Neuroinflammation in Alzheimer‟s disease Focus on NF-κB and C/EBP transcription factors

Veronica Ramberg

i Cover photo: James Lauritz, Getty Images

©Veronica Ramberg, Stockholm 2011

ISBN 978-91-7447-356-8

Printed in Sweden by Universitetsservice US-AB, Stockholm 2011 Distributor: Department of Neurochemistry, Stockholm University

ii

Till Carl, Victor och Erik

iii

iv List of publications

This thesis is based on the following publications, referred to in the text by their corresponding Roman numerals:

I Fisher L., Samuelsson M., Jiang Y., Ramberg V., Figueroa R., Hallberg E., Langel Ü., Iverfeldt K. “Targeting expression in rat primary glial cells by cellular delivery of an NF-κB decoy.” Journal of Molecular Neuroscience, 31:209-219 (2007)

II Samuelsson M., Ramberg V., Iverfeldt K. “Alzheimer Aβ peptides block the activation of C/EBPβ and C/EBPδ in glial cells.” Biochemical and Biophysical Research Communica- tions, 370:619-622 (2008)

III Ramberg V., Tracy L., Samuelsson M., Nilsson L.N.G., Iverfeldt K. “The CCAAT/enhancer binding protein (C/EBP) δ is differently regulated by fibrillar and oligomeric forms of the Alzheimer amyloid-β peptide.” Journal of Neuroinflammation, 8:34 (2011)

IV Figueroa R.*, Ramberg V.*, Gatsinzi T., Samuelsson M., Zhang M., Iverfeldt K., Hallberg E. “Anchored FRET sensors detect local caspase activation prior to neuronal degeneration.” Molecular , 6:35 (2011)

*The first two authors made equal contributions.

v

vi Abstract

Alzheimer‟s disease (AD) is the most common form of dementia among elderly. The disease is characterized by amyloid-β (Aβ) plaques, neurofibrillary tangles, loss of synapses and neurons and chronic neuroinflammation. The significance of neuroinflammatory processes in disease on-set and progression has been debated since activated and reactive have been attributed both protective and damaging properties. However, patients systematically treated with anti-inflammatory drugs have been shown to develop AD to a lesser extent than average. This indicates an important role of neuroinflammation in AD. This thesis focuses on two inflammatory related transcription fac- tors, nuclear factor κB (NF-κB) and CCAAT/enhancer binding protein (C/EBP). Both NF-κB and C/EBP are known regulators of many pro- inflammatory genes and may during certain circumstances dimerize with each other. In paper I we use a new strategy to inhibit NF-κB DNA binding ac- tivity in primary astro-microglial cell cultures treated with Aβ and IL- 1β. By coupling the NF-κB decoy to a transport peptide both concen- tration and incubation time can be shortened in comparison to previ- ous studies. Moreover, using the same in vitro model in paper II and III, we show members of the C/EBP family to be dysregulated during AD mimicking conditions. Additional focus was directed towards C/EBPδ, which was shown to respond differently to oligomeric and fibrillar forms of Aβ. Results were also confirmed in vivo using an AD mouse model characterized by high levels of fibrillar Aβ deposits. Finally, in order to get further insight in neurodegenerative processes, induced by Aβ or microglial activation, we present in paper IV a new set of anchored sensors for detection of locally activated caspases in neuronal cells. By anchoring the sensors to tau they become less dy- namic and caspase activation can be detected early on in the apoptotic process, in a spatio-temporal and reproducible manner.

vii

viii Contents

Abbreviations xi Introduction 1 Alzheimer’s disease 1 Pathological hallmarks 2 Neuroinflammation in AD 3 The immune system of the 4 Microglia 4 Astrocytes 8 Microglia- cross talk 10 Inflammaging 11 Transcription factors involved in the inflammatory response 12 Transcription and the role of transcription factors 13 Nuclear factor-B 14 CCAAT/enhancer binding protein 15 - effector molecules in AD 19 Interleukin-1 19 Interleukin-6 20 Non-steroidal anti-inflammatory drugs and AD 21 Neurodegeneration in AD 22 Roles of caspases in apoptotic cell death 22 The involvement of caspases in AD 23 Methods to determine caspase activation and cell death 24 Aims of the thesis 26 Methodological considerations 27 In vitro and in vivo model systems 27 Primary astro-microglial cultures (paper I-III) 27 Arc-Swe transgenic mice (paper III) 27 Cell lines (paper IV) 28 Cell treatments 28 IL-1β and LPS (paper I-III) 28 Amyloid-β (paper I-IV) 29 Cell penetrating peptides (CPP) (paper I) 30 KillerRed (paper IV) 30

