30.12.2014 Views

Biophysical studies of membrane proteins/peptides. Interaction with ...

Biophysical studies of membrane proteins/peptides. Interaction with ...

Biophysical studies of membrane proteins/peptides. Interaction with ...

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.

UNIVERSIDADE TÉCNICA DE LISBOA<br />

INSTITUTO SUPERIOR TÉCNICO<br />

BIOPHYSICAL STUDIES OF MEMBRANE<br />

PROTEINS/PEPTIDES. INTERACTION WITH<br />

LIPIDS AND DRUGS.<br />

Fábio Monteiro Fernandes<br />

(Licenciado)<br />

Dissertação para obtenção do Grau de Doutor em Química<br />

Orientador:<br />

Co-orientador:<br />

Doutor Manuel José Estevez Prieto<br />

Doutor Luís Miguel Santos Loura<br />

Presidente:<br />

Vogais:<br />

Júri<br />

Reitor da Universidade Técnica de Lisboa<br />

Doutor Horst Vogel<br />

Doutor José Manuel Gaspar Martinho<br />

Doutor Miguel Augusto Rico Botas Castanho<br />

Doutor Manuel José Estevez Prieto<br />

Doutor Mário Nuno de Matos Sequeira Berberan e Santos<br />

Doutora Maria Raquel Murias dos Santos Aires de Barros<br />

Doutor Luís Miguel Santos Loura<br />

3 de Setembro de 2007


PREFÁCIO<br />

PREFÁCIO<br />

Biomembranas são estruturas de complexa composição química. As diversas classes<br />

de moléculas presentes, em especial lípidos e proteínas, apresentam propriedades e<br />

afinidades muito diferenciadas. Esta diferenciação resulta na ocorrência de<br />

heterogeneidades na distribuição lateral dos componentes na membrana. De facto, crêse<br />

que a célula controla certas actividades celulares operadas em biomembranas, através<br />

da mediação da distribuição destes componentes. Um maior conhecimento acerca das<br />

interacções estabelecidas entre os componentes de membranas biológicas é portanto de<br />

importância crucial para a compreensão das funções celulares baseadas nestas<br />

estruturas. Por outro lado as drogas são frequentemente dirigidas à interacção com<br />

proteínas inseridas em biomembranas. Assim, a caracterização das interacções destas<br />

moléculas com a bicamada lipídica e com os seus alvos proteicos é de grande relevância<br />

no processo de desenvolvimento de fármacos.<br />

Os estudos que compõem esta tese resultaram da aplicação de técnicas físicas,<br />

nomeadamente metodologias de fluorescência, à caracterização de interacções entre<br />

diversos componentes de membranas biológicas e de interacções proteína-fármacos.<br />

Sistemas modelo mimetizadores de membranas biológicas (lipossomas) foram<br />

utilizados de forma a trabalhar sobre uma base de estudo simplificada, mantendo ainda<br />

as propriedades fundamentais das biomembranas.<br />

Esta tese foi objecto dos seguintes artigos já publicados ou submetidos para<br />

publicação:<br />

Fernandes, F., Loura, L. M. S., Prieto, M., Koehorst, R., Spruijt, R., and Hemminga, M. (2003).<br />

Dependence <strong>of</strong> M13 major coat protein oligomerization and lateral segregation on bilayer<br />

composition. Biophys. J. 85: 2430-2431.<br />

Fernandes, F., Loura, L. M. S., Koehorst, R., Spruijt, R., Hemminga, M., Fedorov, A., and<br />

Prieto, M. (2004). Quantification <strong>of</strong> protein-lipid selectivity using FRET. Application to the<br />

M13 major coat protein. Biophys. J. 87: 344-352.<br />

Fernandes, F., Loura, L. M. S., Koehorst, R., Dixon, N., Kee, T. P., Hemminga, M., and Prieto,<br />

M. (2006). <strong>Interaction</strong> <strong>of</strong> the indole class <strong>of</strong> V-ATPase inhibitors <strong>with</strong> lipid bilayers.<br />

Biochemistry 45: 5271-5279.<br />

Fernandes, F., Loura, L. M. S., Koehorst, R., Dixon, N., Kee, T. P., Prieto, M., and Hemminga,<br />

M. (2006). Binding assays <strong>of</strong> inhibitors towards selected V-ATPase domains. Biochim Biophys<br />

Acta. 1758: 1777-1786.<br />

i


Fernandes, F., Loura, L. M. S., Fedorov, A., and Prieto, M. (2006). Absence <strong>of</strong> clustering <strong>of</strong><br />

phosphatidylinositol-(4,5)-bisphosphate in fluid phosphatidylcholine. J. Lipid Res. 47: 1521-<br />

1525.<br />

Fernandes, F., Neves, P., Gameiro, P., Loura, L. M. S., and Prieto, M. Cipr<strong>of</strong>loxacin<br />

<strong>Interaction</strong>s With Bacterial Protein OmpF: Modelling <strong>of</strong> FRET from a Multi-Tryptophan<br />

Protein Trimer (aceite para publicação).<br />

Fernandes, F., Loura, L. M. S., Matos, A. P., Fedorov, A., and Prieto, M. Role <strong>of</strong> Helix-0 <strong>of</strong> the<br />

N-BAR domain in <strong>membrane</strong> curvature generation. (submetido).<br />

Uma vez que a maior parte do trabalho já foi aceite para publicação em revistas<br />

internacionais, optou-se pela sua apresentação na forma de artigo científico e pela<br />

redacção desta tese em língua inglesa.<br />

Este manuscrito está dividido em nove partes. No primeiro Capítulo é feita uma<br />

introdução às biomembranas e às interacções proteína-lípido. Na segunda parte, os<br />

trabalhos respeitantes ao estudo da interacção da proteína principal da cápside do<br />

baacteriófago M13 são apresentados. O Capítulo III é dedicado aos estudos de<br />

interacção do potencial fármaco SB242784 com sistemas modelo de membranas e da<br />

ligação de inibidores da V-ATPase a um péptido contendo um putativo local de ligação<br />

para estes inibidores. No Capítulo IV é apresentado o trabalho respeitante aos estudos<br />

de ligação do antibiótico cipr<strong>of</strong>loxacina à proteína membranar OmpF. Seguidamente, no<br />

Capítulo V, os estudos da interacção do péptido N-terminal do domínio N-BAR com<br />

sistemas modelo de membranas são introduzidos e o Capítulo VI é dedicado aos estudos<br />

de caracterização da distribuição lateral de fosfoinositol-(4,5)-bisfosfato em liposomas<br />

de fosfocolinas. Os Capítulos II a VI, são sempre iniciados com uma breve introdução<br />

aos sistemas estudados.<br />

No Capítulo VII são apresentadas as conclusões globais referentes aos estudos de<br />

cada um dos sistemas investigados. Finalmente, o Capítulo VIII é dedicado à<br />

apresentação de considerações finais e perspectivas futuras e o Capítulo IX compreende<br />

a bibliografia.<br />

Gostaria ainda de finalizar agradecendo sinceramente a todos que de uma maneira<br />

ou outra contribuiram para a realização dos trabalhos aqui descritos:<br />

Agradeço ao Centro de Química-Física Molecular, na pessoa do Pr<strong>of</strong>essor José<br />

Manuel Gaspar Martinho, pelas facilidades concedidas para a realização deste trabalho.


PREFÁCIO<br />

Aos Pr<strong>of</strong>essores Manuel Prieto e Luís Loura, meus orientadores dos trabalhos de<br />

doutoramento, pela dedicação, apoio incansável, pela oportunidade de partilhar da sua<br />

paixão pela ciência, pelo muito que me ensinaram, mas principalmente pela amizade.<br />

À Pr<strong>of</strong>essora Ana Coutinho pelas muito úteis discussões científicas, pela<br />

disponibilidade constante e pela amizade.<br />

Aos meus colegas do grupo de Bi<strong>of</strong>ísica Molecular: Liana Silva (companheira de<br />

doutoramento), Rodrigo de Almeida, Bruno Castro e Sandra Pinto pela amizade,<br />

ambiente de camaradagem e apoio no dia-a-dia do laboratório.<br />

Aos antigos membros do grupo de Bi<strong>of</strong>ísica Molecular (Miguel Castanho, Sílvia<br />

Lopes, Ana Silva e Renske Hesselink) pela sua amizade. A Sílvia Lopes merece um<br />

agradecimento especial e com muito carinho, pela paciência, ajuda inestimável na<br />

realização de medidas de dicroísmo linear, mas principalmente pela amizade e pelo<br />

prazer da sua companhia! Um agradecimento especial também ao Pr<strong>of</strong>. Miguel<br />

Castanho pelo encorajamento e disponibilidade que sempre demonstrou.<br />

Aos restantes membros do Laboratório de Bi<strong>of</strong>ísica Molecular da Faculdade de<br />

Ciências da Universidade de Lisboa pela amizade e pelos momentos de diversão.<br />

Ao Doutor Alexander Fedorov, pela contribuição essencial para este trabalho<br />

através da realização das medidas de fluorescência resolvidas no tempo.<br />

À todos os colegas do Centro de Química-Física Molecular pelo companheirismo.<br />

Ao Pr<strong>of</strong>. Marcus Hemminga, Doutor Ruud Sprüijt e Rob Koehorst por todo o apoio<br />

concedido durante as minhas visitas ao seu laboratório. Um agradecimento especial ao<br />

Rob Koehorst pela sua amizade, pela preocupação e pelo seu entusiasmo contagiante<br />

pela ciência.<br />

Ao Doutor António Pedro Matos pela realização das medidas de microscopia<br />

electrónica de transmissão.<br />

Ao Pr<strong>of</strong>essor Benedito Cabral pela realização dos cálculos sobre o momento de<br />

transição do fármaco SB24784.<br />

Ao Pr<strong>of</strong>essor João Pessoa por me ter facultado a utilização do aparelho de dicroísmo<br />

circular.<br />

À Fundação para a Ciência e Tecnologia, pelo apoio financeiro concedido (bolsa<br />

SFRH/BD/14282/2003 e vários projectos no âmbito do POCTI e POCI).<br />

À Comissão Europeia pelo apoio concedido através do contracto número QLG-CT-<br />

2000-01801 (MIVASE consortium).<br />

iii


Ao Pedro e Hugo, amigos de longa data, por me conseguirem arrancar de casa e<br />

sempre ajudarem a manter a vida em perspectiva.<br />

E finalmente gostava de agradecer aos meus irmãos e meus pais, pelo exemplo,<br />

encorajamento e amor, sem os quais esta tese não seria certamente produzida.


CONTENTS<br />

CONTENTS<br />

Prefácio<br />

Contents<br />

Abbreviations and Symbol List<br />

Resumo<br />

Palavras-Chave<br />

Abstract<br />

Keywords<br />

Sinopse<br />

Outline<br />

I - INTRODUCTION<br />

I<br />

V<br />

IX<br />

XI<br />

XI<br />

XIII<br />

XIII<br />

XV<br />

XXI<br />

1<br />

1.- Bio<strong>membrane</strong>s<br />

1.1. – Function and architecture <strong>of</strong> bio<strong>membrane</strong>s<br />

1.2. – Molecular composition <strong>of</strong> bio<strong>membrane</strong>s<br />

1.2.1. – Lipid composition<br />

1.2.1.1. – Glycerolipids<br />

1.2.1.2. – Sphingolipids<br />

1.2.1.3. – Glycolipids<br />

1.2.1.4. – Sterols<br />

1.3. – Lipid asymmetry across the bilayer<br />

1.4. – Lipid structure and curvature<br />

1.5. – Lamellar phase transitions in lipid bilayers<br />

1.6. – Membrane thickness<br />

1.7. – Lateral heterogeneity in lipid bilayer<br />

1.8. – Membrane <strong>proteins</strong><br />

1.9. – Lateral dynamics in bio<strong>membrane</strong>s<br />

1.10. – Membrane model systems<br />

1<br />

1<br />

3<br />

3<br />

3<br />

5<br />

6<br />

6<br />

7<br />

8<br />

10<br />

12<br />

14<br />

17<br />

20<br />

21<br />

2. – Lipid-Protein <strong>Interaction</strong>s<br />

2.1. – Membrane protein reconstitution<br />

23<br />

23<br />

v


2.2. – Peptides as models<br />

2.3. – Amphipatic helix<br />

2.4. – Peptide partitioning to the <strong>membrane</strong><br />

2.5. – Anchoring <strong>of</strong> trans<strong>membrane</strong> domains<br />

2.6. – Lipid-protein hydrophobic mismatch<br />

2.7. – Trans<strong>membrane</strong> protein-lipid interface<br />

2.8. – Lipid selectivity at protein interfaces<br />

2.9. – Lipid phase preferential partition <strong>of</strong> <strong>membrane</strong> <strong>proteins</strong><br />

2.10. – Lipid sorting by <strong>proteins</strong> and formation <strong>of</strong> lipid domains<br />

2.11. – Lipid-mediation <strong>of</strong> protein-protein interactions<br />

25<br />

26<br />

28<br />

30<br />

31<br />

35<br />

37<br />

41<br />

43<br />

44<br />

II – PROTEIN-PROTEIN AND PROTEIN-LIPID INTERACTIONS<br />

OF M13 MAJOR COAT PROTEIN<br />

1. – Introduction<br />

2. – Dependence <strong>of</strong> M13 Major Coat Protein Oligomerization and Lateral<br />

Segregation on Bilayer Composition<br />

3. – Quantification <strong>of</strong> Protein-Lipid Selectivity Using FRET: Application to<br />

the M13 MCP<br />

47<br />

47<br />

53<br />

67<br />

III – BINDING OF INHIBITORS TO A PUTATIVE BINDING<br />

DOMAIN OF V-ATPase<br />

1. – Introduction<br />

2. – <strong>Interaction</strong> <strong>of</strong> the Indole Class <strong>of</strong> Vacuolar H + -ATPase Inhibitors <strong>with</strong><br />

Lipid Bilayers<br />

3. – Binding Assays <strong>of</strong> Inhibitors Towards Selected V-ATPase Domains<br />

79<br />

79<br />

83<br />

95<br />

IV – BINDING OF A QUINOLONE ANTIBIOTIC TO BACTERIAL<br />

PORIN OmpF<br />

1. – Introduction<br />

2. – Cipr<strong>of</strong>loxacin <strong>Interaction</strong>s With Bacterial Protein OmpF: Modelling <strong>of</strong><br />

FRET From a Multi-Tryptophan Protein Trimer<br />

107<br />

107<br />

109


CONTENTS<br />

V – INTERACTION OF HELIX-0 OF THE N-BAR DOMAIN WITH<br />

LIPID MEMBRANES<br />

1. – Introduction<br />

2. – Role <strong>of</strong> Helix-0 <strong>of</strong> the N-BAR Domain in Membrane Curvature<br />

Generation<br />

137<br />

137<br />

141<br />

VI – CLUSTERING OF PI(4,5)P 2 IN FLUID PC BILAYERS<br />

1. – Introduction<br />

2. – Absence <strong>of</strong> Clustering <strong>of</strong> Phosphatidylinositol-(4,5)-bisphosphate in<br />

Fluid Phosphatidylcholine<br />

169<br />

169<br />

173<br />

VII – CONCLUSIONS<br />

181<br />

VIII – FINAL CONSIDERATIONS AND PROSPECTS<br />

189<br />

IX – BIBLIOGRAPHY<br />

191<br />

vii


ABBREVIATIONS AND SYMBOL LIST<br />

ABBREVIATIONS AND SYMBOL LIST<br />

C<br />

cmc<br />

C p<br />

D<br />

DLPC<br />

DMPC<br />

DMPE<br />

DMPA<br />

DOPC<br />

DOPG<br />

DPPC<br />

DSC<br />

DSPC<br />

∆G 0<br />

ESR<br />

F<br />

GUV<br />

H0-NBAR<br />

H I<br />

H II<br />

k<br />

K p<br />

K b<br />

L<br />

L i<br />

L α<br />

L β<br />

L β’<br />

L c<br />

Specific heat<br />

Critical micelle concentration<br />

Heat capacity at constant pressure<br />

Lateral diffusion coefficient<br />

1,2-Dilauroyl-sn-glycero-3-phosphatidylcholine<br />

1,2-Dimyristoyl-sn-glycero-3-phosphatidylcholine<br />

1,2-Dimyristoyl-sn-glycero-3-phosphatidylethanolamine<br />

1,2-Dimyristoyl-sn-glycero-3-phosphatidic acid<br />

1,2-Dioleoyl-sn-glycero-3-phosphatidylcholine<br />

1,2-Dioleoyl-sn-glycero-3-phosphatidylglycerol<br />

1,2-Dipalmitoyl-sn-glycero-3-phosphatidylcholine<br />

Differential scanning calorimetry<br />

1,2-Distearoyl-sn-glycero-3- phosphatidylcholine<br />

Free energy <strong>of</strong> binding<br />

Electron spin resonance<br />

Lipid perturbation energy<br />

Giant unilamellar vesicle<br />

N-terminal amphipatic helix <strong>of</strong> N-BAR domain<br />

Hexagonal type I phase<br />

Hexagonal type II (or inverted) phase<br />

Boltzmann constant<br />

Partition coefficient<br />

Binding constant<br />

Lipid phase<br />

Lipid species i<br />

Fluid liquid-disordered phase<br />

Gel solid-ordered phase<br />

Tilted gel phase<br />

Crystalline phase<br />

ix


L o<br />

L/P<br />

LUV<br />

MARCKS<br />

mcp<br />

MLV<br />

n S,i<br />

PI(4)P<br />

PI(5)P<br />

PI(4,5)P<br />

2<br />

PI(3,4,5)P<br />

3<br />

PL i<br />

P β’<br />

PA<br />

PC<br />

PE<br />

PG<br />

PI<br />

PS<br />

R<br />

SDS<br />

sn<br />

s o<br />

SUV<br />

V i<br />

T<br />

TFE<br />

T m<br />

TM<br />

T p<br />

W<br />

Liquid-ordered phase<br />

Lipid to protein/peptide ratio<br />

Large unilamellar vesicle<br />

Myristoylated alanine-rich C kinase substrate<br />

Major coat protein<br />

Multilamellar vesicle<br />

Moles <strong>of</strong> solute in phase i<br />

Phosphatidylinositol-(4)-phosphate<br />

Phosphatidylinositol-(5)-phosphate<br />

Phosphatidylinositol-(4,5)-bisphosphate<br />

Phosphatidylinositol-(3,4,5)-trisphosphate<br />

Complex <strong>of</strong> protein and lipid species i<br />

Rippled Gel or Periodic Gel Phase<br />

Phosphatidic acid<br />

Phosphatidylcholine<br />

Phosphatidylethanolamine<br />

Phosphatidylglycerol<br />

Phosphatidylinositol<br />

Phosphatidylserine<br />

Ideal gas constant<br />

Sodium dodecyl sulfate<br />

Stereospecific number<br />

Gel or solid-ordered phase<br />

Small unilamellar vesicle<br />

Volume <strong>of</strong> phase i<br />

Temperature<br />

Trifluoroethanol<br />

Main transition temperature<br />

Trans<strong>membrane</strong><br />

Pre-transition temperature<br />

Aqueous phase


RESUMO<br />

RESUMO<br />

As biomembranas são responsáveis pela operação de diversas funções celulares e a<br />

multiplicidade destas interacções requer uma composição química complexa. Apesar da<br />

estrutura básica das biomembranas ser a bicamada lipídica, outros componentes estão<br />

presentes, em particular proteínas de membrana. Estes diferentes componentes<br />

apresentam energias de interacção distintas, induzindo frequentemente heterogenidades<br />

na distribuição lateral de biomembranas.<br />

Neste trabalho foram estudadas as interacções entre componentes de biomembranas<br />

em diferentes sistemas. É demonstrado que a distribuição lateral da proteína principal da<br />

cápside do bacteriófago M13 (mcp) é dependente da composição lipídica da membrana.<br />

Uma nova metodologia para quantificação de selectividades proteína-lípido foi<br />

desenvolvida e aplicada com sucesso à mcp.<br />

Estudos de ligação entre drogas e proteínas de membrana foram conduzidos<br />

seguindo diferentes aproximações. No estudo da ligação de inibidores de V-ATPase a<br />

um domínio transmembranar do enzima, uma aproximação reducionista foi escolhida,<br />

enquanto que no estudo da ligação do cipr<strong>of</strong>loxacin para a OmpF, a proteína foi<br />

utilizada na forma intacta.<br />

A interacção entre um segmento N-terminal anfipático de um domínio BAR com<br />

membranas foi investigada e conclusões foram obtidas relativamente ao papel deste<br />

segmento na remodelação de membranas por domínios BAR.<br />

Finalmente, a hipótese de agregação espontânea de fosfoinositol-(4,5)-bisfosfato foi<br />

testada e excluída.<br />

PALAVRAS‐CHAVE<br />

Interacções Proteína‐Lípido<br />

Proteína Principal da Cápside do M13<br />

Inibição do V‐ATPase<br />

PI(4,5)P 2<br />

Domínios BAR<br />

FRET<br />

xi


ABSTRACT<br />

ABSTRACT<br />

Bio<strong>membrane</strong>s are responsible for the performance <strong>of</strong> multiple cellular functions<br />

and the multiplicity <strong>of</strong> these tasks requires a complex chemical composition. While the<br />

structural framework <strong>of</strong> bio<strong>membrane</strong>s is the lipid bilayer, other components, notably<br />

<strong>proteins</strong>, are present. These distinct components present different energies <strong>of</strong><br />

interaction, leading to heterogeneous lateral distribution in the <strong>membrane</strong>.<br />

In this work, the interactions between the components <strong>of</strong> bio<strong>membrane</strong>s in several<br />

different systems were studied. The lateral distribution <strong>of</strong> major coat protein (mcp) <strong>of</strong><br />

the M13 bacteriophage was shown to be highly dependent on the lipid composition <strong>of</strong><br />

the <strong>membrane</strong>. A new methodology for quantification <strong>of</strong> protein-lipid selectivity was<br />

also developed and applied <strong>with</strong> success to the mcp.<br />

Protein-drug binding <strong>studies</strong> were performed following different approaches. To the<br />

study <strong>of</strong> binding <strong>of</strong> a V-ATPase inhibitor to a selected trans<strong>membrane</strong> domain <strong>of</strong> the<br />

enzyme, a reductionist approach was considered, while in the case <strong>of</strong> binding <strong>of</strong><br />

cipr<strong>of</strong>loxacin to OmpF, the intact protein was used.<br />

The interaction <strong>of</strong> an N-terminal amphipatic segment <strong>of</strong> a BAR domain <strong>with</strong><br />

<strong>membrane</strong>s was also investigated and insight was achieved on the possible role <strong>of</strong> this<br />

segment in <strong>membrane</strong> remodelling.<br />

Finally, the hypothesis <strong>of</strong> spontaneous clustering <strong>of</strong> phosphatidylinositol-(4,5)-<br />

bisphophate was tested and ruled out.<br />

KEYWORDS<br />

Protein‐Lipid <strong>Interaction</strong>s<br />

M13 Major Coat Protein<br />

V‐ATPase Inhibition<br />

PI(4,5)P 2<br />

BAR domains<br />

FRET<br />

xiii


SINOPSE<br />

SINOPSE<br />

Nas últimas duas décadas, um esforço colectivo massivo permitiu um conhecimento<br />

mais detalhado acerca da natureza das interacções entre os componentes de<br />

biomembranas e as importantes consequências biológicas dessas mesmas interacções. O<br />

modelo do mosaico fluído de Singer-Nicolson para biomembranas (1972) descrevia um<br />

sistema em que os lípidos se apresentavam como pouco mais do que uma matriz onde<br />

proteínas se insiriam. A visão actualmente aceite é muito mais complexa. Não só as<br />

interacções proteína-proteína podem ser responsáveis pela mediação de funções<br />

celulares, como as interacções proteína-lípido são consideradas muito mais complexas e<br />

relevantes do que anteriormente. Estudos bi<strong>of</strong>ísicos detalhados como os aqui descritos,<br />

permitem a recolha de informação valiosa relativamente às características destas<br />

interacções.<br />

As biomembranas traçam os limites para a vida celular. Elas permitem a<br />

diferenciação de ambientes moleculares através do controle da permeabilidade, criando<br />

as condições para a especialização no interior da célula. Esta tarefa requer o consumo de<br />

quantidades de energia massivas, enfatizando a importância dramática destas estruturas.<br />

A base estrutural das biomembranas é a bicamada lipídica. No entanto, os lípidos<br />

estão longe de corresponder a uma classe homogénea de moléculas. Quando as energias<br />

de interacção entre moléculas lipídicas são significativamente diferentes, separação de<br />

fases em grande escala pode ocorrer e mesmo para pequenas diferenças entre energias<br />

de interacção lípido-lípido, heterogenidades em pequena escala são expectáveis. Assim,<br />

a distribuição lateral de uma bicamada lipídica é frequentemente heterogénea.<br />

Esta representação das biomembranas torna-se ainda mais complexa tendo em conta<br />

que os lípidos não são os únicos componentes presentes. De facto,as biomembranas são<br />

uma mistura de diferentes classes de moléculas com diferentes funções. Além dos<br />

lípidos, os componentes mais importantes são as proteínas, que se encontram altamente<br />

concentradas nas biomembranas e são largamente responsáveis pelas funções celulares<br />

operadas nestas estruturas.<br />

As interacções entre estas diversas espécies moleculares são bastante complexas.<br />

Ligações de hidrogénio, interacções electrostáticas, e de van der Waals, operam<br />

simultaneamente na ordenação do empacotamento de equilíbrio de proteínas e lípidos.<br />

Proteínas e lípidos organizam-se de forma a proteger os seus domínios hidrófobos do<br />

xv


ambiente aquoso e assim prevenir as perdas entrópicas resultantes da interacção com as<br />

moléculas de água. Como consequência, proteínas de membrana e lípidos interactuam<br />

de forma diferenciada, dependendo da distribuição dos resíduos polares e hidrófobos na<br />

proteína e do comprimento do domínio hidrófobo no lípido.<br />

Estas alterações nas selectividades de associação podem levar a alterações na<br />

distribuição lateral dos componentes de membrana, de uma forma similar à observada<br />

para misturas lipídicas. Desta forma, a distribuição heterógenea de lípidos pode induzir<br />

uma distribuição heterogénea de proteínas e vice-versa. De facto, é conhecida<br />

actualmente a tendência de proteínas e lípidos para se distribuir de forma heterogénea<br />

em biomembranas e este fenómeno é essencial para garantir o carácter localizado de<br />

determinados eventos celulares.<br />

As drogas são agentes que frequentemente também demonstram afinidade por<br />

membranas lipídicas. Os alvos de diversas drogas são proteínas de membrana e a sua<br />

acção é frequentemente realizada no ambiente membranar. Assim, as interacções de<br />

drogas com componentes das biomembranas são de grande relevância biológica. Da<br />

mesma forma, para drogas dirigidas a proteínas no interior da célula, a partição para a<br />

membrana plasmática é o primeiro passo no processo de entrada no ambiente celular e<br />

os seus coeficientes de partição água/lípido são deste modo parâmetros biologicamente<br />

relevantes.<br />

As interacções entre os componentes das biomembranas são utilizadas pelas células<br />

para mediar e localizar diversas funções nas membranas. Uma melhor compreensão<br />

destes mecanismos é o primeiro passo na compreensão das biomembranas e das suas<br />

funções. No entanto, as membranas celulares são muitas vezes sistemas demasiado<br />

complexos para estudar em detalhe as características das interacções estabelecidas pelos<br />

seus componentes. Uma alternativa viável é o estudo destas interacções em sistemas<br />

modelo de membranas. No presente trabalho, a maioria dos estudos foram efectuados<br />

em vesículas unilamelares com diâmetro aproximado de 100 nm. Este sistema modelo<br />

permite uma base de estudo simplificada e a manutenção das propriedades fundamentais<br />

das biomembranas.<br />

As técnicas de fluorescência são ferramentas ideais no estudo da distribuição lateral<br />

em membranas. Através da escolha de um aminoácido fluorescente intrínseco (Trp ou<br />

Tyr), ou através da marcação de proteínas e/ou lípidos, é possível monitorar as<br />

alterações de fluorescência (intensidade, tempos de vida e anisotropia) resultantes da<br />

interacção com outros componentes membranares. Estudos de extinção de fluorescência


SINOPSE<br />

podem fornecer informação acerca das eficiências de colisão, e esta informação pode<br />

dar indicações sobre: i) a localização do grupo fluorescente; ii) a concentração do<br />

agente responsável pela extinção de fluorescência; iii) as propriedades difusivas do<br />

agente responsável pela extinção de fluorescência e da molécula fluorescente. Uma<br />

técnica que é particularmente útil na avaliação das distribuições laterais na membrana é<br />

a transferência de energia electrónica por ressonância (FRET). A eficicência de FRET<br />

para um único par dador-aceitante depende fortemente da distância entre eles. No caso<br />

de membranas lipídicas com uma distribuição de dadores e aceitantes, FRET é sensível<br />

às alterações na distribuição de aceitantes próximos ao dador, numa escala espacial<br />

comparável à distância de Förster (até 10 nm dependendo do par dador-aceitante).<br />

Vários sistemas foram aqui alvo de estudo. No Capítulo II, uma proteína integral<br />

também pertencente à cápside do bacteriófago M13 (mcp), foi estudada em sistemas<br />

modelo de membranas. O estado de oligomerização da proteína s<strong>of</strong>re alterações<br />

repetidas durante o ciclo reproductivo do bacteriófago M13. No entanto, crê-se que o<br />

estado da proteína quando incorporada na membrana do hospedeiro é essencialmente<br />

monomérico, pois tal é necessário para a correcta incorporação da proteína numa<br />

partícula de bacteriófago em formação. Ainda assim, esta questão é alvo de debate,<br />

uma vez que estados oligomerizados da proteína foram observados em certas condições.<br />

Aqui é demonstrado que mcp é um monómero quando incorporado em membranas<br />

compostas por lípidos com espessura hidrófoba idêntica à da proteína (cadeias<br />

monoinsaturadas com 18 carbonos – conformidade hidrófoba). Por outro lado, quando<br />

foram utilizadas bicamadas apresentando uma espessura hidrófoba maior que a da<br />

proteína (cadeias monoinsaturadas com 22 carbonos), foi detectada agregação. Quando<br />

foram utilizadas misturas de lípidos com espessuras hidrófobas em conformidade e<br />

descomformidade com a espessura hidrófoba da proteína, foi observada uma<br />

segregação de mcp para domínios enriquecidos em lípidos com conformidade<br />

hidrófoba. A mcp nestes domínios manteve um estado monomérico. Um aumento de 4<br />

carbonos nas cadeias acilo da matriz lipídica pode induzir agregação proteica, e no caso<br />

da presença de uma fracção significativa de lípido em conformidade hidrófoba,<br />

segregação lípido-lípido pode ser induzida pela presença da proteína. Este resultado é<br />

demonstrativo da enorme relevância da conformidade hidrófoba para a organização de<br />

proteínas e lípidos.<br />

Uma metodologia de FRET foi desenvolvida com o intuito de quantificar o grau de<br />

selectividade exibida pela mcp relativamente a diferentes lípidos. Os resultados obtidos<br />

xvii


para diferentes grupos polares de lípidos, são practicamente idênticos aos obtidos com<br />

espectroscopia de ressonância paramagnética eletrónica (ESR) para a mesma proteína,<br />

ainda que num estado agregado. Esta concordância estabelece esta metodologia para<br />

quantificação da selectividade de proteínas para lípidos como uma alternativa ao uso de<br />

ESR. Esta metodologia de FRET foi também aplicada ao estudo da selectividade da<br />

mcp para diferentes comprimentos de cadeias acilo. Surpreendentemente, os resultados<br />

revelaram constantes de selectividade baixas para um lípido de conformidade hidrófoba,<br />

em bicamadas mais finas e mais espessas que o domínio hidrófobo da proteína.<br />

Duas estratégias diferentes foram utilizadas no estudo da ligação de drogas a<br />

proteínas de membrana. No Capítulo III, uma aproximação reducionista foi escolhida.<br />

Um péptido correspondente ao quarto domínio transmembranar da subunidade c do V-<br />

ATPase foi utilizado em estudos de ligação aos inibidores do V-ATPase, bafilomicina e<br />

SB242784. Era esperado que este domínio transmembranar compreendesse o local de<br />

ligação a estes inibidores. SB242784 foi desenvolvido para utilização terapêutica em<br />

pacientes com osteoporose. Este inibidor apresenta grande selectividade para a forma<br />

osteoclática do V-ATPase que é parcialmente responsável pela degradação óssea. A<br />

bafilomicina apresenta níveis de toxicidade muito elevados devido à ausência de<br />

selectividade para uma forma particular do enzima. FRET foi novamente utilizada para<br />

determinar ligação entre inibidor e péptido (a tirosina do péptido foi o dador e os<br />

inibidores os aceitantes) e os resultados indiciam uma interacção de baixa afinidade<br />

entre a bafilomicina e o péptido. Já a ligação de SB242784 ao péptido não foi detectada.<br />

Globalmente, os resultados indicam que os requesitos moleculares para ligação aos<br />

inibidores na subunidade c incluem provavelmente contribuições dos restantes domínios<br />

transmembranares, reflectindo uma visão mais complexa do mecanismo de inibição do<br />

V-ATPase do que inicialmente considerado.<br />

No Capítulo IV, descreve-se um estudo de ligação entre o antibiótico cipr<strong>of</strong>loxacin<br />

(CP) e um trímero purificado de uma porina da <strong>membrane</strong> externa, OmpF. O<br />

mecanismo de entrada do CP na célula, requer a interacção do CP com a OmpF. Neste<br />

caso, a estratégia escolhida envolveu o uso da estrutura nativa da proteína. FRET foi<br />

novamente aplicada, desta vez fazendo uso da presença de dois tipos de Trp no<br />

monómero de OmpF (como dadores) e das propriedades de absorção no ultravioleta do<br />

CP (como aceitante). A análise da extinção de fluorescência de aminoácidos intrínsecos<br />

de proteínas de grandes dimensões por FRET pode revelar-se difícil, uma vez que cada<br />

uma das populações de dadores pode estar sujeita a diferentes distribuições de


SINOPSE<br />

aceitantes. Dadores mais próximos da periferia da proteína podem estar mais próximos<br />

de aceitantes do que dadores no centro da proteína. Tendo em conta esta preocupação,<br />

um modelo de FRET foi desenvolvido para aplicação em sistemas contendo proteínas<br />

com múltiplos resíduos de Trp, considerando as restrições geométricas da OmpF. Esta<br />

metodologia permitiu a recuperação de um intervalo possível para a constante de<br />

ligação entre a OmpF e o CP. Em virtude das grandes dimensões do trímero de OmpF,<br />

dependendo do local de ligação ao antibiótico assumido no modelo utilizado na análise<br />

dos dados de FRET, a eficiência de ligação pode ser diferente. Dois locais de ligação ao<br />

CP correspondendo a situações limite (ligação no centro do trímero e ligação à periferia<br />

do trímero) foram considerados. Comparando os resultados obtidos segundo estes<br />

modelos e os resultados obtidos através de um método independente, foi possível<br />

concluir que o local de ligação do CP à OmpF está provavelmente distante do centro do<br />

trímero e mais próximo da periferia.<br />

No Capítulo V, a interacção de um péptido anfipático N-terminal de um domínio N-<br />

BAR com sistemas modelo de membranas foi estudada. Diversos autores consideraram<br />

a hipótese de que este segmento contribui para as propriedades remodeladoras de<br />

membranas do domínio N-BAR (tubulação de liposomas esféricos), através da inserção<br />

pr<strong>of</strong>unda no interior da bicamada. Isto teria como consequência a indução de assimetria<br />

entre as monocamadas dos lipossomas, que por fim levaria à alterações da curvatura na<br />

membrana. Esta hipótese foi aqui testada através do estudo das interacções de um<br />

péptido correspondendo a este segmento com sistemas modelo de membranas.<br />

Novamente, metodologias de fluorescência foram utilizadas e complementadas com<br />

microscopia electrónica. Conclui-se que a ligação do péptido às membranas é<br />

essencialmente electrostática, afastando a hipótese frequentemente levantada de este<br />

segmento dos domínios N-BAR agir através da estabilização da conformação do<br />

domínio associado com membranas, mediante fortes interacções hidrófobas com o<br />

interior da bicamada lipídica. De qualquer forma, os resultados de um estudo de FRET<br />

apontam para um papel alternativo do segmento N-terminal do domínio N-BAR. Os<br />

resultados demonstram claramente que o péptido se encontra na forma dimerizada<br />

depois da ligação à membrana. Estes resultados estão em concordância com uma<br />

proposta recente de que o segmento anfipático N-terminal dos domínios N-BAR poderia<br />

agir através da mediação da associação entre dímeros de domínios BAR. A microscopia<br />

electrónica não permitiu identificar alterações morfológicas significativas dos<br />

lipossomas após interacção com o péptido.<br />

xix


Finalmente, a agregação de fosfoinositol-4,5-bisfosfato (PI(4,5)P 2 ) em bicamadas de<br />

fosfocolinas foi estudada (Capítulo VI), de forma a testar a hipótese de agregação<br />

espontânea deste lípido numa matriz de fosfocolinas. Os resultados de extinção de<br />

fluorescência colisional obtidos com um derivado de PI(4,5)P 2 marcado com uma sonda<br />

fluorescente e de um estudo de FRET utilizando este lípido derivatizado como dador<br />

para uma sonda distribuida homogeneamente na membrana, são inequívocos na<br />

confirmação de uma distribuição homogénea de PI(4,5)P 2 na bicamada lipídica. No<br />

mesmo estudo, é demonstrado que a diferentes estados de protonação de PI(4,5)P 2<br />

correspondem diferentes eficiências de partição para o ambiente lipídico, um fenómeno<br />

que pode apresentar relevância biológica significativa.


OUTLINE<br />

OUTLINE<br />

The last two decades, through a massive collective effort, allowed some insight into<br />

the nature <strong>of</strong> the interactions between the components <strong>of</strong> bio<strong>membrane</strong>s and also the<br />

highly relevant biological consequences <strong>of</strong> these interactions. The Singer-Nicolson fluid<br />

mosaic model for bio<strong>membrane</strong>s (1972), described a system where lipids were little else<br />

than a matrix where <strong>proteins</strong> were dispersed. The picture currently accepted is much<br />

more complex. Not only protein-protein interactions can be responsible for mediation <strong>of</strong><br />

cellular functions, but also protein-lipid interactions are much more complex and<br />

important than originally thought. Detailed biophysical <strong>studies</strong> such as the ones<br />

described here, allow gathering <strong>of</strong> valuable information regarding the character <strong>of</strong> such<br />

interactions.<br />

Bio<strong>membrane</strong>s delineate the boundaries <strong>of</strong> cellular life. They allow the<br />

differentiation <strong>of</strong> molecular environments through the control <strong>of</strong> permeability, creating<br />

the conditions for specialization <strong>with</strong>in the cell. This task demands the expenditure <strong>of</strong> a<br />

massive amount <strong>of</strong> energy, emphasizing the dramatic importance <strong>of</strong> bio<strong>membrane</strong>s.<br />

The foremost structural framework <strong>of</strong> bio<strong>membrane</strong>s is the lipid bilayer, but lipids<br />

are far from being a homogeneous class <strong>of</strong> molecules. When the energies <strong>of</strong> interaction<br />

between lipid molecules are significantly different, large scale phase separation <strong>of</strong> lipid<br />

molecules in the <strong>membrane</strong> can occur, and even for small differences in lipid-lipid<br />

interaction energies, short-range heterogeneities are expected. Therefore, the lateral<br />

distribution <strong>of</strong> a multicomponent lipid bilayer is <strong>of</strong>ten heterogeneous.<br />

This portrait <strong>of</strong> bio<strong>membrane</strong>s is made even more complex by the notion that lipids<br />

are not the only components found there. In fact, bio<strong>membrane</strong>s are a mixture <strong>of</strong><br />

different classes <strong>of</strong> molecules <strong>with</strong> different functions. Apart from lipids, the most<br />

notable component are <strong>proteins</strong>, which are highly concentrated in bio<strong>membrane</strong>s and<br />

are largely responsible for the cellular functions operated there.<br />

<strong>Interaction</strong>s between these quite different molecular species are very complex.<br />

Hydrogen bonds, electrostatic, and van der Waals forces, all operate simultaneously in<br />

dictating the equilibrium packing <strong>of</strong> <strong>proteins</strong> and lipids in bilayers. Notably, <strong>proteins</strong><br />

and lipids will arrange in such a manner that their hydrophobic domains are shielded<br />

from the aqueous environment in order to prevent the entropy loss resulting from<br />

interaction <strong>with</strong> water molecules. As a consequence, <strong>membrane</strong> <strong>proteins</strong> and lipids will<br />

xxi


interact differentially, depending on the distribution <strong>of</strong> polar and hydrophobic residues<br />

on the protein and on the size <strong>of</strong> the hydrophobic lipid domain.<br />

These changes in selectivities <strong>of</strong> association can lead to changes in the lateral<br />

distribution <strong>of</strong> the <strong>membrane</strong> components in a similar manner as observed for lipid<br />

mixtures. In this way, a heterogeneous distribution <strong>of</strong> lipids can drive <strong>membrane</strong> protein<br />

heterogeneity and vice-versa. In fact, it is known that <strong>proteins</strong> and lipids are <strong>of</strong>ten<br />

distributed in a heterogeneous manner in the <strong>membrane</strong>, and this is essential in ensuring<br />

the localized character <strong>of</strong> some events in cellular <strong>membrane</strong>s.<br />

Drugs are another type <strong>of</strong> agents which <strong>of</strong>ten display affinity for lipid <strong>membrane</strong>s.<br />

The targets <strong>of</strong> several drugs are <strong>membrane</strong> <strong>proteins</strong>, and their action is frequently<br />

expected to be exerted in the <strong>membrane</strong> environment. Therefore, interaction <strong>of</strong> drugs<br />

<strong>with</strong> bio<strong>membrane</strong> components is <strong>of</strong> great biological relevance. In addition, for drugs<br />

targeting <strong>proteins</strong> inside the cell, partition to the plasma <strong>membrane</strong> is the first step to<br />

cell entry, and water/lipid partition coefficients are in this way, relevant parameters in<br />

drug research.<br />

It is clear that interactions between bio<strong>membrane</strong> components are used by cells to<br />

mediate and localize diverse functions in the cell <strong>membrane</strong>. A better understanding <strong>of</strong><br />

these mechanisms is the first step in rationalizing bio<strong>membrane</strong>s and their functions.<br />

However, cellular <strong>membrane</strong>s are <strong>of</strong>ten too intricate for the properties <strong>of</strong> some<br />

<strong>membrane</strong> component interactions to be resolved, and a valuable alternative is the use<br />

<strong>of</strong> <strong>membrane</strong> model systems. Here, the majority <strong>of</strong> the <strong>studies</strong> were carried out in large<br />

unilamellar vesicles, <strong>with</strong> a diameter around 100 nm. This model system granted us a<br />

simplified basis for the study <strong>of</strong> the interactions between <strong>membrane</strong> components, at the<br />

same time maintaining all the fundamental properties <strong>of</strong> the bio<strong>membrane</strong>.<br />

Fluorescence techniques are ideal tools to study lateral distribution <strong>of</strong> <strong>membrane</strong>s.<br />

By choosing an intrinsic fluorescent amino acid (Trp or Tyr), or fluorescent labelling <strong>of</strong><br />

<strong>proteins</strong> and lipids, we can monitor fluorescence changes (intensity, lifetime,<br />

anisotropy) upon interaction <strong>with</strong> other <strong>membrane</strong> components. Fluorescence quenching<br />

<strong>studies</strong> can give information on the collision efficiency, and this can give us indications<br />

on: i) the location <strong>of</strong> the fluorescent group; ii) on the concentration <strong>of</strong> quencher; and iii)<br />

on the diffusion properties <strong>of</strong> both quencher and fluorescent molecule. One technique<br />

that was particularly useful in evaluating lateral distributions in the <strong>membrane</strong> was<br />

Förster resonance energy transfer (FRET). FRET efficiency for a single donor-acceptor<br />

pair is strongly dependent on the distance between the two. In the case <strong>of</strong> lipid


OUTLINE<br />

<strong>membrane</strong>s <strong>with</strong> a distribution <strong>of</strong> donors and acceptors, FRET will be sensitive to<br />

changes in the distribution <strong>of</strong> acceptors around the donor population in the space range<br />

<strong>of</strong> the Förster radius (up to ~10 nm depending on the donor-acceptor pair).<br />

Several different systems were investigated. In Chapter II, an integral protein from<br />

the coat <strong>of</strong> the bacteriophage M13 (mcp) was studied in model <strong>membrane</strong>s. The<br />

oligomerization state <strong>of</strong> the protein changes repeatedly during the reproductive cycle <strong>of</strong><br />

bacteriophage M13, but the state <strong>of</strong> the protein in the <strong>membrane</strong> was thought to be<br />

monomeric as this is apparently required for correct insertion in nascent bacteriophages.<br />

However, this was still matter <strong>of</strong> debate, as in the past, oligomeric mcp was shown to<br />

exist in some conditions. Here, it is shown that mcp is a monomer when incorporated in<br />

<strong>membrane</strong>s composed by lipids <strong>with</strong> hydrophobic thickness that matched the<br />

hydrophobic thickness <strong>of</strong> the protein (monounsatured chains C 18 chains). On the other<br />

hand, when bilayers <strong>with</strong> thicker hydrophobic thickness (monounsatured chains C 22<br />

chains) were used, aggregation was observed. Surprisingly, when mixtures <strong>of</strong> both<br />

matching and mismatching lipids were used, segregation <strong>of</strong> mcp to domains enriched in<br />

matching lipid was apparently induced, and while in these domains, mcp remained<br />

monomeric. This result is a dramatic demonstration <strong>of</strong> the relevance <strong>of</strong> hydrophobic<br />

matching for protein-lipid organization. Only a small increase <strong>of</strong> 4 carbons in the acylchain<br />

<strong>of</strong> the matrix lipid can drive protein aggregation, and in case that a significant<br />

fraction <strong>of</strong> matching lipid is present, lipid-lipid segregation can be induced by the<br />

presence <strong>of</strong> the protein.<br />

A FRET methodology was developed to quantify the degree <strong>of</strong> selectivity exhibited<br />

by mcp to different lipids. The results obtained for lipids <strong>with</strong> different headgroups are<br />

almost perfectly matched to previous results obtained <strong>with</strong> electron spin resonance<br />

(ESR) for the same protein, even though in an aggregated state. This agreement<br />

establishes the FRET methodology for quantification <strong>of</strong> protein-lipid selectivity<br />

described here as an alternative to the use <strong>of</strong> ESR. This FRET methodology was also<br />

applied to the study <strong>of</strong> selectivity for acyl-chain length, which is as already mentioned,<br />

essential in dictating the lateral distribution <strong>of</strong> the protein in liposomes. Also<br />

surprisingly, the results revealed weak selectivity constants for a hydrophobically<br />

matching lipid, both in thinner as in thicker bilayers.<br />

Drug binding to <strong>membrane</strong> <strong>proteins</strong> was studied following two different strategies.<br />

In chapter III, a reductionist approach was used. A peptide expected to correspond to the<br />

putative 4 th trans<strong>membrane</strong> domain <strong>of</strong> the <strong>membrane</strong> bound c-subunit <strong>of</strong> V-ATPase,<br />

xxiii


was used in binding <strong>studies</strong> for the V-ATPase inhibitors bafilomycin and SB242784.<br />

This trans<strong>membrane</strong> domain is expected to comprise the binding site for the studied<br />

inhibitors. SB242784 was developed to be used as a potential drug for treatment <strong>of</strong><br />

osteoporosis. This inhibitor presents selectivity for the osteoclastic form <strong>of</strong> the enzyme,<br />

while bafilomycin is extremely toxic due to its lack <strong>of</strong> selectivity. FRET was again used<br />

(a tyrosine residue <strong>of</strong> the peptide was used as a donor and the inhibitors as acceptors) to<br />

determine binding, and the results indicate a weak binding <strong>of</strong> the chosen peptide <strong>with</strong><br />

bafilomycin whereas binding <strong>of</strong> the peptide to SB242784 was not detected. Overall, the<br />

results indicate that the V-ATPase inhibitor binding site is likely not formed only by the<br />

4 th trans<strong>membrane</strong> segment <strong>of</strong> the c-subunit <strong>of</strong> V-ATPase, but is the result <strong>of</strong><br />

contributions from other trans<strong>membrane</strong> domains, reflecting a more complex view <strong>of</strong><br />

the inhibitory mechanism <strong>of</strong> V-ATPase than originally proposed.<br />

In Chapter IV, a binding study was conducted between cipr<strong>of</strong>loxacin (CP), a<br />

quinolone antibiotic, and a purified trimer <strong>of</strong> the outer <strong>membrane</strong> porin, OmpF. CP<br />

requires interactions <strong>with</strong> OmpF for efficient entry into the cell. In this case, the native<br />

protein structure was used instead <strong>of</strong> the minimalist approach described in Chapter III.<br />

FRET was again applied, this time making use <strong>of</strong> the presence <strong>of</strong> two types <strong>of</strong><br />

tryptophans <strong>of</strong> OmpF (as donors) and the UV absorbing properties <strong>of</strong> CP (acceptor).<br />

Fluorescence from intrinsic amino acids <strong>of</strong> large <strong>proteins</strong> in FRET can be difficult to<br />

analyze, since each donor (fluorescent amino acid) population is expected to sense a<br />

different population <strong>of</strong> acceptors. Donors closer to the protein periphery, can e.g., be in<br />

closer proximity to acceptors than donors in the core <strong>of</strong> the protein. With this in mind, a<br />

FRET methodology suitable for application to FRET <strong>studies</strong> <strong>with</strong> multi-donor <strong>proteins</strong><br />

was developed according to the geometric restrictions <strong>of</strong> the OmpF system. This model<br />

allowed the recovery <strong>of</strong> a range for the binding constants <strong>of</strong> the OmpF-CP association<br />

process. Due to the large dimensions <strong>of</strong> the OmpF trimer, depending on the site <strong>of</strong><br />

inhibitor binding assumed in the model for FRET data analysis, the retrieved binding<br />

efficiency can be different. Two limiting binding sites were considered, and comparing<br />

the results from our analysis <strong>with</strong> the results obtained <strong>with</strong> an independent method, it<br />

was possible to conclude that the binding site for CP in OmpF is likely to be displaced<br />

from the center <strong>of</strong> the trimer and closer to the periphery.<br />

In Chapter V, the interaction <strong>of</strong> the N-terminal amphipatic peptide <strong>of</strong> a N-BAR<br />

domain <strong>with</strong> model <strong>membrane</strong>s was studied. Several authors hypothesized that this<br />

segment <strong>of</strong> N-BAR domain contributed to the <strong>membrane</strong> remodelling properties <strong>of</strong> N-


OUTLINE<br />

BAR domains (tubulation <strong>of</strong> spherical liposomes) by inserting deeply into the bilayer<br />

core, driving monolayer asymmetry and thereby forcing a change in the normal<br />

curvature <strong>of</strong> the bilayer. We tested this hypothesis by following the interaction <strong>of</strong> the<br />

corresponding peptide <strong>with</strong> <strong>membrane</strong> model systems. Again, fluorescence<br />

spectroscopy methodologies were used and complemented <strong>with</strong> the application <strong>of</strong><br />

electron microscopy to investigate changes in liposome morphology upon binding <strong>of</strong> the<br />

peptide. It is concluded that binding to liposomes is essentially electrostatic, ruling out<br />

the frequently hypothesized role <strong>of</strong> this protein segment to operate through the<br />

stabilization <strong>of</strong> the <strong>membrane</strong> bound conformation <strong>of</strong> BAR domains via hydrophobic<br />

short-range interactions <strong>with</strong> the bilayer core. Nevertheless, the results <strong>of</strong> a FRET study<br />

clearly indicate the tendency <strong>of</strong> the peptide to oligomerize in the lipid <strong>membrane</strong><br />

environment, supporting a recently proposed function <strong>of</strong> this segment as a mediator <strong>of</strong><br />

aggregation between BAR domain dimers. Electron microscopy measurements were not<br />

able to detect significant morphology changes in the liposomes upon addition <strong>of</strong><br />

<strong>peptides</strong>.<br />

Finally, clustering behaviour <strong>of</strong> phosphatidylinositol-4,5-bisphophate (PI(4,5)P 2 ) in<br />

phosphatidylcholine bilayers was studied (Chapter VI) in order to test the hypothesis <strong>of</strong><br />

spontaneous aggregation <strong>of</strong> this lipid in a PC matrix. Both fluorescence self-quenching<br />

data obtained <strong>with</strong> a fluorescently labelled PI(4,5)P 2 , and a FRET study using this lipid<br />

as a donor to a homogeneously distributed <strong>membrane</strong> probe, unequivocally confirmed a<br />

homogeneous distribution <strong>of</strong> PI(4,5)P 2 in the lipid bilayer. In the same study, it is<br />

demonstrated that different protonation states <strong>of</strong> PI(4,5)P 2 correspond to different<br />

water/lipid partition coefficients, a phenomenon that can be <strong>of</strong> significant biological<br />

relevance.<br />

xxv


INTRODUCTION: BIOMEMBRANES<br />

I<br />

INTRODUCTION<br />

1.BIOMEMBRANES<br />

1.1. Function and architecture <strong>of</strong> bio<strong>membrane</strong>s<br />

Bio<strong>membrane</strong>s delineate the boundaries <strong>of</strong> cells as plasma <strong>membrane</strong>s, and in the<br />

eukaryotic cell, enclose organelles in the form <strong>of</strong> intracellular <strong>membrane</strong>s, allowing the<br />

subdivision <strong>of</strong> cellular activities and the more diverse and specialized functions found in<br />

eukaryotes. The bio<strong>membrane</strong> is the passive permeability barrier that allows the<br />

maintenance <strong>of</strong> different molecular environments in the inside and outside <strong>of</strong> the cell or<br />

organelle. The importance <strong>of</strong> this task is emphasized by the fact that a significant<br />

fraction <strong>of</strong> the energy required for life is expended in the preservation <strong>of</strong> the differences<br />

in molecular environments across bio<strong>membrane</strong>s.<br />

The foremost structural framework <strong>of</strong> the bio<strong>membrane</strong> is the lipid bilayer. The<br />

lipid bilayer is formed by spontaneous self-assembly <strong>of</strong> lipid molecules. During this<br />

process, the decrease in entropy resulting from hydrocarbon-water interaction, which is<br />

also called the hydrophobic effect (Tanford, 1980), acts to organize lipid molecules so<br />

that acyl-chains are screened from the water environment. This can also lead to the<br />

formation <strong>of</strong> micelles or other types <strong>of</strong> organization depending on the structure <strong>of</strong> the<br />

lipid, as will be discussed in chapter 1.4.<br />

Due to the amphipatic nature <strong>of</strong> most lipids, the lipid bilayer presents a hydrocarbon<br />

environment in the interior core screened from water molecules by the polar groups <strong>of</strong><br />

lipids (see Fig. I.1). The hydrophobic acyl-chains are in a fluid state whereas the polar<br />

groups <strong>of</strong> the lipids are assembled in an orderly array as in a liquid crystal. This<br />

hydrophobic core <strong>of</strong> the lipid bilayer is the most important structural feature in the role<br />

<strong>of</strong> bio<strong>membrane</strong>s as barriers to passive molecular diffusion. It allows the passage <strong>of</strong><br />

water and other small uncharged molecules but presents great impediment to passive<br />

1


ion diffusion, as the energy required to move a hydrated ion through it, is extremely<br />

high (for a monovalent ion <strong>with</strong> 2 Å radius, it takes about 40 Kcal/mol to transfer it into<br />

the <strong>membrane</strong> core – Gennis, 1989).<br />

Figure I.1 – a) Electron micrograph <strong>of</strong> a section from the plasma <strong>membrane</strong> <strong>of</strong> a erythrocyte <strong>membrane</strong>.<br />

b) Schematic depiction <strong>of</strong> a lamellar arrangement <strong>of</strong> a lipid bilayer. Polar headgroups face outwards and<br />

shield the interior hydrophobic tails (from Lodish et al., 2000).<br />

The bilayer can be regarded as a two-dimensional solvent that provides the<br />

hydrophobic anchor to <strong>membrane</strong> <strong>proteins</strong> in bio<strong>membrane</strong>s. This two-dimensional<br />

fluid character <strong>of</strong> the lipid bilayer is central to lipid-lipid, lipid-protein and proteinprotein<br />

interactions. It is responsible for an increase <strong>of</strong> the efficiency <strong>of</strong> <strong>membrane</strong>bound<br />

enzymes whose substrates are located in the plane <strong>of</strong> the bilayer. The average<br />

time for a reactant to reach a target site depends drastically on the dimensionality.<br />

Reactions that require collisions <strong>of</strong> three bodies are very rare in three dimensions, but<br />

can be very effective in two dimensions (Hardt, 1979; Sackmann, 1995).<br />

The bio<strong>membrane</strong> is not only a physical barrier between the inside and outside <strong>of</strong><br />

the cell or between intracellular compartments and the cytosol. As the cell needs to get<br />

nutrients in and out, the <strong>membrane</strong> must accommodate this. Some <strong>membrane</strong>-bound<br />

2


INTRODUCTION: BIOMEMBRANES<br />

<strong>proteins</strong> are responsible by the regulation <strong>of</strong> the transport <strong>of</strong> ions and molecules across<br />

the <strong>membrane</strong>. Thus, the bio<strong>membrane</strong> is actually a selective barrier. To this effect, the<br />

distributions <strong>of</strong> both lipids and <strong>proteins</strong> exhibit asymmetry in each side <strong>of</strong> the bilayer.<br />

Another consequence <strong>of</strong> the fluid character <strong>of</strong> the lipid bilayer is its high flexibility<br />

that allows for an easy adaptation to exterior perturbations. In bio<strong>membrane</strong>s, the lipid<br />

bilayer also interacts <strong>with</strong> the cytoskeleton. This interaction is responsible for the<br />

mechanical stability <strong>of</strong> cells that combined <strong>with</strong> the flexibility <strong>of</strong> the lipid bilayer<br />

confers the bio<strong>membrane</strong> <strong>with</strong> unique mechanical properties.<br />

1.2. Molecular composition <strong>of</strong> bio<strong>membrane</strong>s<br />

1.2.1. Lipid composition<br />

Bio<strong>membrane</strong>s are typically composed by four classes <strong>of</strong> lipids: glycerolipids,<br />

sphingolipids, glycolipids and sterols. Their common characteristic is the amphipatic<br />

character which determines the lipid bilayer structure.<br />

1.2.1.1. Glycerolipids<br />

The hydrophobic section <strong>of</strong> glycerolipids is composed <strong>of</strong> linear hydrocarbon chains<br />

esterified to sn-1 and sn-2 (stereospecific numbers) positions <strong>of</strong> a glycerol moiety.<br />

These hydrocarbon chains exhibit great diversity <strong>of</strong> length (commonly from 14 to 24<br />

carbon atoms) and unsaturation (from 0 to 4) in bio<strong>membrane</strong>s. A list <strong>of</strong> some <strong>of</strong> the<br />

most common acyl-chains and their nomenclature is shown in Table I.1. A short<br />

notation normally used for describing them is based on the number <strong>of</strong> carbons and<br />

double bonds. Thus 18:1 stands for an acyl-chain <strong>with</strong> 18 carbons in the hydrocarbon<br />

chain and one double bond. In nature, virtually all double bonds in fatty-acids present a<br />

cis configuration and generally the longer the hydrocarbon chain, the more double<br />

bounds are present. The cis configuration has drastic consequences in the packing <strong>of</strong><br />

lipids (Berg et al., 2002) as will be discussed later. The number <strong>of</strong> carbons in the<br />

hydrophobic chain is nearly always even. In bio<strong>membrane</strong>s, lipids generally contain two<br />

different acyl-chains and frequently one <strong>of</strong> them is unsaturated (Mouritsen, 2005).<br />

The predominant class <strong>of</strong> lipids in bio<strong>membrane</strong>s are the phosphoglycerolipids.<br />

These molecules are glycerolipids in which the sn-3 position is esterified to phosphoric<br />

3


acid. If no more groups are linked to the molecule, the resulting compound is<br />

phospatidic acid (PA), the simplest phosphoglycerolipid. In general, the phosphate<br />

group is linked to an alcohol, and together, the phosphate group and the bound alcohol<br />

shape the polar headgroup section <strong>of</strong> the molecule. Phosphoglycerolipids are classified<br />

according to the nature <strong>of</strong> the alcohol moiety in the headgroup. The most common<br />

alcohols found in phospholipids are serine, ethanolamine, choline, glycerol, and inositol<br />

(Figure I.2). The corresponding phosphoglycerolipids are named phosphatidylserine<br />

(PS), phosphatidylethanolamine (PE), phosphatidylcholine (PC), phosphatidylglycerol<br />

(PG), and phosphatidylinositol (PI). The latter can also be phosphorylated at three<br />

different positions in the inositol ring resulting in seven different combinations that<br />

correspond to seven different species as will be discussed in chapter VI.<br />

Table I.1 - Nomenclature <strong>of</strong> relevant acyl-chains<br />

Short notation<br />

Name<br />

14:0 Myristoyl<br />

14:1 (9-cis) Myristoleoyl<br />

16:0 Palmitoyl<br />

16:1 (9-cis) Palmitoleoyl<br />

18:0 Stearoyl<br />

18:1 (9-cis) Oleoyl<br />

18:2 (9,12 cis) Linoleoyl<br />

18:3 (6,9,12–cis) γ-Linolenoyl<br />

18:3 (9,12,15–cis) α-Linolenoyl<br />

20:0 Arachidoyl<br />

20:4 (5,8,11,14–cis) Arachidonoyl<br />

22:0 Behenoyl<br />

22:1 (13-cis) Erucoyl<br />

The phosphate group in phosphoglycerolipids is always negatively charged, and the<br />

net charge <strong>of</strong> the molecule is dictated by the charge <strong>of</strong> the alcohol moiety. Thus PA, PG,<br />

PS, PI, and phosphorylated PI are negatively charged, while PC and PE present a net<br />

neutral charge due to the positively charged amine in choline and ethanolamine. PC is<br />

the most abundant phosphoglycerolipid in the plasma <strong>membrane</strong>, whereas PE is in<br />

general the major component in bacterial <strong>membrane</strong>s.<br />

4


INTRODUCTION: BIOMEMBRANES<br />

Figure I.2 – a) Schematic representation <strong>of</strong> the phosphoglycerolipid structure. b) Most common alcohols<br />

found in the headgroups <strong>of</strong> phosphoglycerolipids. From Berg et al., 2002). c) Structure <strong>of</strong> a<br />

phosphatidylcholine - 1,2-dimyristoyl-sn-glycero-3-phosphatidylcholine (DMPC).<br />

The acyl-chains found in each class <strong>of</strong> phosphoglycerolipids is also different.<br />

PC’s are mainly composed <strong>of</strong> short saturated chains (16 to 18 carbons) or 18:1 and 18:2<br />

unsaturated chains. PE’s have a large fraction <strong>of</strong> polyunsaturated chains, in particular<br />

20:4 (arachidonoyl). Charged phosphoglycerolipids also present a significant content <strong>of</strong><br />

unsaturated acyl-chains (Sackmann, 1995).<br />

Other types <strong>of</strong> glycerolipids are found in bio<strong>membrane</strong>s, namely the<br />

plasmalogens, in which one <strong>of</strong> the acyl-chains is bound to glycerol by a vinyl ether<br />

linkage. Plasmalogens constitute about 20 % <strong>of</strong> the total content <strong>of</strong> phosphoglycerides<br />

in humans, although their abundance varies dramatically among tissues and species.<br />

Human brain and heart tissues are particularly enriched in this class <strong>of</strong> lipids (Lodish et<br />

al., 2000). Cardiolipin is a dimer lipid since it contains four acyl chains. It constitutes<br />

about 20 % <strong>of</strong> the inner mitochondrial <strong>membrane</strong> and is also found in the <strong>membrane</strong> <strong>of</strong><br />

plant chloroplasts and <strong>of</strong> certain bacterias.<br />

1.2.1.2. Sphingolipids<br />

Sphingolipids are based on sphingosine instead <strong>of</strong> glycerol. Sphingosine is an<br />

aminoalcohol <strong>with</strong> a long unsaturated hydrocarbon chain. A long fatty-acid is attached<br />

to sphingosine by an amide bound. This basic structure is a ceramide, which is involved<br />

in cell death and is an essential component <strong>of</strong> skin (Mouritsen, 2005). The linkage <strong>of</strong> a<br />

choline to the terminal hydroxyl group <strong>of</strong> sphingosine in ceramide leads to<br />

5


sphingomyelin, the most abundant sphingolipid, which is a significant component <strong>of</strong><br />

animal plasma <strong>membrane</strong>s.<br />

1.2.1.3. Glycolipids<br />

Both glycerolipids and sphingolipids can be linked to a carbohydrate in their<br />

headgroups, giving rise to a glycolipid. Glycolipids play important roles in the<br />

interactions <strong>of</strong> the cell <strong>with</strong> its surroundings. Glycoglycerolipids are very abundant in<br />

chloroplast <strong>membrane</strong>s, as well as in blue algae and bacteria. They are however rarely<br />

found in animals (Gennis, 1989). In glycosphingolipids, one or more carbohydrates are<br />

linked to the terminal hydroxyl group <strong>of</strong> the sphingosine backbone. The simplest form<br />

<strong>of</strong> glycosphingolipid is a cerebroside that contains a single carbohydrate residue, either<br />

glucose or galactose. Complex glycosphingolipids called gangliosides present one or<br />

two branched carbohydrate chains containing sialic acid groups. Glycosphingolipids are<br />

as a rule located in the outer surface <strong>of</strong> the plasma <strong>membrane</strong> <strong>with</strong> varying<br />

concentrations (2-10 %), and are particularly abundant in the nervous system.<br />

1.2.1.4. Sterols<br />

The basic structure <strong>of</strong> sterols is a four ring hydrocarbon. This structure confers<br />

the molecule much more rigidity than other lipids. The most important sterol in animal<br />

<strong>membrane</strong>s is cholesterol (Fig. I.3). In cholesterol, a hydrocarbon chain is linked to one<br />

end <strong>of</strong> the ring system and a hydroxyl group is attached to the other end <strong>of</strong> the structure,<br />

this being the polar headgroup <strong>of</strong> the molecule. In the lipid bilayer the molecule is<br />

oriented parallel to the fatty-acid chains <strong>of</strong> other lipids while the hydroxyl group<br />

interacts <strong>with</strong> nearby headgroups. Cholesterol is only found in eukaryotes, particularly<br />

in animal plasma <strong>membrane</strong>s, lysosomes, endosomes and Golgi apparatus. The<br />

concentration <strong>of</strong> cholesterol is especially high in the animal plasma <strong>membrane</strong> (20 - 30<br />

%). In plants cholesterol exists in low amounts, and other sterols are present, namely<br />

sitoesterol and stigmasterol. Ergosterol is found in yeast and other eukaryotic<br />

microorganisms.<br />

6


INTRODUCTION: BIOMEMBRANES<br />

Figure 1.3 – Structure <strong>of</strong> cholesterol (From Berg et al., 2002).<br />

1.3. Lipid asymmetry across the bilayer<br />

Lipid composition in the cytoplasmatic and exoplasmatic leaflets <strong>of</strong> the bilayer is<br />

asymmetric. In plasma <strong>membrane</strong>s, the cytoplasmatic leaflet is enriched in PE, PS, and<br />

PI, while in the exoplasmatic leaflet lipids like PC, glycolipids, sphingomyelin and<br />

cholesterol exist in higher concentrations.<br />

One <strong>of</strong> the factors contributing to this lipid asymmetry is the site <strong>of</strong> synthesis <strong>of</strong> the<br />

lipids in the <strong>membrane</strong>s <strong>of</strong> the endoplasmic reticulum and Golgi, as it dictates the<br />

plasma <strong>membrane</strong> leaflet where the lipid is to be inserted. In this way, addition <strong>of</strong><br />

carbohydrates to glycolipids is done through the luminal side <strong>of</strong> the Golgi apparatus that<br />

is topologically equivalent to the exterior <strong>of</strong> the cell. This however does not account for<br />

all asymmetry observed and the action <strong>of</strong> ATP-powered transport <strong>proteins</strong> called<br />

translocases is essential (Lodish et al., 2000).<br />

However, asymmetry would be lost if lipids easily crossed from one leaflet to the<br />

other. In fact, in protein free liposomes, the kinetics <strong>of</strong> this process is extremely slow,<br />

on the order <strong>of</strong> hours/days. This is due to the extremely energetically unfavourable<br />

process <strong>of</strong> inserting the polar headgroup <strong>of</strong> a lipid in the hydrophobic <strong>membrane</strong><br />

interior.<br />

Lipid asymmetry is <strong>of</strong> functional importance, and many cytosolic <strong>proteins</strong> bind to<br />

specific lipid headgroups (like PS or PI) in the cytosolic leaflet <strong>of</strong> the bilayer. Animals<br />

also use phospholipid asymmetry in the plasma <strong>membrane</strong> <strong>of</strong> cells as a control to<br />

distinguish between dead and healthy cells, as when animal cells undergo programmed<br />

cell death, or apoptosis. In this case, PS lipids that are normally found enriched in the<br />

cytosolic leaflet, distribute between both leaflets <strong>of</strong> the bilayer. This functions as a<br />

7


signal for neighbouring cells to phagocytose the dead cell and digest it (Alberts et al.,<br />

2002).<br />

1.4. Lipid structure and curvature<br />

Lipids in hydrated conditions exhibit polymorphic behaviour as they can assemble<br />

in different structures. The structure adopted by a lipid aggregate can be influenced by<br />

the molecular structure <strong>of</strong> the lipid itself and a myriad <strong>of</strong> environmental conditions,<br />

such as water content, pH, ionic strength, temperature and pressure. Lipid molecules can<br />

assemble in an aqueous environment either in a lamellar structure (as in a lipid bilayer)<br />

or in non-lamellar phases (micelle, hexagonal, inverted hexagonal phases, and the cubic<br />

phase (see Fig. I.4)). Because phase transitions between these different aggregates can<br />

be activated by changes in water content, these phases are called lyotropic.<br />

The different possibilities for organizations <strong>of</strong> lipids are the result <strong>of</strong> the intrinsic<br />

shape <strong>of</strong> the lipid molecule. When the headgroup <strong>of</strong> a lipid occupies the same area than<br />

the area occupied by its hydrophobic section, the leaflets or monolayers formed by<br />

association <strong>of</strong> molecules <strong>of</strong> this lipid will present zero spontaneous curvature, forming a<br />

planar <strong>membrane</strong> for which both leaflets (or monolayers) present null curvature. A lipid<br />

bilayer composed <strong>of</strong> this lipid <strong>with</strong> the same number <strong>of</strong> molecules in each monolayer<br />

should be flat. However, in order to eliminate the aqueous exposure <strong>of</strong> the hydrophobic<br />

edges <strong>of</strong> the bilayer, the bilayer can unite the edges and form a curved vesicle. In this<br />

aggregate the exterior monolayer must present a concave (or positive) curvature and the<br />

interior monolayer must present a curvature <strong>with</strong> opposite direction. In this way,<br />

spontaneous curvature <strong>of</strong> a lipid monolayer refers to the curvature observed in the<br />

absence <strong>of</strong> edge conditions (Zimmerberg, 2000).<br />

For monolayers composed <strong>of</strong> lipids presenting headgroups <strong>with</strong> a cross-section<br />

different to that <strong>of</strong> the hydrophobic tails, a spontaneous curvature will be present, and<br />

the packing <strong>of</strong> these lipids in a lipid bilayer <strong>with</strong> a lamellar structure will result in<br />

curvature stress, that can be supported by the lamellar structure only up to a certain<br />

extent. In case the lipid bilayer cannot sustain this curvature stress, the lamellar<br />

structure (L) will be broken and non-lamellar phases will arise. The basic structural<br />

phase <strong>of</strong> biological <strong>membrane</strong>s is nevertheless a lamellar phospholipid bilayer matrix,<br />

and deviations from a lamellar arrangement are generally not desirable in the plasma<br />

<strong>membrane</strong>. The inclusion <strong>of</strong> lipids <strong>with</strong> propensity to non-lamellar structure leads to a<br />

8


INTRODUCTION: BIOMEMBRANES<br />

curvature stress in the <strong>membrane</strong>, and local modulation <strong>of</strong> this stress can be achieved<br />

either by enzymatic action in the lipids, as removal <strong>of</strong> headgroups (Nieva et al., 1993) or<br />

a acyl-chain, or by protein binding, allowing local formation <strong>of</strong> non-lamellar phases.<br />

Micellar<br />

phase<br />

Hexagonal<br />

phase<br />

Lamellar<br />

phase<br />

Cubic<br />

phase<br />

Inverted<br />

hexagonal<br />

phase<br />

Figure I.4 - Representation <strong>of</strong> the most important forms <strong>of</strong> organization <strong>of</strong> lipid aggregates in an aqueous<br />

environment. Different lipid structures correspond to different lipid curvatures and are arranged in<br />

accordance to the value <strong>of</strong> the packing parameter P. Adapted from (Mouritsen, 2005) and (Seddon and<br />

Templer, 1995).<br />

The ability <strong>of</strong> a lipid to fit into a particular type <strong>of</strong> aggregate can be described by a<br />

packing parameter P, which is defined in Figure I.4. A deviation <strong>of</strong> P from 1, either<br />

positive or negative, corresponds to a deviation <strong>of</strong> the lipid shape when in the lipid<br />

aggregate from a cylindrical shape, which is the shape that fits a planar bilayer structure.<br />

Negative deviations (P < 1) correspond to hexagonal (H I ) structures and in the limit P <<br />

1/3 to micelles. Highly phosphorylated PI can induce micelles as will be discussed in<br />

chapter IV. Positive deviations (P > 1) correspond to cubic and inverted hexagonal (H II )<br />

aggregates. These two structures are <strong>of</strong> significant biological relevance, playing a<br />

9


functional role in process such as endo- and exocytosis, <strong>membrane</strong> recycling, protein<br />

trafficking, fat digestion, <strong>membrane</strong> budding and fusion. PE lipids show propensity to<br />

arrange into this structures (Mouritsen, 2005). The cubic phase, which consists <strong>of</strong> short<br />

tubes connected in a hexagonal array, is sometimes found on mixtures <strong>of</strong> lipids that are<br />

in transition from a lamellar phase to an inverted hexagonal phase. The inverted<br />

hexagonal phase consists <strong>of</strong> hexagonally packed water cylinders <strong>with</strong> an outer lining <strong>of</strong><br />

lipid molecules oriented <strong>with</strong> their acyl-chains away from the aqueous cylinders<br />

(Yeagle, 1993).<br />

1.5. Lamellar phase transitions in lipid bilayers<br />

Apart from transitions between different lipid morphologies, lipids also experience<br />

phase transitions <strong>with</strong>out drastic changes in morphology. In a lamellar symmetry, lipids<br />

can experience several packing conditions that correspond to different lipid phases.<br />

These different structures are classified as distinct phases because upon crossing the<br />

transition temperature (or other thermodynamic variable), several physical properties <strong>of</strong><br />

the lipid bilayer change abruptly, including the heat capacity, lateral diffusion,<br />

permeability, thickness, area, vesicle shape, etc.<br />

In Figure I.5, differential scanning calorimetry (DSC) pr<strong>of</strong>iles <strong>of</strong> three<br />

phospholipids are presented. In these, the specific heat <strong>of</strong> the phospholipids is shown as<br />

a function <strong>of</strong> temperature. For dimyristoyl-sn-glycero-3-phosphatidylcholine (DMPC)<br />

two peaks are clearly visible, and each peak corresponds to a phase transition. From<br />

this, it is clear that DMPC can exist as three different phases between 10 and 30 ºC. All<br />

other PC lipids present identical behaviour. For the PE and PA lipids the first transition<br />

is absent, the main transition (T m ) is however generally common to all phospholipids.<br />

This transition is considered to have first order characteristics (Gennis, 1989).<br />

Below the main transition, lipids are tightly packed, the acyl chains are ordered and<br />

extended in an all-trans chain configuration, while the molecules are arranged in a<br />

regular lattice as in a crystalline solid (Figure I.5). This phase is for that reason called<br />

solid-ordered or gel phase (s 0 ). In the gel phase, individual molecules diffuse slowly in<br />

the plane <strong>of</strong> the <strong>membrane</strong>, and the lateral diffusion coefficient (D) <strong>of</strong> a phospholipid<br />

molecule is on the order <strong>of</strong> 10 -10 cm 2 s -1 . Lipids <strong>with</strong> large headgroups as PC must also<br />

tilt in the plane <strong>of</strong> the bilayer while in the gel state (Figure I.5). The cross-sectional area<br />

<strong>of</strong> the headgroup <strong>of</strong> these lipids is larger than the cross-sectional area <strong>of</strong> the acyl-chains,<br />

10


INTRODUCTION: BIOMEMBRANES<br />

and the only way to achieve effective packing is through a tilt <strong>of</strong> the hydrocarbon<br />

chains.<br />

Above the T m <strong>of</strong> the lipid, the acyl chain ordering characteristic <strong>of</strong> the gel phase is<br />

lost (gauche conformations are favoured), and in this new phase, the lateral diffusion <strong>of</strong><br />

the lipid molecules is substantially higher (D is on the order <strong>of</strong> 10 -8 cm 2 s -1 ). This phase<br />

is called fluid or liquid-crystalline phase (L α ). As a result <strong>of</strong> the smaller number <strong>of</strong> trans<br />

conformations in the acyl-chains, bilayers in the fluid phase are considerably thinner<br />

than in the gel phase (up to 6 Å in dipalmitoyl-sn-glycero-3-phosphatidylcholine-DPPC<br />

bilayers), but the cross-sectional area increases dramatically (from 52 Å to 71 Å in<br />

DPPC individual molecules) (Nagle and Tristram-Nagle, 2000). This increase in area is<br />

a consequence <strong>of</strong> the weaker van der Waals attractive interactions between acyl-chains<br />

in the fluid phase (Gennis, 1989). As a rule, a strong coupling exists between the lipid<br />

phases in each monolayer (Bagatolli and Gratton, 2000), exceptions however have<br />

already been observed (Devaux and Morris, 2004).<br />

The acyl-chains <strong>of</strong> a phospholipid dictate to a large extent the stability <strong>of</strong> each <strong>of</strong> the<br />

gel and fluid phases, and as a result they define the T m <strong>of</strong> the lipid. This is due to the<br />

importance <strong>of</strong> the van der Waals interactions in stabilizing the gel phase. Long acylchains<br />

permit stronger van der Waals attractive forces, stabilizing the gel phase and<br />

increasing the lipid T m . Unsaturated chains prevent effective ordered packing into the<br />

gel state, inducing a decrease in lipid T m . As most lipids in bio<strong>membrane</strong>s present<br />

unsaturations, lipid <strong>membrane</strong>s are highly fluid. Headgroup structure can also influence<br />

the transition temperature <strong>of</strong> the lipid. PE lipids present higher T m than PC due to<br />

stabilizing hydrogen bonding in the gel phase. pH, ionic strength and pressure can also<br />

influence T m .<br />

PC lipids, as seen in Figure I.5, experience an additional phase transition. This<br />

transition is called pretransition and occurs between two different gel states, L β’ , and a<br />

ripple or periodic gel phase (P β’ ), at a characteristic temperature (T P ). Due to the<br />

presence <strong>of</strong> the planar ring structure, cholesterol disrupts the packing <strong>of</strong> lipids when<br />

mixed <strong>with</strong> lipids in the gel phase. For lipids in the fluid phase, cholesterol has an<br />

ordering influence (Ipsen et al., 1987). The mixture <strong>of</strong> cholesterol and saturated<br />

phospholipids can give rise to an additional phase called liquid ordered phase (L o ). In<br />

the L o phase, the acyl-chains present a higher degree <strong>of</strong> ordering as compared to the one<br />

11


in the L α phase, while lateral diffusion in the plane <strong>of</strong> the bilayer and freedom <strong>of</strong><br />

rotation remains high.<br />

Figure I.5 – A – Differential scanning calorimetry pr<strong>of</strong>ile <strong>of</strong> DMPC, 1,2-dimyristoyl-sn-glycero-3-<br />

phosphatidylethanolamine (DMPE), and 1,2-dimyristoyl-sn-glycero-3-phosphatidic acid (DMPA). B –<br />

Depiction <strong>of</strong> most common lamellar lipid phases: the fluid L α phase, the gel L β and the ripple P β’ phase<br />

(taken from Gennis, 1989).<br />

1.6. Membrane thickness<br />

Due to the complex chemical structure <strong>of</strong> lipids, the chemical environment along the<br />

normal axis <strong>of</strong> the bilayer is highly stratified. In hydrated bilayers, fluctuations add<br />

increased complexity to the task <strong>of</strong> the structural description <strong>of</strong> the lipid <strong>membrane</strong>. A<br />

description for the position <strong>of</strong> the atoms inside the bilayer using a broad statistical<br />

distribution function can be used. Figure I.6 is a representation <strong>of</strong> the liquidcrystallographic<br />

structure <strong>of</strong> L α -phase dioleoyl-sn-glycero-3-phosphatidylcholine<br />

(DOPC) bilayers along the axis <strong>of</strong> the bilayer normal.<br />

In the presence <strong>of</strong> this dynamical concept for the lipid bilayer structure it becomes<br />

difficult to define certain bilayer properties such as <strong>membrane</strong> thickness. It is reasonable<br />

nevertheless to assume the bilayer thickness as equal to the distance between the peaks<br />

12


INTRODUCTION: BIOMEMBRANES<br />

<strong>of</strong> the distribution functions <strong>of</strong> phosphate groups in each monolayer. In this way, the<br />

bilayer thickness <strong>of</strong> fluid DOPC is 37 Å. Another important concept is the hydrophobic<br />

thickness <strong>of</strong> a bilayer, i.e. the length across the bilayer normal which is occupied by the<br />

hydrocarbon core <strong>of</strong> lipids. Wiener and White (1992) showed that the transbilayer<br />

carbonyl-to-carbonyl distance accurately reported the thickness <strong>of</strong> the hydrocarbon core.<br />

