repositório da universidade de lisboa: página principal -...

99
UNIVERSIDADE DE LISBOA FACULDADE DE CIÊNCIAS DEPARTAMENTO DE BIOLOGIA ANIMAL Microglia demonstrate age-dependent performance and interaction with beta-amyloid peptide Gonçalo Manuel da Costa Lidónio Dissertação Mestrado em Biologia Humana e Ambiente 2013

Upload: others

Post on 25-Nov-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

UNIVERSIDADE DE LISBOA

FACULDADE DE CIÊNCIAS

DEPARTAMENTO DE BIOLOGIA ANIMAL

Microglia demonstrate age-dependent performance and

interaction with beta-amyloid peptide

Gonçalo Manuel da Costa Lidónio

Dissertação

Mestrado em Biologia Humana e Ambiente

2013

Page 2: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é
Page 3: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

UNIVERSIDADE DE LISBOA

FACULDADE DE CIÊNCIAS

DEPARTAMENTO DE BIOLOGIA ANIMAL

Microglia demonstrate age-dependent performance and

interaction with beta-amyloid peptide

Gonçalo Manuel da Costa Lidónio

Dissertação

Mestrado em Biologia Humana e Ambiente

Dissertação orientada por:

Professora Doutora Dora Brites (orientação externa)

iMed.UL – Faculdade de Farmácia da Universidade de Lisboa

Professora Doutora Deodália Dias (orientação interna)

Faculdade de Ciências da Universidade de Lisboa

2013

Page 4: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é
Page 5: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é
Page 6: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é
Page 7: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

I

NOTA PRÉVIA

A realização desta dissertação também contou com a co-orientação da Doutora

Ana Sofia Falcão, investigadora auxiliar da Faculdade de Farmácia da

Universidade de Lisboa.

Parte dos resultados inseridos nesta dissertação foram apresentados nos

seguintes encontros:

Lidónio G, Caldeira C, Vaz AR, Santos M, Moreira R, Falcão AS, Brites D. Exploring a

therapeutic approach to prevent Ab-induced dysfunctional microglia. 5th Postgraduate

iMed.UL Students Meeting, Lisboa, Portugal, 18 Julho, 2013 [Poster]

Caldeira C, Frederico A, Oliveira AF, Lidónio G, Vaz A, Fernandes A, Brites D. Cell

ageing effects on microglia response to Aβ.XIII Reunião da Sociedade Portuguesa de

Neurociências, Luso, Portugal, 31 de Maio - 2 de Junho, 2013 [Poster]

A formatação das referências usada nesta dissertação baseou-se no utilizado

pela revista Neurochemical Research da Springer, conceituada na área e na qual o

grupo de investigação tem vindo a publicar.

Page 8: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

II

[Escre

va um

Page 9: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

III

AGRADECIMENTOS

Em primeiro lugar, naturalmente, quero agradecer à Prof. Doutora Dora Brites,

orientadora externa desta dissertação. Fico muito grato pela oportunidade de realizar a

minha dissertação de mestrado no seu grupo de investigação, Neuron Glia Biology in

Health and Disease, e por todos os conhecimentos que me transmitiu ao longo deste

ano. O seu rigor científico, espírito crítico, empenho e inteligência são admiráveis, e

uma fonte de inspiração.

À Doutra Sofia Falcão, co-orientadora deste trabalho, agradeço todo o apoio,

dedicação, disponibilidade e preocupação revelados ao longo deste projecto. Sem ti,

Sofia, este trabalho não seria possível. Agradeço, profundamente, toda a paciência e

compreensão que tiveste comigo e toda a motivação que me deste. Contigo, muito

aprendi, tanto a nível profissional como a nível pessoal.

À Prof. Doutora Deodália Dias, orientadora interna desta dissertação e

coordenadora do mestrado em Biologia Humana e Ambiente, agradeço por toda a

disponibilidade e pelo apoio e conselhos que me deu, não só ao longo deste ano, mas

durante toda a minha estadia neste mestrado.

Ao Prof. Doutor Rui Silva e à Prof. Doutora Alexandra Brito agradeço toda a

disponibilidade para esclarecer dúvidas que foram surgindo.

Quero também agradecer à Prof. Doutora Adelaide Fernandes e à Doutora Ana

Rita Vaz toda a disponibilidade que sempre demonstraram, o conhecimento que

partilharam comigo e, principalmente, a ajuda no ultrapassar de muitos dos desafios

que encontrei.

Quero também expressar o meu agradecimento para com a Doutora Teresa

Pais por gentilmente nos ter cedido a linha microglial N9, e ao Prof. Doutor Rui Moreira

por nos ter cedido o composto explorado neste trabalho.

Como não poderia deixar de ser, um grande agradecimento para todos os

restantes membros do grupo:

Inês Palmela, ao contrário do que contam, quero salientar a boa disposição que

sempre observei e agradecer toda a ajuda que me deste – mesmo pensando que

adormeci na tua aula. À Filipa agradeço os bons conselhos, o apoio e, claro, os

Page 10: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

IV

conhecimentos para melhorar as minhas imunocitoquímicas. À Andreia Barateiro

agradeço toda a disponibilidade e a sua contribuição ao alegrar os nossos dias com os

seus bolinhos. Em ti, Cláudia, vejo um exemplo. Acho fascinante a tua força de

vontade. Agradeço-te todo o apoio, ajuda e maturidade que partilhaste comigo neste

trabalho que tão próximo está do teu. Cátia e Carolina, apesar das dores de cabeça

que vocês me deram, tenho muito que vos agradecer. Transmitiram-me

conhecimentos preciosos e, passo a passo, contribuíram para que me tornasse um

melhor cientista. A ti, Gisela, agradeço o apoio, a disponibilidade e todas as sugestões

que me deste em muitos dos problemas que encontrei. Agradeço ainda a simpatia, a

motivação que sempre me transmitiste e o ouvido para os meus desabafos nas

situações mais complicadas. E não me esqueço de ti, Inês Figueira, que mesmo que

não tendo os dois pezinhos no grupo, também contribuíste neste meu percurso.

Obrigado.

Um agradecimento também aos mais novos, aos que seguiram ao meu lado

nesta viagem e que mais força me deram para chegar ao destino. À Vera, um enorme

obrigado. Principalmente por todo o teu positivismo, por partilhares comigo os teus

conhecimentos laboratoriais e, acima de tudo, por todos os momentos de convívio que

contribuíram para tornar esta viagem menos difícil. A vocês as duas, Marta e Andreia,

temo que se deixasse passar tudo o que sinto para o papel, a minha dissertação seria

um agradecimento. Desde os dias de faculdade que seguem ao meu lado e que

preenchem as minhas boas memórias. Agradeço-vos por toda a amizade, por me

puxarem sempre por um sorriso e por todos os momentos de drama, comédia e Sci-Fi

com que comigo partilharam. Sem vocês, esta viagem não seria a mesma. Não posso

passar também sem agradecer ao telemóvel da Marta, que muitos desabafos partilhou

comigo e muito me animou nos dias de chuva.

Agradeço ao grupo, como um todo, pela forma como me receberam. Fizeram-

me sentir em casa.

Quero agradecer também aos constituintes do meu grupo de mestrado. Em

especial à Marília e ao Gonçalo, pelo seu companheirismo e amizade, que me

encheram as horas de risos e desabafos.

Um obrigado muito especial aos meus amigos da terrinha, que estão comigo

desde que lembro e me têm acompanhado nos momentos mais importantes da minha

vida. Em especial a vocês, Nuno, Miguel e Sandra. De perto ou de longe, tornaram

este trabalho não apenas meu, mas também vosso.

Page 11: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

V

Por último, mas acima de tudo, quero agradecer à minha família. Aos que cá

estão para presenciar esta etapa e aos que gostaria que cá estivessem. Sem vocês,

não seria o que sou hoje. Agradeço principalmente aos meus pais e à minha irmã, pelo

esforço que têm feito para chegar onde cheguei e por nunca desistirem de mim. Não

poderiam ser melhores. Vocês estão e estarão sempre no meu coração.

Page 12: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

VI

[Escre

va um

Page 13: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

VII

ABSTRACT

Alzheimer’s disease (AD) is the most common neurodegenerative disorder in

the elderly and remains without a cure. One of its prominent features is the appearance

in the brain of amyloid-beta (Aβ) deposits forming amyloid plaques. Microglia, which

are the first line of defense in the brain, are activated by Aβ assuming a phagocytic

phenotype followed by the inflammatory state that is thought to play a role in the

progression of this disease. However, the important role of microglia in Aβ clearance

and neuronal support seems to be decreased with age, through an ensemble of

alterations still not clarified that render cells to function abnormally. Indeed, in these

circumstances, microglia lose responsiveness to their key regulating factors, thus

compromising their neuroprotective or inflammation solving properties. Therefore, in

this work we aimed to: a) understand how ageing influences microglia reactivity to Aβ

and microglia-neuron cross-talk dynamics, using an in vitro ageing model of primary

microglia (isolated or in mixed microglia-neurons cultures) and by evaluating the

release of glutamate, adenosine triphosphate (ATP) and matrix metalloproteinases

(MMPs); b) explore the therapeutical properties of a vinyl sulfone-based compound

(VS) as a modulator of Aβ-induced microglial activation, using the N9 cell line and

assessing cell viability, inflammation-related factors, activation receptors and

phagocytosis.

Our results demonstrate that microglia behave differently accordingly with age,

releasing less glutamate and MMP-9, but more MMP-2. Ageing also seems to render

microglia more irresponsive to neuronal-mediated regulation, particularly regarding

their ability to modulate the extracellular glutamate content. In addition, in our

experimental model of ageing in vitro, Aβ failed to induce any microglia reactivity. In

regard to VS testing, no toxicity was observed for this compound, and it successfully

diminished the Aβ-induced release of MMP-9 and MMP-2, as well as the expression of

high-mobility group box 1 protein, important inflammatory mediators. Moreover, Aβ

reduced the level of microglial toll-like receptor 4 and phagocytosis, effects that were

counteracted by VS.

Overall, these results deepen our knowledge about the role of ageing in

microglial biology and elect VS as a potential new therapeutic approach to modulate

Aβ-induced microglial activation and dysfunction in AD.

Keywords: Ageing, Alzheimer disease, Amyloid-beta peptide, Microglia reactivity,

Neuroinflammation, Neuron-microglia communication, Vinyl sulfone compound

Page 14: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

VIII

[Escre

va um

Page 15: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

IX

SUMÁRIO

A Doença de Alzheimer (DA) é considerada a doença neurodegenerativa mais

comum e a principal causa de demência a nível mundial, afectando cerca de 35

milhões de pessoas. Esta doença é caracterizada por uma perda progressiva das

capacidades cognitivas, o que leva a um completo estado de debilitação e, por último,

à morte. O principal factor de risco é o envelhecimento e permanece sem tratamento

até hoje.

No encéfalo dos pacientes com DA, para além de uma acentuada perda de

neurónios e sinapses, são ainda observadas duas características distintas da doença:

emaranhados neurofibrilares e placas amilóides. Os primeiros consistem em

agregados filamentosos intracelulares de proteína tau hiperfosforilada, enquanto que

as últimas se devem a depósitos extracelulares do péptido β-amilóide (βA), um

pequeno fragmento de 38-43 aminoácidos que resulta da cisão sequencial da proteína

precursora amilóide por duas secretases, a β-secretase e a γ-secretase.

A causa primordial da DA ainda está por desvendar; contudo, o βA, devido aos

seus variados efeitos neurotóxicos é apontado como um dos principais factores

envolvidos no desenvolvimento da doença. Este, principalmente na forma de 42

aminoácidos, possui uma grande tendência para se auto-agregar, formando estruturas

que vão desde oligómeros até fibrilhas, as quais originam as placas amilóides. Apesar

de inicialmente se julgarem ser os agregados fibrilares as formas mediadoras de

toxicidade, nos últimos anos, as formas oligoméricas têm sido apontadas como as

verdadeiramente tóxicas. Porém, além da oligomerização e deposição de βA, outros

mecanismos têm vindo a ser estudados pela sua associação com a patofisiologia da

DA, nomeadamente a hiperfosforilação da proteína tau, o stress oxidativo, a disfunção

mitocondrial e a neuroinflamação.

As células da microglia, consideradas as células imunes do sistema nervoso

central, desempenham um papel fundamental na resposta inflamatória. Estas células,

conhecidas por constantemente sondarem o tecido cerebral e responderem a vários

estímulos patológicos no sentido de reporem a homeostasia, são frequentemente

encontradas num estado activado e associadas às placas amilóides. Estudos já

realizados evidenciaram que o βA interage com a microglia por meio de vários

receptores, induzindo a libertação de diversas moléculas de carácter pro-inflamatório e

citotóxico. Neste sentido, a deposição de βA pode induzir uma permanente activação

da microglia, conduzindo a uma perpetuação da inflamação e, consequentemente,

contribuindo para neurodegeneração. Por outro lado, a microglia também é

responsável pela libertação de vários factores neurotróficos e pela degradação e

Page 16: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

X

fagocitose de βA, funções fundamentais na restrição da acumulação de βA e suporte

dos neurónios, impedindo a progressão da DA.

Vários estudos têm demonstrado que a microglia durante o envelhecimento e

em presença de diversos estímulos assume um fenótipo mais do tipo inflamatório, o

qual na presença de uma nova indução acaba por ocasionar uma resposta mais

exacerbada. Para tal, pode igualmente contribuir qualquer alteração da comunicação

entre a microglia e os neurónios, os quais desempenham um papel fundamental no

controlo do estado de activação da célula glial. Provavelmente na sequência destas

sucessivas activações a microglia pode tornar-se morfologicamente distrófica, sendo

assim frequentemente encontrada no cérebro envelhecido e, principalmente, na DA.

Tal indica que estas células sofrem senescência e degeneração com o avanço da

idade, predispondo à doença, que por sua vez intensifica a disfunção da célula,

possivelmente na sequência da exposição ao βA. Assim, com a idade e/ou à medida

que a doença progride, a microglia tende a degenerar, deixando de dar suporte aos

neurónios, o que contribui para a sua morte. Este ponto de vista também ajuda a

explicar a razão pela qual a abordagem terapêutica com fármacos anti-inflamatórios

não esteróides só produz efeitos benéficos nos estádios mais iniciais da doença,

enfatizando a importância do intervalo de actuação nas abordagens terapêuticas.

Com base no crescente trabalho que tem vindo a ser desenvolvido sobre o

papel da microglia na progressão da DA, esta dissertação teve como primeiro

objectivo compreender de que modo o envelhecimento microglial influencia a

reactividade da microglia ao βA e se a capacidade da microglia responder à

sinalização neuronal fica comprometida em resultado do envelhecimento. Para isto,

recorreu-se a um modelo de envelhecimento in vitro utilizando células microgliais

primárias. Foram escolhidas duas idades celulares diferentes, as de 2 e 15 dias de

cultura, de modo a mimetizar um fenótipo microglial jovem e envelhecido,

respectivamente. Estas células foram então colocadas na presença, ou não, de

neurónios e estimuladas com 50 nM e 1000 nM de βA. Depois da incubação,

determinou-se no meio extracelular a concentração de glutamato, de adenosina

trifosfato (ATP) e das metaloproteinases de matriz extracelular (MMP), factores que

desempenham um papel importante na mediação da resposta inflamatória.

Os resultados obtidos demonstraram que a microglia liberta uma menor

quantidade de glutamato e de MMP-9 à medida que envelhece. Verificou-se ainda

uma ausência de resposta da microglia envelhecida em termos de libertação de

glutamato, quando exposta ao βA na presença de neurónios. Em relação à MMP-2,

por outro lado, foi demonstrado um aumento na sua libertação pela microglia em

resultado do envelhecimento. Quanto à libertação de ATP, em qualquer uma das

Page 17: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

XI

condições estudadas, não foram observadas alterações. Nesta abordagem, a

exposição ao βA também não provocou activação microglial, em qualquer uma das

idades avaliadas. Porém, é de salientar que observámos respostas diferentes em

termos de libertação de glutamato e MMP-9 e MMP-2 com a idade e, de forma geral,

estes resultados fortalecem a ideia de que a biologia microglial se altera como

resultado do envelhecimento. Apesar de o envelhecimento poder condicionar a

resposta da célula glial ao βA, tal não foi evidenciado pelos parâmetros analisados,

pelo que se optará pela determinação de indicadores directos de neuroinflamação no

futuro.

O segundo objectivo desta dissertação foi o de avaliar a capacidade de uma

vinil sulfona (VS) sintetizada pelo grupo de Química Medicinal como modulador da

activação microglial pelo βA. Para isto, a linha microglial N9 foi estimulada com as

duas concentrações de βA (50 nM e 1000 nM), na presença ou ausência de duas

concentrações de VS (10 µM e 20 µM), sendo posteriormente avaliados os seguintes

parâmetros: viabilidade celular, libertação de MMPs, capacidade fagocítica da

microglia e a expressão de proteínas como milk fat globule-EGF factor 8 protein (MFG-

E8), toll-like receptor 4 (TLR-4) e high-mobility group box 1 (HMGB1).

A nível de viabilidade, não foram observadas alterações significativas por parte

deste composto testado, nem pelas concentrações de βA usadas. Tanto quanto à

libertação de MMP-2 e MMP-9 como relativamente à expressão de HMGB1,

importantes mediadores da resposta inflamatória, observou-se uma notória indução

após exposição ao βA, sendo a mesma revertida após tratamento com VS. Quanto à

expressão do TLR-4, verificou-se uma redução na presença de βA, sendo este efeito

invertido na presença da maior concentração de VS usada. Em relação tanto à

capacidade fagocítica microglial como à expressão de MFG-E8, um factor com um

papel importante na fagocitose, foi observada uma redução após indução com βA;

porém, apenas a capacidade fagocítica retomou os valores do controlo após adição de

VS. Assim, de forma geral, o composto por nós estudado demonstrou ser capaz de

reverter alguns dos efeitos causados pelo βA, reduzindo de forma significativa a

expressão de factores inflamatórios. Porém, mais estudos ainda têm que ser

realizados para determinar a verdadeira extensão dos efeitos mediados por este

composto, assim como os mecanismos pelos quais actua.

Em suma, estes resultados demonstraram contribuir para o conhecimento de

como o envelhecimento da microglia condiciona a funcionalidade da célula e

evidenciaram a VS como uma potencial terapêutica a ser explorada na modulação de

outros determinantes, quer da activação, quer da latência microglial, observada na DA.

Page 18: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

XII

Palavras-chave: Beta-Amilóide, Comunicação neurónio-microglia, Doença de

Alzheimer, Envelhecimento celular, Neuroinflamação, Reactividade Microglial, Vinil

sulfona

Page 19: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

XIII

TABLE OF CONTENTS

NOTA PRÉVIA ................................................................................................................. I

AGRADECIMENTOS ....................................................................................................... III

ABSTRACT ................................................................................................................... VII

SUMÁRIO ...................................................................................................................... IX

INDEX OF FIGURES ................................................................................................... XVII

ABBREVIATIONS......................................................................................................... XIX

I. INTRODUCTION ................................................................................................... 1

1. Alzheimer’s disease (AD): An overview .............................................................. 1

1.1. Disease presentation................................................................................... 1

1.2. Pathophysiology .......................................................................................... 2

1.2.1. Amyloid beta peptide ............................................................................... 4

1.2.2. Tau .......................................................................................................... 7

1.2.3. Oxidative Stress ...................................................................................... 8

1.2.4. Mitochondrial Dysfunction ........................................................................ 8

1.2.5. Neuroinflammation................................................................................... 9

1.3. Diagnosis and treatment ........................................................................... 10

2. Microglia in AD: A closer look to the central player in neuroinflammation ......... 12

2.1. Microglia: origin and functions ................................................................... 12

2.2. Activation of microglia in AD ...................................................................... 14

2.2.1. Interaction of amyloid beta with microglia .............................................. 14

2.2.2. Production of inflammation-related factors by amyloid beta-stimulated

microglia .............................................................................................................. 16

2.2.3. Amyloid beta-induced microglial phagocytosis ....................................... 17

2.3. Microglial role in the progression of AD: activation vs. dysfunction ............ 18

2.4. Microglia-neuron interplay in AD ............................................................... 20

2.5. Vinyl sulfones and inflammatory diseases: a possible therapeutic approach

for microglial modulation in AD? .......................................................................... 23

2.6. Microglial in vitro models for studying AD .................................................. 24

3. Aims ................................................................................................................. 27

II. MATERIAL AND METHODS ............................................................................... 29

1. Material ............................................................................................................ 29

1.1. Reactives .................................................................................................. 29

1.1.1. Cell culture media .................................................................................. 29

1.1.2. Supplements and chemicals .................................................................. 29

Page 20: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

XIV

1.1.3. Antibodies .............................................................................................. 30

1.2. Equipment ................................................................................................. 30

1.3. Animals ..................................................................................................... 31

2. Methods ........................................................................................................... 31

2.1. Primary cultures ........................................................................................ 31

2.1.1. Isolation of primary cell cultures ............................................................. 31

2.1.2. In vitro treatment of primary cell cultures with amyloid beta ................... 32

2.1.3. Quantification of extracellular ATP ......................................................... 33

2.1.4. Measurement of extracellular glutamate ................................................ 34

2.1.5. Gelatin zymography ............................................................................... 34

2.2. Cell line ..................................................................................................... 35

2.2.1. N9 microglial cell line culture ................................................................. 35

2.2.2. Treatment of N9 microglial cell line ........................................................ 35

2.2.3. Determination of Cell viability ................................................................. 36

2.2.4. Microglia phagocytic capacity ................................................................ 37

2.2.5. Gelatin zymography ............................................................................... 37

2.2.6. Western blot analysis ............................................................................. 37

2.3. Statistical analysis ..................................................................................... 38

III. RESULTS .......................................................................................................... 39

1. Characterization of young and aged microglia reactivity to Aβ and of modulation

by neurons .............................................................................................................. 39

1.1. Young cells release higher levels of glutamate than older ones when in

monoculture, but not in mixed culture with neurons, and no alterations are

produced by Aβ in each differently aged cell ........................................................ 39

1.2. ATP release by microglia in response to Aβ is not affected by ageing or

neurons ............................................................................................................... 40

1.3. Release of MMP-9 and MMP-2 has opposite profiles in young and aged

microglia and is not altered by Aβ or by the presence of neurons ........................ 41

2. Exploring a new therapeutic approach to prevent Aβ-induced microglial

activation ................................................................................................................. 43

2.1. Loss of viability in Aβ-stimulated microglia is slightly prevented by VS ...... 43

2.2. Release of active MMPs by microglia is enhanced upon Aβ stimulation and

decrease by VS ................................................................................................... 45

2.3. Expression of TLR4 is reduced by the highest concentration of Aβ, an

effect that is counteracted by VS ......................................................................... 46

2.4. Aβ triggers a reduction in microglia phagocytosis, which is prevented by VS

47

Page 21: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

XV

2.5. The expression of MFG-E8 is reduced by 50 nM Aβ and VS alone, while

increases when the highest concentrations of both compounds are concomitantly

used 49

2.6. Aβ induces an increase in the expression of HMGB1 and effect that is

prevented by VS .................................................................................................. 50

IV. DISCUSSION .................................................................................................... 51

Future perspectives ................................................................................................. 59

V. REFERENCES ................................................................................................... 61

Page 22: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

XVI

[Escre

va um

Page 23: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

XVII

INDEX OF FIGURES

I. INTRODUCTION

Fig. I.1: Neuropathology of Alzheimer’s Disease ........................................................... 3

Fig. I.2: Proteolytic processing of amyloid precursor protein ......................................... 5

Fig. I.3: Amyloid beta assembly states .......................................................................... 6

Fig. I.4: Microglia activation states .............................................................................. 14

Fig. I.5: Amyloid beta interaction with microglial receptors .......................................... 16

Fig. I.6: Microglia-neuron communication .................................................................... 21

II. MATERIAL AND METHODS

Fig. II.1: Experimental procedure used in the primary cellular cultures and parameters

evaluated .................................................................................................................... 33

Fig. II.2: Analysis of Aβ aggregation forms present in our experimental conditions ..... 34

Fig. II.3: Experimental procedure used in the N9 microglial cell cultures and parameters

evaluated. .................................................................................................................. 36

III. RESULTS

Fig. III.1: The extracellular levels of glutamate are higher in young microglia than in old

microglia, being reduced in the first by the presence of neurons and no effect by the

presence of Aβ could be observed in any of the cases. ............................................. 40

Fig. III.2: Microglia release of ATP by Aβ stimulation does not change by age or by

incubation with neurons .............................................................................................. 41

Fig. III.3: MMP-2 release from microglia is up-regulated by ageing, while MMP-9 is

instead down-regulated and these features are not modulated by Aβ or neuronal

signaling ..................................................................................................................... 42

Fig. III.4: Cell viability shows a trend to be decreased by Aβ and was slightly prevented

by VS .......................................................................................................................... 44

Fig. III.5: Stimulation of microglia with Aβ increases the release of active MMP-2 and

MMP-9, which is prevent by VS .................................................................................. 46

Fig. III.6: Expression of TLR4 is reduced by both Aβ and VS, but revealed to be

induced by the highest VS concentration when in the presence of 1000 nM Aβ .......... 47

Fig. III.7: Microglia phagocytic ability is decreased upon exposure to Aβ and restored

by VS. ........................................................................................................................ 48

Page 24: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

XVIII

Fig. III.8: The expression of MFG- E8 that is reduced by 50 nM and by VS alone, is

increased when the highest concentration of both compounds are concomitantly used.