ix Staurosporine (paper IV) 30 Studies of protein-DNA interactions 31 Electrophoretic mobility shift assay (paper I-III) 31 Streptavidin-agarose pull-down (paper III) 31 Reverse transcriptase-polymerase chain reaction (paper I) 32 Western blot (paper II-IV) 33 Laser scanning confocal microscopy (paper I and IV) 33 Results and discussion 35 Paper I: Targeting cytokine expression in rat primary glial cells by cellular delivery of an NF-κB decoy. 35 Paper II: Alzheimer Aβ peptides block the activation of C/EBPβ and C/EBPδ in glial cells. 37 Paper III: The CCAAT/enhancer binding protein (C/EBP) δ is differently regulated by fibrillar and oligomeric forms of the Alzheimer amyloid-β peptide. 38 Paper IV: Anchored FRET sensors detect local caspase activation prior to neuronal degeneration. 41 Concluding remarks 45 Populärvetenskaplig sammanfattning på svenska 46 Acknowledgements 48 Referenes 50

x Abbreviations

Aβ Amyloid-β AD Alzheimer‟s disease ADAPT Alzheimer‟s Disease Anti-Inflammatory Prevention Trial Apaf-1 Apoptotic protease-activating factor 1 APP Amyloid precursor protein BACE β-site APP cleaving enzyme BBB Blood brain barrier bZIP Basic leucine zipper CARD Caspase-recruitment domain cDNA Complementary Deoxyribonucleic acid C/EBP CCAAT/enhancer binding protein CNS COX Cyclooxygenase CPP Cell penetrating peptide DAMP Damage-associated molecular pattern DISC Death-inducing signalling complex ECFP Enhanced cyan fluorescent protein EYFP Enhanced yellow fluorescent protein EMSA Electrophoretic mobility shift assay FRET Fluorescence resonance energy transfer GFAP Glial fibrillary acidic protein IAP Inhibitor of protein IFN Interferon IKK IB kinase IL Interleukin IL-1ra IL-1 receptor antagonist IRAK IL-1 receptor-associated kinase IRF Interferon regulatory factor iNOS Inducible LPS Lipopolysaccharide LTP Long-term potentiation MCI Mild cognitive impairment MHCII Major histocompability complex type II mRNA Messenger RNA MyD88 Myeloid differentiation primary response gene 88 NCI Noncognitively impaired

xi NF-κB Nuclear factor κB NES Nuclear export signal NFT Neurofibrillary tangles NLR NOD-like receptor NLS Nuclear localization signal NSAID Non-steroidal anti-inflammatory drug PNA Peptide nucleic acid RA Retinoic acid RAGE Receptor for advanced glycation endproducts RNA Ribonucleic acid rRNA Ribosomal RNA ROS RT-PCR Reverse transcriptase-polymerase chain reaction SASP Senescence-associated secretory phenotype Stat Signal transducer and activators of transcription TAK Transforming growth factor  activated kinase TF Transcription factor Tg-ArcSwe Transgenic mice carrying both the Artic and the Swedish APP mutation TGF Transforming growth factor TIR Toll-interleukin 1 receptor TIRAP TIR-domain-containing adapter protein TLR Toll-like receptor TNF TP Transportan TRAF TNF receptor-associated factor TRAM TRIF-related adaptor molecule TRIF TIR-domain-containing adapter-inducing interferon- TUNEL Terminal deoxynucleotidyl transferase dUTP nick end label- ling WB Western blot

xii Introduction

Neuroinflammation, characterized by activated microglia and reactive astrocytes, has been shown to be involved in several neurodegenera- tive diseases including Alzheimer‟s disease (AD). Even though the neuroinflammatory reaction was documented over 100 years ago, al- most at the same time as Alois Alzheimer defined AD, the role of neu- roinflammation is still debated. Lately, many researchers have started to view neuroinflammation as a double-edged sword. In order to define new therapeutic targets, it is of great importance to delineate the inflammatory mediators and signalling pathways af- fected in AD. In this thesis, the main focus will be on transcription factors (TF) involved in the inflammatory response induced by amy- loid-β (Aβ) peptides. In the last paper a new model system, to study neurodegeneration in a spatio-temporal manner, is presented.

Alzheimer‟s disease Many of us have someone within our circle of friends or relatives that suffer from AD, the most common form of dementia among elderly. It is a devastating illness, first and foremost for the affected person and his or her relatives, but also for society due to enormous costs for community and residential care. Early symptoms of the disease are confusion, mood swings and insomnia. When the disease progresses and the get impaired, the affected person loses the ability to perform activities of daily living. In the very late phase of the disease, the person is unable to perform even the simplest tasks and the cause of death is often a normally harmless infection. In a Delphi consensus study, published in the Lancet 2005 [1], 12 in- ternational experts in the field of AD estimated the prevalence of AD to increase by 100% between the years 2001 and 2040 in developed countries and by more than 300% in less developed countries. The estimation was 4.6 million new AD cases every year, one new case every 7 seconds. Today there is no cure for AD, only symptomatic treatments. One important reason for that is the lack of biomarkers for early detection of the disease in clinical routine. Most patients get their diagnosis after

1 neuropsychological, cognitive and functional tests (like the mini- mental state examination [2]). Since AD is believed to have a long asymptomatic pre-clinical phase, all potential treatments are consid- ered „too late‟.