For fluid DOPC, this value is <strong>of</strong> 27-28 Å (Lewis and Engelman, 1983; Nagle and<br />

Tristram-Nagle, 2000). Interestingly, the presence <strong>of</strong> the double bond in DOPC, which<br />

carries 18 carbons in the acyl-chains, decreases the hydrophobic thickness up to almost<br />

the same value <strong>of</strong> DMPC, which carries only 14 carbons but presents no unsaturation.<br />

As a general rule, the hydrophobic thickness <strong>of</strong> a phospholipid bilayer increases 1,75 Å<br />

per additional carbon in the phospholipid acyl-chain (Lewis and Engelman, 1983).<br />

Figure 1.6 - Crystallographic structure <strong>of</strong> fluid DOPC bilayers along the bilayer normal axis (Wiener and<br />

White, 1992).<br />

Apart from the phospholipid structure, other factors can influence significantly the<br />

thickness <strong>of</strong> a bilayer. Increases in temperature and hydration lead to decreases in the<br />

thickness <strong>of</strong> the bilayer. Cholesterol also has an important impact as it acts by<br />

thickening the <strong>membrane</strong> (Mouritsen, 2005).<br />

13


1.7. Lateral heterogeneity in lipid bilayer<br />

As already pointed out, the lipid composition <strong>of</strong> bio<strong>membrane</strong>s is highly complex.<br />

In the case <strong>of</strong> multicomponent lipid bilayers, no longer a single T m applies to the whole<br />

population <strong>of</strong> lipid molecules. Now the transition takes place over a range <strong>of</strong><br />

temperatures (between the T m <strong>of</strong> the lipid <strong>with</strong> the lower transition temperature and the<br />

T m <strong>of</strong> the lipid <strong>with</strong> the higher transition temperature), in which the lipids phase separate<br />

in gel and fluid phases (Mouritsen, 2005). This phase separation is a direct consequence<br />

<strong>of</strong> the highly cooperative character <strong>of</strong> lipid phase. Immiscibility between lipids can<br />

result from differences in acyl-chain length, unsaturation or headgroup composition<br />

(Shimshick and McConnell,1973; Arnold et al., 1981; Lentz et al., 1976).<br />

Phase diagrams provide an efficient tool for the description <strong>of</strong> phase behaviour <strong>of</strong><br />

multicomponent lipid mixtures. In the phase diagrams, the lines correspond to the<br />

temperatures at which a transition starts or is finished. Three phase diagrams for three<br />

different binary lipid mixtures are presented in Figure I.7. It is clear from Figure I.7 that<br />

as the divergence in acyl-chain length increases, the size <strong>of</strong> the region <strong>of</strong> gel-fluid phase<br />

coexistence also increases due to a more accentuated immiscibility between the two<br />

lipids.<br />

Figure I.7 – Phase diagrams <strong>of</strong> three different binary mixtures <strong>of</strong> saturated PC lipids presenting different<br />

acyl-chain lengths – dilauroyl-sn-glycerol-3-phosphatidylcholine-DLPC (12 carbons); DMPC (14<br />

carbons); DPPC (16 carbons) and distearoyl-sn-glycerol-3- phosphatidylcholine -DSPC (18 carbons). The<br />

fluid phase is represented by an f and the gel phase by g (taken from Mouritsen, 2005).<br />

14


INTRODUCTION: BIOMEMBRANES<br />

Lipid immiscibility gives rise to lateral structuring <strong>of</strong> the lipid bilayer and the<br />

resulting lateral organizations are called domains. Lipid domains are not static<br />

structures. They exhibit dynamics both in space and time, as they can exist as short<br />

(fluctuations) or long-lived structures in the lipid bilayer. Large size domains (in the<br />

order <strong>of</strong> micrometers) have already been detected through imaging techniques (Korlach<br />

et al., 1999; Bagatolli and Gratton, 2000). These techniques make use <strong>of</strong> fluorescent<br />

probes that show preferential partition in one <strong>of</strong> the lipid phases or differential<br />

fluorescent properties in each phase (Klausner and Wolf, 1980; Bagatolli and Gratton,<br />

2000). They are however restricted by the diffraction limit to detect domains larger than<br />

~ 200 nm. An example is shown in Figure I.8.<br />

Figure I.8 - Gel-Fluid coexistence in a giant unilamellar vesicle (GUV) (see 1.10) detected by<br />

confocal microscopy. The system is a DLPC/DSPC (0.4/0.6 mol/mol) mixture. Red areas correspond to<br />

the fluorescence arising from a probe <strong>with</strong> preferential partition to the gel phase (enriched in DSPC) and<br />

green areas correspond to the fluorescence <strong>of</strong> a probe <strong>with</strong> preferential partition to the fluid phase<br />

(enriched in DLPC) (taken from Korlach et al., 1999).<br />

For a situation <strong>of</strong> thermodynamical equilibrium, one should expect complete phase<br />

separation, i.e., only two very large domains inside the vesicle, one in the gel phase, and<br />

the other in the fluid phase. This is due to the interfacial tension between lipid domains.<br />

In the interface, the phases show differences in hydrophobic thickness, and the<br />

consequent exposure <strong>of</strong> the hydrocarbon chains to water molecules create a packing<br />

stress in the interface <strong>of</strong> domains. This tension is the driving force for the fusion <strong>of</strong><br />

small domains into larger lateral structures after phase separation, as larger domains<br />

correspond to a smaller interface area. At infinite time, the complete macroscopic phase<br />

separation should be achieved. However, several domains <strong>of</strong> limited size are observed<br />

(Figure I.8). This is possibly due to the coupling <strong>of</strong> phase separation to the curvature <strong>of</strong><br />

15


the vesicle through bilayer deformation in the interface and to the insertion <strong>of</strong><br />

phospholipids in the interface region <strong>with</strong> intermediate conformations (between gel and<br />

fluid like conformations) in order to substantially decrease the interface stress between<br />

lipid domains (Jorgensen and Mouritsen, 1995). Another possibility is that the finite<br />

size domains correspond to non-equilibrium structures, detected due to the very long<br />

lifetime <strong>of</strong> the process <strong>of</strong> phase separation (de Almeida et al., 2002).<br />

Even when at the same lamellar phase, lipids do not mix ideally. This is particularly<br />

true for the gel phase, as the strict packing restrictions force dissimilar lipids to<br />

immiscibility. As a result, clustering or domain formation is observed and the bilayer<br />

gains lateral heterogeneity. This type <strong>of</strong> behaviour is observed e.g. for the highly<br />

nonideal mixture DLPC-DSPC (Fig. I.7), as two different gel phases (each enriched in<br />

one <strong>of</strong> the lipids) are present below the phase coexistence region.<br />

Deviations from ideal lipid mixing and domain formation also occur in the fluid<br />

phase, and PC lipids <strong>with</strong> a significantly different thickness were already shown to<br />

segregate into domains due to hydrophobic mismatch stress. Also for this situation, the<br />

discrepancy <strong>of</strong> hydrophobic thickness <strong>of</strong> the two PC lipids can create a packing stress in<br />

the bilayer due to exposure <strong>of</strong> hydrocarbons to water, and stimulate lipid segregation<br />

(Lehtonen et al., 1996). The lateral structures created by this type <strong>of</strong> phase separation<br />

are expected to be on the nanometer scale in opposition to gel-fluid phase separation<br />

which can be detected in the micrometer scale. Therefore, these domains are elusive to<br />

most <strong>of</strong> the imaging techniques, and other spectroscopic techniques must be applied,<br />

namely macroscopic fluorescence methodologies.<br />

The ideality degree <strong>of</strong> a lipid mixture dictates not only the phase separation <strong>of</strong> the<br />

mixture but also the size and shape <strong>of</strong> the phase-separated domains. Apart from<br />

temperature, other factors influencing the lateral structuring <strong>of</strong> a multicomponent lipid<br />

bilayer are dehydration and the presence <strong>of</strong> divalent ions. Dehydration can cause<br />

lamellar to inverted hexagonal phase transition and induce phase separation (Webb et<br />

al., 1993). Mixtures containing an anionic phospholipid can experience severe phase<br />

separation in the presence <strong>of</strong> divalent ions like Ca 2+ . Calcium ions induce clustering and<br />

phase separation <strong>of</strong> anionic phospholipids due to intermolecular cross bridges (Silvius,<br />

1990).<br />

A phase coexistence <strong>of</strong> particular biological relevance is that <strong>of</strong> L α and L o phases. In<br />

1997 Simons and Ikonen proposed that small lipid domains in the L o phase (lipid rafts)<br />

composed <strong>of</strong> saturated lipids and cholesterol coexisted <strong>with</strong> a matrix <strong>of</strong> lipids in the<br />

16


INTRODUCTION: BIOMEMBRANES<br />

fluid phase. Since then, several cellular mechanisms have been associated <strong>with</strong> domains<br />

<strong>of</strong> cholesterol and saturated lipids, and several evidences for the existence <strong>of</strong> these<br />

structures have been gathered. There is indication that rafts are involved in cell surface<br />

signalling, intracellular trafficking and sorting <strong>of</strong> lipids and <strong>proteins</strong> (Mouritsen, 2005)<br />

Lateral structuring <strong>of</strong> the lipid bilayer, either through L α -L β , L α -L o or fluid-fluid<br />

domains, implies that <strong>membrane</strong> functions do not require the dependence on random<br />

collisions for interactions between reagents (Mouritsen, 2005). It provides a structuring<br />

principle for lipid bilayer organization, that can permit not only a higher efficiency <strong>of</strong><br />

several <strong>membrane</strong> processes by compartmentalization, but also a mechanism <strong>of</strong><br />

modulating these processes through control <strong>of</strong> the <strong>membrane</strong> lateral structure.<br />

1.8. Membrane <strong>proteins</strong><br />

The lipid bilayer provides the basic architecture <strong>of</strong> the bio<strong>membrane</strong>s. However,<br />

<strong>membrane</strong> <strong>proteins</strong> are responsible for the majority <strong>of</strong> cell functions taking place in the<br />

<strong>membrane</strong>. The distinctive properties and functions <strong>of</strong> different lipid <strong>membrane</strong>s inside<br />

the cell are mainly dictated by their protein content. The concentration <strong>of</strong> <strong>proteins</strong> inside<br />

bio<strong>membrane</strong>s is drastically different among the several cellular <strong>membrane</strong>s. In the<br />

mitochondrial <strong>membrane</strong> the protein content reaches about 76% (w/w), while in the<br />

myelin <strong>membrane</strong> <strong>proteins</strong> are only 18% <strong>of</strong> total <strong>membrane</strong> content. The mapping <strong>of</strong><br />

the yeast genome showed that most <strong>of</strong> the genome code is expected to code for<br />

<strong>membrane</strong> <strong>proteins</strong>, in demonstration <strong>of</strong> their key role for life (Mouritsen, 2005).<br />

Membrane <strong>proteins</strong> are defined by their degree <strong>of</strong> insertion in the lipid bilayer<br />

(Figure I.9). Thus, <strong>membrane</strong> <strong>proteins</strong> deeply buried in the lipid bilayer are called<br />

integral, frequently spanning the <strong>membrane</strong> one or several times, whereas <strong>proteins</strong><br />

bound to the exoplasmic or cytoplasmic periphery <strong>of</strong> the lipid bilayer are entitled<br />

peripheral. Peripheral <strong>membrane</strong> <strong>proteins</strong> are bound to lipid bilayers by interaction<br />

either <strong>with</strong> lipid headgroups or <strong>with</strong> other <strong>membrane</strong> bound <strong>proteins</strong>. Another class <strong>of</strong><br />

<strong>proteins</strong> are bound to the <strong>membrane</strong> through covalent linkage to a lipid molecule. These<br />

<strong>membrane</strong> <strong>proteins</strong> are classified as lipid-anchored.<br />

17


Figure I.9 – Depiction <strong>of</strong> the several possibilities for association <strong>of</strong> <strong>proteins</strong> <strong>with</strong> bio<strong>membrane</strong>s. Protein<br />

domains in the exoplasmic side <strong>of</strong> the <strong>membrane</strong> are frequently glycosilated and can interact <strong>with</strong> the<br />

extracellular matrix. Protein domains on the cytoplasmic side on the other hand are responsible for the<br />

coupling <strong>of</strong> the cytoskeleton network <strong>with</strong> the lipid <strong>membrane</strong> (From Lodish et al., 2000).<br />

Membrane <strong>proteins</strong> on the extracellular side <strong>of</strong> the plasma <strong>membrane</strong> generally bind<br />

other molecules, including external signalling <strong>proteins</strong>, ions, small metabolites, and<br />

adhesion molecules in other cells. Protein domains <strong>with</strong>in the plasma <strong>membrane</strong>, mainly<br />

those that are part <strong>of</strong> channels and pores, are generally responsible for transport <strong>of</strong><br />

molecules across the bilayer, while peripheral protein <strong>membrane</strong>s in the cytosolic side<br />

<strong>of</strong> the plasma <strong>membrane</strong> are responsible for many different functions (Lodish et al.,<br />

2000).<br />

The possibilities <strong>of</strong> secondary structure available for <strong>membrane</strong> spanning protein<br />

domains are very limited. Only two structures are available, the alpha-helix and the<br />

beta-barrel (Figure I.10). These structures have a common characteristic. In both, the<br />

peptide bonds <strong>of</strong> the aminoacids are hydrogen bonded. In fact, if this were not the case,<br />

insertion in the highly apolar environment would be energetically unfavourable, even<br />

for the most hydrophobic amino acids. The energetic cost <strong>of</strong> partitioning a peptide bond<br />

into a highly apolar phase, is significantly larger than the free energy reduction<br />

associated <strong>with</strong> insertion <strong>of</strong> the Trp side chain in the same environment (White and<br />

Wimley, 1999). Therefore, only a polypeptide segment <strong>with</strong> complete backbonebackbone<br />

hydrogen bonding can insert in the highly hydrophobic core <strong>of</strong> the bilayer,<br />

and the alpha-helix and the beta-barrel are the two structures that satisfy this condition.<br />

18


INTRODUCTION: BIOMEMBRANES<br />

A<br />

B<br />

Figure I.10 – Two secondary structures available for <strong>membrane</strong> spanning protein domains, the alphahelix<br />

(A) and the beta-barrel (B).<br />

The amino acids in an alpha-helix are arranged in a right-handed helical structure.<br />

The dimensions <strong>of</strong> the alpha helix are 5,4 Å wide (if accounting for aminoacid side<br />

chains the alpha helix is about 10 Å wide). Each aminoacid correspond to a 100º turn,<br />

and consequently for each turn, 3.6 aminoacids are required. Each aminoacid also<br />

corresponds to a translation across the helical axis <strong>of</strong> 1,5 Å, and each amine group <strong>of</strong> the<br />

peptide bound forms a hydrogen bond <strong>with</strong> the carbonyl group <strong>of</strong> the pepide bound four<br />

aminoacids earlier. Beta barrels are large beta sheet structures for which the last strand<br />

is hydrogen bonded <strong>with</strong> the first, creating in this way a closed structure. These<br />

structures are commonly found in porins (see chapter IV). The beta-sheet structure is<br />

based on the side-by-side hydrogen bonded arrangement <strong>of</strong> stretchs <strong>of</strong> amino acids<br />

linearly extended.<br />

Hydrogen bonding between peptide bonds is essential for formation <strong>of</strong> TM protein<br />

segments, however it is not sufficient. The sum <strong>of</strong> the free energy reduction from<br />

insertion <strong>of</strong> the side chains <strong>of</strong> the polypeptide must surpass the energetic cost <strong>of</strong> burying<br />

the hydrogen bonded peptide bonds. This is only possible in the presence <strong>of</strong> highly<br />

hydrophobic aminoacids, and in the absence <strong>of</strong> to many hydrophilic ones, particularly<br />

charged residues. Octanol is frequently regarded as a suitable model for studying the<br />

insertion <strong>of</strong> molecules in a hydrophobic medium such as the hydrocarbon core <strong>of</strong> the<br />

bilayer. Figure I.11 reproduces the results obtained for the free energy <strong>of</strong> transfer <strong>of</strong><br />

whole aminoacids (including the peptide bond), from water to the octanol phase, and to<br />

the interface region (polar environment) <strong>of</strong> POPC bilayers.<br />

19


Figure I.11 – Experimentally obtained hydrophobicity scales for whole residues (including the peptide<br />

bond). Results obtained for the transfer from water to the polar interface <strong>of</strong> POPC bilayers and from water<br />

to octanol are presented (Wimley and White, 1996; Wimley et al., 1996).<br />

It is clear from Figure 1.11 that residues like Trp (W), Phe (F), Tyr (Y), Leu (L), Ile<br />

(I), Met (M) or Val (V) favour incorporation in hydrocarbon environments, while<br />

charged residues are particularly averse to this. From this information and from the<br />

primary sequence <strong>of</strong> a protein it is possible to make previsions for possible alpha helical<br />

TM segments inside the protein sequence. A calculation is made through summation <strong>of</strong><br />

the free energy required to insert (from water into the bilayer hydrocarbon core)<br />

successive segments <strong>of</strong> the polypeptide chain. Segments tested have a finite size, around<br />

20 aminoacids, as this is just about sufficient to spam the hydrocarbon core <strong>of</strong> a typical<br />

bilayer (this aminoacid length corresponds to 30 Å if the helix is oriented along the<br />

bilayer normal). This procedure allows the creation <strong>of</strong> hydropathy plots that signal the<br />

polypeptide segments which are the most likely candidates for a TM configuration<br />

inside the protein.<br />

Aminoacids also exhibit preferential localization in certain positions along the<br />

bilayer normal. This fact is <strong>of</strong> utmost importance to the final position <strong>of</strong> an alpha-helix<br />

in <strong>membrane</strong>s and will be discussed in more detail in section 2 <strong>of</strong> this chapter.<br />

1.9. Lateral dynamics in bio<strong>membrane</strong>s<br />

The Singer-Nicolson fluid mosaic model for bio<strong>membrane</strong>s (1972) (Figure I.12)<br />

described the lipid <strong>membrane</strong> as a two dimensional liquid where both lipids and<br />

<strong>membrane</strong> <strong>proteins</strong> moved freely. In fact, in pure lipid bilayers in the fluid phase, lipids<br />

experience fast lateral diffusion, as well as rotational diffusion, wobbling, and<br />

movements in the axis <strong>of</strong> the bilayer. Crossing from one monolayer to the other is<br />

20


INTRODUCTION: BIOMEMBRANES<br />

however much more restricted as discussed in 1.3. Lateral dynamics in the gel phase is<br />

approximately 100 times slower (Mouritsen, 2005). The speed <strong>of</strong> integral <strong>membrane</strong><br />

<strong>proteins</strong> is typically 100 times smaller than the speed <strong>of</strong> lipid molecules.<br />

Figure I.12 – The Singer-Nicolson model <strong>of</strong> fluid mosaic (Taken from Singer and Nicolson, 1972)<br />

However, in bio<strong>membrane</strong>s, diffusion <strong>of</strong> lipids and <strong>proteins</strong> is hindered by the<br />

presence <strong>of</strong> compartmentalisations <strong>of</strong> the <strong>membrane</strong> in the form <strong>of</strong> domains.<br />

Additionally a fraction <strong>of</strong> the <strong>membrane</strong> <strong>proteins</strong> is completely immobile due to<br />

interaction <strong>with</strong> the cytoskeleton network. Generally the average diffusion coefficient <strong>of</strong><br />

a protein in the plasma <strong>membrane</strong> <strong>of</strong> intact cells is about 10-30 times smaller than the<br />

one observed in pure lipid liposomes (Mouritsen, 2005). The Singer-Nicolson model is<br />

therefore a simplistic description <strong>of</strong> the complex interactions and dynamics existent in<br />

bio<strong>membrane</strong>s.<br />

1.10. Membrane Model Systems<br />

Cellular <strong>membrane</strong>s are <strong>of</strong>ten too intricate for their properties to be fully resolved as<br />

too many variables are present. A valuable alternative is the use <strong>of</strong> <strong>membrane</strong> model<br />

systems. Model systems grant us a simplified basis for the study <strong>of</strong> bilayer structure and<br />

properties, while at the same time maintaining all the fundamental properties <strong>of</strong> the<br />

bio<strong>membrane</strong>.<br />

Liposomes are the most popular <strong>of</strong> the <strong>membrane</strong> model systems. Preparation is<br />

<strong>of</strong>ten very simple, requiring solubilization <strong>of</strong> lipids in organic solvents, followed by<br />

21


drying into a film and ressuspension in an aqueous environment. Suspensions <strong>of</strong><br />

liposomes prepared in this manner are composed by concentric bilayers or multilamellar<br />

vesicles (MLV’s). Due to this arrangement <strong>of</strong> the lipid vesicles, most <strong>of</strong> the lipid bilayer<br />

surface is buried inside the liposome, as only about 10 % is found in the outside surface<br />

(Yeagle, 1993). The application <strong>of</strong> MLV´s in photophysical <strong>studies</strong> can result in several<br />

difficulties. Due to the size <strong>of</strong> the MLV’s particles, scattering is very significant and for<br />

these reason, MLV´s are seldom used.<br />

There are different strategies for obtaining unilamellar liposomes. By applying<br />

ultrasonic power to suspensions <strong>of</strong> liposomes, unilamellar vesicles <strong>of</strong> small diameter<br />

(~30 Å), also called small unilamellar vesicle (SUV) are obtained. Due to the smaller<br />

size <strong>of</strong> the particles, scattering in suspensions <strong>of</strong> SUV’s is drastically reduced, however<br />

the curvature <strong>of</strong> SUV’s is much higher than the curvature <strong>of</strong> cell <strong>membrane</strong>s, resulting<br />

in poor mimics <strong>of</strong> the properties <strong>of</strong> bio<strong>membrane</strong>s.<br />

Larger unilamellar vesicles (LUV’s) can be produced either by dialysis <strong>of</strong><br />

detergents, reverse phase evaporation, or fast extrusion through polycarbonate filters.<br />

The latter method is particularly useful due to the shorter times required when compared<br />

to the other ones, e.g., dialysis <strong>of</strong> detergents can take several days for efficient removal<br />

<strong>of</strong> detergent from liposomes (for a review, see Gennis, 1989).<br />

Much larger unilamellar vesicles (up to 300 µm) can also be prepared by gentle<br />

hydration and electr<strong>of</strong>ormation (Rodriguez et al., 2005). These vesicles are also known<br />

as giant unilamellar vesicles or GUV’s and are particularly suitable for microscopy<br />

applications as their size enables visualization and micromanipulation.<br />

Still, other types <strong>of</strong> <strong>membrane</strong> model systems exist. Lipid monolayers in water-air<br />

interface allow the study <strong>of</strong> the effect <strong>of</strong> the lateral surface pressure on <strong>membrane</strong><br />

components and on the interactions between them. Black Lipid Membranes have also<br />

proven valuable in the study <strong>of</strong> electrical properties (Yeagle, 1993).<br />

22


INTRODUCTION: LIPID-PROTEIN INTERACTIONS<br />

2.LIPID-PROTEIN INTERACTIONS<br />

2.1. Membrane protein reconstitution<br />

Even for the simplest organisms, <strong>membrane</strong> <strong>proteins</strong> in protein-lipid extracts are<br />

present in a very complex and heterogeneous environment, which can impair the<br />

possibility <strong>of</strong> conducting biophysical <strong>studies</strong> that require a simpler matrix for the<br />

protein (<strong>with</strong> absence <strong>of</strong> possible contaminants), and more controlled conditions.<br />

However, integral <strong>membrane</strong> <strong>proteins</strong> generally cannot be studied in homogeneous<br />

systems such as aqueous or organic solutions due to the complex solubility problems<br />

derived from their bitopic nature. In addition, the study <strong>of</strong> their properties in a isotropic<br />

medium, would not be biologically relevant. Systems are required that satisfy both the<br />

hydrophobicity <strong>of</strong> the <strong>membrane</strong> embedded sections <strong>of</strong> the protein as well as their<br />

hydrophilic domains. Liposomes are frequently used for such <strong>studies</strong>, since, as<br />

discussed in the previous section, they present good mimetic conditions <strong>of</strong><br />

bio<strong>membrane</strong>s. Liposomes containing reconstituted <strong>proteins</strong> are called proteoliposomes.<br />

In order to reconstitute <strong>membrane</strong> <strong>proteins</strong> in liposomes it is first necessary to purify<br />

and solubilize them. Neither <strong>of</strong> these procedures is trivial and detergents have proven<br />

invaluable tools in both. They meet the requirements <strong>of</strong> amphipatic structure necessary<br />

to solubilize the two different environments present in <strong>membrane</strong> <strong>proteins</strong> and these are<br />

frequently soluble in micellar structures. Detergents are classified according to the<br />

charge <strong>of</strong> the headgroup as ionic (cationic or anionic), non-ionic or zwitterionic.<br />

Examples <strong>of</strong> ionic detergents <strong>of</strong>ten used in <strong>membrane</strong> solubilization are sodium<br />

dodecyl sulphate (SDS) and bile acid salts. SDS is a linear chain detergent and a<br />

extremely powerful solubilizing agent for <strong>membrane</strong> <strong>proteins</strong>, however, it is also<br />

frequently denaturing. Protein renaturation is eventually possible under certain<br />

conditions after transfer to other medium (Dong et al., 1997). Bile acid salts are ionic<br />

detergents <strong>with</strong> steroidal groups for backbones. Due to the planar characteristics <strong>of</strong> the<br />

steroidal structure, instead <strong>of</strong> a proper headgroup they present a polar and an apolar<br />

face. These detergents are much weaker than SDS and more efficient in maintaining<br />

protein activity. Examples <strong>of</strong> bile acid salts are sodium cholate and sodium<br />

deoxycholate. Non-ionic detergents are also mild and relatively non-denaturating/non-<br />

23


deactivating agents. Zwitterionic detergents combine the properties <strong>of</strong> ionic and nonionic<br />

detergents (Seddon et al., 2004).<br />

Several methods exist for insertion <strong>of</strong> integral <strong>membrane</strong> <strong>proteins</strong> in liposomes. The<br />

most popular makes use <strong>of</strong> mediation by detergents. The choice <strong>of</strong> the best detergent<br />

varies greatly between <strong>proteins</strong>, confirming the complexity <strong>of</strong> the interactions involved<br />

in the reconstitution process. The simplest <strong>of</strong> the detergent-mediated reconstitution<br />

method is a dilution approach. Liposomes are added to detergent micelles <strong>with</strong> protein<br />

and the detergent is removed. Removal <strong>of</strong> detergent evolves until the concentration <strong>of</strong><br />

the detergent is below the critical micelle concentration (cmc) and micelles become<br />

unstable inducing transfer <strong>of</strong> the protein to the liposomes. Removal <strong>of</strong> detergent can be<br />

achieved by: i) dialysis, making use <strong>of</strong> a <strong>membrane</strong> <strong>with</strong> a pore size smaller than the<br />

size <strong>of</strong> the proteoliposomes; ii) column chromatography; iii) incubation <strong>with</strong> detergent<br />

adsorbing beads. However, pure liposomes are frequently impermeable to <strong>proteins</strong> and<br />

<strong>of</strong>ten a destabilization <strong>of</strong> the bilayer is required for correct insertion <strong>of</strong> the protein<br />

(Rigaud et al., 1995). Destabilization <strong>of</strong> the bilayer can be achieved by using mixtures<br />

<strong>of</strong> lipids and detergents instead <strong>of</strong> pure liposomes. The resulting structures are more<br />

receptive to protein insertion, and after this step the detergent can be removed by one <strong>of</strong><br />

the methods described above.<br />

The required degree <strong>of</strong> detergent-destabilization <strong>of</strong> the liposomes depends on the<br />

nature <strong>of</strong> the detergents. Several <strong>proteins</strong> were reconstituted through mediation <strong>of</strong> nonionic<br />

detergents by destabilization <strong>of</strong> liposomes <strong>with</strong> the concentration <strong>of</strong> the detergent<br />

set at saturation levels for the liposomes. This method frequently leads to asymmetrical<br />

orientation <strong>of</strong> the <strong>proteins</strong> in the liposomes (a large fraction <strong>of</strong> the <strong>proteins</strong> face the<br />

same direction after reconstitution) (Knol et al, 1998). On the other hand, reconstitution<br />

<strong>of</strong> some <strong>proteins</strong> <strong>with</strong> sodium cholate required a concentration <strong>of</strong> detergent above the<br />

saturation point for liposomes, i.e <strong>proteins</strong> in detergent micelles had to be mixed to<br />

detergent micelles containing lipids (Seddon et al., 2004). This latter method usually<br />

results in symmetrical orientations for the protein inside the proteoliposomes (Yeagle,<br />

1993). In the case <strong>of</strong> bile acid salts, dialysis is a good detergent removal procedure.<br />

These detergents present high cmc (high solubility as monomers) so dialysis can be<br />

efficient at reasonable time scales.<br />

Other reconstitution strategies include protein/peptide solubilization in organic<br />

solvents. This procedure must be applied <strong>with</strong> caution as it <strong>of</strong>ten results in aggregation<br />

and denaturation. Some protocols for <strong>proteins</strong> have been developed that make use <strong>of</strong> the<br />

24


INTRODUCTION: LIPID-PROTEIN INTERACTIONS<br />

ability <strong>of</strong> some organic solvents to stimulate hydrogen bonding and alpha-helical<br />

structure (Killian et al., 1994). This is particularly important when working <strong>with</strong> TM<br />

alpha-helical <strong>peptides</strong>, since this treatment can prevent aggregation and/or transition to<br />

different secondary structures that can be irreversible in some cases. Examples <strong>of</strong> such<br />

solvents are trifluoroethanol (TFE) and hexafluoroisopropanol.<br />

Sonication <strong>of</strong> mixed solutions <strong>of</strong> <strong>proteins</strong> in buffer and lipids can also affect the<br />

transfer <strong>of</strong> <strong>proteins</strong> to liposomes. The downfall <strong>of</strong> this technique is that it requires<br />

solubility <strong>of</strong> the protein in water, it can induce protein denaturation and always results<br />

in SUV’s, which can be problematic since some <strong>proteins</strong> reconstituted in high curvature<br />

liposomes have been shown to underwent changes in their activities. Freeze-thawing <strong>of</strong><br />

mixtures <strong>of</strong> sonicated liposomes and <strong>proteins</strong> has also been used in cases where the<br />

protein was sensitive to both detergents and sonication (Seddon et al., 2004).<br />

For a protein to be considered as correctly reconstituted, some requirements must be<br />

fulfilled. The characteristic function or activity <strong>of</strong> the protein in vivo must be regained<br />

(at least partially) after reconstitution. Ideally the process should result in a defined lipid<br />

environment for the protein <strong>with</strong> an also defined lipid to protein ratio (Yeagle, 1993).<br />

2.2. Peptides as models<br />

Membrane <strong>proteins</strong> are extremely complex entities. In addition to protein-lipid<br />

interactions, protein-protein interactions are crucial in dictating final properties as<br />

protein domains interact between themselves in the final structure. In order to study<br />

protein-lipid interactions an alternative is the use <strong>of</strong> peptide models, either TM or<br />

peripherically associated <strong>with</strong> the <strong>membrane</strong>. An additional advantage <strong>of</strong> this<br />

minimalist approach is that synthetic <strong>peptides</strong> can be obtained in large quantities, which<br />

is not feasible for many <strong>membrane</strong> <strong>proteins</strong>. The possibility <strong>of</strong> synthesis also allows an<br />

easy control over the primary sequence and wide mutation possibilities, and this is<br />

essential in understating the role <strong>of</strong> particular amino acids (Wimley and White, 1996)<br />

and peptide segments on the interactions <strong>with</strong> the lipid environment, and on the function<br />

<strong>of</strong> the protein itself. For large <strong>membrane</strong> <strong>proteins</strong> the range <strong>of</strong> possible mutations is<br />

quite limited due to problems related to cell viability <strong>of</strong> mutants.<br />

Putative TM alpha-helices can be estimated from hydropathy plots as described in<br />

Section 1.8, and hydrophobic moment estimations can assist on the search for<br />

25


amphipatic helices (see Section 2.3). Predictions derived from these methods allow the<br />

design <strong>of</strong> synthetic <strong>peptides</strong> for detailed biophysical <strong>studies</strong>.<br />

A number <strong>of</strong> experiments showed that synthetic <strong>peptides</strong> corresponding to the<br />

alpha-helical TM segments <strong>of</strong> some <strong>proteins</strong> can assemble and substitute the native<br />

<strong>proteins</strong> as well as their functions after independent folding, validating the approach <strong>of</strong><br />

studying isolated peptide segments (Marsh, 1996). Popot and Engelman (1990; 1993)<br />

rationalized these results <strong>with</strong> the proposal <strong>of</strong> a two-stage model for <strong>membrane</strong> insertion<br />

<strong>of</strong> the TM segments from a intrinsic <strong>membrane</strong> protein. In the first stage, alpha-helices<br />

are independently folded and inserted in the lipid bilayer, while in the second stage,<br />

alpha-helices interact and assemble in the final protein structure.<br />

2.3. Amphipatic helix<br />

Amphipatic helices are protein alpha-helices for which one face along the helical<br />

axis is hydrophilic whereas the opposite face is hydrophobic. The amphipatic helix is a<br />

common feature <strong>of</strong> biologically active poly<strong>peptides</strong> and <strong>proteins</strong>, such as hormones,<br />

antibiotics, and venoms. In this way, this structure can play a large number <strong>of</strong> roles. A<br />

function <strong>of</strong> great structural importance is the shielding <strong>of</strong> the hydrophobic interior <strong>of</strong><br />

<strong>proteins</strong> exposed to aqueous environments, while for <strong>membrane</strong> <strong>proteins</strong> the roles<br />

played by amphipatic helices are diverse. In the case <strong>of</strong> peripheric <strong>proteins</strong> they can act<br />

as anchors to the <strong>membrane</strong> by tight binding to the hydrocarbon/water interface <strong>of</strong> the<br />

bilayer as the structure <strong>of</strong> this boundary is highly complementary to the amphipatic<br />

helix. The amphipatic structure is also able to promote protein-protein interactions<br />

between hydrophobic segments <strong>of</strong> helices. In some cases this allows amphipatic<br />

<strong>peptides</strong> to create channels and pores in the <strong>membrane</strong> via aggregation into oligomeric<br />

bundles for which the hydrophobic side faces the hydrocarbon chains <strong>of</strong> the lipids and<br />

the hydrophilic faces <strong>of</strong> the helices are exposed to the inside. Thus delineating a pore in<br />

the bilayer (Epand, 1993). Recent <strong>studies</strong> also suggest a role <strong>of</strong> amphipatic helices in<br />

the modulation <strong>of</strong> <strong>membrane</strong> curvature (see chapter V).<br />

Overall, the amphiphilic character is a much more important dictator <strong>of</strong> peptide<br />

conformation than the specific characteristic <strong>of</strong> its constituent amino acids (Taylor,<br />

1990). Parameters like the ratio <strong>of</strong> hydrophobic to hydrophilic amino acids as well as<br />

26


INTRODUCTION: LIPID-PROTEIN INTERACTIONS<br />

the angle <strong>of</strong> the hydrophobic face are crucially important in dictating the location <strong>of</strong> the<br />

peptide in the <strong>membrane</strong> and the type <strong>of</strong> lipid interactions established (Kitamura et al.,<br />

1999; Brasseur, 1991).<br />

Methods exist to predict possible amphipatic helices from primary sequences <strong>of</strong><br />

<strong>proteins</strong>. From the knowledge <strong>of</strong> the periodicity <strong>of</strong> alpha-helices, Shiffer-Edmundson’s<br />

(1967) helical wheels and helical net representations (Dunnil, 1968) can be drawn like<br />

the ones in Figure I.13. In Shiffer-Edmundson’s representation, residues are sketched<br />

around a circle <strong>with</strong> each residue being placed 100º from the preceding one. If the<br />

segment corresponds to an amphipatic helix, then one face <strong>of</strong> the circle should be<br />

enriched in hydrophilic residues while the opposite face should present clustering <strong>of</strong><br />

hydrophobic amino acids. The net helical representation illustrates a hollow cylinder cut<br />

open along a parallel line to its axis and laid flat. The residues are placed in such a way<br />

so that if the cylinder would be closed, the amino acids are positioned as in a alphahelix<br />

(Auger, 1993). These methods are very easy to apply and provide great tools to<br />

visualize the disposition <strong>of</strong> the amino acids in a putative amphipatic helix but are not<br />

quantitative, i.e. they do not provide a quantification <strong>of</strong> the amphipatic character <strong>of</strong> the<br />

protein domain being evaluated.<br />

A<br />

B<br />

Figure I.13 - A- Shiffer-Edmundson’s helical wheel representation for amphipatic domains. B-<br />

Helical net representation for amphipatic domains (Taken from Anantharamaiah et al., 1993).<br />

Some <strong>of</strong> the available quantitative methods work by comparing a sequence <strong>of</strong><br />

hydrophobicities (<strong>of</strong> each amino acid) <strong>with</strong> a sinusoid, reporting the evolution <strong>of</strong> the<br />

amino acid positions in the alpha-helical polypeptide chain (in the axis perpendicular to<br />

the helical axis). The most common quantitative description <strong>of</strong> the amphipatic degree <strong>of</strong><br />

a protein domain is given by the hydrophobic moment (Eisenberg et al., 1982). In this<br />

analysis, the hydrophobic moment <strong>of</strong> a helix is equal to the summation <strong>of</strong> the vectors <strong>of</strong><br />

27


size corresponding to the hydrophobicity <strong>of</strong> each amino acid drawn from the centre <strong>of</strong><br />

the putative helix. Through calculation <strong>of</strong> the hydrophobic moment <strong>of</strong> an appropriate<br />

residue window (typically 11 – three turns in the alpha-helix) for each position in the<br />

protein sequence, it is possible to draw an “amphipatic pr<strong>of</strong>ile” <strong>of</strong> the protein (Fig.<br />

I.14), and by comparison <strong>with</strong> values obtained from known amphipatic helices it is<br />

possible to predict which protein domains are likely to correspond to amphipatic helices<br />

(Auger, 1993).<br />

Figure I.14 – Amphipatic pr<strong>of</strong>ile <strong>of</strong> gp41 from envelope protein <strong>of</strong> HIV variant bh10 (taken from<br />

Auger, 1993).<br />

2.4. Peptide partitioning to the <strong>membrane</strong><br />

Several approaches have been used to study peptide binding to lipid <strong>membrane</strong>s.<br />

The most important are fluorescence spectroscopy (Santos et al., 2003), circular<br />

dichroism, surface plasmon resonance and isothermal titration calorimetry. The binding<br />

<strong>of</strong> a peptide to the lipid bilayer is not a conventional reaction <strong>with</strong> a defined<br />

stoichiometry. In reality, the process is better described as a water/bilayer partition<br />

problem. Although stable complex formation between <strong>proteins</strong> and specific lipids have<br />

been proved and bound lipid molecules have been observed in the crystal structures <strong>of</strong><br />

some <strong>proteins</strong> (see Section 2.8) (Lee, 2003), this type <strong>of</strong> interaction is not generally<br />

found for most <strong>membrane</strong> <strong>proteins</strong>.<br />

28


INTRODUCTION: LIPID-PROTEIN INTERACTIONS<br />

Partitioning <strong>of</strong> TM <strong>peptides</strong> to lipid bilayers is <strong>of</strong>ten a very difficult problem to<br />

study. A TM configuration for a peptide implies a large number <strong>of</strong> strongly<br />

hydrophobic residues in the sequence, as the size <strong>of</strong> the hydrophobic segment must be<br />

large enough to span the hydrocarbon (apolar) region <strong>of</strong> the bilayer (see Section 2.8). As<br />

a result their solubilities in water are frequently very low or null, and the partition<br />

equilibrium <strong>of</strong> these <strong>peptides</strong> is completely shifted to the lipid phase <strong>with</strong> almost no<br />

monomeric peptide in the aqueous environment (the available strategy to shield<br />

hydrophobic residues is non-specific aggregation). For these reasons the study <strong>of</strong><br />

lipid/water partition <strong>of</strong> TM <strong>peptides</strong> is frequently inaccessible apart from the<br />

considerations in Section 1.8. As for peripheral <strong>peptides</strong>, i.e. <strong>peptides</strong> that bind to the<br />

surface <strong>of</strong> lipid bilayers, the hydrophobicity requirements are not so severe and<br />

solubility in aqueous environments is <strong>of</strong>ten observed.<br />

The interaction <strong>of</strong> the peptide <strong>with</strong> the <strong>membrane</strong> is mainly dictated by the<br />

following effects:<br />

1) Electrostatic contributions are <strong>of</strong> key importance when basic residues are present in<br />

the peptide and the lipid surface is rich in negatively charged lipids. However, if the<br />

lipids are zwitterionic, the electrostatic influence can be negligible, and if both lipids<br />

and <strong>peptides</strong> present negative charges, the electrostatic contribution becomes negative<br />

(repulsion). In case <strong>of</strong> effective partition due to electrostatic attraction, this contribution<br />

is no longer independent <strong>of</strong> peptide concentration at very small lipid to peptide ratios<br />

(L/P). In this range <strong>of</strong> concentrations, the potential <strong>of</strong> the bilayer surface becomes<br />

dependent on the peptide concentration as a consequence <strong>of</strong> screening <strong>of</strong> lipid negative<br />

charges by the peptide (Seelig, 2004).<br />

2) Classical hydrophobic contributions are entropically driven, and result from the<br />

energetic requirements for insertion <strong>of</strong> apolar residues in the aqueous environment<br />

(rotational restriction <strong>of</strong> water molecules in the hydration shell). They are also<br />

dependent on the final positioning <strong>of</strong> the peptide in the lipid bilayer (bilayer surface or<br />

core).<br />

3) Perturbation <strong>of</strong> the lipid structure and packing.<br />

4) Hydrophobic matching <strong>of</strong> TM <strong>peptides</strong> (see Section 2.6).<br />

5) Hydrogen bonding properties <strong>of</strong> the lipid bilayer interface are highly complex, since<br />

both hydrogen bond donors and acceptors are present. As an example, PC bilayers<br />

establish stronger hydrogen bonds <strong>with</strong> phenolic compounds than PE bilayers due to the<br />

presence <strong>of</strong> the choline group. Additionally, the dielectric constant <strong>of</strong> the lipid bilayer<br />

29


changes abruptly in the interfacial region, varying from the value in water (ε = 78) to<br />

the value in the hydrocarbon core (ε = 2) (McIntosh, 2002).<br />

5) After adsorption <strong>of</strong> the peptide to the bilayer surface, conformational changes<br />

generally take place that entail changes in the thermodynamic properties <strong>of</strong> the binding<br />

process. One common scenario is that <strong>peptides</strong> are unstructured (random coil) while in<br />

the aqueous environment and after binding to the bilayer take on an amphipatic<br />

structure. The bound structure can also be dependent on the L/P in the <strong>membrane</strong><br />

(transition from an alpha-helical to beta-sheet structure). The change to an amphipatic<br />

structure leads to multiple intermolecular hydrogen bonds and results in a largely<br />

favourable contribution to partition. Isothermal titration calorimetry <strong>studies</strong> <strong>with</strong> some<br />

amphipatic <strong>peptides</strong> revealed that the favourable enthalpic contributions for peptide<br />

binding arising from alpha-helix formation amounted to half <strong>of</strong> the free energy <strong>of</strong><br />

binding (McIntosh, 2002).<br />

The thermodynamic parameter that relates directly to the partition <strong>of</strong> the peptide<br />

between the lipid and water phases is the partition coefficient (K p ):<br />

K =<br />

p<br />

n<br />

n<br />

S,L<br />

S,W<br />

/ V<br />

/ V<br />

L<br />

W<br />

2.1<br />

where V i are the volumes <strong>of</strong> the phases, and n S,i are the moles <strong>of</strong> solute in each phase (i<br />

=W, aqueous phase; i=L, lipid phase) (Mateo et al., 2006). The partition coefficient is<br />

converted in free energy <strong>of</strong> binding to the <strong>membrane</strong> (∆G 0 ) by:<br />

o<br />

∆G = -RT ln<br />

( Kp)<br />

2.2<br />

After binding, amphipatic <strong>peptides</strong> are expect to reside in the lipid/water interface <strong>of</strong><br />

the <strong>membrane</strong>. The exact position is nevertheless dependent on both the <strong>membrane</strong> and<br />

the peptide sequence. From X-ray diffraction measurements on DOPC bilayers, the axis<br />

<strong>of</strong> an ideally amphipatic peptide was located between the mean positions <strong>of</strong> the glycerol<br />

and carbonyl groups (Fig. I.6), exactly at the interface between the polar headgroups<br />

and the hydrocarbon core (Mishra et al., 1994).<br />

30


INTRODUCTION: LIPID-PROTEIN INTERACTIONS<br />

2.5. Anchoring <strong>of</strong> trans<strong>membrane</strong> domains<br />

TM <strong>proteins</strong> are enriched in hydrophobic amino acids such as Leu, Ile, Val, Trp,<br />

Tyr, and Phe. However, the occurrence <strong>of</strong> the aromatic amino acids Trp and Tyr are<br />

strongly dependent on the position inside the TM segment. These aromatic amino acids<br />

are particularly enriched at the end <strong>of</strong> TM sequences, close to the headgroup region <strong>of</strong><br />

the lipid bilayer (Wallin et al., 1997). Apparently, these aromatic residues exhibit strong<br />

preferential interactions for the complex electrostatic carbonyl-glycerol environment <strong>of</strong><br />

the bilayer (see Fig. I.6). In this region the aromatic rings are stabilized, and,<br />

simultaneously, penetration in the hydrocarbon core <strong>of</strong> the bilayer is somewhat<br />

unfavourable due to exclusion <strong>of</strong> the flat aromatic rings (Yau et al., 1988, Braun and<br />

von Heijne, 1999). Interestingly, another aromatic residue, Phe, does not exhibit<br />

significant preferential localization in the <strong>membrane</strong> environment (Braun and von<br />

Heijne, 1999).<br />

Another important characteristic <strong>of</strong> TM domains is that they are generally flanked<br />

by polar regions, rich in charged residues. Positively charged amino acids are found<br />

close to the hydrophobic domain <strong>of</strong> the protein and in some cases they can even be<br />

deeply inserted in this region (Caputo and London, 2003). The energetic cost <strong>of</strong> burying<br />

<strong>of</strong> a charge in the hydrophobic environment <strong>of</strong> the lipid bilayer is very large, and in that<br />

situation the long charged side chains <strong>of</strong> these residues protrude towards the headgroup<br />

region in a phenomenon called “snorkling”. For acidic residues (Asp and Glu), the<br />

absence <strong>of</strong> long side chains present a greater barrier for insertion in the hydrophobic<br />

core <strong>of</strong> the bilayer. This is however observed for TM sequences <strong>with</strong> uninterrupted<br />

hydrophobic domains <strong>of</strong> at least 12 residues (Caputo and London, 2004), and in these<br />

conditions it was already proposed that the buried acidic residues can become<br />

protonated (Monné et al., 1998).<br />

2.6. Lipid-protein hydrophobic matching<br />

Hydrophobic matching is not solely relevant for lipid-lipid interactions (Section<br />

1.7), and the problem exists also for protein-lipid interactions. In this case, the<br />

hydrophobic length <strong>of</strong> the TM domain <strong>of</strong> the protein should match to the hydrophobic<br />

31


thickness <strong>of</strong> the bilayer (Section 1.8). This principle was first introduced by Bloom and<br />

Mouritsen in 1984.<br />

Lipid-protein hydrophobic mismatching can have two origins (Figure I.15). For<br />

positive hydrophobic mismatch, the protein hydrophobic thickness (d P ) is longer than<br />

the hydrophobic thickness <strong>of</strong> adjacent lipid molecules (d L ) and exposure <strong>of</strong> hydrophobic<br />

residues <strong>of</strong> the protein to the aqueous environment results in an energetically<br />

unfavourable term to the protein-lipid interaction. In the case <strong>of</strong> opposite situation, i.e.<br />

negative mismatch, when lipid chains are longer than the hydrophobic domain <strong>of</strong> the<br />

protein, the energetically unfavourable term to protein-lipid interaction originates from<br />

the exposure <strong>of</strong> the hydrocarbon chains to water.<br />

Both situations <strong>of</strong> hydrophobic mismatch create an energetic stress in the proteinlipid<br />

interface that must elicit a response from the system. Most models for protein-lipid<br />

mismatch assume a response in the form <strong>of</strong> stretching or compression <strong>of</strong> lipid acylchains<br />

(Mall et al., 2002).<br />

Figure I.15 – Schematic representation <strong>of</strong> positive and negative hydrophobic mismatch in proteinlipid<br />

interfaces (From Dumas et al. 1999).<br />

Several theoretical models have been developed to estimate the energetic penalties<br />

in protein-lipid hydrophobic mismatch. Fattal and Ben-Shaul (1993) considered that the<br />

total free energy <strong>of</strong> the protein-lipid interaction was a summation <strong>of</strong> the contributions<br />

from the chain and headgroup terms. The presence <strong>of</strong> a rigid protein body (as it was<br />

assumed in the model) in contact <strong>with</strong> lipid acyl-chains, implied an energetic penalty for<br />

the interaction. The stretching or compression <strong>of</strong> lipid-chains to fit the hydrophobic<br />

thickness <strong>of</strong> the <strong>membrane</strong> protein also translated into a positive contribution to the<br />

free-energy <strong>of</strong> interaction. The contributions from interactions in the headgroup region<br />

included an attractive term related to the exposure <strong>of</strong> the hydrocarbon core to the<br />

aqueous environment as well as a repulsive term resulting from electrostatic and<br />

32


INTRODUCTION: LIPID-PROTEIN INTERACTIONS<br />

excluded-volume interactions between headgroups. The resulting free-energy pr<strong>of</strong>ile for<br />

lipid-protein mismatch was nearly symmetrical about the point <strong>of</strong> hydrophobic<br />

matching, and the lipid perturbation energy F (in units <strong>of</strong> kT per Å 2 <strong>of</strong> protein<br />

circumference) is described by (Ben-Shaul, 1995):<br />

F<br />

( ) 2<br />

= 0,37<br />

+ 0,005 ⋅ d P<br />

− d L<br />

2.3<br />

This model assumed concentration <strong>of</strong> the hydrophobic mismatch perturbation on the<br />

first shell <strong>of</strong> lipids around the protein (see 2.7). For a hydrophobic mismatch <strong>of</strong> 7 Å and<br />

assuming that each lipid adjacent to the protein should occupy at least 4 Å 2 <strong>of</strong> the<br />

protein perimeter, the free-energy for lipid-protein interaction would increase by 3.6 kJ<br />

mol -1 . This change in lipid-protein interaction energy corresponds to a decrease factor in<br />

lipid-protein binding constant <strong>of</strong> 4.3. If propagation <strong>of</strong> perturbation spreaded to the<br />

more external shells, the effects <strong>of</strong> hydrophobic mismatch on the adjacent lipid<br />

molecules would be decreased. For example, if effects were averaged over three shells<br />

<strong>of</strong> lipids, the change in binding constant would decrease by a factor <strong>of</strong> 1.6 (Mall et al.,<br />

2002). Alternative models for treating hydrophobic mismatch came to very similar<br />

conclusions regarding interaction energies (Nielsen et al., 1998; Mouritsen and Bloom,<br />

1993).<br />

Energies <strong>of</strong> this magnitude are sufficient to induce changes not only in the<br />

interacting lipid molecules but in the protein itself. Experimental <strong>studies</strong> allowed to<br />

conclude that β-barrel <strong>proteins</strong> are not easily susceptible to structural/conformational<br />

changes due to hydrophobic mismatch and the energetic penalty for protein-lipid<br />

interaction was only affecting lipid structure (O´Keeffe et al, 2000). However, in the<br />

cases <strong>of</strong> alpha-helical TM <strong>proteins</strong> such as the potassium channel KcsA, and the<br />

mechanosensitive channel MscL, the experimentally obtained dependence <strong>of</strong> binding<br />

constants on the extent <strong>of</strong> hydrophobic mismatch was smaller than expected<br />

(Williamson et al., 2002; Powl et al., 2003). This was rationalized as due to the smaller<br />

rigidity <strong>of</strong> the α-helical <strong>proteins</strong> that makes them susceptible to distortion when<br />

hydrophobic mismatch was present and therefore create an additional degree <strong>of</strong> freedom<br />

for relaxation <strong>of</strong> lipid-protein hydrophobic mismatch. Distortion <strong>of</strong> α-helical <strong>membrane</strong><br />

<strong>proteins</strong> by hydrophobic mismatch is supported by the observation that the activities <strong>of</strong><br />

33


some α-helical <strong>membrane</strong> <strong>proteins</strong> are dependent on the <strong>membrane</strong> thickness (East and<br />

Lee, 1982).<br />

Most likely, the most common distortion available for α-helical <strong>membrane</strong> <strong>proteins</strong><br />

as a response to positive hydrophobic mismatch is a change <strong>of</strong> tilt <strong>of</strong> the TM helices<br />

relative to the bilayer normal. This mechanism allows for an easy modulation <strong>of</strong> the<br />

length along the bilayer normal occupied by the hydrophobic domains <strong>of</strong> the protein.<br />

Indeed, tilting has been observed and measured for various TM <strong>proteins</strong> (Harzer and<br />

Bechinger, 2000).<br />

Rotation <strong>of</strong> amino acid residues at the end <strong>of</strong> TM α-helices was also proposed as<br />

another distortion mechanism available for <strong>membrane</strong> <strong>proteins</strong> as this results in changes<br />

in the effective hydrophobic thickness <strong>of</strong> the protein (Lee, 2003).<br />

However, the ability <strong>of</strong> <strong>proteins</strong> and lipids to adapt to situations <strong>of</strong> hydrophobic<br />

mismatch is limited, and drastic changes in the <strong>membrane</strong> distribution <strong>of</strong> these elements<br />

is possible. In cases <strong>of</strong> severe hydrophobic mismatch, the TM orientation <strong>of</strong> TM helices<br />

might not be maintained anymore and transitions to non-TM orientations are possible<br />

(Ren et al., 1997; Ren et al., 1999) as well as decrease <strong>of</strong> partition to bilayers (Figure<br />

I.16). Aggregation <strong>of</strong> TM domains is another possibility in hydrophobic mismatch as<br />

this reduces the lipid-protein interface, minimizing stress. Protein aggregation will be<br />

discussed in greater detail in section 2.11.<br />

A<br />

B<br />

Figure I.16 – Possible adaptations to hydrophobic mismatch between a TM peptide and the lipid bilayer.<br />

A – Response to positive hydrophobic mismatch (a): acyl-chain ordering (b), peptide backbone<br />

deformation (c), peptide oligomerization (d), peptide tilt (e), non-TM orientation or absence <strong>of</strong> binding to<br />

the bilayer (f). B – Response to negative hydrophobic mismatch (a): acyl-chain disordering (b), peptide<br />

backbone deformation (c), peptide oligomerization (d), non-lamellar phase formation (e), non-TM<br />

orientation or absence <strong>of</strong> binding to the bilayer (f). (Adapted from de Planque and Killian, 2003).<br />

34


INTRODUCTION: LIPID-PROTEIN INTERACTIONS<br />

In the presence <strong>of</strong> lipids <strong>with</strong> tendency to establish nonlamellar structures, cubic and<br />

nonlamellar lipid phases can be formed in the interface <strong>with</strong> the protein as a response to<br />

hydrophobic mismatch. Nonlamellar arrangements <strong>of</strong> lipids change the curvature <strong>of</strong> the<br />

bilayer in the vicinity <strong>of</strong> the protein and allow for a better fitting <strong>of</strong> the hydrophobic<br />

sections <strong>of</strong> both <strong>proteins</strong> and lipids (Killian, 2003).<br />

Peptide backbone deformation, such as transition from α-helix to π-helix (helices<br />

<strong>with</strong> a wider helical pitch), could be a possible mechanism for adjustment to situations<br />

<strong>of</strong> hydrophobic mismatch. However, several <strong>studies</strong> have confirmed that the α-helical<br />

structure is insensitive to hydrophobic mismatch and this strategy to minimize<br />

hydrophobic mismatch is unlikely (de Planque and Killian, 2003). Nevertheless a recent<br />

study suggests that TM α-helices may flex in the lipid bilayer in order to submerge most<br />

<strong>of</strong> the hydrophobic domain inside the bilayer core (Yeagle et al., 2007).<br />

The hydrophobic matching principle might be a mechanism by which the activity <strong>of</strong><br />

some <strong>membrane</strong> <strong>proteins</strong> can be modulated. Thickness variations induced either<br />

internally or externally can trigger some <strong>proteins</strong> in or out <strong>of</strong> an active status. This has<br />

already been observed for some ion pumps such as Ca 2+ -ATPase and Na + ,K + -ATPase<br />

that exhibit maximum activity in bilayers <strong>with</strong> a specific number <strong>of</strong> carbon atoms (Lee,<br />

2003).<br />

2.7. Trans<strong>membrane</strong> protein-lipid interface<br />

As already stated, the lipids on the vicinity <strong>of</strong> the protein can either stretch or<br />

compress their acyl-chains in response to hydrophobic mismatch. In pure lipids this<br />

<strong>of</strong>ten results in the observation <strong>of</strong> two different lipid populations in protein-lipid<br />

systems. One <strong>of</strong> these is very similar to the lipid in the absence <strong>of</strong> the protein (but not<br />

necessarily identical), and its contribution to the total lipid population is inversely<br />

dependent on protein concentration. The other population is more significant at higher<br />

protein/peptide concentrations, frequently presenting a different gel-fluid phase<br />

transition temperature (smaller T m for bilayers thinner than the protein and higher T m for<br />

thicker bilayers) as well as a smaller cooperativity for this transition (Liu et al., 2002;<br />

Morein et al., 2002). These two components are expected to correspond to free and<br />

protein-associated lipids, respectively.<br />

35


The formation <strong>of</strong> a lipid population associated <strong>with</strong> <strong>proteins</strong> or <strong>peptides</strong> <strong>with</strong><br />

properties different from the bulk or free lipid was predicted from results <strong>of</strong> different<br />

techniques, namely differential scanning calorimetry and electron spin resonance (ESR).<br />

From ESR experiments, a fraction <strong>of</strong> the spin labelled lipids was found to be<br />

dynamically restricted by interaction <strong>with</strong> <strong>membrane</strong> <strong>proteins</strong> or <strong>peptides</strong>. From the<br />

number <strong>of</strong> lipid molecules affected by the presence <strong>of</strong> a single TM domain it was<br />

possible to retrieve the stoichiometry for the interaction <strong>of</strong> lipids and TM <strong>peptides</strong>. The<br />

value recovered was 12, meaning that only the first shell <strong>of</strong> lipids around the TM<br />

domain is significantly affected by its presence (a hexagonal-type arrangement <strong>of</strong> lipids<br />

around a TM peptide is expected, six in each monolayer – see Figure I.17) (Marsh et al.,<br />

2002). This shell around the TM domain is entitled the annulus and the lipids <strong>with</strong>in are<br />

called annular lipids.<br />

Figure I.17 – Hexagonal lipid packing <strong>of</strong> annular lipids around a TM protein domain. Annular lipids<br />

are shown in dark grey and bulk lipids are depicted in light grey.<br />

The fact that only the first shell <strong>of</strong> lipids around the TM domain is significantly<br />

affected explains why single TM <strong>peptides</strong> are generally not able to induce considerable<br />

physical changes in the bulk <strong>membrane</strong> (except when the lipid/peptide ratio is large<br />

enough so that a significant fraction <strong>of</strong> the lipids are annular lipids). However, multipass<br />

TM <strong>proteins</strong> (containing multiple TM domains) were already shown to be able to affect<br />

the bulk lipids and change the hydrophobic thickness <strong>of</strong> the bilayer (Harroun et al.,<br />

1999). These results seem to indicate that the packing <strong>of</strong> lipid chains around a single α-<br />

helix (which has a diameter ~ 10 Å, comparable to a lipid molecule) is significantly<br />

different from the way lipids pack against a large protein surface (Weiss et al., 2003).<br />

Still, when multipass TM <strong>proteins</strong> were studied by ESR <strong>with</strong> spin labelled-lipids, the<br />

ratio <strong>of</strong> immobilized lipids/TM domains is <strong>of</strong>ten no longer 12 but smaller. This is<br />

obviously due to interactions and contacts between the TM domains that decrease the<br />

area available for contact <strong>with</strong> lipids (Marsh et al., 2002). However, if lipids outside the<br />

36


INTRODUCTION: LIPID-PROTEIN INTERACTIONS<br />

annulus were considerably restricted by the presence <strong>of</strong> protein, the immobilized<br />

lipids/TM domains ratio should come larger that for a single α-helix. Likely, the<br />

influence <strong>of</strong> the TM <strong>proteins</strong> on the annular lipids is much stronger than on lipids on the<br />

exterior shells, and the effects on the latter are not detectable by ESR.<br />

The lipid exchange rate at the interface <strong>of</strong> the protein body was determined to be in<br />

the region <strong>of</strong> 10 7 s -1 , which is on the same order but significantly lower than the value<br />

obtained for lipid diffusion in the absence <strong>of</strong> protein (Marsh, 1990). This means that<br />

although temporarily restricted to the protein surface, the lipid molecules interacting<br />

<strong>with</strong> the protein body are not immobile and are still in exchange <strong>with</strong> other lipids.<br />

Of relevance is the fact that lipids display different properties when at the surface <strong>of</strong><br />

<strong>proteins</strong>. This indicates that optimal hydrophobic matching might not necessarily occur<br />

when the hydrophobic thickness <strong>of</strong> the protein matches the hydrophobic thickness <strong>of</strong> the<br />

unperturbed bilayer (Dumas et al., 1999).<br />

2.8. Lipid selectivity at protein interfaces<br />

For <strong>membrane</strong>s containing multiple lipid species, the hydrophobic matching<br />

principle can result in even more complex behaviour at the lipid-protein interface.<br />

Different lipid species are expected to present different free energies for interaction <strong>with</strong><br />

the presented TM domain and as a result, the lipid composition around the protein can<br />

be different from the bulk lipid composition (Figure I.18).<br />

Selectivity <strong>of</strong> <strong>membrane</strong> <strong>proteins</strong> for lipids in specific phases was observed in<br />

various cases. In the presence <strong>of</strong> a mixture <strong>of</strong> DLPC and DSPC, a combination <strong>of</strong><br />

experimental and theoretical approaches indicated that bacteriorhodopsin exhibited<br />

preferential interaction <strong>with</strong> DLPC at low temperatures, when both lipids where in the<br />

gel phase, and at the fluid phase, bacteriorhodopsin had a preference for long chain<br />

DSPC (Dumas et al., 1997). DLPC and DSPC present a very large difference in<br />

hydrophobic length <strong>of</strong> acyl-chains and strongly nonideal mixing. Consequently, it is not<br />

surprising that the affinity <strong>of</strong> the protein for the lipids differs drastically, and that the<br />

protein is preferentially solvated by the lipids that are the best hydrophobic match in<br />

each phase. Still, selectivity <strong>of</strong> <strong>proteins</strong> for lipids in binary mixtures <strong>with</strong> a much more<br />

ideal mixing is also experimentally verified (Lee, 2003). Nevertheless, deviation <strong>of</strong> a<br />

37


homogeneous distribution <strong>of</strong> lipids in the <strong>membrane</strong> does have an energetic entropic<br />

cost and the degree <strong>of</strong> protein-lipid selectivity is a balance between the need to satisfy<br />

hydrophobic matching conditions in the protein-lipid interface and the intrinsic<br />

tendency <strong>of</strong> the system for mixing.<br />

Figure I.18 – Illustration <strong>of</strong> a lipid bilayer <strong>with</strong> an embedded protein and two different lipid species.<br />

The hydrophobic matching principle implies an accumulation <strong>of</strong> the lipid species that is hydrophobically<br />

best matched to the protein (taken from Jensen and Mouritsen, 2004).<br />

The hydrophobic surface <strong>of</strong> a <strong>membrane</strong> protein is not smooth. The interface<br />

between the protein and the lipids surrounding it is likely to be heterogeneous, and the<br />

interactions taking place there complex (Lee, 2003). Still the results described above,<br />

concerning the presence <strong>of</strong> a fixed stoichiometry for annular lipids, denote some degree<br />

<strong>of</strong> ordering in the TM alpha helix-lipid interface and in this way, processes <strong>of</strong> lipid<br />

binding to simpler systems such as single TM domains are suitable to be described in<br />

terms <strong>of</strong> a uniform surface for which several (12) identical binding sites are available<br />

(Marsh et al., 2002). With total coverage <strong>of</strong> the protein surface by lipids, one lipid<br />

molecule must leave the surface before another can enter. In this sense, the process can<br />

be depicted as competitive binding <strong>of</strong> lipids at the binding sites on the protein surface<br />

(Lee, 2003) (Figure I.19). This method <strong>of</strong> analysis has been applied <strong>with</strong> success to<br />

even more complex <strong>membrane</strong> <strong>proteins</strong> (O’ Keefe et al., 2000; Williamson et al., 2002;<br />

Powl et al., 2003).<br />

38


INTRODUCTION: LIPID-PROTEIN INTERACTIONS<br />

Figure I.19 – View from above <strong>of</strong> lipid binding sites on a TM domain surface. Two lipid species (L 1<br />

and L 2 ) are shown exchanging at one site.<br />

At each site an equilibrium exists:<br />

PL 1 + L 2 PL 2 + L 1<br />

where PL 1 and PL 2 are the protein-lipid complexes <strong>with</strong> lipid species 1 and 2. This<br />

equilibrium can be described by a binding constant (K b ):<br />

b<br />

[ PL<br />

2<br />

] ⋅[ L1]<br />

[ PL ] ⋅[ L ]<br />

K =<br />

2.4<br />

1<br />

2<br />

If L 2 is present at small amounts in the system (L 2 > [PL 2 ], then the probability (µ) <strong>of</strong> a lipid<br />

annular site to be occupied by L 2 is:<br />

[ PL<br />

2<br />

]<br />

[ PL ] + [ PL ]<br />

2<br />

1<br />

[ PL<br />

2<br />

]<br />

[ PL ]<br />

1<br />

b<br />

[ L<br />

2<br />

]<br />

[ L ]<br />

µ = = = K ⋅<br />

2.5<br />

1<br />

Through knowledge <strong>of</strong> [PL 2 ] ([PL 1 ]=[P]-[PL 2 ]), and the concentration <strong>of</strong> the lipid<br />

species 1 and 2 it is then possible to recover a binding constant for protein-lipid<br />

selectivity.<br />

The most popular methodologies for quantifying protein-lipid selectivity have been<br />

the already described ESR measurements <strong>with</strong> spin labelled lipids, and measurements <strong>of</strong><br />

protein fluorescence quenching by brominated lipids (East and Lee, 1982; O’ Keeffe et<br />

al., 2000; Williamson et al., 2002; Powl et al., 2003). From these <strong>studies</strong>, binding<br />

constants were retrieved as function <strong>of</strong> fatty-acid chain for a series <strong>of</strong> <strong>proteins</strong> (Fig.<br />

I.20).<br />

39


Figure I.20 – Relative binding constants for OmpF (ο), KscA (∆) and Ca 2+ -ATPase (□) for<br />

monounsaturated phosphatidylcholines <strong>with</strong> different acyl-chains relative to that for DOPC (Taken from<br />

Lee, 2003).<br />

As already discussed, the amplitude <strong>of</strong> changes in the relative binding constants <strong>of</strong><br />

the α-helical <strong>proteins</strong>, KscA and Ca 2+ -ATPase are not as high as one should expect<br />

from equation 2.3 due to adaptation <strong>of</strong> <strong>proteins</strong> to more hydrophobically matching<br />

configurations, likely by tilting and changes in the packing <strong>of</strong> α-helices (<strong>proteins</strong> are not<br />

rigid bodies) whereas KscA exhibits preference for lipids <strong>with</strong> 22 carbons in the acylchain,<br />

lipid binding constants for Ca 2+ -ATPase are hardly dependent on acyl-chain, and<br />

OmpF is preferentially bound to shorter lipids (14 carbons).<br />

Membrane <strong>proteins</strong> also exhibit selectivity for lipid headgroups. Strong interactions<br />

are <strong>of</strong>ten found between positively charged TM domains (rich in Lys and Arg) and<br />

negatively charged phospholipids. These selectivities can be the result <strong>of</strong> charge<br />

interactions, and in these cases, after binding <strong>of</strong> the anionic lipid to the surface <strong>of</strong> the<br />

protein, the positive potential in the vicinity <strong>of</strong> the protein would be slightly reduced. If<br />

enough anionic lipid molecules bind to the vicinity <strong>of</strong> the protein, the positive potential<br />

will be neutralized and remaining interactions <strong>with</strong> anionic phospholipids will be<br />

nonspecifical. However, <strong>membrane</strong> <strong>proteins</strong> <strong>of</strong>ten present different selectivities for<br />

different anionic phospholipids. This is likely to result from different balances in<br />

charges inside each class <strong>of</strong> phospholipids. In this way, PS lipids due to the presence <strong>of</strong><br />

the positively charged ammonium group as well as the negatively charged carboxyl<br />

group in the headgroup region may present reduced affinity to basic protein domains<br />

when compared to phosphatidic acid. Also, PG, <strong>with</strong> a single charge in the phosphate<br />

40


INTRODUCTION: LIPID-PROTEIN INTERACTIONS<br />

group interacts differently <strong>with</strong> charged <strong>membrane</strong> <strong>proteins</strong>, and <strong>proteins</strong> <strong>of</strong>ten present<br />

smaller affinities for PG than for PS and PA (Mall et al., 2002).<br />

The complexity to the protein-charged lipid interaction problem is raised by the<br />

position <strong>of</strong> amino acid residues in the protein surface. Basic helices <strong>with</strong> different<br />

hydrophobic thickness were shown to present different selectivities to negatively<br />

charged lipids. This was rationalized as the result <strong>of</strong> tilting <strong>of</strong> helices in the <strong>membrane</strong>,<br />

that located basic residues away from the headgroup region where strong interactions<br />

between <strong>proteins</strong> and lipid headgroups are more likely (Mall et al., 2002). Additionally,<br />

experiments <strong>with</strong> model <strong>peptides</strong> showed that insertion <strong>of</strong> Tyr decreased binding<br />

affinities to anionic phospholipids, suggesting that the presence <strong>of</strong> Tyr residues<br />

prevented close association <strong>of</strong> anionic phospholipids and cationic residues (Mall et al.,<br />

2000).<br />

These results suggest that the effects <strong>of</strong> charge on the interactions between lipids<br />

and <strong>membrane</strong> <strong>proteins</strong>, though important are not determinant, and will be strongly<br />

dependent on the detailed structure <strong>of</strong> the peptide and its orientation in the <strong>membrane</strong><br />

(Mall et al., 2002). Binding affinities <strong>of</strong> <strong>proteins</strong> for lipids are generally presented<br />

relative to the binding affinity <strong>of</strong> PC lipids. PC and PE lipids frequently present the<br />

lowest affinities for <strong>membrane</strong> <strong>proteins</strong>.<br />

Binding sites for hydrophobic molecules have been shown to exist for several<br />

<strong>proteins</strong> distinct from the sites for interaction <strong>with</strong> annular lipids. These are called nonannular<br />

sites. They are generally located between TM α-helices or at protein-protein<br />

interfaces. For some cases, these sites are available for interaction <strong>with</strong> some lipids but<br />

not for others (Lee, 2003). Some lipid molecules were shown in crystal structures <strong>of</strong><br />

<strong>proteins</strong> to be bound to such sites. These observations suggest that the binding affinities<br />

for these sites are much higher than the ones observed for annular lipids (Lee et al.,<br />

2003).<br />

Notably, several protein domains bind specifically to one or more forms <strong>of</strong><br />

phosphorylated PIP molecules through interactions that are very specific. Such domains<br />

are <strong>of</strong>ten found on <strong>proteins</strong> involved in signal transduction. Other <strong>proteins</strong> bind<br />

specifically to some components <strong>of</strong> raft domains, establishing a mechanism for<br />

targeting the protein to specific structures in the <strong>membrane</strong> (Epand, 2005; Pérez-Gil et<br />

al., 2005).<br />

41


2.9. Lipid phase preferential partition <strong>of</strong><br />

<strong>membrane</strong> <strong>proteins</strong><br />

Many <strong>membrane</strong> <strong>proteins</strong> and <strong>peptides</strong> show a preference for bilayers in the fluid<br />

phase over the gel phase. This results in a difference in partition between domains when<br />

both phases coexist (Dumas et al., 1997). With even greater biological relevance,<br />

<strong>proteins</strong> exhibit also differential partition between liquid ordered and liquid disordered<br />

phases. In this way, the lateral structure <strong>of</strong> the lipid <strong>membrane</strong> is able to impose major<br />

restrictions to the lateral distribution <strong>of</strong> <strong>membrane</strong> <strong>proteins</strong>. Preferential partition <strong>of</strong> a<br />

<strong>membrane</strong> protein to a given region or phase <strong>of</strong> the <strong>membrane</strong> is likely to define<br />

important structural features, such as protein densities or the probability <strong>of</strong> protein<br />

encountering homologous or heterologous counterparts (see Section 2.11) (Pérez-Gil et<br />

al., 2005).<br />

Acylated <strong>proteins</strong> or lipid-anchored <strong>proteins</strong> will also preferentially partition to<br />

certain domains. Acylation <strong>with</strong> palmitic acid or myristic acid promotes preferential<br />

partition to more ordered domains such as <strong>membrane</strong> rafts. On the other hand,<br />

prenylated <strong>proteins</strong> are commonly excluded from liquid ordered domains (Epand,<br />

2005).<br />

Changes <strong>of</strong> the lateral lipid <strong>membrane</strong> structure would thus have the ability to<br />

induce pr<strong>of</strong>ound rearrangements <strong>of</strong> the protein lateral distribution. This would have<br />

major implications on the establishment <strong>of</strong> specific protein-protein interaction networks.<br />

In fact, it has been proposed that the natural <strong>membrane</strong> composition is optimized to<br />

maintain <strong>membrane</strong>s on the threshold <strong>of</strong> a phase transition. Membranes in this situation<br />

would present a highly dynamical behaviour able to respond through redistribution <strong>of</strong><br />

the protein network to environmental stimulus (Pérez-Gil et al., 2005).<br />

Hydrophobic matching might also be the mechanism by which <strong>proteins</strong> are targeted<br />

to their final destinations in the <strong>membrane</strong>. Sorting <strong>of</strong> <strong>proteins</strong> along the secretory<br />

pathway might be performed by means <strong>of</strong> a gradient <strong>of</strong> hydrophobic thickness <strong>of</strong> the<br />

<strong>membrane</strong> systems that the <strong>proteins</strong> have to pass on their way to their target. The<br />

concentrations <strong>of</strong> cholesterol and sphingomyelin, which have a thickening effect on the<br />

<strong>membrane</strong>, increase going from the endoplasmic reticulum, via the Golgi to the plasma<br />

<strong>membrane</strong>. Also, the hydrophobic domains <strong>of</strong> <strong>proteins</strong> that reside in the Golgi are<br />

shorter by about 5 amino acids when compared to those <strong>of</strong> the plasma <strong>membrane</strong>.<br />

42


INTRODUCTION: LIPID-PROTEIN INTERACTIONS<br />

However, this proposed mechanism for sorting <strong>of</strong> <strong>membrane</strong> <strong>proteins</strong> is still<br />

controversial (Mouritsen, 2005).<br />

Concerning <strong>membrane</strong> protein partition to specific regions <strong>of</strong> the bilayer, <strong>of</strong><br />

particular interest is the possibility that certain <strong>membrane</strong> <strong>proteins</strong> prefer insertion in the<br />

boundaries between lipid phases in the <strong>membrane</strong>s. These boundaries present a higher<br />

probability <strong>of</strong> lipid packing defects that can be used by certain <strong>proteins</strong> to access deeper<br />

regions in the <strong>membrane</strong> (Pérez-Gil et al., 2005).<br />

2.10. Lipid sorting by <strong>proteins</strong> and formation <strong>of</strong><br />

lipid domains<br />

The presence <strong>of</strong> very stable lipid domains in the <strong>membrane</strong> is only expected if the<br />

free energies for interaction between phase separated lipids is very high. In<br />

bio<strong>membrane</strong>s this is rarely the case, and therefore lateral structures that arise must have<br />

a reasonably dynamical nature. In model <strong>membrane</strong>s, the differences in free-energies<br />

for lipid-lipid interactions are typically on the order <strong>of</strong> a few calories per mole. In<br />

contrast, the free-energies involved in protein-lipid interactions are much larger<br />

(Hinderliter et al., 2001). Therefore <strong>membrane</strong> <strong>proteins</strong> when incorporated in lipid<br />

<strong>membrane</strong>s can contribute decisively to the final lipid phase equilibria.<br />

In the presence <strong>of</strong> a multicomponent lipid mixture in the fluid phase, insertion <strong>of</strong> a<br />

<strong>membrane</strong> protein can lead, as already described, to enrichment <strong>of</strong> one <strong>of</strong> the lipid<br />

species around the protein due to more efficient packing in the protein-lipid interface.<br />

This preferential association can stabilize a lipid phase around the protein, even under<br />

conditions where it is metastable in the protein-free system. This preferred phase is said<br />

to wet the protein (Gil et al., 1998). The wetting layer <strong>of</strong> lipids only extends over a<br />

finite (nonmacroscopic) distance. Theoretical <strong>studies</strong> have predicted that effects on<br />

lipids by inclusion <strong>of</strong> <strong>proteins</strong> may extend up to 20 lipid shells around the protein (Fattal<br />

and Ben-Shaul, 1993; Sperotto and Mouritsen, 1991), and results from some<br />

experimental <strong>studies</strong> were rationalized on the basis <strong>of</strong> wetting <strong>of</strong> <strong>membrane</strong> <strong>proteins</strong> and<br />

formation <strong>of</strong> lipid phases induced by protein-lipid interactions (Lehtonen and Kinnunen,<br />

1997; Dumas et al., 1997; Hinderliter et al, 2004).<br />

43


In Figure I.21, theoretical simulations are shown that predict the wetting<br />

phenomenon. For Figure I.21c), sharing <strong>of</strong> a wetting layer by two or more <strong>proteins</strong> can<br />

give rise to colocalization and aggregation phenomena as described in detail on Section<br />

2.11.<br />

A B B<br />

Figure I.21 – Computer simulation snapshots <strong>of</strong> a large protein embedded in a binary lipid matrix in<br />

which one <strong>of</strong> the lipids is preferentially bound to the protein. (a) – Selective binding <strong>of</strong> one <strong>of</strong> the lipid<br />

species at the protein-lipid interface. (b) – Wetting <strong>of</strong> one <strong>of</strong> the lipids around the protein. (c) – Wetting<br />

and formation <strong>of</strong> a capillary condensate between two adjacent <strong>proteins</strong> (taken from Gil et al., 1998).<br />

These types <strong>of</strong> lateral structures initiated by <strong>proteins</strong> can function in bio<strong>membrane</strong>s<br />

as the signal for recruitment <strong>of</strong> other molecules to those domains. This has been<br />

proposed as the molecular mechanism behind some signalling platforms (Pérez-Gil et<br />

al., 2005).<br />

2.11. Lipid-mediation <strong>of</strong> protein-protein<br />

interactions<br />

The principle <strong>of</strong> hydrophobic matching, and the possibility <strong>of</strong> <strong>membrane</strong> <strong>proteins</strong> to<br />

change the lipid environment around them, <strong>of</strong>fer <strong>membrane</strong>s a mechanism for lipidmediation<br />

<strong>of</strong> protein-protein interactions.<br />

When <strong>proteins</strong> change the order parameter <strong>of</strong> lipids around them, overlap <strong>of</strong><br />

disturbed lipids can again lead to colocalization <strong>of</strong> <strong>proteins</strong> in the same area <strong>of</strong> the<br />

<strong>membrane</strong> or aggregation (Gil et al., 1998) (Figure I.22). The system, when moving to<br />

equilibrium, will have the tendency to decrease the extent <strong>of</strong> presented interface, and<br />

this is achievable by clustering <strong>of</strong> <strong>proteins</strong> and ultimately by protein aggregation. Of<br />

course, protein aggregation can have its energetic costs that may rise from helix-dipole<br />

moment and charge repulsion or entropic factors.<br />

44


INTRODUCTION: LIPID-PROTEIN INTERACTIONS<br />

Figure I.22 - Schematic representation <strong>of</strong> protein-protein interactions mediated by hydrophobic<br />

mismatch. (a) – Hydrophobically matched <strong>membrane</strong>. (b, c) - Lipid-mediated protein-protein attraction by<br />

same type <strong>of</strong> hydrophobic mismatch. (d) - Lipid-mediated protein-protein repulsion by opposite type <strong>of</strong><br />

hydrophobic mismatch (Taken from Gil et al., 1998).<br />

When more than one lipid species is present, wetting <strong>of</strong> <strong>membrane</strong> <strong>proteins</strong> also<br />

implies formation <strong>of</strong> an interface between lipids in the wetting phase and lipids in the<br />

protein-poor phase. Again the tendency to minimize the interface between these two<br />

phases can lead to coalescence or capillary condensation <strong>of</strong> wetting domains and<br />

stimulation <strong>of</strong> protein-protein interactions (Fig. I.21.c)). In this way, wetting occurs in<br />

addition to direct protein-protein (Van der Waals and electrostatic) and the above<br />

described lipid-mediated interactions, by selecting the environment in which they take<br />

place, i.e., wetting provides the mechanism for modulation <strong>of</strong> <strong>membrane</strong> protein<br />

organization (clustering and aggregation) on larger length scales (Gil et al., 1998).<br />

45


PROTEIN-PROTEIN AND PROTEIN-LIPID INTERACTIONS<br />

OF M13 MCP<br />

II<br />

1. Introduction<br />

PROTEIN-PROTEIN AND<br />

PROTEIN-LIPID<br />

INTERACTIONS OF M13<br />

MAJOR COAT PROTEIN<br />

The M(unich)13 bacteriophage was discovered in 1963 by P. H. H<strong>of</strong>schneider in an<br />

incubation <strong>of</strong> waste water samples <strong>with</strong> Escherichia coli bacteria. It belongs to the large<br />

family <strong>of</strong> filamentous phages (Inoviridae), which comprise some <strong>of</strong> the smallest<br />

bacteriophages known (Spruijt et al., 1999).<br />

These viruses are able to infect bacteria <strong>with</strong>out great disturbance <strong>of</strong> the host cells.<br />