................................................................................................................................... 49

Fig. III.9: Expression of HMGB1 greatly increases in microglia treated with Aβ, but VS

shows ability to prevent such effect from occurring ..................................................... 50

Page 25: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

XIX

ABBREVIATIONS

ABAD Amyloid-binding alcohol dehydrogenase

Ab/Am Antibiotic-antimycotic

AD Alzheimer’s disease

ADAM A desintegrin and metalloproteinase

APOE Apoliprotein-E

APP Amyloid precursor protein

ATP Adenosine triphosphate

Aβ Amyloid beta

Aβ40 Amyloid beta with 40 amino acids

Aβ42 Amyloid beta with 42 amino acids

BACE1 β-site APP cleaving enzyme 1

BBB Blood-brain barrier

BSA Bovine serum albumin

CaCl2 Calcium chloride

CD Cluster of differentiation

CNS Central nervous system

CSF Cerebrospinal fluid

DIV Days in vitro

DMEM Dulbecco’s modified Eagle’s medium

EDTA Ethylenediamine tetraacetic acid

fAβ Fibrillar amyloid beta

FBS Fetal bovine serum

GLT1 Glutamate transporter 1

HMGB1 High-mobility group box 1

HRP HorseRadish Peroxidase

Iba1 Ionized calcium-binding adaptor molecule 1

IDE Insulin-degrading enzyme

IL Interleukin

KOH Potassium hydroxide

LPS Lipopolysaccharide

MAP Microtubule associated protein

MCI Mild cognitive impairment

MEM Minimum essential medium

MFG-E8 Milk fat globule EGF-like factor 8

Page 26: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

XX

MHC Major histocompatibility complex

MMP Matrix metalloproteinase

NEAA Non-essential amino acids

NEP Neprilysin

NF-κB Nuclear factor-κB

NGF Nerve growth factor

NMDA N-methyl-D-aspartate

NO Nitric oxide

NSAID Non-steroidal anti-inflammatory drugs

oAβ Oligomeric amyloid beta

PBS Phosphate buffered saline

PCR Polymerase Chain Reaction

PDL Poly-D-lysine

PET Positron emission tomography

PI Propidium iodide

PMSF Phenylmethylsulfonyl fluoride

PRR Pattern recognition receptors

PS Phosphatidylserine

PSEN1/2 Presinilin 1/2

RAGE Receptor for advanced glycation end-products

RNS Reactive nitrogen species

ROS Reactive oxygen species

RPMI Roswell Park Memorial Institute

SDS Sodium dodecyl sulphate

SDS-PAGE SDS- polyacrilamide gel electrophoresis

TLR Toll-like receptor

TNF Tumor necrosis factor

VR Vibronectin receptor αvβ3

VS Vinyl sulfone-based inhibitor

Page 27: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

1

I. INTRODUCTION

1. Alzheimer’s disease (AD): An overview

1.1. Disease presentation

In 1901, Alois Alzheimer, a german physician-pathologist, faced the case of

August Deter, a 51 years-old woman who suffered from an intriguing set of symptoms,

namely, progressive memory loss, confusion, disorientation and delusions [1]. During

almost five years, her mental capacities continued to deteriorate rapidly, until she finally

died in a completely demented state [2]. After patient’s death, Alzheimer performed an

autopsy on her brain, finding significant cerebral atrophy and the presence of unusual

plaques and tangles [1]. His characterization and notes on the case led to the

discovery of a new neurodegenerative disease, a disorder that now bears his name –

Alzheimer’s disease (AD).

One hundred years after its identification, AD has become one of the most

serious preoccupations of our society. AD is considered to be the most common

neurodegenerative disorder and the leading cause of dementia, accounting for 50-60%

of all dementia cases [3]. Presently, the number of people affected by AD worldwide is

estimated to be around 35 million [4]; however, this number is expected to increase,

reaching more than 115 million by 2050 [5]. Like most dementias, AD’s clinical course

begins insidiously, with an unusual gradual decline in memory - classified as mild

cognitive impairment (MCI) - and then progresses to involve the deterioration of other

cognitive functions, such as thinking and reasoning, and also behavior abnormalities,

leaving the individual in a completely debilitated state and totally dependent on others.

[6]. Symptoms unroll from several years to a decade and mortality is frequently related

to resulting secondary issues, like multi-organ failure or opportunistic infections

[7]..There is still no effective treatment for the disease, which in association with the

huge need for patients’ care, makes AD an enormous financial burden to families and

healthcare systems [8].

The elderly population is the main group at risk of developing AD, since

advanced age is the factor most strongly associated with the disease [8] . Based on the

age of onset, AD can be classified in two types: late-onset AD and early-onset AD. The

late-onset AD is the most common form of the disease, accounting for around 90-95%

of cases, and has its onset after age 65. On the other hand, the other 5-10% of AD

Page 28: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

2

cases occur between ages 30-65, being defined as early-onset AD [4]. Some of early-

onset AD cases, termed familial AD, are inherited in an autosomal dominant manner

and are caused by mutations in three genes: the amyloid precursor protein (APP) gene

on chromosome 21, the presinilin 1 (PSEN1) gene on chromosome 14 and the

presinilin 2 (PSEN2) gene on chromosome 1 [8]. Nonetheless, most of AD cases are

spontaneous - or within familial clusters that present no clear Mendelian inheritance

pattern -, representing a complex disorder likely caused by an interplay between

multiple susceptibility genes, environmental factors and ageing [5]. In the last years,

several genome-wide association studies have been done, leading to the identification

of different susceptibility genes for AD. Of the genes identified, Apoliprotein-E (APOE)

on chromosome 19 presents, by far, the strongest correlation with AD [5]. There are

three common variants of this gene: ε4, which is the high risk form, increasing the risk

in a dose-dependent manner; ε3, the most common in humans and the neutral allele;

and ε2, which appears to be of low risk [9].

1.2. Pathophysiology

On a macroscopic level, AD pathology can be characterized by a progressive

atrophy of the brain, affecting some regions more than others [10] (Fig. I.1). At a

microscopic level, it is observed that atrophy mainly reflects the loss of neurons and

synaptic contacts [11]. This happens in a defined pattern, with neurons in the

enterohinal cortex being first affected, followed by hippocampal neurons and,

ultimately, surrounding neurons in extra-hippocampal regions, such as neocortical

areas [12]. Nonetheless, not all neuronal populations are equally susceptible to AD. For

instance, neurons that are capable of projecting over long distance (projection

neurons), possessing disproportionately long in relation to cell body size and sparsely

myelinated axons seem to be particularly prone to the disease process [13]. It is not yet

fully elucidated why this happens; however, these neurons have a large cell surface, a

high energy requirement (due to the lack of myelinization) and are very reliant on

axonal transport (anterograde and retrograde) for sustained function and trophic

support, which might increase their susceptibility to damage [14]. In the case of AD,

damage is believed to occur owning to mechanisms outside as well as inside the

neuron and it is characterized by the appearance of intracellular filamentous

aggregates of tau protein, called neurofibrillary tangles, and by extracellular deposits

consisting largely of aggregated amyloid beta (Aβ) peptide, present typically between

neurons with dystrophic neurites, called amyloid plaques [15,16]. These changes that

where firstly described by Dr. Alzheimer are considered the major histopathological

Page 29: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

3

hallmarks of the disease [1]. Besides these, gliosis, the accumulation of reactive

astrocytes and microglia near lesions, is also seen in AD [17]. Of the various

pathological features of AD, synaptic loss and selective neuronal death (in the limbic

system and neocortex) are the ones that correlate best with the progression of

cognitive decline, representing the main substrates of dementia [18].

Fig. I.1: Neuropathology of Alzheimer’s Disease. (A) Extensive atrophy is observed in later stages of

Alzheimer’s Disease (AD), which is accompanied by enlarged ventricles, narrowed cortical gyri and widened sulci. Temporal lobes are most severely affected, while occipital lobes and motor cortex are usually spared (adapted from Holtzman [19]). (B) High number of degenerated neurons containing neurofibrillary tangles are found in AD, as well as extracellular deposits of amyloid beta protein (amyloid plaques), which are frequently surrounded by reactive glia cells. (C) In healthy brain, although some characteristics of the disease may be found as a result of the ageing process, they are usually not only present or present in a lower extent.

One of the most important issues for unraveling the AD mystery is finding out

the mechanisms underlying the disease process. In this regard, several mechanisms

are under study, including Aβ aggregation, deposition and toxicity, tau

hyperphosphorylation with tangle formation, as well as oxidative stress, mitochondrial

dysfunction and neuroinflammation.

Page 30: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

4

1.2.1. Amyloid beta peptide

Aβ, the principal component of amyloid plaques, is considered a central player

in AD and therefore it has been a major focus of research in the last 30 years. The

main source of Aβ is the processing of APP, a transmembrane protein located on the

plasmatic membrane or in intracellular compartments of diverse cells, like neurons and

glia cells, whose main function still remains unknown [20,21].

APP can undergo several different endoproteolytic cleavage events, which

culminate, or not, in the formation of the Aβ peptide (Fig. I.2). This proteolyic cleavage

is sequential, starting with the cleavage by α-secretase or β-secretase, followed in both

cases by cleavage by γ-secretase. Under normal conditions, APP is preferentially

processed by the so-called non-amyloidogenic pathway, which starts with cleavage by

α-secretase [22]. The enzyme is one of the ADAM9, 10 or 17, all a desintegrin and

metalloproteinase (ADAM) family members [18], and cleaves within the Aβ domain,

resulting in the production of a small peptide presumably non-pathogenic dominated

P3, instead of Aβ . Alternatively, APP may be a substrate of β-secretase, a role that

has been essentially attributed to the β-site APP cleaving enzyme 1, BACE1 , initiating

the amyloidogenic pathway, which results in the formation of Aβ [22]. α-secretase and

β-secretase cleave APP at single sites, whereas γ-secretase, an enzyme complex for

which PSEN1 and PSEN2 act as catalytic subunits, performs a sequential series of

intramembranous cuts, giving rise to products of varying length [23]. Aβ peptide length

varies between 38 to 43 amino acids. The most common isoforms within the brain are

the ones with 40 (Aβ40) and 42 (Aβ42) residues, representing ~90% and ~10%,

respectively of the total [24]. Both Aβ40 and Aβ42 can spontaneously self-aggregate into

higher order structures, ranging from low molecular oligomers (oAβ) to insoluble

aggregates of fibrils (fAβ) [25] ; however, A42 is more hydrophobic and therefore more

prone to aggregation [22] (Fig. I.3).

Page 31: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

5

Fig. I.2: Proteolytic processing of amyloid precursor protein. There are two different pathways by

which amyldoid precursor protein (APP) can be processed: the non-amyloidogenic pathway and the amyloidogenic pathway. In the non-amyloidogenic pathway, APP is initially cleaved by α-secretase, originating a secreted fragment of APP (sAPPα) and a membrane-bound carboxy-terminal fragment of 83 amino acids (C83). Subsequently, C83 is further processed by γ-secretase, giving rise to P3, precluding amyloid beta (Aβ) formation. Amiloidogenic processing, alternatively, is initiated by β-secretase and results in the release of sAPPβ. The retained membrane-bound fragment of 99 amino acids (C99) is also a γ-secretase substrate, leading to the production of Aβ. APP intracellular domain (AICD) is an end-product in both pathways, being translocated into the nucleous to regulate gene transcription.

Several lines of investigation support the notion that the pathogenesis of AD is

directly related to a progressive accumulation of Aβ, resulting from an imbalance

between the levels of Aβ production, aggregation and clearance in the brain [26]. The

strongest evidence for this is seen in familial AD, where all cases identified carry

mutations in APP or presenilin genes, leading to an increase of total Aβ production or a

shift in the Aβ40/Aβ42 ratio toward Aβ42. In addition, individuals with Down’s syndrome,

who possess an extra APP gene, also develop an AD-like pathology [27]. In sporadic

form of AD, the Aβ accumulation is thought to be more related with a dysfunction of its

clearance mechanisms [26]. Indeed there is an association between APOE isoforms

and Aβ clearance, with the ε4 isoform showing to be the less efficient in clearing Aβ.

[11]. Furthermore, the activity of major Aβ-degrading enzymes, like neprilysin (NEP)

and insulin-degrading enzyme (IDE), has also been reported to be reduced with age

and in brain regions affected by AD [28]. The identification of these AD-related genes

has led to the creation of diverse animal models, which although not completely

mimicking AD, are a valuable tool to study the mechanisms involved in Aβ

accumulation [29].

Page 32: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

6

How Aβ contributes to neurodegeneration in AD is still not fully understood, but

aggregation seems to be essential for A toxicity [27]. Since extracellular deposition of

A is a major hallmark of AD, attention was given to the insoluble fA, the main

aggregated forms in the amyloid plaques [22]. Some studies demonstrated that fAβ are

neurotoxic in vitro [30] and in vivo [31] and it was then hypothesized that they could

represent a primary cause of neurodegeneration in AD. Nevertheless, the number and

temporal progression of plaques do not correlate well with the local extent of the

neuronal death and synaptic loss, nor with the disease progression [22]. Moreover,

some individuals show high levels of plaques without revealing any cognitive

impairment [18]. On the other hand, the levels of extracellular soluble oA show a

robust correlation with the extent of synapse loss and cognitive impairment [27]. A great

number of different oA, natural or synthetic, possessing various sizes and shapes,

have been reported: dimers and trimers, A-derived diffusible ligands, A56* (56 kDA)

and others [32]. These intermediate forms are considerably more toxic than fA, and

are appointed as the main culprits for neurodegeneration in AD [33] .

Fig. I.3: Amyloid beta assembly states. Amyloid beta (Aβ) can exist in diverse assembly states, which

include monomers, oligomers or fibrils, with the latter being the state deposited as amyloid plaques. Formation of fibrils is a complex multi-step process that is preceded by the oligomerization and aggregation of monomeric Aβ. The mechanism driving this process is not fully understood, but may be related to protein misfolding

Besides extracellular accumulation, studies indicate that A also deposits inside

the cells and that this may be an early event in AD, since it precedes extracellular

accumulation and has been reported on brain regions that are more prone to the

development of early AD pathology, such as entorhinal cortex and hippocampus.

Moreover, it is inside the cells that oligomerization process is thought to begin [34]. The

intracellular localization of Aβ is a result of both APP cleavage inside the cell and new

uptake of extracellular Aβ through receptors and transporters, such as the receptor for

advanced glycation end-products (RAGE) [33].

The mechanisms by which Aβ induce neuron and synaptic degeneration are

complex and far from being resolved, but it seems to be influenced by differences in

aggregation forms and localization (intracellular or extracellular). Some of the proposed

mechanisms are: interaction with diverse cell-surface receptors, conducting to inhibition

Aβ monomers Aβ oligomers Aβ fibrils Amyloid plaques

Page 33: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

7

or aberrant activation of signal transduction pathways; insertion in membrane and

formation of ion channels or pores, disrupting ionic homeostasis; disruption of

lysosomal membrane; proteasome impairment [35,33].

It is important to have in mind that Aβ is a physiological product of the organism

and as so it may have a physiological role. In this regard, some authors suggest a vital

function in modulation of synaptic plasticity, neuronal survival and neuronal

development, particularly at low concentrations, when in the absence of oligomerization

[36].

1.2.2. Tau Almost concomitantly with the identification of Aβ in amyloid plaques,

neurofibrillary tangles were demonstrated to be composed of abnormally

hyperphosphorylated tau protein. Tau is a member of the family of microtubule-

associated proteins (MAPs) that is particularly abundant in neuronal axons [37]. The

best-established functions of tau are the stabilization of microtubules and the regulation

of motor-driven axonal transport [38]. The biological activity of tau is dependent on its

degree of phosphorylation, which modulates the capacity to bind to microtubules [39].

Several kinases (e.g, glycogen synthase kinase 3 beta) and phosphatases (eg, protein

phosphatase 1) are responsible for the fine tuning of tau phosphorylation state [37].

During AD pathogenesis, tau becomes increasingly phosphorylated, which

causes its detachment from microtubules, resulting in a loss of its normal function [40].

Unbound hyperphosphorylated tau is able to sequester normal tau and other MAPs,

further compromising microtubule structure. Moreover, it possesses an inherent

capacity to self-assembly, forming paired helical filaments that will eventually give rise

to neurofibrillary tangles [39]. Tau aggregates have been reported to be toxic primarily

in the form of oligomers, like in the case of Aβ [41]. Therefore, it is though that the loss

of normal functions and the toxic gain of function of tau, resulting from its

hyperphosphorylation, might contribute to synaptic dysfunction and neurodegeneration

[42]. The causes of abnormal phosphorylation are unclear, but altered function of

kinases or phosphatases, or maladaptation by cellular and microenvironmental stress,

are thought to be implicated [43] .

The involvement of tau in neurodegenerative process is also supported by the

existing correlation between neurofibrillary tangles and the AD onset and progression

[39]. Nonetheless, tau mutations have not been connected to AD. Indeed they are used

to distinct phenotypic diseases, like frontotemporal dementia, that do not display

plaques [42]. This suggests that tau pathology in AD is downstream of Aβ

Page 34: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

8

accumulation, and is supported by experimental evidence indicating that Aβ

accumulation precedes and drives tau aggregation. Some studies also indicate that the

presence of tau might be necessary for Aβ neurotoxicity [37].

1.2.3. Oxidative Stress

Oxidative stress derives from an imbalance between the production of reactive

oxygen/nitrogen species (ROS/RNS) and the cellular anti-oxidant mechanisms [44].

ROS and RNS, resulting from hydrogen peroxide and nitric oxide (NO) accumulation,

may damage nucleic acids, carbohydrates, lipids (lipid peroxidation) and proteins

(protein oxidation) [45]. Neurons are considered highly susceptible to an increase of

ROS/NOS due to their low levels of antioxidants mechanisms when compared with

other neural cells [44].

Extensive evidence of increased oxidative damage, including lipid peroxidation,

protein oxidation and nucleic acid oxidation, has been found in AD vulnerable neurons

and suggested to precede any other feature of the disease [46,47]. In addition, low

levels of naturally occurring anti-oxidants, such as α-tocopherol, excessive iron and

copper deposits, intense microglial activation and mitochondrial abnormalities, are all

known to be involved in the production of ROS/RNS and reported in AD [48,49].

Aβ is able to promote the oxidation of compounds like phospholipids and

cholesterol, if an intact methionine in position 35 is present [48]. Furthermore, Aβ also

promotes ROS production by directly associating with transition metals such as iron

and copper, by stimulating microglial activation or through its effects on mitochondria

dynamics [50,51]. Despite promoting oxidative stress, some authors suggest that

accumulation of Aβ might also be a protective response against oxidative stress. For

instance, Aβ has demonstrated to follow the appearance of oxidative stress markers in

AD and to have anti-oxidant properties in cerebrospinal fluid (CSF) and plasma,

protecting lipoproteins from oxidation [49].

Oxidative stress further potentiates Aβ oligomerization and aggregation, which

in turn produces neuroinflammation, mitochondrial damage, and thus, further ROS

generation, leading to a vicious cycle of more oxidative stress and more damage to the

cells [49]. Consequently, there is an enhanced cell dysfunction and demise.

1.2.4. Mitochondrial Dysfunction

Mitochondria are the main producers of both energy and free radicals, through

oxidative phosphorylation and endogenous ROS species, playing a key role in

neuronal function and survival in the brain, the third most energy-expensive organ in

Page 35: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

9

the human body [52,53].

Increasing evidence in patients and disease models indicates that mitochondrial

dysfunction accompanies ageing and plays an important role in AD. Indeed, it seems to

be present at all stages of AD and worsens as the disease progresses [54].

Abnormalities on mitochondrial function that are observed in AD include a decreased in

activity of respiratory chain enzymes (e.g. cytochrome c oxidase) and some Krebs

cycle enzymes (eg. α-ketoglutarate dehydrogenase). Moreover, a reduction in the

mitochondrial membrane potential and in the levels of adenosine triphosphate (ATP),

together with increased oxidative modifications in mitochondrial DNA are observed in

the disease [55,56].

The recent intracellular existence of Aβ has led researchers to consider it as a

potential cause of mitochondrial dysfunction. Indeed, evidence suggests that Aβ

crosses mitochondrial membranes, accumulating in the interior of mitochondria [57],

thus influencing the activity of electron transport chain complexes and disturbing

calcium storage, leading to apoptotic pathways. Moreover, Aβ interacts with

mitochondrial matrix components, causing a decrease in the membrane potential and

impairing ATP formation [58]. amyloid-binding alcohol dehydrogenase (ABAD) is one of

Aβ targets. Interaction with ABAD leads to an inactivation of the enzyme activity and

promotes mitochondrial generation of free radical species [59]. It is also important to

refer that mitochondria-derived ROS seem to be able to trigger amyloidogenic APP-

processing, possibly by inducing BACE1 activation, which further implies an existence

of a vicious cycle that contributes to AD pathogenesis [60].

It is now accepted that mitochondria are dynamic organelles and maintain their

homeostasis and function by constantly undergoing fission (or splitting) and fusion (or

combining) [52]. In AD this dynamic equilibrium is disrupted in favor of fission, which

may compromise cellular integrity [61]. Recently, it has also been appointed that APP

overexpression, likely through Aβ-mediated effects, could be responsible for abnormal

mitochondrial dynamics in AD, including decreased mitochondrial mobility and

alterations in fission/fusion processes [54]

1.2.5. Neuroinflammation

The term neuroinflammation refers to the inflammatory-like process that occurs

in the central nervous system (CNS), which is usually a transient, well controlled

process, resulting from the response to a toxic insult [62]. In general, an acute

neuroinflammatory response is beneficial to the CNS, since it contributes to repair the

damaged tissue and to minimize further injury; in contrast, chronic neuroinflammation

Page 36: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

10

that persists after an initial toxic insult may produce degenerative changes in neurons

and alter brain function [63,64].

Although contribution of chronic neuroinflammation to AD is a controversy

issue, namely in late AD stages [65], it may exacerbate the course of the disease, while

leading to microglia senescence [66]. This process seems to have a genetic relation

with the disease, since AD patients are more represented with pro-inflammatory

genotypes than anti-inflammatory genotypes [67]. Signs of inflammation are particularly

localized in brain areas exhibiting high levels of AD pathology. Furthermore, high

pathology controls (that is, individuals presenting Aβ and tau aggregates at levels

similar to AD patients, but that do not develop dementia) show lower signs of

inflammation. In addition, the inflammatory mechanisms present in AD patients are

comparable to those present in peripheral inflammatory reactions, which are

established to be cytotoxic, and therefore are likely to have cytotoxic effects on

neurons [68]. AD patients who develop a short-term peripheral infection present a

sudden decline in cognitive state, rarely returning to previous level, even after the

eradication of the infection [69]. It is worthwhile to mention that although some studies

demonstrate that individuals treated (+ 24 months of cumulative use) with non-steroidal

anti-inflammatory drugs (NSAIDs) present a 60-80% reduction in the risk of developing

AD [70,71], anti-inflammatory drugs are ineffective or even harmful when used in

advanced stages of the disease [72,73].