Pathological hallmarks The pathological hallmarks of AD are:

 Extracellular plaques - consisting of A. A is produced by sequential cleavage of the Aβ precursor protein (APP) by - and -secretase. The most commonly occurring A pep- tides are 40 or 42 amino acids long. They are naturally secreted but their biological function is not fully clear. In AD tg-mice, A has also been shown to accumulate intraneuronally before extracellular depos- its are apparent [3]. Furthermore, A(1-42) has gained more attention than A(1-40), due to higher propensity to aggregate and more pro- nounced cytotoxic properties.

 Intracellular neurofibrillary tangles (NFT) - consisting of the hyperphosphorylated protein tau. Tau is normally involved in the assembly and stabilization of microtu- bule and exists in six different isoforms. When tau becomes hyper- phosphorylated, as in AD, it detaches from microtubule and gains tox- ic functions presumably by interfering with axonal transport and form- ing cytotoxic aggregates [4]. The distribution and abundance of tau are well correlated with neurodegeneration and clinical symptoms in AD progression [5].

 Loss of synapses and neurons. Neurons in entorhinal cortex are first affected, followed by hippocam- pus, amygdala and neocortical areas [4].

 Chronic neuroinflammation. There are several signs of in AD ; activation of complement-, cytokine-, chemokine- and acute phase pathways [6].

Today, the general view on disease progression is based on the „amy- loid cascade hypothesis‟ [7, 8], stating that A is the causative agent in AD and that the rest of the AD characteristics are results of improper deposition of the peptide. In support of this hypothesis are mutations causing dominant inheritable forms of AD („early-onset‟ AD, 5% of all AD cases) which all affect the properties or production of A (e.g. the „Swedish‟ mutation [9] and the „artic‟ mutation [10]). Additionally,

2 people with Down syndrome having a trisomy in chromosome 21, which contains the gene encoding for APP, nearly exclusively develop AD in their middle age. However, the majority of AD patients are diagnosed after the aged of 65 and is thereby said to have „sporadic late-onset‟ AD. The link between A deposition and tau phosphorylation is not ful- ly understood, and thereby a hotspot in AD research. Quite recently, two different signal transduction pathways for A induced hyperphos- phorylation of tau were suggested [3]: I) Intracellular A causes mito- chondrial dysfunction, which results in oxidative stress and production of reactive oxygen species (ROS). ROS activate c-Jun N-terminal ki- nase, which induces phosphorylation of tau indirectly. II) Extracellular A binds to the 7 nicotinic acetylcholine receptor, which leads to activation of an intracellular signalling pathway and activation of gly- cogen synthase kinase-3. This kinase has been shown to be highly involved in tau phosphorylation [11]. However, further research is needed to clarify this matter.

Neuroinflammation in AD Neuroinflammation associated with AD often is viewed as a second- ary response to A deposition and neuronal cell death. Yet, this possi- bly later neuroinflammatory response can be as pathologically rele- vant as the inducer of it. Furthermore, inflammaging (discussed on page 11) has been proposed to be an important contributing factor in the development of late-onset AD. As mentioned earlier, neuroinflammation in AD is often viewed as a double-edged sword where activated microglia and reactive astro- cytes, the immune cells of the brain, both have beneficial effects by degrading A [12, 13] and adverse effects by producing cytotoxic sub- stances that contribute to disease progression [6]. A certain degree of activation is important for microglial of A [14], whereas overstimulation ultimately may lead to neuronal cell death. Overpro- duction of inflammatory molecules can affect various aspects of dis- ease progression. Firstly, senile plaques contain many other substitu- ents besides A, e.g. acute phase proteins [15] and complement factors [16]. These A-associated proteins have been shown to affect transport, fibrillogenesis and deposition of A [17][18] and thereby disturb the balance between A deposition and removal. Secondly, the cytokine interleukin-1 (IL-1) has been reported to affect A production and deposition in vitro, by promoting synthesis [19] and processing [20] of APP. Thirdly, late-onset AD has been proposed to be a „storage dis-

3 ease‟, partly due in ineffective removal of A by chronically activated glial cells [21, 22]. From a pathological point of view, there are few studies analysing inflammatory markers related to late-onset AD even though at least 60% of variation in cognitive decline has been proposed to be heredi- tary [23]. In a study analysing middle-aged offspring with a parental history of late-onset AD it was clearly shown that increased levels of inflammatory markers was a risk factor for developing AD in old age [24]. Compared to aged matched offspring without a parental history of AD, offspring to AD patients showed increased expression of several different cytokines including IL-1β, tumour necrosis factor (TNF) α and IL-6 when whole blood samples were stimulated with lipopoly- saccharide (LPS) ex vivo. Although this study was performed in whole blood samples, it has been proposed that cytokine response in blood can induce a similar response across the blood brain barrier (BBB) [25]. Furthermore, metabolic diseases, such as diabetes [26], are known risk factors for developing AD. In a study with elderly having a metabolic syndrome, it was shown that high levels of inflammation, as measured by plasma IL-6 levels and serum C-reactive protein levels, had a nega- tive effect on cognitive performance [27]. In contrast, elders with a metabolic syndrome but with low levels of inflammation did not have an increased risk of cognitive impairment. Thereby it is tempting to suggest that inflammation has a significant role in metabolic diseases being risk factors for developing AD.