In this way, reproduction <strong>of</strong> the virus inside the host goes on as cells continue to grow<br />

and divide even if at a reduced rate. For efficient infection, filamentous phages require<br />

expression <strong>of</strong> pili on the surface <strong>of</strong> the host bacteria. The pili acts as a receptor site, and<br />

depending on the specificity for pili interaction, filamentous bacteriophages are divided<br />

in different classes. The M13 bacteriophage belongs to the Ff group, which is only able<br />

to infect E. coli cells carrying sex pili (Spruijt et al., 1999). Genetic manipulation <strong>of</strong><br />

plasmids derived from the M13 bacteriophage proved to be relatively easy, and this<br />

allowed the M13 genome to be widely used as a cloning vector (Spruijt et al., 1999).<br />

One <strong>of</strong> the factors that contributed to the enormous popularity <strong>of</strong> the M13<br />

bacteriophage in biotechnology was its relatively simple architecture. The phage<br />

47


particle is 900 nm in length and 6.5 nm wide. It is formed by a single strand <strong>of</strong> DNA,<br />

about 6400 nucleotides long, encapsulated by a cylindrical protein coat. The protein<br />

coat presents about 3000 copies <strong>of</strong> the major coat protein (mcp or gp8), the product <strong>of</strong><br />

gene VIII. In addition to mcp, four other coat <strong>proteins</strong> are present in few copies, capping<br />

the ends <strong>of</strong> the bacteriophage.<br />

During the infection process, parental coat <strong>proteins</strong> are embedded in the <strong>membrane</strong><br />

while the viral DNA is released in the cytoplasm (see Figure II-1). New copies <strong>of</strong> mcp<br />

are synthesized in the host and inserted in the <strong>membrane</strong>, where they are processed by a<br />

leader peptidase (Kuhn et al., 1986). After peptidase action, the C-terminus <strong>of</strong> mcp is<br />

oriented to the cytoplasm and the N-terminus is located in the periplasm. In the mature<br />

form, mcp is a polypeptide chain 50 aminoacids long and presents a single<br />

trans<strong>membrane</strong> domain. During the reproductive cycle <strong>of</strong> the bacteriophage, coat<br />

<strong>proteins</strong> are stored at very high local levels and finally assembled around viral DNA<br />

into new phage particles (Hemminga et al., 1992).<br />

Figure II-1: Representation <strong>of</strong> the reproductive cycle <strong>of</strong> bacteriophage M13 (taken from Spruijt et<br />

al., 1999).<br />

mcp presents three distinct domains which are expected to be related to the distinct<br />

type <strong>of</strong> interactions that this protein is involved in during the reproductive cycle <strong>of</strong> the<br />

bacteriophage. In this way, the negatively charged region in the N-terminal domain <strong>of</strong><br />

the protein (residues 1-20), rich in glutamate and aspartate, forms the outer hydrophilic<br />

surface <strong>of</strong> the virus particle. The hydrophobic central domain (residues 21-39) is<br />

responsible for the tight <strong>membrane</strong> interaction <strong>of</strong> the <strong>membrane</strong> form <strong>of</strong> mcp and for the<br />

48


PROTEIN-PROTEIN AND PROTEIN-LIPID INTERACTIONS<br />

OF M13 MCP<br />

protein-protein interactions between mcp units inside the phage coat, which are essential<br />

in providing stability to the particle. Finally, the C-terminus <strong>of</strong> mcp (residues 40-50) is<br />

enriched in positively charged lysines. This domain lines the inside <strong>of</strong> the phage particle<br />

and is responsible by non-specific interactions <strong>with</strong> DNA. The amino acid sequence <strong>of</strong><br />

mcp and the structure <strong>of</strong> the <strong>membrane</strong> bound form is shown in Figure II-2.<br />

The N-terminus <strong>of</strong> the <strong>membrane</strong> bound form <strong>of</strong> mcp is a flexible α-helix <strong>with</strong><br />

amphipatic character. A loop connects this amphipatic domain <strong>with</strong> the trans<strong>membrane</strong><br />

helix (Figure II-2). The trans<strong>membrane</strong> domain <strong>of</strong> mcp is tilted by 20 ± 10º <strong>with</strong> respect<br />

to the lipid bilayer normal (Glaubitz et al., 2000; Koehorst et al., 2004). ESR and<br />

fluorescence <strong>studies</strong> making use <strong>of</strong> site-directed labelling <strong>of</strong> mcp (Spruijt et al., 1996;<br />

Stopar et al., 1997a) allowed to conclude that Thr 36 is located in the center <strong>of</strong> the bilayer<br />

in DOPC and 1,2-dioleoyl-sn-glycerol-3-phosphatidylglycerol (DOPG) bilayers, while<br />

aminoacids 25 and 46 delineate the edges <strong>of</strong> the trans<strong>membrane</strong> domain <strong>of</strong> mcp. DOPC<br />

is expected to present ideal hydrophobic matching conditions to mcp.<br />

A<br />

B<br />

Figure II-2: A - Amino acid sequence <strong>of</strong> mcp (taken from Hemminga et al., 1992). B – Structure <strong>of</strong><br />

the <strong>membrane</strong> bound form <strong>of</strong> mcp (taken from Stopar et al., 2003).<br />

mcp was also found to be strongly anchored to the lipid bilayer through its C-<br />

terminus, namely by the two phenylalanines found there and by three lysine residues.<br />

When mcp was incorporated in bilayers <strong>with</strong> slightly thinner or thicker hydrophobic<br />

49


thickness than mcp, the N-terminus <strong>of</strong> the protein shifted its position in order to<br />

accommodate the differences in hydrophobic lengths <strong>of</strong> protein and lipid, but the C-<br />

terminus remains at approximately the same position in the bilayer interface due to<br />

strong anchoring (Meijer et al., 2001). Another adaptation <strong>of</strong> mcp to hydrophobic<br />

mismatch is changing the tilt angle <strong>of</strong> the trans<strong>membrane</strong> domain, as well as the rotation<br />

<strong>of</strong> the helix at its long axis so as to optimize the hydrophobic and electrostatic<br />

interactions at the C-terminus (Koehorst et al., 2004).<br />

Several questions concerning lipid-protein and protein-protein interactions <strong>of</strong> mcp<br />

are still subject <strong>of</strong> debate. During the life-cycle <strong>of</strong> bacteriophage M13, the aggregation<br />

state <strong>of</strong> mcp changes repeatedly. In the phage particle it is aggregated in the coat, and in<br />

the <strong>membrane</strong> it is believed to be monomeric (Russel, 1991; Spruijt and Hemminga,<br />

1991), although a structural dimer might be formed as an intermediate species during<br />

phage disruption (Henry and Sykes, 1992; Stopar et al., 1997b; Stopar et al., 1998). The<br />

mcp structure is predominantly α-helical when inserted in the <strong>membrane</strong>, but depending<br />

on the protein purification and <strong>membrane</strong> reconstitution procedure a irreversibly<br />

aggregated β-sheet conformation <strong>of</strong> the protein appeared (Hemminga et al., 1992). Also,<br />

factors such as L/P ratio, phospholipid headgroup type, length and degree <strong>of</strong><br />

unsaturation <strong>of</strong> the acyl-chains <strong>of</strong> lipids used, and the ionic strength <strong>of</strong> the buffer were<br />

shown to be able to influence the α-helical to β-sheet transition. It is found that saturated<br />

lipid chains and very high protein concentration resulted inevitably in high levels <strong>of</strong><br />

irreversible β-sheet aggregation (Spruijt and Hemminga, 1991). Nevertheless, the<br />

irreversible β-sheet conformation <strong>of</strong> mcp is expected to be an artefact resulting from<br />

inappropriate conditions <strong>of</strong> protein purification and reconstitution, as α-helical structure<br />

is necessary for insertion in the phage particle (Hemminga et al., 1992).<br />

The lipid composition <strong>of</strong> the inner <strong>membrane</strong> <strong>of</strong> non infected E. coli is about 70% <strong>of</strong><br />

PE, 25% <strong>of</strong> PG and 5% cardiolipin. During the infection <strong>of</strong> E. coli by the M13<br />

bacteriophage, the levels <strong>of</strong> anionic lipids in the cell <strong>membrane</strong> are slightly increased<br />

and blocking <strong>of</strong> phage assembly resulted in significant increases in anionic lipid content<br />

in the inner <strong>membrane</strong> <strong>of</strong> bacteria (Pluschke et al., 1978). This phenomenon might<br />

indicate that anionic lipids are required to maintain a functional state <strong>of</strong> the <strong>membrane</strong>bound<br />

form <strong>of</strong> the protein (Hemminga et al., 1992).<br />

The in situ aggregational state <strong>of</strong> α-helical mcp is however still a matter <strong>of</strong><br />

discussion, as reversible oligomeric alpha-helical mcp was shown to exist in some<br />

conditions (Hemminga et al., 1992). In vitro, aggregation behaviour is even more likely<br />

50


PROTEIN-PROTEIN AND PROTEIN-LIPID INTERACTIONS<br />

OF M13 MCP<br />

than in vivo, as the reconstitution procedures normally result in randomized orientations<br />

for mcp instead <strong>of</strong> the single orientation observed in vivo. mcp in the <strong>membrane</strong> <strong>with</strong><br />

identical orientations are expected to present smaller protein-protein electrostatic<br />

attractive forces due to the stronger repulsion between the heavily charged C-terminus,<br />

and consequently a higher free energy is expected for the oligomeric condition. In order<br />

to achieve high local concentrations <strong>of</strong> mcp in the host <strong>membrane</strong> prior to phage<br />

assembly, a dynamic protein-lipid network should be present, characterized by absence<br />

<strong>of</strong> preferential association between coat <strong>proteins</strong> and/or lipids (Hemminga et al., 1992).<br />

Concerning the protein-lipid interactions established by mcp, it is interesting to note<br />

that the monomeric form <strong>of</strong> mcp was not able to induce sufficient immobilization <strong>of</strong> a<br />

fraction <strong>of</strong> lipid molecules so that it could be detected by ESR <strong>with</strong> spin labelled lipids<br />

(Sanders et al., 1992). On the other hand, oligomeric mcp immobilized a population <strong>of</strong><br />

lipids, especially at very high protein concentrations (Peelen et al., 1992). Thus, it can<br />

be concluded that monomeric M13 mcp is not able to induce a long living lipid<br />

boundary shell around it. Possible explanations are that the hydrophobic surface <strong>of</strong> mcp<br />

exposed to the lipids is too small or that the diffusion rate <strong>of</strong> the protein is not<br />

sufficiently lower than that <strong>of</strong> the lipids.<br />

51


DEPENDENCE OF M13 MCP OLIGOMERIZATION AND LATERAL<br />

SEGREGATION ON BILAYER COMPOSITION<br />

2. DEPENDENCE OF M13 MAJOR COAT<br />

PROTEIN OLIGOMERIZATION AND LATERAL<br />

SEGREGATION ON BILAYER COMPOSITION<br />

53


2430 <strong>Biophysical</strong> Journal Volume 85 October 2003 2430–2441<br />

Dependence <strong>of</strong> M13 Major Coat Protein Oligomerization and Lateral<br />

Segregation on Bilayer Composition<br />

Fábio Fernandes,* Luís M.S.Loura,* y Manuel Prieto,* Rob Koehorst, z<br />

Ruud B. Spruijt, z and Marcus A. Hemminga z<br />

*Centro de Química-Física Molecular, Instituto Superior Técnico, Lisboa, Portugal; y Departamento de Química, Universidade de Évora,<br />

Évora, Portugal; and z Laboratory <strong>of</strong> Biophysics, Wageningen University, Wageningen, The Netherlands<br />

ABSTRACT M13 major coat protein was derivatized <strong>with</strong> BODIPY (n-(4,4-difluoro-5,7-dimethyl-4-bora-3a,4a-diaza-sindacene-3-yl)methyl<br />

iodoacetamide), and its aggregation was studied in 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC)<br />

and DOPC/1,2-dioleoyl-sn-glycero-3-[phospho-rac-(1-glycerol)] (DOPG) or 1,2-dioleoyl-sn-glycero-3-phosphoethanolamine<br />

(DOPE)/DOPG (model systems <strong>of</strong> <strong>membrane</strong>s <strong>with</strong> hydrophobic thickness matching that <strong>of</strong> the protein) using photophysical<br />

methodologies (time-resolved and steady-state self-quenching, absorption, and emission spectra). It was concluded that the<br />

protein is essentially monomeric, even in the absence <strong>of</strong> anionic phospholipids. The protein was also incorporated in pure<br />

bilayers <strong>of</strong> lipids <strong>with</strong> a strong mismatch <strong>with</strong> the protein trans<strong>membrane</strong> length, 1,2-dierucoyl-sn-glycero-3-phosphocholine<br />

(DEuPC, longer lipid) and 1,2-dimyristoleoyl-sn-glycero-3-phosphocholine (DMoPC, shorter lipid), and in lipidic mixtures<br />

containing DOPC and one <strong>of</strong> these lipids. The protein was aggregated in the pure vesicles <strong>of</strong> mismatching lipid but remained<br />

essentially monomeric in the mixtures as detected from BODIPY fluorescence emission self-quenching. From fluorescence<br />

resonance energy transfer (FRET) measurements (donor-n-(iodoacetyl)aminoethyl-1-sulfonaphthylamine (IAEDANS)-labeled<br />

protein; acceptor-BODIPY labeled protein), it was concluded that in the DEuPC/DOPC and DMoPC/DOPC lipid mixtures,<br />

domains enriched in the protein and the matching lipid (DOPC) are formed.<br />

INTRODUCTION<br />

M13 coat protein has been shown to exist in many aggregation<br />

states, depending on factors like isolation, reconstitution<br />

procedure, pH, ionic strength, and amphiphile composition<br />

(Hemminga et al., 1993; Stopar et al., 1997). The mechanism<br />

<strong>of</strong> phage assembly in the Escherichia coli <strong>membrane</strong> is not<br />

yet completely understood, but the assembly site has been<br />

proposed to be composed <strong>of</strong> a dynamic protein-lipid network,<br />

characterized by absence <strong>of</strong> preferential association between<br />

M13 coat protein and/or lipids (Hemminga et al., 1993),<br />

which allows storage <strong>of</strong> monomeric coat protein at very high<br />

local concentrations, as the insertion <strong>of</strong> the protein in the<br />

assembling phage particle is only possible on the monomeric<br />

form (Russel, 1991). The type <strong>of</strong> interactions between lipids<br />

and coat protein which allows for formation <strong>of</strong> this structure is<br />

largely unknown and in this study it is intended to obtain<br />

information on the influence <strong>of</strong> lipid bilayer composition on<br />

the lateral distribution and oligomerization properties <strong>of</strong> M13<br />

coat protein in <strong>membrane</strong> model systems, to better understand<br />

the phage particle assembly mechanism.<br />

It has been proposed that self-association behavior <strong>of</strong><br />

trans<strong>membrane</strong> <strong>proteins</strong> while incorporated in lipid <strong>membrane</strong>s<br />

is influenced by hydrophobic matching conditions on<br />

the protein-lipid interface (Mouritsen and Bloom, 1984;<br />

Submitted January 21, 2003, and accepted for publication June 24, 2003.<br />

Address reprint requests to Manuel Prieto, Centro de Química-Física<br />

Molecular, Complexo I, Instituto Superior Técnico, Av. Rovisco Pais,<br />

1049-001 Portugal. Tel.: 35-121-841-9219; Fax: 35-121-846-4455; E-mail:<br />

prieto@alfa.ist.utl.pt.<br />

Ó 2003 by the <strong>Biophysical</strong> Society<br />

0006-3495/03/10/2430/12 $2.00<br />

Killian, 1998). The monomeric protein is expected to be<br />

stable under perfect matching conditions <strong>with</strong> the phospholipids<br />

surrounding it. In case <strong>of</strong> hydrophobic mismatch at the<br />

protein-lipid interface, it is possible that the boundary lipids<br />

reorganize themselves, to lower the tension created by exposure<br />

<strong>of</strong> hydrophobic acyl chains or amino-acid residues,<br />

which can be achieved by ordering/disordering <strong>of</strong> the phospholipids<br />

(Nezil and Bloom, 1992). Moreover, if the hydrophobic<br />

mismatch is too high for correction <strong>with</strong> small<br />

adjustments <strong>of</strong> bilayer hydrophobic thickness, this might<br />

result in protein aggregation to obtain minimization <strong>of</strong> the<br />

protein-lipid contacts (Ren et al., 1999; Lewis and Engelman,<br />

1983, Mobashery et al., 1997; Mall et al., 2001). Other<br />

responses to hydrophobic mismatch might be trans<strong>membrane</strong><br />

helix tilt (for long trans<strong>membrane</strong> hydrophobic protein<br />

domains), change in conformation or decrease <strong>of</strong> bilayer<br />

partitioning <strong>of</strong> the protein, transition <strong>of</strong> phospholipids to<br />

nonlamellar phases, and molecular sorting <strong>of</strong> <strong>proteins</strong> and<br />

lipids in the presence <strong>of</strong> a binary lipid system (for reviews<br />

see Killian, 1998; Dumas et al., 1999).<br />

Some <strong>studies</strong> have already related the oligomerization<br />

state <strong>of</strong> <strong>proteins</strong> to the hydrophobic mismatch extent in the<br />

bilayer. Aggregation was shown to depend on hydrophobic<br />

mismatch for bacteriorhodopsin at extreme hydrophobic<br />

mismatch conditions (Lewis and Engelman, 1983) and<br />

gramicidin (Mobashery et al., 1997). Mall and co-workers<br />

(Mall et al., 2001) concluded that, for a synthetic peptide, the<br />

free energy <strong>of</strong> dimerization increased linearly <strong>with</strong> each<br />

additional carbon <strong>of</strong> the phospholipid fatty acyl chain, <strong>with</strong><br />

a slope <strong>of</strong> 0.5 kJ mol ÿ1 . Apparently, this effect is much more<br />

significant upon negative hydrophobic thickness (thicker<br />

hydrophobic section for the bilayer than for the protein; Ren


M13 Coat Protein Lateral Distribution 2431<br />

et al., 1999; Lewis and Engelman, 1983; Mall et al., 2001).<br />

Meijer et al. (2001) found by electron spin resonance that the<br />

M13 major coat protein incorporated in di(22:1)PC (1,2-<br />

dierucoylphosphatidylcholine) aggregated or existed in several<br />

orientations/conformations, whereas in di(14:1)PC (1,2-<br />

dimyristoleoylphosphatidylcholine) no indication was found<br />

toward aggregation.<br />

In addition, for <strong>proteins</strong> incorporated in binary phospholipid<br />

mixtures, strong selectivity to one lipid component, and<br />

phase separation or lipid sorting effects (depending on their<br />

miscibility), was predicted to occur as a result <strong>of</strong> hydrophobic<br />

mismatch (Sperotto and Mouritsen, 1993). This phenomenon<br />

has also been observed experimentally in a mixture <strong>of</strong><br />

di(12:0)PC (dilauroylphosphatidylcholine)/di(18:0)PC(distearoylphosphatidylcholine),<br />

in which bacteriorhodopsin was<br />

shown to preferentially associate <strong>with</strong> the short chain lipid in<br />

the gel-gel and gel-fluid coexistence regions (Dumas et al.,<br />

1997). Although the preferential association <strong>of</strong> bacteriorhodopsin<br />

<strong>with</strong> short chain lipid in the gel-fluid coexistence<br />

region could be understood by an eventual preference for the<br />

disordered phase, these results were rationalized as being a<br />

consequence <strong>of</strong> lipid-protein hydrophobic matching interactions.<br />

A similar effect was observed for E. coli lactose<br />

permease (Lehtonen and Kinnunen, 1997), and for gramicidin<br />

(Fahsel et al., 2002). In the case <strong>of</strong> bacteriorhodopsin, Dumas<br />

and co-workers (Dumas et al., 1997), using a theoretical<br />

model, concluded that the protein was preferentially associated<br />

<strong>with</strong> the longer chain lipid on the mixed fluid lipid phase.<br />

In this mixture no macroscopic phase separation was occurring,<br />

but only preference <strong>of</strong> protein association <strong>with</strong> the phospholipid<br />

which allowed for more favorable energetic<br />

interactions, resulting in bacteriorhodopsin being surrounded<br />

by di(18:0)PC. This phenomenon is also denominated by<br />

wetting, and may extend to multiple layers <strong>of</strong> phospholipids<br />

(Gil et al., 1998). The interfacial stress, which is likely to<br />

occur between the wetting phase and the lipid matrix, can<br />

lead to sharing <strong>of</strong> these microdomains by many <strong>proteins</strong>,<br />

to minimize the interface area. As a result, formation <strong>of</strong> protein-enriched<br />

domains would occur in the bilayer. This<br />

process has been proposed as a mechanism for protein<br />

aggregation inducement (Gil et al., 1998).<br />

Some <strong>studies</strong> have also dealt <strong>with</strong> trans<strong>membrane</strong> protein/<br />

peptide interaction selectivity toward anionic phospholipids,<br />

based on electrostatic interactions <strong>of</strong> the lipid headgroup and<br />

basic residues on the protein (Horvàth et al., 1995a,b). A<br />

glucosyltransferase from Acholeplasma laidlawii was found<br />

to exhibit preference for localization on PG domains formed<br />

in a PC matrix (Karlsson et al., 1996).<br />

The purpose <strong>of</strong> this work is to study the influence <strong>of</strong> the<br />

bilayer composition on the aggregation state <strong>of</strong> the M13 coat<br />

protein. In addition, the protein was also incorporated in<br />

binary lipidic systems, where there is strong hydrophobic<br />

mismatch <strong>of</strong> the protein <strong>with</strong> one <strong>of</strong> the components. The<br />

possibility <strong>of</strong> protein segregation to lipidic domains in this<br />

situation was also investigated.<br />

For these purposes, several fluorescence methodologies<br />

(fluorescence self-quenching, absorption and emission spectra,<br />

and energy transfer) were applied, using the protein<br />

derivatized <strong>with</strong> n-(4,4-difluoro-5,7-dimethyl-4-bora-3a,4adiaza-s-indacene-3-yl)methyl<br />

iodoacetamide (BODIPY FL<br />

C 1 -IA) or n-(iodoacetyl)aminoethyl-1-sulfonaphthylamine<br />

(IAEDANS), as described in more detail below.<br />

Through the use <strong>of</strong> the self-quenching <strong>of</strong> the BODIPY<br />

fluorescence, it is expected to obtain information about the<br />

aggregation/oligomerization state <strong>of</strong> protein. With the same<br />

objective, the presence <strong>of</strong> BODIPY ground-state dimers is<br />

investigated. These techniques allow us to check for molecular<br />

contacts between BODIPY groups labeled on the trans<strong>membrane</strong><br />

section <strong>of</strong> the mutant protein.<br />

To solve the question <strong>of</strong> whether the presence <strong>of</strong> M13<br />

major coat protein is capable <strong>of</strong> inducing lipid domain<br />

formation through electrostatic interactions or hydrophobic<br />

mismatch, FRET measurements are carried out on different<br />

lipid mixtures <strong>with</strong> M13 major coat protein incorporated. As<br />

the C-terminal <strong>of</strong> the M13 major coat protein is heavily<br />

basic, it is intended to know if the presence <strong>of</strong> protein would<br />

lead to formation <strong>of</strong> domains enriched in anionic phospholipids<br />

and protein. Additionally, by using mixtures <strong>of</strong><br />

phospholipids <strong>with</strong> different acyl-chain lengths, formation<br />

<strong>of</strong> protein-enriched domains due to preferential binding to<br />

hydrophobically matching lipids is checked.<br />

In reconstituted systems we can have two different<br />

orientations for the M13 coat protein in the bilayer (<strong>with</strong><br />

the N-terminus sticking to the inside or the outside <strong>of</strong> the<br />

lipid vesicle), and in this way, interactions between <strong>proteins</strong><br />

might involve parallel or antiparallel protein orientations.<br />

For this reason, the mutants chosen for the present work were<br />

T36C and A35C, because for the coat protein in vesicles <strong>of</strong><br />

DOPC, the Thr36 and Ala35 residues are located near the<br />

center <strong>of</strong> the bilayer, as was shown by AEDANS wavelength<br />

<strong>of</strong> maximum emission (Spruijt et al., 2000). This increases<br />

the possibility <strong>of</strong> self-quenching due to the formation <strong>of</strong><br />

complexes or from collisions between fluorophores, and<br />

allowed us to ignore situations <strong>with</strong> complex oligomers<br />

containing <strong>proteins</strong> <strong>with</strong> parallel and antiparallel orientations<br />

as well as to simplify the energy transfer analysis to the twodimensional<br />

case, as, for both orientations, both residues<br />

should be located at approximately the same position.<br />

MATERIALS AND METHODS<br />

1,2-Dioleoyl-sn-glycero-3-phosphocholine (DOPC), 1,2-dioleoyl-sn-glycero-3-[phospho-rac-(1-glycerol)]<br />

(DOPG), 1,2-dioleoyl-sn-glycero-3-phosphoethanolamine<br />

(DOPE), 1,2-dierucoyl-sn-glycero-3-phosphocholine<br />

(DEuPC) and 1,2-dimyristoleoyl-sn-glycero-3-phosphocholine (DMoPC),<br />

were obtained from Avanti Polar Lipids (Birmingham, AL). N-(iodoacetyl)aminoethyl-1-sulfonaphthylamine<br />

(1,5-IAEDANS) and N-(4,4-difluoro-<br />

5,7-dimethyl-4-bora-3a,4a-diaza-s-indacene-3-yl)methyl) iodoacetamide<br />

(BODIPY FL C 1 -IA) were obtained from Molecular Probes (Eugene,<br />

OR). Fine chemicals were obtained from Merck (Darmstadt, Germany). All<br />

materials were used <strong>with</strong>out further purification.<br />

<strong>Biophysical</strong> Journal 85(4) 2430–2441


2432 Fernandes et al.<br />

Coat protein isolation and labeling<br />

The wild-type M13 major coat protein and the T36C mutant were grown,<br />

purified from the phage (Spruijt et al., 1996) and the A35C mutant was<br />

obtained from transformed cells <strong>of</strong> E. coli B21 (DE3) (Spruijt et al., 2000).<br />

Both mutants were labeled <strong>with</strong> BODIPY and IAEDANS as described<br />

previously (Spruijt et al., 1996). The remaining impurities were extracted<br />

using size exclusion chromatography and the protein was eluted <strong>with</strong> 50 mM<br />

sodium cholate, 150 mM NaCl, and 10 mM Tris-HCl pH 8.<br />

Coat protein reconstitution in lipid vesicles<br />

The labeled protein mutant was reconstituted in the DOPC/DOPG (80/20<br />

mol/mol), DOPC, DOPE/DOPG (70/30 mol/mol), DMoPC/DOPC (60/40<br />

mol/mol), DEuPC/DOPC (60/40 mol/mol), DMoPC, and DEuPC vesicles<br />

using the cholate-dialysis method (Spruijt et al., 1989). The phospholipid<br />

vesicles were produced as follows: the chlor<strong>of</strong>orm from solutions containing<br />

the desired phospholipid amount was evaporated under a stream <strong>of</strong> dry N 2<br />

and last traces removed by a further evaporation under vacuum. The lipids<br />

were then solubilized in 50 mM sodium cholate buffer (150 mM NaCl, 10<br />

mM Tris-HCl, and 1 mM EDTA) at pH 8 by brief sonication (Branson 250<br />

cell disruptor) until a clear opalescent solution was obtained, and then mixed<br />

<strong>with</strong> the wild-type and labeled protein. Samples had a phospholipid<br />

concentration <strong>of</strong> 1 mM and the lipid-to-protein ratio (L/P) was always kept at<br />

50, <strong>with</strong> addition <strong>of</strong> wild-type protein when necessary, except for the <strong>studies</strong><br />

<strong>with</strong> pure DMoPC and DEuPC bilayers in which the L/P was always \50<br />

(due to a smaller labeling efficiency obtained in the preparation <strong>of</strong> A35C<br />

mutant) and no wild-type protein was added. Dialysis was carried at room<br />

temperature and in the dark (to prevent IAEDANS degradation), <strong>with</strong> a 100-<br />

fold excess buffer containing 150 mM NaCl, 10 mM Tris-HCl, and 1 mM<br />

EDTA at pH 8. The buffer was replaced 53 every 12 h.<br />

Fluorescence spectroscopy<br />

Absorption spectroscopy was carried out <strong>with</strong> a Jasco V-560 spectrophotometer<br />

(Tokyo, Japan). The absorption <strong>of</strong> the samples was kept \0.1 at the<br />

wavelength used for excitation.<br />

CD spectroscopy was performed on a Jasco J-720 spectropolarimeter<br />

<strong>with</strong> a 450 W Xe lamp (Easton, MD).<br />

Steady-state fluorescence measurements were carried out <strong>with</strong> an SLM-<br />

Aminco 8100 Series 2 spectr<strong>of</strong>luorimeter (Rochester, NY; <strong>with</strong> double<br />

excitation and emission monochromators, MC400) in a right-angle<br />

geometry. The light source was a 450-W Xe arc lamp and for reference<br />

a Rhodamine B quantum counter solution was used. Correction <strong>of</strong> emission<br />

spectra was performed using the correction s<strong>of</strong>tware <strong>of</strong> the apparatus. 5 3 5<br />

mm quartz cuvettes were used. All measurements were performed at room<br />

temperature.<br />

In the fluorescence quenching measurements, the BODIPY emission<br />

spectra were recorded <strong>with</strong> an excitation wavelength <strong>of</strong> 460 nm and<br />

a bandwidth <strong>of</strong> 4 nm for both excitation and emission.<br />

AEDANS-labeled protein quantum yield was determined using quinine<br />

sulfate (f ¼ 0.55) (Eaton, 1988) as reference.<br />

The excitation wavelength for the energy transfer measurements was 340<br />

nm and the emission spectra were recorded <strong>with</strong> an excitation and emission<br />

bandwidth <strong>of</strong> 4 nm.<br />

Fluorescence decay measurements <strong>of</strong> IAEDANS were carried out <strong>with</strong><br />

a time-correlated single-photon counting system, which is described elsewhere<br />

(Loura et al., 2000). Measurements were performed at room<br />

temperature. Excitation and emission wavelengths were 340 and 440 nm,<br />

respectively. The timescales used were between 12 and 67 ps/ch, depending<br />

on the amount <strong>of</strong> BODIPY-labeled protein present in the sample. Data<br />

analysis was carried out using a nonlinear, least-squares iterative convolution<br />

method based on the Marquardt algorithm (Marquardt, 1963). The goodness<br />

<strong>of</strong> the fit was judged from the experimental x 2 -value, weighted residuals, and<br />

autocorrelation plot.<br />

In all cases, the probes florescence decay were complex and described by<br />

a sum <strong>of</strong> exponentials,<br />

where a i are the normalized amplitudes.<br />

The average lifetimes are defined by<br />

IðtÞ ¼+ a i expðÿt=t i Þ; (1)<br />

i<br />

hti ¼<br />

+ a i t 2 i<br />

i<br />

+<br />

i<br />

a i t i<br />

; (2)<br />

and the amplitude average or lifetime-weighted quantum yield (Lakowicz,<br />

1999), is obtained as<br />

t ¼ + a i t i : (3)<br />

i<br />

THEORETICAL BACKGROUND<br />

Fluorescence self-quenching<br />

Collisional quenching is responsible for a decrease in both<br />

the quantum yield and lifetime <strong>of</strong> a fluorophore, whereas<br />

static quenching only affects the fluorescence intensity due to<br />

the dark (‘‘nonfluorescent’’) character <strong>of</strong> the self-quenching<br />

complex. The effects <strong>of</strong> the collisional contribution <strong>of</strong> selfquenching<br />

on the fluorescence lifetime are described by the<br />

Stern-Volmer equation,<br />

t 0 =t ¼ 1 1 k q 3 t 0 3 ½FŠ; (4)<br />

where t 0 is the lifetime <strong>of</strong> the fluorophore in the absence <strong>of</strong><br />

quenching, t is the lifetime <strong>of</strong> the fluorophore in presence <strong>of</strong><br />

quenching, k q is the bimolecular diffusion rate constant, and<br />

[F] is the concentration <strong>of</strong> the fluorophore.<br />

The diffusion coefficient <strong>of</strong> the fluorophore (D) can be<br />

calculated from k q using the Smoluchowski equation<br />

(Lakowicz, 1999), taking into account transient effects<br />

(Umberger and Lamer, 1945),<br />

k q ¼ 4 3 p 3 N a 3 ð2 3 R c Þ 3 ð2 3 DÞ<br />

3 ½1 1 2 3 R c =ð2 3 t 0 3 DÞ 1=2 Š; (5)<br />

where N a is the Avogadro number and R c is the collisional<br />

radius. Diffusion in <strong>membrane</strong>s is here considered to take<br />

place in an isotropic three-dimensional medium. If the<br />

<strong>membrane</strong> were strictly bidimensional, different boundary<br />

conditions for the Smoluchowski formalism should be<br />

applied (Razi-Naqvi, 1974). The best approach to the<br />

specific situation <strong>of</strong> probe diffusion in a <strong>membrane</strong> is the<br />

one used by Owen (1975), in which the finite bilayer width is<br />

considered (cylindrical geometry). Owen introduced the<br />

parameter t s , which defines the crossover from the spherical<br />

(three-dimensional) to the cylindrical geometry, its value<br />

being t s ffi 42 ns when considering the bilayer and the<br />

peptide. This value is much longer than the longest<br />

fluorescence lifetime <strong>of</strong> the probes in the peptide (6.1 ns),<br />

<strong>Biophysical</strong> Journal 85(4) 2430–2441


M13 Coat Protein Lateral Distribution 2433<br />

and longer than the experimental time-window (28 ns ¼ 28<br />

ps/channel 3 1000 channels) so the three-dimensional<br />

framework approximation is essentially correct. Almgren<br />

(1991), in a comparative study <strong>of</strong> quenching in restricted<br />

dimensionality, also states that deviations from the threedimensional<br />

occur only for very long fluorescence lifetimes.<br />

The static quenching component can be described through<br />

a sphere <strong>of</strong> action, which accounts for statistical contact pairs<br />

formed at the moment <strong>of</strong> excitation. These contact pairs are<br />

nonfluorescent, although they do not form a complex, i.e.,<br />

the interaction energy is \k 3 T. The combined contributions<br />

<strong>of</strong> the collisional and the sphere-<strong>of</strong>-action effects on the<br />

fluorescence intensity are given by Loura et al. (2000) as<br />

I F ¼<br />

C 3 ½FŠ 3 expðÿV<br />

1<br />

s 3 N A 3 ½FŠÞ: (6)<br />

1 k q 3 ½FŠ<br />

t 0<br />

Here, I F is the fluorescence intensity, C is a constant, and V s<br />

is the sphere-<strong>of</strong>-action volume. The sphere-<strong>of</strong>-action radius<br />

is obtained by<br />

<br />

R s ¼ V s = 4 1=3<br />

3 3 p ; (7)<br />

and for a collisional quenching mechanism it should be close<br />

to the sum <strong>of</strong> the Van der Waals radii.<br />

In case that a complex is formed, the model to describe<br />

static quenching effects should take into account its equilibrium<br />

constant. For a monomer/dimer equilibrium <strong>of</strong> only<br />

one molecular species, the fluorescence intensity is given by<br />

I F ¼<br />

C 3 ½FŠ<br />

pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi<br />

ÿ1 1 1 1 8 3 K a 3 ½FŠ<br />

3 ; (8)<br />

1<br />

4 3 K<br />

1 k q 3 ½FŠ<br />

a<br />

t 0<br />

where K a is the oligomerization constant.<br />

However, for our system, in which there are two different<br />

protein species (labeled and unlabeled, where the unlabeled<br />

class includes both nonlabeled mutant and wild-type protein),<br />

there will be several combinations <strong>of</strong> protein species<br />

<strong>with</strong>in an aggregate. For a dimer, there will be three different<br />

combinations available—labeled protein/labeled protein,<br />

labeled protein/unlabeled protein, and unlabeled protein/unlabeled<br />

protein—but only formation <strong>of</strong> the first one induces<br />

self-quenching <strong>of</strong> BODIPY. A complexation model describing<br />

fluorescence static quenching in our system will have to<br />

account for this fact. From the knowledge <strong>of</strong> the concentration<br />

<strong>of</strong> each species, the fraction <strong>of</strong> labeled protein participating<br />

in oligomers (dimers/trimers) containing more than<br />

one BODIPY labeled protein (and as a result nonfluorescent),<br />

can be obtained for a given K a . The resulting set <strong>of</strong><br />

nonlinear equations was solved (for a given total protein<br />

concentration, labeling efficiency, labeled protein concentration,<br />

aggregation number, and K a ) using Maple V (Waterloo,<br />

Ontario).<br />

Fluorescence resonance energy transfer<br />

The energy transfer between fluorophores can be used to<br />

characterize the lateral distribution <strong>of</strong> labeled coat protein<br />

mutant in the bilayer. The degree <strong>of</strong> fluorescence emission<br />

quenching <strong>of</strong> the donor caused by the presence <strong>of</strong> acceptors<br />

is used to calculate the experimental energy transfer efficiency<br />

(E):<br />

E ¼ 1 ÿ I DA =I D : (9)<br />

Here, I DA is the donor fluorescence emission intensity in the<br />

presence <strong>of</strong> acceptor, and I D is the donor fluorescence<br />

emission intensity in the absence <strong>of</strong> acceptor.<br />

Wolber and Hudson (1979) obtained the analytical solution<br />

for energy transfer efficiencies in a random distribution <strong>of</strong><br />

acceptors in a bidimensional space,<br />

<br />

<br />

‘<br />

E ¼ 1 ÿ + ÿp 3 G 2 j G j <br />

3 R 2 0<br />

j¼0 3<br />

3 n 3 1 1<br />

2 3 ;<br />

j!<br />

(10)<br />

where R 0 is the Förster radius, G is the complete gamma<br />

function, and n 2 is the acceptor numerical density (number <strong>of</strong><br />

acceptors per unit area).<br />

The Förster radius is given by<br />

R 0 ¼ 0:2108 3 ðJ 3 k 2 3 n ÿ4 3 f D Þ 1=6 ; (11)<br />

where J is the spectral overlap integral, k 2 is the orientation<br />

factor, n is the refractive index <strong>of</strong> the medium, and f D is the<br />

donor quantum yield. J is calculated from<br />

ð<br />

J ¼ f ðlÞ 3 eðlÞ 3 l 4 3 dl; (12)<br />

where f(l) is the normalized emission spectra <strong>of</strong> the donor<br />

and e(l) is the absorption spectra <strong>of</strong> the acceptor. If the<br />

l-units in Eq. 12 are nm, the calculated R 0 in Eq. 11 has<br />

Å-units (Berberan-Santos and Prieto, 1987).<br />

RESULTS<br />

The M13 coat protein is known to form large irreversible<br />

aggregates under specific conditions. These aggregates are<br />

not found in vivo, and therefore are regarded as an artifact<br />

(Hemminga et al., 1993). Due to their large percentage <strong>of</strong><br />

b-sheet conformation, they can be detected by CD spectroscopy.<br />

Both the wild-type and labeled mutant protein<br />

conformation were checked by CD spectroscopy in all lipid<br />

systems used, and all spectra obtained were typical for<br />

a-helices (data not shown), indicating the absence <strong>of</strong> significant<br />

b-sheet protein conformation.<br />

The fluorescence decay <strong>of</strong> BODIPY in the labeled mutant<br />

<strong>proteins</strong> incorporated in DOPC, DMoPC/DOPC, DEuPC/<br />

DOPC, DMoPC, and DEuPC bilayers was described by two<br />

components t 1 ¼ 6.23 ns (a 1 ¼ 0.9) and t 2 ¼ 3.27 ns, which<br />

<strong>Biophysical</strong> Journal 85(4) 2430–2441


2434 Fernandes et al.<br />

leads to an average lifetime <strong>of</strong> ht 0 i¼6.1 ns (Eq. 2), as measured<br />

in samples <strong>with</strong> [BD-M13 coat protein] eff \10 ÿ3 M<br />

(BODIPY-labeled protein effective <strong>membrane</strong> concentration).<br />

For the determination <strong>of</strong> protein effective concentration<br />

in the <strong>membrane</strong>s, the lipid molar volumes were calculated<br />

from the reported lipid areas (72 Å 2 ) and <strong>membrane</strong> thicknesses<br />

(Tristam-Nagle et al., 1998; Lewis and Engelman,<br />

1983).<br />

Considering a random protein distribution in the bilayers,<br />

and fitting Eq. 4 to the experimental average lifetimes, linear<br />

Stern-Volmer plots are obtained (Fig. 1) and k q values are<br />

recovered (Table 1).<br />

The fluorescence steady-state quenching pr<strong>of</strong>iles for BD-<br />

M13 coat protein in DOPC, DMoPC/DOPC, DEuPC/DOPC,<br />

DMoPC, and DEuPC are presented in Fig. 2. This data can<br />

be fitted to Eq. 6 using the already-obtained k q -values to<br />

retrieve the fluorescence quenching contribution <strong>of</strong> the<br />

sphere-<strong>of</strong>-action effect (Table 1).<br />

Many <strong>studies</strong> on protein and peptide aggregation make<br />

use <strong>of</strong> fluorophore spectral changes induced by ground or<br />

excited state dimer formation (Otoda et al., 1993; Liu et al.,<br />

1998; Melnyk et al., 2002; Shigematsu et al., 2002). Recently,<br />

under restricted geometry, BODIPY was shown to<br />

form ground-state dimers in two different conformations (D j<br />

and D k ), <strong>with</strong> distinct spectroscopic properties from the<br />

monomer. The D j BODIPY dimer (parallel transition dipoles)<br />

is characterized by a strong excitation peak at 477 nm<br />

and a null quantum yield. The D k dimer (planar transition<br />

dipoles) exhibits an absorption peak at 570 nm and a broad<br />

fluorescence emission band centered at 630 nm (Bergström<br />

et al., 2001). In our BODIPY-derivatized protein none <strong>of</strong> the<br />

spectral properties assigned to the D j and D k dimers are<br />

observed, and both the excitation and fluorescence emission<br />

spectra obtained for the samples <strong>with</strong> higher concentration <strong>of</strong><br />

TABLE 1<br />

Systems k q /(10 9 M ÿ1 s ÿ1 ) D/(10 ÿ7 cm 2 s ÿ1 ) R s (Å)<br />

T36C in DOPC 2.3 0.7 14<br />

A35C in DOPC 1.7 0.4 14<br />

A35C in DMoPC 2.6 0.8 23<br />

A35C in DEuPC 5.0 2.6 27<br />

T36C in DMoPC/DOPC 6.6 4.2 14<br />

T36C in DEuPC/DOPC 20 22 14<br />

Bimolecular diffusion rate constants (k q ), diffusion coefficients (D), and<br />

apparent sphere-<strong>of</strong>-action radii (R s ) recovered from BODIPY fluorescence<br />

emission self-quenching from BODIPY-labeled T36C and A35C mutants.<br />

Values <strong>of</strong> R s [ 14 Å are evidence <strong>of</strong> molecular aggregation (see text for<br />

details).<br />

BODIPY-labeled protein were that expected for BODIPY<br />

monomers and identical in all lipid systems (Fig. 3).<br />

The donor-acceptor pair chosen for the energy transfer<br />

study shows a wide overlap <strong>of</strong> donor (AEDANS) fluorescence<br />

emission and acceptor (BODIPY) absorption (Fig. 3).<br />

Energy transfer <strong>studies</strong> were performed for the labeled<br />

protein incorporated in DOPC, DOPC/DOPG (80/20 mol/<br />

mol), DOPE/DOPG (70/30 mol/mol), DEuPC/DOPC (60/40<br />

mol/mol), and DMoPC/DOPC (60/40 mol/mol). It was intended<br />

to study the influence <strong>of</strong> electrostatic interactions,<br />

hydrophobic mismatch, and the presence <strong>of</strong> nonlamellar<br />

lipids in the aggregational and compartmentalization properties<br />

<strong>of</strong> the M13 coat protein. Due to the nonlamellar<br />

character <strong>of</strong> phosphatidylethanolamines, it was necessary to<br />

include a fraction <strong>of</strong> lamellar lipids (PG) in the lipid mixture<br />

for bilayer stabilization.<br />

The fluorescence emission spectra <strong>of</strong> the T36C mutant<br />

labeled <strong>with</strong> IAEDANS and reconstituted in these lipid<br />

systems were identical in the DOPC, DOPC/DOPG, DOPE/<br />

DOPG, DEuPC/DOPC, and DMoPC/DOPC lipid systems<br />

FIGURE 1 Transient state Stern-Volmer plots, describing BODIPY fluorescence self-quenching at different labeled protein concentrations. The lines are fits<br />

<strong>of</strong> Eq. 4 to the data. (A) Labeled protein incorporated in DOPC (m); DMoPC/DOPC (60/40 mol/mol) (); and DEuPC/DOPC (60/40 mol/mol) (d). (B) Labeled<br />

protein incorporated in DMoPC (n); DEuPC (n); and Stern-Volmer fit to DOPC data (see A) (- - -). The BODIPY-labeled mutant used in the DMoPC and<br />

DEuPC lipid systems was A35C, and for the other three lipid compositions the mutant used was T36C.<br />

<strong>Biophysical</strong> Journal 85(4) 2430–2441


M13 Coat Protein Lateral Distribution 2435<br />

FIGURE 2 Fluorescence steady-state data for BODIPY fluorescence self-quenching at different labeled protein concentrations. (A) Protein incorporated in<br />

DOPC (m); DMoPC/DOPC (60/40 mol/mol) (); and DEuPC/DOPC (60/40 mol/mol) (d). Eq. 6 is fitted to the data on the basis <strong>of</strong> dynamical quenching and<br />

a sphere-<strong>of</strong>-action quenching model (14.4 Å <strong>of</strong> radius) (—) for the protein in all lipid systems. (B) Protein incorporated in DOPC (m); DMoPC (); and DEuPC<br />

(). Eq. 6 fit <strong>of</strong> data from DOPC bilayers using a sphere-<strong>of</strong>-action radii <strong>of</strong> 14 Å (–), from DEuPC bilayers <strong>with</strong> a sphere-<strong>of</strong>-action radii <strong>of</strong> 27 Å (- - -), and from<br />

DMoPC bilayers using a sphere-<strong>of</strong>-action <strong>of</strong> 23 Å (–). These higher values are evidence <strong>of</strong> aggregation. See text for details.<br />

(results not shown), their wavelength <strong>of</strong> maximum fluorescence<br />

emission (477 nm) being characteristic <strong>of</strong> a very apolar<br />

environment (Hudson and Weber, 1973) and very similar to<br />

the one obtained by Spruijt and co-workers for the same<br />

mutant (478 nm; Spruijt et al., 2000). The wavelengths <strong>of</strong><br />

maximum emission for the IAEDANS-labeled protein in<br />

DMoPC and DEuPC were slightly different, 478 nm and 475<br />

nm, respectively.<br />

Donor fluorescence intensities (I D and I DA in Eq. 9)<br />

obtained by steady-state measurements and by integrated<br />

donor decays were identical. The IAEDANS-labeled protein<br />

quantum yield determined by us was f ¼ 0.64. Using Eqs.<br />

11 and 12, assuming k 2 ¼ 2/3 (the isotropic dynamic limit)<br />

and n ¼ 1.4 (Davenport et al., 1985), we obtain R 0 ¼ 48.8 Å<br />

FIGURE 3 Corrected excitation spectra <strong>of</strong> IAEDANS-labeled protein<br />

(—), and <strong>of</strong> BD-M13 coat protein (- - -). Corrected emission spectra <strong>of</strong><br />

IAEDANS-labeled protein (–), and <strong>of</strong> BD-M13 coat protein (- --).<br />

for this FRET pair. The value k 2 ¼ 2/3 was considered<br />

because for fluorophores in the center <strong>of</strong> a fluid bilayer, the<br />

rotational freedom should be sufficiently high to randomize<br />

orientations. This is supported by the reasonably low steadystate<br />

anisotropy values obtained for the IAEDANS and<br />

BODIPY probes labeled on the T36C M13 coat protein<br />

mutant (hri AEDANS ¼ 0.14, hri BODIPY ¼ 0.23; for a detailed<br />

discussion, see Loura et al., 1996).<br />

The results for BD-M13 coat protein in DOPC, DOPC/<br />

DOPG, DOPE/DOPG, DMoPC/DOPC, and DEuPC/DOPC<br />

bilayers are presented in Fig. 4.<br />

DISCUSSION<br />

For <strong>membrane</strong> <strong>proteins</strong> incorporated in lipid bilayers, the<br />

net tendency for protein aggregation should be weaker<br />

under good lipid-protein hydrophobic matching conditions<br />

(Mouritsen and Bloom, 1984). As the hydrophobic a-helix<br />

<strong>of</strong> the M13 coat protein is composed by 20 amino-acid<br />

residues, its length is ;30 Å. The lipids used in this work,<br />

<strong>with</strong> the exception <strong>of</strong> DMoPC (21 Å) and DEuPC (35 Å),<br />

form bilayers <strong>with</strong> almost the same hydrophobic thickness<br />

(28 Å) (Lewis and Engelman, 1983). Therefore, the oligomerization<br />

properties <strong>of</strong> the M13 major coat protein in<br />

DMoPC and DEuPC should differ from those in the hydrophobically<br />

matching phospholipid DOPC.<br />

The conditions <strong>of</strong> hydrophobic mismatch considered in<br />

this study do not seem to be able to induce any protein<br />

conformational change, as checked by CD spectroscopy in<br />

both lipid systems (no spectral change found). The absence<br />

<strong>of</strong> b-sheet conformation for the M13 coat protein implies no<br />

irreversible aggregation in the bilayer, and consequently any<br />

self-association observed should be due to reversible interactions<br />

between the a-helices.<br />

<strong>Biophysical</strong> Journal 85(4) 2430–2441


2436 Fernandes et al.<br />

FIGURE 4 (A) Donor (AEDANS) fluorescence quenching by energy transfer acceptor (BODIPY). Experimental data for DOPC (m); DOPC/DOPG (80/20<br />

mol/mol) (d); and DOPE/DOPG bilayers (70/30 mol/mol) (). Theoretical expectation (Eq. 10) for energy transfer in a random distribution <strong>of</strong> acceptors (—).<br />

Energy transfer simulation for a total co-localization <strong>of</strong> M13 coat protein in 20% (- - -), and 30% (— - —) <strong>of</strong> the surface area available. (B) Donor (AEDANS)<br />

fluorescence quenching by energy transfer acceptor (BODIPY). Theoretical expectation for energy transfer in a random distribution <strong>of</strong> acceptors (—). Energy<br />

transfer simulation for a segregation <strong>of</strong> M13 coat protein (mcp) to 60% <strong>of</strong> the surface area available (- - -). Experimental data for DEuPC/DOPC (60/40 mol/<br />

mol) (d); experimental data for DMoPC/DOPC (60/40 mol/mol) (). I DA and I D were obtained by integration <strong>of</strong> donor decays.<br />

The monitoring <strong>of</strong> IAEDANS maximum fluorescence<br />

emission (l max ) <strong>of</strong> the labeled mutant also rules out any<br />

change in conformation or exclusion from the bilayer <strong>of</strong> the<br />

M13 coat protein when incorporated in hydrophobically<br />

mismatching phospholipid, for the IAEDANS l max in these<br />

samples (475–478 nm) are almost identical to that observed<br />

<strong>with</strong> DOPC and the other hydrophobic matching phospholipids,<br />

and are typical for the fluorophore located near the<br />

center <strong>of</strong> the bilayer (Spruijt et al., 2000).<br />

Although BODIPY fluorescence decay is essentially<br />

monoexponential, the complex decay we obtained (dominated<br />

by a component <strong>of</strong> 6.3 ns) is also reported for derivatized<br />

<strong>proteins</strong> (Karolin et al., 1994).<br />

The BODIPY fluorescence emission self-quenching<br />

<strong>studies</strong> were performed <strong>with</strong> two different mutants. For the<br />

lipid mixtures T36C was used, and the A35C mutant was<br />

employed in the <strong>studies</strong> <strong>with</strong> pure mismatching lipid. As a<br />

control, the experiments were performed for both mutants in<br />

pure DOPC bilayers and the results were identical (Table 1).<br />

The bimolecular quenching constants calculated from the<br />

self-quenching results for BD-M13 coat protein incorporated<br />

in DOPC, DMoPC/DOPC (60/40 mol/mol) DEuPC/DOPC<br />

(60/40 mol/mol), DEuPC, and DMoPC allow the estimation<br />

<strong>of</strong> the labeled protein molecular diffusion coefficient through<br />

Eq. 5. Considering for BODIPY a collisional radius <strong>of</strong> 6 Å,<br />

we obtain for D BD-M13 coat protein in DOPC bilayers a value <strong>of</strong><br />

7.0 3 10 ÿ8 cm 2 s ÿ1 , which is the same order <strong>of</strong> magnitude <strong>of</strong><br />

the values <strong>of</strong> D for the M13 coat protein incorporated in fluid<br />

bilayers reported in the literature (Smith et al., 1979, 1980).<br />

However, for the lipid mixtures in which the predominant<br />

lipid does not hydrophobically match <strong>with</strong> the hydrophobic<br />

core <strong>of</strong> the M13 coat protein, the values obtained for D are<br />

unreasonably high. For the DMoPC/DOPC bilayers, a value<br />

<strong>of</strong> 4.2 3 10 ÿ7 cm 2 s ÿ1 is obtained, whereas in DEuPC/<br />

DOPC it is even higher (2.2 3 10 ÿ6 cm 2 s ÿ1 ). If D BD-M13 coat<br />

protein in pure vesicles <strong>of</strong> DOPC is considered to report a<br />

random distribution in the bilayer, the values in these mixtures<br />

are likely to be reporting protein segregation effects in<br />

the bilayer. We believe that this is caused by the hydrophobic<br />

mismatch constraints the protein finds when incorporated in<br />

bilayers that have too-long, or too-short phospholipids in<br />

their composition, probably leading to formation <strong>of</strong> localized<br />

areas <strong>with</strong> increased content <strong>of</strong> DOPC and protein. In this<br />

way, the effective apparent concentration in Eq. 4, should be<br />

higher than the one assumed on the basis <strong>of</strong> a random distribution,<br />

leading to an overestimation <strong>of</strong> k q , and so <strong>of</strong> D.<br />

However, the increase in local protein concentration arising<br />

from this effect would still be insufficient to explain the one<br />

order-<strong>of</strong>-magnitude increase <strong>of</strong> D from pure DOPC bilayers<br />

to the studied DEuPC/DOPC mixture (see below for further<br />

discussion).<br />

This rationalization is supported by D values obtained<br />

from BODIPY labeled protein in the pure mismatching lipid<br />

DMoPC and DEuPC (Table 1). These values are smaller than<br />

the ones obtained from the mixtures, and the diffusion<br />

coefficient in pure DMoPC is almost identical to the value in<br />

pure DOPC. For pure vesicles <strong>of</strong> DEuPC, D BD-M13 coat protein<br />

is larger than in DOPC, but it is still much smaller than the<br />

value obtained from the DEuPC/DOPC mixture. The results<br />

from dynamical self-quenching indicate therefore that although<br />

in pure vesicles <strong>of</strong> DEuPC there are already more<br />

collisions between BODIPY groups than what could be<br />

expected from a random distribution <strong>of</strong> labeled protein in the<br />

bilayer (probably due to aggregation), when DOPC is added<br />

the probability <strong>of</strong> collision greatly increases, and this can in<br />

part be explained in terms <strong>of</strong> protein segregation to DOPC-<br />

<strong>Biophysical</strong> Journal 85(4) 2430–2441


M13 Coat Protein Lateral Distribution 2437<br />

FIGURE 5 Fluorescence steady-state data for BODIPY fluorescence selfquenching<br />

at different labeled protein concentrations. (A) Protein incorporated<br />

in DOPC (m) and simulations considering protein oligomerization<br />

including static quenching: due to trimerization <strong>of</strong> the protein <strong>with</strong> a K a<br />

¼ 1000 ( 10% total mcp oligomerization) (—); due to dimerization <strong>of</strong> the<br />

protein <strong>with</strong> a K a ¼ 10 ( 25% total mcp oligomerization) (- - -). (B) Protein<br />

incorporated in DMoPC () and simulations for dimerization <strong>of</strong> protein <strong>with</strong><br />

a K a ¼ 20 ( 13% total mcp oligomerization for the most concentrated data<br />

point) (- - -) and trimerization <strong>of</strong> the protein <strong>with</strong> a K a ¼ 1300 ( 13% total<br />

mcp oligomerization for the most concentrated data point) (—). (C) Protein<br />

incorporated in DEuPC (d) and simulations for dimerization <strong>of</strong> protein <strong>with</strong><br />

a K a ¼ 30 ( 17% total mcp oligomerization for the most concentrated data<br />

enriched microdomains. In DMoPC/DOPC the effect is<br />

similar, but the bimolecular quenching constant is smaller<br />

than in DEuPC/DOPC.<br />

In Fig. 2, the obtained steady-state quenching pr<strong>of</strong>iles are<br />

presented together <strong>with</strong> the theoretical expectation for a<br />

sphere-<strong>of</strong>-action quenching model (Eq. 6). For the BODIPYlabeled<br />

protein in the DOPC-containing lipid systems<br />

(DOPC, DMoPC/DOPC, and DEuPC/DOPC) the results<br />

are well described using a sphere-<strong>of</strong>-action radius <strong>of</strong> 14 Å.<br />

For the pure mismatching lipids it is necessary to use larger<br />

radii to describe the results using Eq. 6.<br />

In Fig. 5, a simulation was included for a small degree <strong>of</strong><br />

protein aggregation in DOPC, DMoPC, and DEuPC<br />

bilayers. In these simulations it was considered that, due to<br />

the small degree <strong>of</strong> self-association considered, there was no<br />

change in M13 coat protein distribution and dynamics, and<br />

that in an oligomer, the fluorescence intensity <strong>of</strong> a BODIPY<br />

group is reduced to zero by the presence <strong>of</strong> another BODIPY<br />

group in the same aggregate. For DOPC bilayers, the<br />

prediction using a low fraction <strong>of</strong> aggregation (25% for<br />

dimerization and 10% for trimerization) clearly overestimates<br />

the extent <strong>of</strong> aggregation at the high labeled protein<br />

concentration, the range where this methodology is more<br />

sensitive. In agreement, from Fig. 2, it is clear that the data<br />

for the three DOPC-containing lipid systems are rationalized<br />

on the basis <strong>of</strong> dynamic quenching and a sphere <strong>of</strong> action,<br />

<strong>with</strong>out need for assumption <strong>of</strong> aggregation. The recovered<br />

radius (R s ¼ 14 Å) is close to the sum <strong>of</strong> the Van der Waals<br />

radii. These results indicate that BD-M13 coat protein in the<br />

studied DOPC-containing bilayers does not oligomerize.<br />

This conclusion is further supported by the absence <strong>of</strong><br />

BODIPY dimers in our samples, which would be revealed in<br />

the absorption/emission spectra.<br />

It could be possible that in a labeled protein oligomer,<br />

there is no contact between the BODIPY groups, and<br />

therefore the formation <strong>of</strong> oligomers would not necessarily<br />

induce fluorescence self-quenching. Some mutational <strong>studies</strong><br />

actually include the Thr36 residue in a lipid-interactive<br />

face <strong>of</strong> the trans<strong>membrane</strong> segment <strong>of</strong> the M13 coat protein.<br />

The M13 coat protein amino acids in this lipid-interactive<br />

face would only have contact <strong>with</strong> the phospholipid acyl<br />

chains inside the bilayer, and would be excluded from the<br />

trans<strong>membrane</strong> domain face involved in protein-protein interactions,<br />

where the amino acids responsible for the affinity<br />

<strong>of</strong> identical trans<strong>membrane</strong> segments are located (Deber<br />

et al., 1993; Webster and Haigh, 1998). That positioning <strong>of</strong><br />

the mutated Cys36 residue would make the contact between<br />

BODIPY groups from labeled <strong>proteins</strong> participating in an<br />

oligomer unlikely, and therefore exclude the formation <strong>of</strong><br />

‘‘dark’’ (nonfluorescent) dimers. Localization <strong>of</strong> the BOD-<br />

IPY group on the lipid-interactive face <strong>of</strong> the trans<strong>membrane</strong><br />

point) (- - -) and trimerization <strong>of</strong> the protein <strong>with</strong> a K a ¼ 7500 ( 25% total<br />

mcp oligomerization for the most concentrated data point) (—). See text for<br />

details on the simulations.<br />

<strong>Biophysical</strong> Journal 85(4) 2430–2441


2438 Fernandes et al.<br />

section <strong>of</strong> the protein can be seen as an advantage, inasmuch<br />

as the mutation <strong>of</strong> an amino-acid residue to cysteine after the<br />

labeling <strong>with</strong> fluorophore, at the protein-protein interactive<br />

face, could cause dramatic changes in the oligomerization<br />

behavior <strong>of</strong> the M13 coat protein. The T36C mutant labeled<br />

<strong>with</strong> BODIPY could be seen, for that reason, as a more ideal<br />

model for wild-type protein aggregation <strong>studies</strong> than the<br />

M13 coat protein mutant A35C, in which the mutated aminoacid<br />

residue is located in the protein-interactive face <strong>of</strong> the<br />

hydrophobic core domain. However, as can be concluded<br />

from the results obtained for both mutants labeled <strong>with</strong><br />

BODIPY in DOPC (Table 1), the positioning <strong>of</strong> the<br />

BODIPY group is not critical. Probably the fluorophore<br />

has sufficient mobility to extend from the surface <strong>of</strong> the<br />

trans<strong>membrane</strong> helix and probe a large extent <strong>of</strong> the area<br />

surrounding the protein.<br />

The BODIPY self-quenching results obtained in the present<br />

study clearly indicate that there is no protein aggregation<br />

in presence <strong>of</strong> hydrophobic matching phospholipids for all<br />

hydrophobic headgroup compositions used, and points to<br />

a large monomer stability under those conditions, which had<br />

already been suggested in other <strong>studies</strong> (Stopar et al., 1997;<br />

Spruijt et al., 1989; Sanders et al., 1991).<br />

Still, the results from self-quenching on pure bilayers<br />

<strong>of</strong> hydrophobic mismatching lipids (DEuPC and DMoPC)<br />

point to some aggregation, as the sphere-<strong>of</strong>-action radii<br />

recovered from the data fit to Eq. 6 were very unrealistic<br />

(27 and 23 Å for DEuPC and DMoPC, respectively).<br />

Simulations for BODIPY emission self-quenching due to<br />

aggregation are compared <strong>with</strong> the experimental data in these<br />

lipid systems in Fig. 5. DMoPC data could be reasonably<br />

described using a K a ¼ 20 for dimerization and a K a ¼ 1300<br />

for trimerization (13% <strong>of</strong> aggregated protein at the protein<br />

concentration <strong>of</strong> the most concentrated data point in both<br />

simulations).<br />

For the data obtained in DEuPC, the degree <strong>of</strong> aggregation<br />

is higher than in DMoPC (recovered sphere-<strong>of</strong>-action radius<br />

on fit to Eq. 6 was larger), and the fitting <strong>of</strong> the aggregation<br />

models to the data points proved more difficult. Probably the<br />

M13 major coat protein in DEuPC bilayers forms somewhat<br />

larger aggregates than in DMoPC and that could be an explanation<br />

for the increased BODIPY dynamical self-quenching<br />

observed in the longer lipid. This result is in agreement<br />

<strong>with</strong> the observations <strong>of</strong> Meijer et al. (2001), who, from electron<br />

spin resonance <strong>studies</strong>, reported that the protein appeared<br />

to exist in several orientations/conformations or in an aggregated<br />

form while incorporated in DEuPC bilayers. The larger<br />

extent <strong>of</strong> coat protein aggregation observed in the longer<br />

lipid bilayers can be explained by the fact that negative hydrophobic<br />

mismatch is considered to be energetically less<br />

favorable than positive mismatch (Killian, 1998; Mall et al.,<br />

2001).<br />

As mentioned above, for the DEuPC/DOPC and DMoPC/<br />

DOPC lipid mixtures used in the present study, no change in<br />

conformation or orientation was found (CD spectroscopy/<br />

AEDANS fluorescence emission spectra), and even for the<br />

protein in DEuPC/DOPC (60/40 mol/mol) no aggregation<br />

was detected (BODIPY self-quenching). These results point<br />

to stabilization <strong>of</strong> the protein by the hydrophobic matching<br />

phospholipid (DOPC), which was probably achieved by<br />

protein segregation to domains enriched in that phospholipid,<br />

partly explaining the high M13 coat protein apparent<br />

molecular diffusion coefficients obtained for the protein<br />

incorporated in DEuPC/DOPC and DMoPC/DOPC bilayers.<br />

Although an increase in BODIPY emission dynamical selfquenching<br />

was already visible for pure DEuPC bilayers<br />

(probably due to the large dimensions <strong>of</strong> the aggregates,<br />

which will have a similar effect as the segregation <strong>of</strong> protein<br />

to microdomains on inducing co-localization <strong>of</strong> protein in<br />

the bilayer), the bimolecular diffusion rate constant value in<br />

DEuPC/DOPC bilayers is much higher, and still, this can<br />

only be explained by segregation to DOPC-enriched microdomains.<br />

Being that M13 coat protein preferably locates in DOPCenriched<br />

domains, an interesting question is whether these<br />

domains are induced by the protein or exist even in the<br />

absence <strong>of</strong> protein (and the latter merely distributes differently<br />

among the pre-existing DOPC-rich and DOPC-poor<br />

regions). In this regard we obtained preliminary results <strong>of</strong><br />

1,6-diphenylhexatriene fluorescence anisotropy upon varying<br />

compositions for both DMoPC/DOPC and DEuPC/<br />

DOPC mixtures in the absence <strong>of</strong> protein (data not shown),<br />

pointing to the existence <strong>of</strong> ‘‘phase coexistence regions’’<br />

containing the compositions studied in this work. This<br />

experiment is not informative regarding the extent <strong>of</strong> phase<br />

separation, but it is compatible <strong>with</strong> the existence <strong>of</strong> small<br />

lipid clusters enriched in one <strong>of</strong> the components, because<br />

fluorescence anisotropy only senses the immediate vicinity<br />

<strong>of</strong> the fluorophore. Regarding the DEuPC/DOPC mixtures,<br />

these measurements imply that for 60:40 (mol/mol) DEuPC/<br />

DOPC, DEuPC-rich domains should coexist <strong>with</strong> DOPCrich<br />

domains, and the latter should account for approximately<br />

one-third <strong>of</strong> the mixture (not shown). From this<br />

result, we could expect at most a threefold increase in protein<br />

local concentration for this mixture, and an identical factor<br />

for the increase <strong>of</strong> the recovered k q value relative to that in<br />

DOPC. As seen in Table 1, this factor is significantly higher.<br />

This suggests that whereas protein segregation into DOPCrich<br />

domains should be occurring, it is not the sole cause for<br />

the increase in the apparent diffusion coefficient in 60:40<br />

(mol/mol) DEuPC/DOPC. In this regard the other contributing<br />

factors are not clear. It should be pointed out, however,<br />

that the estimate presented above for the fraction <strong>of</strong> DOPCrich<br />

domains was made from experiments performed in<br />

the absence <strong>of</strong> protein, whereas the quenching results refer to<br />

L/P ¼ 50.<br />

The centered position <strong>of</strong> BODIPY in the bilayer allows for<br />

a simplification <strong>of</strong> the energy transfer analysis for a twodimensional<br />

situation, as described by Eq. 10, i.e., there is no<br />

need to consider bilayer FRET geometry (Loura et al., 2001).<br />

<strong>Biophysical</strong> Journal 85(4) 2430–2441


M13 Coat Protein Lateral Distribution 2439<br />

Simulations <strong>of</strong> energy transfer for random distribution <strong>of</strong><br />

acceptors using Eq. 10 can therefore be compared <strong>with</strong> our<br />

experimental results (Fig. 4). The energy transfer efficiencies<br />

obtained for BD-M13 coat protein in the DMoPC/DOPC and<br />

DEuPC/DOPC bilayers support the other results discussed<br />

above for these mixtures, as they can only be explained by<br />

protein segregation in the bilayer or severe aggregation (Fig.<br />

4 B). However, as discussed above, the data obtained by<br />

fluorescence emission self-quenching indicate that segregation<br />

into DOPC-enriched domains (rather than aggregation)<br />

is the major phenomenon in these lipid mixtures.<br />

Although the results can be reasonably explained on the<br />

basis <strong>of</strong> protein segregation to 60% <strong>of</strong> the total bilayer area<br />

(Fig. 5 B), this rationalization should be considered an oversimplification,<br />

and is presented as an illustration. Indeed, the<br />

measured efficiencies are only reporting the average BD-<br />

M13 coat protein surface density that each IAEDANSlabeled<br />

protein is sensing. Probably there will be M13 coat<br />

protein interacting <strong>with</strong> the hydrophobically mismatching<br />

phospholipids, but the majority <strong>of</strong> the <strong>proteins</strong> will be<br />

preferentially surrounded by DOPC, and microdomains<br />

enriched in DOPC and M13 coat protein should be formed.<br />

The lipid mixtures used in this work for hydrophobic<br />

matching <strong>studies</strong> (DMoPC/DOPC and DEuPC/DOPC) are<br />

thought to be considerably closer to ideality than the ones<br />

used in the <strong>studies</strong> <strong>of</strong> Dumas and co-workers (gel/fluid<br />

coexistence; Dumas et al., 1997) and Lehtonen and Kinunnen<br />

(natural and pyrene-derivatized lipids; Lehtonen and<br />

Kununnen, 1997). However, protein segregation was observed<br />

in our study for both DMoPC/DOPC and DEuPC/<br />

DOPC, whereas in DOPC random distribution <strong>of</strong> protein<br />

in the bilayer was confirmed. Also interestingly, the degree<br />

<strong>of</strong> segregation appears to be similar for both mixtures (Fig.<br />

5 B), which was not expected due to the observed larger<br />

aggregation degree <strong>of</strong> coat protein in DEuPC, and therefore<br />

to an apparent larger packing difficulty <strong>with</strong> the longer<br />

lipid.<br />

The formation <strong>of</strong> local structure <strong>with</strong>in the thermodynamic<br />

fluid phase for a mixture <strong>of</strong> two PCs <strong>with</strong> a 4-carbon<br />

difference in acyl-chain lengths (DMPC/DSPC) was theoretically<br />

predicted by Mouritsen and Jørgensen (1994). The<br />

microdomains size in the Monte Carlo configurations<br />

obtained by these authors appears to be very small (10–20<br />

molecules at most). However, significant alterations in FRET<br />

efficiency as measured by Lehtonen et al. (1996) in mixtures<br />

<strong>of</strong> unsaturated PCs, and indeed our study <strong>of</strong> FRET between<br />

IAEDANS-labeled protein and BODIPY-labeled protein,<br />

require that the domain size should be <strong>of</strong> the order <strong>of</strong> magnitude<br />

<strong>of</strong> R 0 , that is ffi5 nm. In our system, this large domain<br />

size might be a consequence <strong>of</strong> protein-induced phase separation.<br />

In this work we also studied whether similar heterogeneities<br />

could be induced by the presence <strong>of</strong> positively<br />

charged M13 coat protein in bilayers composed <strong>of</strong> mixtures<br />

<strong>of</strong> anionic and neutral phospholipids. Due to the basic character<br />

<strong>of</strong> M13 coat protein C-terminal, it is reasonable to<br />

consider the possibility <strong>of</strong> anionic phospholipid-enriched<br />

domain induction by M13 coat protein incorporation in the<br />

bilayer. The formation <strong>of</strong> these domains could actually help<br />

explain some <strong>of</strong> the mechanisms involved in the creation <strong>of</strong><br />

the phage assembly site.<br />

Assuming that the hypothetical domains were composed<br />

by all the protein and DOPG content in the sample, and<br />

that the protein would be randomly distributed inside them,<br />

we can, using Eq. 10, obtain theoretical curves describing<br />

the energy transfer <strong>with</strong>in these domains. These plots are<br />

compared <strong>with</strong> the experimental data points for M13 coat<br />

protein incorporated in DOPC/DOPG (80/20 mol/mol) and<br />

DOPE/DOPG (70:30 mol/mol) in Fig. 5 A.<br />

It is concluded that the segregation <strong>of</strong> M13 coat protein to<br />

a PG-rich phase in the mixed systems, induced by electrostatic<br />

interactions between the positively charged protein and<br />

the negatively charged phospholipid, is ruled out on the basis<br />

<strong>of</strong> the obtained data, for this process would lead locally to<br />

greater surface densities <strong>of</strong> acceptor, and therefore, a very<br />

significant increase <strong>of</strong> energy transfer efficiencies should be<br />

expected in these conditions.<br />

In addition to being the predominant phospholipids <strong>of</strong> the<br />

E. coli inner <strong>membrane</strong> (Woolford et al., 1974; Burnell et al.,<br />

1980), nonlamellar phospholipids (such as DOPE) are<br />

known to interact distinctly from lamellar lipids <strong>with</strong> <strong>proteins</strong>,<br />

and in some cases to influence their conformation and<br />

activity (Hunter et al., 1999; Ahn and Kim, 1998). Our<br />

results suggest that these phospholipids have apparently no<br />

effect on the lateral distribution (Fig. 4 A) <strong>of</strong> the protein,<br />

which is kept random at the total protein concentration used.<br />

Higher protein concentrations than L/P 50 were not tested<br />

due to the risk <strong>of</strong> inducing nonlamellar phases, but it is likely<br />

that at smaller L/P ratios than those used in this work<br />

attractive interactions between monomers <strong>of</strong> M13 coat<br />

protein could take place, especially as we have two protein<br />

orientations in these reconstituted systems (parallel/antiparallel),<br />

and the repulsive electrostatic forces between the<br />

heavily basic C-terminal <strong>of</strong> the protein would be eliminated.<br />

In vivo, as there is only one orientation, the achievement <strong>of</strong><br />

high protein concentrations in the monomeric state at the<br />

assembly site would be more favorable (Hemminga et al.,<br />

1993). The presence <strong>of</strong> anionic phospholipids, as shown, is<br />

not essential for monomer stabilization, and the increase<br />

in phosphatidylglycerol and cardiolipin production during<br />

virus infection (Chamberlain et al., 1978) should only play<br />

a stabilizing role at the very high M13 coat protein concentrations<br />

that are expected to exist at the phage assembly site.<br />

CONCLUSIONS<br />

From this study, it is concluded that the M13 coat protein<br />

monomeric state is highly stable when incorporated in<br />

bilayers containing hydrophobic matching phospholipids.<br />

The lack <strong>of</strong> anionic phospholipids has no effect on the pro-<br />

<strong>Biophysical</strong> Journal 85(4) 2430–2441


2440 Fernandes et al.<br />

tein oligomerization properties at the protein concentrations<br />

used in this study. When the protein is incorporated in pure<br />

vesicles <strong>of</strong> mismatching lipid there is evidence for protein<br />

aggregation but for mixtures <strong>of</strong> lipids containing<br />

both hydrophobically matching (DOPC) and mismatching<br />

(DEuPC or DMoPC) phospholipids, the protein probably<br />

segregates to domains enriched in DOPC, which can explain<br />

the stability <strong>of</strong> the monomeric species <strong>of</strong> the protein in these<br />

lipid systems. This segregation effect is only observed when<br />

hydrophobic mismatching phospholipids are present, suggesting<br />

that the hydrophobic matching conditions on the<br />

protein-lipid interface are more important than electrostatic<br />

interactions between the M13 coat protein and the<br />

phospholipids, for the protein lateral distribution on the<br />

bilayer.<br />

The authors thank Dr. A. Fedorov for assistance in the time-resolved<br />

measurements.<br />

F.F. acknowledges financial support from Fundação para a Ciência e<br />

Tecnologia (FCT), project POCTI/36458/QUI/2000 and COST Action<br />

D:22; L.M.S.L. and M.P. acknowledge financial support from FCT,<br />

projects POCTI/36458/QUI/2000 and POCTI/36389/FCB/2000.<br />

REFERENCES<br />

Ahn, T., and H. Kim. 1998. Effects <strong>of</strong> non-lamellar-prone lipids on the<br />

ATPase activity <strong>of</strong> SecA bound to model <strong>membrane</strong>s. J. Biol. Chem.<br />

273:21692–21698.<br />

Almgren, M. 1991. Kinetics <strong>of</strong> excited states processes in micellar media.<br />

In Kinetics and Catalysis in Microheterogeneous Systems. M. Gratzel,<br />

and K. Kalyanasundaram, editors. Marcel Dekker, New York. 63–113.<br />

Berberan-Santos, M. N., and M. J. E. Prieto. 1987. Energy transfer in<br />

spherical geometry. Application to micelles. J. Chem. Soc. Faraday<br />

Trans. 283:1391–1409.<br />

Bergström, F., I. Mikhalyov, P. Hägglöf, R. Wortmann, T. Ny, and L. B.-Å.<br />

Johansson. 2001. Dimers <strong>of</strong> dipyrrometheneboron difluoride (BODIPY)<br />

<strong>with</strong> light spectroscopic applications in chemistry and biology. J. Am.<br />

Chem. Soc. 124:196–204.<br />

Burnell, E., L. Van Alphen, A. Verkleij, and B. De Kruijff. 1980. 31P<br />

Nuclear magnetic resonance and freeze-fracture electron microscopy<br />

<strong>studies</strong> on Escherichia coli. I. Cytoplasmic <strong>membrane</strong> and total<br />

phospholipids. Biochim. Biophys. Acta. 597:492–501.<br />

Chamberlain, B. K., Y. Nozaki, C. Tanford, and R. E. Webster. 1978.<br />

Association <strong>of</strong> the major coat protein <strong>of</strong> FD bacteriophage <strong>with</strong><br />

phospholipid vesicles. Biochim. Biophys. Acta. 510:18–37.<br />

Davenport, L., R. E. Dale, R. H. Bisby, and R. B. Cundall. 1985.<br />

Transverse location <strong>of</strong> the fluorescent probe 1,6-diphenyl-1,3,5-hexatriene<br />

in model lipid bilayer <strong>membrane</strong> systems by resonance energy<br />

transfer. Biochemistry. 24:4097–4108.<br />

Deber, C. M., A. R. Khan, Z. Li, C. Joensson, and M. Glibowicka. 1993.<br />

Val ! Ala mutations selectively alter helix-helix packing in the<br />

trans<strong>membrane</strong> segment <strong>of</strong> phage M13 coat protein. Proc. Natl. Acad.<br />

Sci. USA. 90:11648–11652.<br />

Dumas, F., M. M. Sperotto, M.-C. Lebrun, J.-F. Tocanne, and O. G.<br />

Mouritsen. 1997. Molecular sorting <strong>of</strong> lipids by bacteriorhodopsin in<br />

dilauroylphosphatidylcholine/distearoylphosphatidylcholine lipid bilayers.<br />

Biophys. J. 73:1940–1953.<br />

Dumas, F., M. C. Lebrun, and J.-F. Tocanne. 1999. Is the protein/lipid<br />

hydrophobic matching principle relevant to <strong>membrane</strong> organization and<br />

functions FEBS Lett. 458:271–277.<br />

Eaton, D. F. 1988. Reference materials for fluorescence measurement. Pure<br />

Appl. Chem. 60:1107–1114.<br />

Fahsel, S., E. M. Pospiech, M. Zein, T. L. Hazlet, E. Gratton, and<br />

R. Winter. 2002. Modulation <strong>of</strong> concentration fluctuations in phaseseparated<br />

lipid <strong>membrane</strong>s by polypeptide insertion. Biophys. J. 83:<br />

334–344.<br />

Gil, T., J. H. Ipsen, O. G. Mouritsen, M. C. Sabra, M. M. Sperotto, and<br />

M. J. Zuckermann. 1998. Theoretical analysis <strong>of</strong> protein organization in<br />

lipid bilayers. Biochim. Biophys. Acta. 1376:245–266.<br />

Hemminga, M. A., J. C. Sanders, C. J. A. M. Wolfs, and R. B. Spruijt.<br />

1993. Lipid-protein interactions involved in bacteriophage M13 infection.<br />

In Protein-Lipid <strong>Interaction</strong>s, New Comprehensive Biochemistry,<br />

Vol. 25. A. Watts, editor. Elsevier, Amsterdam. 191–212.<br />

Horvàth, L. I., T. Heimburg, P. Kovachev, J. B. Findlay, K. Hideg, and<br />

D. Marsh. 1995a. Integration <strong>of</strong> a K 1 channel-associated peptide in a<br />

lipid bilayer: conformation, lipid-protein interactions, and rotational<br />

diffusion. Biochemistry. 34:3893–3898.<br />

Horvàth, L. I., P. F. Knowles, P. Kovachev, J. B. Findlay, and D. Marsh.<br />

1995b. A single-residue deletion alters the lipid selectivity <strong>of</strong> a K 1<br />

channel-associated peptide in the b-conformation; spin-label electron<br />

spin resonance <strong>studies</strong>. Biophys. J. 73:2588–2594.<br />

Hudson, E. N., and G. Weber. 1973. Synthesis and characterization <strong>of</strong> two<br />

fluorescent sulfhydryl reagents. Biochemistry. 12:4154–4161.<br />

Hunter, G. W., S. Negash, and T. C. Squier. 1999. Phosphatidylethanolamine<br />

modulates Ca-ATPase function and dynamics. Biochemistry.<br />

38:1356–1364.<br />

Karlsson, O. P., M. Rytomaa, A. Dahqvist, P. K. Kinnunnen, and<br />

A. Wieslander. 1996. Correlation between bilayer dynamics and activity<br />

<strong>of</strong> the diglucosyldiacylglycerol synthase from Acholeplasma laidlawii<br />

