Quant Biol 59,85-95.<br />

Eloranta, J. J., and Hurst, H. C. (2002). Transcription factor AP-2 interacts with the<br />

SUMO-conjugating enzyme UBC9 and is sumolated in vivo. J Biol Chem 277,<br />

30798-30804.<br />

Elsasser, S., Gali, R. R., Schwickart, M., Larsen, C. N., Leggett, D. S., Muller, B.,<br />

Feng, M. T., Tubing, F., Dittmar, G. A., and Finley, D. (2002). Proteasome subunit<br />

Rpnl binds ubiquitin-like protein domains. Nat Cell Biol 4,725-730.<br />

Endter, C., Kzhyshkowska, J., <strong>St</strong>auber, R., and Dobner, T. (2001). SUMO-1<br />

modification required for transformation by adenovirus type 5 early region 1B 55-<br />

kDa oncoprotein. Proc Natl Acad Sci USA 98,11312-11317.<br />

Espinosa, J. M., and Emerson, B. M. (2001). Transcriptional regulation by p53<br />

through intrinsic DNA/chromatin binding and site-directed c<strong>of</strong>actor recruitment.<br />

Mol Cell 8,57-69.<br />

Everett, R. D. (2001). DNA viruses and viral proteins that interact with PML nuclear<br />

bodies. Oncogene 20,7266-7273.<br />

Everett, R. D., Earnshaw, W. C., Findlay, J., and Lomonte, P. (1999a). Specific<br />

destruction <strong>of</strong> kinetochore protein CENP-C and disruption <strong>of</strong> cell division by herpes<br />

simplex virus immediate-early protein Vmwl 10. Embo J 18,1526-1538.<br />

Everett, R. D., Meredith, M., and Orr, A. (1999b). The ability <strong>of</strong> herpes simplex<br />

virus type 1 immediate-early protein Vmw110 to bind to a ubiquitin-specific<br />

203


protease contributes to its roles in the activation <strong>of</strong> gene expression and stimulation<br />

<strong>of</strong> virus replication.<br />

J Virol 73,417-426.<br />

Fabunmi, R. P., Wigley, W. C., Thomas, P. J., and DeMartino, G. N. (2000).<br />

Activity and regulation <strong>of</strong> the centrosome-associated proteasome. J Biol Chem 275,<br />

409-413.<br />

Fabunmi, R. P., Wigley, W. C., Thomas, P. J., and DeMartino, G. N. (2001).<br />

Interferon gamma regulates accumulation <strong>of</strong> the proteasome activator PA28 and<br />

immunoproteasomes<br />

at nuclear PML bodies. J Cell Sci 114,29-36.<br />

Fan, W., Cai, W., Parimoo, S., Schwarz, D. C., Lennon, G. G., and Weissman, S. M.<br />

(1996). Identification <strong>of</strong> seven new human MHC class I region genes around the<br />

HLA-F locus. Immunogenetics 44,97-103.<br />

Felsenfeld, G., and Groudine, M. (2003). Controlling the double helix. Nature 421,<br />

448-453.<br />

Fichtner, L., Jablonowski, D., Schierhorn, A., Kitamoto, H. K., <strong>St</strong>ark, M. J., and<br />

Schaffrath, R. (2003). Elongator's toxin-target (TOT) function is nuclear localization<br />

sequence dependent and suppressed by post-translational modification. Mol<br />

Microbiol 49,1297-1307.<br />

Finley, D. (2001). Signal transduction. An alternative to destruction. Nature 412,<br />

283,285-286.<br />

Finley, D., and<br />

Chau, V. (1991). Ubiquitination. Annu Rev Cell Biol 7,25-69.<br />

Finley, D., Ozkaynak, E., and Varshavsky, A. (1987). The yeast polyubiquitin gene<br />

'is essential for resistance to high temperatures, starvation, and other stresses. Cell<br />

48,1035-1046.<br />

204<br />

lýiý


Forler, D., Rabut, G., Ciccarelli, F. D., Herold, A., Kocher, T., Niggeweg, R., Bork,<br />

P., Ellenberg, J., and Izaurralde, E. (2004). RanBP2/Nup358 provides a major<br />

binding site for NXF1-p15 dimers at the nuclear pore complex and<br />

nuclear mRNA export. Mol Cell Biol 24,1155-1167.<br />

functions in<br />

Francis, N. J., Saurin, A. J., Shao, Z., and Kingston, R. E. (2001). Reconstitution <strong>of</strong> a<br />

functional core polycomb repressive complex. Mol Cell 8,545-556.<br />

Fraser, A. G., Kamath, R. S., Zipperlen, P., Martinez-Campos, M., Sohrmann, M.,<br />

and Ahringer, J. (2000). Functional genomic analysis <strong>of</strong> C. elegans chromosome I by<br />

systematic RNA interference. Nature 408,325-330.<br />

Freiman, R. N., and Tjian, R. (2003). Regulating the regulators: lysine modifications<br />

make their mark. Cell 112,11-17.<br />

Furukawa, K., Mizushima, N., Noda, T., and Ohsumi, Y. (2000). A protein<br />

conjugation system in yeast with homology to biosynthetic enzyme reaction <strong>of</strong><br />

prokaryotes. J Biol Chem 275,7462-7465.<br />

Gan-Erdene, T., Nagamalleswari, K., Yin, L., Wu, K., Pan, Z. Q., and Wilkinson, K.<br />

D. (2003). Identification and characterization <strong>of</strong> DEN!, a deneddylase <strong>of</strong> the ULP<br />

family. J Biol Chem 278,28892-28900.<br />

Gar<strong>of</strong>f, H., Hewson, R., and Opstelten, D. J. (1998). Virus maturation by budding.<br />

Microbiol Mol Biol Rev 62,1171-1190.<br />

Gerasimova, T. I., and Corces, V. G. (1998). Polycomb and trithorax group proteins<br />

mediate the function <strong>of</strong> a chromatin insulator. Cell 92,511-521.<br />

205


Giorgino, F., de Robertis, 0., Laviola, L., Montrone, C., Perrini, S., McCowen, K.<br />

C., and Smith, R. J. (2000). The sentrin-conjugating enzyme mUbc9 interacts with<br />

GLUT4 and GLUT1 glucose transporters and, regulates transporter levels in skeletal<br />

muscle cells. Proc Natl Acad Sci USA 97,1125-1130.<br />

Giraud, M. F., Desterro, J. M., and Naismith, J. H. (1998). <strong>St</strong>ructure <strong>of</strong> ubiquitin-<br />

conjugating enzyme 9 displays significant differences with other ubiquitin-<br />

conjugating enzymes which may reflect its specificity for sumo rather than ubiquitin.<br />

Acta Crystallogr D Biol Crystallogr 54 (Pt 5), 891-898.,<br />

<strong>Girdwood</strong>, D., Bumpass, D., Vaughan, 0. A., Thain, A., Anderson, L. A., Snowden,<br />

A. W., 'Garcia-Wilson, E., Perkins, N. D., and Hay, R. T. (2003). P300<br />

transcriptional repression is mediated by SUMO modification. Mol Cell 11,1043-<br />

1054.<br />

Glickman, M. H., Rubin, D. M., Coux, 0., Wefes, I., Pfeifer, G., Cjeka, Z.,<br />

Baumeister, W., Fried, V. A., and Finley, D. (1998). A subcomplex <strong>of</strong> the<br />

proteasome regulatory particle required for ubiquitin-conjugate degradation and<br />

related to the COP9-signalosome and eIF3. Cell 94,615-623.<br />

Goldknopf, I. L., and Busch, H. (1977). Isopeptide linkage between nonhistone and<br />

histone 2A polypeptides <strong>of</strong> chromosomal conjugate-protein A24. Proc Natl Acad Sci<br />

USA 74,864-868.<br />

Goldknopf, I. L., Taylor, C. W., Baum, R. M., Yeoman, L. C., Olson, M. 0.,<br />

Prestayko, A. W., and Busch, H. (1975). Isolation and characterization <strong>of</strong> protein<br />

A24, a "histone-like" non-histone chromosomal protein. J Biol Chem 250,7182-<br />

7187.<br />

Goldstein, G., Scheid, M., Hammerling, U., Schlesinger, D. H., Niall, H. D., and<br />

Boyse, E. A. (1975). Isolation <strong>of</strong> a polypeptide that has lymphocyte-differentiating<br />

206


properties and is probably represented universally in living cells. Proc Natl Acad Sci<br />

USA 72,11-15.<br />

Gong, L., Kamitani, T., Fujise, K., Caskey, L. S., and Yeh, E. T. (1997). Preferential<br />

interaction <strong>of</strong> sentrin with a ubiquitin-conjugating enzyme, Ubc9. J Biol Chem 272,<br />

28198-28201.<br />

Gong, L., Li, B., Millas, S., and Yeh, E. T. (1999). Molecular cloning and<br />

characterization <strong>of</strong> human AOS 1 and UBA2, components <strong>of</strong> the sentrin-activating<br />

enzyme complex. FEBS Lett 448,185-189.<br />

Gong, L., Millas, S., Maul, G. G., and Yeh, E. T. (2000). Differential regulation <strong>of</strong><br />

sentrinized proteins by a novel sentrin-specific protease. J Biol Chem 275,3355-<br />

3359.<br />

Gong, L., and Yeh, E. T. (1999). Identification <strong>of</strong> the activating and conjugating<br />

enzymes <strong>of</strong> the NEDD8 conjugation pathway. J Biol Chem 274,12036-12042.<br />

Gongora, C., <strong>David</strong>, G., Pintard, L., Tissot, C., Hua, T. D., Dejean, A., and Mechti,<br />

N. (1997). Molecular cloning <strong>of</strong> a new interferon-induced PML nuclear body-<br />

associated protein. J Biol Chem 272,19457-19463.<br />

Gonzalez, F., Delahodde, A., Kodadek, T., and Johnston, S. A. (2002). Recruitment<br />

<strong>of</strong> a 19S proteasome subcomplex to an activated promoter. Science 296,548-550.<br />

Goodman, R. H., and Smolik, S. (2000). CBP/p300 in cell growth, transformation,<br />

and development. Genes Dev 14,1553-1577.<br />

Goodson, M. L., Hong, Y., Rogers, R., Matunis, M. J., Park-Sarge, 0. K., and Sarge,<br />

K. D. (2001). Sumo-1 modification regulates the DNA binding activity <strong>of</strong> heat shock<br />

207


transcription factor 2, a promyelocytic leukemia nuclear body associated<br />

transcription factor. J Biol Chem 276,18513-18518.<br />

Gorlich, D., and Kutay, U. (1999). Transport between the cell nucleus and the<br />

cytoplasm. Annu Rev Cell Dev Biol 15,607-660.<br />

Gosink, M. M., and Vierstra, R. D. (1995). Redirecting the specificity <strong>of</strong><br />

ubiquitination by modifying ubiquitin-conjugating enzymes. Proc Natl Acad Sci US<br />

A 92,9117-9121.<br />

Gostissa, M., Hengstermann, A., Fogal, V., Sandy, P., Schwarz, S. E., Scheffner, M.,<br />

and Del Sal, G. (1999). Activation <strong>of</strong> p53 by conjugation to the ubiquitin-like<br />

protein SUMO-I. Embo J 18,6462-6471.<br />

Gregory, P. D., Schmid, A., Zavari, M., Munsterkotter, M., and Horz, W. (1999).<br />

Chromatin remodelling at the PHO8 promoter requires SWI-SNF and SAGA at a<br />

step subsequent to activator binding. Embo J 18,6407-6414.<br />

Groettrup, M., Ruppert, T., Kuehn, L., Seeger, M., <strong>St</strong>andera, S., Koszinowski, U.,<br />

and Kloetzel, P. M. (1995). The interferon-gamma-inducible 11 S regulator (PA28)<br />

and the LMP2/LMP7 subunits govern the peptide production by the 20 S proteasome<br />

in vitro. J Biol Chem 270,23808-23815.<br />

Groll, M., Ditzel, L., Lowe, J., <strong>St</strong>ock, D., Bochtler, M., Bartunik, H. D., and Huber,<br />