The immune system of the brain The brain was for a long time considered an immune privileged organ separated from the rest of the body by the BBB. That perception has been shown to be incorrect and today it is generally accepted that pe- ripheral monocytes/ have the capacity to enter the brain in response to injury. However, it should be noted that the primary defence against infection or injury is still provided by the CNS resi- dent immune cells microglia and astrocytes.

Microglia The origin of microglia has been hard to determine. Tentatively they derive from mesenchymal myeloid precursor cells, which migrate into the brain during early embryogenesis. In the CNS, the microenviron- ment induces their maturation into fully developed microglia [28, 29]. At a later phase of embryogenesis, fetal macrophages arising from mono- cytes but also of myeloid origin are believed to enter the developing

4 brain and become a part of the mature microglia population [30]. Due to the lack of specific lineage markers these possible types of micro- glia are indistinguishable although differences are believed to exist. It has also been hard to determine how the very long-lived microglia renews them self. Most probably the major part of the self-renewal is via cell division. Replacement of senescent cells by infiltrating mye- loid precursor cells is a rarely occurring event [31]. Microglial cells are very plastic and display several different phe- notypes depending on e.g. age [32, 33] and brain region [34]. They are commonly referred to as „resting‟ or „activated‟. However these phrases are somewhat misleading. Microglial cells are never resting. In their „resting‟ state, morphologically characterized by long, rami- fied and dynamic processes, they are constantly moving, scanning the environment for agents that could be harmful for the tissue. In vivo imaging of live „resting‟ microglia showed that the entire brain under- goes surveillance every few hours [35]. Their processes can be extend- ed and retracted with a radius of ~80 m [35] and the speed of migra- tion can be 20-35 m/h [36]. They are kept in their „resting‟ state via interactions with neurons, which express cell-surface ligands that in- teract with microglial receptors [29]. Interestingly, one of these ligand- receptor interactions, CD200-CD200R, has been shown to be reduced in cortex and in human AD brain tissues [37]. The „activated‟ state, characterized by an amoeboid morphology and short processes, is achieved in response to chemical or cell-based cues. There is not only one „activated‟ microglial phenotype, but sev- eral. Also, the activation has been proposed to occur gradually from a partially activated to a fully activated state [31]. The different pheno- types are distinguished with respect to cytokine and cell surface recep- tor expression. However, if generalized, microglia can be classically activated (M1) via toll-like receptors (TLR) or interferon , expressing pro-inflammatory molecules, or alternatively activated (M2) by the anti-inflammatory cytokines IL-4 or IL-13, expressing anti- inflammatory molecules. However, the outcome of the CNS injury will depend on both the stimuli causing insult and secondary events that influence the microglial phenotype.

Microglia in AD In AD brains it seems like the balance between microglial phenotypes is shifted towards a pro-inflammatory phenotype with less phagocytic abilities. Since the activating stimuli, A, cannot be removed properly it puts a highly coordinated self-limiting system out of function. This hypothesis has been supported by an in vivo study using the ligand 11C-R-PK11195 for the peripheral benzodiazepine receptor (shown to be up-regulated in activated microglial cells), in combination with

5 positron emission tomography [21]. This study clearly showed activat- ed microglia in cortex of AD patients and that the 11C-R-PK11195 signals could be inversely correlated with cognitive function. Alt- hough most importantly, when the cognitive performance declined over time, the 11C-R-PK11195 signals increased whereas the plaque depositions remained the same. Later, a correlation between cognitive decline and microglial activation was suggested for patients with mild cognitive impairment (MCI) as well [38]. Furthermore, Koenigsknecht- Talboo and Landreth showed that pro-inflammatory cytokines could inhibit microglial phagocytosis of A in vitro [22]. The inhibitory effect was achieved by activation of NF-B and subsequent production of binding to the E prostanoid receptor. Interestingly, the nonsteroidal anti-inflammatory drug (NSAID) ibuprofen could rescue A-elicited phagocytosis in a pro-inflammatory milieu. Microglial TLR signalling in AD A has been shown to bind to several microglial surface receptors, such as: TLRs, receptor for advanced glycation endproducts (RAGE) and NOD-like receptors (NLRs) (reviewed in [39]). Lately, TLRs have gained most attention and the A-TLR signalling pathway is an area of intense investigation (reviewed in [40]). Briefly, TLR signalling in general (further illustrated in figure 1 page 7) can be divided into MyD88-dependent (myeloid differentiation primary response gene 88) „early-phase response‟ and TRIF-dependent (toll-interleukin 1 recep- tor (TIR)-domain-containing adapter-inducing interferon- (IFN-)) „late-phase response‟ pathways, both characterized by the recruitment of numerous adaptor proteins. The MyD88-dependent pathway starts with TIRAP (TIR-domain-containing adapter protein) binding to the TLR complex and thereby recruiting MyD88 via its C-terminal TIR domain. The C-terminal of MyD88 consists of a death domain, which then via homotypic interaction activates IRAK-4 (IL-1 receptor- associated kinase 4). IRAK-4 is then, via IRAK-1 and -2, forming a complex and activating TRAF6 (TNF receptor-associated factor 6), which leads to TAK1 (transforming growth factor  activated kinase 1) activation and finally initiation of MAPK and nuclear factor B (NF-B). The TRIF-dependent pathway has been suggested to start with endocytosis of the TLR/TIRAP/MyD88 complex following an exchange in adapter proteins to TRAM (TRIF-related adaptor mole- cule) and TRIF [41]. Thereafter, TRIF can recruit and activate TRAF6 following the MyD88-dependent pathway or activate TRAF3 and sub- sequently IRF3 (interferon regulatory factor 3), which can translocate to the nucleus and initiate transcription of IFN-.