The neuroinflammatory process involves primarily the activation of microglia,

the immune cells of the CNS, but also astrocytes, and even neurons. [74]. All these

cells are able to produce/release several inflammatory mediators, which include pro-

inflammatory cytokines, mainly tumor necrosis factor (TNF)-α, interleukin (IL)-1β and

IL-6, as well as ROS/NOS, complement proteins and various proteolytic enzymes [62].

Some of these inflammatory factors by enhancing Aβ deposition and tau

phosphorylation, concur to further activation and proliferation of glial cells and to the

chronic inflammatory scenario seen in AD, at least in the earlier stages [75,76]. All

these factors, alone or in concert, can then contribute to the neuronal dysfunction seen

in the disease.

1.3. Diagnosis and treatment

Diagnosis of AD can reach about 95% confidence when realized by highly

experienced clinicians, but a full confirmation can only be obtained at autopsy [77,78].

Until recently, the criteria for the diagnosis of AD only considered the late stages of the

disease, when dementia is already present [79]. In 2011, however, a new form for

Page 37: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

11

diagnosis was proposed [80]. This new diagnostic criteria expanded the definition of

AD into 3 phases: a pre-symptomatic phase; a symptomatic, pre-dementia phase (or

MCI); and a dementia phase. Moreover, it also incorporated the use of biomarkers to

improve diagnosis, which are mainly used for research purposes.

These biomarkers are physiological, biochemical or anatomical parameters that

can be measured in vivo and reflect specific features of disease-related

pathophysiologic processes [81]. Five biomarkers were incorporated in the new

criterion proposed, being divided in two classes. The first set is constituted by

biomarkers reflecting Aβ deposition in brain, comprising both CSF evaluation of Aβ42

and positron emission tomography (PET) with the amyloid binding tracer N-methyl-

11C-2-(4-methylaminophenyl)-6-hydroxybenzothiazole. The second group reflects

neuronal degeneration or injury and includes tau (total and phosphorylated) level in

CSF, fluorodeoxyglucose uptake on PET (used to measure brain metabolic energy)

and regional brain atrophy (hippocampus, medial, basal and lateral lobes, and the

parietal lobe) on structural magnetic resonance imaging [81]. Besides these, other

promising biomarkers under investigation are the level of BACE1 [82] and isoprostanes

(markers of oxidative stress) in CSF [83]. Plasma or serum biomarkers have also been

explored, but none has shown the diagnostic accuracy of CSF biomarkers [84].

Biomarkers might certainly improve diagnosis of AD, principally at early stages.

Nonetheless, deficient standardization or their limited access still needs to be solved

[80].

After diagnosed, there is still no way to effectively deal with AD. Currently

available treatments include cholinesterase inhibitors (donepezil, galanthamine,

rivastigmine), which target the impairment of cholinergic function due to loss of basal

forebrain cholinergic neurons, and memantine, a N-methyl-D-aspartate (NMDA)

receptor antagonist, that aims to reduce excessive glutamate activation [85]. However,

although these treatments show some clinical benefit, they are just symptomatic, not

stopping disease progression [3]. Therefore, several studies for the development of

new treatments are under way. Drugs aiming to prevent Aβ aggregation by decreasing

Aβ generation and stimulating Aβ clearance are under investigation, and some of them

are even in phases II and III of clinical trials [86]. Treatments to target tauopathy in AD

have received less attention than amyloid therapies, but drugs aiming to inhibit tau

hyperphosphorylation, tau oligomerization or to promote the degradation of

hyperphosphorylated tau are also under study [87]. Additional approaches to treatment

intend to improve neuronal function by using trophic factors like nerve growth factor

(NGF). Moreover, due to the importance of oxidative stress and inflammation in AD,

neuroprotective approaches may also include anti-oxidants (e.g. vitamin E) and anti-

Page 38: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

12

inflammatory drugs (e.g. NSAID), although clear positive effects are still missing. In

addition, nondrug approaches for cognitive rehabilitation such as cognitive training and

cognitive stimulation have shown modest, but promising results [88].

2. Microglia in AD: A closer look to the central player in

neuroinflammation

The inflammatory process is unlikely to be the cause of AD [89]; nevertheless,

this does not mean that it is not important. In fact, like it was appointed before, several

evidences indicate that it may play a major role in the early disease stage. Of the

different cells involved in the inflammatory process, microglia by their immune nature

are certainly the most important ones.

2.1. Microglia: origin and functions

Microglia comprise about 12% of all the cells in the brain, occupying both grey

and white matter, with a higher density in hippocampus, substantia nigra, basal ganglia

and olfactory telencephalon [90]. Although still in debate, the general consensus is that

microglia are of hematopoietic origin, being derived from the myeloid precursor cells

from yolk sac that enter the developing CNS during embryogenesis [91].

Usually, under normal conditions, microglia are in a surveying state,

characterized by a ramified morphology with long and thin processes [92]. These

processes are highly mobile, being constantly extending and retracting to monitor

microglia’s surroundings [93]. Each microglia covers a defined territory, non-

overlapping with neighboring microglial cells [94]. Microglia are considered to be long-

lived cells, but might be replaced by low level of local self-renewal [95].

Microglia play numerous important roles through life. In the developing brain,

microglia have seen to be involved in the removal of the massive number of cells that

undergo apoptosis during developmental programmed cell death [96], in the control of

the number of synapses through the process of synaptic pruning [97] and also in CNS

vascularization [96]. Nonetheless, within the adult CNS, the primary role of microglia is

similar to that of macrophages in other tissues, representing the first line of defense

against injury or pathogens[98]. In order to detect and interpret potential insults that

disturb CNS homeostasis, a plethora of different surface receptors are expressed by

microglia, including neurotransmitter receptors and pattern recognition receptors

(PRRs), which represent a vast array of conserved receptors with ability to recognize

Page 39: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

13

and bind small molecular motifs present in pathogens (pathogen-associated molecular

patterns) or factors associated with tissue damage (alarmins or damage-associated

molecular patterns) [99,94]. These signals evoke the activation of microglia, leading to

a change of their ramified morphology to an amoeboid shape. Microglia became highly

mobile and able to migrate to site of the signal. Furthermore, the activation process is

accompanied by the up-regulation of several surface receptors, as well as by the

production of a plethora of bioactive molecules. [94]. The type of activation is

dependent on factors like the intrinsic properties of the stimulus, duration and exposure

to a prior or another existing stimulus [98].

Microglia are highly plastic cells and while displaying similar morphology they

may show distinct activated phenotypes. Inspired by studies in peripheral

macrophages, the activation of microglia has been generally classified in i) classical

activation, ii) alternative activation and iii) acquired deactivation [100]. Classical

activation (commonly called M1 phenotype) is induced by pro-inflammatory mediators

such as lipopolysaccharide (LPS) or interferon-γ and is characterized by the production

of numerous pro-inflammatory cytokines (e.g. TNF-α, IL-1β and IL-6), proteases and

ROS/NOS. These molecules play an important role in the defense of the organism to

pathogens, but can also damage neurons and glial cells [101]. Alternative activation

(M2a phenotype) is induced by IL-4 or IL-13 and is associated with the production of

anti-inflammatory cytokines and trophic factors such as insulin-like growth factor,

playing a role in tissue repair. [100]. Acquired deactivation (M2c phenotype) is

produced by inducing agents like transforming growth factor β and IL-10, and it is

associated with a robust suppression of the innate immune system, mainly through an

elevated production of the anti-inflammatory cytokine IL-10. [102,100]. Furthermore, the

acquired deactivation is also elicited by the exposure to apoptotic cells, permitting the

phagocytosis of these cells without triggering an immune response, which is important

to tissue maintenance [102]. Both alternative activation and acquired deactivation

down-regulate innate immune responses and demonstrate similar, but not identical,

gene profiles; therefore, some authors treat them as subgroups of a single category,

commonly called M2 or alternative activation [100] . It is also important to refer that

these phenotypes are not static. In fact, microglia is able to switch between

phenotypes, and may exist in many intermediate states (Fig. I.4).

Page 40: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

14

Fig. I.4: Microglia activation states. Depending on the stimulus, surveying microglia might respond with

different activation profiles. Lipopolysaccharide (LPS) is responsible for inducing a classical activation of microglia, mainly through toll-like-receptor (TLR) 4, resulting in a pro-inflammatory profile characterized by an increase in the production of interleukin (IL) -6, 1β, tumor necrosis factor (TNF)-α and nitric oxide (NO). Although playing an important role in organism defense, the production of these cytokines is commonly related to neurotoxicity. In contrast, alternative activation and acquired deactivation are both associated with an anti-inflammatory, neuroprotective profile, with the former being characterized by the up-regulation of insulin-like growth factor (IGF)-1, arginine 1 and Ym1 and related to tissue repair and extracellular matrix reconstruction, while the later through the production of high levels of IL-10, is more related to a robust suppression of the immune system. Described inducer agents include IL-4 or IL-3 and IL-10 or TGF-β, respectively, which act through different cytokine receptors (CRs).

2.2. Activation of microglia in AD

2.2.1. Interaction of amyloid beta with microglia

Studies evidence the presence of activated microglia clustered near amyloid

plaques in AD patients [103], as well as in AD mouse models [104]. This clustering

might be explained by chemotactic signaling mediated by A, or by other molecules

found to be associated with plaques, such as complement factors or chemokines [105].

Recently, it was visualized by in vivo imaging techniques that microglia migrate to

newly formed plaques within 1-2 days [106]. The number and the dimension of

Page 41: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

15

microglia around amyloid plaques increase in proportion to the size of plaques, and it

was suggested that plaque-associated microglia can regulate plaque dynamics [107].

Accordingly, microglia activation has been considered an early event in the

pathogenesis of AD [108].

Several triggers for microglial activation are present in AD, which might include

either molecules released from damaged neurons or even Aβ [109]. Indeed, Aβ was

demonstrated to induce microglial activation and to be one of the main triggers of

microglial response in AD [110]. Both oA and fA are able to bind and activate

microglia through several receptors, including a number of different PPRs [111] (Fig.

I.5). In this context, toll-like receptors (TLR) appear to be particularly important. Among

the cell-surface TLRs present in microglia, TLR2 and TLR4 were shown to be involved

in Aβ-induced microglial activation [111]. Furthermore, several studies in vitro and in

vivo have confirmed that TLR4 mediates the neurotoxicity induced by microglia

[112,113]. Besides TLRs, it was demonstrated that receptors like RAGE, scavenger

receptors (e.g. cluster of differentiation 36, CD36) or formyl peptide receptors (FPR; eg.

FPR-like 1) are also involved in Aβ-induced microglial activation[111]. The binding of

Aβ to these receptors is not only responsible for the secretion of inflammatory

molecules through the activation of signaling cascades such as nuclear factor-κB (NF-

κB), but also for inducing phagocytosis [114,111]. Worth to note that, although they

both stimulate a pro-inflammatory activation, oAβ and fAβ present different activation

profiles [115], and oAβ revealed to be a stronger M1-inducter than fAβ [116].

Page 42: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

16

Fig. I.5: Amyloid beta interaction with microglial receptors. Microglia cells express a panoply of

surface receptors with which they are able to sense their environment and respond to stimuli. Amyloid beta (Aβ) is able to interact with several of these receptors, inducing microglial activation. Among the array of receptors described to be involved in Aβ binding are: toll-like receptors (TLR), like TLR2 and TLR4; scavenger receptors (SRs), like CD36; formyl peptide receptors (FPR); and receptor for advanced glycation endproducts (RAGE). The binding of Aβ to microglial receptors elicits a pro-inflammatory response, with the release of inflammatory cytokines like Interleukin (IL)-1β, IL-6 and tumor necrosis factor (TNF)-α, probably through activation of the nuclear factor- κB (NF-κB) signaling pathway, but also seems to induce phagocytosis, which is mediated, at least in some cases, by initiating the Src-Vav-Rac signaling cascade

2.2.2. Production of inflammation-related factors by amyloid beta-stimulated microglia

As aforementioned, challenging of microglia with A results in a pro-

inflammatory response. Inflammatory cytokines like IL-1β, IL-6 and TNF-α have all

been reported to be released by A-induced microglia [117]. IL-1β and TNF-α, for

instance, can directly injure neurons when at elevated levels [118]. These cytokines

may act in an autocrine manner further inducing cytokine production, while also

activating astrocytes [17]. Indeed, inflammatory cytokines have been reported to be

increased in vulnerable brain regions during AD and where found to be associated with

plaques [62]. Interaction of microglia with Aβ also leads to the secretion of chemokines,

such as monocyte chemotactic protein-1 (MCP-1, also known as CCL2) and

macrophage inflammatory protein-1α (MIP-1α, also known as CCL3) [119,120].

Chemokines play an important role in the recruitment of microglia and astrocytes, and

an increased expression of the chemokines receptors CCR3 and CCR5 was found in

microglia associated with Aβ deposits [121].

Other inflammatory-related molecules produced by microglia are ROS and

NOS. In this regard, stimulation with Aβ, for instance, was demonstrated to lead to the

Page 43: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

17

release of NO and superoxide radical [90]. Although these molecules have an

important role in the defense against pathogens, high levels are detrimental and might

contribute to AD, as it was discussed before. Derivatives of APP, including Aβ, were

also reported to induce the release of glutamate from microglia cells [122,123].

Glutamate is an excitatory amino acid that plays an important role in synaptic plasticity,

memory and learning, but it can also be highly toxic to neurons, inducing excitotoxicity,

which consists on overstimulation of glutamate receptors like NMDA receptors, leading

to a massive influx of extracellular Ca2+ that damages cell structures and components

[124,125]. However, these reports are questioned by recent findings indicating that Aβ

might not affect glutamate release [109]. Also an important molecule that was recently

reported to be released by Aβ-stimulated microglia is ATP, which was linked to ROS

production by microglia [126].

Several proteases are released by microglia, contributing to their functions.

MMPs, for example, which are essential to the maintenance and reconstruction of

extracellular matrix, were indicated to be involved in AD [127]. MMP-3, -12, -13 and -9

have all been reported to be released by Aβ-stimulated microglia [128,129]. Moreover,

both MMP-9 and MMP-2 seem to be able to degrade Aβ in vitro [130,131], and their

relevance to amyloid deposition was further corroborated by knockout mice [132];

however, these proteases also have a role as mediators of tissue degradation and

inflammation, damaging the blood-brain barrier (BBB) and processing pro-inflammatory

cytokines [127] .

2.2.3. Amyloid beta-induced microglial phagocytosis

Phagocytosis represents a central housekeeping function played by microglia

that is important to the elimination of neurotoxic compounds, cellular debris and

pathogens. Accordingly to the receptor that is stimulated, it may be or not coupled with

inflammation [133]. Aβ has been documented by several studies to be phagocytized by

microglia. For instance, examination of microglia at amyloid plaques using electron

microscopy has shown that they are able to engulf Aβ with their processes and indeed

Aβ was observed in endosome-like cellular compartments [134]. Moreover, in vitro

studies using labeled Aβ [135] in combination with direct injection of fAβ into rat brain

[136] further demonstrated the capability of microglia to phagocytize Aβ. The

mechanism by which Aβ is phagocytized upon interaction with microglia is dependent

on its physical and biochemical properties. Insoluble Aβ aggregates are internalized

through receptor mediated phagocytosis (e.g TLR2 and 4), while soluble Aβ has been

suggested to not be truly internalized by phagocytosis, but through macropinocytosis

Page 44: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

18

[137,138,99]. However, the question on whether or not microglia restricts Aβ growth in

AD has remained controversial, since the recruitment of microglia to plaques does not

seem to result in their degradation [139,140]. In fact, other in vitro studies indicate that

although microglia is able to phagocytize Aβ the cells may have a limited capacity to

degrade the peptide [141]. Recently, it was even demonstrated that ablation of

microglia does not significantly affect the formation and maintenance of amyloid

plaques and that this might be mediated through the recruitment to the brain of

monocyte-derived macrophages [142], which present a better capacity for Aβ

degradation [143]. The infiltration of monocyte-derived macrophages is thought to only

occur under exceptional conditions, as in the case of CNS disorders such as AD

[144,139]. Nevertheless, this feature is still controversial and to that accounts the

difficulty in distinguishing macrophages from activated microglia based on known

immunophenotypic markers due to their similarity [139].

Recently, besides a possible role in Aβ clearance, microglia was demonstrated

to be able to phagocytize viable neurons and synapses when low (at the range of low

nM) concentrations of Aβ are present in the culture [145]. In line with these findings, Aβ

has also been shown to induce the extracellular exposure of phosphatidylserine (PS)

on neurons [146]. The PS is the best known of “eat-me” signals, which represent a set

of indicators recognized by microglia and evidenced by cells needing to be

phagocytized. Usually, under normal conditions, PS is found in the inner leaflet of cell

plasmatic membrane. However, as a result of several factors, PS may become

externalized and, therefore, it might bind to specific microglial receptors, either directly

or indirectly, inducing the phagocytosis of the cell [147]. One example is the binding to

vibronectin receptor αvβ3 (VR) through the bridging protein milk fat globule EGF-like

factor 8, also known as MFG-E8. [147]. Interestingly, both in AD and MCI it seems to

occur an increase in the number of PS-exposing neurons [148]. Furthermore, the level

of MFG-E8 was indicated to be reduced in AD [149], possibly due to phagocytosis and

degradation, while the β3 subunit of VR was shown to be up-regulated on active

microglia in AD [150]. Nonetheless, MFG-E8 also contributes to Aβ phagocytosis and

its reduction was found more accentuated near amyloid plaques [151]. Thus, it remains

to be determined whether in AD MFG-E8 mediates the phagocytosis of Aβ or of PS-

exposing neurons.

2.3. Microglial role in the progression of AD: activation vs. dysfunction

The involvement of microglia in AD progression is still enigmatic. Aβ is a pro-

inflammatory stimulator of microglial activation. This, in association with the presence

Page 45: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

19

of activated microglia near amyloid plaques, led to the proposition that Aβ deposition

might act as a persistent activator for microglia [152]. As a result, microglia might then

become over-activated and produce pro-inflammatory mediators that act directly on

neurons, leading to their damage, while also activating more glial cells, which further

amplify inflammatory signals [153]. The uncontrolled inflammation may drive the

progression of AD, exacerbating and stimulating neuronal death [154]. However, it

might be that in early stages of the disease the role of microglia is protective and not

detrimental, since microglia assume a more phagocytic phenotype and might be able

mediate Aβ clearance by phagocytosis or by secreting enzymes involved in its

degradation, like IDE and NEP [155,62]. In another point of view proposed by Streit

and Xue [156], instead of a gain in inflammatory function, microglia, which above all

things are beneficial cells, might become dysfunctional. In such circumstances,

microglia lose their capacity to support neurons and do not produce trophic factors, like

bone-derived neurotrophic factor, NGF and glial-derived neurotrophic factor. Moreover,

they are not able to clear out neuronal debris and waste residues, or to eliminate toxic

substances, such as Aβ or glutamate in excess. All this findings together will trigger

neurodegeneration [156]. One possible explanation these alterations in normal

microglia function that contribute to the disease might result from ageing [157]. Indeed,

in many ways, the aged brain differs from the young one and this also includes

microglia.

Microglia in the aged brain have been described as presenting less and smaller

branches [158], with an increase in activation markers, such as major histocompatibility

complex (MHC) II and CD11b, as well as an elevation in the basal production of pro-

inflammatory cytokines [159]. Accumulating evidence indicate that microglial

responses to perturbations in CNS also seem to be dysregulated as a result of ageing.

For instance, studies using several animal models of inflammatory challenge

demonstrate that responses by aged animals are in general larger and increasingly

sustained, leading to more pronounced and deleterious effects when compared to

young animals [92]. It has been indicated that this is might be due to a microglial

priming by inflammatory stimuli along life, causing a phenotypic shift of these cells to a

more sensitized state, which results in prolonged and exaggerate responses to

subsequent inflammatory stimuli [159,160]. These features are paralleled in human

aged brain by the appearance of microglia presenting a “dystrophic” morphology, which

was characterized by deramification and abnormalities at cytoplasmic level, such as

atrophy and cytoplasmic fragmentation. Although dystrophic microglia may be

observed in aged brain, they are rarely found in young brain, what conducts to the

hypothesis of microglia senescence and, consequently, degeneration, impairing their

Page 46: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

20

capacity to support neurons [161]. The finding that telomere shortening, which is a sign

of senescence, occurs in microglia overtime, both in vitro and in vivo further

corroborates such idea [162,163].

The discovery that microglia suffer alterations with ageing, although adding

another degree of complexity to the subject, might help to clarify their role in the

progression of AD. Dystrophic microglia have been reported to be more prevalent in

AD, accompanying the disease progression and to be associated to its pathological

features [164]. Such finding indicates that microglia senescence is related to AD and

that microglia suffers an accentuated deterioration during the disease process

[165,164]. Indeed, Aβ seems to be able of inducing microglia degeneration [166];

therefore, it may be that the loss of neuroprotection by microglia during ageing is

aggravated by the presence of Aβ, and that the lack of neuronal support leads to

neuronal demise [167]. On the other hand, due to lifelong induced priming, some

microglia might also respond abnormally to the pathogenic stimuli that arise during AD,

over-expressing neurotoxic factors, further contributing to neuron degeneration [165].

These age-related alterations might also help to elucidate why the use of NSAIDs,

which have microglia as one their targets, only produce significant effects at early

stages of AD[168]. It may be then expected that the microglia already primed and

producing an exacerbated response to stimuli, may be restrained by the use of

NSAIDs, while the senescent microglial and cell latency during disease progression will

lead to a point where therapeutical intervention has no benefits.

2.4. Microglia-neuron interplay in AD

In the last years, the idea that neurons are just passive targets of microglia

uncontrolled actions suffered an important change. In fact, studies indicate that

neurons play a very active role in microglial regulation by releasing “on” and “off”

signals [169]. Furthermore, it was proposed that an imbalance of this regulation might

be a possible mechanism for the age-related derangements and vulnerabilities to

neurodegenerative processes [159] (Fig. I.6).

Page 47: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

21

Fig. I.6: Microglia-neuron communication. (A) Under steady-state conditions, microglia are maintained

in a surveillant state, presenting an extensively ramified morphology. This phenotype is maintained, in part, through “off” signals constitutively expressed by neurons, such as CX3CL1 and CD200, which act by interacting with their correspondent receptors expressed by microglia. In case of damage, these signaling pathways might become disrupted, inducing microglial activation. (B) Damaged neurons are also responsible for secreting several factors to signal microglia on the need of assistance, and inducing their recruitment and activation. Signals like high-mobility group box 1 (HMGB1) act by binding to toll-like receptors (TLR)2 and 4 or receptor for advanced glycation products (RAGE) to induce the release of pro-inflammatory compounds. Adenosine triphosphate (ATP) is another released factor that may bind to receptors like P2X7 and induce the transcription of pro-inflammatory cytokines. Glutamate, on the other hand, acts on ionotropic and metabotropic glutamate receptors (mGluR) in microglia, leading to a neurotoxic or neuroprotective role, which is dependent on the receptor stimulated.

