09.02.2013 Views

Section 2 – Methods - ResearchSpace@Auckland - The University ...

Section 2 – Methods - ResearchSpace@Auckland - The University ...

Section 2 – Methods - ResearchSpace@Auckland - The University ...

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.

http://researchspace.auckland.ac.nz<br />

<strong>ResearchSpace@Auckland</strong><br />

Copyright Statement<br />

<strong>The</strong> digital copy of this thesis is protected by the Copyright Act 1994 (New<br />

Zealand).<br />

This thesis may be consulted by you, provided you comply with the<br />

provisions of the Act and the following conditions of use:<br />

• Any use you make of these documents or images must be for<br />

research or private study purposes only, and you may not make<br />

them available to any other person.<br />

• Authors control the copyright of their thesis. You will recognise the<br />

author's right to be identified as the author of this thesis, and due<br />

acknowledgement will be made to the author where appropriate.<br />

• You will obtain the author's permission before publishing any<br />

material from their thesis.<br />

To request permissions please use the Feedback form on our webpage.<br />

http://researchspace.auckland.ac.nz/feedback<br />

General copyright and disclaimer<br />

In addition to the above conditions, authors give their consent for the<br />

digital copy of their work to be used subject to the conditions specified on<br />

the Library <strong>The</strong>sis Consent Form and Deposit Licence.<br />

Note : Masters <strong>The</strong>ses<br />

<strong>The</strong> digital copy of a masters thesis is as submitted for examination and<br />

contains no corrections. <strong>The</strong> print copy, usually available in the <strong>University</strong><br />

Library, may contain alterations requested by the supervisor.


Unravelling the interactions of the β7<br />

integrin cytoplasmic domain<br />

Yih-Chih Chan<br />

A thesis submitted in partial fulfillment of the requirements for the degree of Doctor of<br />

Philosophy, <strong>The</strong> <strong>University</strong> of Auckland, 2009.


Abstract<br />

Integrins are dynamic cell adhesion molecules expressed on the cell-surface which transmit<br />

signals through the plasma membrane, and play vital physiological roles in multicellular<br />

organisms. <strong>The</strong> β7 integrins α4β7 and αEβ7 are leukocyte-specific, and evolved to form and<br />

maintain the mucosal immune system of the intestine, which is an important immune barrier.<br />

<strong>The</strong>y participate in antigen presentation, leukocyte migration and homing, and other<br />

leukocyte functions. α4β7 binds the mucosal vascular addressin MAdCAM-1, which is<br />

expressed on specialised high endothelial venules at sites of chronic inflammation. <strong>The</strong> β7<br />

integrins contribute to the initiation and maintenance of inflammation, in particular chronic<br />

inflammatory disease, playing a key role in the migration of leukocytes into inflamed tissue.<br />

It is important to understand integrin signalling mechanisms in order to be able to devise<br />

therapeutic agents that can inhibit integrin function, and thereby attenuate inflammation. <strong>The</strong><br />

host laboratory identified a novel cell adhesion regulatory domain (CARD) within the β7<br />

subunit cytoplasmic domain, which plays a key role in regulating β7-mediated cell adhesion.<br />

<strong>The</strong> CARD comprises the small pseudosymmetrical motif YDRREY, which forms the basis<br />

of the present study.<br />

This study aimed to identify key signalling pathways involved in regulating the cell adhesion<br />

function of the β7 integrins, to identify intracellular ligands that interact with the β7 CARD,<br />

and to characterize their interaction. <strong>The</strong> adhesion of the thymic lymphoma TK-1 (α4β7 +<br />

α4β1 - ) to MAdCAM-1 was used to investigate signalling via α4β7. A panel of chemical<br />

inhibitors of cell signalling molecules was employed in TK-1 cell adhesion assays, revealing<br />

that JNK, the src family of kinases, and myosin light chain kinases all potentially participate<br />

in controlling α4β7-mediated cell adhesion.<br />

Cell-permeable synthetic peptide technology was employed to investigate structure-function<br />

relationships of the YDRREY CARD motif. Substitution of the flanking tyrosine residues of<br />

the YDRREY motif with phenylalanines revealed that the tyrosines were critical for the<br />

function of the CARD. However, the state of phosphorylation of the flanking tyrosines was<br />

inconsequential. <strong>The</strong> core DRRE region was also important for CARD function. Mutagenesis<br />

studies involving substitution of the three tyrosine residues in the β7 subunit cytoplasmic<br />

ii


domain to phenylalanines disrupted α4β7-mediated adhesion of TK-1 cells to MAdCAM-1,<br />

reinforcing the role of tyrosine residues in β7 activation.<br />

<strong>The</strong> tyrosine kinases src and FAK were shown to bind directly to the YDRREY peptide. FAK<br />

and autophosphorylated src was shown to bind to both phosphorylated and non-<br />

phosphorylated forms of YDRREY, while non-phosphorylated src bound only the non-<br />

phosphorylated form. <strong>The</strong> cytoskeletal proteins α-actinin, and paxillin formed complexes<br />

with FAK and bound to the YDRREY peptide, whereas filamin disrupted src binding. FAK<br />

and src were both present in immunoprecipitates of α4β7. Immunofluorescence studies<br />

combined with confocal microscopy suggested both kinases co-localise with the β7 integrin<br />

within the trailing edge uropod and at sites of pseudo supra-molecular complex (SMAC)<br />

formation.<br />

Heat shock protein (hsp) 70 was also identified as a potential intracellular ligand that<br />

associates indirectly with the β7 subunit. Thus, a synthetic peptide encompassing the fulllength<br />

β7 cytoplasmic domain bound to recombinant hsp70, but only in the presence of a TK-<br />

1 cell lysate. Heat shock using fever-range temperatures activated α4β7-mediated cell<br />

adhesion. A chemical inhibitor of hsp70 prevented Mn 2+ -induced adhesion of TK-1 cells to<br />

MAdCAM-1, suggesting that hsp70 plays a key role in regulating signalling by α4β7. <strong>The</strong><br />

latter results suggested that other stressors might activate α4β7-mediated cell adhesion, and<br />

in accord serum starvation of cultured cells also induced α4β7-mediated cell adhesion to<br />

MAdCAM-1.<br />

In summary, this study has provided novel insights into signalling pathways that regulate<br />

signalling by β7 integrins. Structure-function relationships of the β7 CARD have been<br />

investigated, and kinases that interact with the CARD have been identified. Activation of β7<br />

integrin signalling by stressors including fever-range temperatures and serum starvation, and<br />

identification of hsp70 as a molecule which associates indirectly with α4β7 are novel<br />

discoveries. <strong>The</strong> information gained provides a platform of results which provide the basis for<br />

generating agents that might have therapeutic potential for treating inflammatory diseases.<br />

iii


Acknowledgements<br />

Firstly, I would like to express my greatest appreciation to my chief supervisor, Associate<br />

Professor Geoffrey W. Krissansen, for providing me the opportunity to conduct my doctoral<br />

research project within his research group, his ever important advice and guidance throughout<br />

these years, and his help in putting this thesis together.<br />

I also extend my appreciation to my past and present co-supervisors Drs J-Zhong Bai,<br />

Euphemia Leung, and Katherine Woods for their advice on the direction of my project.<br />

Additionally I would like to acknowledge Dr Euphemia Leung for her help and her<br />

tremendous knowledge of molecular biology techniques and experimental trouble shooting.<br />

Further I would like to extend my appreciation to our laboratory manager Yi Yang for her<br />

help in the preparation of recombinant proteins, molecular biology and tissue culture<br />

techniques and management of the laboratory.<br />

I would also like to acknowledge past and present research fellows in the laboratory, Drs<br />

Kevin Sun, Jagat Kanwar, and Rupinder Kanwar for their input into this thesis.<br />

To fellow students Antonio Cheung and Jiwon Hong, I would like to acknowledge their<br />

invaluable support and friendship throughout these years.<br />

To all my friends, I would also like to thank them for their encouragement and friendship<br />

throughout this project.<br />

I would also like to thank the Marsden Fund for their support of the project, and the Maurice<br />

and Phyllis Paykel Trust, the Sir John Logan Campbell Trust, Department of Molecular<br />

Medicine and Pathology, and the Faculty of Medical and Health Sciences Postgraduate<br />

Student Association for supporting my attendance at the Queenstown Signal Transduction<br />

Meeting in Queenstown, New Zealand and the Signaling by Adhesion Receptors Conference<br />

in Massachusetts, United States.<br />

Finally I would like to thank my parents Paul and Judy, my brothers Kai and Woei, my sisterin-law<br />

Ying-Ying, and my wife Enid for their wonderful love, support and encouragement<br />

throughout my life. I dedicate this thesis to them.<br />

iv


Table of Contents<br />

Abstract......................................................................................................................................ii<br />

Acknowledgements...................................................................................................................iv<br />

List of Figures...........................................................................................................................xi<br />

List of Tables ..........................................................................................................................xiv<br />

Abbreviations...........................................................................................................................xv<br />

Chapter 1. Introduction ..............................................................................................................1<br />

1.1. Integrins ....................................................................................................................................... 1<br />

1.1.1 Integrin structures .................................................................................................................. 3<br />

1.1.2 Integrins and their extracellular ligands............................................................................... 11<br />

1.2. β7 integrins ................................................................................................................................ 13<br />

1.2.1 <strong>The</strong> β7 subunit cytoplasmic domain .................................................................................... 13<br />

1.2.2 Integrin α4β7........................................................................................................................ 14<br />

1.2.3 Integrin αEβ7 ....................................................................................................................... 16<br />

1.3. Integrin signalling ...................................................................................................................... 17<br />

1.3.1 Models of integrin activation ............................................................................................... 19<br />

1.3.2 Integrin regulation by inside-out signalling ......................................................................... 22<br />

1.3.3 Intracellular ligands of the α-subunit ................................................................................... 27<br />

1.3.4 Integrin outside-in signalling ............................................................................................... 28<br />

1.3.5 Force sensing and focal adhesions....................................................................................... 30<br />

1.3.6 Integrin crosstalk.................................................................................................................. 30<br />

1.4. Integrins and immunological synapses, unique signalling structures ........................................ 31<br />

1.4.1 Synapses formed between communicating immune cells.................................................... 31<br />

1.4.2 Synapses termed kinapses are formed on migrating T cells ................................................ 33<br />

1.5. Tyrosine kinases and cell signalling .......................................................................................... 33<br />

1.5.1 <strong>The</strong> focal adhesion kinase (FAK) family............................................................................. 33<br />

1.5.2 Src kinase family (SFK)....................................................................................................... 36<br />

1.5.3 Integrin signalling, FAK and src.......................................................................................... 38<br />

1.6. Heat shock proteins.................................................................................................................... 39<br />

1.6.1 Hsp70 family........................................................................................................................ 40<br />

v


1.7. Integrins and disease .................................................................................................................. 44<br />

1.7.1 Hereditary disease................................................................................................................ 45<br />

1.7.2 Inflammatory disease ........................................................................................................... 46<br />

1.7.3 Other diseases ...................................................................................................................... 46<br />

1.7.4 β7 integrins and disease ....................................................................................................... 47<br />

1.8. Aims of this thesis...................................................................................................................... 49<br />

Chapter 2. Materials and <strong>Methods</strong>...........................................................................................50<br />

<strong>Section</strong> 1 <strong>–</strong> Materials ...............................................................................................................50<br />

2.1. Chemicals, reagents, buffers and media..................................................................................... 50<br />

2.1.1 Suppliers .............................................................................................................................. 50<br />

2.1.2 Chemicals/Molecular biology reagents................................................................................ 52<br />

2.1.3 Buffers/Solutions ................................................................................................................. 53<br />

2.1.4 Tissue culture reagents......................................................................................................... 56<br />

2.2. Molecular biology reagents........................................................................................................ 57<br />

2.2.1 Mammalian cell lines........................................................................................................... 57<br />

2.2.2 Insect cell line ...................................................................................................................... 57<br />

2.2.3 Bacterial cell line ................................................................................................................. 57<br />

2.2.4 DNA cloning vectors ........................................................................................................... 58<br />

2.2.5 Enzymes/isotopes................................................................................................................. 58<br />

2.2.6 Antibodies............................................................................................................................ 59<br />

2.2.7 PCR oligonucleotide primers............................................................................................... 60<br />

2.2.8 cDNA for cloning/protein expression .................................................................................. 60<br />

2.2.9 Recombinant proteins .......................................................................................................... 61<br />

2.2.10 Inhibitors of intracellular signalling pathways................................................................... 61<br />

2.2.11 Peptides.............................................................................................................................. 62<br />

<strong>Section</strong> 2 <strong>–</strong> <strong>Methods</strong>.................................................................................................................64<br />

2.3. Molecular biology techniques.................................................................................................... 64<br />

2.3.1 Preparation of competent cells............................................................................................. 64<br />

2.3.2 Small scale preparation of plasmid DNA (miniprep) .......................................................... 64<br />

2.3.3 Large scale preparation of plasmid DNA............................................................................. 65<br />

vi


2.3.4 Linearization and dephosphorylation of plasmid vectors .................................................... 66<br />

2.3.5 Isolation of plasmid inserts .................................................................................................. 66<br />

2.3.6 Agarose gel electrophoresis ................................................................................................. 66<br />

2.3.7 Ligation of DNA inserts into vectors................................................................................... 67<br />

2.3.8 Transformation..................................................................................................................... 67<br />

2.3.9 RNA isolation ...................................................................................................................... 67<br />

2.3.10 Synthesis of cDNA by reverse transcriptase...................................................................... 68<br />

2.3.11 Polymerase chain reaction (PCR) ...................................................................................... 68<br />

2.4. Protein chemistry ....................................................................................................................... 69<br />

2.4.1 Production and purification of GST fusion proteins............................................................ 69<br />

2.4.2 Sodium dodecylsulphate polyacrylamide gel electrophoresis ............................................. 69<br />

2.4.3 Staining of SDS gels ............................................................................................................ 71<br />

2.4.4 Western blot analysis ........................................................................................................... 71<br />

2.4.5 Stripping of nitrocellulose and PVDF membranes .............................................................. 72<br />

2.4.6 Kinase assays ....................................................................................................................... 72<br />

2.4.7 Coupling of proteins and antibodies to Sepharose............................................................... 74<br />

2.4.8 Immunoprecipitation / pull-down assay............................................................................... 74<br />

2.5. Cell biology................................................................................................................................ 75<br />

2.5.1 Cell culture........................................................................................................................... 75<br />

2.5.2 Production of recombinant proteins..................................................................................... 75<br />

2.5.3 Spleen cell isolation ............................................................................................................. 76<br />

2.5.4 Cell adhesion assay .............................................................................................................. 76<br />

2.5.5 Enzyme-linked soluble E-cadherin-Fc-mediated adhesion assay. ....................................... 78<br />

2.5.6 Cell transfection ................................................................................................................... 78<br />

2.5.7 Immunofluorescence staining .............................................................................................. 78<br />

2.6. Phylogenetic analysis................................................................................................................. 79<br />

2.7. Statistical analysis...................................................................................................................... 79<br />

Chapter 3. Results ....................................................................................................................80<br />

3.1. Expression and cell adhesion properties of β7-integrins............................................................ 80<br />

3.1.1 Confirmation of the expression of α4β7 and αEβ7 on TK-1 and MTC-1 cells ................... 80<br />

vii


3.1.2 Production of recombinant forms of E-cadherin and MAdCAM-1 ..................................... 84<br />

3.1.3 Comparison of activators of β7 integrin-mediated cell adhesion......................................... 85<br />

3.2. <strong>The</strong> effect of pathway inhibitors on β7 integrin mediated cell adhesion ................................... 89<br />

3.2.1 Genistein, a PTK inhibitor, inhibits the binding of TK-1 cells to MAdCAM-1 .................. 89<br />

3.2.2 Involvement the MAP kinase kinase (MEK) pathway in TK-1 cell binding to<br />

MAdCAM-1 ......................................................................................................................... 90<br />

3.2.3 Jun N-terminal kinase (JNK) pathway................................................................................. 93<br />

3.2.4 Epidermal growth factor receptor (EGFR) tyrosine kinase ................................................. 96<br />

3.2.5 Src family of tyrosine kinases (SFK)................................................................................... 97<br />

3.2.6 Myosin light chain kinase (MLCK) ................................................................................... 100<br />

3.3. Properties of a cell adhesion regulatory domain in the cytoplasmic tail of the integrin β7<br />

subunit .................................................................................................................................... 102<br />

3.3.1 A cell-permeable YDRREY peptide inhibits β7 integrin-mediated adhesion to<br />

MAdCAM-1 and E-cadherin .............................................................................................. 102<br />

3.3.2 Defining the features of the YDRREY peptide required to inhibit β7-mediated<br />

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

3.3.3 Effect of NPLY on TK-1 cell adhesion ............................................................................. 111<br />

3.4. Tyrosine kinases interact with and phosphorylate the YDRREY motif .................................. 113<br />

3.4.1 Production of GST-β integrin cytoplasmic domain fusion proteins .................................. 113<br />

3.4.2 Phosphorylation of GST-β7 cytoplasmic domain fusion proteins ..................................... 114<br />

3.4.3 Tyrosine phosphorylation of the YDRREY peptide .......................................................... 115<br />

3.4.4 Kinases in multiple cell types can phosphorylate the YDRREY motif ............................. 116<br />

3.4.5 Phosphorylation of the YDRREY peptide by FAK, src and lck........................................ 117<br />

3.4.6 Direct binding of FAK and src to the YDRREY motif...................................................... 119<br />

3.4.7 Binding of YDRREY peptide variants by FAK and src .................................................... 120<br />

3.4.8 Does src remain active after binding to the YDRREY peptide?........................................ 121<br />

3.4.9 FAK and YDRREY interactions........................................................................................ 124<br />

3.4.10 FAK and src bind synergistically to the YDRREY peptide............................................. 125<br />

3.5. <strong>The</strong> impact of cytoskeletal proteins on the interaction of the YDRREY peptide with kinases 126<br />

3.5.1 Filamin disrupts the binding of src to the YDRREY peptide ............................................ 126<br />

3.5.2 Interaction of paxillin with the β7 subunit cytoplasmic domain........................................ 127<br />

viii


3.5.3 FAK, src, and α-actinin bind the YDRREY peptide ......................................................... 128<br />

3.6. In vivo interactions of β7 integrins with kinases ..................................................................... 129<br />

3.6.1 Co-localisation of a fluoresceinated YDRREY peptide in living cells.............................. 129<br />

3.6.2 Phosphorylation of the integrin β7 subunit in vivo............................................................ 131<br />

3.6.3 Confirmation of FAK and src involvement in β7 integrin signalling ................................ 134<br />

3.6.4 Detection of src in β7 integrin immunoprecipitates........................................................... 136<br />

3.6.5 Detection of FAK in β7 integrin immunoprecipitates........................................................ 137<br />

3.6.6 Localisation of interacting kinases in living cells.............................................................. 138<br />

3.7. Localization of α4β7 with kinases in supra-molecular activation complexes (SMAC)........... 140<br />

3.7.1 TK-1 cells bind to MAdCAM-1-coated magnetic beads ................................................... 141<br />

3.7.2 α4β7 colocalizes with src in SMACs................................................................................. 142<br />

3.7.3 α4β7 colocalizes with FAK in SMACs.............................................................................. 143<br />

3.7.4 α4β7 colocalizes with lck in SMACs................................................................................. 144<br />

3.8. Mutation of the β7 cytoplasmic domain and effects on cell adhesion ..................................... 146<br />

3.8.1 Cloning of mutated variants of the integrin β7 subunit into mammalian expression<br />

vectors................................................................................................................................. 146<br />

3.8.2 Cloning of a wild-type integrin α4 construct ..................................................................... 148<br />

3.8.3 Transfection of HEK-293T cells to express the α4 and β7 subunit plasmids .................... 149<br />

3.8.4 Testing the ability of transfectants to adhere to integrin ligands ....................................... 153<br />

3.9. Identification of binding partners of the β7 cytoplasmic domain ............................................ 155<br />

3.9.1 Pull-down assay ................................................................................................................. 155<br />

3.9.2 Mass spectrometry identifies heat shock proteins as ligands of the β7 subunit................. 156<br />

3.9.3 A pull-down assay estabilishes that hsp70 interacts indirectly with the β7 subunit .......... 158<br />

3.9.4 Coimmunoprecipitation of α4β7 and hsp70....................................................................... 159<br />

3.9.5 Localization of hsp70 and hsp90 in TK-1 cells ................................................................. 161<br />

3.9.6 Hsp70 colocalizes with intracellular β7 ligands following ligand-induced clustering of<br />

α4β7 .................................................................................................................................... 163<br />

3.9.7 Effect of heat shock on expression of hsp70 and the β7 subunit ....................................... 164<br />

3.9.8 Effect of heat shock on activation of α4β7 is time dependent ........................................... 165<br />

3.9.9 Fever-range temperatures can activate α4β7...................................................................... 166<br />

ix


3.9.10 Heat shock leads to prolonged activation of α4β7 even after Mn 2+ -treatment ................ 167<br />

3.9.11 An inhibitor of hsp70 blocks the binding of TK-1 cells to MAdCAM-1 ........................ 168<br />

3.9.12 KNK-437 has no detectable effect on hsp70 and β7 expression...................................... 169<br />

3.9.13 Serum deprivation activates α4β7.................................................................................... 170<br />

Chapter 4. Discussion ............................................................................................................173<br />

4.1. Preliminary characterization of α4β7 and αEβ7 expression and adhesion............................... 173<br />

4.2. Analysis of β7 integrin signalling pathways............................................................................ 174<br />

4.3. <strong>The</strong> YDRREY motif ................................................................................................................ 180<br />

4.4. Phylogenetic study of the sequence of the YDRREY motif in the β7 subunit ........................ 181<br />

4.5. <strong>The</strong> NPLY motif ...................................................................................................................... 183<br />

4.6. Interaction of the YDRREY motif with tyrosine kinases ........................................................ 185<br />

4.7. Probing the function of β7 subunit cytoplasmic domain tyrosine phosphorylation sites by<br />

mutational analysis ................................................................................................................. 187<br />

4.8. Interaction of the YDRREY motif with cytoskeletal proteins ................................................. 188<br />

4.9. In vivo interactions of the YDRREY peptide with cellular kinases......................................... 191<br />

4.10. α4β7 clusters within SMACs ................................................................................................. 193<br />

4.11. <strong>The</strong> affect of stress on the function of β7 integrins................................................................ 193<br />

4.12. Summary................................................................................................................................ 195<br />

4.13. Future directions .................................................................................................................... 196<br />

References..............................................................................................................................200<br />

x


List of Figures<br />

Figure 1.1 Integrin subunits and pairings................................................................................................3<br />

Figure 1.2 Structure of the integrin heterodimer.....................................................................................4<br />

Figure 1.3 Domains of the integrin α-subunit.........................................................................................5<br />

Figure 1.4 Domains of the integrin β-subunit.........................................................................................7<br />

Figure 1.5 Structure of the αβ integrin heterodimer................................................................................9<br />

Figure 1.6 <strong>The</strong> amino acid sequence of the cytoplasmic domain of the human β7 subunit. .................13<br />

Figure 1.7 Multistep model of T cell transmigration ............................................................................15<br />

Figure 1.8 Switch blade model of integrin activation...........................................................................20<br />

Figure 1.9 Deadbolt model of integrin activation .................................................................................21<br />

Figure 1.10 Sequence alignments of the human integrin β subunit (a), and α subunit (b)<br />

cytoplasmic domains. ........................................................................................................23<br />

Figure 1.11 Models of the immunological synapse and kinapse ..........................................................32<br />

Figure 1.12 Structural features and binding partners of FAK...............................................................34<br />

Figure 1.13 Structural features and activation of src ............................................................................37<br />

Figure 1.14 Domain structure of human hsp70 family members..........................................................42<br />

Figure 3.1 Immunoblot analysis of an MTC-1 cell lysate with antibodies against the αE and β7<br />

subunits..............................................................................................................................81<br />

Figure 3.2 MTC-1 cells immunostained with an anti-β7 antibody. ......................................................82<br />

Figure 3.3 TK-1 cells stained with the DATK32 antibody against the α4β7 complex.........................83<br />

Figure 3.4 Analysis of purified recombinant E-cadherin-Fc and MAdCAM-1-Fc by SDS-<br />

PAGE. ...............................................................................................................................85<br />

Figure 3.5 Activation of TK-1 cell adhesion to MAdCAM-1 ..............................................................86<br />

Figure 3.6 Activation of MTC-1 cell adhesion to E-cadherin ..............................................................87<br />

Figure 3.7 Binding of E-cadherin to MTC-1 cells is specifically mediated by αEβ7...........................88<br />

Figure 3.8 Effect of genistein on TK-1 cell adhesion to MAdCAM-1 .................................................90<br />

Figure 3.9 Effect of PD98059 on TK-1 cell adhesion to MAdCAM-1.................................................91<br />

Figure 3.10 Effect of SB203580 on TK-1 cell adhesion to MAdCAM-1.............................................92<br />

Figure 3.11 Effect of JNK-I-1 on TK-1 cell adhesion to MAdCAM-1 ................................................94<br />

Figure 3.12 Effect of JNK-I-2 on TK-1 cell adhesion to MAdCAM-1 ................................................95<br />

Figure 3.13 Effect of AG99 on TK-1 cell adhesion to MAdCAM-1....................................................96<br />

Figure 3.14 Effect of PP2 on TK-1 cell adhesion to MAdCAM-1 .......................................................98<br />

Figure 3.15 Effect of damnacanthal on TK-1 cell adhesion to MAdCAM-1 .......................................99<br />

Figure 3.16 Effect of radicicol on TK-1 cell adhesion to MAdCAM-1..............................................100<br />

Figure 3.17 Effect of ML-7 on TK-1 cell adhesion to MAdCAM-1 ..................................................101<br />

Figure 3.18 <strong>The</strong> amino acid sequence of the human β7 integrin cytoplasmic domain .......................102<br />

Figure 3.19 Effect of the YDRREY peptide on TK-1 cell adhesion to MAdCAM-1.........................103<br />

Figure 3.20 Effect of YDRREY peptide on MTC-1 cell adhesion to E-cadherin ..............................104<br />

Figure 3.21 Effect of the phosphorylated xDRREx peptide on TK-1 cell adhesion to<br />

MAdCAM-1 ....................................................................................................................105<br />

Figure 3.22 Effect of the YDRGGGGREY peptide on TK-1 cell adhesion to MAdCAM-1 .............106<br />

Figure 3.23 Effect of the YDGGEY peptide on TK-1 cell adhesion to MAdCAM-1 ........................107<br />

Figure 3.24 Effect of the YEEEEY peptide on TK-1 cell adhesion to MAdCAM-1 .........................109<br />

Figure 3.25 Effect of the pYDRREY peptide on TK-1 cell adhesion to MAdCAM-1.......................110<br />

Figure 3.26 Effect of the pFDRREF peptide on TK-1 cell adhesion to MAdCAM-1........................111<br />

xi


Figure 3.27 Effect of the NPLY peptide on TK-1 cell adhesion to MAdCAM-1...............................112<br />

Figure 3.28 Production of GST-β subunit cytoplasmic domain fusion proteins, and analysis by<br />

SDS-PAGE......................................................................................................................114<br />

Figure 3.29 Phosphorylation of GST-β subunit cytoplasmic domain fusion proteins........................115<br />

Figure 3.30 <strong>The</strong> YDRREY peptide is phosphorylated by a tyrosine kinase(s) in a TK-1 cell<br />

lysate................................................................................................................................116<br />

Figure 3.31 Kinases in multiple cell types can phosphorylate the YDRREY motif...........................117<br />

Figure 3.32 Phosphorylation of the GGYDRREY peptide by recombinant FAK, src, and lck..........118<br />

Figure 3.33 Sequence recognition of the YDRREY peptide by FAK, src, and lck ............................119<br />

Figure 3.34 Analysis of the binding and phosphorylation of the YDRREY peptide by FAK, src,<br />

and lck .............................................................................................................................120<br />

Figure 3.35 FAK and src binding and phosphorylation of variants of the YDRREY peptide............121<br />

Figure 3.36 Src bound to the YDRREY motif remains active............................................................122<br />

Figure 3.37 Src binds to YDRREY but not to xDRREx.....................................................................123<br />

Figure 3.38 Tyrosine phosphorylation of the YDRREY peptide affects the degree of FAK<br />

binding.............................................................................................................................124<br />

Figure 3.39 Do FAK and src bind synergistically to the YDRREY peptide?.....................................125<br />

Figure 3.40 Filamin disrupts the binding of src to the YDRREY peptide..........................................127<br />

Figure 3.41 Erk-phosphorylated paxillin binds to FAK and forms a complex with the YDRREY<br />

peptide. ............................................................................................................................128<br />

Figure 3.42 Binding of src, FAK, and α-actinin to the YDRREY peptide. ........................................129<br />

Figure 3.43 <strong>The</strong> YDRREY peptide co-localises with FAK and src at focal adhesions in vivo ..........130<br />

Figure 3.44 Tyrosine phosphorylation of the β7 subunit in vivo is dependent on the activation<br />

status of cells ...................................................................................................................133<br />

Figure 3.45 <strong>The</strong> Fib504 mAb specifically immunoprecipitates the β7 subunit from TK-1 cells .......135<br />

Figure 3.46 Src is coimmunoprecipitated with the β7 integrin from TK1 cell lysates .......................136<br />

Figure 3.47 FAK is coimmunoprecipitated with the β7 integrin from TK1 lysates ...........................137<br />

Figure 3.48 FAK and α4β7 do not strongly colocalize on TK-1 cells ................................................138<br />

Figure 3.49 Src appears to colocalize with α4β7 in the uropod..........................................................139<br />

Figure 3.50 Lck and α4β7 do not appear to colocalize on TK-1 cells ................................................140<br />

Figure 3.51 TK-1 cells bind to MAdCAM-1-coated magnetic beads.................................................141<br />

Figure 3.52 α4β7 clusters with src in pseudo-SMAC .........................................................................143<br />

Figure 3.53 α4β7 clusters with FAK in pseudo-SMAC......................................................................144<br />

Figure 3.54 Clusters of integrin β7 and lck.........................................................................................145<br />

Figure 3.55 Cloning strategy for constructing plasmids encoding the full-length β7 subunit with<br />

mutated cytoplasmic tyrosines ........................................................................................147<br />

Figure 3.56 Schematic diagram of the strategy used to generate a pcDNA6-V5his vector<br />

encoding the full-length α4 subunit ................................................................................148<br />

Figure 3.57 Transfection of HEK-293T cells with plasmids encoding the integrin α4 and β7<br />

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

Figure 3.58 Western blot analysis of HEK-293T cells transfected with plasmids encoding α4<br />

and β7 varients.................................................................................................................150<br />

Figure 3.59 α4β7 is expressed at the surface of HEK-293T cells transfected with α4 and β7<br />

plasmids...........................................................................................................................152<br />

Figure 3.60 Testing the ability of HEK 293T transfectants to adhere to MAdCAM-1 ......................154<br />

Figure 3.61 SDS-PAGE analysis of proteins precipitated with a synthetic peptide comprising<br />

the complete cytoplasmic domain of the β7 subunit .......................................................156<br />

xii


Figure 3.62 Mass spectrometry identifies four heat shock proteins as potential β7 ligands...............157<br />

Figure 3.63 A synthetic β7 cytoplasmic domain peptide precipitates recombinant hsp70 in a<br />

pull down assay ...............................................................................................................158<br />

Figure 3.64 Hsp70 is coimmunoprecipitated from a TK-1 cell lysate with the β7 subunit ................159<br />

Figure 3.65 <strong>The</strong> β7 subunit is coimmunoprecipitated from a TK-1 cell lysate with hsp70................160<br />

Figure 3.66 Localisation of hsp70 and α4β7 on TK-1 cells attached and spread on MAdCAM-1....161<br />

Figure 3.67 Localisation of hsp90 and α4β7 on TK-1 cells attached and spread on MAdCAM-1....162<br />

Figure 3.68 Hsp70 colocalizes with intracellular β7 ligands within SMAC formed with<br />

MAdCAM-1-coated microspheres ..................................................................................163<br />

Figure 3.69 Effect of heat shock on the expression of hsp70 and the β7 subunit by TK-1 cells........164<br />

Figure 3.70 Effect of heat shock treatment on the binding of TK-1 cells to MAdCAM-1.................166<br />

Figure 3.71 Fever-range temperatures induce TK-1 cell binding to MAdCAM-1 .............................167<br />

Figure 3.72 <strong>The</strong> prolonged activation of TK-1 cells by heat shocking is not affected by Mn 2+<br />

treatment..........................................................................................................................168<br />

Figure 3.73 KNK-437 blocks TK-1 cell adhesion to MAdCAM-1 ....................................................169<br />

Figure 3.74 Effect of KNK-437 on the expression of hsp70 and the β7 subunit................................170<br />

Figure 3.75 Effect of serum deprivation on TK-1 cell adhesion to MAdCAM-1...............................171<br />

Figure 3.76 Expression levels of the β7 subunit and hsp70 remain unchanged after serum<br />

depletion ..........................................................................................................................172<br />

Figure 4.1 Pathways which lead to integrin activation. ......................................................................176<br />

Figure 4.2 Proteins which may interact with the β7 cytoplasmic domain ..........................................189<br />

Figure 4.3 <strong>The</strong> location of zones on migrating TK-1 cells. ................................................................192<br />

xiii


List of Tables<br />

Table 1.1 Gene locations and structures of human integrin subunits......................................................2<br />

Table 1.2 A list of integrins and their known extracellular ligands ......................................................12<br />

Table 1.3 A list of integrins and their known intracellular ligands.......................................................18<br />

Table 1.4 Overview of the proteins/pathways that are involved in FAK-mediated signalling.............36<br />

Table 1.5 Essential features of members of the human Hsp70 family..................................................41<br />

Table 1.6 Phenotypes of Hsp70 knockout mice....................................................................................43<br />

Table 1.7 Ablation of integrin genes in mice and their resulting phenotypes.......................................45<br />

Table 4.1 Summary of the effects of chemical inhibitors on β7-mediated cell adhesion ...................177<br />

Table 4.2 Alignment of the first 20 amino acid residues of the β7 subunit cytoplasmic domain<br />

of various animal species.................................................................................................182<br />

Table 4.3 Alignment of the protein sequences from the distal region of the β7 subunit<br />

cytoplasmic of various animal species ............................................................................184<br />

xiv


Abbreviations<br />

× g g-force<br />

© copyright<br />

°C degree Celsius<br />

µ (prefix) micro-<br />

A ampere<br />

Å Ångströms<br />

aa amino acid<br />

Ab antibody<br />

ABTS 2,2'-azino-bis (3-ethylbenzthiazoline-6-sulfonic acid)<br />

ADMIDAS adjacent to the MIDAS<br />

AF Alexa Fluor®<br />

APC antigen presenting cell<br />

Arg arginine<br />

ATCC American Type Culture Collection<br />

ATP adenosine triphosphate<br />

bp base pair<br />

BSA bovine serum albumin<br />

CARD cell adhesion regulatory domain<br />

CCL CC chemokine ligand<br />

CD cluster of differentiation<br />

cDNA complementary DNA<br />

CO collagen<br />

Corp corporation<br />

CXCL CXC chemokine ligand<br />

DAG diacylglycerol<br />

DEPC diethylpyrocarbonate<br />

DMSO dimethylsulfoxide<br />

DNA deoxyribonucleic acid<br />

DTT dithiothreitol<br />

ECM extracellular matrix<br />

EDTA ethylenediaminetetra-acetic acid<br />

eed embryonic ectoderm development<br />

EGF(R) epidermal growth factor (receptor)<br />

EM electron microscopy<br />

ERK extracellular signal-regulated kinases<br />

FA focal adhesion<br />

FAK focal adhesion kinase<br />

FITC fluorescein isothiocyanate<br />

FLN filamin<br />

fMLP formyl-methionyl-leucyl-phenylalanine<br />

FN fibronectin<br />

FRET Förster resonance energy transfer<br />

g gram<br />

GALT gut associated lymphoid tissue<br />

GAPDH glyceraldehyde-3-phosphate dehydrogenase<br />

HIMEC human intestinal microvascular endothelial cells<br />

HIV human immunodeficiency virus<br />

hr(s) hour(s)<br />

Hsp heat shock protein<br />

IBD inflammatory bowel disease<br />

xv


ICAM intercellular cell adhesion molecule<br />

IL interleukin<br />

ILK integrin-linked kinase<br />

IPTG isopropyl-β-D-thiogalactopyranoside<br />

IS Immunological synapse<br />

JNK c-Jun NH2-terminal kinase<br />

kb kilobase<br />

kDa kilodalton<br />

L liter<br />

LB Luria-Bertani<br />

LFA-1 lymphocyte function-associated antigen 1<br />

LIMBS ligand-induced metal ion-binding site<br />

LN laminin<br />

LPAM lymphocyte Peyer’s patch adhesion molecule<br />

LPL lamina propria lymphocytes<br />

m meters<br />

M molar<br />

m (prefix) milli-<br />

mAb monoclonal antibody<br />

MAdCAM mucosal addressin cell adhesion molecule<br />

MAPK mitogen-activated protein kinase<br />

MIDAS metal ion-dependent adhesive site<br />

min minute<br />

MLCK myosin light chain kinase<br />

mRNA messenger RNA<br />

n (prefix) nano-<br />

NK natural killer<br />

NMR nuclear magnetic resonance<br />

OD optical density<br />

p p-value<br />

PBM perivascular basement membrane<br />

PCR polymerase chain reaction<br />

PFA paraformaldehyde<br />

PIP2 phosphatidylinositol 4,5-biphosphate<br />

PKA protein kinase A<br />

PKC protein kinase C<br />

PMA phorbol 12-myristate 13-acetate<br />

PP Peyer’s patches<br />

PSI plexin-semaphorin integrin<br />

PTB phosphotyrosine binding<br />

R9 polymer of 9 arginines<br />

RGD arginine-glycine-aspartic acid<br />

RIAM Rap1-GTP interacting adapter molecule<br />

rpm revolutions per minute<br />

RTK receptor tyrosine kinase<br />

s second<br />

SDS sodium dodecyl sulphate<br />

SDS-PAGE sodium dodecyl sulfate polyacrylamide gel electrophoresis<br />

SFK src family of kinases<br />

SH src homology<br />

SIEL small intestine intraepithelial lymphocytes<br />

siRNA small interference RNA<br />

SL spleen lymphocytes<br />

SMAC supramolecular activation complex<br />

xvi


TCR T-cell receptor<br />

Tfb transformation buffer<br />

TGF transforming growth factor<br />

TIED ten β-integrin EGF-like repeat domains<br />

TNF tumor necrosis factor<br />

V volts<br />

VCAM vascular cell adhesion molecule<br />

vWFA von Willebrand Factor type A<br />

wt wild-type (native form)<br />

www world wide web<br />

β-TD β-terminal domain<br />

xvii


Chapter 1. Introduction<br />

1.1. Integrins<br />

Integrins are a family of cell adhesion molecules that mediate cell-cell, cell-extracellular<br />

matrix (ECM) and cell-pathogen interactions (reviewed in Hynes 2002; Krissansen et al.<br />

2007). <strong>The</strong>y are widely distributed with diverse functions, and play a central role in cell<br />

adhesion and migration. Integrins play vital physiological roles during the lifespan of an<br />

organism, from embryogenesis and organogenesis to maintaining haemostasis by mediating<br />

cell proliferation, survival, the immune response, and wound healing (reviewed in Hynes<br />

2002; Krissansen et al. 2007).<br />

Integrins have been found to exist in metazoan organisms, from the simplest metazoan<br />

sponges to the most complex eukaryotes (reviewed in Hynes 2002). No homologues have<br />

been detected in prokaryotes, plants or fungi (Whittaker et al. 2002). Integrins are formed<br />

from two non-covalently associated type I transmembrane glycoprotein subunits, named the<br />

α- and β-subunits. <strong>The</strong> α- and β- subunits are totally distinct with no detectable homology<br />

between them, and are encoded by two separate supergene families located on various<br />

chromosomes (Table 1.1; reviewed in Lam and Loftus 2005, Krissansen and Danen 2007,<br />

Takada et al. 2007).<br />

In addition to their adhesive functions, integrins are dynamic cell-surface receptors for cell<br />

signalling that can propagate signals across the plasma membrane in both directions, termed<br />

“inside-out” and “outside-in” signalling (reviewed in Hynes 2002). Inside-out signalling<br />

refers to the transmission of signals from within a cell to the outside extracellular<br />

environment, altering the affinity of integrins for their ligands. Conversely, outside-in<br />

signalling is the transmission of signals from outside of the cell by integrin-ligand interactions<br />

leading to changes inside the cell. Signals delivered from inside and outside of the cell<br />

impinge on the integrin cytoplasmic regions, which serve as important regulators of integrin<br />

signalling and function (reviewed in Hynes 2002; Krissansen et al. 2007).<br />

1


Table 1.1 Gene locations and structures of human integrin subunits<br />

Integrin subunit Synonym Gene location<br />

α1 CD49a 5q11.2<br />

α2 CD49b 5q23<strong>–</strong>q31<br />

α3 CD49c 17q21.33<br />

α4 CD49d 2q31.3<br />

α5 CD49e 12q11<strong>–</strong>q13<br />

α6 CD49f 2q31.1<br />

α7 - 12q13<br />

α8 - 10p.13<br />

α9 - 3p21.3<br />

α10 - 1q21<br />

α11 - 15q23<br />

αIIb CD41b, GPIIb 17q21.32<br />

αv CD51 2q31<strong>–</strong>q32<br />

αE CD103 17q13<br />

αL CD11a 16p11.2<br />

αM CD11b, Mac-1 16p11.2<br />

αX CD11c 16p11.2<br />

αD - 16p11.2<br />

β1 CD29 10p11.2<br />

β2 CD18 21q22.3<br />

β3 CD61 17q21<strong>–</strong>q23<br />

β4 CD104 17q25<br />

β5 - 3q21.2<br />

β6 - 2q24<strong>–</strong>q31<br />

β7 - 12q13.13<br />

β8 - 7p15.3<br />

(Adapted from Lam et al. 2005; Krissansen et al. 2007; Takada et al. 2007)<br />

Integrin-ligand binding is regulated by changes in integrin affinity and avidity, which allows<br />

cells to change shape and migrate (reviewed in Hynes 2002). It involves integrin diffusion<br />

and clustering within the plasma membrane, which ultimately affects cytoskeletal connections<br />

with the extracellular milieu. Extracellular interactions involving integrins include cell-matrix<br />

adhesion where integrins bind to ECM proteins (Geiger et al. 2001), cell-cell adhesion which<br />

leads to “immunological synapse” formation in the case of immune cells (Sims et al. 2002),<br />

transendothelial migration of leukocytes which is important for leukocyte infiltration of<br />

inflamed tissues (Barreiro et al. 2007), and cell-pathogen adhesions (Forsyth et al. 1998).<br />

Disruption of integrin function and signalling contributes to a wide variety of diseases<br />

including, cancer, inflammatory disease, thrombosis and infection (reviewed in Hynes 2002;<br />

Krissansen et al. 2007). In order to understand integrin involvement and contribution to<br />

2


various diseases, it is necessary to understand integrin signalling pathways to elucidate the<br />

control of integrin function. Of particular interest are the integrin cytoplasmic domains, and<br />

their binding/interacting partners which regulate integrin function.<br />

1.1.1 Integrin structures<br />

<strong>The</strong>re are 18 different integrin α chains and 8 different β chains (Table 1.1) which associate<br />

non-randomly in a divalent cation-dependent fashion forming at least 24 different αβ hetero-<br />

dimeric integrin molecules, each with discrete receptor binding functions (Figure 1.1;<br />

reviewed in Hynes 2002). Alternative splicing of certain integrin subunits gives rise to even<br />

greater numbers of different integrin structures and parings, with the different alternatively<br />

spliced forms having divergent functional attributes (reviewed in Hynes 2002). Integrins can<br />

be broadly grouped by ligand specificity into laminin (LN)-binding, collagen (CO)-binding,<br />

and Arginine-Glycine-Aspartic acid (RGD)-binding integrins, and leukocyte-specific<br />

receptors (Figure 1.1; reviewed in Hynes 2002). <strong>The</strong>y can be further defined on the basis that<br />

some integrin α-subunits have an inserted domain called the I-domain (reviewed in<br />

Krissansen et al. 2007).<br />

Figure 1.1 Integrin subunits and pairings.<br />

A diagram depicting the pairing of different integrin α and β subunits. A heterodimer is formed by the<br />

association of particular α and β subunits as denoted by the lines connecting the α and β subunits. Integrins are<br />

broadly grouped into those that bind laminin (red), collagen (purple), and the RGD motif (blue), or by restriction<br />

to leukocytes (yellow). (Adapted from Hynes 2002; Krissansen and Danen 2007).<br />

3


Integrins, being type-1 transmembrane proteins, contain three domains; namely an N-terminal<br />

extracellular domain, a single transmembrane domain, and a C-terminal cytoplasmic domain<br />

(Figure 1.2).<br />

Figure 1.2 Structure of the integrin heterodimer<br />

A cartoon image of an integrin heterodimer, depicting the different domains present in the integrin α and β<br />

subunits. (Adapted from Luo et al. 2007)<br />

Extracellular domain: <strong>The</strong> extracellular domain comprises two separate regions that form a<br />

large globular head and a long stalk region (reviewed in Luo et al. 2007). <strong>The</strong> globular head<br />

of the α and β subunits contains multiple domains and mediates ligand binding. <strong>The</strong> stalk is<br />

also comprised of multiple domains and is important for integrin activation. Details of these<br />

domains will be discussed below.<br />

Transmembrane domain: <strong>The</strong> single transmembrane segment connects the extracellular and<br />

cytoplasmic domains. It is thought to mediate the transmission of signals between the<br />

cytoplasmic and extracellular domains arising from conformational changes in either domain<br />

(reviewed in Luo et al. 2007).<br />

Cytoplasmic domain: Integrin cytoplasmic domains are typically short (13-70 amino acids,<br />

aa), with the exception of the cytoplasmic domain of the β4 subunit (1072 aa; (Hogervorst et<br />

al. 1990). <strong>The</strong>y mediate integrin-cytoskeletal interactions and signal transmission, and play a<br />

central role in integrin activation (reviewed in Krissansen and Danen 2007). Overexpression<br />

of particular proteins that bind to the cytoplasmic domains can result in integrin activation<br />

(Vinogradova et al. 2002).<br />

4


Integrin α-subunits<br />

<strong>The</strong> 18 different α-subunits share 18 to 63% aa identity and are up to 1104 aa residues in<br />

length, with a mature size of 120-210 kDa (reviewed in Krissansen et al. 2007). As mentioned<br />

above, each integrin subunit contains an extracellular domain, a transmembrane domain, and<br />

a cytoplasmic domain. <strong>The</strong> extracellular domain is comprised of five distinctive domains<br />

based on their structure, refered to as the β-propeller, I-domain, thigh domain, Genu domain,<br />

and the calf domains (Figure 1.3; reviewed in Luo et al. 2007)<br />

Figure 1.3 Domains of the integrin α-subunit<br />

A cartoon depiction of the integrin α-subunit, showing the different domains in a linear form. Indicated are the<br />

β-propeller domain, I domain, thigh domain, genu region, calf domains, transmembrane domain, and the<br />

cytoplasmic domain. (Adapted from Luo et al. 2007).<br />

<strong>The</strong> β-propeller domain: <strong>The</strong> N-terminal half of the α subunit contains a globular head<br />

termed the β-propeller domain, which is made up of seven β-sheet repeats (~60 aa each)<br />

formed by four anti-parallel β-strands, folded around a pseudo-symmetric axis similar to the<br />

structure of the heterotrimeric G-protein β-subunit (Springer 1997; Xiong et al. 2001; Xiao et<br />

al. 2004). In addition, there are four Ca 2+ binding motifs in the propeller β-sheet repeats 4-7.<br />

In integrins which lack the I-domain (described below), the β-propeller domain appears to<br />

directly participate in ligand binding (Humphries 2000). In contrast, the β-propeller domain<br />

may cooperate in ligand binding (Yalamanchili et al. 2000) or play no role in integrins which<br />

possess the I-domain (Shimaoka et al. 2001).<br />

5


<strong>The</strong> I-domain: Almost half of integrin α-subunits contain an I (insertion or interaction)<br />

domain [also referred to as an αA, αI or von Willebrand Factor type A (vWFA) domain]<br />

(Humphries 2000). <strong>The</strong> subunits which possess an I-domain include α1, α2, α10, α11, αM,<br />

αL, αD, αX, and αE. <strong>The</strong> I-domain is an inserted sequence of ~190 aa residues comprising<br />

three EF-like motifs at the top of the β-propeller, between the second and third β-sheet<br />

repeats. <strong>The</strong> I-domains fold independently of the β-propeller; and adopt the dinucleotide-<br />

binding or Rossmann fold, with seven α-helices surrounding six hydrophobic β-strands<br />

arranged as a central β-sheet (Lee et al. 1995). At the top of the β-sheet is a metal-ion-<br />

dependent adhesive site (MIDAS) motif. Within the I-domain are two conserved serines, two<br />

aspartic acid residues, and a threonine, which coordinate a Mg 2+ cation to an acidic residue in<br />

the ligand. Divalent cations are known to be universally required for ligand-binding by<br />

integrins, and the MIDAS motif is important for ligand binding. <strong>The</strong> I-domain exists in either<br />

an open (high affinity) or a closed (low affinity) conformation (Humphries 2000). For non-I-<br />

domain integrin α-subunits (α3, α4, α5, α6, α7, α8, α9, αV, and αIIb) the third β-turn of the β-<br />

propeller is directly involved in ligand binding (Humphries 2000).<br />

Stalk region: <strong>The</strong> stalk region is formed from three β-sandwich domains, and links the<br />

globular head and transmembrane domain (Xiong et al. 2001). <strong>The</strong> upper part of the stalk<br />

region is designated the thigh domain and the lower part the calf-1 and calf-2 domains (Xiong<br />

et al. 2001). A small Ca 2+ -binding loop is located between the thigh and calf-1 domains,<br />

named the genu domain (Xiong et al. 2001). <strong>The</strong> genu domain is the key pivot point for the<br />

switchblade extension of the α subunit (described below).<br />

Transmembrane domain: <strong>The</strong> α-subunit transmembrane domain, formed from single-pass α-<br />

helices, anchors the subunit to the plasma membrane. Its association with the β-subunit<br />

transmembrane domain is important in integrin activation (Lau et al. 2008a; Lau et al. 2008b).<br />

Cytoplasmic domain: Integrin α-subunit cytoplasmic domains are short sequences of 15-58 aa<br />

residues. <strong>The</strong>y are highly divergent, except for a membrane proximal GFF(R/K)R motif,<br />

which appears essential for β-subunit association and stability (described below; reviewed in<br />

Krissansen et al. 2007).<br />

6


Integrin β-subunits<br />

<strong>The</strong> 8 different integrin β-subunits share 28 to 55% aa identity and range in size up to 778 aa<br />

residues in length, with a mature size of 90-130 kDa (reviewed in Krissansen et al. 2007). <strong>The</strong><br />

extracellular region contains a globular head and a stalk region. <strong>The</strong> globular head contains<br />

three highly conserved domains, the I-like domain, the PSI domain, and the hybrid domain<br />

(Luo et al. 2007). <strong>The</strong> stalk region consists of cysteine-rich integrin epidermal growth factor<br />

(EGF)-like domains (I-EGF), and a C-terminal β-tail domain (Figure 1.4).<br />

Figure 1.4 Domains of the integrin β-subunit<br />

A cartoon picture depicting the integrin β-subunit and the different domains in a linear form. Indicated are the<br />

PSI domain, hybrid domain, I-like-domain (β I-domain), I-EGF domains 1-4, β-tail, transmembrane domain, and<br />

the cytoplasmic domain. (Adapted from Luo et al. 2007).<br />

<strong>The</strong> PSI domain: <strong>The</strong> extreme N-terminal region of integrin β-subunits termed the plexinsemaphorin-integrin<br />

(PSI) domain is rich in cysteine residues (Zang et al. 2001). It shares<br />

sequence homology with the cysteine-rich membrane proteins plexins and semaphorins<br />

(Figure 1.4; Zang et al. 2001). <strong>The</strong> PSI region contains seven cysteines which cooperate to<br />

restrain the integrin in an inactive conformation (Zang et al. 2001).<br />

I-like domain: <strong>The</strong> N-terminal region of integrin β-subunits contains a highly conserved<br />

domain located between aa residues 100-340, termed the I-like-domain (β I-domain; Figure<br />

1.4; Krissansen et al. 2007). <strong>The</strong> I-like-domain contains a conserved metal-binding DXSXS<br />

sequence (where X denotes any amino acid other than cysteine) which resembles the MIDAS<br />

motif of integrin α-subunits (Lee et al. 1995). Distinguishing features of I-like-domain<br />

regions are the presence of additional cation binding sites, termed adjacent to the MIDAS<br />

7


domain (ADMIDAS), and the ligand-induced metal ion-binding site(s) (LIMBS; Xiong et al.<br />

2001). <strong>The</strong> I-like-domain of the integrin β-subunit is physically associated with the α-subunit<br />

β-propeller domain. <strong>The</strong> β-subunit I-like-domain of integrins which lack the α subunit I-<br />

domain appears to bind ligands directly. It indirectly regulates ligand-binding in integrins that<br />

contain an α-subunit I-domain (Lu et al. 2001b). Several non-conserved residues within this<br />

region determine ligand specificity.<br />

Mutational studies have shown that the LIMBS functions as a positive regulatory site, and the<br />

ADMIDAS as a negative regulatory site (Chen et al. 2003; Mould et al. 2003a; Chen et al.<br />

2006a). <strong>The</strong> addition of Mn 2+ or removal of Ca 2+ can increase ligand-binding affinity and<br />

adhesiveness of almost all integrins (Shimaoka et al. 2002). In most integrins, Ca 2+ can have<br />

both positive and negative regulatory effects, in that high concentrations of Ca 2+ inhibit<br />

adhesion, whereas low concentrations of Ca 2+ synergize with suboptimal Mg 2+ concentrations<br />

to support adhesion (Shimaoka et al. 2002). <strong>The</strong> LIMBS mediates the synergistic effects of<br />

low Ca 2+ concentrations, whereas the ADMIDAS mediates the negative regulatory effects of<br />

higher Ca 2+ concentrations, which are competed by Mn 2+ (Shimaoka et al. 2002).<br />

<strong>The</strong> hybrid domain: An Ig-like hybrid domain is composed of sequences from either side of<br />

the I-like-domain (Figure 1.4; Luo et al. 2007). It forms a β-sandwich structure looped out<br />

between the I-like domain and the PSI domain. <strong>The</strong> hybrid domain plays a central role in<br />

conformational activation of integrins, as described in more detail below (Luo et al. 2007).<br />

<strong>The</strong> stalk region: <strong>The</strong> stalk region contains four internally disulfide-bonded repeats of a<br />

cysteine-rich motif located between aa residues 340-700, termed I-EGF1-4 (Figure 1.4; Luo<br />

et al. 2007). <strong>The</strong> cysteine-rich EGF-like repeats are reminiscent of the EGF repeats of the<br />

protein Ten β integrin EGF-like repeat domains (TIED) protein (Berg et al. 1999a). No<br />

specific functions have been attributed to the stalk region apart from the fact that it may<br />

facilitate ligand binding by ensuring the globular head extends beyond the glycocalyx of a<br />

cell (Kim et al. 2003a; Calderwood 2004).<br />

Transmembrane domain: Like the integrin α-subunit mentioned above, the integrin β-subunit<br />

transmembrane domain contains single-pass transmembrane α-helices which anchor the<br />

integrin β-subunit to the plasma membrane. Its interaction with the α-subunit is thought to be<br />

important in integrin activation (described below; Lau et al. 2008a; Lau et al. 2008b).<br />

8


Cytoplasmic domain: Integrin β-subunit cytoplasmic domains are typically short sequences of<br />

47-66 aa residues, with the exception of the β4 integrin subunit (1018 aa residues). <strong>The</strong> β4<br />

integrin cytoplasmic domain contains two pairs of C-terminal fibronectin (FN) type III<br />

repeats for binding to the keratin cytoskeleton (de Pereda et al. 1999). Most integrin β subunit<br />

cytoplasmic domains contain one or two NPxY motifs (where x is any amino acid), which<br />

upon tyrosine phosphorylation are phosphotyrosine-binding (PTB) domains that interact with<br />

a wide variety of signalling and cytoskeletal proteins. <strong>The</strong> integrin β subunit cytoplasmic<br />

domain recruits several signalling and cytoskeletal proteins, which are discussed in detail<br />

below.<br />

<strong>The</strong> integrin αβ heterodimer:<br />

Integrins appear to assume multiple different conformations (Ginsberg et al. 2005), in<br />

particular bent and extended conformations as evidenced by crystallography, nuclear<br />

magnetic resonance (NMR), electromicroscopy (EM) and Förster resonance energy transfer<br />

(FRET) studies (Figure 1.5).<br />

Figure 1.5 Structure of the αβ integrin heterodimer<br />

A cartoon representation of the integrin αβ heterodimer in an inactive state on the left and the active<br />

confirmation on the right. <strong>The</strong> inactive bent conformation is a schematic representation of the resolved crystal<br />

structure of αVβ3 (Xiong et al. 2001). <strong>The</strong> active extended conformation has been drawn to depict the different<br />

integrin subdomains. (Figure adapted from Wegener et al. 2007).<br />

9


Crystallography uncovered the structures of the large extracellular portions of the integrin<br />

heterodimers αvβ3 (Xiong et al. 2001; Xiong et al. 2002) and αIIbβ3 (Xiao et al. 2004). αVβ3<br />

was found to exist in a bent formation, in which the N-terminus was folded back towards the<br />

cell membrane (Xiong et al. 2001). A sharp bend was located between the thigh and calf-1<br />

domains within the genu domain of the αV subunit, and approximately between EGF domains<br />

1 and 2 of the β3 subunit. It was proposed that this bent conformation represented a low-<br />

affinity state, while the extended conformation was a high-affinity state (Jin et al. 2004;<br />

Nishida et al. 2006). However, the bent form had also been found to bind ligands with high<br />

affinity, as EM images show bent αVβ3 in complex with a fragment of fibronectin (Adair et<br />

al. 2005) and crystalised αVβ3 integrin in the bent conformation can bind to cyclic RGDsequence-containing<br />

peptide (Xiong et al. 2002). In addition, some integrins are thought to<br />

exist in an intermediate-affinity state (Chigaev et al. 2001).<br />

<strong>The</strong> membrane proximal regions of the α and β integrin subunit cytoplasmic domains in the<br />

low-affinity state are believed to be stabilised by a salt bridge between a conserved arginine<br />

in the α-subunit cytoplasmic domain, and an aspartic acid in the β-subunit cytoplasmic<br />

domain, and by hydrophobic contacts (Arnaout et al. 2005). Mutational analysis and<br />

computational modelling suggest the transmembrane domain stabilizes the inactive state, such<br />

that disruption of the transmembrane domain leads to integrin activation (Luo et al. 2005;<br />

Partridge et al. 2005). Several models of integrin activation propose orientational changes in<br />

the stalk region, however recent studies support a model which involves separation of the<br />

transmembrane stalk region (Kim et al. 2003a; Calderwood 2004), as seen in Figure 1.5.<br />

Interactions of the membrane-proximal regions of the α and β subunit cytoplasmic domains<br />

are believed to regulate integrin activation by stabilising integrins in low affinity states,<br />

whereas separation of the cytoplasmic domains and their stabilisation with cytoskeletal<br />

proteins is thought to cause integrin activation, as discussed further below (Luo et al. 2007).<br />

<strong>The</strong> three-dimensional structure of the integrin αIIb and β3 transmembrane segments of<br />

activated αIIbβ3 was recently resolved. <strong>The</strong> αβ extracellular domains were shown to connect<br />

to the transmembrane domains at different crossing angles (Lau et al. 2008a; Lau et al.<br />

2008b).<br />

10


1.1.2 Integrins and their extracellular ligands<br />

Integrins and their extracellular ligands are listed in Table 1.2. Of note, several integrins bind<br />

to many ligands via the tripeptide RGD motif, which is common to many ECM proteins<br />

(Elices et al. 1991). Hence, the complexity of integrin-ligand recognition can be partially<br />

attributed to the ability of certain integrins to interact with multiple ligands. Other integrins<br />

such as αEβ7 have only one identified ligand so far, namely E-cadherin in the case of αEβ7<br />

(reviewed in Krissansen et al. 2007). Several integrins bind in common to the same ligand.<br />

For example, the matrix proteins FN, LN, and CO; and the immunoglobulin superfamily cell<br />

adhesion molecules vascular cell adhesion molecule-1 (VCAM-1) and intercellular cell<br />

adhesion molecule-1 (ICAM-1) are each recognized by more than one integrin (reviewed in<br />

Krissansen et al. 2007). Integrins are either widely distributed or restricted to certain cell<br />

types, hence the environment in which they reside and need to function may account for the<br />

various ligand specificities.<br />

11


Table 1.2 A list of integrins and their known extracellular ligands<br />

Integrin Ligands<br />

Laminin receptors α3β1 LN; TSP; CO; FN; epiligrin; α3 80kDa fragment; fibrillin-2; MMP-1;<br />

ADAMs<br />

α6β1 LN; CO; ADAMs; epiligrin; midkine<br />

α7β1 LN, ADAMs<br />

α6β4 LN<br />

Collagen receptors α1β1 CO; LN<br />

α2β1 CO; LN; MMP-1; chondroadherin<br />

α10β1 CO<br />

α11β1 CO<br />

RGD receptors α5β1 FN; FG; ADAMs<br />

α8β1 FN; NN; TN; LAP; OP; VN; POEM; MAEG; QBRICK<br />

αvβ1 FN; VN; FG; OP; LAP; fibrillin-2, agrin; L1-CAM<br />

αvβ3 FN; VN; FG; VWF; OP; TN; BSP; TSP; LAP; CD31; ADAMs; MMP-2;<br />

uPA; uPAR; ICAM-4; CD31; fibrillin-1; osteoadherin; L1-CAM;<br />

developmental endothelial locus-1; agrin; lactadherin; metargidin; MMP-2;<br />

fibrin; Cyr61; Fisp12; microfibril-associated glycoprotein2; tropoelastin<br />

αIIbβ3 FN; VN; FG; OP; CO; TSP; VWF; CD40L; prothrombin; ICAM-4; VWF;<br />

L1-CAM; factor H<br />

αvβ5 VN; BSP; OP; LAP; FN; CO; lactadherin; sialoprotein; PAS-6/7, betaig-h3;<br />

ADAMs; MFG-E8, P2Y2 receptors<br />

αvβ6 FN; VN; TN; LAP<br />

αvβ8 FN; VN; CO; LN; LAP; fibrin<br />

α4/α9 receptors α4β1 FN; VCAM-1; MAdCAM-1; OP; FG; TSP; uPA; VWF; JAM-2; annexin;<br />

coagulation factor XIII, transglutaminase; ADAM28; ADAMs<br />

α9β1 TN; FN; OP; LN; CO; VWF; VCAM-1; ADAMs; uPA; transglutaminase,<br />

coagulation factor XIII; VEGF-C and VEGF-D<br />

Leukocyte-specific α4β7<br />

integrins<br />

FN; MAdCAM-1; VCAM-1; ADAMs<br />

αEβ7 E-cadherin<br />

αDβ2 ICAM-3; VCAM-1; FN; FG; VN; plasminogen; Cyr61<br />

αLβ2 ICAM1-5, CO; JAM-1<br />

αMβ2 ICAM-1; VCAM-1; FG; FN; VN; LN; CO; iC3b; factor X, galectin-3,<br />

complement factor H, CD23, ICAM-4; neutrophil inhibitory factor; Cyr61;<br />

GP1bα; serum factor x; JAM-3; Thy-1; RAGE; lipoprotein a; DC-SIGN;<br />

oligodeoxynucleotides; heparin; elastase; plasminogen; MMPs; β-glucans;<br />

high molecular weight kininogen; myeloperoxidase; azurocidin; haptoglobin;<br />

denatured proteins; lipopolysaccharides; pertussis toxin<br />

αXβ2 FG; ICAM-1; CO; iC3b; CD23, lipopolysaccharides, RAGE; Thy-1<br />

Notes: ADAM, a disintegrin and metalloprotease; BSP, bone sialic protein; CO, collagen(s); FG, fibrinogen; FN,<br />

fibronectin; iC3b, inactivated complement component 3b; ICAM, intercellular adhesion molecule; LAP, TGFβ latency<br />

associated protein; LN, laminin(s); MAdCAM-1, mucosal addressin cell adhesion molecule-1; MMP, matrix<br />

metalloprotease; NN, nephronectin; OP, osteopontin; RAGE, receptor for advanced glycation end-products; TG, tiggrin;<br />

TN, tenascin; TSP, thrombospondin; VCAM-1, vascular cell adhesion molecule-1; VN, vitronectin; VWF, von Willebrand<br />

factor. (Adapted from Krissansen et al. 2007)<br />

12


1.2. β7 integrins<br />

<strong>The</strong> integrin β7 (βp) subunit is made up of 798 aa with a predicted mass of 87 kDa (Yuan et<br />

al. 1990). When resolved by sodium dodecyl sulfate polyacrylamide gel electrophoresis<br />

(SDS-PAGE) it has a reduced size of 105-110 kDa (Parker et al. 1992). <strong>The</strong> gene is located at<br />

12q13.13, and is 10 kb in length with 14 exons (Table 1.1; Baker et al. 1992).<br />

Complementary DNA (cDNA) encoding the human β7 subunit was first isolated from T<br />

lymphocytes (Yuan et al. 1992). <strong>The</strong> β7 subunit forms two integrins by heterodimerisation<br />

with the integrin α4 and αE subunits, forming α4β7 and αEβ7 respectively (reviewed by<br />

Krissansen et al. 2007). <strong>The</strong> integrin α4β7 is expressed by mucosal lymphocytes, natural<br />

killer (NK) cells, and eosinophils (Holzmann et al. 1989), while αEβ7 is highly expressed by<br />

intraepithelial lymphocytes, and only a small minority of peripheral blood lymphocytes (Cerf-<br />

Bensussan et al. 1987).<br />

1.2.1 <strong>The</strong> β7 subunit cytoplasmic domain<br />

<strong>The</strong> integrin β7 subunit cytoplasmic domain is 52 aa in length (Figure 1.6; Yuan et al. 1992).<br />

It contains a conserved 2-LLv-iHDR-9 motif and an NPxY motif at positions 29-32.<br />

1 11 21 31 41 51<br />

| | | | | |<br />

β7 KLSVEIYDRR EYSRFEKEQQ QLNWKQDSNP LYKSAITTTI NPRFQEADSP TL<br />

Figure 1.6 <strong>The</strong> amino acid sequence of the cytoplasmic domain of the human β7 subunit.<br />

<strong>The</strong> amino acid sequence of the integrin β7 subunit cytoplasmic domain with key residues indicated. In bold<br />

letters are potentially phosphorylatable tyrosine/serine/threonine aa residues, and underlined is the conserved<br />

NPxY and NPRF motifs. <strong>The</strong> conserved LLv-iHDR motif is also present from residues 2-9. Highlighted in blue<br />

is the β7 cell adhesion regulatory domain.<br />

It also contains several potentially phosphorylatable serine, threonine and tyrosine residues as<br />

indicated in Figure 1.6. <strong>The</strong> β7 subunit cytoplasmic domain binds filamin more tightly than<br />

does the β1 cytoplasmic domain (Pfaff et al. 1998). Increased filamin (FLN) binding was<br />

found to restrict integrin-dependent cell migration by inhibiting transient membrane<br />

protrusion and cell polarization (Calderwood et al. 2001). <strong>The</strong> filamin binding site in the β7<br />

cytoplasmic domain was mapped to residues 35-40, which are necessary for high affinity<br />

13


inding of FLNa (one of the two most widely expressed members of the filamin family). <strong>The</strong><br />

amino acid isoleucine at positions 36 and 40 was found to be critical for filamin binding<br />

(Calderwood et al. 2001). Filamin binding supported low levels of cell migration of β7<br />

integrins, in contrast, filamin binding to β1 integrins was weaker with higher levels of cell<br />

migration (Calderwood et al. 2001). Splice variants of filamin lacking a 41 aa segment<br />

interacted more strongly with the β7 tail than did full-length filamin, suggesting alternative<br />

splicing of filamin may negatively regulate cell migration (Travis et al. 2004).<br />

<strong>The</strong> β7 cytoplasmic domain contains the NPxY motif that is important for integrin activation<br />

and signalling by talin and filamin (Pfaff et al. 1998). An additional motif “YDRREY”<br />

uniquely present in the β7 cytoplasmic domain (see Figure 1.6) was identified as a cell<br />

adhesion regulatory domain (CARD) of the β7 subunit (Krissansen et al. 2006b). A peptide<br />

containing the YDRREY motif was found to specifically block the binding of α4β7 to<br />

MAdCAM-1.<br />

Various PTB domain-containing proteins including Dok1, Numb, Dab1, Dab2, and tensin<br />

were shown to associate with the β7 cytoplasmic domain (Calderwood et al. 2003). <strong>The</strong><br />

human WD repeat protein WAIT-1 which is homologous to embryonic ectoderm<br />

development (eed; a member of the superfamily of polycomb group proteins), was found to<br />

specifically bind to the cytoplasmic domain of the β7 subunit but not to the β1 or β2<br />

cytoplasmic domains (Rietzler et al. 1998). However, its functional significance is not fully<br />

known. In studies of the human immunodeficiency virus (HIV) nef protein it was found that<br />

nef recruited eed (a repressor of transcription) from the nucleus to the plasma membrane,<br />

enabling transcriptional derepression (Witte et al. 2004). Similarly, activated integrin<br />

receptors recruited eed from the nucleus, suggesting a link between membrane-associated<br />

activation and transcriptional derepression (Witte et al. 2004).<br />

1.2.2 Integrin α4β7<br />

α4β7 (lymphocyte Peyer’s patch adhesion molecule-1, or LPAM-1) preferentially binds to the<br />

mucosal addressin cell adhesion molecule-1 (MAdCAM-1; Holzmann et al. 1989). It<br />

mediates the binding of lymphocytes to high endothelial venules of Peyer’s patches (PPs) in<br />

the small intestine, but not to peripheral lymph nodes (Holzmann et al. 1989). It is important<br />

in the homing of lymphocytes specifically to intestinal sites where MAdCAM-1 is expressed<br />

on the endothelial cells of PPs, mesenteric lymph nodes, and capillaries of the lamina propria<br />

14


(Berlin et al. 1993; Briskin et al. 1993). Additionally, α4β7 mediates the adhesion of T-<br />

lymphocytes within the lamina propria to microvessels at the tips of the intestinal villi, but<br />

not to the postcapillary venules of the PPs (Berlin et al. 1993; Briskin et al. 1993). α4β7 binds<br />

to the first Ig-domain of MAdCAM-1 with assistance from the second Ig domain. It also<br />

binds to RGD-containing peptides, VCAM-1, and the CS-1 fragment of fibronectin (Chan et<br />

al. 1992; Ruegg et al. 1992).<br />

α4β7 plays an essential role in lymphocyte extravasation across the endothelium (Hogg et al.<br />

2003). After selectin-mediated tethering and fast rolling of lymphocytes on the vessel wall,<br />

the integrin α4β7 mediates slow rolling and firm attachment of lymphocytes in conjunction<br />

with the β2 integrins prior to transendothelial migration (refer to Figure 1.7). α4β7 is also<br />

critical for the transmigration of mast cell progenitors into the small intestine, but not the<br />

lungs (Gurish et al. 2001); haematopoietic progenitor cell homing to bone marrow (Masellis-<br />

Smith et al. 1997); and the migration of eosinophils to the intestine (Brandt et al. 2006).<br />