<strong>membrane</strong>s. Biochemistry. 35:10094–10102.<br />

Karolin, J., J. B.-Å. Johansson, L. Strandberg, and T. Ny. 1994.<br />

Fluorescence and absorption spectroscopic properties <strong>of</strong> dipyrrometheneboron<br />

difluoride (BODIPY) derivatives in lipids, lipid <strong>membrane</strong>s, and<br />

<strong>proteins</strong>. J. Am. Chem. Soc. 116:7801–7806.<br />

Killian, J. A. 1998. Hydrophobic mismatch between <strong>proteins</strong> and lipids in<br />

<strong>membrane</strong>s. Biochim. Biophys. Acta. 1376:401–416.<br />

Lakowicz, J. R. 1999. Principles <strong>of</strong> Fluorescence Spectroscopy, 2nd Ed.<br />

Kluwer Academic/Plenum Press, New York. Chap. 8.<br />

Lewis, B. A., and D. M. Engelman. 1983. Bacteriorhodopsin remains<br />

dispersed in fluid phospholipid bilayers over a wide range <strong>of</strong> bilayer<br />

thickness. J. Mol. Biol. 166:203–210.<br />

Lehtonen, J. Y. A., J. M. Holopainen, and P. K. J. Kinnunen. 1996.<br />

Evidence for the formation <strong>of</strong> microdomains in liquid crystalline large<br />

unilamellar vesicles caused by hydrophobic mismatch <strong>of</strong> the constituent<br />

phospholipids. Biophys. J. 70:1753–1760.<br />

Lehtonen, J. Y. A., and P. K. J. Kinnunen. 1997. Evidence for phospholipid<br />

microdomain formation in liquid crystalline liposomes reconstituted <strong>with</strong><br />

Escherichia coli lactose permease. Biophys. J. 72:1247–1257.<br />

Liu, W., Y. Chen, H. Watrob, S. G. Bartlett, L. Jen-Jacobson, and M. D.<br />

Barkley. 1998. N-termini <strong>of</strong> EcoRI restriction endonuclease dimers are in<br />

close proximity on the protein surface. Biochemistry. 37:15457–15465.<br />

Loura, L. M. S., A. Fedorov, and M. Prieto. 1996. Resonance energy<br />

transfer in a model system <strong>of</strong> <strong>membrane</strong>s: application to gel and liquid<br />

crystalline phases. Biophys. J. 71:1823–1836.<br />

Loura, L. M. S., A. Fedorov, and M. Prieto. 2000. Membrane probe<br />

distribution heterogeneity: a resonance energy transfer study. J. Phys.<br />

Chem. B. 104:6920–6931.<br />

Loura, L. M. S., R. F. M. de Almeida, and M. Prieto. 2001. Detection and<br />

characterization <strong>of</strong> <strong>membrane</strong> microheterogeneity by resonance energy<br />

transfer. J. Fluoresc. 11:197–209.<br />

Mall, S., R. Broadbridge, R. P. Sharma, J. M. East, and A. G. Lee. 2001.<br />

Self-association <strong>of</strong> model trans<strong>membrane</strong> helix is modulated by lipid<br />

structure. Biochemistry. 40:12379–12386.<br />

Marquardt, D. W. 1963. An algorithm for least-squares estimation <strong>of</strong> nonlinear<br />

parameters. J. Soc. Ind. Appl. Math. 11:431–441.<br />

Meijer, A. B., R. B. Spruijt, C. J. A. M. Wolfs, and M. A. Hemminga. 2001.<br />

Membrane-anchoring interactions <strong>of</strong> M13 major coat protein. Biochemistry.<br />

40:8815–8820.<br />

<strong>Biophysical</strong> Journal 85(4) 2430–2441


M13 Coat Protein Lateral Distribution 2441<br />

Melnyk, R., A. Partridge, and C. Deber. 2002. Trans<strong>membrane</strong> domain<br />

mediated self-assembly <strong>of</strong> major coat protein subunits from FF<br />

bacteriophage. J. Mol. Biol. 315:63–72.<br />

Mobashery, N., C. Nielsen, and O. S. Andersen. 1997. The conformational<br />

preference <strong>of</strong> gramicidin channels is a function <strong>of</strong> lipid bilayer thickness.<br />

FEBS Lett. 412:15–20.<br />

Mouritsen, O. G., and M. Bloom. 1984. Mattress model <strong>of</strong> lipid-protein<br />

interactions in <strong>membrane</strong>s. Biophys. J. 46:141–153.<br />

Mouritsen, O. G., and K. Jørgensen. 1994. Dynamical order and disorder in<br />

lipid bilayers. Chem. Phys. Lipids. 73:3–25.<br />

Nezil, F. A., and M. Bloom. 1992. Combined influence <strong>of</strong> cholesterol and<br />

synthetic amphiphilic <strong>peptides</strong> upon bilayer thickness in model<br />

<strong>membrane</strong>s. Biophys. J. 61:1176–1183.<br />

Otoda, K., S. Kimura, and Y. Imanishi. 1993. Orientation and aggregation<br />

<strong>of</strong> hydrophobic helical <strong>peptides</strong> in phospholipid bilayer <strong>membrane</strong>.<br />

Biochim. Biophys. Acta. 1150:1–8.<br />

Owen, C. S. 1975. Two-dimensional diffusion theory: cylindrical diffusion<br />

model applied to fluorescence quenching. J. Chem. Phys. 62:3204–3207.<br />

Razi-Naqvi, K. 1974. Diffusion-controlled reactions in two-dimensional<br />

fluids: discussion <strong>of</strong> measurements <strong>of</strong> lateral diffusion <strong>of</strong> lipids in<br />

biological <strong>membrane</strong>s. Chem. Phys. Lett. 28:280–284.<br />

Ren, J., S. Lew, J. Wang, and E. London. 1999. Control <strong>of</strong> the<br />

trans<strong>membrane</strong> orientation and interhelical interactions <strong>with</strong>in <strong>membrane</strong>s<br />

by hydrophobic helix length. Biochemistry. 38:5905–5912.<br />

Russel, M. 1991. Filamentous phage assembly. Mol. Microbiol. 5:1607–<br />

1613.<br />

Sanders, J. C., N. A. J. van Nuland, O. Edholm, and M. Hemminga. 1991.<br />

Conformational and aggregation <strong>of</strong> M13 coat protein studied by<br />

molecular dynamics. Biophys. Chem. 41:193–202.<br />

Shigematsu, D., M. Matsutani, T. Furuya, T. Kiyota, S. Lee, G. Sugihara,<br />

and S. Yamashita. 2002. Roles <strong>of</strong> peptide-peptide charge interaction and<br />

lipid phase separation in helix-helix association in lipid bilayer. Biochim.<br />

Biophys. Acta. 1564:271–280.<br />

Smith, L. M., B. A. Smith, and H. M. McConnell. 1979. Lateral diffusion <strong>of</strong><br />

M-13 coat protein in model <strong>membrane</strong>s. Biochemistry. 18:2256–2259.<br />

Smith, L. M., J. L. R. Rubenstein, J. W. Parce, and H. M. McConnel. 1980.<br />

Lateral diffusion <strong>of</strong> M-13 coat protein in mixtures <strong>of</strong> phosphatidylcholine<br />

and cholesterol. Biochemistry. 19:5907–5911.<br />

Sperotto, M. M., and O. G. Mouritsen. 1993. Lipid enrichment and<br />

selectivity <strong>of</strong> integral <strong>membrane</strong> <strong>proteins</strong> in two-component lipid<br />

bilayers. Eur. Biophys. J. 22:323–328.<br />

Spruijt, R. B., C. J. A. M. Wolfs, and M. A. Hemminga. 1989. Aggregation<br />

related conformational change <strong>of</strong> the <strong>membrane</strong>-associated coat protein<br />

<strong>of</strong> bacteriophage M13. Biochemistry. 28:9158–9165.<br />

Spruijt, R. B., C. J. A. M. Wolfs, J. W. G. Verver, and M. A. Hemminga.<br />

1996. Accessibility and environment probing using cysteine residues<br />

introduced along the putative trans<strong>membrane</strong> domain <strong>of</strong> the major coat<br />

protein <strong>of</strong> bacteriophage M13. Biochemistry. 35:10383–10391.<br />

Spruijt, R. B., A. B. Meijer, C. J. A. M. Wolfs, and M. A. Hemminga. 2000.<br />

Localization and rearrangement modulation <strong>of</strong> the N-terminal arm <strong>of</strong> the<br />

<strong>membrane</strong>-bound major coat protein <strong>of</strong> the bacteriophage M13. Biochim.<br />

Biophys. Acta. 1509:311–323.<br />

Stopar, D., R. B. Spruijt, C. J. A. M. Wolfs, and M. A. Hemminga. 1997. In<br />

situ aggregational state <strong>of</strong> M13 bacteriophage major coat protein in<br />

sodium cholate and lipid bilayers. Biochemistry. 36:12268–12275.<br />

Tristam-Nagle, S., H. I. Petrache, and J. F. Nagle. 1998. Structure and<br />

interactions <strong>of</strong> fully hydrated dioleoylphosphatidylcholine bilayers.<br />

Biophys. J. 75:917–925.<br />

Umberger, J. Q., and V. K. Lamer. 1945. The kinetics <strong>of</strong> diffusioncontrolled<br />

molecular and ionic reactions in solution as determined by<br />

measurements <strong>of</strong> the quenching <strong>of</strong> fluorescence. J. Am. Chem. Soc. 67:<br />

1099–1109.<br />

Webster, R. E., and N. G. Haigh. 1998. The major coat protein <strong>of</strong><br />

filamentous bacteriophage F1 specifically pairs in the bacterial<br />

cytoplasmic <strong>membrane</strong>. J. Mol. Biol. 279:19–29.<br />

Wolber, P. K., and B. S. Hudson. 1979. An analytical solution to the Förster<br />

energy transfer problem in two dimensions. Biophys. J. 28:197–210.<br />

Woolford, J. L., Jr., J. S. Cashman, and R. E. Webster. 1974. F1 Coat<br />

protein synthesis and altered phospholipid metabolism in F1-infected<br />

Escherichia coli. Virology. 58:544–560.<br />

<strong>Biophysical</strong> Journal 85(4) 2430–2441


QUANTIFICATION OF PROTEIN-LIPID SELECTIVITY USING FRET:<br />

APPLICATION TO THE M13 MCP<br />

3. QUANTIFICATION OF PROTEIN-LIPID<br />

SELECTIVITY USING FRET: APPLICATION<br />

TO THE M13 MAJOR COAT PROTEIN.<br />

67


344 <strong>Biophysical</strong> Journal Volume 87 July 2004 344–352<br />

Quantification <strong>of</strong> Protein-Lipid Selectivity using FRET: Application<br />

to the M13 Major Coat Protein<br />

Fábio Fernandes,* Luís M.S.Loura,* y Rob Koehorst, z Ruud B. Spruijt, z Marcus A. Hemminga, z<br />

Alexander Fedorov,* and Manuel Prieto*<br />

*Centro de Química-Física Molecular, Instituto Superior Técnico, Lisbon, Portugal; y Departamento de Química,<br />

Universidade de Évora, Évora, Portugal; and z Laboratory <strong>of</strong> Biophysics, Wageningen University, Wageningen, The Netherlands<br />

ABSTRACT Quantification <strong>of</strong> lipid selectivity by <strong>membrane</strong> <strong>proteins</strong> has been previously addressed mainly from electron spin<br />

resonance <strong>studies</strong>. We present here a new methodology for quantification <strong>of</strong> protein-lipid selectivity based on fluorescence<br />

resonance energy transfer. A mutant <strong>of</strong> M13 major coat protein was labeled <strong>with</strong> 7-diethylamino-3((4#iodoacetyl)amino)phenyl-<br />

4-methylcoumarin to be used as the donor in energy transfer <strong>studies</strong>. Phospholipids labeled <strong>with</strong> N-(7-nitro-2-1,3-<br />

benzoxadiazol-4-yl) were selected as the acceptors. The dependence <strong>of</strong> protein-lipid selectivity on both hydrophobic mismatch<br />

and headgroup family was determined. M13 major coat protein exhibited larger selectivity toward phospholipids which allow for<br />

a better hydrophobic matching. Increased selectivity was also observed for anionic phospholipids and the relative association<br />

constants agreed <strong>with</strong> the ones already presented in the literature and obtained through electron spin resonance <strong>studies</strong>. This<br />

result led us to conclude that fluorescence resonance energy transfer is a promising methodology in protein-lipid selectivity<br />

<strong>studies</strong>.<br />

INTRODUCTION<br />

For integral <strong>membrane</strong> <strong>proteins</strong>, the interaction <strong>with</strong> lipids is<br />

dependent, among other factors, on the hydrophobic segments<br />

and interface properties <strong>of</strong> both lipids and protein.<br />

Minimal bilayer perturbation is achieved when the hydrophobic<br />

length <strong>of</strong> the protein matches that <strong>of</strong> the surrounding<br />

lipid (Mouritsen and Bloom, 1984). Deviations from perfect<br />

hydrophobic matching at the protein-lipid interface create<br />

a tension arising from the exposure <strong>of</strong> hydrophobic residues<br />

or acyl-chains to the hydrophilic medium. It is considered<br />

that the protein-lipid system adapts to hydrophobic mismatch<br />

conditions through several alternative and nonexclusive<br />

strategies such as ordering or disordering <strong>of</strong> perturbed<br />

phospholipids (change in bilayer thickness), lipid phase<br />

transition to nonlamellar phases, protein oligomerization/<br />

aggregation (minimization <strong>of</strong> interface area), helices tilting<br />

or side-chain rotation <strong>of</strong> a helical terminal residue (reduction<br />

in effective hydrophobic length), change in protein orientation,<br />

and decrease in the bilayer partitioning <strong>of</strong> the protein<br />

(for reviews see Killian, 1998; Dumas et al., 1999).<br />

In the case <strong>of</strong> <strong>proteins</strong> incorporated in lipid systems<br />

containing lipids <strong>with</strong> different electrostatic properties or<br />

hydrophobic lengths, selectivity to one lipid component at<br />

the protein-lipid interface or preferential phase partitioning<br />

(depending on the lipid miscibility) may occur (Dumas et al.,<br />

1997, Lehtonen and Kinnunen, 1997; Fahsel et al., 2002).<br />

In this study, we focused on the process <strong>of</strong> lipid selectivity<br />

at the protein-lipid interface. The problem <strong>of</strong> protein-lipid<br />

selectivity quantification has been addressed mainly from<br />

electron spin resonance (ESR) <strong>studies</strong> (see Marsh and<br />

Horváth, 1998, for a review) or other techniques which focus<br />

only on the protein-lipid interface, like tryptophan fluorescence<br />

quenching by brominated phospholipids (Everett et al.,<br />

1986; Williamson et al., 2002; O’Keeffe et al., 2000). The<br />

results obtained from ESR <strong>studies</strong> agree well <strong>with</strong> an annular<br />

model for protein-lipid selectivity, in which only the first<br />

shell <strong>of</strong> lipids around the integral protein, and in direct<br />

contact <strong>with</strong> it, is significantly disturbed by the protein<br />

incorporation in the bilayer (Lee, 2003; Marsh and Horváth,<br />

1998).<br />

ESR results report the fraction <strong>of</strong> motionally restricted<br />

lipids, whereas fluorescence collisional quenching depends<br />

on molecular contact. On the other hand, fluorescence<br />

resonance energy transfer (FRET) only depends on donoracceptor<br />

distances and is an alternative technique to quantify<br />

lipid selectivity. Gutierrez-Merino derived approximate<br />

analytical expressions for the average rate <strong>of</strong> FRET (hk T i)<br />

in <strong>membrane</strong>s undergoing phase separation or protein<br />

aggregation (Gutierrez-Merino, 1981a,b) and extended this<br />

formalism to the study <strong>of</strong> protein-lipid selectivity (Gutierrez-<br />

Merino et al., 1987). His model has proved to be useful to the<br />

study <strong>of</strong> the lipid annulus around the oligomeric acetylcholine<br />

receptor (Bonini et al., 2002; Antollini et al., 1996).<br />

However, there are some limitations to the model, namely,<br />

the simplification that underlies the formalism, which<br />

consists <strong>of</strong> considering resonance energy transfer (RET)<br />

only to neighboring acceptor molecules. On the other hand,<br />

<strong>with</strong> the experimental observable being the average RET<br />

efficiency given by<br />

Submitted January 21, 2004, and accepted for publication March 5, 2004.<br />

Address reprint requests to Luís M.S. Loura, Centro de Química-Física<br />

Molecular, Complexo I, Instituto Superior Técnico, Av. Rovisco Pais,<br />

1049-001 Lisbon, Portugal. Tel.: 35-121-841-9219; Fax: 35-121-846-4455;<br />

E-mail: pclloura@alfa.ist.utl.pt.<br />

Ó 2004 by the <strong>Biophysical</strong> Society<br />

0006-3495/04/07/344/09 $2.00 doi: 10.1529/biophysj.104.040337


FRET Study <strong>of</strong> Protein-Lipid Selectivity 345<br />

<br />

k T<br />

hEi ¼ ; (1)<br />

k T 1 k D<br />

where k D is the donor intrinsic decay rate coefficient, the<br />

relation <strong>with</strong> hk T i is not straightforward. It is proposed that if<br />

the setting <strong>of</strong> experimental conditions is such that hEi is low<br />

(namely, hk T i is much smaller than k D ), then hEi ffihk T i/k D<br />

(Gutierrez-Merino, 1981a). However, accurate low RET<br />

efficiencies are difficult to measure experimentally.<br />

In the present work a new FRET formalism for an annular<br />

model <strong>of</strong> protein-lipid selectivity is proposed, and used in the<br />

quantification <strong>of</strong> M13 major coat protein selectivity toward<br />

different phospholipids. M13 major coat protein is the main<br />

protein component <strong>of</strong> the filamentous bacteriophage M13<br />

<strong>with</strong> ;2800 copies. It contains a single hydrophobic trans<strong>membrane</strong><br />

segment <strong>of</strong> ;20 amino-acid residues, apart from<br />

an amphipathic N-terminal arm and a heavily basic<br />

C-terminus <strong>with</strong> a high density <strong>of</strong> lysines (for reviews see<br />

Stopar et al., 2003; Hemminga et al., 1993).<br />

The present study is separated in two sections. The first<br />

section focuses on the effect <strong>of</strong> hydrophobic length, and<br />

the selectivity <strong>of</strong> M13 toward 1,2-dioleoyl-sn-glycero-3-<br />

phosphoethanolamine-N-(7-nitro-2-1,3-benzoxadiazol-4-yl)<br />

((18:1) 2 -PE-NBD) was determined in unsaturated phosphatidylcholine<br />

bilayers <strong>of</strong> different acyl chain lengths (14:1,<br />

18:1, and 22:1). Whereas for 18:1 chains the chain length<br />

matches the hydrophobic length <strong>of</strong> the protein, there is<br />

significant hydrophobic mismatch for the other lipids used.<br />

The second part deals <strong>with</strong> specificity <strong>of</strong> M13 major coat<br />

protein to different phospholipid headgroups, some zwitterionic<br />

and other negatively charged. The results are compared<br />

to the results from the other methodologies for quantification<br />

<strong>of</strong> protein-lipid selectivity. Conclusions on the validity <strong>of</strong> the<br />

annular model for the M13 coat protein interaction <strong>with</strong><br />

lipids are obtained.<br />

MATERIALS AND METHODS<br />

1,2-Dioleoyl-sn-glycero-3-phosphocholine (DOPC; (18:1) 2 -PC), 1,2-dierucoyl-sn-glycero-3-phosphocholine<br />

(DEuPC; (22:1) 2 -PC), 1,2-dimyristoleoyl-snglycero-3-phosphocholine<br />

(DMoPC; (14:1) 2 -PC), 1,2-dioleoyl-sn-glycero-<br />

3-phosphoethanolamine-N-(7-nitro-2-1,3-benzoxadiazol-4-yl) ((18:1) 2 -PE-<br />

NBD), 1-Oleoyl-2-[12-[(7-nitro-2-1,3-benzoxadiazol-4-yl)amino]dodecanoyl]-sn-Glycero-Phosphocholine<br />

(18:1-(12:0-NBD)-PC), 1-Oleoyl-2-[12-[(7-<br />

nitro-2-1,3-benzoxadiazol-4-yl)amino]dodecanoyl]-sn-Glycero-Phosphoethanolamine<br />

(18:1-(12:0-NBD)-PE), 1-Oleoyl-2-[12-[(7-nitro-2-1,3-benzoxadiazol-4-yl)amino]dodecanoyl]-sn-Glycero-Phosphoserine<br />

(18:1-(12:0-NBD)-<br />

PS) (Sodium salt), 1-Oleoyl-2-[12-[(7-nitro-2-1,3-benzoxadiazol-4-yl)amino]dodecanoyl]-sn-Glycero-Phosphate<br />

(18:1-(12:0-NBD)-PA) (Monosodium<br />

salt), and 1-Oleoyl-2-[12-[(7-nitro-2-1,3-benzoxadiazol-4-yl)amino]dodecanoyl]-sn-Glycero-3-[Phospho-rac-(1-glycerol)]<br />

(18:1-(12:0-NBD)-PG) (Sodium<br />

salt), were obtained from Avanti Polar Lipids (Birmingham, AL).<br />

7-diethylamino-3((4#iodoacetyl)amino)phenyl-4-methylcoumarin (DCIA)<br />

was from Molecular Probes (Eugene, OR). Fine chemicals were obtained<br />

from Merck (Darmstadt, Germany). All materials were used <strong>with</strong>out further<br />

purification.<br />

Coat protein isolation and labeling<br />

The T36C mutant <strong>of</strong> the M13 major coat protein was grown, purified from<br />

the phage and labeled <strong>with</strong> DCIA as described previously (Spruijt et al.,<br />

1996). For the removal <strong>of</strong> free label, DNA and other coat <strong>proteins</strong>, the<br />

mixture was applied to a Superdex 75 prep-grade HR 16/50 column<br />

(Pharmacia, Amersham Biosciences, Piscataway, NJ) and eluted <strong>with</strong> 50<br />

mM sodium cholate, 150 mM NaCl, and 10 mM Tris-HCl pH 8. Fractions<br />

<strong>with</strong> an A 280 /A 260 absorption ratio .1.5 were collected and concentrated by<br />

Amicon filtration (Amicon, Millipore, Bedford, MA).<br />

Coat protein reconstitution in lipid vesicles<br />

The labeled protein mutant was reconstituted in DOPC ((18:1) 2 -PC),<br />

DMoPC ((14:1) 2 -PC), and DEuPC ((22:1) 2 -PC) vesicles using the cholatedialysis<br />

method (Spruijt et al., 1989). The phospholipid vesicles were<br />

produced as follows: the chlor<strong>of</strong>orm from solutions containing the desired<br />

NBD labeled and unlabeled phospholipid amount was evaporated under<br />

a stream <strong>of</strong> dry N 2 and last traces were removed by further evaporation under<br />

vacuum. The lipids were then solubilized in 50 mM sodium cholate buffer<br />

(150 mM NaCl, 10 mM Tris-HCl, 1 mM EDTA) at pH 8 by brief sonication<br />

(Branson 250 cell disruptor, Branson Ultrasonics, Danbury, CT) until a clear<br />

opalescent solution was obtained, and then mixed <strong>with</strong> the wild-type and<br />

labeled protein. Samples had a phospholipid concentration between 0.5 and<br />

1 mM (phospholipid concentration was determined through the analysis <strong>of</strong><br />

inorganic phosphate according to McClare, 1971) and the lipid to protein<br />

ratio (L/P) was always kept at 700. Dialysis was carried out at room<br />

temperature and in the dark, <strong>with</strong> a 100-fold excess buffer containing<br />

150 mM NaCl, 10 mM Tris-HCl, 1 mM EDTA at pH 8. The buffer was<br />

replaced five times every 12 h.<br />

Fluorescence spectroscopy<br />

Absorption spectroscopy was carried out <strong>with</strong> a Jasco V-560 spectrophotometer<br />

(Tokyo, Japan). The absorption <strong>of</strong> the samples was kept ,0.1 at the<br />

wavelength used for excitation.<br />

Steady-state fluorescence measurements were obtained <strong>with</strong> an SLM-<br />

Aminco 8100 Series 2 spectr<strong>of</strong>luorimeter (Rochester, NY; <strong>with</strong> double<br />

excitation and emission monochromators, MC400) in a right-angle<br />

geometry. The light source was a 450-W Xe arc lamp and for reference<br />

a Rhodamine B quantum counter solution was used. 5 3 5 mm quartz<br />

cuvettes were used. All measurements were performed at room temperature.<br />

The quantum yield <strong>of</strong> DCIA-labeled protein was determined using<br />

quinine bisulfate dissolved in 1 N H 2 SO 4 (f ¼ 0.55; Eaton, 1988) as<br />

a reference.<br />

Fluorescence decay measurements <strong>of</strong> DCIA were carried out <strong>with</strong> a timecorrelated<br />

single-photon timing system, which is described elsewhere<br />

(Loura et al., 2000). Measurements were performed at room temperature.<br />

Excitation and emission wavelengths were 340 and 450 nm, respectively.<br />

The timescales used were between 3 and 12 ps/ch, depending on the amount<br />

<strong>of</strong> NBD-labeled phospholipid present in the sample. Data analysis was<br />

carried out using a nonlinear, least-squares iterative convolution method<br />

based on the Marquardt algorithm (Marquardt, 1963). The goodness <strong>of</strong> the<br />

fit was judged from the experimental x 2 value, weighted residuals, and<br />

autocorrelation plot.<br />

In all cases, the probe florescence decay was complex and described by<br />

a sum <strong>of</strong> exponentials,<br />

IðtÞ ¼+ a i expðÿt=t i Þ; (2)<br />

i<br />

where a i are the normalized amplitudes and t i are the fluorescence lifetimes.<br />

<strong>Biophysical</strong> Journal 87(1) 344–352


346 Fernandes et al.<br />

THEORETICAL BACKGROUND<br />

Fluorescence resonance energy transfer<br />

Fluorescence resonance energy transfer (FRET) can be used<br />

to characterize the lateral distribution <strong>of</strong> labeled coat protein<br />

mutants in the bilayer. In the case <strong>of</strong> energy heterotransfer,<br />

the degree <strong>of</strong> fluorescence emission quenching <strong>of</strong> the donor<br />

caused by the presence <strong>of</strong> acceptors is used to calculate the<br />

experimental energy transfer efficiency (E):<br />

E ¼ 1 ÿ t DA =t D : (3)<br />

Here t DA is the donor lifetime-weighted quantum yield in the<br />

presence <strong>of</strong> acceptor and t D is the donor lifetime-weighted<br />

quantum yield in the absence <strong>of</strong> acceptor. In turn, lifetimeweighted<br />

quantum yields are defined by Lakowicz (1999) as<br />

The Förster radius is given by<br />

t ¼ + a i t i : (4)<br />

i<br />

R 0 ¼ 0:2108ðJk 2 n ÿ4 f D Þ 1=6 ; (5)<br />

where J is the spectral overlap integral, k 2 is the orientation<br />

factor, n is the refractive index <strong>of</strong> the medium, and f D is the<br />

donor quantum yield. J is calculated as<br />

Z<br />

J ¼ f ðlÞeðlÞl 4 dl; (6)<br />

where f(l) is the normalized emission spectrum <strong>of</strong> the donor<br />

and e(l) is the absorption spectrum <strong>of</strong> the acceptor. If the<br />

l-units in Eq. 6 are nm, the calculated R 0 in Eq. 5 has Å units<br />

(Berberan-Santos and Prieto, 1987).<br />

The acceptors in the annular shell (Fig. 1) are at a constant<br />

distance (d) to the coumarin fluorophore located in the center<br />

<strong>of</strong> the trans<strong>membrane</strong> domain, and therefore we can assume<br />

that the energy transfer to each <strong>of</strong> these acceptors is described<br />

by the rate constant<br />

k r ¼ 1 6<br />

R 0<br />

; (8)<br />

t D d<br />

where t D is the donor lifetime (in the absence <strong>of</strong> acceptor).<br />

The NBD fluorophores in the acceptor probes used in this<br />

study (phospholipids labeled <strong>with</strong> NBD in the headgroup or<br />

in the acyl-chain) are assumed to be located in the bilayer<br />

surface. For the chain-labeled lipids, this is justified because<br />

the NBD group ‘‘loops up’’ to the surface when attached to<br />

the end <strong>of</strong> the phospholipids acyl-chain (Chattopadhyay,<br />

1990). The donor fluorophore is labeled in the M13 major<br />

coat protein 36th residue, located near the center <strong>of</strong> the<br />

bilayer (Spruijt et al., 1996). Therefore, to calculate d it is<br />

necessary to estimate the average distance (l) between the<br />

donor plane (center <strong>of</strong> bilayer) and the acceptors planes (both<br />

leaflets), as well as the lateral separation between both probes<br />

inside the trans<strong>membrane</strong> protein-annular shell lipids<br />

complex. For DOPC bilayers the position <strong>of</strong> the NBD<br />

fluorophore (l) in the derivatized phospholipids has been<br />

calculated through the parallax method (Abrams and<br />

London, 1993), and it was 18.9 and 19.8 Å from the bilayer<br />

center for the phospholipids labeled at the headgroup and at<br />

the acyl-chain, respectively. These values agree <strong>with</strong> other<br />

<strong>studies</strong> which employed different techniques to obtain the<br />

fluorophore position (Wolf et al., 1992; Màzeres et al.,<br />

1996). The reason for a position <strong>of</strong> NBD closer to the surface<br />

<strong>of</strong> the <strong>membrane</strong> while labeled at the acyl-chain is probably<br />

the increase in flexibility that the C12 chain allows. The<br />

Annular model for M13 coat protein selectivity<br />

toward phospholipids<br />

To analyze the FRET results, a model for trans<strong>membrane</strong><br />

protein selectivity toward phospholipids was derived. The<br />

model assumes two populations <strong>of</strong> energy transfer acceptors,<br />

one located in the annular shell around the protein and the<br />

other outside it. The donor fluorescence decay curve will<br />

have energy transfer contributions from both populations,<br />

i DA ðtÞ ¼i D ðtÞr annular ðtÞr random ðtÞ: (7)<br />

Here i D and i DA are the donor fluorescence decay in the<br />

absence and presence <strong>of</strong> acceptors respectively, and r annular<br />

and r random are the FRET contributions arising from energy<br />

transfer to annular labeled lipids and to randomly distributed<br />

labeled lipids outside the annular shell, respectively.<br />

FIGURE 1 Molecular model for the FRET analysis ((A) side view; (B) top<br />

view). Protein-lipid organization presents a hexagonal geometry. Donor<br />

fluorophore from the mutant protein is located in the center <strong>of</strong> the bilayer,<br />

whereas the acceptors are distributed in the bilayer surface. Two different<br />

environments are available for the labeled lipids (acceptors), the annular<br />

shell surrounding the protein and the bulk lipid. Energy transfer to acceptors<br />

in direct contact <strong>with</strong> the protein has a rate coefficient dependent on the<br />

distance between donor and annular acceptor (Eq. 8). Energy transfer toward<br />

acceptors in the bulk lipid is given by Eq. 11 (see text for details).<br />

<strong>Biophysical</strong> Journal 87(1) 344–352


FRET Study <strong>of</strong> Protein-Lipid Selectivity 347<br />

lateral separation between the probes was assumed to be 8 Å.<br />

The estimated d was then 20.5 and 21.4 Å for NBD<br />

derivatized in the headgroup and acyl-chain, respectively.<br />

For the DMoPC and DEuPC bilayers the values used for l<br />

were 15.4 and 22.4 Å, respectively, as for each additional<br />

carbon in the phospholipid chain in the liquid crystalline<br />

phase the bilayer thickness increases 1.75 Å (Lewis and<br />

Engelman, 1983).<br />

Considering a hexagonal-type geometry for the proteinlipid<br />

arrangement (Fig. 1 b), each protein will be surrounded<br />

by 12 annular lipids. In bilayers composed by both labeled<br />

and unlabeled phospholipids, these 12 sites will be available<br />

for both <strong>of</strong> them. The probability (m) <strong>of</strong> one <strong>of</strong> these sites to<br />

be occupied by labeled phospholipid is given by<br />

m ¼ K S<br />

n NBD<br />

n NBD 1 n lipid<br />

: (9)<br />

Here, n NBD is the concentration <strong>of</strong> labeled lipid, and n lipid is<br />

the concentration <strong>of</strong> unlabeled lipid. K S is the relative<br />

association constant, which reports the relative affinity <strong>of</strong><br />

the labeled and unlabeled phospholipid. Using a binomial<br />

distribution we can calculate the probability <strong>of</strong> each occupation<br />

number (0–12 sites occupied simultaneously by<br />

labeled lipid), and finally the FRET contribution arising from<br />

energy transfer to annular lipids,<br />

<br />

n¼12<br />

r annular ðtÞ ¼ + e ÿnk Tt 12<br />

m n ð1 ÿ mÞ 12ÿn : (10)<br />

n¼0<br />

n<br />

The FRET contribution from energy transfer to acceptors<br />

randomly distributed outside the annular region in two<br />

different planes at the same distance to the donor plane (from<br />

the center <strong>of</strong> the bilayer to both leaflets) is given by<br />

Davenport et al. (1985) as<br />

RESULTS<br />

M13 coat protein selectivity toward phospholipids<br />

<strong>with</strong> different hydrophobic thickness<br />

The DCIA-labeled protein quantum yield was determined<br />

(f ¼ 0.41). Using Eqs. 5 and 6, and assuming k 2 ¼ 2/3<br />

(the isotropic dynamic limit) and n ¼ 1.4 (Davenport et al.,<br />

1985), R 0 ¼ 39.3 Å is obtained for the DCIA-NBD FRET<br />

pair (Fig. 2). The value k 2 ¼ 2/3 was used, because for<br />

fluorophores in the center <strong>of</strong> a liquid crystalline bilayer, the<br />

rotational freedom should be sufficiently high to randomize<br />

orientations (for a detailed discussion see Loura et al., 1996).<br />

FRET selectivity <strong>studies</strong> were performed in bilayers <strong>of</strong><br />

one lipid component (DOPC, DMoPC, or DEuPC) using<br />

T36C M13 major coat protein mutant labeled <strong>with</strong> DCIA as<br />

the donor and (18:1) 2 -PE-NBD (1,2-dioleoyl-sn-glycero-<br />

3-phosphoethanolamine derivatized <strong>with</strong> NBD at the headgroup)<br />

as the acceptor.<br />

The donor fluorescence intensities ratio (t DA =t D ), which<br />

is related to the energy transfer efficiency, decreases upon<br />

increasing the acceptor (Eq. 3). The results are presented in<br />

Fig. 3. The results <strong>of</strong> fitting the derived formalism to the data<br />

are also shown in this figure, and the corresponding K S<br />

values are summarized in Table 1.<br />

M13 coat protein selectivity toward phospholipids<br />

<strong>with</strong> different headgroups<br />

Energy transfer <strong>studies</strong> were also performed to determine the<br />

selectivity properties <strong>of</strong> M13 coat protein toward phospholipids<br />

<strong>with</strong> different headgroups. Again, the donor was T36C<br />

coat protein mutant labeled <strong>with</strong> coumarin (DCIA), but<br />

various probes were used as acceptors, all <strong>studies</strong> being<br />

made in DOPC vesicles. The probes used as acceptors were<br />

r random<br />

( Z pffiffiffiffiffiffiffiffi<br />

)<br />

l<br />

¼ exp ÿ4n 2 pl 2 l 2 1 R 2 1 ÿ expðÿtb 3 a 6 Þ<br />

e<br />

a 3 da ;<br />

0<br />

(11)<br />

where b ¼ðR 2 0 =lÞ2 t ÿ1=3<br />

D ; n 2 is the acceptor density in each<br />

leaflet, l is the distance between the plane <strong>of</strong> the donors and<br />

the planes <strong>of</strong> acceptors, and R e is the distance between the<br />

protein axis and the second lipid shell (exclusion distance for<br />

bulk-located acceptors). In the present system, l is the<br />

unlabeled lipid bilayer thickness, and the exclusion distance<br />

is 16 Å assuming a radii <strong>of</strong> 5 Å and 4.5 Å for the protein and<br />

the phospholipid, respectively; see Fig. 1 b). The value n 2<br />

must be corrected for the presence <strong>of</strong> labeled lipid in the<br />

annular region, which therefore is not part <strong>of</strong> the randomly<br />

distributed acceptors pool.<br />

FIGURE 2 Corrected emission spectrum <strong>of</strong> DCIA-labeled M13 major<br />

coat protein (—), and corrected excitation spectrum <strong>of</strong> NBD-derivatized<br />

phospholipid (---).<br />

<strong>Biophysical</strong> Journal 87(1) 344–352


348 Fernandes et al.<br />

FIGURE 3 Donor (DCIA-labeled protein) fluorescence quenching by energy transfer acceptor ((18:1) 2 -PE-NBD) in pure phosphatidylcholine bilayers <strong>with</strong><br />

different hydrophobic thickness. (d), Experimental energy transfer efficiencies; (—), theoretical simulations obtained from the annular model for protein-lipid<br />

interaction using the fitted K S ; and (---), simulations for random distribution <strong>of</strong> acceptors (K S ¼ 1.0). (A) Labeled protein incorporated in DOPC (fitted K S ¼<br />

1.4); (B) labeled protein incorporated in DMoPC (fitted K S ¼ 2.9); and (C) labeled protein incorporated in DEuPC (fitted K S ¼ 2.1).<br />

phospholipids <strong>of</strong> identical acyl-chains (18:1 and 12:0) and<br />

different headgroups (PC, PE, PS, PG, and PA) classes,<br />

derivatized <strong>with</strong> NBD at the 12:0 chain. The results are<br />

shown in Fig. 4, together <strong>with</strong> the model fits. Table 1<br />

summarizes the recovered K S values.<br />

M13 coat protein could be submitted to irreversible<br />

aggregation. This hypothesis has been ruled out for both<br />

lipids in recent <strong>studies</strong> (Meijer et al., 2001; Fernandes et al.,<br />

2003). Nevertheless, M13 major coat protein was shown<br />

recently by us to aggregate reversibly while in these con-<br />

DISCUSSION<br />

M13 coat protein selectivity toward phospholipids<br />

<strong>with</strong> different hydrophobic thickness<br />

The M13 coat protein is known to form large irreversible<br />

aggregates under specific conditions. These aggregates are<br />

not found in vivo, and therefore are regarded as an artifact<br />

(Hemminga et al., 1993). Due to their b-sheet conformation,<br />

they are detected by CD spectroscopy, and because <strong>of</strong> the<br />

hydrophobic mismatch between the protein and DEuPC<br />

(longer) or DMoPC (shorter) lipids, it was possible that<br />

while incorporated in pure bilayers <strong>of</strong> these components, the<br />

TABLE 1 Labeled phospholipids’ relative association<br />

constants toward M13 major coat protein<br />

Labeled phospholipid Bilayer composition K S K S /K S (PC)*<br />

((18:1) 2 -PE-NBD) DOPC (18:1) 2 PC 1.4 —<br />

((18:1) 2 -PE-NBD) DEuPC (22:1) 2 PC 2.1 —<br />

((18:1) 2 -PE-NBD) DMoPC (14:1) 2 PC 2.9 —<br />

(18:1-(12:0-NBD)-PE) DOPC 2.0 1.0<br />

(18:1-(12:0-NBD)-PC) DOPC 2.0 1.0<br />

(18:1-(12:0-NBD)-PG) DOPC 2.3 1.1<br />

(18:1-(12:0-NBD)-PS) DOPC 2.7 1.3<br />

(18:1-(12:0-NBD)-PA) DOPC 3.0 1.5<br />

*K S (PC) is the relative association constant <strong>of</strong> (18:1-(12:0-NBD)-PC).<br />

<strong>Biophysical</strong> Journal 87(1) 344–352


FRET Study <strong>of</strong> Protein-Lipid Selectivity 349<br />

FIGURE 4 Donor (DCIA-labeled protein) fluorescence quenching by energy transfer acceptor (18:1-(12:0-NBD)-PX), where X stands for the different<br />

headgroup structures, in pure bilayers <strong>of</strong> DOPC. (d), Experimental energy transfer efficiencies; (—), theoretical simulations obtained from the annular model<br />

for protein-lipid interaction using the fitted K S ; and (---), simulations for random distribution <strong>of</strong> acceptors (K S ¼ 1.0). (A) PC-labeled phospholipid (fitted K S ¼<br />

2.0); (B) PE-labeled phospholipid (fitted K S ¼ 2.0); (C) PG-labeled phospholipid (fitted K S ¼ 2.3); (D) PS-labeled phospholipid (fitted K S ¼ 2.7); and (E) PAlabeled<br />

phospholipid (fitted K S ¼ 3).<br />

ditions from a study <strong>of</strong> BODIPY labeled protein steady-state<br />

fluorescence emission (Fernandes et al., 2003). However, in<br />

that study much higher concentrations <strong>of</strong> protein were used<br />

when compared <strong>with</strong> the present one, and assuming the aggregation<br />

constants obtained in that work, only up to 5% at<br />

the very most <strong>of</strong> protein could be aggregated at the protein<br />

concentration used throughout this study. Therefore we can<br />

consider that our results report the phospholipid selectivity<br />

properties <strong>of</strong> the monomeric M13 major coat protein.<br />

The fitting <strong>of</strong> the annular model for protein-lipid<br />

interactions to the FRET data (Figs. 3 and 4), converged<br />

always to K S values above 1 (Table 1), as the energy transfer<br />

<strong>Biophysical</strong> Journal 87(1) 344–352


350 Fernandes et al.<br />

efficiencies (1 ÿ t DA =t D ) are above the expected value for<br />

random distribution <strong>of</strong> the labeled phospholipids. As our<br />

annular model assumes a random distribution outside the<br />

protein-lipid interface (which should be true for onecomponent<br />

bilayers in the liquid crystalline phase; Loura<br />

et al., 1996), this result is rationalized as an increase in local<br />

concentration <strong>of</strong> probe in the lipid annular shell around the<br />

protein. For (18:1) 2 -PE-NBD probe in DOPC ((18:1) 2 -PC)<br />

bilayers (Fig. 3 A), the value <strong>of</strong> K S was 1.4, pointing to<br />

almost complete randomization <strong>of</strong> the probe distribution in<br />

the bilayer, and therefore identical selectivity to the DOPC<br />

lipid. This was expected, because the probe acyl-chains are<br />

identical to the unlabeled lipid and allow a perfect hydrophobic<br />

matching <strong>of</strong> the protein.<br />

The results from Fig. 3, A–C, all report energy transfer<br />

efficiencies to the (18:1) 2 -PE-NBD probe, but in different<br />

bilayers <strong>of</strong> one lipid component. In DOPC bilayers the value<br />

<strong>of</strong> K S was 1.4 as discussed above, but in DMoPC ((14:1) 2 -<br />

PC) and DEuPC ((22:1) 2 -PC) bilayers the relative association<br />

constant values were 2.9 and 2.1, respectively,<br />

confirming a greater selectivity toward the hydrophobic<br />

matching unlabeled phospholipid (DOPC).<br />

M13 coat protein selectivity toward phospholipids<br />

<strong>with</strong> different headgroups<br />

Results from analysis <strong>of</strong> the data on M13 coat protein<br />

selectivity toward phospholipids headgroups are presented in<br />

Fig. 4. Clearly the anionic-labeled phospholipids exhibit<br />

larger selectivity to the lipid annular region around the<br />

protein, especially the 18:1-(12:0-NBD)-PA and 18:1-(12:<br />

0-NBD)-PS probes (K S ¼ 3.0 and 2.7, respectively). The<br />

18:1-(12:0-NBD)-PG probe presents an intermediate selectivity<br />

(K S ¼ 2.3), whereas 18:1-(12:0-NBD)-PC and 18:<br />

1-(12:0-NBD)-PE have identical relative association constants<br />

(K S ¼ 2.0). The selectivity for anionic phospholipids<br />

must be a consequence <strong>of</strong> electrostatic interaction <strong>of</strong> these<br />

<strong>with</strong> the highly basic C-terminal domain <strong>of</strong> the protein,<br />

which contains four lysines.<br />

Overall the selectivity <strong>of</strong> annular lipid-M13 coat protein is<br />

not large, which is common for intrinsic <strong>membrane</strong> <strong>proteins</strong><br />

(Lee, 2003). Additionally, it has been shown that selectivity<br />

<strong>of</strong> some <strong>proteins</strong> toward anionic lipids is significantly decreased<br />

in the presence <strong>of</strong> increasing ionic strength (Marsh<br />

and Horváth, 1998). In our case the ionic strength was kept<br />

high, because it is necessary to keep the protein in the<br />

monomeric state (Spruijt and Hemminga, 1991), and this<br />

further explains our results.<br />

Phospholipid selectivity ESR <strong>studies</strong> have already been<br />

performed <strong>with</strong> aggregated forms <strong>of</strong> M13 major coat protein<br />

(Peelen et al., 1992; Wolfs et al., 1989, Datema et al., 1987),<br />

resulting in similar selectivity patterns <strong>of</strong> M13 coat protein<br />

to phospholipid headgroups. Peelen et al. (1992), using<br />

a identical buffer type, ionic strength, and pH to that used in<br />

the present study, obtained the following relative association<br />

constants ratios (K S (PX)/K S (PC)) <strong>of</strong> M13 coat protein incorporated<br />

in 1,2-dimiristoyl-sn-glycero-3-phosphocholine<br />

(DMPC): K S (PA)/K S (PC) ¼ 1.6, K S (PS)/K S (PC) ¼ 1.2,<br />

K S (PG)/K S (PC) ¼ 1.1, and K S (PE)/K S (PC) ¼ 1. Overall the<br />

selectivity pattern is the same, and the relative association<br />

constants ratios are almost identical. The M13 coat protein<br />

was aggregated in that study, and according to the authors<br />

the number <strong>of</strong> first shell sites was five, that is, for each<br />

protein only a maximum <strong>of</strong> five lipids could be motionally<br />

restricted due to contact <strong>with</strong> the protein surface. For<br />

a monomeric helix, however, a value <strong>of</strong> 12 should be<br />

expected (Marsh and Horváth, 1998), and that was the<br />

number used in our model for the data analysis. Therefore it<br />

is particularly interesting that the ratios <strong>of</strong> the relative<br />

association constants remain almost identical. Apparently<br />

protein aggregation lowers the selectivity degree <strong>of</strong> each<br />

protein for phospholipids only through a decrease <strong>of</strong><br />

available area for protein-lipid contacts but the relative<br />

association ratio <strong>with</strong> phospholipids <strong>of</strong> different headgroups<br />

remains the same. Even though the protein presents higher<br />

selectivity for the NBD-labeled phospholipids than for<br />

DOPC (K S (PC) ¼ 2.0) (possibly due to electrostatic<br />

interactions <strong>with</strong> the NBD at the bilayer interface), the result<br />

presented above clearly shows that the presence <strong>of</strong> NBD at<br />

the bilayer interface does not change significantly the relative<br />

association ratios <strong>of</strong> the phospholipids.<br />

Sanders et al. (1992) were not able to determine the<br />

selectivity <strong>of</strong> the M13 coat protein monomeric species<br />

toward phospholipids using ESR, because the monomer was<br />

not able to produce a sufficiently long-living boundary shell<br />

<strong>of</strong> lipids that could be detected by ESR spectroscopy. The<br />

fact that it was possible to clearly quantify relative association<br />

constants using FRET in the present study presents<br />

this technique as an alternative to ESR in protein-lipid<br />

<strong>studies</strong>.<br />

One important difference between the ESR and FRET<br />

techniques is that the latter is not restricted to the lipids<br />

adjacent to a given protein molecule. Not only labeled lipids<br />

in the first shell <strong>of</strong> lipids will be potential acceptors to<br />

a donor-labeled integral protein, but also the acceptors in the<br />

other lipid shells surrounding the protein will contribute to<br />

the final result. For that reason, this study also seems to<br />

confirm the hypothesis <strong>of</strong> selectivity to anionic phospholipids<br />

by the protein to largely confine itself to an annular shell<br />

<strong>of</strong> lipids in direct contact <strong>with</strong> the protein, in the case <strong>of</strong> the<br />

M13 major coat protein.<br />

In case that the annular region would extend beyond this<br />

first shell, our FRET analysis methodology (based in transfer<br />

to a single annular shell and also to the bulk) would recover<br />

substantially larger values for the relative association<br />

constants. Moreover, as commented above, our recovered<br />

K S /K S (PC) match those obtained from ESR measurements,<br />

which only detects immobilization <strong>of</strong> annular lipids upon<br />

incorporation <strong>of</strong> protein. The existence <strong>of</strong> a single annular<br />

<strong>Biophysical</strong> Journal 87(1) 344–352


FRET Study <strong>of</strong> Protein-Lipid Selectivity 351<br />

lipid layer for this protein might be related to the fact that it<br />

has a sole trans<strong>membrane</strong> segment.<br />

The FRET methodology has three interesting features.<br />

First, by choosing donor-acceptor pairs <strong>with</strong> different Förster<br />

radii it is possible to specifically study mainly the first-shell<br />

<strong>of</strong> lipids or also the outside shells, as was the case in the<br />

present study. The joint analysis <strong>of</strong> results coming from these<br />

different donor-acceptor pairs could allow for an even more<br />

detailed description <strong>of</strong> the protein-lipid arrangement in more<br />

complex systems. In our study, the relatively large R 0 value<br />

for the used donor-acceptor pair meant that the experimental<br />

quenching curves shown in both Figs. 3 and 4 look similar at<br />

first sight. Nevertheless it is impressive that the analysis<br />

methodology is able to retrieve significant K s values. Of<br />

course, this methodology could still be improved by the use<br />

<strong>of</strong> a donor-acceptor pair <strong>with</strong> a smaller R 0 value, closer to the<br />

distances under measurement. Second, the more economic<br />

character <strong>of</strong> fluorescence <strong>studies</strong>, which requires much<br />

smaller amounts <strong>of</strong> material than ESR, should be stressed.<br />

And third, although this model leads to a somewhat complex<br />

decay law (Eqs. 7, 10, and 11), it is actually not necessary to<br />

analyze the decay curves <strong>with</strong> this law to recover the relevant<br />

parameters, unlike in other FRET <strong>studies</strong> (e.g., <strong>of</strong> lipid phase<br />

separation; see Loura et al., 2001). The theoretical curves are<br />

conveniently simulated and integrated in a worksheet to<br />

calculate the theoretical FRET efficiencies. These can be<br />

matched to experimental values by varying the K S value (the<br />

sole unknown parameter). The experimental FRET efficiencies<br />

could also be obtained from steady-state data. In our<br />

case, we obtained them from integration <strong>of</strong> donor decay<br />

curves because these are less prone to artifacts (e.g., light<br />

scattering, inner filter effects, measurement <strong>of</strong> absolute<br />

intensities), which in any case, could in principle be corrected<br />

for in a steady-state experiment.<br />

CONCLUSIONS<br />

In the present study FRET has been applied <strong>with</strong> success in the<br />

characterization <strong>of</strong> the M13 major coat protein selectivity<br />

toward phospholipids. As expected, the protein has no<br />

significant selectivity for the lipid probe containing two<br />

oleoyl acyl-chains while in DOPC bilayers, but exhibits larger<br />

selectivity for the same probe while in bilayers <strong>with</strong> different<br />

hydrophobic thickness due to hydrophobic mismatch stress.<br />

The protein also presents larger selectivity for anionic lipids,<br />

particularly for phosphatidic acid and phosphatidylserine<br />

phospholipids. FRET was shown here to be a promising<br />

methodology in protein-lipid selectivity <strong>studies</strong>.<br />

F.F. acknowledges financial support from Fundacxão para a Ciência e<br />

Tecnologia (FCT), project POCTI/36458/QUI/2000, and COST-European<br />

Cooperation in the Field <strong>of</strong> Scientific and Technical Research Action D:22.<br />

L.M.S.L. and M.P. acknowledge financial support from FCT, projects<br />

POCTI/36458/QUI/2000, and POCTI/36389/FCB/2000. A.F. acknowledges<br />

a research grant (BPD/11488/2002) from Programa Operacional<br />

‘‘Ciência, Tecnologia, Inovacxão’’ (POCTI)/FCT.<br />

REFERENCES<br />

Abrams, F. S., and E. London. 1993. Extension <strong>of</strong> the parallax analysis <strong>of</strong><br />

<strong>membrane</strong> penetration depth to the polar region <strong>of</strong> model <strong>membrane</strong>s:<br />

use <strong>of</strong> fluorescence quenching by a spin-label attached to the<br />

phospholipids polar headgroup. Biochemistry. 32:10826–10831.<br />

Antollini, S. S., M. A. Soto, I. B. de Romanelli, C. Gutiérrez-Merino, P.<br />

Sotomayor, and F. J. Barrantes. 1996. Biophys. J. 70:1275–1284.<br />

Berberan-Santos, M. N., and M. J. E. Prieto. 1987. Energy transfer in<br />

spherical geometry. Application to micelles. J. Chem. Soc. Faraday<br />

Trans. 283:1391–1409.<br />

Bonini, I. C., S. S. Antollini, C. Gutiérrez-Merino, and F. J. Barrantes.<br />

2002. Sphingomyelin composition and physical asymmetries in native<br />

acetylcholine receptor-rich <strong>membrane</strong>s. Eur. Biophys. J. 31:417–427.<br />

Chattopadhyay, A. 1990. Chemistry and biology <strong>of</strong> N-(7-nitrobenz-2-oxa-<br />

1,3-diazol-4-yl)-labeled lipids: fluorescent probes <strong>of</strong> biological and<br />

model <strong>membrane</strong>s. Chem. Phys. Lipids. 53:1–15.<br />

Datema, K. P., C. J. A. M. Wolfs, D. Marsh, A. Watts, and M. A.<br />

Hemminga. 1987. Spin label electron spin resonance study <strong>of</strong><br />

bacteriophage M13 coat protein incorporation into mixed lipid bilayers.<br />

Biochemistry. 26:7571–7574.<br />

Davenport, L., R. E. Dale, R. H. Bisby, and R. B. Cundall. 1985.<br />

Transverse location <strong>of</strong> the fluorescent probe 1,6-diphenyl-1,3,5-hexatriene<br />

in model lipid bilayer <strong>membrane</strong> systems by resonance energy<br />

transfer. Biochemistry. 24:4097–4108.<br />

Dumas, F., M. C. Lebrun, and J.-F. Tocanne. 1999. Is the protein/lipid<br />

hydrophobic matching principle relevant to <strong>membrane</strong> organization and<br />

functions FEBS Lett. 458:271–277.<br />

Dumas, F., M. M. Sperotto, M.-C. Lebrun, J.-F. Tocanne, and O. G.<br />

Mouritsen. 1997. Molecular sorting <strong>of</strong> lipids by bacteriorhodopsin<br />

in dilauroylphosphatidylcholine/distearoylphosphatidylcholine lipid bilayers.<br />

Biophys. J. 73:1940–1953.<br />

Eaton, D. F. 1988. Reference materials for fluorescence measurement. Pure<br />

Appl. Chem. 60:1107–1114.<br />

Everett, J., A. Zlotnick, J. Tennyson, and P. W. Holloway. 1986.<br />

Fluorescence quenching <strong>of</strong> cytochrome b5 in vesicles <strong>with</strong> an<br />

asymmetric transbilayer distribution <strong>of</strong> brominated phosphatidylcholine.<br />

J. Biol. Chem. 261:6725–6729.<br />

Fahsel, S., E. M. Pospiech, M. Zein, T. L. Hazlet, E. Gratton, and R.<br />

Winter. 2002. Modulation <strong>of</strong> concentration fluctuations in phaseseparated<br />

lipid <strong>membrane</strong>s by polypeptide insertion. Biophys. J. 83:<br />

334–344.<br />

Fernandes, F., L. M. S. Loura, M. Prieto, R. Koehorst, R. Spruijt, and M. A.<br />

Hemminga. 2003. Dependence <strong>of</strong> M13 major coat protein oligomerization<br />

and lateral segregation on bilayer composition. Biophys. J. 85:2430–<br />

2441.<br />

Gutierrez-Merino, C. 1981a. Quantitation <strong>of</strong> the Förster energy transfer for<br />

two-dimensional systems. I. Lateral phase separation in unilamellar<br />

vesicles formed by binary phospholipids mixtures. Biophys. Chem.<br />

14:247–257.<br />

Gutierrez-Merino, C. 1981b. Quantitation <strong>of</strong> the Förster energy transfer for<br />

two-dimensional systems. II. Protein distribution and aggregation state in<br />

biological <strong>membrane</strong>s. Biophys. Chem. 14:259–266.<br />

Gutierrez-Merino, C., F. Munkonge, A. M. Mata, J. M. East, B. L.<br />

Levinson, R. M. Napier, and A. G. Lee. 1987. The position <strong>of</strong> the ATP<br />

binding site on the (Ca 21 1 Mg 21 )-ATPase. Biochim. Biophys. Acta.<br />

897:207–216.<br />

Hemminga, M. A., J. C. Sanders, C. J. A. M. Wolfs, and R. B. Spruijt.<br />

1993. Lipid-protein interactions involved in bacteriophage M13 infection.<br />

Protein Lipid Inter. New Compr. Biochem. 25:191–212.<br />

Killian, J. A. 1998. Hydrophobic mismatch between <strong>proteins</strong> and lipids in<br />

<strong>membrane</strong>s. Biochim. Biophys. Acta. 1376:401–416.<br />

Lakowicz, J. R. 1999. Principals <strong>of</strong> Fluorescence Spectroscopy, 2nd Ed.<br />

Kluwer Academic/Plenum Press, New York. Chap. 8.<br />

Lee, A. G. 2003. Lipid-protein interactions in biological <strong>membrane</strong>s:<br />

a structural perspective. Biochim. Biophys. Acta. 1612:1–40.<br />

<strong>Biophysical</strong> Journal 87(1) 344–352


352 Fernandes et al.<br />

Lehtonen, J. Y. A., and P. K. J. Kinnunen. 1997. Evidence for phospholipid<br />

microdomain formation in liquid crystalline liposomes reconstituted <strong>with</strong><br />

Escherichia coli lactose permease. Biophys. J. 72:1247–1257.<br />

Lewis, B. A., and D. M. Engelman. 1983. Lipid bilayer thickness varies<br />

linearly <strong>with</strong> acyl chain length in fluid phosphatidylcholine vesicles.<br />

J. Mol. Biol. 166:211–217.<br />

Loura, L. M. S., A. Fedorov, and M. Prieto. 1996. Resonance energy<br />

transfer in a model system <strong>of</strong> <strong>membrane</strong>s: application to gel and liquid<br />

crystalline phases. Biophys. J. 71:1823–1836.<br />

Loura, L. M. S., R. F. M. de Almeida, and M. Prieto. 2001. Detection and<br />

characterization <strong>of</strong> <strong>membrane</strong> microheterogeneity by resonance energy<br />

transfer. J. Fluorescence. 11:197–209.<br />

Loura, L. M. S., A. Fedorov, and M. Prieto. 2000. Membrane probe<br />

distribution heterogeneity: a resonance energy transfer study. J. Phys.<br />

Chem. B. 104:6920–6931.<br />

Marquardt, D. W. 1963. An algorithm for least-squares estimation <strong>of</strong> nonlinear<br />

parameters. J. Soc. Ind. Appl. Math. (SIAM J.). 11:431–441.<br />

Marsh, D., and L. I. Horváth. 1998. Structure, dynamics and composition <strong>of</strong><br />

the lipid-protein interface. Perspectives from spin-labelling. Biochim.<br />

Biophys. Acta. 1376:267–296.<br />

Màzeres, S., V. Schram, J.-F. Tocanne, and A. Lopez. 1996. 7-Nitrobenz-2-<br />

oxa-1,3-diazole-4-yl-labeled phospholipids in lipid <strong>membrane</strong>s: differences<br />

in fluorescence behavior. Biophys. J. 71:327–335.<br />

McClare, C. 1971. An accurate and convenient organic phosphorus assay.<br />

Anal. Biochem. 39:527–530.<br />

Meijer, A. B., R. B. Spruijt, C. J. A. M. Wolfs, and M. A. Hemminga. 2001.<br />

Membrane-anchoring interactions <strong>of</strong> M13 major coat protein. Biochemistry.<br />

40:8815–8820.<br />

Mouritsen, O. G., and M. Bloom. 1984. Mattress model <strong>of</strong> lipid-protein<br />

interactions in <strong>membrane</strong>s. Biophys. J. 46:141–153.<br />

O’Keeffe, A. H., J. M. East, and A. G. Lee. 2000. Selectivity in lipid<br />

binding to the bacterial outer <strong>membrane</strong> protein OmpF. Biophys. J.<br />

79:2066–2074.<br />

Peelen, S. J. C. J., J. C. Sanders, M. A. Hemminga, and D. Marsh. 1992.<br />

Biochemistry. Stoichiometry, selectivity and exchange dynamics <strong>of</strong> lipidprotein<br />

interaction <strong>with</strong> bacteriophage M13 coat protein studied by spin<br />

label electron spin resonance. Effects <strong>of</strong> protein secondary structure.<br />

Biochemistry. 31:2670–2677.<br />

Sanders, J. C., M. F. Ottaviani, A. van Hoek, and A. J. W. G. Visser. 1992.<br />

A small protein in model <strong>membrane</strong>s: a time-resolved fluorescence and<br />

ESR study on the interaction <strong>of</strong> M13 coat protein <strong>with</strong> lipid bilayers. Eur.<br />

Biophys. J. 21:305–311.<br />

Spruijt, R. B., and M. A. Hemminga. 1991. The in situ aggregational and<br />

conformational state <strong>of</strong> the major coat protein <strong>of</strong> bacteriophage M13 in<br />

phospholipid bilayers mimicking the inner <strong>membrane</strong> <strong>of</strong> host Escherichia<br />

coli. Biochemistry. 30:11147–11154.<br />

Spruijt, R. B., C. J. A. M. Wolfs, and M. A. Hemminga. 1989. Aggregation<br />

related conformational change <strong>of</strong> the <strong>membrane</strong>-associated coat protein<br />

<strong>of</strong> bacteriophage M13. Biochemistry. 28:9158–9165.<br />

Spruijt, R. B., C. J. A. M. Wolfs, J. W. G. Verver, and M. A. Hemminga.<br />

1996. Accessibility and environment probing using cysteine residues<br />

introduced along the putative trans<strong>membrane</strong> domain <strong>of</strong> the major coat<br />

protein <strong>of</strong> bacteriophage M13. Biochemistry. 35:10383–10391.<br />

Stopar, D., R. B. Spruijt, C. J. A. M. Wolfs, and M. A. Hemminga. 2003.<br />

Biochim. Biophys. Acta. 1611:5–15.<br />

Williamson, I. M., S. J. Alvis, J. M. East, and A. G. Lee. 2002. <strong>Interaction</strong>s<br />

<strong>of</strong> phospholipids <strong>with</strong> the potassium channel KcsA. Biophys. J. 83:2026–<br />

2038.<br />

Wolf, D. E., A. P. Winiski, A. E. Ting, K. M. Bocian, and R. E. Pagano.<br />

1992. Determination <strong>of</strong> the transbilayer distribution <strong>of</strong> fluorescent lipid<br />

analogues by nonradiative fluorescence resonance energy transfer.<br />

Biochemistry. 31:2865–2873.<br />

Wolfs, C. L. A. M., L. I. Horváth, D. Marsh, A. Watts, and M. A.<br />

Hemminga. 1989. Spin label ESR <strong>of</strong> bacteriophage M13 coat protein in<br />

mixed lipid bilayers. Characterization <strong>of</strong> molecular selectivity <strong>of</strong> charged<br />

phospholipids for the bacteriophage M13 coat protein in lipid bilayers.<br />

Biochemistry. 28:995–1001.<br />

<strong>Biophysical</strong> Journal 87(1) 344–352


BINDING OF INHIBITORS TO A PUTATIVE BINDING<br />

DOMAIN OF V-ATPase<br />

III<br />

BINDING OF INHIBITORS TO A<br />

PUTATIVE BINDING DOMAIN<br />

OF V-ATPase<br />

1. Introduction<br />

Osteoporosis, a disease endemic in Western society, is characterized by a net loss <strong>of</strong><br />

bone mass which results from an imbalance <strong>of</strong> the bone-remodelling process, <strong>with</strong> bone<br />

resorption exceeding bone formation. Bone demineralization involves acidification <strong>of</strong><br />

the isolated extracellular microenvironment. The lowering <strong>of</strong> pH is necessary both to<br />

dissolve the bone material and to provide the acidic environment required by the<br />

collagenases to degrade the bone matrix (Teitelbaum, 2000). The acidification is due to<br />

proton translocation by the vacuolar-type H + -ATPases (V-ATPases) that are present in<br />

large numbers on the ruffled border <strong>of</strong> the osteoclast while the counterions diffuse<br />

passively through a chloride channel (Fig. III-1). Although there are a number <strong>of</strong> drugs<br />

directed to osteoporosis treatment through indirect inhibition <strong>of</strong> osteoclasts (Lark and<br />

James, 2002; Rodan and Martin, 2000), they allow only for a limited therapy as bone<br />

loss is only temporarily reduced due to weak inhibition <strong>of</strong> these cells. The osteoclast V-<br />

ATPase is, therefore, an attractive alternative target for the development <strong>of</strong> better<br />

therapeutic agents that effectively and safely reduce osteoclast activity.<br />

V-ATPases are large, multisubunit complexes organized in two domains (Fig. III.2)<br />

(Nishi and Forgac, 2002; Kawasaki-Nishi et al., 2003). The V 1 domain, which is<br />

responsible for ATP hydrolysis, comprises eight different subunits (A-H) that can be<br />

removed from the <strong>membrane</strong> in soluble form. The V o domain, which is tightly<br />

associated <strong>with</strong> the <strong>membrane</strong> and composed <strong>of</strong> five different subunits (a, c, c’, c’’, and<br />

d), is responsible for proton translocation. The enzyme functions as a rotary motor<br />

(Yokoyama et al., 2003; Harrison et al., 1997). By analogy <strong>with</strong> the mechanism<br />

proposed for the F-ATPase, ATP hydrolysis at the catalytic sites drives rotation <strong>of</strong> the<br />

79


central D-subunit, which in turn drives rotation <strong>of</strong> the hexameric ring <strong>of</strong> c subunits<br />

immersed in the <strong>membrane</strong>. Subunit a is held fix relative to the A3B3 hexamer by the<br />

peripherical stalk, which functions as a stator. Proton transport is hypothesized to occur<br />

at the interface between the subunit a and the subunits c, c’ and c’’: rotation <strong>of</strong> the ring<br />

<strong>of</strong> c subunits relative to subunit a, brings the buried acidic residues into sequential<br />

contact <strong>with</strong> two hemichannels in subunit a, that provide access to each side <strong>of</strong> the<br />

<strong>membrane</strong>, driving unidirectional proton transport (Fig. III-2) (Grabe et al., 2000).<br />

Figure III-1: Representation <strong>of</strong> the resorbing osteoclast and acidification <strong>of</strong> the resorption lacunae<br />

(taken from Väänanen, 2005).<br />

Figure III-2: Structural Model <strong>of</strong> V-ATPase and proposed rotary mechanism <strong>of</strong> ATP-driven proton<br />

translocation. The most important amino acid residues for proton translocation are marked (adapted from<br />

Nishi and Forgac, 2002)<br />

80


BINDING OF INHIBITORS TO A PUTATIVE BINDING<br />

DOMAIN OF V-ATPase<br />

In eukaryotic cells, V-ATPase, apart from being present in the plasma <strong>membrane</strong> <strong>of</strong><br />

osteoclasts and kidney acid secreting cells, is also found in many organelles such as<br />

vacuoles, lysossomes, coated vesicles, Golgi, and secretory vesicles. Numerous<br />

physiological processes depend on the activity <strong>of</strong> V-ATPases, including intracellular<br />

targeting, protein processing, transport <strong>of</strong> metabolites, receptor-mediated endocytosis,<br />

neurotransmitter uptake, and apoptosis (Finbow and Harrison, 1997; Nishi and Forgac,<br />

2002).<br />

The macrolide antibiotics bafilomycins and concanamycins are specific, highly<br />

potent inhibitors <strong>of</strong> all eukaryotic V-ATPases in vitro and in vivo (Fig. III-3). However,<br />

their unselective action towards V-ATPases turns them to be extremely toxic agents and<br />

unsuitable for therapeutical use in osteoporosis treatment (Dröse and Altendorf, 1997).<br />

Recently, novel and selective inhibitors <strong>of</strong> osteoclasts V-ATPase have been synthesized<br />

according to structure-function relationships <strong>of</strong> bafilomycins derivatives (Gagliardi et<br />

al., 1998a; Gagliardi et al. 1998b), suggesting that selective modulation <strong>of</strong> different V-<br />

ATPases might be possible by this new indole class <strong>of</strong> inhibitors, namely by its most<br />

promising representative SB 242784 (INH-3) (see Fig. III.3) (Nadler et al., 1998). SB<br />

242784 <strong>studies</strong> <strong>with</strong> animal models <strong>of</strong> bone resorption showed a remarkable efficiency<br />

in the prevention <strong>of</strong> ovariectomy-induced bone loss on rats, and no adverse effects <strong>of</strong> its<br />

administration on the animals was observed. In situ <strong>studies</strong> revealed that SB 242784<br />

inhibited osteoclastic V-ATPase activity at 1000 times smaller concentrations than<br />

enzymes from any other tissue evaluated (Visentin et al., 2000).<br />

Figure III-3: V-ATPase inhibitors.<br />

81


Our understanding <strong>of</strong> the mechanism <strong>of</strong> V-ATPase inhibition is rudimentary.<br />

Nevertheless, site-directed mutagenic <strong>studies</strong> strongly supported the hypothesis <strong>of</strong> the<br />

location <strong>of</strong> the bafilomycin A 1 binding site in the c-subunit <strong>of</strong> V-ATPase (Bowman and<br />

Bowman, 2002; Bowman et al., 2004). These <strong>studies</strong> indicated several amino acid<br />

residues in the c-subunit <strong>of</strong> V-ATPase for which mutations induced resistance to<br />

bafilomycin A 1 . The majority <strong>of</strong> these residues were located at the 4 th putative TM<br />

helix, and four <strong>of</strong> them (L132, F136, L140, and Y143) are positioned in the same helix<br />

side, which is expected to be exposed to the lipid environment (Harrison et al., 2000).<br />

Some residues found in other domains <strong>of</strong> the c-subunit are apparently also relevant for<br />

bafilomycin binding. However, it is not clear, if they are part <strong>of</strong> the bafilomycin binding<br />

site, as a structure formed by these residues would imply a helix organization for the c-<br />

subunit which is in disagreement <strong>with</strong> the one suggested by cysteine cross-linking<br />

experiments (Harrison et al., 1999). Another study detected immobilization <strong>of</strong> spinlabeled<br />

residues <strong>of</strong> c-subunit in the presence <strong>of</strong> concanamycin A, in agreement <strong>with</strong> a<br />

binding site location on this region <strong>of</strong> V-ATPase (Páli et al., 2004).<br />

Whyteside et al. (2005) found that SB242784 bound competitively <strong>with</strong><br />

concanamycin to the c-subunit <strong>of</strong> V-ATPase. However, a mutation on Y143 from the c-<br />

subunit <strong>of</strong> V-ATPase on yeast conferred different degrees <strong>of</strong> resistance for<br />

concanamycin and SB242784, suggesting that the binding sites for these inhibitors are<br />

overlapping but not identical (Páli et al., 2004).<br />

Subunit c and c´ present very similar sequences. However, mutations in subunit c´<br />

homologous to the ones in subunit c that conferred resistance to bafilomycin have no<br />

effect (Bowman et al., 2004). This fact, led to the proposal that either subunit c´ does<br />

not present bafilomycin binding sites or that the affinity <strong>of</strong> the macrolide antibiotic<br />

binding site was much smaller in subunit c´.<br />

One strategy to resolve the problem <strong>of</strong> the binding site location for the above<br />

mentioned inhibitors is the use <strong>of</strong> a reductionist approach. Reductionist approaches to<br />

protein-ligand problems through use <strong>of</strong> <strong>peptides</strong> have been used in the past <strong>with</strong> success<br />

and <strong>peptides</strong> comprising entire binding sites were already shown to present ligand<br />

binding affinities identical to the ones observed <strong>with</strong> the intact protein (Lentz et al.,<br />

1998; Wilson et al., 1985; Neumann et al., 1986; Taylor et al., 1995; Ried et al., 1996;<br />

Scatigno et al., 2004).<br />

82


INTERACTION OF THE INDOLE CLASS OF V-ATPase<br />

INHIBITORS WITH LIPID BILAYERS<br />

2. INTERACTION OF THE INDOLE CLASS OF<br />

VACUOLAR H + -ATPase INHIBITORS WITH<br />

LIPID BILAYERS<br />

83


Biochemistry 2006, 45, 5271-5279<br />

5271<br />

<strong>Interaction</strong> <strong>of</strong> the Indole Class <strong>of</strong> Vacuolar H + -ATPase Inhibitors <strong>with</strong> Lipid<br />

Bilayers †<br />

F. Fernandes, ‡ L. Loura, ‡,§ R. B. M. Koehorst, | N. Dixon, ⊥ T. P. Kee, ⊥ M. A. Hemminga, | and M. Prieto* ,‡<br />

Centro de Química-Física Molecular, Instituto Superior Técnico, Lisbon, Portugal, Departamento de Química, UniVersidade de<br />

EÄVora, EÄVora, Portugal, Laboratory <strong>of</strong> Biophysics, Wageningen UniVersity, Wageningen, The Netherlands, and<br />

Department <strong>of</strong> Chemistry, UniVersity <strong>of</strong> Leeds, Leeds LS2 9JT, United Kingdom<br />

ReceiVed NoVember 7, 2005; ReVised Manuscript ReceiVed March 8, 2006<br />

ABSTRACT: The selective inhibitor <strong>of</strong> osteoclastic V-ATPase (2Z,4E)-5-(5,6-dichloro-2-indolyl)-2-methoxy-<br />

N-(1,2,2,6,6-pentamethylpiperidin-4-yl)-2,4-pentadienamide (SB 242784), member <strong>of</strong> the indole class <strong>of</strong><br />

V-ATPase inhibitors, is expected to target the <strong>membrane</strong>-bound domain <strong>of</strong> the enzyme. A structural study<br />

<strong>of</strong> the interaction <strong>of</strong> this inhibitor <strong>with</strong> the lipidic environment is an essential step in the understanding<br />

<strong>of</strong> the mechanism <strong>of</strong> inhibition. In this work, a comprehensive study <strong>of</strong> the relevant features <strong>of</strong> this<br />

interaction was performed. Inhibitor partition coefficients to lipid vesicles as well as its transverse location,<br />

orientation (order parameters), and dynamics while bound to bilayers were determined through<br />

photophysical techniques, taking advantage <strong>of</strong> the intrinsic fluorescence <strong>of</strong> the molecule. To better evaluate<br />

the functionally relevant features <strong>of</strong> SB 242784, a second inhibitor, INH-1, from the same class and<br />

having a reduced activity was also examined. It is shown that regarding <strong>membrane</strong> interaction their<br />

properties remain very similar for both molecules, suggesting that the differences in inhibition efficiencies<br />

are solely a consequence <strong>of</strong> the molecular recognition processes <strong>with</strong>in the inhibition site in the V-ATPase.<br />

Bone degradation requires a lowering <strong>of</strong> pH in the<br />

resorption lacuna (extracellular space between the osteoclast<br />

cell and bone surface) (1). This is achieved through pumping<br />

<strong>of</strong> protons by V-ATPases 1 located in the ruffled border <strong>of</strong><br />

osteoclasts (2). Inhibitors <strong>of</strong> osteoclast V-ATPase activity<br />

could therefore be used in new therapeutics for treatment <strong>of</strong><br />

bone diseases related to excess bone resorption, namely,<br />

osteoporosis.<br />

The macrolide antibiotic bafilomycin is a powerful inhibitor<br />

<strong>of</strong> V-ATPases (3) and was shown to prevent bone<br />

resorption (4). Nevertheless, due to the lack <strong>of</strong> specificity<br />

<strong>of</strong> bafilomycin toward osteoclastic V-ATPase, several other<br />

important V-ATPases are inhibited, and this is responsible<br />

for the extremely high toxicity and, consequently, for the<br />

clinical inadequacy <strong>of</strong> this compound. Recently, some<br />

†<br />

This work was supported by Contract QLG-CT-2000-01801 <strong>of</strong> the<br />

European Commission (MIVase Consortium) and by FCT (Portugal)<br />

under the Program POCTI. M.A.H. and M.P. are members <strong>of</strong> the COST<br />

D22 Action <strong>of</strong> the European Union. F.F. acknowledges Grant SFRH/<br />

BD/14282/2003 from FCT (Portugal).<br />

* Corresponding author. E-mail: prieto@alfa.ist.utl.pt. Telephone:<br />

(+351) 218413248. Fax: (+351) 218464455.<br />

‡<br />

Instituto Superior Técnico.<br />

§<br />

Universidade de EÄ vora.<br />

|<br />

Wageningen University.<br />

⊥<br />

University <strong>of</strong> Leeds.<br />

1<br />

Abbreviations: V-ATPase, vacuolar H + -ATPase; DOPC, 1,2-<br />

dioleoyl-sn-glycero-3-phosphocholine; DOPG, 1,2-dioleoyl-sn-glycero-<br />

3-[phospho-rac-(1-glycerol)]; DMPC, 1,2-dimyristoyl-sn-glycero-3-<br />

phosphocholine; 5-DOX-PC, 1-palmitoyl-2-stearoyl(5-DOXYL)-snglycero-3-phosphocholine;<br />

12-DOX-PC, 1-palmitoyl-2-stearoyl(12-<br />

DOXYL)-sn-glycero-3-phosphocholine; SB 242784, (2Z,4E)-5-(5,6-<br />

dichloro-2-indolyl)-2-methoxy-N-(1,2,2,6,6-pentamethylpiperidin-4-yl)-<br />

2,4-pentadienamide, INH-1, methyl (2Z,4E)-5-(5,6-dichloro-2-indolyl)-<br />

2-methoxy-2,4-pentadienoate.<br />

FIGURE 1: Chemical structures <strong>of</strong> SB 242784 and INH-1.<br />

derivatives <strong>of</strong> bafilomycin have been synthesized which<br />

maintained high inhibitory activity toward V-ATPases and<br />

exhibited high selectivity for the osteoclast form <strong>of</strong> the<br />

enzyme (5, 6). The most promising <strong>of</strong> these compounds,<br />

(2Z,4E)-5-(5,6-dichloro-2-indolyl)-2-methoxy-N-(1,2,2,6,6-<br />

pentamethylpiperidin-4-yl)-2,4-pentadienamide (SB 242784)<br />

(Figure 1), was shown to have more than 1000-fold selectivity<br />

for the osteoclast V-ATPase compared <strong>with</strong> the enzyme<br />

measured in kidney, liver, spleen, stomach, brain, or endothelial<br />

cells and to have no effect on other cellular ATPases.<br />

SB 242784 was also extremely effective in preventing bone<br />

loss in ovariectomized rats (7). This inhibitor is therefore a<br />

promising candidate for osteoporosis treatment.<br />

The site <strong>of</strong> action <strong>of</strong> bafilomycin in the enzyme was shown<br />

to be located in the <strong>membrane</strong>-bound domain (8-11).<br />

Therefore, the type and extent <strong>of</strong> the interaction <strong>of</strong> the<br />

bafilomycin derivative SB 242784 <strong>with</strong> the lipid environment<br />

and the properties <strong>of</strong> the inhibitor molecule while in this<br />

medium are relevant, since it is expected that the first step<br />

prior to the interaction <strong>with</strong> the enzyme is the <strong>membrane</strong><br />

10.1021/bi0522753 CCC: $33.50 © 2006 American Chemical Society<br />

Published on Web 04/05/2006


5272 Biochemistry, Vol. 45, No. 16, 2006 Fernandes et al.<br />

incorporation <strong>of</strong> the inhibitor <strong>with</strong> the concomitant increase<br />

<strong>of</strong> its effective concentration.<br />

Recently, spin-labeled derivatives <strong>of</strong> SB 242784 have been<br />

studied by electron spin resonance (ESR) (12). It was then<br />

possible to acquire information regarding the location in<br />

the <strong>membrane</strong> <strong>of</strong> the piperidine ring where the nitroxide label<br />

was inserted. However, a photophysical approach would<br />

enable a report on the whole molecule, since the indole<br />

ring is conjugated <strong>with</strong> the double bond system, and this is<br />

largely expected to dictate the inhibitor properties in the lipid<br />

phase.<br />

In this work we report a thorough study on the interaction<br />

<strong>of</strong> SB 242784 <strong>with</strong> lipid vesicles together <strong>with</strong> a detailed<br />

study <strong>of</strong> the compound behavior in aqueous solution, using<br />

for this effect its intrinsic photophysical properties (absorption,<br />

fluorescence, fluorescence anisotropy, linear dichroism),<br />

thus avoiding the derivatization <strong>of</strong> the molecule. The waterlipid<br />

partition coefficient <strong>of</strong> SB 242784 was determined, and<br />

its aggregation was studied. The position and orientation <strong>of</strong><br />

the molecule when interacting <strong>with</strong> lipid bilayers were also<br />

obtained. The same <strong>studies</strong> were performed <strong>with</strong> another less<br />

powerful inhibitor <strong>of</strong> the same class (INH-1) to compare the<br />

results and retrieve information on the relevance <strong>of</strong> specific<br />

structural features <strong>of</strong> SB 242784 on these properties.<br />

MATERIALS AND METHODS<br />

DOPC, DOPG, and DMPC were obtained from Avanti<br />

Polar Lipids (Birmingham, AL). 5-DOX-PC and 12-DOX-<br />

PC were from Avanti Polar Lipids (Birmingham, AL). Fine<br />

chemicals were obtained from Merck (Darmstadt, Germany).<br />

All materials were used <strong>with</strong>out further purification.<br />

SB 242784 and INH-1 were synthesized according to<br />

Nadler et al. (6) and Gagliardi et al. (5).<br />

Inhibitor Incorporation in Lipid Vesicles. The V-ATPase<br />

inhibitors were incorporated in lipid vesicles by two different<br />

methods: In the cosolubilization method the phospholipid<br />

and inhibitor solutions were mixed in chlor<strong>of</strong>orm and dried<br />

under a flow <strong>of</strong> dry N 2 . The last traces <strong>of</strong> solvent were<br />

removed under vacuum. Multilamellar vesicles were then<br />

obtained through solubilization in Tris buffer, and large<br />

unilamellar vesicles (LUVs) were produced by extrusion<br />

through polycarbonate filters <strong>with</strong> a pore size <strong>of</strong> 100 nm<br />