“On” signals are released by damaged neurons and are responsible for

inducing a defined microglial activation program [169]. One of the most interesting “on”

signals is the high-mobility group box 1 (HMGB1). This alarmin is usually present in the

nucleus, where it functions to stabilize DNA structure and modulate transcriptional

activity. Nevertheless, during activation or cell necrosis, HMGB1 elevates in the

cytoplasm and is released to the extracellular space, mediating inflammatory

responses through binding to receptors such as RAGE and TLR2-4, which seems to

conduct to memory impairment [170,171]. In AD, the levels of HMGB1 are increased,

primarily associating with amyloid plaques [172]. In addition, this protein was reported

Page 48: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

22

to inhibit microglial phagocytosis of Aβ [173], suggesting that it may play a role in

preventing Aβ clearance. Other signals also released by damaged neurons are purines

and neurotransmitters [169]. Purines, such as ATP, are detected by the nucleotide

receptors P2Y and P2X in microglia [174]. Different studies indicate that purines are

able to trigger various microglial responses, playing an important role in the activation

of microglia and inducing their migration to the lesion [175,176]. Moreover, it was

recently demonstrated that microglia reacts to extracellular ATP by releasing ATP

themselves, thus providing a positive feedback mechanism to attract more microglia to

the site of injury [177]. The neurotransmitter glutamate also seems to play an important

role in neuron-microglia communication. In fact, although glutamate directly leads to

neuronal death, it also activates microglia, which express both ionotropic and

metabotropic glutamate receptors [94]. Depending on the receptor stimulated in

microglia, glutamate binding might result in neurotoxicity or neuroprotection [169]. Both

ATP and glutamate are intimately connected to neuroinflammation and, therefore,

might also play an important role in AD.

However, probably more important to the context of microglial ageing are the

“off” signals. This type of signals is constitutively expressed in the healthy, normal

brain, being responsible for down-regulating microglial activation [169]. Two “off”

signals that have received great attention in the literature are CD200 and CX3CL1

(also known as fractalkine). CD200 is a neuronal membrane protein that interacts with

its cognate receptor CD200R which is expressed by microglia to deliver regulatory

signals that result in the inhibition of microglia inflammatory phenotype [178].

Interestingly, the axis CD200-CD200R seems to be decreased in the AD brain [179].

Similarly to the CD200-CD200R pair, CX3CL1, a transmembrane chemokine

expressed by neurons in the brain, is also able to interact with its specific

correspondent receptor CX3CR1, which is expressed mainly by microglia in the CNS

and also able to restraining microglial activation [178]. CX3CL1 can be cleaved from

the cell membrane by the ADAM10 and ADAM17, as well as by cathepsin S [180]. Its

soluble form is of crucial importance for chemotaxis, being involved in the recruitment

of microglia [178]. In relation to AD, the role of CX3CL1-CX3CR1 signaling is still

controversial. For instance, some studies in AD mice indicate that the disruption of this

signaling pathway might reduce amyloid deposition, probably by altering microglial

activation and their phagocytic capability [181].On the other hand, it was also reported

that the absence of CX3CR1 protects against neuronal loss [182]. Still, other studies

indicate that the disruption of this signaling enhances tau pathology [183], as well as

inflammation and plaque-independent neuronal dysfunction [184]. Therefore, the

precise contribution of this signaling to AD is yet to be determined. Noteworthy, CD200

Page 49: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

23

and CX3CL1, as well as CX3CR1, are all decreased by ageing, representing a loss of

communication between neurons and microglia and may be associated with the

exaggerated immune responses observed in the elderly [159].

2.5. Vinyl sulfones and inflammatory diseases: a possible therapeutic

approach for microglial modulation in AD?

Sulfones are a core functional group of interest in both organic and medicinal

chemistry because of their versatile synthetic utility and due to their role as inhibitors of

various types of enzymatic processes. More specifically, vinyl sulfones are well known

for their ability to irreversible inhibit cysteine proteases, including cysteine cathepsins

like cathepsin B and S [185,186]. Cathepsins are essentially endosomal/lysosomal

proteases involved in physiological protein turnover. Nonetheless, increasing evidence

indicates that some cathepsins also possess other functions [187]. For instance,

cathepsins S is preferentially expressed by cells of mononuclear phagocytic origin, like

microglia, and plays a role in MHCII antigen presentation [188]. Moreover, some

members, such as cathepsins S and cathepsins B, are also secreted upon activation of

microglia with inflammatory mediators and seem to be involved in extracellular matrix

degradation and neuronal death [189]. Notably, cathepsin S is particularly stable at

neutral pH [180].

Increasing evidence indicates that disturbance of normal balance and

extralysosomal localization of cathepsins contributes to neurodegenerative diseases,

like AD [190]. Cathepsin S that plays an important role in microglial activation

maintenance was found overexpressed in AD [191]. Furthermore, cathepsins such as

cathepsins S and cathepsins B have also been suggested to play a role in Aβ

formation, since specific inhibitors for these cathpesins were shown to reduce the level

of Aβ in cells and animals [192], connecting these proteases with AD. Cathepsin S was

recently reported to be directly associated with the ageing process [188], which may be

relate with the exacerbated inflammatory state observed in early stages of AD,

resulting from its role in modulating CX3CL1-CX3CR1 signaling. Taken these concepts

together, cathepsins deserve further evaluation as therapeutic targets to develop

disease modifying drugs to treat AD.

Page 50: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

24

2.6. Microglial in vitro models for studying AD

Cell cultures systems have been an appealing method for studying microglial

biology. These systems present a valuable tool to study the activation state, releasable

factors, mobility and other important characteristics of microglia functionally, which

cannot be appropriately examined in vivo [193]. Between the most used cell culture

systems, are the primary microglial cultures and microglial cell lines like BV2 and N9.

The primary cultures are a common choice in inflammatory research due to

their similarity with in vivo cells. They are often derived from the cortex of mouse or rat,

either before or shortly after birth [193]. This type of experimental model is the one that

most closely mimics the endogenous cells and has been greatly used to study

microglial function in AD, mainly at the level of Aβ-related effects [194,195]. However,

due to the time consuming techniques and reduced yield, many studies have utilized

microglia cell lines instead of primary cells to investigate the role of microglia in

physiological and pathological conditions [193]. Interestingly, studies in our lab have

evidenced that cells undergo different characteristics, from young to mature and

senescent, accordingly to the time in culture (in vitro ageing) [196].These findings were

recently observed by us in regard to microglia [197] and reveal to be of high relevance

to evaluate age-related microglial changes. Indeed, up to now there is no suitable

model to obtain reproducible age-related microglia, even from murine models at

different ages, due to the isolation processes limitations

The microglial cell line N9 is derived from mouse and was immortalized using

an oncogene-carrying retrovirus [198]. The usage of a cell line is less expensive and

less time consuming than the use of primary microglia. In addition, there is also less

variations across different cultures [199]. Nonetheless, microglial cell lines are

genetically modified, which determines changes in their adhesion and proliferative

capabilities, and thus inflammatory responses not truly reproducing that of primary

cultures [200]. Still, these cultures represent a useful advantage to investigate

microglia-related phenomena in AD.

Another important approach in the study of microglia dynamics is the use of co-

culture systems, which allows the evaluation of the interactions between microglia and

other cell types. The understanding of the interaction between microglia and neurons is

of prime interest to establish the link between neuroinflammation and

neurodegenerative diseases. Therefore, combined cultures of primary neuronal and

microglial cells have been useful tools for assessing the effect of an inflammatory

response at the level of neurons from diverse brain regions. In fact, there are numerous

published studies using different co-culture approaches [201,202]. Mixed co-culture

Page 51: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

25

systems are probably one of the approaches that more closely mimic what happens in

vivo, permitting the study of both physical and soluble cell-to-cell communication

pathways [203].

Page 52: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

26

[Escre

va um

Page 53: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

27

3. Aims

The main goals of the present work are a) to understand the role of ageing in

microglial responses and b) to explore the use of a vinyl sulfone-based compound (VS)

as a modulator for microglial activation - both in the context of AD.

More specifically, the aims are to:

a) Evaluate if young and aged microglia are differently stimulated by Aβ and if ageing

also results on a differential response to neuronal derived signals. For this, we will

use an in vitro model of microglial ageing in which primary microglial cell cultures will

be cultivated for 2 and 15 days in vitro (DIV), representing respectively young and

aged phenotypes. Behavior of these different aged microglia in the presence of Aβ

will be assessed by using monocultures and mixed cultures with 17 DIV

hippocampal neurons. The parameters to be determined will be the release of

glutamate, ATP and MMPs.

b) Explore the benefits of VS to modulate microglial activation by Aβ. This will be done

by stimulating the N9 microglial cell line with Aβ, in the presence or absence of VS,

and by evaluating cell viability, the release of inflammation-related factors,

phagocytic capacity and the expression of activation-related receptors.

Collectively, the results to be obtained in this project aim to give further insights

to microglial biology in the context of AD, and hopefully, find a possible new therapeutic

approach for this still untreatable neurodegenerative disorder.

Page 54: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

28

[Escre

va um

Page 55: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

29

II. MATERIAL AND METHODS

1. Material

1.1. Reactives

1.1.1. Cell culture media

Dulbecco’s modified Eagle’s medium (DMEM)–Ham’s F12 and Minimum

Essential Medium (MEM) with Earle’s salts were acquired from Biochrom AG (Berlin,

Germany). Neurobasal medium was from Invitrogen Corporation™ (Carlsbad, CA,

USA). Roswell Park Memorial Institute (RPMI) 1640 medium was purchased from

Sigma Aldrich (St. Louis, MO, USA).

1.1.2. Supplements and chemicals

Fetal bovine serum (FBS), L-glutamine, sodium pyruvate, non-essential

aminoacids (NEAA), penicillin/streptomycin and HEPES (4-(2-hydroxyethyl)-1-

piperazineethanesulfonic acid) were purchased from Biochrom AG (Berlin, Germany).

Trypsin-EDTA (Ethylenediamine tetraacetic acid) solution (1x) Hoechst 33258, bovine

serum albumin (fraction V, fatty acid free) (BSA), tris-base, β- nicotinamide adenine

dinucleotide 2’-phosphate, ATP, phenylmethylsulfonyl fluoride (PMSF), poly-D-lysine

(PDL), fluorescent latex beads 1 µm (2.5%), Coomassie Brilliant Blue R-250, propidium

iodide [PI; 3,8-diamino-5-(3-(diethylmethylamino) propyl)-6-phenyl phenanthridinium

diiodide], antibiotic-antimycotic (Ab/Am) and Dimethyl sulfoxide (DMSO) were from

Sigma Aldrich (St. Louis, MO, USA). Trypsin-EDTA solution (0.25% trypsin and 1 nM

EDTA in Hank’s balanced salt solution), laminin and B-27 supplement were acquired

from Invitrogen Corportation™ (Carlsbad, CA, USA). L-glutamic acid kit, Triton X-100,

6-phosphateglucose dehydrogenase and hexokinase were from Roche Diagnostics

(Indianapolis, USA). Aβ42 was from EZBiolab (Carmel, IN, USA). DPX mountant for

microscopy was obtained from BDH Prolabo (Bangkok, Thailand). Cell lysis buffer®

and LumiGLO® were from Cell Signaling (Beverly, MA, USA). Sodium dodecyl

sulphate (SDS) was acquired from VWR-Prolabo. Nitrocellulose membrane was

obtained from Amersham Biosciences (Piscataway, NJ, USA). Acrylamide, bis-

acrylamide, Tween 20, glycerol, absolute ethanol, acetic acid, potassium hydroxide

(KOH), gelatin, calcium chloride (CaCl2) and perchloric acid (70%) were obtained from

Page 56: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

30

Merck (Darmstadt, Germany). BioRad’s Proteins Assay Reagent was obtained from

BioRad Laboratories (Hercules, CA, USA). All the other common chemicals used were

purchased either from Sigma-Aldrich or Merck. The VS was synthetized in Rui

Moreira’s Medicinal Chemistry unit of iMed.UL.

1.1.3. Antibodies

Primary antibodies: Rabbit anti-ionized calcium-binding adaptor molecule 1

(Iba1) was purchased from Wako (Japan). Rabbit polyclonal anti-MFG-E8 and rabbit

polyclonal anti-TLR4 were from Santa Cruz Biotechonlogy® (Santa Cruz, CA, USA).

Mouse monoclonal anti-HMGB1 was acquired from BioLegend® (San Diego, CA,

USA). Mouse anti-β-actin was obtained from Sigma-Aldrich (St. Louis, MO, USA).

Secondary antibodies: Alexa Fluor® 594 goat anti-rabbit was obtained from

Invitrogen Corporation™ (Carlsbad, CA, USA). HorseRadish Peroxidase (HRP)-labeled

goat anti-rabbit IgG was from Santa Cruz Biotechonlogy® (Santa Cruz, CA, USA).

HRP-labeled goat anti-mouse IgG was from Amersham Biosciences (Piscataway, NJ,

USA).

1.2. Equipment

Basic equipment included an Axio Scope A1 fluorescent microscope with and

adapted camera Leica DFC4900 (both from Zeiss, Gottingen, Germany). For Western

Blot and Zymography assays it was used the Mini-Protean P3 Multi-Casting Chamber

from BioRad (Hercules, CA, USA). Microplate reader (PR 2100) was also from BioRad

and it was used for spectrophotometric measurements of protein and glutamate

content. GloMax® Multi Detection System (Sunnyvale, CA, USA) was used for ATP

quantification. For MMPs gels photos and immunodetection in Western Blot

nitrocellulose membranes it was used the Chemidoc™ system, also from BioRad

Laboratories. Samples sonication for a well homogenization was performed in the

Ultrasonic Processor UP100H (Hielscher-Ultrasound Technology, Teltow, Germany).

To ensure a stable environment to optimal cell growth (37ºC and 5% CO2), cell cultures

were maintained in HERAcell 150 incubators (Thermo Scientific, Waltham, MA, USA).

The work was performed in sterile conditions in a Holten Lamin Air HVR 2460 (Allerod,

Denmark). Eppendorf 580R (Eppendorf, Hamburg, Germany) and Sigma 3K30

centrifuges were used for different experimental procedures.

Page 57: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

31

1.3. Animals

The animals used were CD1 mice. Animal care followed the recommendations

of European Convention for the Protection of Vertebrate Animals Used for

Experimental and other Scientific Purposes (Council Directive 86/609/EEC) and

National Law 1005/92 (rules for protection of experimental animals). All animal

procedures were approved by the Institutional Animal Care and Use Committee. The

best efforts were made to minimize the number of animals used and their suffering.

2. Methods

2.1. Primary cultures

2.1.1. Isolation of primary cell cultures

Primary cultures of hippocampal neurons: Primary hippocampal neurons were

isolated from E16 CD1 mice fetuses, as previously described by Lanier and colleagues

[204]. Cells (approximately 1x105 cells/ml) were plated on 24-well tissue plates coated

with PDL (100 µg/mL) and laminin (4µg/mL) in plating medium (MEM with Earle’s salts

supplemented with 10 mM HEPES, 10 mM sodium pyruvate, 0.5 mM L-glutamine, 12.5

µM glutamate, 10% FBS and 0.6% glucose). After 3 hours, the media was replaced

with neuronal growth medium (Neurobasal media supplemented with B-27 supplement

and 0.5 mM L-glutamine). Cultured hippocampal neurons were maintained under

stable growth conditions and the first replacement of the culture media was performed

after 4 DIV, with the procedure being repeated every 7 days until neurons achieve 17

DIV.

Microglia primary cultures: Mixed glial cultures, containing microglia and

astrocytes, were prepared from 1-to-2 day-old CD1 mice as previously described by us

[205]. Cells (4x104 cells/cm2) were plated on uncoated 6-well tissue culture plates in

culture medium (DMEM-Ham’s F-12 medium supplemented with 1% L-glutamine, 1%

sodium pyruvate, 1% NEAA, 10% FBS and 1% Ab/Am solution) and maintained at

37ºC in a humidified atmosphere of 5% CO2. The old medium was removed every 7

days and replaced with fresh culture medium. Pure microglia cells were then isolated

as previously described by us [206]. In brief, after 21 days in culture, microglia were

obtained by mild trypsinization with trypsin-EDTA solution diluted 1:3 in DMEM-Ham’s

F-12 medium for 45 to 60 minutes. The trypsinization resulted in a detachment of an

upper layer of cells containing all the astrocytes, whereas the microglia remained

Page 58: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

32

attached to the bottom of the well. The medium containing detached cells was removed

and replaced with the initial mixed glial-conditioned medium. After 4 days, the mixed

glial-conditioned medium was removed and replaced with fresh culture medium, being

again replaced with fresh media at every 4 days. At 2 or 15 DIV post isolation, cells

were removed by total trypsinization with trypsin-EDTA solution for 10 minutes and

seeded on 24-well tissue plates, either alone or on top of 17 DIV neurons, at a density

of 4.0x104 cells/ml. Cultures were kept for 24h before being treated.

The intent of maintaining cells for 21 days in culture before isolation is to

achieve the maximal yield and purity of microglial cultures. In fact, the astrocyte

contamination was previously verified to be less than 2%, as assessed by

immunocytochemical staining with a primary antibody against glial fibrillary acidic

protein, a specific astrocyte marker. Furthermore, neuron contamination was also

excluded, as it was previously assessed by immunocytochemical staining with a

primary antibody against MAP2, a specific marker for neurons [206]. The choose of 2

DIV and 15 DIV as time points intends to mimic a young and aged microglial

phenotype, respectively, and it is based on observations obtained in our group that

indicate that microglia experiences age-related alterations along time in culture [197].

2.1.2. In vitro treatment of primary cell cultures with amyloid beta

Microglia cells, alone or in the presence of 17 DIV primary hippocampal

neurons, were incubated at 2 or 15 DIV in the absence (control) or in the presence of

50 nM or 1000 nM of Aβ42, during 24 h at 37ºC (Fig. II.1). Same procedure was

performed on 17 DIV hippocampal neurons alone, for control purposes. We chose 50

nM since it was the lowest concentration value demonstrated to be sufficient to elicit

microglial activation [194]. The 1000 nM, on the other hand, corresponds to a

concentration range reported in the brain of AD patients [145]. The Aβ42 stock solution

(111 µM) was prepared dissolving Aβ42 to 0.5 mg/ml in DMEM-Ham’s F12 medium.

Before cell treatment, Aβ42 peptides were incubated for 24 h at 37°C to form

aggregates (Fig.II.2).

.

Page 59: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

33

Fig. II.1: Experimental procedure used in the primary cellular cultures and parameters evaluated.

Mixed glial cultures were isolated from newborn CD1 mice and maintained for 21 days in culture. After this time, microglia cells were isolated by mild trypsinization and maintained until 2 or 15 DIV. To obtain the mixed neuronal-microglial cultures, hippocampal neurons were isolated from E16 CD1 mice fetuses and maintained to 17 DIV, time when they were co-cultured with microglia, either with 2 or 15DIV. Neuronal hippocampal 17 DIV cultures were also used for control purposes. Microglial cultures, neuronal cultures and neuron-microglial cultures were maintained for 24 h before treatment. During this time, amyloid beta 1-42 (Aβ42) peptides were also allowed to aggregate. Treatment was done with 50 nM or 1000 nM Aβ42, or no addition (control), for 24h. After this period, cell media was removed and evaluated for the release of glutamate, adenosine triphosphate (ATP) and active matrix metalloproteinase (MMP)-2 and MMP-9. ATP was quantified by an enzymatic assay, glutamate by L-glutamic acid kit and MMP-2 and MMP-9 activities by gelatin zymography.

2.1.3. Quantification of extracellular ATP

The assessment of extracellular ATP levels was performed as previously

described by us [207]. The incubation media was collected and treated with 2 M

perchloric acid for 10 minutes, followed by centrifugation for 5 minutes at 10000 g, at

4ºC, to remove cellular debris. Subsequently, 4 M KOH was added for 10 minutes to

neutralize pH value and the centrifugation step was repeated. All the procedure was

performed on ice to prevent ATP degradation. ATP levels were measured indirectly by

the production of NADPH using an enzymatic assay, which consisted in the addition of

a solution containing NADP+, glucose and 6P-glucose dehydrogenase, followed by

hexokinase, the reaction starting enzyme. Fluorescence intensity was quantified using

Page 60: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

34

a fluorimeter at ʎem = 450 nm and ʎex = 340 nm. A calibration curve of ATP was used

for each assay.

Fig. II.2: Analysis of Aβ aggregation forms present in our experimental conditions. We have

previously performed this study by western blot using a monoclonal anti-amyloid beta (Aβ) antibody 6E10 and separation on 4-16% Tris-Sodium dodecyl sulphate (SDS)-polyacrilamide gel electrophoresis (PAGE). It was observed that after the aggregation procedure and incubation time our Aβ solution contains a mixture of both oligomeric and fibrillar forms[197].

2.1.4. Measurement of extracellular glutamate

The glutamate content in the incubation media of primary cultures was

determined using the L-glutamic acid kit, as previously described [205]. The reaction

was performed in a 96-well microplate and the absorbance measured at 490 nm. A

calibration curve of glutamic acid was used for each assay.

2.1.5. Gelatin zymography

The quantification of the release of active MMP-9 and MMP-2 in cell media of

primary cell cultures was performed through gelatin zymography method, in which the

protease activity is directly observed in the running gel based in the absence of color

(white bands) at the particular site of protease action. For this determination, aliquots of

the treated culture supernatant were analyzed by SDS-polyacrilamide gel

electrophoresis (PAGE) zymography in 0.1% gelatin/10% acrylamide gels under non-

reducing conditions, at 30 mA/gel. After electrophoresis, gels were washed for one

hour with 2.5% Triton X-100 (in 5 mM Tris pH 7.4; 5 mM CaCl2; 1 µM ZnCl2) to remove

SDS and renature the MMP species in gel. Posteriorly, gels were incubated overnight

in developing buffer (5 mM Tris pH 7.4; 5 mM CaCl2; 1 µM ZnCl2) to induce gelatin lysis.

For enzyme activity analysis, the gels were stained with 0.5% Coomassie Brilliant Blue

R-250 and distained with 30% ethanol/ 10% acetic acid/ H2O. Gelatinase activity,

Page 61: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

35

detected as a white band on a blue background, was photographed using Chemidoc

and analyzed using Image LabTM analysis software (BioRad laboratories, version

3.0.1). The digested bands promoted by the different MMPs are distinguished based on

the different molecular weights presented by active MMP-9 (82 kDa) and active MMP-2

(62 kDa).

2.2. Cell line

2.2.1. N9 microglial cell line culture

The N9 microglial cell results from immortalization of microglial cells obtained

from CD1 mouse cortex [198] and it was a kindly gift from Teresa Pais, Institute of

Molecular Medicine (IMM), Lisboa, Portugal. Cells were cultured (4x104 cell/ml) in

RPMI supplemented with 10% FBS, 1% L-Glutamine, 1% penicillin/streptomycin and

grown to confluence. The cells were splitted every 2 to 3 days. For each new

experiment, cells were plated (4x104 cell/ml) on uncoated 6-well or 24-well (with 12 mm

coverslips) tissue plates in the referred RPMI supplemented medium and were

maintained at 37ºC in a humidified atmosphere of 5% CO2 for 2 days before being

treated.

2.2.2. Treatment of N9 microglial cell line

Cells at 2 DIV (platting was considered day 0) were treated without or with Aβ42,

50 nM or 1000 nM for 24 hours at 37ºC, in the absence or presence of VS that was

synthetized at Rui Moreira lab from the Medicinal Chemistry unit at iMed.UL, and which

is currently indicated as a cathepsins S inhibitor (Fig.II.3). Aβ42 was aggregated as

previously mentioned for primary cell cultures. A 20 mM VS stock solution was

prepared by dissolving VS in DMSO. For treatment, VS was diluted in RPMI medium to

10 µM or 20 µM concentrations and added 10 minutes before incubation with Aβ42. We

based our concentrations choice in the values used for a known vinyl sulfone-based

cathepsin S inhibitor, termed LHVS (morpholinurea-leucine-homophenylalanine-

vinylsulfone phenyl) [208].

Page 62: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

36

Fig. II.3: Experimental procedure used in the N9 microglial cell cultures and parameters evaluated.