Figure 1.7 Multistep model of T cell transmigration<br />

A cartoon representation of the multistep model of T cell migration across the endothelium (above), and a table<br />

showing the adhesion molecules involved in each step of the transmigration process (below). (Figure adapted<br />

from Hogg et al. 2003).<br />

15


Chemokines can activate α4β7 to adhere and migrate on MAdCAM-1. <strong>The</strong> immobilised CC<br />

chemokine ligands (CCL), 21, 25, 28, and CXC-chemokine ligand (CXCL) 12 all converted<br />

the rolling adhesion of human lymphocytes on MAdCAM-1 to static arrest (Hogg et al.<br />

2003). However, only CCL21 and CXCL12 also triggered a motile phenotype characterized<br />

by lamellipodia and uropod formation. α4β1-VCAM-1 and α4β7-MAdCAM-1 interactions<br />

were reported to operate independently to support lymphocyte adhesion under flow (Hogg et<br />

al. 2003). Different chemokines may act in concert in triggering integrin-mediated arrest and<br />

promoting motility and transendothelial migration (Miles et al. 2008). In addition to<br />

activation by chemokines, α4β7 is also regulated by transforming growth factor (TGF)-β1<br />

(Lim et al. 1998; Bartolome et al. 2003), as well as the chemoattractants interleukin (IL)-8<br />

and formyl-methionyl-leucyl-phenylalanine (fMLP; Sadhu et al. 1998).<br />

α4β7 is critical for the homing of lymphocytes to the gut and mucosal sites which are<br />

normally chronically inflamed (Butcher et al. 1999). <strong>The</strong> formation of the gut-associated<br />

lymphoid tissue (GALT) is severely impaired in β7 gene knockout mice due to the failure of<br />

the β7 -/- lymphocytes to adhere to blood vessel walls at the site of transmigration into the<br />

GALT (Wagner et al. 1996). β7-deficient mice also show aberrant distribution of mast cells in<br />

the small intestine, and are more susceptible to Cryptosporidium parvum at the early stages of<br />

infection (Mancassola et al. 2004).<br />

1.2.3 Integrin αEβ7<br />

<strong>The</strong> second α subunit that associates with the β7 subunit is the αE subunit (CD103), which<br />

was discovered in the late 1980’s (Cerf-Bensussan et al. 1987). It forms the αEβ7 integrin<br />

(HML-1, αIEL or M290 in mice) whose expression is restricted almost exclusively to T cells<br />

and dendritic antigen presenting cells (APC) in the mucosal immune system (Kilshaw 1999).<br />

It is particularly prominent on intraepithelial lymphocytes in the gut (Kilshaw 1999). <strong>The</strong><br />

distribution of αEβ7 is largely attributed to the restricted expression of the αE subunit,<br />

because the β7 subunit in association with α4 is widely distributed on T and B cells, mast<br />

cells, activated macrophages, and eosinophils (Kilshaw 1999). In inflammatory diseases, αE<br />

is detectable on T cells in the rheumatoid joint (Baumgart et al. 1996), CD8 T cells<br />

infiltrating the kidney tubular epithelium during acute rejection of renal allografts (Yuan et al.<br />

2005), and in lachrymal glands in Sjogren’s syndrome (Fujihara et al. 1999).<br />

16


αEβ7 null mice show a reduction in the numbers of mucosal T cells and display T cell-<br />

dependent dermatitis of unknown etiology (Lefrancois et al. 1999; Schon et al. 1999).<br />

Interestingly αEβ7 null mice transplanted with pancreatic islet cell allografts did not reject the<br />

implants, suggesting impaired systemic immunity (Feng et al. 2002).<br />

<strong>The</strong> αE subunit uniquely contains an X-domain N-terminal to the I-domain (Shaw et al.<br />

1994), which is important for ligand binding. <strong>The</strong> only ligand for αEβ7 discovered so far is E-<br />

cadherin, which is an epithelial homophilic adhesion molecule (Kilshaw 1999). However,<br />

αEβ7-transfected K562 cells were found to bind to human intestinal microvascular<br />

endothelial cells (HIMEC) independently of E-cadherin, suggesting a role for αEβ7 in<br />

lymphocyte homing or interaction with the intestinal endothelium, mediated by a ligand that<br />

is yet to be identified (Strauch et al. 2001).<br />

Expression of αEβ7 is induced by TGF-β, which is prominent in all tissue microenvironments<br />

in which αE-expressing cells are found (Kilshaw et al. 1990). TGF-β caused a rapid increase<br />

in the expression of αE messenger RNA (mRNA) in mouse T lymphoma cells (Robinson et<br />

al. 2001). TNFβ was found to differentially regulate αEβ7 in mouse intestinal lymphocytes<br />

(Ni et al. 1995). TNFβ was found to have no effect on αEβ7 expression in small intestine<br />

intraepithelial lymphocytes (SIEL), partially downregulated αEβ7 expression in lamina<br />

propria lymphocytes (LPL) and strongly upregulated in spleen lymphocytes (SL). Thus,<br />

TNFβ regulated expression of αEβ7 depends on the microenvironments in which cells reside,<br />

where the lymphocytes may play different roles (Ni et al. 1995).<br />

1.3. Integrin signalling<br />

Extracellular ligand-binding generates intracellular signals (outside-in signalling), and<br />

conversely ligand-binding is regulated by signals from within the cell (inside-out signaling;<br />

reviewed in Krissansen et al. 2007). Hence, integrins serve as a transmembrane bridge<br />

between the extracellular milieu and actin microfilaments of the cytoskeleton. <strong>The</strong> short<br />

integrin cytoplasmic tails lack intrinsic enzymatic activity, therefore there is the necessity for<br />

the recruitment of adaptor molecules and signalling proteins for effector function and integrin<br />

regulation (reviewed in Krissansen et al. 2007). <strong>The</strong> integrin cytoplasmic domain therefore<br />

plays a crucial role in signalling, as it serves as the interface through which many signalling<br />

and cytoskeletal molecules exert their actions. A list of the many intracellular ligands<br />

identified for integrins is given in Table 1.3.<br />

17


Table 1.3 A list of integrins and their known intracellular ligands<br />

Integrin tail<br />

α Calreticulin, caveolin, GTP-bound G(h)/TG2, ancient ubiquitous protein 1 (Aup1)<br />

α1 Paxillin, talin, α-actinin, FAK, actin<br />

α2 F-actin<br />

α3 DRAL/FHL2, nucleotide exchange factor Mss4, putative tumour suppressor protein BIN1<br />

α4 Paxillin<br />

α5 Tap-20, TIP-2/GIPC, nischarin<br />

α6 TIP-2/GIPC<br />

α7 DRAL/FHL2<br />

α8 ?<br />

α9 Paxillin, spermidine/spermine N (1)-acetyltransferase (SSAT)<br />

α10, α11 ?<br />

αv ?<br />

αL Rap1?<br />

αM, αX, αD ?<br />

αIIb Talin, CIB, ICln<br />

β Myosin-X<br />

β1 Talin, filamin, α-actinin, paxillin, ILK, FAK, skelemin, melusin, MIBP, ICAP-1, CD98<br />

Dab1, X11a, EPS8, EB-1, GAPCenA, tensin, merlin, phospholipase Cγ (PLCγ), Nckinteracting<br />

kinase (NIK), 14-3-3β, Kindlin-2<br />

β2 Talin, filamin, α-actinin, FAK, cytohesin-1, cytohesin-3, Rack-1, JAB1, Dab1, Dok-1,<br />

ILK Rack-1,<br />

β3 Talin, filamin, paxillin, FAK, GRb2, Shc, ERK1, myosin, skelemin, β3-endonexin,<br />

CD98, Pyk2, Numb, Dab1, Dab2, X11a, X11g, Lin10, JIP, EPS8, RGS12, CED6, EB1,<br />

GAPCenA, tensin, IRS-1, Dok-1, ILK, Syk, ZAP-70, Kindlin-2<br />

β4 F-actin, plectin, p27(BBP), ERBIN, 12-lipoxygenase<br />

β5 Talin, Rack-1, p21-activated kinase 1, Numb, Dab1, Dab2, EPS8, tensin, Dok-1, annexin-<br />

V<br />

β6 ERK-2<br />

β7 Talin, filamin, WAIT-1, Shc, Numb, Dab1, Dab2, X11a, X11g, JIP, CED6, EB-1,<br />

GAPCenA, tensin, Dok-1<br />

β8 ?<br />

(Adapted from Krissansen et al. 2006a)<br />

18


<strong>The</strong> binding of integrins to their ligands leads to integrin clustering and changes in integrin-<br />

avidity, and the recruitment of actin filaments through interactions of the cytoplasmic<br />

domains with cytoskeletal proteins such as talin, vinculin, α-actinin and filamin (reviewed in<br />

Krissansen et al. 2006a). Integrin activation first came to light from studies of integrin αIIbβ3<br />

on platelets, where thrombin binding to its receptor led to increased fibrinogen binding<br />

(reviewed in Kieffer et al. 1990; Phillips et al. 1991; Krissansen et al. 2006a). It was then<br />

shown that deletions of the cytoplasmic domain of α- or β- tails resulted in constitutively<br />

active integrins (O'Toole et al. 1991; Crowe et al. 1994). Over the years, there have been<br />

extensive studies on integrin activation and signalling.<br />

1.3.1 Models of integrin activation<br />

As mentioned, it is thought that integrins in their inactive state are in a bent conformation.<br />

<strong>The</strong> integrin must be extended to enable binding to its ligand some 200 Å away. Currently<br />

there are two prevailing models of integrin activation, the switch-blade and deadbolt models.<br />

Switch-blade model: In the switch-blade model, inside-out activation causes a separation of<br />

the cytoplasmic and transmembrane domains of the two subunits (Kim et al. 2003b), causing<br />

a jack-knife-like swing-out extension at the “knee” of the integrin β-subunit hybrid domain<br />

which occurs concurrent with or after ligand binding (Figure 1.8). <strong>The</strong> swing-out motion has<br />

been proposed to participate in the initiation of outside-in signalling induced by ligand<br />

binding. This is supported by the calculated Stokes radius based on gel elution profiles which<br />

increases from 56 Å (in 1 mM Ca 2+ ), to 60 Å (in Mn 2+ ), to 63 Å upon addition of cyclic RGD<br />

tripeptide (Takagi et al. 2002). In addition, X-ray solution scattering, EM imaging and crystal<br />

structures show an outward swing of the β-subunit hybrid domain from 45˚ to 80˚ (Mould et<br />

al. 2003b; Takagi et al. 2003; Xiao et al. 2004).<br />

19


Figure 1.8 Switch blade model of integrin activation<br />

A cartoon depiction of the switch blade model of integrin activation. <strong>The</strong> different domains of the integrin α and<br />

β-subunits are shown in different colours. <strong>The</strong> integrin α-subunit is shown without (A), and with the α subunit Idomain<br />

(B). <strong>The</strong> illustrations depict the conformations of the inactive state in which the head piece is bent<br />

towards the plasma membrane (left), the extended conformation with a closed headpiece (middle), and the active<br />

state with an open headpiece for ligand binding (right). <strong>The</strong> dotted lines show the flexibility of the integrin βsubunit<br />

stalk region. (Adapted from Luo et al. 2007).<br />

Deadbolt model: <strong>The</strong> deadbolt model of integrin activation was coined by taking account of<br />

the discrepancies in the switchblade model (Arnaout et al. 2007), such as the ability of<br />

integrins in the bent integrin conformation to bind to ligand (Xiong et al. 2003). In the<br />

deadbolt model, the site of ligand-binding is thought to be tightly bound (deadbolt) when in<br />

an inactive form. When integrins are activated the head region does not extend out straight in<br />

response to straightening of the “knee” region, but rather changes in the transmembrane angle<br />

disengage the hybrid domain from the I-domain or the β-subunit I-like-domain (Figure 1.9).<br />

20


<strong>The</strong> β-terminal domain (β-TD) releases the β-subunit I-like-domain which is responsible for<br />

ligand binding, allowing it to undergo a subtle conformational change and adopt a high-<br />

affinity ligand-binding state. Extension of the receptor at both the genu domain of the α-<br />

subunit and the interface between the integrin epidermal growth factor-1 and IEGF-2 domains<br />

of the β subunit is proposed to occur after ligand-binding in this model (Arnaout et al. 2007).<br />

This is supported by the ability of the bent integrin to bind its ligands in solution, not by leg<br />

straightening, but through the Mn 2+ -induced shift of the ADMIDAS metal ion coordination to<br />

a ligand bound state, causing tertiary changes as seen in the active state (Mould et al. 2003a).<br />

In support, deletion of integrin transmembrane and cytoplasmic segments leads to partial<br />

disengagement of the deadbolt (Humphries 2000). Furthermore, recent cryoelectron<br />

tomography of integrin αIIbβ3 revealed membrane heights remained the same after Mn 2+<br />

activation (Ye et al. 2008).<br />

Figure 1.9 Deadbolt model of integrin activation<br />

A cartoon representation of the deadbolt model of integrin activation. <strong>The</strong> α-subunit without an I-domain is<br />

shown in the upper diagram, and with an I domain in the lower diagram. (A) and (D) denote the inactive integrin<br />

conformation, while (B) and (E) shows the release of the deadbolt to an active state, and (C) indicates a ligand<br />

bound state. (Adapted from Arnaout et al. 2007).<br />

21


1.3.2 Integrin regulation by inside-out signalling<br />

Inside-out integrin signalling refers to the transmission of signals from within the cell to the<br />

extracellular domain of the integrin, affecting the affinity of integrins for their ligands<br />

(reviewed in Hynes 2002). This form of regulation is precisely controlled in time and space,<br />

as exemplified in platelet aggregation and leukocyte transmigration (reviewed in Hynes<br />

2002). <strong>The</strong> switching from low- to high-affinity states is rapid, occurring in subsecond time<br />

frames, and is reversible in less then a minute. Leukocyte integrins generally exist in a lowaffinity<br />

state, and are activated intracellularly by signals from within the cell in response to<br />

extracellular chemicals and factors (Lollo et al. 1993; Constantin et al. 2000), or mechanical<br />

stress (Zwartz et al. 2004).<br />

As mentioned above, integrin α- and β-subunit cytoplasmic domains contain conserved<br />

sequence motifs which interact with intracellular proteins, and are important for integrin<br />

regulation and function. Sequence motifs within the α- and β- subunits that are recognized by<br />

intracellular signalling proteins have been mapped (Figure 1.10).<br />

<strong>The</strong> GFF(R/K)/R motif in the α-subunits is recognised by calreticulin, and appears to be<br />

essential for subunit association and stability (Leung-Hagesteijn et al. 1994). Calreticulin<br />

transiently interacts with integrins during cell attachment and spreading. <strong>The</strong>re are many<br />

cytoplasmic proteins that interact with the β-subunit, as discussed below.<br />

22


Figure 1.10 Sequence alignments of the human integrin β subunit (a), and α subunit (b) cytoplasmic<br />

domains.<br />

Comparison of the primary sequences of the cytoplasmic domains of the integrin α and β subunits. Highly<br />

conserved amino acid residues are highlighted, and provided as a consensus below the alignments. (Adapted<br />

from Krissansen et al. 2006b).<br />

Intracellular ligands of the integrin β-subunit<br />

This thesis will attempt to summarise the features of some important intracellular binding<br />

partners for the integrin β-subunit, with specific attention to binding partners relevant to the<br />

aims of the thesis. <strong>The</strong> current model of integrin activation places the binding of the<br />

cytoskeletal protein talin to the integrin β-subunit cytoplasmic domain as the common last<br />

step in the pathway of inside-out signalling (Tadokoro et al. 2003; Tanentzapf et al. 2006).<br />

Talin: <strong>The</strong> cytoskeletal adaptor protein talin is critical for integrin activation (Tadokoro et al.<br />

2003; Tanentzapf et al. 2006). Talin consists of a large C-terminal rod domain and an Nterminal<br />

FERM (band 4.1, ezrin, radixin, and moesin) domain with three subdomains F1, F2,<br />

and F3 (Garcia-Alvarez et al. 2003). Talin plays a crucial role in regulating integrin affinity.<br />

<strong>The</strong> binding of the talin head region to the integrin cytoplasmic domain causes the<br />

dissociation of the α- and β- cytoplasmic domains and induces conformational changes in the<br />

23


extracellular regions to increase affinity for ligand (Tadokoro et al. 2003). In addition, talin is<br />

essential in the early coupling of integrins to the cytoskeleton following the binding of<br />

integrins to their extracellular ligands (Jiang et al. 2003). Talin reinforces the integrin-<br />

cytoskeleton linkage through recruitment of other cytoskeletal proteins such as paxillin,<br />

vinculin, α-actinin, tensin, and zyxin (Giannone et al. 2003). Talin depletion studies revealed<br />

that talin is not important in the initial signalling processes required for spreading of cells, but<br />

rather it provides an important mechanical linkage between the ligand-bound integrin and the<br />

actin cytoskeletal network (Zhang et al. 2008a).<br />

Talin binds to the central NPxY motif of several integrin β subunit cytoplasmic domains<br />

(Calderwood et al. 2002). Using structure-based mutagenesis, the talin F3 domain was shown<br />

to complex with the membrane proximal region of the cytoplasmic tail of the β3 integrin<br />

subunit (Wegener et al. 2007). This result led to a proposed model of talin-induced integrin<br />

activation. In this model, chemokine or growth factor receptor activation of cells leads to the<br />

activation of the small G-protein Rap1 by protein kinase C (PKC) or Rap1-GEF, which is<br />

then recruited to the plasma membrane. RIAM (Rap1<strong>–</strong>GTP-interacting adapter molecule)<br />

then associates with talin, freeing talin from its autoinhibitory interactions and rendering it<br />

available for binding to integrins (Lafuente et al. 2004; Han et al. 2006). Other mediators of<br />

talin activation include phosphatidylinositol 4,5-bisphosphate (PIP2; Martel et al. 2001) and<br />

calpain (Yan et al. 2001). <strong>The</strong> F3 subdomain of talin, which resembles a phosphotyrosine-<br />

binding domain, binds to the NPxY motif in the membrane distal region of integrin β subunit<br />

cytoplasmic domains. While still maintaining an interaction with the membrane distal region,<br />

talin then engages the membrane proximal portion, which destabilises the salt bridge between<br />

the α- and β-subunits, and stabilises the helical structure of the membrane proximal region.<br />

This changes the conformation of the transmembrane helix causing a separation or reorientation<br />

of the integrin tails leading to integrin activation. It should be noted, however, that<br />

the exact role of the salt bridge in vivo has been questioned from the results of genetic studies<br />

of β1 integrins in mice, in which mutations of the aspartic acid to alanine in the membrane<br />

proximal region had no apparent function under physiological conditions (Czuchra et al.<br />

2006). Talin binding to integrin β-subunits is thought to be the final step leading to integrin<br />

activation, but exactly how this step is regulated still remains to be solved. It is possible that<br />

phosphorylation (Tapley et al. 1989), proteolysis (Yan et al. 2001) and phosphoinositide<br />

binding (Martel et al. 2001) all play roles.<br />

24


Cytoplasmic proteins<br />

Other cytoplasmic proteins that influence integrin activation are described briefly below.<br />

Kindlin-2: Kindlin-2 (mitogen-inducible gene-2, MIG-2), another adaptor protein which<br />

possesses a FERM domain, has recently been shown to bind to β1 and β3 integrins to enhance<br />

talin-mediated integrin activation (Shi et al. 2007; Montanez et al. 2008). Kindlin-2 binds to<br />

the membrane distal NxxY motif of integrin β subunits through its PTB site. Loss of kindlin-<br />

2 results in peri-implantation lethality and impaired integrin activation in mice and embryonic<br />

stem cells respectively. Kindlin-2-deficiency also prevents the formation of multiple focal<br />

adhesions (FAs) and recruitment of integrin-linked kinase (ILK) to integrin adhesion sites<br />

(Dowling et al. 2008; Montanez et al. 2008).<br />

Filamin: Filamin is an actin binding protein widely expressed in animal cells, which is a<br />

potent actin filament cross-linker (Stossel et al. 2001). Filamin subunits of 280 kDa form noncovalently<br />

associated dimers. Filamin contains an actin-binding domain at its N-terminus<br />

followed by 24 Ig-like domains (IgFLN1-24). It was found that IgFLNa21 binds to the β7<br />

integrin subunit cytoplasmic domain. Filamin binds to the same region as talin within the<br />

integrin β-subunit cytoplasmic domain. It is thought that filamin acts as a negative regulator<br />

of talin-induced integrin activation, which may explain the known inhibitory role of filamin<br />

on cell migration (Calderwood et al. 2001). Threonine phosphorylation at aa 758 in the triple<br />

threonine motif (TTT) immediately downstream of the central NPxY motif of β2 integrins<br />

prevents filamin binding. In contrast, threonine phosphorylation enables the molecular<br />

adaptor protein 14-3-3 to bind to the β2 subunit cytoplasmic domain (Kiema et al. 2006;<br />

Takala et al. 2008). This suggests that phosphorylation of threonine 758 may provide a<br />

molecular switch for filamin and 14-3-3 binding.<br />

Migfilin: Migfilin possesses three LIM domains at its C-terminus, a central proline-rich<br />

region, and an N-terminal portion which lacks identifiable structural domains (Tu et al. 2003).<br />

<strong>The</strong> LIM domains mediate binding to kindlin-2 and are required for localisation to focal<br />

adhesions (Tu et al. 2003). Migfilin is thought to provide a link between the actin<br />

cytoskeleton and integrin-extracellular matrix contact sites. Migfilin binds preferentially to<br />

the filamin IgFLNa21 domain, similarly to the β7 integrin-filamin (IgFLNa21) interaction,<br />

suggesting it may provide a mechanism for switching between different integrin-cytoskeleton<br />

linkages (Lad et al. 2008).<br />

25


Dok1: Dok1 was found to bind weakly to the central NPxY in the β3 subunit cytoplasmic<br />

domain, where the affinity of interaction was strongly increased by phosphorylation of<br />

tyrosine residue Y747 (Oxley et al. 2008). Phosphorylation of Y747 inhibited talin binding,<br />

suggesting that phosphorylation at this position can act as a switch by reversing the<br />

preference for talin to Dok1 to bind the β3 subunit cytoplasmic domain (Oxley et al. 2008).<br />

Paxillin: Paxillin is an adaptor protein found in focal adhesions, which is highly tyrosine<br />

phosphorylated (Schaller et al. 1995b). It regulates focal adhesion dynamics and cell<br />

migration . Paxillin binds to the tyrosine kinases focal adhesion kinase (FAK) and src, as well<br />

as other cytoskeletal proteins such as vinculin (Turner 1998; Wade et al. 2006). It binds with<br />

high affinity to the α4 integrin subunit cytoplasmic domain (Liu et al. 1999; Liu et al. 2000).<br />

α-actinin: α-actinin is an actin binding protein consisting of two rod-shaped anti-parallel 100<br />

kDa monomers (Ylanne et al. 2001). It has actin-binding motifs at either ends of the rod<br />

enabling α-actinin to cross-link actin filaments into bundles (Ylanne et al. 2001). α-actinin<br />

binds to the integrin β1 subunit cytoplasmic domain in vitro (Otey et al. 1990) and binds to<br />

activated β2 integrins, but not to inactivated integrins in vivo (Pavalko et al. 1993).<br />

Vinculin: Vinculin is an adaptor protein that is a key regulator of FAs (Zamir et al. 2001;<br />

Ziegler et al. 2006). Vinculin is comprised of three domains, namely an N-terminal head, a<br />

flexible proline-rich hinge (neck), and a C-terminal tail domain (Winkler et al. 1996).<br />

Intramolecular interaction between the head and tail domains constrain vinculin in an inactive<br />

conformation (Bakolitsa et al. 2004). Upon recruitment to FAs, vinculin switches to an open<br />

active conformation allowing direct interactions with other cytoskeletal proteins such as the<br />

binding of talin and α-actinin to the head domain, and actin and paxillin to the tail domain<br />

(Zamir et al. 2001; Ziegler et al. 2006). <strong>The</strong> interaction of the N-terminal head of vinculin<br />

with talin drives the clustering of integrins in cell-matrix adhesions. It also leads to the<br />

recruitment of paxillin by a mechanism that is independent of the paxillin binding site of<br />

vinculin. <strong>The</strong> binding of actin to the tail of vinculin is a major link in the interaction of focal<br />

adhesions with the actin cytoskeleton (Humphries et al. 2007).<br />

Radixin: Radixin is another FERM domain-containing protein that activates integrins (Tang<br />

et al. 2007). Radixin enhances the adhesive activity of the integrin αMβ2, by binding to a<br />

membrane proximal region (possibly the LxxLxxYRRF sequence) of the β2 subunit (Tang et<br />

al. 2007).<br />

26


PTB proteins: Several PTB proteins are known to bind to the integrin β-subunit cytoplasmic<br />

domain with high affinity, including syk and Numb, however none of them are able to<br />

activate integrins (Calderwood et al. 2002). <strong>The</strong> PTB-domain containing protein tensin binds<br />

to the β subunit cytoplasmic domain in a talin-like manner, but with lower affinity<br />

(McCleverty et al. 2007). Many other PTB proteins interact with the NPXY motif in β-<br />

subunits including Shc, Dab1, Dab2, X11α, X11γ, CeLin10, JIP, EPS8, RGS12, CED6, EB-1<br />

and GAPCenA, IRS-1 and ICAP-1, as listed in Table 1.3 (Krissansen et al. 2006a). It is<br />

possible that one or a number of the latter proteins can activate or regulate integrins.<br />

1.3.3 Intracellular ligands of the α-subunit<br />

So far this thesis has focussed largely on intracellular ligands that interact with the<br />

cytoplasmic domain of the β-subunit. However, it should be noted that the integrin α-subunit<br />

cytoplasmic domain is not a passive player in integrin activation and/or signalling. As can be<br />

seen in Table 1.3, the integrin α-subunit has several intracellular ligands. Of note,<br />

interactions between talin and the αIIb cytoplasmic domain have been described (Knezevic et<br />

al. 1996). Calcium and integrin binding protein 1 (CIB1) interacts with the membrane<br />

proximal region of the αIIb cytoplasmic domain and interferes with talin binding (Yamniuk et<br />

al. 2006), suggesting it acts as a potential negative regulator of talin induced-integrin<br />

activation. <strong>The</strong> integrin α4 subunit cytoplasmic domain has been showed to bind to paxillin<br />

(Liu et al. 1999), which allows talin to bind indirectly to the cytoplasmic domain. Spatial α4<br />

phosphorylation at the leading edge of cell migration limits the interaction of α4 with paxillin.<br />

However the α4-paxillin complex at the side and trailing edge can recruit ADP-ribosylation<br />

factor GTPase-activating protein (Arf-GAP) which inhibits Rac and blocks lamellipodia<br />

formation (Nishiya et al. 2005).<br />

Certain integrins including α1β1, α5β1, and αvβ3 can signal through the Fyn/Shc pathway<br />

(Wary et al. 1996; Guo et al. 2004). <strong>The</strong> α-subunits of theses integrins are coupled to Fyn<br />

through association with the oligomeric transmembrane protein caveolin-1 (Wary et al. 1996;<br />

Guo et al. 2004). Upon integrin ligand-binding, Fyn is activated and subsequently recruits and<br />

phosphorylates Shc at Tyr317, leading to Grb2-SOS complex binding and the activation of<br />

the Ras-Raf-MEK-Erk pathway which promotes cell cycle progression (Wary et al. 1996;<br />

Guo et al. 2004).<br />

27


1.3.4 Integrin outside-in signalling<br />

Changes in integrin avidity (clustering) complement changes in integrin affinity, and are<br />

critical for cell signalling (Hato et al. 1998). Integrin clustering, which is RhoA-dependent,<br />

takes place within seconds when induced by chemokines (reviewed in Hynes 2002). It leads<br />

to the activation of signalling cascades and recruitment of multiprotein complexes to FAs<br />

(reviewed in Hynes 2002). FA formation at the cytoplasmic face of the plasma membrane<br />

connects integrins through adaptor proteins to the actin cytoskeleton, as well as receptor<br />

tyrosine kinases (RTKs).<br />

Integrin-mediated adhesion controls multiple aspects of cellular behaviour. It can trigger<br />

calcium influx and an increase in intracellular pH, activate tyrosine and serine/threonine<br />

protein kinases, regulate the activity of the Rho family of small GTPases, activate inositol<br />

lipid metabolism, and trigger gene expression leading to cytokine and metalloprotease<br />

production (reviewed in Krissansen et al. 2006a).<br />

A prominent biochemical event in integrin outside-in signalling is protein tyrosine<br />

phosphorylation due to activation of the src and FAK family of protein kinases (Arias-<br />

Salgado et al. 2003). Src and its inhibitor Csk are constitutively bound to the C-terminus of<br />

the integrin β subunit cytoplasmic domain. Ligand-binding induces src activation by<br />

dissociation of Csk leading to src transphosphorylation in clustered integrins and/or<br />

recruitment of tyrosine phosphatases such as receptor tyrosine phosphatase alpha (RPTPα), or<br />

PTP-1B (von Wichert et al. 2003b; Arias-Salgado et al. 2005). Src-induced tyrosine<br />

phosphorylation of phosphatidylinositol (4) phosphate 5 kinase type Iγ (PIPKIγ) at tyrosine<br />

644 increases its affinity for the talin head domain, which dissociates the integrin β-tails from<br />

the talin F3 domain (Ling et al. 2003). PIP2, a lipid second messenger, is generated that coactivates<br />

a number of focal adhesion proteins, including vinculin and talin. An increase in<br />

PIP2 may enhance talin-integrin association and displace PIPKIγ from talin, leading to a<br />

reduction in PIP2, suggesting a self-limiting dynamic process. Integrins may also activate Syk<br />

in focal complexes (Woodside et al. 2001). Syk binds directly to the last 4 amino acids of the<br />

C-terminal portion of the β3 tail. It binds tensin, converting focal adhesions into fibrillar<br />

adhesions (Woodside et al. 2001). Src can phosphorylate and activate FAK and p130Cas,<br />

where the latter is essential in converting mechanical force into biochemical signals (Sawada<br />

et al. 2006). This creates a binding site for the adaptor protein Crk either through the<br />

association of an activator of Ras, son of sevenless (SOS), or through the association of Rap<br />

28


guanine nucleotide exchange factor (GEF) 1 (also known as C3G) which subsequently<br />

activates Erk and Raf respectively (Sawada et al. 2006). <strong>The</strong> adaptor protein Nck is<br />

stimulated upon integrin-mediated adhesion, leading to association with p130Cas which can<br />

lead to Erk activation. In other possible pathways, the protein tyrosine phosphatase Shp2 is<br />

recruited, which downregulates FAK and thereby permits the incorporation of α-actinin into<br />

focal adhesions (von Wichert et al. 2003a).<br />

FAK is autophosphorylated at Tyr397 upon integrin-mediated adhesion, which creates a<br />

binding site for Src (Schlaepfer et al. 1998). Src subsequently phosphorylates FAK at Tyr925,<br />

creating a binding site for the adapter protein Grb2 that links FAK to SOS, a guanine<br />

nucleotide exchange factor for Ras which activates the Ras-Raf-MEK-ERK pathway<br />

(Schlaepfer et al. 1998). In addition, the adaptor protein Shc also binds to the phosphorylated<br />

Tyr397 residue of FAK. Shc is then phosphorylated by FAK and Src and provides an<br />

alternate link to Grb2.<br />

Phosphoinositide 3-kinases (PI-3-kinase) binds to phosphorylated FAK at Tyr397 upon<br />

integrin-mediated adhesion (Chen et al. 1996). Active PI-3-kinase may activate Erk through<br />

its protein kinase activity or through SOS and phosphatidylinositol-3,4,5-trisphosphate (PIP3;<br />

Chen et al. 1996). PI-3-kinase is recruited to the plasma membrane by Cbl binding (Ojaniemi<br />

et al. 1997; Zell et al. 1998; Feigelson et al. 2001), which can stimulate ubiquitin-mediated<br />

turnover of both Src-family kinases and PI-3-kinase itself (Fang et al. 2001). PI-3-kinase is<br />

activated by integrins in a FAK and src independent pathway, resulting in the<br />

phosphorylation of Akt, the generation of PIP2 and PIP3, and the recruitment of signalling<br />

cascade proteins (Velling et al. 2004).<br />

<strong>The</strong> serine/threonine (ser/thr) kinase ILK binds directly to and may phosphorylate integrin β1<br />

and β3 subunit cytoplasmic domains (Legate et al. 2006). ILK binds to the LIM adaptor<br />

proteins PINCH-1, and PINCH-2 (Tu et al. 1999). It links the actin binding proteins α, β, and<br />

γ parvins to the integrin β cytoplasmic domain, hence linking integrins to the cytoskeleton<br />

and signalling pathways (Legate et al. 2006). <strong>The</strong> assembly of an ILK-PINCH-parvin<br />

complex is essential for the localisation of FAs (Zhang et al. 2002), while the binding of ILK<br />

with PINCH stabilises these proteins and protects them from proteasome-mediated<br />

degradation (Fukuda et al. 2003; Legate et al. 2006).<br />

29


1.3.5 Force sensing and focal adhesions<br />

Movement of focal adhesion proteins has been visualised by total internal reflection<br />

fluorescence microscopy of fluorescently tagged proteins (Brown et al. 2006; Hu et al. 2007).<br />

<strong>The</strong>se studies found that active cells display vigorous retrograde movement of the cell<br />

membrane resulting from continuous pulling of myosin II while ECM-integrin bonds remain<br />

stationary. <strong>The</strong> transmission of motion from F-actin to proteins in the focal adhesion (FAK,<br />

paxillin, zyxin) to integrin decreased. <strong>The</strong> actin binding protein α-actinin displayed the<br />

greatest motion, due to its late incorporation into focal adhesions. Talin and vinculin were<br />

only partially coupled to the F-actin motion which prevents transmission of forces from<br />

actomyosin contractions to the ECM through integrins (Evans et al. 2007).<br />

As mentioned above, p130Cas is a primary sensor and transducer of externally applied force<br />

into biochemical signals (Sawada et al. 2006). p130Cas binds to FAK through the N-terminal<br />

src homology (SH)3 domain, possibly linking the β-integrin cytoplasmic domain and/or α-<br />

integrin cytoplasmic domain through paxillin. Mechanical extension of the central domain in<br />

p130Cas enhances tyrosine phosphorylation by src, which leads to the association of Crk and<br />

C3G, leading to Rap1 activation and ERK signalling (Tamada et al. 2004; Defilippi et al.<br />

2006).<br />

<strong>The</strong> small GTPases Rac and RhoA also play a role in integrin activation (Nobes et al. 1995;<br />

Laudanna et al. 1996). Both Rac1 and its downstream effector molecule, RhoA, induce the<br />

adhesion of the α4β7 integrin to its physiological ligand MAdCAM-1 (Zhang et al. 1999).<br />

<strong>The</strong> Rho-related guanosine triphosphatase, Cdc42, induces filopodia, whereas Rac induces<br />

lamellipodia, and Rho induces focal adhesions and associated stress fibers (Hall 1998).<br />

<strong>The</strong>refore each of these Rho-related proteins interact with multiple downstream effectors to<br />

control the actin cytoskeleton.<br />

1.3.6 Integrin crosstalk<br />

To add to the complexity of integrin signalling, the signalling cascades mediated by integrins<br />

are also linked to those of other receptors including growth factor receptors, receptor tyrosine<br />

kinases (RTKs), cytokine receptors, and ion channels (Schwartz 2001). <strong>The</strong> signalling<br />

pathways downstream of growth factor receptors and integrins can synergize. For instance,<br />

the combined effects of cell adhesion and growth factors can transform a weak transient<br />

activation of Erk to a strong and sustained Erk activity (Schwartz 2001). Integrin-mediated<br />

30


adhesion leads to the clustering and transactivation of several RTKs, including platelet-<br />

derived growth factor receptor, epidermal growth factor receptor, Ron, Met, and vascular<br />

endothelial growth factor receptor. This transactivation of RTKs can occur in the absence of<br />

their cognate growth factors (Schwartz 2001).<br />

Crosstalk can also occur between different integrin receptors. Integrins bound to a particular<br />

ligand can actively suppress the activation of other integrins on the same cell (Orr et al.<br />

2006). For instance, α2β1 bound to its ligand suppressed the α5β1 and αvβ3 integrins through<br />

PKA, and ligated α5β1 suppressed α2β1 through PKCα (Orr et al. 2006). In contrast,<br />

activation of one integrin can lead to the activation of another integrin. For instance, the<br />

binding of T-cells to VCAM-1 via integrin α4β1 activates β2 integrin function (Hyduk et al.<br />

2002), while binding of αLβ2 to ICAM-1 will activate α4β1 (Porter et al. 1997).<br />

1.4. Integrins and immunological synapses, unique signalling structures<br />

1.4.1 Synapses formed between communicating immune cells<br />

<strong>The</strong> immunological synapse (IS) is a supramacromolecular structure that forms at the<br />

membrane interface between a lymphoid effector cell and an APC or target cell through<br />

which information passes. <strong>The</strong> IS enables sustained signalling and T-cell activation, which<br />

can be stable for many hours (Sims et al. 2002). Formation of an IS is a multistep process,<br />

such that a nascent or immature IS initially forms and then develops into a mature IS,<br />

otherwise referred to as a supramolecular activation complex, (SMAC). α4β1 is initially<br />

engaged in the central region of the nascent SMAC (cSMAC; Figure 1.11A), and the T-cell<br />

receptors (TCRs) are in the peripheral SMAC (pSMAC). In the mature IS, the localisation of<br />

the α4β1 integrins and TCR are reversed, such that integrins form a ring around the centrally<br />

located signalling molecules. Talin remains associated with the active form of α4β1 within<br />

the pSMAC ring (Monks et al. 1998), while the TCR and its associating proteins such as<br />

CD2, CD28, PKC, lck, Fyn, CD4 and CD8 are localised in the cSMAC (Lin et al. 2005).<br />

Integrin α4β1-mediated formation of the pSMAC enhances the accumulation of TCR<br />

complexes and the exclusion of the phosphatase CD45 from the cSMAC (Graf et al. 2007),<br />

and thereby α4β1 helps to organize the proteins present in the IS. CD45 inactivates lck by<br />

dephosphorylation of Y394 which reduces TCR signalling within the microclusters, hence its<br />

exclusion from the cSMAC serves an important purpose.<br />

31


Figure 1.11 Models of the immunological synapse and kinapse<br />

(A) <strong>The</strong> model of an IS comprised of three concentric zones. <strong>The</strong> cSMAC (central SMAC) in the center contains<br />

concentrated TCRs where signaling is terminated. <strong>The</strong> pSMAC (peripheral SMAC) which surrounds the<br />

cSMAC contains LFA-1 and talin. <strong>The</strong> dSMAC (distal SMAC) is a second outer ring which contains the<br />

phosphatase CD45 to prevent dephosphorylation of key kinases. (B) A polarized migrating cell is divided into<br />

three regions, the lamellipodium at the leading edge, the focal zone in the mid section, and the uropod at the<br />

trailing edge. <strong>The</strong> migrating cell forms transient synapses or kinapses at the cell surface. Signalling<br />

microclusters form in the dSMAC and move through the pSMAC to the cSMAC, corresponding to the three<br />

regions of a polarized cell. LFA-1 which is expressed at different levels in each of the three zones varies in its<br />

affinity for ligand depending on the particular zone, as indicated. (Figure is adapted from information published<br />

by Evans et al. 2009, and Dustin 2008a).<br />

<strong>The</strong> roles of the different substructures of the SMAC are still somewhat controversial. It was<br />

initially proposed that TCR signalling occurred within the cSMAC, however at least one<br />

study suggests that the cSMAC is the site where TCRs are downregulated (Lee et al. 2003).<br />

Sustained TCR signalling is thought to take place in transient microclusters that form at the<br />

outside edges of pSMAC, also known as distal SMAC, or dSMAC (Campi et al. 2005; Varma<br />

et al. 2006). <strong>The</strong> microclusters only transmit signals while they are being transported through<br />

the pSMAC to the cSMAC (Campi et al. 2005; Varma et al. 2006). However, it is not known<br />

whether all IS form pSMAC and cSMAC, and it is possible that in T-cell signalling the<br />

formation of these SMACs is not a physiological requirement. Improved resolution of<br />

membrane structures by in vivo microscopy will be required to unravel the substructures of<br />

SMAC in vivo (Lin et al. 2005).<br />

32


1.4.2 Synapses termed kinapses are formed on migrating T cells<br />

T cells are highly motile when scanning APCs. Synapses are formed at the surface of<br />

migratory T cells, which have been termed kinapses (Dustin 2008a). Migrating T cells are<br />

polarized and travel in a particular direction towards a chemoattractant. <strong>The</strong>y are speedsters<br />

travelling at speeds of 5 to 40 μm/minute. In contrast, fibroblasts move at approximately 15<br />

μm/hour (reviewed by Evans et al. 2009). <strong>The</strong> distribution of αLβ2 (LFA-1) on the surface of<br />

migrating T cells is not uniform. LFA-1 is present at low levels at the leading edge and at<br />

high levels at the rear of the cell (Smith et al. 2005). A migrating T cell can be separated into<br />

three different zones with a leading lamellipodium which expresses low levels of intermediate<br />

affinity LFA-1, a mid-cell focal zone which expresses higher levels of high affinity LFA-1,<br />

and a trailing edge uropod with the highest expression of LFA-1 of unknown affinity (Figure<br />

1.11B; reviewed by Evans et al. 2009).<br />

<strong>The</strong> lamellipodial membrane rapidly protrudes and retracts to rapidly scan its target, and<br />

therefore expresses intermediate-affinity LFA-1 (reviewed by Evans et al. 2009). <strong>The</strong> focal<br />

zone provides firm attachment, which requires high affinity LFA-1 (Figure 1.11B; reviewed<br />

by Evans et al. 2009). <strong>The</strong> uropod at the rear of the cell anchors the cells to a substrate while<br />

the front leading edge is able to scan its surroundings (reviewed by Evans et al. 2009). Dustin<br />

proposed that kinapses contain a dSMAC at the leading edge of the cell, pSMAC at the mid<br />

region and a cSMAC at the uropod of the polarized cell Figure 1.11B (Dustin 2008a).<br />

1.5. Tyrosine kinases and cell signalling<br />

As mentioned above, tyrosine kinases play an important role in integrin activation and<br />

signalling. This thesis largely focusses on two tyrosine kinase families, namely the FAK and<br />

src families.<br />

1.5.1 <strong>The</strong> focal adhesion kinase (FAK) family<br />

FAK is a 125 kDa non-receptor protein tyrosine kinase, first identified as a phosphorylated<br />

protein whose expression became increased after v-src transformation of chicken embryo<br />

cells (Schaller et al. 1992). FAK is ubiquitously expressed in virtually all tissues and cell<br />

types, and plays a crucial role in mediating signal transduction pathways initiated either at the<br />

sites of cell attachment or through growth factor receptors. Hence, FAK plays an important<br />

role in cell attachment, migration, invasion, proliferation and survival. It is involved in<br />

33


important processes in cell, tissue, and organ structuring, functioning and remodelling<br />

(reviewed in van Nimwegen et al. 2007). In addition, FAK is essential during the early stages<br />

of embryonic development as FAK-null mice embryos die at day 8.5 (Ilic et al. 1995).<br />

Conversely, increased FAK signalling may result in uncontrolled proliferation, survival and<br />

migration of cells (reviewed in van Nimwegen et al. 2007). FAK exists as various isoforms<br />

due to alternative splicing and usage of alternative promoters (Schaller et al. 1993). <strong>The</strong> most<br />

well known isoform lacks the catalytic domain, and is called FAK related non-kinase (FRNK;<br />

Schaller et al. 1993). FRNK competes with FAK for localisation at focal adhesions. It acts as<br />

an inhibitor of FAK due to the lack of a kinase domain.<br />

<strong>The</strong> only other FAK family member is proline-rich tyrosine kinase 2 (PYK2), also called cell<br />

adhesion kinase β (CAKβ), related adhesion focal tyrosine kinase (RAFTK), or calciumdependent<br />

proteins-tyrosine kinase (CADTK; reviewed in van Nimwegen et al. 2007). PYK2<br />

is slightly smaller then FAK, at 112 kDa, and its highest expression levels are in cells of the<br />

nervous system (reviewed in van Nimwegen et al. 2007).<br />

FAK structure and signalling<br />

<strong>The</strong>re is more then 90% aa similarity between the human, chicken, mouse and frog<br />

homologues of FAK. FAK has a central kinase domain flanked by a large N-terminal region<br />

containing a FERM domain, and a C-terminal domain containing a focal adhesion targeting<br />

(FAT) sequence (Figure 1.12; reviewed in van Nimwegen et al. 2007).<br />

Figure 1.12 Structural features and binding partners of FAK<br />

Structural features of FAK include an N-terminal region containing a FERM domain, a central kinase domain,<br />

and a C-terminal focal adhesion targeting (FAT) site. <strong>The</strong>re are various tyrosine phosphorylation sites for<br />

autophosphorylation and activation. <strong>The</strong> known interacting sites for other cytoplasmic signalling proteins are<br />

indicated. (Figure adapted from van Nimwegen et al. 2007).<br />

34


Integrin clustering leads to autophosphorylation of FAK on tyrosine 397 (Y397), which<br />

increases the catalytic activity of FAK (Calalb et al. 1995). <strong>The</strong> autophosphorylation of Y397<br />

leads to the binding of src, p85, shc, or Grb7 via their src-homology 2 (SH2) domains. <strong>The</strong><br />

kinase domain contains two tyrosines at aa positions 576 and 577 (Y577 and Y577,<br />

respectively), which are phosphorylated in response to src binding and induce the full<br />

activation of FAK (Calalb et al. 1995). <strong>The</strong> kinase domain of FAK phosphorylates several<br />

focal adhesion associated proteins such as paxillin, Grb2, and p130CAS (Thomas et al. 1999).<br />

<strong>The</strong> FERM domain mediates protein-proteins interactions, where the most well known<br />

partner is the integrin β subunit cytoplasmic domain (Cooper et al. 2003). <strong>The</strong> FERM domain<br />

also functions as a regulator of FAK activity, as it can inhibit the autophosphorylation of<br />

Y397, which is necessary for FAK activation (Cooper et al. 2003). Upon integrin clustering,<br />

the FERM domain binds to the integrin β-subunit cytoplasmic tail, which releases the FERM-<br />

mediated autoinhibiton of FAK, and allows autophosphorylation of Y397, and activation of<br />

FAK (Cooper et al. 2003). Conversely, the FERM domain can also interact with full-length<br />

FAK, optimise its kinase activity, and promote FAK signalling (Dunty et al. 2004).<br />

<strong>The</strong> FAT sequence of FAK is required for the localisation of FAK to focal adhesions, as the<br />

fusion of the FAT sequence with other proteins is sufficient for their localisation to focal<br />

adhesions (Hildebrand et al. 1993). <strong>The</strong> FAT domain has also been shown to interact with<br />

integrin β subunit cytoplasmic domains. However, whether the interaction is direct or via<br />

other integrin-associated proteins like talin, and paxillin remains unclear (Schaller et al.<br />

1995b). Phosphorylation of FAK residues tyrosine 861 (Y861) and/or tyrosine 925 (Y925)<br />

early during cancer cell adhesion by association with the src family kinase (SFK) member,<br />

fyn, helps FAK translocate from lipid rafts in which SFKs usually reside, and leads to the<br />

activation of Akt-mediated signalling (Baillat et al. 2008).<br />

FAK has been implicated in a variety of biological processes, including cell survival,<br />

proliferation, migration, invasion, and angiogenesis, through various signalling cascades<br />

summarised in Table 1.4. In respect of cell survival, apoptotic proteins such as Bad, GSK3<br />

and p53 are inactivated by FAK either directly as with p53, or indirectly through signalling<br />

pathways such as the PI-3 kinase pathway as with Bad and GSK3 (Datta et al. 1997; Sonoda<br />

et al. 1999; Almeida et al. 2000; Golubovskaya et al. 2005). Overexpression of FAK induces<br />

the up-regulation of Cyclin D1 and Cyclin D3, causing accelerated cell cycle transition from<br />

the G1 phase to the S phase (Sonoda et al. 2000; Yamamoto et al. 2003).<br />

35


Table 1.4 Overview of the proteins/pathways that are involved in FAK-mediated signalling<br />

Biological<br />

Proteins of FAK-mediated signalling cascades<br />

process<br />

Survival FAK→PI-3 kinase → PKB → Bad → GSK3 (Sonoda et al. 1999)<br />

FAK→p53 (Golubovskaya et al. 2005)<br />

FAK→ Src→p130CAS → Ras → → JNK (Almeida et al. 2000)<br />

Proliferation FAK→Grb2 → MAPK → Cyclin D1 (Sonoda et al. 2000)<br />

FAK→PKC/PI-3 kinase → Rb → Cyclin D3 (Yamamoto et al. 2003)<br />

Migration FAK↔Calpains (Giannone et al. 2004)<br />

FAK→Src/p130CAS → MAPK (Cary et al. 1998)<br />

FAK→Src/p130CAS/PI-3 kinase → MAPK (Reiske et al. 1999)<br />

GSK3/PP1→ regulate FAK Ser 722 phosphorylation (Bianchi et al. 2005)<br />

(Adapted from van Nimwegen et al. 2007)<br />

FAK is both an activator and a substrate of calpains (Giannone et al. 2004). <strong>The</strong> interplay<br />

between FAK and calpains is important in the disassembly of focal adhesions at the rear of<br />

cells during cell migration (Giannone et al. 2004). As mentioned above, FAK also activates<br />

p130CAS through src, which leads to the activation of the mitogen-activated protein kinases<br />

(MAPK) pathway and cell migration (Cary et al. 1998; Reiske et al. 1999). Phosphorylation<br />

of FAK on serine 722 (S722) through the competing actions of the serine/threonine kinase<br />

GSK3, and the protein phosphatase PP1 serves to regulate cell migration (Bianchi et al.<br />

2005).<br />

1.5.2 Src kinase family (SFK)<br />

<strong>The</strong> non-receptor tyrosine kinase src is important in many aspects of cellular physiology<br />

(Sandilands et al. 2008) . <strong>The</strong> src kinase family (SFK) consists of nine members, namely c-<br />

src, c-yes, c-fyn, fgr, lyn, hck, lck, blk and yrc (Sandilands et al. 2008). <strong>The</strong>y share a similar<br />

structure, with molecular masses ranging from 52-62 kDa. <strong>The</strong>y are each comprised of 6<br />

functional domains: a myristylation domain, which enables plasma membrane interactions; a<br />

unique domain; SH2 and SH3 domains for binding to other proteins; a kinase domain for<br />

phosphorylation of tyrosines which also contains an autophosphorylation site at tyrosine 416<br />

(Y416); and a C-terminal regulatory region which contains a phosphorylation site at tyrosine<br />

527 (Y527) which is crucial for regulating the tyrosine kinase activity of src (Sandilands et al.<br />

2008). V-src is a constitutively active form of src, which lacks several amino acids at the Cterminus,<br />

notably Y527 (Y527; reviewed by Martin 2001).<br />

36


Src is negatively regulated by phosphorylation of Y527, which is catalysed by the tyrosine<br />

kinase Csk (c-Src terminal kinase; Cooper et al. 1986; Okada et al. 1989). Phosphorylation of<br />

Y527 confines src in an inactive state (Roussel et al. 1991). <strong>The</strong> SH2 domain of src interacts<br />

with Y527 (Roussel et al. 1991) allowing the SH3 domain of src to become engaged with the<br />

SH2 kinase linker region such that Y416 in the kinase region becomes unphosphorylated<br />

(Figure 1.13). <strong>The</strong> crystal structure of src showed that the SH2 and SH3 domains were<br />

positioned at the back of the kinase domain, and interactions between these domains locked<br />

the molecule in an inactive state (Xu et al. 1997).<br />

Figure 1.13 Structural features and activation of src<br />

A cartoon picture depicting the features of src and its active and inactive conformations. In the inactive<br />

conformation on the left-hand side, the SH2 domain is bound to the phosphorylated tyrosine residue 527 (Y527)<br />

and bends back towards the kinase domain. <strong>The</strong> SH3 domain is likewise bound to the kinase domain. Upon<br />

activation, tyrosine residue 416 (Y416) is phosphorylated, and tyrosine residue 527 is dephosphorylated as<br />

shown in the middle drawing. Activated src is extended with a phosphorylated tyrosine residue at 416 as shown<br />

in the right-hand cartoon. <strong>The</strong> essential phosphorylation sites for activation are indicated. (Adapted from Martin<br />

2001).<br />

Most wild-type c-src is localised at the perinuclear regions of cells, where it directly<br />

associates with the microtubule-organising centre and co-localises with endosomal markers<br />

(Kaplan et al. 1992). However, active c-src is found in focal adhesions near the plasma<br />

membrane (Kaplan et al. 1994). This change in localisation of src is due to endosomal<br />

trafficking of src, which is regulated by Rho GTPases in concert with the actin cytoskeleton<br />

(Sandilands et al. 2008).<br />

37


Upon activation, src is autophosphorylated at Y416 (Y419 in humans), and dephosphorylated<br />

at Y527 by protein tyrosine phosphatases (PTPs; Pallen 2003). <strong>The</strong> most likely candidate is<br />

the transmembrane phosphatase PTPα (Pallen 2003), as ptpα-/- cells were found to have<br />

decreased src and fyn kinase activity, which correlated with enhanced tyrosine<br />

phosphorylation of the negative regulatory domain of src (Ponniah et al. 1999). Furthermore,<br />

overexpression of PTPα leads to increased tyrosine phosphorylation of focal adhesion<br />

associated proteins in cells, and was shown to activate src in vivo (den Hertog et al. 1993;<br />

Harder et al. 1998; Zheng et al. 2000). Other tyrosine phosphatases that might regulate src in<br />

response to cell adhesion include PTP1B and Shp2. Overexpression of a catalytically<br />

defective mutant of PTP1B in mouse L cells reduced src activity, and PTP1B purified from<br />

breast cancer cells was able to dephosphorylate a src peptide phosphorylated on Y527<br />

(Arregui et al. 1998; Bjorge et al. 2000). shp2-/- cells deficient in Shp2 were found to have<br />

increased numbers of immature focal adhesions, and displayed defective phosphorylation of<br />

focal adhesion-associated proteins upon cell adhesion (Oh et al. 1999).<br />

Src and cancer: Many human tumours exhibit elevated src kinase activity (Irby et al. 2000),<br />

including human mammary carcinomas, colon cancer and pancreatic cancer (Playford et al.<br />

2004). <strong>The</strong> increased activity of src in cancers has been proposed to be due to tyrosine<br />

phosphatase-mediated dephosphorylation of the C-terminal Y527, an increase in src protein<br />

levels, altered src protein stability, an increase in upstream RTKs, or a loss of key proteins<br />

that negatively regulate src.<br />

1.5.3 Integrin signalling, FAK and src<br />

As mentioned briefly above, integrin cytoplasmic domains have no intrinsic kinase activity,<br />

therefore the signals initiated by integrin-ECM interactions are transduced into cells through<br />

association of the cytoplasmic domains with intracellular signalling proteins, in particular<br />

non-receptor tyrosine kinases such as FAK and SFKs (Mitra et al. 2006).<br />

Integrin clustering recruits FAK via its C-terminal FAT domain to integrin-associated<br />

proteins such as talin and paxillin, or through its FERM domain to the integrin β cytoplasmic<br />

tail (Mitra et al. 2006). <strong>The</strong> latter interactions are thought to be required to stably localise and<br />

autophosphorylate FAK at the focal adhesions (Mitra et al. 2006). As mentioned above, the<br />

binding of the FERM domain to the integrin β cytoplasmic tail releases the autoinhibiton of<br />

FAK, and triggers the autophosphorylation of FAK on Y397. <strong>The</strong> autophosphorylation of<br />

38


Y397 creates a high affinity binding site for the SH2 domain of SFKs, and the binding of src<br />

leads to conformational activation of src. Src then transphosphorylates FAK on residues Y576<br />

and Y577 within the kinase domain activation loop, and on residues Y861 and Y925 within<br />

the FAK C-terminal domain. <strong>The</strong> FAK-src complex acts to control cell shape and focal<br />

contact turnover events during cell motility (Mitra et al. 2005). <strong>The</strong> phosphorylation of Y925<br />

creates a Grb2 binding site, thereby linking FAK to the MAP kinase pathway (Schlaepfer et<br />

al. 1994). <strong>The</strong> FAK-src complex phosphorylates the adaptor proteins paxillin at residues<br />

tyrosine 31 (Y31) and tyrosine 118 (Y118), and p130Cas within the Cas substrate domain.<br />

This creates binding sites for the SH2 domain of the adaptor protein Crk. <strong>The</strong> overexpression<br />

of Crk can lead to p130Cas-dependent activation of FAK (Iwahara et al. 2004). <strong>The</strong><br />

overexpression of p130Cas can promote enhanced tyrosine phosphorylation of FAK and<br />

paxillin (Brabek et al. 2004). <strong>The</strong> association of p130Cas with src may preferentially trap src<br />

in the active state (Nasertorabi et al. 2006).<br />

Src can also be activated by integrins via other pathways, as briefly mentioned above. Src<br />

binds directly to the β3 integrin cytoplasmic tail through its SH3 domain, and it was proposed<br />

that this primes src for maximal activation during integrin clustering (Arias-Salgado et al.<br />

2003; Shattil 2005). α4β1 can activate src independently of FAK, and rescue the motility<br />

defects of FAK-null fibroblasts through increased phosphorylation of p130Cas (Hsia et al.<br />

2005).<br />

p130Cas-Crk signalling is negatively regulated by the Abl RTK (Holcomb et al. 2006). Abl is<br />

activated by integrin clustering, and localises to focal adhesions (Holcomb et al. 2006). It<br />

phosphorylates Crk at Y221 and prevents Crk from binding to p130Cas (Holcomb et al.<br />

2006). Other proteins found to regulate p130Cas function include p130Cas associated protein<br />

(p140Cap; Di Stefano et al. 2004), and the LIM protein Ajuba (Pratt et al. 2005).<br />

1.6. Heat shock proteins<br />

Heat shock proteins (Hsps) are a superfamily of highly conserved molecular chaperones<br />

which are induced in plants, yeast, bacterial and mammalian cells in response to sub-lethal<br />

physiological and environmental stresses (reviewed in Daugaardet al., 2007). Heat shock<br />

proteins were first identified as inducible proteins whose expression was induced by elevated<br />

temperatures, however it became clear that Hsp synthesis was also initiated by many<br />

chemical and physical stress stimuli (Lindquist et al. 1988). Physiological events such as<br />

39


antigen presentation, cell growth, differentiation, development and aging can also induce Hsp<br />

synthesis (Calderwood et al. 2006).<br />

Hsps protect cells from stress by assisting in protein folding and stabilisation, and they help to<br />

sequester damaged proteins for degradation (reviewed in Daugaardet al., 2007). <strong>The</strong>y are<br />

overexpressed in a wide range of cancers, and have been implicated in cancer growth by<br />

promoting tumour cell proliferation and inhibiting cellular death pathways.<br />

Extracellular and membrane-bound Hsps have important immunological roles (reviewed in<br />

Daugaardet al., 2007). In particular they stimulate the immune system, in part by presenting<br />

antigenic peptides. Hsp70 and Hsp90 families are frequently located on the plasma membrane<br />

of tumour cells (reviewed in Daugaardet al., 2007). In addition, elevated levels of Hsp70 and<br />

Hsp90 have been detected in the medium of tumour cell lines and the serum of cancer<br />

patients.<br />

Mammalian hsps have been divided into 5 different subfamilies according to their molecular<br />

weights, namely Hsp20, Hsp60, Hsp70, Hsp90 and Hsp100 (reviewed in Daugaardet al.,<br />

2007). This thesis will primarily concentrate on the Hsp70 family of heat shock proteins and<br />

their role in integrin regulation/function.<br />

1.6.1 Hsp70 family<br />

<strong>The</strong> human Hsp70 family consists of at least 8 different members, encoded by different<br />

genes, which differ in amino acid sequence, expression and stress induction (Table 1.5;<br />

reviewed in Daugaardet al., 2007). Hsp70 proteins mainly reside in the cytosol and nucleus,<br />

however some are confined to the lumen of the endoplasmic reticulum (ER; Hsp70-5), and<br />

the mitochondrial matrix (Hsp70-9; Daugaard et al. 2007). <strong>The</strong> different compartmental<br />

localisation of the Hsp70 members suggests they may display specificity for client proteins,<br />

or serve other chaperone-independent functions. Three members of the Hsp70 family, namely<br />

Hsp70-1a, Hsp70-1b and Hsp70-6 are stress-inducible, whereas the other members Hsc70,<br />

Hsp70-1t, Hsp70-2, Hsp70-5 and Hsp70-9 are not stress-inducible (reviewed in Daugaardet<br />

al., 2007). Hsp70-1a and Hsp70-1b may function primarily in cellular stress situations, while<br />

the others may perform tissue-specific and housekeeping functions. Hsp70 family members<br />

are involved in a large range of processes, including the folding of newly synthesized<br />

proteins, the transport of proteins across membranes, the refolding of misfolded and<br />

40


aggregated proteins, and the control of activity of regulatory proteins (reviewed in<br />

Daugaardet al., 2007).<br />

Table 1.5 Essential features of members of the human Hsp70 family<br />

Protein<br />

Hsp70-1a<br />

Hsp70-1b<br />

Alternative<br />

names<br />

Hsp70, Hsp72,<br />

Hsp70-1<br />

Hsp70, Hsp72,<br />

Hsp70-1<br />

Homology to<br />

Hsp70-1a (%)<br />

Locus Localization<br />

100 HSPA1A 6p21.3<br />

99 HSPA1B 6p21.3<br />

Hsp70-1t Hsp70-hom 91 HSPA1L 6p21.3<br />

41<br />

Cellular<br />

localization Stress-induced<br />

Cytosol,<br />

Nucleus, Yes<br />

Lysosomes<br />

Cytosol,<br />

Nucleus, Yes<br />

Lysosomes<br />

Cytosol,<br />

No<br />

Nucleus<br />

Cytosol,<br />

No<br />

Nucleus<br />

Hsp70-2<br />

Hsp70-3,<br />

HspA2<br />

84 HSPA2 14q24.1<br />

Hsp70-5 Bib, Grp78 64 HSPA5 9q33-q34.1 ER No<br />

Hsp70-6 Hsp70B′ 85 HSPA6 1cen-qter<br />

Cytosol,<br />

Nucleus<br />

Yes<br />

Hsc70 a Hsp70-8,<br />

Hsp73<br />

Grp75,<br />

86 HSPA8 11q23.3-q25<br />

Cytosol,<br />

Nucleus<br />

No<br />

Hsp70-9 mtHsp75,<br />

Mortalin<br />

52 HSPA9 5q31.1 Mitochondria No<br />

a<br />

A 54 kDa Hsc70 splice variant (NM_153201) has been reported, but its functional significance remains<br />

unclear. (Adapted from Daugaard et al. 2007).<br />

<strong>The</strong> Hsp70 family, like other Hsps, show a high level of conservation in their amino acid<br />

sequences and domain structures (reviewed in Daugaardet al., 2007). <strong>The</strong>y all have a<br />

conserved ATPase domain, a central protease sensitive site, a peptide binding domain, and a<br />

G/P-rich C-terminal region which contains an EEVD-motif that binds co-chaperones and<br />

other Hsps (Figure 1.14). <strong>The</strong> various conserved domains enable Hsp70 proteins to bind and<br />

release, in an ATP-dependent manner, extended hydrophobic amino acids exposed by<br />

incorrectly folded globular proteins (Bukau et al. 2006).


Figure 1.14 Domain structure of human hsp70 family members<br />

A cartoon picture of the human hsp70 family members showing their different structural and functional<br />

domains. (Adapted from Daugaard et al. 2007).<br />

<strong>The</strong> importance of the Hsp70 family is illustrated by the phenotypes of Hsp70 knockout mice<br />

(Table 1.6). Knockout of the genes of different family members produces a variety of<br />

phenotypes ranging from embryonic lethality, to viable and fertile mice that are susceptible to<br />

various stressors such as UV and heat, and predisposed to develop disease, including<br />

pancreatitis and infections (reviewed in Daugaard et al. 2007). <strong>The</strong> range of phenotypes<br />

illustrates the various functions of different Hsp70 family members.<br />

Hsp70 is expressed by tumour cells, and is associated with chemotherapeutic resistance in the<br />

case of leukemias (Sliutz et al. 1996) and breast cancers (Vargas-Roig et al. 1998). Inhibition<br />

of Hsp70 by the heat shock protein inhibitor KNK-437 was shown to reduce the adhesion of<br />

myeloma cells to fibronectin (FN) or stromal cells, causing them to undergo apoptosis, and<br />

sensitizing them to chemotherapeutic agents (Nimmanapalli et al. 2008).<br />

This thesis will concentrate on the Hsp70-1(a-b), Hsc70, and Hsp-9 proteins, hence a fuller<br />

description of each of these Hsps is provided below.<br />

42


Table 1.6 Phenotypes of Hsp70 knockout mice<br />

Protein<br />

Gene<br />

locus<br />

Chromosomal<br />

localization<br />

Analogous<br />

human gene<br />

43<br />

Phenotype<br />

Hsp70.1 Hspa1a 17 B1 HSPA1A<br />

Viable and fertile. Susceptible to UV,<br />

osmotic stress, ischemia, TNF, pancreatitis<br />

and heat.<br />

Hsp70.3 Hspa1b 17 B1 HSPA1B Viable and fertile. Susceptible to heat.<br />

Hsp70.1 +<br />

Hsp70.3<br />

Hspa1a +<br />

Hspa1b<br />

17 B1<br />

HSPA1A +<br />

HSPA1B<br />

Viable and fertile. Susceptible to radiation<br />

and sepsis. Increased genomic instability.<br />

Hsp70.2 Hspa2 17 B1 HSPA2<br />

Viable and females fertile. Meiotic defects in<br />

male germ cells.<br />

Hsc70 Hspa8 9 A5.1 HSPA8 Not applicable. Knockout cells are non-viable<br />

Bip, GRP78 Hspa5 2 B HSPA5 Lethal at embryonic day 3.5<br />

(Adapted from Daugaard et al. 2007)<br />

Hsp70-1a and Hsp70-1b<br />

Hsp70-1a and Hsp70-1b (Hsp70-1) are encoded by the genes HSPA1A and HSPA1B,<br />

respectively, which are closely related, sharing 99% aa identity (reviewed in Daugaard et al.<br />

2007). <strong>The</strong>y are the major stress-inducible members of the Hsp70 family. Basal levels of<br />

mRNAs encoding Hsp70-1a and Hsp70-1b are detectable at a similar level in many tissues<br />

and cell types, however HSPA1A expression is dramatically higher in blood (Su et al. 2004).<br />

In normal conditions, Hsp70-1 proteins are expressed in a cell type and cell cycle dependent<br />

manner. However, during stress conditions, the hsp70-1 genes are activated by the binding of<br />

a stress inducible transcription factor, heat shock factor 1 (HSF1), to upstream regulatory<br />

regions called heat shock elements (HSE; reviewed in Daugaard et al. 2007). Stress-induced<br />

Hsp70-1 enables the cell to cope with denatured protein aggregates, and hence confers<br />

protection against protein damage caused by heat, ischemia, and oxidative stress (Li et al.<br />

1991; Bellmann et al. 1996; Chong et al. 1998).<br />

Knockout of the murine Hsp70-1 homologues Hsp70.1 and Hsp70.3 produces mice that are<br />

viable and fertile, but which however are more sensitive to pancreatitis, UV light (epidermis),<br />

osmotic stress (renal medulla), and brain ischemia (Table 1.6; reviewed in Daugaardet al.,<br />

2007).<br />

Hsp70-1 is abundantly expressed in malignant tumours of various origins, where its<br />

expression correlates with increased cell proliferation, poor differentiation, and lymph node<br />

metastases (Mosser et al. 2004).