(13). In an alternative method <strong>of</strong> incorporation, a very small<br />

volume (


V-ATPase Indole Inhibitor <strong>Interaction</strong> <strong>with</strong> Bilayers Biochemistry, Vol. 45, No. 16, 2006 5273<br />

The order parameters (〈P 2 〉 and 〈P 4 〉) are obtained from<br />

the equations:<br />

sin(ω)A ω<br />

Α ω)(π/2)<br />

) 1 +<br />

3〈P 2 〉<br />

(1 - 〈P 2 〉)n 2 cos2 ω (3)<br />

I VH /I VV ) a sin 2 R+b (4)<br />

A ω is the absorbance at angle ω, n is the refraction index [n<br />

) 1.5 (21)], R is the angle at which the fluorescence is<br />

measured, and the a and b parameters depend on both 〈P 2 〉<br />

and 〈P 4 〉. Complete formalisms and further details are<br />

described in Lopes et al. (22).<br />

Dichroic ratios were determined using Glan-Thompson<br />

polarizers. In fluorescence emission measurements for determination<br />

<strong>of</strong> 〈P 4 〉, excitation and emission wavelengths were<br />

kept the same as in the remaining steady-state measurements.<br />

From the knowledge <strong>of</strong> 〈P 2 〉 and 〈P 4 〉, a combination <strong>of</strong><br />

the maximum entropy method together <strong>with</strong> the formalism<br />

<strong>of</strong> the Lagrange multipliers is used to describe the single<br />

particle distribution function f(Ψ), where Ψ is the angle<br />

between the transition moment <strong>of</strong> the molecule and the<br />

director <strong>of</strong> the system (normal to the bilayer plane). This<br />

distribution is the broadest one that is compatible <strong>with</strong> the<br />

experimental order parameters, and to obtain the population<br />

density probability function, which is the probability <strong>of</strong><br />

finding the transition moment between Ψ and Ψ + dΨ, it<br />

should be multiplied by sin Ψ (22).<br />

The determination <strong>of</strong> the transition moment <strong>of</strong> SB 242784<br />

was performed for an optimized geometry <strong>of</strong> the molecule<br />

based on the density functional theory (DFT) (23), which<br />

recovered the 2Z,4E isomer that is the preponderant species<br />

(6). The theoretical level applied to the calculations was<br />

Becke3LYP/6-31G(d) (24-26).<br />

RESULTS<br />

Indole Inhibitor BehaVior in Aqueous EnVironment. Since<br />

the interaction <strong>with</strong> the <strong>membrane</strong> proceeds from the<br />

inhibitors in aqueous solution, their characterization in an<br />

aqueous environment was first carried out. Preliminary results<br />

showed both SB 242784 and INH-1 (Figure 1) to be unstable<br />

under aqueous condition (27). The fluorescence quantum<br />

yield <strong>of</strong> the two molecules in aqueous buffer was extremely<br />

low (0.01 and 0.004 for SB 242784 and INH-1, respectively)<br />

when compared <strong>with</strong> the quantum yield observed in other<br />

environments (e.g., 0.06 and 0.05 for SB 242784 and INH-1<br />

in ethanol, respectively). The large decrease in fluorescence<br />

intensity <strong>of</strong> the inhibitors is the result <strong>of</strong> formation <strong>of</strong><br />

nonfluorescent or “dark” aggregates, as the quantum yield<br />

values determined from transient state fluorescence measurements<br />

were larger than the ones obtained from steady-state<br />

data. This aggregation is due to the hydrophobic character<br />

<strong>of</strong> the inhibitors, and the fluorescence quenching effect is<br />

enhanced by the presence <strong>of</strong> chlorine atoms in the molecule.<br />

In the present work we measured the extent and kinetics <strong>of</strong><br />

aggregation <strong>of</strong> both inhibitors in an aqueous environment.<br />

For this purpose, fluorescence intensity measurements were<br />

started immediately after addition <strong>of</strong> 10 µL <strong>of</strong> a ethanol stock<br />

solution to a cuvette containing 3 mL <strong>of</strong> buffer under<br />

permanent stirring, and the final concentration <strong>of</strong> inhibitor<br />

in buffer was 4 µM. The buffer used was 10 mM Tris and<br />

FIGURE 2: SB 242784 fluorescence emission self-quenching<br />

induced by aggregation in an aqueous environment. The buffer used<br />

was 10 mM Tris and 150 mM NaCl, pH ) 7.4.<br />

150 mM NaCl, pH 7.4. The result for SB 242784 is presented<br />

in Figure 2. The kinetics for the aggregation in water <strong>of</strong><br />

INH-1 was very fast, and the phenomenon could not be<br />

resolved as well as it was for SB 242784. In addition, the<br />

extent <strong>of</strong> aggregation, as taken from the remaining fluorescence<br />

intensity after the process reaches equilibrium, was<br />

larger for INH-1 than for SB 242784. For both inhibitors<br />

the fluorescence intensity increased dramatically upon addition<br />

<strong>of</strong> sodium dodecyl sulfate to the samples (results not<br />

shown), this suggesting disruption <strong>of</strong> aggregates and emission<br />

<strong>of</strong> monomeric inhibitors dispersed in the micellar medium<br />

and ruling out a significant contribution <strong>of</strong> photobleaching.<br />

The anisotropies <strong>of</strong> INH-1 and SB 242784 in water were<br />

0.20 and 0.22, respectively, and did not change during the<br />

process <strong>of</strong> aggregation. This indicates that the aggregates<br />

being formed are nonfluorescent, and the residual emission<br />

<strong>of</strong> monomeric species is the one contributing to anisotropy.<br />

In an attempt to determine the existence <strong>of</strong> a critical<br />

micellar concentration (cmc) for the inhibitors in an aqueous<br />

environment, we measured fluorescence emission intensities<br />

for inhibitor solutions <strong>with</strong> concentrations up to 4 µM. A<br />

cmc should appear as a change in linearity <strong>of</strong> fluorescence<br />

intensity vs inhibitor concentration. However, a completely<br />

linear plot was obtained up to 10 nM, the limit <strong>of</strong> detection<br />

<strong>of</strong> the fluorometer (results not shown). Therefore, the<br />

aggregation <strong>of</strong>fset, if existing at all, should occur for both<br />

inhibitors in water at concentrations below 10 nM, and<br />

therefore no evidence was obtained for micellar-type aggregates<br />

<strong>of</strong> these inhibitors.<br />

Inhibitor Partition to Lipid Vesicles. The fluorescence<br />

emission spectra <strong>of</strong> SB 242784 and INH-1 in the presence<br />

<strong>of</strong> DOPC vesicles are presented in Figure 3. DOPC and<br />

DOPG lipids were chosen for the reconstitution <strong>of</strong> the<br />

inhibitors because SB242784 was used by us in binding<br />

assays to selected trans<strong>membrane</strong> peptide fragments <strong>of</strong><br />

V-ATPase that required DOPC (and in some cases a small<br />

fraction <strong>of</strong> DOPG) for correct peptide conformation while<br />

inserted in liposomes (unpublished experiments). Using the<br />

cosolubilization method for incorporation <strong>of</strong> the inhibitors<br />

(see Materials and Methods section), the fluorescence<br />

emission maximum for SB 242784 is 436 nm whereas for<br />

INH-1 it is 461 nm. On the other hand, in DOPC-DOPG<br />

(80:20 mol/mol) bilayers, the SB 242784 fluorescence


5274 Biochemistry, Vol. 45, No. 16, 2006 Fernandes et al.<br />

FIGURE 3: Absorption and fluorescence emission spectra <strong>of</strong> the<br />

indole class <strong>of</strong> V-ATPase inhibitors. Fluorescence emission spectra<br />

were obtained using excitation wavelengths corresponding to the<br />

maxima <strong>of</strong> each inhibitor: (thin dashed line) absorption spectrum<br />

<strong>of</strong> INH-1; (thin solid line) absorption spectrum <strong>of</strong> SB 242784; (thick<br />

dashed line) fluorescence spectrum <strong>of</strong> INH-1; (thick solid line)<br />

fluorescence spectrum <strong>of</strong> SB 242784.<br />

emission maximum was 432 nm. For the alternative incorporation<br />

method <strong>of</strong> adding the inhibitor to the preformed<br />

vesicles, the fluorescence emission maxima were slightly blue<br />

shifted (≈3 nm) when comparing <strong>with</strong> the previous values,<br />

but after a small period <strong>of</strong> time (≈5 min) they became<br />

identical to the ones obtained through the cosolubilization<br />

method, thus suggesting that the initially formed aggregates<br />

are disrupted upon <strong>membrane</strong> interaction. The quantum<br />

yields <strong>of</strong> SB 242784 and INH-1 determined in the presence<br />

<strong>of</strong> 2 mM DOPC LUVs were 0.15 and 0.10, respectively.<br />

These values are not biased by the fraction <strong>of</strong> inhibitors in<br />

the aqueous phase, because at this lipidic concentration it<br />

can be concluded from the determined partition coefficients<br />

(see below) that the inhibitors are almost completely<br />

incorporated in the <strong>membrane</strong> (>95%), and in addition, the<br />

quantum yield in water is negligible.<br />

Although an extremely large blue shift in the inhibitor<br />

fluorescence is observed upon incorporation in lipid vesicles<br />

(e.g., the maximum emission wavelength <strong>of</strong> SB 242784 in<br />

water is 505 nm), this is not per se evidence for a large extent<br />

<strong>of</strong> partitioning <strong>of</strong> the two molecules to the lipid environment,<br />

because, as mentioned above, the quantum yield in the<br />

aqueous medium is very low due to aggregation. To obtain<br />

a quantitative description, both the extent and kinetics <strong>of</strong><br />

the partition <strong>of</strong> both inhibitors to DOPC vesicles were<br />

studied. Partition coefficients were determined through<br />

monitoring <strong>of</strong> the change in fluorescence emission intensity<br />

(I) vs lipid concentration [L] (Figure 4) and applying the<br />

equation (28):<br />

I ) I w + K p γ L [L]I L<br />

1 + K p γ L [L]<br />

(5)<br />

where I w is the fluorescence <strong>of</strong> the fluorophore in water,<br />

I L is the fluorescence <strong>of</strong> the fluorophore incorporated in the<br />

lipid vesicles, K p is the partition coefficient defined in terms<br />

<strong>of</strong> solute concentrations in each phase, and γ L is the lipid<br />

molar volume [0.78 dm -3 mol -1 for DOPC (29)]. The results<br />

FIGURE 4: Increase in steady-state fluorescence emission intensity<br />

<strong>of</strong> indole V-ATPase inhibitors <strong>with</strong> lipid concentration: (b) SB<br />

242784 incorporated in DOPC vesicles by the cosolubilization<br />

method; (O) INH-1 incorporated in DOPC vesicles by the method<br />

<strong>of</strong> addition after lipid solubilization. The buffer used was 10 mM<br />

Tris and 150 mM NaCl, pH ) 7.4. The lines represent the fits <strong>of</strong><br />

eq 5 to the data, and the partition coefficients are shown in Table<br />

1.<br />

Table 1: V-ATPase Inhibitor Partition Coefficients to Lipid<br />

Vesicles<br />

system<br />

inhibitor<br />

cosolubilization<br />

<strong>with</strong> DOPC<br />

cosolubilization<br />

<strong>with</strong> DOPC-DOPG<br />

addition to<br />

liposomes<br />

<strong>with</strong> DOPC<br />

SB 242784 (1.20 ( 0.08) × 10 4 (6.64 ( 1.76) × 10 3 (1.29 ( 0.23) × 10 4<br />

INH-1 (1.49 ( 0.26) × 10 4 (2.84 ( 0.99) × 10 4<br />

from a nonlinear fitting <strong>of</strong> eq 5 to the experimental data are<br />

shown in Table 1.<br />

For SB 242784, the anisotropy obtained while adding the<br />

inhibitor to preformed vesicles was 〈r〉 ) 0.33, the same<br />

value being obtained through the cosolubilization method.<br />

In vesicles composed <strong>of</strong> DOPC-DOPG (4:1 mol/mol), SB<br />

242784 anisotropy was virtually identical (〈r〉 ) 0.32). The<br />

INH-1 anisotropy in samples <strong>with</strong> preformed vesicles (〈r〉<br />

) 0.24) was lower than observed for the same molecule<br />

incorporated by cosolubilization (〈r〉 ) 0.31). Again, these<br />

experiments were carried out at a lipidic concentration for<br />

which the fraction <strong>of</strong> light emitted from inhibitor molecules<br />

in water is negligible.<br />

Inhibitor TransVerse Location in Lipid Vesicles. The<br />

effects on the inhibitors’ fluorescence emission spectral shape<br />

and intensity caused by the presence <strong>of</strong> nitroxide-labeled lipid<br />

in DOPC vesicles were investigated. A 10% ratio <strong>of</strong> labeled<br />

(5- or 12-DOX-PC) to nonlabeled lipids was used. For SB<br />

242784, there was a very slight shift on the fluorescence<br />

emission maximum, depending on the nitroxide-labeled lipid<br />

present. In the absence <strong>of</strong> quenchers the fluorescence<br />

emission maximum was at 436 nm, whereas in the presence<br />

<strong>of</strong> 5-DOX-PC this wavelength was 435 nm. Using 12-DOX-<br />

PC a red shift was obtained, the maximum being now 438<br />

nm (Figure 5A). For INH-1, there was no difference in the<br />

shape <strong>of</strong> the emission spectrum in the presence <strong>of</strong> 12-DOX-<br />

PC (maxima were at 461 nm), but using 5-DOX-PC a very<br />

small blue shift (2 nm) was obtained (Figure 5B). These shifts<br />

are evidence that the two nitroxide labels, which are localized<br />

at different depths in the <strong>membrane</strong>, are selectively quenching<br />

inhibitor populations which lie at different transverse<br />

positions.


V-ATPase Indole Inhibitor <strong>Interaction</strong> <strong>with</strong> Bilayers Biochemistry, Vol. 45, No. 16, 2006 5275<br />

FIGURE 6: Orientation <strong>of</strong> the transition moment <strong>of</strong> SB 242784,<br />

obtained from TD-DFT calculations (see text for details).<br />

Table 2: Order Parameters and Lagrange Coefficients <strong>of</strong> the Two<br />

Inhibitors in DOPC Multilayers Obtained from Linear Dichroism<br />

Studies a 〈P 2〉 〈P 4〉 λ 2 λ 4<br />

SB 242784 0.15 -0.21 0.99 -2.31<br />

INH-1 0.18 -0.31 1.27 -4.67<br />

a<br />

See text for details.<br />

FIGURE 5: Fluorescence emission spectra <strong>of</strong> indole V-ATPase<br />

inhibitors in the absence (s) and presence <strong>of</strong> different nitroxidelabeled<br />

lipids: (- - -) 5-DOX-PC and (‚‚‚) 12-DOX-PC). (A) SB<br />

242784. (B) INH-1.<br />

Quantitative information on the inhibitor location in the<br />

<strong>membrane</strong> can be obtained via the so-called parallax method<br />

(30) using the equation:<br />

z cf ) L cl + (-ln(F 1 /F 2 )/πC - L 21 2 )/2L 21 (6)<br />

where z cf is the distance <strong>of</strong> the fluorophore from the center<br />

<strong>of</strong> the bilayer, F 1 is the fluorescence intensity in the presence<br />

<strong>of</strong> the surface-located quencher (5-DOX-PC), F 2 is the<br />

fluorescence intensity in the presence <strong>of</strong> the quencher located<br />

deeper in the <strong>membrane</strong> (12-DOX-PC), L c1 is the distance<br />

<strong>of</strong> the shallow quencher from the center <strong>of</strong> the bilayer, L c2<br />

is the distance <strong>of</strong> the deep quencher from the center <strong>of</strong> the<br />

bilayer, L 21 is the distance between the shallow and deep<br />

quenchers (L c1 - L c2 ), and C is the concentration <strong>of</strong> the<br />

quencher in molecules/Å 2 . Using the distances <strong>of</strong> the<br />

nitroxide quenchers from the center <strong>of</strong> the bilayer given by<br />

Abrams and London (31), the obtained position for INH-1<br />

was <strong>of</strong> 11.9 Å from the center <strong>of</strong> the bilayer, whereas for<br />

SB 242784 this value was 12.8 Å.<br />

To gain further information regarding the inhibitors’<br />

location in the <strong>membrane</strong>, acrylamide quenching assays were<br />

also performed. Acrylamide is a hydrophilic quencher, and<br />

therefore differences in the extent <strong>of</strong> quenching by acrylamide<br />

allow a direct comparison <strong>of</strong> the bilayer penetration<br />

by the fluorophores. Acrylamide quenched both inhibitors<br />

<strong>with</strong> very similar efficiencies (results not shown), and a very<br />

slightly increased quenching observed for INH-1 is due to<br />

the also slight difference in average lifetimes (16) for the<br />

two inhibitors while incorporated in DOPC vesicles (〈τ〉 INH-1<br />

) 0.60 ns and 〈τ〉 SB 242784 ) 0.55 ns), because the Stern-<br />

Volmer rate constant is the product between the effective<br />

bimolecular rate constant (which can be decreased due to<br />

fluorophore shielding from the quencher) and the intrinsic<br />

lifetime <strong>of</strong> the fluorophore.<br />

Inhibitor Orientation in Lipid Vesicles. From the linear<br />

dichroism methodology, information regarding the orientation<br />

<strong>of</strong> the transition moment is obtained. In this way, to obtain<br />

information about the orientation <strong>of</strong> the molecule regarding<br />

the director <strong>of</strong> the system (normal to the <strong>membrane</strong><br />

interface), it is necessary to know the orientation <strong>of</strong> the<br />

transition moment relative to the molecular axes. To this<br />

effect, a quantum chemical calculation was carried out.<br />

Applying time-dependent density functional theory (TD-<br />

DFT) to an energy-optimized geometry <strong>of</strong> SB 242784, the<br />

orientation <strong>of</strong> the transition dipole moment relative to the<br />

molecule structure was obtained. As shown in Figure 6, the<br />

transition dipole moment is almost parallel to the molecular<br />

axis defined by the double bond conjugated system, and an<br />

approximately identical transition dipole moment orientation<br />

can be considered for INH-1, as both molecules have very<br />

similar fluorophores and the molecular differences are not<br />

expected to largely influence this property.<br />

The orientation parameters 〈P 2 〉 and 〈P 4 〉 and Lagrange<br />

coefficients obtained for INH-1 and SB 242784 from the<br />

linear dichroism measurements are presented in Table 2. The<br />

corresponding orientational density probability functions are<br />

shown in Figure 7.<br />

DISCUSSION<br />

Inhibitor Aggregation and Partition to Lipid Vesicles. It<br />

is surprising that INH-1 appears to aggregate in aqueous<br />

solution more readily and to a larger extent than SB 242784.<br />

The presence <strong>of</strong> the piperidine ring in the latter was expected


5276 Biochemistry, Vol. 45, No. 16, 2006 Fernandes et al.<br />

FIGURE 7: Orientation distribution function in DMPC <strong>of</strong> SB 242784<br />

(- - -) and INH-1 (s).<br />

FIGURE 8: Proposed topography and orientation <strong>of</strong> SB 242784 in<br />

a DOPC bilayer. The molecule is depicted from the estimated most<br />

probable angle and depth (see Discussion).<br />

to result in a smaller hydrophilicity, but this does not seem<br />

to be the case. Regarding the extent <strong>of</strong> aggregation, one<br />

possibility is that the piperidine ring induces a different type<br />

<strong>of</strong> packing in SB 242784 aggregates which quenches the<br />

molecule fluorescence less efficiently than in INH-1 aggregates,<br />

resulting in partially fluorescent aggregates. For<br />

INH-1 aggregates it is reasonable to assume that these would<br />

consist <strong>of</strong> stacked molecules <strong>with</strong> π-π interactions, but for<br />

SB 242784, this type <strong>of</strong> π-π stacking would not be able to<br />

exclude the piperidine ring from the aqueous environment<br />

and another aggregation geometry would have to come into<br />

play. The larger self-quenching <strong>of</strong> INH-1 in water would<br />

then not be solely dictated by the extent <strong>of</strong> aggregation but<br />

also related to a more effective geometry <strong>of</strong> packing<br />

regarding the quenching <strong>of</strong> fluorescence. The absence <strong>of</strong> a<br />

micellar type <strong>of</strong> aggregate for these inhibitors is most<br />

probably related to the absence <strong>of</strong> significative amphiphilic<br />

character.<br />

Both inhibitors partition <strong>with</strong> a high efficiency to lipid<br />

vesicles, and for SB 242784 the partition coefficient is<br />

independent <strong>of</strong> the method <strong>of</strong> incorporation (Table 1).<br />

Nevertheless, the partition to the lipid phase is only complete<br />

at moderately high lipid concentrations (for a K p ≈ 12000,<br />

95% <strong>of</strong> the total inhibitor population is incorporated only at<br />

a lipid concentration <strong>of</strong> 2 mM).<br />

For INH-1, the molecule partitions slightly more effectively<br />

when added to the preformed liposomes, but the<br />

differences in the partition coefficients are close to the<br />

uncertainty <strong>of</strong> the measurements. On thermodynamic grounds<br />

the two partition coefficients should in fact be identical upon<br />

reaching equilibrium. However, in most situations, the one<br />

from cosolubilization is higher than adding the solute from<br />

the aqueous solution, this being the case among other<br />

situations when stable aggregates (e.g., micelles) are formed<br />

in water. The identical values observed for the two methodologies<br />

<strong>of</strong> incorporation are consistent <strong>with</strong> the absence<br />

<strong>of</strong> a cmc previously described. The lower value for the<br />

partition coefficient <strong>of</strong> SB 242784 in DOPC-DOPG bilayers<br />

indicates a decreased partition to the bilayer upon the change<br />

in headgroup composition. This effect will be addressed in<br />

detail below when the fluorescence emission shifts <strong>of</strong> the<br />

inhibitors are discussed.<br />

INH-1 anisotropy was smaller when it was added to<br />

preformed vesicles (〈r〉 ) 0.24) when compared <strong>with</strong> the<br />

one observed for the same molecule incorporated via<br />

cosolubilization (〈r〉 )0.31), and this is likely to be related<br />

to energy migration inside aggregates in the <strong>membrane</strong>. As<br />

there is an overlap between the absorption and fluorescence<br />

emission spectra <strong>of</strong> the inhibitors (Figure 3), energy migration<br />

(energy homotransfer) is operative for two molecules<br />

<strong>of</strong> inhibitor lying in close proximity as in the case <strong>of</strong> an<br />

aggregate. The INH-1 Förster radius (R 0 )(16) for energy<br />

migration while in lipid <strong>membrane</strong>s is 11 Å (result not<br />

shown), and therefore for INH-1 in close proximity fluorescence<br />

depolarization is expected. The effects <strong>of</strong> energy<br />

migration can only be observed via the fluorescence anisotropy<br />

which is expected to decrease, since the fluorescence<br />

lifetime remains constant in this type <strong>of</strong> photophysical<br />

interaction. At variance <strong>with</strong> the aggregates in water, these<br />

aggregates must be fluorescent; i.e., there is no selfquenching,<br />

as in order to change the steady-state anisotropy<br />

they have to contribute to a significant fraction <strong>of</strong> the total<br />

fluorescence. This feature is probably related to the specific<br />

geometry <strong>of</strong> the aggregate in the <strong>membrane</strong>. In addition, in<br />

the case that quantitative dimer formation is invoked, the<br />

intermolecular distance would be 12.7 Å as concluded from<br />

the theoretical formalism (32); this is slightly larger than the<br />

distance for a collisional complex, so this would mean that<br />

no contact (static) fluorescence quenching would happen.<br />

Possibly, INH-1, when added to preformed vesicles, either<br />

(i) segregates to a <strong>membrane</strong>-perturbed region, and what is<br />

being observed is not strictly aggregate formation but<br />

increased probability <strong>of</strong> insertion in a <strong>membrane</strong>-perturbed<br />

area, leading to a higher local monomer concentration, such<br />

as observed for micelles (33) and <strong>membrane</strong>s (34), or (ii) is<br />

only partially aggregated (<strong>with</strong> intermolecular distance


V-ATPase Indole Inhibitor <strong>Interaction</strong> <strong>with</strong> Bilayers Biochemistry, Vol. 45, No. 16, 2006 5277<br />

structural differences <strong>of</strong> the two molecules (presence <strong>of</strong> the<br />

piperididine ring in SB 242784).<br />

Inhibitor TransVerse Location and Orientation in Lipid<br />

Vesicles. The large blue shifts observed upon the incorporation<br />

<strong>of</strong> the inhibitors in lipid vesicles are an indication that<br />

the inhibitors are not adsorbed to the surface <strong>of</strong> the bilayer<br />

but are somehow buried, having some contact <strong>with</strong> the acyl<br />

chain region. The wavelengths <strong>of</strong> maximum fluorescence<br />

emission for both inhibitors (436 and 461 nm) are almost<br />

the same as the ones observed for the same molecules in<br />

acetone (27). Taking as a reference the dielectric constant<br />

<strong>of</strong> acetone (ɛ ) 20.7), the corresponding region <strong>of</strong> PC<br />

bilayers <strong>with</strong> the same polarity properties lies in the<br />

headgroup region, outside the hydrocarbon core <strong>of</strong> the bilayer<br />

(35, 36), which for DOPC can be roughly located at >14 Å<br />

from the center <strong>of</strong> the bilayer (29). On the other hand, SB<br />

242784 in solution only exhibited such a blue-shifted<br />

emission when solubilized in nonprotic solvents, and as such<br />

this inhibitor is likely not to be in contact <strong>with</strong> a high density<br />

<strong>of</strong> water molecules when in lipid bilayers, excluding the<br />

possibility <strong>of</strong> a strictly superficial position, and therefore a<br />

positioning in the polar/apolar interface <strong>of</strong> the bilayer is more<br />

likely. Although other factors such as hydrogen bonding<br />

might affect the emission properties <strong>of</strong> the fluorophores, it<br />

is worthwhile noting that in addition to both inhibitors having<br />

fluorescence properties almost identical to those observed<br />

in acetone, they are located almost at the same distance to<br />

the center <strong>of</strong> the bilayer, as observed by acrylamide quenching<br />

and the parallax method, the latter positioning SB 242784<br />

and INH-1 at 12.8 and 11.9 Å from the center <strong>of</strong> the bilayer<br />

(corresponding approximately to the position <strong>of</strong> the fourth<br />

to third carbon <strong>of</strong> the DOPC acyl chain in the fluid state),<br />

respectively. The difference in the position between the two<br />

inhibitors is


5278 Biochemistry, Vol. 45, No. 16, 2006 Fernandes et al.<br />

wavelengths. At variance, the fraction <strong>of</strong> molecules deeper<br />

inside the <strong>membrane</strong> is emitting more in the blue. This would<br />

explain the observed shifts in fluorescence emission maximum<br />

in the presence <strong>of</strong> quenchers at different depths in the<br />

bilayer for SB 242784. The tilted configuration <strong>of</strong> SB<br />

242784, and the resulting poor packing <strong>with</strong> the bilayer<br />

lipids, can be the explanation for the perturbations on lipid<br />

mobility induced by the presence <strong>of</strong> this V-ATPase inhibitor<br />

in the bilayer as detected by ESR (40).<br />

The orientational probability density function (Figure 7)<br />

<strong>of</strong> SB 242784 has a small contribution near 90°. This<br />

contribution is very likely to be spurious and probably results<br />

from the data analysis method (15), as it seems unlikely that<br />

a population <strong>of</strong> the inhibitors lies parallel to the bilayer<br />

surface in view <strong>of</strong> the other results presented in this study,<br />

and bilayer defects that could induce a closer to random<br />

orientation <strong>of</strong> the molecule were absent in the INH-1 samples<br />

in the same lipid system. On the other hand, the contribution<br />

to the orientational probability density function <strong>of</strong> SB 242784<br />

for smaller angles relative to the bilayer normal could<br />

originate from a fraction <strong>of</strong> SB 242784 presenting orientations<br />

closer to the bilayer normal, due to the presence <strong>of</strong> the<br />

piperidine ring, which, as already mentioned, is expected to<br />

prefer the hydrocarbon core <strong>of</strong> the bilayer and therefore to<br />

dislocate the inhibitor toward the bilayer center.<br />

CONCLUSIONS<br />

Summarizing our results, both inhibitors readily aggregate<br />

in an aqueous environment and present relatively large<br />

partition coefficients to lipid bilayers. It is reasonable to place<br />

the fluorophore section <strong>of</strong> the inhibitors close to the polar/<br />

apolar interface <strong>of</strong> the bilayer, in contact <strong>with</strong> the headgroup<br />

region but still protected from the aqueous environment. The<br />

conjugated double bonds are likely to be in contact <strong>with</strong> the<br />

first carbons <strong>of</strong> the lipid acyl chains, and for the SB 242784<br />

V-ATPase inhibitor the piperidine ring should insert deeper<br />

among the acyl chains. The localization <strong>of</strong> the molecules in<br />

bilayers seems to be almost completely dictated by the<br />

fluorophore moiety, shared to a large extent by the two<br />

molecules. In this way, and assuming that the first step is<br />

the inhibitor incorporation in the <strong>membrane</strong> followed by<br />

diffusion to the protein, it would be likely that the site for<br />

interaction in the protein would be at a similar position inside<br />

the <strong>membrane</strong> to maximize the diffusional encounters.<br />

The two inhibitors studied here are samples <strong>of</strong> the entire<br />

indole class <strong>of</strong> V-ATPase inhibitors. SB 242784 exhibits very<br />

high inhibitory activities against osteoclastic V-ATPase, <strong>with</strong><br />

efficient inhibition in nanomolar concentrations, whereas<br />

INH-1 is capable <strong>of</strong> inhibition <strong>of</strong> the enzyme only in<br />

micromolar concentrations (5-7). According to our results,<br />

it can be inferred that this high difference in inhibitory<br />

efficiency inside the class <strong>of</strong> V-ATPase indole inhibitors is<br />

essentially due to specific molecular recognition processes<br />

<strong>with</strong> the enzyme inhibition site. The behavior in <strong>membrane</strong>s<br />

<strong>of</strong> the two molecules is very similar, namely, its effective<br />

concentration in the <strong>membrane</strong>, since the partition coefficients<br />

do not vary significantly. This strong discrimination<br />

by the enzyme reflected on their IC 50 values (30 µM and 26<br />

nM for INH-1 and SB 242784, respectively) is interesting,<br />

considering the close similarity <strong>of</strong> the two molecules.<br />

Experiments are under way to study the interactions <strong>of</strong><br />

V-ATPase inhibitors <strong>with</strong> selected trans<strong>membrane</strong> segments<br />

<strong>of</strong> the enzyme to gain further insight into the proteininhibitor<br />

recognition process.<br />

ACKNOWLEDGMENT<br />

We thank Benedito Cabral for the TD-DFT calculations<br />

<strong>with</strong> SB 242784.<br />

REFERENCES<br />

1. Baron, R., Neff, L., Louvard, D., and Courtoy, P. J. (1986) Cellmediated<br />

extracellular acidification and bone resorption: evidence<br />

for a low pH in resorbing lacunae and localization <strong>of</strong> a 100-kD<br />

lysosomal <strong>membrane</strong> protein at the osteoclast ruffled border, J.<br />

Cell Biol. 101, 2210-2222.<br />

2. Blair, H. C., Teitelbaum, S. L., Ghiselli, R., and Gluck, S. (1989)<br />

Osteoclastic bone resorption by a polarized vacuolar proton pump,<br />

Science 245, 855-857.<br />

3. Bowman, E. J., Siebers, A., and Altendorf, K. (1988) Bafilomycins:<br />

a class <strong>of</strong> inhibitors <strong>of</strong> <strong>membrane</strong> ATPases from microorganisms,<br />

animal cells, and plant cells, Proc. Natl. Acad. Sci. U.S.A.<br />

85, 7972-7976.<br />

4. Sundquist, K., Lakkakorpi, P., Wallmark, B., and Väänänen, K.<br />

(1990) Inhibition <strong>of</strong> osteoclast proton transport by bafilomycin<br />

A1 abolishes bone resorption, Biochem. Biophys. Res. Commun.<br />

168, 309-313.<br />

5. Gagliardi, S., Nadler, G., Consolandi, E., Parini, C., Morvan, M.,<br />

Legave, M. N., Belfiore, P., Zocchetti A., Clarke, G. D., James,<br />

I., Nambi, P., Gowen, M., and Farina, C. (1998) 5-(5,6-Dichloro-<br />

2-indolyl)-2-methoxy-2,4-pentadienamides: Novel and selective<br />

inhibitors <strong>of</strong> the vacuolar H + -ATPase <strong>of</strong> osteoclasts <strong>with</strong> bone<br />

antiresorptive activity, J. Med. Chem. 41, 1568-1573.<br />

6. Nadler, G., Morvan, M., Delimoge, I., Belfiore, P., Zocchetti, A.,<br />

James, I., Zembryki, D., Lee-Rycakzewski, E., Parini, C., Consolandi,<br />

E., Gagliardi, S., and Farina, C. (1998) (2Z,4E)-5-(5,6-<br />

Dichoro-2-indolyl)-2-methoxy-N-(1,2,2,6,6-pentamethylpiperidin-<br />

4-yl)-2,4-pentadienamide, a novel, potent and selective inhibitor<br />

<strong>of</strong> the osteoclast V-ATPase, Bioorg. Med. Chem. Lett. 8, 3621-<br />

3626.<br />

7. Visentin, L., Dodds, R. A., Valente, M., Misiano, P., Bradbeer, J.<br />

N., Oneta, S., Liang, X., Gowen, M., and Farina, C. (2000) A<br />

selective inhibitor <strong>of</strong> the osteoclastic V-H + -ATPase prevents bone<br />

loss in both thyroparathyroidectomized and ovariectomized rats,<br />

J. Clin. InVest. 106, 309-318.<br />

8. Mattsson, J. P., and Keeling, D. J. (1996) [ 3 H]Bafilomycin as a<br />

probe for the trans<strong>membrane</strong> proton channel <strong>of</strong> the osteoclast<br />

vacuolar H(+)-ATPase, Biochim. Biophys. Acta 1280, 98-106.<br />

9. Zhang, J., Feng, Y., and Forgac, M. (1994) Proton conduction<br />

and bafilomycin binding by the V o domain <strong>of</strong> the coated vesicle<br />

V-ATPase, J. Biol. Chem. 269, 23518.<br />

10. Crider, B. P., Xie, X. S., and Stone, D. K. (1994) Bafilomycin<br />

inhibits proton flow through the H + channel <strong>of</strong> vacuolar proton<br />

pumps, J. Biol. Chem. 269, 17379-17381.<br />

11. Wang, Y., Inoue, T., and Forgac, M. (2005) Subunit a <strong>of</strong> the yeast<br />

V-ATPase participates in binding <strong>of</strong> bafilomycin, J. Biol. Chem.<br />

280, 40481-40488.<br />

12. Dixon, N., Páli, T., Kee, T. P., and Marsh, D. (2004) Spin labelled<br />

V-ATPase inhibitors in lipid <strong>membrane</strong>s, Biochim. Biophys. Acta<br />

1665, 177-183.<br />

13. Mayer, L. D., Hope, M. J., and Cullis, P. R. (1986) Vesicles <strong>of</strong><br />

variable sizes produced by a rapid extrusion procedure, Biochim.<br />

Biophys. Acta 858, 161-168.<br />

14. McClare, C. (1971) An accurate and convenient organic phosphorus<br />

assay, Anal. Biochem. 39, 527-530.<br />

15. Castanho, M. A. R. B., Lopes, S., and Fernandes, M. (2003) Using<br />

UV-Vis. linear dichroism to study the orientation <strong>of</strong> molecular<br />

probes and biomolecules in lipidic <strong>membrane</strong>s, Spectroscopy 17,<br />

377-398.<br />

16. Lakowicz, J. R. (1999) Principles <strong>of</strong> Fluorescence Spectroscopy,<br />

2nd ed., Kluwer Academic/Plenum Publishers, New York.<br />

17. Chen, R., and Bowman, R. L. (1965) Fluorescence polarization:<br />

measurements <strong>with</strong> ultraviolet-polarizing filters in a spectrophot<strong>of</strong>luorometer,<br />

Science 147, 729-732.<br />

18. Eaton, D. F. (1988) Reference materials for fluorescence measurement,<br />

Pure Appl. Chem. 60, 1107-1114.


V-ATPase Indole Inhibitor <strong>Interaction</strong> <strong>with</strong> Bilayers Biochemistry, Vol. 45, No. 16, 2006 5279<br />

19. Loura, L. M. S., Fedorov, A., and Prieto, M. (2000) Membrane<br />

probe distribution heterogeneity: a resonance energy transfer<br />

study, J. Phys. Chem. B 104, 6920-6931.<br />

20. Marquardt, D. W. (1963) An algorithm for least squares estimation<br />

<strong>of</strong> non linear parameters, J. Soc. Ind. Appl. Math. 11, 431-441.<br />

21. Kooyman, R. P. H., Levine, Y. K., and Van der Meer, B. W. (1981)<br />

Measurement <strong>of</strong> second and fourth rank order parameters by<br />

fluorescence polarization experiments in a lipid <strong>membrane</strong> system,<br />

Chem. Phys. 60, 317-326.<br />

22. Lopes, S., and Castanho, M. A. R. B. (2002) Revealing the<br />

orientation <strong>of</strong> nystatin and amphotericin B in lipidic multibilayers<br />

by UV-Vis linear dichroism, J. Phys. Chem. B 106, 7278-7282.<br />

23. Parr, R. G., and Yang, W. (1989) Density Functional Theory <strong>of</strong><br />

Atoms and Molecules, Oxford University Press, Oxford.<br />

24. Becke, A. D. (1993) Density-functional thermochemistry. 3. The<br />

role <strong>of</strong> exact exchange, J. Chem. Phys. 98, 5648.<br />

25. Lee, C., Yang, W., and Parr, R. G. (1988) Development <strong>of</strong> the<br />

Colle-Salvetti correlation-energy formula into a functional <strong>of</strong> the<br />

electron density, Phys. ReV. 37, 785-789.<br />

26. Hariharan P. C., and Pople, J. A. (1974) Accuracy <strong>of</strong> AHn,<br />

equilibrium geometries by single determinant molecular orbital<br />

theory, Mol. Phys. 27, 209.<br />

27. Fernandes, F., Loura, L. M. S., Koehorst, R., Hemminga, M., and<br />

Prieto, M. (2004) <strong>Interaction</strong> <strong>of</strong> the indole class <strong>of</strong> V-ATPase<br />

inhibitors <strong>with</strong> lipid bilayers, Biophys. J. 88, 240a.<br />

28. Santos, N. C., Prieto, M., and Castanho, M. A. R. B. (2003)<br />

Quantifying molecular partition into model systems <strong>of</strong> bio<strong>membrane</strong>s:<br />

an emphasis on optical spectroscopic methods, Biochim.<br />

Biophys. Acta 1612, 123-135.<br />

29. Tristram-Nagle, S., Petrache, H. I., and Nagle, J. F. (1998)<br />

Structure and interactions <strong>of</strong> fully hydrated dioleoylphosphatidylcholine<br />

bilayers, 75, 917-925.<br />

30. Chattopadhyay, A., and London, E. (1987) Parallax method for<br />

direct measurement <strong>of</strong> <strong>membrane</strong> penetration depth utilizing<br />

fluorescence quenching by spin-labeled phospholipids, Biochemistry<br />

26, 39-45.<br />

31. Abrams, F. S., and London, E. (1992) Calibration <strong>of</strong> the parallax<br />

fluorescence quenching method for determination <strong>of</strong> <strong>membrane</strong><br />

penetration depth: refinement and comparison <strong>of</strong> quenching by<br />

spin-labeled and brominated lipids, Biochemistry 31, 5312-5322.<br />

32. Runnels, L. W., and Scarlata, S. F. (1995) Theory and application<br />

<strong>of</strong> fluorescence homotransfer to melittin oligomerization, Biophys.<br />

J. 69, 1569-1583.<br />

33. Berberan-Santos, M. N., and Prieto, M. J. E. (1987) Energy transfer<br />

in spherical geometry. Application to micelles, J. Chem. Soc.,<br />

Faraday Trans. 83, 1391-1409.<br />

34. de Almeida, R. F. M., Loura, L. M. S., Prieto, M., Watts, A.,<br />

Fedorov, A., and Barrantes, F. J. (2004) Cholesterol modulates<br />

the organization <strong>of</strong> the M4 trans<strong>membrane</strong> domain <strong>of</strong> the muscle<br />

nicotinic acetylcholine receptor, Biophys. J. 86, 2261-2272.<br />

35. Koehorst, R. B. M., Spruijt, R. B., Vergeldt, F. J., and Hemminga,<br />

M. A. (2004) Lipid bilayer topology <strong>of</strong> the trans<strong>membrane</strong> alfahelix<br />

<strong>of</strong> M13 major coat protein and bilayer polarity pr<strong>of</strong>ile by<br />

site-directed fluorescence spectroscopy, Biophys. J. 87, 1445-<br />

1455.<br />

36. White, S. H., and Wimley, W. C. (1998) Hydrophobic interactions<br />

<strong>of</strong> <strong>peptides</strong> <strong>with</strong> <strong>membrane</strong> interfaces, Biochim. Biophys. Acta<br />

1376, 339-352.<br />

37. Contreras, L. M., de Almeida, R. F. M., Villalaín, J., Fedorov,<br />

A., and Prieto, M. (2001) <strong>Interaction</strong> <strong>of</strong> -melanocyte stimulating<br />

hormone <strong>with</strong> binary phospholipid <strong>membrane</strong>s: structural changes<br />

and relevance <strong>of</strong> phase behaviour, Biophys. J. 80, 2273-2283.<br />

38. Yau, W., Wimley, W. C., Gawrisch, K., and White, S. H. (1998)<br />

The Preference <strong>of</strong> tryptophan for <strong>membrane</strong> interfaces, Biochemistry<br />

37, 14713-14718.<br />

39. de Planque, M. R. R., Kruijtzer, J. A. W., Liskamp, R. M. J.,<br />

Marsh, D., Greathouse, D. V., Koeppe II, de Kruijff, B., and<br />

Killian, J. A. (1999) Different <strong>membrane</strong> anchoring positions <strong>of</strong><br />

tryptophans and lysine in synthetic trans<strong>membrane</strong> alpha-helical<br />

<strong>peptides</strong>, J. Biol. Chem. 274, 20839-20846.<br />

40. Páli, T., Dixon, N., Kee, T. P., and Marsh, D. (2004) Incorporation<br />

<strong>of</strong> the V-ATPase inhibitors concanamycin and indole pentadiene<br />

in lipid <strong>membrane</strong>s. Spin-label EPR <strong>studies</strong>, Biochim. Biophys.<br />

Acta 1663, 14-18.<br />

BI0522753


BINDING ASSAYS OF INHIBITORS TOWARDS SELECTED<br />

V-ATPase DOMAINS<br />

3. BINDING ASSAYS OF INHIBITORS TOWARDS<br />

SELECTED V-ATPase DOMAINS<br />

95


Biochimica et Biophysica Acta 1758 (2006) 1777–1786<br />

www.elsevier.com/locate/bbamem<br />

Binding assays <strong>of</strong> inhibitors towards selected V-ATPase domains<br />

F. Fernandes a , L.M.S. Loura a,b, ⁎ , A. Fedorov a , N. Dixon c , T.P. Kee c ,<br />

M. Prieto a , M.A. Hemminga d<br />

a Centro de Química-Física Molecular, Instituto Superior Técnico, Lisbon, Portugal<br />

b Departamento de Química, Universidade de Évora, Évora, Portugal<br />

c Department <strong>of</strong> Chemistry, University <strong>of</strong> Leeds, Leeds LS2 9JT, UK<br />

d Laboratory <strong>of</strong> Biophysics, Wageningen University, Wageningen, The Netherlands<br />

Received 14 March 2006; received in revised form 4 May 2006; accepted 13 July 2006<br />

Available online 21 July 2006<br />

Abstract<br />

The macrolide antibiotic bafilomycin and the related synthetic compound SB 242784 are potent inhibitors <strong>of</strong> the vacuolar H + -ATPases (V-<br />

ATPase). It is currently believed that the site <strong>of</strong> action <strong>of</strong> these inhibitors is located on the <strong>membrane</strong> bound c-subunits <strong>of</strong> V-ATPases. To address<br />

the identification <strong>of</strong> the critical inhibitors binding domain, their specific binding to a synthetic peptide corresponding to the putative 4th<br />

trans<strong>membrane</strong> segment <strong>of</strong> the c-subunit was investigated using fluorescence resonance energy transfer (FRET), and for this purpose a specific<br />

formalism was derived. Another peptide <strong>of</strong> the corresponding domain <strong>of</strong> the c′ is<strong>of</strong>orm, was checked for binding <strong>of</strong> bafilomycin, since it is not<br />

clear if V-ATPase inhibition can also be achieved by interaction <strong>of</strong> the inhibitor <strong>with</strong> the c′-subunit. It was concluded that bafilomycin binds to the<br />

selected <strong>peptides</strong>, whereas SB 242784 was unable to interact, and in addition for bafilomycin, its interaction <strong>with</strong> the <strong>peptides</strong> either corresponding<br />

to the c- or the c′-subunit is<strong>of</strong>orms is identical. Since the observed interactions are however much weaker as compared to the very efficient binding<br />

<strong>of</strong> both bafilomycin and SB 242784 to the whole protein, it can be concluded that assembly <strong>of</strong> all V-ATPase trans<strong>membrane</strong> segments is required<br />

for an efficient interaction.<br />

© 2006 Elsevier B.V. All rights reserved.<br />

Keywords: V-ATPase; Bafilomycin A 1 ; Fluorescence; FRET<br />

1. Introduction<br />

Vacuolar ATPases (V-ATPases) are responsible for proton<br />

pumping in acidic organelles and plasma <strong>membrane</strong>s <strong>of</strong><br />

eukaryotic cells and their activity is essential in a large variety<br />

<strong>of</strong> cellular processes [1,2]. V-ATPases are composed <strong>of</strong> two<br />

domains (Fig. 1A), a soluble domain (V 1 ), where the catalytic<br />

center for ATP hydrolysis is located, and a <strong>membrane</strong> bound<br />

domain (V 0 ), responsible for proton pumping. V 0 contains<br />

multiple subunits including the 16-kDa proteolipids (subunits c)<br />

(4–5 copies) and their is<strong>of</strong>orms c′ and c″ (one copy each) [3],<br />

expected to play a major role in proton translocation. The<br />

⁎ Corresponding author. Centro de Química-Física Molecular, Complexo I,<br />

Instituto Superior Técnico, Lisbon, Portugal.<br />

E-mail address: pclloura@alfa.ist.utl.pt (L.M.S. Loura).<br />

c-subunit contains 4 putative trans<strong>membrane</strong> domains, and an<br />

essential glutamate residue located in the 4th putative<br />

trans<strong>membrane</strong> sequence is expected to be the carrier <strong>of</strong><br />

protons across the <strong>membrane</strong> through protonation/deprotonation<br />

events [4].<br />

Through site-directed mutagenic <strong>studies</strong> it was reported that<br />

mutations in three residues <strong>of</strong> the c-subunit <strong>of</strong> the V-ATPase<br />

from Neurospora crassa conferred resistance to bafilomycin A 1<br />

(Fig. 1B) [5], a macrolide antibiotic inhibitor <strong>of</strong> V-ATPases<br />

produced by Streptomyces griseus [6]. Two <strong>of</strong> these residues<br />

reside in the same side <strong>of</strong> the putative 4th trans<strong>membrane</strong> helix<br />

(F136 and Y143), and this helix side is exposed to the lipid<br />

environment [7]. Immobilization <strong>of</strong> spin-labeled residues <strong>of</strong> c-<br />

subunit by concanamycin A, a macrolide antibiotic very similar<br />

in structure to bafilomycin A 1 , has also been observed [8]. Also<br />

recently, six other sites where mutations induced resistance to<br />

bafilomycin A 1 were reported, and three <strong>of</strong> these are located in<br />

0005-2736/$ - see front matter © 2006 Elsevier B.V. All rights reserved.<br />

doi:10.1016/j.bbamem.2006.07.006


1778 F. Fernandes et al. / Biochimica et Biophysica Acta 1758 (2006) 1777–1786<br />

Fig. 1. (A) V-ATPase structure (adapted from [2]). (B) V-ATPase inhibitors SB 242784 and bafilomycin A 1 .<br />

the putative 4th trans<strong>membrane</strong> segment <strong>of</strong> c-subunit [9]. Of the<br />

latter, two residues are located in the same helix side as F136<br />

and Y143, pointing to the existence <strong>of</strong> an inhibitor binding site<br />

on this protein surface. It is not clear yet if residues from other<br />

helices (1 and 2) could also be part <strong>of</strong> this binding site and form<br />

a binding pocket [9], as this structure would imply a different<br />

helix organization than the one suggested from cysteine crosslinking<br />

experiments [10].<br />

The exact location and structure <strong>of</strong> the inhibitor binding site<br />

remains a matter <strong>of</strong> discussion, and in the present study we aim<br />

to detect the molecular requirements for bafilomycin A 1 binding<br />

through the use <strong>of</strong> a reductionist approach, using a simpler<br />

model system, a synthetic c-subunit derived peptide expected to<br />

mimic the putative 4th trans<strong>membrane</strong> segment <strong>of</strong> the V-<br />

ATPase c-subunit (Fig. 2). Synthetic <strong>peptides</strong> have proven to be<br />

useful tools in the characterization <strong>of</strong> ligand binding-sites on<br />

<strong>proteins</strong> and in some cases these isolated <strong>peptides</strong> were capable<br />

<strong>of</strong> ligand binding efficiencies identical to the intact protein [11–<br />

16]. Additionally, peptide fragments corresponding to trans<strong>membrane</strong><br />

sequences in intact <strong>membrane</strong> <strong>proteins</strong> were shown<br />

to specifically disrupt the intact protein activity (likely by<br />

competing <strong>with</strong> native trans<strong>membrane</strong> domains) [17], and<br />

functional <strong>membrane</strong> <strong>proteins</strong> have already been reformed from<br />

reconstituted isolated trans<strong>membrane</strong> segments [18–20], proving<br />

that isolated <strong>peptides</strong> corresponding to trans<strong>membrane</strong><br />

fragments are able to maintain specific information even when<br />

incorporated in the absence <strong>of</strong> its neighboring trans<strong>membrane</strong><br />

Fig. 2. Primary sequence <strong>of</strong> the model <strong>peptides</strong>. The sequences correspond to the<br />

putative 4th trans<strong>membrane</strong> segment <strong>of</strong> the V-ATPase c-(H4) and c′-(H4 c′ )<br />

subunit from Saccharomyces cerevisiae. Peptide H4 A ← E corresponds to a<br />

modification <strong>of</strong> peptide H4 in which the glutamate residue was replaced by an<br />

alanine. This mutation is underlined. The <strong>peptides</strong> were flanked by lysines to<br />

increase the <strong>membrane</strong> anchoring.<br />

segments. Similar to a previous study on synthetic <strong>peptides</strong> that<br />

mimic the 7th putative helix <strong>of</strong> the Vph1p subunit <strong>of</strong> yeast V-<br />

ATPase [21], flanking lysines were included in the design <strong>of</strong> the<br />

model <strong>peptides</strong> used in the present study to stabilize the<br />

trans<strong>membrane</strong> orientation due to the anchoring effect <strong>of</strong> the<br />

lysine residue.<br />

Another V-ATPase inhibitor, (2Z,4E)-5-(5,6-dichloro-2-<br />

indolyl)-2-methoxy-N-(1,2,2,6,6-pentamethylpiperidin-4-yl)-<br />

2,4-pentadienamide (SB 242784) (Fig. 1B), a synthetic<br />

molecule developed by Farina and coworkers [22] which is<br />

based on the presumptive pharmacophore <strong>of</strong> bafilomycin A 1<br />

[23], and whose mechanism <strong>of</strong> inhibition is therefore expected<br />

to be similar, was proposed as an agent for a new therapeutical<br />

strategy in osteoporosis disease due to its specificity for the<br />

osteoclastic form <strong>of</strong> the enzyme [22,24]. Bafilomycines are<br />

powerful toxins, since they do not present specificity for any<br />

particular type <strong>of</strong> V-ATPases, inhibiting indiscriminately all<br />

enzymes, and thus preventing their therapeutical use. The<br />

reason for the difference in selectivity between the two<br />

inhibitors is still unclear, especially when noting that the<br />

inhibitor acts on the c-subunit that is extremely conserved<br />

between c-subunits from enzymes <strong>of</strong> different origins. A<br />

possible explanation would be the participation <strong>of</strong> the a-subunit<br />

(part <strong>of</strong> the enzyme <strong>membrane</strong> bound domain—V 0 )(Fig. 1A) in<br />

the inhibitory mechanism, since the is<strong>of</strong>orm a3 is expressed<br />

primarily in osteoclasts [25], and recent mutational <strong>studies</strong><br />

support this hypothesis [26]. However, in a recent study using<br />

fluorescence resonance energy transfer (FRET), it was reported<br />

that SB 242784 and concanamycin bound competitively to<br />

isolated c-subunits, and SB 242784 was shown to bind to these<br />

as strongly as to the entire V 0 domain, apparently excluding a<br />

contribution <strong>of</strong> the a-subunit to the inhibitor binding mechanism<br />

[27]. The same authors also detected binding efficiencies <strong>of</strong> c-<br />

subunit for SB 242784 consistent <strong>with</strong> the inhibitor′s IC 50<br />

values, whereas for concanamycin these were slightly lower<br />

than expected [27]. Mutation <strong>of</strong> the tyrosine residue (Y143)<br />

from the 4th putative trans<strong>membrane</strong> segment had different<br />

effects on the conferred degree <strong>of</strong> resistance <strong>of</strong> yeast to<br />

concanamycin and to SB 242784, even though both were


F. Fernandes et al. / Biochimica et Biophysica Acta 1758 (2006) 1777–1786<br />

1779<br />

increased [8]. The binding sites for both inhibitors seem, for this<br />

reason, to be overlapping but not identical.<br />

Subunits c and c′ present a very large degree <strong>of</strong> sequence<br />

similarity, but mutations in subunit c′ homologous to those in<br />

subunit c producing bafilomycin A 1 insensitive strains have no<br />

effect in bafilomycin resistance [9]. On the basis <strong>of</strong> these results,<br />

Bowman and coworkers suggested that if c-subunit is<strong>of</strong>orms<br />

possess bafilomycin-binding sites, then these should have lower<br />

inhibitor affinities than in the c-subunit.<br />

Using the UV absorption properties <strong>of</strong> bafilomycin A 1<br />

(acceptor), we applied FRET in an assay (Tyr is the donor) to<br />

check for the existence <strong>of</strong> binding between the inhibitor and<br />

synthetic <strong>peptides</strong> corresponding to the putative 4th trans<strong>membrane</strong><br />

segment <strong>of</strong> the c and c′-subunits from the V-<br />

ATPase <strong>of</strong> Saccharomyces cerevisiae in a lipidic environment.<br />

The same methodology was applied in binding assays <strong>with</strong><br />

SB 242784. We detected binding <strong>of</strong> bafilomycin A 1 to the<br />

selected <strong>peptides</strong>, while for SB 242784 no interaction took<br />

place.<br />

2. Materials<br />

1,2-Dioleoyl-sn-glycero-3-phosphoethanolamine-N-(7-<br />

nitro-2-1,3-benzoxadiazol-4-yl) (NBD-DOPE), 1,2-Dioleoylsn-glycero-3-phosphocholine<br />

(DOPC), 1,2-Dioleoyl-sn-glycero-3-[Phospho-rac-(1-glycerol)]<br />

(DOPG), 1,2-Dierucoyl-snglycero-3-phosphocholine<br />

(DEuPC), 2-Lauroyl-sn-glycero-3-<br />

phosphocholine (DLPC) and 1,2-Miristoyl-sn-glycero-3-phosphocholine<br />

(DMPC) were obtained from Avanti Polar Lipids<br />

(Birmingham, AL). (2Z,4E)-5-(5,6-dichloro-2-indolyl)-2-methoxy-N-(1,2,2,6,6-pentamethylpiperidin-4-yl)-2,4-pentadienamide<br />

(SB 242784) was synthesized as described elsewhere [22].<br />

Bafilomycin A 1 was obtained from LC Laboratories (Woburn,<br />

MA). Peptides H4, H4 A ← E , and H4c′ (see Fig. 2) were<br />

synthesized by Pepceuticals (Nottingham, UK). Trifluoroacetic<br />

acid (TFA), 16-DOXYL-stearic acid (2-(14-Carboxytetradecyl)-2-ethyl-4,4-dimethyl-3-oxazolidinyloxy)<br />

and 5-DOXYLstearic<br />

acid (2-(3-Carboxypropyl)-4,4-dimethyl-2-tridecyl-3-<br />

oxazolidinyloxy) were obtained from Sigma-Aldrich (St.<br />

Louis, USA). Trifluoroethanol (TFE) was obtained from<br />

Acrōs Organics (Geel, Belgium). Other fine chemicals were<br />

obtained from Merck (Darmstadt, Germany).<br />

3. Experimental procedures<br />

3.1. Sample preparation<br />

The <strong>peptides</strong> (Fig. 2) were solubilized in 100 μl <strong>of</strong> TFA and immediately<br />

dried under a N 2(g) flow. Following that, the peptide was suspended in TFE.<br />

When the TFA solubilization step was not introduced, the solubility in TFE was<br />

greatly reduced, and the peptide aggregation levels after reconstitution in lipid<br />

bilayers were enhanced.<br />

For peptide reconstitution in lipid bilayers, the desired amount <strong>of</strong><br />

phospholipids and solubilized peptide (and <strong>of</strong> the inhibitors in the inhibitor<br />

binding <strong>studies</strong>) were mixed in chlor<strong>of</strong>orm and dried under a N 2(g) flow. The<br />

sample was then kept in vacuum overnight. Liposomes were prepared <strong>with</strong><br />

buffer Tris 10 mM pH 7.4. The hydration step was performed <strong>with</strong> gentle<br />

addition <strong>of</strong> buffer at a temperature above the phospholipid main transition<br />

temperature.<br />

For the fluorescence quenching experiments <strong>with</strong> the N-DOXYL-stearic<br />

acids, the nitroxide labeled fatty acids were 10% <strong>of</strong> the total lipid (molar<br />

fraction).<br />

3.2. CD measurements<br />

CD spectroscopy was performed on a Jasco J-720 spectropolarimeter <strong>with</strong> a<br />

450 W Xe lamp. Samples for CD spectroscopy were extruded 8 times on a<br />

homemade extruder using polycarbonate filters <strong>of</strong> 0.1 μm. Peptide concentration<br />

was always 40 μM.<br />

3.3. Fluorescence spectroscopy<br />

Steady-state fluorescence measurements were carried out <strong>with</strong> an SLM-<br />

Aminco 8100 Series 2 spectr<strong>of</strong>luorimeter (<strong>with</strong> double excitation and emission<br />

monochromators, MC400) in a right angle geometry. The light source was a<br />

450 W Xe arc lamp and for reference a Rhodamine B quantum counter solution<br />

was used. Correction <strong>of</strong> emission spectra was performed using the correction<br />

s<strong>of</strong>tware <strong>of</strong> the apparatus. 5 ×5 mm quartz cuvettes were used. All<br />

measurements were performed at room temperature.<br />

The emission spectrum from the tyrosine <strong>of</strong> peptide H4 was recorded<br />

using an excitation wavelength <strong>of</strong> 270 nm. The tyrosine quantum yield <strong>of</strong><br />

peptide H4 in lipid bilayers was determined using quinine sulfate (ϕ=0.55)<br />

as a reference [28].<br />

4. Results<br />

4.1. Characterization <strong>of</strong> reconstituted <strong>peptides</strong><br />

When using a peptide as a model for a protein section, it is important that it retains the structural properties <strong>of</strong> the latter [21].<br />

Therefore, it was important for the present study to accurately know the behavior <strong>of</strong> the <strong>peptides</strong> when incorporated in lipid<br />

bilayers. The characterization focused on secondary structure, aggregation and position in the bilayer. The 4th trans<strong>membrane</strong><br />

segment from which the peptide is derived is expected to assume a α-helical structure inside the V-ATPase subunit c, and<br />

therefore conditions that favor this type <strong>of</strong> structure should be determined and used throughout the study.<br />

CD spectra <strong>of</strong> peptide H4 incorporated in DOPC at a lipid to protein ratio (L/P-molar ratio) <strong>of</strong> 50 revealed the presence <strong>of</strong><br />

different types <strong>of</strong> secondary structures (Fig. 3). Deconvolution <strong>of</strong> the spectrum <strong>with</strong> the CDNN CD Spectra Deconvolution v.<br />

2.1 s<strong>of</strong>tware recovered a fraction <strong>of</strong> about 40% <strong>of</strong> α-helical structure, 29% was reported to be random coil, while the<br />

remaining was divided between the different types <strong>of</strong> β-sheet structure. When a L/P ratio <strong>of</strong> 25 was used, the fraction <strong>of</strong> α-<br />

helical structure present decreased to 30.5%. The presence <strong>of</strong> increasing amounts <strong>of</strong> β-sheet structure at lower L/P ratios is an<br />

indication <strong>of</strong> peptide aggregation [29]. It was not possible to obtain a reasonable quality CD spectrum <strong>with</strong> samples <strong>of</strong> larger<br />

L/P ratios as those used in the inhibitor binding <strong>studies</strong>, as the lipid concentration became too high and large light scattering<br />

problems arose.


1780 F. Fernandes et al. / Biochimica et Biophysica Acta 1758 (2006) 1777–1786<br />

Fig. 3. CD spectrum <strong>of</strong> peptide H4 incorporated in DOPC at an L/P <strong>of</strong> 50.<br />

Another method to study aggregation <strong>of</strong> <strong>peptides</strong> is to follow the quantum yield changes <strong>of</strong> the aromatic residues present.<br />

Fluorescence from tyrosine residues is known to be quenched by several mechanisms besides FRET, namely interactions <strong>with</strong> other<br />

residues side chains such as glutamate and aspartate, and uncharged forms <strong>of</strong> basic residues such as arginine, lysine and histidine, as<br />

well as electron transfer to peptide α-carbonyl groups [30,31]. Considering this, it can be expected that <strong>peptides</strong> forming aggregates<br />

will have a lower quantum yield than their monomeric counterparts. The quantum yield dependence <strong>of</strong> Tyr15 (corresponding to<br />

Tyr142 in the native protein [32]) from reconstituted <strong>peptides</strong> on the lipid composition is shown in Fig. 4. The presence <strong>of</strong> anionic<br />

phospholipids clearly increases the quantum yield <strong>of</strong> peptide H4. In liposome <strong>of</strong> pure zwitterionic phospholipids, the Tyr15 quantum<br />

yield trend is: DEuPC< DOPC


F. Fernandes et al. / Biochimica et Biophysica Acta 1758 (2006) 1777–1786<br />

1781<br />

Fig. 5. (A): Model used for FRET simulations. The donor (Tyr15) is located 6 Å from the center <strong>of</strong> the bilayer while the acceptors can be located at different planes,<br />

assuming a superficial location such as for NBD-DOPE (A 1 ) or more buried positions such as for the V-ATPase inhibitors (A 2 ). In case <strong>of</strong> aggregation the number <strong>of</strong><br />

protein–protein contacts increases while protein–lipid contacts decrease, as a result, R e (exclusion distance) increases. (B) H4 fluorescence quenching <strong>of</strong> Tyr15 <strong>of</strong><br />

peptide H4 due to FRET to NBD-DOPE in DOPC bilayers using a lipid to protein ratio <strong>of</strong> 50 (●) and 100 (○). Curve corresponds to an exclusion radius fit to the energy<br />

transfer data using equations 1–5 (—). A value <strong>of</strong> 9 Å was recovered for R e .(–––) FRET simulation for a random distribution <strong>of</strong> acceptors around a donor <strong>with</strong> an<br />

exclusion radius <strong>of</strong> 20 Å. n 2 is the superficial concentration <strong>of</strong> labeled phospholipids (molecules per Å 2 ). The range <strong>of</strong> NBD labeled phospholipid to total phospholipid<br />

ratios in this experiment was 0.4% to 2%. Inset: Overlap between tyrosine fluorescence emission ( ) and DOPE-NBD absorption spectrum (––), R 0 (Tyr-NBD) =22Å.<br />

On the basis <strong>of</strong> these results, the lipid composition chosen for the inhibitor-peptide binding <strong>studies</strong> by FRET measurements was<br />

100% DOPC at an L/P <strong>of</strong> 100, as this lipid is able to favor the trans<strong>membrane</strong> orientation for peptide H4. The problem <strong>of</strong> lateral<br />

aggregation in the <strong>membrane</strong> nevertheless potentially remained, as in DOPC the tyrosine quantum yield was significantly lower<br />

than in liposomes containing anionic phospholipids. The possibility <strong>of</strong> aggregation could however also be assessed, as described<br />

below.<br />

A FRET experiment was performed in order to determine the size <strong>of</strong> the possible aggregates. Using the Tyr15 as a donor and<br />

NBD labeled DOPE as acceptors (NBD-DOPE), FRET efficiencies were obtained (Fig. 5B). Fitting the available theoretical models<br />

for energy transfer between donor and acceptors in different planes [34] to our experimental data, we can recover an averaged<br />

exclusion radius (R e ) <strong>of</strong> the donor, defined as the minimum distance between the Tyr15 and the NBD labeled phospholipids (Fig.<br />

5A). For a monomeric peptide this value should be around 10 Å as this is approximately the sum <strong>of</strong> the radius <strong>of</strong> a α-helix backbone<br />

and that <strong>of</strong> a phospholipid molecule. Any significant increase from this value is likely to be reporting an extensive aggregation<br />

phenomenon.<br />

FRET efficiencies (E) are calculated from the degree <strong>of</strong> fluorescence emission quenching <strong>of</strong> the donor caused by the presence <strong>of</strong><br />

acceptors.<br />

E ¼ 1<br />

I DA<br />

I D<br />

¼ 1<br />

Z l<br />

0<br />

Z l<br />

i DA ðÞdt= t i<br />

0 D ðÞdt t<br />

ð1Þ<br />

i DA ðtÞ ¼i D ðtÞ:q interplanar ðtÞ<br />

ð2Þ<br />

where I DA and I D are the steady-state fluorescence intensities <strong>of</strong> the donor in the presence and absence <strong>of</strong> acceptors respectively.<br />

i DA and i D are the donor decays in the presence and absence <strong>of</strong> acceptors. ρ interplanar is the FRET contribution arising from<br />

energy transfer to randomly distributed acceptors in two different planes from the donors (two monolayer leaflets) [34].<br />

( Z l<br />

p<br />

) ( ffiffiffiffiffiffiffi 1<br />

q interplanar ¼ exp 2n 2 pl1<br />

2 l<br />

1 2 1 expð tb 3 Z l þR2 e<br />

1 a6 Þ<br />

p<br />

)<br />

ffiffiffiffiffiffiffi 2<br />

a 3 da exp 2n 2 pl2<br />

2 l<br />

2 2 1 expð tb 3 þR2 e<br />

2 a6 Þ<br />

a 3 da<br />

0<br />

0<br />

ð3Þ


1782 F. Fernandes et al. / Biochimica et Biophysica Acta 1758 (2006) 1777–1786<br />

where b i =(R 0 2 /l i ) 2 τ D −1/3 , R 0 is the Förster radius, n 2 is the acceptor density in each leaflet, and l 1 and l 2 are the distance between<br />

the plane <strong>of</strong> the donors and the two planes <strong>of</strong> acceptors. Using Eqs. (1)–(3), theoretical expectations for FRET efficiency in a<br />

random distribution <strong>of</strong> acceptors can be calculated and converted to I DA /I D ratios. The Förster radius is given by:<br />

R 0 ¼ 0:2108ðJj 2 n 4 / D Þ 1=6<br />

ð4Þ<br />

where J is the spectral overlap integral, κ 2 is the orientation factor, n is the refractive index <strong>of</strong> the medium, and ϕ D is the donor<br />

quantum yield. J is calculated as:<br />

Z<br />

J ¼ f ðkÞeðkÞk 4 dk<br />

ð5Þ<br />

where f(λ) is the normalized emission spectra <strong>of</strong> the donor and ε(λ) is the absorption spectra <strong>of</strong> the acceptor. The numeric factor<br />

in Eq. (4) assumes nm units for the wavelength λ and Å units for R 0 .<br />

Assuming the already determined value for Tyr15 position in the bilayer (6 Å from the center) and the distance from the center <strong>of</strong><br />

the bilayer <strong>of</strong> the NBD moiety in NBD labeled phospholipids (19 Å) [35], l 1 and l 2 were determined to be 13 and 25 Å, respectively.<br />

Using Eqs. (4) and (5) the Förster radius <strong>of</strong> the Tyr-NBD donor–acceptor pair was determined to be R 0 =22 Å. In the fitting<br />

procedure, these values are kept fixed and the only parameter being fitted is R e. The results from the experiment along <strong>with</strong> the curve<br />

fitted from the theoretical model, are presented in Fig. 5B. Both sets <strong>of</strong> data (L/P=50 and 100) were fitted <strong>with</strong> a value <strong>of</strong> R e =9 Å,<br />

matching the expectation for a monomeric peptide. In case there were aggregates at the protein concentrations used these must be<br />

small enough so that the distances between Tyr15 and the surrounding phospholipids are not affected. From Eqs. (1)–(3) and using<br />

the same parameter values described above as well as R e =20 Å (large aggregates), a FRET simulation was performed for the same<br />

system and the obtained donor quenching curve is compared to our experimental values in Fig. 5B.<br />

4.2. Peptide inhibitor binding <strong>studies</strong><br />

Both bafilomycin A 1 and SB 242784 absorb in the UV region and can act as acceptors for tyrosine in energy transfer experiments<br />

(spectra are shown in Fig. 6 and 7). The Förster radius for the Tyr-bafilomycin A 1 donor–acceptor pair is 20 Å, whereas for the Tyr-<br />

SB 242784 pair it is 24 Å (Eq. (4)). A 3 mM concentration <strong>of</strong> lipid was used to ensure an almost complete incorporation <strong>of</strong> inhibitors<br />

in the bilayer. SB 242784 partition coefficient to DOPC bilayers is 1.20×10 4 [36], and at the lipid concentration used 97% <strong>of</strong> the<br />

inhibitor molecules are incorporated in the bilayer. The fraction <strong>of</strong> inhibitors not incorporated in the vesicles was taken into account<br />

on the acceptor concentrations plot in Fig. 7, which depicts the Tyr15 quenching via energy transfer to SB 242784.<br />

In contrast to SB 242784, bafilomycin A 1 is not a fluorescent molecule and partition coefficients could not be determined using<br />

photophysical techniques. Overestimation <strong>of</strong> its partition coefficient could lead to an underestimation <strong>of</strong> bafilomycin-peptide binding<br />

constants. However, it was recently showed by EPR <strong>of</strong> spin-labeled lipids that the macrolide molecule concanamycin A, another<br />

powerful V-ATPase inhibitor <strong>with</strong> very similar structure to bafilomycin A 1 , readily incorporates in lipid <strong>membrane</strong>s [37]. Thus, it is<br />

to be expected from this result and from the high hydrophobic character <strong>of</strong> the molecule that the incorporation <strong>of</strong> bafilomycin A 1 at<br />

the lipid concentrations used is close to 100%.<br />

On the inhibitor-peptide binding assays, the L/P was kept at 100 in order to ensure minimum levels <strong>of</strong> aggregation. Experimental<br />

FRET efficiencies obtained <strong>with</strong> both peptide H4/SB 242784 and peptide H4/bafilomycin A 1 donor/acceptor pairs (Eq. (1)), are<br />

compared to theoretical expectations for a random distribution <strong>of</strong> acceptors (the scenario in which there is an absence <strong>of</strong> binding)<br />

obtained from Eqs. (1) and (2) (see Figs. 6 and 7). The position <strong>of</strong> SB 242784 in DOPC bilayers was already determined from<br />

selective quenching methodology (parallax method) to be 12.8 Å from the center <strong>of</strong> the bilayer [36], but the position <strong>of</strong> the<br />

bafilomycin A 1 chromophore was impossible to determine by the same methodology since this molecule is not fluorescent. In Fig.<br />

6A, the simulations corresponding to different positions <strong>of</strong> the bafilomycin A 1 chromophore inside the bilayer are shown, together<br />

<strong>with</strong> the experimental data. It is clear that even when assuming that the Tyr15 is located on the same bilayer plane as the bafilomycin<br />

A 1 chromophore (6 Å from the center <strong>of</strong> the bilayer), in a situation <strong>of</strong> maximal energy transfer efficiency, the experimental energy<br />

transfer efficiencies cannot be solely explained by the unbound population <strong>of</strong> inhibitor molecules (i.e. random distribution <strong>of</strong><br />

acceptors) and a fraction <strong>of</strong> peptide H4 must be binding bafilomycin A 1 . It is difficult to precisely quantify the binding constant (K b )<br />

for this process due to the uncertainty relative to the bafilomycin A 1 chromophore position in the bilayer, but a lower and a higher<br />

limit can be determined assuming the closest and furthest position possible for bafilomycin A 1 and Tyr15 (corresponding<br />

respectively to a position in the center and in the surface <strong>of</strong> the bilayer for bafilomycin A 1 ), and also that peptide H4/bafilomycin A 1<br />

complexes are completely non-fluorescent (Eqs. (6) and (7)). This last assumption is valid, since for a contact interaction either by<br />

Förster type or other transfer (exchange) mechanism, the efficiency <strong>of</strong> transfer would be 100%. The equations valid for the FRET<br />

efficiency in a scenario <strong>of</strong> peptide H4-bafilomycin A 1 1:1 complex formation are given below:<br />

E EXP ¼ E random þ H4 Baf ð1 E random Þ ð6Þ<br />

½H4Š T


F. Fernandes et al. / Biochimica et Biophysica Acta 1758 (2006) 1777–1786<br />

1783<br />

qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi<br />

½H4 Baf Š<br />

¼ 1 K bð½Baf Š T<br />

þ½H4Š T<br />

Þþ ð1 þ K b ð½Baf Š T<br />

þ½H4Š T<br />

ÞÞ 2 4ðKb 2½Baf Š T ½H4Š T Þ<br />

ð7Þ<br />

½H4Š T<br />

ð 2K b Þ½H4Š T<br />

where [H4-Baf] is the effective concentration (molecules per lipid volume as the binding is confined to the lipid phase) <strong>of</strong> the<br />

peptide–bafilomycin A 1 complex, and, [H4] T and [Baf] T are respectively, the effective analytical concentrations <strong>of</strong> the peptide and <strong>of</strong><br />

bafilomycin A 1 . E EXP is the energy transfer efficiencies obtained experimentally and E random is the theoretical energy transfer<br />

efficiency assuming a random distribution <strong>of</strong> acceptors (Eqs. (1)–(3)). In case binding is detected, the use <strong>of</strong> effective bafilomycin A 1<br />

concentrations will result in lower binding constants as compared to the ones <strong>of</strong> bulk concentrations, but the binding constants<br />

obtained in this way have a larger physical meaning as they will not change <strong>with</strong> lipid concentration. Anyway they can be easily<br />

interconverted.<br />

Assuming a position for bafilomycin A 1 in the center <strong>of</strong> the bilayer, the fitting procedure (Fig. 6A) recovered K b =10.5±1.32 M − 1<br />

(mol per volume <strong>of</strong> lipid). When the bafilomycin A 1 position was fixed to 20 Å from the center <strong>of</strong> the bilayer (i.e. at the surface), the<br />

value recovered was K b =36.2±2.8 M − 1 . Therefore, the binding constant upper and lower bounds for the peptide H4–bafilomycin<br />

A 1 complex are 10.5±1.32 13.72±3.05 M − 1 .<br />

Fig. 6. (A) Fluorescence emission quenching <strong>of</strong> Tyr15 <strong>of</strong> peptide H4 by FRET to bafilomycin A 1 (●). Theoretical expectations (Eqs. (1–3)) for FRET in the absence <strong>of</strong><br />

inhibitorbindingtopeptideH4assumingpositionsforbafilomycinA 1 <strong>of</strong>12.5Å(⋯),8.5Å(–––),and4.5Å(—) fromthecenter<strong>of</strong>thebilayer.( )Fitting<strong>of</strong>equations6 and<br />

7 to FRET data assuming a bafilomycin A 1 position <strong>of</strong> 4.5 Å from the center <strong>of</strong> the bilayer. The interval 10.5±1.34


1784 F. Fernandes et al. / Biochimica et Biophysica Acta 1758 (2006) 1777–1786<br />

Fig. 7. Fluorescence emission quenching <strong>of</strong> Tyr15 <strong>of</strong> peptide H4 by FRET to SB 242784 (●). Curve corresponds to the theoretical expectation for FRET in the absence<br />

<strong>of</strong> inhibitor binding to peptide H4 (see text). The range <strong>of</strong> SB 242784 to total phospholipid ratios in these experiments was 0.18% to 1.8%. Inset: overlap between<br />

tyrosine fluorescence emission ( ) and SB242784 absorption spectrum (–––), R 0 (Tyr-SB 242784)=24 Å.<br />

When the FRET binding assay was applied to the peptide H4/SB 242784 system, the results were substantially different (Fig.<br />

7). With this inhibitor, the obtained FRET efficiencies matched the theoretical expectations for a random distribution <strong>of</strong> acceptors,<br />

suggesting absence <strong>of</strong> binding between SB 242784 and the peptide. No uncertainty exists regarding this result as the position <strong>of</strong><br />

the inhibitor inside the bilayer is accurately know from parallax fluorescence quenching techniques (12.8 Å from the center <strong>of</strong> the<br />

bilayer — [36]).<br />

5. Discussion<br />

5.1. Use <strong>of</strong> peptide H4 as a model peptide for the 4th<br />

trans<strong>membrane</strong> segment <strong>of</strong> subunit c<br />

Aggregation <strong>of</strong> trans<strong>membrane</strong> <strong>peptides</strong> containing a hydrophobic<br />

sequence and flanked by basic residues has already been<br />

reported [38–40], thus detection <strong>of</strong> aggregates <strong>of</strong> peptide H4 at<br />

low L/P ratios (


F. Fernandes et al. / Biochimica et Biophysica Acta 1758 (2006) 1777–1786<br />

1785<br />

[26]. In the same way, c-subunit residues shown to be important<br />

for bafilomycin A 1 inhibitory activity through mutational<br />

<strong>studies</strong> [9] could not be directly involved in the binding <strong>of</strong> the<br />

inhibitor, but still play a role in the inhibition mechanism. Here,<br />

we proposed to further define the structural requirements for<br />

bafilomycin A 1 and SB 242784 binding to the V-ATPase, by<br />

focusing on the protein domain expected to be the primary<br />

contributor to the inhibitor binding site, the putative 4th<br />

trans<strong>membrane</strong> domain <strong>of</strong> the c-subunit. The extent <strong>of</strong> binding<br />

efficiencies achieved using only <strong>peptides</strong> corresponding to this<br />

section <strong>of</strong> the protein is an indicator <strong>of</strong> the relevance <strong>of</strong> the<br />

remaining protein domains to the binding <strong>of</strong> inhibitor, the first<br />

step in the inhibitory mechanism.<br />

No evidence was found supporting SB 242784 association<br />

<strong>with</strong> the peptide H4. On the other hand, bafilomycin A 1 was<br />

shown to possess only moderate affinity for the trans<strong>membrane</strong><br />

<strong>peptides</strong>. Due to uncertainties in bafilomycin A 1 position in<br />

bilayers it was impossible to rigorously quantify a binding<br />

constant for the process, but a reliable range for this parameter<br />

could be presented. The differences in bafilomycin A 1 binding<br />

efficiencies for the peptide H4, H4c′ and H4 A ← E are not<br />

significant as they fall <strong>with</strong>in, or close to the error <strong>of</strong> the fits. For<br />

H4c′ this result is somewhat surprising. Mutations on conserved<br />

residues <strong>of</strong> the c-subunit known to confer resistance to<br />

bafilomycin A 1 and concanamycin were shown to be ineffective<br />

in the c′ and c′ is<strong>of</strong>orms [9]. It was proposed that if a<br />

bafilomycin A 1 binding site was present on these is<strong>of</strong>orms, then<br />

it must possess lower affinity than in the c is<strong>of</strong>orm [9]. If this is<br />

the case, then the differences in binding efficiencies between the<br />

two is<strong>of</strong>orms must be explained by discrepancies in the<br />

interaction <strong>of</strong> the inhibitor <strong>with</strong> parts <strong>of</strong> the protein other than<br />

the highly conserved putative 4th trans<strong>membrane</strong> helix.<br />

Although the contribution from the 4th trans<strong>membrane</strong><br />

segment (TM4) to the binding site affinity for bafilomycin A 1 is<br />

not negligible, it is absolutely unable to explain the extremely<br />

low IC 50 <strong>of</strong> bafilomycin A 1 (0.1 nM). The binding efficiencies<br />

observed for concanamycin (IC 50 close to bafilomycin A 1 ) and<br />

the intact isolated c-subunit [27] were more than 1000 times<br />

larger than the ones recovered in the present study. Therefore, it<br />

is clear that either: (i) the other trans<strong>membrane</strong> segments from<br />

the c-subunit, namely TM1 and TM2, are essential in the<br />

formation <strong>of</strong> an efficient inhibitor binding site, or that (ii)<br />

interactions <strong>of</strong> TM4 <strong>with</strong> other protein segments are required for<br />

this protein section to assume the appropriate folding which<br />

enables the establishment <strong>of</strong> the necessary interactions for an<br />

efficient binding between the residues <strong>of</strong> the 4th trans<strong>membrane</strong><br />

segment <strong>of</strong> the c-subunit and the inhibitors. Absence <strong>of</strong> SB<br />

242784 binding to the peptide H4 could be explained by the<br />

much lower IC 50 <strong>of</strong> this inhibitor (26 nM in chicken osteoclasts<br />

[22]) that reflects a much lower binding site affinity for this<br />

inhibitor, and the absence <strong>of</strong> remaining c-subunit trans<strong>membrane</strong><br />

segments potentiates this effect.<br />

In this work, in addition to the detailed comparison <strong>of</strong> the<br />

different inhibitors and <strong>peptides</strong>, it was concluded that in<br />

contrast to some reported cases [11–16], there is a very<br />

significant difference in binding when comparing the binding to<br />

selected protein segments or to the whole protein. Although<br />

<strong>studies</strong> using <strong>peptides</strong> are very relevant, clearly in the case <strong>of</strong><br />

the interaction between the bafilomycin A 1 or SB 242784 <strong>with</strong><br />

the c-subunit from V-ATPase, the whole protein architecture<br />

and the environment that it provides, are key factors for an<br />

efficient binding.<br />

Acknowledgments<br />

This work was supported by contract no. QLG-CT-2000-<br />

01801 <strong>of</strong> the European Commission (MIVase consortium).<br />

M. H. and M. P. are members <strong>of</strong> the COST D22 Action <strong>of</strong><br />

the European Union. F. F. acknowledges grant SFRH/BD/<br />

14282/2003 from FCT (Portugal). This work was partially<br />

funded by FCT (Portugal) under the program POCTI.<br />

References<br />

[1] M.E. Finbow, M.A. Harrison, The vacuolar H+-ATPase: a universal<br />

proton-pump <strong>of</strong> eukaryotes, Biochem. J. 324 (1997) 697–712.<br />

[2] T. Nishi, M. Forgac, The vacuolar (H+)-ATPases—Nature's most versatile<br />

proton pumps, Mol. Cell. Biol. 3 (2002) 94–103.<br />

[3] B. Powell, L.A. Graham, T.H. Stevens, Molecular characterization <strong>of</strong> the<br />

yeast vacuolar H+-ATPase proton pore, J. Biol. Chem. 275 (2000)<br />

23654–23660.<br />

[4] M.A. Harrison, M.E. Finbow, J.B. Findlay, Postulate for the molecular<br />

mechanism <strong>of</strong> the vacuolar H(+)-ATPase (hypothesis), Mol. Membr. Biol.<br />

14 (1997) 1–3.<br />

[5] B.J. Bowman, E.J. Bowman, Mutations in subunit c <strong>of</strong> the vacuolar<br />

ATPase confer resistance to bafilomycin and identify a conserved<br />

antibiotic binding site, J. Biol. Chem. 277 (2002) 3965–3972.<br />

[6] G. Werner, H. Hagenmaier, H. Drautz, A. Baumgartner, H. Zähner,<br />

Metabolic products <strong>of</strong> microorganisms. Bafilomycins, isolation,<br />

chemical structure and biological activity, J. Antibiot. 37 (1984)<br />

110–117.<br />

[7] M. Harrison, B. Powell, M.E. Finbow, J.B.C. Findlay, Identification <strong>of</strong><br />

lipid-accessible sites on the Nephrops 16-kDa proteolipid incorporated into<br />

a hybrid vacuolar H+-ATPase: site-directed labeling <strong>with</strong> n-(1-pyrenyl)<br />

cyclohexylcarbodiimide and fluorescence quenching analysis, Biochemistry<br />

39 (2000) 7531–7537.<br />

[8] T. Páli, G. Whyteside, N. Dixon, T. Kee, S. Ball, M.A. Harrison, J.B.C.<br />

Findlay, M.E. Finbow, D. Marsh, <strong>Interaction</strong> <strong>of</strong> inhibitors <strong>of</strong> the vacuolar<br />

H+-ATPase <strong>with</strong> the trans<strong>membrane</strong> Vo-sector, Biochemistry 43 (2004)<br />

12297–12305.<br />

[9] E.J. Bowman, L.A. Graham, T.H. Stevens, B.J. Bowman, The bafilomycin/concanamycin<br />

binding site in subunit c <strong>of</strong> the V-ATPases from Neurospora<br />

crassa and Saccharomyces cerevisiae, J. Biol. Chem. 279 (2004)<br />

33131–33138.<br />

[10] M. Harrison, J. Murray, B. Powell, Y. Kim, M.E. Finbow, J.B.C. Findlay,<br />

Helical interactions and <strong>membrane</strong> disposition <strong>of</strong> the 16-kDa proteolipid<br />

subunit <strong>of</strong> the vacuolar H1-ATPase analyzed by cysteine replacement<br />

mutagenesis, J. Biol. Chem. 274 (1999) 25461–25470.<br />

[11] T.L. Lentz, V. Chaturverdi, B.M. Conti-Fine, Amino acids <strong>with</strong>in residues<br />

181–200 <strong>of</strong> the nicotinic acetylcholine receptor alpha 1 subunit in nicotine<br />

binding, Biochem. Pharmacol. 55 (1998) 341–347.<br />

[12] P.T. Wilson, T.L. Lentz, E. Hawrot, Determination <strong>of</strong> the primary amino<br />

acid sequence specifying the alpha-bungarotoxin binding site on the alpha<br />

subunit <strong>of</strong> the acetylcholine receptor from Torpedo californica, Proc. Natl.<br />

Acad. Sci. U. S. A. 82 (1985) 8790–8794.<br />

[13] D. Neumann, D. Barchan, M. Fridkint, S. Fuchs, Analysis <strong>of</strong> ligand<br />

binding to the synthetic dodecapeptide <strong>of</strong> the acetylcholine receptor a<br />

subunit, Proc. Natl. Acad. Sci. U. S. A. 83 (1986) 9250–9253.<br />

[14] A.H. Taylor, G. Heavner, M. Nedelman, D. Sherris, E. Brunt, D. Knight, J.<br />

Ghrayeb, Lipopolysaccharide (LPS) neutralizing <strong>peptides</strong> reveal a lipid A<br />

binding site <strong>of</strong> LPS binding protein, J. Biol. Chem. 270 (1995)<br />

17934–17938.