R. (1997). <strong>St</strong>ructure <strong>of</strong> 20S proteasome from yeast at 2.4 A resolution. Nature 386,<br />

463-471.<br />

Gronroos, E., Hellman, U., Heldin, C. H., and Ericsson, J. (2002). Control <strong>of</strong> Smad7<br />

stability by competition between acetylation and ubiquitination. Mol Cell 10,483-<br />

493.<br />

208


Gross, M., Yang, R., Top, I., Gasper, C., and Shuai, K. (2004). PIASy-mediated<br />

repression <strong>of</strong> the androgen receptor is independent <strong>of</strong> sumoylation. Oncogene.<br />

Grossman, S. R., Deato, M. E., Brignone, C., Chan, H. M., Kung, A. L., Tagami, H.,<br />

Nakatani, Y., and Livingston, D. M. (2003). Polyubiquitination <strong>of</strong> p53 by a ubiquitin<br />

ligase activity <strong>of</strong> p300.<br />

Science 300,342-344.<br />

Gu, W., Shi, X. L., and Roeder, R. G. (1997). Synergistic activation <strong>of</strong> transcription<br />

by CBP and p53.<br />

Nature 387,819-823.<br />

Gunster, M. J., Raaphorst, F. M., Hamer, K. M., den Blaauwen, J. L., Fieret, E.,<br />

Meijer, C. J., and Otte, A. P. (2001). Differential expression <strong>of</strong><br />

human Polycomb<br />

group proteins in various tissues and cell types. J Cell Biochem Suppl Suppl 36,129-<br />

143.<br />

Guo, A., Salomoni, P., Luo, J., Shih, A., Zhong, S., Gu, W., and Paolo Pandolfi, P.<br />

(2000). The function <strong>of</strong> PML in p53-dependent apoptosis. Nat Cell Biol 2,730-736.<br />

Haas, A. L., Ahrens, P., Bright, P. M., and Ankel, H. (1987). Interferon induces a<br />

15-kilodalton protein exhibiting marked homology to ubiquitin. J Biol Chem 262,<br />

11315-11323.<br />

Haas, A. L., and Rose, I. A. (1982). The mechanism <strong>of</strong> ubiquitin activating enzyme.<br />

A kinetic and equilibrium analysis. J Biol Chem 257,10329-10337.<br />

Haas, A. L., and Siepmann, T. J. (1997). Pathways <strong>of</strong> ubiquitin conjugation. Faseb J<br />

11,1257-1268.<br />

209


Hamerman, J. A., Hayashi, F., Schroeder, L. A., Gygi, S. P., Haas, A. L., Hampson,<br />

L., Coughlin, P., Aebersold, R., and Aderem, A. (2002). Serpin 2a is induced in<br />

activated macrophages and conjugates to a ubiquitin homolog. J Immunol 168,<br />

2415-2423.<br />

Hang, J., and Dasso, M. (2002). Association <strong>of</strong> the human SUMO-1 protease SENP2<br />

with the nuclear pore. J Biol Chem 277,19961-19966.<br />

Hardeland, U., <strong>St</strong>einacher, R., Jiricny, J., and Schar, P. (2002). Modification <strong>of</strong> the<br />

human thymine-DNA glycosylase by ubiquitin-like proteins facilitates enzymatic<br />

turnover. Embo J 21,1456-1464.<br />

Hart, M., Concordet, J. P., Lassot, I., Albert, I., del los Santos, R., Durand, H.,<br />

Perret, C., Rubinfeld, B., Margottin, F., Benarous, R., and Polakis, P. (1999). The F-<br />

box protein beta-TrCP associates with phosphorylated beta-catenin and regulates its<br />

activity in the cell. Curr Biol 9,207-210.<br />

Hasan, S., Hassa, P. 0., Imh<strong>of</strong>, R., and Hottiger, M. 0. (2001). Transcription<br />

coactivator p300 binds PCNA and may have a role in DNA repair synthesis. Nature<br />

410,387-391.<br />

Hassan, A. H., Neely, K. E., and Workman, J. L. (2001). Histone acetyltransferase<br />

complexes stabilize swi/snf binding to promoter nucleosomes. Cell 104,817-827.<br />

Hay, R. T. (2001). Protein modification by SUMO. Trends Biochem Sci 26,332-<br />

333.<br />

Hay, R. T. (2004). Modifying NEMO. Nat Cell Biol 6,89-91.<br />

210


Hebbes, T. R., Clayton, A. L., Thorne, A. W., and Crane-Robinson, C. (1994). Core<br />

histone hyperacetylation co-maps with generalized DNase I sensitivity in the<br />

chicken beta-globin chromosomal domain. Embo J 13,1823-1830.<br />

Hicke, 'L. (2001). Protein regulation by monoubiquitin. Nat Rev Mol Cell Biol 2,<br />

195-201.<br />

Hingamp, P. M., Arnold, J. E., Mayer, R. J., and Dixon, L. K. (1992). A ubiquitin<br />

conjugating enzyme encoded by African swine fever virus. Embo J 11,361-366.<br />

Hingamp, P. M., Leyland, M. L., Webb, J., Twigger, S., Mayer, R. J., and Dixon, L.<br />

K. (1995). Characterization <strong>of</strong> a ubiquitinated protein which is externally located in<br />

African swine fever virions.<br />

J Virol 69,1785-1793.<br />

Hirano, Y., Murata, S., Tanaka, K., Shimizu, M., and Sato, R. (2003). <strong>St</strong>erol<br />

regulatory element-binding proteins are negatively regulated through SUMO-1<br />

modification independent <strong>of</strong> the ubiquitin/26 S proteasome pathway. J Biol Chem<br />

278,16809-16819.<br />

Ho, J. C., Warr, N. J., Shimizu, H., and Watts, F. Z. (2001). SUMO modification <strong>of</strong><br />

Rad22, the Schizosaccharomyces<br />

pombe homologue <strong>of</strong> the recombination protein<br />

Rad52. Nucleic Acids Res 29,4179-4186.<br />

Ho, J. C., and Watts, F. Z. (2003). Characterization <strong>of</strong> SUMO-conjugating enzyme<br />

mutants in Schizosaccharomyces pombe identifies a dominant-negative allele that<br />

severely reduces SUMO conjugation. Biochem J 372,97-104.<br />

Hochstrasser, M. (2002). New structural clues to substrate specificity in the<br />

"ubiquitin system". Mol Cell 9,453-454.<br />

211


Hoege, C., Pfander, B., Moldovan, G. L., Pyrowolakis, G., and Jentsch, S. (2002).<br />

RAD6-dependent DNA repair is linked to modification <strong>of</strong> PCNA by ubiquitin and<br />

SUMO. Nature 419,135-141.<br />

H<strong>of</strong>mann, H., Floss, S., and <strong>St</strong>amminger, T. (2000). Covalent modification <strong>of</strong> the<br />

transactivator protein IE2-p86 <strong>of</strong> human cytomegalovirus by conjugation to the<br />

ubiquitin-homologous proteins SUMO-1 and hSMT3b. J Virol 74,2510-2524.<br />

Holmstrom, S., Van Antwerp, M. E., and Iniguez-Lluhi, J. A. (2003). Direct and<br />

distinguishable inhibitory roles for SUMO is<strong>of</strong>orms in the control <strong>of</strong> transcriptional<br />

synergy. Proc Natl Acad Sci USA 100,15758-15763.<br />

Hong, R., and Chakravarti, D. (2003). The human proliferating Cell nuclear antigen<br />

regulates transcriptional coactivator p300 activity and promotes transcriptional<br />

repression. J Biol Chem 278,44505-44513.<br />

Hong, Y., Rogers, R., Matunis, M. J., Mayhew, C. N., Goodson, M. L., Park-Sarge,<br />

0. K., Sarge, K. D., and Goodson, M. (2001). Regulation <strong>of</strong> heat shock transcription<br />

factor 1 by stress-induced SUMO-1 modification. J Biol Chem 276,40263-40267.<br />

Hook, S. S., Orian, A., Cowley, S. M., and Eisenman, R. N. (2002). Histone<br />

deacetylase 6 binds polyubiquitin through its zinc finger (PAZ domain) and<br />

copurifies with deubiquitinating enzymes. Proc Natl Acad Sci USA 99,13425-<br />

13430.<br />

Hori, T., Osaka, F., Chiba,. T., Miyamoto, C., Okabayashi, K., Shimbara, N., Kato,<br />

S., and Tanaka, K. (1999). Covalent modification <strong>of</strong> all members <strong>of</strong> human cullin<br />

family proteins by NEDD8. Oncogene 18,6829-6834.<br />

212


Horie, K., Tomida, A., Sugimoto, Y., Yasugi, T., Yoshikawa, H., Taketani, Y., and<br />

Tsuruo, T. (2002). SUMO-1 conjugation to intact DNA topoisomerase I amplifies<br />

cleavable complex formation induced by camptothecin. Oncogene 21,7913-7922.<br />

Huang, T. T., Wuerzberger-Davis, S. M., Wu, Z. H., and Miyamoto, S. (2003).<br />

Sequential modification <strong>of</strong> NEMO/IKKgamma by SUMO-1 and ubiquitin mediates<br />

NF-kappaB activation by genotoxic stress. Cell 115,565-576.<br />

Ichimura, Y., Kirisako, T., Takao, T., Satomi, Y., Shimonishi, Y., Ishihara, N.,<br />

Mizushima, N., Tanida, I., Kominami, E., Ohsumi, M., et al. (2000). A ubiquitin-<br />

like system mediates protein lipidation. Nature 408,488-492.<br />

Impey, S., Fong, A. L., Wang, Y., Cardinaux, J. R., Fass, D. M., Obrietan, K.,<br />

Wayman, G. A., <strong>St</strong>orm, D. R., Soderling, T. R., and Goodman, R. H. (2002).<br />

Phosphorylation <strong>of</strong> CBP mediates transcriptional activation by neural activity and<br />

CaM kinase IV. Neuron 34,235-244.<br />

Ishov, A. M., Sotnikov, A. G., Negorev, D., Vladimirova, O. V., Neff, N., Kamitani,<br />

T., Yeh, E. T., <strong>St</strong>rauss, J. F., 3rd, and Maul, G. G. (1999). PML is critical for ND10<br />

formation and recruits the PML-interacting protein daxx to this nuclear structure<br />

when modified by SUMO-1. J Cell Biol 147,221-234.<br />

Isogai, Y., and Tjian, R. (2003). Targeting genes and transcription factors to<br />

segregated nuclear compartments. Curr Opin Cell Biol 15,296-303.<br />

Ito, T., Ikehara, T., Nakagawa, T., Kraus, W. L., and Muramatsu, M. (2000). p300-<br />

mediated acetylation facilitates the transfer <strong>of</strong> histone H2A-H2B dimers from<br />

nucleosomes to a histone chaperone. Genes Dev 14,1899-1907.<br />

Jayachandra, S., Low, K. G., Thlick, A. E., Yu, J., Ling, P. D., Chang, Y., and<br />

Moore, P. S. (1999). Three unrelated viral transforming proteins (vIRF, EBNA2, and<br />

213


E1A) induce the MYC oncogene through the interferon-responsive PRF element by<br />

using different transcription coadaptors. Proc Natl Acad Sci USA 96,11566-11571.<br />

Jentsch, S. (1992). The ubiquitin-conjugation system. Annu Rev Genet 26,179-207.<br />

Jentsch, S., and Pyrowolakis, G. (2000). Ubiquitin and its kin: how close are the<br />

family ties? Trends Cell Biol 10,335-342.<br />

Jentsch, S., and Ulrich, H. D. (1998). Protein breakdown. Ubiquitous deja vu. Nature<br />

395,321,323.<br />

Jenuwein, T., and Allis, C. D. (2001). Translating the histone code. Science 293,<br />

1074-1080.<br />

Jiang, W. Q., Szekely, L., Klein, G., and Ringertz, N. (1996). Intranuclear<br />

redistribution <strong>of</strong> SV40T, p53, and PML in a conditionally SV40T-immortalized cell<br />

line. Exp Cell Res 229,289-300.<br />

Joazeiro, C. A., and Weissman, A. M. (2000). RING finger proteins: mediators <strong>of</strong><br />

ubiquitin ligase activity. Cell 102,549-552.<br />

Joazeiro, C. A., Wing, S. S., Huang, H., Leverson, J. D., Hunter, T., and Liu, Y. C.<br />