Hsc70 (Hsp70-8)<br />

Hsc70 is encoded by the gene HSPA8A (Table 1.5) and is expressed in most tissues (Su et al.<br />

2004). It is 86% identical to Hsp70-1a, and is an important housekeeping chaperone which is<br />

involved in the folding of nascent polypeptides, protein translocation across membranes,<br />

prevention of protein aggregation under stress conditions, and disassembly of clathrin-coated<br />

vesicles (Bukau et al. 2006). Hsc70 is essential for cell survival as gene knockout of Hsc70 is<br />

lethal (Table 1.6), and RNA interference (RNAi)-based knock-down of Hsc70 results in<br />

massive cell death (Rohde et al. 2005). Hsc70 participates in cytokine-mediated, posttranscriptional,<br />

co-chaperone-dependent, regulation of the Bcl-2 family member Bim in<br />

human blood cells (Matsui et al. 2007).<br />

Hsp70-9<br />

Hsp70-9, which is encoded by the gene HSPA9 gene (Table 1.5), is 52% identical to Hsp70-<br />

1a but is not inducible by stress. Hsp70-9 is localised to the mitochondrial lumen where it<br />

assists in protein folding after transmembrane transport.<br />

1.7. Integrins and disease<br />

Integrins are implicated in many human diseases, including inflammation, thrombosis,<br />

tumour metastasis and infections, and hence are being considered as major targets for<br />

therapeutic intervention (Horwitz 1997). Knockout of integrin genes in mice has shown<br />

integrins are essential for normal embryonic development (Bouvard et al. 2001). A number of<br />

diseases are associated with corresponding mutations of human integrin genes (Table<br />

1.7Error! Reference source not found.). Gene knockout studies have identified critical roles<br />

for integrins in development (β1 integrins), vasculogenesis (αV integrins), lymphangiogenesis<br />

(α9β1), thrombus formation (αIIbβ3), the integrity of the skin (α6β4), and immune responses<br />

(β2 and β7 integrins; Takada et al. 2007).<br />

44


Table 1.7 Ablation of integrin genes in mice and their resulting phenotypes<br />

Gene Phenotype<br />

α1 V, F Increased collagen synthesis, reduced tumour vascularisation<br />

α2 V, F Reduced branching morphogenesis and platelet adhesion<br />

α3 L, birth Defects in kidneys, lungs, and cerebral cortex; skin blistering<br />

α4 L, E11<strong>–</strong>E14 Defects in chorioallantois fusion and cardiac development<br />

α5 L, E10 Defects in extraembryonic and embryonic vascular development<br />

α6 L, birth Defects in cerebral cortex and retina; skin blistering<br />

α7 V, F Muscular dystrophy<br />

α8 L+V/F Small or absent kidneys; inner ear defects<br />

α9 L, perinatal Bilateral chylothorax<br />

α10 V, F No apparent phenotype<br />

α11 ... ...<br />

αv L, E12<strong>–</strong>birth Defects in placenta and in CNS and GI blood vessels; cleft palate<br />

αD V, F Reduced T-cell response and T-cell phenotypic changes<br />

αL V, F Impaired leukocyte recruitment and tumour rejection<br />

αM V, F Impaired phagocytosis and PMN apoptosis; obesity; mast cell<br />

development<br />

αX ... ...<br />

αE V, F Inflammatory skin lesions<br />

αIIb V, F Defective platelet aggregation<br />

β1 L, E5.5 Inner cell mass deterioration<br />

β2 V, F Impaired leukocyte recruitment; skin infections<br />

β3 V, F Defective platelet aggregation; osteosclerosis<br />

β4 L, perinatal Skin blistering<br />

β5 V, F No apparent phenotype<br />

β6 V, F Skin and lung inflammation and impaired lung fibrosis<br />

β7 V, F Abnormal Peyer’s patches; decreased number of intraepithelial<br />

lymphocytes<br />

β8 L, E12<strong>–</strong>birth Defects in placenta and in CNS and GI blood vessels; cleft palate<br />

V, indicates viable; F, fertile; L, lethal; L+V/F, mutations that disrupt development but<br />

also allow survival in a fraction of mice; CNS, central nervous system; GI,<br />

gastrointestinal; and PMN, polymorphonuclear neutrophil. (Adapted from Bouvard et al.<br />

2001; Takada et al. 2007)<br />

1.7.1 Hereditary disease<br />

Gene mutations or polymorphisms of integrin subunits have been associated with various<br />

hereditary diseases. <strong>The</strong> lack of expression of the α6 integrin subunit in humans results in<br />

junctional epidermolysis bullosa with pyloric or duodenal atresia (Pulkkinen et al. 1997).<br />

Similarly β4 mutations cause generalized skin blistering and aplasia cutis (Vidal et al. 1995).<br />

Mutations in the α7 gene gives rise to congenital myopathy and forms of congenital muscular<br />

dystrophy (Hayashi et al. 1998), while the down regulation of expression of α7β1 contributes<br />

to the pathology of congenital laminin deficiencies (Vachon et al. 1997).<br />

Defective platelet aggregation arising from mutations in the αIIb and β3 genes is the cause of<br />

the inherited bleeding disorder Glanzmann thrombasthenia (Derrick et al. 2001). Low platelet<br />

α2β1 densities due to the polymorphic C807T allele of the α2 subunit of the integrin α2β1<br />

45


leads to an increased tendency to bleed. In contrast, the polymorphic C807T/G873A allele<br />

causes high receptor expression which may contribute to an increased risk of stroke in<br />

younger people (reviewed in Carlsson et al. 1999; Krissansen et al. 2006a).<br />

Leukocyte adhesion deficiency syndrome (LAD), is a hereditary immunodeficiency disease<br />

with high morbidity and mortality due to mutations in the β2 gene (Kishimoto et al. 1987).<br />

1.7.2 Inflammatory disease<br />

Integrins have been implicated in a variety of inflammatory diseases, as listed below. Several<br />

anti-integrin therapeutic reagents are currently being developed to treat these diseases.<br />

Multiple sclerosis: Blockade of the α4, β2, and β7 integrin pathways in mice leads to<br />

successful remission of different forms of experimental autoimmune encephalomyelitis<br />

(EAE), an animal model for multiple sclerosis (Kanwar et al. 2000; Bullard et al. 2005).<br />

Asthma: Blockade of integrin α4 expression prevents the transmigration of eosinophils into<br />

the lungs, and the development of bronchial hyper-responsiveness to agonists in animal<br />

models (reviewed in Krissansen et al. 2006a). Blocking αLβ2-ICAM-1 pathways also has a<br />

protective effect (Taylor et al. 1997; Krissansen et al. 2006a).<br />

Arthritis: Several integrins and their ligands are expressed in the arthritic joint tissue,<br />

including αLβ2, αMβ2, ICAM-1, α4β1 and α4β7 on cells in the synovial infiltrate, αLβ2,<br />

αMβ2 and ICAM-1 on the synovial lining layer, and VCAM-1 on endothelial cells (reviewed<br />

by Krissansen et al. 2006a).<br />

1.7.3 Other diseases<br />

Integrins have also been implicated in a variety of other diseases, including cancer, hepatitis,<br />

skin diseases, Graves disease, glomerulonephritis, muscular dystrophy, sepsis, emphysema,<br />

osteoporosis, wound healing, cardiovascular disease, atherosclerosis, and osteoporosis<br />

(reviewed by Krissansen et al. 2006a).<br />

46


1.7.4 β7 integrins and disease<br />

<strong>The</strong> β7 integrins are involved in the following diseases:<br />

Graft versus host disease (GVHD): β7 -/- donor T cells caused less GVHD morbidity and<br />

mortality than wild-type (WT) donor T cells in a murine hematopoietic stem cell<br />

transplantation model because of selectively decreased T-cell infiltration of the liver and<br />

intestines (Waldman et al. 2006).<br />

Inflammation and cancer: Polycyclic aromatic hydrocarbons (PAHs) are a major class of<br />

chemical pollutants to which humans are widely exposed, that are generated by the burning of<br />

fossil fuels or wood (Burchiel 1999). <strong>The</strong>y are notably found in diesel exhaust particles,<br />

cigarette smoke, charcoal barbecued foods, and industrial waste byproducts (Burchiel 1999).<br />

PAHs were found to upregulate the expression of β7 integrins in human macrophages through<br />

both the aryl hydrocarbon receptor (AhR) and the basic leucine zipper transcription factor cmaf<br />

(Monteiro et al. 2007). This may contribute to the inflammatory effects PAHs have by<br />

altering macrophage homing to epithelial tissues (Haskill et al. 1992). <strong>The</strong> carcinogenic<br />

effects PAHs may be mediated in part by overexpression of c-maf/β7, which increases the<br />

development of myelomas (Hurt et al. 2004) and T-lymphomas (Morito et al. 2006).<br />

Human immunodeficiency virus (HIV): Recently, β7 integrins have been identified as playing<br />

a new role in HIV infection. It was suggested that HIV-1 is able to mediate depletion of gut<br />

CD4+ T-cells in part by binding to the integrin α4β7, which helps direct T-cells to the gut<br />

(Arthos et al. 2008). HIV-1 envelope glycoprotein 120 (gp120) was found to bind to α4β7<br />

independently of the principle HIV receptor CD4, and antibodies against α4β7 could block<br />

the binding of gp120 to α4β7 (Arthos et al. 2008). <strong>The</strong> binding of HIV-1 to α4β7 induced the<br />

activation of LFA-1, which has a key role in immune cell interactions.<br />

Diabetes: Mucosa-associated lymphocytes expressing high levels of β7 integrins associate<br />

and accumulate in pancreatic islets during insulitis in the nonobese diabetic (NOD) mouse<br />

model of type 1 diabetes (Hanninen et al. 1996). Antibody blockade of α4 and β7 integrins<br />

and their ligand MAdCAM-1 successfully interrupted the development of diabetes<br />

development in NOD mice (Kommajosyula et al. 2001).<br />

47


Crohn’s disease and colitis: In the Tnf(DeltaARE) mouse model of Crohn's-like<br />

inflammatory bowel disease, development of intestinal inflammation was found to be<br />

critically dependent on β7 integrin-mediated T-lymphocyte recruitment (Apostolaki et al.<br />

2008). Furthermore, a mAb against αEβ7 prevents immunization-induced colitis in IL-2 -/-<br />

mice (Ludviksson et al. 1999).<br />

<strong>The</strong>rapeutic agents based on α4β7<br />

A specific barbituric acid-based inhibitor that blocks α4β7-MAdCAM-1 interactions, but<br />

spares α4β1-VCAM-1 interactions was recently reported (Harriman et al. 2008). Small<br />

interference RNA (siRNA) against cyclin D1 specifically directed to leukocytes expressing<br />

β7 integrin shows promise for the treatment of inflammatory bowel disease (Peer et al. 2008).<br />

Monoclonal antibodies (mAb) are being developed to treat Crohn’s disease and ulcerative<br />

colitis, and nonpeptide chemicals are being developed to treat inflammatory bowel disease<br />

and multiple sclerosis (Hehlgans et al. 2007). In addition, there are several drugs currently<br />

being developed and in clinical trials that specifically target α4β7, such as Vedolizumab, a<br />

humanised monoclonal antibody for treatment of inflammatory bowel disease (Soler et al.<br />

2009).<br />

Clinical examples of therapeutic agents targeting integrins<br />

Examples of drugs in the clinic are; Tysabri (Natalizumab), a humanised monoclonal<br />

antibody against α4 subunit is used for therapy for multiple sclerosis (Sheremata et al. 1999)<br />

and Crohn’s disease (Ghosh et al. 2003), and Efelizumab, a humanised anti-αL (CD11a)<br />

monoclonal antibody is used for treating psoriasis and tissue rejection (Dedrick et al. 2002;<br />

Schon 2008).<br />

48


1.8. Aims of this thesis<br />

Integrins signal through their relatively short cytoplasmic domains by binding to key<br />

signalling molecules and cytoskeletal elements that change the affinity of an integrin for its<br />

ligand, or induce clustering and changes in avidity. <strong>The</strong> β7 integrins are leukocyte-specific<br />

integrins that play central roles in gut mucosal immunity, but also contribute to the<br />

progression of an array of chronic inflammatory diseases. <strong>The</strong> host laboratory discovered a<br />

cell adhesion regulatory domain (CARD) within the cytoplasmic domain of the β7 subunit<br />

that regulates β7 integrin-mediated cell adhesion. Further investigation of this domain, and<br />

identification of intracellular ligands for the β7 cytoplasmic domain, may provide insights<br />

into signalling pathways controlling β7 integrin-mediated cell adhesion. Armed with this<br />

information, it might be possible to devise therapeutic interventions to prevent β7 integrin-<br />

mediated inflammatory disease.<br />

<strong>The</strong> broad aim of this thesis was to identify β7 integrin signalling pathways and the molecules<br />

involved. One of the specific objectives was to investigate structure-function relationships of<br />

the β7 CARD by identifying residues that are essential for function. Another objective was to<br />

identify intracellular signalling molecules that interact with the CARD, and the β7<br />

cytoplasmic domain, and determine whether they influence α4β7-mediated adhesion of T<br />

cells.<br />

49


Chapter 2. Materials and <strong>Methods</strong><br />

<strong>Section</strong> 1 <strong>–</strong> Materials<br />

2.1. Chemicals, reagents, buffers and media<br />

2.1.1 Suppliers<br />

Abcam Inc., (Cambridge, MA, USA)<br />

Alpha Diagnostic (San Antonio, TX, USA)<br />

American Type Culture Collection, ATCC (Manassas, VA, USA)<br />

Amersham Bioscience <strong>–</strong> now GE Healthcare (Buckingham, United Kingdom)<br />

AppliChem (Darmstadt, Germany)<br />

Bangs Laboratories Inc. (Fishers, IN, USA)<br />

BD Biosciences (Franklin Lakes, NJ, USA)<br />

BDH chemicals (Poole, United Kingdom)<br />

Bio-Rad Laboratories (Hercules, CA, USA)<br />

Biosource <strong>–</strong> now supplied by Invitrogen<br />

Biovision (Mountain View, CA, USA)<br />

Calbiochem <strong>–</strong> now supplied by Merck<br />

Cell Signaling Technology (Danvers, MA, USA)<br />

Chemicon International <strong>–</strong> now supplied by Millipore Corp.<br />

Cytoskeleton Inc. (Denver, CO, USA)<br />

DIFCO <strong>–</strong> now supplied by BD Biosciences<br />

Eppendorf AG (Hamburg, Germany)<br />

GE Healthcare (Chalfont St. Giles, United Kingdom)<br />

GibcoBRL/Life Technologies <strong>–</strong> now supplied by Invitrogen<br />

Immunochemistry Technologies, LLC. (Bloomington, MN, USA)<br />

Invitrogen (Carlsbad, CA, USA)<br />

Merck KGaA. (Darmstadt, Germany)<br />

50


Millipore Corp. (Bedford, MA, USA)<br />

Mimotopes Pty Ltd. (Melbourne, Australia)<br />

Molecular Probes <strong>–</strong> now supplied by Invitrogen<br />

Nunc <strong>–</strong> now part of <strong>The</strong>rmo Fisher Scientific<br />

Peptide 2.0 Inc. (Chantilly, VA, USA)<br />

Pierce (Rockford, IL, USA)<br />

Pharmacia Biotech (Piscataway, NJ, USA)<br />

Progen Biotechnik GmbH (Heidelberg, Germany)<br />

ProSpec-Tany TechnoGene Ltd (Rehovot, Israel)<br />

Promega Corp. (Madison, WI, USA)<br />

Qiagen Inc. (Hilden, Germany)<br />

R&D Systems (Minneapolis, MN, USA)<br />

Roche Applied Science (Roche; Mannheim, Germany)<br />

Santa Cruz Biotechnology Inc., (Santa Cruz, CA, USA)<br />

Scharlau Chemie S.A. (Barcelona, Spain)<br />

SERVA Electrophoresis (Heidelberg, Germany)<br />

Sigma-Aldrich (St. Louis, MO, USA)<br />

<strong>The</strong>rmo Fisher Scientific Inc. (Waltham, MA, USA)<br />

Upstate Cell Signaling Solutions <strong>–</strong> now supplied by Millipore<br />

USB Corporation. (Cleveland, OH, USA)<br />

United States Biological (Swampscott, MA, USA)<br />

Vector laboratories (Peterborough, United Kingdom)<br />

51


2.1.2 Chemicals/Molecular biology reagents<br />

All chemicals were purchased from; AppliChem, BDH chemicals, Bio-Rad Laboratories,<br />

Invitrogen, Scharlau Chemie S.A., or Sigma-Aldrich, unless otherwise stated.<br />

Molecular biology reagents:<br />

ABTS <strong>–</strong> Vector Labs, cat#SK4100<br />

Agarose LE <strong>–</strong> Roche Diagnostic, cat #1685678<br />

Alkaline phosphatase <strong>–</strong> Roche Diagnostic, cat #713023<br />

Ampicillin <strong>–</strong> Roche Diagnostic, cat #835269<br />

Aprotinin <strong>–</strong> Sigma-Aldrich, cat #A6103<br />

Aquacide II <strong>–</strong> Calbiochem, cat #17851<br />

Bovine serum albumin (BSA) <strong>–</strong> Gibco, cat #30036-578<br />

Bromophenol blue <strong>–</strong> BDH chemicals, cat #200152E<br />

CMFDA <strong>–</strong> Invitrogen, cat#C7025<br />

Complete Mini EDTA-free caspase inhibitors <strong>–</strong> Roche Diagnostic, cat #11836171001<br />

Coomassie Brilliant Blue R250 <strong>–</strong> DIFCO, cat#5128-13<br />

Cyanogen bromide-activated Sepharose 4B <strong>–</strong> Amersham Bioscience, cat#17 0430 01<br />

Deoxynucleoside triphosphate set <strong>–</strong> Roche, cat #1969064<br />

Dialysis tubing (cellulose membrane) <strong>–</strong> Sigma-Aldrich, cat #D-9277<br />

Dimethyl sulfoxide (DMSO) <strong>–</strong> Sigma-Aldrich, cat #D-5879<br />

DL-Dithiothreitol (DTT) <strong>–</strong> Sigma-Aldrich, cat #D0632-10G<br />

DNA ladder, 1 kb plus <strong>–</strong> Invitrogen, cat #10787-026<br />

Glutathione, reduced <strong>–</strong> Sigma-Aldrich, cat #G-4251<br />

Hybond-C Extra nitrocellulose membrane <strong>–</strong> Amersham Bioscience, cat #RPN303E<br />

Hybond-P PVDF membrane <strong>–</strong> Amersham Bioscience, cat #RPN303F<br />

Kodak BioMax film <strong>–</strong> Kodak Company, cat #1651454<br />

LB Broth Base <strong>–</strong> Invitrogen, cat #1270-052<br />

Leupeptin <strong>–</strong> Sigma-Aldrich, cat#L9783<br />

Lipofectamine-2000 reagent <strong>–</strong> Invitrogen, cat #11668-019<br />

Lysozyme <strong>–</strong> Roche Diagnostic, cat #107255<br />

Magnetic beads (Dynabeads® M-280 Streptavidin) <strong>–</strong> Invitrogen, cat#112-05D<br />

2-β-mercaptoethanol <strong>–</strong> Sigma-Aldrich, cat #M-7522<br />

Methylene blue <strong>–</strong> USB, Cat#19220<br />

Pepstatin A <strong>–</strong> Sigma-Aldrich, cat#P5318<br />

Peptone 140 <strong>–</strong> GibcoBRL, cat #30392-021<br />

Phenylmethylsulphonyl fluoride (PMSF) <strong>–</strong> Sigma-Aldrich, cat#P7626<br />

Phorbol-12-myristate 13-acetate (PMA) <strong>–</strong> Sigma-Aldrich, cat#P8139<br />

PhosSTOP phosphatase inhibitor cocktail <strong>–</strong> Roche Applied Science, cat#04906845001<br />

Precision Plus protein standard (All Blue) <strong>–</strong> Bio-Rad Laboratories, cat #161-0373<br />

52


Precision Plus protein standard (Unstained) <strong>–</strong> Bio-Rad Laboratories, cat #161-0363<br />

Protease inhibitor cocktail - Roche Applied Science, cat #11836145001<br />

Protein A coated polystyrene microspheres (0.95μm) <strong>–</strong> Bangs Laboratories, cat#CP02N<br />

Protein G Sepharose 4B <strong>–</strong> Sigma-Aldrich, cat#P3296<br />

Random oligonucleotide primers <strong>–</strong> Invitrogen, cat #48190011<br />

RNase A <strong>–</strong> Sigma-Aldrich, cat #R4875<br />

Rubidium chloride <strong>–</strong> Sigma-Aldrich, cat #R-2252<br />

SuperScriptTM First-strand synthesis system - Invitrogen, cat #11904-018<br />

SuperScriptTM RNase H- reverse transcriptase <strong>–</strong> Invitrogen, cat #18064-014<br />

T4 DNA ligase <strong>–</strong> Roche Diagnostic, cat #10481220001<br />

Triton X-100 <strong>–</strong> BDH chemicals, cat #306324N<br />

TRIzol reagent <strong>–</strong> Invitrogen, cat #15596-018<br />

Yeast extract <strong>–</strong> Merck, cat #1.03753<br />

2.1.3 Buffers/Solutions<br />

Buffer/Solution Composition<br />

AGE buffer (10x; DNA Loading<br />

Buffer)<br />

0.25% Bromophenol Blue<br />

0.1% Sodium dodecyl sulphate (SDS)<br />

100 mM Tris(hydroxymethyl)methylamine (Tris)-Cl<br />

100 mM Ethylenediaminetetra-actetic acid disodium salt<br />

(EDTA)<br />

50% (v/v) Glycerol<br />

pH 8.0<br />

Agarose gel (1%) 1% (w/v)Agarose in deionised water<br />

Antibiotics:<br />

Ampicillin (Sigma-Aldrich) was used at a final concentration of<br />

100 μg/mL<br />

Geneticin (G-418 sulphate; Gibco, Invitrogen)<br />

Hygromycin B (Invitrogen)<br />

EDTA 0.5 M EDTA pH 8.0<br />

LB agar 10% (w/v) Bacto-peptone<br />

5% (w/v) Bacto-yeast extract<br />

10% (w/v) Sodium chloride (NaCl)<br />

1.5% (w/v) Agar<br />

LB medium 10% (w/v) Bacto-peptone<br />

5% (w/v) Bacto-yeast extract<br />

10% (w/v) NaCl<br />

4% Paraformaldehyde 4% (w/v) Paraformaldehyde made in PBS, pH 7.2<br />

Phosphate Buffered Saline (1X; 137 mM NaCl<br />

PBS)<br />

2.7 mM Potassium chloride (KCl)<br />

1.4 mM Potassium dihydrogen orthophosphate (KH2PO4)<br />

4.3 mM Disodium hydrogen orthophosphate (Na2HPO4.7H2O)<br />

pH 7.4<br />

STE buffer<br />

10 mM Tris-Cl pH 8<br />

1 mM EDTA<br />

50 mM NaCl<br />

TAE (50x)<br />

2 M Tris<br />

1 M Acetic acid<br />

0.2 M EDTA<br />

pH 8.2<br />

53


Cell culture solutions:<br />

AlF4 - 10mM Sodium fluoride (NaF)<br />

40µM Aluminium chloride (AlCl3)<br />

Magnesium solution 200 mM Magnesium chloride (MgCl2)<br />

200 mM Calcium chloride (CaCl2)<br />

Manganese solution 200 mM Manganese Chloride (MnCl2)<br />

200 mM CaCl2<br />

HBSS buffer w/o Ca/Mg (1x) 5.36 mM KCl<br />

0.44 mM KH2PO4<br />

0.14 mM NaCl<br />

0.34 mM Na2HPO4.2H2O<br />

5.55 mM D-glucose<br />

10 mM 4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid<br />

(HEPES) solution<br />

pH 6.5<br />

Methylene blue 0.1% (w/v) Methylene blue in water<br />

Solutions to prepare competent cells:<br />

Psi broth 5% (w/v) Bacto-yeast extract<br />

20% (w/v) Bacto-peptone<br />

5% (w/v) Magnesium sulphate (MgSO4)<br />

water to 1L, pH to 7.6 with potassium hydroxide (KOH)<br />

TfbI<br />

30 mM Potassium acetate<br />

100 mM Rubidium chloride (RbCl)<br />

10 mM CaCl2<br />

50 mM MnCl2<br />

15% (v/v) Glycerol<br />

pH to 5.8 with dilute acetic acid<br />

TfbII<br />

10 mM 3-[N-Morpholino] propane sulphonic acid (MOPS)<br />

75 mM CaCl2<br />

10 mM RbCl<br />

15% (v/v) Glycerol<br />

pH to 6.5 with dilute NaOH<br />

Western blot solutions:<br />

TTBS 102.7 mM NaCl<br />

24.7 mM Tris-Cl<br />

2.7 mM KCl,<br />

pH 8.0<br />

0.1% Tween-20<br />

Transfer buffer 48 mM Tris<br />

39 mM Glycine<br />

0.037% SDS<br />

20% v/v Methanol<br />

Blocking buffer TBS-T containing 5% (w/v) non-fat milk powder, or 2% ECL<br />

advanced blocking solution (Amersham)<br />

Western blot buffer TBS-T containing 3% (w/v) non-fat milk powder, or 2% ECL<br />

advanced blocking solution (Amersham)<br />

2-ME stripping solution 62.5 mM Tris pH 7.2<br />

100 mM 2-β-mercaptoethanol<br />

2% (w/v) SDS<br />

High-pH-stripping solution 0.2 M NaOH<br />

54


Phosphorylation assay solutions:<br />

Cell lysis buffer<br />

10 mM Tris pH 7.4<br />

50 mM NaCl<br />

50 mM NaF<br />

1 mM Sodium orthovanadate (Na3VO4)<br />

300 μg/mL PMSF<br />

1% NP-40<br />

Protease inhibitors:<br />

20 μg/mL Aprotinin<br />

10 μg/mL Leupeptin<br />

10 μg/mL Pepstatin A<br />

Or 1 x Complete EDTA-free protease inhibitor cocktail (Roche)<br />

Phosphatase inhibitors:<br />

1 x PhosSTOP phosphatase inhibitor cocktail (Roche)<br />

Kinase buffer 10 mM Tris pH 7.4<br />

50 mM NaCl<br />

5 mM MgCl2<br />

5 mM MnCl2<br />

0.1 mM Na3VO4<br />

Plasmid preparation solutions:<br />

Solution I (GET)<br />

Solution II (NaOH/SDS)<br />

Solution III (Potassium acetate)<br />

SDS-PAGE solutions:<br />

50 mM Glucose<br />

10 mM EDTA<br />

25 mM Tris-Cl (pH 8)<br />

0.2 M NaOH<br />

1% (w/v) SDS<br />

3 M Potassium acetate<br />

11.4% (v/v) Acetic acid<br />

pH 5.2<br />

30% Acrylamide mixture 30% (w/v) Acrylamide<br />

0.2% Bis-acrylamide<br />

2 x Crosslinker acrylamide 48% (w/v) Acrylamide<br />

3.0% (w/v) Bis-acrylamide<br />

Separating Tris-glycine buffer 1.5 M Tris-Cl pH 8.8<br />

Stacking Tris-glycine buffer 0.5 M Tris-Cl pH 6.8<br />

Separating/spacing Tris-tricine 3 M Tris base<br />

buffer<br />

0.3% SDS<br />

pH 8.9 with HCl<br />

Stacking Tris-tricine buffer 1 M Tris-HCl pH 6.8<br />

10% SDS 10% (w/v) SDS<br />

10% APS 10% (w/v) Ammonium persulfate<br />

SDS-PAGE Tris-glycine running 274.4 mM Tris<br />

buffer (10x)<br />

2 M Glycine<br />

10% (w/v) SDS<br />

Cathode 10x buffer<br />

1 M Tris base<br />

1 M Tricine, 1% w/v SDS<br />

Anode 10x buffer 2 M Tris-Cl pH 8.9<br />

SDS-PAGE reducing loading 10% (w/v) SDS<br />

buffer (5x)<br />

30% (w/v) Glycine<br />

360 mM Tris-Cl pH 6.8<br />

0.5 M DTT<br />

0.03% Bromophenol blue<br />

55


SDS-PAGE staining solutions:<br />

Coomassie blue staining solutions<br />

Fixing solution 25% (v/v) Propanol<br />

10% (v/v) Acetic acid<br />

Staining solution (Coomassie blue 0.006% Coomassie brilliant blue R-250<br />

stain)<br />

40% (v/v) Methanol<br />

10% (v/v) Acetic acid<br />

Destaining solution 50% (v/v) Methanol<br />

10% (v/v) Acetic acid<br />

Silver Staining solutions<br />

Fixation solution 40% (v/v) Ethanol<br />

10% (v/v) Acetic acid<br />

Sensitizing solution 30% (v/v) Ethanol<br />

0.125% Glutaraldehyde<br />

0.2% Sodium thiosulphate<br />

6.8% (w/v) Sodium acetate<br />

(make in fume hood)<br />

Silver reaction solution 0.25% Silver nitrate<br />

0.04% Formaldehyde<br />

(make in fume hood)<br />

Developing solution 2.5% (w/v) Sodium carbonate<br />

0.02% Formaldehyde<br />

(make in fume hood)<br />

Stopping solution 1.5% (w/v) EDTA<br />

Preserving solution 10% (v/v) Glycerol<br />

2.1.4 Tissue culture reagents<br />

RPMI 1640 / Advanced RPMI 1640 (23400-062 / 12633-012, GibcoBRL ® / Invitrogen)<br />

Dulbecco's Modified Eagle's Medium (DMEM) / Advanced DMEM (30-2002 /12491-015,<br />

GibcoBRL ® / Invitrogen)<br />

Sf-900 II SFM (10902088, GibcoBRL ® / Invitrogen)<br />

Penicillin, streptomycin, glutamine (PSG; 10378016, GibcoBRL ® / Invitrogen) - 2 mM<br />

glutamine, 50 μg/mL penicillin, 50 μg/mL streptomycin.<br />

2-β-mercaptoethanol (M-7522, Sigma-Aldrich) was used at a final concentration of 0.05 mM.<br />

Fetal calf serum / fetal bovine serum (FCS; GibcoBRL ® / Invitrogen) was heat-inactivated for<br />

30 min at 56°C before use. It was added to culture media at a final concentration of 10%<br />

(v/v).<br />

Recombinant human tumour growth factor β1 (TGF-β1; 100-21B, Peprotech) was kept as a<br />

stock solution at 50 µg/mL in PBS containing 2 mg/mL of albumin, and stored at -20 °C.<br />

56


2.2. Molecular biology reagents<br />

2.2.1 Mammalian cell lines<br />

TK-1 (CRL-2396 , ATCC ® ) <strong>–</strong> Mouse CD8+ T lymphoma cell line from the AKR/CUM<br />

mouse strain which expresses high levels of α4β7, but no β1 integrins. Cultured in RPMI-<br />

1640 supplemented with 10% FCS, PSG, 2 ME at 37°C in 5% CO2.<br />

MTC-1 (kindly donated by Dr P Kilshaw, AFRC Babraham Institute, Cambridge, UK) <strong>–</strong> T<br />

cell hybridoma which expresses high levels of αEβ7. Cultured in RPMI-1640 medium<br />

supplemented with 10% FCS, PSG,, and 5 ng/ml TGF-β1 at 37°C in 5% CO2.<br />

HEK-293T (CRL-1573 , ATCC ® ) <strong>–</strong> Human epithelial kidney cell line transformed with the<br />

SV40 large T-antigen. Cultured in DMEM, containing 10% FCS, and PSG at 37°C in 10%<br />

CO2.<br />

H9 (HTB-176 , ATCC ® ) <strong>–</strong> Human T cell lymphoma. Cultured in RPMI, containing 10%<br />

FCS, and PSG at 37°C in 5% CO2.<br />

CHO-K1 (CRL-9618 , ATCC ® ) <strong>–</strong> Chinese hamster ovary cells. Cultured in DMEM,<br />

containing 10% FCS, and PSG at 37°C in 10% CO2.<br />

C2C12 (CRL-1772 , ATCC ® ) <strong>–</strong> Mouse myoblast cells. Cultured in DMEM, containing 10%<br />

FCS, and PSG at 37°C in 10% CO2.<br />

2.2.2 Insect cell line<br />

Sf9 (CRL-1711 , ATCC ® ) <strong>–</strong> Spodoptera frugiperda (fall armyworm) epithelial ovary cells.<br />

Grown in Sf-900 II SFM.<br />

2.2.3 Bacterial cell line<br />

DH5α (Stratagene): This is a derivative of E.coli K12 (hsd R17, supE44, rec A1, endA1,<br />

lacZM5) which provided a source of competent cells for DNA subcloning.<br />

57


2.2.4 DNA cloning vectors<br />

pGEM ® -T (A3600, Promega): <strong>The</strong> pGEM-T vector system was used for cloning PCR<br />

products.<br />

pGEX-2T/2TK (27-4801-01/27-4587-01, GE Healthcare Life Sciences): <strong>The</strong> pGEX plasmids<br />

are designed for inducible, high-level expression of genes or gene fragments as fusions with<br />

Schistosoma japonicum 26kDa glutathione S-transferase (GST).<br />

pcDNA3.1/Hygro © (V870-20, Invitrogen): pcDNA3.1 is derived from the pcDNA vector and<br />

is designed for high-level stable and transient expression of proteins in mammalian host cells,<br />

using the hygromycin resistance gene for selection.<br />

pcDNA TM 6/V5 (V220-01, Invitrogen): pcDNA6/V5 vector is derived from pcDNA3.1(+) and<br />

is designed for high-level stable and transient expression and detection of recombinant<br />

proteins in mammalian host cells. <strong>The</strong> vector contains the blasticidin resistance gene from<br />

Aspergillus terreus for cell selection.<br />

pIRES2-EGFP (6029-1, BD Biosciences): pIRES2-EGFP contains an internal ribosome entry<br />

site of the encephalomyocarditis virus between the multiple cloning site (MCS) and the<br />

enhanced green fluorescent protein (EGFP) allowing the translation of both the gene of<br />

interest and the EGFP gene.<br />

2.2.5 Enzymes/isotopes<br />

Restriction enzymes: BamHI, BglII, EcoRI, EcoRV, HindIII, KpnI, NheI, NotI, SacI, XbaI,<br />

and XhoI. All restriction enzymes were purchased from Roche Diagnostics.<br />

T4-DNA ligase: Manufactured by Promega.<br />

Expand High Fidelity PCR System DNA polymerase: From Roche Diagnostics.<br />

Ribonuclease A (RNase A): Manufactured by Sigma-Aldrich, and resuspended at a stock<br />

concentration of 10 μg/mL in 10 mM Tris (pH 7.5) and 15 mM NaCl. <strong>The</strong> solution was<br />

heated at 100°C for 15 min to inactivate contaminating DNase activity, and aliquots were<br />

stored at <strong>–</strong>20°C.<br />

Isotopes: 32 P-γATP: [γ 32 P] <strong>–</strong> [adenosine 5’-triphosphate]-[tetra(triethylammonium) salt] from<br />

Amersham Biosciences.<br />

58


2.2.6 Antibodies<br />

Antigen recognised Company or source Dilution used<br />

β-actin ab8227, Abcam 1:1000<br />

E-cadherin BTA3, R & D Systems 1:200<br />

E-cadherin ECCD-1, R & D Systems 25 μg/mL<br />

FAK PC314, Calbiochem 1:200<br />

Hsp90 (N-17) sc-1055, Santa Cruz Biotechnology 1:200<br />

Hsp70 OBT1688, AbD Serotec 1:200<br />

integrin α4 (PS/2) Produced from hybridoma cells by host 1:200<br />

laboratory (Purchased from ATCC)<br />

integrin α4 120-65984, Sapphire Biosciences 1:200<br />

integrin α-M290 Produced from hybridoma cells by host<br />

integrin-αE (R-15)<br />

laboratory (Kindly donated by Dr P Kilshaw,<br />

AFRC Babraham Institute, Cambridge, UK)<br />

sc-6607, Santa Cruz Biotechnology 1:200<br />

Integrin-α4β7 (DAKT32) Produced from hybridoma cells by host 1:100<br />

laboratory (Purchased from ATCC)<br />

integrin-β7 (Fib504) Produced from hybridoma cells by host 1:100<br />

integrin-β7-cytoplasmic domain<br />

laboratory (Purchased from ATCC)<br />

Produced by host laboratory (Krissansen et al. 1:400<br />

1992)<br />

integrin-β7-extracellular domain AF4669, R & D Systems 1:1000<br />

lck sc-13, Santa Cruz Biotechnology 1:200<br />

MAdCAM-1 (MECA357) Produced by host laboratory 25 μg/mL<br />

phosphotyrosine Ab-4, Calbiochem 1:200<br />

src ABsrc.1, Oncor<br />

src 2109, Cell Signaling Technology 1:200<br />

src 2110, Cell Signaling Technology 1:200<br />

Secondary antibodies:<br />

Goat anti-human-Fc-IgG A0170, Sigma-Aldrich 1:1000<br />

Goat anti-mouse IgG peroxidase<br />

conjugated<br />

A4416, Sigma-Aldrich 1:80,000<br />

Goat anti-rabbit IgG peroxidase<br />

conjugated<br />

A0545, Sigma-Aldrich 1:10,000<br />

Mouse anti-goat/sheep IgG A9452, Sigma-Aldrich 1:80,000<br />

peroxidase conjugated<br />

Alexa Fluor® 488 goat anti-rat<br />

IgG<br />

A-11006, Molecular Probes<br />

1:250<br />

Alexa Fluor® 594 goat antirabbit<br />

IgG<br />

A-11012, Molecular Probes 1:250<br />

Alexa Fluor® 594 donkey antigoat<br />

IgG<br />

A-11058, Molecular Probes<br />

1:250<br />

Other Antibodies:<br />

Mouse ascites Produced by host laboratory<br />

Rat-IgG I8015, Sigma-Aldrich<br />

59


2.2.7 PCR oligonucleotide primers<br />

All oligonucleotide primers were purchased from Invitrogen.<br />

Primers for integrin α4 cloning:<br />

CC28 primer: 5’ AAGCTTGATATCATGGCTTGGGAAGCGAGG 3’<br />

CC29 primer: 5’ AAACACTCTTCCTTCCTCTC 3’<br />

Primers for integrin β7 cloning:<br />

CC24 primer: 5’ GGATCCGAGCTCCTGGGGATGTGTGCGA 3’<br />

CC25 primer: 5’ GATTTCCACGGAGAGCGGGTAAGCCAGGAC 3’<br />

CC26 primer: 5’ GTCCTGGCTTACCGGCTCTCGGTGGAAATC 3’<br />

CC27 primer: 5’ GAATTCGCGGCCAGAGAGTGGGACTG 3’<br />

2.2.8 cDNA for cloning/protein expression<br />

GST-β7 cytoplasmic domain construct: A pGEX-2T construct encoding the cytoplasmic<br />

domain of the β7 subunit was obtained from Dr Andrew Lazarovits, Imperial Cancer<br />

Research Fund, London, UK.<br />

pGEX-2T vectors encoding mutated forms of the β7 subunit cytoplasmic domain were<br />

constructed by Dr Klaus Lehnert of the host laboratory (<strong>The</strong> <strong>University</strong> of Auckland).<br />

α4 full-length DNA construct in pCDM8 kindly donated by Professor B. Seed (Harvard<br />

Medical School Massachusetts General Hospital, MA, USA).<br />

β7 full-length DNA construct in pCDM8 constructed by Dr Euphemia Leung of the host<br />

laboratory (<strong>The</strong> <strong>University</strong> of Auckland).<br />

60


2.2.9 Recombinant proteins<br />

Name Companies<br />

FAK PKA-312, ProSpec; PHO3145, BioSource<br />

src PKA-317, ProSpec; S5439, Sigma-Aldrich<br />

p56lck (lck) L2792, Sigma-Aldrich<br />

Hsp70a HSP-170, ProSpec<br />

Filamin 62006, Progen<br />

Paxillin P3113-85, United States Biological<br />

α-actinin AT01, Cytoskeleton Inc.<br />

2.2.10 Inhibitors of intracellular signalling pathways<br />

PD98059 <strong>–</strong> C16H13NO3, 2′-Amino-3′-methoxyflavone (Cat. No. 513000, Calbiochem) <strong>–</strong><br />

Inhibitor of MAP kinase kinase (MEK).<br />

SB203580 <strong>–</strong> C21H16N3OSF, 4-(4-Fluorophenyl)-2-(4-methylsulfinyl phenyl)-5-(4-pyridyl)1H-<br />

imidazole, (Cat. No. 559389, Calbiochem) <strong>–</strong> Inhibitor of p38 MAP kinase.<br />

JNK-I <strong>–</strong> C168H293N67O42, (L)-HIV-TAT48-57-PP-JBD20, (L)-JNKI1, c-Jun NH2-terminal kinase<br />

H-GRKKRRQRRRPPRPKRPTTLNLFPQVPRSQDT-NH2, SAPK Inhibitor I (Cat. No.<br />

420116, Calbiochem) <strong>–</strong> Blocks c-Jun NH2-terminal kinase (JNK) signalling.<br />

JNK-II <strong>–</strong> C14H8N2O, Anthra[1,9-cd]pyrazol-6(2H)-one, 1,9-pyrazoloanthrone, SAPK<br />

Inhibitor II, SP600125 (Cat. No. 420119, Calbiochem) <strong>–</strong> Inhibitor of c-Jun N-terminal kinase<br />

(JNK).<br />

Radicicol <strong>–</strong> C18H17ClO6 (Cat. No. 553400, Calbiochem) <strong>–</strong> Inhibits v-src kinase activity.<br />

AG99 <strong>–</strong> C10H8N2O3, a-Cyano-(3,4-dihydroxy)cinnamide, Tyrphostin A46, Tyrphostin B40<br />

(Cat. No. 658430, Calbiochem) <strong>–</strong> Inhibits epidermal growth factor (EGF) receptor tyrosine<br />

kinase.<br />

PP2 <strong>–</strong> C15H16ClN5, AG 1879, 4-Amino-5-(4-chlorophenyl)-7-(t-butyl)pyrazolo[3,4d]pyrimidine<br />

(Cat. No. 529576, Calbiochem) <strong>–</strong> Inhibitor of the src family of protein tyrosine<br />

kinases. Inhibits lck, fyn, hck, and src.<br />

Damnacanthal <strong>–</strong> C16H10O5, 3-Hydroxy-1-methoxyanthraquinone-2-aldehyde (Cat. No.<br />

251650, Calbiochem) <strong>–</strong> Inhibitor of the tyrosine kinase lck.<br />

61


ML-7 <strong>–</strong> C15H17IN2O2S . HCl, 1-(5-Iodonaphthalene-1-sulfonyl)homopiperazine-HCl (Cat. No.<br />

345834 , Calbiochem) <strong>–</strong> Inhibitor of myosin light chain kinase (MLCK).<br />

Genistein <strong>–</strong> C15H10O5, 4′,5,7-Trihydroxyisoflavone (Cat. No. 345834, Calbiochem) <strong>–</strong> Inhibits<br />

protein tyrosine kinases.<br />

KNK-437 <strong>–</strong> C13H11NO4, Heat shock protein inhibitor 1, N-Formyl-3,4-methylenedioxybenzylidine-γ-butyrolactam<br />

(Cat. No. 373260, Calbiochem) <strong>–</strong>Inhibits heat shock proteins,<br />

including HSP70, Hsp72, and HSP105.<br />

2.2.11 Peptides<br />

Biotin-rrrrrrrrr-pptdqsrpvqpflnlttprkpr-OH (Mimotopes) <strong>–</strong> used as a negative control peptide<br />

for poly-arginine containing peptides, it is an inverse sequence of part of the mitogen-<br />

activated protein kinase 8 interacting protein 1 (157-176 aa)<br />

Biotin-APTLPPAWQPFLK-OH (Mimotopes) <strong>–</strong> used as a negative control peptide for non-<br />

poly-arginine containing peptides, it is part of the baculoviral IAP repeat-containing 5 protein<br />

(3-15 aa)<br />

Biotin-GGYDRREY-OH (GGYDRREY; Mimotopes)<br />

Biotin-rrrrrrrrr-YDRREY-OH (YDRREY; Mimotopes)<br />

Biotin-rrrrrrrrr-YDRRE-OH (YDRRE; Mimotopes)<br />

Biotin-rrrrrrrrr-DRREY-OH (DRREY; Mimotopes)<br />

Biotin-rrrrrrrrr-YDGGEY-OH (YDGGEY; Mimotopes)<br />

Biotin-rrrrrrrrr-YEEEEY-OH (YEEEEY; Mimotopes)<br />

Biotin-rrrrrrrrr-YDRGGGGREY-OH (YDRGGGGREY; Mimotopes)<br />

Biotin-rrrrrrrrr-xDRREx-OH (xDRREx; Mimotopes)<br />

Biotin-rrrrrrrr-YDRREYGYDRREYGYDRREYGYRDDEY-OH (pYDRREY; Mimotopes)<br />

Biotin-rrrrrrrr-FDRREFGFDRREFGFDRREFGFDRREF-OH (pFDRREF; Mimotopes)<br />

62


5-6-FAM-rrrrrrrrr-YDRREY-OH (FAM-YDRREY; Mimotopes)<br />

Biotin-rrrrrrrrr-SILQEENRRDSWSYI-OH (Mimotopes) <strong>–</strong> α4-integrin-paxillin binding site.<br />

Biotin-RLSVEIYDRREYSRFEKEQQQLNWKQDSNPLYKSAITTTINPRFQEADSPTL-<br />

OH (Synthesized by Chemistry Department, <strong>The</strong> <strong>University</strong> of Auckland) <strong>–</strong> full length β7<br />

subunit cytoplasmic domain<br />

Biotin-rrrrrrrr-NPLY-OH (NPLY; Peptide 2.0)<br />

Polymer of glutamate and tyrosine (4:1), (PolyEY, Sigma-Aldrich)<br />

63


<strong>Section</strong> 2 <strong>–</strong> <strong>Methods</strong><br />

2.3. Molecular biology techniques<br />

2.3.1 Preparation of competent cells<br />

DH5α E. coli were inoculated into Psi broth and cultured at 37°C overnight. One millilitre of<br />

overnight culture was inoculated into 100 mL of Psi Broth, and incubated at 37°C with<br />

aeration to an optical density of 0.48 at 550 nm. <strong>The</strong> culture was then placed on ice for 15<br />

min, and cells pelleted at 5,000 × g (Sorvall) for 5 min. <strong>The</strong> pellet was resuspended in 0.4<br />

volume of TfbI, and placed on ice for a further 15 min. Cells were pelleted again at 5,000 × g<br />

for 5 min, resuspended in 0.04 volume of TfbII, and placed on ice for a further 15 min.<br />

Aliquots of 100 μL were snap-frozen in ethanol-dry ice, and stored at <strong>–</strong>80°C until needed.<br />

2.3.2 Small scale preparation of plasmid DNA (miniprep)<br />

DH5α E. coli that were transformed with the desired plasmid were grown overnight<br />

(approximately 16 hrs) in 5 mL of LB medium containing 100 μg/mL of ampicillin (amp). A<br />

1.5 mL aliquot of culture was centrifuged for 1 min at 12,000 × g. <strong>The</strong> pellet was resuspended<br />

in 100 μL of GTE buffer (solution 1) by vortexing, and left for 5 min at room temperature<br />

(RT). <strong>The</strong> cells were lysed by the addition of 200 μL of lysis solution (solution 2) with gentle<br />

inversion 5 times, and left to stand for a further 5 min. <strong>The</strong> solution was neutralised with the<br />

addition of 150 μL of ice-cold potassium acetate (solution 3) and mixed immediately by<br />

repeated inversion. <strong>The</strong> mixture was centrifuged for 10 min at maximum speed (12,000 × g)<br />

on a bench top centrifuge and the supernatant transferred into 500 μL of chloroform (CHCl3),<br />

followed by vortexing. <strong>The</strong> mixture was centrifuged a further 10 min at 12,000 × g, and 80-<br />

90% of the top aqueous phase was transferred into a new tube containing 1 mL of 100%<br />

ethanol. This mixture was then vortexed, and centrifuged for 5 min at 12,000 × g. <strong>The</strong> pellet<br />

was washed once with 70% ethanol, air-dried at 37°C, then re-suspended in 30 μL of sterile<br />

water containing RNase (10 μL/mL), and stored at <strong>–</strong>20°C until needed. Plasmids to be<br />

subjected to DNA sequencing were purified using a Qiagen ® Rapid Plasmid Mini Kit<br />

according to the manufacturer’s instructions. All DNA sequencing was carried out at the<br />

DNA sequencing facility in the School of Biological Sciences, Faculty of Science, <strong>University</strong><br />

of Auckland.<br />

64


2.3.3 Large scale preparation of plasmid DNA<br />

Large scale plasmid preparations were carried out using either a midiprep kit from Invitrogen,<br />

or by centrifugation through caesium chloride gradients.<br />

Large scale preparation of plasmid DNA with the midiprep kit<br />

Purification of plasmids was carried out using the PureLink HiPure Midiprep kit<br />

(Invitrogen) by following the manufacturer’s instructions. All required buffers were provided<br />

with the kit. <strong>The</strong> following is a brief description of the method. DH5α E. coli containing the<br />

desired plasmid was cultured overnight in LB media at 37°C. <strong>The</strong> bacteria were pelleted and<br />

resuspended in 4 mL of Resuspension Buffer containing RNase, and then lysed by the<br />

addition of 4 mL of Lysis Buffer with gentle inversion. <strong>The</strong> solution was neutralised using 4<br />

mL of Precipitation Buffer and the mixture was centrifuged at 12,000 × g for 10 min at RT.<br />

<strong>The</strong> soluble fraction was passed through the provided columns and washed twice with 10 mL<br />

of Wash Buffer, and the DNA was eluted with 5 mL of Elution Buffer. <strong>The</strong> DNA was<br />

precipitated with the addition of 3.5 mL of isopropanol and pelleted at 15,000 × g for 30 min<br />

at 4°C. <strong>The</strong> pellet was further washed with 3 mL of 70% ethanol and re-pelleted at 15,000 × g<br />

for 30 min at 4°C. <strong>The</strong> pellet was air-dried for 10 min at 37°C and the DNA pellet was<br />

resuspended in 200 µL of water and stored at -20°C until needed.<br />

Large scale preparation of plasmid DNA with caesium chloride<br />

Bacteria transformed with a desired plasmid were cultured overnight in LB medium with 100<br />

μg/ml of amp. Five millilitres of the resulting culture was inoculated into 500 ml of fresh LB<br />

medium containing 100 μg/ml amp. <strong>The</strong> culture was incubated overnight at 37°C with<br />

vigorous shaking. <strong>The</strong> bacteria were collected by centrifugation at 12,000 × g, and then<br />

resuspended in 40 ml of GTE buffer (solution 1). <strong>The</strong> bacteria were lysed by the addition of<br />

80 ml of 0.2 M NaOH/1% SDS (solution 2) with gentle inversion. <strong>The</strong> solution was<br />

neutralized by the addition of 40 ml of ice-cold potassium acetate (solution 3), mixed<br />

thoroughly by inversion, and placed on ice for 10 min. <strong>The</strong> mixture was centrifuged at 12,000<br />

× g for 20 min at 4°C. <strong>The</strong> supernatant was transferred to a new bottle, and 90 ml of<br />

isopropanol was added followed by centrifugation at 12,000 × g for 20 min at 15°C. <strong>The</strong><br />

pellet was washed with 40 ml of 70% ethanol followed by centrifugation at 12,000 × g for 5<br />

min at 4°C. <strong>The</strong> plasmid DNA was resuspended in 4 ml of water and mixed with 4.5 g of<br />

caesium chloride. To the mixture was added 100 μl of ethidium bromide solution (stock 10<br />

mg/ml) and 51 μl of Triton-X100 (1:100 stock). <strong>The</strong> mixture was transferred into an<br />

65


ultracentrifuge tube, and centrifuged for 22 hrs at 78,000 rpm (40,000 × g) at 20°C under<br />

vacuum. <strong>The</strong> plasmid DNA band, which was coloured a faint pink, was harvested and mixed<br />

with 2x volume of saturated-butanol. <strong>The</strong> mixture was shaken and left standing until two<br />

layers formed. <strong>The</strong> pink DNA layer containing traces of ethidium bromide was removed and<br />

saturated-butanol was again added followed by shaking. <strong>The</strong> pink DNA layer was harvested,<br />

and the process was repeated until the solvent phase became clear. <strong>The</strong> aqueous phase was<br />

transferred to a clean tube and mixed with 7.5 ml of 100% ethanol. <strong>The</strong> mixture was<br />

centrifuged at 12,000 × g for 20 min, the supernatant was removed, and the pellet was washed<br />

with 7.5 ml of 70% ethanol and air-dried. <strong>The</strong> plasmid pellet was resuspended in sterile water<br />

and stored at -80°C.<br />

2.3.4 Linearization and dephosphorylation of plasmid vectors<br />

Plasmid vectors were linearised by digestion with appropriate enzymes for 2 hrs at 37°C,<br />

unless otherwise stated. <strong>The</strong> linearised plasmids were then de-phosphorylated by adding 2 μL<br />

of alkaline phosphatase to a reaction volume of 50 μL and further incubated at 37°C for 45<br />

min. <strong>The</strong> DNA was electrophoresed on 0.8-1% agarose gels, and the desired DNA band(s)<br />

was excised. <strong>The</strong> DNA was isolated from the gel slice using a Wizard ® DNA clean-up kit<br />

(Promega) by following the manufacturer’s instructions.<br />

2.3.5 Isolation of plasmid inserts<br />

DNA inserts were liberated from plasmids by restriction enzyme digestion with appropriate<br />

enzymes and temperatures for 2 hrs. <strong>The</strong> DNA inserts were electrophoresed on 1-2% agarose<br />

gels, and the desired DNA fragment(s) excised, and purified using Wizard ® DNA clean-up kit<br />

(Promega) following manufacturer’s instructions.<br />

2.3.6 Agarose gel electrophoresis<br />

DNA electrophoresis was carried out in 1x TAE solution using 0.8-2% agarose depending on<br />

the expected size of the desired DNA fragments. Electrophoresis was carried out at 50 V for<br />

1-2 hrs. Ethidium bromide (2 μL of 10 mg/mL) was added to the TAE solution<br />

(approximately 50 mL) after electrophoresis and gently mixed on a platform shaker for 10-15<br />

min. <strong>The</strong> gels were viewed using an EagleEye © UV light gel imager (Bio-Rad), and<br />

photographs were taken.<br />

66


2.3.7 Ligation of DNA inserts into vectors<br />

Ligation reactions were mixed as shown in the following table.<br />

Reaction Mixture Background control Reaction (1:3) Reaction (1:6)<br />

10x Buffer 1 μL 1 μL 1 μL<br />

Vector 5 μg 5 μg 5 μg<br />

DNA insert - 15 μg 30 μg<br />

T4 DNA ligase - 1 μL 1 μL<br />

Water to 10 μL 10 μL 10 μL<br />

A background control was carried out in which insert and ligase were omitted, to determine<br />

whether the vector was linearised and dephosphorylated properly. Two ligation reactions<br />

using different ratios of vector to insert were carried out. Where two inserts were to be ligated<br />

into one vector, the overall insert concentration was kept at 30 μg/10 μl with a ratio of<br />

approximately 2:1 (smaller : larger) for the DNA fragments to be ligated. All reaction<br />

mixtures were left at either 4°C overnight or at 16°C for 8 hrs.<br />

2.3.8 Transformation<br />

Frozen competent DH5α cells were thawed on 30% ice-water. A 50 μL aliquot of the<br />

competent cells was mixed gently with 5-10 μL of ligation mixture, and incubated on ice-<br />

water for a 15 min. <strong>The</strong> cell-ligation mixture was then heat-shocked for 45 s at 42°C, then<br />

placed back on ice-water immediately for a further 5 min. <strong>The</strong> mixture was then added to 1<br />

mL of LB medium (without antibiotics) and incubated at 37°C with aeration for 45 min. <strong>The</strong><br />

cells were pelleted at 5,000 × g for 5 min and most of the LB medium was removed leaving<br />

100 μL. <strong>The</strong> cells were resuspended in the 100 μL solution and spread-out on LB-agar plates<br />

containing antibiotics, and grown overnight at 37°C.<br />

2.3.9 RNA isolation<br />

Cells to be used for RNA isolation were washed and pelleted. An aliquot of 10 7 cells was<br />

incubated with 1 mL of TRIzol reagent for 5 min at RT. <strong>The</strong> solution was vortexed for 15 s<br />

with the addition of chloroform (0.2 ml per 1 ml of TRIzol), and then incubated for a further<br />

15 s at RT. <strong>The</strong> mixture was centrifuged at 12,000 × g for 15 min at 4°C, and the aqueous<br />

phase was collected and dispensed into a fresh tube. Propanol (0.5 ml per 1 ml of TRIzol) was<br />

added and the tube was inverted several times in order to mix the solutions. <strong>The</strong> mixture was<br />

centrifuged at 12,000 × g for 10 min at 4°C and the supernatant discarded. <strong>The</strong> pellet was<br />

67


washed once with 75% (v/v) ethanol (diluted in DEPC-treated water) and air-dried. <strong>The</strong> RNA<br />

pellet was resuspended in DEPC-treated water and stored at -80°C.<br />

2.3.10 Synthesis of cDNA by reverse transcriptase<br />

Complementary DNA was synthesized from mRNA using the SuperScript TM First-Strand<br />

Synthesis System from Invitrogen. Total RNA (1-3 μg) was mixed with 1 μl of dNTPs (10<br />

mM), 1 μl of random oligos (50 ng/μl) and DEPC-treated water in a 13 μl reaction mixture.<br />

<strong>The</strong> mixture was incubated at 65°C for 5 min, quick-chilled on ice, and 2 μl of 10X reverse<br />

transcriptase (RT) buffer, 4 μl of MgCl2 (25 mM), and 2 μl of DTT (0.1 M) were added. <strong>The</strong><br />

reaction mixture was incubated at 25°C for 2 min and 1 μl (50 units) of SuperScript TM RT<br />

was added, and the mixture incubated for a further 10 min at 25°C, then at 42°C for 50 min.<br />

<strong>The</strong> reaction was terminated by heating at 70°C for 15 min, and 1 U of RNase H was added<br />

followed by incubation at 37°C for 20 min before storage at -80°C.<br />

2.3.11 Polymerase chain reaction (PCR)<br />

A polymerase chain reaction was prepared by mixing 15 ng of template DNA or 2 μL of<br />

cDNA from RT of mRNA, 2 mM dNTPs, 1x PCR reaction buffer, 0.1 μM of specific forward<br />

and reverse primers (synthesized by Invitrogen), 2 U High fidelity DNA polymerase (Roche)<br />

or 2 U of Taq polymerase (supplied by Professor John Fraser, Department of Molecular<br />

Medicine & Pathology, <strong>The</strong> <strong>University</strong> of Auckland) in sterile MilliQ grade water. <strong>The</strong><br />

reactions were carried out in a DNA thermal cycler (PE Applied Biosystems, Gene Amp ®<br />

PCR system 9700). <strong>The</strong> cycling program consisted of 25 cycles with the following steps: a<br />

denaturation temperature of 95°C for 30 s, an annealing temperature of 50°C to 60°C for 30 s<br />

depending on the annealing properties of the different PCR primer sets, and an elongation<br />

temperature of 72°C for 30 s.<br />

Overlapping PCR<br />

Overlapping PCR was performed in order to join two separate fragments of DNA with<br />

overlapping ends. <strong>The</strong> PCR was preformed as above with equal amounts of two separate<br />

DNA fragments having overlapping complementary ends.<br />

68


2.4. Protein chemistry<br />

2.4.1 Production and purification of GST fusion proteins<br />

An overnight bacterial culture (5 mL) containing a desired plasmid was inoculated into 500<br />

mL of LB medium containing 100 μg/mL amp, and grown on a bacterial shaker at 37°C until<br />

an optical density at 600 nm of between 0.6 and 0.8 was reached. Synthesis of fusion proteins<br />

was induced by the addition of IPTG (0.4-0.8 mM), and the bacteria cultured for a further 2 to<br />

3 hrs at 30°C or 37°C. <strong>The</strong> bacteria were collected by centrifugation at 3000 × g for 10 min at<br />

4°C, washed once with PBS, and frozen at <strong>–</strong>80°C. <strong>The</strong> frozen pellet was thawed on ice and<br />

resuspended in 10 mL of STE containing 1% (v/v) Triton X-100. Cells were lysed by the<br />

addition of 1 % (w/v) lysozyme, and 10 mM DTT, and sonicated for 1 min. <strong>The</strong> sonicate was<br />

then centrifuged at 15,000 × g for 10 min at 4°C, and the supernatant was transferred to a 50<br />

ml tube. One millilitre of glutathione-agarose slurry was washed with several volumes of PBS<br />

and added to the supernatant followed by incubation for 2 to 3 hrs on a rotating wheel at 4°C.<br />

<strong>The</strong> beads and supernatant were transferred to a 10 ml disposable-mini column (Bio-Rad),<br />

washed once with 10 mL of STE, and then with 50 ml of PBS. <strong>The</strong> beads were kept in PBS<br />

with 50% glycerol containing 0.02% NaH3, and stored at 4°C.<br />

2.4.2 Sodium dodecylsulphate polyacrylamide gel electrophoresis<br />

Proteins were analysed under reducing Sodium dodecylsulphate polyacrylamide gel<br />

electrophoresis (SDS-PAGE) unless otherwise stated. <strong>The</strong> protein samples were mixed with<br />

an equal volume of SDS-PAGE reducing loading buffer and heated at 98°C for 5 min.<br />

Samples (1-10 µg) were loaded into the wells of the polyacrylamide SDS-gel. Minielectrophoresis<br />

tanks including the Hoefer Mighty Small SE250/SE260 mini vertical unit<br />

(Amersham Bioscience) and the Mini Protean ® II Cell (Bio-Rad) were used. Two types of<br />

SDS gels were employed: Tris-glycine gels for proteins with molecular weights greater then<br />

20 kDa, and Tris-tricine gels for proteins with molecular weights less than 20 kDa. Trisglycine<br />

gels were run using a Tris-glycine running buffer, whereas for Tris-tricine gels the<br />

proteins were resolved using a cathode and anode buffer system. Electrophoresis was carried<br />

out at a constant current of 20 mA per gel for approximately 2-6 hrs depending on<br />

polyacrylamide composition of the gel. If required, gels were either stained with Coomassie<br />

blue or silver stain, and dried using a gel drying system (Model 583 Gel Dryer, Bio-Rad;<br />

DryEase Gel Drying System, Invitrogen) for storage or autoradiography.<br />

69


Tris-glycine gels<br />

Tris-glycine gels were prepared as follows for a 10% polyacrylamide SDS gel.<br />

Separating gel (10 mL):<br />

Water 4 mL<br />

30% Acrylamide mixture 3.3 mL<br />

Separating Tris-glycine buffer 2.5 mL<br />

10% SDS 0.1 mL<br />

10% APS 100 µL<br />

TEMED 4 µL<br />

Stacking gel (4 mL):<br />

Water 2.7 mL<br />

30% Acrylamide mixture 0.67 mL<br />

Stacking Tris-glycine buffer 0.5 mL<br />

10% SDS 40 µL<br />

10% APS 40 µL<br />

TEMED 4 µL<br />

Tris-tricine gels<br />

Tris-tricine gels (16% polyacrylamide) for the separation of low molecular weight proteins<br />

were cast with a separating gel layer, a 1 cm spacing gel, and a stacking gel to form wells for<br />

protein loading. Gels were made by mixing the following reagents:<br />

Separating gel (30 mL):<br />

Water 6.7 ml<br />

Separating/spacing Tris-tricine buffer 10 ml<br />

2 x Crosslinker acrylamide 10 ml<br />

Glycerol 3.2 ml<br />

TEMED 10 μl<br />

10% APS 100 μl<br />

Spacing gel (15 mL):<br />

Water 6.95 ml<br />

Separating/spacing Tris-tricine buffer 5 ml<br />

30% Acylamide mixture 3 ml<br />

TEMED 5 μl<br />

10% APS 50 μl<br />

70


Stacking gel (15 mL):<br />

Water 10.3 ml<br />

Stacking Tris-tricine buffer 1.9 ml<br />

30% Acrylamide mixture 2.5 ml<br />

0.5M EDTA 150 μl<br />

TEMED 7.5 μl<br />

10% APS 150 μl<br />

2.4.3 Staining of SDS gels<br />

Coomassie blue staining<br />

SDS gels were fixed with 10% (v/v) acetic acid for 10 min, and then stained with Coomassie<br />

blue staining solution for 30 min with gentle agitation. <strong>The</strong> gel was then de-stained with destaining<br />

solution under gentle agitation, with the presence of an absorbent sponge or tissue to<br />

aid dye absorption, until bands were clearly visible.<br />

Silver staining<br />

Each silver staining solution was prepared freshly before use. <strong>The</strong> staining of SDS gels was<br />

carried out at RT. SDS gels were fixed for 30 min with silver staining Fixation buffer,<br />

followed by treatment with Sensitisation buffer for 30 min. <strong>The</strong> gels were washed with<br />

MilliQ-grade water for at least 3 x 5 min, and silver-stained by incubation in Silver staining<br />

solution for 20 min. Excess staining solution was removed and the gel washed again with<br />

MilliQ-grade water for 2 x 1 min washes. Developing solution was added for 3 to 5 min until<br />

the desired intensity of staining of proteins was obtained. Further development was<br />

terminated by incubating the gel with Stop solution for 10 min. Gels were preserved by<br />

placing in Preserving solution for at least 20 min.<br />

2.4.4 Western blot analysis<br />

Proteins separated by SDS-PAGE were transferred onto Hybond -C-extra nitrocellulose<br />

membrane or Hybond -P PVDF membranes (Amersham Biosciences) using a Semi-Dry<br />

Transfer Unit (Pharmacia Biotech). Three transfer buffer-soaked Whatman filter papers were<br />

placed onto the unit, and overlaid with nitrocellulose / PVDF membrane. An acrylamide gel<br />

was placed on top of the membrane, and covered by three more transfer buffer-soaked filter<br />

papers. Transfer was carried out at 0.8 mA/cm 2 for 2 hrs. <strong>The</strong> membrane was then stained<br />

with Ponceau S solution to confirm protein transfer, washed in water, and incubated in<br />

71


locking solution overnight. <strong>The</strong> membrane was then washed thrice in TTBS for 5 min each,<br />

and incubated in Western blot solution containing the primary antibody for 1 <strong>–</strong> 2 hrs at RT or<br />

overnight (16 hrs) at 4°C. <strong>The</strong> membrane was once again washed thrice in TTBS for 5 min,<br />

and incubated with Western blot solution containing the secondary antibody for 1 hr. Finally,<br />

the membrane was washed thrice for 5 min, and developed with Amersham ECL Plus TM or an<br />

ECL Advance TM kit (Amersham Biosciences). Antibody reactivity was visualised using either<br />

the FujiFilm LAS-3000 scanner (FUJIFILM Cooperation) or by exposure to Kodak BioMax<br />

XAR film (165 1454, Eastman Kodak Company) and development using the Kodak M35 X-<br />

OMAT processor (Eastman Kodak Company).<br />

2.4.5 Stripping of nitrocellulose and PVDF membranes<br />

Stripping of PVDF membranes by heat and detergent.<br />

PVDF membranes were washed with deionised water, and incubated with blocking solution<br />

for 1 hr at RT. <strong>The</strong>y were again washed with deionised water, and incubated with 2-ME<br />

stripping solution for 30 min at 50˚C.<br />

Stripping of nitrocellulose membranes with high pH.<br />

Nitrocellulose membranes were washed with deionised water, and incubated with blocking<br />

solution for 1 hr at RT. Membranes were again washed with deionised water, and incubated<br />

with high pH stripping solution for 10 min at RT. After stripping, the membranes were<br />

washed with copious amounts of water for at least 30 min at RT before re-blocking with<br />

blocking solution and antibody probing.<br />

2.4.6 Kinase assays<br />

Phosphorylation of GST-fusion proteins<br />

TK-1 cells grown in RPMI were harvested, and 5x10 7 cells were lysed in 1 mL of cell lysis<br />

buffer on ice for 10 min. <strong>The</strong> cell lysate was centrifuged at 12,000 × g for 15 to 20 min at<br />

4°C. One hundred microlitres of supernatant was pre-cleared once with 10 μL of activated<br />

GSH-Sepharose and twice with 10 μL of GST-2T-Sepharose for 30 min at 4°C. A 1 μL<br />

aliquot of pre-cleared supernatant was mixed and incubated with 10 μL of GST-fusion<br />

proteins coupled to Sepharose beads for 2 hrs at 4°C. GST-fusion protein-beads were then<br />

washed thrice with ice-cold lysis buffer and once with kinase buffer. <strong>The</strong> beads were<br />

suspended in 30 μL of kinase buffer, 1 μCi of 32 P γ-ATP was added, and the mixture was<br />

72


incubated for 20 min at 30°C. Alternatively, 0.4 μg of recombinant kinase(s) was mixed with<br />

10 μL of Sepharose beads in 30 μL of kinase buffer and 1 μCi of 32 P γ-ATP, and incubated<br />

for 20 min at 30°C. <strong>The</strong> phosphorylated GST-fusion protein-Sepharose beads were washed<br />

extensively with kinase buffer to remove all non-specific radioactivity. <strong>The</strong> amount of<br />

phosphorylation was measured with a Wallac 1450 TRILUX Microbeta ® counter or<br />

visualised by separation of proteins via SDS-PAGE followed by gel drying, and<br />

autoradiography.<br />

Phosphorylation of peptides<br />

Biotinylated peptides were bound to either streptavidin-coated Sepharose (Sigma-Aldrich) or<br />

magnetic beads (Dynal, Invitrogen) for 30 min at RT (1 mM peptide/mg of beads) and<br />

resuspended in kinase binding buffer. Recombinant kinases (0.4 µg) or trace amounts of cell<br />

lysate (1 μL of 3 x 10 7 cells/mL diluted 1:100) were added together with 1 µCi of 32 P-γATP<br />

per 30 µL of reaction mixture, and incubated for 30 min at 30°C. Some assays used<br />

autophorphoylated kinases prior to addition to peptides. Phosphorylation of peptides was<br />

either determined by measurement with a Wallac 1450 TRILUX Microbeta ® counter, or<br />

visualised by separation of proteins by Tris-tricine SDS-PAGE followed by autoradiography.<br />

Assays to determine kinase binding to peptides:<br />

Peptides immobilised on beads were mixed with recombinant kinases (0.4 μg) for 2 hrs at 4°C<br />

in cell lysis buffer without detergent. <strong>The</strong> beads were then washed six times with lysis buffer<br />

without detergent and twice with kinase buffer before the addition of 1 μCi of 32 P-γATP as<br />

above.<br />

Autophosphorylation of kinases:<br />

Recombinant kinases were autophosphorylated by incubation of kinases (0.4 µg) with 1 µCi<br />

of 32 P-γATP in 30 µL of kinases buffer, followed by incubation for 30 min at 30°C.<br />

Assays to measure the binding of recombinant proteins to β7 subunit cytoplasmic<br />

domain peptides:<br />

Integrin β7 subunit cytoplasmic domain peptides immobilised on beads were mixed with<br />

recombinant proteins (0.1 μg) and recombinant kinases (0.4 μg) for 2 hrs at 4°C in cell lysis<br />

buffer without detergent. <strong>The</strong> beads were then washed six times with lysis buffer without<br />

detergent and twice with kinase buffer before the addition of 1 μCi of 32 P-γATP as above.<br />

73


2.4.7 Coupling of proteins and antibodies to Sepharose<br />

Proteins and antibodies to be coupled to Sepharose were resuspended in PBS containing 0.5<br />

M NaCl (coupling buffer). Freeze-dried cyanogen bromide-activated Sepharose beads<br />

(Amersham Bioscience) were swelled using 1 mM HCl and washed for 15 min with 1 mM<br />

HCl. About 5 mL of coupling solution containing 25 to 50 mg of protein was mixed with<br />

each gram of freeze-dried powder. <strong>The</strong> mixture was rotated end-over-end overnight at 4°C.<br />

<strong>The</strong> beads were washed with at least 5 volumes of coupling buffer to remove excess ligand.<br />

Remaining active groups were blocked with 1 M ethanolamine for 2 hrs at 4°C. <strong>The</strong> beads<br />

were then washed with three cycles of 0.1 M acetate buffer, pH 4.0 containing 0.5 M NaCl<br />

followed by 0.1 M Tris-HCl, pH 8 containing 0.5 M NaCl. <strong>The</strong> beads were further washed<br />

with 6 x 1 mL of PBS, and stored in PBS containing 0.02% NaH3 at 4°C until needed.<br />

2.4.8 Immunoprecipitation / pull-down assay<br />

Cells were lysed at a concentration of 1 x 10 8 cells/mL in cell lysis buffer for 10 min on ice.<br />

<strong>The</strong> cell lysate was separated from insoluble material by centrifugation in a bench-top<br />

centrifuge at 12,000 × g for 15 min at 4°C. One millilitre of cell lysate was precleared thrice<br />

by mixing with 50 μl of Sepharose beads for 30 min at 4°C. For immunoprecipitation, one<br />

mL of precleared cell lysate was incubated with 50 μL of antibody-Sepharose for 2 hrs at 4°C<br />

with gentle inversion on a rotating platform. For the pull-down assay, one mL of precleared<br />

cell lysate was incubated with 50 μL of peptide-Sepharose overnight at 4°C with gentle<br />

inversion. <strong>The</strong> Sepharose beads were recovered by centrifugation at 1,000 × g for 5 min at<br />

4°C. <strong>The</strong> beads were washed 6 times with 1 mL of PBS for 5 min at 4°C.<br />

An alternative procedure was employed when detecting whether immunoprecipitated proteins<br />

were phosphorylated. Cells were lysed in cell lysis buffer containing phosphatase inhibitors,<br />

and the lysate was denatured immediately in the presence of SDS-PAGE loading dye without<br />

bromophenol blue and SDS by heating at 95°C for 5 min. <strong>The</strong> denatured cell lysate was<br />

incubated overnight at 4°C with antibody-Sepharose beads. <strong>The</strong> beads were washed 6 times<br />

with 1 mL of PBS for 5 min at 4°C.<br />

74


2.5. Cell biology<br />

2.5.1 Cell culture<br />

Suspension cells<br />

Suspension cells: TK-1, H9, and MTC-1 cells were subcultured every 2 to 3 days at<br />

approximately 2 x 10 5 cells/mL with fresh cell culture medium, and grown at 37°C in a<br />

humidified atmosphere of 5% CO2.<br />

Adherent cells<br />

<strong>The</strong> adherent cells HEK-293T and CHO-K1 were cultured in DMEM cell culture medium at<br />

37°C in a humidified atmosphere of 10% CO2. To subculture, the media was removed and<br />

discarded, and the adherent cells were washed once with PBS. Trypsin-EDTA solution<br />

(GibcoBRL) was added and the cells were observed by microscopy until the cell layer was<br />

dispersed (5 to 10 min). Detached cells were collected and centrifuged at 1,000 × g for 5 min<br />

at RT, and washed once with cell culture media. <strong>The</strong> cells were re-cultured in complete<br />

medium at approximately 2 x 10 5 cells/mL, and the media renewed every 2-3 days.<br />

2.5.2 Production of recombinant proteins<br />

Spodoptera frugiperda Sf9 insect cells were grown and adapted to Sf900II SFM. <strong>The</strong>y were<br />

then infected with a baculovirus engineered to produce a particular Fc-tagged recombinant<br />

protein, and cultured at 27°C for 5 days. <strong>The</strong> culture was centrifuged at 1,000 × g for 5 min at<br />