1786 F. Fernandes et al. / Biochimica et Biophysica Acta 1758 (2006) 1777–1786<br />

[15] C. Ried, C. Wahl, T. Miethke, G. Wellnh<strong>of</strong>er, C. Landgraf, J.<br />

Schneider-Mergener, A. Hoess, High affinity endotoxin-binding and<br />

neutralizing <strong>peptides</strong> based on the crystal structure <strong>of</strong> recombinant<br />

Limulus anti-lipopolysaccharide factor, J. Biol. Chem. 271 (1996)<br />

28120–28127.<br />

[16] A.C. Scatigno, S.S. Garrido, R. Marchetto, A 4.2 kDa synthetic peptide as<br />

a potential probe to evaluate the antibacterial activity <strong>of</strong> coumarin drugs,<br />

J. Peptide Sci. 10 (2004) 566–577.<br />

[17] S.R. George, G.Y.K. Ng, S.P. Lee, T. Fan, G. Varghese, C. Wang, C.M.<br />

Deber, P. Seeman, B.F. O'Dowd, Blockade <strong>of</strong> G protein-coupled<br />

receptors and the dopamine transporter by a trans<strong>membrane</strong> domain<br />

peptide: novel strategy for functional inhibition <strong>of</strong> <strong>membrane</strong> <strong>proteins</strong> in<br />

vivo, J. Pharmacol. Exp. Ther. 307 (2003) 481–489.<br />

[18] J.L. Popot, J. Trewhella, D.M. Engelman, Reformation <strong>of</strong> crystalline<br />

purple <strong>membrane</strong> from purified bacteriorhodopsin fragments, EMBO J.<br />

(1986) 3039–3044.<br />

[19] T. Marti, Refolding <strong>of</strong> bacteriorhodopsin from expressed polypeptide<br />

fragments, J. Biol. Chem. 273 (1998) 9312–9322.<br />

[20] S. Ozawa, R. Hayashi, A. Masuda, T. Iio, S. Takahashi, Reconstitution <strong>of</strong><br />

bacteriorhodopsin from a mixture <strong>of</strong> a proteinase V8 fragment and two<br />

synthetic <strong>peptides</strong>, Biochim. Biophys. Acta 1323 (1997) 145–153.<br />

[21] R.W. Hesselink, R.B.M. Koehorst, P.V. Nazarov, M.A. Hemminga,<br />

Membrane-bound <strong>peptides</strong> mimicking trans<strong>membrane</strong> Vph1p helix 7 <strong>of</strong><br />

yeast V-ATPase: a spectroscopic and polarity mismatch study, Biochim.<br />

Biophys. Acta 1716 (2005) 137–145.<br />

[22] G. Nadler, M. Morvan, I. Delimoge, P. Belfiore, A. Zocchetti, I. James, D.<br />

Zembryki, E. Lee-Rycakzewski, C. Parini, E. Consolandi, S. Gagliardi, C.<br />

Farina, (2Z,4E)-5-(5,6-dichloro-2-indolyl)-2-methoxy-n-(1,2,2,6,6-pentamethylpiperidin-4-yl)-2,4-pentadienamide,<br />

a novel, potent and selective<br />

inhibitor <strong>of</strong> the osteoclast V-ATPase, Bioorg. Med. Chem. Lett. 8 (1998)<br />

3621–3626.<br />

[23] S. Gagliardi, G. Nadler, E. Consolandi, C. Parini, M. Morvan, M.N. Legave,<br />

P. Belfiore, A. Zocchetti, G.D. Clarke, I. James, P. Nambi, M. Gowen, C.<br />

Farina, 5-(5,6-Dichloro-2-indolyl)-2-methoxy-2,4-pentadienamides: novel<br />

and selective inhibitors <strong>of</strong> the vacuolar H+-ATPase <strong>of</strong> osteoclasts <strong>with</strong> bone<br />

antiresorptive activity, J. Med. Chem. 41 (1998) 1568–1573.<br />

[24] L. Visentin, L.A. Dodds, M. Valente, P. Misiano, J.N. Bradbeer, S. Oneta,<br />

X. Liang, M. Gowen, C. Farina, A selective inhibitor <strong>of</strong> the osteoclastic<br />

V-H+-ATPase prevents bone loss in both thyroparathyroidectomized and<br />

ovariectomized rats, J. Clin. Invest. 103 (2000) 309–318.<br />

[25] M.F. Manolson, H. Yu, W. Chen, Y. Yao, K. Li, R.L. Lees, J.N.M.<br />

Heersche, The a3 is<strong>of</strong>orm <strong>of</strong> the 100-kDa V-ATPase subunit is highly but<br />

differentially expressed in large (10 nuclei) and small (5 Nuclei)<br />

osteoclasts, J. Biol. Chem. 278 (2003) 49271–49278.<br />

[26] Y. Wang, T. Inoue, M. Forgac, Subunit a <strong>of</strong> the yeast V-ATPase participates<br />

in binding <strong>of</strong> bafilomycin, J. Biol. Chem. 280 (2005) 40481–40488.<br />

[27] G. Whyteside, P.J. Meek, S.K. Ball, N. Dixon, M.E. Finbow, T.P. Kee,<br />

J.B.C. Findlay, M.A. Harrison, Concanamycin and indolyl pentadieneamide<br />

inhibitors <strong>of</strong> the vacuolar H+-ATPase bind <strong>with</strong> high affinity to the<br />

purified proteolipid subunit <strong>of</strong> the <strong>membrane</strong> domain, Biochemistry 44<br />

(2005) 15024–15031.<br />

[28] D.F. Eaton, Reference materials for fluorescence measurement, Pure Appl.<br />

Chem. 60 (1988) 1107–1114.<br />

[29] R. Cerpa, F.E. Cohen, I.D. Kuntz, Conformational switching in designed<br />

<strong>peptides</strong>: the helix/sheet transition, Fold. Des. 1 (1996) 91–101.<br />

[30] R.W. Cogwill, Tyrosil fluorescence in <strong>proteins</strong> and model <strong>peptides</strong>, in: R.F.<br />

Chen, H. Edelhoch (Eds.), Biochemical Fluorescence: Concepts 2, Marcel<br />

Dekker, New York, 1976, pp. 441–486.<br />

[31] J.W. Longworth, Excited states <strong>of</strong> <strong>proteins</strong> and nucleic acid, in: R.F.<br />

Steiner, I. Weinryb (Eds.), Plenum Press, New York, 1971, pp. 319–484.<br />

[32] H. Nelson, N. Nelson, The progenitor <strong>of</strong> ATP synthases was closely related<br />

to the current vacuolar H+-ATPase, FEBS Lett. 247 (1989) 147–153.<br />

[33] A. Chattopadhyay, E. London, Parallax method for direct measurement<br />

<strong>of</strong> <strong>membrane</strong> penetration depth utilizing fluorescence quenching by<br />

spin-labeled phospholipids, Biochemistry 26 (1987) 39–45.<br />

[34] L. Davenport, R.E. Dale, R.H. Bisby, R.B. Cundall, Transverse location <strong>of</strong><br />

the fluorescent probe 1,6-diphenyl-1,3,5-hexatriene in model lipid bilayer<br />

<strong>membrane</strong> systems by resonance energy transfer, Biochemistry 24 (1985)<br />

4097–4108.<br />

[35] F.S. Abrams, E. London, Extension <strong>of</strong> the parallax analysis <strong>of</strong> <strong>membrane</strong><br />

penetration depth to the polar region <strong>of</strong> model <strong>membrane</strong>s: use <strong>of</strong><br />

fluorescence quenching by a spin-label attached to the phospholipid polar<br />

headgroup, Biochemistry 32 (1993) 10826–10831.<br />

[36] F. Fernandes, L. Loura, R.B.M. Koehorst, N. Dixon, T.P. Kee, M.A.<br />

Hemminga, M. Prieto, <strong>Interaction</strong> <strong>of</strong> the indole class <strong>of</strong> V-ATPase<br />

inhibitors <strong>with</strong> lipid bilayers, Biochemistry 45 (2006) 5271–5279.<br />

[37] T. Páli, N. Dixon, T.P. Kee, D. Marsh, Incorporation <strong>of</strong> the V-ATPase<br />

inhibitors concanamycin and indole pentadiene in lipid <strong>membrane</strong>s, Spinlabel<br />

EPR <strong>studies</strong>, Biochim. Biophys. Acta 1663 (2004) 14–18.<br />

[38] J. Ren, S. Lew, J. Wang, E. London, Control <strong>of</strong> the trans<strong>membrane</strong><br />

orientation and interhelical interactions <strong>with</strong>in <strong>membrane</strong>s by hydrophobic<br />

helix length, Biochemistry 38 (1999) 5905–5912.<br />

[39] F.X. Zhou, H.J. Merianos, A.T. Brunger, D.M. Engelman, Polar residues<br />

drive association <strong>of</strong> polyleucine trans<strong>membrane</strong> helices, Proc. Natl. Acad.<br />

Sci. U. S. A. 98 (2000) 2250–2255.<br />

[40] S. Mall, R. Broadbridge, R.P. Sharma, J.M. East, A.G. Lee, Selfassociation<br />

<strong>of</strong> model trans<strong>membrane</strong> helix is modulated by lipid structure,<br />

Biochemistry 40 (2000) 12379–12386.


BINDING OF A QUINOLONE ANTIBIOTIC TO BACTERIAL<br />

PORIN OmpF<br />

IV<br />

BINDING OF A QUINOLONE<br />

ANTIBIOTIC TO BACTERIAL<br />

PORIN OmpF<br />

1. Introduction<br />

Fluoroquinolones (quinolones) have been shown to present broad-spectrum<br />

antimicrobial activity and are currently one <strong>of</strong> the most successful classes <strong>of</strong> drugs.<br />

They are reported as relatively well tolerated drugs <strong>with</strong> reasonably few undesirable<br />

effects (Albini and Monti, 2003; Stahlmann and Lode, 1999), and as a consequence<br />

their therapeutic application is increasing. The antibiotic activity <strong>of</strong> quinolones is the<br />

outcome <strong>of</strong> inhibition <strong>of</strong> a series <strong>of</strong> important enzymes for chromosome function and<br />

topology (homologous type II topoisomerases, DNA gyrase, and DNA topoisomerase<br />

IV) (Pestova et al., 2000).<br />

First generation quinolones (nalidixic acid) were used in the treatment <strong>of</strong><br />

infections caused by Gram-negative bacteria. More recently, new derivatives were<br />

developed <strong>with</strong> enhanced antibiotic activity against Gram-positive bacteria, namely<br />

Streptococus pneumoniae, responsible for infections in the respiratory system and other<br />

pathologies (Quiniliani et al., 1999). Some <strong>of</strong> the new generation quinolones are also<br />

apparently effective against multidrug resistant Mycobacterium tuberculosis and were<br />

shown to be useful in the treatment <strong>of</strong> AIDS patients infected <strong>with</strong> this pathogen<br />

(Houston and Farming, 1994; Hooper and Wolfson, 1995). These bacteria are resistant<br />

to a wide range <strong>of</strong> aggressive agents (acid/alcohol) due to the presence <strong>of</strong> a complex<br />

permeability barrier (Nikaido et al., 1993).<br />

The possible strategies for quinolones entry through the outer and inner<br />

<strong>membrane</strong>s <strong>of</strong> bacteria are via porin channels (Dechené et al., 1990; Chevalier et al.,<br />

2000), or hydrophobic diffusion through the lipid bilayer. The porin pathway is<br />

107


apparently the most significant as the efficiency <strong>of</strong> the quinolones is not directly related<br />

to the water/lipid partition coefficient <strong>of</strong> the molecules (Vazquez et al., 2001). Indeed,<br />

by either not expressing or expressing structurally changed outer <strong>membrane</strong> porins,<br />

entrance <strong>of</strong> quinolones in bacteria is severely impaired (Mascaretti, 2003; Chevalier et<br />

al., 2000).<br />

OmpF is one <strong>of</strong> the porins involved in the permeation process <strong>of</strong> quinolones<br />

across the outer <strong>membrane</strong> <strong>of</strong> bacteria (Chapman and Georgopapadakou, 1988). It was<br />

the first <strong>membrane</strong> protein to yield crystals <strong>of</strong> size and order allowing to be analyzed by<br />

X-ray crystallography (Garavito and Rosenbusch, 1980). This porin forms a homotrimer<br />

in the outer <strong>membrane</strong> and each monomer presents β-barrel structure in which one <strong>of</strong><br />

the loops folds back to the barrel, forming a constriction zone at half the height <strong>of</strong> the<br />

channel (Figure 1 in section 2 <strong>of</strong> this Chapter). This loop is expected to contribute<br />

significantly to the control <strong>of</strong> the permeability <strong>of</strong> OmpF.<br />

It is not yet know whether OmpF operates as a channel or as an enabler <strong>of</strong><br />

quinolones diffusion across the lipid-OmpF interface. Therefore, the knowledge <strong>of</strong> the<br />

exact type <strong>of</strong> interaction established by OmpF and quinolones is <strong>of</strong> great biological<br />

relevance, and could be potentially useful in the development <strong>of</strong> more effective drugs in<br />

the future. In this chapter, the interaction <strong>of</strong> OmpF and a second generation quinolone,<br />

cipr<strong>of</strong>loxacin (CP), was studied. CP is a 6-fluoroquinolone which structure is shown in<br />

Figure IV-1. Recent <strong>studies</strong> suggested a 1:1 stoichiometry for OmpF-CP interaction<br />

(Neves et al., 2005).<br />

Figure IV-1: Chemical structure <strong>of</strong> CP.<br />

108


BINDING OF A QUINOLONE ANTIBIOTIC TO BACTERIAL<br />

PORIN OmpF<br />

2. CIPROFLOXACIN INTERACTIONS WITH<br />

BACTERIAL PROTEIN OMPF: MODELLING<br />

OF FRET FROM A MULTI-TRYPTOPHAN<br />

PROTEIN TRIMER<br />

ABSTRACT<br />

The outer <strong>membrane</strong> protein F (OmpF) is known to play an important role in the uptake <strong>of</strong><br />

fluoroquinolone antibiotics by bacteria. In this study, the degree <strong>of</strong> binding <strong>of</strong> the fluoroquinolone<br />

antibiotic cipr<strong>of</strong>loxacin to OmpF in a lipid <strong>membrane</strong> environment is quantified using a methodology<br />

based on Förster resonance energy transfer (FRET). Analysis <strong>of</strong> the fluorescence quenching <strong>of</strong> OmpF is<br />

complex as each OmpF monomer presents two tryptophans at different positions, thus sensing two<br />

different distributions <strong>of</strong> acceptors in the bilayer plane. Specific FRET formalisms were derived<br />

accounting for the different energy transfer contributions to quenching <strong>of</strong> each type <strong>of</strong> tryptophan <strong>of</strong><br />

OmpF, allowing the recovery <strong>of</strong> upper and lower boundaries for the cipr<strong>of</strong>loxacin-OmpF binding constant<br />

(K B ). log (K B ) was found to lie in the range 3.15-3.62 or 3.58-4.00 depending on the location for the<br />

cipr<strong>of</strong>loxacin binding site assumed in the FRET modelling, closer to the centre or to the periphery <strong>of</strong> the<br />

OmpF trimer, respectively. This methodology is suitable for the analysis <strong>of</strong> FRET data obtained <strong>with</strong><br />

similar protein systems and can be readily adapted to different geometries.<br />

109


110


BINDING OF A QUINOLONE ANTIBIOTIC TO BACTERIAL<br />

PORIN OmpF<br />

CIPROFLOXACIN INTERACTIONS WITH BACTERIAL<br />

PROTEIN OMPF: MODELLING OF FRET FROM A<br />

MULTI-TRYPTOPHAN PROTEIN TRIMER<br />

Fábio Fernandes a , Patrícia Neves b , Paula Gameiro b , Luís M. S. Loura c , Manuel Prieto a, *<br />

a - Centro de Química-Física Molecular, Instituto Superior Técnico, Lisbon, Portugal.<br />

b - Requimte, Departamento de Química, Faculdade de Ciências, Porto, Portugal.<br />

c - Centro de Química and Departamento de Química, Universidade de Évora, Évora,<br />

Portugal.<br />

ABSTRACT<br />

The outer <strong>membrane</strong> protein F (OmpF) is known to play an important role in the uptake<br />

<strong>of</strong> fluoroquinolone antibiotics by bacteria. In this study, the degree <strong>of</strong> binding <strong>of</strong> the<br />

fluoroquinolone antibiotic cipr<strong>of</strong>loxacin to OmpF in a lipid <strong>membrane</strong> environment is<br />

quantified using a methodology based on Förster resonance energy transfer (FRET).<br />

Analysis <strong>of</strong> the fluorescence quenching <strong>of</strong> OmpF is complex as each OmpF monomer<br />

presents two tryptophans at different positions, thus sensing two different distributions<br />

<strong>of</strong> acceptors in the bilayer plane. Specific FRET formalisms were derived accounting<br />

for the different energy transfer contributions to quenching <strong>of</strong> each type <strong>of</strong> tryptophan<br />

<strong>of</strong> OmpF, allowing the recovery <strong>of</strong> upper and lower boundaries for the cipr<strong>of</strong>loxacin-<br />

OmpF binding constant (K B ). log (K B ) was found to lie in the range 3.15-3.62 or 3.58-<br />

4.00 depending on the location for the cipr<strong>of</strong>loxacin binding site assumed in the FRET<br />

modelling, closer to the centre or to the periphery <strong>of</strong> the OmpF trimer, respectively.<br />

This methodology is suitable for the analysis <strong>of</strong> FRET data obtained <strong>with</strong> similar<br />

protein systems and can be readily adapted to different geometries.<br />

Abbreviations<br />

CP-Cipr<strong>of</strong>loxacin; FRET- Förster resonance energy transfer; oPOE- n-<br />

octylpolyoxyethylene; OmpF- Outer Membrane protein F; DMPC-Dimyristoyl-L-αphosphatidylcholine;<br />

Trp- tryptophan; CMC- critical micellar concentration.<br />

111


Introduction<br />

Quinolones are broad-spectrum antibacterial agents which mechanism <strong>of</strong> action is the<br />

inhibition <strong>of</strong> DNA gyrase and DNA topoisomerase IV enzymes that control DNA<br />

topology and are vital for bacterial replication [1-3]. Access to the target site is a major<br />

determinant <strong>of</strong> antibacterial activity, <strong>with</strong> the outer <strong>membrane</strong> being the major<br />

permeability barrier in Gram negative bacteria [1]. In fact, one <strong>of</strong> the mechanisms <strong>of</strong><br />

resistance developed by the bacterial cell is the process <strong>of</strong> making more difficult the<br />

access <strong>of</strong> quinolones to their target <strong>of</strong> action, by either not expressing or expressing<br />

structurally changed outer <strong>membrane</strong> porins [3,4]. One <strong>of</strong> those porins, which<br />

microbiology <strong>studies</strong> related to the permeation <strong>of</strong> some quinolones through the outer<br />

<strong>membrane</strong>, is OmpF [1, 3-6]. Indeed, porin-deficient mutants <strong>of</strong> E. coli are resistant to<br />

fluoroquinolones, although the role <strong>of</strong> OmpF, either as a channel or as an enabler <strong>of</strong><br />

quinolone diffusion at the OmpF/ lipid interface, has not yet been elucidated. The<br />

relative importance and the different areas <strong>of</strong> contact <strong>of</strong> each quinolone <strong>with</strong> OmpF, is a<br />

subject <strong>of</strong> great importance in the context <strong>of</strong> developing new molecules <strong>with</strong> less<br />

resistance problems.<br />

OmpF is a trimer <strong>with</strong>in the <strong>membrane</strong> and it contains just two tryptophan residues per<br />

monomer (Fig.1), Trp 214 at the lipid protein interface and Trp 61 at the trimer interface<br />

[7]. The protein shows a maximum <strong>of</strong> emission at relatively low wavelengths, which<br />

suggests that both tryptophans are in hydrophobic environments. This is confirmed by<br />

experiments involving OmpF mutants [7], which lack one or both Trp’s.<br />

Cipr<strong>of</strong>loxacin (CP) (Fig. 2) is a 6-fluoroquinolone antibiotic currently under clinical use<br />

for which many resistances have been reported in a large number <strong>of</strong> microbial species.<br />

The aim <strong>of</strong> the present study was to investigate the role <strong>of</strong> OmpF as a major pathway for<br />

CP entry through the outer <strong>membrane</strong> into the bacterial cell by analysing the alteration<br />

<strong>of</strong> OmpF fluorescence in presence <strong>of</strong> increasing concentration <strong>of</strong> CP.<br />

In this way, the extent <strong>of</strong> Förster resonance energy transfer (FRET) between OmpF and<br />

CP was measured, which associated <strong>with</strong> a rational modelling <strong>of</strong> the distribution <strong>of</strong><br />

donor and acceptors in the bilayer allowed us to quantify the extent <strong>of</strong> binding between<br />

the protein and the antibiotic. The FRET modelling presented here is also suitable for<br />

application to systems <strong>with</strong> different geometries, and the new FRET formalisms derived<br />

can be readily adapted to the analysis <strong>of</strong> FRET data obtained <strong>with</strong> other large <strong>membrane</strong><br />

<strong>proteins</strong> presenting donor fluorophores in the protein-lipid interface.<br />

112


BINDING OF A QUINOLONE ANTIBIOTIC TO BACTERIAL<br />

PORIN OmpF<br />

Our study assumes significance in the overall context <strong>of</strong> the increasing problem <strong>of</strong><br />

bacterial resistance to the antibiotics currently in use, and the consequent need <strong>of</strong><br />

understanding the processes involved, in order to create new molecules <strong>with</strong> increased<br />

antibacterial activity and less resistance problems.<br />

Materials and methods<br />

Cipr<strong>of</strong>loxacin (CP) was a gift from Bayer (Leverkusen, Germany). N-(2-hydroxyethyl)<br />

piperazine-N’-ethanesulfonic acid (Hepes) was from Sigma (Sigma, St. Louis, MO).<br />

Octylpolyoxyethylene (oPOE) was from Bachem (Bubendorf, Switzerland) and all<br />

other chemicals were from Merck (Darmstadt, Germany). All solutions were prepared<br />

<strong>with</strong> 10 mM Hepes buffer (0.1 M NaCl; pH 7.4). OmpF was purified from E. coli, strain<br />

BL21 (DE3) Omp8, following published procedures (8). OmpF concentration was<br />

estimated using the bicinchoninic acid protein assay against bovine serum albumin as<br />

standard.<br />

All the absorption measurements were carried out <strong>with</strong> a UNICAM UV-300<br />

spectrophotometer equipped <strong>with</strong> a constant-temperature cell holder. Spectra were<br />

recorded at 37ºC in 1 cm quartz cuvettes in the range 230 to 350 nm. Fluorescence<br />

measurements were performed in a Varian spectr<strong>of</strong>luorimeter, model Cary Eclipse,<br />

equipped <strong>with</strong> a constant-temperature cell holder (Peltier single cell holder). All the<br />

spectra were recorded at 37ºC, under constant stirring, <strong>with</strong> slit widths <strong>of</strong> excitation and<br />

emission <strong>of</strong> 10 nm.<br />

Solutions<br />

All the antibiotic solutions and proteoliposomes suspensions were prepared in 10 mM<br />

Hepes buffer (pH 7.4, 0.1 M NaCl).<br />

Reconstitution <strong>of</strong> OmpF in DMPC liposomes<br />

As the final purpose for the proteoliposomes was the study <strong>of</strong> the quinolones interaction<br />

<strong>with</strong> OmpF in a structural perspective, correct orientation <strong>of</strong> protein insertion was<br />

important and best achieved by direct incorporation <strong>of</strong> OmpF into preformed liposomes<br />

113


[9-12]. Having this into consideration, OmpF proteoliposomes were prepared by direct<br />

incorporation into preformed DMPC liposomes, by a well established methodology [10,<br />

13-15]. Briefly, an adequate volume (~ 2.6 ml) <strong>of</strong> DMPC liposome suspension (~ 2.0<br />

mM) in Hepes buffer, prepared according usual procedures [16], is added to OmpF<br />

(0.45 mg) in a Hepes buffer solution <strong>with</strong> 0.4% <strong>of</strong> oPOE. The lipid/protein mole ratio is<br />

always near 1000 and the total volume <strong>of</strong> the mixture assures a final concentration <strong>of</strong><br />

oPOE lower than the value <strong>of</strong> its CMC (0.23%). After an efficient homogenization by<br />

gentle stirring, the mixture is incubated 15 min at room temperature followed by 1 h on<br />

ice. The detergent is then adsorbed onto SM2 Bio-Beads ® from Bio-Rad (Hercules, CA)<br />

at a concentration <strong>of</strong> 0.2 g <strong>of</strong> Bio-Beads/ml, by gently shaking <strong>of</strong> the suspension during<br />

a period <strong>of</strong> 3 h. After this time, a second portion <strong>of</strong> the same amount <strong>of</strong> Bio-Beads is<br />

added and the suspension is again shaken for another 3 h. In the end <strong>of</strong> this period,<br />

proteoliposomes are gently removed by decanting the Bio-Beads. The orientation <strong>of</strong><br />

OmpF by this reconstitution procedure is considered similar to that observed in the<br />

bacterial <strong>membrane</strong>s, based in experimental and molecular dynamics <strong>studies</strong> [17,18].<br />

As the study to be performed is based on the fluorescence <strong>of</strong> OmpF, the presence <strong>of</strong><br />

protein not inserted in the <strong>membrane</strong> would lead to errors in the final results. The<br />

suspension <strong>of</strong> proteoliposomes was therefore submitted to an ultracentrifugation (80000<br />

g, 4 ºC, 2 h) in order to remove any traces <strong>of</strong> protein not incorporated. After this<br />

procedure the supernatant was rejected and the pellet suspended in Hepes buffer.<br />

Protein was quantified in the liposomes and in the supernatant. The percentage <strong>of</strong><br />

incorporation by this methodology was always higher than 76%. In the final step <strong>of</strong> the<br />

procedure the proteoliposomes are sequentially extruded through 200 and 100 nm<br />

polycarbonate <strong>membrane</strong>s from Nucleopore (Kent, WA).<br />

Quenching <strong>of</strong> OmpF fluorescence by CP<br />

The fluorescence quenching <strong>studies</strong> were achieved by successive additions <strong>of</strong> a constant<br />

volume (10 µl) <strong>of</strong> CP solution (~ 296 µM) to the cuvette (final concentration range: 0-<br />

38µM) containing a constant amount <strong>of</strong> OmpF (~ 0.45 µM) incorporated in the<br />

liposomes.<br />

Fluorescence spectra were measured <strong>with</strong> an excitation wavelength <strong>of</strong> 290 nm. Inner<br />

filter effects and dilution <strong>of</strong> the solution were taken into account.<br />

114


BINDING OF A QUINOLONE ANTIBIOTIC TO BACTERIAL<br />

PORIN OmpF<br />

Theoretical Modelling<br />

CP quenches OmpF fluorescence through a Förster resonance energy transfer (FRET)<br />

mechanism. FRET efficiencies (E) are calculated from the extent <strong>of</strong> fluorescence<br />

emission quenching <strong>of</strong> the donor induced by the presence <strong>of</strong> acceptors,<br />

IDA<br />

E = 1− = 1−<br />

I<br />

D<br />

∞<br />

∫<br />

DA<br />

0<br />

∞<br />

0<br />

i<br />

∫<br />

i<br />

D<br />

() t dt<br />

() t dt<br />

where I DA and I D are the steady-state fluorescence intensities <strong>of</strong> the donor in the<br />

presence and absence <strong>of</strong> acceptors respectively. i DA (t) and i D (t) are the donor<br />

fluorescence decays in the presence and absence <strong>of</strong> acceptors respectively. Although<br />

there is a contribution <strong>of</strong> static quenching to FRET in OmpF-CP complexes, Eq.(1) still<br />

holds since this is taken into account as described below (Eqs. 8-12).<br />

As seen in Fig.3, CP absorbance overlaps <strong>with</strong> the OmpF’s fluorescence emission<br />

spectrum (from both tryptophans 214 and 61), and the overlap integral (J) is calculated<br />

as:<br />

(1)<br />

∫<br />

( )<br />

4<br />

J = f λ ⋅ε(λ) ⋅λ ⋅dλ<br />

(2)<br />

where f(λ) is the normalized emission spectrum <strong>of</strong> the donor and ε(λ) is the absorption<br />

spectrum <strong>of</strong> the acceptor.<br />

The Förster radius (R 0 ) is given by:<br />

R<br />

2 −4<br />

1/ 6<br />

0<br />

= 0.2108<br />

⋅ ( J ⋅ κ ⋅ n ⋅ φ<br />

D<br />

)<br />

(3)<br />

where the orientational factor is assumed in the dynamic isotropic regime (κ 2 = 2/3), the<br />

refractive index <strong>of</strong> the medium is n = 1.44, and φ D is the donor quantum yield. The<br />

numeric factor in Eq.(3) assumes nm units for the wavelength λ and Å units for R 0 .<br />

The Förster radius (R 0 ) <strong>of</strong> the Tryptophan(OmpF)-Cipr<strong>of</strong>laxacin donor-acceptor pair<br />

was determined to be 23.3 Å. This is an average value since each <strong>of</strong> the tryptophans has<br />

distinct fluorescence properties and therefore is expected to present different (but, in<br />

115


any case, very similar, probably <strong>with</strong>in 1-2 Å <strong>of</strong> each other) Förster radii for energy<br />

transfer to CP.<br />

Through an analysis <strong>of</strong> the extent <strong>of</strong> fluorescence quenching <strong>of</strong> a donor by the acceptor<br />

in a FRET experiment it is possible to calculate a binding constant (K B ) between these<br />

two molecules. However, care must be taken when working in solutions where<br />

acceptors are highly concentrated, as it is <strong>of</strong>ten the case in experiments performed on<br />

liposomes. In these cases, when both donors and acceptors partition to the lipid bilayer<br />

and interact there, the concentration <strong>of</strong> acceptors around the donors increases<br />

dramatically relatively to a situation where they are both free in solution. FRET can be<br />

efficient at distances up to 100 Å depending on the Förster radius <strong>of</strong> the donor-acceptor<br />

pair [19] and therefore, donors in a medium highly concentrated in acceptors, besides<br />

being susceptible to fluorescence quenching due to formation <strong>of</strong> specific donor-acceptor<br />

complexes, are also quenched by non-bound, nearby acceptors. Only the first<br />

contribution is relevant for the determination <strong>of</strong> the binding affinity <strong>of</strong> the donoracceptor<br />

complex, and as the two contributions are not additive (Eq. 4), formalisms that<br />

accurately account for the contribution <strong>of</strong> non-bound acceptors for the experimentally<br />

measured energy transfer efficiencies must be derived, considering the geometrical<br />

peculiarities <strong>of</strong> the studied system. The donor fluorescence decay in the presence <strong>of</strong><br />

acceptors (i DA (t)) is described as:<br />

i () t = γ ⋅i () t ⋅ρ () t ⋅ρ () t<br />

DA D bound nonbound<br />

( γ )<br />

+ 1- ⋅i ( t) ⋅ρ<br />

() t<br />

D<br />

nonbound<br />

(4)<br />

where γ is the fraction <strong>of</strong> OmpF bound to CP, ρ bound is the FRET contribution from<br />

energy transfer to acceptors bound to Ompf, and ρ nonbound is the FRET contribution<br />

arising from energy transfer to non-bound acceptors, randomly distributed in different<br />

planes (i) from the donors.<br />

Contribution to FRET <strong>of</strong> non-bound acceptors.<br />

ρ nonbound for a cylindrically symmetric donor geometry is given by [20]:<br />

116


BINDING OF A QUINOLONE ANTIBIOTIC TO BACTERIAL<br />

PORIN OmpF<br />

⎡<br />

wi<br />

⎛<br />

⎞⎤<br />

2 2<br />

⎢<br />

wi<br />

+ R<br />

⎜<br />

e 6<br />

2 1−exp( −t⋅b<br />

(5)<br />

i<br />

⋅α ) ⎟⎥<br />

ρnonbound = ∏ ⎢exp⎜−2⋅n2 ⋅π⋅wi<br />

⋅<br />

dα⎟⎥<br />

3<br />

i ⎢<br />

∫<br />

α<br />

0<br />

⎥<br />

⎢ ⎜<br />

⎟<br />

⎣ ⎝<br />

⎠⎦<br />

⎥<br />

where b i =(R 2 0 /w i ) 2 τ −1/3 D , n 2 is the acceptor density in each bilayer leaflet (number <strong>of</strong><br />

acceptors per unit area), w i is the distance between the plane <strong>of</strong> the donors and the i-th<br />

plane <strong>of</strong> acceptors and R e is the donor exclusion radius (which defines the area around<br />

the donor from where the acceptors are excluded). In protein-ligand FRET <strong>studies</strong> the<br />

exclusion radius is particularly important if the size <strong>of</strong> the protein is comparable to the<br />

Förster radius <strong>of</strong> the donor-acceptor pair.<br />

In order to calculate ρ nonbound , a model must be used to describe the positions <strong>of</strong> the<br />

donors and acceptors in the bilayer. OmpF has two clear belts <strong>of</strong> aromatic residues at<br />

the lipid/water interface separated by 25 Å [21]. Both Trp’s (61 and 214) are located in<br />

the same aromatic belt and are assumed to be in the same plane, at 12.5 Å from the<br />

center <strong>of</strong> the bilayer. When incorporated in liposomes, CP is located in the headgroup<br />

region <strong>of</strong> the bilayer [22] and according to Nagle and Tristan-Nagle [23] for DMPC the<br />

headgroups average position is 18 Å away from the centre <strong>of</strong> the bilayer. In the FRET<br />

simulations this is considered to be the position <strong>of</strong> the plane <strong>of</strong> acceptors and w 1 and w 2<br />

are assumed to be 5.5 Å and 30.5 Å (Fig. 4A). In our formalism we consider that there<br />

is no partition <strong>of</strong> non-bound CP in the bilayer area occupied by the OmpF trimer apart<br />

from the specifically bound population <strong>of</strong> antibiotic molecules.<br />

Trp 61 is located at the trimer interface (Fig. 4B) and the exclusion area for non-bound<br />

CP will be significant. R e (Eq. 5) <strong>of</strong> Trp 61 is assumed to be 30 Å which is the<br />

approximate diameter <strong>of</strong> an OmpF monomer in the bilayer plane [24] (Fig. 4B).<br />

On the other hand, Trp 214 is located in the periphery <strong>of</strong> the OmpF trimer and the<br />

distribution <strong>of</strong> non-bound CP around it will be cylindrically asymmetric. This difference<br />

in the distribution <strong>of</strong> acceptors sensed by the two donors (Trp 214 and Trp 61 ) can only be<br />

accounted through the use in the FRET simulations <strong>of</strong> two different ρ nonbound<br />

contributions.<br />

For Trp 61 the FRET contribution is simply given by Eq.5 and setting i = 2, w 1 = 5.5 Å,<br />

w 2 = 30.5 Å and R e = 30 Å. In the case <strong>of</strong> Trp 214 a large section <strong>of</strong> the bilayer is also<br />

occupied by the protein trimer itself and inaccessible to acceptors. However, this<br />

117


placement <strong>of</strong> the protein around the tryptophan is no longer symmetrical as for Trp 61 ,<br />

and energy transfer must be accounted differently:<br />

2 2<br />

⎡ ⎧ ⎡ wi / wi + Rp<br />

6<br />

⎪ 2 ⎪<br />

⎧ ⎡⎛ t ⎞⎛ R ⎞ ⎤<br />

0 6 ⎪<br />

⎫<br />

ρ<br />

nonbound '() t =<br />

⎢<br />

∏ exp −2π n2w ⎢<br />

i<br />

1−exp ⎢ − α ⎥ f( α)<br />

dα<br />

−<br />

⎢ ⎨ ⎨ ⎬<br />

i<br />

⎢ ∫<br />

⎜ ⎟⎜ ⎟<br />

τ<br />

2 '2 ⎢<br />

wi / w D<br />

w<br />

i ⎥<br />

⎢ ⎪<br />

i + R ⎪ ⎝ ⎠⎝ ⎠ ⎪<br />

d<br />

⎣ ⎩ ⎣ ⎩ ⎣<br />

⎦ ⎭<br />

⎡<br />

⎤ ⎤⎫⎤<br />

⎛ ⎞ ⎪⎥<br />

⎜ ⎟ ⎥<br />

⎣ ⎦ ⎦⎭⎦⎥<br />

1<br />

6<br />

t ⎛ R ⎞<br />

0 6 −3<br />

exp α f( α)<br />

α dα⎥<br />

∫ ⎢ − ⎜ ⎟ ⎥<br />

⎬<br />

τ<br />

2 2 ⎢<br />

wi / w D<br />

wi<br />

⎥ ⎥<br />

i + R ⎝ ⎠⎝ ⎠ ⎪<br />

p<br />

(6)<br />

ρ nonbound’ is therefore the FRET contribution to the decay <strong>of</strong> a donor located in the<br />

surface <strong>of</strong> a protein <strong>of</strong> radius R P due to the presence <strong>of</strong> acceptors randomly distributed in<br />

i different planes (for details on the derivation <strong>of</strong> this component and on function f(α)<br />

see Appendix).<br />

Using eqs. (1) and (4)–(6), theoretical expectations for FRET efficiencies for a random<br />

distribution <strong>of</strong> acceptors can be estimated.<br />

In the simulations and data analysis <strong>of</strong> FRET efficiencies it was considered that Trp 61<br />

and Trp 214 contributed equally to the fluorescence intensity <strong>of</strong> the protein. Also, energy<br />

migration among the Trp 61 does not affect the geometry <strong>of</strong> the problem, and energy<br />

migration between Trp 61 and Trp 214 is a very inefficient process due to the distance (30<br />

Å) between the two residues.<br />

Contribution to FRET <strong>of</strong> specifically bound acceptors.<br />

To quantify the contribution <strong>of</strong> specifically bound CP to FRET, it is first necessary to<br />

relate the binding constant <strong>of</strong> this antibiotic and OmpF (K B ) to the factor γ (fraction <strong>of</strong><br />

OmpF bound to CP) in Eq. 4:<br />

( 1 [ OmpF] KB<br />

[ CP]<br />

KB)<br />

⎡ + ⋅ + ⋅ +<br />

⎤<br />

γ = ⎢<br />

⎥ 2⋅K<br />

2<br />

B<br />

⋅ OmpF<br />

⎢<br />

2<br />

( 1+ [ OmpF] ⋅ KB + [ CP]<br />

⋅KB) −4⋅KB<br />

⋅[ OmpF] ⋅[ CP<br />

⎥<br />

⎣<br />

]<br />

⎦<br />

[ ]<br />

Total<br />

(7)<br />

where [OmpF] Total and [OmpF] are the total and unbound concentrations <strong>of</strong> OmpF, and<br />

[CP] is the concentration <strong>of</strong> unbound CP. It is assumed that only 1:1 complexes can be<br />

118


BINDING OF A QUINOLONE ANTIBIOTIC TO BACTERIAL<br />

PORIN OmpF<br />

formed between OmpF and CP. The 1:1 complex model is supported by data from a<br />

previous study [25].<br />

In the event <strong>of</strong> binding <strong>of</strong> CP to Ompf, the fluorescence <strong>of</strong> the several Trp’s present in<br />

one trimer will be quenched differently, as the distances between them and the antibiotic<br />

will be also different. Therefore, also for ρ bound , two different contributions for Trp 214<br />

and Trp 61 must be considered.<br />

ρ = ρ + ρ<br />

bound bound (Trp 214) bound (Trp61)<br />

(8)<br />

Assuming that at least one <strong>of</strong> the tryptophans <strong>of</strong> OmpF is part <strong>of</strong> a hypothetical CP<br />

binding site and subsequently completely quenched, and that no more than one<br />

antibiotic molecule can be specifically bound to a OmpF trimer at a given time, two<br />

models were devised: i) the CP binding site is close to the center <strong>of</strong> the trimer and<br />

binding <strong>of</strong> antibiotic leads to complete quenching <strong>of</strong> the three Trp 61 (distance <strong>of</strong> Trp 61 to<br />

CP


⎛<br />

II t ⎛ R ⎞<br />

0<br />

ρ () t = exp<br />

−<br />

bound(Trp 214)<br />

⎜<br />

⋅⎜ ⎟<br />

τ<br />

D<br />

d<br />

⎝ ⎝ 2 ⎠<br />

6<br />

⎞<br />

⎟<br />

⎠<br />

(12)<br />

Results and Discussion<br />

Successive additions <strong>of</strong> small volumes <strong>of</strong> a CP solution to OmpF proteoliposomes<br />

resulted as expected in a decrease <strong>of</strong> fluorescence from OmpF (Fig.6). Using Eqs 4-6 it<br />

is possible to compare the theoretical expectation for the efficiencies <strong>of</strong> energy transfer<br />

from the tryptophans <strong>of</strong> OmpF to unbound CP (randomly distributed) <strong>with</strong> the data<br />

obtained experimentally. Clearly the extent <strong>of</strong> quenching <strong>of</strong> the tryptophans cannot be<br />

exclusively explained on the basis <strong>of</strong> energy transfer to unbound CP, and a FRET<br />

contribution to specifically associated antibiotic must be considered.<br />

The binding constants that allow for better fits to the data points are different depending<br />

on the model for binding that is considered (I or II). In the case <strong>of</strong> model I, that assumes<br />

binding <strong>of</strong> CP close to the centre <strong>of</strong> the trimer, quenching is much more effective as the<br />

binding site is in the vicinity <strong>of</strong> three tryptophans (Trp 61 ), while for model II only one<br />

tryptophan is located near the bound antibiotic. This results in the recovery <strong>of</strong> larger K B<br />

values for model II than for model I. Upper and lower limits for K B (that allow<br />

respectively for better fits to the data points corresponding to the lower and higher CP<br />

concentration ranges) were determined for each model, and are shown in Table I.<br />

In a previous study, changes in CP absorption spectra were used to estimate a OmpF-CP<br />

binding constant and values <strong>of</strong> log(K B ) = 3.85 ± 0.34 or 4.17 ± 0.03 were obtained<br />

depending on the method chosen to analyze the absorption spectra changes after<br />

addition <strong>of</strong> OmpF in a micellar environment [25]. These values fall <strong>with</strong>in or close to<br />

the range <strong>of</strong> K B obtained by us from the energy transfer data assuming binding <strong>of</strong> CP to<br />

the periphery <strong>of</strong> OmpF (3.58 < K B < 4.00). This agreement <strong>with</strong> an alternative method<br />

can be seen as an indication that the antibiotic binding site is likely to be located away<br />

from the centre <strong>of</strong> the OmpF trimer where FRET would be more efficient and imply a<br />

smaller K B .<br />

120


BINDING OF A QUINOLONE ANTIBIOTIC TO BACTERIAL<br />

PORIN OmpF<br />

Conclusions<br />

In the present study, the CP fluoroquinolone is shown to associate <strong>with</strong> OmpF using<br />

FRET methodologies. The high symmetry <strong>of</strong> the OmpF trimer allowed the analysis <strong>of</strong><br />

FRET data <strong>with</strong> formalisms that accounted for the different distributions <strong>of</strong> acceptors<br />

around the two types <strong>of</strong> tryptophans present in the protein. This methodology is suitable<br />

for application in the analysis <strong>of</strong> FRET data obtained <strong>with</strong> similar protein systems and<br />

can be adapted to different geometries. The new FRET formalisms presented here for a<br />

donor in the surface <strong>of</strong> a cylinder <strong>of</strong> large radius (<strong>with</strong> radius comparable to the Förster<br />

radius) are expected to be <strong>of</strong> great assistance in the analysis <strong>of</strong> FRET data obtained <strong>with</strong><br />

<strong>membrane</strong> <strong>proteins</strong> presenting donor fluorophores in the protein-lipid interface.<br />

The equilibrium constant retrieved for the association event between OmpF and CP<br />

assuming binding <strong>of</strong> the antibiotic in the periphery <strong>of</strong> the porin is in agreement <strong>with</strong> a<br />

value determined previously by an independent method (in a micellar environment),<br />

indicating that the binding site <strong>of</strong> the antibiotic is likely to be located away from the<br />

center <strong>of</strong> OmpF. Also, the FRET methodology can be applied in cases where there are<br />

no spectral shifts, and its sensitivity is higher as compared to absorption methodologies.<br />

121


APPENDIX – Derivation <strong>of</strong> the FRET rate between a donor at the outer radius <strong>of</strong> a<br />

cylindrical trans<strong>membrane</strong> protein to a plane <strong>of</strong> acceptors distributed around the latter<br />

Fig. A1 shows the geometry relevant to this problem. Our derivation <strong>of</strong> the<br />

FRET rate law begins <strong>with</strong> the general expression for the average decay for a donor<br />

located in the centre <strong>of</strong> a finite disk <strong>with</strong> radius R d , N , taking into account all<br />

statistical arrangements <strong>of</strong> the N acceptor molecules located inside the disk (26):<br />

2 2<br />

N<br />

w + R<br />

6<br />

d ⎡ t ⎛ R<br />

0<br />

⎞ ⎤<br />

< ρ(t)<br />

><br />

N<br />

= exp( − t / τ) ⋅∏<br />

∫ exp⎢−<br />

⎜ ⎥W(R<br />

i<br />

)dR<br />

i<br />

i=<br />

1<br />

R<br />

⎟<br />

(A1)<br />

w ⎢ τ<br />

⎣ ⎝ i ⎠ ⎥⎦<br />

where W(R i )dR i is the probability <strong>of</strong> finding acceptor molecule A i in the ring <strong>of</strong> inner<br />

radius R i and outer radius R i +dR i . The acceptor distribution function is normalized, in<br />

the sense that<br />

∫<br />

w<br />

2 2<br />

w + R d<br />

W (R<br />

i<br />

)dR<br />

i<br />

= 1<br />

(A2)<br />

We now proceed as previously done by Davenport et al. (20) and Capeta et al.<br />

(27) for similar (albeit simpler) geometries. Because all acceptors have the same<br />

distribution function, W(R i )dR i = W(R j )dR j = simply W(R)dR, all integrals in Eq. (A1)<br />

are identical, and denoting them by J(t), we can write<br />

⎛ t ⎞<br />

< ρ( t ) ><br />

[ ] N<br />

N<br />

= exp<br />

⎜−<br />

⎟ ⋅ J ( t)<br />

(A3)<br />

⎝ τ<br />

D ⎠<br />

As pointed out by Davenport et al. [20], the probability <strong>of</strong> finding an acceptor at<br />

a distance between R and R + dR to the donor in question, given by W(R)dR, is equal to<br />

that <strong>of</strong> finding an acceptor in the ring between R’ and R’ + dR’ in the acceptors plane,<br />

given by W(R’)dR’. The acceptor distribution function, in terms <strong>of</strong> R’ = Rsinθ (see Fig.<br />

A1A), is given by<br />

122


[ 1+<br />

(2 / π)arcsin(<br />

R'<br />

/ 2R<br />

)]<br />

BINDING OF A QUINOLONE ANTIBIOTIC TO BACTERIAL<br />

PORIN OmpF<br />

⎧<br />

p<br />

R'<br />

⎪<br />

0 < R'<br />

< 2R<br />

'2 2<br />

p<br />

⎪<br />

Rd<br />

− Rp<br />

W ( R'<br />

) = ⎨<br />

(A4)<br />

⎪ 2R'<br />

⎪<br />

'2 2<br />

2Rp<br />

< R'<br />

< R'<br />

d<br />

⎪⎩<br />

Rd<br />

− Rp<br />

Note that, whereas for R’ > 2R p (where R p is the cylindrical protein radius)<br />

W(R’) is as expected for a uniform distribution <strong>of</strong> acceptors in a planar disk (only<br />

corrected in the denominator for the excluded area resulting from the existence <strong>of</strong> a<br />

protein molecule inside R d ), for R’ < 2R p the more complex expression denotes that only<br />

a fraction <strong>of</strong> the region between R’ and R’ + dR is available to acceptors (the shaded<br />

area in Fig. A1B). Insertion <strong>of</strong> this distribution function in the definition <strong>of</strong> J(t) (the<br />

integral in Eq. A1), together <strong>with</strong> the substitution α = cosθ = w/R, leads to<br />

⎡<br />

2 1<br />

6<br />

2w<br />

⎛ t ⎞⎛<br />

R0<br />

⎞ 6<br />

−3<br />

J ( t)<br />

= ∫ exp⎢<br />

⎜−<br />

⎟⎜<br />

⎟ α ⎥ f ( α)<br />

α dα<br />

(A5)<br />

'2 2<br />

Rd<br />

− R<br />

p + ⎢⎣<br />

⎝ τ<br />

2 ' 2 ⎠⎝<br />

⎠<br />

w / w R D<br />

w<br />

d<br />

⎥⎦<br />

⎤<br />

where<br />

⎧<br />

⎪<br />

w<br />

w<br />

⎪<br />

1<br />

< α <<br />

2 '2<br />

2 2<br />

w + Rd<br />

w + 4Rp<br />

⎪<br />

f ( α)<br />

= ⎨<br />

(A6)<br />

⎪<br />

w<br />

⎪ ⎛<br />

2<br />

1 1<br />

⎞<br />

< α <<br />

⎪<br />

⎜ w 1− α<br />

1<br />

+ arcsin ⎟<br />

2 2<br />

+<br />

π ⎜ ⎟<br />

w 4Rp<br />

2<br />

⎩ ⎝<br />

2αRp<br />

⎠<br />

We now rearrange Eq. A5 into<br />

123


2<br />

2w<br />

J ( t)<br />

=<br />

'2<br />

R − R<br />

+<br />

w /<br />

1<br />

∫<br />

d<br />

2 2<br />

w + Rp<br />

2<br />

p<br />

⎛<br />

⎜<br />

⎜<br />

⎝<br />

w /<br />

w /<br />

⎡⎛<br />

t<br />

exp⎢<br />

⎜−<br />

⎢⎣<br />

⎝ τ<br />

D<br />

2 2<br />

w + Rp<br />

−3<br />

∫<br />

α<br />

2 ' 2<br />

w + Rd<br />

⎞⎛<br />

R0<br />

⎟⎜<br />

⎠⎝<br />

w<br />

dα −<br />

6<br />

⎞<br />

⎟<br />

⎠<br />

w /<br />

w /<br />

2 2<br />

w + Rp<br />

∫<br />

2 ' 2<br />

w + Rd<br />

⎤<br />

6<br />

α ⎥ f ( α)<br />

α<br />

⎥⎦<br />

⎪⎧<br />

⎡⎛<br />

t<br />

⎨1<br />

− exp⎢<br />

⎜−<br />

⎪⎩ ⎢⎣<br />

⎝ τ<br />

−3<br />

⎞<br />

⎟<br />

dα⎟<br />

⎠<br />

D<br />

⎞⎛<br />

R0<br />

⎟⎜<br />

⎠⎝<br />

w<br />

6<br />

⎞<br />

⎟<br />

⎠<br />

⎤ ⎪⎫<br />

6<br />

α ⎥ f ( α)<br />

⎬dα +<br />

⎥⎦<br />

⎪⎭<br />

(A7)<br />

and denote the integrals on the right hand side <strong>of</strong> Eq. A7 by J 1 , J 2 (t) and J 3 (t), according<br />

to the order in which they appear in this equation. J 1 is easily calculated, leading to<br />

2<br />

2w<br />

J ( t)<br />

= 1−<br />

[ J<br />

2<br />

( t)<br />

− J<br />

3(<br />

t)<br />

]<br />

(A8)<br />

'2 2<br />

R − R<br />

d<br />

p<br />

Being the average concentration <strong>of</strong> acceptors in the bilayer given by,<br />

n<br />

2<br />

= N<br />

'2 2<br />

π(<br />

R d<br />

− R p<br />

)<br />

(A9)<br />

it follows that<br />

2<br />

2πn2w<br />

J ( t)<br />

= 1−<br />

[ J<br />

2<br />

( t)<br />

− J<br />

3(<br />

t)<br />

]<br />

(A10)<br />

N<br />

Inserting this expression for J(t) in Eq. A1, and taking the limit (N → ∞, R’ d →<br />

∞), one obtains the macroscopic decay law. The result is,<br />

2<br />

{ − 2πn<br />

w [ J ( t)<br />

− J ( )]}<br />

ρ ( t)<br />

= exp<br />

2 2 3<br />

t<br />

(A11)<br />

which is equivalent to each <strong>of</strong> the i items in the product <strong>of</strong> the right hand side <strong>of</strong> Eq. 6.<br />

Acknowledgements:<br />

Financial support for this work was provided by “Fundação para a Ciência e<br />

Tecnologia” (FCT, Lisboa) through projects POCI/SAU-FCF/56003/2004 and<br />

124


BINDING OF A QUINOLONE ANTIBIOTIC TO BACTERIAL<br />

PORIN OmpF<br />

POCI/QUIM/57114/2004. P. N. thanks FCT for a fellowship (SFRH/ BD/ 6369/ 2001).<br />

F. F. acknowledges grant SFRH/BD/14282/2003 from FCT (Portugal).<br />

P.N. thanks Luminita Damian for the help <strong>with</strong> the extraction and purification <strong>of</strong> OmpF.<br />

References<br />

[1] J. S. Chapman, N. H. Georgopapadakou, Routes <strong>of</strong> quinolone permeation in Escherichia coli,<br />

Antimicrob. Agents and Chemother. 32 (1988) 438-442.<br />

[2] E. Pestova, J. J. Millichap, G. A. Noskin, L. R. Peterson, Intracellular targets <strong>of</strong> moxifloxacin: a<br />

comparison <strong>with</strong> other fluoroquinolones, J. Antimicrob. Chemother. 45 (2000) 583-590.<br />

[3] O. A Mascaretti, Bacteria versus Antibacterial Agents: An integrated approuch. ASM Press,<br />

Washington D.C. 2003<br />

[4] J. Chevalier, M. Malléa, J.-M. Pagès, Comparative aspects <strong>of</strong> the diffusion <strong>of</strong> norfloxacin,<br />

cefepime and spermine through the F porin channel <strong>of</strong> Enterobacter cloacae, Biochem. J. 348 (2000)<br />

223-227.<br />

[5] L. J. V. Piddock, Y.–F. Jin, V. Ricci, A. E. Asuquo, Quinolone accumulation by Pseudomonas<br />

aeruginosa, Staphylococcus aureus and Escherichia coli, J. Antimicrob. Chemother. 43 (1999) 61-70.<br />

[6] K. Hirai, H. Aoyama, T. Irikura, S. Iyobe, S. Mitsuhashi, Differences in susceptibility to<br />

quinolones <strong>of</strong> outer <strong>membrane</strong> mutants <strong>of</strong> Salmonella typhimurium and Escherichia coli, Antimicrob.<br />

Agents and Chemother. 29 (1986) 535-538.<br />

[7] A. H. O’Keeffe, J. M. East, A. G. Lee, Selectivity in lipid binding to the bacterial outer<br />

<strong>membrane</strong> protein OmpF, Biophys. J. 79 (2000) 2066-2074.<br />

[8] R. M. Gravito, J.P. Rosenbusch, Isolation and crystallization <strong>of</strong> bacterial porin, Methods<br />

Enzymol.125 (1986) 309-328.<br />

[9] J.-L. Rigaud, B. Pitard, D. Levy, Reconstitution <strong>of</strong> <strong>membrane</strong> <strong>proteins</strong> into liposomes:<br />

application to energy transducing <strong>membrane</strong> <strong>proteins</strong>, Biochim. Biophys. Acta. 123 (1995) 223-246.<br />

[10] J.-L. Rigaud, M. T. Paternostre, A. Bluzat, Mechanisms <strong>of</strong> <strong>membrane</strong> protein insertion into<br />

liposomes during reconstitution procedures involving the use <strong>of</strong> detergents. 2. Incorporation <strong>of</strong> the lightdriven<br />

proton pump bacteriorhodopsin, Biochemistry. 27 (1988) 2677-2688.<br />

[11] P. Richard, J.-L. Rigaud, P. Graber, Reconstitution <strong>of</strong> CF0F1 into liposomes using a new<br />

reconstitution procedure, Eur. J. Biochem. 193 (1990) 921-930.<br />

[12] D. Levy, A. Gulik, A. Bluzat, J-L. Rigaud, Reconstitution <strong>of</strong> the sarcoplasmic reticulum<br />

Ca 2+ ATPase: mechanisms <strong>of</strong> <strong>membrane</strong> protein insertion into liposomes during reconstitution procedures<br />

involving the use <strong>of</strong> detergents, Biochim. Biophys. Acta. 1107 (1992) 283-298.<br />

[13] M. T. Paternostre, M. Roux, J.-L. Rigaud, Mechanisms <strong>of</strong> <strong>membrane</strong> protein insertion into<br />

liposomes during reconstituction procedures involving the use <strong>of</strong> detergents. Solubilization <strong>of</strong> large<br />

unilamellar liposomes (prepared by reverse-phase evaporation) by Triton X-100, octyl glucoside and<br />

sodium cholate, Biochemistry 27 (1988) 2668-2677.<br />

125


[14] L. Plançon, M. Chami, L. Letellier, Reconstitution <strong>of</strong> FhuA, an Escherichia coli outer <strong>membrane</strong><br />

protein, into liposomes, J. Biol. Chem. 272 (1997) 16868-16872.<br />

[15] J. Knol, K. Sjollema, B. Poolman, Detergent- mediated reconstitution <strong>of</strong> <strong>membrane</strong> <strong>proteins</strong>.<br />

Biochemistry 37 (1998) 16410-16415.<br />

[16] A. D. Bangham M. M. Standish, J. C. Watkins, Diffusion <strong>of</strong> univalent ions across the lamellae <strong>of</strong><br />

swollen phospholipids, J. Mol. Biol. 13 (1965) 238-252.<br />

[17] W. Im, B. Roux, Ions and counterions in a biological channel: A molecular dynamics simulation<br />

<strong>of</strong> OmpF porin from Escherichia coli in an explicit <strong>membrane</strong> <strong>with</strong> 1M KCl aqueous salt solution.<br />

Journal <strong>of</strong> Molecular biology, 319 (2002) 1177-1197.<br />

[18] E. M. Nestorovich, C. Danelon, M. Winterhalter, S. M. Bezrukov, Designed to penetrate: timeresolved<br />

interaction <strong>of</strong> single antibiotic molecules <strong>with</strong> bacterial pores. Proc. Nat. Acad. Sci. USA, 99<br />

(2002) 9789-9794.<br />

[19] J. R. Lakowicz, Principles <strong>of</strong> fluorescence spectroscopy. Chapter 13, 443-475. Springer Science,<br />

3 rd edition, New York, 2006.<br />

[20] L. Davenport, R. E. Dale, R. E. Bisby, R. B. Cundall, Transverse location <strong>of</strong> the fluorescent<br />

probe 1,6-diphenyl-1,3,5-hexatriene in model lipid bilayer <strong>membrane</strong> systems by resonance energy<br />

transfer, Biochemistry 24 (1985) 4097-4108.<br />

[21] D. W. Martin, Active unit <strong>of</strong> solubilized sarcoplasmic reticulum calcium adenosinetriphophatase:<br />

an active enzyme centrifugation analysis, Biochemistry 22 (1983) 2276 – 2282.<br />

[22] J. L. Vazquez, M. T. Montero, S. Merino, O. Domenech, M. Berlanga, M. Vinas, J. Hernandez-<br />

Borrel, Location and Nature <strong>of</strong> the Surface Membrane Binding Site <strong>of</strong> Cipr<strong>of</strong>loxacin: A Fluorescence<br />

Study, Langmuir. 17 (2001) 1009-1014.<br />

[23] J. F. Nagle, S. Tristam-Nagle, Structure <strong>of</strong> lipid bilayers, Biochim. Biophys. Acta 1469 (2000)<br />

159-195.<br />

[24] S. W. Cowan, R. M. Garavito, J. N. Jansonius, J. A. Jenkins, R. Karlsson, N. Konig, E. F. Pai,<br />

R. A. Pauptit, P. J. Rizkallah, J. P. Rosenbusch, G. Rummel, T. Schimer, The structure <strong>of</strong> OmpF porin in<br />

a tetragonal crystal form, Structure 10 (1995) 1041-50.<br />

[25] P. Neves, E. Berkane, P. Gameiro, M. Winterhalter, B. de Castro, <strong>Interaction</strong> between<br />

quinolones antibiotics and bacterial outer <strong>membrane</strong> porin OmpF, Biophys.Chem. 113 (2005)123-128.<br />

[26] P. K. Wolber, B. S. Hudson, An analytical solution to the Förster energy transfer problem in two<br />

dimensions, Biophys. J. 28 (1979) 197–210.<br />

[27] R. C. Capeta, J. A. Poveda, L. M. S. Loura, Non-uniform <strong>membrane</strong> probe distribution in<br />

resonance energy transfer. Application to protein-lipid selectivity, J. Fluoresc. 16 (2005) 161-172.<br />

126


BINDING OF A QUINOLONE ANTIBIOTIC TO BACTERIAL<br />

PORIN OmpF<br />

Figure Legends:<br />

Figure 1: The structure <strong>of</strong> the OmpF trimer. The positions <strong>of</strong> the two Trp residues<br />

(Trp 214 and Trp 61 ) in each monomer are shown. Views <strong>of</strong> OmpF organization: top (A)<br />

(8) (Reproduced by permission <strong>of</strong> <strong>Biophysical</strong> Journal) and perpendicular to <strong>membrane</strong><br />

axis (B). Trimer interface in (B) is assigned by T. The image was draw <strong>with</strong> PYMOL<br />

(DeLano Scientific, San Carlos, CA, http://pymol.sourceforge.net) using PDB<br />

coordinates 1OMF1 (22).<br />

Figure 2: Chemical structure <strong>of</strong> cipr<strong>of</strong>loxacin.<br />

Figure 3: Normalized fluorescence emission spectra <strong>of</strong> OmpF (---) and absorption<br />

spectra <strong>of</strong> cipr<strong>of</strong>loxacin (▬).<br />

Figure 4: A: Positions <strong>of</strong> donors (Trp 61 and Trp 214 ) and acceptors (cipr<strong>of</strong>loxacin) in the<br />

DMPC bilayer. In the simulations, the OmpF monomer is approximated to a cylinder <strong>of</strong><br />

radius 30 Å (see text). B: Representation <strong>of</strong> the OmpF trimer as assumed in the FRET<br />

simulations. Trp 61 is located in the trimer interface whereas Trp 214 is located in the<br />

periphery <strong>of</strong> the OmpF trimer.<br />

Figure 5: Representations <strong>of</strong> the OmpF trimer <strong>with</strong> possible cipr<strong>of</strong>loxacin binding sites<br />

specified (diagonal filling) for each <strong>of</strong> chosen binding models. A: Model I) binding site<br />

for cipr<strong>of</strong>loxacin is located near the trimer interface. Quenching <strong>of</strong> Trp 61 is complete<br />

after antibiotic binding, while the three Trp 214 are at a distance <strong>of</strong> 30 Å from<br />

cipr<strong>of</strong>loxacin. B: Model II) binding site for cipr<strong>of</strong>loxacin is located in the trimer<br />

periphery near one <strong>of</strong> the Trp 214 which is completely quenched after binding. The other<br />

Trp 214 are at a distance <strong>of</strong> 52 Å from the binding site, whereas the three Trp 61 are at a<br />

distance <strong>of</strong> 30 Å.<br />

Figure 6: Extent <strong>of</strong> energy transfer (Eq.1) for the OmpF tryptophan-Cipr<strong>of</strong>loxacin,<br />

donor acceptor FRET pair. Simulation for absence <strong>of</strong> binding (random distribution <strong>of</strong><br />

127


acceptors) (Eqs. 4-6) (⋅-⋅-⋅). A: A model assuming binding <strong>of</strong> cipr<strong>of</strong>loxacin near the<br />

trimer interface (Model I) was fitted to the experimental data (Eqs. 4-12). The interval<br />

<strong>of</strong> 3.15 < log(K B ) < 3.62 was recovered for the binding constant (grey area). B: A model<br />

assuming binding <strong>of</strong> cipr<strong>of</strong>loxacin to the OmpF periphery (Model II) was fitted to the<br />

experimental data. The interval <strong>of</strong> 3.58 < log(K B ) < 4.00 was recovered for the binding<br />

constant (grey area). The simulations for the upper and lower bounds <strong>of</strong> K B are<br />

represented in both figures by a solid line (⎯) and by a dotted line (⋅⋅⋅⋅⋅), respectively.<br />

Figure A1 – Schematic side (A) and top (B) views <strong>of</strong> a lipid bilayer containing a<br />

cylindrical <strong>membrane</strong> protein labelled <strong>with</strong> a donor fluorophore (D) at its outer radius,<br />

capable <strong>of</strong> FRET to acceptors (A) located in the lipid bilayer, on a plane parallel to the<br />

bilayer plane. See text for details.<br />

128


BINDING OF A QUINOLONE ANTIBIOTIC TO BACTERIAL<br />

PORIN OmpF<br />

129


FIGURE 1A<br />

FIGURE 1B<br />

130


BINDING OF A QUINOLONE ANTIBIOTIC TO BACTERIAL<br />

PORIN OmpF<br />

FIGURE 2<br />

FIGURE 3<br />

131


FIGURE 4A<br />

FIGURE 4B<br />

132


BINDING OF A QUINOLONE ANTIBIOTIC TO BACTERIAL<br />

PORIN OmpF<br />

FIGURE 5A<br />

FIGURE 5B<br />

133


FIGURE 6A<br />

FIGURE 6B<br />

134


BINDING OF A QUINOLONE ANTIBIOTIC TO BACTERIAL<br />

PORIN OmpF<br />

FIGURE A1<br />

135


136


INTERACTION OF HELIX-0 OF THE N-BAR DOMAIN<br />

WITH LIPID BILAYERS<br />

V<br />

INTERACTION OF HELIX-0 OF<br />

THE N-BAR DOMAIN WITH<br />

LIPID MEMBRANES<br />

1. Introduction<br />

Endocytosis is a crucial phenomenon in several cellular processes, such as<br />

receptor recycling and degradation, delivery <strong>of</strong> <strong>membrane</strong>-bound and soluble cargo to<br />

intracellular organelles, and nutrient uptake. Clathrin-mediated endocytosis is likely to<br />

be the best characterized <strong>of</strong> all endocytic pathways, and is responsible for recycling <strong>of</strong><br />

<strong>proteins</strong> from the plasma <strong>membrane</strong>. It is also important in the recycling <strong>of</strong> synaptic<br />

vesicles. A complex assemble <strong>of</strong> clathrin and several other <strong>proteins</strong> drives invagination<br />

<strong>of</strong> the plasma <strong>membrane</strong> and ultimately fission (Dawson et al., 2006). Therefore,<br />

clathrin mediated endocytosis can be separated in two steps. In the first step, the<br />

<strong>membrane</strong> must bend inwards to create a vesicular structure. Following this, the vesicle<br />

is actively pinched <strong>of</strong>f from the <strong>membrane</strong>. The last step is known to be carried out by a<br />

protein called dynamin that polymerizes upon <strong>membrane</strong> binding and arranges into<br />

coiled coats around the neck <strong>of</strong> the nascent vesicles, twisting (through an unknown<br />

mechanism) in response to GTP hydrolysis, leading to fission <strong>of</strong> the vesicle (McMahon<br />

and Gallop, 2005).<br />

In recent years, <strong>studies</strong> have been unravelling the complex network <strong>of</strong> <strong>proteins</strong><br />

involved in clathrin mediated endocytosis. In the end <strong>of</strong> the last decade, De Camilli and<br />

co-workers (Takei et al, 1998; Takei et al, 1999) observed that a protein called<br />

amphiphysin, known to interact <strong>with</strong> dynamin during endocytosis, bounds to spherical<br />

liposomes and induces formation <strong>of</strong> tubules both in the presence and absence <strong>of</strong><br />

137


dynamin. Soon after, the same group observed a similar behaviour <strong>with</strong> another protein,<br />

endophilin, that was also know to bind dynamin (Farsad et al., 2001).<br />

Amphiphysin and endophilin both share a homologous N-terminal amino acid<br />

stretch which was shown to be crucial for liposome tubulation activity (Farsad et al.,<br />

2001). This domain is highly conserved in both <strong>proteins</strong> and is found in the N-terminal<br />

<strong>of</strong> other <strong>proteins</strong>. Due to its presence in the bin, amphiphysin and Rvs161/167 <strong>proteins</strong><br />

(all amphiphysin family members), it was called BAR domain (Ge and Prendergast,<br />

2000). BAR domains are responsible for the dimerization <strong>of</strong> amphiphysin in lipid<br />

<strong>membrane</strong>s and are able to induce tubulation <strong>of</strong> liposomes isolated from the rest <strong>of</strong> the<br />

protein (Farsad et al., 2001). The structure <strong>of</strong> the BAR domain <strong>of</strong> amphiphysin was<br />

solved by McMahon and co-workers (Peter et al., 2004). It is a crescent-shaped dimer,<br />

and each monomer has three kinked α-helices (Figure V-1).<br />

Figure V-1: Crystal structure <strong>of</strong> the BAR domain <strong>of</strong> amphiphysin from Drosophila (taken from<br />

Peter et al., 2004).<br />

McMahon and co-workers also observed increased amphiphysin binding<br />

efficiency to liposomes <strong>of</strong> higher curvature, and suggested that this was due to a better<br />

fitting <strong>of</strong> the concave surface <strong>of</strong> the BAR domain dimer to the liposome surface. When<br />

a short N-terminal stretch predicted to be an amphipatic helix (Helix-0 or H0) was<br />

removed, liposome binding efficiencies decreased. Tubulation, however, was still<br />

detected, albeit much less efficient. This N-terminal stretch is not present in all BAR<br />

domains, and BAR domains containing it are called N-BAR domains (Peter et al., 2004;<br />

Gallop and McMahon, 2005). The authors also suggested that the function <strong>of</strong> this N-<br />

terminal segment was to insert into the bilayer as an amphipatic helix. This insertion<br />

would lead to an asymmetry between the inner and outer leaflet causing an increase <strong>of</strong><br />

positive curvature, whereas the concave surface <strong>of</strong> the BAR domain acts as a scaffold to<br />

the <strong>membrane</strong> curvature (Gallop et al., 2006; Zimmerberg and Kozlov, 2006). In fact,<br />

the rigidity <strong>of</strong> the crescent dimer shape is crucial for liposome tubulation (Masuda et al.,<br />

138


INTERACTION OF HELIX-0 OF THE N-BAR DOMAIN<br />

WITH LIPID BILAYERS<br />

2006). The N-terminal amphipatic segments were also proposed to act as mediators <strong>of</strong><br />

oligomerization between BAR domain dimers, increasing the efficiency <strong>of</strong> the<br />

tubulation process (Richnau et al., 2004; Gallop and McMahon, 2005)<br />

A model for the function <strong>of</strong> amphiphysin and other BAR <strong>proteins</strong> involved in<br />

clathrin mediated endocytosis is represented in Figure V-2. Amphiphysin can act in<br />

clathrin mediated endocytosis by binding to the nascent vesicle on the early stages <strong>of</strong><br />

invagination, either through the BAR domain itself or through interaction <strong>with</strong> other<br />

<strong>proteins</strong>. Following this, the BAR domain can contribute to the generation <strong>of</strong> the bud<br />

neck <strong>with</strong> the required dimensions, and as the pit matures, more amphiphysin is<br />

recruited to the vesicle neck due to its curvature sensing properties. Dynamin is<br />

assembled together <strong>with</strong> amphiphysin and GTP hydrolysis leads to vesicle fission<br />

(Yoshida et al., 2004).<br />

Figure V-2: Possible mechanism <strong>of</strong> action <strong>of</strong> amphiphysin in clathrin mediated endocytosis.<br />

Coat components (clathrin) were omitted (taken from Yoshida et al., 2004).<br />

139


140


INTERACTION OF HELIX-0 OF THE N-BAR DOMAIN<br />

WITH LIPID BILAYERS<br />

2. ROLE OF HELIX-0 OF THE N-BAR DOMAIN<br />

IN MEMBRANE CURVATURE GENERATION<br />

141


142


Abstract<br />

A group <strong>of</strong> <strong>proteins</strong> <strong>with</strong> cell <strong>membrane</strong> remodeling properties are also able to change<br />

dramatically the morphology <strong>of</strong> liposomes in vitro, frequently inducing tubulation. For a<br />

number <strong>of</strong> these <strong>proteins</strong>, the mechanism by which this effect is exerted has been proposed to<br />

be the embedding <strong>of</strong> amphipathic helices into the lipid bilayer. For <strong>proteins</strong> presenting BAR<br />

domains, removal <strong>of</strong> a N-terminal amphipathic alpha-helix (H0-NBAR) results in much lower<br />

<strong>membrane</strong> tubulation efficiency, pointing to a fundamental role <strong>of</strong> this protein segment. Here,<br />

we studied the interaction <strong>of</strong> a peptide corresponding to H0-NBAR <strong>with</strong> model lipid<br />

<strong>membrane</strong>s. H0-NBAR bound avidly to anionic liposomes but partitionate weakly to<br />

zwitterionic bilayers, suggesting an essentially electrostatic interaction <strong>with</strong> the lipid bilayer.<br />

Interestingly, it is shown that after <strong>membrane</strong> incorporation, the peptide oligomerizes as an<br />

antiparallel dimer, suggesting a potential role <strong>of</strong> H0-NBAR in the mediation <strong>of</strong> BAR domain<br />

oligomerization. Through monitoring the effect <strong>of</strong> H0-NBAR on liposome shape by electron<br />

microscopy it is clear that <strong>membrane</strong> morphology is not radically changed. We conclude that<br />

H0-NBAR alone is not able to induce vesicle curvature, and its function must be related to the<br />

promotion <strong>of</strong> the scaffold effect provided by the concave surface <strong>of</strong> the BAR domain.<br />

Keywords: BAR <strong>proteins</strong>, <strong>membrane</strong> tubulation; amphipatic peptide; FRET; fluorescence<br />

2


Introduction<br />

Control <strong>of</strong> <strong>membrane</strong> remodeling is essential in clathrin-mediated endocytosis (CME)<br />

as different types and levels <strong>of</strong> curvature are required at each stage <strong>of</strong> the budding <strong>of</strong> clathrincoated<br />

vesicles [1,2]. Several <strong>of</strong> the <strong>proteins</strong> thought to play relevant roles in CME (dynamin,<br />

amphiphysin, endophilin and epsin) were recently shown to induce tubulation in protein-free<br />

spherical liposomes, indicating a potential role as mediators in the <strong>membrane</strong> remodeling<br />

observed during CME (3, 4, 5 6).<br />

The BAR (Bin, amphiphysin, Rvs) domain, found in amphiphysin, endophilins, and a<br />

wide variety <strong>of</strong> other <strong>proteins</strong> <strong>with</strong> or <strong>with</strong>out known function in CME (7), is able to bind lipid<br />

<strong>membrane</strong>s, generate tubulation (both in vivo and in vitro) and to sense bilayer curvature (5, 8).<br />

This versatile domain has a banana shape and dimerizes in <strong>membrane</strong>s, giving rise to a<br />

positively charged concave surface that binds to lipid bilayers. This concave surface is likely to<br />

be the reason why the domain presents higher affinities for high curvature liposomes in vitro.<br />

Several BAR domains also present an N-terminal sequence that forms an amphipathic<br />

helix upon <strong>membrane</strong> binding (9). This sequence is here referred to as helix 0 (H0-NBAR).<br />

BAR domains presenting this sequence are called N-BAR and are able to bind to liposomes<br />

and induce tubulation <strong>with</strong> much higher efficiency, even tough sensitivity for curvature is lost<br />

(8). After a point mutation in H0-NBAR <strong>of</strong> a conserved hydrophobic residue (F) to an acidic<br />

residue (E), lipid binding and tubulation were abolished for endophilin (5), and reduced for the<br />

corresponding mutation in amphiphysin1 (8). Conservative mutations <strong>of</strong> the same residue (F to<br />