(1999). The tyrosine kinase negative regulator c-Cbl as a RING-type, E2-dependent<br />

ubiquitin-protein ligase. Science 286,309-312.<br />

Johnson, E. S., and Blobel, G. (1999). Cell cycle-regulated attachment <strong>of</strong> the<br />

ubiquitin-related protein SUMO to the yeast septins. J Cell Biol 147,981-994.<br />

Johnson, E. S., and Gupta, A. A. (2001). An E3-like factor that promotes SUMO<br />

conjugation to the yeast septins. Cell 106,735-744.<br />

214


Joseph, J., Tan, S. H., Karpova, T. S., McNally, J. G., and Dasso, M. (2002).<br />

SUMO-1 targets RanGAP1 to kinetochores and mitotic spindles. J Cell Biol 156,<br />

595-602.<br />

Kadare, G., Toutant, M., Formstecher, E., Corvol, J. C., Carnaud, M., Boutterin, M.<br />

C., and Girault, J. A. (2003). PIAS 1-mediated sumoylation <strong>of</strong> focal adhesion kinase<br />

activates its autophosphorylation. J Biol Chem 278,47434-47440.<br />

Kadoya, T., Yamamoto, H., Suzuki, T., Yukita, A., Fukui, A., Michiue, T., Asahara,<br />

T., Tanaka, K., Asashima, M., Kikuchi, A. (2002). Desumoylation activity <strong>of</strong><br />

Axam, a novel Axin-binding protein, is involved in downregulation <strong>of</strong> beta-catenin.<br />

Mol Cell Biol 11,3803-3819.<br />

Kagey, M. H., Melhuish, T. A., and Wotton, D. (2003). The polycomb protein Pc2 is<br />

a SUMO E3. Cell 113,127-137.<br />

Kahyo, T., Nishida, T., and Yasuda, H. (2001). Involvement <strong>of</strong> PIAS1 in the<br />

sumoylation <strong>of</strong> tumor suppressor p53. Mol Cell 3,713-718.<br />

Kalkhoven, E., Roelfsema, J. H., Teunissen, H., den Boer, A., Ariyurek, Y.,<br />

Zantema, A., Breuning, M. H., Hennekam, R. C., and Peters, D. J. (2003). Loss <strong>of</strong><br />

CBP acetyltransferase activity by PHD finger mutations in Rubinstein-Taybi<br />

syndrome. Hum Mol Genet 12,441-450.<br />

Kamitani, T., Kito, K., Nguyen, H. P., Fukuda-Kamitani, T., and Yeh, E. T. (1998a).<br />

Characterization <strong>of</strong> a second member <strong>of</strong> the sentrin family <strong>of</strong> ubiquitin-like proteins<br />

J Biol Chem 273,11349-11353.<br />

Kamitani, T., Kito, K., Nguyen, H. P., and Yeh, E. T. (1997a). Characterization <strong>of</strong><br />

NEDD8, a developmentally down-regulated ubiquitin-like protein. J Biol Chem 272,<br />

28557-28562.<br />

215


Kamitani, T., Nguyen, H. P., Kito, K., Fukuda-Kamitani, T., and Yeh, E. T. (1998b).<br />

Covalent modification <strong>of</strong> PML by the sentrin family <strong>of</strong> ubiquitin-like proteins. J Biol<br />

Chem 273,3117-3120.<br />

Kamitani, T., Nguyen, H. P., and Yeh, E. T. (1997b). Activation-induced<br />

aggregation and processing <strong>of</strong> the human Fas antigen. Detection with cytoplasmic<br />

domain-specific antibodies. J Biol Chem 272,22307-22314.<br />

Kang, E. S., Park, C. W., and Chung, J. H. (2001). Dnmt3b, de novo DNA<br />

methyltransferase, interacts with SUMO-1 and Ubc9 through its N-terminal region<br />

and is subject to modification by SUMO-1. Biochem Biophys Res Commun 289,<br />

862-868.<br />

Kang, S. I., Chang, W. J., Cho, S. G., and Kim, I. Y. (2003). Modification <strong>of</strong><br />

promyelocytic leukemia zinc finger protein (PLZF) by SUMO-1 conjugation<br />

regulates its transcriptional repressor activity. J Biol Chem 278,51479-51483.<br />

Kas, K., Michiels, L., and Merregaert, J. (1992). Genomic structure and expression<br />

<strong>of</strong> the human fau gene: encoding the ribosomal protein S30 fused to a ubiquitin-like<br />

protein. Biochem Biophys Res Commun 187,927-933.<br />

Kawabe, Y., Seki, M., Seki, T., Wang, W. S., Imamura, 0., Furuichi, Y., Saitoh, H.,<br />

and Enomoto, T. (2000). Covalent modification <strong>of</strong> the Werner's syndrome gene<br />

product with the ubiquitin-related protein, SUMO-1. J Biol Chem 275,20963-<br />

20966.<br />

Kawai, H., Nie, L., Wiederschain, D., and Yuan, Z. M. (2001). Dual role <strong>of</strong> p300 in<br />

the regulation <strong>of</strong> p53 stability. J Biol Chem 276,45928-45932.<br />

216


Kawakami, T., Chiba, T., Suzuki, T., Iwai, K., Yamanaka, K., Minato, N., Suzuki,<br />

H., Shimbara, N., Hidaka, Y., Osaka, F., et al. (2001). NEDD8 recruits E2-ubiquitin<br />

to SCF E3 ligase. Embo J 20,4003-4012.<br />

Kawasaki, H., Eckner, R., Yao, T. P., Taira, K., Chiu, R., Livingston, D. M., and<br />

Yokoyama, K. K. (1998). Distinct roles <strong>of</strong> the co-activators p300 and CBP in<br />

retinoic-acid-induced F9-cell differentiation. Nature 393,284-289.<br />

Kee, B. L., Arias, J., and Montminy, M. R. (1996). Adaptor-mediated recruitment <strong>of</strong><br />

RNA polymerase II to a signal-dependent activator. J Biol Chem 271,2373-2375.<br />

Kennison, J. A. (1995). The Polycomb and trithorax group proteins <strong>of</strong> Drosophila:<br />

trans-regulators <strong>of</strong> homeotic gene function. Annu Rev Genet 29,289-303.<br />

Khochbin, S., Verdel, A., Lemercier, C., Seigneurin-Berny, D. (2001). Functional<br />

significance <strong>of</strong> histone deacetylase diversity. Curr Opin Genet Dev 21,162-166.<br />

Kim, J., Cantwell, C. A., Johnson, P. F., Pfarr, C. M., and <strong>William</strong>s, S. C. (2002).<br />

Transcriptional activity <strong>of</strong> CCAAT/enhancer-binding proteins is controlled by a<br />

conserved inhibitory domain that is a target for sumoylation. J Biol Chem 277,<br />

38037-38044.<br />

Kim, K. I., Baek, S. H., Jeon, Y. J., Nishimori, S., Suzuki, T., Uchida, S., Shimbara,<br />

N., Saitoh, H., Tanaka, K., and Chung, C. H. (2000). A new SUMO-1-specific<br />

protease, SUSP1, that is highly expressed in reproductive organs. J Biol Chem 275,<br />

14102-14106.<br />

Kim, K. I., and Zhang, D. E. (2003). ISG15, not just another ubiquitin-like protein.<br />

Biochem Biophys Res Commun 307,431-434. ,<br />

217


Kim, S. Y., Herbst, A., Tworkowski, K. A., Salghetti, S. E., and Tansey, W. P.<br />

(2003). Skp2 regulates Myc protein stability and activity. Mol Cell 11,1177-1188.<br />

Kim, Y. H., Choi, C. Y., and Kim, Y. (1999). Covalent modification <strong>of</strong> the<br />

homeodomain-interacting protein kinase 2 (HIPK2) by the ubiquitin-like protein<br />

SUMO-1. Proc Natl Acad Sci USA 96,12350-12355.<br />

Kirsh, 0., Seeler, J. S., Pichler, A., Gast, A., Muller, S., Miska, E., Mathieu, M.,<br />

Harel-Bellan, A., Kouzarides, T., Melchior, F., and Dejean, A. (2002). The SUMO<br />

E3 ligase RanBP2 promotes modification <strong>of</strong> the HDAC4 deacetylase. Embo J 21,<br />

2682-2691.<br />

Kishi, A., Nakamura, T., Nishio, Y., Maegawa, H., and Kashiwagi, A. (2003).<br />

Sumoylation <strong>of</strong> Pdxl is associated with its nuclear localization and insulin gene<br />

activation. Am J Physiol Endocrinol Metab 284, E830-840.<br />

Kito, K., Yeh, E. T., and Kamitani, T. (2001). NUB 1, a NEDD8-interacting protein,<br />

is induced by interferon and down-regulates the NEDD8 expression. J. Biol Chem<br />

276,20603-20609.<br />

Klement, I. A., Skinner, PA., Kaytor, M. D., Yi, H., Hersch, S. M., Clark, H. B.,<br />

Zoghbi, H. Y., and Orr, H. T. (1998). Ataxin-1 nuclear localization and aggregation:<br />

role in polyglutamine-induced disease in, SCA1 transgenic mice. Cell 95,41-53.<br />

Kohler, A., Bajorek, M., Groll, M., Moroder, L., Rubin, D. M., Huber, R.,<br />

Glickman, M. H., and Finley, D. (2001). The substrate translocation channel <strong>of</strong> the<br />

proteasome. Biochimie 83,325-332.<br />

Kok, K., H<strong>of</strong>stra, R., Pilz, A., van den Berg, A., Terpstra, P., Buys, C. H., and<br />

Carritt, B. (1993). A gene in the chromosomal region 3p21 with greatly reduced<br />

218


expression in lung cancer is similar to the gene for ubiquitin-activating enzyme. Proc<br />

Natl Acad Sci USA 90,6071-6075.<br />

Komatsu, M., Chiba, T., Tatsumi, K., lemura, S., Tanida, I., Okazaki, N., Ueno, T.,<br />

Kominami, E., Natsume, T., and Tanaka, K. (2004). A novel protein-conjugating<br />

system for Ufml, a ubiquitin-fold modifier. Embo J 23,1977-1986.<br />

Kotaja, N., Karvonen, U., Janne, 0. A., and Palvimo, J. J. (2002). PIAS proteins<br />

modulate transcription factors by functioning as SUMO-1 ligases. Mol Cell Biol 22,<br />

5222-5234.<br />

Kovalenko, 0. V., Plug, A. W., Haaf, T., Gonda, D. K., Ashley, T., Ward, D. C.,<br />

Radding, C. M., and Golub, E. I. (1996). Mammalian ubiquitin-conjugating enzyme<br />

Ubc9 interacts with Rad51 recombination protein and localizes in synaptonemal<br />

complexes. Proc Natl Acad Sci USA 93,2958-2963.<br />

Kumar, S., Tomooka, Y., and Noda, M. (1992). Identification <strong>of</strong> a set <strong>of</strong> genes with<br />

developmentally down-regulated expression in the mouse brain. Biochem Biophys<br />

Res Commun 185,1155-1161.<br />

Kung, A. L., Rebel, V. I., Bronson, R. T., Ch'ng, L. E., Sieff, C. A., Livingston, D.<br />

M., and Yao, T. P. (2000). Gene dose-dependent<br />

control <strong>of</strong> hematopoiesis and<br />

hematologic tumor suppression by CBP. Genes Dev 14,272-277.<br />

Kwok, R. P., Laurance, M. E., Lundblad, J. R., Goldman, P. S., Shih, H., Connor, L.<br />

M., Marriott, S. J., and Goodman, R. H. (1996). Control <strong>of</strong> cAMP-regulated<br />

enhancers by the viral transactivator Tax through CREB and the co-activator CBP.<br />

Nature 380,642-646.<br />

219


Lai, H. K., and Borden, K. L. (2000). The promyelocytic leukemia (PML) protein<br />

suppresses cyclin D1 protein production by altering the nuclear cytoplasmic<br />

distribution <strong>of</strong> cyclin D1 mRNA. Oncogene 19,1623-1634.<br />

Lake, M. W., Wuebbens, M. M., Rajagopalan, K. V., and Schindelin, H. (2001).<br />