RT, and the supernatant stored at -20°C in the presence of 0.02% NaH3 and 1.6 µg/mL<br />

soybean trypsin inhibitor. <strong>The</strong> supernatant was thrice passed through a protein A-Sepharose<br />

column at 4°C to isolate the recombinant Fc-tagged proteins. <strong>The</strong> column was washed<br />

extensively with PBS and the recombinant proteins were eluted with 0.1 M citric acid pH 1.5,<br />

followed by neutralisation with 0.2 M Tris-Cl (pH 8). <strong>The</strong> elutant was concentrated in a<br />

dialysis membrane covered with aquacide II (Calbiochem). <strong>The</strong> resulting solution was<br />

dialysed against PBS, and stored in the presence of 0.02% NaH3 at -80˚C at a concentration<br />

of 1 mg/mL until needed.<br />

75


2.5.3 Spleen cell isolation<br />

Mice were sacrificed and the spleens removed and temporarily placed on ice in DMEM media<br />

containing 10% FCS and PSG. <strong>The</strong> spleens were mashed through a fine gauze with a sterile<br />

syringe plunger, and clusters of cells were broken up by passing the solution 4 to 5 times<br />

through an 18G needle. <strong>The</strong> cells were centrifuged for 5 min at 1,000 × g at 4°C, and then<br />

resuspended in 10 to 20 mL of PBS. <strong>The</strong> cell suspension was layered gently onto 15 to 20 mL<br />

of Ficoll-Hypaque solution and centrifuged at 2,000 × g for 30 min at 20°C with no brake.<br />

<strong>The</strong> buffy coat layer was transferred into a 50 mL tube, topped up with PBS and centrifuged<br />

at 2,000 × g for 5 min. <strong>The</strong> cells were washed thrice with PBS, and cultured at 1 x 10 6<br />

cells/mL in DMEM media containing 10% FCS, PSG, and 1 µg/mL of concanavalin A<br />

(Sigma-Aldrich).<br />

2.5.4 Cell adhesion assay<br />

Sixteen well chamber slides (178599, Nunc) for assays using unlabelled cells, and 96- well<br />

plates (F96 Maxisorp 439454, Nunc) for assays using fluorescein-labelled cells, were coated<br />

overnight with 70 µL (per well) of recombinant MAdCAM-1-Fc (10 µg/mL) at 4°C. <strong>The</strong><br />

wells were washed thrice with PBS, and blocked at RT with 100% FBS for 2 hrs. <strong>The</strong>y were<br />

then washed once with PBS and twice with HBSS binding buffer immediately prior to the<br />

addition of cells.<br />

Cell assays using 96-well plates: Cells were harvested by centrifugation at 1,000 × g for 5<br />

min. <strong>The</strong>y were incubated with the cell tracker CMFDA (Invitrogen) in serum-free media at<br />

37°C for 30 min. All cells were washed thrice with PBS and incubated with 5 mM EDTA in<br />

PBS for 20 min. <strong>The</strong>y were then washed once with PBS and twice with HBSS binding buffer.<br />

<strong>The</strong> cells were placed into aliquots of 5 x 10 5 cells per 300 µL of HBSS binding buffer. <strong>The</strong>y<br />

were activated by the addition of the activators 2 mM Mn 2+ /Ca 2+ , AlF4 - , or 50 ng/mL of PMA<br />

in HBSS binding buffer for 5 to 10 min at RT. Controls cells were left non-activated or<br />

incubated with the non-activating cation control 2 mM Mg 2+ /Ca 2+ in HBSS binding buffer for<br />

5 to 10 min at RT. Each aliquot of cells was then transferred into the ligand<strong>–</strong>coated well of a<br />

96-well plate, and allowed to adhere to the ligand at RT for 30 min. <strong>The</strong> solution in the wells<br />

was gently removed with a multi-channel pipette. Each well was gently filled with PBS and<br />

sealed. <strong>The</strong> plate was centrifuged in an inverted position in a microplate carrier at 500 × g for<br />

5 min, and the solution discarded by pipetting. This step was performed 3 times. <strong>The</strong><br />

76


fluorescence of the bound cells was measured using a fluorescence microplate reader<br />

(VICTOR 1420 multilabel counter from Wallac) at an excitation wavelength of 485 nm and<br />

an absorbance wavelength of 535 nm.<br />

Cell assays using 16-well chamber slides: Cells were harvested and activated as above, but<br />

were not labelled with CMFDA. <strong>The</strong>y were allowed to adhere to the ligand-coated slides at<br />

RT for 30 min. <strong>The</strong> chambers were removed and the slide dipped thrice into PBS to remove<br />

unbound cells. Bounds cells were fixed in a solution of 2% glutaraldehyde in PBS at 4°C<br />

overnight. Fixed cells were dipped twice into PBS, twice into water, and stained with 0.1%<br />

methylene blue for 5 min. Excess dye was removed by dipping cells twice into water, and<br />

then once into each of 50%, 70% and 100% ethanol. <strong>The</strong> dyed cells were counted using a<br />

microplate reader (µQuant, Bio-Tek instruments) at a wavelength of 595 nm.<br />

Effect of cell treatments on cell adhesion<br />

Treatment with inhibitory peptides:<br />

For peptide blocking assays, aliquoted cells were incubated with peptides for 30 min in HBSS<br />

binding buffer prior to the addition of activators, and then transferred into the wells.<br />

Treatment with chemical inhibitors:<br />

For testing the effects of chemical inhibitors, aliquoted cells were incubated with the<br />

inhibitors for 3 hrs in HBSS binding buffer prior to the addition of activators, and then<br />

transferred into the wells.<br />

Heat shock treatment:<br />

For heat shock treatment, cells were subjected to increased temperatures ranging from 37°C<br />

to 41°C for up to 1 hr, and then allowed to recover overnight at 37°C before use.<br />

Serum starvation:<br />

For serum starvation, cells were cultured at 1 x 10 6 cells/mL in media containing decreasing<br />

amounts of serum ranging from 10% to 2.5%, and grown overnight at 37°C before use.<br />

77


2.5.5 Enzyme-linked soluble E-cadherin-Fc-mediated adhesion assay.<br />

MTC-1 cells were grown in RPMI containing 10% FCS, PSG (penicillin, streptomycin,<br />

glutamine), and 5 ng/ml of TGFβ1 to induce the expression of αEβ7. Cells were pelleted by<br />

centrifugation at 1000 × g for 5 min, washed thrice with PBS then incubated in PBS<br />

containing 5 mM EDTA for 20 min. <strong>The</strong> cells were again washed thrice with PBS, then<br />

incubated with goat anti-human IgG (diluted 1:200) in PBS for 30 min at RT as a blocking<br />

step to prevent cell binding to the Fc region of E-cadherin-Fc. <strong>The</strong> cells were washed thrice<br />

with PBS, once with HBSS, and then resuspended in HBSS containing cell activators (final<br />

volume 200 µL), and incubated for 5 min. Recombinant soluble E-cadherin-Fc/goat-anti-<br />

human-Fc-HRP complex (1 μg of E-cadherin-Fc mixed with 1:1000 dilution of goat-anti-<br />

human-Fc-HRP in a 100 µL volume and incubated at RT for 30 min) was added to the cells,<br />

followed by incubation at RT for 30 min. <strong>The</strong> cells were washed thrice with PBS, and ABTS<br />

substrate [2,2'-azino-bis (3-ethylbenzthiazoline-6-sulfonic acid), 100 µL per well] was added.<br />

ABTS produces a water soluble green colour upon reaction with HRP. <strong>The</strong> colorimetric<br />

enzyme reaction was left to develop for 10 to 20 min, and the amount of dye formed was<br />

measured on a Bio-Tek µQuant TM Microplate Reader at 405 nm.<br />

2.5.6 Cell transfection<br />

HEK-293T cells were cultured in 6-well plates in DMEM + 10% FCS until they were 80 to<br />

90% confluent, and then transfected with Lipofectamine 2000 (Invitrogen). To prepare the<br />

transfection solution, 4 µg of DNA and 10 µL of Lipofectamine 2000 were each separately<br />

mixed with 250 µL of OptiMEM media (Invitrogen). <strong>The</strong> DNA and Lipofectamine solutions<br />

were incubated together at RT for 20 min, and then added dropwise to a well. <strong>The</strong> cells in the<br />

well were mixed into the solution by agitating the plates, which were then incubated in a<br />

humidified atmosphere of 10% CO2 at 37°C for 24 to 48 hrs.<br />

2.5.7 Immunofluorescence staining<br />

Cells were fixed with 4% paraformaldehyde for 10 min on ice, and permeabilised with PBS<br />

containing 0.1% Tween-20. <strong>The</strong>y were incubated with primary antibody for 60 min at RT or<br />

overnight at 4°C, followed by washing thrice with PBS for 5 min at RT. Fluorescentconjugated<br />

secondary antibody was added followed by incubation for 1 hr at RT, and then<br />

washing with PBS for 5 min at RT. <strong>The</strong> cells were mounted with mounting media containing<br />

DAPI (H-1200, Vector Laboratories). <strong>The</strong>y were covered with a coverslip, with the edges<br />

78


eing sealed with nail polish. <strong>The</strong> slides were examined by microscopy using a Nikon E600<br />

microscope under epi-fluorescence, and further examined by confocal microscopy using a<br />

Leica SP2 confocal microscope (Biomedical Imaging Research Unit, <strong>The</strong> <strong>University</strong> of<br />

Auckland).<br />

2.6. Phylogenetic analysis<br />

Phylogenetic analysis was performed using the Ensembl project website<br />

(http://www.ensembl.org). <strong>The</strong> site was searched for orthologues of the human integrin β7<br />

gene (itgb7). <strong>The</strong> intracellular domains of the orthologues were compared by alignment of the<br />

deduced protein sequences.<br />

In addition, a blast search using the National Center for Biotechnology Information website<br />

(http://www.ncbi.nlm.nih.gov/) was performed using the cytoplasmic domain sequence of the<br />

human integrin β7 subunit.<br />

2.7. Statistical analysis<br />

Results were expressed as mean values ± standard deviation (SD), and a Student’s t-test was<br />

used for evaluating statistical significance for comparison with control groups. A value less<br />

than 0.05 (p < 0.05) indicated statistical significance. Each experiment was repeated at least<br />

twice.<br />

79


Chapter 3. Results<br />

3.1. Expression and cell adhesion properties of β7-integrins<br />

To investigate the regulation of cell adhesion by the β7 integrins α4β7 and αEβ7, cells which<br />

express high levels of the β7 integrins were required. Cultured cells which were known to<br />

express high levels of α4β7 and αEβ7 were chosen for the study. <strong>The</strong> mouse thymic<br />

lymphoma T cell line TK-1 expresses high levels of α4β7 (Ruegg et al. 1992). Importantly,<br />

TK-1 cells do not express the integrin β1 subunit; hence they represent an ideal model cell<br />

system in which the function of α4β7 can be examined in the absence of α4β1. <strong>The</strong> murine T<br />

cell hybridoma MTC-1 expresses αEβ7 when grown in the presence of TGF-β (Roberts et al.<br />

1993). <strong>The</strong>refore the initial aim of the thesis was to characterize the expression of the β7<br />

integrins on TK-1 and MTC-1 cells.<br />

3.1.1 Confirmation of the expression of α4β7 and αEβ7 on TK-1 and MTC-1 cells<br />

TGF-β1-stimulated MTC-1 cells express the integrin αEβ7<br />

<strong>The</strong> initial experiment was to confirm the expression of αEβ7 by TGF-β-stimulated MTC-1<br />

cells by using immunoblotting. MTC-1 cells were grown in the presence of 5 ng/mL of TGF-<br />

β1 for three days. <strong>The</strong> cells were harvested, lysed and the total cellular lysate was resolved by<br />

SDS-PAGE. <strong>The</strong> proteins were transferred to PVDF membrane and immunoblotted with<br />

rabbit antibodies against the integrin αE and β7 subunits (Figure 3.1). <strong>The</strong> major bands<br />

detected by the anti-β7 and anti-αE antibodies were 120 and 100 kDa in size, respectively,<br />

confirming that MTC-1 cells express both the αE and β7 subunits. Minor bands that appeared<br />

beneath the αE and β7 subunits were non-specific.<br />

80


Figure 3.1 Immunoblot analysis of an MTC-1 cell lysate with antibodies against the αE and β7 subunits<br />

MTC-1 cells were grown in the presence of 5 ng/mL of TGF-β1 for three days. <strong>The</strong> cells were lysed in cell lysis<br />

buffer at 1 x 10 8 cells/mL. <strong>The</strong> cell lysate (1 x 10 6 cells/lane) was resolved on an 8% polyacrylamide SDS-gel,<br />

and proteins transferred onto a PVDF membrane. <strong>The</strong> PVDF membrane was immunoblotted with rabbit<br />

antibodies against the integrin αE subunit (1:1000 dilution) and the β7 subunit (1:400 dilution) as indicated.<br />

Indicated in the right-hand margin are the positions of the αE and β7 integrin subunits. <strong>The</strong> sizes of molecular<br />

weight markers are shown in the left-hand margin. This experiment was repeated twice.<br />

To further confirm the expression of αEβ7 on MTC-1 cells, MTC-1 cells were stained with a<br />

fluorescently-conjugated antibody against the β7 subunit, and staining visualized by<br />

fluorescence microscopy. MTC-1 cells grown in TGF-β were fixed and stained with the rat<br />

mAb antibody Fib504 against the β7 integrin subunit. <strong>The</strong> Fib504 antibody was detected with<br />

an Alexa Fluor (AF) 488-conjugated secondary antibody (green). <strong>The</strong> green staining indicates<br />

the presence of the β7 integrin on MTC-1 cells (marked by yellow arrows), as shown in<br />

Figure 3.2B. DAPI staining of the nuclei (blue) reveals all cells in the field of view. <strong>The</strong><br />

Fib504 and DAPI images were merged, revealing that some MTC-1 cells have little or no<br />

detectable β7 integrin subunit expression, as indicated by the red arrows. Analysis of 5 fields<br />

suggested that approximately 50% of TGF-β1-stimulated MTC-1 cells expressed αEβ7. As a<br />

control, MTC-1 cells were stained with an isotype control rat antibody (rat-IgG; Figure<br />

3.2A). <strong>The</strong> control antibody did not stain the cells green, indicating the absence of nonspecific<br />

antibody binding or autofluorescence.<br />

81


Figure 3.2 MTC-1 cells immunostained with an anti-β7 antibody.<br />

MTC-1 cells grown in the presence of 5 ng/mL of TGF-β1 were fixed with 4% paraformaldehyde (PFA) and<br />

immunostained with either; (A) a control rat-IgG antibody (diluted 1:100), or (B) the rat Fib504 mAb against the<br />

β7 integrin subunit (diluted 1:100). Antibody reactivity was detected with an Alexa Fluor (AF) 488-conjugated<br />

secondary antibody (diluted 1:250), and visualised by epi-fluorescence microscopy. Pictures shown are:<br />

antibody staining (left), where the green colour indicates positive antibody staining; DAPIi staining of the nuclei<br />

(middle); and a merged image of antibody and DAPI staining (right), which reveals the antibody staining with<br />

respect to the nucleus. <strong>The</strong> yellow arrows indicate cells expressing the integrin β7 subunit, and the red arrows<br />

indicate cells that do not express the β7 subunit. <strong>The</strong> bar indicates the length of 20 µm. This experiment was<br />

repeated three times.<br />

82


TK-1 cells express the integrin α4β7<br />

TK-1 cells were immunostained with the DAKT32 antibody which recognises the murine<br />

α4β7 complex, both to confirm that TK-1 cells express α4β7, and to test the activity of the<br />

DAKT32 antibody. TK-1 cells were fixed and stained with either an isotype control antibody<br />

(Figure 3.3, images 1 and 2) or the DAKT32 antibody (Figure 3.3, images 3 and 4), and<br />

stained with DAPI to reveal antibody staining with respect to the nuclei. <strong>The</strong> antibodies were<br />

detected with an AF-488-conjugated secondary antibody, and visualised by fluorescence<br />

microscopy. <strong>The</strong> green staining indicates that almost all TK-1 cells express α4β7 (Figure 3.3,<br />

images 3 and 4), whereas no cells were stained green by the control antibody (Figure 3.3,<br />

images 1 and 2). This result confirmed that TK-1 cells express α4β7, and that the DATK32<br />

antibody was useful for the subsequent studies.<br />

Figure 3.3 TK-1 cells stained with the DATK32 antibody against the α4β7 complex.<br />

TK-1 cells in suspension were fixed in 4% PFA and immunostained with; a rat IgG isotype control antibody<br />

(diluted 1:100; images 1 and 2), or the DAKT32 mAb against the integrin α4β7 complex (diluted 1:100; images<br />

3 and 4). Cell nuclei were stained blue with DAPI. Antibody reactivity was detected with an AF-488-conjugated<br />

secondary antibody (diluted 1:250), and visualised by epi-fluorescence microscopy. <strong>The</strong> pictures shown are<br />

merged images of the antibody staining in green, and the DAPI staining of the nuclei in blue. Images 1 and 3<br />

were taken at 10 x magnification, whereas images 2 and 4 were taken at 40x magnification. <strong>The</strong> bar indicates the<br />

length of 20 µm. This experiment was repeated twice.<br />

83


3.1.2 Production of recombinant forms of E-cadherin and MAdCAM-1<br />

Recombinant MAdCAM-1 and E-cadherin proteins, which are ligands for integrin α4β7 and<br />

αEβ7 respectively, were produced to study β7 integrin-mediated cell adhesion. Expression<br />

systems to produce both proteins had already been established in our laboratory.<br />

Complementary DNAs encoding MAdCAM-1 and E-cadherin without the membrane<br />

domains had been fused to sequences encoding the CH2/CH3 regions of the human IgG<br />

antibody heavy chain Fc domain, and inserted into the pVL1393 baculovirus vector (Yang et<br />

al. 1995; Berg et al. 1999b). <strong>The</strong> baculovirus expression vectors were able to produce<br />

recombinant MAdCAM-1-Fc, and E-cadherin-Fc proteins containing C-terminal Fc tags<br />

which facilitate purification of the recombinant proteins.<br />

<strong>The</strong> baculovirus expression systems were used to produce the recombinant MAdCAM-1-Fc,<br />

and E-cadherin-Fc proteins for the current study of β7 integrin-mediated cell adhesion.<br />

Briefly, Sf9 insect cells adapted for growth in serum-free medium were infected with the<br />

baculoviral vectors containing the cDNAs encoding MAdCAM-1-Fc and E-cadherin-Fc. <strong>The</strong><br />

recombinant E-cadherin-Fc and MAdCAM-1-Fc proteins produced and secreted into the cell<br />

media were purified by passing the cell culture supernatant through a protein-G Sepharose<br />

column. <strong>The</strong> recombinant proteins were eluted at low pH, and were resolved by SDS-PAGE.<br />

Purified E-cadherin was resolved as two bands of approximately 115 and 130 kDa (Figure<br />

3.4A), which represent the processed and unprocessed forms of E-cadherin with or without<br />

the 18 kDa propeptide, respectively (Berg et al. 1999b). Purified MAdCAM-1 was resolved<br />

as a band of approximately 75 kDa, which corresponds to the correct size of MAdCAM-1-Fc<br />

(Figure 3.4B). <strong>The</strong> identity of the protein bands were confirmed by Western blotting with<br />

antibodies against E-cadherin and MAdCAM-1 (data not shown).<br />

84


Figure 3.4 Analysis of purified recombinant E-cadherin-Fc and MAdCAM-1-Fc by SDS-PAGE.<br />

Recombinant E-cadherin-Fc and MAdCAM-1-Fc were produced by a baculovirus expression system. One<br />

microgram of each of the affinity-purified recombinant E-cadherin-Fc (A) and MAdCAM-1-Fc (B) proteins was<br />

resolved on 8% and 10% polyacrylamide SDS-gels, respectively, and stained with Coomassie blue. Purified Ecadherin<br />

yielded a major processed form and a minor unprocessed form. <strong>The</strong> sizes of molecular weight markers<br />

are shown in the left-hand margin.<br />

3.1.3 Comparison of activators of β7 integrin-mediated cell adhesion<br />

Cell adhesion assays provide a useful tool for investigating the effects of chemical inhibitors<br />

or small inhibitory peptides on pathways that control integrin-mediated cell adhesion. Cells<br />

expressing the β7 integrins were activated with a panel of different activators of β7 integrinmediated<br />

cell adhesion, and bound to plates coated with their respective recombinant ligands.<br />

<strong>The</strong> integrin activation agents included the pan-activator Mn 2+ , the protein kinase C (PKC)<br />

activator phorbol-12-myristate 13-acetate (PMA), and the G-protein activator AlF4 - .<br />

85


Activation of α4β7-mediated T cell adhesion to MAdCAM-1<br />

TK-1 cells cultured in normal culture media were harvested and labelled with the fluorescent<br />

dye chloromethyl fluorescein diaacetate (CMFDA). <strong>The</strong> cells were left non-activated, or<br />

treated with 2 mM Mn 2+ , AlF4 - or 50 ng/mL PMA, and allowed to bind to MAdCAM-1-<br />

coated plates. TK-1 cells activated with Mn 2+ , AlF4 - and PMA displayed 2 to 3-fold increased<br />

cell binding to immobilised MAdCAM-1-Fc compared to non-activated cells (Figure 3.5).<br />

Mn 2+ appeared to be the best activator of cell adhesion.<br />

Fluorescence (485/535nm)<br />

100000<br />

80000<br />

60000<br />

40000<br />

20000<br />

0<br />

No activator Mn2+ AlF4- PMA<br />

Activators<br />

Figure 3.5 Activation of TK-1 cell adhesion to MAdCAM-1<br />

TK-1 cells labelled with CMFDA were left non-activated, or activated with Mn 2+ , AlF4-, and PMA. Cells were<br />

allowed to bind (5 x 10 5 cells/well of a 96 well plate) to MAdCAM-1-Fc coated wells at RT for 30 min. Bound<br />

cells were detected using a fluorescence microplate reader with excitation and emission wavelengths of 485 and<br />

535 nm, respectively. <strong>The</strong> increase in fluorescence detected correlates with increased cell adhesion. Data<br />

represent the mean and SD of three separate wells. <strong>The</strong> experiment was performed in triplicate.<br />

Activation of αEβ7-mediated T cell binding to E-cadherin<br />

<strong>The</strong> expression of αEβ7 on MTC-1 cells was variable, as were the activation and subsequent<br />

cell adhesion of MTC-1 cells to E-cadherin-coated plates (data not shown). <strong>The</strong>refore the<br />

MTC-1 cell adhesion assay was modified to bind E-cadherin in solution rather than on coated<br />

surfaces, which produced more consistent data. Briefly, MTC-1 cells were harvested and<br />

either left non-activated or activated in the presence of Mn 2+ , AlF4 - or PMA (5 - 10 mins).<br />

Cells were mixed with recombinant E-cadherin-Fc which had previously been complexed<br />

with a goat-anti-human-Fc antibody conjugated with HRP (E-cadherin-antibody complex).<br />

Cells that bound to the E-cadherin-antibody complex were detected via HRP conversion of<br />

86


the peroxidise substrate ABTS. <strong>The</strong> colormetric products were detected using a microplate<br />

reader. Mn 2+ , AlF4 - and PMA each increased the binding of the E-cadherin-antibody complex<br />

to MTC-1 cells by approximately 5-6 fold compared to non-activated control cells (Figure<br />

3.6). A goat-anti-human-HRP antibody alone did not stain activated MTC-1 cells (data not<br />

shown).<br />

Absorbance (A405nm)<br />

0.2<br />

0.16<br />

0.12<br />

0.08<br />

0.04<br />

0<br />

No activator Mn2+ AlF4- PMA<br />

Activators<br />

Figure 3.6 Activation of MTC-1 cell binding to E-cadherin<br />

MTC-1 cells were cultured in complete growth media in the presence of 5 ng/mL of TGF-β1. <strong>The</strong> cells were<br />

harvested and left non-activated or activated with, Mn 2+ , AlF4-, or PMA. Recombinant E-cadherin-Fc (1 µg)<br />

was mixed with 100 μl of goat-anti-human-Fc antibody conjugated with HRP (diluted 1:1000) to form an “Ecadherin-antibody<br />

complex”. <strong>The</strong> cells were allowed to bind to the E-cadherin-antibody complex at RT for 30<br />

min. Unbound E-cadherin-antibody complexes were removed by washing the cells thrice with PBS. Cell binding<br />

to the E-cadherin-antibody complex was measured by the addition of the peroxidase substrate ABTS, and<br />

analysis in a microplate reader at a wavelength of 405 nm. <strong>The</strong> increase in absorbance detected correlated with<br />

binding of the E-cadherin-antibody complex. Data represent the mean and SD of three wells. <strong>The</strong> experiment<br />

was performed in triplicate.<br />

Antibodies against the αE and β7 subunits block the binding of E-cadherin to MTC-1<br />

cells<br />

Antibodies against the αE and β7 subunits were introduced into the adhesion assay to confirm<br />

that αEβ7 was responsible for mediating the binding of MTC-1 cells to the soluble E-<br />

cadherin-antibody complex. MTC-1 cells were incubated with anti-α4 and anti-αE (M290<br />

mAb) subunit antibodies at RT for 30 min prior to addition of the cell activators. AlF4 - was<br />

chosen to activate MTC-1 cells as it stimulated the binding of the E-cadherin-antibody<br />

87


complex the greatest (Figure 3.6). As controls, antibody was omitted (no Ab) and cells were<br />

incubated with antibodies against the α4 subunit, and MAdCAM-1, which are not expressed<br />

by MTC-1 cells. Equivalent levels of binding of the E-cadherin-antibody complex to MTC-1<br />

cells were obtained with each of the latter controls (Figure 3.7). In contrast, binding of the E-<br />

cadherin-antibody complex was significantly reduced by antibodies against the αE subunit (p<br />

= 7.08 × 10 -4 ), β7 subunit (p = 1.09 × 10 -5 ), and E-cadherin (p = 1.61 × 10 -5 ; Figure 3.7).<br />

Note that the adhesion assays were preformed in the presence of the antibodies, which<br />

explains why the anti-cadherin antibody was inhibitory. <strong>The</strong> results indicate that MTC-1 cell<br />

adhesion to the E-cadherin-antibody complex is specifically mediated by αEβ7.<br />

Figure 3.7 Binding of E-cadherin to MTC-1 cells is specifically mediated by αEβ7.<br />

MTC-1 cells grown in the presence of 5 ng/mL of TGF-β1 were incubated in the absence of antibody (no Ab), or<br />

with antibodies against the α4 subunit (anti-α4), the αE- subunit (anti-αE), β7 subunit (anti-β7), E-cadherin<br />

(anti-E-cadherin) and MAdCAM-1 (anti-MAdCAM-1); and then treated with AlF4 - . <strong>The</strong> soluble E-cadherin-<br />

antibody complex (E-cadherin-Fc-anti-human-IgG-HRP) was added and cells incubated at RT for 30 min.<br />

Unbound E-cadherin-antibody complexes were removed by washing the cells thrice with PBS. Cell binding to<br />

the E-cadherin-antibody complex was measured by the addition of the peroxidase-substrate ABTS, and analysis<br />

in a microplate reader at a wavelength of 405 nm. <strong>The</strong> increase in absorbance detected correlated with binding of<br />

the E-cadherin-antibody complex. Data shown represent the mean and SD of three wells. *** denotes p-value <<br />

0.001 compared to cell binding in the absence of antibody. <strong>The</strong> experiments were performed in triplicate.<br />

88


3.2. <strong>The</strong> effect of pathway inhibitors on β7 integrin mediated cell adhesion<br />

Integrin activation is the result of a cascade of intracellular signalling events which propagate<br />

from the cytoplasm to the plasma membrane (inside-out), or vice-versa are initiated from<br />

extracellular signals (outside-in), as introduced in <strong>Section</strong> 1.3. A panel of inhibitors of<br />

various cellular kinases and cell signalling pathways were tested for their ability to prevent β7<br />

integrin-mediated cell adhesion in order to identify intracellular pathways that regulate β7<br />

integrin signalling. Genistein, a general protein tyrosine kinase (PTK) inhibitor (Akiyama et<br />

al. 1987), was previously found to prevent α4β7-mediated adhesion of TK-1 cells to<br />

MAdCAM-1 (Zhang et al. 1999). Here, genistein was tested for its affects on cell adhesion,<br />

together with a panel of other kinase inhibitors including inhibitors of the mitogen-activated<br />

protein kinase (MAPK) pathway, Jun N-terminal kinase (JNK), epidermal growth factor<br />

receptor (EGFR) tyrosine kinase, the Src family of tyrosine kinases (SFKs), and myosin light<br />

chain kinase (MLCK).<br />

Cells were deactivated by EDTA metal chelation, pretreated with chemical inhibitors at 37°C<br />

for 3 hrs in HBSS buffer, and either left unactivated (treated only with diluents used to<br />

dissolve the activators) or activated with Mn 2+ or PMA. All cells were checked for viability<br />

by trypan blue exclusion prior to assessing cell adhesion. Cells were tested for their ability to<br />

bind to MAdCAM-1-coated plates. Disruption of β7 integrin-mediated cell adhesion by a<br />

chemical inhibitor would implicate a particular pathway/kinase in β7 integrin signalling.<br />

3.2.1 Genistein, a PTK inhibitor, inhibits the binding of TK-1 cells to MAdCAM-1<br />

Genistein is a general inhibitor of PTKs which acts as a competitive inhibitor of ATP<br />

(Akiyama et al. 1987). EGFR tyrosine kinase induces several signal transduction cascades<br />

including the MAPK pathway, Akt pathway, and JNK phosphorylation, which lead to DNA<br />

synthesis and cell proliferation (Oda et al. 2005). TK-1 cells were bound to MAdCAM-1coated<br />

slides in the presence or absence of genistein at a concentration of 10 µM. Genistein<br />

significantly (p = 5x10 -4 ) prevented the binding of TK-1 cells to MAdCAM-1 (Figure 3.8).<br />

Cell binding was inhibited by 60%, being almost reduced to levels seen with unactivated TK-<br />

1 cells. Genistein at 10 µM was subsequently used as positive control in the following celladhesion<br />

assays.<br />

89


Absorbance (A495nm)<br />

Figure 3.8 Effect of genistein on TK-1 cell adhesion to MAdCAM-1<br />

TK-1 cells were resuspended in HBSS with or without 10 µM genistein, and incubated at 37˚C for 3 hrs. <strong>The</strong><br />

cells were then either left non-activated or activated with Mn 2+ , as indicated. Cells were adhered to MAdCAM-<br />

1-coated plates at RT for 30 min. <strong>The</strong> bound cells were fixed with glutaraldehyde, stained with methylene blue<br />

and quantified using a microplate reader at a wavelength of 495 nm. Data shown represent the mean and SD of<br />

three wells. *** denotes a p-value of < 0.001 compared to cell binding of activated control cells not treated with<br />

genistein. <strong>The</strong> experiments were performed in triplicate.<br />

3.2.2 Involvement the MAP kinase kinase (MEK) pathway in TK-1 cell binding to<br />

MAdCAM-1<br />

0.250<br />

0.200<br />

0.150<br />

0.100<br />

0.050<br />

0.000<br />

Non-activated Activated Activated +<br />

genistein 10 μM<br />

Activation/Inhibitor<br />

Many integrin signalling pathways involve the activation of the mitogen-activated protein<br />

kinase (MAPK) cascade, ras → raf → MEK → ERK, which regulates cell growth through<br />

cell cycle progression (Giancotti et al. 1999). <strong>The</strong>refore the involvement of the MEK pathway<br />

in β7 integrin signalling was investigated.<br />

PD98059<br />

PD98059 is an inhibitor of MAP kinase kinase (Dudley et al. 1995). It inhibits the activation<br />

of MAP kinase and the subsequent phosphorylation of MAP kinase substrates. Treatment of<br />

PMA- and Mn 2+ -activated TK-1 cells with 1, 10 and 100 µM of PD98059 did not<br />

significantly affect α4β7-mediated cell adhesion to MAdCAM-1 (Figure 3.9A and B). All<br />

cells were viable as analyzed by trypan blue exclusion. In contrast, genistein inhibited both<br />

PMA and Mn 2+ -induced adhesion.<br />

90<br />

***


Figure 3.9 Effect of PD98059 on TK-1 cell adhesion to MAdCAM-1<br />

TK-1 cells were incubated with 0, 1, 10 or 100 µM PD98059 or 10 µM genistein at 37°C for 3 hrs. <strong>The</strong> cells<br />

were activated with either PMA (A) or Mn 2+ (B), and adhered to MAdCAM-1-Fc-coated plates at RT for 30<br />

min. PMA-activated TK-1 cells (A) were labelled with CMFDA prior to addition of the inhibitors, and cell<br />

binding was analysed using a fluorescence microplate reader with excitation and emission wavelengths of 485<br />

nm and 535 nm respectively. Mn 2+ -activated TK-1 cells (B) were fixed with glutaraldehyde, stained with<br />

methylene blue, and cell binding was analysed using a microplate reader at a wavelength of 495 nm. Data<br />

shown represent the mean and SD of three wells. <strong>The</strong> experiments were performed in triplicate.<br />

91


SB203580<br />

SB203580 is an inhibitor of p38 MAP kinase (Cuenda et al. 1995), the most well<br />

characterised member of the MAP kinase family (Cuenda et al. 2007). p38 MAP kinase is<br />

activated in response to inflammatory cytokines, endotoxins and osmotic stress (Cuenda et al.<br />

2007). Treatment of PMA- (Figure 3.10A) and Mn 2+ -activated (Figure 3.10B) TK-1 cells<br />

with SB203580 at a concentration of 1 to 100 µM did not significantly affect α4β7-mediated<br />

cell adhesion to MAdCAM-1. In contrast, genistein inhibited both PMA and Mn 2+ -induced<br />

cell adhesion. All cells were viable as analyzed by trypan blue exclusion.<br />

Figure 3.10 Effect of SB203580 on TK-1 cell adhesion to MAdCAM-1<br />

TK-1 cells were incubated with 0, 1, 10 and 100 µM SB203580 or 10 µM genistein at 37 °C for 3 hrs. <strong>The</strong> cells<br />

were activated with either PMA (A) or Mn 2+ (B), and adhered to MAdCAM-1-Fc-coated plates at RT for 30<br />

min. PMA-activated TK-1 cells (A) were labelled with CMFDA prior to addition of the inhibitors, and cell<br />

binding was analysed using a fluorescence microplate reader with excitation and emission wavelengths of 485<br />

nm and 535 nm respectively. Mn 2+ -activated TK-1 cells (B) were fixed with glutaraldehyde, stained with<br />

methylene blue, and cell binding was analysed using a microplate reader at a wavelength of 495 nm. Data shown<br />

represent the mean and SD of three wells. <strong>The</strong> experiments were performed in triplicate.<br />

92


3.2.3 Jun N-terminal kinase (JNK) pathway<br />

JNK is a serine-directed protein kinase which is involved in the phosphorylation and<br />

activation of the transcription factors c-jun and activating-transcription-factor-2 (ATF2;<br />

Bonny et al. 2001). JNK is activated in response to inflammation, endotoxins, and<br />

environmental stress, and mediates the expression of proinflammatory genes, cell<br />

proliferation and apoptosis (Bonny et al. 2001). <strong>The</strong>re are three isoforms, namely JNK1 and 2<br />

which exhibit broad tissue expression profiles, and JNK3 which is expressed predominantly<br />

in the central nervous system (Roy et al. 2008). Studies of animal models and humans have<br />

revealed increased activation of JNK pathway in inflammatory bowel diseases (reviewed by<br />

Roy et al. 2008). <strong>The</strong>refore, the involvement of the JNK pathway in β7 integrin signalling<br />

was investigated.<br />

JNK inhibitor 1 (JNK-I-1)<br />

JNK inhibitor 1 is a biologically active cell-permeable peptide consisting of a C-terminal<br />

sequence derived from the JNK-binding domain (JBD) and an N-terminal peptide containing<br />

the HIV-TAT48-57 sequence (Bonny et al. 2001). It blocks phosphorylation of the activation<br />

domain of JNK, which prevents the activation of c-jun (Bonny et al. 2001). It has no effect on<br />

the activity of either ERK 1/2 or p38. TK-1 cells were treated with JNK-I-1 at a concentration<br />

ranging from 0.1 to 10 µM or with 10 µM genistein. All cells were viable as analyzed by<br />

trypan blue exclusion. Treatment of PMA-activated TK-1 cells with 10 µM JNK-I-1<br />

significantly (p = 2.43 × 10 -4 ) decreased cell adhesion to MAdCAM-1-Fc by 20%, whereas<br />

lower concentrations of JNK-I-1 had no statistically significant affect (Figure 3.11A).<br />

Treatment of Mn 2+ -activated TK-1 cells with 0.1, 1, and 10 µM JNK-I-1 slightly reduced cell<br />

adhesion by 11% (p = 2.34 × 10 -3 ), 13% (p = 4.13 × 10 -4 ), and 11% (p = 1.75 × 10 -3 ),<br />

respectively (Figure 3.11B). As before, genistein inhibited both PMA and Mn 2+ -induced cell<br />

adhesion.<br />

93


Figure 3.11 Effect of JNK-I-1 on TK-1 cell adhesion to MAdCAM-1<br />

TK-1 cells were incubated with 0, 0.1, 1 and 10 µM JNK-I or 10 µM genistein at 37 °C for 3 hrs. <strong>The</strong> cells were<br />

activated with either PMA (A) or Mn 2+ (B), and adhered to MAdCAM-1-Fc-coated plates at RT for 30 min.<br />

PMA-activated TK-1 cells (A) were labelled with CMFDA prior to addition of the inhibitors, and cell binding<br />

was analysed using a fluorescence microplate reader with excitation and emission wavelengths of 485 nm and<br />

535 nm respectively. Mn 2+ -activated TK-1 cells (B) were fixed with glutaraldehyde, stained with methylene<br />

blue, and cell binding was analysed using a microplate reader at a wavelength of 495 nm. ** denotes p-value of<br />

< 0.01 and *** denotes p-value of < 0.001 compared to cell binding of activated control cells not treated with<br />

inhibitor. Data shown represent the mean and SD of three wells. <strong>The</strong> experiments were performed in triplicate.<br />

94


JNK inhibitor 2 (JNK-I-2)<br />

JNK-I-2 inhibits the activities of JNK-1, JNK-2, and JNK-3 by competing for ATP (Bennett<br />

et al. 2001). It inhibits the phosphorylation of c-jun and blocks cellular expression of the<br />

cytokines IL-2, IFN-γ, and TNF-α, and the enzyme cycloxygenase-2 (COX-2; Bennett et al.<br />

2001). JNK-I-2 also blocks IL-1-induced accumulation of phospho-jun and induction of c-Jun<br />

transcription (Bennett et al. 2001). Treatment of PMA- (Figure 3.12A) and Mn 2+ -activated<br />

(Figure 3.12B) TK-1 cells with JNK-I-2 at a concentration of 1 to 100 µM had no statistically<br />

significant affect on α4β7-mediated cell adhesion to MAdCAM-1. All cells were viable as<br />

assessed by trypan blue exclusion. As before, genistein inhibited both the PMA and Mn 2+ -<br />

induced cell adhesion.<br />

Figure 3.12 Effect of JNK-I-2 on TK-1 cell adhesion to MAdCAM-1<br />

TK-1 cells were incubated with 0, 1, 10 and 100 µM JNK-I-2 or 10 µM of genistein at 37 °C for 3 hrs. <strong>The</strong> cells<br />

were activated with either PMA (A) or Mn 2+ (B), and adhered to MAdCAM-1-Fc-coated plates at RT for 30<br />

min. TK-1 cells were labelled with CMFDA prior to addition of the inhibitors, and cell binding was analysed<br />

using a fluorescence microplate reader with excitation and emission wavelengths of 485 nm and 535 nm<br />

respectively. Data shown represent the mean and standard deviation of three wells. <strong>The</strong> experiments were<br />

performed in triplicate.<br />

95


3.2.4 Epidermal growth factor receptor (EGFR) tyrosine kinase<br />

AG99<br />

AG99 inhibits EGFR tyrosine kinase and EGF-dependent cell proliferation by inhibiting<br />

EGFR phosphorylation, and the activities of ERK1 and ERK2. Treatment of PMA (Figure<br />

3.13A) and Mn 2+ -activated (Figure 3.13B) TK-1 cells with AG99 at a concentration of 1 to<br />

100 µM had no statistically significant affect on α4β7-mediated cell adhesion to MAdCAM-<br />

1. All cells were viable as assessed by trypan blue exclusion. As before, genistein inhibited<br />

both PMA and Mn 2+ -induced cell adhesion.<br />

Figure 3.13 Effect of AG99 on TK-1 cell adhesion to MAdCAM-1<br />

TK-1 cells were incubated with 0, 1, 10 and 100 µM AG99 or 10 µM of genistein at 37 °C for 3 hrs. <strong>The</strong> cells<br />

were activated with either PMA (A) or Mn 2+ (B), and adhered to MAdCAM-1-Fc-coated plates at RT for 30<br />

min. PMA-activated TK-1 cells (A) were labelled with CMFDA prior to addition of the inhibitors, and cell<br />

binding was analysed using a fluorescence microplate reader with excitation and emission wavelengths of 485<br />

nm and 535 nm respectively. Mn 2+ -activated TK-1 cells (B) were fixed with glutaraldehyde, stained with<br />

methylene blue, and cell binding was analysed using a microplate reader at a wavelength of 495 nm. Data shown<br />

represent the mean and SD of three wells. <strong>The</strong> experiments were performed in triplicate.<br />

96


3.2.5 Src family of tyrosine kinases (SFK)<br />

<strong>The</strong> Src family of tyrosine kinases has been implicated in controlling signal transduction<br />

pathways downstream of a variety of cell-surface receptors including integrins (reviewed in<br />

<strong>Section</strong> 1.5.2); therefore it was important to determine their contribution to α4β7-mediated<br />

cell adhesion.<br />

PP2<br />

PP2 is a potent and selective inhibitor of SFKs, which inhibits lck, fyn, hck, and src (Hanke et<br />

al. 1996). It does not significantly affect the activity of EGFR kinase, the non-receptor<br />

tyrosine kinase JAK2, or the protein kinase ZAP-70 (Hanke et al. 1996). PP2 also inhibits the<br />

activation of FAK and its phosphorylation on Y577 (Hanke et al. 1996), and inhibits anti-<br />

CD3-stimulated tyrosine phosphorylation of human T cells (Hanke et al. 1996). Treatment of<br />

PMA-activated (Figure 3.14A) TK-1 cells with 1 µM PP2 significantly (p = 2.9 × 10 -2 )<br />

decreased cell adhesion to MAdCAM-1-Fc by 15%, whereas lower concentrations of PP2 had<br />

no statistically significant affect. Treatment of Mn 2+ -activated TK-1 cells (Figure 3.14B)<br />

with PP2 at 10 to 1000 nM did not significantly affect α4β7-mediated cell adhesion to<br />

MAdCAM-1. All cells were viable as assessed by trypan blue exclusion. As before, genistein<br />

inhibited both PMA and Mn 2+ -induced cell adhesion.<br />

97


Figure 3.14 Effect of PP2 on TK-1 cell adhesion to MAdCAM-1<br />

TK-1 cells were incubated in HBSS with 0, 10, 100 and 1000 nM PP2 or 10 µM of genistein at 37˚C for 3 hrs.<br />

<strong>The</strong> cells were activated with either PMA (A) or Mn 2+ (B), and adhered to MAdCAM-1-Fc-coated plates at RT<br />

for 30 min. TK-1 cells were labelled with CMFDA prior to addition of the inhibitors, and cell binding was<br />

analysed using a fluorescence microplate reader with excitation and emission wavelengths of 485 nm and 535<br />

nm respectively. Data shown represent the mean and SD of three wells. * denotes a p-value < 0.05 compared to<br />

cell binding of activated control cells not treated with inhibitor. <strong>The</strong> experiments were performed in triplicate.<br />

Damnacanthal<br />

Damnacanthal is a selective inhibitor of the tyrosine kinase lck (Faltynek et al. 1995). It<br />

inhibits lck autophosphorylation and phosphorylation of exogenous peptide by lck (Faltynek<br />

et al. 1995). It exhibits 7 to 20-fold greater selectivity for lck over src and fyn (Faltynek et al.<br />

1995). TK-1 cells were treated with damnacanthal at concentrations ranging from 1 to 100<br />

nM. All cells treated with 1 and 10 nM damnacanthal remained viable for the duration of the<br />

experiment as determined by trypan blue exclusion. In contrast, 100 nM damnacanthal<br />

induced approximately 50% cell death, hence the results using this concentration were<br />

excluded from statistical analysis. Treatment of PMA-activated TK-1 cells (Figure 3.15A)<br />

with 1 and 10 nM damnacanthal had no statistically significant affect on α4β7-mediated cell<br />

adhesion to MAdCAM-1. In contrast, treatment of Mn 2+ -activated TK-1 cells (Figure 3.15B)<br />

with 1 and 10 nM damnacanthal significantly reduced cell adhesion to MAdCAM-1 by 3% (p<br />

= 1.47 × 10 -2 ) and 23% (p = 2.65 × 10 -4 ), respectively.<br />

98


Figure 3.15 Effect of damnacanthal on TK-1 cell adhesion to MAdCAM-1<br />

TK-1 cells were incubated with 0, 1, 10 and 100 µM damnacanthal or 10 µM of genistein at 37 °C for 3 hrs. <strong>The</strong><br />

cells were activated with either PMA (A) or Mn 2+ (B), and adhered to MAdCAM-1-Fc-coated plates at RT for<br />

30 min. PMA-activated TK-1 cells (A) were labelled with CMFDA prior to addition of the inhibitors, and cell<br />

binding was analysed using a fluorescence microplate reader with excitation and emission wavelengths of 485<br />

nm and 535 nm respectively. Mn 2+ -activated TK-1 cells (B) were fixed with glutaraldehyde, stained with<br />

methylene blue, and cell binding was analysed using a microplate reader at a wavelength of 495 nm. Data shown<br />

represent the mean and SD of three wells. <strong>The</strong> experiments were performed in triplicate. # denotes 50% cell<br />

death as assessed by trypan blue exclusion. * denotes a p-value of < 0.05 and *** denotes a p-value of < 0.001<br />

compared to cell binding of activated control cells not treated with inhibitor.<br />

Radicicol<br />

Radicicol is a protein tyrosine kinase inhibitor, which inhibits v-src kinase activity<br />

(Chanmugam et al. 1995). It also disrupts K-Ras-activated signalling pathways by selectively<br />

depleting Raf kinase, and inhibiting tyrosine phosphorylation of lyn (Chanmugam et al.<br />

1995). It suppresses the transformation of NIH/3T3 cells by diverse oncogenes such as src,<br />

ras, and mos, in part by blocking the key signal transduction intermediates MAP kinase and<br />

GAP-associated p62 (Soga et al. 1998). Treatment of PMA- (Figure 3.16A) and Mn 2+ -<br />

activated (Figure 3.16B) TK-1 cells with radicicol at a concentration of 0.3 to 30 µM had no<br />

statistically significant affect on α4β7-mediated cell adhesion to MAdCAM-1. All cells were<br />

viable as assessed by trypan blue exclusion. As before, genistein inhibited both PMA and<br />

Mn 2+ -induced cell adhesion.<br />

99


Figure 3.16 Effect of radicicol on TK-1 cell adhesion to MAdCAM-1<br />

TK-1 cells were incubated with 0, 0.3, 3 or 30µM radicicol or 10 µM of genistein at 37 °C for 3 hrs. <strong>The</strong> cells<br />

were activated with either PMA (A) or Mn 2+ (B), and adhered to MAdCAM-1-Fc-coated plates at RT for 30<br />

min. TK-1 cells were labelled with CMFDA prior to addition of the inhibitors, and cell binding was analysed<br />

using a fluorescence microplate reader with excitation and emission wavelengths of 485 nm and 535 nm<br />

respectively. Data shown represent the mean and SD of three wells. <strong>The</strong> experiments were performed in<br />

triplicate<br />

3.2.6 Myosin light chain kinase (MLCK)<br />

MLCK is a serine/threonine protein kinase that phosphorylates myosin II light chain (Smith et<br />

al. 2003). Phosphorylation of myosin enables myosin to bind to actin filaments and allows<br />

cell contraction to begin (Smith et al. 2003). T cell migration on ICAM-1 mediated by LFA-1<br />

was found to involve MLCK activity, where MLCK was required for cell attachment and<br />

movement of the cell at the leading edge (Smith et al. 2003).<br />

ML-7<br />

ML-7 is a selective inhibitor of myosin light chain kinase, which also inhibits protein kinase<br />

A and protein kinase C, but at much higher concentrations (Saitoh et al. 1987). TK-1 cells<br />

were treated with 0.3, 25, and 50 µM ML-7. Those treated with 0.3 and 25 µM ML-7<br />

remained viable for the duration of the experiment, whereas those treated with 50 µM ML-7<br />

showed more than 70% cell death as assessed by trypan blue exclusion and hence were<br />

100


excluded from statistical analysis. Treatment of PMA-activated TK-1 cells with 0.3 and 25<br />

µM ML-7 (Figure 3.17A) significantly decreased α4β7-mediated cell adhesion to<br />

MAdCAM-1 by 11% (p = 4.88 × 10 -3 ) and 30% (p = 2.24 × 10 -5 ), respectively. Treatment of<br />

Mn 2+ -activated TK-1 cells with 25 µM ML-7 (Figure 3.17B) significantly decreased α4β7-<br />

mediated cell adhesion to MAdCAM-1 by 34% (p = 1.64 × 10 -4 ), whereas statistical<br />

significance was not achieved with the lower concentration. As before, genistein inhibited<br />

both PMA and Mn 2+ -induced cell adhesion.<br />

Figure 3.17 Effect of ML-7 on TK-1 cell adhesion to MAdCAM-1<br />

TK-1 cells were incubated with 0, 0.3, 25 and 50 µM ML-7 or 10 µM of genistein at 37 °C for 3 hrs. <strong>The</strong> cells<br />

were activated with either PMA (A) or Mn 2+ (B), and adhered to MAdCAM-1-Fc-coated plates at RT for 30<br />

min. PMA-activated TK-1 cells (A) were labelled with CMFDA prior to addition of the inhibitors, and cell<br />

binding was analysed using a fluorescence microplate reader with excitation and emission wavelengths of 485<br />

nm and 535 nm respectively. Mn 2+ -activated TK-1 cells (B) were fixed with glutaraldehyde, stained with<br />

methylene blue, and cell binding was analysed using a microplate reader at a wavelength of 495 nm. Data shown<br />

represent the mean and SD of three wells. <strong>The</strong> experiments were performed in triplicate. # denotes cells with<br />

greater then 70% cell death by trypan blue exclusion. *** denotes a p-value < 0.001 which signifies statistical<br />

significance when compared to cell binding of activated cells not treated with inhibitor.<br />

101


3.3. Properties of a cell adhesion regulatory domain in the cytoplasmic tail<br />

of the integrin β7 subunit<br />

α4β7-mediated cell adhesion is regulated in part via the integrin β subunit cytoplasmic<br />

domain. Previously, our laboratory employed cell-permeable peptide technology to scan the<br />

β7 integrin cytoplasmic domain for cell adhesion regulatory domains (CARDs). Short<br />

peptides encompassing the 52 aa sequence of the β7 cytoplasmic domain were synthesized by<br />

fusion to a cell-permeable poly-arginine (r9) sequence. <strong>The</strong> peptides were tested for their<br />

ability to interfere with α4β7-mediated cell adhesion. This approach led to the identification<br />

of the 6 aa motif YDRREY which is proposed to serve as a CARD for β7 integrins (Figure<br />

3.18; Krissansen et al. 2006b). <strong>The</strong> focus turned to an investigation of the YDDREY motif to<br />

determine whether components of the above β7 integrin regulatory pathway(s) identified by<br />

chemical inhibition might interact with the YDRREY peptide. <strong>The</strong> initial aim was to identify<br />

critical residues in the β7 integrin CARD involved in β7 integrin regulation. Peptides fused<br />

N-terminally to the poly-arginine cell permeable peptide were used in all the following<br />

experiments, unless otherwise stated.<br />

731 741 751 761 771 781<br />

| | | | | |<br />

β7 K LSVEIYDRRE YSRFEKEQQQ LNWKQDSNPL YKSAITTTIN PRFQEADSPT L<br />

Figure 3.18 <strong>The</strong> amino acid sequence of the human β7 integrin cytoplasmic domain<br />

<strong>The</strong> amino acid sequence of the integrin β7 subunit cytoplasmic domain with the potential cell adhesion<br />

regulatory domain (CARD) highlighted, and the conserved NPLY motif underlined.<br />

3.3.1 A cell-permeable YDRREY peptide inhibits β7 integrin-mediated adhesion to<br />

MAdCAM-1 and E-cadherin<br />

α4β7-mediated adhesion to MAdCAM-1<br />

TK-1 cells were incubated with a synthetic, biotinylated, cell-permeable YDRREY peptide<br />

and tested for adhesion to MAdCAM-1 to confirm that the YDRREY motif sequence has the<br />

properties of a CARD. Peptides were biotinylated so that peptide uptake by cells could be<br />

confirmed by staining cells with fluoresceinated streptavidin (data not shown). TK-1 cells<br />

were incubated with 0, 25, 50 and 100 µM biotin-r9-YDRREY-OH (YDRREY) peptide for<br />

30 min, and then activated with Mn 2+ and added to plates coated with MAdCAM-1-Fc. All<br />

102


cells were confirmed to be > 98% viable as assessed by trypan blue exclusion. Bound cells<br />

were fixed and stained with methylene blue and quantified in a microplate reader. Non-<br />

activated cells were used as negative controls. Increasing concentrations of the YDRREY<br />

peptide decreased TK-1 cell binding to MAdCAM-1 (Figure 3.19). YDDREY at 50 and 100<br />

μM significantly decreased cell binding by 45% (p = 3.05 × 10 -3 ) and 68% (p = 2.1 × 10 -5 ),<br />

respectively, compared to the binding of non-activated cells. <strong>The</strong> YDRREY peptide at 100<br />

µM served as a positive control in subsequent experiments.<br />

Figure 3.19 Effect of the YDRREY peptide on TK-1 cell adhesion to MAdCAM-1.<br />

TK-1 cells were incubated in HBSS with 0, 25, 50 or 100 µM YDRREY peptide (green) at RT for 30 min, and<br />

activated with Mn 2+ . Non-activated cells not treated with the YDRREY served as negative controls (blue bar).<br />

Cells were adhered to MAdCAM-1-coated plates at RT for 30 min. <strong>The</strong> bound cells were fixed with<br />

glutaraldehyde, stained with methylene blue and analysed using a microplate reader at a wavelength of 495 nm.<br />

Data shown represent the mean and SD of three wells. <strong>The</strong> experiments were performed in triplicate. ** denotes<br />

a p-value of < 0.01 and *** denotes a p-value of < 0.001, compared to cell binding of Mn 2+ -activated cells not<br />

treated with the YDRREY peptide.<br />

αEβ7-mediated adhesion to E-cadherin<br />

<strong>The</strong> cell permeable peptide YDRREY was tested for its ability to disrupt αEβ7-mediated<br />

adhesion to E-cadherin to confirm that the YDRREY motif is a CARD for both members of<br />

the β7 integrin subfamily. TGF-β-differentiated MTC-1 cells were incubated with 0, 10, 50<br />

µM YDRREY peptide for 30 min and activated with AlF4 - to bind to soluble E-cadherin-<br />

antibody complexes (E-cadherin-Fc complexed with anti-human IgG-HRP; Figure 3.20).<br />

Non-activated cells incubated with anti-human IgG-HRP served as negative controls. All<br />

cells were confirmed to be viable as assessed by trypan blue exclusion. Treatment of AlF4 - -<br />

activated MTC-1 cells with the YDRREY peptide at 10 µM and 50 µM significantly<br />

103


decreased αEβ7-mediated adhesion to E-cadherin by 35% (p = 3.16 × 10 -5 ) and 70% (p = 2 ×<br />

10 -7 ), respectively, compared to cell binding in the absence of the YDRREY peptide.<br />

Figure 3.20 Effect of YDRREY peptide on MTC-1 cell adhesion to E-cadherin<br />

TGF-β-differentiated MTC-1 cells were incubated in HBSS with 0, 10, or 50 µM biotin-r9-YDRREY-OH<br />

peptide (green) at RT for 30 min. Cells were activated with AlF4 - and bound to the E-cadherin-complex at RT for<br />

30 min. <strong>The</strong> E-cadherin-antibody complex was formed by incubating 1 µg of E-cadherin with of anti-human-<br />

IgG-HRP (diluted 1:1000) at RT for 30 min. Control MTC-1 cells were stained with anti-human-IgG-HRP in the<br />

absence of E-cadherin (No E-cadherin, blue bar). Cells bound to the E-cadherin-antibody complex were<br />

visualised by the addition of ABTS, and analysed using a microplate reader at a wavelength of 405 nm. Data<br />

shown represent the mean and SD of three wells. <strong>The</strong> experiments were performed in triplicate. *** denotes a pvalue<br />

of < 0.001, compared to binding of the E-cadherin-antibody complex in the absence of YDRREY peptide.<br />

3.3.2 Defining the features of the YDRREY peptide required to inhibit β7-mediated<br />

adhesion<br />

Various aa residues of the YDDREY peptide were substituted in order to identify key<br />

residues and features of the peptide responsible for inhibiting α4β7 integrin-mediated cell<br />

adhesion to MAdCAM-1. A previous study by the host laboratory had revealed that deletion<br />

of either of the flanking tyrosines destroyed the activity of the peptide.<br />

xDRREx, a phosphorylated form of YDRREY<br />

Unphosphorylated tyrosine residues which flanked the YDRREY motif were replaced with<br />

phosphorylated tyrosine residues to give the sequence xDRREx, where x represents a<br />

phosphorylated tyrosine, to determine whether peptide phosphorylation affected the<br />

inhibitory activity of the YDRREY peptide. TK-1 cells were incubated with 0, 25, 50 and 100<br />

µM biotin-r9-xDRREx (xDRREx) peptide for 30 min, activated with Mn 2+ , and added to<br />

MAdCAM-1-Fc-coated plates. Non-activated cells served as negative controls. All cells were<br />

104


confirmed viable as assessed by trypan blue exclusion. Treatment of Mn 2+ -activated TK-1<br />

cells with the xDRREx peptide at 25, 50 and 100 µM decreased α4β7 integrin-mediated cell<br />

adhesion to MAdCAM-1 by 14%, 27% (p = 1.47 × 10 -2 ), and 88% (3.07 × 10 -4 ), respectively,<br />

compared to the binding of activated cells in the absence of peptide (Figure 3.21). <strong>The</strong><br />

xDRREx peptide at 100 µM prevented α4β7 integrin-mediated cell adhesion to the same<br />

extent as the YDRREY peptide at 100 µM, suggesting that the phosphorylation state of the<br />

flanking tyrosines does not adversely influence the activity of the CARD.<br />

Figure 3.21 Effect of the phosphorylated xDRREx peptide on TK-1 cell adhesion to MAdCAM-1<br />

TK-1 cells were incubated in HBSS with 0, 25, 50 or 100 µM xDRREx peptide (green) at RT for 30 min. <strong>The</strong><br />

cells were either left non-activated or activated with Mn 2+ . Non-activated cells not treated with the xDRREx<br />

served as negative controls (blue bar). TK-1 cells incubated with the YDRREY peptide (purple bar) at 100 µM<br />

served as positive controls. Cells were adhered to MAdCAM-1-Fc-coated plates at RT for 30 min. <strong>The</strong> bound<br />

cells were fixed with glutaraldehyde, stained with methylene blue, and analysed using a microplate reader at a<br />

wavelength of 495 nm. Data shown represent the mean and SD of three wells. <strong>The</strong> experiments were performed<br />

in triplicate. * denotes a p-value < 0.05 and *** denotes a p-value of < 0.001; compared to cell binding of Mn 2+ -<br />

activated cells not treated with the xDRREx peptide.<br />

YDRGGGGREY<br />

<strong>The</strong> two tyrosine residues flanking the YDRREY motif were found to be critical for the<br />

ability of the YDRREY peptide to inhibit α4β7 integrin-mediated cell adhesion to<br />

MAdCAM-1 (Krissansen et al. 2006b). However, as shown above, the state of<br />

phosphorylation of the two flanking tyrosine residues did not seem to influence the activity of<br />

the YDRREY peptide. <strong>The</strong> aim here was to determine whether the activity of the YDRREY<br />

peptide was dependent on the proximity of the two flanking tyrosines. Four glycine residues<br />

105


were inserted into the middle of the YDRREY sequence between the arginine residues to<br />

form the biotin-r8-YDRGGGGREY-OH (YDRGGGGREY) peptide. TK-1 cells were<br />

incubated with 0, 25, 50 and 100 µM concentrations of the YDRGGGGREY peptide for 30<br />

min, activated with Mn 2+ , and added to MAdCAM-1-Fc-coated plates. Non-activated cells<br />

were used as negative controls. All cells were confirmed to be viable as assessed by trypan<br />

blue exclusion. Treatment of TK-1 cells with increasing concentrations of the<br />

YDRGGGGREY peptide decreased the amount of TK-1 cell binding to MAdCAM-1 (Figure<br />

3.22). Treatment with 50 and 100 µM YDRGGGGREY peptide significantly decreased cell<br />

adhesion by 23% (p = 1.46 × 10 -2 ) and 95% (p = 3.4 × 10 -6 ), respectively, compared to the<br />

binding of activated cells in the absence of peptide. <strong>The</strong> YDRGGGGREY peptide at 100 µM<br />

prevented α4β7 integrin-mediated cell adhesion to the same extent as the YDRREY peptide<br />

at 100 µM. <strong>The</strong> results demonstrate that precise spacing of the flanking tyrosines may not be<br />

a critical factor in the structure/function of the CARD. It is possible the flanking tyrosines act<br />

independently.<br />

Figure 3.22 Effect of the YDRGGGGREY peptide on TK-1 cell adhesion to MAdCAM-1<br />

TK-1 cells were incubated in HBSS with 0, 25, 50 or 100 µM YDRGGGGREY peptide (green) at RT for 30<br />

min. <strong>The</strong> cells were either left non-activated or activated with Mn 2+ . Non-activated cells not treated with the<br />

YDRGGGGREY peptide served as negative controls (blue bar). TK-1 cells incubated with the YDRREY<br />

peptide (purple bar) at 100 µM served as positive controls. Cells were adhered to MAdCAM-1-Fc-coated plates<br />

at RT for 30 min. <strong>The</strong> bound cells were fixed with glutaraldehyde, stained with methylene blue, and analysed<br />

using a microplate reader at a wavelength of 495 nm. Data shown represent the mean and SD of three wells. <strong>The</strong><br />

experiments were performed in triplicate. * denotes a p-value < 0.05 and *** denotes a p-value of < 0.001; when<br />

compared to cell binding of Mn 2+ -activated cells not treated with the YDRGGGGREY peptide.<br />

106


YDGGEY<br />

<strong>The</strong> two central positively charged arginine residues in the YDRREY motif were replaced by<br />

neutral uncharged glycine residues, to determine whether the central arginines are important<br />

for influencing β7 integrin mediated-cell adhesion. TK-1 cells were incubated with 0, 50, 100,<br />

and 200 µM biotin-r9-YDGGEY (YDGGEY) peptide for 30 min, activated with Mn 2+ , and<br />

added to MAdCAM-1-Fc-coated plates. Non-activated cells served as negative controls. All<br />

cells were confirmed to be viable as assessed by trypan blue exclusion. Figure 3.23 shows<br />

that treatment of TK-1 cells with up to 100 µM YDGGEY peptide had no significant affect<br />

on the binding of TK-1 cells to MAdCAM-1, where 100 µM YDGGEY peptide inhibited<br />

adhesion by just 10%. <strong>The</strong> peptide at the very high concentration of 200 µM significantly<br />

reduced cell adhesion by 60% (p = 1.19 × 10 -3 ), but was not as effective as 100 µM YDRREY<br />

peptide. <strong>The</strong> results suggest that the central arginines of the YDRREY peptide play an<br />

influential role in the function of the CARD. <strong>The</strong> YDGGEY peptide was readily taken up by<br />

cells (data not shown), ruling out the possibility that problems with cell-permeability<br />

accounted for the lack of efficacy of the YDGGEY peptide.<br />

Figure 3.23 Effect of the YDGGEY peptide on TK-1 cell adhesion to MAdCAM-1<br />

TK-1 cells were incubated in HBSS with 0, 50, 100 or 200 µM YDGGEY peptide (green) at RT for 30 min. <strong>The</strong><br />

cells were either left non-activated or activated with Mn 2+ . Non-activated cells not treated with the YDGGEY<br />

peptide served as negative controls (blue bar). TK-1 cells incubated with the YDRREY peptide (purple bar) at<br />

100 µM served as positive controls. Cells were added to MAdCAM-1-Fc-coated plates at RT for 30 min. <strong>The</strong><br />

bound cells were fixed with glutaraldehyde, stained with methylene blue and analysed using a microplate reader<br />

at a wavelength of 495 nm. Data shown represent the mean and SD of three wells. <strong>The</strong> experiments were<br />

performed in triplicate. ** denotes a p-value of < 0.01; when compared to cell binding of Mn 2+ -activated cells<br />

not treated with the YDGGEY peptide.<br />

107


YEEEY<br />

<strong>The</strong> central core DRRE of the YDRREY motif was substituted with four glutamic acid<br />

residues. Poly EY (4:1) is a substrate for many tyrosine kinases (Braun et al. 1984), hence<br />

comparison of the YDRREY and YEEEEY peptides sought to investigate whether the<br />

YDRREY peptide acts in a nonspecific manner as a general substrate sink by providing<br />

phosphorylatable tyrosines, or whether the non-phosphorylatable central core was critical for<br />

activity. TK-1 cells were incubated with 0, 50, 100, and 200 µM biotin-r9-YEEEEY<br />

(YEEEEY) peptide for 30 min, activated with Mn 2+ , and added to MAdCAM-1-Fc-coated<br />

plates. Non-activated cells served as negative controls. All cells were confirmed to be viable<br />

as assessed by trypan blue exclusion. Figure 3.24 shows that treatment of TK-1 cells with up<br />

to 100 µM YEEEEY peptide had no significant affect on the binding of TK-1 cells to<br />

MAdCAM-1, where adhesion was inhibited by just 24% at 100 µM YEEEEY peptide. <strong>The</strong><br />

peptide at the very high concentration of 200 µM significantly reduced cell adhesion by 40%<br />

(p = 5.29 × 10 -3 ), but was not as effective as 100 µM YDRREY peptide. <strong>The</strong> results confirm<br />

that the central arginines in the central core of the YDRREY peptide do play an influential<br />

role in the function of the CARD. <strong>The</strong> YDRREY peptide may be a preferential substrate for<br />

tyrosine kinases that regulate the function of β7 integrins. <strong>The</strong> YEEEEY peptide was readily<br />

taken up by cells (data not shown), ruling out the possibility that problems with cellpermeability<br />

accounted for the lack of efficacy of the YEEEY peptide.<br />

108


Figure 3.24 Effect of the YEEEEY peptide on TK-1 cell adhesion to MAdCAM-1<br />

TK-1 cells were incubated in HBSS with 0, 50, 100 or 200 µM YEEEEY peptide (green) at RT for 30 min. <strong>The</strong><br />

cells were either left non-activated or activated with Mn 2+ . Non-activated cells not treated with the YEEEEY<br />

peptide served as negative controls (blue bar). TK-1 cells incubated with the YDRREY peptide (purple bar) at<br />

100 µM served as positive controls. Cells were added to MAdCAM-1-Fc-coated plates at RT for 30 min. <strong>The</strong><br />

bound cells were fixed with glutaraldehyde, stained with methylene blue, and analysed using a microplate reader<br />

at a wavelength of 495 nm. Data shown represent the mean and SD of three wells. <strong>The</strong> experiments were<br />

performed in triplicate. ** denotes a p-value of < 0.01; compared to cell binding of Mn 2+ -activated cells not<br />

treated with the YEEEEY peptide.<br />

YDRREYGYDRREYGYDRREYGYDRREY<br />

A polymer (4-mer) of the YDRREY motif YDRREYGYDRREYGYDRREYGYDRREY,<br />

was synthesised and tested to determine whether multiple YDRREY motifs potentially<br />

working in a multivalent fashion may more effectively inhibit TK-1 cell adhesion to<br />

MAdCAM-1. TK-1 cells were incubated with 0, 25, 50 and 100 µM biotin-r8-<br />

YDRREYGYDRREYGYDRREYGYDRREY (pYDRREY) peptide for 30 min, activated<br />

with Mn 2+ , and added to MAdCAM-1-Fc-coated plates. Non-activated cells served as<br />

negative controls. All cells were confirmed to be viable as assessed by trypan blue exclusion.<br />

Increasing concentrations of the pYDRREY peptide decreased the amount of TK-1 cell<br />

binding to MAdCAM-1 (Figure 3.25). Treatment with 25, 50 and 100 µM pYDRREY<br />

peptide significantly decreased cell adhesion by 22% (p = 7 × 10 -3 ), 61% (p = 6.16 × 10 -4 ),<br />

and 80% (p = 2.95 × 10 -4 ), respectively, compared to the binding of activated cells in the<br />

absence of peptide. <strong>The</strong> pYDRREY peptide at 100 µM prevented α4β7 integrin-mediated cell<br />

adhesion to the same extent as the YDRREY peptide at 100 µM, but was slightly more<br />

109


effective at 50 µM (refer to Figure 3.19). <strong>The</strong> pYDRREY peptide was used as a positive<br />

control in subsequent experiments given its excellent inhibitory properties.<br />

Figure 3.25 Effect of the pYDRREY peptide on TK-1 cell adhesion to MAdCAM-1<br />

TK-1 cells were incubated in HBSS with 0, 25, 50 or 100 pYDRREY peptide (green) at RT for 30 min. <strong>The</strong><br />

cells were either left non-activated or activated with Mn 2+ . Non-activated cells not treated with the pYDRREY<br />

peptide served as negative controls (blue bar). TK-1 cells incubated with the YDRREY peptide (purple bar) at<br />

100 µM served as positive controls. Cells were added to MAdCAM-1-coated plates at RT for 30 min. <strong>The</strong> bound<br />

cells were fixed with glutaraldehyde, stained with methylene blue and analysed using a microplate reader at a<br />

wavelength of 495 nm. Data shown represent the mean and SD of three wells. <strong>The</strong> experiments were performed<br />

in triplicate. ** denotes a p-value of < 0.01 and *** denotes a p-value of < 0.001; compared to cell binding of<br />

Mn 2+ -activated cells not treated with the pYDRREY peptide.<br />

FDRREFGFDRREFGFDRREFGFDRREF<br />

As mentioned above, the flanking tyrosines of the YDRREY motif are important for the<br />

function of the CARD. However, the central core sequence DRRE was also found to be<br />

important for the activity of the YDRREY peptide (refer to Figure 3.24). Here, a polymer (4mer)<br />

of the core sequence was tested for activity with the expectation that it may have<br />

efficacy due to its multivalency. <strong>The</strong> flanking tyrosines in the 4-mer were substituted with<br />

phenylalanine residues to give the sequence FDRREFGFDRREFGFDRREFGFDRREF. TK-<br />

1 cells were incubated with 0, 25, 50 and 100 µM biotin-r8-<br />

FDRREFGFDRREFGFDRREFGFDRREF (pFDRREF) peptide for 30 min, activated with<br />

Mn 2+ , and added to MAdCAM-1-Fc-coated plates. Non-activated cells served as negative<br />

controls. All cells were confirmed to be viable as assessed by trypan blue exclusion.<br />

Increasing concentrations of the pFDRREF peptide decreased the amount of TK-1 cell<br />

binding to MAdCAM-1 (Figure 3.26). Treatment with 50 µM and 100 µM pFDRREF<br />

110


peptide significantly reduced α4β7 integrin-mediated cell adhesion by 53% (p = 3.84 × 10 -3 )<br />

and 82% (p = 5.19 × 10 -4 ), respectively. <strong>The</strong> pFDRREF peptide at 100 µM showed a similar<br />

level of inhibition of cell adhesion as achieved with the YDRREY peptide at 100 µM. <strong>The</strong><br />

results suggest that the central core sequence in the absence of flanking tyrosines can be<br />

rendered active when synthesized as a polymer to potentially increase the multivalency of<br />

interaction.<br />

Figure 3.26 Effect of the pFDRREF peptide on TK-1 cell adhesion to MAdCAM-1<br />