N9 microglia cells were grown in supplemented RPMI medium and incubations were performed at 2 days in vitro (DIV) after passage with 50 nM or 1000 nM aggregated amyloid beta 1-42 (Aβ42) peptides , or no addition (control), in the absence or presence of 10 μM or 20 μM of a vinyl sulfone-based compound (VS). After 24 h of incubation period, cells were fixed or lysed and the respective cell media was also collected. Cell viability was evaluated by propidium idodide incorporation. Phagocytic capacity was evaluated by the incorporation of fluorescent latex beads followed by immunocytochemistry with an anti-Iba1 antibody. The activity of matrix metalloproteinases (MMP)-2 and -9 was accessed by gelatin zymography. Toll-like-receptor 4 (TLR4), milk fat globule-EGF factor 8 (MFG-E8) and high-mobility group box 1 (HMGB1) were evaluated by western blot using specific counter antibodies.

2.2.3. Determination of Cell viability

Necrotic-like cell death was assessed by monitoring the cellular uptake of the

fluorescent dye PI. PI readily enters and stains non-viable cells, but leaves viable cells

unstained, since it cannot cross their cellular membrane. This dye binds to double-

stranded DNA and emits red fluorescence (emission: 630 nm; excitation 493 nm).

After the incubation period, unpermeabilized adherent cultured cells on

coverslips were incubated with 75 µM PI solution diluted in RPMI for 15 minutes at

37ºC, in the absence of light. Subsequently, the medium was removed and the cells

were fixed with freshly prepared 4% (w/v) paraformaldehyde in phosphate buffered

saline (PBS). The nuclei were stained with Hoechst dye 33258 (1:1000, in PBS) during

2 minutes and coverslips were mounted on DPX. Fluorescence was visualized using

an Axio Scope A1 fluorescent microscope with and adapted camera Leica DFC4900.

Red-fluorescence and U.V. images of ten random microscopic fields were acquired per

sample and the percentage of PI positive cells was calculated.

Page 63: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

37

2.2.4. Microglia phagocytic capacity

After incubation, to evaluate the phagocytic ability of N9 microglia, adherent

cultured cells on coverslips were incubated with 0.0025% (w/w) 1 µm fluorescent latex

beads for 75 minutes at 37ºC, in the absence of light. Subsequently, the medium was

removed and the cells were fixed with freshly prepared 4% (w/v) paraformaldehyde in

PBS. Cells were immunostained with rabbit anti-Iba1 followed by Alexa Fluor 594 and

the nuclei stained with Hoechst dye 33258. Fluorescence was visualized using an Axio

Scope A1 fluorescent microscope with and adapted camera Leica DFC4900. U.V,

green and red-fluorescence images of ten random microscopic fields were acquired.

Total number of beads internalized was counted to determine the mean number of

ingested beads per cell.

For immunostaining, cells were first permeabilized with 0.1% TritonX-100 in

PBS for 30 minutes and then incubated with blocking solution (PBS with 1% BSA, 0.4%

Triton X-100 and 4% FBS) for 30 minutes. Cells were incubated overnight at 4ºC with

rabbit anti-Iba1 (1:250) diluted in blocking solution. The secondary antibody used was

the goat anti-rabbit Alexa Fluor 594 (1:1000), which was diluted in blocking solution

and incubated for 2 h at room temperature. For nuclei staining, cells were stained with

Hoechst dye 33258 (1:1000, in PBS) during 2 minutes.

2.2.5. Gelatin zymography

The evaluation of the release of active MMP-2 and MMP-9 in N9 microglial cell

line was performed by gelatin zymography as previously described to the primary cell

cultures (section 2.1.5).

2.2.6. Western blot analysis

Western blot was carried out as usual in our lab [209]. Briefly, total cell extracts

from the 6-well tissue plates were obtained by lysing the cells with ice-cold Cell Lysis

Buffer plus 1 mM PMSF for 10 minutes on ice and with shaking, followed by sonication

during 40 seconds. The lysate was centrifuged at 14,000g for 10 minutes at 4ºC and

the supernatants were collected and stored at -80ºC. Protein concentration was

determined using the Bradford method [210], with the BioRad’s protein assay reagent.

Equal amounts of protein content were subjected to 12% SDS-PAGE, running with

fixed amperage of 30 mA/gel. After running the gel, proteins were transferred to a

nitrocellulose membrane, and blotted membranes were rinsed once with TBS-T (10 nM

Tris-HCl (pH 7.6), 150 mM NaCl and 0.1% Tween 20) and blocked for 1 hour at room

temperature with 5% milk in TBS-T. Membranes were then incubated overnight at 4ºC

Page 64: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

38

with the following primary antibodies diluted in TBS-T with 5% BSA: anti-HMGB1

mouse antibody (1:500); anti-TLR4 rabbit antibody (1:200); anti-MFG-E8 rabbit

antibody (1:200). After washing with TBS-T, the membranes were incubated for 1 hour

at room temperature with HRP-labeled anti-rabbit or anti-mouse antibody (both

1:5000), as appropriate, diluted in blocking buffer. After washing membranes with TBS-

T, chemiluminescent detection was performed by LumiGLO® and bands were

visualized using Chemidoc. The relative intensities of protein bands were analyzed

using the Image LabTM analysis software (BioRad laboratories, version 3.0.1). All the

results obtained were normalized to β-actin, which was determined using an anti-β-

actin mouse antibody (1:5000 in TBS-T with 5% BSA) followed by a HRP-labeled anti-

mouse antibody (1:5000 in blocking buffer).

2.3. Statistical analysis

Results from at least two different experiments were expressed as mean ± SEM.

Statistical analysis was performed using two-tailed Student’s t test, on the basis of

equal or unequal variance, or one-way ANOVA using GraphPad Prism 5 (GraphPad

Software inc, version 5.00), as appropriate. It was considered p<0.05 as statistically

significant.

Page 65: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

39

III. RESULTS

1. Characterization of young and aged microglia reactivity to Aβ

and of modulation by neurons

Microglia have several fundamental functions in the CNS and may mediate

either beneficial or detrimental effects [101]. Changes in normal microglial function

have been vastly implied in AD. As advanced age is the largest risk factor for the

disease, it has been hypothesized that microglia age-related changes may drive its

pathogenic progression through a diminution of neuroprotective functions and

dysregulated responses to signals and perturbations that increase neurotoxicity

[156,165,167]. To dissect whether microglial ageing might mediate an altered reactivity

to Aβ, we developed an in vitro model using cultures of primary microglial cells grown

till 2 and 15 DIV, in order to mimic young and aged cells, respectively. Since neurons

are major players in modulating the microglial reactivity, we decided to give another

dimension to our model and we also studied how microglia behaves in their presence.

The alterations in microglial reactivity were addressed by determining the release of

glutamate, ATP and MMPs

1.1. Young cells release higher levels of glutamate than older ones when in

monoculture, but not in mixed culture with neurons, and no alterations

are produced by Aβ in each differently aged cell

Glutamate release by activated microglia is a well described event in the

literature, being of special importance to the context of neurodegenerative diseases,

since the excess of glutamate may contribute to the neuronal death observed in these

disorders [109]. As evidenced in Figure III.1, we observed that glutamate released by

aged microglia in control conditions was significantly lower than by young microglia

(0.43-fold, p<0.01). Interestingly, when young microglia was in the presence of neurons

there was a remarkable reduction in glutamate levels (0.54-fold, p<0.01), a feature that

was not observed in mixed cultures of aged microglia with neurons. In all the cultures,

the exposure to Aβ did not lead to significant alterations.

Page 66: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

40

1.2. ATP release by microglia in response to Aβ is not affected by ageing or

neurons

ATP has been reported to be released from damaged neuronal cells, being a

major factor in the migration of microglia to sites of injury [177]. Microglia have also

been reported to release ATP when stimulated with Aβ [211], which may be involved in

mediating the recruitment of more microglia to amyloid deposits. Therefore, next it was

evaluated the release of ATP. Once again, no clear effects were observed by cells

exposed to Aβ (Fig. III.2). Moreover, there were no significant alterations caused by the

presence of neurons in both young and aged microglia and no differences in the ATP

release between differently aged cells.

Fig. III.1: The extracellular levels of glutamate are higher in young microglia than in old microglia, being reduced in the first by the presence of neurons and no effect by the presence of Aβ could be observed in any of the cases. Microglia (MG) with 2 and 15 days in vitro (DIV) were cultured in the

absence or presence of 17 DIV neurons (N) and incubated with 50 nM and 1000 nM of amyloid beta (Aβ), or no addition (control). Aliquots of the culture media were collected at 24 h after incubation, and glutamate content was assessed by the L-glutamic acid kit. Results are mean ± SEM from at least two independent experiments performed in duplicate, and presented as fold change relatively to the control of MG (2 DIV), assumed as 1.

$$p<0.01 and

$p<0.05 vs. 2 DIV MG at the same experimental condition.

Page 67: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

41

Fig. III.2: Microglia release of ATP by Aβ stimulation does not change by age or by incubation with neurons. Microglia (MG) with 2 and 15 days in vitro (DIV) were cultured in the absence or presence of 17

DIV neurons (N) and incubated with 50 nM and 1000 nM of amyloid beta (Aβ), or no addition (control). Aliquots of the culture media were collected at 24 h after incubation, and adenosine triphosphate (ATP) release was measured by an enzymatic assay. Results are expressed as mean ± SEM from at least two independent experiments performed in duplicate, and presented as fold change relatively to the control of 2 DIV MG, which was assumed as 1.

1.3. Release of MMP-9 and MMP-2 has opposite profiles in young and aged

microglia and is not altered by Aβ or by the presence of neurons

Increasing evidence indicates a link between the activation of MMPs and AD

pathophysiology, in which both detrimental and beneficial roles have been suggested

[212]. We decided to evaluate if the release of active MMP-2 and MMP-9 was altered

with microglial ageing in response to Aβ and neuronal signals. By gelatin zymography,

we were able to detect two different bands, at 62 kDa and 82 kDa, corresponding to

MMP-2 and MMP-9, respectively (Fig. III.3). We observed an age-dependent increase

in the activity of MMP-2 (2.9-fold, p<0.05); however, neither young nor aged microglia

presented any responsiveness to Aβ or to neurons. Conversely to MMP-2, the activity

of MMP-9 presented a marked reduction with age (0.4-fold, p<0.01). Nevertheless, no

alterations were produced by the exposure to Aβ or by the presence of neurons.

Page 68: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

42

Fig. III.3: MMP-2 release from microglia is up-regulated by ageing, while MMP-9 is instead down-

regulated and these features are not modulated by Aβ or neuronal signaling. Microglia (MG) with 2 and 15 days in vitro (DIV) were cultured in the absence or presence of 17 DIV neurons (N) and incubated with 50 nM and 1000 nM of amyloid beta (Aβ), or no addition (control). Aliquots of the culture media were collected at 24 h after incubation and used to assess matrix metalloproteinases (MMPs) activity by gelatin zymography assay. (A) Active forms of MMP-2 and MMP-9 were identified by their apparent molecular mass of 62 kDa and 82 kDa, respectively. Representative results from one experiment are shown. (B) and (C) Graph bars represent the intensity of the bands for each MMP quantified by scanning densitometry. Results are expressed as mean ± SEM from at least two independent experiments performed in duplicate, and presented as fold change relatively to the control of 2 DIV MG, which was assumed as 1.

$$p<0.01

and $p<0.05 vs. 2 DIV MG at same the experimental condition.

A

B

C

Page 69: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

43

2. Exploring a new therapeutic approach to prevent Aβ -induced

microglial activation

Microglia react to diverse stimuli by becoming morphologically and functionally

activated, which might lead to more protective or detrimental responses, depending on

the molecules involved [94]. In AD, one of the main microglial activators is Aβ, which

induces a classical, more inflammatory phenotype, that is thought to pave the way to

the chronical inflammatory state observed in the disease [213]. Therefore, in an

exploratory approach, we decided to evaluate the suitability of VS, which is currently

indicated as a cathepsin S inhibitor, to modulate Aβ-induced microglial activation. For

that, N9 microglial cell line was used. This cell line is obtained by immortalization of

CD1 mice cortex microglia and shows diverse features similar to microglia in primary

cultures, such as migration, phagocytosis and inflammation-related features. [214-216].

2.1. Loss of viability in Aβ-stimulated microglia is slightly prevented by VS

Since a new compound will be tested, the evaluation of toxicity is of special

importance. Toxicity of VS was determined by the number of necrotic cells, through

measurement of PI uptake. As evidenced in Figure III.4, we observed a tendency to a

concentration-dependent decrease in the viability of microglia cells exposed to Aβ. The

presence of VS did not increase toxicity and, interestingly, slightly reduced the

neurotoxic effects of the highest concentration of Aβ used. Nevertheless, no statistical

significance was obtained.

Page 70: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

44

Fig. III.4: Cell viability shows a trend to be decreased by Aβ and was slightly prevented by VS. N9

cells were treated for 24 h with 50 nM and 1000 nM amyloid beta (Aβ), as in previous experiments, but now in the presence or absence of 10 μM and 20 μM vinyl sulfone (VS). Cells not-treated with Aβ served as controls. After incubation, cells were treated with 75 µM of the fluorescent dye propidium iodide (PI) for 15 min. (A) Uptake of PI was assessed by fluorescent microscopy and the percentage of PI-positive cells was counted. Results are mean ± SEM from at least four independent experiments. (B) Nuclei were stained with Hoechst dye (blue) and necrotic cells were labeled for PI (red). (C) Representative results of one experiment are shown. Scale bar represents 40 μm.

A

B

C

Page 71: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

45

2.2. Release of active MMPs by microglia is enhanced upon Aβ stimulation

and decrease by VS

MMPs are important inflammatory mediators released by activated microglia

and might play an important role in AD onset and progression, as previously

mentioned. Thus, we evaluated if the release of MMP-2 and MMP-9 were induced by

Aβ and modulated by VS (Fig.III.5). We observed that the release of active MMP-2

increased significantly upon Aβ exposure, in a concentration-dependent manner (1.1-

and 1.4-fold for Aβ 50 and 1000 nM; p<0.05 and p<0.01, respectively). In regard to VS

alone, none of the two concentrations used evidenced to produce significant

alterations. However, the effects caused by 1000 nM of Aβ on microglia were reverted

to near control values in the presence of 20 µM VS (0.9-fold, p<0.05). The same

pattern of results was observed for the release of active MMP-9. The exposure to Aβ

induced a significant increase in the MMP-9 release by microglia in a dose-dependent

manner (1.1- and 1.6-fold for Aβ 50 and 1000 nM, respectively, p<0.01). No alterations

were induced by both concentrations of VS. In addition, considering the promising

modulatory effects by VS on Aβ, the highest dose of 20 µM caused a significant

decrease in MMP-9 (1.1-fold, p<0.05) release in response to 1000 nM Aβ.

Page 72: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

46

Fig. III.5: Stimulation of microglia with Aβ increases the release of active MMP-2 and MMP-9, which is prevent by VS. N9 cells were treated for 24 h with 50 nM and 1000 nM amyloid beta (Aβ), as in

previous experiments, but now in the presence or absence of 10 μM and 20 μM vinyl sulfone (VS). Cells not-treated with Aβ served as controls. (A) Matrix metalloproteinases (MMP)-2 and MMP-9 were identified by their apparent molecular mass of 62 kDa and 82 kDa, respectively. Representative results from one experiment are shown. (B) and (C) Graph bars represent the intensity of the bands quantified by scanning densitometry. Results are expressed as mean ± SEM from at least five independent experiments performed in duplicate, and presented as fold change relatively to control, which was assumed as1. **p<0.01 and *p<0.05 vs. respective control;

$p<0.05 vs. same experimental condition in the absence of

VS.

2.3. Expression of TLR4 is reduced by the highest concentration of Aβ, an

effect that is counteracted by VS

The TLR4 receptor is one of the best described receptors and referred to be

involved in Aβ-induced microglial activation [111]. Therefore, its expression was an

important point to be here addressed. The determination of TLR4 expression was

A

B

C

Page 73: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

47

performed by western blot analysis of N9 cell lysates. We observed a significant

decrease (0.7-fold, p<0.01) in the expression of TLR4 after exposure to Aβ 1000 nM

(Fig. III.6). Surprisingly, also VS revealed to decrease the expression of TLR4,

although only the highest concentration produced significant effect (0.5-fold, p<0.01).

To note, however, that this same concentration was able to even significantly increase

the expression of TLR4 when in the presence of 1000 nM (1.4-fold, p<0.05).

Fig. III.6: Expression of TLR4 is reduced by both Aβ and VS, but revealed to be induced by the highest VS concentration when in the presence of 1000 nM Aβ. N9 cells were treated for 24 h with 50

nM and 1000 nM amyloid beta (Aβ), as in previous experiments, but now in the presence or absence of 10 μM and 20 μM vinyl sulfone (VS). Cells not-treated with Aβ served as controls. After treatment, total cell lysates were analyzed by western blot with an antibody specific for toll-like receptor 4 (TLR4). (A) Representative results from one experiment are shown. (B) The intensity of the bands was quantified by scanning densitometry, standardized with respect to β-actin protein. Results are expressed as mean ± SEM from at least two independents experiments, and presented as fold change relatively to control, which was assumed as 1. **p<0.01 vs. respective control;

$$p<0.01 and

$p<0.05 vs. same experimental

condition in the absence of VS.

2.4. Aβ triggers a reduction in microglia phagocytosis, which is prevented by

VS

Phagocytosis is a fundamental feature of functional microglia, being essential to

the removal of dead cells and microbes, as well as misfolded protein aggregates [217].

To evaluate if the phagocytic ability of microglia cells was modified in the presence of

Aβ and modulated by VS, we determined the capacity of microglia in ingesting latex

beads. Interestingly, as depicted in Figure III.7, the number of latex beads per cell was

A

B

Page 74: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

48

significantly reduced by Aβ, mainly at the highest concentration (0.7-fold, p<0.01). In

regard to VS, although not significantly, the compound revealed ability to reverse the

effects induced by Aβ to close control values

Fig. III.7: Microglia phagocytic ability is decreased upon exposure to Aβ and restored by VS. N9

cells were treated for 24 h with 50 nM and 1000 nM amyloid beta (Aβ), as in previous experiments, but now in the presence or absence of 10 μM and 20 μM vinyl sulfone (VS). Cells not-treated with Aβ served as controls. At the end of incubation, cells were treated with 1 μm green latex beads as described in methods. (A) Microglial cells were counterstained with an antibody raised against Iba-1(red) and nuclei were stained for Hoechst dye (blue). Latex beads (green) ingested are pointed out by a white arrow. Scale bar represents 40 μm. (B) Results are the mean number of ingested beads per cell ± SEM from at least four independent experiments, presented as fold change relatively to control, which was assumed as 1. **p<0.01 and *p<0.05 vs. respective control.

A

B

Page 75: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

49

2.5. The expression of MFG-E8 is reduced by 50 nM Aβ and VS alone, while

increases when the highest concentrations of both compounds are

concomitantly used

The MFG-E8 is a glycoprotein secreted by microglia that has been shown to be

a key factor in the phagocytic removal of apoptotic cells [149] and suggested to be

involved in Aβ clearance [151]. Next we explored if MFG-E8 expression was affected

by Aβ and the potential modulation by VS. As shown in Figure III.8, MFG-E8

expression was significantly reduced by 50 nM of Aβ (0.9-fold, p<0.05). Although not

statistically significant, VS alone resulted in a concentration-dependent decrease in the

expression of MFG-E8. In conjunction with Aβ, although the lower VS concentration did

not lead to alterations, the highest concentrations of both compounds induced MFG-E8

expression; however, also not significantly.

Fig. III.8: The expression of MFG- E8 that is reduced by 50 nM and by VS alone, is increased when the highest concentration of both compounds are concomitantly used. N9 cells were treated for 24 h

with 50 nM and 1000 nM amyloid beta (Aβ), as in previous experiments, but now in the presence or absence of 10 μM and 20 μM vinyl sulfone (VS). Cells not-treated with Aβ served as controls. After treatment, total cell lysates were analyzed by western blot with an antibody specific for milk fat globule-EGF factor 8 protein (MFG-E8). (A) Representative results from one experiment are shown. (B) The intensity of the bands was quantified by scanning densitometry, standardized with respect to β-actin protein. Results are expressed as mean ± SEM from at least one experiment, and presented as fold change relatively to control, which was assumed as 1. *p<0.05 vs. respective control.

A

B

Page 76: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

50

2.6. Aβ induces an increase in the expression of HMGB1 and effect that is

prevented by VS

Finally, a particular attention was given to HMGB1, an ubiquous nuclear protein

released by necrotic cells or stimulated monocytes, which mediates the activation of

immune response, including migration and the release of cytokines [170]. Recent

evidence points to HMGB1 as a risk factor for the progression of neuroinflammation

and chronic degeneration in AD[171]. Therefore, we decided to acquire more insight

into this topic. To study the expression of HMGB1 by N9, western blot assay was

performed. We observed a marked increase in the expression of HMGB1 by microglia

after exposition to Aβ (Fig. III.9; 3.5- and 2.7-fold increase for Aβ 50 nM and 1000 nM,

respectively, p<0.05). We verified, additionally, that VS alone did not change the

expression of HMGB1. However, when microglia cells were exposed to Aβ in the

presence of VS, Aβ-induced effects were reversed in a dose-dependent manner, which

was mainly statistically significant with 20μM VS to both Aβ concentrations (0.8- and

0.6-fold for Aβ 50 and 1000 nM; p<0.01 and p<0.05 respectively).

Fig. III.9: Expression of HMGB1 greatly increases in microglia treated with Aβ, but VS shows ability to prevent such effect from occurring. N9 cells were treated for 24 h with 50 nM and 1000 nM amyloid

beta (Aβ), as in previous experiments, but now in the presence or absence of 10 μM and 20 μM vinyl sulfone (VS). Cells not-treated with Aβ served as controls. After treatment, total cell lysates were analyzed by western blot with an antibody specific for high-mobility group box 1 (HMGB1). (A) Representative results from one experiment are shown. (B) The intensity of the bands was quantified by scanning densitometry, standardized with respect to β-actin protein. Results are expressed as mean ± SEM from at least four independents experiments, and presented as fold change relatively to control, which was assumed as 1. *p<0.05 vs. respective control;

$$p<0.01 and

$p<0.05 vs. same experimental condition in the

absence of VS.

A

B

Page 77: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

51

IV. DISCUSSION

Current research indicates that microglia age-dependent neuroinflammatory

changes play an important role in neurodegenerative diseases such as AD [165].

Microglia are able to respond to different kinds of stimuli, being the primary mediators

of the inflammatory response. Several studies demonstrate that Aβ is able to activate

microglia, inducing the production of several pro-inflammatory and neurotoxic factors

[90]. Moreover, with ageing, microglia has been described to present an exaggerated

response to stimuli that subsequently shift to a loss in responsiveness to the same

stimuli and failure in sustaining the regulatory systems that keep the cell under control

[92]. In the first part of this thesis, we used an in vitro ageing model of primary microglia

cultures to study the differences between young and aged microglia when treated with

Aβ and the responsiveness to neuronal regulation. This model was previous

established by characterizing a pure primary microglial cell culture along different DIV,

which demonstrated that cells present typical alterations of ageing with the progression

of time in vitro. Thus, our preliminary data with aged (senescent/dormant) and young

(reactive microglia) cultures of microglia isolated from mice have evidenced that the

first show a diminished NF-κB activation and reactivity, less migration ability and

decreased phagocytic capacity as compared to the second [197]. Corroborating these

results, several aspects of microglial biology were already reported to be altered with

ageing by other studies, including morphology, migratory function and release of

inflammatory cytokines upon microglial activation [218,219]. Here we have further

deepened this topic by evaluating the release of glutamate, ATP and MMPs after Aβ-

stimulation. These inflammatory mediators released by microglia present important

roles in the homeostasis of brain tissue and disease. We were able to show that the

levels of glutamate and MMPs are altered with microglial ageing, and that the release

of glutamate is age-dependently modulated by neuronal signals, results that further

point to the idea that ageing leads to altered microglial reactivity.