Mechanism <strong>of</strong> ubiquitin activation revealed by the structure <strong>of</strong> a bacterial MoeB-<br />

MoaD complex. Nature 414,325-329.<br />

Lallemand-Breitenbach, V., Zhu, J., Puvion, F., Koken, M., Honore, N.,<br />

Doubeikovsky, A., Duprez, E., Pandolfi, P. P., Puvion, E., Freemont, P., and de The,<br />

H. (2001). Role <strong>of</strong> promyelocytic leukemia (PML) sumolation in nuclear body<br />

formation, 11 S proteasome recruitment, and As203-induced PML or PML/retinoic<br />

acid receptor alpha degradation. J Exp Med 193,1361-1371.<br />

Lamond, A. I., and Sleeman, J. E. (2003). Nuclear substructure and dynamics. Curr<br />

Biol 13, R825-828.<br />

LaMorte, V. J., Dyck, J. A., Ochs, R. L., and Evans, R. M. (1998). Localization <strong>of</strong><br />

nascent RNA and CREB binding protein with the PML-containing nuclear body.<br />

Proc Nat! Acad Sci USA 95,4991-4996.<br />

Lapenta, V., Chiurazzi, P., van der Spek, P., Pizzuti, A., Hanaoka, F., and Brahe, C.<br />

(1997). SMT3A, a human homologue <strong>of</strong> the S. cerevisiae SMT3 gene, maps to<br />

chromosome 2 lgter and defines a novel gene family. Genomics 40,362-366.<br />

Lee, C. G., Ren, J., Cheong, I. S., Ban, K. H., Ooi, L. L., Yong Tan, S., Kan, A.,<br />

Nuchprayoon, I., Jin, R., Lee, K. H., et al. (2003). Expression <strong>of</strong> the FAT10 gene is<br />

highly upregulated in hepatocellular carcinoma and other gastrointestinal and<br />

gynecological cancers. Oncogene 22,2592-2603. "''<br />

....<br />

7ý_3<br />

220


Lee J-S, Galvin, K. M., See, R. H., Eckner, R., Livingston, D., Moran, E., Shi, Y.<br />

(1995). Relief <strong>of</strong> YY1 transcriptional repression by adenovirus E1A is mediated by<br />

EIA-associated protein p300. Genes Dev 9,1188-1198.<br />

Leggett, D. S., Hanna, J., Borodovsky, A., Crosas, B., Schmidt, M., Baker, R. T.,<br />

Walz, T., Ploegh, H., and Finley, D. (2002). Multiple associated proteins regulate<br />

proteasome structure and function. Mol Cell 10,495-507.<br />

Lehembre, F., Badenhorst, P., Muller, S., Travers, A., Schweisguth, F., and Dejean,<br />

A. (2000). Covalent modification <strong>of</strong> the transcriptional repressor tramtrack by the<br />

ubiquitin-related protein Smt3 in Drosophila flies. Mol Cell Biol 20,1072-1082.<br />

Lehembre, F., Muller, S., Pandolfi, P. P., and Dejean, A. (2001). Regulation <strong>of</strong> Pax3<br />

transcriptional activity by SUMO-1-modified PML. Oncogene 20,1-9.<br />

Lehming, N., Le Saux, A., Schuller, J., and Ptashne, M. (1998). Chromatin<br />

components as part <strong>of</strong> a putative transcriptional repressing complex. Proc Natl Acad<br />

Sci USA 95,7322-7326.<br />

Lei, X., Johnson, R. P., and Porco, J. A., Jr. (2003). Total synthesis <strong>of</strong> the ubiquitin-<br />

activating enzyme inhibitor (+)-panepophenanthrin. Angew Chem Int Ed Eng142,<br />

3913-3917.<br />

Leimkuhler, S., Wuebbens, M. M., and Rajagopalan, K. V. (2001). Characterization<br />

<strong>of</strong> Escherichia coli MoeB and its involvement in the activation <strong>of</strong> molybdopterin<br />

synthase for the biosynthesis <strong>of</strong> the molybdenum c<strong>of</strong>actor. J Biol Chem 276,34695-<br />

34701.<br />

Li, H., Leo, C., Zhu, J., Wu, X., O'Neil, J., Park, E. J., and Chen, J. D. (2000a).<br />

Sequestration and inhibition <strong>of</strong> Daxx-mediated transcriptional repression by PML.<br />

Mol Cell Biol 20,1784-1796.<br />

221


Li, M., Brooks, C. L., Wu-Baer, F., Chen, D., Baer, R., and Gu, W. (2003a). Mono-<br />

versus polyubiquitination: differential control <strong>of</strong> p53 fate by Mdm2. Science 302,<br />

1972-1975.<br />

Li, M., Damania, B., Alvarez, X., Ogryzko, V., Ozato, K., and Jung, J. U. (2000b).<br />

Inhibition <strong>of</strong> p300 histone acetyltransferase by viral interferon regulatory factor. Mol<br />

Cell Biol 20,8254-8263.<br />

Li, S. J., and Hochstrasser, M. (1999). A new protease required for cell-cycle<br />

progression in yeast. Nature 398,246-251.<br />

Li, S. J., and Hochstrasser, M. (2000). The yeast ULP2 (SMT4) gene encodes a<br />

novel protease specific for the ubiquitin-like Smt3 protein. Mol Cell Biol 20,2367-<br />

2377.<br />

Li, W., Hesabi, B., Babbo, A., Pacione, C., Liu, J., Chen, D. J., Nickol<strong>of</strong>f, J. A., and<br />

Shen, Z. (2000c). Regulation <strong>of</strong> double-strand break-induced mammalian<br />

homologous recombination by UBL1, a RAD51-interacting protein. Nucleic Acids<br />

Res 28,1145-1153.<br />

Li, Y., Wang, H., Wang, S., Quon, D., Liu, Y. W., and Cordell, B. (2003b). Positive<br />

and negative regulation <strong>of</strong> APP amyloidogenesis by sumoylation. Proc Natl Acad<br />

Sci USA 100,259-264.<br />

Lieberman, A. P. (2004). SUMO, a ubiquitin-like modifier<br />

neurodegeneration.<br />

Exp Neurol 185,204-207.<br />

implicated in<br />

Lin, R. J., <strong>St</strong>ernsdorf, T., Tini, M., and Evans, R. M. (2001). Transcriptional<br />

regulation in acute promyelocytic leukemia. Oncogene 49,7204-7215.<br />

222


Lin, D., Tatham, M. H., Yu, B., Kim, S., Hay, R. T., and Chen, Y. (2002).<br />

Identification <strong>of</strong> a substrate recognition site on Ubc9. J Biol Chem 277,21740-<br />

21748.<br />

Lin, X., Liang, M., Liang, Y. Y., Brunicardi, F. C., and Feng, X. H. (2003a).<br />

SUMO-1/Ubc9 promotes nuclear accumulation and metabolic stability <strong>of</strong> tumor<br />

suppressor Smad4. J Biol Chem 278,31043-31048.<br />

Lin, X., Sun, B., Liang, M., Liang, Y. Y., Gast, A., Hildebrand, J., Brunicardi, F. C.,<br />

Melchior, F., and Feng, X. H. (2003b). Opposed regulation <strong>of</strong> corepressor CtBP by<br />

SUMOylation and PDZ binding. Mol Cell 11,1389-1396.<br />

Ling, Y., Sankpal, U. T., Robertson, A. K., McNally, J. G., Karpova, T., and<br />

Robertson, K. D. (2004). Modification <strong>of</strong> de novo DNA methyltransferase 3a<br />

(Dnmt3a) by SUMO-1 modulates its interaction with histone deacetylases<br />

(HDACs)<br />

and its capacity to repress transcription. Nucleic Acids Res 32,598-610.<br />

Linnen, J. M., Bailey, C. P., and Weeks, D. L. (1993). Two related localized mRNAs<br />

from Xenopus laevis encode ubiquitin-like fusion proteins. Gene 128,181-188.<br />

Liu, B., Liao, J., Rao, X., Kushner, S. A., Chung, C. D., Chang, D. D., and Shuai, K.<br />

(1998). Inhibition <strong>of</strong> <strong>St</strong>atl-mediated gene activation by PIAS 1. Proc Natl Acad Sci<br />

USA 95,10626-10631.<br />

Liu, J., Furukawa, M., Matsumoto, T., and Xiong, Y. (2002). NEDD8 modification<br />

<strong>of</strong> CULl dissociates p120(CAND1), an inhibitor <strong>of</strong> CUL1-SKP1 binding and SCF<br />

ligases. Mol Cell 10,1511-1518.<br />

Liu, Q., Jin, C., Liao, X., Shen, Z., Chen, D. J., and Chen, Y. (1999a). The binding<br />

interface between an E2 (UBC9) and a ubiquitin homologue (UBL1). J Biol Chem<br />

274,16979-16987.<br />

223


Liu, Y. C., Pan, J., Zhang, C., Fan, W., Collinge, M., Bender, J. R., and Weissman,<br />

S. M. (1999b). A MHC-encoded ubiquitin-like protein (FAT10) binds noncovalently<br />

to the spindle assembly checkpoint protein MAD2. Proc Natl Acad Sci USA 96,<br />

4313-4318.<br />

Livengood, J. A., Scoggin, K. E., Van Orden, K., McBryant, S. J.,<br />

Edayathumangalam, R. S., Laybourn, P. J., and Nyborg, J. K. (2002). p53<br />

Transcriptional activity is mediated through the SRC 1-interacting domain <strong>of</strong><br />

CBP/p300. J Biol Chem 277,9054-9061.<br />

Long, X., and Griffith, L. C. (2000). Identification and characterization <strong>of</strong> a SUMO-<br />

1 conjugation system that modifies neuronal calcium/calmodulin-dependent protein<br />

kinase II in Drosophila melanogaster. J Biol Chem 275,40765-40776.<br />

Lonze, B. E., and Ginty, D. D. (2002). Function and regulation <strong>of</strong><br />

transcription factors in the nervous system. Neuron 35,605-623.<br />

CREB family<br />

Luders, J., Pyrowolakis, G., and Jentsch, S. (2003). The ubiquitin-like protein HUB 1<br />

forms SDS-resistant complexes with cellular proteins in the absence <strong>of</strong> ATP. EMBO<br />

Rep 4,1169-1174.<br />

Lyapina, S., Cope, G., Shevchenko, A., Serino, G., Tsuge, T., Zhou, C., Wolf, D. A.,<br />

Wei, N., and Deshaies, R. J. (2001). Promotion <strong>of</strong> NEDD-CUL1 conjugate cleavage<br />

by COPS signalosome. Science 292,1382-1385.<br />

Mahajan, R., Delphin, C., Guan, T., Gerace, L., and Melchior, F. (1997). A small<br />

ubiquitin-related polypeptide involved in targeting RanGAP 1 to nuclear pore<br />

complex protein<br />

RanBP2. Cell 88,97-107.<br />

224


Malakhov, M. P., Kim, K. I., Malakhova, 0. A., Jacobs, B. S., Borden, E. C., and<br />

Zhang, D. E. (2003). High-throughput immunoblotting. Ubiquitiin-like protein<br />

ISG15 modifies key regulators <strong>of</strong> signal transduction. J Biol. Chem 278,16608-<br />

16613.<br />

Malakhova, 0. A., Yan, M., Malakhov, M. P., Yuan, Y., Ritchie, K. J., Kim, K. I.,<br />

Peterson, L. F., Shuai, K., and Zhang, D. E. (2003). Protein ISGylation modulates<br />

the JAK-STAT signaling pathway. Genes Dev 17,455-460.<br />

Mangel, W. F., Toledo, D. L., Brown, M. T., Martin, J. H., and McGrath, W. J.<br />

(1996). Characterization <strong>of</strong> three components <strong>of</strong> human adenovirus proteinase<br />

activity in vitro. J Biol Chem 271,536-543.<br />

Mannen, H., Tseng, H. M., Cho, C. L., and Li, S. S. (1996). Cloning and expression<br />