TK-1 cells were incubated in HBSS with 0, 25, 50 or 100 pFDRREF peptide (green) at RT for 30 min. <strong>The</strong> cells<br />

were either left non-activated or activated with Mn 2+ . Non-activated cells not treated with the pFDRREF peptide<br />

served as negative controls (blue bar). TK-1 cells incubated with the YDRREY peptide (purple bar) at 100 µM<br />

served as positive controls. Cells were added to MAdCAM-1-coated plates at RT for 30 min. <strong>The</strong> bound cells<br />

were fixed with glutaraldehyde, stained with methylene blue and analysed using a microplate reader at a<br />

wavelength of 495 nm. Data shown represent the mean and SD of three wells. <strong>The</strong> experiments were performed<br />

in triplicate. *** denotes a p-value < 0.001; compared to cell binding of Mn 2+ -activated cells not treated with the<br />

pFDRREF peptide.<br />

3.3.3 Effect of NPLY on TK-1 cell adhesion<br />

<strong>The</strong> β7 cytoplasmic domain contains three tyrosine residues, in addition to the two within the<br />

YDRREY motif. <strong>The</strong> third tyrosine is present in the conserved central 758-NPLY-761 motif<br />

(refer to Figure 3.18). Members of the laboratory had previously shown that cell-permeable<br />

forms of the NPKF peptide from the cytoplasmic domain of the integrin β2 subunit were<br />

capable of inhibiting αLβ2-mediated adhesion of T cells to ICAM-1 (unpublished results).<br />

Hence, a cell-permeable form of the biotin-r8-NPLY (NPLY) peptide was synthesized and<br />

tested for its ability to prevent α4β7-mediated adhesion of TK-1 cells to MAdCAM-1. TK-1<br />

111


cells were incubated with 0, 25, 50 and 100 µM NPLY peptide for 30 min, activated with<br />

Mn 2+ , and added to MAdCAM-1-Fc-coated plates. Non-activated cells served as negative<br />

controls. All cells were confirmed to be viable as assessed by trypan blue exclusion.<br />

Increasing concentrations of the NPLY peptide decreased the amount of TK-1 cell binding to<br />

MAdCAM-1 (Figure 3.27). Treatment with 50 and 100 µM NPLY peptide significantly<br />

reduced α4β7-mediated cell adhesion by 21% (3.92 x 10 -3 ) and 72% (1.1 x 10 -3 ), respectively.<br />

<strong>The</strong> NPLY peptide at 100 µM inhibited cell adhesion to a similar extent as that obtained with<br />

the YDRREY peptide at 100 µM. Thus, the NPLY motif represents a second CARD within<br />

the β7 subunit cytoplasmic domain that appears to be equally important for α4β7-mediated<br />

cell adhesion.<br />

Figure 3.27 Effect of the NPLY peptide on TK-1 cell adhesion to MAdCAM-1<br />

TK-1 cells were incubated in HBSS with 0, 10, 50 or 100 NPLY peptide (green) at RT for 30 min. <strong>The</strong> cells<br />

were either left non-activated or activated with Mn 2+ . Non-activated cells not treated with the NPLY peptide<br />

served as negative controls (blue bar). TK-1 cells incubated with the YDRREY peptide (purple bar) at 100 µM<br />

served as positive controls. Cells were added to MAdCAM-1-coated plates at RT for 30 min. <strong>The</strong> bound cells<br />

were fixed with glutaraldehyde, stained with methylene blue and analysed using a microplate reader at a<br />

wavelength of 495 nm. Data shown represent the mean and SD of three wells. <strong>The</strong> experiments were performed<br />

in triplicate. ** denotes a p-value < 0.01; compared to cell binding of Mn 2+ -activated cells not treated with the<br />

NPLY peptide.<br />

112


3.4. Tyrosine kinases interact with and phosphorylate the YDRREY motif<br />

Integrin β subunit cytoplasmic domains do not possess intrinsic kinase activity. However they<br />

contain several tyrosines residues that are potentially phosphorylatable (refer Figure 1.10).<br />

<strong>The</strong> β1 subunit cytoplasmic domain contains two tyrosine residues at position 783 and 795<br />

within the two conserved central and distal NPxY motifs, respectively. <strong>The</strong> β2 subunit<br />

cytoplasmic domain contains only one tyrosine residue at position 736 within the β2<br />

counterpart (SDLREY) of the β7 YDRREY motif. As mentioned above, the β7 subunit<br />

cytoplasmic domain contains two tyrosine residues at positions 736 and 741 within the<br />

YDRREY motif, and at position 761 within the central NPxY motif. <strong>The</strong> initial aim was to<br />

determine whether these sites were phosphorylatable, with a particular interest in the<br />

YDRREY motif of the β7 integrins.<br />

3.4.1 Production of GST-β integrin cytoplasmic domain fusion proteins<br />

Whilst the above experiments (Figure 3.21) suggested that the phosphorylation status of the<br />

cell-permeable YDDREY peptide does not influence the activity of the peptide, it was<br />

possible that the peptide was phosphorylated by an endogenous tyrosine(s) kinase after<br />

uptake into cells. Phosphorylation of the YDRREY peptide would be expected to create SH2<br />

binding domains for intracellular ligands (Krissansen et al. 2006b). <strong>The</strong> focus turned to<br />

studying the phosphorylation status of tyrosine residues within the integrin β7 cytoplasmic<br />

domain. Our laboratory had previously constructed vectors encoding GST-fusion proteins of<br />

the β integrin subunit cytoplasmic domains. DNAs encoding the integrin β1, β2, and β7<br />

cytoplasmic domains had been subcloned into the pGEX-2T vector. In addition to the native<br />

form of the β7 cytoplasmic domain, the β7 cytoplasmic domain was mutated by substituting<br />

phenylalanine residues for each of the tyrosines (Y736F, Y741F, Y736F+Y741F, and<br />

Y761F), either alone or in combination. DNAs encoding the mutated variants were sub-<br />

cloned into pGEX-2T. <strong>The</strong> pGEX-2T vectors containing the integrin β subunit sequences<br />

were transformed into the DH5α strain of E.coli. Protein expression was induced with IPTG<br />

for 180 min, and the bacteria were lysed with 1% Triton X-100. <strong>The</strong> soluble protein fractions<br />

were purified by affinity chromatography using batch binding to glutathione (GSH)-<br />

Sepharose. <strong>The</strong> purified GST-fusion proteins bound to GSH-Sepharose were kept in PBS in<br />

the presence of sodium azide. <strong>The</strong> GST-fusion proteins bound to Sepharose were eluted with<br />

reduced glutathione, resolved by SDS-PAGE under reducing conditions, and stained with<br />

113


Coomassie blue (Figure 3.28). <strong>The</strong> purification system yielded similar amounts (~ 1 mg/L of<br />

bacterial culture) of each of the seven GST-fusion proteins. <strong>The</strong>y migrated with a molecular<br />

weight of ~30 kDa. GST protein alone with a molecular weight of ~25 kDa was also<br />

produced and purified to serve as an experimental negative control (Figure 3.28, lane 3).<br />

Figure 3.28 Production of GST-β subunit cytoplasmic domain fusion proteins, and analysis by SDS-PAGE<br />

E. coli were transformed with GST-fusion constructs encoding integrin β-subunit cytoplasmic domains. Protein<br />

expression was induced with 0.4 mM IPTG at 30°C for 180 min, and bacteria collected and lysed in 1% Triton<br />

X-100. <strong>The</strong> soluble protein fraction was affinity-purified using GSH-Sepharose and eluted with reduced<br />

glutathione. One microgram of each purified protein was analysed by SDS-PAGE under reducing conditions,<br />

and stained with Coomassie blue. Lane 1, GST-β1 cytoplasmic domain; lane 2, GST-β2 cytoplasmic domain;<br />

lane 3, GST; lane 4, GST-β7 cytoplasmic domain; lane 5, GST-β7 (Y736F) cytoplasmic domain; lane 6, GST-β7<br />

(Y741F) cytoplasmic domain; lane 7, GST-β7 (Y736+41F) cytoplasmic domain; lane 8, GST-β7 (Y761F)<br />

cytoplasmic domain. Molecular weights of marker proteins are shown in the left-hand margin in kDa.<br />

3.4.2 Phosphorylation of GST-β7 cytoplasmic domain fusion proteins<br />

Purified GST-β7 cytoplasmic domain fusion proteins bound to Sepharose beads were tested<br />

for their ability to serve as substrates for tyrosine kinases present in TK-1 cells. <strong>The</strong>y were<br />

incubated with trace amounts of a TK-1 cellular lysate in the presence of 32 P-γATP. After<br />

phosphorylation, the GST-fusion proteins bound to Sepharose beads were washed thoroughly,<br />

and the amount of 32 P incorporation was analysed using a liquid scintillation counter (Figure<br />

3.29). <strong>The</strong> level of phosphorylation corresponded to the number of tyrosine residues present<br />

in each of the GST-β subunit cytoplasmic domains fusion proteins. GST-β1 with two tyrosine<br />

residues had almost twice the level of 32 P incorporation (CPM 463,349) of GST-β2 with one<br />

tyrosine residue (CPM 276,522). GST-β7 with three tyrosine residues showed the greatest<br />

level of phosphorylation (CPM 651,641). Mutation of the β7 subunit tyrosine residues to<br />

phenylalanines decreased the level of phosphorylation such that 32 P incorporation into Y736F<br />

(CPM 412,924), and Y741F (CPM 315,989), was similar to that of GST-β1 with two tyrosine<br />

residues. 32 P incorporation into Y736F+Y741F (CPM 56,794) was very low, suggesting that<br />

114


Y761 is poorly phosphorylated. In accord, substitution of Y761 with phenylalanine had little<br />

affect on 32 P incorporation (CPM 565,109) into this cytoplasmic domain variant. This result<br />

indicates that all the tyrosine residues in the cytoplasmic domains of the integrin β1, β2, and<br />

β7 subunits are potentially phosphorylatable. However, Y761 within the NPxY CARD of the<br />

β7 subunit cytoplasmic domain does not appear to be as readily phosphorylated as Y736 and<br />

Y741 within the YDRREY CARD.<br />

Figure 3.29 Phosphorylation of GST-β subunit cytoplasmic domain fusion proteins<br />

Recombinant GST-β subunit cytoplasmic domain fusion proteins were incubated with trace amounts of a TK-1<br />

cell lysate (3 x 10 7 cells/mL of lysis buffer, diluted 1:100) and 1 μCi 32 P-γATP. <strong>The</strong> amount of 32 P incorporated<br />

was determined using a Wallac Trilux microbeta 1450 liquid scintillation counter, measured as counts per<br />

minute (CPM). GST-fusion proteins included those containing the β1 (GST- β1), β2 (GST- β2), and β7 (GST-<br />

β7) cytoplasmic domains, GST- β7 in which tyrosines 736 (Y736F), 741 (Y741F), 761 (Y761F) and both<br />

tyrosines 736 and 741 (Y736F+Y741F) had been substituted for phenylalanines. A GST only control was<br />

processed as for the GST-β subunit cytoplasmic domain fusion proteins to measure non-specific 32 P<br />

incorporation which was subtracted from all the individual readings before plotting the bar graph. Data shown<br />

represent the mean and SD of three wells. <strong>The</strong> experiments were performed in triplicate.<br />

3.4.3 Tyrosine phosphorylation of the YDRREY peptide<br />

<strong>The</strong> aim here was to employ a synthetic YDRREY peptide to confirm that Y736 and Y741<br />

were phosphorylatable as suggested by the results obtained with the GST-β7 subunit<br />

cytoplasmic domain fusion protein. A biotin-GGYDRREY peptide and an unrelated peptide<br />

control (biotin-APTLPPAWQPFLK) were immobilised onto streptavidin-coated Sepharose<br />

beads, and incubated with trace amounts of a TK-1 cell lysate and 32 P-γATP in an in vitro<br />

115


kinase assay. <strong>The</strong> in vitro kinase assay confirmed that the YDRREY peptide was<br />

phosphorylatable, giving a two-fold increase in 32 P incorporation as compared to streptavidin-<br />

coated Sepharose beads alone, or the unrelated peptide control (Figure 3.30).<br />

Figure 3.30 <strong>The</strong> YDRREY peptide is phosphorylated by a tyrosine kinase(s) in a TK-1 cell lysate<br />

A biotin-GGYDRREY peptide and a biotinylated unrelated control peptide (biotin-APTLPPAWQPFLK) were<br />

each bound to streptavidin-Sepharose (1 µg /10 μL) at RT for 30 min. <strong>The</strong> peptide-Sepharose beads were<br />

incubated with a TK-1 cell lysate diluted 1:100, and 1 µCi 32 P-γATP at 30˚C for 30 min. Sepharose beads treated<br />

with the TK-1 cell lysate and 32 P-γATP (no peptide) served as a negative control. <strong>The</strong> amount of 32 P<br />

incorporation (phosphorylation) was determined using the Wallac Trilux microbeta 1450 liquid scintillation<br />

counter. Data shown represent the mean and SD of three wells. <strong>The</strong> experiments were performed in triplicate. **<br />

denotes a p-value < 0.05 which signifies statistical significance, compared to Sepharose beads only (i.e. no<br />

peptide).<br />

3.4.4 Kinases in multiple cell types can phosphorylate the YDRREY motif<br />

<strong>The</strong> YDRREY peptide (biotin-GGYDRREY; GGYDRREY) was phosphorylated by tyrosine<br />

kinase(s) present in TK-1 cells. <strong>The</strong> question asked here was whether tyrosine kinases in<br />

other unrelated cell types were also capable of phosphorylating the GGYDRREY peptide.<br />

Cellular lysates tested included those prepared from TK-1 cells (positive control), a mouse<br />

myoblast cell line (C2C12), human embryonic kidney cells (HEK-293T), and Chinese<br />

hamster ovary cells (CHO-K1). Trace amounts of each cell lysate (1 µL of 1:100 dilution)<br />

were incubated with the GGYDRREY peptide with the addition of 32 P-γATP, and the<br />

incorporation of 32 P was analysed. Lysates of all cells tested were able to phosphorylate the<br />

GGYDRREY peptide (Figure 3.31) with comparable ability, indicating that the tyrosine<br />

kinase(s) able to phosphorylate the GGYDRREY peptide are not cell-type-specific.<br />

116


Figure 3.31 Kinases in multiple cell types can phosphorylate the YDRREY motif<br />

<strong>The</strong> GGYDRREY peptide bound to streptavidin-coated Sepharose was incubated with lysates of TK-1, HEK-<br />

293T, C2C12 and CHO-K1 cells, as indicated, in the presence of 1 µCi 32 P-γATP. <strong>The</strong> amount of 32 P<br />

incorporation was determined using the Wallac Trilux microbeta 1450 liquid scintillation counter. Data shown<br />

represent the mean and SD of three wells. <strong>The</strong> experiments were performed in triplicate.<br />

3.4.5 Phosphorylation of the YDRREY peptide by FAK, src and lck<br />

To identify the tyrosine kinase(s) that may be responsible for phosphorylating the YDRREY<br />

peptide, several potential tyrosine kinase candidates, namely FAK, src, and p56lck (lck), were<br />

chosen to study based on their involvement in integrin signalling. <strong>The</strong> recombinant kinases<br />

were incubated with the GGYDRREY peptide immobilised either on streptavidin-Sepharose<br />

or on streptavidin-magnetic beads, and subjected to an in vitro kinase assay. Increasing<br />

amounts (0.1, 0.2 and 0.4 μg) of each of the recombinant kinases increased the level of<br />

phosphorylation of the GGYDRREY peptide. <strong>The</strong> different kinases produced equivalent<br />

levels of 32 P incorporation (Figure 3.32). This result shows that the kinases FAK, src, and lck<br />

are able to phosphorylate the YDRREY sequence.<br />

117


Figure 3.32 Phosphorylation of the GGYDRREY peptide by recombinant FAK, src, and lck<br />

Biotin-r9-GGYDRREY peptide immobilised on streptavidin-coated Sepharose beads was incubated with 0.1,<br />

0.2, and 0.4 µg of recombinant FAK, src, or lck and 1 µCi 32 P-γATP at 30°C for 30 min. <strong>The</strong> amount of 32 P<br />

incorporation was determined using the Wallac Trilux microbeta 1450 liquid scintillation counter. This<br />

experiment was repeated twice.<br />

Phosphorylation of the YDRREY peptide by recombinant FAK, src and lck was confirmed by<br />

resolving the phosphorylated peptides by SDS-PAGE analysis. Two additional biotinylated<br />

cell-permeable CARD peptides, biotin-r9-YDRRE (YDRRE) and biotin-r9-DRREY<br />

(DRREY), were synthesized which lacked one of the flanking tyrosine residues to determine<br />

whether the kinases phosphorylated both flanking tyrosines. <strong>The</strong> peptides and the intact<br />

YDRREY peptide were immobilised on streptavidin-coated Sepharose beads and incubated<br />

with recombinant FAK, src, and lck, in the presence of 32 P-γATP. <strong>The</strong> phosphorylated<br />

peptides were resolved by SDS-PAGE, and the dried gel subjected to autoradiography<br />

(Figure 3.33). <strong>The</strong> YDRREY, YDRRE and DRREY peptides were each phosphorylated by<br />

FAK and lck. In contrast, src only phosphorylated the YDRREY and DRREY peptides. A<br />

polyEY peptide, comprised of a 4:1 ratio of glutamic acid to tyrosine residues, was included<br />

as a positive control. It was phosphorylated by FAK and src, but not by lck. A polymer of<br />

arginine with an unrelated peptide (biotin-rrrrrrrrr-pptdqsrpvqpflnlttprkpr) not containing a<br />

tyrosine phosphorylation site served as a negative control, and was not phosphorylated by the<br />

kinases.<br />

118


Figure 3.33 Sequence recognition of the YDRREY peptide by FAK, src, and lck<br />

Biotin-r9-YDRREY, biotin-r9-DRREY, biotin-r9-YDRRE, a positive control peptide polyEY, and a negative<br />

control peptide r9 (biotin-rrrrrrrrr-pptdqsrpvqpflnlttprkpr) were immobilised on streptavidin-coated Sepharose.<br />

<strong>The</strong> peptide-Sepharose beads were incubated with 0.4 µg of recombinant FAK, src, and lck in the presence of 1<br />

µCi 32 P-γATP at 30°C for 30 min. <strong>The</strong> peptides were analysed on a 16% polyacrylamide tris-tricine SDS-gel.<br />

<strong>The</strong> gel was dried and autoradiographed. Molecular weights of marker proteins are shown in the left-hand<br />

margin in kDa. This experiment was repeated three times.<br />

3.4.6 Direct binding of FAK and src to the YDRREY motif<br />

Given that FAK, src and lck phosphorylate the YDRREY peptide, as shown above, it was<br />

important to determine whether the kinases physically bind to the peptide. FAK had<br />

previously been shown to bind the KLLMIIHDRREF sequence present in the tail of the<br />

integrin β1 subunit (Schaller et al., 1995), which shares partial similarity with the YDRREY<br />

motif. <strong>The</strong> pYDRREY, pFDRREF, and biotin-r9-pptdqsrpvqpflnlttprkpr (control) peptides<br />

were immobilised on Sepharose beads, and mixed with recombinant FAK, src, and lck to<br />

determine whether the kinases bind to the YDRREY sequence. <strong>The</strong> beads were thoroughly<br />

washed, and subjected to an in vitro kinase assay. Phosphorylated forms of the peptides were<br />

visualised by SDS-PAGE and exposure onto X-ray film. <strong>The</strong> premise was that in order for the<br />

pYDRREY peptide to be phosphorylated the kinases must remain bound to the peptide on<br />

washing. <strong>The</strong> pFDRREF and biotin-r9-control peptides do not contain a phosphorylation site,<br />

hence served as negative controls. FAK and src phosphorylated the pYDRREY peptide after<br />

extensive washing, suggesting that they had bound the peptide (Figure 3.34). In contrast, lck<br />

119


did not phosphorylate the pYDRREY peptide, and hence had not been retained by the<br />

peptide. However, it is also possible lck bound the peptide but was inactivated by the peptide<br />

and hence not detected. As expected no phosphopeptides were detected with the pFDRREF<br />

and control-coated beads.<br />

Figure 3.34 Analysis of the binding and phosphorylation of the YDRREY peptide by FAK, src, and lck<br />

Control (biotin-rrrrrrrrr-pptdqsrpvqpflnlttprkpr), pYDRREY and pFDRREF peptides bound to streptavidincoated<br />

Sepharose beads were mixed with 0.4 µg of recombinant FAK, lck, or src at 4°C for 2 hrs, and<br />

extensively washed. <strong>The</strong> beads were resuspended in kinase buffer with the addition of 1 µCi 32 Pγ-ATP at 30°C<br />

for 30 min for phosphorylation. After phosphorylation, the peptides were resolved on a 16% polyacrylamide tristricine<br />

SDS-gel. <strong>The</strong> gel was dried, and exposed to Kodak X-ray film, and developed. A single pYDRREY<br />

phosphopeptide was detected. Molecular weights of marker proteins are shown in the left-hand margin in kDa.<br />

This experiment was repeated three times.<br />

3.4.7 Binding of YDRREY peptide variants by FAK and src<br />

Peptide variants of the YDRREY motif were employed to define the structural features<br />

required for binding by FAK and src. YDRREY, pYDRREY, YEEEEY and YDGGEY<br />

peptides were immobilised onto Sepharose beads and mixed with recombinant FAK or src.<br />

<strong>The</strong> beads were extensively washed and subjected to an in vitro kinase assay. <strong>The</strong> peptides<br />

were resolved by SDS-PAGE, and peptide phosphorylation was detected by exposure to X-<br />

ray film. As shown in Figure 3.35, src phosphorylated the peptides YDRREY, pYDRREY,<br />

YEEEEY and YDGGEY to similar extents. In contrast, FAK only weakly phosphorylated the<br />

pYDRREY and YEEEEY peptides. This result suggests that the core region of the YDRREY<br />

motif influences binding and/or phosphorylation by FAK, but has little effect on<br />

binding/phosphorylation by src.<br />

120


Figure 3.35 FAK and src binding and phosphorylation of variants of the YDRREY peptide<br />

YDRREY, pYDRREY, YEEEEY, and YDGGEY peptides bound to streptavidin-Sepharose beads were mixed<br />

with 0.4 μg of either FAK or src at 4°C for 2 hrs and extensively washed. <strong>The</strong> beads were subjected to an in<br />

vitro kinase assay, and phosphopeptides were analysed on a 16% polyacrylamide Tris-tricine SDS- gel, dried<br />

and exposed to Kodak X-ray film, and developed. A single phosphopeptide of ~4 kDa was detected in each lane.<br />

Molecular weights of marker proteins are shown in the left-hand margin in kDa. This experiment was repeated<br />

twice.<br />

3.4.8 Does src remain active after binding to the YDRREY peptide?<br />

Src bound to the YDRREY peptide is still able to phosphorylate soluble unbound<br />

YDRREY peptide<br />

To further investigate the interaction between src and the YDRREY peptide, src was tested<br />

for activity when bound to the YDRREY motif. Src was bound to the peptides YDRREY,<br />

pYDRREY, pFDRREF, GGYDRREY and control peptide (biotin-rrrrrrrrrpptdqsrpvqpflnlttprkpr)<br />

immobilised on magnetic beads. <strong>The</strong> beads were washed thoroughly<br />

to remove unbound src, YDRREY peptide in solution was added or omitted, and the mixture<br />

was subjected to an in vitro kinase assay. <strong>The</strong> phosphorylated peptides were analysed by<br />

SDS-PAGE, as above. Src bound to each of the GGYDRREY, pFDRREF, pYDRREY and<br />

YDRREY peptides immobilised on magnetic beads and remained able to phosphorylate the<br />

added YDRREY peptide in solution (Figure 3.36). In the case of the immobilised pYDRREY<br />

and YDRREY peptides, phosphopeptides of pYDRREY and YDRREY could be seen without<br />

the addition of YDRREY in solution due to src phosphorylation of these peptides. In the case<br />

of pYDRREY the YDRREY in solution and bead-immobilised (pYDRREY) were readily<br />

distinguishable based on their differences in size. In the case of bead-immobilised YDRREY,<br />

there was an increase in phosphopeptide due to the addition of soluble YDRREY peptide. Src<br />

did not bind to the control peptide, and consequently the added YDRREY peptide in solution<br />

121


was not phosphorylated. YDRREY was directly phosphorylated by src and served as a<br />

positive experimental control. <strong>The</strong> results demonstrate that src binds to the F/YDRREY/F<br />

sequences, and remains fully active upon binding. <strong>The</strong> GGYDRREY peptide cannot be<br />

resolved by Tris-tricine-SDS-PAGE (data not shown), therefore no phosphopeptides of<br />

GGYDRREY were seen. However, soluble YDRREY was still phosphorylated by src that<br />

was bound to the GGYDRREY peptide.<br />

Figure 3.36 Src bound to the YDRREY motif remains active.<br />

<strong>The</strong> GGYDRREY, pFDRREF, pYDRREY, YDRREY and control (biotin-rrrrrrrrr-pptdqsrpvqpflnlttprkpr)<br />

peptides immobilised on magnetic beads were incubated with recombinant src (0.4 µg/ 10 µl of beads) at 4 °C<br />

for 2 hrs. <strong>The</strong> beads were extensively washed to remove unbound src kinase. YDRREY peptide (0.2 μg) in<br />

solution was added (+), or omitted (-), and the mixture subjected to an in vitro kinase assay at 30°C for 30 min.<br />

<strong>The</strong> beads were centrifuged and the phosphorylated peptides in the supernatant were analysed on 16%<br />

polyacrylamide Tris-tricine SDS-gels, as before. Soluble YDRREY directly phosphorylated by src served as a<br />

positive experimental control. Molecular weights of marker proteins are shown in the left-hand margin in kDa.<br />

This experiment was repeated twice.<br />

Does src remain peptide-bound after phosphorylating the YDRREY peptide?<br />

<strong>The</strong> above results suggest that src binds and phosphorylates the YDRREY peptide and<br />

remains active once bound, but does src remain attached once it has phosphorylated the<br />

peptide? Here this question was addressed. Autophosphorylated src was allowed to bind to<br />

the YDRREY or xDRREx peptides immobilised on magnetic beads, which were washed<br />

thoroughly to remove unbound kinases. Alternatively, unphosphorylated src was allowed to<br />

bind to the peptides immobilised on magnetic beads, which were then subjected to an in vitro<br />

kinase assay, and either left unwashed or washed extensively with PBS. Autophosphorylated<br />

forms of src remaining on the beads were analysed by SDS-PAGE (Figure 3.37).<br />

Autophosphorylated src (60 kDa) bound to both the YDRREY peptide (lane 1) and to the<br />

xDRREx phosphopeptide (lane 4). Similarly unphosphorylated src bound to the YDRREY<br />

122


peptide and was able to be detected in its autophosphorylated form in the absence of washing<br />

after it had phosphorylated the YDRREY peptide (lane 2), whereas in contrast<br />

unphosphorylated src did not bind to xDRREx (lane 5). Once src had phosphorylated the<br />

YDRREY peptide it was readily washed off the beads and could not be detected in its<br />

autophosphorylated form after resolving bead-bound proteins by SDS-PAGE (lane 3 and 6).<br />

Autophosphorylated src was included as a control for comparison (lane 7). <strong>The</strong> results<br />

suggest that src binds and phosphorylates the YDRREY peptide, but having phosphorylated<br />

the peptide it then detaches. Autophosphorylated and unphosphorylated forms of src are able<br />

to bind to unphosphorylated forms of the YDRREY peptide. In contrast, only<br />

autophosphorylated forms of src can bind to the phosphorylated YDRREY peptide.<br />

Figure 3.37 Src binds to YDRREY but not to xDRREx<br />

<strong>The</strong> peptides YDRREY and xDRREx immobilised on magnetic beads were incubated with 32 Pautophosphorylated<br />

(lanes 1 and 4) and unphosphorylated src (lanes 2, 3, 5 and 6). <strong>The</strong> beads were washed<br />

thoroughly to remove unbound src. <strong>The</strong> unphosphorylated src remaining bound to the beads was<br />

autophosphorylated by the addition of 32 P-γATP in kinase buffer and incubation at 30°C for 30 min. <strong>The</strong> beads<br />

were left unwashed (lanes 2 and 5) or washed thrice with PBS (lanes 3 and 6). Src remaining on the beads was<br />

resolved on a 10% polyacrylamide SDS-gel, and autoradiographed. Lane 7, autophosphorylated src. <strong>The</strong><br />

position of phosphorylated src of 60 kDa is indicated in the right-hand margin. Molecular weights of marker<br />

proteins are indicated in the left-hand margin in kDa. This experiment was repeated three times.<br />

123


3.4.9 FAK and YDRREY interactions<br />

Unphosphorylated and autophosphorylated forms of src bound to the YDRREY peptide, but<br />

only the autophosphorylated form bound to the phosphopeptide xDRREx (Figure 3.37). Here<br />

the aim was to determine whether unphosphorylated FAK would bind the xDRREx peptide.<br />

To test this notion, recombinant FAK was allowed to bind to the YDRREY and xDRREx<br />

peptides immobilised on magnetic beads. <strong>The</strong> beads were washed to remove unbound FAK,<br />

subjected to an in vitro kinase assay, and either left unwashed or washed with PBS. <strong>The</strong><br />

presence of autophosphorylated FAK at 125 kDa was analysed by SDS-PAGE (Figure 3.38).<br />

Unphosphorylated FAK bound to both the YDRREY and xDRREx peptides, albeit FAK<br />

bound substantially less well to the xDRREx peptide. Washing of the beads after FAK-<br />

mediated phosphorylation of the YDRREY peptide had little effect on the amount of<br />

phosphorylated FAK detected (compare lanes 1 and 3 with lanes 2 and 4). <strong>The</strong> results suggest<br />

that FAK can bind to the YDRREY peptide in both its tyrosine phosphorylated and<br />

unphosphorylated forms, albeit binding to the nonphosphorylated from is more substantial.<br />

Unlike src, FAK does not detach from the peptide once the peptide has been phosphorylated.<br />

Figure 3.38 Tyrosine phosphorylation of the YDRREY peptide affects the degree of FAK binding<br />

<strong>The</strong> peptides YDRREY and xDRREx immobilised on magnetic beads were incubated with unphosphorylated<br />

recombinant FAK, and washed thoroughly to remove unbound FAK. <strong>The</strong> beads were subjected to an in vitro<br />

kinase assay at 30°C for 30 min, and either left unwashed (lanes 1 and 3) or washed thrice with PBS (lanes 2 and<br />

4). FAK was resolved on a 10% polyacrylamide SDS-gel, and autoradiographed. <strong>The</strong> position of phosphorylated<br />

FAK at 125 kDa is indicated in the right-hand margin. Molecular weights of marker proteins are indicated in the<br />

left-hand margin in kDa. This experiment was repeated twice.<br />

124


3.4.10 FAK and src bind synergistically to the YDRREY peptide<br />

FAK and src both bind to the YDRREY peptide, and are known to bind and phosphorylate<br />

each other (Mitra et al. 2005), raising the hypothesis that they might bind synergistically to<br />

the YDRREY peptide. To test this hypothesis, FAK and src were added alone or together to<br />

the YDRREY peptide immobilised on beads. <strong>The</strong> beads were thoroughly washed, and<br />

subjected to an in vitro kinase assay. <strong>The</strong> phosphoproteins were resolved by SDS-PAGE and<br />

autoradiographed (Figure 3.39). <strong>The</strong> binding of src and FAK to the YDRREY peptide<br />

appeared to be enhanced when they were added together rather than singly. An alternative<br />

explanation to account for the increased intensity of the FAK and src bands when the kinases<br />

are added together is simply that the two kinases have phosphorylated one another. Further<br />

experiments are required to distinguish the two possibilities.<br />

Figure 3.39 Do FAK and src bind synergistically to the YDRREY peptide?<br />

<strong>The</strong> YDRREY peptide immobilised on magnetic beads was mixed with 0.4 µg of recombinant FAK and src,<br />

either alone or in combination. <strong>The</strong> beads were washed thoroughly, and subjected to an in vitro kinase assay.<br />

<strong>The</strong> phosphorylated kinases were resolved by SDS-PAGE and autoradiographed. <strong>The</strong> positions of the FAK and<br />

src phosphoproteins of 125 and 60 kDa, respectively, are indicated in the right-hand margin. Molecular weights<br />

of marker proteins are indicated in the left-hand margin in kDa. This experiment was repeated twice.<br />

125


3.5. <strong>The</strong> impact of cytoskeletal proteins on the interaction of the YDRREY<br />

peptide with kinases<br />

Here the aim was to determine the impact of cytoskeletal proteins on the interaction of the<br />

YDRREY peptide with FAK and src. Cytoskeletal proteins, including the actin-binding<br />

protein filamin, are phosphorylated by lck (Pal Sharma et al. 2004). Filamin binds to the β1,<br />

β2, and β7 integrin subunit cytoplasmic domains (Sharma et al. 1995; Calderwood et al.<br />

2001). Filamin binds to the β1 and β7 integrin subunits at a region which overlaps the central<br />

conserved NPxY motif. In contrast, filamin binds to the β2 subunit at a site within the<br />

membrane proximal region, overlapping the region where the YDRREY motif resides in the<br />

β7 subunit. Paxillin also binds to a membrane proximal region (Schaller et al. 1995a) in the<br />

β1 subunit, which also overlaps the region containing the YDRREY motif in the β7 subunit.<br />

3.5.1 Filamin disrupts the binding of src to the YDRREY peptide<br />

As several signalling proteins bind to one another and to overlapping regions of the integrin β<br />

subunit cytoplasmic domains, the aim here was to investigate whether the cytoskeletal<br />

proteins filamin and paxillin synergize or antagonize the binding and/or phosphorylation of<br />

the YDRREY peptide by src. <strong>The</strong> peptides pYDRREY and pFDRREF were immobilised on<br />

magnetic beads and mixed with a combination of recombinant filamin and src, paxillin and<br />

src, or src alone. <strong>The</strong> beads were washed thoroughly to remove unbound proteins, and<br />

subjected to an in vitro kinase assay to determine whether src bound to the peptides. Bead-<br />

bound protein was resolved by SDS-PAGE to detect the presence of src, which would be<br />

expected to be present in its autophosphorylated form (Figure 3.40). Filamin appeared to<br />

completely disrupt the binding of src to pYDRREY and pFDRREF, as autophosphorylated<br />

src was not detected. In contrast src bound to the pYDRREY and pFDRREF peptides in the<br />

presence of paxillin, albeit binding was slightly decreased. <strong>The</strong> apparent failure of src to bind<br />

to the pYDRREY and pFDRREF peptides in the presence of filamin was not due to filamin<br />

disruption of src autophosphorylation. Src incubated with filamin or paxillin in solution and<br />

subjected to an in vitro kinase assay still underwent autophosphorylation. <strong>The</strong>re is also the<br />

possibility that the band at approximately 60 kDa represents paxillin, or a combination of<br />

paxillin and src, as both molecules are of similar molecular weight. Distinguishing the various<br />

possibilities will require further investigation.<br />

126


Figure 3.40 Filamin disrupts the binding of src to the YDRREY peptide<br />

<strong>The</strong> peptides pFDRREF and pYDRREY immobilised on magnetic beads were mixed with a combination of 1 μg<br />

of recombinant filamin and 0.4 μg of src, a combination of 1 μg of recombinant paxillin and 0.4 μg of src, or 0.4<br />

μg of src alone at 4°C for 2 hrs. Unbound filamin, paxillin and src were removed by washing thrice with kinase<br />

buffer, and the beads subjected to an in vitro kinase assay at 30°C for 30 min. Src was incubated with filamin<br />

and paxillin in solution and subjected to an in vitro kinase assay to determine whether filamin and paxillin<br />

interfered with autophosphorylation of src (left-hand panel). <strong>The</strong> presence of autophosphorylated src, and src<br />

incubated with both filamin and paxillin were analysed by SDS-PAGE. Molecular weights of marker proteins<br />

are indicated in the left-hand margin in kDa. This experiment was repeated twice.<br />

3.5.2 Interaction of paxillin with the β7 subunit cytoplasmic domain<br />

As shown above, paxillin did not affect src binding to the YDRREY motif. Paxillin has been<br />

shown to bind to FAK after it is phosphorylated by ERK-2 (Liu et al. 2002). <strong>The</strong>refore, the<br />

aim here was to determine whether ERK-2-phosphorylated paxillin and FAK would form a<br />

complex and bind to the YDRREY peptide, or whether paxillin would antagonize the binding<br />

of FAK to the YDRREY peptide. Recombinant ERK-2, paxillin, and FAK were added<br />

together to the pYDRREY peptide or the control peptide (biotin-r9-pptdqsrpvqpflnlttprkpr)<br />

immobilised on magnetic beads. As controls, ERK-2 alone, and ERK-2 in combination with<br />

paxillin, were incubated with the pYDRREY and control peptides immobilised on magnetic<br />

beads. <strong>The</strong> beads were subjected to an in vitro kinase assay and thoroughly washed. <strong>The</strong><br />

bead-bound phosphoproteins were resolved by SDS-PAGE and autoradiographed (Figure<br />

3.41). Autophosphorylated ERK-2 (44 kDa) was not detected in any of the lanes.<br />

Recombinant FAK and paxillin when added in combination with ERK-2 bound to the-<br />

pYDRREY peptide, and were visualized as phosphoproteins of 125 kDa, and 60 kDa,<br />

respectively. Binding of FAK was much weaker when ERK-2 was omitted, suggesting either<br />

that ERK-2 is required to facilitate the binding of paxillin. Thus, the experiment shows that a<br />

trimer of FAK bound to pYDRREY and paxillin bound to FAK is formed.<br />

127


Figure 3.41 Erk-phosphorylated paxillin binds to FAK and forms a complex with the YDRREY peptide.<br />

<strong>The</strong> pYDRREY peptide and the control peptide (biotin-r9-pptdqsrpvqpflnlttprkpr) were immobilised on<br />

magnetic beads and mixed with recombinant ERK-2, paxillin and FAK, either alone or in combination as<br />

indicated. <strong>The</strong> beads were subjected to an in vitro kinase assay, washed thoroughly, and the bound<br />

phosphoproteins resolved by SDS-PAGE. <strong>The</strong> SDS-PAGE gel was dried and exposed at -80°C to x-ray film<br />

overnight (left panel) or for 6 hrs (right panel). <strong>The</strong> positions of FAK (125 kDa) and paxillin (60 kDa) are<br />

indicated in the right-hand margin. Molecular weights of marker proteins are indicated in the left-hand margin in<br />

kDa. This experiment was repeated twice.<br />

3.5.3 FAK, src, and α-actinin bind the YDRREY peptide<br />

<strong>The</strong> cytoskeletal protein α-actinin has previously been reported to bind the cytoplasmic<br />

domains of the β1 and β2 integrin subunits (Otey et al. 1990; Pavalko et al. 1993; Sampath et<br />

al. 1998). α-actinin binds to src, and is phosphorylated by FAK (Izaguirre et al. 2001), raising<br />

the possibility that the kinases indirectly mediate the interaction of α-actinin with the β7<br />

cytoplasmic domain. In this experiment, α-actinin was subjected to phosphorylation by FAK<br />

and src and tested for binding to the YDRREY peptide or the control peptide (biotin-r9-<br />

pptdqsrpvqpflnlttprkpr) which had been coated on magnetic beads. <strong>The</strong> beads were washed<br />

thoroughly, and the bound phosphoproteins were resolved by SDS-PAGE (Figure 3.42). α-<br />

actinin was phosphorylated by both FAK and src, as evidenced by the presence of a 100 kDa<br />

α-actinin band when either FAK (125 kDa) or src (60 kDa) were present. <strong>The</strong> complex of<br />

FAK in combination with α-actinin bound weakly to the YDRREY peptide, compared to the<br />

complex of src and α-actinin. When FAK, src, and actinin were added in combination, it<br />

appeared that the complex that bound the YDRREY peptide was predominantly composed of<br />

src and α-actinin. However, the presence of FAK may have been obscured by the background.<br />

128


FAK and α-actinin did not bind to the control peptide. A faint 60 kDa protein bound the<br />

control peptide, which may correspond to peptide-bound src or more likely represents trace<br />

amounts of src from incompletely washed beads. <strong>The</strong> identity of the band which migrated<br />

with a molecular weight of ~40 kDa is not known. <strong>The</strong> results suggest that α-actinin can<br />

potentially bind indirectly to the β7 CARD via FAK or src.<br />

Figure 3.42 Binding of src, FAK, and α-actinin to the YDRREY peptide.<br />

Recombinant FAK and src were incubated either alone or together with 1 µg of recombinant α-actinin and the<br />

mix subjected to an in vitro kinase assay. <strong>The</strong> phosphoproteins were mixed with either the control peptide<br />

(biotin-r9-pptdqsrpvqpflnlttprkpr) or the YDRREY peptide immobilised on magnetic beads, and incubated at<br />

4°C for 2 hrs. <strong>The</strong> beads were washed thoroughly, and the bound phosphoproteins were resolved by SDS-<br />

PAGE. <strong>The</strong> positions of FAK (125 kDa), α-actinin (100 kDa), and src (60 kDa) are indicated in the right-hand<br />

margin. <strong>The</strong> identity of the ~40 kDa band is not known. Molecular weights of marker proteins are indicated in<br />

the left-hand margin in kDa.<br />

3.6. In vivo interactions of β7 integrins with kinases<br />

3.6.1 Co-localisation of a fluoresceinated YDRREY peptide in living cells<br />

After showing that FAK and src interact with the YDRREY motif in an in vitro system, the<br />

next step was to investigate these interactions in living cells. <strong>The</strong> r9-YDRREY peptide was<br />

fluoresceinated by attachment of a 5-FAM green fluorescent molecule to the 5’-end of the<br />

peptide (FAM-YDRREY). TK-1 cells were incubated with a low concentration of the FAM-<br />

YDRREY peptide (1 µM) in order to prevent disruption by the peptide of α4β7-mediated<br />

binding of TK-1 cells to MAdCAM-1. <strong>The</strong> cells were washed, activated with Mn 2+ , incubated<br />

129


with MAdCAM-1-coated microspheres, fixed with paraformaldehyde and immunostained<br />

with anti-kinase antibodies to observe whether the FAM-YDRREY peptide would localize to<br />

focal adhesions (Figure 3.43). <strong>The</strong> FAM-YDRREY peptide (green) localised to focal<br />

adhesions at the cell-surface (Figure 3.43A, B). Src (Figure 3.43A) and FAK (Figure 3.43B;<br />

both red), also were concentrated at focal adhesions, and when the images were merged it was<br />

clear that both FAK and src colocalized with the FAM-YDRREY peptide at the focal<br />

adhesions (indicated by the arrows in the merged images). Thus, the YDRREY peptide may<br />

bind FAK and src in vivo.<br />

Figure 3.43 <strong>The</strong> YDRREY peptide co-localises with FAK and src at focal adhesions in vivo<br />

TK-1 cells were pre-incubated with a fluoresceinated cell-permeable FAM-YDRREY (1 μM) peptide for 30<br />

min, activated with Mn 2+ , and bound to MAdCAM-1-Fc-coated polystyrene microspheres (0.95 μm). <strong>The</strong> cells<br />

were fixed with paraformaldehyde, and immunostained with rabbit anti-src (A) and rabbit anti-FAK (B)<br />

antibodies (both used at 1:100 dilutions). <strong>The</strong> primary antibodies were detected using an anti-rabbit-AF594<br />

secondary antibody (1:250), and visualised by confocal microscopy. <strong>The</strong> green fluorescence indicates the<br />

location of the FAM-YDRREY peptide; the red fluorescence denotes the locations of FAK and src. <strong>The</strong> merged<br />

photo shows the colocalisation of the peptide with src (A) and FAK (B) as indicated by the yellow coloration.<br />

Focal adhesions are donoted by blue arrows. <strong>The</strong> bar represents 10 µm.<br />

130


3.6.2 Phosphorylation of the integrin β7 subunit in vivo<br />

<strong>The</strong> above results have shown that the YDRREY motif of the cytoplasmic domain of the<br />

integrin β7 subunit specifically binds to FAK and src in vitro, and possibly in vivo. In<br />

addition, a variety of cytoskeletal proteins were found to influence the binding and<br />

phosphorylation of the YDRREY motif by FAK and src. <strong>The</strong> NPxY motif within the<br />

cytoplasmic domain of the β3 subunit of αIIbβ3 is phosphorylated during platelet aggregation<br />

(Law et al. 1996; Phillips et al. 2001). <strong>The</strong> question addressed here is whether the β7<br />

cytoplasmic domain is phosphorylated on tyrosine in vivo.<br />

<strong>The</strong> YDRREY peptide and a peptide encompassing the full-length β7 cytoplasmic domain<br />

were subjected to an in vitro kinase assay with trace amounts of cell lysate prepared from<br />

activated and non-activated TK-1 cells. Surprisingly the peptides were phosphorylated to the<br />

same extent by the two lysates (data not shown). To partly address the question raised above,<br />

a comparison was made of the tyrosine phosphorylation of non-activated β7 integrins,<br />

activated β7 integrins, and activated ligand-bound β7 integrins. <strong>The</strong> β7 integrins from<br />

activated and non-activated TK-1 cells were immunoprecipitated and immunoblotted with<br />

anti-phosphotyrosine antibodies. TK-1 cells were initially inactivated by treatment with the<br />

metal ion chelator, EDTA, and the cells were either left unactivated or were activated with<br />

Mn 2+ . <strong>The</strong> activated TK-1 cells were incubated with MAdCAM-1-coated magnetic beads to<br />

generate ligand-bound β7 integrins. <strong>The</strong> above cells representing three different states of β7<br />

integrin activation were lysed in the presence of sodium orthovanadate and phosphatase<br />

inhibitors (PhosSTOP, Roche). <strong>The</strong> soluble fraction was immediately denatured and<br />

immunoprecipitated with anti-β7 antibodies recognising the β7 cytoplasmic domain that had<br />

been crosslinked to Sepharose beads. <strong>The</strong> immunoprecipitated β7 integrin subunit was<br />

resolved by SDS-PAGE and immunoblotted with an anti-phosphotyrosine antibody (Figure<br />

3.44A, upper panel). β7 immunoprecipitated from unactivated cells was weakly<br />

phosphorylated, giving a single band of approximately 100 kDa (lane 1). In contrast, β7<br />

immunoprecipitated from activated cells, including cells incubated with MAdCAM-1 was<br />

more strongly phosphorylated, giving a major band of 100 kDa, and a more diffuse band of<br />

130 kDa (lanes 2 and 3, respectively).<br />

131


<strong>The</strong> membrane was stripped and re-probed with the anti-β7 cytoplasmic domain antibody that<br />

had been used for the immunoprecipitation. Despite the fact that the antibody had been<br />

covalently cross-linked to Sepharose beads, there is always trace amounts that leak off the<br />

beads when boiled in reducing SDS-loading buffer. Hence the secondary antibody used to<br />

detect the β7 antibody recognised trace amounts of the antibody from the<br />

immunoprecipitation, as evidenced by the reduced antibody bands at 50 and 70 kDa (Figure<br />

3.44A, lower panel). For comparison, an aliquot of the β7 antibody was resolved on SDS-<br />

PAGE, confirming the bands at 50 and 70 kDa represent antibody heavy chains (Figure<br />

3.44B).<br />

<strong>The</strong> β7 subunit appeared to migrate as two bands of similar intensity of 100 and 130 kDa<br />

irrespective of the activation status of the cells (Figure 3.44A, lanes 1, 2 and 3, respectively).<br />

<strong>The</strong> presence of two β7 subunit bands is in accord with the fact that the β7 subunit exists as<br />

an immature unglycosylated form (Yuan et al. 1990) and as a larger mature cell-surface<br />

glycosylated form. It is interesting therefore that cells have to be activated in order for<br />

phosphorylation of the mature form to occur. <strong>The</strong> immature form appears to be constitutively<br />

phosphorylated. <strong>The</strong>re may have been a slight increase in phosphorylation of both forms upon<br />

ligand binding/clustering of α4β7 integrins, but the increase was negligible.<br />

132


Figure 3.44 Tyrosine phosphorylation of the β7 subunit in vivo is dependent on the activation status of<br />

cells<br />

TK-1 cells were treated with EDTA to inactivate α4β7and were either left non-activated, or activated with Mn 2+ .<br />

Activated cells were either left unbound or bound to MAdCAM-1 immobilised on magnetic beads. <strong>The</strong> TK-1<br />

cells were lysed in the presence of phosphatase inhibitors, and the soluble protein fractions were collected and<br />

denatured at 95°C for 5 min. <strong>The</strong> denatured protein cell lysates from TK-1 cells that were inactivated (lane 1),<br />

activated (lane 2) and activated in the presence of the ligand MAdCAM-1 (lane 3) were immunoprecipitated<br />

with a rabbit antibody against the cytoplasmic domain of the integrin β7 subunit which had been crosslinked to<br />

Sepharose beads. <strong>The</strong> immunoprecipitates were resolved by SDS-PAGE, and immunoblotted with a mouse antiphosphotyrosine<br />

(1:200) antibody. Antibody reactivity was detected using a goat anti-mouse-HRP secondary<br />

antibody (1:80,000; A, upper panel). Phosphotyrosine bands of approximately 100 and 130 kDa as indicated<br />

are the approximate size of immature and mature forms of the integrin β7 subunit.<br />

<strong>The</strong> membrane was stripped with NaOH and probed with the rabbit anti-β7 cytoplasmic domain antibody<br />

(1:200), which was detected using a goat anti-rabbit-HRP antibody (1:20,000; A, lower panel). <strong>The</strong> integrin β7<br />

subunit was resolved as its immature and mature forms of 100 and 130 kDa, respectively, as indicated. Antibody<br />

chains of 50 and 70 kDa, which presumably leached from the antibody-Sepharose matrix were also detected by<br />

the secondary goat anti-rabbit-HRP antibody. <strong>The</strong> latter confirms that equal amounts of antibody beads were<br />

used in the immunoprecipitations. B) shows β7 antibody alone resolved by SDS-PAGE and Western blotted as<br />

in A (lower panel). Molecular weights of marker proteins are indicated in the left-hand margin in kDa. <strong>The</strong><br />

experiments were performed in duplicate.<br />

133


3.6.3 Confirmation of FAK and src involvement in β7 integrin signalling<br />

<strong>The</strong> tyrosine kinases FAK and src bind and phosphorylate the YDRREY peptide in vitro and<br />

possibly in vivo. To unequivocally determine whether FAK and src bind the β7 subunit in<br />

living cells, the β7 subunit was immunoprecipitated from TK-1 cells, and immunoblotted for<br />

the presence of the tyrosine kinases. Thus the soluble fraction from a TK-1 cell lysate was<br />

incubated with the anti-β7 antibody Fib504 (rat-IgG2a) cross-linked to Sepharose beads, and<br />

with a control rat IgG antibody cross-linked to Sepharose beads. <strong>The</strong> immunoprecipitated<br />

proteins were resolved by SDS-PAGE, transferred onto a PVDF membrane, and<br />

immunoblotted with a rabbit antibody against the β7 subunit cytoplasmic domain (Figure<br />

3.45).<br />

<strong>The</strong> Fib504 mAb immunoprecipitated the β7 subunit from the soluble fraction of the TK-1<br />

cell lysate (Figure 3.45, lane 5), as evidenced by an intense band of approximately 150 kDa<br />

detected by the rabbit polyclonal antibody against the β7 cytoplasmic domain. Direct<br />

immunoblotting of the soluble fraction of the TK-1 lysate (lane 1) gave bands at ~150 kDa,<br />

which may also represent the β7 subunit. No bands of 150 kDa were detected in control<br />

samples, including rat-IgG-Sepharose alone (lane 2), the immunoprecipitate formed with rat-<br />

IgG-Sepharose (lane 3), Fib504-conjugated Sepharose beads alone (lane 4), non-conjugated<br />

Fib504 antibody (lane 6), non-conjugated rat-IgG antibody (lane 7), Sepharose beads alone<br />

(lane 8), or the insoluble fraction of the TK-1 cell lysate (lane 9). <strong>The</strong> results indicate that the<br />

β7 integrin was specifically immunoprecipitated by the Fib504 mAb.<br />

134


Figure 3.45 <strong>The</strong> Fib504 mAb specifically immunoprecipitates the β7 subunit from TK-1 cells<br />

<strong>The</strong> β7 integrin subunit was immunoprecipitated from the soluble fraction of a TK-1 cell lysate (10 7 cells/mL)<br />

with the anti-β7 subunit mAb Fib504 cross-linked to Sepharose. <strong>The</strong> immunoprecipitate was resolved by SDS-<br />

PAGE and immunoblotted with a rabbit polyclonal antibody against the β7 cytoplasmic domain (1:400). A goat<br />

anti-rabbit-IgG-HRP secondary antibody (1:20,000) was used to detect immunoreactivity. Lane 1, soluble<br />

fraction of the TK-1 lysate; lane 2, control rat-IgG antibody-conjugated Sepharose beads; lane 3, control rat-IgG<br />

antibody-conjugated Sepharose beads incubated with TK-1 lysate; lane 4, Fib504-conjugated Sepharose beads;<br />

lane 5, Fib504-conjugated Sepharose incubated with TK-1 lysate; lane 6, non-conjugated Fib504 antibody; lane<br />

7, non-conjugated control rat-IgG antibody; lane 8, Sepharose beads; lane 9, insoluble fraction of the TK-1<br />

lysate. <strong>The</strong> band in lane 5 at approximately 150 kDa is the β7 integrin subunit. Molecular weights of marker<br />

proteins are indicated in the left-hand margin in kDa. This experiment was repeated twice.<br />

135


3.6.4 Detection of src in β7 integrin immunoprecipitates<br />

After establishing the presence of the β7 integrin in the immunoprecipitate formed with the<br />

Fib504 mAb as shown in Figure 3.45, the membrane was stripped of protein with NaOH, and<br />

immunoblotted with an anti-src antibody (Figure 3.46). A prominent band at 60 kDa was<br />

detected in the β7 immunoprecipitate (Figure 3.46, lane 5). It was also present in the soluble<br />

(lane 1) and insoluble (lane 9) fractions of TK-1 cell lysate, respectively. In contrast, the 60<br />

kDa band was not present in any of the control samples. Additional minor bands appear to<br />

bind non-specifically to antibody as evidenced by their immunoprecipitation with rat-IgG-<br />

Sepharose (lane 3). This result establishes that src binds the β7 integrin expressed in cells.<br />

Figure 3.46 Src is coimmunoprecipitated with the β7 integrin from TK1 cell lysates<br />

<strong>The</strong> blot in Figure 3.45 was stripped of protein with 0.2 mM NaOH, and probed with a rabbit anti-src antibody<br />

(1:200), followed by a goat anti-rabbit-HRP secondary mAb. <strong>The</strong> region of the gel between 75 and 50 kDa<br />

where src should migrate was magnified (inset). <strong>The</strong> lanes are as specified in Figure 3.45. Molecular weights of<br />

marker proteins are indicated in the left-hand margin in kDa. This experiment was repeated twice.<br />

136


3.6.5 Detection of FAK in β7 integrin immunoprecipitates<br />

<strong>The</strong> membrane in Figure 3.45 was re-stripped with NaOH and probed for FAK with an anti-<br />

FAK antibody (Figure 3.47). FAK was detected as a doublet of bands at approximately 100<br />

and 75 kDa (Figure 3.47, lane 5). Bands of similar size were present in the soluble fraction,<br />

but not the insoluble fraction, of a TK-1 cell lysate. A number of lower molecular weight<br />

doublet bands in the soluble fraction of the lysate may signify that FAK is partially degraded.<br />

FAK is known to be cleaved by the protease calpain and caspases into smaller products of<br />

approximately 95 kDa and 77 kDa (Wen et al. 1997; Carragher et al. 2001). This result<br />

establishes that FAK binds the β7 integrin expressed in cells.<br />

Figure 3.47 FAK is coimmunoprecipitated with the β7 integrin from TK1 lysates<br />

<strong>The</strong> blot in Figure 3.46 was stripped of protein with 0.2 mM NaOH, and probed with a rabbit anti-FAK<br />

antibody (1:200), followed by a goat anti-rabbit-HRP secondary antibody. <strong>The</strong> region of the gel between 60 and<br />

140 kDa where FAK should migrate was magnified (inset). <strong>The</strong> lanes are as specified in Figure 3.45. Molecular<br />

weights of marker proteins are indicated in the left-hand margin in kDa. This experiment was repeated twice.<br />

137


3.6.6 Localisation of interacting kinases in living cells<br />

FAK and src were detected in immunoprecipitates of the β7 integrin, suggesting that FAK<br />

and src complex with β7 integrins in vivo. <strong>The</strong> aim here was to colocalize FAK and src with<br />

α4β7 in living TK-1 cells. Activated TK-1 cells allowed to bind to MAdCAM-1-coated<br />

surfaces were polarised and migrated randomly (data not shown). TK-1 cells were activated<br />

with Mn 2+ and bound to glass cover slips coated with MAdCAM-1 at a low concentration to<br />

enable the formation of focal adhesions and cell migration. <strong>The</strong>y were allowed to bind and<br />

spread, then fixed and stained with an antibody against the β7 integrin and antibodies<br />

recognizing either FAK (Figure 3.48), src (Figure 3.49), or lck (Figure 3.50). <strong>The</strong> β7<br />

integrin was detected with the green fluorophore AF-488, whereas the kinases were detected<br />

with the red fluorophore AF-594. <strong>The</strong> cells were stained with isotype control antibodies, but<br />

they failed to stain the cells (data not shown).<br />

Figure 3.48 FAK and α4β7 do not strongly colocalize on TK-1 cells<br />

TK-1 cells were activated with Mn 2+ and adhered to coverslips coated with MAdCAM-1 at 2.5 µg/mL. <strong>The</strong> cells<br />

were permeabilised and co-stained with anti-β7 (1:100) and anti-FAK (1:100) antibodies. <strong>The</strong> β7 subunit was<br />

visualised with the green fluorophore AF-488 (1:250), while FAK was visualised with the red fluorophore AF-<br />

594 (1:250). <strong>The</strong> images were merged to detect colocalization (merged). Uropods are indicated by the blue<br />

arrow. Images were captured using epi-fluorescence microscopy. Bar indicates the length of 10 µm. This<br />

experiment was repeated three times.<br />

Both α4β7 and FAK were diffusely expressed throughout cells, but with multiple points of<br />

intense staining giving a punctate pattern of expression (Figure 3.48). α4β7 was strongly<br />

expressed on the uropod while FAK was only very weakly expressed, and largely at pinpoints<br />

of expression. <strong>The</strong> results suggest that α4β7 does not colocalize with FAK on the uropod, and<br />

does not strongly colocalize with FAK on the cell body<br />

138


Src was diffusely expressed throughout the cells, but was also expressed in a punctate<br />

fashion, albeit not as marked as for the β7 integrin. Unlike FAK, src was expressed on the<br />

uropod, where expression largely overlapped that of the α4β7 (Figure 3.49). In contrast,<br />

expression of both molecules in the cell body was mostly punctate, and did not overlap as in<br />

the uropod.<br />

Figure 3.49 Src appears to colocalize with α4β7 in the uropod<br />

TK-1 cells were activated with Mn 2+ and adhered to coverslips coated with MAdCAM-1 at 2.5 µg/mL. <strong>The</strong> cells<br />

were permeabilised and stained with anti-β7 (1:100) and anti-src (1:100) antibodies. <strong>The</strong> β7 integrin was<br />

visualised with the green fluorophore AF-488 (1:250), while src was visualised with the red fluorophore AF-594<br />

(1:250). <strong>The</strong> images were merged to detect colocalization (merged). Uropods are indicated by the blue arrow.<br />

Images were captured using epi-fluorescence microscopy. Bar indicates the length of 10 µm. This experiment<br />

was repeated three times.<br />

Lck was diffusely expressed throughout the cells, and like FAK was poorly expressed on the<br />

uropod in a punctate fashion (Figure 3.50). Lck was also expressed in a fine punctate<br />

expression in the cell body. <strong>The</strong>re was no apparent overlap with α4β7.<br />

139


Figure 3.50 Lck and α4β7 do not appear to colocalize on TK-1 cells<br />

TK-1 cells were activated with Mn 2+ and adhered to coverslips coated with MAdCAM-1 at 2.5 µg/mL. <strong>The</strong> cells<br />

were permeabilised and stained with anti-β7 (1:100) and anti-lck (1:100) antibodies. <strong>The</strong> β7 integrin was<br />

visualised with the green fluorophore AF-488 (1:250), whereas lck was visualised with the red fluorophore AF-<br />

594 (1:250). <strong>The</strong> images were merged to detect colocalization (merged). Panel B shows an enlargement of the<br />

uropod of a cell. Uropods are indicated by the blue arrow. Images were captured using epi-fluorescence<br />

microscopy. Bar indicates the length of 20 µm. This experiment was repeated three times.<br />

3.7. Localization of α4β7 with kinases in supra-molecular activation<br />

complexes (SMAC)<br />

SMACs are clusters of receptors and signalling molecules formed on cells when TCR and<br />

integrins on lymphocytes bind their ligands expressed on APCs. Preliminary experiments in<br />

the laboratory had led to the discovery that integrin ligands immobilised on polystyrene<br />

microspheres could cause the clustering of β7 integrins at the cell-surface and the formation<br />

of SMAC-like structures when incubated with activated TK-1 cells (Zhang 1999, unpublished<br />

data). <strong>The</strong> pseudo-SMAC could be visualised by staining of the β7 integrins with a<br />

fluorescent anti-β7 subunit mAb.<br />

140


<strong>The</strong> aim here was to determine whether src, FAK and lck might co-localise with α4β7 in the<br />

pseudo-SMAC structures. Integrin clustering serves as a useful platform to detect co-<br />

localisation of proteins, as the complexes formed are readily detectable by light microscopy.<br />

For this study, TK-1 cells were activated and incubated with MAdCAM-1-coated magnetic<br />

beads. <strong>The</strong> SMAC that formed were stained with antibodies against the β7 integrin subunit,<br />

and the tyrosine kinases src, FAK and lck, and analyzed by confocal microscopy.<br />

3.7.1 TK-1 cells bind to MAdCAM-1-coated magnetic beads<br />

Mn 2+ -activated TK-1 cells were incubated with MAdCAM-1-Fc coated magnetic beads and<br />

binding analyzed by light microscopy. <strong>The</strong> cells were seen to bind to the beads, where some<br />

cells appeared to bind to more than one magnetic bead (Figure 3.51B). Similar results were<br />

achieved with beads ranging in size from 0.95 μm to 5 μm in diameter (data not shown).<br />

Magnetic beads that had not been coated with MAdCAM-1 did not bind to Mn 2+ -activated<br />

TK-1 cells (Figure 3.51A). In addition, non-activated TK-1 cells did not bind to MAdCAM-<br />

1-coated magnetic beads (data not shown).<br />

Figure 3.51 TK-1 cells bind to MAdCAM-1-coated magnetic beads<br />

TK-1 cells were activated with Mn 2+ and incubated with (A) non-coated and (B) MAdCAM-1-Fc-coated (100 ng<br />

MAdCAM-1/μL beads) magnetic beads at RT for 30 min. <strong>The</strong> cells were fixed in 4% paraformaldehyde, placed<br />

on glass slides and visualised under light microscopy. 40x magnification. This experiment was repeated three<br />

times.<br />

141


3.7.2 α4β7 colocalizes with src in SMACs<br />

Mn 2+ -activated TK-1 cells were incubated with 0.9 μm diameter polystyrene microspheres<br />

coated with MAdCAM-1 for molecular colocalization experiments as 0.9 μm diameter<br />

microspheres provide a smaller point of contact for clustering of integrins. In addition,<br />

polystyrene microspheres produce less auto-fluorescence as compared to magnetic beads or<br />

Sepharose beads (data not shown). <strong>The</strong> cells were fixed and stained with green and red<br />

fluorescent antibodies against the β7 subunit and src, respectively, and analyzed by confocal<br />

microscopy. α4β7 (green) clustered in distinct focal adhesions around the plasma membrane,<br />

whereas src (red) occupied the entire plasma membrane but also clustered in focal adhesions<br />

within the membrane (Figure 3.52A). Merger of the two images clearly illustrated colocalisation<br />

(yellow) of src and β7 integrins in the adhesion clusters. <strong>The</strong> image of one<br />

representative cell was magnified to better visualize the SMAC (Figure 3.52B).<br />

Several of the β7 clusters appeared to have a circular structure resembling pSMAC-like<br />

structures Figure 3.52B. In contrast, src appeared to occupy both the central SMAC<br />

(cSMAC) and peripherial SMAC (pSMAC).<br />

142


Figure 3.52 α4β7 clusters with src in pseudo-SMAC<br />

Polystyrene microspheres (0.95 μm diameter) coated with MAdCAM-1 (100 ng/μL) were incubated with Mn 2+ -<br />

activated TK-1 cells at RT for 30 min. <strong>The</strong> cells were stained with the rat-anti-β7 mAb Fib504 (1:100) and a<br />

rabbit-anti-src antibody (1:100). <strong>The</strong> anti-β7 antibody was detected with a secondary anti-rat-AF488 (green)<br />

antibody, and the anti-src antibody was detected with a secondary anti-rabbit-AF594 (red) antibody. Stained<br />

cells were analysed by confocal microscopy. (A) A representative population of TK-1 cells incubated with<br />

MAdCAM-1-coated microspheres and stained with anti-β7 (green) and anti-src (red) antibodies. <strong>The</strong> two images<br />

were merged to aid visualization of colocalization (merged). (B) A representative single cell was enlarged for<br />

clearer visualisation. Pseudo-SMACs are indicated by blue arrows. Bar indicates the length of 20 µm. This<br />

experiment was repeated three times.<br />

3.7.3 α4β7 colocalizes with FAK in SMACs<br />

TK-1 cells were activated with Mn 2+ , incubated with MAdCAM-1-coated microspheres to<br />

cluster α4β7 integrins, and stained with antibodies against the β7 subunit and FAK.<br />

Fluorescent green and red secondary antibodies were used to detect the primary anti-β7 and<br />

anti-FAK antibodies, respectively, and visualized by confocal microscopy (Figure 3.53).<br />

Both α4β7 (green) and FAK (red) clustered in distinct focal adhesions around the plasma<br />

143


membrane (Figure 3.53A). Colocalization of α4β7 with FAK was obvious even without<br />

having to merge the two images. Nevertheless merger confirmed colocalization.<br />

High magnification of a representative single cell revealed that clusters of α4β7 formed<br />

distinct circular pSMAC-like structures (Figure 3.52B) FAK weakly localised to the<br />

pSMAC-like structures indicating co-localisation with α4β7. It was present at much higher<br />

levels within the cSMAC.<br />

Figure 3.53 α4β7 clusters with FAK in pseudo-SMAC<br />

Polystyrene microspheres (0.95 μm diameter) coated with MAdCAM-1 (100 ng/μL) were incubated with Mn 2+ -<br />

activated TK-1 cells at RT for 30 min. <strong>The</strong> cells were stained with the rat-anti-β7 mAb Fib504 (1:100) and a<br />

rabbit-anti-FAK antibody (1:100). <strong>The</strong> anti-β7 antibody was detected with a secondary anti-rat-AF488 (green)<br />

antibody, and the anti-src antibody was detected with a secondary anti-rabbit-AF594 (red) antibody. Stained<br />

cells were analysed by confocal microscopy. (A) A representative population of TK-1 cells incubated with<br />

MAdCAM-1-coated microspheres and stained with anti-beta7 (green) and anti-FAK (red) antibodies. <strong>The</strong> two<br />

images were merged to aid visualization of colocalization (merged). (B) A representative single cell was<br />

enlarged for clearer visualisation. Pseudo-SMACs are indicated by blue arrows. Bar indicates the length of 20<br />

µm. This experiment was repeated three times.<br />

3.7.4 α4β7 colocalizes with lck in SMACs<br />

TK-1 cells were activated with Mn 2+ , incubated with MAdCAM-1-coated microspheres to<br />

cluster α4β7 integrins, and stained with antibodies against the β7 subunit and lck. Fluorescent<br />

144


green and red secondary antibodies were used to detect the primary anti-β7 and anti-lck<br />

antibodies, respectively, and visualized by confocal microscopy (Figure 3.54). Both α4β7<br />

(green) and lck (red) clustered in distinct focal adhesions around the plasma membrane<br />

(Figure 3.54A).<br />

High magnification of a representative single cell revealed that clusters of α4β7 formed<br />

distinct circular pSMAC-like structures (Figure 3.54B). Lck predominantly localised to the<br />

cSMAC, suggesting it was unlikely to associate with α4β7.<br />

Figure 3.54 Clusters of integrin β7 and lck<br />

Polystyrene microspheres (0.95 μm diameter) coated with MAdCAM-1 (100 ng/μL) were incubated with Mn 2+ -<br />

activated TK-1 cells at RT for 30 min. <strong>The</strong> cells were stained with the rat-anti-β7 mAb Fib504 (1:100) and a<br />

rabbit-anti-lck antibody (1:100). <strong>The</strong> anti-β7 antibody was detected with a secondary anti-rat-AF488 (green)<br />

antibody, and the anti-lck antibody was detected with a secondary anti-rabbit-AF594 (red) antibody. Stained<br />

cells were analysed by confocal microscopy. (A) A representative population of TK-1 cells incubated with<br />

MAdCAM-1-coated microspheres and stained with anti-beta7 (green) and anti-lck (red) antibodies. <strong>The</strong> two<br />

images were merged to aid visualization of colocalization (merged). (B) A representative single cell was<br />

enlarged for clearer visualisation. Pseudo-SMACs are indicated by blue arrows. Bar indicates the length of 20<br />

µm. This experiment was repeated three times.<br />

145


3.8. Mutation of the β7 cytoplasmic domain and effects on cell adhesion<br />

<strong>The</strong> above results suggested that the β7 subunit CARD associates with the tyrosine kinases<br />

src and FAK both in vitro and in vivo. To further investigate structure-function properties of<br />

the β7 CARD in vivo, the next aim was to examine the effect of mutated tyrosines 736 and<br />

741 in the β7 CARD on α4β7 integrin adhesion. <strong>The</strong> third downstream tyrosine residue at<br />

position 761 was also mutated for comparison. <strong>The</strong> tyrosine residues were substituted with<br />

phenylalanines in the full-length β7 sequence as described previously in <strong>Section</strong> 3.4.1.<br />

Plasmids encoding the mutant β7 subunit were transfected together with a plasmid encoding<br />

the wild-type α4 integrin subunit into HEK-293T cells to analyze α4β7-mediated cell<br />

adhesion to MAdCAM-1.<br />

3.8.1 Cloning of mutated variants of the integrin β7 subunit into mammalian expression<br />

vectors<br />

<strong>The</strong> initial aim was to clone wild-type integrin β7 cDNA and mutated variants into a<br />

mammalian expression vector. <strong>The</strong> cloning strategies are summarised in Figure 3.55. <strong>The</strong> full<br />

length wild-type human β7 (β7wt) cDNA was excised from the pcDM8-huβ7 vector with<br />

HindIII and NotI and ligated it into a pcDNA3.1 H+ vector (Figure 3.54, Step 1). DNAs<br />

encoding variant forms of the integrin β7 cytoplasmic domains carrying the mutations Y736F,<br />

Y741F, Y761F and Y736F+741F had previously been cloned into pGEX-2T vectors by lab<br />

colleagues (refer to <strong>Section</strong> 3.4.1). <strong>The</strong>y were amplified by PCR (Step 2), and the resultant<br />

products were used in an overlap PCR reaction with a fragment of β7wt cut from the<br />

pcDNA3.1 H+ vector by the restriction enzymes SacI and NotI (Step 3). <strong>The</strong> overlap PCR<br />

products carrying the mutated tyrosines (Step 4) were ligated in a 3-way ligation with an<br />

empty pcDNA3.1 H+ vector and a cDNA fragment containing the remaining portion of wild-<br />

type β7 cDNA to form a plasmid carrying the complete mutated full-length β7 cDNA (Step<br />