W) had no effect (5). These results point to an important role <strong>of</strong> H0-NBAR in <strong>membrane</strong><br />

remodeling by N-BAR domains, and this role is likely to be dependent on <strong>membrane</strong><br />

embedding <strong>of</strong> H0-NBAR. The exact function <strong>of</strong> H0-NBAR in <strong>membrane</strong> tubulation is however<br />

still elusive.<br />

The H0 fragment from BRAP (breast-cancer-associated protein)/Bin2, one <strong>of</strong> the first<br />

BAR domain containing <strong>proteins</strong> to be identified (10, 11), presents great homology to other N-<br />

terminal amphipathic fragments <strong>of</strong> BAR domain-containing <strong>proteins</strong> (Fig.1). The N-BAR<br />

domain <strong>of</strong> BRAP was already shown to tubulate liposomes in an identical fashion to other N-<br />

BAR domains. The mutations performed on the N-BAR domain from BRAP had also<br />

analogous effects in lipid tubulation as the corresponding mutants <strong>of</strong> the N-BAR domain <strong>of</strong><br />

amphiphysin, corroborating that the same liposome tubulation mechanism was shared by the<br />

two <strong>proteins</strong>. Here we investigated the interaction <strong>of</strong> a peptide comprising the H0-NBAR<br />

fragment <strong>of</strong> BRAP <strong>with</strong> model lipid <strong>membrane</strong>s. We performed a thorough study <strong>of</strong> the effects<br />

<strong>of</strong> partition <strong>of</strong> the N-BAR N-terminal domain to lipid <strong>membrane</strong>s on both structure and<br />

dynamics <strong>of</strong> H0-NBAR itself and the interacting lipid <strong>membrane</strong>s. We show that the N-<br />

terminal fragment <strong>of</strong> the N-BAR domain assumes in effect an alpha-helical structure upon<br />

<strong>membrane</strong> binding and that <strong>membrane</strong> binding is dependent on the presence <strong>of</strong> anionic<br />

phospholipids but virtually insensitive to both anionic lipid structure and liposome curvature.<br />

Through FRET (Förster resonance energy transfer) it is demonstrated that H0-NBAR dimerizes<br />

after incorporation in lipid <strong>membrane</strong>s, providing a possible mechanism for generation <strong>of</strong> highorder<br />

oligomers <strong>of</strong> N-BAR domains. Monitoring the fluorescence <strong>of</strong> different <strong>membrane</strong><br />

probes, <strong>membrane</strong> insertion <strong>of</strong> H0-NBAR is show to increase the packing <strong>of</strong> lipids both in the<br />

hydrophobic and headgroup regions <strong>of</strong> the bilayer. Finally, we also find that H0-NBAR is<br />

effective in inducing liposome fusion but has no liposome tubulation activity. Our results rule<br />

out the insertion <strong>of</strong> H0-NBAR in the exposed outer <strong>membrane</strong> leaflet as the mechanism <strong>of</strong><br />

tubulation induced by N-BAR domains and point to a likely interplay between the <strong>membrane</strong><br />

binding <strong>of</strong> H0-NBAR and the scaffold provided by the concave surface <strong>of</strong> the BAR domain.<br />

3


Methods<br />

Materials<br />

Peptides H0-NBAR, H0-NBAR-EDANS(5-((2-aminoethyl)amino)naphthalene-1-<br />

sulfonic acid), H0-NBAR-FITC(fluorescein isothiocyanate), and H0-ENTH(Epsin N-terminal<br />

homology domain) were synthesized by Genemed Synthesis (San Francisco, CA). Labeling<br />

was achieved by conjugation on N-terminal end <strong>of</strong> the peptide. The purity was always higher<br />

than 95%.<br />

1-Palmitoyl-2-oleoyl-sn-phosphocholine (POPC), 1-Palmitoyl-2-oleoyl-sn-glycero-3-<br />

[Phospho-rac-(1-glycerol)] (POPG), 1-Palmitoyl-2-oleoyl-sn-glycero-3-(phospho-l-serine)<br />

(POPS), 1,2-Dioleoyl-sn-glycero-3-phosphoethanolamine-N-(7-nitro-2-1,3-benzoxadiazol-4-<br />

yl) (NBD-DOPE), 1-Oleoyl-2-[12-[(7-nitro-2-1,3-benzoxadiazol-4-yl)amino]dodecanoyl]-snglycero-phosphocholine<br />

(PC-NBD), 1-Oleoyl-2-[12-[(7-nitro-2-1,3-benzoxadiazol-4-<br />

yl)amino]dodecanoyl]-sn-glycero-3-[phospho-rac-(1-glycerol)] (PG-NBD), 1-Oleoyl-2-[12-<br />

[(7-nitro-2-1,3-benzoxadiazol-4-yl)amino]dodecanoyl]-sn-glycero-phosphoserine (PS-NBD),<br />

1-Oleoyl-2-[12-[(7-nitro-2-1,3-benzoxadiazol-4-yl)amino]dodecanoyl]-sn-glycero-phosphate<br />

(PA-NBD), 1,2-Dioleoyl-sn-glycero-3-phosphoethanolamine-N-(Lissamine Rhodamine B<br />

Sulfonyl) (Rho-DOPE) were from Avanti Polar-Lipids (Alabaster, AL).<br />

1,6-Diphenyl-1,3,5-hexatriene (DPH) and 5,6-carboxyfluorescein (5,6-CF) were from<br />

Molecular Probes (Eugene, OR). Brain extract from bovine brain (Folch fraction I) was from<br />

Sigma-Aldrich (St. Louis, MO).<br />

Fine chemicals were obtained from Merck (Darmstadt, Germany). All materials were used<br />

<strong>with</strong>out further purification.<br />

Liposomes preparation<br />

The desired amount <strong>of</strong> phospholipids were mixed in chlor<strong>of</strong>orm and dried under a N 2(g)<br />

flow. The sample was then kept in vacuum overnight. Liposomes were prepared <strong>with</strong> buffer<br />

Hepes 20 mM, NaCl 150mM pH 7.4. The hydration step was performed <strong>with</strong> gentle addition<br />

<strong>of</strong> buffer followed by freeze-thaw cycles. Large unilamellar vesicles (LUV) were produced by<br />

extrusion through polycarbonate filters (12) in a Avestin extruder. Liposomes <strong>with</strong> 5,6-CF<br />

encapsulated were prepared by hydration <strong>with</strong> buffer Hepes 20 mM, NaCl 150mM pH 8.4 <strong>with</strong><br />

5mM 5,6-CF. After extrusion the suspension was passed through a 10ml Econo-Pac 10DG<br />

column Bio-Gel P-6DG gel <strong>with</strong> 6 kDa molecular weight exclusion) and eluted <strong>with</strong> buffer<br />

Hepes 20 mM, NaCl 150mM pH 8.4.<br />

CD spectroscopy<br />

CD spectroscopy was performed on a Jasco J-720 spectropolarimeter <strong>with</strong> a 450 W Xe<br />

lamp. Lipid suspensions were extruded using polycarbonate filters <strong>of</strong> 0.1 μm. Peptide<br />

concentration was 40 μM.<br />

Fluorescence Spectroscopy<br />

4


Steady-state fluorescence measurements were carried out <strong>with</strong> an SLM-Aminco 8100<br />

Series 2 spectr<strong>of</strong>luorimeter described in detail elsewhere (13).<br />

Fluorescence anisotropies were determined as described in (14). Time-resolved<br />

fluorescence measurements <strong>of</strong> H0-BAR-EDANS were carried out <strong>with</strong> a time-correlated<br />

single-photon timing system, which is also described elsewhere (13). For steady-state<br />

fluorescence anisotropies and time resolved fluorescence measurements <strong>with</strong> H0-NBAR-<br />

EDANS, the excitation and emission wavelengths were 340 and 460 nm, respectively. For<br />

time-resolved fluorescence measurements <strong>with</strong> H0-NBAR-FITC, the excitation and emission<br />

wavelengths were 470 and 525 nm, respectively. Analysis <strong>of</strong> fluorescence intensity and<br />

anisotropy decays were carried out as previously described (15,16). All measurements were<br />

performed at room temperature.<br />

Transmission electron microscopy (TEM)<br />

TEM was carried out using a Jeol (Tokyo, Japan) Electron Microscope 100SX,<br />

operated at 60kV. Samples were placed over copper grids covered <strong>with</strong> carbon <strong>membrane</strong>.<br />

Negative staining was performed <strong>with</strong> uranyl acetate 1% in water. Total lipid concentration <strong>of</strong><br />

the mixtures was varied (ranging from 0.05 mM to 0.5 mM).<br />

Results<br />

The N-terminal segment <strong>of</strong> the N-BAR domain is 100% alpha-helical in anionic<br />

liposomes.<br />

CD measurements were performed on the unlabeled H0-NBAR peptide while in the<br />

absence <strong>of</strong> liposomes, in the presence <strong>of</strong> zwitterionic liposomes (POPC) and anionic liposomes<br />

(POPG) (Fig.2). CD measurements <strong>of</strong> the labeled H0-NBAR <strong>peptides</strong> (both H0-NBAR-<br />

EDANS and H0-NBAR-FITC) produced the same results, confirming that labeling did not<br />

change the structure <strong>of</strong> H0-NBAR (results not shown). Deconvolution <strong>of</strong> the obtained spectrum<br />

<strong>with</strong> the CDNN CD Spectra Deconvolution v. 2.1 s<strong>of</strong>tware revealed a predominantly<br />

unstructured peptide in the absence <strong>of</strong> lipids and in the presence <strong>of</strong> 1mM <strong>of</strong> POPC. After<br />

addition <strong>of</strong> 1mM <strong>of</strong> POPG the peptide presented more than 80 % alpha-helical structure,<br />

compatible <strong>with</strong> the expected amphipathic nature <strong>of</strong> the N-terminal segment <strong>of</strong> the N-BAR<br />

domain while interacting <strong>with</strong> lipid <strong>membrane</strong>s.<br />

<strong>Interaction</strong> <strong>of</strong> H0-NBAR <strong>with</strong> lipid bilayers is only dependent on lipid charge.<br />

The fluorescence emission spectrum from H0-NBAR-EDANS is practically unaffected<br />

upon interaction <strong>with</strong> lipid bilayers, as the wavelength <strong>of</strong> maximum emission intensity remains<br />

the same and only a small broadening in the lower wavelength range <strong>of</strong> the spectrum is visible<br />

(results not shown). EDANS emission spectrum is extremely sensitive to the environment (17)<br />

and this result is an indication that the EDANS fluorophore is located in the hydrophilic face <strong>of</strong><br />

the amphipathic helix and that it remains fully exposed to the aqueous environment. There was<br />

however a subtle difference in quantum yields (20% higher in the presence <strong>of</strong> anionic<br />

phospholipids) and a significant change in fluorescence anisotropy, from = 0.022 in buffer<br />

to > 0.06 in the presence <strong>of</strong> anionic phospholipids. This change in fluorescence anisotropy<br />

5


is the result <strong>of</strong> the immobilization <strong>of</strong> the peptide in the lipid bilayer and can be used to quantify<br />

the partition <strong>of</strong> H0-NBAR to bilayers from equation 1 (18),<br />

where W is the fluorescence anisotropy in water, L is the fluorescence anisotropy in the<br />

bilayer, φ W is the fluorescence quantum yield in water, φ L is the fluorescence quantum yield in<br />

the <strong>membrane</strong>, γ L is the molar volume <strong>of</strong> the lipid, [L] is the lipid concentration and K p is the<br />

lipid/water partition coefficient.<br />

Figure 3A shows the dependence <strong>of</strong> on the lipid concentration for different<br />

phospholipids. It is clear that partition to POPC liposomes is extremely low when compared to<br />

the partition observed for both anionic phospholipids POPS and POPG.<br />

The results <strong>of</strong> fitting Equation 1 to the data in Figure 3 are presented in Table I. Partition <strong>of</strong><br />

H0-NBAR to POPS and POPG liposomes is identical suggesting that the nature <strong>of</strong> the anionic<br />

phospholipid is irrelevant for H0-NBAR interaction <strong>with</strong> liposomes. Further evidence for the<br />

absence <strong>of</strong> a specific interaction <strong>with</strong> a class <strong>of</strong> lipids, can be obtained from a FRET assay,<br />

where lipids (e.g. PS, PA, and PG) are derivatized <strong>with</strong> an acceptor (NBD) for Förster<br />

resonance energy transfer (FRET) from EDANS, and dispersed (2% mol/mol concentration) in<br />

a POPG matrix.<br />

FRET efficiencies (E) are obtained from the extent <strong>of</strong> fluorescence emission quenching<br />

<strong>of</strong> the donor induced by the presence <strong>of</strong> acceptors,<br />

[ ]<br />

[ L]<br />

φw⋅ < r > + K ⋅γ<br />

⋅ L ⋅φ<br />

⋅< r ><br />

< r >=<br />

1+ K ⋅γ<br />

⋅<br />

w P L L L<br />

P<br />

L<br />

(1)<br />

IDA<br />

E = 1− = 1−<br />

I<br />

D<br />

∞<br />

∫<br />

DA<br />

0<br />

∞<br />

0<br />

i<br />

∫<br />

i<br />

D<br />

() t<br />

() t<br />

(2)<br />

where I DA and I D are the the steady-state fluorescence intensities <strong>of</strong> the donor in the presence<br />

and absence <strong>of</strong> acceptors respectively, and i DA (t) and i D (t) are the donor fluorescence decays in<br />

the presence and absence <strong>of</strong> acceptors respectively.<br />

In case there was a specific interaction between H0-NBAR-EDANS and the NBDlabeled<br />

phospholipid, the peptide would be incorporated in the vicinity <strong>of</strong> that lipid and FRET<br />

would increase. Since there is also liposome fusion induced by H0-NBAR-EDANS (see later),<br />

the available two dimensional FRET formalisms may not be applicable for retrieving<br />

quantitative information about H0-NBAR-EDANS selectivity for a specific phospholipid (19),<br />

and a more qualitative approach was used. The results are shown in the inset <strong>of</strong> Fig. 3A. It is<br />

clear that no significant difference exists between E obtained <strong>with</strong> any <strong>of</strong> the anionic<br />

phospholipids. The small differences in FRET efficiencies are a result <strong>of</strong> the lack <strong>of</strong> selectivity<br />

<strong>of</strong> H0-NBAR-EDANS for specific phospholipids.<br />

Partition <strong>of</strong> H0-NBAR-EDANS to lipid <strong>membrane</strong>s is also insensitive to the<br />

dimensions <strong>of</strong> the liposomes and hence insensitive to the degree <strong>of</strong> curvature <strong>of</strong> the bilayer in<br />

this range <strong>of</strong> liposome size (Fig. 3B). Even though the values <strong>of</strong> K p recovered for larger<br />

liposomes are slightly higher (as seen in Table I), the differences are <strong>with</strong>in the errors <strong>of</strong> the<br />

fits and thus non significant.<br />

Residual partition to POPC liposomes was detected. After long incubation times<br />

(overnight), an increase <strong>of</strong> fusion <strong>of</strong> POPC liposomes loaded <strong>with</strong> H0-NBAR was evident<br />

relative to blank POPC vesicles. This was clear from a FRET experiment carried out <strong>with</strong> a<br />

mixture <strong>of</strong> liposomes loaded <strong>with</strong> 2% NBD-DOPE (donor) and liposomes loaded <strong>with</strong> 2%<br />

6


Rho-DOPE (acceptor). The increase <strong>of</strong> FRET efficiencies after the mixture reports fusion <strong>of</strong><br />

the liposomes. H0-NBAR stimulated fusion for both POPG (considerably) and POPC<br />

liposomes (slightly) (Fig. 4) and was more effective in doing so than a control peptide (H0-<br />

ENTH), corresponding to the helix 0 <strong>of</strong> the epsin N-terminal homology domain that is also<br />

related to <strong>membrane</strong> curvature induction (albeit in the case <strong>of</strong> epsin the presence <strong>of</strong><br />

phosphatidylinositol-(4,5)-bisphosphate PIP(4,5) 2 is required for <strong>membrane</strong> remodeling) [20].<br />

It was checked that under these conditions H0-ENTH is completely bound to liposomes<br />

(results not shown).<br />

H0-NBAR is an antiparallel dimer in the <strong>membrane</strong><br />

The function <strong>of</strong> H0-NBAR in liposome tubulation mediated by the BAR domain was<br />

recently proposed to be related to BAR domain oligomerization (7).<br />

With the intent to verify if H0-NBAR oligomerizes in the <strong>membrane</strong>, FRET<br />

measurements were performed using H0-NBAR-EDANS as a donor and H0-NBAR-FITC as<br />

an acceptor (emission and absorption spectra in Supplementary Data).<br />

For this system, the high Förster radius (R 0 ) <strong>of</strong> the EDANS-FITC donor-acceptor pair<br />

(R 0 (EDANS-FITC) = 40 Å) entails a small contribution <strong>of</strong> energy transfer between nonoligomerized<br />

H0-NBAR and this was taken into account in our analysis (see FRET simulation<br />

for the monomeric hypothesis in Fig.5A). The donor fluorescence decay in the presence <strong>of</strong><br />

acceptors is described as,<br />

n−1<br />

nk -<br />

i () t =<br />

⎡<br />

f ( k)ρ () t<br />

⎤<br />

()ρ<br />

DA<br />

⎢∑<br />

⋅ ⋅i t ⋅<br />

Dq nonbound<br />

() t +<br />

bound D<br />

⎣<br />

⎥<br />

k=<br />

1<br />

⎦<br />

n−1<br />

⎛<br />

1 f ( k) ⎞<br />

⎜ −∑<br />

i ( t) ρ t<br />

Dq ⎟⋅ ⋅<br />

D nonbound<br />

⎝ k=<br />

1 ⎠<br />

()<br />

[3]<br />

where ρ bound is the FRET contribution from energy transfer to each acceptor in the oligomer<br />

containing the donor, ρ nonbound is the FRET contribution arising from energy transfer to<br />

acceptors that are not integrated in the same oligomer as the donor, and f Dq (k) is the fraction <strong>of</strong><br />

donors bound to k acceptors,<br />

Here n is the number <strong>of</strong> units in the oligomer, k counts the number <strong>of</strong> donors, P D is the<br />

mole ratio <strong>of</strong> donors and P A is the mole ratio <strong>of</strong> acceptors.<br />

Through measurement <strong>of</strong> the dependence <strong>of</strong> FRET efficiencies on the acceptor/donor<br />

ratio it is possible to conclude about the presence <strong>of</strong> oligomerization and to gather information<br />

on the type <strong>of</strong> oligomerization found (21). The contribution for FRET from each acceptor in<br />

the same oligomer as the donor is given by,<br />

ρ<br />

n<br />

( k ) D<br />

( ) k n k<br />

fDq k = k⋅ ⋅P<br />

P<br />

⎛<br />

6<br />

1 ⎛ R ⎞<br />

0<br />

bound<br />

= exp<br />

−<br />

⎜<br />

⋅⎜ ⎟<br />

τ0 rD-A<br />

⎝<br />

−<br />

A<br />

⎝<br />

⎞<br />

⎠ ⎟<br />

⎠<br />

[4]<br />

[5]<br />

7


where τ 0 is the lifetime <strong>of</strong> the donor in the absence <strong>of</strong> acceptors, and r D-A is the distance<br />

between the donor and the acceptor inside the oligomer.<br />

Equation 6 gives ρ nonbound for a two dimensional system (15).<br />

nonbound 2<br />

[<br />

ρ = exp −n<br />

⋅π<br />

⋅Γ(2 / 3) ⋅<br />

R<br />

2<br />

0<br />

⎛ t ⎞<br />

⋅⎜ ⎟<br />

⎝τ<br />

0 ⎠<br />

1<br />

3<br />

]<br />

[6]<br />

where n 2 is the acceptor density in each bilayer leaflet, and Γ is the gamma function.<br />

From Eqs. 2-6, simulations for FRET efficiencies are obtained (Fig. 5A). In the FRET<br />

analysis, the different oligomerization models were fitted to two series <strong>of</strong> data simultaneously.<br />

In the first series the lipid to protein ratio (L/P) equals 1000 while in the second the protein was<br />

diluted two fold (L/P = 2000). With this methodology, uncertainty resulting from erroneous<br />

quantification <strong>of</strong> the two FRET contributions (to acceptors <strong>with</strong>in the same oligomers and to<br />

unbound acceptors) is eliminated. The oligomerization model that provided the best fit to the<br />

two data series was one that considered dimerization <strong>of</strong> the peptide and a distance <strong>of</strong> 43 Å<br />

between donor and acceptor inside the dimer. This distance is in good agreement <strong>with</strong> an<br />

antiparallel dimer as the size <strong>of</strong> a rigid and fully alpha-helical H0-NBAR is expected to be<br />

around 49.5 Å (1.5 Å per residue). The sensitivity <strong>of</strong> the methodology chosen is clear from the<br />

comparison <strong>with</strong> the simulations for FRET efficiencies arising from a parallel dimer (r D-A < 20<br />

Å) and other oligomerization numbers (see Fig. 5A).<br />

The oligomerization <strong>of</strong> H0-NBAR is dependent on <strong>membrane</strong> binding, as the peptide<br />

exists as a monomer while in buffer. This was concluded from anisotropy decays <strong>of</strong> H0-<br />

NBAR-FITC (Supplementary Data). The anisotropy decays were well described by two<br />

rotational correlation times, φ 1 and φ 2 .<br />

rt ()<br />

⎡ ⎛−t<br />

⎞<br />

r ⋅ ⎢ β ⋅ exp⎜ ⎟+<br />

⎣<br />

0 1<br />

= ⎝ φ1<br />

⎠<br />

⎛−t<br />

⎞⎤<br />

β2<br />

⋅exp⎜<br />

⎟⎥<br />

⎝φ2<br />

⎠⎦<br />

[7]<br />

Here r 0 is the fundamental anisotropy, β 1 and β 2 are the component amplitudes. The<br />

methodology for analysis <strong>of</strong> time-resolved anisotropy decays is described in detail elsewhere<br />

(16).<br />

φ 1 was typically smaller than 250 ps and is expected to be related to independent and<br />

fast movement <strong>of</strong> the FITC fluorophore. The recovered φ 2 values were between 1,45 and 1,69<br />

ns (Fig. 5B) and are expected to correspond to the motion <strong>of</strong> the entire peptide. The rotational<br />

correlation time can be estimated from the following equation (strictly valid for spherical<br />

molecules),<br />

η ⋅V<br />

φ=<br />

R ⋅T<br />

[8]<br />

8


where η is the viscosity <strong>of</strong> the medium, V is the molar volume <strong>of</strong> the molecule, R is the ideal<br />

gas constant and T is the temperature. From Equation 8 and taking into consideration the<br />

hydration volume for a protein (22) the expected φ 2 <strong>of</strong> monomeric H0-NBAR is 1.54 ns (Fig.<br />

5B). Hence, H0-NBAR dynamics in solution is as expected for a monomer and is independent<br />

<strong>of</strong> concentration up to 20 μM <strong>of</strong> peptide.<br />

H0-NBAR does not translocate efficiently across the bilayer<br />

When bound to the full N-BAR domain, H0-NBAR is expected to interact <strong>with</strong> only<br />

one monolayer. The structural changes induced in the exposed monolayer and the resulting<br />

monolayer asymmetry could be responsible for <strong>membrane</strong> bending mediated by N-BAR, in a<br />

similar mechanism as the one observed for the ENTH domain <strong>of</strong> epsin after PIP(4,5) 2 binding<br />

(23). In order to study if the H0-NBAR is able to induce by itself the sort <strong>of</strong> <strong>membrane</strong> bending<br />

observed <strong>with</strong> full N-BAR domains it is necessary to determine if the peptide, after interaction<br />

<strong>with</strong> anionic liposomes, remains bound to the outer monolayer, or if it exhibits fast<br />

translocation across the bilayer. If H0-NBAR translocated efficiently across the bilayer, effects<br />

on both monolayers would limit monolayer asymmetry and consequently H0-NBAR would not<br />

behave in a comparable manner as the helix 0 in the N-terminal segment <strong>of</strong> N-BAR.<br />

To evaluate this, the following methodology was applied. The iodide ion exhibits low<br />

permeability across lipid bilayers (Fig. 6- Inset), and by measuring the degree <strong>of</strong> fluorescence<br />

quenching induced by I - on H0-NBAR-EDANS it is possible to estimate the fraction <strong>of</strong> H0-<br />

NBAR-EDANS that translocated across the bilayer. Two sets <strong>of</strong> samples were measured. In the<br />

first, 1mM POPG liposomes were incubated <strong>with</strong> 2μM <strong>of</strong> H0-NBAR-EDANS for only 3<br />

minutes, as this amount <strong>of</strong> time guarantees almost complete binding <strong>of</strong> the peptide to the<br />

<strong>membrane</strong> (results not shown). In the second set <strong>of</strong> samples, the same concentrations <strong>of</strong> POPG<br />

liposomes and H0-NBAR-EDANS were incubated for 40 min. The fluorescence intensities<br />

were measured immediately before and after addition <strong>of</strong> KI and the Stern-Volmer plots for<br />

fluorescence quenching were obtained (Fig. 6). It is clear that the difference between the<br />

degree <strong>of</strong> exposure <strong>of</strong> H0-NBAR to KI is minimal between the two data sets. Assuming that<br />

H0-NBAR-EDANS in the inside <strong>of</strong> vesicles was 100% inaccessible to KI, after 40 minutes <strong>of</strong><br />

incubation, there was only a maximum <strong>of</strong> 4 % translocated H0-NBAR.<br />

H0-NBAR increases packing <strong>of</strong> anionic phospholipids at a higher degree than a control<br />

peptide<br />

The effect <strong>of</strong> unlabeled H0-NBAR in the packing and dynamics <strong>of</strong> phospholipids was<br />

measured through the fluorescence anisotropy <strong>of</strong> two lipid <strong>membrane</strong> probes, NBD-DOPE and<br />

DPH. The first is a good probe for the headgroup region <strong>of</strong> the bilayer while the second is a<br />

well known probe <strong>of</strong> acyl chain packing. The effects <strong>of</strong> adding increasing amounts <strong>of</strong> H0-<br />

NBAR to POPG liposomes loaded <strong>with</strong> 1 % (mol/mol) <strong>of</strong> fluorescent probe are shown in<br />

Figure 7. The results obtained <strong>with</strong> H0-ENTH are also shown for comparison. ENTH is<br />

thought to be able to tubulate liposomes in the presence <strong>of</strong> PIP(4,5) 2 through insertion <strong>of</strong> H0-<br />

ENTH in the acyl-chain. In the absence <strong>of</strong> this lipid, H0-ENTH is unable to penetrate the<br />

<strong>membrane</strong> (6). It should be stressed that the induced probe anisotropy increase is not related to<br />

a decrease <strong>of</strong> its lifetime (Perrin equation, [14]), since the probe fluorescence intensity is<br />

invariant (there is no fluorescence quenching), upon peptide addition (results not shown).<br />

H0-NBAR clearly rigidifies both the acyl-chains and the headgroup region <strong>of</strong> the<br />

bilayer. H0-ENTH has only minor effects on the fluorescence anisotropy <strong>of</strong> both probes even<br />

9


at very high concentrations, confirming that H0-NBAR is particularly rigidifying for<br />

phospholipid packing.<br />

H0-NBAR is not able to tubulate lipid <strong>membrane</strong>s in the absence <strong>of</strong> the scaffold domain<br />

<strong>of</strong> N-BAR.<br />

TEM was used to monitor the effects <strong>of</strong> H0-NBAR on liposome morphology (Fig. 8).<br />

H0-NBAR is clearly not able to induce tubulation <strong>of</strong> liposomes composed <strong>of</strong> pure synthetic<br />

phospholipids (POPG) (Fig. 8) or <strong>of</strong> lipid brain extracts (results not shown). The only<br />

noticeable effect <strong>of</strong> H0-NBAR was the fusion <strong>of</strong> the liposomes, <strong>with</strong>out a significant change in<br />

their morphology or size. Different P/L ratios produced similar results.<br />

Discussion<br />

Proteins are believed to be able to induce <strong>membrane</strong> bending by means <strong>of</strong> three<br />

mechanisms, namely the scaffold, local spontaneous curvature and the bilayer-couple<br />

mechanism (24). In the scaffold mechanism, <strong>proteins</strong> present and force their intrinsic curvature<br />

to the lipid <strong>membrane</strong>. This intrinsic curvature can be the result <strong>of</strong> tertiary structure or <strong>of</strong> the<br />

surface from a protein network. In the local spontaneous curvature hypothesis, a shallow<br />

insertion in the lipid <strong>membrane</strong> <strong>of</strong> an amphipathic moiety from a protein induces local<br />

perturbation <strong>of</strong> the packing <strong>of</strong> lipid headgroups resulting in higher local curvature. Finally in<br />

the bilayer-couple mechanism, the insertion <strong>of</strong> an amphipathic helix in the lipid bilayer could<br />

result in an increase <strong>of</strong> the area <strong>of</strong> the monolayer where the protein is inserted that is<br />

compensated by an increase <strong>of</strong> bilayer curvature.<br />

The presence <strong>of</strong> H0-NBAR greatly enhances the liposome tubulation activity <strong>of</strong> BAR<br />

domains. Several theories have been recently presented to explain the relevance <strong>of</strong> H0-NBAR<br />

in this phenomenon. It is feasible that the sole function <strong>of</strong> H0-NBAR is the increase in<br />

residence time <strong>of</strong> the BAR domain in the <strong>membrane</strong> by tight binding to the hydrophobic<br />

interior <strong>of</strong> the bilayer, but it is also possible that this segment is important in the<br />

oligomerization <strong>of</strong> BAR domains in the <strong>membrane</strong>, since cross-linking experiments suggest<br />

that BAR domains exist in <strong>membrane</strong>s as high-order oligomers [5, 25].<br />

Therefore, N-BAR domains could induce remodeling <strong>of</strong> <strong>membrane</strong>s by the scaffold, the<br />

local spontaneous curvature and the bilayer-couple mechanisms or by a combination <strong>of</strong> them.<br />

The concave shape <strong>of</strong> the BAR dimer can act as a scaffold, forcing the <strong>membrane</strong> curvature to<br />

adapt to its own curvature. Insertion <strong>of</strong> the amphipathic helix could also act through either the<br />

bilayer-couple or the local spontaneous curvature mechanism.<br />

Here we showed that insertion <strong>of</strong> H0-NBAR in the lipid bilayer is unable to induce<br />

significant changes in lipid <strong>membrane</strong> morphology. Recently, some authors reported that<br />

highly amphipathic <strong>peptides</strong> induced liposomes or supported lipid bilayers to adopt tubular<br />

structures when present at very high concentrations (L/P


ound conformation <strong>of</strong> the BAR domain through strong hydrophobic interactions <strong>with</strong> the<br />

<strong>membrane</strong>. Although partition <strong>of</strong> H0-NBAR to anionic liposomes is very efficient, it is<br />

unlikely to be higher than the partition <strong>of</strong> the BAR domain itself, as the concave surface <strong>of</strong> the<br />

BAR domain dimer is already strongly positively charged. Partition <strong>of</strong> H0-NBAR to anionic<br />

bilayers disturbs both the headgroup as the acyl-chain regions <strong>of</strong> the bilayer. H0-NBAR is<br />

more disturbing to the bilayer than H0-ENTH, and this can be the result <strong>of</strong> H0-NBAR<br />

dimerization in the <strong>membrane</strong> environment, as dimerization <strong>of</strong> amphipathic <strong>peptides</strong> was<br />

previously shown to be particularly disturbing to bilayer structure (28). The dimerization <strong>of</strong><br />

H0-NBAR explains the detection <strong>of</strong> high-order oligomers <strong>of</strong> N-BAR domains (5,25), and if<br />

effective in the full BAR domain, can be the mechanism by which H0-NBAR provides a more<br />

favorable framework for tubulation mediated by N-BAR (7).<br />

A recent molecular dynamics study (29) showed binding <strong>of</strong> N-BAR domains to a lipid<br />

<strong>membrane</strong> resulting in generation <strong>of</strong> <strong>membrane</strong> curvature through the scaffold mechanism.<br />

Results suggested that the main role <strong>of</strong> the N-terminal amphipathic helix <strong>of</strong> N-BAR was to<br />

favor the orientation <strong>of</strong> the N-BAR domain that allowed direct interaction between the<br />

<strong>membrane</strong> and the protein concave face. In effect, from our result, it is clear that the <strong>membrane</strong><br />

bending activity <strong>of</strong> the N-BAR domain must be achieved through the scaffold mechanism in<br />

which the BAR domain presents its concave surface to the <strong>membrane</strong>, and forces the<br />

<strong>membrane</strong> to adopt the same curvature, while H0-NBAR only plays a promoting role in this<br />

process, either by: i) enhancing the <strong>membrane</strong> affinity <strong>of</strong> the full N-BAR domain; ii) mediating<br />

N-BAR high order oligomerization and stimulating the increase <strong>of</strong> local concentrations <strong>of</strong> the<br />

protein; iii)increasing lipid packing and facilitating curvature generation; iv) forcing the protein<br />

to present its concave face to the <strong>membrane</strong> or by a combination <strong>of</strong> these mechanisms.<br />

Acknowledgments<br />

F. F. acknowledges grant SFRH/BD/14282/2003 from FCT (Portugal). A.F. acknowledges<br />

grant SFRH/BPD/26150/2005 from FCT (Portugal) This work was funded by FCT<br />

(Portugal) under the program POCI.<br />

11


References<br />

1. Farsad, K., and P. De Camilli. 2003. Mechanisms <strong>of</strong> <strong>membrane</strong> deformation. Curr.<br />

Opin. Cell Biol. 15:372-381.<br />

2. McMahon, H. T., and J. L. Gallop. 2005. Membrane curvature and mechanisms <strong>of</strong><br />

dynamic cell <strong>membrane</strong> remodeling. Nature 438:590-596.<br />

3. Takei, K., V. Haucke, V. Slepnev, K. Farsad, M. Salazar, H. Chen, and P. De Camilli.<br />

1998. Generation <strong>of</strong> coated intermediates <strong>of</strong> clathrin-mediated endocytosis on protein-free<br />

liposomes. Cell 94:131-141.<br />

4. Takei, K., V. I. Slepnev, V. Haucke, P. De Camilli. 1999. Functional partnership<br />

between amphiphysin and dynamin in clathrin-mediated endocytosis. Nat. Cell Biol. 1:33-<br />

39.<br />

5. Farsad, K., N. Ringstad, K. Takei, S. R. Floyd, K. Rose, and P. De Camilli. 2001.<br />

Generation <strong>of</strong> high curvature <strong>membrane</strong>s mediated by direct endophilin bilayer<br />

interactions. J. Cell Biol. 155:193-200.<br />

6. Ford, M. G. J., I. G. Mills, B. J. Peter, Y. Vallis, G. J. K. Praefcke, P. R. Evans, and H.<br />

T. McMahon, 2002. Curvature <strong>of</strong> clathrin-coated pits driven by epsin. Nature 419:361-<br />

366.<br />

7. Gallop, J. L., and H. T. McMahon. 2005. BAR domains and <strong>membrane</strong> curvature:<br />

bringing your curves to the BAR. Biochem. Soc. Symp. 72:223-31.<br />

8. Peter, B. J., H. M. Kent, I.G. Mills, Y. Vallis, P. J. G. Butler, P. R. Evans, and H. T.<br />

McMahon. 2004. BAR domains as sensors <strong>of</strong> <strong>membrane</strong> curvature: the amphiphysin BAR<br />

structure. Science 303:495-499.<br />

9. Gallop, J. L., C. C. Jao, H. M. Kent, P. J. G. Butler, P. R. Evans, R. Langen, and H. T.<br />

McMahon. 2006. Mechanism <strong>of</strong> endophilin N-BAR domain-mediated <strong>membrane</strong><br />

curvature. EMBO J. 25:2898-2910.<br />

10. Ge, K., and G. C. Prendergast. 2000. Bin2, a functionally nonredundant member <strong>of</strong> the<br />

BAR adaptor gene family. Genomics 67: 210-220.<br />

11. Greiner, J., M. Ringh<strong>of</strong>fer, M. Taniguchi, T. Hauser, A. Schmitt, H. Döhner, and M.<br />

Schmitt. 2003. Characterization <strong>of</strong> several leukemia-associated antigens inducing humoral<br />

immune responses in acute and chronic myeloid leukemia. Int. J. Cancer 106:224-231.<br />

12. Mayer, L.D., M. J. Hope, and P. R. Cullis. 1986. Vesicles <strong>of</strong> variable sizes produced<br />

by a rapid extrusion procedure. Biochim. Biophys. Acta 858:161-168.<br />

13. Loura, L. M. S., A. Fedorov, and M. Prieto. 2001. Fluid-fluid <strong>membrane</strong><br />

microheterogeneity: a fluorescence resonance energy transfer study. Biophys. J. 80: 776-<br />

788.<br />

14. Lakowicz J.R. 1999. Principles <strong>of</strong> fluorescence spectroscopy. Kluwer<br />

Academic/Plenum Publishers, NY.<br />

15. Loura, L.M.S., A. Fedorov, and M. Prieto. 1996. Resonance energy transfer in a model<br />

system <strong>of</strong> <strong>membrane</strong>s: application to gel and liquid crystalline phases. Biophys. J.<br />

71:1823-1836.<br />

16. de Almeida, R. F. M. , L. S. M. Loura, M. Prieto, A. Watts, A. Fedorov, and F. J.<br />

Barrantes. 2006. Structure and dynamics <strong>of</strong> the γ-M4 trans<strong>membrane</strong> domain <strong>of</strong> the<br />

acetylcholine receptor in lipid bilayers: insights into receptor assembly and function Mol.<br />

Membr. Biol. 23: 305-315.<br />

17. Hudson, E. N., and G. Weber. 1973. Synthesis and characterization <strong>of</strong> two fluorescent<br />

sulfhydryl reagents. Biochemistry 12:4154-4161.<br />

18. Santos, N. C., M. Prieto, and M. A. R. B. Castanho. 2003. Quantifying molecular<br />

partition into model systems <strong>of</strong> bio<strong>membrane</strong>s: an emphasis on optical spectroscopic<br />

methods. Biochimica et Biophysica Acta 1612: 123-135.<br />

12


19. Fernandes, F., L. M. S. Loura, R. Koehorst, R. B. Spruijt, M. A. Hemminga, A.<br />

Fedorov, and M. Prieto. 2004. Quantification <strong>of</strong> protein-lipid selectivity using FRET:<br />

application to the M13 major coat protein. Biophys. J. 87:344-352.<br />

20. Itoh, T., K. Koshiba, T. Kigawa, A. Kikuchi, S. Yokoyama, and T. Takenawa. 2001.<br />

Role <strong>of</strong> the ENTH domain in phosphatidylinositol-4,5-bisphosphate binding and<br />

endocytosis. Science 291:1047-1051.<br />

21. Adair, B. D., and D. M. Engelman. 1994. Glycophorin A helical trans<strong>membrane</strong><br />

domains dimerize in phospholipid bilayers: a resonance energy transfer study.<br />

Biochemistry 33:5539-5544.<br />

22. Durchschlag, H., and P. Zipper. 2001. Comparative investigations <strong>of</strong> biopolymer<br />

hydration by physicochemical and modeling techniques. Biophys. Chem. 93:141-157.<br />

23. Stahelin, R. V., F. Long, B. J. Peter, D. Murray, P. De Camilli, H. T. McMahon, and<br />

W. Cho. 2003. Contrasting <strong>membrane</strong> interaction mechanisms <strong>of</strong> AP180 N-terminal<br />

homology (ANTH) and epsin N-terminal homology (ENTH) domains. J. Biol. Chem.<br />

278:28993-28999.<br />

24. Zimmerberg, J., and M. M. Kozlov. 2006. How <strong>proteins</strong> produce cellular <strong>membrane</strong><br />

curvature. Nat. Rev. Mol. Cell Biol. 7: 9-19.<br />

25. Richnau, N., A. Fransson, K. Farsad, and P. Aspenstrom. 2004. RICH-1 has a<br />

BIN/Amphiphysin/Rvsp domain responsible for binding to <strong>membrane</strong> lipids and<br />

tubulation <strong>of</strong> liposomes. Biochem. Biophys. Res. Commun. 320: 1034-1042.<br />

26. Furuya, T., T. Kiyota, S. Lee, T. Inoue, G. Sugihara, A. Logvinova, P. Goldsmith, and<br />

H. M. Ellerby. 2003. Nanotubules formed by highly hydrophobic amphiphilic alphahelical<br />

<strong>peptides</strong> and natural phospholipids. Biophys. J. 84:1950-1959.<br />

27. Domanov, Y. A., and P. K. J. Kinnunen. 2006. Antimicrobial <strong>peptides</strong> temporins B and<br />

L induce formation <strong>of</strong> tubular lipid protrusions from supported phospholipid bilayers.<br />

Biophys. J. 91:4427-4439.<br />

28. Hristova, K., C. E. Dempsey, and S. H. White. 2001. Structure, location, and lipid<br />

perturbations <strong>of</strong> melittin at the <strong>membrane</strong> interface. Biophys. J. 80:801-811.<br />

29. Blood, P. D., and G. A. Voth. 2006. Direct observation <strong>of</strong> Bin/amphiphysin/Rvs<br />

(BAR) domain-induced <strong>membrane</strong> curvature by means <strong>of</strong> molecular dynamics simulations<br />

Pro. Natl. Acad. Sci. U S A. 103:15068-15072.<br />

13


TABLE I<br />

Membrane/Water partition coefficients (K p ) for H0-NBAR-EDANS determined from<br />

fluorescence anisotropies (Figure 3)<br />

Vesicles<br />

K p<br />

POPC * (5.6 ± 2.7) ×10 1<br />

POPS (4.2 ± 1.0) ×10 4<br />

POPG (diameter 30 nm) (3.2 ± 1.1) ×10 4<br />

POPG (diameter 100 nm) (4.4 ± 0.9) ×10 4<br />

POPG (diameter 400 nm) (5.5 ± 1.5) ×10 4<br />

* This K p value is a lower bound assuming that the anisotropy <strong>of</strong> the peptide population in<br />

interaction <strong>with</strong> POPC is identical as the one obtained <strong>with</strong> anionic phospholipids ( ~ 0.06).<br />

14


Figure Legends<br />

Fig.1: N-terminal amphipatic helix <strong>of</strong> the BAR domain <strong>of</strong> BRAP. A – Alignments <strong>of</strong> N-<br />

terminal sequences (H0) <strong>of</strong> BRAP/Bin2 and several amphiphysins (Amph) show great degree <strong>of</strong><br />

homology (h-Human, d-Drosophila, r-Rat). B – Helical wheel representation <strong>of</strong> H0 from BRAP.<br />

Numbers indicate position <strong>of</strong> the aminoacid. Hydrophobic residues are show in grey.<br />

Fig. 2: The N-terminal sequence <strong>of</strong> the BAR domain is alpha-helical in anionic liposomes. CD<br />

spectrum <strong>of</strong> H0 in buffer (---), in the presence <strong>of</strong> POPC (●) and POPG liposomes (―). Lipid<br />

concentration was 1mM.<br />

Fig. 3: Partition <strong>of</strong> H0 to bilayers is only sensitive to charge. A - Increase <strong>of</strong> fluorescence<br />

emission anisotropy <strong>of</strong> H0-EDANS <strong>with</strong> lipid concentration. H0-EDANS in the presence <strong>of</strong> POPC<br />

(▲), POPG (○) and POPS (■) liposomes <strong>of</strong> 100 nm diameter. INSET: Efficiencies <strong>of</strong> energy<br />

transfer from H0-NBAR-EDANS to NBD-labeled phospholipids (PX-NBD) in POPG liposomes.<br />

Bars show the difference in FRET efficiency (E) relative to the value obtained for PG-NBD (E PG-<br />

NBD = 0.38). Concentration <strong>of</strong> PX-NBD in POPG bilayers was 2% (mol/mol). B - Increase <strong>of</strong><br />

fluorescence emission anisotropy <strong>of</strong> H0-EDANS <strong>with</strong> POPG concentration for different liposome<br />

sizes. H0-EDANS in the presence <strong>of</strong> POPG liposomes <strong>with</strong> 30 (∆), 100 (○) and 400 (■) nm radius.<br />

Fluorescence emission anisotropies were measured <strong>with</strong> excitation wavelength <strong>of</strong> 340 nm and<br />

emission wavelength <strong>of</strong> 460 nm. In both panels, the lines are the fits <strong>of</strong> Equation 1 to the data, as<br />

described in the text.<br />

Fig. 4: H0-NBAR promotes liposome fusion in both POPC and POPG. Decrease in<br />

fluorescence emission intensities <strong>of</strong> NBD-DOPE 15 min. after mixing <strong>of</strong> liposomes <strong>with</strong> 2% NBD-<br />

DOPE and liposomes <strong>with</strong> 2% Rho-DOPE in the presence and in the absence <strong>of</strong> H0-NBAR.<br />

Results obtained <strong>with</strong> H0-ENTH are shown for comparison. Fluorescence intensity in the absence<br />

<strong>of</strong> acceptor was measured and no significant donor fluorescence bleaching was detected.<br />

Fluorescence intensities were obtained <strong>with</strong> excitation and emission wavelengths set to 460 and<br />

510 nm respectively.<br />

Fig. 5: H0 forms an antiparallel dimer in POPG bilayers. A- FRET efficiencies determined<br />

from the integration <strong>of</strong> H0-EDANS fluorescence decays in the presence <strong>of</strong> increasing fraction <strong>of</strong><br />

acceptors (H0-FITC) (■) at a L/P = 1000 and 2000. Simulations for FRET between monomers (···),<br />

parallel dimers (---) or trimers (-⋅-) do not describe the data accurately. The simulation which<br />

provided a more accurate description <strong>of</strong> the data was obtained for an antiparallel dimer (⎯) in<br />

which EDANS and FITC are separated by 43 Å (the estimated size <strong>of</strong> H0-NBAR is 49.5 Å).<br />

Insensitivity to L/P ratios further supports the derived formalisms. B – Dependence <strong>of</strong> the longer<br />

rotational correlation time (φ 2 ) <strong>of</strong> the fluorescence anisotropy decay <strong>of</strong> H0-NBAR-FITC on the<br />

concentration <strong>of</strong> peptide in buffer. The dotted line is the expected φ 2 for a H0-NBAR monomer.<br />

The dynamics <strong>of</strong> H0-NBAR-FITC is virtually unaffected by the increase in concentration and the<br />

recovered longer correlation time is in agreement <strong>with</strong> the rotation <strong>of</strong> a H0 monomer (see text). All<br />

samples contained the same concentration <strong>of</strong> H0-NBAR-FITC, and different concentrations <strong>of</strong> H0-<br />

NBAR were obtained through addition <strong>of</strong> unlabelled peptide.<br />

Fig. 6: Translocation <strong>of</strong> H0-BAR in POPG liposomes is very slow. Stern-Volmer plots for<br />

fluorescence emission quenching <strong>of</strong> H0-NBAR-EDANS by iodide. H0-EDANS in buffer (○), in<br />

the presence <strong>of</strong> POPG liposomes <strong>with</strong> a 3 min. (●) and 40 min. (▲) incubation time prior to<br />

addition <strong>of</strong> iodide quencher. Permeation <strong>of</strong> iodide across the bilayer is slow (see Inset) and after 40<br />

min. the amount <strong>of</strong> H0-NBAR-EDANS that translocated through the bilayer is estimated to be a<br />

maximum <strong>of</strong> 4%. INSET: Permeation <strong>of</strong> I - across POPG liposomes as measured by fluorescence<br />

quenching <strong>of</strong> encapsulated 5,6-CF.<br />

15


Fig. 7: H0-NBAR increases phospholipid packing. A – Effect <strong>of</strong> unlabeled H0-NBAR on the<br />

fluorescence emission anisotropy <strong>of</strong> NBD-DOPE (•). NBD-DOPE fluorescence is sensitive to<br />

changes in the headgroup region <strong>of</strong> the bilayer. B - Effect <strong>of</strong> unlabelled H0-NBAR on the<br />

fluorescence emission anisotropy <strong>of</strong> DPH (•). DPH fluorescence is sensitive to changes in the acylchain<br />

region <strong>of</strong> the bilayer. For comparison, the effect on both probes <strong>of</strong> an amphipathic peptide<br />

which is not expected to insert in the bilayer (H0-ENTH) is also presented (○). Lines are mere<br />

guides to the eye.<br />

Fig. 8: H0-NBAR promotes liposome fusion but does not alter liposome morphology. A –<br />

TEM micrograph <strong>of</strong> POPG liposomes. B – TEM micrograph <strong>of</strong> POPG liposomes after incubation<br />

<strong>with</strong> H0-NBAR (L/P = 10).<br />

16


FIG. 1A<br />

FIG.1B<br />

17


FIG. 2<br />

18


FIG. 3A<br />

FIG. 3B<br />

19


FIG.4<br />

20


FIG.5A<br />

21


Fig. 5B<br />

22


FIG 6.<br />

23


FIG.7A<br />

FIG. 7B<br />

24


FIG. 8<br />

A<br />

B<br />

25


CLUSTERING OF PI(4,5)P 2 IN FLUID PC BILAYERS<br />

VI<br />

CLUSTERING OF PI(4,5)P2 IN<br />

FLUID PC BILAYERS<br />

1. Introduction<br />

Phosphatidylinositol-(4,5)-bisphosphate (PI(4,5)P ) (Figure VI-1) is the most<br />

2<br />

abundant polyphosphatidylinositide in mammalian cells. It is however only a minor<br />

component <strong>of</strong> the plasma <strong>membrane</strong>. In human erythrocytes PI(4,5)P comprises about<br />

2<br />

1% <strong>of</strong> the total lipid content <strong>of</strong> the plasma <strong>membrane</strong> (McLaughlin et al., 2002),<br />

nevertheless it is responsible for the regulation <strong>of</strong> many essential cellular functions. Its<br />

role as a precursor <strong>of</strong> two second messengers, diacylglycerol and inositol 1,4,5-<br />

trisphosphate is well established (van Rheenen et al., 2005), and it is now recognized to<br />

act as signal for channel gating and for establishing sites for vesicular trafficking,<br />

<strong>membrane</strong> remodelling and movement, and actin cytoskeletal assembly (Martin, 2001).<br />

PI(4,5)P is also expected to act merely as a protein anchor to the plasma <strong>membrane</strong> in<br />

2<br />

some cases (McLaughlin, 2002).<br />

Figure VI-1: Structure <strong>of</strong> 1,2-dioleoyl-sn-glycero-3-phosphatidylinositol-(4,5)-bisphosphate.<br />

169


The signalling functions <strong>of</strong> PI(4,5)P<br />

2<br />

are performed via interactions <strong>with</strong> signalling<br />

<strong>proteins</strong>. Several <strong>proteins</strong> <strong>with</strong> known actin regulatory properties have been shown to<br />

interact <strong>with</strong> PI(4,5)P . N-WASP, a protein responsible for the stimulation <strong>of</strong> actin<br />

2<br />

nucleation is activated by binding to PI(4,5)P<br />

2<br />

through a short polybasic domain<br />

(Papayannopoulos et al., 2005). Overall, it seems that free (non protein-bound)<br />

PI(4,5)P in the plasma <strong>membrane</strong> acts as a signal for anchoring <strong>of</strong> actin to the<br />

2<br />

<strong>membrane</strong> (McLaughlin et al., 2002). Several <strong>proteins</strong> related to the exocytosis and<br />

clathrin mediated endocytosis mechanisms have already been shown to specifically bind<br />

PI(4,5)P , most notably epsin, a protein <strong>with</strong> <strong>membrane</strong> remodelling properties that was<br />

2<br />

shown to bind lipid <strong>membrane</strong>s through interaction <strong>with</strong> PI(4,5)P (Ford et al., 2002).<br />

2<br />

The question <strong>of</strong> how can a single lipid species regulate so many different functions<br />

in the cell <strong>with</strong> an apparent spatial resolution is still open. It has been hypothesized that<br />

the distribution <strong>of</strong> PI(4,5)P in the plasma <strong>membrane</strong> is non-uniform and that pools <strong>of</strong><br />

2<br />

spatially confined PI(4,5)P must exist (Martin, 2001). Several authors detected<br />

2<br />

evidence <strong>of</strong> cholesterol dependent localization <strong>of</strong> PI(4,5)P<br />

2<br />

<strong>with</strong>in domains in the<br />

plasma <strong>membrane</strong> (Pike and Miller, 1998; Liu et al., 1998; Laux et al., 2000; Botelho et<br />

al., 2000; Kwik et al., 2003; Epand et al., 2004; Epand et al., 2005; Gokhale et al.,<br />

2005). These observations led to the conclusion that PI(4,5)P were preferentially bound<br />

2<br />

to raft domains. However, some authors still disagree <strong>with</strong> the hypothesis <strong>of</strong> PI(4,5)P<br />

2<br />

enrichment in rafts (van Rheenen et al., 2005).<br />

Spontaneous partition <strong>of</strong> PI(4,5)P to lipid rafts is highly unlikely, as this lipid<br />

2<br />

almost always presents polyunsaturated acyl-chains which are not expected to favour<br />

incorporation in the more rigid environment found in rafts (McLaughlin et al., 2002).<br />

One possible explanation for the detection <strong>of</strong> PI(4,5)P in these structures would be its<br />

2<br />

local synthesis. However, this cannot account for the levels detected, as diffusion is<br />

expected to occur at faster rates than synthesis (Ghambir et al., 2004). The most likely<br />

mechanism for PI(4,5)P concentration in lateral domains is that some <strong>proteins</strong> <strong>with</strong><br />

2<br />

favoured partition to rafts can act as buffers, binding and passively concentrating these<br />

lipids. A family <strong>of</strong> <strong>proteins</strong> (GMC <strong>proteins</strong>) have been identified that are likely to play<br />

this role. GAP43, myristoylated alanine-rich C kinase substrate (MARCKS), CAP23,<br />

170


CLUSTERING OF PI(4,5)P 2 IN FLUID PC BILAYERS<br />

are substrates <strong>of</strong> protein kinase C that bind strongly to PI(4,5)P (Laux et al., 2000;<br />

2<br />

Wang et al., 2001). Although these <strong>proteins</strong> present saturated acyl-chains, this alone is<br />

not expected to result in accumulation in rafts. It is possible that the <strong>proteins</strong> are crosslinked<br />

via actin and this would increase significantly the preference for the cholesterol<br />

enriched phase (McLaughlin et al., 2002). In model lipid <strong>membrane</strong>s, the basic domain<br />

<strong>of</strong> MARCKS responsible for binding to PI(4,5)P was able to sequester this<br />

2<br />

phospholipid into clusters through electrostatic interactions (Denisov et al., 1998; Rauch<br />

et al., 2002; Gambhir et al., 2004).<br />

Recently, Redfern and Gericke (2004,2005) suggested an alternative method for<br />

PI(4,5)P clustering. According to these authors, phosphatidylinositol lipids<br />

2<br />

spontaneously clustered at and above physiological pH, and both in the gel and fluid<br />

phases, when incorporated in zwitterionic bilayers. The mechanism proposed to achieve<br />

non-uniform distribution <strong>of</strong> PI(4,5)P in the bilayer, was the establishment <strong>of</strong> hydrogen<br />

2<br />

bonds between phosphatidylinositol headgroups, and no external agent (cholesterol or<br />

<strong>proteins</strong>) was required. According to the authors, the same type <strong>of</strong> clustering behaviour<br />

was observed for phosphatidylinositol monophosphates and polyphosphates. This latter<br />

proposal is very controversial, as the presence <strong>of</strong> large charges in the headgroups <strong>of</strong><br />

phosphatidylinositol phosphates must result in strong repulsion between the molecules<br />

(Pap et al., 1995), and consequently a clustering phenomena is expected to be unlikely.<br />

171


172


CLUSTERING OF PI(4,5)P 2 IN FLUID PC BILAYERS<br />

2. Absence <strong>of</strong> clustering <strong>of</strong><br />

phosphatidylinositol-(4,5)-bisphosphate in<br />

fluid phosphatidylcholine<br />

173


174


Absence <strong>of</strong> clustering <strong>of</strong> phosphatidylinositol-(4,5)-<br />

bisphosphate in fluid phosphatidylcholine<br />

Fábio Fernandes, 1, * Luís M. S. Loura,* ,† Alexander Fedorov,* and Manuel Prieto*<br />

Centro de Química-Física Molecular,* Instituto Superior Técnico, Lisbon, Portugal; and Centro de<br />

Química e Departamento de Química, † Universidade de Évora, Évora, Portugal<br />

Abstract Phosphatidylinositol-(4,5)-bisphosphate [PI(4,5)<br />

P 2 ] plays a key role in the modulation <strong>of</strong> actin polymerization<br />

and vesicle trafficking. These processes seem to depend<br />

on the enrichment <strong>of</strong> PI(4,5)P 2 in plasma <strong>membrane</strong><br />

domains. Here, we show that PI(4,5)P 2 does not form domains<br />

when in a fluid phosphatidylcholine matrix in the pH<br />

range <strong>of</strong> 4.8–8.4. This finding is at variance <strong>with</strong> the spontaneous<br />

segregation <strong>of</strong> PI(4,5)P 2 to domains as a mechanism<br />

for the compartmentalization <strong>of</strong> PI(4,5)P 2 in the<br />

plasma <strong>membrane</strong>. Water/bilayer partition <strong>of</strong> PI(4,5)P 2<br />

is also shown to be dependent on the protonation state <strong>of</strong><br />

the lipid.—Fernandes, F., L. M. S. Loura, A. Fedorov, and M.<br />

Prieto. Absence <strong>of</strong> clustering <strong>of</strong> phosphatidylinositol-(4,5)-<br />

bisphosphate in fluid phosphatidylcholine. J. Lipid Res.<br />

2006. 47: 1521–1525.<br />

Supplementary key words PIP2 . lipid domains . fluorescence .<br />

fluorescence resonance energy transfer<br />

Manuscript received 13 March 2006 and in revised form 18 April 2006.<br />

Published, JLR Papers in Press, April 21, 2006.<br />

DOI 10.1194/jlr.M600121-JLR200<br />

Phosphatidylinositol-(4,5)-bisphosphate [PI(4,5)P 2 ] is<br />

found mainly in the plasma <strong>membrane</strong>, where it is a critical<br />

regulator <strong>of</strong> several cellular functions. It plays a fundamental<br />

role, particularly in actin polymerization and vesicle<br />

trafficking (1, 2). These processes seem to depend on<br />

large transient and spatially localized increases <strong>of</strong> PI(4,5)P 2<br />

concentration, as this phospholipid constitutes only z1%<br />

<strong>of</strong> the lipids in the plasma <strong>membrane</strong> (3).<br />

Significant evidence indicates that the <strong>membrane</strong><br />

patches where localized enrichment in PI(4,5)P 2 is observed<br />

are cholesterol-rich rafts (4–6), but partition <strong>of</strong><br />

these phospholipids to liquid-ordered domains is difficult<br />

to explain, as the sn-2 acyl chain <strong>of</strong> PI(4,5)P 2 is mainly<br />

arachidonic acid, a polyunsaturated acyl chain that is not<br />

expected to favor partition into rafts. Local enrichment in<br />

PI(4,5)P 2 was shown to overlap <strong>with</strong> enrichment <strong>of</strong> phosphatidylinositol-4-monophosphate<br />

kinases in the same<br />

<strong>membrane</strong> patches, possibly leading to localized synthesis<br />

<strong>of</strong> the inositol (7). However, as argued previously, this effect<br />

alone cannot explain the degree <strong>of</strong> PI(4,5)P 2 compartmentalization<br />

observed, as diffusion away from the site <strong>of</strong> synthesis<br />

is faster than the synthesis itself (8). However, several<br />

<strong>proteins</strong> [myristoylated alanine-rich C-kinase substrate<br />

(MARCKS), growth associated protein 43 (GAP43), cytoskeleton-associated<br />

protein 23 (CAP23), and neuronal<br />

axonal <strong>membrane</strong> protein (NAP22)] have been shown to<br />

be responsible for the lateral sequestration <strong>of</strong> PI(4,5)P 2 to<br />

specific domains in the <strong>membrane</strong> in a process that for some<br />

cases was dependent on cholesterol (9–11). This phenomenon,<br />

together <strong>with</strong> localized synthesis, provides a<br />

plausible mechanism for PI(4,5)P 2 compartmentalization,<br />

as large concentrations <strong>of</strong> these <strong>proteins</strong> could prevent free<br />

diffusion <strong>of</strong> the phospholipid.<br />

In a recent study, it was proposed that PI(4,5)P 2 compartmentalization<br />

can be achieved simply through hydrogen<br />

bonding between PI(4,5)P 2 head groups at or slightly<br />

above physiological pH (corresponding to partial or complete<br />

deprotonation <strong>of</strong> the phosphomonoester group)<br />

(12), <strong>with</strong>out the contribution <strong>of</strong> any external agent (cholesterol<br />

or <strong>proteins</strong>). Through fluorescence resonance<br />

energy transfer (FRET) experiments in a phosphatidylcholine<br />

(PC) matrix in the fluid state, the authors claimed<br />

to have detected PI(4,5)P 2 domain formation, and similar<br />

behaviors were observed for PI(3,4)P 2 and PI(3,4,5)P 3 as<br />

well as for phosphatidylinositol monophosphates in<br />

another study (13). Because <strong>of</strong> the large biological relevance<br />

<strong>of</strong> the problem and the controversy <strong>of</strong> these conclusions,<br />

we propose in this work to investigate this issue<br />

using a detailed experimental approach. Our results show<br />

clearly that PI(4,5)P 2 does not form domains when in a PC<br />

matrix in the pH range <strong>of</strong> 4.8–8.4 and that partition <strong>of</strong><br />

PI(4,5)P 2 to PC bilayers is largely dependent on the pH.<br />

MATERIALS AND METHODS<br />

POPC, 1-palmitoyl-2-oleoylphosphatidylglycerol (POPG), 1,2-<br />

dipalmitoylphosphatidycholine (DPPC), and 7-nitro-2-1,3-ben-<br />

Abbreviations: DPH, diphenylhexatriene; DPPC, 1,2-dipalmitoylphosphatidycholine;<br />

FRET, fluorescence resonance energy transfer;<br />

NBD, 7-nitro-2-1,3-benzoxadiazol; PC, phosphatidylcholine; PI(4,5)P 2 ,<br />

phosphatidylinositol-(4,5)-bisphosphate; POPG, 1-palmitoyl-2-oleoylphosphatidylglycerol.<br />

1 To whom correspondence should be addressed.<br />

e-mail: fernandesf@ist.utl.pt<br />

Downloaded from www.jlr.org by on September 3, 2007<br />

Copyright D 2006 by the American Society for Biochemistry and Molecular Biology, Inc.<br />

This article is available online at http://www.jlr.org Journal <strong>of</strong> Lipid Research Volume 47, 2006 1521


zoxadiazol (NBD)-PC were obtained from Avanti Polar Lipids<br />

(Birmingham, AL). NBD-PI(4,5)P 2 was from Echelon Biosciences<br />

(Salt Lake City, UT). Diphenylhexatriene (DPH) was obtained<br />

from Molecular Probes (Leiden, The Netherlands). Other<br />

fine chemicals were from Merck (Darmstadt, Germany).<br />

Liposome reconstitution procedure<br />

The desired amounts <strong>of</strong> phospholipids were mixed in chlor<strong>of</strong>orm-methanol<br />

(1:2) and dried under an N 2(g) flow. The sample<br />

was then kept in a vacuum overnight. Liposomes were prepared<br />

<strong>with</strong> 20 mM HEPES, 100 mM NaCl buffer at pH 8.4 and 7.1 and<br />

<strong>with</strong> 20 mM sodium citrate, 100 mM NaCl buffer at pH 4.8. The<br />

hydration step was performed <strong>with</strong> the addition <strong>of</strong> buffer followed<br />

by freeze-thaw cycles. Anisotropy measurements were performed<br />

<strong>with</strong> large unilamellar vesicles produced by extrusion<br />

through polycarbonate filters <strong>with</strong> a pore size <strong>of</strong> 100 nm (14).<br />

to quantify the effective fraction <strong>of</strong> NBD-PI(4,5)P 2 incorporated<br />

in liposomes. NBD quantum yield decreases <strong>with</strong><br />

hydration, and clustering <strong>of</strong> NBD-PI(4,5)P 2 leads to NBD selfquenching<br />

(20). Consequently, it was not surprising that<br />

NBD-PI(4,5)P 2 in buffer (micellar state) exhibited quantum<br />

yields z100 times smaller than when incorporated in PC<br />

liposomes (lamellar state) (Fig. 1). This difference allowed<br />

us to use fluorescence intensity as a tool to estimate the extent<br />

<strong>of</strong> water/bilayer partition for NBD-PI(4,5)P 2 . From<br />

equation 3 (21), the extent <strong>of</strong> partition can be quantified<br />

into partition coefficients (K p )<br />

I 5 I w 1 K P 3 g L 3 [L] 3 I L<br />

1 1 K P 3 g L 3 [L]<br />

(Eq: 3)<br />

Fluorescence measurements<br />

Steady-state fluorescence measurements were carried out <strong>with</strong><br />

an SLM-Aminco 8100 Series 2 spectr<strong>of</strong>luorimeter described in<br />

detail elsewhere (15). Steady-state anisotropies were determined<br />

according to Lakowicz (16). NBD-PC and NBD-PI(4,5)P 2 anisotropies<br />

were recorded <strong>with</strong> excitation and emission wavelengths<br />

<strong>of</strong> 460 and 540 nm, respectively, <strong>with</strong> spectral bandwidths<br />

<strong>of</strong> 4 nm. All measurements were performed at room temperature.<br />

Fluorescence decay measurements <strong>of</strong> DPH (donor) were carried<br />

out <strong>with</strong> a time-correlated single-photon timing system, which is<br />

described elsewhere (15). All measurements were performed at<br />

room temperature. Excitation and emission wavelengths were 340<br />

and 430 nm, respectively.<br />

In the FRET experiments, the donor decays in the presence <strong>of</strong><br />

acceptors [i DA (t)] can be described as<br />

i DA (t) 5 i D (t) 3r interplanar (t) (Eq: 1)<br />

where i DA (t) and i D (t) are the fluorescence decays <strong>of</strong> the donor in<br />

the presence and absence <strong>of</strong> acceptors, respectively. r interplanar is<br />

the FRET contribution arising from energy transfer to randomly<br />

distributed acceptors in two different planes from the donors<br />

(two monolayer leaflets) (17)<br />

8<br />

l<br />

9<br />

qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi<br />

l 2 1 R ><<br />

2 e<br />

r interplanar 5 exp 24 n 2 3p3l 2 1 2 exp(2t 3 b<br />

#<br />

3 3a 6 )<br />

>=<br />

a 3 da<br />

>:<br />

0<br />

>;<br />

Downloaded from www.jlr.org by on September 3, 2007<br />

(Eq: 2)<br />

where b 5 (R 0 2 /l) 2 H D 21/3 ,R 0 is the Förster radius, n 2 is the<br />

acceptor density in each leaflet, and l is the distance between the<br />

plane <strong>of</strong> the donor and the two planes <strong>of</strong> acceptors. The Förster<br />

radius was calculated as described elsewhere (18).<br />

RESULTS<br />

NBD-PI(4,5)P 2 partition to lipid bilayers<br />

When using a labeled phospholipid as the acceptor in a<br />

FRET study, the energy transfer efficiency will depend<br />

strongly on its concentration inside these vesicles. As<br />

PI(4,5)P 2 is known to generally form micelles in the aqueous<br />

medium (19), it is essential in the type <strong>of</strong> measurements<br />

mentioned above to determine whetherallNBD-PI(4,5)P 2<br />

incorporates in the liposomes and, if that is not the case,<br />

Fig. 1. Fluorescence intensity (I F ) <strong>of</strong> phosphatidylinositol-(4,5)-<br />

bisphosphate [PI(4,5)P 2 ] as a function <strong>of</strong> POPC (A) and 1-palmitoyl-<br />

2-oleoylphosphatidylglycerol (POPG) (B) concentration. Closed<br />

circles, pH 8.4; open circles, pH 7.1; closed squares, pH 4.8. The<br />

curves are fits <strong>of</strong> equation 3 to the experimental data, allowing the<br />

determination <strong>of</strong> water/bilayer partition coefficients (K p ). 7-Nitro-<br />

2-1,3-benzoxadiazol (NBD)-PI(4,5)P 2 concentration 5 1 mM. a.u.,<br />

arbitrary units.<br />

1522 Journal <strong>of</strong> Lipid Research Volume 47, 2006


where I is the fluorescence intensity, I W and I l are the fluorescence<br />

intensities in buffer and in liposomes, respectively,<br />

g l is the lipid molar volume, and [l] is the lipid concentration.<br />

The curves obtained when the NBD-PI(4,5)P 2 fluorescence<br />

intensities were plotted versus the lipid concentrations<br />

are shown in Fig. 1. Three different pH values<br />

(8.4, 7.1, and 4.8) corresponding to three different<br />

protonation states <strong>of</strong> PI(4,5)P 2 (for the micellar state<br />

Pk a2 5 6.7 for the 49 position and 7.6 for the 59 position)<br />

were studied. The results for the different pH values were<br />

clearly different when using POPC bilayers: a larger extent<br />

<strong>of</strong> partition was observed when pH 8.4 was used (corresponding<br />

to a total charge <strong>of</strong> 25), whereas partition for<br />

pH 7.1 (24) was almost identical to that for pH 4.8 (23).<br />

For POPG, the partition was much less dependent on pH.<br />

The K p values obtained at pH 8.4, 7.1, and 4.8 for water/<br />

POPC partition were (5.38 6 0.60) 3 10 4 , (1.84 6 0.14) 3<br />

10 4 , and (2.18 6 0.18) 3 10 4 , respectively, whereas for<br />

micelle/POPG partition, these were (1.84 6 0.33) 3 10 4 ,<br />

(2.29 6 0.38) 3 10 4 , and (1.47 6 0.18) 3 10 3 , respectively.<br />

Clustering <strong>of</strong> NBD-PI(4,5)P 2 in POPC<br />

Clustering <strong>of</strong> NBD-PI(4,5)P 2 inside the POPC matrix<br />

would result in a decrease <strong>of</strong> fluorescence intensity attributable<br />

to self-quenching <strong>of</strong> NBD and in a decrease in<br />

fluorescence anisotropy attributable to energy migration<br />

(energy homotransfer) inside the clusters (20). The two<br />

parameters were studied for increasing NBD-PI(4,5)P 2<br />

concentrations inside a POPC matrix, as shown in Fig. 2.<br />

NBD-PI(4,5)P 2 concentrations were corrected for the partition<br />

coefficients determined above. As a control, the<br />

fluorescence intensity and anisotropies <strong>of</strong> NBD-PC in<br />

POPC and <strong>of</strong> NBD-PI(4,5)P 2 in DPPC were compared <strong>with</strong><br />

the data for NBD-PI(4,5)P 2 in POPC. NBD-PC in POPC<br />

is homogeneously distributed, whereas in DPPC, clustering<br />

is expected for NBD-PI(4,5)P 2 as a result <strong>of</strong> gel-fluid<br />

phase separation.<br />

It is clear that NBD-PI(4,5)P 2 and NBD-PC have identical<br />

clustering behavior in POPC. The degree <strong>of</strong> selfquenching<br />

and energy migration <strong>of</strong> both lipids is almost<br />

identical and much smaller than for NBD-PI(4,5)P 2 in<br />

DPPC at room temperature, at which clustering is observed<br />

as a result <strong>of</strong> packing restraints in the DPPC gel<br />

matrix (22).<br />

We also performed a FRET experiment, using for this<br />

purpose the fluorescence decay <strong>of</strong> the donor. The reason<br />

for a time-dependent approach instead <strong>of</strong> a steady-state<br />

approach is that the kinetics <strong>of</strong> the donor decay gives<br />

information on both the distribution (homogeneity/heterogeneity)<br />

and concentration <strong>of</strong> acceptors (15), whereas<br />

the steady-state approach is unable to distinguish between<br />

the two. Again, we chose POPC as the matrix lipid, the<br />

donor was DPH, which is known to have no preference<br />

between lipid phases (23), and the acceptor was NBD-<br />

PI(4,5)P 2 . In this system, in the case <strong>of</strong> NBD-PI(4,5)P 2<br />

clustering, two populations <strong>of</strong> DPH would be detectable,<br />

one residing in a NBD-PI(4,5)P 2 -rich site and the other<br />

in a NBD-PI(4,5)P 2 -poor area <strong>of</strong> the bilayer.<br />

We globally fitted the FRET model for a single discrete<br />

concentration <strong>of</strong> acceptors (homogeneous distribution <strong>of</strong><br />

acceptors) (equations 2, 3) to the data in Fig. 3, and for all<br />

pH values used (only results for pH 8.4 are shown), good<br />

quality fits were obtained (global Chi-square , 1.5). The<br />

acceptor concentrations were the only free parameter<br />

during these fits, and the recovered values matched closely<br />

(610%) the bilayer concentrations expected from the<br />

partition coefficients determined above, again confirming<br />

the absence <strong>of</strong> NBD-PI(4,5)P 2 clustering.<br />

DISCUSSION<br />

Compartmentalization <strong>of</strong> PI(4,5)P 2 has been reported<br />

several times, but its origin had always been attributed to<br />

Downloaded from www.jlr.org by on September 3, 2007<br />

Fig. 2. Fluorescence intensity (I F ) <strong>of</strong> NBD-PI(4,5)P 2 in<br />

POPC (closed circles), POPG (open circles), and 1,2-dipalmitoylphosphatidycholine<br />

(DPPC; closed diamonds)<br />

bilayers and <strong>of</strong> phosphatidylcholine NBD-PC in POPC<br />

(closed squares). Inset: Fluorescence anisotropy (,r.) <strong>of</strong><br />

NBD-PI(4,5)P 2 in POPC (closed circles) and DPPC (open<br />

diamonds) bilayers and <strong>of</strong> NBD-PC in POPC (closed<br />

squares). Error bars represent SD. All measurements were<br />

performed at pH 8.4. Total lipid concentration 5 0.2 mM.<br />

a.u., absorbance units.<br />

Absence <strong>of</strong> PI(4,5)P 2 clustering in fluid PC 1523


Fig. 3. Fluorescence decays <strong>of</strong> diphenylhexatriene in the absence and presence <strong>of</strong> increasing concentrations<br />

<strong>of</strong> NBD-PI(4,5)P 2 in POPC bilayers at pH 8.4 (thin lines). Global analysis <strong>of</strong> the data according to<br />

the model for homogeneous distribution <strong>of</strong> NBD-PI(4,5)P 2 (equation 3) resulted in the fitted curves (thick<br />

lines). The concentration <strong>of</strong> PI(4,5)P 2 was kept constant (5%) by the addition <strong>of</strong> unlabeled PI(4,5)P 2 . Total<br />

lipid concentration 5 0.2 mM. t, time.<br />

the interaction <strong>with</strong> the cytoskeleton and/or <strong>with</strong> specific<br />

<strong>proteins</strong> exhibiting preferences for particular phases<br />

inside the <strong>membrane</strong>. The recent proposal for a spontaneous<br />

demixing <strong>of</strong> completely deprotonated PI(4,5)P 2<br />

when in a PC matrix even at very low concentrations as a<br />

result <strong>of</strong> intermolecular hydrogen bonding (12) suggested<br />

an additional mechanism for localized enrichments <strong>of</strong> this<br />

phospholipid. In this work, we tested this hypothesis<br />

through a carefully designed approach that considered<br />

and tested the existence <strong>of</strong> possible artifacts that could go<br />

undetected in a phenomenological approach (mere observation<br />

<strong>of</strong> increases or decreases in energy transfer as a<br />

result <strong>of</strong> pH variations).<br />

Redfern and Gericke (12) showed by differential scanning<br />

calorimetry and Fourier transform infrared spectroscopy<br />

that demixing occurred for mixtures <strong>of</strong> PI(4,5)P 2<br />

and PC lipids in the gel phase. Nevertheless, in vivo,<br />

PI(4,5)P 2 contains polyunsaturated acyl chains, and a gel<br />

state is highly unlikely for this lipid. With this in mind,<br />

these authors used FRET as a tool to detect the immiscibility<br />

<strong>of</strong> PI(4,5)P 2 and PC in the fluid state at low<br />

PI(4,5)P 2 concentrations. They used POPC as the lipid<br />

matrix and a NBD-labeled PC as the FRET donor to a<br />

short-chain (C6) Dipyrrometheneboron difluoride labeled<br />

PI(4,5)P 2 . In this system, they observed variations<br />

<strong>of</strong> energy transfer efficiency as the pH was scanned. The<br />

authors interpreted the decrease <strong>of</strong> energy transfer efficiencies<br />

at high pH [.Pk a2 <strong>of</strong> PI(4,5)P 2 ] as a demixing<br />

between PI(4,5)P 2 and POPC. However, the use <strong>of</strong> a very<br />

short-chain PI(4,5)P 2 as the acceptor in this FRET experiment<br />

is likely to have resulted in a significant fraction<br />

<strong>of</strong> acceptors not being incorporated in the <strong>membrane</strong>.<br />

The lipid/water partition coefficients <strong>of</strong> this short-chainlabeled<br />

PI(4,5)P 2 is expected to be significantly smaller<br />

than those determined here for NBD-PI(4,5)P 2 , because<br />

K p varies dramatically <strong>with</strong> the number <strong>of</strong> carbons in the<br />

acyl chains <strong>of</strong> a phospholipid (24). In this case, energy<br />

transfer efficiencies would decrease not because <strong>of</strong> the<br />

formation <strong>of</strong> PI(4,5)P 2 clusters but as a result <strong>of</strong> changes<br />

in the lipid/water partition coefficient <strong>of</strong> the probe, as the<br />

PI(4,5)P 2 group becomes fully deprotonated. Redfern<br />

and Gericke (12) acknowledge that the extent <strong>of</strong> <strong>membrane</strong><br />

incorporation changed considerably for different<br />

labeled phosphatidylinositides.<br />

Analysis <strong>of</strong> the kinetics <strong>of</strong> donor decay in a FRET experiment<br />

was the method chosen by us to probe the<br />

eventual clustering <strong>of</strong> PI(4,5)P 2 in a zwitterionic bilayer,<br />

because it is capable <strong>of</strong> retrieving information on the<br />

homogeneity <strong>of</strong> the fluorescently labeled PI(4,5)P 2 independently<br />

<strong>of</strong> any quantitative knowledge <strong>of</strong> the probe’s<br />

bilayer concentration. The good agreement between the<br />

model considering a homogeneous distribution <strong>of</strong> NBD-<br />

PI(4,5)P 2 and the data is solid pro<strong>of</strong> <strong>of</strong> the absence <strong>of</strong><br />

PI(4,5)P 2 clustering at low PI(4,5)P 2 concentrations<br />

(,5%) even in a completely deprotonated state. The result<br />

<strong>of</strong> this experiment alone cannot exclude the formation<br />

<strong>of</strong> very small PI(4,5)P 2 clusters (zR 0 for this Förster<br />

pair 5 40 Å), but, together <strong>with</strong> the NBD energy migration<br />

and self-quenching <strong>studies</strong>, it is clear that PI(4,5)P 2<br />

has no lateral organization in a fluid POPC matrix. It is<br />

Downloaded from www.jlr.org by on September 3, 2007<br />

1524 Journal <strong>of</strong> Lipid Research Volume 47, 2006


certainly feasible that PI(4,5)P 2 /PC demixing exists in the<br />

gel phase, as packing restraints are severely increased (22).<br />

The difference in the lipid/water partition coefficients <strong>of</strong><br />

the NBD-PI(4,5)P 2 fully and partially deprotonated states<br />

is interesting, and the same trend is expected for nonlabeled<br />

PI(4,5)P 2 . This difference might be the result <strong>of</strong><br />

destabilization <strong>of</strong> the fully deprotonated PI(4,5)P 2 micellar<br />

structure. However, it is very likely that the labeling<br />

<strong>with</strong> NBD decreases the extent <strong>of</strong> partition to some extent,<br />

and further <strong>studies</strong> will be necessary to determine whether<br />

this phenomenon has physiological relevance.<br />

F.F. acknowledges Grant SFRH/BD/14282/2003 from Fundação<br />

para a Ciência e a Tecnologia (FCT) (Portugal). A.F. acknowledges<br />

Grant SFRH/BPD/26150/2005 from FCT (Portugal).<br />

This work was funded by FCT (Portugal) under the program<br />

Programa Operacional Ciência e Inovação (POCI).<br />

REFERENCES<br />

1. Payrastre, B., K. Missy, S. Giuriato, S. Bodin, M. Plantavid, and M-P.<br />

Gratacap. 2001. Phosphoinositides: key players in cell signalling, in<br />

time and space. Cell. Signal. 13: 377–387.<br />

2. Czech, M. P. 2003. Dynamics <strong>of</strong> phosphoinositides in <strong>membrane</strong><br />

retrieval and insertion. Annu. Rev. Physiol. 65: 791–815.<br />

3. Ferrell, J., Jr., and W. Huestis. 1984. Phosphoinositide metabolism<br />

and the morphology <strong>of</strong> human erythrocytes. J. Cell Biol. 98:<br />

1992–1998.<br />

4. Pike, L. J., and L. Casey. 1996. Localization and turnover <strong>of</strong><br />

phosphatidylinositol 4,5-bisphosphate in caveolin-enriched <strong>membrane</strong><br />

domains. J. Biol. Chem. 271: 26453–26456.<br />

5. Pike, L. J., and J. M. Miller. 1998. Cholesterol depletion delocalizes<br />

phosphatidylinositol bisphosphate and inhibits hormone-stimulated<br />

phosphatidylinositol turnover. J. Biol. Chem. 273: 22298–22304.<br />

6. Rozelle, A. L., L. M. Machesky, M. Yamamoto, M. H. E. Driessens,<br />

R. H. Insall, M. G. Roth, K. Luby-Phelps, G. Marriott, A. Hall, and<br />

H. L. Yin. 2000. Phosphatidylinositol 4,5-bisphosphate induces<br />

actin-based movement <strong>of</strong> raft-enriched vesicles through WASP-<br />

Arp2/3. Curr. Biol. 10: 311–320.<br />

7. Botelho, R. J., M. Teruel, R. Dierckman, R. Anderson, A. Wells, J. D.<br />

York, T. Meyer, and S. Grinstein. 2000. Localized biphasic changes<br />

in phosphatidylinositol-4,5-bisphosphate at sites <strong>of</strong> phagocytosis.<br />