6

Figure 1. Illustration of TLR and IL-1R signalling cascades. Upon ligand binding, adaptor proteins are recruited to the intracellular domain of the receptor via homo- typic interactions. The signalling pathway for the IL-1R and the ‘early-phase re- sponse’ of TLR is similar down-stream of MyD88, finally resulting in MAPK and NF-κB activation and expression of pro-inflammatory molecules. The TLR ‘late- phase response’ is characterized by internalization of the receptor complex, a change in adaptor proteins and dimerization and activation of IRF3 and expression of IFN-β.

In AD, there are several studies pointing towards important roles of TLRs in microglial activation. For instance, it has been shown that microglial activation by fibrillar A is dependent on TLR2, TLR4 and CD14 (a TLR co-receptor) [42]. Furthermore, all three receptors have been shown to be up-regulated in animal models of AD, AD brains and AD patients and also in mice brains [43-46]. Moreover, it has

7 been reported that AD tg mice expressing non-functional TLR4, showed less production of the pro-inflammatory cytokines TNF, IL- 1 and IL-17 [47]. Thereby, the authors suggested TLR4 as a future therapeutic target for AD. Finally, an Italian study showed that a TLR4 polymorphism, resulting in blunted TLR signalling, decreased the risk for late onset AD by 2.7 fold [48]. The mutation, where an as- partic acid in position 299 was changed to a glycine, was significantly higher in non-demented age matched controls. However, conflicting results exist. For instance, TLR2 has been reported as an important receptor for clearance of A, where activation resulted in delayed cognitive decline in a mouse model of AD [14]. Though, the contradic- tory result might be explained by the age of the animals assessed in the study. Suggestively, microglia have phagocytic capabilities at ear- ly stages of the disease but when the disease progresses, and the genes involved in A clearance are down-regulated, they display a more cytotoxic phenotype [49].

Astrocytes In addition to microglia, astrocytes (astroglia) are also activated in response to chemical or cell-based cues. Even though astrocytes have been shown to have phagocytic capabilities as well as ability to pro- duce harmful cytotoxic molecules, they have not received as much attention as microglia. For a long time they were considered as the „brain glue‟, not able to communicate with each other and there just to organize the neuronal network. However, the present view of astro- cytes has changed dramatically and today astrocyte dysfunction has been shown to be involved in numerous different neuropathologies (reviewed in [50]). Astrocytes, which are of neuroectoderm origin, are the most com- mon cell type in the CNS representing about 50% of the cells. They are highly plastic, star-shaped cells with long processes that com- municate with each other via gap junctions. In order to sense their surrounding and rapidly respond to changes in the environment they are highly organized into „astrocytic domains‟ showed to have mini- mal overlap. Thereby a single astrocyte can cover a specific territory and it can act as a bridge between other kinds of cells in the CNS [51, 52]. Astrocytes are of great importance for neuronal function and sur- vival and also for overall brain homeostasis. The roles of astrocytes seem innumerable; providing energy to neurons by the astrocyte- neuron lactate shuttle system, control glutamate, ion and water home- ostasis, modulate synaptic formation, activity and remodelling, be a part of the defence system and formation and maintenance of the BBB (reviewed in [50, 53]).

8 When astrocytes become activated, they withdraw their processes and a „calcium wave‟ is initiated where intracellular Ca2+ is increased and propagated to neighbouring astrocytes. This results in activation of intracellular signalling pathways ultimately leading to further mor- phological changes and rapid motility of the cells. However, activated astrocytes are most often referred to as „reactive‟ astrocytes. This is due to the conventional immunologic criterion of activation, which demands expression of the major histocompatibility complex type II (MHCII) [6]. Astrocytes can be activated without expressing the MHCII and additionally, activation is not always synonymous with proliferation. The activation status of astrocytes is usually measured by the expression of glial fibrillary acidic protein (GFAP), which is an intermediate filament protein shown to be up-regulated in response to activation. Once activated, astrocytes are capable of expressing sever- al inflammatory mediators like complement components, IL-1, IL-6, cyclooxygenase (COX) -2 and iNOS [6]. Astrocytes in AD ‘Quite possibly saving astrocytes from dying in neurological disease would be a far more effective strategy than trying to save neurons, glia already know how to save neurons, whereas neuroscientists still have no clue’. Ben Barres [54]