5). All clones were sequenced to confirm the integrity of the cloned sequence (data not<br />

shown).<br />

In addition, the full-length β7 cDNAs were subcloned between the restriction sites NheI and<br />

XhoI of the pIRES2-EGFP vector (Step 6). <strong>The</strong> pIRES2-EGFP vector expresses the enhanced<br />

green fluorescent protein (EGFP), and also contains an internal ribosomal targeting sequence<br />

to co-express a desired protein, such as the β7 integrin cDNAs. <strong>The</strong> pIRES2-EGFPcontaining<br />

β7 cDNA vectors were primarily used to control for transfection efficiency.<br />

146


Figure 3.55 Cloning strategy for constructing plasmids encoding the full-length β7 subunit with mutated<br />

cytoplasmic tyrosines<br />

Complementary DNA encoding the full-length wild-type human β7 subunit (β7wt) was excised from the clone<br />

pCDM8-β7wt and cloned into the plasmid pcDNA3.1 H+ to form the pcDNA3.1-β7wt vector (Step 1).<br />

Complementary DNA encoding the C-terminal region of the wild-type β7 subunit was amplified from the<br />

pcDNA3.1-β7wt vector with the primers cc24 and cc25. Complementary DNAs encoding the cytoplasmic<br />

domain of the β7 subunit carrying mutations in tyrosines 736, 741, and 761 were PCR-amplified from the<br />

respective pGEX-2T vectors with the primers cc26 and cc27 (Step 2). <strong>The</strong> PCR products that resulted from the<br />

latter two amplifications were joined together by overlap PCR using the primers cc24 and cc27. <strong>The</strong> PCR<br />

product, which bore the restriction enzyme site SacI at the 5’-end and NotI at the 3’-end (Step 4), was digested<br />

with SacI and NotI for subcloning. Complementary DNA encoding the N-terminal region of the wild-type fulllength<br />

β7 cDNA was excised from the pcDNA3.1-β7wt vector with the restriction enzymes HindIII and SacI<br />

(Step 3), and ligated together with PCR products encoding the C-terminal region that carried tyrosine mutations.<br />

<strong>The</strong> ligated fragments were subcloned into the pcDNA3.1 vector to form full-length mutated variants of the β7<br />

subunit (Step 5). <strong>The</strong> full-length β7 subunit mutants were excised with NheI and XhoI and cloned into the<br />

pIRES2-EGFP vector (Step 6). <strong>The</strong> boxes show keys to the diagram, indicating the START and STOP codons,<br />

the start of the cytoplasmic domain, and the positions of the tyrosines (Y) that were mutated to phenylalanines<br />

(F). <strong>The</strong> expression plasmids generated carried the substitutions Y736F, Y741F, Y761F, and Y736+741F in the<br />

cytoplasmic domain.<br />

147


3.8.2 Cloning of a wild-type integrin α4 construct<br />

<strong>The</strong> aim was to clone a full-length cDNA encoding the integrin α4 subunit into the pcDNA6<br />

mammalian expression vector. <strong>The</strong> cloning strategy is summarised in Figure 3.56. <strong>The</strong><br />

available α4 subunit cDNA in the pCDM8 plasmid lacked the first 114 amino acid residues at<br />

the 5’-end of the α4 subunit cDNA (data not shown). mRNA from the human T-lymphocyte<br />

cell line H9 (Step 1) was used as a template to RT-PCR amplify the missing N-terminal<br />

region of the α4 integrin cDNA (Step 2). <strong>The</strong> PCR product was cloned into pGEM-T and<br />

excised with the restriction enzymes BamHI and EcoRV (Step 3). <strong>The</strong> integrin α4 cDNA in<br />

the pCDM8 vector was excised with EcoRV and NotI, where the EcoRV site was located 3’<br />

of the missing region (Step 4). <strong>The</strong> PCR product encoding the missing N-terminus of the α4<br />

subunit was then ligated in a 3-way ligation with the rest of the α4 subunit cDNA into a<br />

pCDNA6-V5his vector (Step 5). <strong>The</strong> resulting construct was sequenced to confirm its<br />

integrity (data not shown).<br />

Figure 3.56 Schematic diagram of the strategy used to generate a pcDNA6-V5his vector encoding the full-<br />

length α4 subunit<br />

<strong>The</strong> α4 subunit cDNA provided in a pCDM8 vector lacked an extreme N-terminal sequence (denoted in purple).<br />

RNA was isolated from H9 cells (Step 1) and used as a template for RT-PCR with the primers cc28 and cc29 to<br />

repair the cDNA to obtain a full-length α4 subunit cDNA (Step 2). <strong>The</strong> resulting PCR product (red) was ligated<br />

into pGEM-T and excised with BamHI and EcoRV (Step 3). <strong>The</strong> C-terminus of the α4 subunit was excised from<br />

pCDM8 with EcoRV and NotI digestion (4). <strong>The</strong> latter two α4 subunit fragments were ligated together into the<br />

pcDNA6-V5his vector, producing a construct encoding a full-length wild-type α4 subunit.<br />

148


3.8.3 Transfection of HEK-293T cells to express the α4 and β7 subunit plasmids<br />

Confirmation of high transfection efficiency of integrin plasmids<br />

HEK-293T cells were transfected with pcDNA6-V5his and pIRES-2-EGFP plasmids<br />

encoding the α4 and variant β7 subunits, respectively, using the Lipofectamine 2000<br />

transfection reagent. Cells transfected with the α4 expression plasmid were cotransfected<br />

with a pIRES2-EGFP vector to measure transfection efficiency. <strong>The</strong> transfectants were grown<br />

in normal growth media containing of 5 μg/mL of blasticidin for 48-72 hrs, and the success of<br />

transfection was visualised by viewing GFP fluorescence by using epi-fluorescence<br />

microscopy (Figure 3.57). Transfection efficiencies were similar for all transfections, and<br />

estimated to be approximately 80-90%. GFP was uniformly expressed in cells.<br />

Figure 3.57 Transfection of HEK-293T cells with plasmids encoding the integrin α4 and β7 subunits.<br />

HEK-293T cells were cotransfected with 4 µg of the pcDNA6-V5his vector expressing the native α4 subunit and<br />

4 µg of the pIRES2-EGFP vector (Panel 1). <strong>The</strong>y were also transfected with 4 µg of each of the pIRES-2-EGFP<br />

vectors encoding native β7 (Panel 2), and the β7 subunit with the mutations Y736F (Panel 3), Y741F (Panel 4),<br />

Y761F (Panel 5), and Y736F+741F (Panel 6). Images were captured using epi-fluorescence microscopy. Bar<br />

indicates the length of 20 µm. This experiment was repeated three times.<br />

149


Confirmation of integrin expression by immunoblotting<br />

<strong>The</strong> transfectants were analysed by immunoblotting for expression of the introduced integrin<br />

subunits. Lysates of each of the above transfectants were resolved by SDS-PAGE and<br />

immunoblotted with an antibody recognising the β7 subunit extracellular domain (Figure<br />

3.58). <strong>The</strong> β7 subunit was expressed at similar levels by each transfectant (lanes 2 to 6). In<br />

contrast, HEK-293T cells transfected with the α4 expression plasmid did not express the β7<br />

subunit. <strong>The</strong> membrane was probed with an antibody against β-actin to show that equal<br />

amounts of each lysate had been loaded on the gel.<br />

Figure 3.58 Western blot analysis of HEK-293T cells transfected with plasmids encoding α4 and β7<br />

varients<br />

Transfected HEK-293T cells (5 x 10 5 ) were lysed in the presence of NP40, resolved by SDS-PAGE, and<br />

immunoblotted with an antibody against the β7 subunit extracellular domain (sheep anti-huβ7ext-IgG, used at<br />

1:1000), which was detected with an anti-sheep-HRP antibody (1:80,000). <strong>The</strong> HEK-293T cells were transfected<br />

with plasmids encoding the α4 subunit (lane 1); α4 and β7wt subunits (lane 2); α4 and β7(Y736F) subunits (lane<br />

3); α4 and β7(Y741F) subunits (lane 4); α4 and β7(Y761F) subunits (lane 5); and α4 and β7(Y736+741F)<br />

subunits (lane 6). <strong>The</strong> membrane was also immunoblotted for β-actin as a loading control. Molecular weights of<br />

marker proteins are indicated in the left-hand margin in kDa. This experiment was repeated three times.<br />

150


Staining of transfectants with anti-integrin antibodies to confirm integrin expression<br />

Expression of the α4 and β7 integrins by transfected HEK-293T cells was confirmed by<br />

staining of cells with fluoresceinated anti-integrin antibodies (Figure 3.59). <strong>The</strong> upper row of<br />

photographs in Figure 3.59 show transfected cells stained with a red fluorescent antibody<br />

against the β7 subunit cytoplasmic domain. Cells transfected with the α4 plasmid did not stain<br />

with the anti-β7 subunit antibody, indicating that HEK-293T cells do not express β7<br />

integrins, as expected. Cells cotransfected with the α4 and β7wt plasmids were brightly<br />

stained, whereas cells cotransfected with the α4 plasmid and one of the mutated β7 plasmids<br />

stained less brightly. <strong>The</strong> observed difference in staining of β7wt compared to the β7 mutants<br />

may be explained by the fact that the antibody used to detect expression was raised against<br />

the β7 subunit cytoplasmic domain such that mutations of the tyrosine residues to<br />

phenylalanines may cause a decrease in antibody affinity. To examine this possibility,<br />

transfected HEK-293T cells were stained with a green fluorescent Fib504 mAb against the<br />

extracellular domain of the β7 subunit (Figure 3.59, middle row of photographs). As<br />

expected, cells transfected with the α4 plasmid did not stain with the Fib504 mAb. In contrast<br />

to the above results, cells cotransfected with the α4 plasmid and either wild-type or mutant β7<br />

plasmids stained green and showed similar levels of staining.<br />

<strong>The</strong> transfectants were also stained with a green fluorescent DATK32 antibody against α4β7,<br />

which is complex-specific. Cells transfected with the α4 plasmid did not stain with the<br />

DATK32 mAb. Cells cotransfected with the α4 plasmid and either wild-type or mutant β7<br />

plasmids were stained green and showed similar levels of staining. Not all cells expressed<br />

α4β7, as the transfection efficiency was not 100%.<br />

Taken together, the results show that HEK-293T cells were successfully transfected with the<br />

plasmids encoding the α4 and β7 integrin subunits, leading to detectable levels of cell-surface<br />

expression of α4β7 and mutated forms.<br />

151


Figure 3.59 α4β7 is expressed at the surface of HEK-293T cells transfected with α4 and β7 plasmids<br />

HEK-293T cells were transfected with plasmids encoding the wild-type α4 subunit (pcDNA6-α4wt) alone or together with either wild-type (pcDNA3.1-β7wt) or mutant forms<br />

(pcDNA3.1-β7Y736F, -β7Y741F, -β7Y761F, or -β7Y736+741F) of the integrin β7 subunit. Transfectants were grown for 48 hrs, fixed, and stained with fluorescent antibodies.<br />

<strong>The</strong> cDNAs used for transfection are indicated at the top of each column of photographs, whereas the antibodies used to detect the encoded integrin subunit are indicated in the<br />

right-hand margin of each row. <strong>The</strong> top row shows transfectants immunostained with a rabbit polyclonal antibody against the cytoplasmic domain of the β7 subunit, which was<br />

detected using the red fluorescent anti-rabbit-antibody AF594. <strong>The</strong> middle row shows immunostaining with the Fib504 mAb against the extracellular domain of the β7 subunit,<br />

which was detected with a green fluorescent anti-rat AF488 antibody. <strong>The</strong> bottom row shows transfectants stained with the DATK32 mAb against the α4β7 complex, which was<br />

detected with a green fluorescent anti-rat AF488 antibody. Cell nuclei were stained blue with DAPI. Images were captured using epi-fluorescence microscopy. Bar indicates a<br />

length of 20 µm. This experiment was repeated three times.<br />

152


3.8.4 Testing the ability of transfectants to adhere to integrin ligands<br />

<strong>The</strong> aim here was to determine whether HEK-293T cells transfected with plasmids encoding<br />

wild-type α4β7 and forms containing mutated cytoplasmic tyrosines (Y736F, Y741F, Y761F,<br />

and Y736+741F) were able to bind to MAdCAM-1 via α4β7. <strong>The</strong> transfectants were<br />

activated with Mn 2+ , adhered to MAdCAM-1-coated plates, fixed and stained with methylene<br />

blue, and the adherent cells either visualised by light microscopy (Figure 3.60A), or<br />

quantified using a microplate reader (Figure 3.60C). In parallel, Mn 2+ -activated TK-1 cells<br />

were adhered to MAdCAM-1-coated slides for comparison as positive controls (Figure<br />

3.60B and D).<br />

HEK-293T cells transfected with the α4 plasmid that lacked α4β7 expression adhered very<br />

poorly to MAdCAM-1 (Figure 3.60C). In contrast, cells cotransfected with the α4 and β7<br />

plasmids showed a 1.5 fold increase in cell adhesion compared to cells transfected with the<br />

α4 plasmid. Single tyrosine mutations Y736F, Y741F, and Y761F in the β7 cytoplasmic<br />

domain significantly reduced the level of cell binding by 24% to 41% (p = 2.49 × 10 -2 to 3.26<br />

× 10 -2 ), whereas the double mutant Y736+741F significantly reduced cell adhesion by 53%<br />

(p = 1.88 × 10 -3 ), compared to adhesion mediated by wild-type α4β7 (Figure 3.60C).<br />

Visualization of attached cells by light microscopy confirmed the above results (Figure<br />

3.60A).<br />

For comparison, non-activated and Mn 2+ -activated TK-1 cells were bound to MAdCAM-1.<br />

<strong>The</strong>re was a 2-fold increase in binding of activated TK-1 cells compared to non-activated<br />

cells (Figure 3.60D). Under light microscopy, activated TK-1 cells densely covered the<br />

entire surface of MAdCAM-1-coated slides, whereas adherent non-activated TK-1 cells were<br />

more sparsely distributed (Figure 3.60B).<br />

<strong>The</strong> above results suggest that all three tyrosine residues in the cytoplasmic domain of the β7<br />

subunit are important in mediating the cell adhesion properties of α4β7 integrin.<br />

153


Figure 3.60 Testing the ability of HEK 293T transfectants to adhere to MAdCAM-1<br />

HEK-293T cells transfected with plasmids encoding α4, α4β7wt, α4β7(Y736F), α4β7(Y741F), α4β7(Y761F),<br />

and α4β7(Y736+741F) were tested for their ability to bind to MAdCAM-1-coated slides and plates. Slides and<br />

plates were coated with 10 μg/mL of MAdCAM-1 overnight at 4 °C. Cells (5 x 10 5 to 1 x 10 6 per well) were left<br />

to adhere to plates for 30 min at RT, and were then fixed and stained with methylene blue. <strong>The</strong>y were visualised<br />

by light microscopy (A) and quantified using a microplate reader at the wavelength 405 nm (C). For<br />

comparison, non-activated and activated TK-1 cells were adhered to MAdCAM-1-coated slides, visualised by<br />

light microscopy (B), and quantified in a microplate reader (D). Data shown represent the mean and standard<br />

deviation of three wells. * denotes a p-value < 0.05 and ** denotes a p-value of < 0.01. <strong>The</strong> experiments were<br />

performed twice in triplicate.<br />

154


3.9. Identification of binding partners of the β7 cytoplasmic domain<br />

Integrin cytoplasmic domains play a key role in regulating integrin functions. <strong>The</strong> results of<br />

this thesis so far have shown that a CARD within the β7 cytoplasmic domain binds src and<br />

FAK, and interacts with the cytoskeletal proteins paxillin, filamin and α-actinin. <strong>The</strong> aim here<br />

was to identify additional binding partners for the β7 cytoplasmic domain, and examine their<br />

involvement in the signalling of β7 integrins.<br />

3.9.1 Pull-down assay<br />

A biotin-labelled synthetic peptide comprising the full-length 52 aa β7 cytoplasmic domain<br />

was bound to streptavidin-coated magnetic beads, and incubated with TK-1 cell lysates. <strong>The</strong><br />

beads were washed thoroughly, and the pulled-down proteins resolved by SDS-PAGE, and<br />

stained with Coomassie blue (Figure 3.61, lanes 2 and 4) and by silver staining (data not<br />

shown). A control synthetic peptide (biotin-APTLPPAWQPFLK-OH) was incubated with a<br />

TK-1 cell lysate for comparison (Figure 3.61, lanes 1 and 3). <strong>The</strong> β7 cytoplasmic domain<br />

peptide specifically precipitated two proteins of approximately 70 kDa when compared to the<br />

control peptide. <strong>The</strong> 70 kDa proteins were equally precipitated from the lysates of Mn 2+ -<br />

activated and non-activated cells (Figure 3.61, lane 2 compared to lane 4).<br />

155


Figure 3.61 SDS-PAGE analysis of proteins precipitated with a synthetic peptide comprising the complete<br />

cytoplasmic domain of the β7 subunit<br />

A synthetic peptide comprising the complete cytoplasmic domain of the β7 subunit was conjugated to magnetic<br />

beads, and incubated with TK-1 cell lysates for 2 hrs at 4°C. <strong>The</strong> lysates were prepared from non-activated<br />

(lanes 1 and 2) and Mn 2+ -activated (lanes 3 and 4) cells. <strong>The</strong> pull-down assays were performed with an<br />

unrelated control peptide (biotin-APTLPPAWQPFLK; lanes 1 and 3) and the β7 cytoplasmic domain peptide<br />

(lanes 2 and 4). <strong>The</strong> affinity-purified proteins were resolved by SDS-PAGE, and stained with Coomassie blue.<br />

Two protein bands of 70 kDa that were specifically precipitated by the β7 cytoplasmic domain peptide<br />

(indicated by arrow in lane 4) were analysed by mass spectrometry. <strong>The</strong> sizes of molecular weight markers are<br />

indicated in the left-hand margin. This experiment was repeated twice.<br />

3.9.2 Mass spectrometry identifies heat shock proteins as ligands of the β7 subunit<br />

<strong>The</strong> 70 kDa proteins isolated in the pull-down assay (Figure 3.61, lane 4) by the β7<br />

cytoplasmic domain peptide were sent for mass spectrometry analysis. <strong>The</strong> whole gel was<br />

sent to the Centre for Genomics and Proteomics at the <strong>University</strong> of Auckland. <strong>The</strong> two<br />

distinguishable bands were separately and analysed. <strong>The</strong> resulting spectra were compared<br />

with a Mus musculus genomic database. Mass spectrometry analysis identified four hsp70<br />

family members as potential ligands of the β7 cytoplasmic domain. <strong>The</strong>y included heat shock<br />

protein 9, heat shock protein 8a, heat shock 70 kDa protein 1-like, and heat shock 70 kDa<br />

protein 1a (Figure 3.62A and B).<br />

156


Figure 3.62 Mass spectrometry identifies four heat shock proteins as potential β7 ligands<br />

Mass spectrometry analysis identified four hsp70 family members as potential ligands of the β7 cytoplasmic<br />

domain. <strong>The</strong>y include heat shock protein 9 (A), heat shock protein 8a, heat shock 70 kDa protein 1-like, and<br />

heat shock 70 kDa protein 1a (B). Mass spectrometry analysis was performed by the Centre for Genomics and<br />

Proteomics at the <strong>University</strong> of Auckland.<br />

157


3.9.3 A pull-down assay estabilishes that hsp70 interacts indirectly with the β7 subunit<br />

<strong>The</strong> synthetic β7 cytoplasmic domain peptide was tested for its ability to precipitate a<br />

recombinant hsp70 protein to determine whether hsp70 interacts directly or indirectly with<br />

the β7 cytoplasmic domain. <strong>The</strong> β7 cytoplasmic domain peptide and a control peptide<br />

(control peptide 1) were immobilised on magnetic beads and incubated with recombinant<br />

hsp70-1a (ProSpec-Tany TechnoGene Ltd) in the presence and absence of a TK-1 cell lysate.<br />

<strong>The</strong> beads were washed thoroughly, bound proteins were resolved by SDS-PAGE, and the<br />

gel was silver-stained (Figure 3.63). <strong>The</strong> control peptide did not precipitate recombinant<br />

hsp70 in the absence or presence of a TK-1 cell lysate (lanes 1 and 2). <strong>The</strong> β7 cytoplasmic<br />

domain peptide precipitated recombinant hsp70, but only in the presence of a TK-1 cell lysate<br />

(compare lanes 3 and 4). <strong>The</strong> precipitated doublet of 70 kDa proteins co-migrated with<br />

recombinant hsp70 (lane 5), and were not detectable in the crude TK-1 cell lysate (lane 6).<br />

<strong>The</strong> results suggest that hsp70 does not bind directly to the β7 cytoplasmic domain, but does<br />

so after facilitation by other proteins.<br />

Figure 3.63 A synthetic β7 cytoplasmic domain peptide precipitates recombinant hsp70 in a pull down<br />

assay<br />

A β7 cytoplasmic domain peptide and a control peptide bound to magnetic beads were incubated with 4 μg of<br />

recombinant human hsp70-1a in the absence or presence of a TK-1 lysate. Lanes 1 and 2, control peptide<br />

incubated with hsp70 in the absence or presence of a TK-1 cell lysate, respectively. Lanes 3 and 4, β7<br />

cytoplasmic domain peptide incubated with hsp70 in the absence or presence of a TK-1 cell lysate, respectively.<br />

Lane 5, 2 μg of recombinant hsp70-1a protein. Lane 6, TK-1 cell lysate (10 4 cells/lane). <strong>The</strong> arrow indicates the<br />

location of the recombinant hsp70-1a. <strong>The</strong> left-hand margin indicates the size of the molecular weight markers.<br />

This experiment was repeated three times.<br />

158


3.9.4 Coimmunoprecipitation of α4β7 and hsp70<br />

<strong>The</strong> aim here was to determine whether α4β7 and hsp70 could be coimmunoprecipitated with<br />

each other from a TK-1 cell lysate. <strong>The</strong> β7 integrin was immunoprecipitated from a TK-1 cell<br />

lysate, the immunoprecipitate resolved by SDS-PAGE, transferred to PVDF membrane, and<br />

immunoblotted with antibodies against hsp70 (Figure 3.64A) and the integrin β7 subunit<br />

(Figure 3.64B). Immunoprecipitates formed with rat-IgG were used as controls. Hsp70 was<br />

immunoprecipitated by antibodies against the β7 subunit, indicating that hsp70 complexes<br />

with the integrin β7 subunit (Figure 3.64A, 70 kDa band indicated by arrow). <strong>The</strong> specificity<br />

of the immunoprecipitating Fib504 anti-β7 integrin subunit antibody was confirmed by<br />

immunoblotting the membrane with a rabbit polyclonal antibody against the β7 cytoplasmic<br />

domain (Figure 3.64B, 120 kDa band indicated by arrow).<br />

Figure 3.64 Hsp70 is coimmunoprecipitated from a TK-1 cell lysate with the β7 subunit<br />

TK-1 cell lysates were incubated with the rat anti-β7 subunit antibody Fib504, and a control rat IgG. Both<br />

antibodies had been cross-linked to Sepharose beads. <strong>The</strong> immunoprecipitates were resolved on 10%<br />

polyacrylamide SDS-gels, and immunoblotted with antibodies against hsp70 (A) and the β7 integrin subunit (B).<br />

A band at approximately 70 kDa in (A) indicates the presence of hsp70 in the β7 immunoprecipitate (right lane;<br />

refer to arrow), and a band at approximately 120 kDa in (B) indicates the presence of the β7 subunit (refer to<br />

arrow). No bands of the corresponding sizes were present in the control immunoprecipitates (middle lanes). For<br />

comparison, the unfractionated TK-1 cell lysate was immunoblotted (left lanes). <strong>The</strong> sizes of the molecular<br />

weight markers are indicated in the left-hand margin. This experiment was repeated twice.<br />

159


Hsp70 was immunoprecipitated from a TK-1 cell lysate using anti-hsp70 antibodies, and the<br />

immunoprecipitate immunoblotted with the rabbit anti-β7 subunit antibody to further confirm<br />

that hsp70 forms a complex with α4β7 (Figure 3.65). <strong>The</strong> TK-1 cell lysate was<br />

immunoprecipitated with the Fib504 mAb against the β7 subunit (lane 3) and an antibody<br />

against hsp70 (lane 7). An immunoprecipitate formed with mouse ascites served as a control<br />

(lane 5). For comparison, the TK-1 cell lysate was loaded into lane 1, and the<br />

immunoprecipitating antibodies in the absence of lysate were included in lanes 2, 4 and 6.<br />

Both the Fib504 mAb and the anti-hsp70 mAb immunoprecipitated the β7 subunit, as<br />

indicated by a protein band at 120 kDa (lanes 3 and 7, respectively). <strong>The</strong> β7 subunit may<br />

have been partially degraded as indicated by bands of lesser molecular weight. In contrast,<br />

the 120 kDa protein was not immunoprecipitated by the control antibody (lane 5).<br />

Figure 3.65 <strong>The</strong> β7 subunit is coimmunoprecipitated from a TK-1 cell lysate with hsp70<br />

TK-1 cell lysates were incubated with the anti-β7 mAb Fib504, a mouse ascites to anti-human IgG (control<br />

antibody), and with an anti-hsp70 mAb (mouse ascites). <strong>The</strong> immunoprecipitates were resolved by SDS-PAGE<br />

and immunoblotted with a rabbit anti-β7 polyclonal antibody. Lane 1, TK-1 cell lysate; lane 2, anti-β7 (Fib504)<br />

antibody alone; lane 3, lysate immunoprecipitated with anti-β7 mAb (Fib504); lane 4, mouse ascites alone; lane<br />

5, lysate immunoprecipitated with mouse ascites; lane 6, anti-hsp70 antibody alone; lane 7, lysate<br />

immunoprecipitated with the anti-hsp70 antibody. Arrow in the right-hand margin indicates the size and location<br />

of the mature β7 subunit. <strong>The</strong> sizes of the molecular weight markers are indicated in the left-hand margin. This<br />

experiment was repeated twice.<br />

160


3.9.5 Localization of hsp70 and hsp90 in TK-1 cells<br />

<strong>The</strong> results above indicate that hsp70 complexes with the β7 integrin in vivo. <strong>The</strong> aim here<br />

was to visualise the localisation of endogenous hsp70 and β7 in cells. TK-1 cells were<br />

adhered to MAdCAM-1-coated coverslips, fixed, and co-stained with a red fluorescent<br />

antibody against hsp70 and a green fluorescent antibody against the β7 integrin subunit<br />

Figure 3.66. Staining was visualised by fluorescence microscopy. α4β7 strongly localized to<br />

the uropod (Figure 3.66, refer to arrow), but was also present at what appear to be small focal<br />

points. In contrast, hsp70 was more diffusely expressed throughout the cytoplasm. Merger of<br />

the images failed to reveal co-localisation of α4β7 and hsp70, but it was clear that hsp70 was<br />

not expressed with α4β7 on the uropod.<br />

Figure 3.66 Localisation of hsp70 and α4β7 on TK-1 cells attached and spread on MAdCAM-1<br />

Mn 2+ -activated TK-1 cells were adhered to MAdCAM-1-coated coverslips, fixed with paraformaldehyde, and<br />

stained with rat anti-β7 (Fib504, 1:100) and rabbit anti-hsp70 (1:100) antibodies. <strong>The</strong> primary antibodies were<br />

detected with anti-rat-AF488 (green, 1:250) and anti-rabbit-AF594 (red, 1:250) antibodies, respectively. Nuclei<br />

were stained with DAPI, and the images were merged to measure co-localisation of hsp70 and β7. <strong>The</strong> magenta<br />

arrow shows the localisation of the uropod. Images were captured using confocal fluorescence microscopy. Bar<br />

indicates the length of 20 μm. This experiment was repeated three times.<br />

161


In addition to localization of hsp70 in T cells, the location of hsp90 was investigated for<br />

comparison. Hsp90 had been included as a control for hsp70 in the above experiment. Hsp90<br />

represents another major family of heat shock proteins involved in chaperone activity<br />

required to protect signalling and adaptor proteins from degradation. TK-1 cells were adhered<br />

to MAdCAM-1-coated coverslips, fixed, and co-stained with a red fluorescent antibody<br />

against hsp90 and a green fluorescent antibody against the β7 integrin subunit (Figure 3.67).<br />

As before, α4β7 was shown to be strongly expressed by the uropod and within small focal<br />

adhesions (Figure 3.67, refer to magnified image B). Hsp90 was diffusely expressed<br />

throughout the cytoplasm (Figure 3.67A), and was weakly expressed by the uropod (Figure<br />

3.67, refer to magnified image B). Merger of the images revealed that α4β7 and hsp90 do not<br />

colocalize within the uropod.<br />

Figure 3.67 Localisation of hsp90 and α4β7 on TK-1 cells attached and spread on MAdCAM-1<br />

Mn 2+ -activated TK-1 cells were adhered to MAdCAM-1-coated coverslips, fixed with paraformaldehyde, and<br />

stained with rat anti-β7 (Fib504, 1:100) and goat anti-hsp90 (1:100) antibodies. <strong>The</strong> primary antibodies were<br />

detected with anti-rat-AF488 (green, 1:250) and anti-goat-AF594 (red, 1:250) antibodies, respectively (A).<br />

Images were merged to measure co-localisation of hsp90 and β7. <strong>The</strong> uropod of one cell was magnified (B).<br />

Images were captured using epi-fluorescence microscopy. Bar indicates the length of 20 μm. This experiment<br />

was repeated three times.<br />

162


3.9.6 Hsp70 colocalizes with intracellular β7 ligands following ligand-induced clustering<br />

of α4β7<br />

<strong>The</strong> aim here was to determine whether hsp70 might colocalize with intracellular ligands for<br />

β7 within SMAC formed by MAdCAM-1-induced clustering of α4β7. A green<br />

fluoresceinated FAM-YDRREY peptide was employed which is expected to interact with<br />

intracellular ligands for α4β7. TK-1 cells were pre-incubated with the FAM-YDRREY<br />

peptide, and then activated with Mn 2+ to induce binding to MAdCAM-1-coated microspheres<br />

and clustering of α4β7 at the cell surface. <strong>The</strong> cells were fixed and stained with antibodies<br />

against hsp70 to look for co-localisation with the FAM-YDRREY peptide (Figure 3.68). <strong>The</strong><br />

FAM-YDRREY peptide localized largely within a small number of clusters which formed<br />

near or at the plasma membrane. In contrast, hsp70 was more diffusely scattered throughout<br />

the cells, but nevertheless it was clearly present within large clusters that appeared to have<br />

similar positions as those containing the FAM-YDRREY peptide (Figure 3.68, refer to blue<br />

arrows). Colocalization of hsp70 and the FAM-YDRREY peptides was confirmed by<br />

merging the images. Interestingly, there were several clusters where hsp70 and the FAM-<br />

YDRREY peptides did not colocalize (Figure 3.68, refer to white arrows)<br />

Figure 3.68 Hsp70 colocalizes with intracellular β7 ligands within SMAC formed with MAdCAM-1coated<br />

microspheres<br />

TK-1 cells were incubated with 1 μM of the green fluorescent peptide FAM-YDRREY and then activated with<br />

Mn 2+ . Clustering of α4β7 was induced by incubating the cells with MAdCAM-1-coated microspheres. <strong>The</strong> cells<br />

were fixed and stained with a rabbit anti-hsp70 antibody (1:100) and then with a red fluorescent anti-rabbit-<br />

AF594 antibody. <strong>The</strong> merged images reveal clusters where hsp70 and the FAM-YDRREY peptide co-localize<br />

(blue arrows), and clusters where they do not colocalize (white arrows). Images were captured using epifluorescence<br />

microscopy. Bar indicates the length of 20 μm. This experiment was repeated three times.<br />

163


3.9.7 Effect of heat shock on expression of hsp70 and the β7 subunit<br />

Given that hsp70 interacts with α4β7, albeit indirectly, it was decided to determine whether<br />

hsp70 might influence α4β7 expression and/or function. Here, the effect of heat shock on the<br />

expression of hsp70 and the β7 subunit was examined, given that hsp70 is expected to protect<br />

cells against heat shock. TK-1 cells were heat shocked at 42ºC for 45 min and allowed to<br />

recover overnight at 37ºC. Non-heat shocked cells were cultured overnight at 37ºC, and<br />

served as controls. <strong>The</strong> cells were lysed and the lysates were analysed by SDS-PAGE and<br />

immunoblot analysis. <strong>The</strong> membrane was separated into three sections according to the<br />

molecular weight ladder, whereby the larger and intermediate sized proteins were<br />

immunoblotted with anti-β7 and anti-hsp70 antibodies, respectively (Figure 3.69). <strong>The</strong>re was<br />

no detectable change in the level of the β7 subunit in the soluble (S) and insoluble (P)<br />

fractions after heat shock treatment. In contrast, the level of hsp70 was increased in both the<br />

soluble and insoluble fractions after heat shock treatment (Figure 3.69). <strong>The</strong> membrane<br />

containing the smaller proteins was probed an antibody against β-actin to serve as a loading<br />

control.<br />

Figure 3.69 Effect of heat shock on the expression of hsp70 and the β7 subunit by TK-1 cells<br />

164


TK-1 cells were heat shocked at 42ºC for 45 min and allowed to recover overnight at 37ºC. Non-heat shocked<br />

cells were cultured overnight at 37ºC, and served as controls. <strong>The</strong> soluble (S) and insoluble (P) fractions of<br />

lysates of the latter cells were resolved by SDS-PAGE, and the proteins transferred to PVDF membrane. <strong>The</strong><br />

membrane was immunoblotted with antibodies against the integrin β7 subunit (rabbit anti-β7 cytoplasmic<br />

domain), hsp70 (rabbit anti-hsp70) and β-actin (rabbit anti-β-actin). <strong>The</strong> primary antibodies were detected using<br />

goat anti-rabbit-HRP antibodies (1:20,000). <strong>The</strong> positions of the β7 subunit, hsp70 and β-actin are indicated by<br />

the arrows in the right-hand margin. <strong>The</strong> sizes of the molecular weight markers are indicated in the left-hand<br />

margin. This experiment was repeated twice.<br />

3.9.8 Effect of heat shock on activation of α4β7 is time dependent<br />

Here the aim was to determine whether the activation of α4β7 by Mn 2+ might be affected by<br />

heat shock treatment. TK-1 cells were heat shocked at 42˚C for 15, 30 and 45 min, and then<br />

allowed to recover overnight at 37ºC. Non-heat shocked cells were cultured overnight at<br />

37ºC, and served as controls. <strong>The</strong> cells were either left unactivated or were activated with<br />

Mn 2+ and compared in their ability to adhere to MAdCAM-1-coated slides. As can be seen in<br />

Figure 3.70, α4β7 integrins on resting TK-1 cells are normally inactive and are only able to<br />

marginally support cell binding to MAdCAM-1. In the presence of Mn 2+ , integrins rapidly<br />

switch to an active state. Interestingly, prolonged heat shock treatment for 15 to 45 minutes<br />

activated α4β7 such that cells were able to bind to MAdCAM-1 without prior activation with<br />

Mn 2+ . <strong>The</strong> degree of cell binding depended on the length of heat shock treatment, where<br />

maximal activation of α4β7 was achieved by heat shocking cells for 45 min. Heat shock<br />

treatment of cells for 45 min produced levels of cell binding that were similar to those<br />

induced by Mn 2+ (Figure 3.70A). While all cells in the experiment were viable as determined<br />

by trypan blue exclusion, heat shock treatment for times greater than 45 min led to increasing<br />

cell death (data not shown). Cells subjected to heat shock for 45 min were also incubated<br />

with antibodies against α4 and β7 prior to cell adhesion (Figure 3.70B). In the presence of<br />

antibodies, there was a decrease in cell adhesion, confirming heat shock activated cell<br />

binding to MAdCAM-1 was mediated by α4β7. <strong>The</strong> results indicate that heat shock treatment<br />

very gradually induces the activation of α4β7 integrins on TK-1 cells, and that the degree of<br />

activation of α4β7 is dependent on the length of heat shock treatment. <strong>The</strong> ability of α4β7 to<br />

remain active overnight following heat shock treatment is unusual, as it was previously<br />

reported that Mn 2+ -induced activation of α4β7 is very transient i.e. once activated, α4β7<br />

slowly deactivates over period of to 2 hrs (Zhang et al. 1999).<br />

165


Figure 3.70 Effect of heat shock treatment on the binding of TK-1 cells to MAdCAM-1<br />

TK-1 cells were heat shocked at 42˚C for 0, 15, 30 or 45 min, and incubated overnight at 37˚C. Non-heat<br />

shocked cells were cultured overnight at 37ºC, and served as controls. Cells were either left non-activated or<br />

activated with Mn 2+ . Cells were adhered to MAdCAM-1-coated slides at RT for 30 min (A). B) Cells were also<br />

heat shocked at 42˚C for 45 min incubated overnight at 37˚C, and then incubated with antibodies against α4 and<br />

β7 prior to adhesion to MAdCAM-1-coated slides. Bound cells were fixed, stained with 1% methylene blue, and<br />

cell attachment analysed using a microplate reader at a wavelength of 495 nm. Data shown represent the mean<br />

and standard deviation of three wells. <strong>The</strong> experiment was repeated three times.<br />

3.9.9 Fever-range temperatures can activate α4β7<br />

TK-1 cells were incubated at 37, 39, 42 and 44°C for 45 min to assess whether the degree of<br />

activation of α4β7 was temperature-dependent. In particular would fever-range temperatures<br />

of ~39°C be able to activate α4β7? <strong>The</strong> cells were allowed to recover overnight at 37°C and<br />

tested for their ability to bind to MAdCAM-1-coated slides (Figure 3.71). TK-1 cells<br />

incubated at 39 and 42°C showed similar levels of binding to MAdCAM-1 i.e. about 1-fold<br />

greater binding than non-heat shocked cells. <strong>The</strong> level of cell binding was comparable to that<br />

166


of TK-1 cells activated with Mn 2+ . Cells treated for 45 min at 44 °C did not survive (data not<br />

shown).<br />

Figure 3.71 Fever-range temperatures induce TK-1 cell binding to MAdCAM-1<br />

TK-1 cells were incubated at 37, 39, and 42˚C for 45 min. <strong>The</strong> cells were then incubated overnight at 37˚C.<br />

Non-heat shocked cells were cultured overnight at 37ºC, and served as controls. <strong>The</strong> Ccells were either left nonactivated<br />

or activated with Mn 2+ and adhered to MAdCAM-1-coated surfaces at RT for 30 min. Bound cells<br />

were fixed, stained with 1% methylene blue, and cell attachment analysed using a microplate reader at a<br />

wavelength of 495 nm. Data shown represent the mean and standard deviation of three wells. <strong>The</strong> experiment<br />

was repeated three times.<br />

3.9.10 Heat shock leads to prolonged activation of α4β7 even after Mn 2+ -treatment<br />

As mentioned above, it was previously reported that Mn 2+ -induced α4β7 activation is very<br />

transient i.e. once activated, α4β7 slowly deactivates over a period of to 2 hrs (Zhang et al.<br />

1999). However, heat shock treatment appeared to cause the prolonged activation of α4β7,<br />

where α4β7 remained active after overnight culture for at least 16 hrs (Figure 3.70). Here, it<br />

was determined whether treatment with Mn 2+ after heat shocking would render α4β7<br />

activation more labile.<br />

TK-1 cells were heat shocked at 42˚C and allowed to recover overnight at 37˚C. Non-heatshocked<br />

cells were cultured overnight at 37ºC, and served as controls. Cells were harvested<br />

and either left untreated or treated with Mn 2+ for 0, 1 or 2 hrs. <strong>The</strong>y were then allowed to<br />

attach to MAdCAM-1-coated slides (Figure 3.72). As before, heat shocked cells bound to<br />

MAdCAM-1 even without Mn 2+ -induced activation. <strong>The</strong> cells retained their ability to bind to<br />

MAdCAM-1 even after 2 hrs treatment with Mn 2+ . In contrast, TK-1 cells that had not been<br />

heat shocked were unable to bind to MAdCAM-1 unless they were activated with Mn 2+ .<br />

167


Further, their ability to bind to MAdCAM-1 was transient, where binding was decreased by<br />

54% after 1 hr of Mn 2+ treatment. After a further hour the cells had completely lost their<br />

ability to bind to MAdCAM-1.<br />

Figure 3.72 <strong>The</strong> prolonged activation of TK-1 cells by heat shocking is not affected by Mn 2+ treatment<br />

TK-1 cells were heat shocked at 42˚C for 45 min, or left untreated, and cultured overnight at 37˚C. Both heat<br />

shocked and untreated cells were activated with Mn 2+ at RT for 0, 1 or 2 hrs before testing their ability to adhere<br />

to MAdCAM-1-coated slides. Cells were left to adhere at RT for 30 min, and then fixed, stained with 0.1%<br />

methylene blue, and attachment analysed using a microplate reader at a wavelength of 495nm. All cells were<br />

confirmed to be viable by trypan blue exclusion. Data shown represent the mean and standard deviation of three<br />

wells. <strong>The</strong> experiment was repeated three times.<br />

3.9.11 An inhibitor of hsp70 blocks the binding of TK-1 cells to MAdCAM-1<br />

<strong>The</strong> effects of the hsp70 inhibitor KNK-437 on TK-1 cell binding to MAdCAM-1 was<br />

examined to further study the effects of hsp70 on the function of β7 integrins. KNK-437<br />

inhibits heat shock factor 1 (HSF1) binding to heat shock element (HSE) thereby preventing<br />

expression of inducible hsps (Ohnishi et al. 2004). TK-1 cells were incubated with increasing<br />

concentrations (10, 50 to 100 µM) of KNK-437 at RT for 2 hrs, activated with Mn 2+ , and<br />

tested for binding to MAdCAM-1-coated slides. KNK-437 decreased the binding of TK-1<br />

cells to MAdCAM-1 in a dose dependent manner. It caused significant reductions in cell<br />

binding of 72% (p = 1.53 × 10 -3 ), and 94% (p = 1.48 × 10 -4 ) at 50 and 100 µM, respectively<br />

(Figure 3.73). Treatment of TK-1 cells with DMSO, the dilutent for KNK-437, had no effect<br />

on TK-1 cell adhesion (data not shown). Inhibition of TK-1 cell binding to MAdCAM-1 by<br />

168


KNK-437 suggests that hsp70 is involved in the activation and/or signalling pathways of<br />

α4β7.<br />

Figure 3.73 KNK-437 blocks TK-1 cell adhesion to MAdCAM-1<br />

TK-1 cells were incubated in the presence or absence of 0, 10, 50, or 100 µM KNK-437 at 37˚C for 2 hrs and<br />

activated with Mn 2+ . Cells not treated with KNK-437 and left unactivated served as controls. Cells were adhered<br />

to MAdCAM-1-coated slides at RT for 30 min. <strong>The</strong> bound cells were fixed, stained with 0.1% methylene blue,<br />

and cell attachment analysed using a microplate reader at a wavelength of 495 nm. Data shown represent the<br />

mean and standard deviation of three wells. <strong>The</strong> experiment was repeated three times. ** denotes a p-value <<br />

0.01 and *** denotes a p-value of < 0.001; compared to the number of cells bound in the absence of peptide.<br />

3.9.12 KNK-437 has no detectable effect on hsp70 and β7 expression.<br />

Here the aim was to determine whether KNK-437 alters the level of expression of hsp70 and<br />

the β7 subunit, which might explain how it blocks cell adhesion. TK-1 cells were incubated<br />

with 0, 25, 50 or 100 μM KNK-437, and cell lysates were subjected to Western blot analysis<br />

with antibodies against the β7 subunit and hsp70 (Figure 3.74). <strong>The</strong> membrane was also<br />

immunoblotted for β-actin as a loading control. KNK-437 had no detectable effect on the<br />

expression of either hsp70 or the β7 subunit, when levels were compared with those of βactin<br />

169


Figure 3.74 Effect of KNK-437 on the expression of hsp70 and the β7 subunit<br />

TK-1 cells were incubated with 0, 25, 50 and 100 µM KNK-437 at 37˚C overnight. <strong>The</strong> cells were lysed and the<br />

soluble protein fraction was resolved by SDS-PAGE, and transferred onto PVDF membrane. <strong>The</strong> membrane<br />

was immunoblotted with antibodies against the β7 subunit, hsp70, and β-actin as a loading control. <strong>The</strong><br />

positions of the β7 subunit (120 kDa), hsp70 (70 kDa), and β-actin (40 kDa) are indicated in the right-hand<br />

margin. <strong>The</strong> sizes of the molecular weight markers are indicated in the left-hand margin. <strong>The</strong> experiment was<br />

repeated three times.<br />

3.9.13 Serum deprivation activates α4β7<br />

<strong>The</strong> effect of heat shock on TK-1 cell adhesion to MAdCAM-1 is a stress-related response.<br />

<strong>The</strong> aim here was to determine whether other stressors might induce the activation of α4β7.<br />

<strong>The</strong> response of TK-1 cells to serum deprivation which leads to nutrient starvation, and<br />

decreased levels of growth factors was examined. TK-1 cells are normally grown in the<br />

presence of 10% serum. For serum deprivation the serum concentration was decreased to 5<br />

and 2.5%, and the cells were incubated overnight. <strong>The</strong> cells were either left unactivated or<br />

were activated by Mn 2+ and tested for binding to MAdCAM-1-coated slides. Cells grown in<br />

10% FCS were unable to bind to MAdCAM-1 without activation by Mn 2+ . In contrast, cells<br />

grown in 5% and 2.5% serum bound to MAdCAM-1 without prior activation with Mn 2+ . Cell<br />

binding was increased by 2-fold and 3-fold, respectively (Figure 3.75A). <strong>The</strong> results show<br />

that activation of α4β7 can be induced by stress caused by low serum levels. Cells were also<br />

grown in 10%, 5% and 2.5% serum and incubated with antibodies against α4 and β7 prior to<br />

cell adhesion (Figure 3.75B). In the presence of antibodies cell adhesion decreased,<br />

confirming that α4β7 mediates cell adhesion in response to serum deprivation.<br />

170


Figure 3.75 Effect of serum deprivation on TK-1 cell adhesion to MAdCAM-1<br />

TK-1 cells were cultured in normal conditions of 10% serum, or stressed by culturing in 5% serum and 2.5%<br />

serum overnight (A). For comparison, an aliquot of the cells was heat shocked at 42˚C for 45 min and cultured<br />

overnight in 10% FCS. <strong>The</strong> cells were either left unactivated or were activated with Mn 2+ , and bound to<br />

MAdCAM-1-coated slides at RT for 30 min. Cells cultured in 10%, 5% and 2.5% serum were also incubated<br />

with antibodies against α4 and β7 for 30 min prior to cell binding assay (B). <strong>The</strong> bound cells were fixed, stained<br />

with methylene blue, and cell attachment analysed using a microplate reader at a wavelength of 495nm. All cells<br />

were confirmed to be viable by trypan blue exclusion. Data shown represent the mean and standard deviation of<br />

three wells. <strong>The</strong> experiment was repeated three times.<br />

<strong>The</strong> level of expression of the β7 subunit and hsp70 by cells deprived of serum was<br />

examined. TK-1 cells were grown overnight in media containing 2.5%, 5%, 7.5% and 10%<br />

serum, and the soluble fraction of the cell lysates was analyzed by SDS-PAGE, and subjected<br />

to Western blot analysis with antibodies against the β7 subunit and hsp70 (Figure 3.76).<br />

Serum deprivation did not lead to any detectable change the levels of the β7 subunit and<br />

hsp70, suggesting that increased binding of TK-1 cells to MAdCAM-1 in response to serum<br />

deprivation was not due to increased expression of α4β7. All cells were confirmed to be<br />

viable by trypan blue exclusion before cell lysis (data not shown).<br />

171


Figure 3.76 Expression levels of the β7 subunit and hsp70 remain unchanged after serum depletion<br />

TK-1 cells were cultured overnight in media containing 2.5%, 5%, 7.5%, and 10% serum. <strong>The</strong> soluble protein<br />

fraction of the cell lysates was resolved by SDS-PAGE, and transferred onto PVDF membrane. <strong>The</strong> membrane<br />

was cut into two pieces between the 75 and 50 kDa protein markers. <strong>The</strong> piece containing the higher molecular<br />

weight proteins was immunoblotted with antibodies against the β7 subunit cytoplasmic domain (rabbit anti-β7<br />

cytoplasmic domain, 1:400) and then stripped and probed with an antibody against hsp70 (rabbit anti-hsp70,<br />

1:200). <strong>The</strong> piece containing the lower molecular proteins was immunoblotted with an antibody against β-actin<br />

(rabbit-anti-β-actin, 1:1000), in order to measure protein loading. <strong>The</strong> positions of the β7 subunit (120 kDa),<br />

hsp70 (70 kDa), and β-actin (40 kDa) are indicated in the right-hand margin. <strong>The</strong> sizes of the molecular weight<br />

markers are indicated in the left-hand margin. <strong>The</strong> multiple bands beneath the major actin band are suggestive of<br />

degradation products, indicating that the proteins analyzed have been partially degraded. <strong>The</strong> experiment was<br />

repeated three times.<br />

172


Chapter 4. Discussion<br />

4.1. Preliminary characterization of α4β7 and αEβ7 expression and<br />

adhesion<br />

Development and characterization of reagents and methods for the study<br />

<strong>The</strong> initial aim of this thesis was to confirm the expression of β7 integrins on T-cell<br />

lymphoma TK-1 (Ruegg et al. 1992) and T-cell hybridoma MTC-1 (Roberts et al. 1993) cells<br />

which had previously been reported to express the α4β7 and αEβ7 integrins, respectively. <strong>The</strong><br />

expression of both integrins on the latter cell lines was detected by immunoblotting and<br />

fluorescence microscopy, confirming previous reports. <strong>The</strong> principle ligands of α4β7 and<br />

αEβ7 are MAdCAM-1 and E-cadherin respectively, and were required as recombinant forms<br />

in order to serve as substrates to analyze α4β7 and αEβ7-mediated cell adhesion. <strong>The</strong> cDNAs<br />

of MAdCAM-1 and E-cadherin had previously been fused by lab colleagues, to the CH2/CH3<br />

regions of the human IgG antibody heavy chain Fc domain to produce recombinant<br />

MAdCAM-1-Fc, and E-cadherin-Fc proteins, respectively (Yang et al. 1995; Berg et al.<br />

1999b). <strong>The</strong> latter reagents were used to produce recombinant forms of MAdCAM-1-Fc and<br />

E-cadherin-Fc, whose purity was confirmed by SDS-PAGE.<br />

In order to assess β7 integrin signalling pathways it was necessary to choose specific integrin<br />

activating agents. Integrins can be activated by a variety of stimuli, including the integrin<br />

pan-activator Mn 2+ (Dransfield et al. 1992), the protein kinase C (PKC) activator phorbol-12myristate<br />

13-acetate (PMA; Rothlein et al. 1986), and the G-protein activator AlF4 - . AlF4 -<br />

mimics the γ phosphate of GTP by promoting the active conformation of heterotrimeric Gproteins<br />

(Coleman et al. 1994) and was found to activate α4β7 on TK-1 cells to bind<br />

MAdCAM-1 (Driessens et al. 1997; Zhang et al. 1999).This thesis confirmed that the binding<br />

of TK-1 cells to MAdCAM-1-coated slides could be induced by activation of α4β7 with<br />

Mn 2+ , PMA and ALF4 - .<br />

While MTC-1 cells were confirmed to express cell-surface αEβ7, the cells did not bind to Ecadherin-coated<br />

slides. However, MTC-1 cells activated with either Mn 2+ , PMA or AlF4 -<br />

bound to E-cadherin in solution via αEβ7, as confirmed by antibody blockade with anti-αE,<br />

anti-β7, and anti-E-cadherin antibodies. This result is unusual as MTC-1 cells have<br />

173


previously been shown to bind to E-cadherin-coated slides (Berg et al. 1999b). Further, it<br />

would be expected that cell binding to E-cadherin-coated slides would involve a multivalent<br />

interaction which would thereby increase the affinity of interaction. A possible explanation<br />

for this is that the recombinant E-cadherin became denatured on the slides, or assumed a<br />

conformation that rendered it inaccessible to αEβ7. In any event, an adhesion assay based on<br />

the binding of E-cadherin in solution to αEβ7 was established, and used in the subsequent<br />

study.<br />

4.2. Analysis of β7 integrin signalling pathways<br />

Integrins are implicated in many different intracellular signalling pathways (Hynes 2002;<br />

Krissansen et al. 2006a). Various chemical inhibitors were tested for their ability to prevent<br />

TK-1 cell adhesion to MAdCAM-1 in order to identify signalling pathways and molecules<br />

responsible for activating the adhesion and/or clustering of α4β7. Genistein, a general<br />

tyrosine kinase inhibitor, was shown to inhibit the G-protein-mediated activation of αLβ2.<br />

Thus, it inhibited AlF4 - -activated, but not PMA or Mn 2+ -activated adhesion of TAM2D2 T-<br />

cell hybridoma cells to ICAM-1 (Driessens et al. 1997). Similarly, α4β7-mediated adhesion<br />

of TK-1 cells to MAdCAM-1 in response to activation with AlF4 - was inhibited by genistein<br />

(Zhang et al. 1999). In contrast to αLβ2, α4β7-mediated adhesion of Mn 2+ -activated TK-1<br />

cells and PMA-differentiated HL60 cells was also inhibited by genistein (Walsh 1996). This<br />

thesis confirmed that genistein inhibits Mn 2+- , PMA- and AlF4 - activation of α4β7-mediated<br />

adhesion. <strong>The</strong> fact that genistein does not disrupt Mn 2+ and PMA-mediated activation of<br />

αLβ2 is evidence that different integrins utilize different signalling pathways (Driessens et al.<br />

1997).<br />

It has been shown that different integrins within the same family differ in their response to<br />

agonists. For example, it is much more difficult to activate αXβ2 than other β2 integrins (Lu<br />

et al. 2001a), and αIIbβ3 is more resistant to the effects of Mn 2+ than is αVβ3 (Kamata et al.<br />

2005). Several studies have looked into the dose-dependency of activation of integrins by<br />

divalent cations. Conformation studies of the α1 helix of the β1 integrin subunit A-domain<br />

reveal that Mn 2+ is a potent activator of β1 integrins. Mn 2+ binds to the β1 integrin A-domain<br />

MIDAS and promotes a shift in position of the α1 helix (Mould et al. 2002). It can activate<br />

integrins at subnanomolar ranges, with activation being maximal at about 2 mM<br />

concentration. Mg 2+ also binds to the MIDAS and activates integrin ligand-binding, however<br />

at a much lower level. In contrast, Ca 2+ is unable to cause the same conformational changes,<br />

174


however it is likely to also bind to the MIDAS site because Ca 2+ can support low affinity<br />

ligand-binding to α5β1, and high affinity binding of activation-independent ligands to α4β1<br />

(Chen et al. 2001). However, the binding of Ca 2+ to sites other than MIDAS appears to favour<br />

an inactive conformation. Increasing doses of Ca 2+ up to 2 mM cause increased reduction in<br />

ligand-binding (Mould et al. 2002). Questions have been raised whether Mn 2+ -induced<br />

activation accurately reflects physiological integrin activation. <strong>The</strong> shift of the α1 helix by<br />

Mn 2+ parallels the shift of the α1 helix by inside-out signalling of β2 integrins in response to<br />

activating cytoplasmic domain mutations and mAb binding (Lu et al. 2001c). Recently, a<br />

trivalent cation gadolinium (Ga 3+ ) was shown to activate β2 integrins at concentration doses<br />

in the 10 µM range, and promote extended conformation of integrins as does Mn 2+ . However<br />

it did not alter intracellular Ca 2+ levels of neutrophils which is a feature exhibited by Mn 2+<br />

(Zhang et al. 2008b). Mn 2+ stimulation has also been found to activate integrins in a PLC-<br />

dependent fashion (Driessens et al. 1997).<br />

Doses of PMA ranging from 0.1 to 16 nM have been shown to induce integrin-mediated<br />

binding of gastric carcinoma cells to fibronectin. PMA induces the tyrosine phosphorylation<br />

of FA proteins such as FAK on Y397, paxillin on Y118, and c-src on Y416 (Lee et al. 2006).<br />

AlF4 - mimics the γ-phosphate of GTP and activates the heterotrimeric G-proteins (Coleman<br />

et al. 1994). It is thought that G-proteins activate PLCγ through phosphorylation by the<br />

tyrosine kinase syk, which leads to the hydrolysis of PIP2 and formation of IP3 and DAG.<br />

IP3 mobilises intracellular Ca 2+ intracellular stores, and together with DAG activates PKC<br />

and Rap1-GEF (Figure 4.1; Driessens et al. 1997). <strong>The</strong> activation of Rap1-GTP activates<br />

RIAM and recruits talin to the β integrin cytoplasmic domain leading to integrin activation<br />

(Han et al. 2006). G-proteins can also be activated by agonist G-protein coupled receptors or<br />

tyrosine kinase coupled receptors, which leads to either the activation of PKC through RhoA,<br />

or PKC and DAG by PI3-kinase and PLC. Additionally, PMA also activates DAG, which<br />

leads to the activation of PKC and Rap1-GTP, and so forth (Han et al. 2006). As mentioned<br />

above, DAG together with Ca 2+ activates PKC and Rap1-GEF which leads to Rap1-GTP-<br />

RIAM activation and the recruitment of talin (Figure 4.1).<br />

175


Figure 4.1 Pathways which lead to integrin activation.<br />

Depicted are some of the signalling pathways which lead to integrin activation. <strong>The</strong> Mn 2+ cation can bind<br />

directly to the integrin extracellular domain to activate integrins which mimics inside-out integrin activation.<br />

Alternatively it can activate integrins through a PLC-dependent pathway, which may signal directly through<br />

PKC or through DAG, which recruits Rap1-GEF leading to Rap1-GTP activation. PMA stimulation can directly<br />

activate DAG which signals through PKC or Rap1-GEF leading to Rap1-GTP activation. TCR signalling can<br />

lead to Rap1-GTP activation. AlF4 - activates G-proteins which leads to either PI3-kinase signalling to PKC and<br />

Rap1-GTP activation, or through PLC leading to Rap1-GTP activation. Agonists such as cytokines can activate<br />

G-protein coupled receptors or tyrosine kinase receptors leading to the activation of G-proteins, and the<br />

subsequent activation of RhoA which signals through PKC or through PI3-kinase, or tyrosine kinase syk leading<br />

to the activation of Rap1-GTP. Activated Rap1-GTP interacts with RIAM which recruits talin to the β subunit<br />

cytoplasmic domain and initiates integrin activation. This figure was generated from information gathered from<br />

published data (Driessens et al. 1997; Kinashi 2005; Han et al. 2006)<br />

Stimulation of TK-1 cells via the α4β7 integrin results in the activation of src and MAP-<br />

kinases (Uhlemann et al. 1997). <strong>The</strong> three MAP kinase cascades are well known signalling<br />

pathways that are activated in response to signals originating from growth factor receptors,<br />

integrins, src and fyn, and G-protein coupled receptors (Ramos 2008). <strong>The</strong> ERK, JNK and<br />

p38 MAPK pathways have an important role in cell migration, and may be activated via the<br />

Ras-Raf-MEK cascade triggered by integrin engagement (Desban et al. 2006). In addition,<br />

p38 MAPK has been reported to mediate the downregulation of the integrin α4 subunit in<br />

human mammary epithelial cells transformed to overexpress the small G-protein erbB-2,<br />

176


which is responsible for remodelling of the cytoskeleton (Woods Ignatoski et al. 2003). This<br />

thesis investigated the involvement of the MAP kinase cascade and src kinases in α4β7<br />

activation, and the consequent adhesion to MAdCAM-1. Specific signalling/kinase inhibitors<br />

were introduced into cells prior to performing the cell adhesion assays. A 3 hour incubation<br />

of TK-1 cells with PD98059 and SB203580, inhibitors of MEK and p38 MAP kinase,<br />

respectively (Cuenda et al. 1995; Dudley et al. 1995), had no observable effect on preventing<br />

Mn 2+ or PMA activated α4β7-mediated cell adhesion (Table 4.1). A brief incubation with the<br />

inhibitors was chosen in order to prevent any changes to gene transcription. This finding is<br />

consistent with previously published data in which treatment of TK-1 cells with SB203580<br />

and PD98059 for 2 hours had no effect on TK-1 cell adhesion to MAdCAM-1. In contrast,<br />

incubation of TK-1 cells with SB203580 for 48 hours decreased cell adhesion to MAdCAM-<br />

1 by approximately 50% (Shafiei 2004). This same study reported that the surface expression<br />

of the β7 integrin subunit was decreased after 48 hours of treatment with SB203580, and in<br />

accord α4 and β7 gene transcription was downregulated.<br />

Table 4.1 Summary of the effects of chemical inhibitors on β7-mediated cell adhesion<br />

Inhibitors Inhibit β7-mediated cell adhesion? Inhibition of other integrins<br />

genistein yes (###) αLβ2 (Driessens et al. 1997).<br />

PD98059 no -<br />

SB203580 no -<br />

JNK-I-1 yes (#) -<br />

JNK-I-2 no -<br />

AG99 no -<br />

PP2 yes (#) -<br />

damnacanthal yes (##) -<br />

radicicol no -<br />

ML-7 yes (##) αLβ2 (Smith et al. 2003)<br />

# denotes inhibition of less then 15% at the highest concentration tested<br />

## denotes inhibition of between 10 and 50% at the highest concentration tested<br />

### denotes inhibition of greater then 50% at the highest concentration tested<br />

This thesis did not examine the role of the JNK pathway in the activation of β7 integrins.<br />

However, recent published findings provide some information on this subject. Jun N-terminal<br />

kinases (JNKs) are a subfamily of the MAPKs that have been implicated in many integrinmediated<br />

signalling events and processes (Dolfi et al. 1998; Oktay et al. 1999; Pankov et al.<br />

2003). <strong>The</strong> JNK inhibitor, SP600125, was previously found to downregulate β7 gene<br />

transcription after 1 hour of incubation, and decreased the expression of the β7 subunit on the<br />

cell surface of TK-1 cells after 48 hours of treatment (Shafiei 2004). This study showed that<br />

177


one inhibitor of the JNK pathway, JNK inhibitor-1 (JNK-I-1), could significantly decrease<br />

Mn 2+ and PMA-induced β7 integrin activation (Table 4.1). However another inhibitor, JNK<br />

inhibitor-2 (JNK-I-2, SP600125) had no detectable effect. JNK-I-1 is a peptide which blocks<br />

JNK phosphorylation, and thereby blocks the transactivation of c-jun, but has no effect on the<br />

activities of ERK1/2 and p38 (Barr et al. 2002). JNK-I-2 inhibits JNK by competing for ATP,<br />

and thereby inhibits the phosphorylation of c-jun, and blocks the cellular expression of<br />

cytokines such as IL-2, IFN-γ, and TNF-α (Bennett et al. 2001). <strong>The</strong> reason for the<br />

differences in the abilities of JNK-I-1 and -2 to inhibit β7 integrin activation is not clear.<br />

RTKs which mediate signalling by growth factors and cytokines, are involved in a variety of<br />

biological processes including cell migration (Friedman et al. 2006). <strong>The</strong>re is considerable<br />

evidence demonstrating that integrins co-operate with RTKs. Thus, integrins associate with<br />

RTKs to stimulate downstream signalling pathways, including MAPK that is necessary for<br />

cell migration (Guo et al. 2004). Integrin-dependent adhesion triggers ligand-independent<br />

activation of the RTK EGFR (Cabodi et al. 2004). In the present study, an inhibitor of EGFR<br />

tyrosine kinase, namely AG99, which inhibits the phosphorylation of EGFR and the<br />

activation of ERK1 and ERK2 (Pai et al. 1998), had no effect on α4β7-mediated adhesion of<br />

TK-1 cells (Table 4.1). This result suggests that EGFR, ERK1 and ERK2 are not involved in<br />

the activation/clustering of α4β7 induced by Mn 2+ or PMA. It does not rule out the possibility<br />

that EGFR, ERK1 and ERK2 are involved in another function of α4β7 signalling, such as cell<br />

migration, spreading, etc.<br />

<strong>The</strong> src family of tyrosine kinases (SFK) are intrinsically involved in many signalling<br />

transduction pathways, including integrin signalling (Playford et al. 2004; Mitra et al. 2006).<br />

<strong>The</strong> SFK inhibitor PP2 (Hanke et al. 1996) and the lck inhibitor damnacanthal (Faltynek et al.<br />

1995) only slightly inhibited α4β7-mediated cell adhesion by PMA-activated TK-1 cells. In<br />

contrast, the src inhibitor radicicol (Kwon et al. 1992) did not appear to prevent α4β7mediated<br />

adhesion of TK-1 cells to MAdCAM-1 (Table 4.1). <strong>The</strong> only minor inhibition of<br />

α4β7-mediated cell adhesion by PP2 (PMA activation) and damnacanthal (Mn 2+ activation)<br />

suggests that SFKs are not responsible for integrin activation in response to the agents used in<br />

this study. Alternatively, SFKs may show redundancy in their ability to mediate integrin<br />

activation, such that it may be necessary to inactivate several SFKs in order to demonstrate<br />

the importance of SFKs for α4β7 activation. For example, combinational targeting of SFKs<br />

by gene deletion leads to an increased frequency of perinatal lethality (Stein et al. 1994).<br />

178


Myosin light chain kinase (MLCK) phosphorylates and regulates myosin binding to actin<br />

filaments which is required for actin contraction (Bresnick 1999). MLCK was found to<br />

contribute to αLβ2-mediated attachment and migration of T cells on ICAM-1 (Smith et al.<br />

2003). In the current study, inhibition of MLCK by ML-7 (Saitoh et al. 1987) significantly<br />

decreased PMA and Mn 2+ mediated activation of TK-1 cell adhesion to MAdCAM-1 (Table<br />

4.1). This result suggests that MLCK is required for the activation/clustering of α4β7 and<br />

attachment of TK-1 cells to MAdCAM-1. Inhibition of cell adhesion by ML-7 was not<br />

complete indicating that other components are involved.<br />

Blockade of α4β7-mediated TK-1 cell adhesion to MAdCAM-1 with chemical inhibitors has<br />

identified several signalling molecule candidates that participate in α4β7 signalling.<br />

Nevertheless, the above approach is confounded by two factors. Integrin signalling can<br />

involve multiple redundant pathways, and hence when one pathway or molecule is inhibited,<br />

another pathway or molecule may compensate. Hence, it cannot be assumed that the failure<br />

of an inhibitor to inhibit α4β7-mediated TK-1 cell adhesion necessarily means that the<br />

pathway it inhibits does not play a role in α4β7 activation/clustering. Another factor concerns<br />

the specificity of the chemical inhibitors. Chemical inhibitors are not specific, and in<br />

particular can inhibit other molecules when used at high concentrations. Hence it cannot be<br />

assumed that the ability of an inhibitor to inhibit α4β7-mediated TK-1 cell adhesion<br />

necessarily means that the nominated target molecule is responsible. <strong>The</strong> above results can<br />

only give clues to potential candidates, and further work is required using other types of<br />

inhibitor e.g. using RNAi- knockdown, cell-permeable inhibitor peptide technology etc.<br />

In this study, the effectiveness or viability of the inhibitors was not tested in parallel, but<br />

relied on the manufacturers’ product certification. <strong>The</strong> inhibitors that showed no activity<br />

should be tested in parallel in assays where they are known to exert activity.<br />

<strong>The</strong> length of incubation with the inhibitors needs further consideration. A 3 hour incubation<br />

of cells with the inhibitor was employed, according to a previously published thesis from the<br />

host laboratory (Walsh 1996). Shorter or longer incubations with the inhibitors may produce<br />

different results depending on the pathways involved and how the inhibitors interact with<br />

them. <strong>The</strong>refore a time-course monitoring of inhibitor activity may be useful.<br />

179


4.3. <strong>The</strong> YDRREY motif<br />

<strong>The</strong> YDRREY motif within the β7 subunit cytoplasmic domain was previously found to be a<br />

cell adhesion regulatory domain (CARD; Krissansen et al. 2006b). Cell-permeable peptides<br />

containing the YDRREY motif prevented T-cells from binding to MAdCAM-1 and VCAM-1<br />

by preventing α4β7 activation and clustering. <strong>The</strong> tyrosine residues flanking the motif were<br />

deemed to be critical for the peptide’s activity, as substitution of the tyrosines with<br />

phenylalanines abrogated the ability of the peptide to prevent cell adhesion. This thesis<br />

confirmed that a cell-permeable form of the YDRREY peptide could prevent activated TK-1<br />

cells from binding to MAdCAM-1. In addition, the peptide was found to prevent αEβ7-<br />

mediated adhesion of H9 T-cells from binding to E-cadherin. Thus, the YDRREY motif is a<br />

CARD for both α4β7 and αEβ7.<br />

<strong>The</strong> YDRREY motif is unique to the β7 integrins. <strong>The</strong> β1 and β5 subunits contain the core<br />

DRRE sequence but lack the flanking tyrosines. <strong>The</strong> β2 subunit, with its partial motif<br />

SDLREY, is the only other integrin β subunit that contains a flanking tyrosine. A key<br />

question was whether the YDRREY peptide acted as a kinase “sink” i.e. whether it competed<br />

with the natural kinase substrates by being a scaffold for phosphorylation. <strong>The</strong> YDRREY<br />

peptide was synthesized as an already phosphorylated form with phosphorylated tyrosines.<br />

<strong>The</strong> YDRREY phosphopeptide showed a similar level of activity in preventing α4β7-<br />

mediated adhesion of TK-1 cells to MAdCAM-1, suggesting that the YDRREY peptide does<br />

not serve as a kinase “sink”. Interestingly, the proximity of the tyrosines also had no effect on<br />

the activity of the peptide. Thus, the YDRGGGGREY peptide containing a spacer of four<br />

glycine residues was as active as the native peptide. A peptide in which the internal DRRE<br />

sequence was modified to either DGGE or EEEE still retained some activity, albeit the<br />

concentration of peptide required to achieve similar levels of inhibition was much greater<br />

(approximately 2 fold compared to the YDRREY peptide). A 4-mer multimer of the<br />

YDRREY peptide, YDRREYGYDRREYGYDRREYGYDRREY, prevented α4β7-mediated<br />

adhesion of TK-1 cells to MAdCAM-1 as effectively as the native peptide. It was anticipated<br />

that the multimer by acting in a multivalent fashion would be substantially more active than<br />

the native peptide, but this was not the case. Surprisingly, the 4-mer<br />

FDRREFGFDRREFGFDRREFGFDRREF in which the tyrosines had been substituted with<br />

phenylalanines, retained activity which was surprising given that the flanking tyrosines had<br />

been found to be essential for peptide activity. Taking all of the above data into account, the<br />

180


esults suggest that the tyrosine residues are important for the activity of the YDRREY<br />

peptide, but the proximity of the tyrosines and their state of phosphorylation is not critical.<br />

<strong>The</strong> core region DRRE also contributes to the activity of the CARD where the acidic<br />

residues, but not the internal basic residues, are essential. A multimer of FDRREF appears to<br />

provide a sequence of amino acids that can compensate for the lack of tyrosines.<br />

Interestingly, substitution of the SDLREY motif in the β2 subunit for SDRREY induces<br />

clustering of LFA-1 in K562 cells (Bleijs et al. 2001).<br />

4.4. Phylogenetic study of the sequence of the YDRREY motif in the β7<br />

subunit<br />

Further insights into the structure-function relationships of the YDRREY motif can be gained<br />

by studying the motif present in the β7 subunits of different species. <strong>The</strong> Ensembl website<br />

(www.ensembl.org) was searched for nucleotide sequences encoding all orthologues of the<br />