It is known that microglia cells release high quantities of glutamate when

activated [220], which has been demonstrated to be an important factor in microglia

neurotoxicity by NMDA receptor signaling [221]. In our work, the presence of Aβ did

not lead to significant differences in the level of released glutamate at both ages. These

findings are in accordance with recent reports indicating that sAPP, but not Aβ, is able

to induce glutamate release from microglia [109]. On the other hand, we observed a

reduction on the basal level of extracellular glutamate with microglial ageing. It is

known that microglia possesses glutamate transporters [94] and, therefore, microglia

Page 78: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

52

are both involved in glutamate release and glutamate uptake, being the extracellular

pool a reflection of this balance. Interestingly, it was reported that microglia ability to

uptake glutamate is increased with age [218], which might have a connection with the

decrease we observed in the extracellular media. Furthermore, the levels of glutamate

are also reduced in mice hippocampus as a result of ageing [222]. Although microglia

may not be the primary responsible, our results point these cells as possible

contributors to the observed decrease. In regard to co-cultures with neurons, the

extracellular level of glutamate in young microglia culture presents a marked reduction.

Indeed, Nakajima and colleagues [223] have already demonstrated that neuronal

stimulation induces an increase in glutamate uptake by microglia, primarily through the

up-regulation of the GLT1 (glutamate transporter 1) glutamate transporter.

Nevertheless, is important to note that the neuronal factors involved were not still

identified. In our case, this neuron-glia crosstalk seems to be only occurring in the

presence of young microglia, and to be absent in aged microglial cells, thus suggesting

a lower ability of these cells to react to neuronal signaling. This might be translated to a

loss of neuroprotection by microglia with ageing, which would lead to a reduced

capacity to cope with excessive glutamate, facilitating neurodegenerative events.

Extracellular ATP is a prominent signaling molecule usually present at low

levels in the extracellular space, but presenting a considerable increase in sites of

inflammation or intense cell stimulation [224]. It was already observed that Aβ induces

the release of ATP to the extracellular medium by microglia [211]. However, our results

are in contrast with these findings, since we did not observe differences regarding the

release of ATP in both isolated young and aged microglia, or even in the presence of

neurons, after exposure to Aβ. We think that this might be due to an increased

hydrolysis by ecto-ATPases or to CTFR inhibitors [225]. In addition, Kim and

colleagues [211] were able to detect an increase in ATP release after Aβ stimulation for

up to 2,5 hours, presenting its maximal peak at 10 minutes after stimulation. This

indicates that Aβ effects might be lost after our 24 h-incubation period. Unexpectedly,

no alterations were visible in the basal level of extracellular ATP between young and

aged microglia, suggesting that the energy metabolism is not impaired with ageing.

Nonetheless, these results refer to the level of extracellular ATP, and we cannot

exclude that differences might be occurring at the intracellular level.

Microglia cells can produce several MMPs that can degrade the extracellular

matrix and contribute to inflammation, and are thought to play a role in

neurodegenerative diseases, as in the case of AD [127,212]. In respect to the release

of active MMP-2 and MMP-9, the MMPs evaluated in this work, we observed no

significant variations after stimulation with Aβ, in either young or aged microglia,

Page 79: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

53

isolated or in the presence of neurons. Nevertheless, notorious differences are

observed in the basal release of these two MMPs with age. Our results show that the

release of active MMP-2 is higher in aged microglia, which is in line with results

obtained in an in vitro model of senescence using keratocytes [226], as well as with our

previous work [197]. Although MMP-2 was reported to play a role in Aβ catabolism

[132], studies also indicate that this protease is involved in myelin degradation [227]. A

reduction of myelin proteins has been observed with ageing, being associated with an

increase in the activation state of microglia [228], pointing to a possible relation

between MMP-2 and myelin loss. This breakdown in myelin observed with ageing,

might contribute to the progressive nonrandom neuronal damage characteristic of AD

[229]. In regard to MMP-9, the results are the opposite, with isolated young microglia

releasing more active MMP-9 than aged cells, results that once more replicate our

previous findings [197]. Interestingly, MMP-9 presents α-secretase activity [230], and a

reduction in its levels with ageing might be linked with the appearance of β-amyloid

pathology. Moreover, MMP-9 is also able to degrade Aβ even in the form of fibrils and

plaques [131], which further suggests that a reduction of is level may have a role in

disease onset. In the presence of neurons, we observed no significant differences in

the release of active MMPs by young and aged cells, which suggests that the

expression of these proteases by microglia is not under neuronal modulation.

In the second part of this Master Thesis we followed a different approach and

aimed to explore the ability of VS to modulate the response of microglia cells to an

inflammatory stimulus. This aspect is particularly relevant since no therapy has to now

prove efficacy in halting AD progression. Once more, we used Aβ as a stimulant, since,

as it was already referred, it has been widely described that it activates microglia,

rendering them more inflammatory and cytotoxic [115]. Furthermore, this also helps to

give a better understanding of the role of Aβ in the immune processes in AD. Due to

the fact that this is an exploratory approach, we opted to use our N9 microglial cell line,

which was recently characterized in our group [231]. Here, we were able to

demonstrate that Aβ induces a more inflammatory microglial phenotype, characterized

by an increased release of the inflammatory mediators MMP-2, MMP-9 and HMGB1. In

addition, VS reveals to reduce these effects close to basal levels, presenting an anti-

inflammatory action. We also demonstrated that Aβ reduces TLR4 levels and

phagocytic capability of microglial cell line N9, and that again VS reveals aptitude to

counteract such effects.

We first evaluated cell viability after exposure to Aβ and in the presence of VS.

Our results indicate that VS does not present significant toxicity to the cells at both

concentrations tested, which points for a use in in vivo settings without expecting

Page 80: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

54

significant direct side effects. Exposure to Aβ, although it also did not present

significant toxicity, demonstrated a slightly reduction in cell viability in a dose-

dependent manner. In fact, studies indicating Aβ-induced toxicity in microglia cells only

found significant reduction in cell viability at concentrations greater than 5 µM [232],

values higher than the ones here used. We also observed no significant protection by

VS at this level. However, it is worth to note that VS slightly increases the viability of

cells when exposed to the highest Aβ concentration. Thus, we cannot exclude that VS

may even have protective effects at concentrations of Aβ above the ones used in this

study.

To address the impact of VS in modulating microglial inflammatory properties,

we evaluated the release of MMPs, which are known to be involved in the modulation

of inflammatory processes [233]. In this study, Aβ was able to induce the release of

both active MMP-9 and MMP-2 in a dose-dependent manner. Indeed, the release of

MMP-9 [234] and MMP-2 [235] were already reported to be induced by Aβ,

corroborating our findings. The excessive production of these proteases might be

linked with various detrimental roles in neuroinflammatory-related diseases, as in the

case of AD. Both MMP-9 and MMP-2 were reported to be able to activate TNF-α in

vitro, indicating that they may contribute to the inflammatory state in AD [233].

Furthermore, MMP-2 is increased in AD mouse models and was found to correlate with

inflammatory injury [236]. On the other hand, MMP-9 was reported to be involved in

Aβ-induced cognitive impairment and neurotoxicity [237], suggesting that it may also

play a role in disease progression. In addition, MMP-2 and MMP-9 can regulate

endothelial permeability by cleaving occludin and may contribute to the opening of the

BBB [238], which has been reported to be impaired in AD [239]. In this regard,

inhibition of these proteases might represent a therapeutical approach to AD.

Consistent with this idea, in this study we have shown that VS, at 20 µM, is able to

counteract the release of MMP-9 and MMP-2 by the highest concentration of Aβ,

suggesting its utility for therapeutical intervention. Nonetheless, these results should be

carefully taken. As it was aforementioned, both MMP-9 and MMP-2 are involved in Aβ

degradation and may play different contributions in different states of the disease

development. Thus, further studies need to be done to understand the consequences

of inhibiting MMPs in the context of AD.

Several receptors have been reported to be responsible for the microglial Aβ

stimulation. Of these, TLR4 deserves special attention. It has been demonstrated that a

mutation in the TLR4 gene strongly inhibits microglial activation by Aβ, resulting in a

significant lower release of inflammatory products like IL-6, TNF-α and NO, and there is

an increased expression of TLR4 in brain tissue associated with amyloid plaques in AD

Page 81: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

55

[113]. Interestingly, it was also recently reported that LPS, another pro-inflammatory

stimulant, induces the expression of MMP-9 through TLR4 [240]. However, our results

regarding this receptor were unexpected. We detected a decrease in the levels of

TLR4 after Aβ treatment, particularly at the highest concentration. Nevertheless, this

might be a result of the existence of two forms of TLR4. In our observations, we could

only detect a band of 93 kDa, which corresponds to the non-glycosylated form of TLR4,

the form that does not translocate to the membrane and that is not available for

binding. We could not detect the 130 kDa band, which is responsible for the

glycosylated form of TLR that is found on cell surface [241]. Thus, we speculate that

the decrease in the non-glycosylated form may be a consequence of an increase in the

glycosylated one, but that our detection method is not sensible enough to detect it. In

further studies we plan to use different methodology, such as immunocytochemistry or

Real Time-Polymerase Chain Reaction (PCR), to quantify TLR4 and better clarify the

involvement of this receptor in Aβ-induced activation. In respect to the effects of VS, we

observed that at 20 µM and in the presence of 1000 nM Aβ, VS increases the

expression of the non-glycosylated form. Moreover, VS alone appears to reduce the

non-glycosylated form of TLR4. Nevertheless, it is difficult to draw conclusions of these

results without further confirmation. Indeed, at the light of our speculation, these results

might indicate that the inflammatory effects of Aβ may be mediated through this

receptor, pointing for a positive role of 20 μM and a mechanism for its anti-inflammatory

effects. However, further studies need to be performed to obtain more conclusive

results, since other receptors beside TLR4 might also be involved in Aβ-induced

activation [111].

When activated, microglia can become highly phagocytic, a function critical for

the uptake and degradation of infectious agents and degenerated cells, contributing for

the homeostasis of the brain tissue [217]. Aβ has been described to not only incite an

inflammatory response, but also to be able to induce microglial phagocytosis [114].

Surprisingly, our results show a decrease in the phagocytic capacity of microglia in a

dose-dependent manner after exposure to Aβ. Although this is in discordance with

previous findings reporting a marked increase of phagocytic activity after Aβ-stimulation

[242], other studies indicate a frustrated phagocytosis [157]. To note that on one hand

microglia is able to phagocytize the aggregates of Aβ, but is not able to efficiently

perform their degradation [141], and that on the other hand it may result from an

increased formation of oligomers, which showed to be less capable of inducing

phagocytosis when compared with fibrils [243]. Decreased phagocytic capacity may

also derive from inflammation resulting from the Aβ-stimulation itself, since expression

of inflammatory cytokines has been demonstrated to attenuate phagocytosis induced

Page 82: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

56

by Aβ [244,243]. In fact, it was previously demonstrated that microglial reactivity

changes from phagocytic to inflammatory along time of exposure to unconjugated

bilirubin, a well characterized pro-inflammatory stimulant, stressing the importance of

the time of incubation [206]. In regard to VS, both concentrations tested slightly

reversed the effects of microglia exposure to Aβ. Although this was not statistically

significant, it suggests that this compound may be an inducer of phagocytosis.

MFG-E8 is a bridging protein that is known to mediate the phagocytic removal

of apoptotic cells by binding to PS [147]. Due to its connection with phagocytosis, we

also evaluated how MFG-E8 is modulated by Aβ and VS. Here, we shown a significant

decrease in its content in microglia stimulated with the low nM concentration of Aβ. It

was previously reported that Aβ small concentrations (250 nM) is able to induce the

phagocytosis of neurons and that this is mediated by the release of MFG-E8, which

may account for the loss of neurons in AD [245]. Thus, we may assume that the

reduction we observed is related to a release of MFG-E8 to the extracellular space.

Unfortunately, we were not able to evaluate the extracellular content of this protein and

this still remains to be determined. To further add that VS, although not producing

significant changes, can be envisaged as a phagocytic inducer.

We also evaluated HMGB1, a nuclear protein that is released as an

inflammatory factor and related with several neurodegenerative diseases [171]. This

protein has been found in elevated levels in AD, co-localizing with amyloid plaques in

association with microglia [172]. Here we show, for the first time, that Aβ might present

a link to these associations. Our results demonstrate that Aβ is able to markedly induce

the expression of HMGB1 by microglial cells. Surprisingly, we also noticed that VS is

able to prevent the effects produced by Aβ, mainly at 20 µM. This is of particular

interest in the context of AD, since HMGB1 contributes to Aβ accumulation by

preventing its phagocytosis by microglia [173]. HMGB1 is also able to bind directly to

Aβ, stabilizing the oligomer conformations [172], and to interact with receptors like

RAGE and TLR4, leading to an increase in toxicity and inflammation [170].

Furthermore, HMGB1 was reported to mediate cognitive impairment [246], which raises

the possibility that it may be involved in cognitive deficit observed in AD. Therefore, a

reduction of the levels of HMGB1 might represent a valuable therapeutical approach to

AD, aiming to reduce the effects of Aβ accumulation and neuroinflammation.

Nonetheless, it is important to emphasize that only levels of HMGB1 inside the cells

were investigated and that further studies should search the existence of HMGB1 in the

extracellular milieu.

Overall, although VS presents some promising properties, it has to be referred

that further studies need to be done to better understand its effects, as well as its

Page 83: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

57

mechanism of action, which is currently unknown. VS was indicated as a cathepsin S

inhibitor and indeed the inhibition of cathepsin S has been described to reduce de

activity of MMP-9 and MMP-2 [247]. Furthermore, it was also suggested that inhibition

of cathepsin S provides neuroprotective effects by suppressing the inflammatory

response [248]. Nevertheless, it is not safe to assume that all these results are just

mediated by the inhibition of cathepsin S at this point, since this same inhibitor might

be acting on other cathepsins as well. For instance, the inhibition of cathepsin L was

recently shown to be able to reduce the activation of NF-κB [249] which is involved in

the regulation of several genes linked to inflammation [250].

In conclusion, in the first part of this Master Thesis, we have shown that

microglia behaves differently with age, releasing lower levels of glutamate, MMP-9 and

higher levels of MMP-2. Furthermore, based on our glutamate results, we also

observed that aged microglia loses, at least to some extent, the ability to respond to

neuronal signaling. These results are indeed interesting in the context of AD, and in

age-related neurodegenerative disease in general, as well, since they converge to the

idea that microglia becomes dysfunctional with age, contributing to the progression of

the disease. Besides, these results also strength our model as an alternative approach

to study microglial senescence. In the second part of this Thesis, we showed that Aβ

activates microglia, inducing a more inflammatory phenotype, characterized by an

increase of the levels of MMP-2, MMP-9 and HMGB-1 and that VS, our promising test

compound, is able to at least partially prevent the Aβ-induced effects, which elects it to

further investigation as a possible modulator of microglial activity in inflammatory

diseases like AD.

Page 84: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

58

[Escre

va um

Page 85: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

59

Future perspectives

In his work, we have provided some important clues about the effect of

microglial ageing on these cells biology, using an innovative approach – an in vitro

model of microglial ageing. This model might be a valuable asset to further study

ageing in microglia and to evaluate at what extent it contributes to AD onset and

progression, which to date remains poorly understood. Nevertheless, some

considerations might still need to be taken in the future when evaluating the effects of

Aβ, like assessing the best time of incubation or the evaluation of effects using higher

concentrations. Furthermore, in regard to co-culture studies, cell viability should also be

assessed, since microglia might directly eliminate neurons in the presence of Aβ. In

this regard, the use of conditioned media studies, although not permitting the

evaluation of cell-to-cell signaling, might present a useful alternative. There are also

some parameters that may be interesting to assess in the future, to better characterize

this model and understand our results. Evaluation of the level of glutamate receptors

like GLutamate ASpartate Transporter and GLT1, might give some insight to the

differences observed in extracellular glutamate as a result of microglial ageing.

Moreover, evaluating the release of inflammatory cytokines would also help to better

characterized how ageing affects the immune state of microglial cells in our model. The

evaluation of CX3CL1-CX3CR1 was one of the initial aims of this Thesis, which we,

unfortunately, were not able to accomplish in due time. However, both CX3CL1-

CX3CR1 and CD200L-CD200R signaling are interesting parameters to be addressed

in order to understand how microglial ageing affects neuron-microglia communication.

In regard to our approach to explore VS as a modulator of microglia activation,

we presented some interesting preliminary results, which indicate that this compound

has anti-inflammatory properties. To further confirm this, it would be important to

directly measure the level of inflammatory cytokines. The receptor that mediates the

inflammatory effects of Aβ remains, however, to identify. Although we assessed the

expression of TLR4 by western blot we believe that it will be necessary its

determination by complementary methods, namely immunocytochemistry or Real

Time-PCR, to dissect the relevance that TLR4 and probably TLR2 represent in the

process of ageing, microglia irresponsiveness or over-activation, and susceptibility to

the appearance of age-related disorders. In respect to the phagocytic capability, some

unexpected results, in a form of a reduced phagocytosis of beads in response to Aβ,

were also observed. Nevertheless, repeating the phagocytic assay using fluorescently

labeled Aβ might represent a way to clarify the reason behind the results observed.

Another point of vital importance that we plan to address in the future is the evaluation

Page 86: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

60

of extracellular MFG-E8 and HMGB1, since these molecules play most of their roles

outside the cell. Finally, to understand if the effects of VS are indeed mediated by a

specific inhibition of cathepsins S it would be important to perform specific tests in

regard to different cathepsin activity in the presence of this inhibitor. In case of

specificity to cathepsins S, this might also reveal new biological pathways in which this

cathepsin may be involved.

In a last note, although we did not address this question in the present study,

since oAβ and fAβ are the main forms of Aβ characterized in AD studies and might

possess different effects, the induction of microglia with purified oAβ and fAβ would

allow to identify if the effects we obtained may be appointed to any particular

aggregation form, giving further insight to the role of Aβ aggregation state in the onset

and progression of such a devastating disease.

Page 87: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

61

V. REFERENCES

1. Kidd PM (2008) Alzheimer's disease, amnestic mild cognitive impairment, and age-associated memory impairment: current understanding and progress toward integrative prevention. Altern Med Rev 13:85-115. 2. Allsop D (2000) Introduction to Alzheimer's disease. Methods Mol Med 32:1-21. 3. Khairallah MI, Kassem LA (2011) Alzheimer's disease: current status of etiopathogenesis and therapeutic strategies. Pak J Biol Sci 14:257-272. 4. Isik AT (2010) Late onset Alzheimer's disease in older people. Clin Interv Aging 5:307-311. 5. Bettens K, Sleegers K, Van Broeckhoven C (2013) Genetic insights in Alzheimer's disease. Lancet Neurol 12:92-104. 6. Rao AT, Degnan AJ, Levy LM (2013) Genetics of Alzheimer Disease. AJNR Am J Neuroradiol. (doi:10.3174/ajnr.A3545) 7. Siciliano R, Barone E, Calabrese V, Rispoli V, Butterfield DA, Mancuso C (2011) Experimental research on nitric oxide and the therapy of Alzheimer disease: a challenging bridge. CNS Neurol Disord Drug Targets 10:766-776. 8. Salawu FK, Umar JT, Olokoba AB (2011) Alzheimer's disease: a review of recent developments. Ann Afr Med 10:73-79. 9. Liu CC, Kanekiyo T, Xu H, Bu G (2013) Apolipoprotein E and Alzheimer disease: risk, mechanisms and therapy. Nat Rev Neurol 9:106-118. 10. Svetkovic-Dozic D, Skender-Gazibara M, Dozic S (2001) Neuropathological hallmarks of Alzheimer's disease. Arch Oncol 9:195-199. 11. Huang Y, Mucke L (2012) Alzheimer mechanisms and therapeutic strategies. Cell 148:1204-1222. 12. Avila J (2012) Spreading of Tau Pathology. J Alzheimers Dis Parkinsonism 2:1-2. 13. Braak H, Rub U, Schultz C, Del Tredici K (2006) Vulnerability of cortical neurons to Alzheimer's and Parkinson's diseases. J Alzheimers Dis 9:35-44. 14. Mattson MP, Magnus T (2006) Ageing and neuronal vulnerability. Nat Rev Neurosci 7:278-294. 15. Sheng M, Sabatini BL, Sudhof TC (2012) Synapses and Alzheimer's disease. Cold Spring Harb Perspect Biol. (doi: 10.1101/cshperspect.a005777) 16. Petrella JR, Coleman RE, Doraiswamy PM (2003) Neuroimaging and early diagnosis of Alzheimer disease: a look to the future. Radiology 226:315-336. 17. Rubio-Perez JM, Morillas-Ruiz JM (2012) A review: inflammatory process in Alzheimer's disease, role of cytokines. ScientificWorldJournal 2012:756357. 18. Bi X (2010) Alzheimer disease: update on basic mechanisms. J Am Osteopath Assoc 110:S3-9. 19. Holtzman DM, Morris JC, Goate AM (2011) Alzheimer's disease: the challenge of the second century. Sci Transl Med 3:77sr71. 20. Puzzo D, Arancio O (2013) Amyloid-beta peptide: Dr. Jekyll or Mr. Hyde? J Alzheimers Dis 33:S111-120. 21. O'Brien RJ, Wong PC (2011) Amyloid precursor protein processing and Alzheimer's disease. Annu Rev Neurosci 34:185-204. 22. Cavallucci V, D'Amelio M, Cecconi F (2012) Abeta toxicity in Alzheimer's disease. Mol Neurobiol 45:366-378. 23. Sheng M, Sabatini BL, Sudhof TC (2012) Synapses and Alzheimer's disease. Cold Spring Harb Perspect Biol 4:a005777. 24. Wesson DW, Nixon RA, Levy E, Wilson DA (2011) Mechanisms of neural and behavioral dysfunction in Alzheimer's disease. Mol Neurobiol 43:163-179. 25. Shankar GM, Walsh DM (2009) Alzheimer's disease: synaptic dysfunction and Abeta. Mol Neurodegener 4:48.