J. Cell Biol. 151: 1353–1368.<br />

8. McLaughlin, S., J. Wang, A. Gambhir, and D. Murray. 2002. PIP2<br />

and <strong>proteins</strong>: interactions, organization, and information flow.<br />

Annu. Rev. Biophys. Biomol. Struct. 31: 151–175.<br />

9. Laux, T., K. Fukami, M. Thelen, T. Golub, D. Frey, and P. Caroni.<br />

2000. GAP43, MARCKS, and CAP23 modulate PI(4,5)P2 at<br />

plasmalemmal rafts, and regulate cell cortex actin dynamics<br />

through a common mechanism. J. Cell Biol. 149: 1455–1472.<br />

10. Rauch, M. E., C. G. Ferguson, G. D. Prestwich, and D. S. Cafiso.<br />

2002. Myristoylated alanine-rich C kinase substrate (MARCKS)<br />

sequesters spin-labeled phosphatidylinositol 4,5-bisphosphate in<br />

lipid bilayers. J. Biol. Chem. 277: 14068–14076.<br />

11. Epand, R. M., P. Vuong, C. M. Yip, S. Maekawa, and R. F. Epand.<br />

2004. Cholesterol-dependent partitioning <strong>of</strong> PtdIns(4,5)P2 into<br />

<strong>membrane</strong> domains by the N-terminal fragment <strong>of</strong> NAP-22 (neuronal<br />

axonal myristoylated <strong>membrane</strong> protein <strong>of</strong> 22 kDa). Biochem.<br />

J. 379: 527–532.<br />

12. Redfern, D. A., and A. Gericke. 2005. pH-dependent domain<br />

formation in phosphatidylinositol polyphosphate/phosphatidylcholine<br />

mixed vesicles. J. Lipid Res. 46: 504–515.<br />

13. Redfern, D. A., and A. Gericke. 2004. Domain formation in<br />

phosphatidylinositol monophosphate/phosphatidylcholine mixed<br />

vesicles. Biophys. J. 86: 2980–2992.<br />

14. Mayer, L. D., M. J. Hope, and P. R. Cullis. 1986. Vesicles <strong>of</strong> variable<br />

sizes produced by a rapid extrusion procedure. Biochim. Biophys.<br />

Acta. 858: 161–168.<br />

15. Loura, L. M. S., A. Fedorov, and M. Prieto. 2001. Fluid-fluid<br />

<strong>membrane</strong> microheterogeneity: A fluorescence resonance energy<br />

transfer study. Biophys. J. 80: 776–788.<br />

16. Lakowicz, J. R. 1999. Principles <strong>of</strong> Fluorescence Spectroscopy.<br />

Kluwer Academic/Plenum Publishers, New York.<br />

17. Davenport, L., R. E. Dale, R. H. Bisby, and R. B. Cundall. 1985.<br />

Transverse location <strong>of</strong> the fluorescent probe 1,6-diphenyl-1,3,5-<br />

hexatriene in model lipid bilayer <strong>membrane</strong> systems by resonance<br />

energy transfer. Biochemistry. 24: 4097–4108.<br />

18. Berberan-Santos, M. N., and M. Prieto. 1987. Energy transfer in<br />

spherical geometry. J. Chem. Soc. Faraday Trans. 83: 1391–1409.<br />

19. Sugiura, Y. 1981. Structure <strong>of</strong> molecular aggregates <strong>of</strong> 1-(3-snphosphatidyl)-L-myo-inositol<br />

3,4-bis(phosphate) in water. Biochim.<br />

Biophys. Acta. 641: 149–159.<br />

20. Prieto, M. J., M. Castanho, A. Coutinho, A. Ortiz, F. J. Aranda, and<br />

J. C. Gomez-Fernandez. 1994. Fluorescence study <strong>of</strong> a derivatized<br />

diacylglycerol incorporated in model <strong>membrane</strong>s. Chem. Phys.<br />

Lipids. 69: 75–85.<br />

21. Santos, N. C., M. Prieto, and M. A. R. B. Castanho. 2003.<br />

Quantifying molecular partition into model systems <strong>of</strong> bio<strong>membrane</strong>s:<br />

an emphasis on optical spectroscopic methods. Biochim.<br />

Biophys. Acta. 1612: 123–135.<br />

22. Loura, L. M. S., A. Fedorov, and M. Prieto. 2000. Membrane probe<br />

distribution heterogeneity: a resonance energy transfer study. J.<br />

Phys. Chem. B. 104: 6920–6931.<br />

23. Lentz, B. R., Y. Barenholz, and T. E. Thompson. 1976. Fluorescence<br />

depolarization <strong>studies</strong> <strong>of</strong> phase transitions and fluidity in phospholipid<br />

bilayers. II. Two-component phosphatidylcholine liposomes.<br />

Biochemistry. 15: 4529–4537.<br />

24. Nichols, J. W. 1985. Thermodynamics and kinetics <strong>of</strong> phospholipid<br />

monomer-vesicle interaction. Biochemistry. 24: 6390–6398.<br />

Downloaded from www.jlr.org by on September 3, 2007<br />

Absence <strong>of</strong> PI(4,5)P 2 clustering in fluid PC 1525


CONCLUSIONS<br />

VII<br />

CONCLUSIONS<br />

Chapter II ‐ Protein‐Protein and Protein‐Lipid <strong>Interaction</strong>s<br />

<strong>of</strong> M13 Major Coat.<br />

In this chapter an extensive study <strong>of</strong> the protein-protein and protein-lipid<br />

interactions <strong>of</strong> M13 mcp was conducted. The fluorescence self-quenching data obtained<br />

from the study <strong>with</strong> a BODIPY labelled mcp allowed us to conclude that mcp is in fact<br />

monomeric in DOPC bilayers at the conditions used for protein purification and protein<br />

reconstitution. The use <strong>of</strong> two different sites for labelling, resulted in identical results<br />

excluding any possible influence <strong>of</strong> labelling on the oligomerization efficiency <strong>of</strong> mcp.<br />

Nevertheless, as the <strong>studies</strong> <strong>with</strong> lipid bilayers presenting a different hydrophobic<br />

thickness than mcp shows, mcp can aggregate is certain conditions. In this way, it is<br />

possible that at much higher protein concentrations than the one used here (L/P < 50),<br />

some aggregation takes place, although in vivo the presence <strong>of</strong> only one orientation <strong>of</strong><br />

mcp should contribute to keep it monomeric. This could explain the increase in the<br />

levels <strong>of</strong> anionic lipids (cardiolipin and phosphatidylglycerol) in the lipid <strong>membrane</strong> <strong>of</strong><br />

E. coli during the process <strong>of</strong> infection by the bacteriophage M13. Therefore, although<br />

anionic lipids are not required to keep mcp monomeric in normal conditions, at the very<br />

high mcp concentrations found in the sites <strong>of</strong> phage assembly in the lipid <strong>membrane</strong> <strong>of</strong><br />

the host, they might contribute to stabilize the monomeric state <strong>of</strong> the protein, which is<br />

required for correct assembly.<br />

Fluorescence self-quenching <strong>of</strong> BODIPY labelled mcp is more efficient when mcp<br />

is incorporated in bilayers <strong>with</strong> a thinner (positive hydrophobic mismatch) or thicker<br />

(negative hydrophobic mismatch) hydrophobic length, than the hydrophobic domain <strong>of</strong><br />

mcp. This is likely to result from aggregation <strong>of</strong> mcp, and in this case aggregation in<br />

thicker bilayers is significantly more effective than in thinner bilayers. This discrepancy<br />

can be the result <strong>of</strong> the availability <strong>of</strong> alternative methods for protein adaptation in<br />

positive hydrophobic mismatch conditions. In this case, the protein is able to increase<br />

181


the tilt angle, and maintain a larger section <strong>of</strong> the hydrophobic domain inside the bilayer<br />

(Figure I.16). When the bilayer thickness is larger than the protein hydrophobic domain,<br />

the TM protein might find that compensation by change <strong>of</strong> tilt angle is not possible or it<br />

is insufficient, and the only route available to decrease exposure <strong>of</strong> hydrophobic<br />

aminoacid side-chains is aggregation. Recently, mcp was shown to change its tilt angle<br />

when incorporated in bilayers <strong>with</strong> different hydrophobic thickness (Koehorst et al.,<br />

2004).<br />

Results obtained for fluorescence self-quenching <strong>of</strong> BODIPY labelled mcp<br />

incorporated in binary lipid mixtures composed by lipids <strong>with</strong> different acyl-chain<br />

lengths are dramatically different from the results obtained <strong>with</strong> pure lipids. The<br />

apparent diffusion coefficients retrieved from the analysis <strong>of</strong> this data are clearly not<br />

justified on the basis <strong>of</strong> normal diffusion <strong>of</strong> monomeric species (as it was for mcp in<br />

DOPC), or on the basis <strong>of</strong> protein oligomerization/aggregation (as it as for mcp in<br />

DMoPC and in DEuPC). In this case, only segregation <strong>of</strong> mcp into domains enriched in<br />

the hydrophobically matching lipid can provide rationalization <strong>of</strong> the data. This<br />

hypothesis is confirmed by FRET <strong>studies</strong> that are also consistent <strong>with</strong> mcp segregation<br />

in the presence <strong>of</strong> binary mixtures <strong>of</strong> lipids <strong>with</strong> different acyl-chains. Importantly, the<br />

lipid components used in the binary mixtures are not dramatically different, the only<br />

difference being the presence <strong>of</strong> 4 additional carbons in the acyl-chain <strong>of</strong> one <strong>of</strong> the<br />

lipids. This difference is not expected to result in large scale phase separation <strong>of</strong> the<br />

lipid components such as it was detected in the FRET <strong>studies</strong> (domain size larger than<br />

or comparable to the Förster radius <strong>of</strong> AEDANS-BODIPY, 48.8 Å). In this way,<br />

formation <strong>of</strong> large domains enriched in the hydrophobically matching lipid must be a<br />

consequence <strong>of</strong> incorporation <strong>of</strong> mcp. Nevertheless, experiments making use <strong>of</strong> the<br />

properties <strong>of</strong> 1,6-diphenylhexatriene, indicate that even in the absence <strong>of</strong> protein, short<br />

scale clustering <strong>of</strong> the lipids in the binary mixtures is expected to occur.<br />

The same type <strong>of</strong> segregation <strong>of</strong> mcp was not detected when the acyl-chains’ lengths<br />

<strong>of</strong> the lipid components in the binary mixtures were identical and the only difference<br />

resides in the headgroups, confirming the crucial relevance <strong>of</strong> hydrophobic matching for<br />

protein organization.<br />

Concerning the quantification <strong>of</strong> the lipid selectivity <strong>of</strong> mcp, FRET proved to be a<br />

valuable tool, and an important alternative to ESR in these type <strong>of</strong> <strong>studies</strong>. FRET was<br />

able to detect lipid enrichment around monomeric mcp in a binary lipid mixture, while<br />

ESR was unable to resolve this problem (Sanders et al., 1992). Relative association<br />

182


CONCLUSIONS<br />

constants recovered by FRET were almost identical to the ones recovered for an<br />

aggregated form <strong>of</strong> mcp, confirming the reliability and sensitivity <strong>of</strong> the technique. mcp<br />

showed affinity for lipids <strong>with</strong> acyl chain lengths that matched the hydrophobic length<br />

<strong>of</strong> its TM domain, and for anionic lipids, particularly PA and PS.<br />

Only the lipids in the immediate vicinity <strong>of</strong> the protein are sufficiently immobilized<br />

to be detected by ESR, and as a consequence the binding affinity determined by ESR<br />

only refers to this shell <strong>of</strong> lipids. On the other hand, FRET is not limited to this, and for<br />

high Förster radii such as the one for the donor-acceptor pair used in this study, the<br />

efficiency <strong>of</strong> FRET from the protein to an acceptor in the second shell <strong>of</strong> lipids is<br />

similar to the efficiency <strong>of</strong> FRET to an acceptor in the first shell around the protein. In<br />

this way, the compatibility <strong>of</strong> the results from FRET and ESR points clearly to the<br />

validity <strong>of</strong> the annular model for protein-lipid interaction (single TM protein only<br />

induces changes in lipid distribution on the first shell <strong>of</strong> lipids around it).<br />

This FRET methodology for the quantification <strong>of</strong> protein-lipid selectivity also <strong>of</strong>fers<br />

the possibility <strong>of</strong> probing different shells <strong>of</strong> lipids around the protein by choosing<br />

donor-acceptor pairs <strong>with</strong> different Förster radii. To increase the sensitivity <strong>of</strong> the<br />

method to selectivity in the first shell <strong>of</strong> lipids around the protein, a small Förster radius<br />

would be advisable, that allowed to concentrate the majority <strong>of</strong> the contributions to<br />

donor quenching on the acceptors located there.<br />

183


Chapter III ‐ Binding <strong>of</strong> Inhibitors to a Putative Binding<br />

Domain <strong>of</strong> V‐ATPase.<br />

The <strong>studies</strong> conducted allowed us to know in detail the properties <strong>of</strong> the indole type<br />

<strong>of</strong> inhibitors in lipid <strong>membrane</strong>s. The molecules studied, which correspond to the initial<br />

(low efficiency) and final stage (high efficiency) in the development <strong>of</strong> these inhibitors<br />

as potential drugs in osteoporosis treatment, present almost identical properties in the<br />

liposomes environment. The preference <strong>of</strong> the lipid phase over the aqueous phase, the<br />

position in the bilayer and the orientation <strong>of</strong> the molecules were all almost identical for<br />

both molecules studied.<br />

According to the results <strong>of</strong> Gagliardi et al. (1998a), the change in inhibition<br />

efficiency <strong>of</strong> the V-ATPase inhibitors <strong>with</strong>in the indole class presented some extent <strong>of</strong><br />

correlation to the hydrophobicity <strong>of</strong> the molecules, i.e. the more hydrophobic indole<br />

inhibitor molecules had also a tendency to be more efficient inhibitors <strong>of</strong> V-ATPase. It<br />

was possible that the interaction <strong>of</strong> the molecules <strong>with</strong> the lipid milieu was significantly<br />

different <strong>with</strong>in this class <strong>of</strong> inhibitors, and this could correspond to dramatically<br />

different efficiencies <strong>of</strong> inhibition, explaining the 10 3 difference in IC 50 values <strong>of</strong> INH-1<br />

and SB242784. In this way, the process <strong>of</strong> inhibitor binding to the lipid bilayer would<br />

be crucial to the efficiency <strong>of</strong> the inhibition event. However, the results obtained point<br />

to a non-determinant role <strong>of</strong> inhibitor interaction <strong>with</strong> the lipid bilayer, and the specific<br />

molecular recognition processes <strong>with</strong> the enzyme’s inhibition site are very likely to be<br />

responsible for dictating the large differences <strong>of</strong> inhibition efficiencies <strong>with</strong>in the V-<br />

ATPase indole inhibitors class.<br />

The binding <strong>studies</strong> <strong>of</strong> bafilomycin and SB242784 to the peptide corresponding to<br />

the 4 th TM helix <strong>of</strong> subunit c <strong>of</strong> V-ATPase revealed no affinity at all for SB242784 and<br />

very low affinities for bafilomycin. These results can be seen as an indicator that the<br />

binding pocket for the inhibitors is not solely composed <strong>of</strong> the lipid exposed face <strong>of</strong> the<br />

4 th TM helix <strong>of</strong> subunit c, but also involves contributions <strong>of</strong> amino acids from other<br />

domains <strong>of</strong> the protein. This proposal is in agreement <strong>with</strong> the hypothesis <strong>of</strong> the<br />

inhibitors operating as a stone in a gear, blocking movements <strong>of</strong> the TM domains <strong>of</strong><br />

subunit c and interrupting proton transport.<br />

184


CONCLUSIONS<br />

Chapter IV ‐ Binding <strong>of</strong> a Quinolone Antibiotic to Bacterial<br />

Porin OmpF.<br />

In this case, the FRET analysis allowed us to conclude on a likely location for the<br />

binding site <strong>of</strong> CP in OmpF. Two limiting positions for the binding site were<br />

considered: i) in the trimer interface, leading to a maximum efficiency <strong>of</strong> donor<br />

quenching; ii) in the periphery <strong>of</strong> the trimer, causing minimum quenching <strong>of</strong> donor<br />

fluorescence due to FRET. For each <strong>of</strong> these two limiting conditions, intervals for<br />

binding constants were retrieved. By comparison <strong>of</strong> the binding constants retrieved <strong>with</strong><br />

the results <strong>of</strong> an independent method that made use <strong>of</strong> changes in the absorption spectra<br />

<strong>of</strong> CP upon binding to OmpF (Neves et al., 2005), it is concluded that the binding site<br />

for CP is likely to be located away from the trimer interface and close to the periphery<br />

<strong>of</strong> OmpF.<br />

The FRET methodology developed for the analysis <strong>of</strong> the interactions between<br />

OmpF and CP is suitable to be used in the analysis <strong>of</strong> FRET data for similar systems,<br />

containing two populations <strong>of</strong> donors (or acceptors) separated by a distance larger or<br />

comparable to the Förster distance <strong>of</strong> the donor-acceptor pair used, so that the analysis<br />

is sensitive enough to be able to distinguish the contributions from these two<br />

populations.<br />

185


Chapter V – <strong>Interaction</strong> <strong>of</strong> Helix‐0 <strong>of</strong> the N‐BAR Domain<br />

With Lipid Membranes<br />

The <strong>membrane</strong> modelling properties <strong>of</strong> N-BAR domains are generally attributed to<br />

two structural features <strong>of</strong> the BAR domain containing <strong>proteins</strong>: the concave surface <strong>of</strong><br />

the crescent shaped dimer, that is expected to act as a scaffold for curvature and an N-<br />

terminal amphipatic helix or H0, that is expected to be deeply buried <strong>with</strong>in the lipid<br />

bilayer, driving monolayer asymmetry and finally curvature. This latter strategy for<br />

curvature generation should be effective in the absence <strong>of</strong> the rest <strong>of</strong> the protein, and a<br />

peptide comprising this amphipatic helix would be expected to produce the same<br />

results. The results reproduced here clearly demonstrate that this is not the case. H0 is<br />

unable to drive significant changes in liposome morphology. Therefore we can conclude<br />

that N-BAR domains do not use their N-terminal amphipatic helices to produce bilayer<br />

curvature in the manner <strong>of</strong>ten proposed (Gallop and McMahon, 2005; Gallop et al.,<br />

2006; Masuda et al., 2006). Nevertheless, the results point to different roles <strong>of</strong> Helix-0<br />

on <strong>membrane</strong> remodelling by N-BAR domains. The high water/lipid partition<br />

coefficients determined for H0 are expected to have a dramatic effect in the partition<br />

properties <strong>of</strong> the entire protein, and this alone could explain the great decreases in<br />

liposome tubulation efficiency <strong>of</strong> BAR domains <strong>with</strong>out H0. A role as a mediator <strong>of</strong><br />

interactions between dimers <strong>of</strong> BAR domains was also proposed for H0 (Gallop and<br />

McMahon, 2005), and here it is shown that H0 does indeed oligomerize into dimers in<br />

the lipid environment. Oligomerization <strong>of</strong> H0 might provide BAR domain-containing<br />

<strong>proteins</strong> <strong>with</strong> a more efficient mechanism to colocalize, and <strong>membrane</strong> curvature<br />

induction by scaffolding by <strong>proteins</strong> is only expected to be efficient at high local<br />

concentrations <strong>of</strong> scaffold <strong>proteins</strong>.<br />

186


CONCLUSIONS<br />

Chapter VI – Clustering <strong>of</strong> PI(4,5)P2 in Fluid PC Bilayers.<br />

Domain enrichment <strong>of</strong> PI(4,5)P 2 is expected to occur in vivo. Nevertheless the<br />

process must require strong binding to agents (<strong>proteins</strong>) that present a strong affinity for<br />

incorporation in cholesterol enriched domains. Clustering <strong>of</strong> charged phospholipids in<br />

the gel state is certainly possible as energetic requirements resulting from packing<br />

restrictions in this state can surpass the energetic penalty resulting from electrostatic<br />

repulsion between the clustered lipids. However, the same type <strong>of</strong> phenomena is<br />

expected to be highly unlikely in fluid bilayers, where packing restrictions are not so<br />

severe.<br />

The results from this study allowed us to conclude that PI(4,5)P 2 is homogeneously<br />

distributed in fluid PC bilayers. FRET <strong>studies</strong> ruled out the existence <strong>of</strong> any large scale<br />

domain (dimensions larger than 40 Å) enriched in PI(4,5)P 2 , while energy migration<br />

and fluorescence self-quenching <strong>studies</strong> <strong>of</strong> a fluorescently labelled PI(4,5)P 2 clearly<br />

indicate absence <strong>of</strong> strong short-range clustering <strong>of</strong> the lipid.<br />

Another important result from this study are the differences in water/bilayer<br />

partition efficiencies obtained for different protonation states <strong>of</strong> PI(4,5)P 2 . One possible<br />

explanation for this observation is that the micellar structure <strong>of</strong> PI(4,5)P 2 is destabilized<br />

at the completely deprotonated state (at low lamellar lipid concentrations, a micellar<br />

population <strong>of</strong> PI(4,5)P 2 is expected to be in equilibrium <strong>with</strong> a bilayer inserted<br />

population). These differences in partition behaviour can result in erroneous<br />

rationalization <strong>of</strong> fluorescent data, specially taking into account that the most popular<br />

fluorescently labelled PI(4,5)P 2 in the market presents very short acyl-chains, and will<br />

present even smaller partition efficiencies that the ones determined here for a long chain<br />

fluorescently labelled PI(4,5)P 2 .<br />

187


188


FINAL CONSIDERATIONS AND PROSPECTS<br />

VIII<br />

FINAL CONSIDERATIONS AND<br />

PROSPECTS<br />

<strong>Biophysical</strong> <strong>studies</strong> are likely the only possibility <strong>of</strong> direct structural and dynamic<br />

characterization <strong>of</strong> the type <strong>of</strong> interactions addressed in this work. The recurrent<br />

difficulties in <strong>membrane</strong> protein expression and purification were responsible for the<br />

popularity <strong>of</strong> the application <strong>of</strong> highly sensitive fluorescence techniques in <strong>studies</strong> <strong>of</strong><br />

<strong>membrane</strong> <strong>proteins</strong>. The presence <strong>of</strong> aromatic amino acids in <strong>proteins</strong> allows the<br />

application <strong>of</strong> fluorescence techniques <strong>with</strong>out requirement for labelling <strong>with</strong> a extrinsic<br />

fluorophore. Still, when necessary, the developments in synthesis <strong>of</strong> fluorophores and<br />

labelling techniques, provide the researcher <strong>with</strong> great flexibility in the process <strong>of</strong><br />

choosing the best conditions for fluorescence methodologies. Overall, the <strong>studies</strong><br />

presented here are a demonstration <strong>of</strong> the potential <strong>of</strong> biophysical <strong>studies</strong>, particularly<br />

fluorescence spectroscopy <strong>studies</strong>, in the characterization <strong>of</strong> interactions between<br />

components <strong>of</strong> bio<strong>membrane</strong>s.<br />

In this work, the use <strong>of</strong> <strong>membrane</strong> model systems, allowed for the simplification <strong>of</strong><br />

the number <strong>of</strong> variables in the experiments. This is essential in <strong>studies</strong> that focus on the<br />

characterization <strong>of</strong> very specific interactions. Apart from the technical difficulties<br />

inherent to in vivo <strong>studies</strong> (huge heterogeneity <strong>of</strong> samples, background fluorescence,<br />

light scattering, low fluorescence intensity, etc.), the presence <strong>of</strong> a large number <strong>of</strong><br />

uncontrolled (and sometimes unknown) variables is highly likely to hinder the<br />

possibility <strong>of</strong> data rationalization. The information recovered from <strong>studies</strong> <strong>with</strong> less<br />

complex model systems can then be used for a more clear rationalization <strong>of</strong> data<br />

gathered from in vivo <strong>studies</strong>.<br />

The new FRET methodologies developed here, notably the methodology for<br />

quantification <strong>of</strong> protein-lipid selectivity, are expected to be <strong>of</strong> great use in the future as<br />

they are readily applicable to different systems. The FRET methodologies for<br />

quantification <strong>of</strong> drug-protein binding will require only some straightforward<br />

adaptations to the geometric constraints <strong>of</strong> the systems to be studied.<br />

189


Some <strong>of</strong> the biological questions debated here are still open. Recently, a model for<br />

the binding site <strong>of</strong> V-ATPase inhibitors in the c-subunit was proposed (Bowman et al.,<br />

2006), confirming that residues in at least three different trans<strong>membrane</strong> domains were<br />

required for effective bafilomycin inhibition. However the picture was not completely<br />

clear, as residues in opposite sides <strong>of</strong> the same helix were found to be relevant for<br />

bafilomycin binding pointing to indirect effects <strong>of</strong> some residues in the interaction <strong>with</strong><br />

the inhibitor. Although, clinical trials <strong>with</strong> SB242784 were not started due to possible<br />

adverse effects (Nikura et al., 2005), a derivative <strong>of</strong> this molecule was shown recently to<br />

potentiate the effect <strong>of</strong> an cytotoxic/anti-tumor drug likely due to inhibition <strong>of</strong> V-<br />

ATPases in tumor cells (Petrangoline et al., 2006).<br />

The study <strong>of</strong> the mechanisms behind clustering <strong>of</strong> PI(4,5)P 2 into cholesterol<br />

dependent domains in bio<strong>membrane</strong>s is a very promising field. It is clear that PI(4,5)P 2<br />

should not spontaneously cluster or partitionate to liquid ordered domains, and this<br />

behaviour must be connected to specific protein-lipid interactions. In this way, the study<br />

<strong>of</strong> the lipid <strong>membrane</strong> partition properties <strong>of</strong> <strong>proteins</strong> known to bind PI(4,5)P 2 might<br />

give us some insight into these mechanisms. The relevance <strong>of</strong> acylation characteristics<br />

(number and length <strong>of</strong> acyl-chains in the protein), for <strong>membrane</strong> protein organization in<br />

bio<strong>membrane</strong>s, is a problem <strong>of</strong> great biological relevance.<br />

The use <strong>of</strong> fluorescence imaging techniques in the study <strong>of</strong> protein-lipid interactions<br />

can add additional layers <strong>of</strong> information to the results <strong>of</strong> non-space resolution<br />

biophysical <strong>studies</strong>. The biophysical <strong>studies</strong> presented here relied mostly on<br />

macroscopic observables, and although this type <strong>of</strong> data is, as demonstrated, highly<br />

enriched in information, some <strong>of</strong> it is lost by averaging <strong>of</strong> fluorophore populations,<br />

while imaging techniques allow for resolution <strong>of</strong> different populations <strong>of</strong> fluorophores.<br />

Therefore, the application <strong>of</strong> the methodologies derived and described here to<br />

fluorescence imaging data could help identify features <strong>of</strong> interactions between<br />

<strong>membrane</strong> components which go undetected in the macroscopic data. In the limit,<br />

single-molecule techniques could also be used. Imaging techniques are obviously very<br />

useful in the observation <strong>of</strong> large scale changes (µm scale) in the lateral distribution <strong>of</strong><br />

<strong>membrane</strong> components and are a useful complement to FRET methodologies, which are<br />

able to probe deviations to homogeneity in the nanometer scale. Additionally, FRET<br />

<strong>studies</strong> under the microscope are also carried out, allowing the direct comparison<br />

between macroscopic and microscopic data.<br />

190


BIBLIOGRAPHY<br />

IX<br />

BIBLIOGRAPHY<br />

Alberts, B., Johnson, A., Lewis, J., Raff, M., Roberts, K., and Walter, P. (2002) Molecular Biology <strong>of</strong> the<br />

Cell. 4 th Ed. Garland Science, Oxford, UK.<br />

Albini, A., and Monti, S. (2003). Photophysics and Photochemistry <strong>of</strong> Fluoroquinolones. Chem. Soc.<br />

Rev. 32: 238-250.<br />

Anantharamaiah, G. M., Jones, M. K., and Segrest, J. P. (1993) An Atlas <strong>of</strong> the Amphipatic Helical<br />

Domains <strong>of</strong> Human Exchangeable Plasma Apolipo<strong>proteins</strong>. In The Amphipatic Helix. Ed. Epand, R. M.<br />

CRC Press, Boca-Raton, U.S.A.<br />

Arnold, K., Losche, A., and Gawrisch, K. (1981). 31p-NMR Investigations <strong>of</strong> Phase Separation in<br />

Phosphatidylcholine/Phosphatidylethanolamine Mixtures. Biochem. Biophys. Acta. 645:143–148.<br />

Auger, I. E. (1993) Computational Techniques to Predict Amphipatic Helical Segments. in The<br />

Amphipatic Helix. Ed. Epand, R. M. CRC Press, Boca-Raton, U.S.A.<br />

Bagatolli, L. A., and Gratton, E. (2000). Two Photon Fluorescence Microscopy <strong>of</strong> Coexisting Llipid<br />

Domains in Giant Unilamellar Vesicles <strong>of</strong> Binary Phospholipid Mixtures. Biophys. J. 78: 290-305.<br />

Ben-Shaul, A. (1995). Molecular Theory <strong>of</strong> Chain Packing, Elasticity and Lipid-Protein <strong>Interaction</strong>s in<br />

Lipid Bilayers. In “Structure and Dynamics <strong>of</strong> Membranes. . From Cells to Vesicles.” (Eds Lipowsky, R.,<br />

and Sackmann, E.) North Holland, Amsterdam, The Netherlands.<br />

Berg, J. M, Tymoczko, J. L., and Stryer, L. (2002). Biochemistry. 5 th Ed. W. H. Freeman, New York,<br />

USA.<br />

Bloom, M. E., and Mouritsen, O. G. (1984). Mattress Model <strong>of</strong> Lipid-Protein <strong>Interaction</strong>s in Membranes.<br />

Biophys. J. 46: 141-153.<br />

Botelho, R. J., Teruel, M., Dierckman, R., Anderson, R., Wells, A., York, J. D., Meyer, T., and Gristein,<br />

S. (2000). Localized Biphasic Changes in Phosphatidylinositol-4,5-Bisphosphate at Sites <strong>of</strong> Phagocytosis.<br />

J. Cell. Biol. 151: 1353-1367.<br />

Bowman, B. J., and Bowman, E. J. (2002). Mutations in Subunit c <strong>of</strong> the Vacuolar ATPase Confer<br />

Resistance to Bafilomycin and Identify a Conserved Antibiotic Binding Site. J. Biol. Chem. 277: 3965-<br />

3972.<br />

191


Bowman, E. J., Graham, L. A., Stevens, T. H., and Bowman, B. J. (2004). The<br />

Bafilomycin/Concanamycin Binding Site in Subunit c <strong>of</strong> the V-ATPase from Neurospora crassa and<br />

Sacharomyces cerevisiae. J. Biol. Chem. 279: 33131-33138.<br />

Bowman, B. J., McCall, M. E., Baertsch, R., and Bowman, E. J. (2006). A Model for the Proteolipid Ring<br />

and Bafilomycin/Concanamycin-binding Site in the Vacuolar ATPase <strong>of</strong> Neurospora crassa. J. Biol.<br />

Chem. 42: 31885-31893.<br />

Brasseur, R. (1991). Differentiation <strong>of</strong> lipid-associating helices by use <strong>of</strong> three-dimensional molecular<br />

hydrophobicity potential calculations. J. Biol. Chem. 266: 16120-16127.<br />

Braun, P., and von Heijne, G. (1999). The Aromatic Residues Trp and Phe Have Different Effects on the<br />

Positioning <strong>of</strong> a Trans<strong>membrane</strong> Helix in the Microsomal Membrane. Biochemistry 38: 9778-9782.<br />

Caputo, G. A., and London, E. (2003). Cumulative Effects <strong>of</strong> Amino-Acid Substitution and Hydrophobic<br />

Mismatch Upon the Trans<strong>membrane</strong> Stability and Conformation <strong>of</strong> Hydrophobic α-Helices.<br />

Biochemistry 42: 3275-3285.<br />

Caputo, G. A., and London, E. (2004). Position and Ionization State <strong>of</strong> Asp in the Core <strong>of</strong> Membrane-<br />

Inserted α-Helices Control Both the Equilibrium between Trans<strong>membrane</strong> and Nontrans<strong>membrane</strong> Helix<br />

Topography and Trans<strong>membrane</strong> Helix Positioning. Biochemistry 42: 3275-3285.<br />

Chapman, J. S., and Georgopapadakou, N. H. (1988). Routes <strong>of</strong> Quinolone Permeation in Escherichia<br />

coli. Antimicrob. Agents and Chemoter. 32: 438-442.<br />

Chevalier, J., Malléa, M., and Pagés, J. M. (2000). Comparative Aspects <strong>of</strong> the Diffusion <strong>of</strong> Norfloxacin,<br />

Cefepime and Spermine through the F Porin Channel <strong>of</strong> Enterobacter cloacae. J. Biochem. 348: 223-227.<br />

Dawson, J. C., Legg, J. A., and Machesky, L. M. (2006). Bar Domain Proteins: A Role in Tubulation,<br />

Scission and Actin Assembly in Clathrin-Mediated Endocytosis. Trends Cell Biol. 16: 493-498.<br />

de Almeida, R. F. M., Loura, L. M. S., Fedorov, A., and Prieto, M. (2002). Nonequilibrium Phenomena in<br />

the Phase Separation <strong>of</strong> a Two-Component Lipid Bilayer. Biophys. J. 82: 823-834.<br />

de Planque, M. R. R., Killian, J. A. (2003). Protein-Lipid <strong>Interaction</strong>s Studied With Designed<br />

Trans<strong>membrane</strong> Peptides: Role <strong>of</strong> Hydrophobic Matching and Interfacial Anchoring. Mol. Memb. Biol.<br />

20: 271-284.<br />

Dechené, M., Leying, H., and Cullman, W. (1990). Role <strong>of</strong> the Outer Membrane for Quinolone<br />

Resistance in Enterobacteria. Chemoterapy 36: 13-23.<br />

Denisov, G., Wanaski, S., Luan, P., Glaser, M., and McLaughlin, S. (1998). Binding <strong>of</strong> Basic Peptides to<br />

Membranes Produces Lateral Domains Enriched in the Acidic Lipids Phosphatidylserine and<br />

192


BIBLIOGRAPHY<br />

Phosphatidylinositol 4,5-bisphosphate: An Electrostatic Model and Experimental Results. Biophys. J. 74:<br />

731-744.<br />

Devaux, P. F., and Morris, R. (2004). Trans<strong>membrane</strong> Asymmetry and Lateral Domains in Biological<br />

Membranes. Traffic 5: 241-246.<br />

Dong, M., Bagetto, L. G., Falson, P., Le Maire, M., and Penin, F. (1997). CompleteRemoval and<br />

Exchange <strong>of</strong> Sodium Dodecyl Sulfate Bound to Soluble and Membrane Proteins and Restoration <strong>of</strong> Their<br />

Activities, Using Ceramic Hydroxypatite Chromatography. Anal. Biochem. 247: 333-341.<br />

Dröse, S. and Altendorf, K. (1997). Bafilomycins and Concanamycins as Inhibitors <strong>of</strong> V-ATPases and P-<br />

ATPases. J. Exp. Biol. 200, 1-8.<br />

Dumas, F., Sperotto, M. M., Lebrun, M.-C., Tocanne, J.-F., and Mouritsen, O. G. (1997). Molecular<br />

Sorting <strong>of</strong> Lipids by Bacteriorhodopsin in Dilauroylphosphatidylcholine/Distearoylphophatidylcholine<br />

Lipid Bilayers. Biophys. J. 73: 1940-1953.<br />

Dumas, F., Lebrun, C., and Tocanne, J-F. (1999). Is the Protein/Lipid Hydrophobic Matching Principle<br />

Relevant to Membrane Organization and Function. FEBS Lett. 458: 271-277.<br />

Dunnil, P. (1968). The Use <strong>of</strong> Helical Net-Diagrams to Represent Protein Structures. Biophys. J. 8: 865-<br />

875<br />

East, J. M., and Lee, A. G. (1982). Lipid Selectivity <strong>of</strong> the Calcium and Magnesium Ion Dependent<br />

Adenosinetriphosphatase, Studied With Fluorescence Quenching by a Brominated Phospholipid.<br />

Biochemistry 21: 4144-4151.<br />

Eisenberg, D., Weiss, R. M., Terwillinger, T. C. (1982). The Helical Hydrophobic Moment: A Measure <strong>of</strong><br />

the Amphiphilicity <strong>of</strong> Helix. Nature 299: 371-374.<br />

Epand, R. M. (1993) Introduction. in The Amphipatic Helix. Ed. Epand, R. M. CRC Press, Boca-Raton,<br />

U.S.A.<br />

Epand, R. M., Vuong, P., Yip, C. M., Maekawa, S., and Epand, R. F. (2004). Cholesterol-Dependent<br />

Partitioning <strong>of</strong> PtdIns(4,5)P 2 into Membrane Domains by the N-terminal Fragment <strong>of</strong> NAP-22 (Neuronal<br />

Axonal Myristoylated Membrane Protein <strong>of</strong> 22 kDa). Biochem. J. 379: 527-532.<br />

Epand, R. M. (2005) The Role <strong>of</strong> Proteins in the Formation <strong>of</strong> Domains in Membranes. In In “Protein-<br />

Lipid <strong>Interaction</strong>s: New Approaches and Emerging Concepts.” (Eds. Mateo, C. R., Gómez, J., Villalaín,<br />

J., Ros, J. M. G.) Springer-Verlag. Berlin, Germany.<br />

Epand, R. F., Sayer, B. G., and Epand, R. M. (2005). Induction <strong>of</strong> Raft-Like Domains by a Myristoilated<br />

NAP-22 Peptide and its Tyr Mutant. FEBS J. 272: 1792-1803.<br />

193


Farsad, K., Ringstad, N., Takei, K., Floyd, S. R., Rose, K., and De Camilli, P. (2001). Generation <strong>of</strong> High<br />

Curvature Membranes Mediated by Direct Endophilin Bilayer <strong>Interaction</strong>s. J. Cell Biol. 155: 193-200.<br />

Fattal, D. R., and Ben-Shaul, A. (1993). A Molecular Model for Lipid-Protein <strong>Interaction</strong> in Membranes:<br />

The Role <strong>of</strong> Hydrophobic Mismatch. Biophys.J. 65: 1795-1809.<br />

Finbow, M. E., and Harrison, M. A. (1997). The Vacuolar H + -ATPase: A Universal Proton Pump <strong>of</strong><br />

Eukaryotes. Biochem. J. 324: 697-712.<br />

Ford, M.G.J., Mills, I.G., Peter, B.J., Vallis, Y., Praefcke, G.J.K., Evans, P.R. and McMahon, H.T. (2002)<br />

Curvature <strong>of</strong> Clathrin-Coated Pits Driven by Epsin. Nature 419: 361-366.<br />

Gagliardi, S., Nadler, G., Consolandi, E., Parini, C., Morvan, M., Legave, M. N., Belfiore, P., Zocchetti,<br />

A., Clarke, G. D., James, I., Nambi, P., Gowen, M., and Farina, C. (1998a). 5-(5,6-Dichloro-2-indolyl)-2-<br />

methoxy-2,4-pentadienamides: Novel and Selective Inhibitors <strong>of</strong> the Vacuolar H + -ATPase <strong>of</strong> Osteoclasts<br />

<strong>with</strong> Bone Antiresorptive Activity. J. Med. Chem. 41: 1568-1573.<br />

Gagliardi, S., Gatti, P. A., Belfiore, P., Zochetti, A., Clarke, G. D., and Farina, C. (1998b). Synthesis and<br />

Structure-Activity Relationships <strong>of</strong> Bafilomycin A 1 Derivatives as Inhibitors <strong>of</strong> Vacuolar H + -ATPase. J.<br />

Med. Chem. 41: 1883-1893.<br />

Gallop, J. L., and McMahon, H. T. (2005). BAR Domains and Membrane Curvature: Bringing your<br />

Curves to the BAR. Biochem. Soc. Symp. 72: 223-31.<br />

Gallop, J. L., Jao, C. C., Kent, H. M., Butler, P. J. G., Evans, P. R., Langen, R., and McMahon, H. T.<br />

(2006). Mechanism <strong>of</strong> Endophilin N-BAR Domain-Mediated Membrane Curvature. EMBO J. 25: 2898-<br />

2910.<br />

Garavito, R. M., and Rosenbusch, J. P. (1980). Three Dimensional Crystals <strong>of</strong> an Integral Membrane<br />

Protein: An Initial X-Ray Analysis. J. Cell. Biol. 86: 327-329.<br />

Ge, K., and Prendergast, G. C. (2000). Bin2, a Functionally Nonredundant Member <strong>of</strong> the BAR Adaptor<br />

Gene Family. Genomics 67: 210-220.<br />

Gennis, RB (1989). Bio<strong>membrane</strong>s: Molecular Structure and Function. Springer-Verlag, New York,<br />

USA.<br />

Ghambir, A., Hangyás-Milhályné, G., Zaitseva, I., Cafiso, D. S., Wang, J., Murray, D., Pentyala, S. N.,<br />

Smith, S. O., and McLaughlin, S. (2004). Electrostatic Sequestration <strong>of</strong> PIP 2 on Phospholipid Membranes<br />

by Basic/Aromatic Regions <strong>of</strong> Proteins. Biophys. J. 86: 2188-2207.<br />

Gil, T., Ipsen, J. H., Mouritsen, O. G. , Sabra, M. C., Sperotto, M. M., and Zuckermann, M. J. (1998).<br />

Theoretical Analysis <strong>of</strong> Protein Organization in Lipid Membranes. Biochim. Biophys. Acta 1376: 245-<br />

266.<br />

194


BIBLIOGRAPHY<br />

Glaubitz, C., Grobner, G., Watts, A. (2000). Structural and Orientational Information <strong>of</strong> the Membrane<br />

Embedded M13 Coat Protein by 13C-MAS NMR Spectroscopy. Biochim. Biophys. Acta 1463: 151-161.<br />

Gokhale, N. A., Abraham, A., Digman, M. A., Gratton, E., and Cho, W. (2005). Phosphoinositide<br />

Specificity <strong>of</strong> and Mechanism <strong>of</strong> Lipid Domain Formation by Annexin A2-p11 Heterotetramer. J. Biol.<br />

Chem. 280: 42831-42840.<br />

Grabe, M., Wang, H., and Oster, G. (2000). The Mechanochemistry <strong>of</strong> V-ATPase Proton Pumps. Biophys.<br />

J. 78: 2798-2813.<br />

Harrison, M. A., Finbow M. E., and Findlay, J. B. (1997) Postulate for the Molecular Mechanism <strong>of</strong> the<br />

Vacuolar H(+)-ATPase (Hypothesis). Mol. Membr. Biol. 14: 1-3.<br />

Harrison, M. A., Murray, J., Powell, B., Kim, Y.-I., Finbow, M. E., and Findlay, J. B. C. (1999). Helical<br />

<strong>Interaction</strong>s and Membrane Disposition <strong>of</strong> the 16-kDa Proteolipid Subunit <strong>of</strong> the Vacuolar H + -ATPase<br />

Analyzed by Cysteine Replacement Mutagenesis. J. Biol. Chem. 274: 25461-25470.<br />

Harrison, M., Powell, B., Finbow, M. E., and Findlay, J. B. C. (2000). Identification <strong>of</strong> Lipid-Acessible<br />

Sites on the Nephrops 16-kDa Proteolipid Incorporated into a Hybrid Vacuolar H + -ATPase: Site<br />

Directed Labeling <strong>with</strong> N-(1-Pyrenyl)cyclohexylcarbodiimide and Fluorescence Quenching Analysis.<br />

Biochemistry 39: 7531-7537.<br />

Hardt, S. L. (1979). Rates <strong>of</strong> Diffusion Controlled Reactions in One, Two, Three Dimensions. Biophys.<br />

Chem. 10: 239-243.<br />

Harroun, T. A., Heller, W. T., Weiss, T. M., Yang, L., and Huang, H. W. (1999). Experimental Evidence<br />

for Hydrophobic Matching and Membrane-Mediated <strong>Interaction</strong>s in Lipid Bilayers Containing<br />

Gramicidin. Biophys. J. 76: 937-945.<br />

Harzer, U. and Bechinger, B. (2000). Alignment <strong>of</strong> Lysine-Anchored Membrane Peptides Under<br />

Conditions <strong>of</strong> Hydrophobic Mismatch: a CD, 15N and 31P Solid-State NMR Spectroscopy Investigation.<br />

Biochemistry 39: 13106-13114.<br />

Hemminga, M. A., Sanders, J. C., and Spruijt, R. B. (1992). Spectroscopy <strong>of</strong> Lipid-Protein <strong>Interaction</strong>s:<br />

Structural Aspects <strong>of</strong> Two Different Forms <strong>of</strong> the Coat Protein <strong>of</strong> Bacteriophage M13 Incorporated in<br />

Model Membranes. Prog. Lipid. Res. 31: 301-333.<br />

Henry, G. D., and Sykes, B. D. (1992). Assignment <strong>of</strong> Amide 1 H and 15 N NMR Resonances in Detergent-<br />

Solubilized M13 Coat Protein: A Model for the Coat Protein Dimer. Biochemistry 31: 5284-5297.<br />

Hinderliter, A., Almeida, P. F. F., Creutz, C. E., and Biltonen, R. L. (2001). Domain Formation in a Fluid<br />

Mixed Lipid Bilayer Modulated Through Binding <strong>of</strong> the C2 Protein Motif. Biochemistry 40: 4181-4191.<br />

195


Hinderliter, A., Biltonen, R. L., and Almeida, P. F. F. (2004). Lipid Modulation <strong>of</strong> Protein-Induced<br />

Membrane Domains as a Mechanism for Controlling Signal Transduction. Biochemistry 43: 7102-7110.<br />

Hooper, D. C., and Wolfson, J. S. (Eds.), (1995). Quinolone Antimicrobial Agents, 2 nd Ed. American<br />

Society from Microbiology, Washington D.C, U.S.A.<br />

Houston, S., and Farming, A. (1994). Current and Potential Treatment <strong>of</strong> Tuberculosis. Drugs 48: 689-<br />

708.<br />

Ipsen, J. H., Karlstrom, G., Mouritsen, O. G., Wennerstrom, H., and Zuckermann, M. J. (1987). Phase<br />

Equilibria in the Phosphatidylcholine-Cholesterol System. Biochem. Biophys. Acta 905:162–72.<br />

Jensen, M. O, and Mouritsen, O. G. (2004). Lipids do Influence Protein Function – The Hydrophobic<br />

Matching Hypothesis Revisited. Biochim. Biophys. Acta 1666: 205-226.<br />

Jorgensen, K., and Mouritsen, O. G. (1995). Phase Separation Dynamics and Lateral Organization <strong>of</strong><br />

Two-Component Lipid Membranes. Biophys. J. 95: 942-954.<br />

Kawasaki-Nishi, S. Nishi, T., and Forgac, M. (2003). Proton Translocation Driven by ATP Hydrolisis in<br />

V-ATPases. FEBS Letters 545: 76-85.<br />

Killian, J. A., Trouard, T. P., Greathouse, D. V., Chupin, V., Lindblom, G. (1994). A General Method for<br />

the Preparation <strong>of</strong> Mixed Micelles <strong>of</strong> Hydrophobic Peptides and Sodium Dodecyl Sulphate. FEBS Lett.<br />

348: 161-165.<br />

Killian, J. A. (2003). Synthetic Peptides as Models for Intrinsic Membrane Proteins. FEBS Lett. 555:<br />

134-138.<br />

Kitamura, A., Kiyota, T., Tomohiro, M., Umeda, A., Lee, S., Inoue, T., and Sugihara, G. (1999).<br />

Morphological Behaviour <strong>of</strong> Acidic and Neutral Liposomes Induced by Basic Amphiphilic Alpha-Helical<br />

Peptides <strong>with</strong> Systematically Varied Hydrophobic-Hydrophilic Balance. Biophys. J. 76: 1457-1468.<br />

Kuhn, A., Wickner, W., and Kreil, G. (1986). The Cytoplasmic Carboxy Terminus <strong>of</strong> M13 Procoat is<br />

Required for the Membrane insertion <strong>of</strong> its Central Domain. Nature 322: 335-339.<br />

Klausner, R. D., and Wolf, D. E. (1980). Selectivity <strong>of</strong> Fluorescent Lipid Analogues for Lipid Domains.<br />

Biochemistry 19: 6199-6203.<br />

Knol, J., Sjollema, K., and Poolman, B. (1998). Detergent-Mediated Reconstitution <strong>of</strong> Membrane<br />

Proteins. Biochemistry 37: 16410-16415.<br />

Korlach, J., Schwille, P., Webb, W.W., Feigenson, G.W. (1999). Characterization <strong>of</strong> Lipid Bilayer<br />

Phases by Confocal Microscopy and Fluorescence Correlation Spectroscopy. Proc. Natl. Acad. Sci.<br />

U.S.A. 96: 8461-8466.<br />

196


BIBLIOGRAPHY<br />

Koehorst, R. B. M., Spruijt, R. B., Vergeldt, F. J. and Hemminga, M. A. (2004). Lipid Bilayer Topology<br />

<strong>of</strong> the Trans<strong>membrane</strong> α-Helix <strong>of</strong> M13 Major Coat Protein and Bilayer Polarity Pr<strong>of</strong>ile by Site-Directed<br />

Fluorescence Spectroscopy. Biophys. J. 87: 1445-1455.<br />

Kwik, J., Boyle, S., Fooksman, Margolis, L., Sheetz, M. P., and Edidin, M. (2003). Membrane<br />

Cholesterol, Lateral Mobility, and the Phosphatidylinositol 4,5-Bisphosphate-Dependent Organization <strong>of</strong><br />

Cell Actin. PNAS USA. 100: 13964-13969.<br />

Lark, M. W., James, I. E. (2002). Novel Bone Antiresorptive Approaches. Curr. Opin. Pharmac. 2: 330-<br />

337.<br />

Laux, T., Fukami, K., Thelen, M., Golub, T., Frey, D., and Caroni, P. (2000). GAP43, MARCKS, and<br />

CAP23 Modulate PI(4,5)P 2 at Plasmalemmal Rafts, and Regulate Cell Cortex Actin Dynamics Through a<br />

Common Mechanism. J. Cell Biol. 149: 1455-1471.<br />

Lee, A.G. (2003). Lipid-Protein <strong>Interaction</strong>s in Biological Membranes: a Structural Perspective.<br />

Biophys. Biochim. Acta 1612: 1-40.<br />

Lehtonen, J. Y. A., Holopainen, J. M., and Kinnunen, P. K. (1996). Evidence for the Formation <strong>of</strong><br />

Microdomains in Liquid Crystalline Large Unilamellar Vesicles Caused by Hydrophobic Mismatch <strong>of</strong> the<br />

Constituent Phospholipids. Biophys. J. 70: 1753-1760.<br />

Lehtonen, J. Y. A., and Kinnunen, P. K. J. (1997). Evidence for Phospholipid Microdomain Formation in<br />

Liquid Crystalline Liposomes Reconstituted With Eschericia coli Lactose Permease. Biophys. J. 72:<br />

1247-1257.<br />

Liu, F., Lewis, R. N. A. H., Hodges, R. S., and McElhaney, R. N. (2002). Effect <strong>of</strong> Variations in the<br />

Structure <strong>of</strong> a Polyleucine-Based α-Helical Trans<strong>membrane</strong> Peptide on its <strong>Interaction</strong>s <strong>with</strong><br />

Phosphatidylcholine Bilayers. Biochemistry 87: 2470-82<br />

Lentz, B. R., Barenholz, Y. and Thompson, T. E. (1976). Fluorescence Depolarization Studies <strong>of</strong> Phase<br />

Transitions and Fluidity in Phospholipid Bilayers. 2. Two-component Phosphatidylcholine Liposomes.<br />

Biochemistry 15:4529–4537.<br />

Lentz, T. L., Chaturverdi, V., and Conti-Fine, B. M. (1998). Amino Acids Within Residues 181-200 <strong>of</strong> the<br />

Nicotinic Acetylcholine Receptor Alpha1 Subunit Involved in Nicotine Binding. Biochem. Pharmacol. 55:<br />

341-347.<br />

Lewis, B. A., and Engelman, D. M. (1983). Bacteriorhodopsin Remains Dispersed in Fluid Phospholipid<br />

Bilayers Over a Wide Range <strong>of</strong> Bilayer Thickness. J. Mol. Biol. 166:203–210.<br />

Liu, Y., Casey, L., and Pike, L. J. (1998). Compartmentalization <strong>of</strong> Phosphatidylinositol 4,5-<br />

Bisphosphate in Low-Density Membrane Domains in the Absence <strong>of</strong> Caveolin. Biochem. Biophys. Res.<br />

Commun. 245: 684-690.<br />

197


Lodish, H., Berk, A., Zipursky, S. L., Matsudaira, P., Baltimore, D., and Darnell, J. E. (2000). Molecular<br />

Cell Biology. 4 th Ed. W. H. Freeman, New York, USA<br />

Mall, S., Broadbridge, R., Sharma, R. P., Lee, A. G., and East J. M. (2000). Effects <strong>of</strong> Aromatic Residues<br />

at the Ends <strong>of</strong> Trans<strong>membrane</strong> Alpha-Helices on Helix <strong>Interaction</strong>s With Bilayers. 39: 2071-2078.<br />

Mall, S., East, J. M., and Lee, A. G. (2002). Trans<strong>membrane</strong> α-Helices. in Current Topics in Membranes<br />

52: 339-371<br />

Marsh, D. (1990). Lipid-Protein <strong>Interaction</strong>s in Membranes. FEBS Lett. 268: 271-375.<br />

Marsh, D. (1996). Peptide Models for Membrane Channels. Biochem, J. 315: 345-361.<br />

Marsh, D., Horváth, L. I., Swamy, M. J., Mantripragada, S., and Kleinschmidt, J. H. (2002). <strong>Interaction</strong>s<br />

<strong>of</strong> Membrane-Spanning Proteins With Peripheral and Lipid-Anchored Membrane Proteins: Perspectives<br />

From Protein-Lipid <strong>Interaction</strong>s. Mol. Memb. Biol. 19: 247-255.<br />

Martin, T. F. J. (2001). PI(4,5)P 2 Regulation <strong>of</strong> Surface Membrane Traffic. Curr. Opin. Cell Biol. 13: 493-<br />

499.<br />

Mascaretti, O. (2003). Bacteria Versus Antibacterial Agents: An Integrated Approach. ASM Press,<br />

Washington D. C.<br />

Masuda, M., Takeda, S., Sone, M., Ohki, T., Mori, H., Kamioka, Y., and Mochizuki, N. (2006).<br />

Endophilin BAR Domain Drives Membrane Curvature By Two Newly Identified Structure-Based<br />

Mechanisms. EMBO J. 25: 2889-2897.<br />

Mateo, C. R., de Almeida, R. F. M., Loura, L. M. S., and Prieto, M. (2006). From Lipid Phases to<br />

Membrane Protein Organization: Fluorescence Methodologies in the Study <strong>of</strong> Lipid-Protein <strong>Interaction</strong>s.<br />

in “Protein-Lipid <strong>Interaction</strong>s: New Approaches and Emerging Concepts.” (Eds. Mateo, C. R., Gómez,<br />

J., Villalaín, J., Ros, J. M. G.) Springer-Verlag. Berlin, Germany.<br />

McIntosh, T. J., Vidal, A., and Simon, S. A. (2002). The Energetics <strong>of</strong> Peptide-Lipid <strong>Interaction</strong>s:<br />

Modulation by Interfacial Dipoles and Cholesterol. in Current Topics in Membranes 52: 205-253.<br />

McLaughlin, S., Wang, J., Gambhir, A., and Murray, D. (2002). PIP 2 and Proteins: <strong>Interaction</strong>s,<br />

Organization, and Information Flow. Annu. Rev. Biophys. Struct. 31: 151-175.<br />

McMahon, H. T., and Gallop, J. L. (2005). Membrane Curvature and Mechanisms <strong>of</strong> Dynamic Cell<br />

Membrane Remodeling. Nature 438: 590-596.<br />

Meijer, A. B., Spruijt, R. B., Wolfs, C. J. A. M., and Hemminga, M. A. (2001). Membrane-Anchoring<br />

<strong>Interaction</strong>s <strong>of</strong> M13 Major Coat Protein. Biochemistry 40: 8815-8820.<br />

198


BIBLIOGRAPHY<br />

Mishra, V. K., Palgunachari, M. N., Segrest, J. P., and Anantharamaiah, G. M. (1994). <strong>Interaction</strong>s <strong>of</strong><br />

Synthetic Peptide Analogs <strong>of</strong> the Class A Amphipathic Helix With Lipids. Evidence for the Snorkel<br />

Hypothesis. J. Biol. Chem. 269: 7185-7191.<br />

Monné, M., Nilsson, I., Johansson, M., Elmhed, N., and von Heijne, G. (1998). Positively and Negatively<br />

Charged Residues Have Different Effects on the Position in the Membrane <strong>of</strong> a Model Trans<strong>membrane</strong><br />

Helix. J. Mol. Biol. 284: 1177-1183.<br />

Morein, S., Killian, J. A., and Sperotto M. (2002). Characterization <strong>of</strong> the Thermotropic Behaviour and<br />

Lateral Organization <strong>of</strong> Lipid-Peptide Mixtures by a Combined Experimental and Theoretical Approach:<br />

Effects <strong>of</strong> Hydrophobic Mismatch and Role <strong>of</strong> Flanking Residues. Biophys. J. 82: 1405-1417.<br />

Mouritsen, O. G., and Bloom, M. (1984). Mattress Model <strong>of</strong> Lipid-Protein <strong>Interaction</strong>s in Membranes.<br />

Biophys. J. 46: 141-153.<br />

Mouritsen, O. G., and Bloom, M. (1993). Models <strong>of</strong> Lipid-Protein <strong>Interaction</strong>s in Membranes. Annu.<br />

Rev. Biophys. Bioeng. 22: 145-171.<br />

Mouritsen, O.G. (2005). Life – As a Matter <strong>of</strong> Fat. The Emerging Science <strong>of</strong> Lipidomics. Springer,<br />

Heildelberg, Germany.<br />

Nadler, G., Morvan, M., Delimoge, I., Belfiore, P., Zocchetti, A., James, I., Zembryki, D., Lee-<br />

Rycakzewski, E., Parini, C., Consolandi, E., Gagliardi, S., and Farina, C. (1998). (2Z,4E)-5-(5,6-dichloro-<br />

2-indolyl)-2-methoxy-n-(1,2,2,6,6-pentamethylpiperidin-4-yl)-2,4-pentadienamide, a Novel, Potent and<br />

Selective Inhibitor <strong>of</strong> the Osteoclast V-ATPase. Bioorg. Medic. Chem. Lett. 8: 3621-3626.<br />

Nagle, J. F., and Tritram-Nagle, S. (2000). Structure <strong>of</strong> Lipid Bilayers. Biochim. Biophys. Acta 1469:<br />

159-195.<br />

Neumann, D., Barchan D., Fridkint, M., and Fuchs, S. (1986). Analysis <strong>of</strong> Ligand Binding to the Synthetic<br />

Dodecapeptide <strong>of</strong> the Acetylcholine Receptor a Subunit. Proc. Natl. Acad. Sci. USA 83: 9250-9253.<br />

Neves, P., Berkane, E., Gameiro, P., Winterhalter, M., and de Castro, B. (2005). <strong>Interaction</strong> Between<br />

Quinolones Antibiotics and Bacterial Outer Membrane Porin OmpF. Biophys. Chem. 113: 123-128.<br />

Nielsen, C., Goulian, M., and Andersen, O. S. (1998). Energetics <strong>of</strong> Inclusion-Induced Bilayer<br />

Deformations. Biophys. J. 74: 1966-1983.<br />

Nieva, J. L., Goni, F. M., and Alonso, A. (1993). Phospholipase C-Promoted Membrane Fusion.<br />

Retroinhibition by the End-Product Diacylglycerol. Biochemistry 32: 1054-1058.<br />

Nikaido, H., Kim, S. H., and Rosenberg, E. Y. (1993). Physical Organization <strong>of</strong> Lipids in the Cell Wall <strong>of</strong><br />

Mycobacterium Chelonae. Mol. Microbiol. 8: 1025-1030.<br />

199


Nikura, K., Takeshita, N., and Takano, M. (2005). A Vacuolar ATPase inhibitor, FR167356 Prevents<br />

Bone Resorption in Ovariectomized Rats With High Potency and Specificity: Potential for Clinical<br />

Application. J. Bone Miner. Res. 20: 1579-1588.<br />

Nishi, T., and Forgac, M. (2002). The Vacuolar (H + ) - ATPases – Nature’s Most Versatile Proton Pumps.<br />

Mol. Cell. Biol. 3: 94-103.<br />

O´Keeffe, A. H., East, J. M., and Lee, A. G. (2000). Selectivity in Lipid Binding to the Bacterial Outer<br />

Membrane Protein OmpF. Biophys. J. 79: 2066-2074.<br />

Páli, T., Whyteside, G., Dixon, N., Kee, T. P., Ball, S., Harrison, M. A., Findlay, J. B. C., Finbow, M. E.,<br />

and Marsh, D. (2004). <strong>Interaction</strong> <strong>of</strong> Inhibitors <strong>of</strong> the Vacuolar H + -ATPase With the Trans<strong>membrane</strong> V o -<br />

sector. Biochemistry. 43: 12297-12305.<br />

Pap, E. H. W., van Hoek, A., Wirtz, K. W. A., and Visser, A. J: W. G. (1995). Quantitative Analysis <strong>of</strong><br />

Lipid-Lipid and Lipid-Protein <strong>Interaction</strong>s in Membranes by Use <strong>of</strong> Pyrene-Labeled Phosphoinositides.<br />

Biochemistry 34: 9118-9125.<br />

Papayannopoulos, V., Co, C., Prehoda, K. E., Snapper, S., Taunton, J., and Lim, W. A. (2005). A<br />

Polybasic Motif Allows N-WASP to Act as a Sensor <strong>of</strong> PIP 2 Density. Moll. Cell 17: 181-191.<br />

Peelen, S. J. C. J., Sanders, J. C., Hemminga, M. A. and Marsh, D. (1992). Stoichiometry, Selectivity and<br />

Exchange Dynamics <strong>of</strong> Lipid-Protein <strong>Interaction</strong> <strong>with</strong> Bacteriophage M13 Coat Protein Studied by Spin<br />

Label Electron Spin Resonance. Effects <strong>of</strong> Protein Secondary Structure. Biochemistry 31: 2670-2677.<br />

Pérez-Gil, J., Cruz, A., and de la Serra, J. B. (2005). Lateral Membrane Structure and Lipid-Protein<br />

<strong>Interaction</strong>s. in “Protein-Lipid <strong>Interaction</strong>s: New Approaches and Emerging Concepts.” (Eds. Mateo, C.<br />

R., Gómez, J., Villalaín, J., Ros, J. M. G.) Springer-Verlag. Berlin, Germany.<br />

Pestova, E., Millichap, J. J., Noskin, G. A., and Peterson, L. R. (2000). Intracellular Targets <strong>of</strong><br />

Moxifloxacin: A Comparison With Other Fluoroquinolones. J. Antimicrob. Chemother. 45: 583-590.<br />

Peter, B. J., Kent, H. M., Mills, I.G., Vallis, Y., Butler, P. J. G., Evans, P. R., and McMahon, H. T.<br />

(2004). BAR Domains as Sensors <strong>of</strong> Membrane Curvature: the Amphiphysin BAR Structure. Science 303:<br />

495-499.<br />

Petrangoline, G., Supino, R., Pratesi, G., Bo, L. D., Tortoreto, M., Croce, A. C., Misiano, P., Belfiore, P.,<br />

Farina, C., and Zunino, F. (2006). Effect <strong>of</strong> a Novel Vacuolar-H + -ATPase Inhibitor on Cell and Tumor<br />

Response to Camptothecins. Chemot. Antib. Gene Ther. 318: 939-946.<br />

Pike, L. J., and Miller, J. M. (1998). Cholesterol Depletion Delocalizes Phosphatidylinositol<br />

Bisphosphate and Inhibits Hormone-Stimulated Phosphatidylinositol Turnover. J. Biol. Chem. 273:<br />

22298-22304.<br />

200


BIBLIOGRAPHY<br />

Pluschke, G., Hirota, Y, and Overath, P. (1978). Function <strong>of</strong> Phospholipids in Escherichia Coli.<br />

Characterization <strong>of</strong> a Mutant Deficient in Cardiolipin Synthesis. J. Biol. Chem. 253: 5048-5055.<br />

Popot, J.-L., and Engelman, D. M. (1990). Membrane Protein Folding and Oligomerization: The Ttwo-<br />

Stage Model. Biochemistry 29: 4031-4037.<br />

Popot, J. L. (1993). Integral Membrane Protein Structure: Trans<strong>membrane</strong> Alpha-Helices as<br />

Autonomous Folding Domains.Curr. Opin. Struct. Biol. 3: 532-540.<br />

Powl, A. M., East, J. M., and Lee, A. G. (2003). Lipid-Protein <strong>Interaction</strong>s Studied by Introduction <strong>of</strong> a<br />

Tryptophan Residue: The Mechanosensitive Channel MscL. Biochemistry 42: 14306-14317.<br />

Quintiliani, R., Owens, R. C., and Grant, E. M. (1999). Clinical Role <strong>of</strong> Fluoroquinolones in Patients<br />

With Respiratory Tract Infections. Infect. Dis. Clin. Pract. 8: 28-41.<br />

Rauch, M. E., Ferguson, C. G., Prestwich, G. D., and Cafiso, D. S. (2002). Myristoylated Alanine-Rich C<br />

Kinase Substrate (MARCKS) Sequesters Spin-labeled Phosphatidylinositol 4,5-Bisphosphate in Lipid<br />

Bilayers. J. Biol. Chem. 277: 14068-14076.<br />

Redfern, D. A., and Gericke, A. (2004). Domain Formation in Phosphatidylinositol<br />

Monophosphate/Phosphatidylcholine Mixed Vesicles. Biophys. J. 86: 2980-2992.<br />

Redfern, D.A., and Gericke, A. (2005). pH-Dependent Domain Formation in Phosphatidylinositol<br />

Polyphosphate/Phosphatidylcholine Mixed Vesicles. J. Lipid Res. 46: 504-515.<br />

Ren, J., Lew, S., Wang, J., and London, E. (1997). Trans<strong>membrane</strong> Orientation <strong>of</strong> Hydrophobic α-<br />

Helices is Regulated Both by the Relationship <strong>of</strong> Helix Length to Bilayer Thickness and by the<br />

Cholesterol Concentration. Biochemistry 36: 10213-10220.<br />

Ren, J., Lew, S., Wang, J., and London, E. (1999). Control <strong>of</strong> the Trans<strong>membrane</strong> Orientation and<br />

Interhelical <strong>Interaction</strong>s Within Membranes by Hydrophobic Length. Biochemistry 38: 5905-5912<br />

Richnau, N., Fransson, A., Farsad, K., and Aspenstrom, P. (2004). RICH-1 has a BIN/Amphiphysin/Rvsp<br />

Domain Responsible for Binding to Membrane Lipids and Tubulation <strong>of</strong> Liposomes. Biochem. Biophys.<br />

Res. Commun. 320: 1034-1042.<br />

Ried, C., Wahl, C., Miethke, T., Wellnh<strong>of</strong>er, G., Landgraf, C., Schneider-Mergener, J., and Hoess, A.<br />

(1996). High Affinity Endotoxin-Binding and Neutralizing Peptides Based on the Crystal Structure <strong>of</strong><br />

Recombinant Limulus Anti-Lipopolysaccharide Factor. J. Biol. Chem. 271: 28120-28127.<br />

Rigaud, J.-L., Paternoste, M.-T., and Pitard, B. (1995). Reconstitution <strong>of</strong> Membrane Proteins Into<br />

Liposomes: Application to Energy Transducing Membrane Proteins. Biochim. Biophys. Acta. 1231: 223-<br />

246.<br />

201


Rodan, G. A. and Martin, T. J. (2000). Therapeutic Approaches to Bone Diseases. 289: 1508-1414.<br />

Rodriguez, N., Pincet, F., and Cribier, S. (2005) Giant Vesicles Formed by Gentle Hydration and<br />

Electr<strong>of</strong>ormation: A Comparison by Fluorescence Microscopy. Coll. Surf. B: Biointerfaces 42: 125-130.<br />

Russel, M. (1991). Filamentous Phage Assembly. Mol. Microbiol. 5: 1607-1613.<br />

Sackmann, E. (1995). Biological Membranes Architecture and Function in Structure and Dynamics <strong>of</strong><br />

Membranes. From Cells to Vesicles. Eds Lipowsky, R., and Sackmann, E. North Holland, Amsterdam,<br />

The Netherlands.<br />

Sanders, J. C., Ottaviani, M. F., van Hoek, A., Visser, A. J. W. G., and Hemminga, M. A. (1992). A Small<br />

Protein in Model Membranes: A Time Resolved Fluorescence and ESR Study on the <strong>Interaction</strong> <strong>of</strong> M13<br />

Coat Protein With Lipid Bilayers. Eur. Biophys. J. 21: 305-311.<br />

Santos, N. C., Prieto, M., and Castanho, M. A. R. B. (2003). Quantifying Molecular Partition into Model<br />

Systems <strong>of</strong> Bio<strong>membrane</strong>s: an Emphasis on Optical Spectroscopic Methods. Biophys. Biochim. Acta<br />

1612: 123-135.<br />

Scatigno, A. C., Garrido, S. S., and Marchetto, R. (2004). A 4.2 kDa Synthetic Peptide as a Potential<br />

Probe to Evaluate the Antibacterial Activity <strong>of</strong> Coumarin Drugs. J. Peptide Sci. 10: 566-577.<br />

Schiffer, M., and Edmundson, A. B. (1967). Use <strong>of</strong> Helical Wheels to Represent the Structures <strong>of</strong><br />

Proteins and to Identify Segments <strong>with</strong> Helical Potential. Biophys. J. 7: 121-135.<br />

Seddon, J. M., and Templer, R. H. (1995). Polymorphism <strong>of</strong> Lipid-Water Systems in Structure and<br />

Dynamics <strong>of</strong> Membranes. From Cells to Vesicles. Eds Lipowsky, R., and Sackmann, E. North Holland,<br />

Amsterdam, The Netherlands.<br />

Seddon, A. M., Curnow, P., and Booth, P. J. (2004). Membrane Proteins, Lipids and Detergents: Not Just<br />

a Soap Opera. Biochim. Biophys. Acta 1666: 105-117.<br />

Seelig, J. (2004). Thermodynamics <strong>of</strong> Lipid-Peptide <strong>Interaction</strong>s. Biochim. Biophys. Acta 1666: 40-50.<br />

Shimshick, E. J., and McConnell, H. M. (1973). Lateral Phase Separation in Phospholipid Membranes.<br />

Biochemistry. 12:2351–2360.<br />

Silvius, J. R. (1990). Calcium-induced Lipid Phase Separations and <strong>Interaction</strong>s <strong>of</strong><br />

Phosphatidylcholine/Anionic Phospholipid Vesicles. Fluorescence Studies Using Carbazole-Labeled and<br />

Brominated Phospholipids. Biochemistry. 24: 2930-2938.<br />

Simons, K., and Ikonen, E. (1997). Functional Rafts in Cell Membranes. Nature 387:569-572.<br />

Singer, S. J., and Nicolson, G. L. (1972). The Fluid Mosaic Model <strong>of</strong> the Structure <strong>of</strong> Cell Membranes.<br />

Science 175: 720-731.<br />

202


BIBLIOGRAPHY<br />

Sperotto, M. M., and Mouritsen, O. G.(1991). Monte Carlo Simulation Studies <strong>of</strong> Lipid Order Parameter<br />

Pr<strong>of</strong>iles Near Integral Membrane Proteins. Biochemistry 59: 261-270.<br />

Spruijt, R., and Hemminga, M. A. (1991). The in Situ Aggregational and Conformational State <strong>of</strong> the<br />

Major Coat Protein <strong>of</strong> Bacteriophage M13 in Phospholipid Bilayers Mimicking the Inner Membrane <strong>of</strong><br />

Host Escherichia Coli. Biochemistry 30: 11147-11154.<br />

Spruijt, R., Wolfs, C., and Hemminga, M. (1999). M13 Phage. in “Encyclopedia <strong>of</strong> Molecular Biology.”<br />

(Ed. Creighton, T. E.), Wiley-Interscience, New York, U.S.A.<br />

Spuijt, R. B., Wolfs, C. J. A. M., Verveer, J. W. G., and Hemminga, M. (1996). Putative Trans<strong>membrane</strong><br />

Domain <strong>of</strong> the Major Coat Protein <strong>of</strong> Bacteriophage M13. Biochemistry 35: 10383-10391.<br />

Stahlmannn, R., and Lode, H. (1999). Toxicity <strong>of</strong> Quinolones. Drugs 58: 37-42.<br />

Stopar, D., Jansen, K. A. J., Páli, T., Marsh, D., and Hemminga, M. A. (1997a). Membrane Location <strong>of</strong><br />

Spin-Labeled M13 Major Coat Proteins Determined by Paramagnetic Relaxation Agents. Biochemistry<br />

36: 8261-8268.<br />

Stopar, D., Spruijt, R., Wolfs, C. J. A. M., and Hemminga, M. A. (1997b). In Situ Aggregational State <strong>of</strong><br />

M13 Bacteriophage Major Coat Protein in Sodium Cholate and Lipid Bilayers. Biochemistry 36: 12268-<br />

12275.<br />

Stopar, D., Spruijt, R., Wolfs, C. J. A. M., and Hemminga, M. A. (1998). Mimicking Initial <strong>Interaction</strong>s<br />

<strong>of</strong> Bacteriophage M13 Coat Protein Disassembly in Model Membrane Systems. Biochemistry 37: 10181-<br />

10187.<br />

Stopar, D., Spruijt, R., Wolfs, C. J. A. M., and Hemminga, M. A. (2003). Protein-Lipid <strong>Interaction</strong>s <strong>of</strong><br />

Bacteriophage M13 Major Coat Protein. Biochim. Biophys. Acta 78451: 1-11.<br />

Takei, K., Haucke, V., Slepnev, V., Farsad, K., Salazar, M., Chen, H. and De Camilli, P. (1998).<br />

Generation <strong>of</strong> Coated Intermediates <strong>of</strong> Clathrin-Mediated Endocytosis on Protein-Free Liposomes. Cell<br />

94: 131-141.<br />

Takei, K., Slepnev, V. I., Haucke, V., and De Camilli, P. (1999). Functional Partnership Between<br />

Amphiphysin and Dynamin in Clathrin-Mediated Endocytosis. Nat. Cell Biol. 1:33-39.<br />

Tanford, C. (1980). The Hydrophobic Effect. Wiley, New York, USA.<br />

Taylor, J. W. (1990). Peptide Models <strong>of</strong> dynorphin A(1-17) Incorporating Minimally Homologous<br />

Substitutes for the Potential Amphiphilic β-Strand in Residues 7-15. Biochemistry 29: 5364-5373<br />

Taylor, A. H., Heavner, G., Nedelman, M., Sherris, D., Brunt, E., Knight, D., and Ghrayeb, J. (1995).<br />

Lipopolysaccharide (LPS) Neutralizing Peptides Reveal a Lipid A Binding Site <strong>of</strong> LPS Binding Protein.<br />

203


J. Biol. Chem. 270: 17934-17938.<br />

Teitelbaum, S. L. (2000). Bone Resorption by Osteoclasts. Science 289: 1504-1508.<br />

van Rheenen, J., Achame, M. E., Janssen, H., Calafat, J., and Jalink, K. (2005). PIP 2 Signaling in Lipid<br />

Domains: A Critical Re-evaluation. EMBO J. 24: 1664-1673.<br />

Väänänen, K. (2005). Mechanism <strong>of</strong> osteoclast mediated bone resorption—rationale for the design <strong>of</strong> new<br />

therapeutics. Adv. Drug Deliv. Rev. 57: 959-971<br />

Vázquez, J. L., Merino, S., Domenech, Ò, Berlanga, M., Viñas, M., Montero, M. T., and Hernández-<br />

Borrell, J. (2001). Determination <strong>of</strong> the Partition Coefficients <strong>of</strong> a Homologous Series <strong>of</strong> Cipr<strong>of</strong>loxacin:<br />

Influence <strong>of</strong> the N-4 Piperazinyl Alkylation on the Antimicrobial Activity. Int. J. Pharm. 220: 53-62.<br />

Visentin, L., Dodds, L. A., Valente, M., Misiano, P., Bradbeer, J. N., Oneta, S., Liang, X., Gowen, M.,<br />

and Farina, C. (2000). A Selective Inhibitor <strong>of</strong> the Osteoclastic V-H + -ATPase Prevents Bone Loss in Both<br />

Thyroparathyroidectomized and Ovariectomized Rats. J. Clin. Invest. 103: 309-318.<br />

Wallin, E., Tsukihara, T., Yoshikawa, S., von Heijne, G., and El<strong>of</strong>sson, A. (1997). Architecture <strong>of</strong> Helix<br />

Bundle Membrane Proteins. An Analysis <strong>of</strong> Cytochrome c Oxidase from Bovine Mitochondria. Protein<br />

Sci. 6: 808-815.<br />

Wang, J., Arbuzova, A., Hangyás-Milhályné, G., and McLaughlin, S. (2001). The Effector Domain <strong>of</strong><br />

Myristoylated Alanine-rich C Kinase Substrate Binds Strongly to Phosphatidylinositol 4,5-Bisphosphate.<br />

J. Biol. Chem. 276: 5012-5019.<br />

Webb, M. S., Hui, S. W., and Steponkus, P. L. (1993). Dehydration Induced Lamellar-to-Hexagonal-II<br />

Phase Transitions in DOPE/DOPC Mixtures. Biochim. Biophys. Acta. 1145:93-104.<br />

Weiss, T. M., van der Wel, P. C. A., Killian, J. A., Koeppe II, R. E., and Huang, H. W. (2003).<br />

Hydrophobic Mismatch between Helices and Lipid Bilayers. Biophys. J. 84: 379-385.<br />

White, S. H., and Wimley, W. C. (1999). Membrane Protein Folding and Stability: Physical Principles.<br />

Annu. Rev. Biophys. Biomol. Struct. 28: 319-365.<br />

Whyteside, G., Meek, P. J., Ball, S. K., Dixon, N., Finbow, M. E., Kee, T. P., Findlay, J. B. C., and<br />

Harrison, M .A. (2005). Concanamycin and Indolyl Pentadieneamide Inhibitors <strong>of</strong> the Vacuolar H+-<br />

ATPase Bind <strong>with</strong> High Affinity to the Purified Proteolipid Subunit <strong>of</strong> the Membrane Domain.<br />

Biochemistry. 44: 15024-15031.<br />

Wiener, M. C., and White, S. H. (1992). Structure <strong>of</strong> a Fluid Dioleoylphosphatidylcholine Bilayer<br />

Determined by Joint Refinement <strong>of</strong> X-Ray and Neutron Diffraction Data. III. Complete Structure.<br />

Biophys. J. 61:434-447.<br />

204


BIBLIOGRAPHY<br />

Williamson, I. M., Alvis, S. J., East, J. M., and Lee, A. G. (2002). <strong>Interaction</strong> <strong>of</strong> Phospholipids <strong>with</strong> the<br />

Potassium Channel KscA. Biophys. J. 83: 2026-2038.<br />

Wilson, P. T., Lentz, T.L., and Hawrot, E. (1985). Determination <strong>of</strong> the Primary Amino Acid Sequence<br />

Specifying the Alpha-Bungarotoxin Binding Site on the Alpha-Subunit <strong>of</strong> the Acetylcholine Receptor<br />

From Torpedo Californica. Proc. Natl. Acad. Sci. USA 82: 8790-8794.<br />

Wimley, W. C., and White, S. H. (1996). Experimentally Determined Hydrophobicity Scale for Proteins<br />

at Membrane Interfaces. Nat. Struct. Biol. 3: 842-848.<br />

Wimley, W.C., Creamer, T.P., and White, S.H. (1996) Solvation Energies <strong>of</strong> Amino Acid Residues and<br />

Backbone in a Family <strong>of</strong> Host Guest Penta<strong>peptides</strong>. Biochemistry 35: 5109-5124.<br />

Yau, W.-M., Wimley, W. C., Gawrisch, K., and White, S. H. (1998). The Preference <strong>of</strong> Tryptophan for<br />

Membrane Interfaces. Biochemistry 37: 14713-14718.<br />

Yeagle, P. L. (1993). The Membranes <strong>of</strong> Cells. 2 nd Ed. Academic Press, San Diego, USA.<br />

Yeagle, P. L., Bennet, M., Lemaître, V., and Watts, A. (2007). Trans<strong>membrane</strong> Helices <strong>of</strong> Membrane<br />

Proteins May Flex to Satisfy Hydrophobic Mismatch. Biochim. Biophys. Acta 1768: 530-537.<br />

Yokoyama, K., Nakano, M., Imamura, H., Yoshida, M., and Tamakoshi, M. (2003). Rotation <strong>of</strong> the<br />

Proteolipid Ring in the V-ATPase. J. Biol. Chem. 278: 24255-24258.<br />

Yoshida, Y., Kinuta, M., Abe, T., Liang, S., Araki, K., Cremona, O., Di Paolo, G., Moriyama, Y.,<br />

Yasuda, T., De Camilli, P., and Takei, K. (2004). The Stimulatory Action <strong>of</strong> Amphiphysin on Dynamin<br />

Function is Dependent on Lipid Bilayer Curvature. EMBO J. 23: 3483-3491.<br />

Zimmerberg, J. (2000). Are the Curves in All the Right Places Traffic 1: 366-368.<br />

Zimmerberg, J., and Kozlov, M. M. (2006). How Proteins Produce Cellular Membrane Curvature. Nat.<br />

Rev. Mol. Cell Biol. 7: 9-19.<br />

205

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!