β7 subunit. Mammalian orthologues were identified for chimpanzee, gorilla, orangutan,<br />

megabat, horse, hyrax, guinea pig, dolphin, mouse, mouse lemur, macaque, elephant,<br />

microbat, and armadillo. Non-mammaliam orthologues were identified for frog, anole lizard,<br />

zebrafish, and hagfish. <strong>The</strong> deduced protein sequences of the cytoplasmic domains for each<br />

orthologue are listed in Table 4.2. <strong>The</strong> β7 integrin cytoplasmic domain is well conserved in<br />

mammals, with protein identity ranging from 99% for human, chimpanzee, gorilla, and<br />

orangutan sequences, to 53% and 48% for cats and armadillo, respectively, in comparison<br />

with the human sequence (data not shown).<br />

181


Species<br />

Human (Homo sapiens)<br />

Chimpanzee (Pan troglodytes)<br />

Gorilla (Gorilla gorilla)<br />

Orangutan (Pongo pygmaeus)<br />

Megabat (Pteropus vampyrus)<br />

Horse (Equus caballus)<br />

Cow (Bos taurus)<br />

Hyrax (Procavia capensis)<br />

Rat (Rattus norvegicus)<br />

Guinea Pig (Cavia porcellus)<br />

Dolphin (Tursiops truncatus)<br />

Mouse (Mus musculus)<br />

Mouse Lemur (Microcebus murinus)<br />

Macaque (Macaca mulatta)<br />

Elephant (Loxodonta africana)<br />

Kangaroo rat (Dipodomys ordii)<br />

Pika (Ochotona princeps)<br />

Rabbit (Oryctolagus cuniculus)<br />

Microbat (Myotis lucifugus)<br />

Dog (Canis familiaris)<br />

Cat (Felis catus)<br />

Armadillo (Dasypus novemcinctus)<br />

Frog (Xenopus tropicalis)<br />

Anole Lizard (Anolis carolinensis)<br />

Zebrafish (Danio rerio)<br />

Hagfish (Eptatretus burgeri)<br />

182<br />

Cytoplasmic domain sequence (1-20)<br />

RLSVEIYDRR EYSRFEKEQQ<br />

RLSVEIYDRR EYSRFEKEQQ<br />

RLSVEIYDRR EYSRFEKEQQ<br />

RLSVEIYDRR EYSRFEKEQQ<br />

RLSVEIYDRR EYNRFEKERQ<br />

RLSVEIYDRR EYRRFEKERQ<br />

RLSVEIYDRR EFHRFEKERQ<br />

RISVEIYDRR EYSRFEKEQK<br />

RLSVEVYDRL EYSRFEKERQ<br />

RLLVEIYDRR EYSRFEKEQL<br />

RLSVEIYDRR EYSRFEKEQQ<br />

RLSVEIYDRR EYRRFEKEQQ<br />

RLSVEIYDRR EYRRFEKEQQ<br />

RLSVEIYDRR EYSRFEKEQQ<br />

RISVEIYDRR EYSRFEKEQQ<br />

RLLVEIYDRK EYIRFEKEQQ<br />

RVLVEIYDRR EFSRFEKERQ<br />

RLLVEIYDRR EFNRFEKERQ<br />

RLSVEIYDRR EYKRFEKERQ<br />

RLSVEIYDRR EFSRFEKEQK<br />

RLSVEIYDRQ EYNRFEKERQ<br />

RLSVEIYDRR EYNRFEKEQR<br />

SITVEIYDRQ EYNRFQKERS<br />

RVVVDVYDRR EFNRFEKECQ<br />

RLLLELYDYR EYQSFVKMQN<br />

KLLTTLYDRR EYXKFEMERS<br />

Table 4.2 Alignment of the first 20 amino acid residues of the β7 subunit cytoplasmic domain of various<br />

animal species<br />

Comparison of protein sequences comprising the first 20 amino acid residues of the β7 subunit cytoplasmic<br />

domains of different animal species. <strong>The</strong> conserved YDRREY motif is highlighted in green, with differences<br />

between humans and other mammals highlighted in red. <strong>The</strong> table was generated using the Ensembl website<br />

(www.ensembl.org), which is a joint project between European Bioinformatics Institute (EBI), an outstation of<br />

the European Molecular Biology Laboratory (EMBL), and the Wellcome Trust Sanger Institute (WTSI).<br />

<strong>The</strong> core region of the YDRREY motif in rat (YDRLEY), kangaroo rat (YDRKEY), and cat<br />

(YDRQEY) has a substitution of the second arginine residue for either L, K, or Q,<br />

respectively. <strong>The</strong> terminal tyrosine residue in cow, pika, rabbit, and dog is substituted with<br />

phenylalanine giving the YDRREF sequence (Table 4.2). <strong>The</strong>re is no example where both


tyrosines are substituted, where an acidic residue is substituted, or where both arginines are<br />

substituted. Thus, in some species subtle changes in the YDRREY motif are permitted,<br />

presumably because structure-function relationships are slightly different, for example the<br />

way in which the YDRREY motif engages its target signalling molecules which are<br />

presumably orthologues of those in humans and will have slightly different structures. <strong>The</strong><br />

YDRREY motif is completely conserved in some non-mammalian species including the<br />

hagfish (Eptatretus burgeri). <strong>The</strong> motif exists as YDRQEY in the frog (Xenopus tropicalis)<br />

and as YDYREY in zebrafish (Danio rerio). In the anole lizard (Anolis carolinensis) the Cterminal<br />

tyrosine is substituted with phenylalanine (YDRREF).<br />

4.5. <strong>The</strong> NPLY motif<br />

A cell-permeable peptide representing the NPLY motif also prevented α4β7 from binding to<br />

MAdCAM-1. <strong>The</strong> NPLY motif is found in several integrin β subunits including the β3, β5,<br />

β6 and β7 subunits. <strong>The</strong> β1 subunit has an NPIY motif at the corresponding region, whereas<br />

the β2 subunit has an NPLF motif, being the only integrin β subunit lacking the terminal<br />

tyrosine. Thus, the NPLY motif, unlike the YDRREY motif, is not specific to the β7 subunit.<br />

<strong>The</strong> membrane proximal NPxY motif is known to bind several phosphotyrosine binding<br />

(PTB) proteins including talin and DOK1. <strong>The</strong> NPLY motif may bind to talin and thereby<br />

prevent talin from binding to β integrin subunits to initiate integrin activation. <strong>The</strong> NPLY<br />

motif is completely conserved in the β7 subunits of all species examined (Table 4.3) with the<br />

exception of the anole lizard (Anolis carolinensis) and zebrafish (Danio rerio) for which the<br />

terminal tyrosine is substituted with a phenylalanine residue. <strong>The</strong> armadillo sequence<br />

terminates before the NPLY motif, however this sequence may be incomplete. An NPRF<br />

motif (Table 4.3) which was not examined in this thesis is conserved in all species except<br />

guinea pig, mouse lemur, kangaroo rat, pika, dog, frog, anole lizard, zebrafish and hagfish.<br />

183


Species<br />

Human (Homo sapiens)<br />

Chimpanzee (Pan troglodytes)<br />

Gorilla (Gorilla gorilla)<br />

Orangutan (Pongo pygmaeus)<br />

Megabat (Pteropus vampyrus)<br />

Horse (Equus caballus)<br />

Cow (Bos taurus)<br />

Hyrax (Procavia capensis)<br />

Rat (Rattus norvegicus)<br />

Guinea Pig (Cavia porcellus)<br />

Dolphin (Tursiops truncatus)<br />

Mouse (Mus musculus)<br />

Mouse Lemur (Microcebus murinus)<br />

Macaque (Macaca mulatta)<br />

Elephant (Loxodonta africana)<br />

Kangaroo rat (Dipodomys ordii)<br />

Pika (Ochotona princeps)<br />

Rabbit (Oryctolagus cuniculus)<br />

Microbat (Myotis lucifugus)<br />

Dog (Canis familiaris)<br />

Cat (Felis catus)<br />

Armadillo (Dasypus novemcinctus)<br />

Frog (Xenopus tropicalis)<br />

Anole Lizard (Anolis carolinensis)<br />

Zebrafish (Danio rerio)<br />

Hagfish (Eptatretus burgeri)<br />

Cytoplasmic domain sequence (21-end)<br />

QLNWKQDSNP LYKSAITTTI NPRFQEADSP TL<br />

QLNWKQDSNP LYKSAITTTI NPRFQEADSP TL<br />

QLNWKQDSNP LYKSAITTTI NPRFQEADSP TL<br />

QLNWKQDSNP LYKSAITTTI NPRFQEADSP TL<br />

QLNWKQDSNP LYKSAITTTV NPRFQETESP LL<br />

QLNWKQDSNP LYKSAITTTV NPRFQEADSP PL<br />

HLNWKQDHNP LYQSAITTTV NPRFQEADSP VL<br />

QLKWKQDNNP LYKSAITTTI NPRFQTESPS L<br />

QLNWKQDSNP LYKSAVTTTV NPRFQGGNKQ SLSLPLTQEA D<br />

QLNWKQDSNP LYKSAITTTI NPQFQRAESP VL<br />

QLHWKQENNP LYKSAITTTV NPRFQEAEGA PL<br />

QLNWKQDNNP LYKSAITTTV NPRFQGTNGR SPSLSLTREA D<br />

QLNWKQDNNP LYKSAITTTI NPCFQDADRP AL<br />

QLNWKQDSNP LYKSAITTTI NPRFQEADSP IL<br />

QLNWNQDNNP LYKSAITTTI NPRFQGRESP LL<br />

QLNWKQDNNP LYQSAITTTV NPNFQGTDGS SL<br />

QLNWNKDNNP LYQSAITTTV NPFYQEAEKT PLR<br />

QLNWNQDNNP LYRSAVTTTI NPRFQEPERP LL<br />

QPNWKQDRNP LYKSAVTTII NPRFQAPESP L<br />

HLNWKQENNP LYRSAITTTV NPQFQETGSL LS<br />

QLNWKQDSNP LYRSAITTTV NPRFQEARSP FL<br />

HLNWKQ<br />

NAQWNELNNP LYKSATTTVK NPHFIEYVRP NV<br />

SAKWNEMSNP LFRSATTTIM NPRY<br />

QTQWKEAQNP LFKGATTTVM NPLHMQNEA<br />

QAKWNKDDNP LYKSATTTVV NPKFDEN<br />

Table 4.3 Alignment of the protein sequences from the distal region of the β7 subunit cytoplasmic of<br />

various animal species<br />

Comparison of protein sequences comprising the distal region (21 st amino acid residue to last residue) of the β7<br />

subunit cytoplasmic domains of different animal species. Conserved regions are highlighted, the NPLY motif in<br />

blue, and the NPRF motif in yellow. Differences in the conserved regions between humans and other mammals<br />

are highlighted red. <strong>The</strong> table was generated using the Ensembl website (www.ensembl.org), which is joint<br />

project between European Bioinformatics Institute (EBI), an outstation of the European Molecular Biology<br />

Laboratory (EMBL), and the Wellcome Trust Sanger Institute (WTSI).<br />

184


4.6. Interaction of the YDRREY motif with tyrosine kinases<br />

Tyrosine residues in the YDRREY motif are important for the function of the β7 CARD as<br />

evidenced by the fact that substitution of the tyrosine residues with phenylalanines destroys<br />

the ability of peptide to disrupt α4β7-mediated cell adhesion. Notably, the activity of the<br />

peptide was not dependent on its state of tyrosine phosphorylation, as both synthetic<br />

phosphorylated and unphosphorylated forms of the peptide were active. Nevertheless, it is<br />

conceivable that the peptide has to be phosphorylated, and that the unphosphorylated cell-<br />

permeable peptide becomes phosphorylated by cellular kinases once it enters the cell. Indeed,<br />

TK-1 cell lysates were shown to phosphorylate the YDRREY peptide. In this thesis, the study<br />

of the tyrosines phosphorylated in the β7 cytoplasmic domain was extended to include all 3<br />

cytoplasmic tyrosines at positions 736, 741, 761. <strong>The</strong> aim was to identify which of the<br />

tyrosine residues is phosphorylatable, and to identify the kinase(s) responsible. A GST-fusion<br />

protein of the β7 cytoplasmic domain was phosphorylatable using a TK-1 cell lysate. <strong>The</strong><br />

identity of the tyrosine residues that were phosphorylatable was determined by creating GST-<br />

fusion proteins of the β7 cytoplasmic domain carrying tyrosine to phenylalanine substitutions<br />

at positions 736, 741, 761 and 736+741 combined. Substitution of each tyrosine with<br />

phenylalanine reduced the degree of phosphorylation of the cytoplasmic domain, as<br />

evidenced by reduced incorporation of 32 P. <strong>The</strong> data clearly showed that all three tyrosines in<br />

the β7 cytoplasmic domain are potentially phosphorylatable.<br />

Immunoprecipitation of α4β7 from activated and non-activated TK-1 cells revealed<br />

differences in the phosphorylation state of the β7 subunit as detected with an antiphosphotyrosine<br />

antibody. <strong>The</strong> β7 subunit isolated from activated cells was more highly<br />

phosphorylated compared to that isolated from non-activated cells. <strong>The</strong>re was no detectable<br />

difference in the degree of tyrosine phosphorylation of the β7 subunit isolated from activated<br />

TK-1 cells as compared to activated TK-1 cells containing bound ligand. <strong>The</strong> results suggest<br />

that the tyrosine phosphorylation state of the β7 subunit is increased upon cell activation,<br />

involving either the YDRREY or NPLY motif or both.<br />

Phosphorylation of conserved motifs within the cytoplasmic domains of integrins can<br />

modulate the binding of intracellular proteins. Phosphorylation of the β-integrin cytoplasmic<br />

domain on threonine residues 783, 783, 785, or β2 residue 758, potentially mediated by<br />

protein kinase C (PKC), inhibits the binding of filamin, but not talin. Phosphorylation of the<br />

β2 subunit cytoplasmic domain on threonine residue 758 enables a 14-3-3 protein to bind<br />

185


(Fagerholm et al. 2005; Kiema et al. 2006; Takala et al. 2008). 14-3-3 proteins have the<br />

ability to bind a multitude of functionally diverse signalling proteins (Morrison 2009).<br />

Tyrosine phosphorylation of the conserved NP(I/L)Y motif, possibly mediated by src,<br />

inhibits talin binding to the β3 subunit but enhances the binding of other PTB-domain-<br />

containing proteins (Fagerholm et al. 2004; Kiema et al. 2006; Oxley et al. 2008). In this<br />

thesis, the tyrosine kinases FAK, src and lck were shown to phosphorylate the YDRREY<br />

motif. Src may not phosphorylate the initial tyrosine residue as the peptide YDRRE missing<br />

the C-terminal tyrosine was not phosphorylated by src. FAK and src were found to bind<br />

directly to the YDRREY multimer. Surprisingly, substitution of the internal arginines with<br />

glycines or glutamic acid residues had a little effect on FAK and src binding, suggesting that<br />

the internal arginines may serve largely as a spacer. FAK was previously reported to bind to<br />

the integrin β1 (KLLMIIHDRREFA), β2 (KALIHLSDLREYR) and β3 (KLLITIHDRKEFA)<br />

cytoplasmic domains at regions corresponding to the YDRREY motif. Replacement of the<br />

aspartic and glutamic acid residues in the β1 subunit DRRE motif with alanines diminished<br />

FAK-binding (Schaller et al. 1995a). Src and fyn were reported to bind to the extreme C-<br />

terminal region 760-RGT-763 of the integrin β3 cytoplasmic domain via their SH3 domains<br />

(Arias-Salgado et al. 2003). Recently fyn was reported to bind to the membrane proximal<br />

region 721-IHDRK-725 of the β3 subunit, a region that overlaps the YDRREY motif (Reddy<br />

et al. 2008). In the present study, src also bound to the FDRREFDRREFDRREFDRREF<br />

peptide, a multimeric form of the FDRREF peptide, suggesting that src can bind to both the<br />

DRRE core and to an YDXXEY peptide. It is possible that both motifs work together to<br />

increase the affinity of binding. As regards the YDXXEY motif, it is interesting that synthetic<br />

polymers of tyrosine and glutamic acid residues serve as substrates for PTK (Braun et al.<br />

1984).<br />

<strong>The</strong> present study found that autophosphorylated and unphosphorylated src binds to the nontyrosine<br />

phosphorylated form of the YDRREY peptide. In contrast, only autophosphorylated<br />

forms of src can bind to the phosphorylated YDRREY peptide. This was expected as<br />

autophosphorylation of src activates the src kinase freeing the SH2 and SH3 domains. <strong>The</strong><br />

SH2 domains are known to bind preferentially to the phosphorylated pY-E-E-I motif due to<br />

the polar and electrostatic interactions of the glutamic acid residues at pY +1 and +2 positions<br />

(Songyang et al. 1993). As other residues can be accommodated in these positions, it has<br />

been predicted that src would bind to the phosphorylated form of the YDRREY motif. In<br />

186


contrast, FAK bound equally to both the tyrosine phosphorylated and non-phosphorylated<br />

forms.<br />

FAK and src appeared to bind synergistically to the YDRREY motif. FAK and src can transphosphorylate<br />

each other to enable their activation (Mitra et al. 2005; Mitra et al. 2006).<br />

Whether the YDRREY motif preferentially binds src over FAK or vice versa would be<br />

interesting to determine. Taken together, FAK and src show differences in their abilities to<br />

phosphorylate and bind to the YDRREY motif, yet can cooperate to increase binding.<br />

4.7. Probing the function of β7 subunit cytoplasmic domain tyrosine<br />

phosphorylation sites by mutational analysis<br />

<strong>The</strong> role of the three tyrosine phosphorylation sites in the β7 subunit cytoplasmic domain in<br />

regulating α4β7-mediated cell adhesion was studied by transfection of plasmids expressing<br />

mutant forms of the β7 subunit in which the tyrosines had been substituted with<br />

phenylalanines. <strong>The</strong> cDNA of the wild-type β7 subunit, and mutant forms bearing the<br />

substitutions Y736F, Y741F, Y761F, and Y736+41F were successfully cloned into a<br />

mammalian expression vector and transfected into HEK-293T cells engineered to express the<br />

integrin α4 subunit. <strong>The</strong> transfectants expressed good levels of expression of α4β7, despite<br />

the fact that HEK-293T cells don’t normally express the leukocyte-restricted α4β7 receptor.<br />

HEK-293T cells were used because of their ease of transfectability, whereas leukocytes are<br />

known to be notoriously difficult to transfect at high efficiency. Another study also utilized<br />

HEK-293T cells to successfully express α4β7 (Arthos et al. 2008). A concern in using HEK-<br />

293T cells is that they are derived from kidney cells, whereas the β7 integrins are restricted in<br />

their expression to leukocytes. It was not known whether HEK-293T cells would express all<br />

the signalling molecules required to activate ectopically-expressed α4β7. Fortunately, Mn 2+ -<br />

activation of HEK-293T cells induced the adhesiveness of α4β7, and its binding to<br />

MAdCAM-1. Individual substitution of the three tyrosines with phenylalanines decreased the<br />

binding of the α4β7 integrin to MAdCAM-1 in each case. Not surprisingly therefore,<br />

combined substitution of both tyrosines 736 and 741 caused the greatest decrease in α4β7mediated<br />

cell binding. <strong>The</strong>se results suggest that each tyrosine is potentially involved in<br />

signalling by β7 integrins, and confirm that the β7 CARD is critical for β7 integrin activation.<br />

While the precise mechanism for the decrease in α4β7-mediated cell adhesion was not<br />

identified in this study, several possibilities can be gleaned from the literature. Talin is known<br />

187


to bind to the NPxY motif. Tyrosine to alanine substitution of this motif in the β1A (Y788A),<br />

β3 (Y747A), and β7 (Y761A) subunits disrupts talin and other PTB proteins from binding to<br />

the integrin subunits (Calderwood et al. 2003). In accord, in the current thesis, substitution of<br />

Y761 with phenylalanine decreased α4β7-mediated cell adhesion possibly due to disruption<br />

of the binding of talin and other PTB proteins. Similarly, mutations of Y736F and Y741F<br />

may disrupt the binding of either talin, src, FAK or other signalling proteins such as WAIT-1<br />

which control β7 integrin activation.<br />

4.8. Interaction of the YDRREY motif with cytoskeletal proteins<br />

It has been established that the binding of talin to the β subunit cytoplasmic domain of<br />

integrins is critical for integrin activation (Tadokoro et al. 2003). In this interaction, the<br />

phosphotyrosine binding domain of talin binds to the membrane proximal region of the β<br />

subunit cytoplasmic domain (Wegener et al. 2007). This thesis examined the interaction<br />

between the β7 cytoplasmic domain motif YDRREY and adaptor proteins such as paxillin,<br />

filamin and α-actinin.<br />

Paxillin is an essential scaffolding protein which is recruited early to integrin adhesion<br />

plaques (Deakin et al. 2008). Paxillin binds to the β1 and β3 subunits at a membrane-<br />

proximal site (β3, 742-KLLITIHDRKE-752) that overlaps the YDRREY motif (Schaller et<br />

al. 1995a; Chen et al. 2000). Paxillin also binds to the α4 subunit cytoplasmic domain (Liu et<br />

al. 1999). <strong>The</strong> results of this thesis demonstrated that paxillin does not affect src binding to<br />

the YDRREY peptide, suggesting either that paxillin does not bind directly to the YDRREY<br />

motif or it binds with lesser affinity than src. Paxillin binds to FAK (Hayashi et al. 2002), and<br />

is phosphorylated by ERK-2 (Liu et al. 2002). In the present study, a complex of paxillin,<br />

FAK, and ERK-2 was able to bind to the YDRREY peptide (Figure 4.2). Paxillin can recruit<br />

FAK to focal adhesions (Hildebrand et al. 1995; Tachibana et al. 1995). This provides the<br />

possibility for paxillin to bind to the α4 subunit, and recruit FAK to the cytoplasmic interface,<br />

where it might be in a position to bind and/or phosphorylate the β7 subunit.<br />

188


Figure 4.2 Proteins which may interact with the β7 cytoplasmic domain<br />

Depicted is the sequence of the β7 cytoplasmic domain, and underlined in black are the regions where<br />

cytoskeletal proteins or kinases may interact directly with the β7 cytoplasmic domain or the cytoplasmic<br />

domains of other integrins. <strong>The</strong> green arrow indicates the interaction of FAK and src with the YDRREY motif,<br />

as reported in this thesis. <strong>The</strong> blue lines indicate complexes formed between cytoskeletal elements and kinases<br />

which bind to the YDRREY motif. <strong>The</strong> red line denotes the antagonistic effect filamin has on src binding to<br />

YDRREY. This figure was constructed from information in the published literature and the results reported in<br />

this thesis.<br />

Filamin is an actin binding protein. Tight binding of filamin to the β1 and β7 subunit tails<br />

decreased cell migration (Calderwood et al. 2001). <strong>The</strong> present study found that filamin<br />

inhibited src binding to the YDRREY motif, but the mechanism for this inhibition was not<br />

determined. Filamin is known to bind to the region 35-AITTTI-40 (refer to Table 4.3, and<br />

Figure 4.2) between the NPLY and NPRF motifs in the β7 subunits, and to corresponding<br />

regions in the β1 and β2 subunits (Calderwood et al. 2001). It has also been reported to bind<br />

to a membrane-proximal region corresponding to where the YDRREY motif would reside in<br />

the β1 integrin subunit (Schaller et al. 1995a). <strong>The</strong>refore it is possible that filamin prevents<br />

src binding to the YDRREY motif by binding to the YDRREY sequence itself. <strong>The</strong> SFK<br />

member lck phosphorylates filamin which modulates actin filament cross-linking (Pal<br />

Sharma et al. 2004). An alternative possibility is that filamin binds to src and thereby<br />

prevents it from binding to the YDRREY peptide.<br />

189


α-actinin is a cytoskeletal protein which has previously been found to bind the cytoplasmic<br />

domains of the β1 and β2 integrin subunits at a site immediately downstream of the region<br />

corresponding to the YDRREY motif (Otey et al. 1990; Pavalko et al. 1993; Sampath et al.<br />

1998). α-actinin binds to src and is phosphorylated by FAK (Izaguirre et al. 2001). <strong>The</strong><br />

present study showed that complexes of α-actinin with src or FAK can bind to the YDRREY<br />

peptide. This result suggests that src or FAK could recruit α-actinin to the β7 subunit<br />

cytoplasmic domain, and thereby provide a link to the cytoskeletal network. It is unlikely that<br />

α-actinin binds directly to the YDRREY motif, as the binding sites for α-actinin within the β1<br />

and β2 integrin subunits were downstream of the YDRREY motif as mentioned above. It is<br />

possible, however, that α-actinin binds to a site downstream of the YDRREY motif in the β7<br />

subunit cytoplasmic domain which could lead to the recruitment of kinases such as src and<br />

FAK (shown diagrammatically in Figure 4.2). In this fashion, multiple assemblies of kinases<br />

could accumulate at the cytoplasmic domain, potentially enhancing signalling.<br />

This thesis did not investigate the roles of other β subunit interacting proteins such as talin<br />

and kindlin, which have recently been reported to activate integrins (Shi et al. 2007;<br />

Montanez et al. 2008; Tadokoro et al. 2003; Tanentzapf et al. 2006). As shown in the diagram<br />

in Figure 4.2, talin binds to the central NPLY motif and to the region around the terminal<br />

tyrosine of the YDRREY motif of the β3 integrin (Wegener et al. 2007; Rodius et al. 2008).<br />

Pull-down assays with the YDRREY peptide to isolate interacting proteins from TK-1 cell<br />

lysates did not show talin binding (data not shown). However, it cannot be excluded that talin<br />

binds to part of the YDRREY motif. <strong>The</strong>refore it would be interesting to investigate which<br />

flanking residues might be required to allow talin to interact with the YDRREY motif.<br />

Kindlin binds to the membrane distal region of the cytoplasmic domains of the β1 and β3<br />

subunits (Shi et al. 2007; Montanez et al. 2008), therefore it is unlikely to interact with the<br />

YDRREY motif, however it may still influence the interactions of YDRREY-binding<br />

proteins.<br />

In summary, this thesis has investigated potential interactions between the YDRREY motif,<br />

cytoplasmic adaptor proteins, and the kinases src and FAK. It has not determined the exact<br />

sequence of interaction, but has provided an insight into potential molecular interactions that<br />

may impact upon intracellular signalling by β7 integrins. In addition, these experiments were<br />

conducted with an isolated YDRREY peptide. Answers to whether the identified interactions<br />

190


hold true in vivo, and for which states of integrin activation, binding, cell migration and so<br />

forth, will only come from undertaking considerable further research.<br />

4.9. In vivo interactions of the YDRREY peptide with cellular kinases<br />

A fluorescently labelled cell-permeable YDRREY peptide was introduced into TK-1 cells to<br />

determine where the peptide would localize once inside the cells. <strong>The</strong> peptide was found to<br />

cluster near the plasma membrane, possibly in FAs. FAK and src also co-localised within<br />

these regions, suggesting that these regions may be sites of interaction between FAK and src<br />

and the YDRREY peptide.<br />

<strong>The</strong> above discussion has considered the interaction of FAK and src with the YDRREY<br />

peptide, which is a contrived interaction. <strong>The</strong> following discussion will examine the evidence<br />

that FAK and src kinase associate with the β7 subunit in vivo. FAK and src were both present<br />

in α4β7 immunoprecipitates, indicating that they associate either directly or indirectly with<br />

the α4 and/or β7 subunits.<br />

Integrin expression on migrating leukocytes can be crudely separated into three different<br />

zones, namely the lamellipodium/lamella at the leading edge, the focal zone at the midsection,<br />

and the uropod at the trailing edge (Figure 4.3; Evans et al. 2009). Adhesive<br />

junctions maintained as migrating cells move over each other, for example immune cells<br />

moving over antigen presenting cells, are termed kinapses [combining roots indicating<br />

movement (kin- from kinesis) and fastening (-apse from haptein)] (Dustin 2008a; Dustin<br />

2008b). Kinapses are asymmetric and can also be separated into three different zones where<br />

SMAC form. <strong>The</strong> lamellipodium is analogous to the dSMAC, the focal zone to the pSMAC,<br />

and the uropod to cSMAC (cSMAC; Dustin 2008a; Evans et al. 2009). Focal adhesions form<br />

at the pSMAC, travel to and mature within the dSMAC, then to the cSMAC. TCR<br />

microclusters can form at the leading lamellipodium. TCR microclusters and integrin clusters<br />

intermix within the integrin-rich lamella, and adhesion and TCR structures disengage at the<br />

trailing uropod as the T cell moves. In this thesis, immunofluorescence studies were used to<br />

localize FAK and src with respect to α4β7 on the surface of TK-1 cells bound to MAdCAM-<br />

1-coated slides. <strong>The</strong> results suggested that FAK and src, but not lck, colocalize with α4β7.<br />

MAdCAM-1 was coated on the slides at low concentrations to facilitate cell spreading and<br />

migration. Fluorescence visualisation revealed that the uropod of migrating cells highly<br />

expressed the β7 subunit, as previously reported for LFA-1 (reviewed in Evans et al. 2009).<br />

191


Src was also strongly expressed by the uropod, whereas FAK was weakly expressed, and lck<br />

was undetectable. <strong>The</strong> β7 subunit and the kinases were also diffusely expressed throughout<br />

the cytoplasm. FAK is thought to perform a variety of roles within a migrating cell from<br />

participating in leading edge formation, FA turnover, and trailing edge retraction (Tomar et<br />

al. 2009). <strong>The</strong>refore FAK is expected to have a homogenous distribution as described in this<br />

thesis. <strong>The</strong> strong presence of both β7 integrin and src in the uropod gives the potential for<br />

molecular interactions between the two molecules. <strong>The</strong> uropod functions to tether LFA-1-<br />

expressing T cells when they are confronted by chemokine-expressing APC. This attachment<br />

at the rear allows the cells to scan their immediate surroundings at the leading edge. In<br />

addition, during leukocyte transmigration the trailing edge can maintain a grip on the surface<br />

as the leading edge penetrates through the endothelial junction (Evans et al. 2009). It is<br />

possible that kinapses play a similar role in α4β7-expressing T-cells. A summary of the<br />

expression profiles of α4β7 and the signalling molecules src, FAK, and lck on migrating TK-<br />

1 cells is represented schematically in Figure 4.3.<br />

Figure 4.3 <strong>The</strong> location of zones on migrating TK-1 cells.<br />

A schematic of a TK-1 cell migrating on MAdCAM-1 showing the different kinapse zones and the expression<br />

levels of α4β7, src, FAK, and lck. <strong>The</strong> leading edge (lamellipodium) expresses low levels of α4β7 and src, and<br />

the mid-zone (focal zone) expresses higher levels of α4β7 and src. <strong>The</strong> trailing edge (uropod) expresses the<br />

highest levels of α4β7 and src. FAK is expressed evenly throughout the different zones, whereas lck is<br />

expressed evenly through the lamellipodium and focal zones, but weakly in the uropod. This figure was adapted<br />

from (Evans et al. 2009) based on the results from this thesis.<br />

192


4.10. α4β7 clusters within SMACs<br />

α4β7 can be induced to cluster on the surface of TK-1 cells by the binding of cells to<br />

MAdCAM-1-coated microspheres (Zhang et al. 1999). TK-1 cells bound to both MAdCAM-<br />

1-coated microspheres and Sepharose/magnetic beads. Confocal microscopy of α4β7 clusters<br />

revealed they were part of multifocal immunological SMACs. This study found that α4β7<br />

localized within a ring-like structure similar to the pSMAC. A parallel can be drawn with the<br />

localization of α4β1 to the pSMAC of mature immunological synapses (Freiberg et al. 2002).<br />

Confocal microscopy revealed that src co-localised with α4β7 within the pSMAC-like<br />

regions, but was also located in the cSMAC, as well as around the entire plasma membrane.<br />

FAK strongly localized within the cSMAC and weakly in the pSMAC with α4β7. Lck has<br />

been reported to colocalize with the TCR within cSMAC (Freiberg et al. 2002). In accord, lck<br />

was predominantly located within the cSMAC in the present study.<br />

Increased filamin binding to integrins restricts integrin-dependent cell migration by inhibiting<br />

transient memebrane protusions and cell polarization (Calderwood et al. 2001). <strong>The</strong>refore,<br />

during various stages of cell adhesion, migration and spreading, a dynamic balance must exist<br />

between the binding of integrin cytoplasmic domains to cytoplasmic proteins which inhibit<br />

and promote integrin activation. Taken together, the results reported in this study indicate that<br />

α4β7 forms part of the structure of SMACs, as does α4β1. It may help to recruit src, but not<br />

FAK which was predominantly located in the cSMAC. Whether these results are<br />

representative of naturally occurring phenomena will need further analysis.<br />

4.11. <strong>The</strong> affect of stress on the function of β7 integrins<br />

Pull-down assays using a synthetic full-length β7 cytoplasmic domain peptide identified four<br />

proteins which belong to the hsp70 family as potential intracellular ligands for the β7<br />

cytoplasmic domain. Confirmation of an interaction between the β7 cytoplasmic domain and<br />

hsp70 was obtained by demonstrating an interaction between recombinant hsp70-1a and the<br />

synthetic full-length β7 cytoplasmic domain peptide. However, the interaction between hsp70<br />

and the β7 cytoplasmic domain peptide was not direct, but rather was facilitated by the<br />

presence of an unknown factor in a TK-1 cell lysate. This result suggests that the interaction<br />

of hsp70 and the β7 cytoplasmic domain requires an adaptor protein, or some other factor to<br />

be present.<br />

193


Hsp70 is a chaperone which assists in the folding of misfolded proteins (Daugaard et al.<br />

2007), therefore it was possible that the binding of hsp70 to the β7 peptide might be<br />

nonspecific i.e. hsp70 was just fulfilling its role as a chaperone. However this was not the<br />

case, as the interaction of hsp70 and β7 integrin was confirmed to take place in vivo, as<br />

evidenced by immunoprecipitation analysis of TK-1 cell lysates. Thus, hsp70 was present in<br />

immune-complexes formed with an anti-β7 antibody, and vice versa, the β7 subunit was<br />

present in immune-complexes formed with an anti-hsp70 antibody. Hsp70 did not co-localise<br />

with β7 integrin in the uropod, but was detected at sites where fluoresceinated YDRREY<br />

clustered in TK-1 cells bound to MAdCAM-1-coated microspheres.<br />

This thesis addressed potential functional relationships between hsp70 and α4β7. Firstly,<br />

hsp70 and α4β7 expression was examined in resting cells and in cells heat-shocked at 42°C.<br />

Whilst α4β7 expression was not changed by heat-shocking, hsp70 expression was increased.<br />

It was discovered that heat-shock treatment of TK-1 cells induced the activation of α4β7<br />

adhesiveness, and delayed its deactivation. <strong>The</strong> dependence of activation on the temperature<br />

of the heat shock treatment was examined. Physiological fever-range temperatures of 39°C<br />

found at inflammatory sites (Chen et al. 2006b) produced similar levels of integrin activation<br />

as induced by heat shocking at 42°C. Fever range hyperthermia has been found to stimulate<br />

α4β7 integrin-dependent lymphocyte adhesion to high endothelial venules in the Peyer’s<br />

patch and mesenteric lymph node.(Evans et al. 2000). In addition, cultured endothelial cells<br />

cultured in fever range hyperthermia conditions augmented actin polymerisation and<br />

enhanced endothelial-derived factors to transactivate α4β7 (Shah et al. 2002). This suggests<br />

that febrile temperatures associated with infection may promote lymphocyte targeting to<br />

those sites. Myocardial stress associated with heat shock, was reported to cause an increased<br />

interaction between integrins and FAK, an increase of paxillin in the membrane fraction of<br />

cell lysates, and protection against lethal ischemic injury (Wei et al. 2003). <strong>The</strong> same group<br />

reported that heat stress led to the activation of AKT via FAK mediated pathways in rat<br />

myocytes (Wei et al. 2008).<br />

<strong>The</strong> hsp70 inhibitor, KNK437, causes a decrease in the adhesion of multiple myeloma cells to<br />

stromal cells and fibronectin (Nimmanapalli et al. 2008). Similarly, this study showed that<br />

KNK437 prevented the binding of TK-1 cells to MAdCAM-1, apparently due to a decrease in<br />

hsp70 expression. Thus, hsp70 is critical for the α4β7-mediated adhesion of TK-1 cells to<br />

MAdCAM-1, where it may potentiate α4β7 activation and/or clustering.<br />

194


Given that heat shocking activated α4β7-mediated adhesion, the possibility that other<br />

stressors might do the same was explored. Cells were stressed by culture in low serum media,<br />

which led to increased activation of α4β7-mediated adhesion. Whether hsp70 was involved<br />

was not explored.<br />

4.12. Summary<br />

Cell adhesion studies of the binding of TK-1 cells to MAdCAM-1-Fc using specific<br />

signalling inhibitors revealed potential intracellular pathways that may control α4β7<br />

activation and clustering. This study suggests the involvement of JNK, SFK, lck and MLCK.<br />

Intracellular signalling pathways are complex, involving many molecules, which are often<br />

further complicated due to crosstalk between the pathways. <strong>The</strong>refore inhibition of one<br />

molecule in one pathway may not be enough to identify the signalling pathways involved<br />

unless it is one which is critical or one in which many pathways converge.<br />

Studies of the YDRREY CARD motif, and its importance to TK-1 cell adhesion to<br />

MAdCAM-1-Fc, revealed that the flanking tyrosines residues, the core DRRE, and the acidic<br />

residues were all functionally important, depending on the overall composition of the motif.<br />

Changes to the central DRRE core such as substitution of the internal arginines was<br />

allowable, and even the tyrosines were redundant in the case of the FDRREF multimer,<br />

which retained the ability to block TK-1 cell adhesion. <strong>The</strong> proximity of the flanking<br />

tyrosines was found not to be a crucial factor. A cell-permeable peptide based on the NPLY<br />

motif was also able to prevent TK-1 cell adhesion to MAdCAM-1, suggesting multiple<br />

interactions between intracellular signalling molecules and the β7 cytoplasmic domain<br />

control α4β7-mediated cell adhesion. Phylogenetic comparison of the sequence of the β7<br />

cytoplasmic domain revealed that the YDRREY motif was well conserved between species.<br />

Only single amino acid substitutions in the motif, in either the core or flanking tyrosines,<br />

were evident in some species.<br />

<strong>The</strong> tyrosine residues in the motif were phosphorylatable by tyrosines kinases such as FAK<br />

and src. FAK and src were also able to bind to the YDRREY motif. FAK could bind both<br />

phosphorylated and unphosphorylated forms, whereas src was found to bind to only the<br />

unphosphorylated form. Cytoskeletal adaptor proteins affected the binding of FAK and src to<br />

the YDRREY motif. Filamin prevented src from binding to the YDRREY motif, whereas<br />

paxillin had no effect.<br />

195


Tyrosine residues within the β7 integrin cytoplasmic domain were found to be important for<br />

activation of α4β7 and its adhesive function. Substitution of the tyrosines with phenylalanines<br />

resulted in decreased α4β7-mediated adhesion of HEK293T cells to MAdCAM-1.<br />

Immunoprecipitation and immunofluorescence studies confirmed that α4β7 interacted with<br />

FAK and src. Thus, FAK and src were present in immunoprecipitates formed with an anti-β7<br />

subunit antibody. FAK and src both co-localised with the α4β7 at the cell-surface following<br />

ligand-induced clustering. SMAC-like formations developed following clustering of α4β7<br />

with ligand-coated microspheres, providing a useful tool to identify proteins present in focal<br />

adhesions in future studies.<br />

Hsp70 was unexpectedly identified as a potential intracellular binding partner of the β7<br />

subunit. It was found to indirectly associate with the β7 subunit. Blocking hsp70 activity<br />

decreased α4β7-mediated cell adhesion to MAdCAM-1. This finding then led to the<br />

discovery that certain stressors could induce α4β7-mediated cell adhesion. Both heat shock<br />

and serum-depletion stress induced the activation of α4β7, which was prolonged compared to<br />

that induced by Mn 2+ -activation. This discovery provides a potential link between the heat<br />

generated during inflammation and the activation of T cells, where heat may lower the<br />

threshold of activation.<br />

<strong>The</strong> YDRREY motif which in part controls β7 integrin-mediated cell adhesion could<br />

potentially be a useful target for therapeutic intervention in the treatment of chronic<br />

inflammatory diseases, including IBD, and multiple sclerosis.<br />

4.13. Future directions<br />

This thesis has generated several interesting findings that could be further explored. Outlined<br />

below are a number of experiments which could help advance the aims of the thesis and our<br />

understanding of the signalling pathways of β7 integrins.<br />

Unravelling β7 integrin-mediated intracellular signalling pathways<br />

<strong>The</strong> chemical inhibition of TK-1 cell adhesion to MAdCAM-1 was not as fruitful as had been<br />

expected, however potential signalling pathways were identified. As briefly mentioned<br />

above, use of a combination of inhibitors might be more productive, however the choice of<br />

inhibitors would need to be carefully considered. To further these experiments it might be<br />

useful to look into the effect of chemical inhibitors on cell migration or cell spreading, which<br />

196


involve downstream signalling pathways after integrin activation, including the MAPK or<br />

JNK pathways. <strong>The</strong> experiments could be carried out using time lapse microscopy to observe<br />

the effects of chemical inhibitors on cell function in real time for a prolonged period. <strong>The</strong><br />

findings could be validated by use of siRNAs targeted against key components of signaling<br />

pathways, given the potential non-specificity of chemical inhibition.<br />

Identification of kinases that regulate β7 integrins and investigation of their roles in<br />

signaling by β7 integrins<br />

<strong>The</strong> roles of the kinases src and FAK, which bind and phosphorylate the YDRREY motif in<br />

the β7 subunit cytoplasmic domain need to be further explored. Are src and FAK involved<br />

independently, or do they have synergistic or antagonistic effects? <strong>The</strong> results in this thesis<br />

suggests synergy, however further experiments need to be done to determine the<br />

spatiotemporal binding and phosphorylation of the YDRREY motif by each kinase. Recent<br />

advances in integrin signalling have provided a clearer picture of integrin activation, such as<br />

the ability of the talin head domain to bind and activate integrins. <strong>The</strong> results of this thesis<br />

raise several questions regarding the activation of β7 integrins. How do src and FAK fit into<br />

the pathways which lead to β7 integrin activation? Do the kinases act before talin binding or<br />

after? What are the roles of cytoskeletal or adaptor proteins in signaling by β7 integrins, and<br />

how do they interact with the kinases. Addressing these questions would lead to a more<br />

complete picture of the signalling pathways which regulate theactivation of β7 integrins.<br />

Fluorescently-tagged integrins and kinases coupled with live cell confocal microscopy could<br />

be used to visualise the interactions between the kinases and integrins in real-time during cell<br />

activation, adhesion, cell spreading, cell migration and integrin clustering. Such studies could<br />

include the β7 cytoplasmic domain mutants described in this study.<br />

Small interference RNA (siRNA) would be a useful tool for studying the roles src and FAK<br />

play in regulating the functions of β7 integrins. It would be interesting to observe the effect of<br />

knockdown of src and FAK on β7 integrin-mediated cell adhesion, migration and spreading.<br />

Potential problems to be surmounted would be the difficulty of transfecting siRNA into<br />

leukocytes and the fact that src and FAK have complex roles in other signalling pathways<br />

thereby making it difficult to observe specific integrin-related effects. <strong>The</strong> former could be<br />

overcome by using viral expression of antisense constructs against src and FAK.<br />

197


It would be useful to screen a library of recombinant kinases with the YDRREY peptide as a<br />

substrate to identify those that are capable of phosphorylating the CARD. Are src and Fak<br />

just two of several kinases that can bind and phosphorylate the YDRREY motif?<br />

SMACs<br />

<strong>The</strong> study of β7 integrins in SMACs would be useful to elucidate the roles of β7 integrins in<br />

antigen presentation or other forms of cell communication. T cells expressing β7 integrins<br />

could be bound to ligand-coated surfaces (plates, microspheres) or to the ligand-expressing<br />

endothelial cells. <strong>The</strong> integrins would be fluorescently tagged and visualized using real-time<br />

confocal microscopy. Identifying changes in integrin localisation in lamellipodia, the focal<br />

zone and uropod during migration could be worthwhile.<br />

Heat shock and stress<br />

As reported here, heat shock and serum depletion caused integrin activation, however the<br />

mechanism which led to integrin activation remains to be determined. Pathways involved in<br />

integrin activation could be elucidated by immunoprecipitation of β7 complexes and<br />

identification of proteins in the complexes by mass spectrometry. siRNA and chemical<br />

inhibitors targeted against components of intracellular signalling pathways could be used to<br />

inhibit activation in response to heat shock, and serum starvation, with Mn 2+ cation activation<br />

as a control. This study only investigated the involvement of hsp70 in β7 integrin activation.<br />

Whether hsp70 regulates the activation of other integrins in response to heat shock warrants<br />

investigation.<br />

<strong>The</strong>rapeutic agents based on the YDRREY motif<br />

<strong>The</strong> activity of the β7 CARD in its peptide form in preventing β7 integrin activation has<br />

therapeutic potential as β7 integrins contribute to the development and/or pathogenesis of<br />

several inflammatory diseases. Peptidomimetic modification of the YDRREY sequence could<br />

be used to create a more powerful small molecular inhibitor that is specific for β7 integrins.<br />

Modifications would include those to increase stability, and overall pharmacodynamics.<br />

Modified peptide sequences would initially be tested for their ability to block β7 integrinmediated<br />

adhesion in an in vitro cell adhesion model. Peptides with improved properties<br />

would then be tested for their ability to prevent the development and/or attenuate the severity<br />

of selected inflammatory diseases (e.g. multiple sclersosis, inflammatory bowel disease),<br />

198


using known animal models. In addition, recent studies have found that the HIV protein<br />

gp120 binds to α4β7, similarly to the way in which α4β7 binds MAdCAM-1, causing rapid<br />

activation of LFA-1 (αLβ2) and cell-to-cell spreading of HIV. <strong>The</strong>refore it would be<br />

interesting to see if the YDRREY peptide can prevent the binding of gp120 to α4β7, and<br />

either reduce cell infection by HIV or reduce cell-to-cell spread of the virus.<br />

199


References<br />

Adair, B. D., J. P. Xiong, C. Maddock, S. L. Goodman, M. A. Arnaout and M. Yeager<br />

(2005). "Three-dimensional EM structure of the ectodomain of integrin<br />

{alpha}V{beta}3 in a complex with fibronectin." J Cell Biol 168(7): 1109-18.<br />

Akiyama, T., J. Ishida, S. Nakagawa, H. Ogawara, S. Watanabe, N. Itoh, M. Shibuya and Y.<br />

Fukami (1987). "Genistein, a specific inhibitor of tyrosine-specific protein kinases." J<br />

Biol Chem 262(12): 5592-5.<br />

Almeida, E. A., D. Ilic, Q. Han, C. R. Hauck, F. Jin, H. Kawakatsu, D. D. Schlaepfer and C.<br />

H. Damsky (2000). "Matrix survival signaling: from fibronectin via focal adhesion<br />

kinase to c-Jun NH(2)-terminal kinase." J Cell Biol 149(3): 741-54.<br />

Apostolaki, M., M. Manoloukos, M. Roulis, M. A. Wurbel, W. Muller, K. A. Papadakis, D.<br />

L. Kontoyiannis, B. Malissen and G. Kollias (2008). "Role of beta7 integrin and the<br />

chemokine/chemokine receptor pair CCL25/CCR9 in modeled TNF-dependent<br />

Crohn's disease." Gastroenterology 134(7): 2025-35.<br />

Arias-Salgado, E. G., S. Lizano, S. Sarkar, J. S. Brugge, M. H. Ginsberg and S. J. Shattil<br />

(2003). "Src kinase activation by direct interaction with the integrin beta cytoplasmic<br />

domain." Proc Natl Acad Sci U S A 100(23): 13298-302.<br />

Arias-Salgado, E. G., S. Lizano, S. J. Shattil and M. H. Ginsberg (2005). "Specification of the<br />

direction of adhesive signaling by the integrin beta cytoplasmic domain." J Biol Chem<br />

280(33): 29699-707.<br />

Arnaout, M. A., S. L. Goodman and J. P. Xiong (2007). "Structure and mechanics of integrinbased<br />

cell adhesion." Curr Opin Cell Biol 19(5): 495-507.<br />

Arnaout, M. A., B. Mahalingam and J. P. Xiong (2005). "Integrin structure, allostery, and<br />

bidirectional signaling." Annu Rev Cell Dev Biol 21: 381-410.<br />

Arregui, C. O., J. Balsamo and J. Lilien (1998). "Impaired integrin-mediated adhesion and<br />

signaling in fibroblasts expressing a dominant-negative mutant PTP1B." J Cell Biol<br />

143(3): 861-73.<br />

Arthos, J., C. Cicala, E. Martinelli, K. Macleod, D. Van Ryk, D. Wei, Z. Xiao, T. D.<br />

Veenstra, T. P. Conrad, R. A. Lempicki, S. McLaughlin, M. Pascuccio, R. Gopaul, J.<br />

McNally, C. C. Cruz, N. Censoplano, E. Chung, K. N. Reitano, S. Kottilil, D. J.<br />

Goode and A. S. Fauci (2008). "HIV-1 envelope protein binds to and signals through<br />

integrin alpha4beta7, the gut mucosal homing receptor for peripheral T cells." Nat<br />

Immunol 9(3): 301-9.<br />

Baillat, G., C. Siret, E. Delamarre and J. Luis (2008). "Early adhesion induces interaction of<br />

FAK and Fyn in lipid domains and activates raft-dependent Akt signaling in SW480<br />

colon cancer cells." Biochim Biophys Acta.<br />

Baker, E., G. R. Sutherland, W. M. Jiang, Q. Yuan, E. Leung, J. D. Watson and G. W.<br />

Krissansen (1992). "Mapping of the human integrin beta 7 gene (ITG beta 7) to<br />

12q13.13 by non-isotopic in situ hybridization." Mamm Genome 2(4): 272-3.<br />

Bakolitsa, C., D. M. Cohen, L. A. Bankston, A. A. Bobkov, G. W. Cadwell, L. Jennings, D.<br />

R. Critchley, S. W. Craig and R. C. Liddington (2004). "Structural basis for vinculin<br />

activation at sites of cell adhesion." Nature 430(6999): 583-6.<br />

Barr, R. K., T. S. Kendrick and M. A. Bogoyevitch (2002). "Identification of the critical<br />

features of a small peptide inhibitor of JNK activity." J Biol Chem 277(13): 10987-97.<br />

Barreiro, O., H. de la Fuente, M. Mittelbrunn and F. Sanchez-Madrid (2007). "Functional<br />

insights on the polarized redistribution of leukocyte integrins and their ligands during<br />

leukocyte migration and immune interactions." Immunol Rev 218: 147-64.<br />

200


Bartolome, R. A., F. Sanz-Rodriguez, M. M. Robledo, A. Hidalgo and J. Teixido (2003).<br />

"Rapid up-regulation of alpha4 integrin-mediated leukocyte adhesion by transforming<br />

growth factor-beta1." Mol Biol Cell 14(1): 54-66.<br />

Baumgart, M., C. Witt, W. Huge and B. Muller (1996). "Increase in the expression of alpha E<br />

beta 7, characteristic of intestinal intraepithelial lymphocytes, on T cells in the lung<br />

epithelium of patients with interstitial lung diseases and in synovial fluid of patients<br />

with rheumatic diseases." Immunobiology 196(4): 415-24.<br />

Bellmann, K., M. Jaattela, D. Wissing, V. Burkart and H. Kolb (1996). "Heat shock protein<br />

hsp70 overexpression confers resistance against nitric oxide." FEBS Lett 391(1-2):<br />

185-8.<br />

Bennett, B. L., D. T. Sasaki, B. W. Murray, E. C. O'Leary, S. T. Sakata, W. Xu, J. C. Leisten,<br />

A. Motiwala, S. Pierce, Y. Satoh, S. S. Bhagwat, A. M. Manning and D. W. Anderson<br />

(2001). "SP600125, an anthrapyrazolone inhibitor of Jun N-terminal kinase." Proc<br />

Natl Acad Sci U S A 98(24): 13681-6.<br />

Berg, R. W., E. Leung, S. Gough, C. Morris, W. P. Yao, S. X. Wang, J. Ni and G. W.<br />

Krissansen (1999a). "Cloning and characterization of a novel beta integrin-related<br />

cDNA coding for the protein TIED ("ten beta integrin EGF-like repeat domains") that<br />

maps to chromosome band 13q33: A divergent stand-alone integrin stalk structure."<br />

Genomics 56(2): 169-78.<br />

Berg, R. W., Y. Yang, K. Lehnert and G. W. Krissansen (1999b). "Mouse M290 is the<br />

functional homologue of the human mucosal lymphocyte integrin HML-1:<br />

antagonism between the integrin ligands E-cadherin and RGD tripeptide." Immunol<br />

Cell Biol 77(4): 337-44.<br />

Berlin, C., E. L. Berg, M. J. Briskin, D. P. Andrew, P. J. Kilshaw, B. Holzmann, I. L.<br />

Weissman, A. Hamann and E. C. Butcher (1993). "Alpha 4 beta 7 integrin mediates<br />

lymphocyte binding to the mucosal vascular addressin MAdCAM-1." Cell 74(1): 185-<br />

95.<br />

Bianchi, M., S. De Lucchini, O. Marin, D. L. Turner, S. K. Hanks and E. Villa-Moruzzi<br />

(2005). "Regulation of FAK Ser-722 phosphorylation and kinase activity by GSK3<br />

and PP1 during cell spreading and migration." Biochem J 391(Pt 2): 359-70.<br />

Bjorge, J. D., A. Pang and D. J. Fujita (2000). "Identification of protein-tyrosine phosphatase<br />

1B as the major tyrosine phosphatase activity capable of dephosphorylating and<br />

activating c-Src in several human breast cancer cell lines." J Biol Chem 275(52):<br />

41439-46.<br />

Bleijs, D. A., G. C. van Duijnhoven, S. J. van Vliet, J. P. Thijssen, C. G. Figdor and Y. van<br />

Kooyk (2001). "A single amino acid in the cytoplasmic domain of the beta 2 integrin<br />

lymphocyte function-associated antigen-1 regulates avidity-dependent inside-out<br />

signaling." J Biol Chem 276(13): 10338-46.<br />

Bonny, C., A. Oberson, S. Negri, C. Sauser and D. F. Schorderet (2001). "Cell-permeable<br />

peptide inhibitors of JNK: novel blockers of beta-cell death." Diabetes 50(1): 77-82.<br />

Bouvard, D., C. Brakebusch, E. Gustafsson, A. Aszodi, T. Bengtsson, A. Berna and R.<br />

Fassler (2001). "Functional consequences of integrin gene mutations in mice." Circ<br />

Res 89(3): 211-23.<br />

Brabek, J., S. S. Constancio, N. Y. Shin, A. Pozzi, A. M. Weaver and S. K. Hanks (2004).<br />

"CAS promotes invasiveness of Src-transformed cells." Oncogene 23(44): 7406-15.<br />

Brandt, E. B., N. Zimmermann, E. E. Muntel, Y. Yamada, S. M. Pope, A. Mishra, S. P.<br />

Hogan and M. E. Rothenberg (2006). "<strong>The</strong> alpha4bbeta7-integrin is dynamically<br />

expressed on murine eosinophils and involved in eosinophil trafficking to the<br />

intestine." Clin Exp Allergy 36(4): 543-53.<br />

201


Braun, S., W. E. Raymond and E. Racker (1984). "Synthetic tyrosine polymers as substrates<br />

and inhibitors of tyrosine-specific protein kinases." J Biol Chem 259(4): 2051-4.<br />

Bresnick, A. R. (1999). "Molecular mechanisms of nonmuscle myosin-II regulation." Curr<br />

Opin Cell Biol 11(1): 26-33.<br />

Briskin, M. J., L. M. McEvoy and E. C. Butcher (1993). "MAdCAM-1 has homology to<br />

immunoglobulin and mucin-like adhesion receptors and to IgA1." Nature 363(6428):<br />

461-4.<br />

Brown, C. M., B. Hebert, D. L. Kolin, J. Zareno, L. Whitmore, A. R. Horwitz and P. W.<br />

Wiseman (2006). "Probing the integrin-actin linkage using high-resolution protein<br />

velocity mapping." J Cell Sci 119(Pt 24): 5204-14.<br />

Bukau, B., J. Weissman and A. Horwich (2006). "Molecular chaperones and protein quality<br />

control." Cell 125(3): 443-51.<br />

Bullard, D. C., X. Hu, T. R. Schoeb, R. C. Axtell, C. Raman and S. R. Barnum (2005).<br />

"Critical requirement of CD11b (Mac-1) on T cells and accessory cells for<br />

development of experimental autoimmune encephalomyelitis." J Immunol 175(10):<br />

6327-33.<br />

Burchiel, S. W. (1999). "<strong>The</strong> effects of environmental and other chemicals on the human<br />

immune system: the emergence of immunotoxicology." Clin Immunol 90(3): 285-6.<br />

Butcher, E. C., M. Williams, K. Youngman, L. Rott and M. Briskin (1999). "Lymphocyte<br />

trafficking and regional immunity." Adv Immunol 72: 209-53.<br />

Cabodi, S., L. Moro, E. Bergatto, E. Boeri Erba, P. Di Stefano, E. Turco, G. Tarone and P.<br />

Defilippi (2004). "Integrin regulation of epidermal growth factor (EGF) receptor and<br />

of EGF-dependent responses." Biochem Soc Trans 32(Pt3): 438-42.<br />

Calalb, M. B., T. R. Polte and S. K. Hanks (1995). "Tyrosine phosphorylation of focal<br />

adhesion kinase at sites in the catalytic domain regulates kinase activity: a role for Src<br />

family kinases." Mol Cell Biol 15(2): 954-63.<br />

Calderwood, D. A. (2004). "Integrin activation." J Cell Sci 117(Pt 5): 657-66.<br />

Calderwood, D. A., Y. Fujioka, J. M. de Pereda, B. Garcia-Alvarez, T. Nakamoto, B.<br />

Margolis, C. J. McGlade, R. C. Liddington and M. H. Ginsberg (2003). "Integrin beta<br />

cytoplasmic domain interactions with phosphotyrosine-binding domains: a structural<br />

prototype for diversity in integrin signaling." Proc Natl Acad Sci U S A 100(5): 2272-<br />

7.<br />

Calderwood, D. A., A. Huttenlocher, W. B. Kiosses, D. M. Rose, D. G. Woodside, M. A.<br />

Schwartz and M. H. Ginsberg (2001). "Increased filamin binding to beta-integrin<br />

cytoplasmic domains inhibits cell migration." Nat Cell Biol 3(12): 1060-8.<br />

Calderwood, D. A., B. Yan, J. M. de Pereda, B. G. Alvarez, Y. Fujioka, R. C. Liddington and<br />

M. H. Ginsberg (2002). "<strong>The</strong> phosphotyrosine binding-like domain of talin activates<br />

integrins." J Biol Chem 277(24): 21749-58.<br />

Calderwood, S. K., M. A. Khaleque, D. B. Sawyer and D. R. Ciocca (2006). "Heat shock<br />

proteins in cancer: chaperones of tumorigenesis." Trends Biochem Sci 31(3): 164-72.<br />

Campi, G., R. Varma and M. L. Dustin (2005). "Actin and agonist MHC-peptide complexdependent<br />

T cell receptor microclusters as scaffolds for signaling." J Exp Med 202(8):<br />

1031-6.<br />

Carlsson, L. E., S. Santoso, C. Spitzer, C. Kessler and A. Greinacher (1999). "<strong>The</strong> alpha2<br />

gene coding sequence T807/A873 of the platelet collagen receptor integrin<br />

alpha2beta1 might be a genetic risk factor for the development of stroke in younger<br />

patients." Blood 93(11): 3583-6.<br />

Carragher, N. O., V. J. Fincham, D. Riley and M. C. Frame (2001). "Cleavage of focal<br />

adhesion kinase by different proteases during SRC-regulated transformation and<br />

apoptosis. Distinct roles for calpain and caspases." J Biol Chem 276(6): 4270-5.<br />

202


Cary, L. A., D. C. Han, T. R. Polte, S. K. Hanks and J. L. Guan (1998). "Identification of<br />

p130Cas as a mediator of focal adhesion kinase-promoted cell migration." J Cell Biol<br />

140(1): 211-21.<br />

Cerf-Bensussan, N., A. Jarry, N. Brousse, B. Lisowska-Grospierre, D. Guy-Grand and C.<br />

Griscelli (1987). "A monoclonal antibody (HML-1) defining a novel membrane<br />

molecule present on human intestinal lymphocytes." Eur J Immunol 17(9): 1279-85.<br />

Chan, B. M., M. J. Elices, E. Murphy and M. E. Hemler (1992). "Adhesion to vascular cell<br />

adhesion molecule 1 and fibronectin. Comparison of alpha 4 beta 1 (VLA-4) and<br />

alpha 4 beta 7 on the human B cell line JY." J Biol Chem 267(12): 8366-70.<br />

Chanmugam, P., L. Feng, S. Liou, B. C. Jang, M. Boudreau, G. Yu, J. H. Lee, H. J. Kwon, T.<br />

Beppu, M. Yoshida and et al. (1995). "Radicicol, a protein tyrosine kinase inhibitor,<br />

suppresses the expression of mitogen-inducible cyclooxygenase in macrophages<br />

stimulated with lipopolysaccharide and in experimental glomerulonephritis." J Biol<br />

Chem 270(10): 5418-26.<br />

Chen, H. C., P. A. Appeddu, H. Isoda and J. L. Guan (1996). "Phosphorylation of tyrosine<br />

397 in focal adhesion kinase is required for binding phosphatidylinositol 3-kinase." J<br />

Biol Chem 271(42): 26329-34.<br />

Chen, J., A. Salas and T. A. Springer (2003). "Bistable regulation of integrin adhesiveness by<br />

a bipolar metal ion cluster." Nat Struct Biol 10(12): 995-1001.<br />

Chen, J., W. Yang, M. Kim, C. V. Carman and T. A. Springer (2006a). "Regulation of<br />

outside-in signaling and affinity by the beta2 I domain of integrin alphaLbeta2." Proc<br />

Natl Acad Sci U S A 103(35): 13062-7.<br />

Chen, L. L., A. Whitty, D. Scott, W. C. Lee, M. Cornebise, S. P. Adams, R. C. Petter, R. R.<br />

Lobb and R. B. Pepinsky (2001). "Evidence that ligand and metal ion binding to<br />

integrin alpha 4beta 1 are regulated through a coupled equilibrium." J Biol Chem<br />

276(39): 36520-9.<br />

Chen, L. M., D. Bailey and C. Fernandez-Valle (2000). "Association of beta 1 integrin with<br />

focal adhesion kinase and paxillin in differentiating Schwann cells." J Neurosci<br />

20(10): 3776-84.<br />

Chen, Q., D. T. Fisher, K. A. Clancy, J. M. Gauguet, W. C. Wang, E. Unger, S. Rose-John,<br />

U. H. von Andrian, H. Baumann and S. S. Evans (2006b). "Fever-range thermal stress<br />

promotes lymphocyte trafficking across high endothelial venules via an interleukin 6<br />

trans-signaling mechanism." Nat Immunol 7(12): 1299-308.<br />

Chigaev, A., A. M. Blenc, J. V. Braaten, N. Kumaraswamy, C. L. Kepley, R. P. Andrews, J.<br />

M. Oliver, B. S. Edwards, E. R. Prossnitz, R. S. Larson and L. A. Sklar (2001). "Real<br />

time analysis of the affinity regulation of alpha 4-integrin. <strong>The</strong> physiologically<br />

activated receptor is intermediate in affinity between resting and Mn(2+) or antibody<br />

activation." J Biol Chem 276(52): 48670-8.<br />

Chong, K. Y., C. C. Lai, S. Lille, C. Chang and C. Y. Su (1998). "Stable overexpression of<br />

the constitutive form of heat shock protein 70 confers oxidative protection." J Mol<br />

Cell Cardiol 30(3): 599-608.<br />

Coleman, D. E., A. M. Berghuis, E. Lee, M. E. Linder, A. G. Gilman and S. R. Sprang<br />

(1994). "Structures of active conformations of Gi alpha 1 and the mechanism of GTP<br />

hydrolysis." Science 265(5177): 1405-12.<br />

Constantin, G., M. Majeed, C. Giagulli, L. Piccio, J. Y. Kim, E. C. Butcher and C. Laudanna<br />

(2000). "Chemokines trigger immediate beta2 integrin affinity and mobility changes:<br />

differential regulation and roles in lymphocyte arrest under flow." Immunity 13(6):<br />

759-69.<br />

203


Cooper, J. A., K. L. Gould, C. A. Cartwright and T. Hunter (1986). "Tyr527 is<br />

phosphorylated in pp60c-src: implications for regulation." Science 231(4744): 1431-<br />

4.<br />

Cooper, L. A., T. L. Shen and J. L. Guan (2003). "Regulation of focal adhesion kinase by its<br />

amino-terminal domain through an autoinhibitory interaction." Mol Cell Biol 23(22):<br />

8030-41.<br />

Crowe, D. T., H. Chiu, S. Fong and I. L. Weissman (1994). "Regulation of the avidity of<br />

integrin alpha 4 beta 7 by the beta 7 cytoplasmic domain." J Biol Chem 269(20):<br />

14411-8.<br />

Cuenda, A., J. Rouse, Y. N. Doza, R. Meier, P. Cohen, T. F. Gallagher, P. R. Young and J. C.<br />

Lee (1995). "SB 203580 is a specific inhibitor of a MAP kinase homologue which is<br />

stimulated by cellular stresses and interleukin-1." FEBS Lett 364(2): 229-33.<br />

Cuenda, A. and S. Rousseau (2007). "p38 MAP-kinases pathway regulation, function and<br />

role in human diseases." Biochim Biophys Acta 1773(8): 1358-75.<br />

Czuchra, A., H. Meyer, K. R. Legate, C. Brakebusch and R. Fassler (2006). "Genetic analysis<br />

of beta1 integrin "activation motifs" in mice." J Cell Biol 174(6): 889-99.<br />

Datta, S. R., H. Dudek, X. Tao, S. Masters, H. Fu, Y. Gotoh and M. E. Greenberg (1997).<br />

"Akt phosphorylation of BAD couples survival signals to the cell-intrinsic death<br />

machinery." Cell 91(2): 231-41.<br />

Daugaard, M., M. Rohde and M. Jaattela (2007). "<strong>The</strong> heat shock protein 70 family: Highly<br />

homologous proteins with overlapping and distinct functions." FEBS Lett 581(19):<br />

3702-10.<br />

de Pereda, J. M., G. Wiche and R. C. Liddington (1999). "Crystal structure of a tandem pair<br />

of fibronectin type III domains from the cytoplasmic tail of integrin alpha6beta4."<br />

EMBO J 18(15): 4087-95.<br />

Deakin, N. O. and C. E. Turner (2008). "Paxillin comes of age." J Cell Sci 121(Pt 15): 2435-<br />

44.<br />

Dedrick, R. L., P. Walicke and M. Garovoy (2002). "Anti-adhesion antibodies efalizumab, a<br />

humanized anti-CD11a monoclonal antibody." Transpl Immunol 9(2-4): 181-6.<br />

Defilippi, P., P. Di Stefano and S. Cabodi (2006). "p130Cas: a versatile scaffold in signaling<br />

networks." Trends Cell Biol 16(5): 257-63.<br />

den Hertog, J., C. E. Pals, M. P. Peppelenbosch, L. G. Tertoolen, S. W. de Laat and W.<br />

Kruijer (1993). "Receptor protein tyrosine phosphatase alpha activates pp60c-src and<br />

is involved in neuronal differentiation." Embo J 12(10): 3789-98.<br />

Derrick, J. M., S. J. Shattil, M. Poncz, R. A. Gruppo and T. K. Gartner (2001). "Distinct<br />

domains of alphaIIbbeta3 support different aspects of outside-in signal transduction<br />

and platelet activation induced by LSARLAF, an alphaIIbbeta3 interacting peptide."<br />

Thromb Haemost 86(3): 894-901.<br />

Desban, N., J. C. Lissitzky, P. Rousselle and J. L. Duband (2006). "alpha1beta1-integrin<br />

engagement to distinct laminin-1 domains orchestrates spreading, migration and<br />

survival of neural crest cells through independent signaling pathways." J Cell Sci<br />

119(Pt 15): 3206-18.<br />

Di Stefano, P., S. Cabodi, E. Boeri Erba, V. Margaria, E. Bergatto, M. G. Giuffrida, L.<br />

Silengo, G. Tarone, E. Turco and P. Defilippi (2004). "P130Cas-associated protein<br />

(p140Cap) as a new tyrosine-phosphorylated protein involved in cell spreading." Mol<br />

Biol Cell 15(2): 787-800.<br />

Dolfi, F., M. Garcia-Guzman, M. Ojaniemi, H. Nakamura, M. Matsuda and K. Vuori (1998).<br />

"<strong>The</strong> adaptor protein Crk connects multiple cellular stimuli to the JNK signaling<br />

pathway." Proc Natl Acad Sci U S A 95(26): 15394-9.<br />

204


Dowling, J. J., E. Gibbs, M. Russell, D. Goldman, J. Minarcik, J. A. Golden and E. L.<br />

Feldman (2008). "Kindlin-2 is an essential component of intercalated discs and is<br />

required for vertebrate cardiac structure and function." Circ Res 102(4): 423-31.<br />

Dransfield, I., C. Cabanas, A. Craig and N. Hogg (1992). "Divalent cation regulation of the<br />

function of the leukocyte integrin LFA-1." J Cell Biol 116(1): 219-26.<br />

Driessens, M. H., P. E. van Hulten, E. A. van Rijthoven, R. D. Soede and E. Roos (1997).<br />

"Activation of G-proteins with AIF-4 induces LFA-1-mediated adhesion of T-cell<br />

hybridoma cells to ICAM-1 by signal pathways that differ from phorbol ester- and<br />

manganese-induced adhesion." Exp Cell Res 231(2): 242-50.<br />

Dudley, D. T., L. Pang, S. J. Decker, A. J. Bridges and A. R. Saltiel (1995). "A synthetic<br />

inhibitor of the mitogen-activated protein kinase cascade." Proc Natl Acad Sci U S A<br />

92(17): 7686-9.<br />

Dunty, J. M., V. Gabarra-Niecko, M. L. King, D. F. Ceccarelli, M. J. Eck and M. D. Schaller<br />

(2004). "FERM domain interaction promotes FAK signaling." Mol Cell Biol 24(12):<br />

5353-68.<br />

Dustin, M. L. (2008a). "T-cell activation through immunological synapses and kinapses."<br />

Immunol Rev 221: 77-89.<br />

Dustin, M. L. (2008b). "Visualization of cell-cell interaction contacts-synapses and<br />

kinapses." Adv Exp Med Biol 640: 164-82.<br />

Elices, M. J., L. A. Urry and M. E. Hemler (1991). "Receptor functions for the integrin VLA-<br />

3: fibronectin, collagen, and laminin binding are differentially influenced by Arg-Gly-<br />

Asp peptide and by divalent cations." J Cell Biol 112(1): 169-81.<br />

Evans, E. A. and D. A. Calderwood (2007). "Forces and bond dynamics in cell adhesion."<br />

Science 316(5828): 1148-53.<br />

Evans, R., I. Patzak, L. Svensson, K. De Filippo, K. Jones, A. McDowall and N. Hogg<br />

(2009). "Integrins in immunity." J Cell Sci 122(Pt 2): 215-25.<br />

Evans, S. S., M. D. Bain and W. C. Wang (2000). "Fever-range hyperthermia stimulates<br />

alpha4beta7 integrin-dependent lymphocyte-endothelial adhesion." Int J<br />

Hyperthermia 16(1): 45-59.<br />

Fagerholm, S. C., T. J. Hilden and C. G. Gahmberg (2004). "P marks the spot: site-specific<br />

integrin phosphorylation regulates molecular interactions." Trends Biochem Sci<br />