Page 88: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

62

26. Crews L, Masliah E (2010) Molecular mechanisms of neurodegeneration in Alzheimer's disease. Hum Mol Genet 19:R12-20. 27. Walsh DM, Selkoe DJ (2007) Abeta oligomers - a decade of discovery. J Neurochem 101:1172-1184. 28. Wang YJ, Zhou HD, Zhou XF (2006) Clearance of amyloid-beta in Alzheimer's disease: progress, problems and perspectives. Drug Discov Today 11:931-938. 29. Bekris LM, Yu CE, Bird TD, Tsuang DW (2010) Genetics of Alzheimer disease. J Geriatr Psychiatry Neurol 23:213-227. 30. Lorenzo A, Yankner BA (1994) Beta-amyloid neurotoxicity requires fibril formation and is inhibited by congo red. Proc Natl Acad Sci U S A 91:12243-12247. 31. Geula C, Wu CK, Saroff D, Lorenzo A, Yuan M, Yankner BA (1998) Aging renders the brain vulnerable to amyloid beta-protein neurotoxicity. Nat Med 4:827-831. 32. Larson ME, Lesne SE (2012) Soluble Abeta oligomer production and toxicity. J Neurochem 120:125-139. 33. Kayed R, Lasagna-Reeves CA (2013) Molecular mechanisms of amyloid oligomers toxicity. J Alzheimers Dis 33:S67-78. 34. LaFerla FM, Green KN, Oddo S (2007) Intracellular amyloid-beta in Alzheimer's disease. Nat Rev Neurosci 8:499-509. 35. Lorenzo A, Yuan M, Zhang Z, Paganetti PA, Sturchler-Pierrat C, Staufenbiel M, Mautino J, Vigo FS, Sommer B, Yankner BA (2000) Amyloid beta interacts with the amyloid precursor protein: a potential toxic mechanism in Alzheimer's disease. Nat Neurosci 3:460-464. 36. Chasseigneaux S, Allinquant B (2012) Functions of Abeta, sAPPalpha and sAPPbeta : similarities and differences. J Neurochem 120:99-108. 37. Duan Y, Dong S, Gu F, Hu Y, Zhao Z (2012) Advances in the pathogenesis of Alzheimer's disease: focusing on tau-mediated neurodegeneration. Transl Neurodegener 1:24. 38. Gotz J, Ittner LM, Kins S (2006) Do axonal defects in tau and amyloid precursor protein transgenic animals model axonopathy in Alzheimer's disease? J Neurochem 98:993-1006. 39. Gong CX, Grundke-Iqbal I, Iqbal K (2010) Targeting tau protein in Alzheimer's disease. Drugs Aging 27:351-365. 40. Iqbal K, Alonso Adel C, Grundke-Iqbal I (2008) Cytosolic abnormally hyperphosphorylated tau but not paired helical filaments sequester normal MAPs and inhibit microtubule assembly. J Alzheimers Dis 14:365-370. 41. Flach K, Hilbrich I, Schiffmann A, Gartner U, Kruger M, Leonhardt M, Waschipky H, Wick L, Arendt T, Holzer M (2012) Tau oligomers impair artificial membrane integrity and cellular viability. J Biol Chem 287:43223-43233. 42. Ballatore C, Lee VM, Trojanowski JQ (2007) Tau-mediated neurodegeneration in Alzheimer's disease and related disorders. Nat Rev Neurosci 8:663-672. 43. Wippold FJ, 2nd, Cairns N, Vo K, Holtzman DM, Morris JC (2008) Neuropathology for the neuroradiologist: plaques and tangles. AJNR Am J Neuroradiol 29:18-22. 44. Higgins GC, Beart PM, Shin YS, Chen MJ, Cheung NS, Nagley P (2010) Oxidative stress: emerging mitochondrial and cellular themes and variations in neuronal injury. J Alzheimers Dis 20:S453-473. 45. von Bernhardi R, Eugenin J (2012) Alzheimer's disease: redox dysregulation as a common denominator for diverse pathogenic mechanisms. Antioxid Redox Signal 16:974-1031. 46. Nunomura A, Perry G, Aliev G, Hirai K, Takeda A, Balraj EK, Jones PK, Ghanbari H, Wataya T, Shimohama S, Chiba S, Atwood CS, Petersen RB, Smith MA (2001) Oxidative damage is the earliest event in Alzheimer disease. J Neuropathol Exp Neurol 60:759-767. 47. Castellani RJ, Harris PL, Sayre LM, Fujii J, Taniguchi N, Vitek MP, Founds H, Atwood CS, Perry G, Smith MA (2001) Active glycation in neurofibrillary pathology of

Page 89: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

63

Alzheimer disease: N(epsilon)-(carboxymethyl) lysine and hexitol-lysine. Free Radic Biol Med 31:175-180. 48. Axelsen PH, Komatsu H, Murray IV (2011) Oxidative stress and cell membranes in the pathogenesis of Alzheimer's disease. Physiology (Bethesda) 26:54-69. 49. Bonda DJ, Wang X, Perry G, Nunomura A, Tabaton M, Zhu X, Smith MA (2010) Oxidative stress in Alzheimer disease: a possibility for prevention. Neuropharmacology 59:290-294. 50. Pagani L, Eckert A (2011) Amyloid-Beta interaction with mitochondria. Int J Alzheimers Dis 2011:925050. 51. Su B, Wang X, Nunomura A, Moreira PI, Lee HG, Perry G, Smith MA, Zhu X (2008) Oxidative stress signaling in Alzheimer's disease. Curr Alzheimer Res 5:525-532. 52. Yan MH, Wang X, Zhu X (2013) Mitochondrial defects and oxidative stress in Alzheimer disease and Parkinson disease. Free Radic Biol Med 62:90-101. 53. Herculano-Houzel S (2011) Scaling of brain metabolism with a fixed energy budget per neuron: implications for neuronal activity, plasticity and evolution. PLoS One 6:e17514. 54. Maruszak A, Zekanowski C (2011) Mitochondrial dysfunction and Alzheimer's disease. Prog Neuropsychopharmacol Biol Psychiatry 35:320-330. 55. Silva DF, Selfridge JE, Lu J, E L, Cardoso SM, Swerdlow RH (2012) Mitochondrial abnormalities in Alzheimer's disease: possible targets for therapeutic intervention. Adv Pharmacol 64:83-126. 56. Eckert GP, Renner K, Eckert SH, Eckmann J, Hagl S, Abdel-Kader RM, Kurz C, Leuner K, Muller WE (2012) Mitochondrial dysfunction--a pharmacological target in Alzheimer's disease. Mol Neurobiol 46:136-150. 57. Tillement L, Lecanu L, Papadopoulos V (2011) Alzheimer's disease: effects of beta-amyloid on mitochondria. Mitochondrion 11:13-21. 58. Spuch C, Ortolano S, Navarro C (2012) New insights in the amyloid-Beta interaction with mitochondria. J Aging Res 2012:324968. 59. Lustbader JW, Cirilli M, Lin C, Xu HW, Takuma K, Wang N, Caspersen C, Chen X, Pollak S, Chaney M, Trinchese F, Liu S, Gunn-Moore F, Lue LF, Walker DG, Kuppusamy P, Zewier ZL, Arancio O, Stern D, Yan SS, Wu H (2004) ABAD directly links Abeta to mitochondrial toxicity in Alzheimer's disease. Science 304:448-452. 60. Leuner K, Schutt T, Kurz C, Eckert SH, Schiller C, Occhipinti A, Mai S, Jendrach M, Eckert GP, Kruse SE, Palmiter RD, Brandt U, Drose S, Wittig I, Willem M, Haass C, Reichert AS, Muller WE (2012) Mitochondrion-derived reactive oxygen species lead to enhanced amyloid beta formation. Antioxid Redox Signal 16:1421-1433. 61. Swerdlow RH (2012) Mitochondria and cell bioenergetics: increasingly recognized components and a possible etiologic cause of Alzheimer's disease. Antioxid Redox Signal 16:1434-1455. 62. Agostinho P, Cunha RA, Oliveira C (2010) Neuroinflammation, oxidative stress and the pathogenesis of Alzheimer's disease. Curr Pharm Des 16:2766-2778. 63. Eikelenboom P, van Exel E, Hoozemans JJ, Veerhuis R, Rozemuller AJ, van Gool WA (2010) Neuroinflammation - an early event in both the history and pathogenesis of Alzheimer's disease. Neurodegener Dis 7:38-41. 64. Lucas SM, Rothwell NJ, Gibson RM (2006) The role of inflammation in CNS injury and disease. Br J Pharmacol 147:S232-240. 65. Wojtera M, Sobow T, Kloszewska I, Liberski PP, Brown DR, Sikorska B (2012) Expression of immunohistochemical markers on microglia in Creutzfeldt-Jakob disease and Alzheimer's disease: morphometric study and review of the literature. Folia Neuropathol 50:74-84. 66. Njie EG, Boelen E, Stassen FR, Steinbusch HW, Borchelt DR, Streit WJ (2012) Ex vivo cultures of microglia from young and aged rodent brain reveal age-related changes in microglial function. Neurobiol Aging 33:195.e1-12.

Page 90: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

64

67. Vasto S, Candore G, Duro G, Lio D, Grimaldi MP, Caruso C (2007) Alzheimer's disease and genetics of inflammation: a pharmacogenomic vision. Pharmacogenomics 8:1735-1745. 68. Krause DL, Muller N (2010) Neuroinflammation, microglia and implications for anti-inflammatory treatment in Alzheimer's disease. Int J Alzheimers Dis 2010:732806. 69. Perry VH, Cunningham C, Holmes C (2007) Systemic infections and inflammation affect chronic neurodegeneration. Nat Rev Immunol 7:161-167. 70. Szekely CA, Thorne JE, Zandi PP, Ek M, Messias E, Breitner JC, Goodman SN (2004) Nonsteroidal anti-inflammatory drugs for the prevention of Alzheimer's disease: a systematic review. Neuroepidemiology 23:159-169. 71. in t' Veld BA, Ruitenberg A, Hofman A, Launer LJ, van Duijn CM, Stijnen T, Breteler MM, Stricker BH (2001) Nonsteroidal antiinflammatory drugs and the risk of Alzheimer's disease. N Engl J Med 345:1515-1521. 72. Martin BK, Szekely C, Brandt J, Piantadosi S, Breitner JC, Craft S, Evans D, Green R, Mullan M (2008) Cognitive function over time in the Alzheimer's Disease Anti-inflammatory Prevention Trial (ADAPT): results of a randomized, controlled trial of naproxen and celecoxib. Arch Neurol 65:896-905. 73. Weggen S, Eriksen JL, Das P, Sagi SA, Wang R, Pietrzik CU, Findlay KA, Smith TE, Murphy MP, Bulter T, Kang DE, Marquez-Sterling N, Golde TE, Koo EH (2001) A subset of NSAIDs lower amyloidogenic Abeta42 independently of cyclooxygenase activity. Nature 414:212-216. 74. Wyss-Coray T, Rogers J (2012) Inflammation in Alzheimer disease-a brief review of the basic science and clinical literature. Cold Spring Harb Perspect Med 2:a006346. 75. Yamamoto M, Kiyota T, Horiba M, Buescher JL, Walsh SM, Gendelman HE, Ikezu T (2007) Interferon-gamma and tumor necrosis factor-alpha regulate amyloid-beta plaque deposition and beta-secretase expression in Swedish mutant APP transgenic mice. Am J Pathol 170:680-692. 76. Arnaud L, Robakis NK, Figueiredo-Pereira ME (2006) It may take inflammation, phosphorylation and ubiquitination to 'tangle' in Alzheimer's disease. Neurodegener Dis 3:313-319. 77. Alves L, Correia AS, Miguel R, Alegria P, Bugalho P (2012) Alzheimer's disease: a clinical practice-oriented review. Front Neurol 3:63. 78. Fletcher LC, Burke KE, Caine PL, Rinne NL, Braniff CA, Davis HR, Miles KA, Packer C (2013) Diagnosing Alzheimer's disease: are we any nearer to useful biomarker-based, non-invasive tests? GMS Health Technol Assess 9:Doc01. 79. Kimchi E, Desai EK, Grossberg GT (2012) New Alzheimer’s disease guidelines: Implications for clinicians. Current Psychiatry 11:15-20. 80. McKhann GM, Knopman DS, Chertkow H, Hyman BT, Jack CR, Jr., Kawas CH, Klunk WE, Koroshetz WJ, Manly JJ, Mayeux R, Mohs RC, Morris JC, Rossor MN, Scheltens P, Carrillo MC, Thies B, Weintraub S, Phelps CH (2011) The diagnosis of dementia due to Alzheimer's disease: recommendations from the National Institute on Aging-Alzheimer's Association workgroups on diagnostic guidelines for Alzheimer's disease. Alzheimers Dement 7:263-269. 81. Jack CR, Jr., Albert MS, Knopman DS, McKhann GM, Sperling RA, Carrillo MC, Thies B, Phelps CH (2011) Introduction to the recommendations from the National Institute on Aging-Alzheimer's Association workgroups on diagnostic guidelines for Alzheimer's disease. Alzheimers Dement 7:257-262. 82. Zhong Z, Ewers M, Teipel S, Burger K, Wallin A, Blennow K, He P, McAllister C, Hampel H, Shen Y (2007) Levels of beta-secretase (BACE1) in cerebrospinal fluid as a predictor of risk in mild cognitive impairment. Arch Gen Psychiatry 64:718-726. 83. Montine TJ, Peskind ER, Quinn JF, Wilson AM, Montine KS, Galasko D (2011) Increased cerebrospinal fluid F2-isoprostanes are associated with aging and latent Alzheimer's disease as identified by biomarkers. Neuromolecular Med 13:37-43. 84. Demarin V, Zavoreo I, Kes VB, Simundic AM (2011) Biomarkers in Alzheimer's disease. Clin Chem Lab Med 49:773-778.

Page 91: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

65

85. Parsons CG, Danysz W, Dekundy A, Pulte I (2013) Memantine and Cholinesterase Inhibitors: Complementary Mechanisms in the Treatment of Alzheimer's Disease. Neurotox Res 24:358–369. 86. Stone JG, Casadesus G, Gustaw-Rothenberg K, Siedlak SL, Wang X, Zhu X, Perry G, Castellani RJ, Smith MA (2011) Frontiers in Alzheimer's disease therapeutics. Ther Adv Chronic Dis 2:9-23. 87. Jiang T, Yu JT, Tan L (2012) Novel disease-modifying therapies for Alzheimer's disease. J Alzheimers Dis 31:475-492. 88. Corbett A, Smith J, Ballard C (2012) New and emerging treatments for Alzheimer's disease. Expert Rev Neurother 12:535-543. 89. Rogers J, Shen Y (2000) A perspective on inflammation in Alzheimer's disease. Ann N Y Acad Sci 924:132-135. 90. Block ML, Zecca L, Hong JS (2007) Microglia-mediated neurotoxicity: uncovering the molecular mechanisms. Nat Rev Neurosci 8:57-69. 91. Leavy O (2010) Macrophages: Microglial cell origins. Nat Rev Immunol 10:808-809. 92. Norden DM, Godbout JP (2013) Review: microglia of the aged brain: primed to be activated and resistant to regulation. Neuropathol Appl Neurobiol 39:19-34. 93. Nimmerjahn A, Kirchhoff F, Helmchen F (2005) Resting microglial cells are highly dynamic surveillants of brain parenchyma in vivo. Science 308:1314-1318. 94. Kettenmann H, Hanisch UK, Noda M, Verkhratsky A (2011) Physiology of microglia. Physiol Rev 91:461-553. 95. Ajami B, Bennett JL, Krieger C, Tetzlaff W, Rossi FM (2007) Local self-renewal can sustain CNS microglia maintenance and function throughout adult life. Nat Neurosci 10:1538-1543. 96. Streit WJ (2001) Microglia and macrophages in the developing CNS. Neurotoxicology 22:619-624. 97. Fourgeaud L, Boulanger LM (2007) Synapse remodeling, compliments of the complement system. Cell 131:1034-1036. 98. Lue LF, Kuo YM, Beach T, Walker DG (2010) Microglia activation and anti-inflammatory regulation in Alzheimer's disease. Mol Neurobiol 41:115-128. 99. Lucin KM, Wyss-Coray T (2009) Immune activation in brain aging and neurodegeneration: too much or too little? Neuron 64:110-122. 100. Colton CA (2009) Heterogeneity of microglial activation in the innate immune response in the brain. J Neuroimmune Pharmacol 4:399-418. 101. Czeh M, Gressens P, Kaindl AM (2011) The yin and yang of microglia. Dev Neurosci 33:199-209. 102. Cameron B, Landreth GE (2010) Inflammation, microglia, and Alzheimer's disease. Neurobiol Dis 37:503-509. 103. McGeer PL, Itagaki S, Tago H, McGeer EG (1987) Reactive microglia in patients with senile dementia of the Alzheimer type are positive for the histocompatibility glycoprotein HLA-DR. Neurosci Lett 79:195-200. 104. Frautschy SA, Yang F, Irrizarry M, Hyman B, Saido TC, Hsiao K, Cole GM (1998) Microglial response to amyloid plaques in APPsw transgenic mice. Am J Pathol 152:307-317. 105. Sokolowski JD, Mandell JW (2011) Phagocytic clearance in neurodegeneration. Am J Pathol 178:1416-1428. 106. Meyer-Luehmann M, Spires-Jones TL, Prada C, Garcia-Alloza M, de Calignon A, Rozkalne A, Koenigsknecht-Talboo J, Holtzman DM, Bacskai BJ, Hyman BT (2008) Rapid appearance and local toxicity of amyloid-beta plaques in a mouse model of Alzheimer's disease. Nature 451:720-724. 107. Bolmont T, Haiss F, Eicke D, Radde R, Mathis CA, Klunk WE, Kohsaka S, Jucker M, Calhoun ME (2008) Dynamics of the microglial/amyloid interaction indicate a role in plaque maintenance. J Neurosci 28:4283-4292.

Page 92: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

66

108. Cagnin A, Brooks DJ, Kennedy AM, Gunn RN, Myers R, Turkheimer FE, Jones T, Banati RB (2001) In-vivo measurement of activated microglia in dementia. Lancet 358:461-467. 109. Mizuno T (2012) The biphasic role of microglia in Alzheimer's disease. Int J Alzheimers Dis 2012:737846. 110. El Khoury J, Luster AD (2008) Mechanisms of microglia accumulation in Alzheimer's disease: therapeutic implications. Trends Pharmacol Sci 29:626-632. 111. Salminen A, Ojala J, Kauppinen A, Kaarniranta K, Suuronen T (2009) Inflammation in Alzheimer's disease: amyloid-beta oligomers trigger innate immunity defence via pattern recognition receptors. Prog Neurobiol 87:181-194. 112. Lehnardt S, Massillon L, Follett P, Jensen FE, Ratan R, Rosenberg PA, Volpe JJ, Vartanian T (2003) Activation of innate immunity in the CNS triggers neurodegeneration through a Toll-like receptor 4-dependent pathway. Proc Natl Acad Sci U S A 100:8514-8519. 113. Walter S, Letiembre M, Liu Y, Heine H, Penke B, Hao W, Bode B, Manietta N, Walter J, Schulz-Schuffer W, Fassbender K (2007) Role of the toll-like receptor 4 in neuroinflammation in Alzheimer's disease. Cell Physiol Biochem 20:947-956. 114. Schlachetzki JC, Hull M (2009) Microglial activation in Alzheimer's disease. Curr Alzheimer Res 6:554-563. 115. Sondag CM, Dhawan G, Combs CK (2009) Beta amyloid oligomers and fibrils stimulate differential activation of primary microglia. J Neuroinflammation 6:1. 116. Michelucci A, Heurtaux T, Grandbarbe L, Morga E, Heuschling P (2009) Characterization of the microglial phenotype under specific pro-inflammatory and anti-inflammatory conditions: Effects of oligomeric and fibrillar amyloid-beta. J Neuroimmunol 210:3-12. 117. Rogers J, Lue LF (2001) Microglial chemotaxis, activation, and phagocytosis of amyloid beta-peptide as linked phenomena in Alzheimer's disease. Neurochem Int 39:333-340. 118. Weitz TM, Town T (2012) Microglia in Alzheimer's Disease: It's All About Context. Int J Alzheimers Dis 2012:314185. 119. El Khoury JB, Moore KJ, Means TK, Leung J, Terada K, Toft M, Freeman MW, Luster AD (2003) CD36 mediates the innate host response to beta-amyloid. J Exp Med 197:1657-1666. 120. Cho H, Hashimoto T, Wong E, Hori Y, Wood LB, Zhao L, Haigis KM, Hyman BT, Irimia D (2013) Microfluidic chemotaxis platform for differentiating the roles of soluble and bound amyloid-beta on microglial accumulation. Sci Rep 3:1823. 121. Xia MQ, Qin SX, Wu LJ, Mackay CR, Hyman BT (1998) Immunohistochemical study of the beta-chemokine receptors CCR3 and CCR5 and their ligands in normal and Alzheimer's disease brains. Am J Pathol 153:31-37. 122. Barger SW, Basile AS (2001) Activation of microglia by secreted amyloid precursor protein evokes release of glutamate by cystine exchange and attenuates synaptic function. J Neurochem 76:846-854. 123. Noda M, Nakanishi H, Akaike N (1999) Glutamate release from microglia via glutamate transporter is enhanced by amyloid-beta peptide. Neuroscience 92:1465-1474. 124. Koutsilieri E, Riederer P (2007) Excitotoxicity and new antiglutamatergic strategies in Parkinson's disease and Alzheimer's disease. Parkinsonism Relat Disord 13:S329-331. 125. Esposito Z, Belli L, Toniolo S, Sancesario G, Bianconi C, Martorana A (2013) Amyloid beta, Glutamate, Excitotoxicity in Alzheimer's Disease: Are We on the Right Track? CNS Neurosci Ther 19:549-555. 126. Kim SY, Moon JH, Lee HG, Kim SU, Lee YB (2007) ATP released from beta-amyloid-stimulated microglia induces reactive oxygen species production in an autocrine fashion. Exp Mol Med 39:820-827. 127. Nakanishi H (2003) Microglial functions and proteases. Mol Neurobiol 27:163-176.

Page 93: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

67

128. Ito S, Kimura K, Haneda M, Ishida Y, Sawada M, Isobe K (2007) Induction of matrix metalloproteinases (MMP3, MMP12 and MMP13) expression in the microglia by amyloid-beta stimulation via the PI3K/Akt pathway. Exp Gerontol 42:532-537. 129. Gottschall PE (1996) beta-Amyloid induction of gelatinase B secretion in cultured microglia: inhibition by dexamethasone and indomethacin. Neuroreport 7:3077-3080. 130. Roher AE, Kasunic TC, Woods AS, Cotter RJ, Ball MJ, Fridman R (1994) Proteolysis of A beta peptide from Alzheimer disease brain by gelatinase A. Biochem Biophys Res Commun 205:1755-1761. 131. Yan P, Hu X, Song H, Yin K, Bateman RJ, Cirrito JR, Xiao Q, Hsu FF, Turk JW, Xu J, Hsu CY, Holtzman DM, Lee JM (2006) Matrix metalloproteinase-9 degrades amyloid-beta fibrils in vitro and compact plaques in situ. J Biol Chem 281:24566-24574. 132. Yin KJ, Cirrito JR, Yan P, Hu X, Xiao Q, Pan X, Bateman R, Song H, Hsu FF, Turk J, Xu J, Hsu CY, Mills JC, Holtzman DM, Lee JM (2006) Matrix metalloproteinases expressed by astrocytes mediate extracellular amyloid-beta peptide catabolism. J Neurosci 26:10939-10948. 133. Napoli I, Neumann H (2009) Microglial clearance function in health and disease. Neuroscience 158:1030-1038. 134. Frackowiak J, Wisniewski HM, Wegiel J, Merz GS, Iqbal K, Wang KC (1992) Ultrastructure of the microglia that phagocytose amyloid and the microglia that produce beta-amyloid fibrils. Acta Neuropathol 84:225-233. 135. Paresce DM, Ghosh RN, Maxfield FR (1996) Microglial cells internalize aggregates of the Alzheimer's disease amyloid beta-protein via a scavenger receptor. Neuron 17:553-565. 136. Pluta R, Barcikowska M, Misicka A, Lipkowski AW, Spisacka S, Januszewski S (1999) Ischemic rats as a model in the study of the neurobiological role of human beta-amyloid peptide. Time-dependent disappearing diffuse amyloid plaques in brain. Neuroreport 10:3615-3619. 137. Reed-Geaghan EG, Savage JC, Hise AG, Landreth GE (2009) CD14 and toll-like receptors 2 and 4 are required for fibrillar A{beta}-stimulated microglial activation. J Neurosci 29:11982-11992. 138. Mandrekar S, Jiang Q, Lee CY, Koenigsknecht-Talboo J, Holtzman DM, Landreth GE (2009) Microglia mediate the clearance of soluble Abeta through fluid phase macropinocytosis. J Neurosci 29:4252-4262. 139. Lee CY, Landreth GE (2010) The role of microglia in amyloid clearance from the AD brain. J Neural Transm 117:949-960. 140. Liu Z, Condello C, Schain A, Harb R, Grutzendler J (2010) CX3CR1 in microglia regulates brain amyloid deposition through selective protofibrillar amyloid-beta phagocytosis. J Neurosci 30:17091-17101. 141. Paresce DM, Chung H, Maxfield FR (1997) Slow degradation of aggregates of the Alzheimer's disease amyloid beta-protein by microglial cells. J Biol Chem 272:29390-29397. 142. Simard AR, Soulet D, Gowing G, Julien JP, Rivest S (2006) Bone marrow-derived microglia play a critical role in restricting senile plaque formation in Alzheimer's disease. Neuron 49:489-502. 143. Majumdar A, Chung H, Dolios G, Wang R, Asamoah N, Lobel P, Maxfield FR (2008) Degradation of fibrillar forms of Alzheimer's amyloid beta-peptide by macrophages. Neurobiol Aging 29:707-715. 144. Ginhoux F, Lim S, Hoeffel G, Low D, Huber T (2013) Origin and differentiation of microglia. Front Cell Neurosci 7:45. 145. Neniskyte U, Neher JJ, Brown GC (2011) Neuronal death induced by nanomolar amyloid beta is mediated by primary phagocytosis of neurons by microglia. J Biol Chem 286:39904-39913. 146. Mohmmad Abdul H, Butterfield DA (2005) Protection against amyloid beta-peptide (1-42)-induced loss of phospholipid asymmetry in synaptosomal membranes by