In AD brains there are clear signs of (increased number of astrocytes) in plaque-containing areas [6]. Additionally, synchronous hyperactivity and intercellular Ca2+ waves have been observed in as- trocytes in response to A, both in vivo and in vitro [55, 56]. However, there are several different suggestions of how astrocytes contribute to AD pathogenesis. The perhaps most obvious contribution to disease progression is the release of cytotoxic compounds in response to A, which might result in microglia activation and a vicious cycle of chronic inflammation (this will be discusses in the next subsection). Moreover, a recently published study showed that rat primary astro- cytes in co-culture with neurons directly contributed to A-induced and tau phosphorylation [57]. The negative effect of as- trocytes was suggested to depend on a factor released into the media resulting in increased neuronal caspase-3 activity. There is also the „neuro-neglect hypothesis‟ coined by Fuller and colleagues, which rather highlights the failure in astrocyte-neuron in- teraction [58]. This means that A-activated astrocytes neglect their neuro-supportive roles, making the neurons even more vulnerable to toxic insults. There are especially three areas of astrocyte dysfunction reported from post-mortem and in vivo studies of AD brains: in astro-

9 cytic energy metabolism, glutamate recycling and glutathione supply [58]. Furthermore, in contrast to microglia, which do not seem to express detectable levels of APP mRNA [59] or the -secretase BACE1 (β-site APP cleaving enzyme) [60], it is not determined if astrocytes produce A or not. Representing at least 50% of the cells in the brain paren- chyma, it is of great importance to find out if they also directly con- tribute to the senile plaque deposition. In two subsequent studies, Rossner and colleagues showed that BACE1 is expressed and synthe- sized in astrocytes in close proximity to A-plaques in tg2576 mice [60, 61]. However, the expression seemed to be limited to chronic gliosis and influenced by pro-inflammatory cytokines released by microglia. Additionally, astrocytic BACE1 expression could be observed in the neocortex of AD patients.

Microglia-astrocyte cross talk It is well established that both microglia and astrocytes are activated in response to A and that they communicate with each other in a bi- directional manner. Microglial release of IL-1 has been shown to lead to astrocyte acti- vation and astrogliosis, followed by release of inflammatory mole- cules. In contrast, IL-1 may also stimulate by mediat- ing astrocytic release of the neurosupportive molecule transforming growth factor  (TGF-) [52]. In turn, TGF- may pursue inhibitory effects on microglia. An in vitro study (using murine microglial BV-2 and primary microglial cells) showed less clustering and chemotactic migration of microglia towards A aggregates, when cells were pre- treated with TGF- [62]. When astrocytes are activated first, calcium waves can spread the signal to microglia. It has been shown that astrocytes in response to injury can release ATP that will activate microglia, resulting in micro- glial proliferation, migration and release of cytotoxic compounds [52]. In AD, the microglial-astrocyte cross talk may generate a vicious cycle. For instance, A-activated microglia have the capacity to accel- erate the activation of astrocytes, following further activation of dis- tant microglia. Whether a negative cycle is initiated or not, might de- pend on the degree of inflammation and the duration of the A stimu- li.

10

Figure 2. Illustration of the dual role of neuroinflammation in AD. Upon activation, microglia and astrocytes can display different phenotypes, phagocytic or neurotoxic, depending on the stimuli. In AD, there are strong indications that the cytotoxic phe- notype is favourable, since the inducing Aβ stimuli cannot be removed properly. The figure is also illustrating the changes in morphology upon activation and the cross- talk between microglia and astrocytes. The figure is inspired by [31, 63].

Inflammaging Age is by far the single most important risk factor for developing AD. In the year of 2000, Franceschi and colleagues coined the term „in- flamm-aging‟ to describe an imbalance between inflammatory and anti-inflammatory networks observed in elderly individuals [64]. The starting point of the project, which has now resulted in numerous pub- lications regarding aging and longevity, was the observation that pe- ripheral blood mononuclear cells from healthy elderly individuals produced higher amounts of pro-inflammatory cytokines than cells from younger people [65]. The characteristics of inflammaging differ significantly from the general inflammatory signs, in that it is a low- grade, controlled, asymptomatic, chronic and systemic state of in- flammation [66]. The importance of inflammaging in AD has been shown in studies investigating „high pathology controls‟. These are non-demented per- sons exhibiting enough A and NFT to qualify for an AD diagnosis. The thing that distinguishes them the most from AD patients is the