29(9): 504-12.<br />

Fagerholm, S. C., T. J. Hilden, S. M. Nurmi and C. G. Gahmberg (2005). "Specific integrin<br />

alpha and beta chain phosphorylations regulate LFA-1 activation through affinitydependent<br />

and -independent mechanisms." J Cell Biol 171(4): 705-15.<br />

Faltynek, C. R., J. Schroeder, P. Mauvais, D. Miller, S. Wang, D. Murphy, R. Lehr, M.<br />

Kelley, A. Maycock, W. Michne and et al. (1995). "Damnacanthal is a highly potent,<br />

selective inhibitor of p56lck tyrosine kinase activity." Biochemistry 34(38): 12404-10.<br />

Fang, D., H. Y. Wang, N. Fang, Y. Altman, C. Elly and Y. C. Liu (2001). "Cbl-b, a RINGtype<br />

E3 ubiquitin ligase, targets phosphatidylinositol 3-kinase for ubiquitination in T<br />

cells." J Biol Chem 276(7): 4872-8.<br />

Feigelson, S. W., V. Grabovsky, E. Winter, L. L. Chen, R. B. Pepinsky, T. Yednock, D.<br />

Yablonski, R. Lobb and R. Alon (2001). "<strong>The</strong> Src kinase p56(lck) up-regulates VLA-<br />

4 integrin affinity. Implications for rapid spontaneous and chemokine-triggered T cell<br />

adhesion to VCAM-1 and fibronectin." J Biol Chem 276(17): 13891-901.<br />

Feng, Y., D. Wang, R. Yuan, C. M. Parker, D. L. Farber and G. A. Hadley (2002). "CD103<br />

expression is required for destruction of pancreatic islet allografts by CD8(+) T cells."<br />

J Exp Med 196(7): 877-86.<br />

205


Forsyth, C. B., E. F. Plow and L. Zhang (1998). "Interaction of the fungal pathogen Candida<br />

albicans with integrin CD11b/CD18: recognition by the I domain is modulated by the<br />

lectin-like domain and the CD18 subunit." J Immunol 161(11): 6198-205.<br />

Freiberg, B. A., H. Kupfer, W. Maslanik, J. Delli, J. Kappler, D. M. Zaller and A. Kupfer<br />

(2002). "Staging and resetting T cell activation in SMACs." Nat Immunol 3(10): 911-<br />

7.<br />

Friedman, A. and N. Perrimon (2006). "A functional RNAi screen for regulators of receptor<br />

tyrosine kinase and ERK signalling." Nature 444(7116): 230-4.<br />

Fujihara, T., H. Fujita, K. Tsubota, K. Saito, K. Tsuzaka, T. Abe and T. Takeuchi (1999).<br />

"Preferential localization of CD8+ alpha E beta 7+ T cells around acinar epithelial<br />

cells with apoptosis in patients with Sjogren's syndrome." J Immunol 163(4): 2226-<br />

35.<br />

Fukuda, T., K. Chen, X. Shi and C. Wu (2003). "PINCH-1 is an obligate partner of integrinlinked<br />

kinase (ILK) functioning in cell shape modulation, motility, and survival." J<br />

Biol Chem 278(51): 51324-33.<br />

Garcia-Alvarez, B., J. M. de Pereda, D. A. Calderwood, T. S. Ulmer, D. Critchley, I. D.<br />

Campbell, M. H. Ginsberg and R. C. Liddington (2003). "Structural determinants of<br />

integrin recognition by talin." Mol Cell 11(1): 49-58.<br />

Geiger, B., A. Bershadsky, R. Pankov and K. M. Yamada (2001). "Transmembrane crosstalk<br />

between the extracellular matrix--cytoskeleton crosstalk." Nat Rev Mol Cell Biol<br />

2(11): 793-805.<br />

Ghosh, S., E. Goldin, F. H. Gordon, H. A. Malchow, J. Rask-Madsen, P. Rutgeerts, P.<br />

Vyhnalek, Z. Zadorova, T. Palmer and S. Donoghue (2003). "Natalizumab for active<br />

Crohn's disease." N Engl J Med 348(1): 24-32.<br />

Giancotti, F. G. and E. Ruoslahti (1999). "Integrin signaling." Science 285(5430): 1028-32.<br />

Giannone, G., G. Jiang, D. H. Sutton, D. R. Critchley and M. P. Sheetz (2003). "Talin1 is<br />

critical for force-dependent reinforcement of initial integrin-cytoskeleton bonds but<br />

not tyrosine kinase activation." J Cell Biol 163(2): 409-19.<br />

Giannone, G., P. Ronde, M. Gaire, J. Beaudouin, J. Haiech, J. Ellenberg and K. Takeda<br />

(2004). "Calcium rises locally trigger focal adhesion disassembly and enhance<br />

residency of focal adhesion kinase at focal adhesions." J Biol Chem 279(27): 28715-<br />

23.<br />

Ginsberg, M. H., A. Partridge and S. J. Shattil (2005). "Integrin regulation." Curr Opin Cell<br />

Biol 17(5): 509-16.<br />

Golubovskaya, V. M., R. Finch and W. G. Cance (2005). "Direct interaction of the Nterminal<br />

domain of focal adhesion kinase with the N-terminal transactivation domain<br />

of p53." J Biol Chem 280(26): 25008-21.<br />

Graf, B., T. Bushnell and J. Miller (2007). "LFA-1-mediated T cell costimulation through<br />

increased localization of TCR/class II complexes to the central supramolecular<br />

activation cluster and exclusion of CD45 from the immunological synapse." J<br />

Immunol 179(3): 1616-24.<br />

Guo, W. and F. G. Giancotti (2004). "Integrin signalling during tumour progression." Nat<br />

Rev Mol Cell Biol 5(10): 816-26.<br />

Gurish, M. F., H. Tao, J. P. Abonia, A. Arya, D. S. Friend, C. M. Parker and K. F. Austen<br />

(2001). "Intestinal mast cell progenitors require CD49dbeta7 (alpha4beta7 integrin)<br />

for tissue-specific homing." J Exp Med 194(9): 1243-52.<br />

Hall, A. (1998). "Rho GTPases and the actin cytoskeleton." Science 279(5350): 509-14.<br />

Han, J., C. J. Lim, N. Watanabe, A. Soriani, B. Ratnikov, D. A. Calderwood, W. Puzon-<br />

McLaughlin, E. M. Lafuente, V. A. Boussiotis, S. J. Shattil and M. H. Ginsberg<br />

206


(2006). "Reconstructing and deconstructing agonist-induced activation of integrin<br />

alphaIIbbeta3." Curr Biol 16(18): 1796-806.<br />

Hanke, J. H., J. P. Gardner, R. L. Dow, P. S. Changelian, W. H. Brissette, E. J. Weringer, B.<br />

A. Pollok and P. A. Connelly (1996). "Discovery of a novel, potent, and Src familyselective<br />

tyrosine kinase inhibitor. Study of Lck- and FynT-dependent T cell<br />

activation." J Biol Chem 271(2): 695-701.<br />

Hanninen, A., M. Salmi, O. Simell and S. Jalkanen (1996). "Mucosa-associated (beta 7integrinhigh)<br />

lymphocytes accumulate early in the pancreas of NOD mice and show<br />

aberrant recirculation behavior." Diabetes 45(9): 1173-80.<br />

Harder, K. W., N. P. Moller, J. W. Peacock and F. R. Jirik (1998). "Protein-tyrosine<br />

phosphatase alpha regulates Src family kinases and alters cell-substratum adhesion." J<br />

Biol Chem 273(48): 31890-900.<br />

Harriman, G. C., M. Brewer, R. Bennett, C. Kuhn, M. Bazin, G. Larosa, P. Skerker, N.<br />

Cochran, D. Gallant, D. Baxter, D. Picarella, B. Jaffee, J. R. Luly and M. J. Briskin<br />

(2008). "Selective cell adhesion inhibitors: Barbituric acid based alpha4beta7--<br />

MAdCAM inhibitors." Bioorg Med Chem Lett 18(7): 2509-12.<br />

Haskill, S., A. D. Yurochko and K. L. Isaacs (1992). "Regulation of macrophage infiltration<br />

and activation in sites of chronic inflammation." Ann N Y Acad Sci 664: 93-102.<br />

Hato, T., N. Pampori and S. J. Shattil (1998). "Complementary roles for receptor clustering<br />

and conformational change in the adhesive and signaling functions of integrin<br />

alphaIIb beta3." J Cell Biol 141(7): 1685-95.<br />

Hayashi, I., K. Vuori and R. C. Liddington (2002). "<strong>The</strong> focal adhesion targeting (FAT)<br />

region of focal adhesion kinase is a four-helix bundle that binds paxillin." Nat Struct<br />

Biol 9(2): 101-6.<br />

Hayashi, Y. K., F. L. Chou, E. Engvall, M. Ogawa, C. Matsuda, S. Hirabayashi, K. Yokochi,<br />

B. L. Ziober, R. H. Kramer, S. J. Kaufman, E. Ozawa, Y. Goto, I. Nonaka, T.<br />

Tsukahara, J. Z. Wang, E. P. Hoffman and K. Arahata (1998). "Mutations in the<br />

integrin alpha7 gene cause congenital myopathy." Nat Genet 19(1): 94-7.<br />

Hehlgans, S., M. Haase and N. Cordes (2007). "Signalling via integrins: implications for cell<br />

survival and anticancer strategies." Biochim Biophys Acta 1775(1): 163-80.<br />

Hildebrand, J. D., M. D. Schaller and J. T. Parsons (1993). "Identification of sequences<br />

required for the efficient localization of the focal adhesion kinase, pp125FAK, to<br />

cellular focal adhesions." J Cell Biol 123(4): 993-1005.<br />

Hildebrand, J. D., M. D. Schaller and J. T. Parsons (1995). "Paxillin, a tyrosine<br />

phosphorylated focal adhesion-associated protein binds to the carboxyl terminal<br />

domain of focal adhesion kinase." Mol Biol Cell 6(6): 637-47.<br />

Hogervorst, F., I. Kuikman, A. E. von dem Borne and A. Sonnenberg (1990). "Cloning and<br />

sequence analysis of beta-4 cDNA: an integrin subunit that contains a unique 118 kd<br />

cytoplasmic domain." EMBO J 9(3): 765-70.<br />

Hogg, N., M. Laschinger, K. Giles and A. McDowall (2003). "T-cell integrins: more than just<br />

sticking points." J Cell Sci 116(Pt 23): 4695-705.<br />

Holcomb, M., A. Rufini, D. Barila and R. L. Klemke (2006). "Deregulation of proteasome<br />

function induces Abl-mediated cell death by uncoupling p130CAS and c-CrkII." J<br />

Biol Chem 281(5): 2430-40.<br />

Holzmann, B. and I. L. Weissman (1989). "Integrin molecules involved in lymphocyte<br />

homing to Peyer's patches." Immunol Rev 108: 45-61.<br />

Horwitz, A. F. (1997). "Integrins and health." Sci Am 276(5): 68-75.<br />

Hsia, D. A., S. T. Lim, J. A. Bernard-Trifilo, S. K. Mitra, S. Tanaka, J. den Hertog, D. N.<br />

Streblow, D. Ilic, M. H. Ginsberg and D. D. Schlaepfer (2005). "Integrin alpha4beta1<br />

207


promotes focal adhesion kinase-independent cell motility via alpha4 cytoplasmic<br />

domain-specific activation of c-Src." Mol Cell Biol 25(21): 9700-12.<br />

Hu, K., L. Ji, K. T. Applegate, G. Danuser and C. M. Waterman-Storer (2007). "Differential<br />

transmission of actin motion within focal adhesions." Science 315(5808): 111-5.<br />

Humphries, J. D., P. Wang, C. Streuli, B. Geiger, M. J. Humphries and C. Ballestrem (2007).<br />

"Vinculin controls focal adhesion formation by direct interactions with talin and<br />

actin." J Cell Biol 179(5): 1043-57.<br />

Humphries, M. J. (2000). "Integrin structure." Biochem Soc Trans 28(4): 311-39.<br />

Hurt, E. M., A. Wiestner, A. Rosenwald, A. L. Shaffer, E. Campo, T. Grogan, P. L.<br />

Bergsagel, W. M. Kuehl and L. M. Staudt (2004). "Overexpression of c-maf is a<br />

frequent oncogenic event in multiple myeloma that promotes proliferation and<br />

pathological interactions with bone marrow stroma." Cancer Cell 5(2): 191-9.<br />

Hyduk, S. J. and M. I. Cybulsky (2002). "Alpha 4 integrin signaling activates<br />

phosphatidylinositol 3-kinase and stimulates T cell adhesion to intercellular adhesion<br />

molecule-1 to a similar extent as CD3, but induces a distinct rearrangement of the<br />

actin cytoskeleton." J Immunol 168(2): 696-704.<br />

Hynes, R. O. (2002). "Integrins: bidirectional, allosteric signaling machines." Cell 110(6):<br />

673-87.<br />

Ilic, D., Y. Furuta, S. Kanazawa, N. Takeda, K. Sobue, N. Nakatsuji, S. Nomura, J. Fujimoto,<br />

M. Okada and T. Yamamoto (1995). "Reduced cell motility and enhanced focal<br />

adhesion contact formation in cells from FAK-deficient mice." Nature 377(6549):<br />

539-44.<br />

Irby, R. B. and T. J. Yeatman (2000). "Role of Src expression and activation in human<br />

cancer." Oncogene 19(49): 5636-42.<br />

Iwahara, T., T. Akagi, Y. Fujitsuka and H. Hanafusa (2004). "CrkII regulates focal adhesion<br />

kinase activation by making a complex with Crk-associated substrate, p130Cas." Proc<br />

Natl Acad Sci U S A 101(51): 17693-8.<br />

Izaguirre, G., L. Aguirre, Y. P. Hu, H. Y. Lee, D. D. Schlaepfer, B. J. Aneskievich and B.<br />

Haimovich (2001). "<strong>The</strong> cytoskeletal/non-muscle isoform of alpha-actinin is<br />

phosphorylated on its actin-binding domain by the focal adhesion kinase." J Biol<br />

Chem 276(31): 28676-85.<br />

Jiang, G., G. Giannone, D. R. Critchley, E. Fukumoto and M. P. Sheetz (2003). "Twopiconewton<br />

slip bond between fibronectin and the cytoskeleton depends on talin."<br />

Nature 424(6946): 334-7.<br />

Jin, M., I. Andricioaei and T. A. Springer (2004). "Conversion between three conformational<br />

states of integrin I domains with a C-terminal pull spring studied with molecular<br />

dynamics." Structure 12(12): 2137-47.<br />

Kamata, T., M. Handa, Y. Sato, Y. Ikeda and S. Aiso (2005). "Membrane-proximal<br />

{alpha}/{beta} stalk interactions differentially regulate integrin activation." J Biol<br />

Chem 280(26): 24775-83.<br />

Kanwar, J. R., J. E. Harrison, D. Wang, E. Leung, W. Mueller, N. Wagner and G. W.<br />

Krissansen (2000). "Beta7 integrins contribute to demyelinating disease of the central<br />

nervous system." J Neuroimmunol 103(2): 146-52.<br />

Kaplan, K. B., K. B. Bibbins, J. R. Swedlow, M. Arnaud, D. O. Morgan and H. E. Varmus<br />

(1994). "Association of the amino-terminal half of c-Src with focal adhesions alters<br />

their properties and is regulated by phosphorylation of tyrosine 527." Embo J 13(20):<br />

4745-56.<br />

Kaplan, K. B., J. R. Swedlow, H. E. Varmus and D. O. Morgan (1992). "Association of p60csrc<br />

with endosomal membranes in mammalian fibroblasts." J Cell Biol 118(2): 321-<br />

33.<br />

208


Kieffer, N. and D. R. Phillips (1990). "Platelet membrane glycoproteins: functions in cellular<br />

interactions." Annu Rev Cell Biol 6: 329-57.<br />

Kiema, T., Y. Lad, P. Jiang, C. L. Oxley, M. Baldassarre, K. L. Wegener, I. D. Campbell, J.<br />

Ylanne and D. A. Calderwood (2006). "<strong>The</strong> molecular basis of filamin binding to<br />

integrins and competition with talin." Mol Cell 21(3): 337-47.<br />

Kilshaw, P. J. (1999). "Alpha E beta 7." Mol Pathol 52(4): 203-7.<br />

Kilshaw, P. J. and S. J. Murant (1990). "A new surface antigen on intraepithelial lymphocytes<br />

in the intestine." Eur J Immunol 20(10): 2201-7.<br />

Kim, M., C. V. Carman and T. A. Springer (2003a). "Bidirectional transmembrane signaling<br />

by cytoplasmic domain separation in integrins." Science 301(5640): 1720-5.<br />

Kim, S., A. K. Chamberlain and J. U. Bowie (2003b). "A simple method for modeling<br />

transmembrane helix oligomers." J Mol Biol 329(4): 831-40.<br />

Kinashi, T. (2005). "Intracellular signalling controlling integrin activation in lymphocytes."<br />

Nat Rev Immunol 5(7): 546-59.<br />

Kishimoto, T. K., N. Hollander, T. M. Roberts, D. C. Anderson and T. A. Springer (1987).<br />

"Heterogeneous mutations in the beta subunit common to the LFA-1, Mac-1, and<br />

p150,95 glycoproteins cause leukocyte adhesion deficiency." Cell 50(2): 193-202.<br />

Knezevic, I., T. M. Leisner and S. C. Lam (1996). "Direct binding of the platelet integrin<br />

alphaIIbbeta3 (GPIIb-IIIa) to talin. Evidence that interaction is mediated through the<br />

cytoplasmic domains of both alphaIIb and beta3." J Biol Chem 271(27): 16416-21.<br />

Kommajosyula, S., S. Reddy, K. Nitschke, J. R. Kanwar, M. Karanam and G. W. Krissansen<br />

(2001). "Leukocytes infiltrating the pancreatic islets of nonobese diabetic mice are<br />

transformed into inactive exiles by combinational anti-cell adhesion therapy." J<br />

Leukoc Biol 70(4): 510-7.<br />

Krissansen, G. W. and E. H. J. Danen (2006a). Integrin: Signalling and Disease, John Wiley<br />

& Sons, Ltd.<br />

Krissansen, G. W. and E. H. J. Danen (2007). Integrin Superfamily, Encyclopedia of Life<br />

Sciences, John Wiley & Sons, Ltd.<br />

Krissansen, G. W., J. Singh, R. K. Kanwar, Y. C. Chan, E. Leung, K. B. Lehnert, J. R.<br />

Kanwar and Y. Yang (2006b). "A pseudosymmetric cell adhesion regulatory domain<br />

in the beta7 tail of the integrin alpha4beta7 that interacts with focal adhesion kinase<br />

and src." Eur J Immunol 36(8): 2203-14.<br />

Kwon, H. J., M. Yoshida, Y. Fukui, S. Horinouchi and T. Beppu (1992). "Potent and specific<br />

inhibition of p60v-src protein kinase both in vivo and in vitro by radicicol." Cancer<br />

Res 52(24): 6926-30.<br />

Lad, Y., P. Jiang, S. Ruskamo, D. S. Harburger, J. Ylanne, I. D. Campbell and D. A.<br />

Calderwood (2008). "Structural Basis of the Migfilin-Filamin Interaction and<br />

Competition with Integrin {beta} Tails." J Biol Chem 283(50): 35154-63.<br />

Lafuente, E. M., A. A. van Puijenbroek, M. Krause, C. V. Carman, G. J. Freeman, A.<br />

Berezovskaya, E. Constantine, T. A. Springer, F. B. Gertler and V. A. Boussiotis<br />

(2004). "RIAM, an Ena/VASP and Profilin ligand, interacts with Rap1-GTP and<br />

mediates Rap1-induced adhesion." Dev Cell 7(4): 585-95.<br />

Lam, S. C.-T. and J. C. Loftus (2005). Integrins, John Wiley & Sons, Ltd.<br />

Lau, T. L., V. Dua and T. S. Ulmer (2008a). "Structure of the integrin alphaIIb<br />

transmembrane segment." J Biol Chem 283(23): 16162-8.<br />

Lau, T. L., A. W. Partridge, M. H. Ginsberg and T. S. Ulmer (2008b). "Structure of the<br />

integrin beta3 transmembrane segment in phospholipid bicelles and detergent<br />

micelles." Biochemistry 47(13): 4008-16.<br />

Laudanna, C., J. J. Campbell and E. C. Butcher (1996). "Role of Rho in chemoattractantactivated<br />

leukocyte adhesion through integrins." Science 271(5251): 981-3.<br />

209


Law, D. A., L. Nannizzi-Alaimo and D. R. Phillips (1996). "Outside-in integrin signal<br />

transduction. Alpha IIb beta 3-(GP IIb IIIa) tyrosine phosphorylation induced by<br />

platelet aggregation." J Biol Chem 271(18): 10811-5.<br />

Lee, J. O., P. Rieu, M. A. Arnaout and R. Liddington (1995). "Crystal structure of the A<br />

domain from the alpha subunit of integrin CR3 (CD11b/CD18)." Cell 80(4): 631-8.<br />

Lee, K. H., A. R. Dinner, C. Tu, G. Campi, S. Raychaudhuri, R. Varma, T. N. Sims, W. R.<br />

Burack, H. Wu, J. Wang, O. Kanagawa, M. Markiewicz, P. M. Allen, M. L. Dustin,<br />

A. K. Chakraborty and A. S. Shaw (2003). "<strong>The</strong> immunological synapse balances T<br />

cell receptor signaling and degradation." Science 302(5648): 1218-22.<br />

Lee, M. S., Y. B. Kim, S. Y. Lee, J. G. Kim, S. H. Kim, S. K. Ye and J. W. Lee (2006).<br />

"Integrin signaling and cell spreading mediated by phorbol 12-myristate 13-acetate<br />

treatment." J Cell Biochem 99(1): 88-95.<br />

Lefrancois, L., C. M. Parker, S. Olson, W. Muller, N. Wagner, M. P. Schon and L.<br />

Puddington (1999). "<strong>The</strong> role of beta7 integrins in CD8 T cell trafficking during an<br />

antiviral immune response." J Exp Med 189(10): 1631-8.<br />

Legate, K. R., E. Montanez, O. Kudlacek and R. Fassler (2006). "ILK, PINCH and parvin:<br />

the tIPP of integrin signalling." Nat Rev Mol Cell Biol 7(1): 20-31.<br />

Leung-Hagesteijn, C. Y., K. Milankov, M. Michalak, J. Wilkins and S. Dedhar (1994). "Cell<br />

attachment to extracellular matrix substrates is inhibited upon downregulation of<br />

expression of calreticulin, an intracellular integrin alpha-subunit-binding protein." J<br />

Cell Sci 107 ( Pt 3): 589-600.<br />

Li, G. C., L. G. Li, Y. K. Liu, J. Y. Mak, L. L. Chen and W. M. Lee (1991). "<strong>The</strong>rmal<br />

response of rat fibroblasts stably transfected with the human 70-kDa heat shock<br />

protein-encoding gene." Proc Natl Acad Sci U S A 88(5): 1681-5.<br />

Lim, S. P., E. Leung and G. W. Krissansen (1998). "<strong>The</strong> beta7 integrin gene (Itgb-7)<br />

promoter is responsive to TGF-beta1: defining control regions." Immunogenetics<br />

48(3): 184-95.<br />

Lin, J., M. J. Miller and A. S. Shaw (2005). "<strong>The</strong> c-SMAC: sorting it all out (or in)." J Cell<br />

Biol 170(2): 177-82.<br />

Lindquist, S. and E. A. Craig (1988). "<strong>The</strong> heat-shock proteins." Annu Rev Genet 22: 631-77.<br />

Ling, K., R. L. Doughman, V. V. Iyer, A. J. Firestone, S. F. Bairstow, D. F. Mosher, M. D.<br />

Schaller and R. A. Anderson (2003). "Tyrosine phosphorylation of type Igamma<br />

phosphatidylinositol phosphate kinase by Src regulates an integrin-talin switch." J<br />

Cell Biol 163(6): 1339-49.<br />

Liu, S. and M. H. Ginsberg (2000). "Paxillin binding to a conserved sequence motif in the<br />

alpha 4 integrin cytoplasmic domain." J Biol Chem 275(30): 22736-42.<br />

Liu, S., S. M. Thomas, D. G. Woodside, D. M. Rose, W. B. Kiosses, M. Pfaff and M. H.<br />

Ginsberg (1999). "Binding of paxillin to alpha4 integrins modifies integrin-dependent<br />

biological responses." Nature 402(6762): 676-81.<br />

Liu, Z. X., C. F. Yu, C. Nickel, S. Thomas and L. G. Cantley (2002). "Hepatocyte growth<br />

factor induces ERK-dependent paxillin phosphorylation and regulates paxillin-focal<br />

adhesion kinase association." J Biol Chem 277(12): 10452-8.<br />

Lollo, B. A., K. W. Chan, E. M. Hanson, V. T. Moy and A. A. Brian (1993). "Direct evidence<br />

for two affinity states for lymphocyte function-associated antigen 1 on activated T<br />

cells." J Biol Chem 268(29): 21693-700.<br />

Lu, C., M. Ferzly, J. Takagi and T. A. Springer (2001a). "Epitope mapping of antibodies to<br />

the C-terminal region of the integrin beta 2 subunit reveals regions that become<br />

exposed upon receptor activation." J Immunol 166(9): 5629-37.<br />

Lu, C., M. Shimaoka, Q. Zang, J. Takagi and T. A. Springer (2001b). "Locking in alternate<br />

conformations of the integrin alphaLbeta2 I domain with disulfide bonds reveals<br />

210


functional relationships among integrin domains." Proc Natl Acad Sci U S A 98(5):<br />

2393-8.<br />

Lu, C., J. Takagi and T. A. Springer (2001c). "Association of the membrane proximal regions<br />

of the alpha and beta subunit cytoplasmic domains constrains an integrin in the<br />

inactive state." J Biol Chem 276(18): 14642-8.<br />

Ludviksson, B. R., W. Strober, R. Nishikomori, S. K. Hasan and R. O. Ehrhardt (1999).<br />

"Administration of mAb against alpha E beta 7 prevents and ameliorates<br />

immunization-induced colitis in IL-2-/- mice." J Immunol 162(8): 4975-82.<br />

Luo, B. H., C. V. Carman and T. A. Springer (2007). "Structural basis of integrin regulation<br />

and signaling." Annu Rev Immunol 25: 619-47.<br />

Luo, B. H., C. V. Carman, J. Takagi and T. A. Springer (2005). "Disrupting integrin<br />

transmembrane domain heterodimerization increases ligand binding affinity, not<br />

valency or clustering." Proc Natl Acad Sci U S A 102(10): 3679-84.<br />

Mancassola, R., S. Lacroix-Lamande, M. Barrier, M. Naciri, H. Salmon and F. Laurent<br />

(2004). "Increased susceptibility of beta7-integrin-deficient neonatal mice in the early<br />

stage of Cryptosporidium parvum infection." Infect Immun 72(6): 3634-7.<br />

Martel, V., C. Racaud-Sultan, S. Dupe, C. Marie, F. Paulhe, A. Galmiche, M. R. Block and<br />

C. Albiges-Rizo (2001). "Conformation, localization, and integrin binding of talin<br />

depend on its interaction with phosphoinositides." J Biol Chem 276(24): 21217-27.<br />

Martin, G. S. (2001). "<strong>The</strong> hunting of the Src." Nat Rev Mol Cell Biol 2(6): 467-75.<br />

Masellis-Smith, A., A. R. Belch, M. J. Mant and L. M. Pilarski (1997). "Adhesion of multiple<br />

myeloma peripheral blood B cells to bone marrow fibroblasts: a requirement for<br />

CD44 and alpha4beta7." Cancer Res 57(5): 930-6.<br />

Matsui, H., H. Asou and T. Inaba (2007). "Cytokines direct the regulation of Bim mRNA<br />

stability by heat-shock cognate protein 70." Mol Cell 25(1): 99-112.<br />

McCleverty, C. J., D. C. Lin and R. C. Liddington (2007). "Structure of the PTB domain of<br />

tensin1 and a model for its recruitment to fibrillar adhesions." Protein Sci 16(6):<br />

1223-9.<br />

Miles, A., E. Liaskou, B. Eksteen, P. F. Lalor and D. H. Adams (2008). "CCL25 and CCL28<br />

promote alpha4 beta7-integrin-dependent adhesion of lymphocytes to MAdCAM-1<br />

under shear flow." Am J Physiol Gastrointest Liver Physiol 294(5): G1257-67.<br />

Mitra, S. K., D. A. Hanson and D. D. Schlaepfer (2005). "Focal adhesion kinase: in command<br />

and control of cell motility." Nat Rev Mol Cell Biol 6(1): 56-68.<br />

Mitra, S. K. and D. D. Schlaepfer (2006). "Integrin-regulated FAK-Src signaling in normal<br />

and cancer cells." Curr Opin Cell Biol 18(5): 516-23.<br />

Monks, C. R., B. A. Freiberg, H. Kupfer, N. Sciaky and A. Kupfer (1998). "Threedimensional<br />

segregation of supramolecular activation clusters in T cells." Nature<br />

395(6697): 82-6.<br />

Montanez, E., S. Ussar, M. Schifferer, M. Bosl, R. Zent, M. Moser and R. Fassler (2008).<br />

"Kindlin-2 controls bidirectional signaling of integrins." Genes Dev 22(10): 1325-30.<br />

Monteiro, P., D. Gilot, E. Le Ferrec, V. Lecureur, M. N'Diaye, M. Le Vee, N. Podechard, C.<br />

Pouponnot and O. Fardel (2007). "AhR- and c-maf-dependent induction of beta7integrin<br />

expression in human macrophages in response to environmental polycyclic<br />

aromatic hydrocarbons." Biochem Biophys Res Commun 358(2): 442-8.<br />

Morito, N., K. Yoh, Y. Fujioka, T. Nakano, H. Shimohata, Y. Hashimoto, A. Yamada, A.<br />

Maeda, F. Matsuno, H. Hata, A. Suzuki, S. Imagawa, H. Mitsuya, H. Esumi, A.<br />

Koyama, M. Yamamoto, N. Mori and S. Takahashi (2006). "Overexpression of c-Maf<br />

contributes to T-cell lymphoma in both mice and human." Cancer Res 66(2): 812-9.<br />

Morrison, D. K. (2009). "<strong>The</strong> 14-3-3 proteins: integrators of diverse signaling cues that<br />

impact cell fate and cancer development." Trends Cell Biol 19(1): 16-23.<br />

211


Mosser, D. D. and R. I. Morimoto (2004). "Molecular chaperones and the stress of<br />

oncogenesis." Oncogene 23(16): 2907-18.<br />

Mould, A. P., J. A. Askari, S. Barton, A. D. Kline, P. A. McEwan, S. E. Craig and M. J.<br />

Humphries (2002). "Integrin activation involves a conformational change in the alpha<br />

1 helix of the beta subunit A-domain." J Biol Chem 277(22): 19800-5.<br />

Mould, A. P., S. J. Barton, J. A. Askari, S. E. Craig and M. J. Humphries (2003a). "Role of<br />

ADMIDAS cation-binding site in ligand recognition by integrin alpha 5 beta 1." J<br />

Biol Chem 278(51): 51622-9.<br />

Mould, A. P., E. J. Symonds, P. A. Buckley, J. G. Grossmann, P. A. McEwan, S. J. Barton, J.<br />

A. Askari, S. E. Craig, J. Bella and M. J. Humphries (2003b). "Structure of an<br />

integrin-ligand complex deduced from solution x-ray scattering and site-directed<br />

mutagenesis." J Biol Chem 278(41): 39993-9.<br />

Nasertorabi, F., K. Tars, K. Becherer, R. Kodandapani, L. Liljas, K. Vuori and K. R. Ely<br />

(2006). "Molecular basis for regulation of Src by the docking protein p130Cas." J Mol<br />

Recognit 19(1): 30-8.<br />

Ni, J., A. G. Porter and D. Hollander (1995). "Beta 7 integrins and other cell adhesion<br />

molecules are differentially expressed and modulated by TNF beta in different<br />

lymphocyte populations." Cell Immunol 161(2): 166-72.<br />

Nimmanapalli, R., E. Gerbino, W. S. Dalton, V. Gandhi and M. Alsina (2008). "HSP70<br />

inhibition reverses cell adhesion mediated and acquired drug resistance in multiple<br />

myeloma." Br J Haematol 142(4): 551-61.<br />

Nishida, N., C. Xie, M. Shimaoka, Y. Cheng, T. Walz and T. A. Springer (2006). "Activation<br />

of leukocyte beta2 integrins by conversion from bent to extended conformations."<br />

Immunity 25(4): 583-94.<br />

Nishiya, N., W. B. Kiosses, J. Han and M. H. Ginsberg (2005). "An alpha4 integrin-paxillin-<br />

Arf-GAP complex restricts Rac activation to the leading edge of migrating cells." Nat<br />

Cell Biol 7(4): 343-52.<br />

Nobes, C. D. and A. Hall (1995). "Rho, rac, and cdc42 GTPases regulate the assembly of<br />

multimolecular focal complexes associated with actin stress fibers, lamellipodia, and<br />

filopodia." Cell 81(1): 53-62.<br />

O'Toole, T. E., D. Mandelman, J. Forsyth, S. J. Shattil, E. F. Plow and M. H. Ginsberg<br />

(1991). "Modulation of the affinity of integrin alpha IIb beta 3 (GPIIb-IIIa) by the<br />

cytoplasmic domain of alpha IIb." Science 254(5033): 845-7.<br />

Oda, K., Y. Matsuoka, A. Funahashi and H. Kitano (2005). "A comprehensive pathway map<br />

of epidermal growth factor receptor signaling." Mol Syst Biol 1: 2005 0010.<br />

Oh, E. S., H. Gu, T. M. Saxton, J. F. Timms, S. Hausdorff, E. U. Frevert, B. B. Kahn, T.<br />

Pawson, B. G. Neel and S. M. Thomas (1999). "Regulation of early events in integrin<br />

signaling by protein tyrosine phosphatase SHP-2." Mol Cell Biol 19(4): 3205-15.<br />

Ohnishi, K., A. Takahashi, S. Yokota and T. Ohnishi (2004). "Effects of a heat shock protein<br />

inhibitor KNK437 on heat sensitivity and heat tolerance in human squamous cell<br />

carcinoma cell lines differing in p53 status." Int J Radiat Biol 80(8): 607-14.<br />

Ojaniemi, M., S. S. Martin, F. Dolfi, J. M. Olefsky and K. Vuori (1997). "<strong>The</strong> protooncogene<br />

product p120(cbl) links c-Src and phosphatidylinositol 3'-kinase to the<br />

integrin signaling pathway." J Biol Chem 272(6): 3780-7.<br />

Okada, M. and H. Nakagawa (1989). "A protein tyrosine kinase involved in regulation of<br />

pp60c-src function." J Biol Chem 264(35): 20886-93.<br />

Oktay, M., K. K. Wary, M. Dans, R. B. Birge and F. G. Giancotti (1999). "Integrin-mediated<br />

activation of focal adhesion kinase is required for signaling to Jun NH2-terminal<br />

kinase and progression through the G1 phase of the cell cycle." J Cell Biol 145(7):<br />

1461-9.<br />

212


Orr, A. W., M. H. Ginsberg, S. J. Shattil, H. Deckmyn and M. A. Schwartz (2006). "Matrixspecific<br />

suppression of integrin activation in shear stress signaling." Mol Biol Cell<br />

17(11): 4686-97.<br />

Otey, C. A., F. M. Pavalko and K. Burridge (1990). "An interaction between alpha-actinin<br />

and the beta 1 integrin subunit in vitro." J Cell Biol 111(2): 721-9.<br />

Oxley, C. L., N. J. Anthis, E. D. Lowe, I. Vakonakis, I. D. Campbell and K. L. Wegener<br />

(2008). "An integrin phosphorylation switch: the effect of beta3 integrin tail<br />

phosphorylation on Dok1 and talin binding." J Biol Chem 283(9): 5420-6.<br />

Pai, R., M. Ohta, R. M. Itani, I. J. Sarfeh and A. S. Tarnawski (1998). "Induction of mitogenactivated<br />

protein kinase signal transduction pathway during gastric ulcer healing in<br />

rats." Gastroenterology 114(4): 706-13.<br />

Pal Sharma, C. and W. H. Goldmann (2004). "Phosphorylation of actin-binding protein<br />

(ABP-280; filamin) by tyrosine kinase p56lck modulates actin filament crosslinking."<br />

Cell Biol Int 28(12): 935-41.<br />

Pallen, C. J. (2003). "Protein tyrosine phosphatase alpha (PTPalpha): a Src family kinase<br />

activator and mediator of multiple biological effects." Curr Top Med Chem 3(7): 821-<br />

35.<br />

Pankov, R., E. Cukierman, K. Clark, K. Matsumoto, C. Hahn, B. Poulin and K. M. Yamada<br />

(2003). "Specific beta1 integrin site selectively regulates Akt/protein kinase B<br />

signaling via local activation of protein phosphatase 2A." J Biol Chem 278(20):<br />

18671-81.<br />

Parker, C. M., K. L. Cepek, G. J. Russell, S. K. Shaw, D. N. Posnett, R. Schwarting and M.<br />

B. Brenner (1992). "A family of beta 7 integrins on human mucosal lymphocytes."<br />

Proc Natl Acad Sci U S A 89(5): 1924-8.<br />

Partridge, A. W., S. Liu, S. Kim, J. U. Bowie and M. H. Ginsberg (2005). "Transmembrane<br />

domain helix packing stabilizes integrin alphaIIbbeta3 in the low affinity state." J Biol<br />

Chem 280(8): 7294-300.<br />

Pavalko, F. M. and S. M. LaRoche (1993). "Activation of human neutrophils induces an<br />

interaction between the integrin beta 2-subunit (CD18) and the actin binding protein<br />

alpha-actinin." J Immunol 151(7): 3795-807.<br />

Peer, D., E. J. Park, Y. Morishita, C. V. Carman and M. Shimaoka (2008). "Systemic<br />

leukocyte-directed siRNA delivery revealing cyclin D1 as an anti-inflammatory<br />

target." Science 319(5863): 627-30.<br />

Pfaff, M., S. Liu, D. J. Erle and M. H. Ginsberg (1998). "Integrin beta cytoplasmic domains<br />

differentially bind to cytoskeletal proteins." J Biol Chem 273(11): 6104-9.<br />

Phillips, D. R., I. F. Charo and R. M. Scarborough (1991). "GPIIb-IIIa: the responsive<br />

integrin." Cell 65(3): 359-62.<br />

Phillips, D. R., K. S. Prasad, J. Manganello, M. Bao and L. Nannizzi-Alaimo (2001).<br />

"Integrin tyrosine phosphorylation in platelet signaling." Curr Opin Cell Biol 13(5):<br />

546-54.<br />

Playford, M. P. and M. D. Schaller (2004). "<strong>The</strong> interplay between Src and integrins in<br />

normal and tumor biology." Oncogene 23(48): 7928-46.<br />

Ponniah, S., D. Z. Wang, K. L. Lim and C. J. Pallen (1999). "Targeted disruption of the<br />

tyrosine phosphatase PTPalpha leads to constitutive downregulation of the kinases Src<br />

and Fyn." Curr Biol 9(10): 535-8.<br />

Porter, J. C. and N. Hogg (1997). "Integrin cross talk: activation of lymphocyte functionassociated<br />

antigen-1 on human T cells alters alpha4beta1- and alpha5beta1-mediated<br />

function." J Cell Biol 138(6): 1437-47.<br />

213


Pratt, S. J., H. Epple, M. Ward, Y. Feng, V. M. Braga and G. D. Longmore (2005). "<strong>The</strong> LIM<br />

protein Ajuba influences p130Cas localization and Rac1 activity during cell<br />

migration." J Cell Biol 168(5): 813-24.<br />

Pulkkinen, L., V. E. Kimonis, Y. Xu, E. N. Spanou, W. H. McLean and J. Uitto (1997).<br />

"Homozygous alpha6 integrin mutation in junctional epidermolysis bullosa with<br />

congenital duodenal atresia." Hum Mol Genet 6(5): 669-74.<br />

Ramos, J. W. (2008). "<strong>The</strong> regulation of extracellular signal-regulated kinase (ERK) in<br />

mammalian cells." Int J Biochem Cell Biol 40(12): 2707-19.<br />

Reddy, K. B., D. M. Smith and E. F. Plow (2008). "Analysis of Fyn function in hemostasis<br />

and alphaIIbbeta3-integrin signaling." J Cell Sci 121(Pt 10): 1641-8.<br />

Reiske, H. R., S. C. Kao, L. A. Cary, J. L. Guan, J. F. Lai and H. C. Chen (1999).<br />

"Requirement of phosphatidylinositol 3-kinase in focal adhesion kinase-promoted cell<br />

migration." J Biol Chem 274(18): 12361-6.<br />

Rietzler, M., M. Bittner, W. Kolanus, A. Schuster and B. Holzmann (1998). "<strong>The</strong> human WD<br />

repeat protein WAIT-1 specifically interacts with the cytoplasmic tails of beta7integrins."<br />

J Biol Chem 273(42): 27459-66.<br />

Roberts, K. and P. J. Kilshaw (1993). "<strong>The</strong> mucosal T cell integrin alpha M290 beta 7<br />

recognizes a ligand on mucosal epithelial cell lines." Eur J Immunol 23(7): 1630-5.<br />

Robinson, P. W., S. J. Green, C. Carter, J. Coadwell and P. J. Kilshaw (2001). "Studies on<br />

transcriptional regulation of the mucosal T-cell integrin alphaEbeta7 (CD103)."<br />

Immunology 103(2): 146-54.<br />

Rodius, S., O. Chaloin, M. Moes, E. Schaffner-Reckinger, I. Landrieu, G. Lippens, M. Lin, J.<br />

Zhang and N. Kieffer (2008). "<strong>The</strong> talin rod IBS2 alpha-helix interacts with the beta3<br />

integrin cytoplasmic tail membrane-proximal helix by establishing charge<br />

complementary salt bridges." J Biol Chem 283(35): 24212-23.<br />

Rohde, M., M. Daugaard, M. H. Jensen, K. Helin, J. Nylandsted and M. Jaattela (2005).<br />

"Members of the heat-shock protein 70 family promote cancer cell growth by distinct<br />

mechanisms." Genes Dev 19(5): 570-82.<br />

Rothlein, R. and T. A. Springer (1986). "<strong>The</strong> requirement for lymphocyte function-associated<br />

antigen 1 in homotypic leukocyte adhesion stimulated by phorbol ester." J Exp Med<br />

163(5): 1132-49.<br />

Roussel, R. R., S. R. Brodeur, D. Shalloway and A. P. Laudano (1991). "Selective binding of<br />

activated pp60c-src by an immobilized synthetic phosphopeptide modeled on the<br />

carboxyl terminus of pp60c-src." Proc Natl Acad Sci U S A 88(23): 10696-700.<br />

Roy, P. K., F. Rashid, J. Bragg and J. A. Ibdah (2008). "Role of the JNK signal transduction<br />

pathway in inflammatory bowel disease." World J Gastroenterol 14(2): 200-2.<br />

Ruegg, C., A. A. Postigo, E. E. Sikorski, E. C. Butcher, R. Pytela and D. J. Erle (1992). "Role<br />

of integrin alpha 4 beta 7/alpha 4 beta P in lymphocyte adherence to fibronectin and<br />

VCAM-1 and in homotypic cell clustering." J Cell Biol 117(1): 179-89.<br />

Sadhu, C., B. Masinovsky and D. E. Staunton (1998). "Differential regulation of<br />

chemoattractant-stimulated beta 2, beta 3, and beta 7 integrin activity." J Immunol<br />

160(11): 5622-8.<br />

Saitoh, M., T. Ishikawa, S. Matsushima, M. Naka and H. Hidaka (1987). "Selective inhibition<br />

of catalytic activity of smooth muscle myosin light chain kinase." J Biol Chem<br />

262(16): 7796-801.<br />

Sampath, R., P. J. Gallagher and F. M. Pavalko (1998). "Cytoskeletal interactions with the<br />

leukocyte integrin beta2 cytoplasmic tail. Activation-dependent regulation of<br />

associations with talin and alpha-actinin." J Biol Chem 273(50): 33588-94.<br />

Sandilands, E. and M. C. Frame (2008). "Endosomal trafficking of Src tyrosine kinase."<br />

Trends Cell Biol 18(7): 322-9.<br />

214


Sawada, Y., M. Tamada, B. J. Dubin-Thaler, O. Cherniavskaya, R. Sakai, S. Tanaka and M.<br />

P. Sheetz (2006). "Force sensing by mechanical extension of the Src family kinase<br />

substrate p130Cas." Cell 127(5): 1015-26.<br />

Schaller, M. D., C. A. Borgman, B. S. Cobb, R. R. Vines, A. B. Reynolds and J. T. Parsons<br />

(1992). "pp125FAK a structurally distinctive protein-tyrosine kinase associated with<br />

focal adhesions." Proc Natl Acad Sci U S A 89(11): 5192-6.<br />

Schaller, M. D., C. A. Borgman and J. T. Parsons (1993). "Autonomous expression of a<br />

noncatalytic domain of the focal adhesion-associated protein tyrosine kinase<br />

pp125FAK." Mol Cell Biol 13(2): 785-91.<br />

Schaller, M. D., C. A. Otey, J. D. Hildebrand and J. T. Parsons (1995a). "Focal adhesion<br />

kinase and paxillin bind to peptides mimicking beta integrin cytoplasmic domains." J<br />

Cell Biol 130(5): 1181-7.<br />

Schaller, M. D. and J. T. Parsons (1995b). "pp125FAK-dependent tyrosine phosphorylation<br />

of paxillin creates a high-affinity binding site for Crk." Mol Cell Biol 15(5): 2635-45.<br />

Schlaepfer, D. D., S. K. Hanks, T. Hunter and P. van der Geer (1994). "Integrin-mediated<br />

signal transduction linked to Ras pathway by GRB2 binding to focal adhesion<br />

kinase." Nature 372(6508): 786-91.<br />

Schlaepfer, D. D. and T. Hunter (1998). "Integrin signalling and tyrosine phosphorylation:<br />

just the FAKs?" Trends Cell Biol 8(4): 151-7.<br />

Schon, M. P. (2008). "Efalizumab in the treatment of psoriasis: mode of action, clinical<br />

indications, efficacy, and safety." Clin Dermatol 26(5): 509-14.<br />

Schon, M. P., A. Arya, E. A. Murphy, C. M. Adams, U. G. Strauch, W. W. Agace, J. Marsal,<br />

J. P. Donohue, H. Her, D. R. Beier, S. Olson, L. Lefrancois, M. B. Brenner, M. J.<br />

Grusby and C. M. Parker (1999). "Mucosal T lymphocyte numbers are selectively<br />

reduced in integrin alpha E (CD103)-deficient mice." J Immunol 162(11): 6641-9.<br />

Schwartz, M. A. (2001). "Integrin signaling revisited." Trends Cell Biol 11(12): 466-70.<br />

Shafiei, F. (2004). <strong>The</strong> role of MAP kinase signaling pathways in integrin beta7 gene<br />

regulation. Molecular Medicine and Pathology. Auckland, <strong>University</strong> of Auckland.<br />

Doctor of Philosophy.<br />

Shah, A., E. Unger, M. D. Bain, R. Bruce, J. Bodkin, J. Ginnetti, W. C. Wang, B. Seon, C. C.<br />

Stewart and S. S. Evans (2002). "Cytokine and adhesion molecule expression in<br />

primary human endothelial cells stimulated with fever-range hyperthermia." Int J<br />

Hyperthermia 18(6): 534-51.<br />

Sharma, C. P., R. M. Ezzell and M. A. Arnaout (1995). "Direct interaction of filamin (ABP-<br />

280) with the beta 2-integrin subunit CD18." J Immunol 154(7): 3461-70.<br />

Shattil, S. J. (2005). "Integrins and Src: dynamic duo of adhesion signaling." Trends Cell Biol<br />

15(8): 399-403.<br />

Shaw, S. K., K. L. Cepek, E. A. Murphy, G. J. Russell, M. B. Brenner and C. M. Parker<br />

(1994). "Molecular cloning of the human mucosal lymphocyte integrin alpha E<br />

subunit. Unusual structure and restricted RNA distribution." J Biol Chem 269(8):<br />

6016-25.<br />

Sheremata, W. A., T. L. Vollmer, L. A. Stone, A. J. Willmer-Hulme and M. Koller (1999).<br />

"A safety and pharmacokinetic study of intravenous natalizumab in patients with<br />

MS." Neurology 52(5): 1072-4.<br />

Shi, X., Y. Q. Ma, Y. Tu, K. Chen, S. Wu, K. Fukuda, J. Qin, E. F. Plow and C. Wu (2007).<br />

"<strong>The</strong> MIG-2/integrin interaction strengthens cell-matrix adhesion and modulates cell<br />

motility." J Biol Chem 282(28): 20455-66.<br />

Shimaoka, M., C. Lu, R. T. Palframan, U. H. von Andrian, A. McCormack, J. Takagi and T.<br />

A. Springer (2001). "Reversibly locking a protein fold in an active conformation with<br />

215


a disulfide bond: integrin alphaL I domains with high affinity and antagonist activity<br />

in vivo." Proc Natl Acad Sci U S A 98(11): 6009-14.<br />

Shimaoka, M., J. Takagi and T. A. Springer (2002). "Conformational regulation of integrin<br />

structure and function." Annu Rev Biophys Biomol Struct 31: 485-516.<br />

Sims, T. N. and M. L. Dustin (2002). "<strong>The</strong> immunological synapse: integrins take the stage."<br />

Immunol Rev 186: 100-17.<br />

Sliutz, G., J. Karlseder, C. Tempfer, L. Orel, G. Holzer and M. M. Simon (1996). "Drug<br />

resistance against gemcitabine and topotecan mediated by constitutive hsp70<br />

overexpression in vitro: implication of quercetin as sensitiser in chemotherapy." Br J<br />

Cancer 74(2): 172-7.<br />

Smith, A., M. Bracke, B. Leitinger, J. C. Porter and N. Hogg (2003). "LFA-1-induced T cell<br />

migration on ICAM-1 involves regulation of MLCK-mediated attachment and<br />

ROCK-dependent detachment." J Cell Sci 116(Pt 15): 3123-33.<br />

Smith, A., Y. R. Carrasco, P. Stanley, N. Kieffer, F. D. Batista and N. Hogg (2005). "A talindependent<br />

LFA-1 focal zone is formed by rapidly migrating T lymphocytes." J Cell<br />

Biol 170(1): 141-51.<br />

Soga, S., T. Kozawa, H. Narumi, S. Akinaga, K. Irie, K. Matsumoto, S. V. Sharma, H.<br />

Nakano, T. Mizukami and M. Hara (1998). "Radicicol leads to selective depletion of<br />

Raf kinase and disrupts K-Ras-activated aberrant signaling pathway." J Biol Chem<br />

273(2): 822-8.<br />

Soler, D., T. Chapman, L. L. Yang, T. Wyant, R. Egan and E. R. Fedyk (2009). "<strong>The</strong> Binding<br />

Specificity and Selective Antagonism of Vedolizumab, an Anti-{alpha}4{beta}7<br />

Integrin <strong>The</strong>rapeutic Antibody in Development for Inflammatory Bowel Diseases." J<br />

Pharmacol Exp <strong>The</strong>r.<br />

Songyang, Z., S. E. Shoelson, M. Chaudhuri, G. Gish, T. Pawson, W. G. Haser, F. King, T.<br />

Roberts, S. Ratnofsky, R. J. Lechleider and et al. (1993). "SH2 domains recognize<br />

specific phosphopeptide sequences." Cell 72(5): 767-78.<br />

Sonoda, Y., Y. Matsumoto, M. Funakoshi, D. Yamamoto, S. K. Hanks and T. Kasahara<br />

(2000). "Anti-apoptotic role of focal adhesion kinase (FAK). Induction of inhibitorof-apoptosis<br />

proteins and apoptosis suppression by the overexpression of FAK in a<br />

human leukemic cell line, HL-60." J Biol Chem 275(21): 16309-15.<br />

Sonoda, Y., S. Watanabe, Y. Matsumoto, E. Aizu-Yokota and T. Kasahara (1999). "FAK is<br />

the upstream signal protein of the phosphatidylinositol 3-kinase-Akt survival pathway<br />

in hydrogen peroxide-induced apoptosis of a human glioblastoma cell line." J Biol<br />

Chem 274(15): 10566-70.<br />

Springer, T. A. (1997). "Folding of the N-terminal, ligand-binding region of integrin alphasubunits<br />

into a beta-propeller domain." Proc Natl Acad Sci U S A 94(1): 65-72.<br />

Stein, P. L., H. Vogel and P. Soriano (1994). "Combined deficiencies of Src, Fyn, and Yes<br />

tyrosine kinases in mutant mice." Genes Dev 8(17): 1999-2007.<br />

Stossel, T. P., J. Condeelis, L. Cooley, J. H. Hartwig, A. Noegel, M. Schleicher and S. S.<br />

Shapiro (2001). "Filamins as integrators of cell mechanics and signalling." Nat Rev<br />

Mol Cell Biol 2(2): 138-45.<br />

Strauch, U. G., R. C. Mueller, X. Y. Li, M. Cernadas, J. M. Higgins, D. G. Binion and C. M.<br />

Parker (2001). "Integrin alpha E(CD103)beta 7 mediates adhesion to intestinal<br />

microvascular endothelial cell lines via an E-cadherin-independent interaction." J<br />

Immunol 166(5): 3506-14.<br />

Su, A. I., T. Wiltshire, S. Batalov, H. Lapp, K. A. Ching, D. Block, J. Zhang, R. Soden, M.<br />

Hayakawa, G. Kreiman, M. P. Cooke, J. R. Walker and J. B. Hogenesch (2004). "A<br />

gene atlas of the mouse and human protein-encoding transcriptomes." Proc Natl Acad<br />

Sci U S A 101(16): 6062-7.<br />

216


Tachibana, K., T. Sato, N. D'Avirro and C. Morimoto (1995). "Direct association of<br />

pp125FAK with paxillin, the focal adhesion-targeting mechanism of pp125FAK." J<br />

Exp Med 182(4): 1089-99.<br />

Tadokoro, S., S. J. Shattil, K. Eto, V. Tai, R. C. Liddington, J. M. de Pereda, M. H. Ginsberg<br />

and D. A. Calderwood (2003). "Talin binding to integrin beta tails: a final common<br />

step in integrin activation." Science 302(5642): 103-6.<br />

Takada, Y., X. Ye and S. Simon (2007). "<strong>The</strong> integrins." Genome Biol 8(5): 215.<br />

Takagi, J., B. M. Petre, T. Walz and T. A. Springer (2002). "Global conformational<br />

rearrangements in integrin extracellular domains in outside-in and inside-out<br />

signaling." Cell 110(5): 599-11.<br />

Takagi, J., K. Strokovich, T. A. Springer and T. Walz (2003). "Structure of integrin<br />

alpha5beta1 in complex with fibronectin." Embo J 22(18): 4607-15.<br />

Takala, H., E. Nurminen, S. M. Nurmi, M. Aatonen, T. Strandin, M. Takatalo, T. Kiema, C.<br />

G. Gahmberg, J. Ylanne and S. C. Fagerholm (2008). "Beta2 integrin phosphorylation<br />

on Thr758 acts as a molecular switch to regulate 14-3-3 and filamin binding." Blood<br />

112(5): 1853-62.<br />

Tamada, M., M. P. Sheetz and Y. Sawada (2004). "Activation of a signaling cascade by<br />

cytoskeleton stretch." Dev Cell 7(5): 709-18.<br />

Tanentzapf, G. and N. H. Brown (2006). "An interaction between integrin and the talin<br />

FERM domain mediates integrin activation but not linkage to the cytoskeleton." Nat<br />

Cell Biol 8(6): 601-6.<br />

Tang, P., C. Cao, M. Xu and L. Zhang (2007). "Cytoskeletal protein radixin activates integrin<br />

alpha(M)beta(2) by binding to its cytoplasmic tail." FEBS Lett 581(6): 1103-8.<br />

Tapley, P., A. Horwitz, C. Buck, K. Duggan and L. Rohrschneider (1989). "Integrins isolated<br />

from Rous sarcoma virus-transformed chicken embryo fibroblasts." Oncogene 4(3):<br />

325-33.<br />

Taylor, B. M., K. P. Kolbasa, J. E. Chin, I. M. Richards, W. E. Fleming, R. L. Griffin, S. F.<br />

Fidler and F. F. Sun (1997). "Roles of adhesion molecules ICAM-1 and alpha4<br />

integrin in antigen-induced changes in microvascular permeability associated with<br />

lung inflammation in sensitized brown Norway rats." Am J Respir Cell Mol Biol<br />

17(6): 757-66.<br />

Thomas, J. W., M. A. Cooley, J. M. Broome, R. Salgia, J. D. Griffin, C. R. Lombardo and M.<br />

D. Schaller (1999). "<strong>The</strong> role of focal adhesion kinase binding in the regulation of<br />

tyrosine phosphorylation of paxillin." J Biol Chem 274(51): 36684-92.<br />

Tomar, A. and D. D. Schlaepfer (2009). "Focal adhesion kinase: switching between GAPs<br />

and GEFs in the regulation of cell motility." Curr Opin Cell Biol.<br />

Travis, M. A., A. van der Flier, R. A. Kammerer, A. P. Mould, A. Sonnenberg and M. J.<br />

Humphries (2004). "Interaction of filamin A with the integrin beta 7 cytoplasmic<br />

domain: role of alternative splicing and phosphorylation." FEBS Lett 569(1-3): 185-<br />

90.<br />

Tu, Y., F. Li, S. Goicoechea and C. Wu (1999). "<strong>The</strong> LIM-only protein PINCH directly<br />

interacts with integrin-linked kinase and is recruited to integrin-rich sites in spreading<br />

cells." Mol Cell Biol 19(3): 2425-34.<br />

Tu, Y., S. Wu, X. Shi, K. Chen and C. Wu (2003). "Migfilin and Mig-2 link focal adhesions<br />

to filamin and the actin cytoskeleton and function in cell shape modulation." Cell<br />

113(1): 37-47.<br />

Turner, C. E. (1998). "Paxillin." Int J Biochem Cell Biol 30(9): 955-9.<br />

Uhlemann, A. C., B. Brenner, E. Gulbins, O. Linderkamp and F. Lang (1997). "Stimulation<br />

of TK1 lymphoma cells via alpha 4 beta 7 integrin results in activation of srctyrosine-<br />

and MAP-kinases." Biochem Biophys Res Commun 239(1): 68-73.<br />

217


Vachon, P. H., H. Xu, L. Liu, F. Loechel, Y. Hayashi, K. Arahata, J. C. Reed, U. M. Wewer<br />

and E. Engvall (1997). "Integrins (alpha7beta1) in muscle function and survival.<br />

Disrupted expression in merosin-deficient congenital muscular dystrophy." J Clin<br />

Invest 100(7): 1870-81.<br />

van Nimwegen, M. J. and B. van de Water (2007). "Focal adhesion kinase: a potential target<br />

in cancer therapy." Biochem Pharmacol 73(5): 597-609.<br />

Vargas-Roig, L. M., F. E. Gago, O. Tello, J. C. Aznar and D. R. Ciocca (1998). "Heat shock<br />

protein expression and drug resistance in breast cancer patients treated with induction<br />

chemotherapy." Int J Cancer 79(5): 468-75.<br />

Varma, R., G. Campi, T. Yokosuka, T. Saito and M. L. Dustin (2006). "T cell receptorproximal<br />

signals are sustained in peripheral microclusters and terminated in the<br />

central supramolecular activation cluster." Immunity 25(1): 117-27.<br />

Velling, T., S. Nilsson, A. Stefansson and S. Johansson (2004). "beta1-Integrins induce<br />

phosphorylation of Akt on serine 473 independently of focal adhesion kinase and Src<br />

family kinases." EMBO Rep 5(9): 901-5.<br />

Vidal, F., D. Aberdam, C. Miquel, A. M. Christiano, L. Pulkkinen, J. Uitto, J. P. Ortonne and<br />

G. Meneguzzi (1995). "Integrin beta 4 mutations associated with junctional<br />

epidermolysis bullosa with pyloric atresia." Nat Genet 10(2): 229-34.<br />

Vinogradova, O., A. Velyvis, A. Velyviene, B. Hu, T. Haas, E. Plow and J. Qin (2002). "A<br />

structural mechanism of integrin alpha(IIb)beta(3) "inside-out" activation as regulated<br />

by its cytoplasmic face." Cell 110(5): 587-97.<br />

von Wichert, G., B. Haimovich, G. S. Feng and M. P. Sheetz (2003a). "Force-dependent<br />

integrin-cytoskeleton linkage formation requires downregulation of focal complex<br />

dynamics by Shp2." Embo J 22(19): 5023-35.<br />

von Wichert, G., G. Jiang, A. Kostic, K. De Vos, J. Sap and M. P. Sheetz (2003b). "RPTPalpha<br />

acts as a transducer of mechanical force on alphav/beta3-integrin-cytoskeleton<br />

linkages." J Cell Biol 161(1): 143-53.<br />

Wade, R. and S. Vande Pol (2006). "Minimal features of paxillin that are required for the<br />

tyrosine phosphorylation of focal adhesion kinase." Biochem J 393(Pt 2): 565-73.<br />

Wagner, N., J. Lohler, E. J. Kunkel, K. Ley, E. Leung, G. Krissansen, K. Rajewsky and W.<br />

Muller (1996). "Critical role for beta7 integrins in formation of the gut-associated<br />

lymphoid tissue." Nature 382(6589): 366-70.<br />

Waldman, E., S. X. Lu, V. M. Hubbard, A. A. Kochman, J. M. Eng, T. H. Terwey, S. J.<br />

Muriglan, T. D. Kim, G. Heller, G. F. Murphy, C. Liu, O. Alpdogan and M. R. van<br />

den Brink (2006). "Absence of beta7 integrin results in less graft-versus-host disease<br />

because of decreased homing of alloreactive T cells to intestine." Blood 107(4): 1703-<br />

11.<br />

Walsh, N. R. (1996). Regulation of the function of the integrins LPAM-1 and VLA-4; and<br />

soluble forms of their ligands MAdCAM-1 and VCAM-1. Molecular Medicine and<br />

Pathology. Auckland, <strong>University</strong> of Auckland. Masters of Science.<br />

Wary, K. K., F. Mainiero, S. J. Isakoff, E. E. Marcantonio and F. G. Giancotti (1996). "<strong>The</strong><br />

adaptor protein Shc couples a class of integrins to the control of cell cycle<br />

progression." Cell 87(4): 733-43.<br />

Wegener, K. L., A. W. Partridge, J. Han, A. R. Pickford, R. C. Liddington, M. H. Ginsberg<br />

and I. D. Campbell (2007). "Structural basis of integrin activation by talin." Cell<br />

128(1): 171-82.<br />

Wei, H., T. L'Ecuyer and R. S. Vander Heide (2003). "Effect of increased expression of<br />

cytoskeletal protein vinculin on ischemia-reperfusion injury in ventricular myocytes."<br />

Am J Physiol Heart Circ Physiol 284(3): H911-8.<br />

218


Wei, H. and R. S. Vander Heide (2008). "Heat stress activates AKT via focal adhesion<br />

kinase-mediated pathway in neonatal rat ventricular myocytes." Am J Physiol Heart<br />

Circ Physiol 295(2): H561-8.<br />

Wen, L. P., J. A. Fahrni, S. Troie, J. L. Guan, K. Orth and G. D. Rosen (1997). "Cleavage of<br />

focal adhesion kinase by caspases during apoptosis." J Biol Chem 272(41): 26056-61.<br />

Whittaker, C. A. and R. O. Hynes (2002). "Distribution and evolution of von<br />

Willebrand/integrin A domains: widely dispersed domains with roles in cell adhesion<br />

and elsewhere." Mol Biol Cell 13(10): 3369-87.<br />

Winkler, J., H. Lunsdorf and B. M. Jockusch (1996). "<strong>The</strong> ultrastructure of chicken gizzard<br />

vinculin as visualized by high-resolution electron microscopy." J Struct Biol 116(2):<br />

270-7.<br />

Witte, V., B. Laffert, O. Rosorius, P. Lischka, K. Blume, G. Galler, A. Stilper, D. Willbold,<br />

P. D'Aloja, M. Sixt, J. Kolanus, M. Ott, W. Kolanus, G. Schuler and A. S. Baur<br />

(2004). "HIV-1 Nef mimics an integrin receptor signal that recruits the polycomb<br />

group protein Eed to the plasma membrane." Mol Cell 13(2): 179-90.<br />

Woods Ignatoski, K. M., N. K. Grewal, S. Markwart, D. L. Livant and S. P. Ethier (2003).<br />

"p38MAPK induces cell surface alpha4 integrin downregulation to facilitate erbB-2mediated<br />

invasion." Neoplasia 5(2): 128-34.<br />

Woodside, D. G., A. Obergfell, L. Leng, J. L. Wilsbacher, C. K. Miranti, J. S. Brugge, S. J.<br />

Shattil and M. H. Ginsberg (2001). "Activation of Syk protein tyrosine kinase through<br />

interaction with integrin beta cytoplasmic domains." Curr Biol 11(22): 1799-804.<br />

Xiao, T., J. Takagi, B. S. Coller, J. H. Wang and T. A. Springer (2004). "Structural basis for<br />

allostery in integrins and binding to fibrinogen-mimetic therapeutics." Nature<br />

432(7013): 59-67.<br />

Xiong, J. P., T. Stehle, B. Diefenbach, R. Zhang, R. Dunker, D. L. Scott, A. Joachimiak, S. L.<br />

Goodman and M. A. Arnaout (2001). "Crystal structure of the extracellular segment<br />

of integrin alpha Vbeta3." Science 294(5541): 339-45.<br />

Xiong, J. P., T. Stehle, S. L. Goodman and M. A. Arnaout (2003). "New insights into the<br />

structural basis of integrin activation." Blood 102(4): 1155-9.<br />

Xiong, J. P., T. Stehle, R. Zhang, A. Joachimiak, M. Frech, S. L. Goodman and M. A.<br />

Arnaout (2002). "Crystal structure of the extracellular segment of integrin alpha<br />

Vbeta3 in complex with an Arg-Gly-Asp ligand." Science 296(5565): 151-5.<br />

Xu, W., S. C. Harrison and M. J. Eck (1997). "Three-dimensional structure of the tyrosine<br />

kinase c-Src." Nature 385(6617): 595-602.<br />

Yalamanchili, P., C. Lu, C. Oxvig and T. A. Springer (2000). "Folding and function of I<br />

domain-deleted Mac-1 and lymphocyte function-associated antigen-1." J Biol Chem<br />

275(29): 21877-82.<br />

Yamamoto, D., Y. Sonoda, M. Hasegawa, M. Funakoshi-Tago, E. Aizu-Yokota and T.<br />

Kasahara (2003). "FAK overexpression upregulates cyclin D3 and enhances cell<br />

proliferation via the PKC and PI3-kinase-Akt pathways." Cell Signal 15(6): 575-83.<br />

Yamniuk, A. P., H. Ishida and H. J. Vogel (2006). "<strong>The</strong> interaction between calcium- and<br />

integrin-binding protein 1 and the alphaIIb integrin cytoplasmic domain involves a<br />

novel C-terminal displacement mechanism." J Biol Chem 281(36): 26455-64.<br />

Yan, B., D. A. Calderwood, B. Yaspan and M. H. Ginsberg (2001). "Calpain cleavage<br />

promotes talin binding to the beta 3 integrin cytoplasmic domain." J Biol Chem<br />

276(30): 28164-70.<br />

Yang, Y., M. Sammar, J. E. Harrison, K. Lehnert, C. G. Print, E. Leung, R. Prestidge and G.<br />

W. Krissansen (1995). "Construction and adhesive properties of a soluble MadCAM-<br />

1-Fc chimera expressed in a baculovirus system: phylogenetic conservation of<br />

receptor-ligand interaction." Scand J Immunol 42(2): 235-47.<br />

219


Ye, F., J. Liu, H. Winkler and K. A. Taylor (2008). "Integrin alpha IIb beta 3 in a membrane<br />

environment remains the same height after Mn2+ activation when observed by<br />

cryoelectron tomography." J Mol Biol 378(5): 976-86.<br />

Ylanne, J., K. Scheffzek, P. Young and M. Saraste (2001). "Crystal structure of the alphaactinin<br />

rod reveals an extensive torsional twist." Structure 9(7): 597-604.<br />

Yuan, Q., W. M. Jiang, E. Leung, D. Hollander, J. D. Watson and G. W. Krissansen (1992).<br />

"Molecular cloning of the mouse integrin beta 7 subunit." J Biol Chem 267(11): 7352-<br />

8.<br />

Yuan, Q. A., W. M. Jiang, G. W. Krissansen and J. D. Watson (1990). "Cloning and<br />

sequence analysis of a novel beta 2-related integrin transcript from T lymphocytes:<br />

homology of integrin cysteine-rich repeats to domain III of laminin B chains." Int<br />

Immunol 2(11): 1097-108.<br />

Yuan, R., R. El-Asady, K. Liu, D. Wang, C. B. Drachenberg and G. A. Hadley (2005).<br />

"Critical role for CD103+CD8+ effectors in promoting tubular injury following<br />

allogeneic renal transplantation." J Immunol 175(5): 2868-79.<br />

Zamir, E. and B. Geiger (2001). "Components of cell-matrix adhesions." J Cell Sci 114(Pt<br />

20): 3577-9.<br />

Zang, Q. and T. A. Springer (2001). "Amino acid residues in the PSI domain and cysteinerich<br />

repeats of the integrin beta2 subunit that restrain activation of the integrin<br />

alpha(X)beta(2)." J Biol Chem 276(10): 6922-9.<br />

Zell, T., C. S. Warden, A. S. Chan, M. E. Cook, C. L. Dell, S. W. Hunt, 3rd and Y. Shimizu<br />

(1998). "Regulation of beta 1-integrin-mediated cell adhesion by the Cbl adaptor<br />

protein." Curr Biol 8(14): 814-22.<br />

Zhang, W. V., Y. Yang, R. W. Berg, E. Leung and G. W. Krissansen (1999). "<strong>The</strong> small<br />

GTP-binding proteins Rho and Rac induce T cell adhesion to the mucosal addressin<br />

MAdCAM-1 in a hierarchical fashion." Eur J Immunol 29(9): 2875-85.<br />

Zhang, X., G. Jiang, Y. Cai, S. J. Monkley, D. R. Critchley and M. P. Sheetz (2008a). "Talin<br />

depletion reveals independence of initial cell spreading from integrin activation and<br />

traction." Nat Cell Biol.<br />

Zhang, Y., K. Chen, Y. Tu, A. Velyvis, Y. Yang, J. Qin and C. Wu (2002). "Assembly of the<br />

PINCH-ILK-CH-ILKBP complex precedes and is essential for localization of each<br />

component to cell-matrix adhesion sites." J Cell Sci 115(Pt 24): 4777-86.<br />

Zhang, Y., H. N. Hayenga, M. R. Sarantos, S. I. Simon and S. Neelamegham (2008b).<br />

"Differential regulation of neutrophil CD18 integrin function by di- and tri-valent<br />

cations: manganese vs. gadolinium." Ann Biomed Eng 36(4): 647-60.<br />

Zheng, X. M., R. J. Resnick and D. Shalloway (2000). "A phosphotyrosine displacement<br />

mechanism for activation of Src by PTPalpha." Embo J 19(5): 964-78.<br />

Ziegler, W. H., R. C. Liddington and D. R. Critchley (2006). "<strong>The</strong> structure and regulation of<br />

vinculin." Trends Cell Biol 16(9): 453-60.<br />

Zwartz, G. J., A. Chigaev, D. C. Dwyer, T. D. Foutz, B. S. Edwards and L. A. Sklar (2004).<br />

"Real-time analysis of very late antigen-4 affinity modulation by shear." J Biol Chem<br />

279(37): 38277-86.<br />

220

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!