<strong>of</strong> human homolog HSMT3 to yeast SMT3 suppressor <strong>of</strong> MIF2 mutations in a<br />

centromere protein gene. Biochem Biophys Res Commun 222,178-180.<br />

Mao, Y., Desai, S. D., and Liu, L. F. (2000a). SUMO-1 conjugation to human DNA<br />

topoisomerase II isozymes. J Biol Chem 275,26066-26073.<br />

Mao, Y., Sun, M., Desai, S. D., and Liu, L. F. (2000b). SUMO-1 conjugation to<br />

topoisomerase I: A possible repair response to topoisomerase-mediated DNA<br />

damage. Proc Natl Acad Sci USA 97,4046-4051.<br />

Matsuzaki, K., Minami, T., Tojo, M., Honda, Y., Uchimura, Y., Saitoh, H., Yasuda,<br />

H., Nagahiro, S., Saya, H., and Nakao, M. (2003). Serum response factor is<br />

modulated by the SUMO-1 conjugation system. Biochem Biophys Res Commun<br />

306,32-38.<br />

Matunis, M. J., Coutavas, E., and Blobel, G. (1996). A novel ubiquitin-like<br />

modification modulates the partitioning <strong>of</strong> the Ran-GTPase-activating protein<br />

225


RanGAP1 between the cytosol and the nuclear pore complex. J Cell Biol 135,1457-<br />

1470.<br />

Maul, G. G., Yu, E., Ishov, A. M., and Epstein, A. L. (1995). Nuclear domain 10<br />

(ND10) associated proteins are also present in nuclear bodies and redistribute to<br />

hundreds <strong>of</strong> nuclear sites after stress. J Cell Biochem 59,498-513.<br />

McLaughlin, P. M., Helfrich, W., Kok, K., Mulder, M., Hu, S. W., Brinker, M. G.,<br />

Ruiters, M. H., de Leij, L. F., and Buys, C. H. (2000). The ubiquitin-activating<br />

enzyme El-like protein in lung cancer cell lines. Int J Cancer 85,871-876.<br />

Melchior, F. (2000). SUMO--nonclassical ubiquitin. Annu Rev Cell Dev Biol 16,<br />

591-626.<br />

Meluh, P. B., and Koshland, D. (1995). Evidence that the MIF2 gene <strong>of</strong><br />

Saccharomyces<br />

cerevisiae encodes a centromere protein with homology to the<br />

mammalian centromere protein CENP-C. Mol Biol Cell 6,793-807.<br />

Mendoza, H. M., Shen, L. N., Botting, C., Lewis, A., Chen, J., Ink, B., and Hay, R.<br />

T. (2003). NEDP1, a highly conserved cysteine protease that deNEDDylates Cullins.<br />

J Biol Chem 278,25637-25643.<br />

Mimnaugh, E. G., Kayastha, G., McGovern, N. B., Hwang, S. G., Marcu, M. G.,<br />

Trepel, J., Cai, S. Y., Marchesi, V. T., and Neckers, L. (2001). Caspase-dependent<br />

deubiquitination <strong>of</strong> monoubiquitinated nucleosomal histone H2A induced by diverse<br />

apoptogenic stimuli. Cell Death Differ 8,1182-1196.<br />

Minty, A., Dumont, X., Kaghad, M., and Caput, D. (2000). Covalent modification <strong>of</strong><br />

p73alpha by SUMO-1. Two-hybrid screening with p73 identifies novel SUMO-1-<br />

interacting proteins and a SUMO-1 interaction motif. J Biol Chem 275,36316-<br />

36323.<br />

226


Mizushima, N., Noda, T., Yoshimori, T., Tanaka, Y., Ishii, T., George, M. D.,<br />

Klionsky, D. J., Ohsumi, M., and Ohsumi, Y. (1998). A protein conjugation system<br />

essential for autophagy. Nature 395,395-398.<br />

Moore, M. S., and Blobel, G. (1994). Purification <strong>of</strong> a Ran-interacting protein that is<br />

required for protein import into the nucleus. Proc Natl Acad Sci USA 91,10212-<br />

10216.<br />

Mu, Z. M., Chin, K. V., Liu, J. H., Lozano, G., and Chang, K. S. (1994). PML, a<br />

growth suppressor disrupted in acute promyelocytic leukemia. Mol Cell Biol 14,<br />

6858-6867.<br />

Muller, S., Berger, M., Lehembre, F., Seeler, J. S., Haupt, Y., and Dejean, A. (2000).<br />

c-Jun and p53 activity is modulated by SUMO-1 modification. J Biol Chem 275,<br />

13321-13329.<br />

Muller, S., Ledl, A., and Schmidt, D. (2004). SUMO: a regulator <strong>of</strong> gene expression<br />

and genome integrity. Oncogene 23,1998-2008.<br />

Muraoka, M., Konishi, M., Kikuchi-Yanoshita, R., Tanaka, K., Shitara, N., Chong,<br />

J. M., Iwama, T., Miyaki, M. (1996). p300 gene alterations in colorectal and gastric<br />

carcinomas 12,1565-1569.<br />

Naar, A. M., Lemon, B. D., and Tjian, R. (2001). Transcriptional coactivator<br />

complexes. Annu Rev Biochem 70,475-501.<br />

Nakagawa, K., and Yokosawa, H. (2002). PIAS3 induces SUMO-1 modification and<br />

transcriptional repression <strong>of</strong> IRF-1. FEBS Lett 530,204-208.<br />

227


Narasimhan, J., Potter, J. L., and Haas, A. L. (1996). Conjugation <strong>of</strong> the 15-kDa<br />

interferon-induced ubiquitin homolog is distinct from that <strong>of</strong> ubiquitin. J Biol Chem<br />

271,324-330.<br />

Narlikar, G. J., Fan, H. Y., and Kingston, R. E. (2002). Cooperation between<br />

complexes that regulate chromatin structure and transcription. Cell 108,475-487.<br />

Negorev, D., Ishov, A. M., and Maul, G. G. (2001). Evidence for separate ND 10-<br />

binding and homo-oligomerization domains <strong>of</strong> SplOO. J Cell Sci 114,59-68.<br />

Negorev, D., and Maul, G. G. (2001). Cellular proteins localized at and interacting<br />

within ND1O/PML nuclear bodies/PODs suggest functions <strong>of</strong> a nuclear depot.<br />

Oncogene 20,7234-7242.<br />

Nelson, D. C., Lasswell, J., Rogg, L. E., Cohen, M. A., and Bartel, B. (2000). FKF1,<br />

a clock-controlled gene that regulates the transition to flowering in Arabidopsis. Cell<br />

101,331-340.<br />

Nishida, T., Kaneko, F., Kitagawa, M., and Yasuda, H. (2001). Characterisation <strong>of</strong> a<br />

novel mammalian SUMO-1/Smt3-specific isopeptidase, a homologue <strong>of</strong> rat axam,<br />

which is an axin-binding protein promoting beta-catenin degradation. J Bio Chem<br />

42,39060-39066.<br />

Ogawa, H., Ishiguro, K., Gaubatz, S., Livingston, D. M., and Nakatani, Y. (2002). A<br />

complex with chromatin modifiers that occupies E2F- and Myc-responsive genes in<br />

GO cells. Science 296,1132-1136.<br />

Ohsumi, Y. (2001). Molecular dissection <strong>of</strong> autophagy: two ubiquitin-like systems.<br />

Nat Rev Mol Cell Biol 2,211-216.<br />

228


Ohtsubo, M., Kai, R., Furuno, N., Sekiguchi, T., Sekiguchi, M., Hayashida, H.,<br />

Kuraa, K., Miyata, T., Fukushige, S., Murotsu, T., and et al. (1987). Isolation and<br />

characterization <strong>of</strong> the active cDNA <strong>of</strong> the human cell cycle gene (RCC1) involved<br />

in the regulation <strong>of</strong> onset <strong>of</strong> chromosome condensation. Genes Dev 1,585-593.<br />

Okura, T., Gong, L., Kamitani, T., Wada, T., Okura, I., Wei, C. F., Chang, H. M.,<br />

and Yeh, E. T. (1996). Protection against Fas/APO-1- and tumor necrosis factor-<br />

mediated cell death by a novel protein, sentrin. J Immunol 157,4277-4281.<br />

Osaka, F., Kawasaki, H., Aida, N., Saeki, M., Chiba, T., Kawashima, S., Tanaka, K.,<br />

and Kato, S. (1998). A new NEDD8-ligating system for cullin-4A. Genes Dev 12,<br />

2263-2268.<br />

Ottosen, S., Herrera, F. J., and Triezenberg, S. J. (2002). Transcription. Proteasome<br />

parts at gene promoters. Science 296,479-481.<br />

Owhashi, M., Taoka, Y., Ishii, K., Nakazawa, S., Uemura, H., and Kambara, H.<br />

(2003). Identification <strong>of</strong> a ubiquitin family protein as a novel neutrophil chemotactic<br />

factor. Biochem Biophys Res Commun 309,533-539.<br />

Ozkaynak, E., Finley, D., and Varshavsky, A. (1984). The yeast ubiquitin gene:<br />

head-to-tail repeats encoding a polyubiquitin precursor protein. Nature 312,663-<br />

666.<br />

Palombella, V. J., Rando, 0. J., Goldberg, A. L., and Maniatis, T. (1994). The<br />

ubiquitin-proteasome pathway is required for processing the NF-kappa B1 precursor<br />

protein and the activation <strong>of</strong> NF-kappa B. Cell 78,773-785.<br />

Peng, J., Schwartz, D., Elias, J. E., Thoreen, C. C., Cheng, D., Marsischky, G.,<br />

Roel<strong>of</strong>s, J., Finley, D., and Gygi, S. P. (2003). A proteomics approach to<br />

understanding protein ubiquitination. Nat Biotechnol 21,921-926.<br />

229


Perkins, N. D., Felzien, L. K., Betts, J. C., Leung, K., Beach, D. H., and Nabel, G. J.<br />

(1997). Regulation <strong>of</strong> NF-kappaB by cyclin-dependent kinases associated with the<br />

p300 coactivator. Science 275,523-527.<br />

Petrij, F., Giles, R. H., Dauwerse, H. G., Saris, J. J., Hennekam, R. C., Masuno, M.,<br />

Tommerup, N., van Ommen, G. J., Goodman, R. H., Peters, D. J., and et al. (1995).<br />

Rubinstein-Taybi syndrome caused by mutations in the transcriptional co-activator<br />

CBP. Nature 376,348-351.<br />

Pham, A. D., and Sauer, F. (2000). Ubiquitin-activating/conjugating activity <strong>of</strong><br />

TAFII250, a mediator <strong>of</strong> activation <strong>of</strong> gene expression in Drosophila. Science 289,<br />

2357-2360.<br />

Pichler, A., Gast, A., Seeler, J. S., Dejean, A., and Melchior, F. (2002). The<br />

nucleoporin RanBP2 has SUMO1 E3 ligase activity. Cell 108,109-120.<br />

Pickart, C. M. (2002). DNA repair: right on target with ubiquitin. Nature 419,120-<br />

121.<br />

Pitterle, D. M., and Rajagopalan, K. V. (1993). The biosynthesis <strong>of</strong> molybdopterin in<br />

Escherichia coli. Purification and characterization <strong>of</strong> the converting factor. J Biol<br />

Chem 268,13499-13505.<br />

Potter, J. L., Narasimhan, J., Mende-Mueller, L., and Haas, A. L. (1999). Precursor<br />

processing <strong>of</strong> pro-ISG 15IUCRP, an interferon-beta-induced ubiquitin-like protein. J<br />

Biol Chem 274,25061-25068.<br />

Poukka, H., Karvonen, U., Janne, 0. A., and Palvimo, J. J. (2000). Covalent<br />

modification <strong>of</strong> the androgen receptor by small ubiquitin-like modifier 1 (SUMO-1).<br />

Proc Natl Acad Sci USA 97,14145-14150.<br />

230


Pountney, D. L., Huang, Y., Bums, R. J., Haan, E., Thompson, P. D., Blumbergs, P.<br />