11 dramatically lower production of pro-inflammatory molecules [67]. Furthermore, there is an IL-10 promoter polymorphism resulting in higher expression of the anti-inflammatory cytokine IL-10 in males carrying this mutation, which is underrepresented in patients affected by late-onset AD [68, 69]. Astrocytes in the normal aging brain have been shown to express characteristics of a senescence-associated secretory phenotype (SASP) (review in [70]). This phenotype is linked to an increased expression and secretion of pro-inflammatory factors such as ILs, chemokines and nitric oxide. Both NF-B and CCAAT/enhancer binding protein  (C/EBP) have been proposed to be the major transcriptional inducers of SASP-related genes [71]. Interestingly, the expression of GFAP, the astrocytic marker of activation, is increasing radically after the age of 65 exactly as the probability to develop AD. Microglial cells also undergo age-dependent changes such as in- creased cell sensitivity, increased pro-inflammatory cytokine expres- sion and signalling and decreased anti-inflammatory cytokine expres- sion and signalling (review in [72]). Noteworthy is that even though there is an increase of MHCII immunoreactive microglia in the aging brain, there are no studies that demonstrate increased proliferation of microglia with age [6]. Furthermore, cell cultures of adult murine corti- cal microglia were shown to respond differently to A stimulation than postnatal derived microglia [73]. Whereas both oligomeric and fibrillar forms of A stimulated postnatal microglial secretion of TNF, adult microglia only secreted TNF in response to oligomeric A. Furthermore, the ability to phagocytise Aβ fibrils was markedly reduced in adult microglia compared to postnatal microglia. A better understanding on how inflammaging could contribute to the onset and/or progression of AD could have important implications for development of new therapeutic strategies.

Transcription factors involved in the inflammatory response In a review written by Glass and colleagues [63], the inflammatory pro- cesses involved in neurodegenerative diseases is described in terms of inducers, sensors, transducers and effectors. In AD, the primary in- ducer is obviously A, the sensors are the proposed A binding recep- tors: TLRs, RAGE and NLRs, the transducers are signalling pathways downstream of these receptors together with transcription factors and finally, the effectors are newly transcribed molecules such as cyto- kines and chemokines. The publications, which this thesis is based on, are focused on AD transducers, i.e. transcription factors, and more specifically NF-B and C/EBP.

12 Transcription and the role of transcription factors ‘The human genome has been called ‘the blueprint for life’. This mas- ter plan is realized through the process of gene expression’. Or- phanides and Reinberg [74]

It is remarkable that all cells in our body holds the same set of genes but still can be highly specialized in their function and in response to different stimuli. It is the tight regulation of gene expression that ena- bles this. Important players in the transcriptional machinery, i.e. trans- forming DNA to RNA, are TFs. It has been predicted that at least 5% of our genes codes for TFs [75], which highlights the importance of this protein family. There are two types of TFs, general (or basal) TFs and TFs that acts as transcriptional activators or repressors. The general TFs are highly conserved and represent the core of transcription by forming the pre-initiation complex. The complex consists of at least six different general TFs: TFIID, TFIIA, TFIIB, TFIIE, TFIIF and TFIIH and it binds to the TATA box (or initiator element) upstream of the transcription start site [75]. The primary functions of the general transcription factors are to recruit RNA polymerase II and to promote access to the tightly packed DNA. However, activation of general transcription factors exclusively results in very low levels of transcrip- tion. Thereby, in order to control the rate of transcription the other type of TFs is most often activated as well. They bind in a DNA se- quence specific manner to promoter or enhancer regions of the gene and interact with the pre-initiation complex. Depending on their DNA binding domain, they are classified into different families, such as zinc fingers, basic leucine zipper (bZIP) and helix-loop-helix proteins. The promoter region of a single gene usually contains several binding sites for different TFs. Thereby each gene can respond to multiple signal- ling pathways and in addition, different interactions among the acti- vated TFs can alter the outcome [74]. Moreover, TFs are also stabiliz- ing the pre-initiating complex resulting in more efficient and stable DNA binding [76]. There are also co-regulators recruited to the pro- moter by sequence-specific DNA binding TFs. These enzymes key assignment is to modulate chromatin structure making DNA more or less accessible. This can be achieved through ATP-dependent nucleo- some remodelling complexes or catalysing post-translational modifi- cations of histones. Both activation and inactivation of TFs are fast in order to screen the environment and quickly respond to different stimulus. In general, promoter-bound activated TFs have a rapid turnover involving the ubiquitin protease system [74].