Page 94: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

68

tricyclodecan-9-xanthogenate (D609) and ferulic acid ethyl ester: implications for Alzheimer's disease. Biochim Biophys Acta 1741:140-148. 147. Neher JJ, Neniskyte U, Brown GC (2012) Primary phagocytosis of neurons by inflamed microglia: potential roles in neurodegeneration. Front Pharmacol 3:27. 148. Bader Lange ML, Cenini G, Piroddi M, Abdul HM, Sultana R, Galli F, Memo M, Butterfield DA (2008) Loss of phospholipid asymmetry and elevated brain apoptotic protein levels in subjects with amnestic mild cognitive impairment and Alzheimer disease. Neurobiol Dis 29:456-464. 149. Fuller AD, Van Eldik LJ (2008) MFG-E8 regulates microglial phagocytosis of apoptotic neurons. J Neuroimmune Pharmacol 3:246-256. 150. Akiyama H, Kawamata T, Dedhar S, McGeer PL (1991) Immunohistochemical localization of vitronectin, its receptor and beta-3 integrin in Alzheimer brain tissue. J Neuroimmunol 32:19-28. 151. Boddaert J, Kinugawa K, Lambert JC, Boukhtouche F, Zoll J, Merval R, Blanc-Brude O, Mann D, Berr C, Vilar J, Garabedian B, Journiac N, Charue D, Silvestre JS, Duyckaerts C, Amouyel P, Mariani J, Tedgui A, Mallat Z (2007) Evidence of a role for lactadherin in Alzheimer's disease. Am J Pathol 170:921-929. 152. Heneka MT, Kummer MP, Stutz A, Delekate A, Schwartz S, Vieira-Saecker A, Griep A, Axt D, Remus A, Tzeng TC, Gelpi E, Halle A, Korte M, Latz E, Golenbock DT (2013) NLRP3 is activated in Alzheimer's disease and contributes to pathology in APP/PS1 mice. Nature 493:674-678. 153. Glass CK, Saijo K, Winner B, Marchetto MC, Gage FH (2010) Mechanisms underlying inflammation in neurodegeneration. Cell 140:918-934. 154. Solito E, Sastre M (2012) Microglia function in Alzheimer's disease. Front Pharmacol 3:14. 155. Hickman SE, Allison EK, El Khoury J (2008) Microglial dysfunction and defective beta-amyloid clearance pathways in aging Alzheimer's disease mice. J Neurosci 28:8354-8360. 156. Streit WJ, Xue QS (2012) Alzheimer's disease, neuroprotection, and CNS immunosenescence. Front Pharmacol 3:138. 157. Streit WJ (2006) Microglial senescence: does the brain's immune system have an expiration date? Trends Neurosci 29:506-510. 158. Olah M, Biber K, Vinet J, Boddeke HW (2011) Microglia phenotype diversity. CNS Neurol Disord Drug Targets 10:108-118. 159. Wong WT (2013) Microglial aging in the healthy CNS: phenotypes, drivers, and rejuvenation. Front Cell Neurosci 7:22. 160. Harry GJ, Kraft AD (2012) Microglia in the developing brain: a potential target with lifetime effects. Neurotoxicology 33:191-206. 161. Streit WJ, Sammons NW, Kuhns AJ, Sparks DL (2004) Dystrophic microglia in the aging human brain. Glia 45:208-212. 162. Flanary BE, Streit WJ (2004) Progressive telomere shortening occurs in cultured rat microglia, but not astrocytes. Glia 45:75-88. 163. Flanary BE, Streit WJ (2003) Telomeres shorten with age in rat cerebellum and cortex in vivo. J Anti Aging Med 6:299-308. 164. Streit WJ, Braak H, Xue QS, Bechmann I (2009) Dystrophic (senescent) rather than activated microglial cells are associated with tau pathology and likely precede neurodegeneration in Alzheimer's disease. Acta Neuropathol 118:475-485. 165. Luo XG, Ding JQ, Chen SD (2010) Microglia in the aging brain: relevance to neurodegeneration. Mol Neurodegener 5:12. 166. Korotzer AR, Pike CJ, Cotman CW (1993) beta-Amyloid peptides induce degeneration of cultured rat microglia. Brain Res 624:121-125. 167. Flanary BE, Sammons NW, Nguyen C, Walker D, Streit WJ (2007) Evidence that aging and amyloid promote microglial cell senescence. Rejuvenation Res 10:61-74. 168. Imbimbo BP, Solfrizzi V, Panza F (2010) Are NSAIDs useful to treat Alzheimer's disease or mild cognitive impairment? Front Aging Neurosci 2.

Page 95: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

69

169. Biber K, Neumann H, Inoue K, Boddeke HW (2007) Neuronal 'On' and 'Off' signals control microglia. Trends Neurosci 30:596-602. 170. Harris HE, Andersson U, Pisetsky DS (2012) HMGB1: a multifunctional alarmin driving autoimmune and inflammatory disease. Nat Rev Rheumatol 8:195-202. 171. Fang P, Schachner M, Shen YQ (2012) HMGB1 in development and diseases of the central nervous system. Mol Neurobiol 45:499-506. 172. Takata K, Kitamura Y, Kakimura J, Shibagaki K, Tsuchiya D, Taniguchi T, Smith MA, Perry G, Shimohama S (2003) Role of high mobility group protein-1 (HMG1) in amyloid-beta homeostasis. Biochem Biophys Res Commun 301:699-703. 173. Takata K, Takada T, Ito A, Asai M, Tawa M, Saito Y, Ashihara E, Tomimoto H, Kitamura Y, Shimohama S (2012) Microglial Amyloid-beta1-40 Phagocytosis Dysfunction Is Caused by High-Mobility Group Box Protein-1: Implications for the Pathological Progression of Alzheimer's Disease. Int J Alzheimers Dis 2012:685739. 174. Inoue K (2006) The function of microglia through purinergic receptors: neuropathic pain and cytokine release. Pharmacol Ther 109:210-226. 175. Sperlagh B, Illes P (2007) Purinergic modulation of microglial cell activation. Purinergic Signal 3:117-127. 176. Koizumi S, Ohsawa K, Inoue K, Kohsaka S (2013) Purinergic receptors in microglia: functional modal shifts of microglia mediated by P2 and P1 receptors. Glia 61:47-54. 177. Dou Y, Wu HJ, Li HQ, Qin S, Wang YE, Li J, Lou HF, Chen Z, Li XM, Luo QM, Duan S (2012) Microglial migration mediated by ATP-induced ATP release from lysosomes. Cell Res 22:1022-1033. 178. Gemma C, Bachstetter AD, Bickford PC (2010) Neuron-Microglia Dialogue and Hippocampal Neurogenesis in the Aged Brain. Aging Dis 1:232-244. 179. Walker DG, Dalsing-Hernandez JE, Campbell NA, Lue LF (2009) Decreased expression of CD200 and CD200 receptor in Alzheimer's disease: a potential mechanism leading to chronic inflammation. Exp Neurol 215:5-19. 180. Clark AK, Malcangio M (2012) Microglial signalling mechanisms: Cathepsin S and Fractalkine. Exp Neurol 234:283-292. 181. Lee S, Varvel NH, Konerth ME, Xu G, Cardona AE, Ransohoff RM, Lamb BT (2010) CX3CR1 deficiency alters microglial activation and reduces beta-amyloid deposition in two Alzheimer's disease mouse models. Am J Pathol 177:2549-2562. 182. Fuhrmann M, Bittner T, Jung CK, Burgold S, Page RM, Mitteregger G, Haass C, LaFerla FM, Kretzschmar H, Herms J (2010) Microglial Cx3cr1 knockout prevents neuron loss in a mouse model of Alzheimer's disease. Nat Neurosci 13:411-413. 183. Bhaskar K, Konerth M, Kokiko-Cochran ON, Cardona A, Ransohoff RM, Lamb BT (2010) Regulation of tau pathology by the microglial fractalkine receptor. Neuron 68:19-31. 184. Cho SH, Sun B, Zhou Y, Kauppinen TM, Halabisky B, Wes P, Ransohoff RM, Gan L (2011) CX3CR1 protein signaling modulates microglial activation and protects against plaque-independent cognitive deficits in a mouse model of Alzheimer disease. J Biol Chem 286:32713-32722. 185. Meadows DC, Gervay-Hague J (2006) Vinyl sulfones: synthetic preparations and medicinal chemistry applications. Med Res Rev 26:793-814. 186. Santos MM, Moreira R (2007) Michael acceptors as cysteine protease inhibitors. Mini Rev Med Chem 7:1040-1050. 187. Conus S, Simon HU (2010) Cathepsins and their involvement in immune responses. Swiss Med Wkly 140:w13042. 188. Wendt W, Lubbert H, Stichel CC (2008) Upregulation of cathepsin S in the aging and pathological nervous system of mice. Brain Res 1232:7-20. 189. Nakanishi H (2003) Neuronal and microglial cathepsins in aging and age-related diseases. Ageing Res Rev 2:367-381.

Page 96: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

70

190. Haque A, Banik NL, Ray SK (2008) New insights into the roles of endolysosomal cathepsins in the pathogenesis of Alzheimer's disease: cathepsin inhibitors as potential therapeutics. CNS Neurol Disord Drug Targets 7:270-277. 191. Lemere CA, Munger JS, Shi GP, Natkin L, Haass C, Chapman HA, Selkoe DJ (1995) The lysosomal cysteine protease, cathepsin S, is increased in Alzheimer's disease and Down syndrome brain. An immunocytochemical study. Am J Pathol 146:848-860. 192. Schechter I, Ziv E (2011) Cathepsins S, B and L with aminopeptidases display beta-secretase activity associated with the pathogenesis of Alzheimer's disease. Biol Chem 392:555-569. 193. Stansley B, Post J, Hensley K (2012) A comparative review of cell culture systems for the study of microglial biology in Alzheimer's disease. J Neuroinflammation 9:115. 194. Maezawa I, Zimin PI, Wulff H, Jin LW (2011) Amyloid-beta protein oligomer at low nanomolar concentrations activates microglia and induces microglial neurotoxicity. J Biol Chem 286:3693-3706. 195. Takata K, Kitamura Y, Saeki M, Terada M, Kagitani S, Kitamura R, Fujikawa Y, Maelicke A, Tomimoto H, Taniguchi T, Shimohama S (2010) Galantamine-induced amyloid-{beta} clearance mediated via stimulation of microglial nicotinic acetylcholine receptors. J Biol Chem 285:40180-40191. 196. Falcao AS, Fernandes A, Brito MA, Silva RF, Brites D (2005) Bilirubin-induced inflammatory response, glutamate release, and cell death in rat cortical astrocytes are enhanced in younger cells. Neurobiol Dis 20:199-206. 197. Caldeira C, Frederico A, Vaz AR, Fernandes A, Brites D (2013) Microglia reactivity to amyloid-beta oligomers (AbetaO) changes according to experimental ageing. Glia 61:T06-13A. 198. Righi M, Mori L, De Libero G, Sironi M, Biondi A, Mantovani A, Donini SD, Ricciardi-Castagnoli P (1989) Monokine production by microglial cell clones. Eur J Immunol 19:1443-1448. 199. Ni M, Aschner M (2010) Neonatal rat primary microglia: isolation, culturing, and selected applications. Curr Protoc Toxicol Chapter 12: Unit 12.17. 200. Horvath RJ, Nutile-McMenemy N, Alkaitis MS, Deleo JA (2008) Differential migration, LPS-induced cytokine, chemokine, and NO expression in immortalized BV-2 and HAPI cell lines and primary microglial cultures. J Neurochem 107:557-569. 201. Gresa-Arribas N, Vieitez C, Dentesano G, Serratosa J, Saura J, Sola C (2012) Modelling neuroinflammation in vitro: a tool to test the potential neuroprotective effect of anti-inflammatory agents. PLoS One 7:e45227. 202. Culbert AA, Skaper SD, Howlett DR, Evans NA, Facci L, Soden PE, Seymour ZM, Guillot F, Gaestel M, Richardson JC (2006) MAPK-activated protein kinase 2 deficiency in microglia inhibits pro-inflammatory mediator release and resultant neurotoxicity. Relevance to neuroinflammation in a transgenic mouse model of Alzheimer disease. J Biol Chem 281:23658-23667. 203. Carreira BP, Morte MI, Carvalho CM, Araujo IM (2012) Assessing the influence of neuroinflammation on neurogenesis: in vitro models using neural stem cells and microglia as valuable research tools. In: Sun T (ed) Neural stem cells and therapy. InTech, 204. Lanier LM, Gates MA, Witke W, Menzies AS, Wehman AM, Macklis JD, Kwiatkowski D, Soriano P, Gertler FB (1999) Mena is required for neurulation and commissure formation. Neuron 22:313-325. 205. Gordo AC, Falcao AS, Fernandes A, Brito MA, Silva RF, Brites D (2006) Unconjugated bilirubin activates and damages microglia. J Neurosci Res 84:194-201. 206. Silva SL, Vaz AR, Barateiro A, Falcao AS, Fernandes A, Brito MA, Silva RF, Brites D (2010) Features of bilirubin-induced reactive microglia: from phagocytosis to inflammation. Neurobiol Dis 40:663-675. 207. Vaz AR, Delgado-Esteban M, Brito MA, Bolanos JP, Brites D, Almeida A (2010) Bilirubin selectively inhibits cytochrome c oxidase activity and induces apoptosis in

Page 97: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

71

immature cortical neurons: assessment of the protective effects of glycoursodeoxycholic acid. J Neurochem 112:56-65. 208. Clark AK, Wodarski R, Guida F, Sasso O, Malcangio M (2010) Cathepsin S release from primary cultured microglia is regulated by the P2X7 receptor. Glia 58:1710-1726. 209. Fernandes A, Falcao AS, Silva RF, Gordo AC, Gama MJ, Brito MA, Brites D (2006) Inflammatory signalling pathways involved in astroglial activation by unconjugated bilirubin. J Neurochem 96:1667-1679. 210. Bradford MM (1976) A rapid and sensitive method for the quantitation of microgram quantities of protein utilizing the principle of protein-dye binding. Anal Biochem 72:248-254. 211. Kim HJ, Ajit D, Peterson TS, Wang Y, Camden JM, Gibson Wood W, Sun GY, Erb L, Petris M, Weisman GA (2012) Nucleotides released from Abeta(1)(-)(4)(2) -treated microglial cells increase cell migration and Abeta(1)(-)(4)(2) uptake through P2Y(2) receptor activation. J Neurochem 121:228-238. 212. Li W, Poteet E, Xie L, Liu R, Wen Y, Yang SH (2011) Regulation of matrix metalloproteinase 2 by oligomeric amyloid beta protein. Brain Res 1387:141-148. 213. Cai Z, Hussain MD, Yan LJ (2013) Microglia, neuroinflammation, and beta-amyloid protein in Alzheimer's disease. Int J Neurosci. 214. Bruce-Keller AJ, Keeling JL, Keller JN, Huang FF, Camondola S, Mattson MP (2000) Antiinflammatory effects of estrogen on microglial activation. Endocrinology 141:3646-3656. 215. Cui YH, Le Y, Gong W, Proost P, Van Damme J, Murphy WJ, Wang JM (2002) Bacterial lipopolysaccharide selectively up-regulates the function of the chemotactic peptide receptor formyl peptide receptor 2 in murine microglial cells. J Immunol 168:434-442. 216. Fleisher-Berkovich S, Filipovich-Rimon T, Ben-Shmuel S, Hulsmann C, Kummer MP, Heneka MT (2010) Distinct modulation of microglial amyloid beta phagocytosis and migration by neuropeptides (i). J Neuroinflammation 7:61. 217. Neumann H, Kotter MR, Franklin RJ (2009) Debris clearance by microglia: an essential link between degeneration and regeneration. Brain 132:288-295. 218. Lai AY, Dibal CD, Armitage GA, Winship IR, Todd KG (2013) Distinct activation profiles in microglia of different ages: A systematic study in isolated embryonic to aged microglial cultures. Neuroscience 254:185-195. 219. Damani MR, Zhao L, Fontainhas AM, Amaral J, Fariss RN, Wong WT (2011) Age-related alterations in the dynamic behavior of microglia. Aging Cell 10:263-276. 220. Barger SW, Goodwin ME, Porter MM, Beggs ML (2007) Glutamate release from activated microglia requires the oxidative burst and lipid peroxidation. J Neurochem 101:1205-1213. 221. Takeuchi H, Mizuno T, Zhang G, Wang J, Kawanokuchi J, Kuno R, Suzumura A (2005) Neuritic beading induced by activated microglia is an early feature of neuronal dysfunction toward neuronal death by inhibition of mitochondrial respiration and axonal transport. J Biol Chem 280:10444-10454. 222. Saransaari P, Oja SS (1995) Age-related changes in the uptake and release of glutamate and aspartate in the mouse brain. Mech Ageing Dev 81:61-71. 223. Nakajima K, Yamamoto S, Kohsaka S, Kurihara T (2008) Neuronal stimulation leading to upregulation of glutamate transporter-1 (GLT-1) in rat microglia in vitro. Neurosci Lett 436:331-334. 224. Sanz JM, Chiozzi P, Ferrari D, Colaianna M, Idzko M, Falzoni S, Fellin R, Trabace L, Di Virgilio F (2009) Activation of microglia by amyloid {beta} requires P2X7 receptor expression. J Immunol 182:4378-4385. 225. Schneider SW, Egan ME, Jena BP, Guggino WB, Oberleithner H, Geibel JP (1999) Continuous detection of extracellular ATP on living cells by using atomic force microscopy. Proc Natl Acad Sci U S A 96:12180-12185.

Page 98: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

72

226. Sandeman SR, Faragher RG, Allen MC, Liu C, Lloyd AW (2001) Does MMP-2 expression and secretion change with increasing serial passage of keratocytes in culture? Mech Ageing Dev 122:157-167. 227. Ihara M, Tomimoto H, Kinoshita M, Oh J, Noda M, Wakita H, Akiguchi I, Shibasaki H (2001) Chronic cerebral hypoperfusion induces MMP-2 but not MMP-9 expression in the microglia and vascular endothelium of white matter. J Cereb Blood Flow Metab 21:828-834. 228. Xie F, Zhang JC, Fu H, Chen J (2013) Age-related decline of myelin proteins is highly correlated with activation of astrocytes and microglia in the rat CNS. Int J Mol Med 32:1021-1028. 229. Bartzokis G, Cummings JL, Sultzer D, Henderson VW, Nuechterlein KH, Mintz J (2003) White matter structural integrity in healthy aging adults and patients with Alzheimer disease: a magnetic resonance imaging study. Arch Neurol 60:393-398. 230. Filippov MA, Dityatev A (2012) Matrix metalloproteinase-9 and non-amyloidogenic pathway of amyloid precursor protein processing. J Neurochem 121:181-183. 231. Cunha J (2012) Exploring motor neuron degeneration in ALS: prevention by glycoursodeoxycholic acid and signaling to microglia. Master thesis, Universidade Nova de Lisboa, Lisboa 232. Shang YC, Chong ZZ, Wang S, Maiese K (2012) Prevention of beta-amyloid degeneration of microglia by erythropoietin depends on Wnt1, the PI 3-K/mTOR pathway, Bad, and Bcl-xL. Aging (Albany NY) 4:187-201. 233. Manicone AM, McGuire JK (2008) Matrix metalloproteinases as modulators of inflammation. Semin Cell Dev Biol 19:34-41. 234. Park SY, Jin ML, Kim YH, Kim Y, Lee SJ (2012) Anti-inflammatory effects of aromatic-turmerone through blocking of NF-kappaB, JNK, and p38 MAPK signaling pathways in amyloid beta-stimulated microglia. Int Immunopharmacol 14:13-20. 235. Wilcock DM, Morgan D, Gordon MN, Taylor TL, Ridnour LA, Wink DA, Colton CA (2011) Activation of matrix metalloproteinases following anti-Abeta immunotherapy; implications for microhemorrhage occurrence. J Neuroinflammation 8:115. 236. Du H, Li P, Wang J, Qing X, Li W (2012) The interaction of amyloid beta and the receptor for advanced glycation endproducts induces matrix metalloproteinase-2 expression in brain endothelial cells. Cell Mol Neurobiol 32:141-147. 237. Mizoguchi H, Takuma K, Fukuzaki E, Ibi D, Someya E, Akazawa KH, Alkam T, Tsunekawa H, Mouri A, Noda Y, Nabeshima T, Yamada K (2009) Matrix metalloprotease-9 inhibition improves amyloid beta-mediated cognitive impairment and neurotoxicity in mice. J Pharmacol Exp Ther 331:14-22. 238. Reijerkerk A, Kooij G, van der Pol SM, Khazen S, Dijkstra CD, de Vries HE (2006) Diapedesis of monocytes is associated with MMP-mediated occludin disappearance in brain endothelial cells. FASEB J 20:2550-2552. 239. Desai BS, Monahan AJ, Carvey PM, Hendey B (2007) Blood-brain barrier pathology in Alzheimer's and Parkinson's disease: implications for drug therapy. Cell Transplant 16:285-299. 240. Li H, Xu H, Sun B (2012) Lipopolysaccharide regulates MMP-9 expression through TLR4/NF-kappaB signaling in human arterial smooth muscle cells. Mol Med Rep 6:774-778. 241. Ohnishi T, Muroi M, Tanamoto K (2003) MD-2 is necessary for the toll-like receptor 4 protein to undergo glycosylation essential for its translocation to the cell surface. Clin Diagn Lab Immunol 10:405-410. 242. Kopec KK, Carroll RT (1998) Alzheimer's beta-amyloid peptide 1-42 induces a phagocytic response in murine microglia. J Neurochem 71:2123-2131. 243. Pan XD, Zhu YG, Lin N, Zhang J, Ye QY, Huang HP, Chen XC (2011) Microglial phagocytosis induced by fibrillar beta-amyloid is attenuated by oligomeric beta-amyloid: implications for Alzheimer's disease. Mol Neurodegener 6:45.

Page 99: Repositório da Universidade de Lisboa: Página principal - …repositorio.ul.pt/bitstream/10451/10326/1/ulfc103181... · 2018. 10. 26. · à morte. O principal factor de risco é

73

244. Koenigsknecht-Talboo J, Landreth GE (2005) Microglial phagocytosis induced by fibrillar beta-amyloid and IgGs are differentially regulated by proinflammatory cytokines. J Neurosci 25:8240-8249. 245. Neniskyte U, Brown GC (2013) Lactadherin/MFG-E8 is essential for microglia-mediated neuronal loss and phagoptosis induced by amyloid beta. J Neurochem 126:312-317. 246. Chavan SS, Huerta PT, Robbiati S, Valdes-Ferrer SI, Ochani M, Dancho M, Frankfurt M, Volpe BT, Tracey KJ, Diamond B (2012) HMGB1 mediates cognitive impairment in sepsis survivors. Mol Med 18:930-937. 247. Qin Y, Cao X, Guo J, Zhang Y, Pan L, Zhang H, Li H, Tang C, Du J, Shi GP (2012) Deficiency of cathepsin S attenuates angiotensin II-induced abdominal aortic aneurysm formation in apolipoprotein E-deficient mice. Cardiovasc Res 96:401-410. 248. Xu J, Wang H, Ding K, Lu X, Li T, Wang J, Wang C, Wang J (2013) Inhibition of cathepsin S produces neuroprotective effects after traumatic brain injury in mice. Mediators of inflammation 2013:187873. 249. Wang YR, Qin S, Han R, Wu JC, Liang ZQ, Qin ZH, Wang Y (2013) Cathepsin L Plays a Role in Quinolinic Acid-Induced NF-Kappab Activation and Excitotoxicity in Rat Striatal Neurons. PLoS One 8:e75702. 250. Tak PP, Firestein GS (2001) NF-kappaB: a key role in inflammatory diseases. J Clin Invest 107:7-11.