C., and Gai, W. P. (2003). SUMO-1 marks the nuclear inclusions in familial<br />

neuronal intranuclear inclusion disease. Exp Neurol 184,436-446.<br />

Raasi, S., Schmidtke, G., de Giuli, R., and Groettrup, M. (1999). A ubiquitin-like<br />

protein which is synergistically inducible by interferon-gamma and tumor necrosis<br />

factor-alpha. Eur J Immunol 29,4030-4036.<br />

Rangasamy, D., and Wilson, V. G. (2000). Bovine papillomavirus El protein is<br />

sumoylated by the host cell LJbc9 protein.<br />

J Biol Chem 275,30487-30495.<br />

Read, M. A., Brownell, J. E., Gladysheva, T. B., Hottelet, M., Parent, L. A.,<br />

Coggins, M. B., Pierce, J. W., Podust, V. N., Luo, R. S., Chau, V., and Palombella,<br />

V. J. (2000). Nedd8 modification <strong>of</strong> cul-1 activates SCF(beta(TrCP))-dependent<br />

ubiquitination <strong>of</strong> IkappaBalpha. Mol Cell Biol 20,2326-2333.<br />

Rechsteiner, M. (1988). Regulation <strong>of</strong> enzyme levels by proteolysis: the role <strong>of</strong> pest<br />

regions. Adv Enzyme Regul 27,135-151.<br />

Reid, G., Hubner, M. R., Metivier, R., Brand, H., Denger, S., Manu, D., Beaudouin,<br />

J., Ellenberg, J., and Gannon, F. (2003). Cyclic, proteasome-mediated turnover <strong>of</strong><br />

unliganded and liganded ERalpha on responsive promoters is an integral feature <strong>of</strong><br />

estrogen signaling. Mol Cell 11,695-707.<br />

Robzyk, K., Recht, J., and Osley, M. A. (2000). Rad6-dependent ubiquitination <strong>of</strong><br />

histone H2B in yeast. Science 287,501-504.<br />

Rochat-<strong>St</strong>einer, V., Becker, K., Micheau, 0., Schneider, P., Burns, K., and Tschopp,<br />

J. (2000). FIST/HIPK3: a Fas/FADD-interacting serine/threonine kinase that induces<br />

FADD phospho y1ation and inhibits fas-mediated Jun NH(2)-terminal kinase<br />

activation. J Exp Med 192,1165-1174.<br />

231


Rodriguez, J. M., Salas, M. L., and Vinuela, E. (1992). Genes homologous to<br />

ubiquitin-conjugating proteins and eukaryotic transcription factor SII in African<br />

swine fever virus.<br />

Virology 186,40-52.<br />

Rodriguez, M. S., Dargemont, C., and Hay, R. T. (2001). SUMO-1 conjugation in<br />

vivo requires both a consensus modification motif and nuclear targeting. J Biol<br />

Chem 276,12654-12659.<br />

Rodriguez, M. S., Desterro, J. M., Lain, S., Midgley, C. A., Lane, D. P., and Hay, R.<br />

T. (1999). SUMO-1 modification activates the transcriptional response <strong>of</strong> p53.<br />

Embo J 18,6455-6461.<br />

Rodriguez, M. S., Wright, J., Thompson, J., Thomas, D., Baleux, F., Virelizier, J. L.,<br />

Hay, R. T., and Arenzana-Seisdedos, F. (1996). Identification <strong>of</strong> lysine residues<br />

required for signal-induced ubiquitination and degradation <strong>of</strong> I kappa B-alpha in<br />

vivo. Oncogene 12,2425-2435.<br />

Ross, S., Best, J. L., Zon, L. I., and Gill, G. (2002). SUMO-1 modification represses<br />

Spa transcriptional activation and modulates its subnuclear localization. Mol Cell<br />

10,831-842.<br />

Rouaux, C., Jokic, N., Mbebi, C., Boutillier, S., Loeffler, J. P., and Boutillier, A. L.<br />

(2003). Critical loss <strong>of</strong> CBP/p300 histone acetylase activity by caspase-6 during<br />

neurodegeneration. Embo J 22,6537-6549.<br />

Rudolph, M. J., Wuebbens, M. M., Rajagopalan, K. V., and Schindelin, H. (2001).<br />

Crystal structure <strong>of</strong> molybdopterin synthase and its evolutionary relationship to<br />

ubiquitin activation. Nat <strong>St</strong>ruct Biol 8,42-46.<br />

232


Rui, H. L., Fan, E., Zhou, H. M., Xu, Z., Zhang, Y., and Lin, S. C. (2002). SUMO-1<br />

modification <strong>of</strong> the C-terminal KVEKVD <strong>of</strong> Axin is required for JNK activation but<br />

has no effect on Wnt signaling. J Biol Chem 277,42981-42986.<br />

Ruthardt, M., Orleth, A., Tomassoni, L., Puccetti, E., Riganelli, D., Alcalay, M.,<br />

Mannucci, R., Nicoletti, I., Grignani, F., Fagioli, M., and Pelicci, P. G. (1998). The<br />

acute promyelocytic leukaemia specific PML and PLZF proteins localize to adjacent<br />

and functionally distinct nuclear bodies. Oncogene 16,1945-1953.<br />

Ryu, S. W., Chae, S. K., and Kim, E. (2000). Interaction <strong>of</strong> Daxx, a Fas binding<br />

protein, with sentrin and Ubc9. Biochem Biophys Res Commun 279,6-10.<br />

Sachdev, S., Bruhn, L., Sieber, H., Pichler, A., Melchior, F., and Grosschedl, R.<br />

(2001). PIASy, a nuclear matrix-associated SUMO E3 ligase, represses LEF1<br />

activity by sequestration into nuclear bodies. Genes Dev 15,3088-3103.<br />

Saitoh, H., and Hinchey, J. (2000). Functional heterogeneity <strong>of</strong> small ubiquitin-<br />

related protein modifiers SUMO-1 versus SUMO-2/3. J Biol Chem 275,6252-6258.<br />

Salina, D., Enarson, P., Rattner, J. B., and Burke, B. (2003). Nup358 integrates<br />

nuclear envelope breakdown with kinetochore assembly. J Cell Biol 162,991-1001.<br />

Santos-Rosa, H., Schneider, R., Bannister, A. J., Sherriff, J., Bernstein, B. E., Emre,<br />

N. C., Schreiber, S. L., Mellor, J., and Kouzarides, T. (2002). Active genes are tri.<br />

methylated at K4 <strong>of</strong> histone H3. Nature 419,407-411.<br />

Sapetschnig, A., Rischitor, G., Braun, H., Doll, A., Schergaut, M., Melchior, F., and<br />

Suske, G. (2002). Transcription factor Sp3 is silenced through SUMO modification<br />

by PIAS 1. EMBO J 21,5206-5215.<br />

233


Satijn, D. P., Hamer, K. M., den Blaauwen, J., and Otte, A. P. (2001). The polycomb<br />

group protein EED interacts with YY1, and both proteins induce neural tissue in<br />

Xenopus embryos. Mol Cell Biol 21,1360-1369.<br />

Saurin, A. J., Shiels, C., <strong>William</strong>son, J., Satijn, D. P., Otte, A. P., Sheer, D., and<br />

Freemont, P. S. (1998). The human polycomb group complex associates with<br />

pericentromeric heterochromatin to form a novel nuclear domain. J Cell Biol 142,<br />

887-898.<br />

Schaeper, U., Subramanian, T., Lim, L., Boyd, J. M., and Chinnadurai, G. (1998).<br />

Interaction between a cellular protein that binds to the C-terminal region <strong>of</strong><br />

adenovirus E1A (CtBP) and a novel cellular protein is disrupted by E1A through a<br />

conserved PLDLS motif. J Biol Chem 273,8549-8552.<br />

Scherer, D. C., Brockman, J. A., Chen, Z., Maniatis, T., and Ballard, D. W. (1995).<br />

Signal-induced degradation <strong>of</strong> I kappa B alpha requires site-specific ubiquitination.<br />

Proc Natl Acad Sci USA 92,11259-11263.<br />

Schmidt, D. and Muller, S. (2003). PIAS/SUMO: new partners in transcriptional<br />

' regulation. Cell Mol Life Sci 12,2561-2574.<br />

Schumacher, A., and Magnuson, T. (1997). Murine Polycomb- and trithorax-group<br />

genes regulate homeotic pathways and beyond. Trends Genet 13,167-170.<br />

Schwartz, D. C., and Hochstrasser, M. (2003). A superfamily <strong>of</strong> protein tags:<br />

ubiquitin, SUMO and related modifiers.<br />

Trends Biochem Sci 28,321-328.<br />

Schwarz, S. E., Matuschewski, K., Liakopoulos, D., Scheffner, M., and Jentsch, S.<br />

(1998). The ubiquitin-like proteins SMT3 and SUMO-1 are conjugated by the UBC9<br />

E2 enzyme. Proc Natl Acad Sci USA 95,560-564.<br />

234


Schwechheimer, C., and Deng, X. W. (2001). COP9 signalosome revisited: a novel<br />

mediator <strong>of</strong> protein degradation. Trends Cell Biol 11,420-426.<br />

Seale, R. L. (1981). Site <strong>of</strong> histone assembly. Prog Nucleic Acid Res Mol Biol 26,<br />

123-134.<br />

Seeler, J. S., and Dejean, A. (1999). The PML nuclear bodies: actors or extras? Curr<br />

Opin Genet Dev 9,362-367.<br />

Seeler, J. S., Marchio, A., Losson, R., Desterro, J. M., Hay, R. T., Chambon, P., and<br />

Dejean, A. (2001). Common properties <strong>of</strong> nuclear body protein SP100 and<br />

TIFlalpha chromatin factor: role <strong>of</strong> SUMO modification. Mol Cell Biol 21,3314-<br />

3324.<br />

Seeler, J. S., Marchio, A., Sitterlin, D., Transy, C., and Dejean, A. (1998).<br />

Interaction <strong>of</strong> SP100 with HP! proteins: a link between the promyelocytic leukemia-<br />

associated nuclear bodies and the chromatin compartment. Proc Natl Acad Sci US<br />

A 95,7316-7321.<br />

Seigneurin-Berny, D., Verdel, A., Curtet, S., Lemercier, C., Rousseaux, S., and<br />

Knochbin, S. (2001). Identification <strong>of</strong> components <strong>of</strong> the murine histone<br />

deacetylase 6 complex: link between acetylation and ubiquitination signaling<br />

pathways. Mol Cell Bio 23,8035-8044.<br />

Sekizawa, R., Ikeno, S., Nakamura, H., Naganawa, H., Matsui, S., Iinuma, H., and<br />

Takeuchi, T. (2002). Panepophenanthrin, from a mushroom strain, a novel inhibitor<br />

<strong>of</strong> the ubiquitin-activating enzyme. J Nat Prod 65,1491-1493.<br />

Seufert, W., and Jentsch, S. (1990a). Nucleotide sequence <strong>of</strong> two tRNA(Arg)-<br />

tRNA(Asp) tandem genes linked to duplicated UBC genes in Saccharomyces<br />

cerevisiae. Nucleic Acids Res 18,1638.,<br />

235


Seufert, W., and Jentsch, S. (1990b). Ubiquitin-conjugating enzymes UBC4 and<br />

UBC5 mediate selective degradation <strong>of</strong> short-lived and abnormal proteins. Embo J<br />

9,543-550.<br />

Sewalt, R. G., van der Vlag, J., Gunster, M. J., Hamer, K. M., den Blaauwen, J. L.,<br />

Satijn, D. P., Hendrix, T., van Driel, R., and Otte, A. P. (1998). Characterization <strong>of</strong><br />

interactions between the mammalian polycomb-group proteins Enx1/EZH2 and EED<br />

suggests the existence <strong>of</strong> different mammalian polycomb-group protein complexes.<br />

Mol Cell Biol 18,3586-3595.<br />

Shang, F., Gong, X., and Taylor, A. (1997). Activity <strong>of</strong> ubiquitin-dependent pathway<br />

in response to oxidative stress. Ubiquitin-activating enzyme is transiently up-<br />

regulated. J Biol Chem 272,23086-23093.<br />

Shen, Z., Pardington-Purtymun, P. E., Comeaux, J. C., Moyzis, R. K., and Chen, D.<br />

J. (1996a). Associations <strong>of</strong> UBE2I with RAD52, UBL1, p53, and RAD51 proteins in<br />

a yeast two-hybrid system. Genomics 37,183-186.<br />

Shen, Z., Pardington-Purtymun, P. E., Comeaux, J. C., Moyzis, R. K., and Chen, D.<br />

J. (1996b). UBL1, a human ubiquitin-like protein associating with human RAD<br />

51/RAD52 proteins. Genomics 36,271-279., E<br />

Sherr, C. J. (1998). Tumor surveillance via the ARF-p53 pathway. Genes Dev 12,<br />

2984-2991.<br />

Shi, Y., Seto, E., Chang, L. S., and Shenk, T. (1991). Transcriptional repression by<br />

YY1, a human GLI-Kruppel-related protein, and relief <strong>of</strong> repression by adenovirus<br />

EIA protein. Cell 67,377-388.<br />

Shiio, Y., and Eisenman, R. N. (2003). Histone sumoylation is associated with<br />

transcriptional repression. Proc Natl Acad Sci USA 100,13225-13230.<br />

236


Simon, J., Bornemann, D., Lunde, K., and Schwartz, C. (1995). The extra sex combs<br />

product contains WD40 repeats and its time <strong>of</strong> action implies a role<br />

other Polycomb group products. Mech Dev 53,197-208.<br />

distinct from<br />

Skinner, P. J., Koshy, B. T., Cummings, C. J., Klement, I. A., Helin, K., Servadio,<br />

A., Zoghbi, H. Y., and Orr, H.. T. (1997). Ataxin-1 with an expanded glutamine tract<br />

alters nuclear matrix-associated structures.<br />

Nature 389,971-974.<br />

Smits, P. H., de Wit, L., van der Eb, A. J., Zantema, A. (1996). The adenovirus E1A-<br />

associated 300 kDa adaptor counteracts the inhibition <strong>of</strong> the collagenase promoter<br />

by E1A and represses transformation. Oncogene 12,1529-1535.<br />

Snowden, A. W., Anderson, L. A., Webster, G. A., and Perkins, N. D. (2000). A<br />

novel transcriptional repression domain mediates p21(WAF1/CIP1) induction <strong>of</strong><br />

p300 transactivation. Mol Cell Biol 20,2676-2686.<br />

Sobko, A., Ma, H., and Firtel, R. A. (2002). Regulated SUMOylation and<br />

ubiquitination <strong>of</strong> DdMEKI is required for proper chemotaxis.<br />

Dev Cell 2,745-756.<br />

Sommer, T., and Jentsch, S. (1993). A protein translocation defect linked to<br />

ubiquitin conjugation at the endoplasmic reticulum. Nature 365,176-179.<br />

Spence, J., Gali, R. R., Dittmar, G., Sherman, F., Karin, M., and Finley, D. (2000).<br />

Cell cycle-regulated modification <strong>of</strong> the ribosome by a variant multiubiquitin chain.<br />

Cell 102,67-76.<br />

Spengler, M. L., Kurapatwinski, K., Black, A. R., and Azizkhan-Clifford, J. (2002).<br />

SUMO-1 modification <strong>of</strong> human cytomegalovirus IE1/IE72. J Virol 76,2990-2996.<br />

237


<strong>St</strong>ead, K., Aguilar, C., Hartman, T., Drexel, M., Meluh, P., and Guacci, V. (2003).<br />

Pds5p regulates the maintenance <strong>of</strong> sister chromatid cohesion and is sumoylated to<br />

promote the dissolution <strong>of</strong> cohesion. J Cell Biol 163,729-741.<br />

<strong>St</strong>effan, J. S., Agrawal, N., Pallos, J., Rockabrand, E., Trotman, L. C., Slepko, N.,<br />

Illes, K., Lukacsovich, T., Zhu, Y. Z., Cattaneo, E., et al. (2004). SUMO<br />

modification <strong>of</strong> Huntingtin and Huntington's disease pathology. Science 304,100-<br />

104.<br />

<strong>St</strong>ein, R. W., Corrigan, M., Yaciuk, P., Whelan, J., and Moran, E. (1990). Analysis<br />

<strong>of</strong> EIA-mediated growth regulation functions: binding <strong>of</strong> the 300-kilodalton cellular<br />

product correlates with E1A enhancer repression function and DNA synthesis-<br />

inducing activity. J Virol 64,4421-4427.<br />

<strong>St</strong>elter, P., and Ulrich, H. D. (2003). Control <strong>of</strong> spontaneous and damage-induced<br />

mutagenesis by SUMO and ubiquitin conjugation. Nature 425,188-191.<br />

<strong>St</strong>ephen, A. G., Trausch-Azar, J. S., Ciechanover, A., and Schwartz, A. L. (1996).<br />

The ubiquitin-activating enzyme El is phosphorylated and localized to the nucleus<br />

in a cell cycle-dependent manner. J Biol Chem 271,15608-15614.<br />

<strong>St</strong>ephen, A. G., Trausch-Azar, J. S., Handley-Gearhart, P. M., Ciechanover, A., and<br />

Schwartz, A. L. (1997). Identification <strong>of</strong> a region within the ubiquitin-activating<br />

enzyme required for nuclear targeting and phosphorylation. J Biol Chem 272,<br />

10895-10903.<br />

<strong>St</strong>ernsdorf, T., Jensen, K., and Will, H. (1997). Evidence for covalent modification<br />

<strong>of</strong> the nuclear dot-associated proteins PML and SplOO by PIC1/SUMO-1. J Cell<br />

Biol 139,1621-1634.<br />

238


<strong>St</strong>ickle, N. H., Chung, J., Klco, J. M., Hill, R. P., Kaelin, W. G., Jr., and Ohh, M.<br />

(2004). pVHL modification by NEDD8 is required for fibronectin matrix assembly<br />

and suppression <strong>of</strong> tumor development. Mol Cell Biol 24,3251-3261.<br />

<strong>St</strong>ott, F. J., Bates, S., James, M. C., McConnell, B. B., <strong>St</strong>arborg, M., Brookes, S.,<br />

Palmero, I., Ryan, K., Hara, E., Vousden, K. H., and Peters, G. (1998). The<br />

alternative product from the human CDKN2A locus, p14(ARF), participates in a<br />

regulatory feedback loop with p53 and MDM2. Embo J 17,5001-5014.<br />

<strong>St</strong>rahl, B. D., and Allis, C. D. (2000). The language <strong>of</strong> covalent histone<br />

modifications. Nature 403,41-45.<br />

Sullivan, M. L., and Vierstra, R. D. (1991). Cloning <strong>of</strong> a 16-kDä ubiquitin carrier<br />

protein from wheat and Arabidopsis thaliana. Identification <strong>of</strong> functional domains by<br />

in vitro mutagenesis. J Biol Chem 266,23878-23885.<br />

Sun, Z. W., and Allis, C. D. (2002). Ubiquitination <strong>of</strong> histone H2B regulates H3<br />

methylation and gene silencing in yeast. Nature 418,104-108.<br />

Suzuki, H., Seki, M., Kobayashi, T., Kawabe, Y., Kaneko, H., Kondo, N., Harata,<br />

M., Mizuno, S., Masuko, T., and Enomoto, T. (2001). The N-terminal internal<br />

region <strong>of</strong> BLM is required for the formation <strong>of</strong> dots/rod-like structures which are<br />

associated with SUMO-1. Biochem Biophys Res Commun 286,322-327.<br />

Szostecki, C., Guldner, H. H., Netter, H. J., and Will, H. (1990). Isolation and<br />

characterization <strong>of</strong> cDNA encoding a human nuclear antigen predominantly<br />

recognized by autoantibodies from patients with primary biliary cirrhosis. J<br />

Immunol 145,4338-4347.<br />

Takahashi, Y., Iwase, M., Konishi, M., Tanaka, M., Toh-e, A., and Kikuchi, Y.<br />

(1999). Smt3, a SUMO-1 homolog, is conjugated to Cdc3, a component <strong>of</strong> septin<br />

239


ings at the mother-bud neck in budding yeast. Biochem Biophys Res Commun 259,<br />

582-587.<br />

Takahashi, Y., Kahyo, T., Toh, E. A., Yasuda, H., and Kikuchi, Y. (2001). Yeast<br />

Ulll/Sizl is a novel SUMO1/Smt3ligase for septin components and functions as an<br />

adaptor between conjugating enzyme and substrates. J Biol Chem 276,48973-<br />

48977.<br />

Tanaka, K., Nishide, J., Okazaki, K., Kato, H., Niwa, 0., Nakagawa, T., Matsuda,<br />

H., Kawamukai, M., and Murakami, Y. (1999). Characterization <strong>of</strong> a fission yeast<br />

SUMO-1 homologue, pmt3p, required for multiple nuclear events, including the<br />

control <strong>of</strong> telomere length and chromosome segregation. Mol Cell Biol 19,8660-<br />

8672.<br />

Tanaka, K., Suzuki, T., and Chiba, T. (1998). The ligation systems for ubiquitin and<br />

ubiquitin-like proteins. Mol Cells 8,503-512.<br />

Tatham, M. H., Jaffray, E., Vaughan, O. A., Desterro, J. M., Botting, C. H.,<br />

Naismith, J. H., and Hay, R. T. (2001). Polymeric chains <strong>of</strong> SUMO-2 and SUMO-3<br />

are conjugated to protein substrates by SAE1/SAE2 and Ubc9. J Biol Chem 276,<br />

35368-35374.<br />

Taylor, D. L., Ho, J. C., Oliver, A., and Watts, F. Z. (2002). Cell-cycle-dependent<br />

localisation <strong>of</strong> Ulpl, a Schizosaccharomyces<br />

pombe Pmt3 (SUMO)-specific<br />

protease. J Cell Sci 115,1113-1122.<br />

Terashima, T., Kawai, H., Fujitani, M., Maeda, K., and Yasuda, H. (2002). SUMO-1<br />

co-localized with mutant atrophin-1 with expanded polyglutamines accelerates<br />

intranuclear aggregation and cell<br />

death. Neuroreport 13,2359-2364.<br />

240


Tian, S., Poukka, H., Palvimo, J. J., and Janne, O. A. (2002). Small ubiquitin-related<br />

modifier-1 (SUMO-1) modification <strong>of</strong> the glucocorticoid receptor. Biochem J 367,<br />

907-911.<br />

Tojo, M., Matsuzaki, K., Minami, T., Honda, Y., Yasuda, H., Chiba, T., Saya, H.,<br />

Fujii-Kuriyama, Y., and Nakao, M. (2002). The aryl hydrocarbon receptor nuclear<br />

transporter is modulated by the SUMO-1 conjugation system. J Biol Chem 277,<br />

46576-46585.<br />

Tong, H., Hateboer, G., Perrakis, A., Bernards, R., and Sixma, T. K. (1997). Crystal<br />

structure <strong>of</strong> murine/human Ubc9 provides insight into the variability <strong>of</strong> the<br />

ubiquitin-conjugating system. J Biol Chem 272,21381-21387.<br />

Trost, M., Kochs, G., and Haller, 0. (2000). Characterization <strong>of</strong> a novel<br />

serine/threonine kinase associated with nuclear bodies. J Biol Chem 275,7373-7377.<br />

Tsytsykova, A. V., Tsitsikov, E. N., Wright, D. A., Futcher, B., and Geha, R. S.<br />

(1998). The mouse genome contains two expressed intronless retroposed<br />

pseudogenes for the sentrin/sumo-1/PIC1 conjugating enzyme Ubc9. Mol Immunol<br />

35,1057-1067.<br />

Ueda, H., Goto, J., Hashida, H., Lin, X., Oyanagi, K., Kawano, H., Zoghbi, H. Y.,<br />

Kanazawa, I., and Okazawa, H. (2002). Enhanced SUMOylation in polyglutamine<br />

diseases. Biochem Biophys Res Commun 293,307-313.<br />

Ulrich, H. D., and Jentsch, S. (2000). Two RING finger proteins mediate<br />

cooperation between ubiquitin-conjugating enzymes in DNA repair. Embo J 19,<br />

3388-3397.<br />

241


Vallian, S., Chin, K. V., and Chang, K. S. (1998). The promyelocytic leukemia<br />

protein interacts with Spl and inhibits its transactivation <strong>of</strong> the epidermal growth<br />

factor receptor promoter. Mol Cell Biol 18,7147-7156.<br />

Vallian, S., Gaken, J. A., Trayner, I. D., Gingold, E. B., Kouzarides, T., Chang, K.<br />

S., and Farzaneh, F. (1997). Transcriptional repression by the promyelocytic<br />

leukemia protein, PML. Exp Cell Res 237,371-382.<br />

VanDemark, A. P., and Hill, C. P. (2003). Two-stepping with El. Nat <strong>St</strong>ruct Biol 10,<br />

244-246.<br />

Verma, R., Aravind, L., Oania, R., McDonald, W. H., Yates, J. R., 3rd, Koonin, E.<br />

V., and Deshaies, R. J. (2002). Role <strong>of</strong> Rpnl 1 metalloprotease<br />

and degradation by the 26S proteasome. Science 298,611-615.<br />

in deubiquitination<br />

Vierstra, R. D., and Callis, J. (1999). Polypeptide tags, ubiquitous modifiers for plant<br />

protein regulation. Plant Mol Biol 41,435-442.<br />

Vijay-Kumar, S., Bugg, C. E., Wilkinson, K. D., Vierstra, R. D., Hatfield, P. M., and<br />

Cook, W. J. (1987). Comparison <strong>of</strong> the three-dimensional structures <strong>of</strong> human,<br />

yeast, and oat ubiquitin. J Biol Chem 262,6396-6399.<br />

von der Lehr, N., Johansson, S., Wu, S., Bahram, F., Castell, A., Cetinkaya, C.,<br />

Hydbring, P., Weidung, I., Nakayama, K., Nakayama, K. I., et al. (2003). The F-box<br />

protein Skp2 participates in c-Myc proteosomal degradation and acts as a c<strong>of</strong>actor<br />

for c-Myc-regulated transcription. Mol Cell 11,1189-1200.<br />

von Mikecz, A., Zhang, S., Montminy, M., Tan, E. M.,; and Hemmerich, P. (2000).<br />

CREB-binding protein (CBP)/p300 and RNA polymerase II colocalize in<br />

transcriptionally active domains in the nucleus. J Cell Biol 150,265-273.,<br />

.., ýý.,<br />

242


Wada, H., Yeh, E. T., and Kamitani, T. (1999). The von Hippel-Lindau tumor<br />

suppressor gene product promotes, but is not essential for, NEDD8 conjugation to<br />

cullin-2. J Biol Chem 274,36025-36029.<br />

Walden, H., Podgorski, M. S., and Schulman, B. A. (2003). Insights into the<br />

ubiquitin transfer cascade from the structure <strong>of</strong> the activating enzyme for NEDD8.<br />

Nature 422,330-334.<br />

Wang, C., Deng, L., Hong, M., Akkaraju, G. R., Inoue, J., and Chen, Z. J. (2001a).<br />

TAK1 is a ubiquitin-dependent kinase <strong>of</strong> MKK and IKK. Nature 412,346-351.<br />

Wang, C., Xi, J., Begley, T. P., and Nicholson, L. K. (2001b). Solution structure <strong>of</strong><br />

ThiS and implications for the evolutionary roots <strong>of</strong> ubiquitin. Nat <strong>St</strong>ruct Biol 8,47-<br />

51.<br />

Waterman, H., Levkowitz, G., Alroy, I., and Yarden, Y. (1999). The RING finger <strong>of</strong><br />

c-Cbl mediates desensitization <strong>of</strong> the epidermal growth factor receptor. J Biol Chem<br />

274,22151-22154.<br />

Webster, A., Hay, R. T., and Kemp, G. (1993). The adenovirus protease is activated<br />

by a virus-coded disulphide-linked peptide. Cell 72,97-104.<br />

Weissman, A. M. (2001). Themes and variations on ubiquitylation. Nat Rev Mol<br />

Cell Bio13,169-178.<br />

Wiborg, 0., Pedersen, M. S., Wind, A., Berglund, L. E., Marcker, K. A., and Vuust,<br />

J. (1985). The human ubiquitin multigene family: some genes contain multiple<br />

directly repeated ubiquitin coding sequences. Embo J 4,755-759. °.<br />

Wiebel, F. F., and Kunau, W. H. (1992). The Pas2 protein essential for peroxisome<br />

biogenesis is related to ubiquitin-conjugating enzymes. Nature 359,73-76.<br />

243


Wilkinson, C. R. (2002). New tricks for ubiquitin and friends. Trends Cell Biol 12,<br />

545-546.<br />

Wilkinson, C. R., Seeger, M., Hartmann-Petersen, R., <strong>St</strong>one, M., Wallace, M.,<br />

Semple, C., and Gordon, C. (2001). Proteins containing the UBA domain are able to<br />

bind to multi-ubiquitin chains. Nat Cell Bio13,939-943.<br />

Wu, P. Y., Hanlon, M., Eddins, M., Tsui, C., Rogers, R. S., Jensen, J. P., Matunis,<br />

M. J., Weissman, A. M., Wolberger, C. P., and Pickart, C. M. (2003). A conserved<br />

catalytic residue in the ubiquitin-conjugating enzyme family. Embo J 22,5241-5250.<br />

Xie, Y., and Varshavsky, A. (2001). RPN4 is a ligand, substrate, and transcriptional<br />

regulator <strong>of</strong> the 26S proteasome: a negative feedback circuit. Proc Natl Acad Sci U<br />

SA 98,3056-3061.<br />

Xirodimas, D. P., Chisholm, J., Desterro, J. M., Lane, D. P., and Hay, R. T. (2002).<br />

P14ARF promotes accumulation <strong>of</strong> SUMO-1 conjugated (H)Mdm2. FEBS Lett 528,<br />

207-211.<br />

Yamamoto, H., Ihara, M., Matsuura, Y., and Kikuchi, A. (2003). Sumoylation is<br />

involved in beta-catenin-dependent<br />

activation <strong>of</strong> Tcf-4. Embo J 22,2047-2059.<br />

Yang, S. H., Jaffray, E., Hay, R. T., and Sharrocks, A. D. (2003). Dynamic interplay<br />

<strong>of</strong> the SUMO and ERK pathways in regulating Elk-1 transcriptional activity. Mol<br />

Cell 12,63-74.<br />

Yang, S. H., and Sharrocks, A. D. (2004). SUMO promotes HDAC-mediated,<br />

transcriptional repression. Mol Cell 13,611-617. .<br />

.ý<br />

244


Yang, X. J., Ogryzko, V. V., Nishikawa, J., Howard, B. H., and Nakatani, Y. (1996).<br />

A p300/CBP-associated factor that competes with the adenoviral oncoprotein E1A.<br />

Nature 382,319-324.<br />

Yankiwski, V., Marciniak, R. A., Guarente, L., and Neff, N. F. (2000). Nuclear<br />

structure in normal and Bloom syndrome cells. Proc Nat! Acad Sci USA 97,5214-<br />

5219.<br />

Yao, T., and Cohen, R. E. (2002). A cryptic protease couples deubiquitination and<br />

degradation by the proteasome. Nature 419,403-407.<br />

Yao, T. P., Oh, S. P., Fuchs, M., Zhou, N. D., Ch'ng, L. E., Newsome, D., Bronson,<br />

R. T., Li, E., Livingston, D. M., and Eckner, R. (1998). Gene dosage-dependent<br />

embryonic development and proliferation defects in mice lacking the transcriptional<br />

integrator p300. Cell 93,361-372.<br />

Yashiroda, H., and Tanaka, K. (2003). Butl and But2 proteins bind to Uba3, a<br />

catalytic subunit <strong>of</strong> El for neddylation, in fission yeast. Biochem Biophys Res<br />

Commun 311,691-695.<br />

Yasugi, T., and Howley, P. M. (1996). Identification <strong>of</strong> the structural and functional<br />

human homolog <strong>of</strong> the yeast ubiquitin conjugating enzyme UBC9. Nucleic Acids<br />

Res 24,2005-2010.<br />

Yeh, E. T., Gong, L., and Kamitani, T. (2000). Ubiquitin-like proteins: new wines in<br />

new bottles. Gene 248,1-14.<br />

Yuan, W., Aramini, J. M., Montelione, G. T., and Krug, R. M. (2002). <strong>St</strong>ructural<br />

basis for ubiquitin-like ISG 15 protein binding to the NS 1 protein <strong>of</strong> influenza B<br />

virus: a protein-protein interaction function that is not shared by the corresponding<br />

N-terminal domain <strong>of</strong> the NS1 protein <strong>of</strong> influenza A virus. Virology 304,291-301.<br />

245


Yuan, W., Condorelli, G., Caruso, M., Felsani, A., and Giordano, A. (1996). Human<br />

p300 protein is a coactivator for the transcription factor MyoD. J Biol Chem 271,<br />

9009-9013.<br />

Yuan, W., and Krug, R. M. (2001). Influenza B virus NS1 protein inhibits<br />

conjugation <strong>of</strong> the interferon (IFN)-induced ubiquitin-like ISG15 protein. Embo J<br />

20,362-371.<br />

Yuan, Z. M., Huang, Y., Ishiko, T., Nakada, S., Utsugisawa, T., Shioya, H.,<br />

Utsugisawa, Y., Shi, Y., Weichselbaum, R., and Kufe, D. (1999). Function for p300<br />

and not CBP in the apoptotic response to DNA damage. Oncogene 18,5714-5717.<br />

Zhang, H., Saitoh, H., and Matunis, M. J. (2002). Enzymes <strong>of</strong> the SUMO<br />

modification pathway localize to filaments <strong>of</strong> the nuclear pore complex. Mol Cell<br />

Biol 22,6498-6508.<br />

Zhang, H., Smolen, G. A., Palmer, R., Christ<strong>of</strong>orou, A., Van Den Heuvel, S., and<br />

Haber, D. A. (2004). SUMO modification is required for in vivo Hox gene<br />

regulation by the Caenorhabditis elegans Polycomb group protein SOP-2. Nat Genet.<br />

Zhang, Q., Vo, N., and Goodman, R. H. (2000). Histone binding protein RbAp48<br />

interacts with a complex <strong>of</strong> CREB binding protein and phosphorylated CREB. Mol<br />

Cell Bio120,4970-4978.<br />

Zhang, Y., and Xiong, Y. (2001). Control <strong>of</strong> p53 ubiquitination and nuclear export<br />

by MDM2 and ARE Cell Growth Differ 12,175-186.<br />

Zhang, Y., Xiong, Y., and Yarbrough, W. G. (1998). ARF promotes MDM2<br />

degradation and stabilizes p53: ARF-INK4a locus deletion impairs both the Rb and<br />

p53 tumor suppression pathways. Cell 92,725-734.<br />

246


Zheng, J., Yang, X., Harrell, J. M., Ryzhikov, S., Shim, E. H., Lykke-Andersen, K.,<br />

Wei, N., Sun, H., Kobayashi, R., and Zhang, H. (2002). CAND1 binds to<br />

unneddylated CUL1 and regulates the formation <strong>of</strong> SCF ubiquitin E3 ligase<br />

complex. Mol Cell 10,1519-1526.<br />

247


7. PUBLICATIONS<br />

<strong>Girdwood</strong> D, Bumpass D, Vaughan OA, Thain A, Anderson LA, Snowden AW,<br />

Garcia-Wilson E, Perkins ND, and Hay RT. (2003) p300 Transcriptional<br />

Repression Is Mediated by SUMO modification. Molecular Cell 11,1043-1054.<br />

<strong>Girdwood</strong> DWII, Tatham MH, and Hay RT. (2004) SUMO and transcriptional<br />

regulation. Seminars in Cell and Developmental Biology 15,201-210.<br />

<strong>Girdwood</strong> D, Tatham MH, and Hay RT. (2004) Control <strong>of</strong> Gene Expression by<br />

SUMO, Chapter 10. Van G. Wilson. (Ed) Sumoylation: Molecular Biology and<br />

Biochemistry.<br />

248

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!