13 Nuclear factor-B As a key player in the innate immunity and due to the up-regulated levels of cytokines, chemokines and acute phase proteins, it is not sur- prising that up-regulated levels of the TF NF-B have been found in AD brains [77-79]. Furthermore, several in vitro studies have showed increased NF-B activity in response to A [80-82]. Within the CNS, NF-B is expressed both in microglia and astrocytes and among the target genes are many pro-inflammatory molecules such as TNF, IL- 1, IL-6 and COX-2 [83]. NF-B was discovered in mature B-cells in 1986 by Sen and Balti- more [84]. The protein was named after the observation that it bound to a kappa enhancer element of the immunoglobulin kappa light chain promoter. The NF-B family consists of five members: p65 (RelA), p50;p105 (NF-B1), p52;p100 (NF-B2), c-Rel and RelB. The p50 and p52 proteins are generated by proteolytic cleavage by the precur- sor proteins p105 and p100, respectively. All family members share a structurally conserved N-terminal consisting of a dimerization and a DNA binding domain as well as a nuclear localization signal (NLS). The C-terminal of p65, c-Rel and RelB contains a transactivation do- main necessary for enhanced transcriptional activity. In order to be active they have to homo- or heterodimerize within the family or with other compatible TFs. Homodimers of p50 and p52 and p50:p52 com- plex works as transcriptional repressors since they do not contain a transactivation domain. However, p50:p65 seems to be the most commonly occurring activating complex [85]. In glial cells, the constitutive NF-B activity is very low. The NF- B dimers are kept in an inactive state in the cytoplasm via interaction with regulatory proteins called inhibitors of B (IB). The IB that regulates transient activation, as in glial cells, is IB that can be rap- idly degraded and rapidly resynthesized. However, the inactive NF- B-IB is not really kept in the cytoplasm all time but shuttle in and out of the nucleus. When IB binds to the NF-B dimer, only one of the NLS is hidden and since there is a nuclear export signal (NES) in the C-terminus of IB the complex is constantly moving between the nucleus and cytoplasm. However, the NES is stronger than the NLS hence the view that the inactive complex is located in the cytoplasm. Upon activation, by pathogens or pro-inflammatory cytokines for in- stance, IB becomes phosphorylated by the IB kinase complex (IKK). Once phosphorylated, IB becomes ubiquinated and targeted for degradation by the 26S proteasome releasing NF-B that can enter the nucleus and activate transcription. When bound to DNA there are several factors that influence the transcriptional efficiency of NF-B: phosphorylation, acetylation and co-activators, like CREB-binding

14 protein and p300 [85]. As mentioned earlier, IB is rapidly resynthe- sized after degradation since it has an NF-B site in the promoter re- gion. This is part of an inhibitory feedback mechanism, where IB via an intrinsic NLS enter the nucleus and detaches NF-B from the enhancer region and transport it back via the NES to the cytoplasm in an inactive form [86]. Interestingly, in an earlier study from our labora- tory it was shown that primary astro-microglial cultures treated with IL-1 and A induced a persistent NF-B activation [80]. Thereby it was suggested that glial cells exposed to A in an inflammatory mi- lieu possess a dysfunctional feedback mechanism.

NF-B in neurons In contrast to glial cells, there is a high degree of constitutive active NF-B in neurons during normal conditions. The proteins are present in all parts of the cell: in soma, axon, dendrites and synaptic terminals. The most commonly occurring complex is still p50:p65 (and IB), but in neurons it is associated with positive effects on neuronal sur- vival and neurite plasticity. Signals that have been shown to activate neuronal NF-B are many, e.g.: TNF, glutamate and nerve growth factor. However, the target genes are less well characterized (reviewed in [87]). The importance of NF-B in cognitive function [88], growth factor signalling [89] and acute CNS trauma [90] is well established and its role in chronic neurodegenerative disorders is an area of intense research. In AD brains, increased levels of neuronal NF-B activity has been reported [77]. The up-regulation is believed to be a neuroprotective defence response to protect cells against A induced cell death, in similar ways as in other CNS traumas.

CCAAT/enhancer binding protein Lately, another TF involved in the inflammatory response has gained attention due to its altered expression in AD brains, namely C/EBP. The C/EBP family (originally discovered in the lab of S. McKnight [91]) has six members, -. All C/EBPs share a highly conserved C- terminal (90%) consisting of a basic-leucine zipper (bZIP) dimeriza- tion domain and a basic DNA binding domain (which also includes an NLS [92]). Due to the high homology in the bZIP domain, all members within the family are capable to dimerize with each other, a necessary event in order to activate transcription. Furthermore, all C/EBP com- plex combinations have been shown to interact with the same DNA recognition sequence in vitro. On the other hand, the N-terminal of different C/EBPs, containing transactivation domains and negative regulatory regions, are very dissimilar (20% sequence identity) (re-

15 viewed in [93]). C/EBP,  and  have been shown to be expressed within the CNS [94]. C/EBP exist as two different isoforms, p42 and p30, and C/EBP exist in at least three different isoforms, full-length (38 kDa), LAP (35 kDa) and LIP (20 kDa). The shorter protein vari- ants created by a „leaky ribosome scanning‟ mechanism or regulated processing [95, 96], show less transactivation properties and LIP even works as a transcriptional repressor. C/EBP,  and  are all expressed in different amounts in glial cells [97-99] and C/EBP recognition sequences have been found in the pro- moters of several pro-inflammatory cytokines including IL-1, IL-6 and TNF [93]. However, suggestions of how the inflammatory re- sponse is regulated in brain most often come from studies of acute phase response in other tissues (like liver) or other cell types (like macrophages). During normal conditions, C/EBP complexes bound to enhancer regions are suggested to consist of C/EBP homodimers or C/EBP/ heterodimers [100]. However, when cells are challenged with inflammatory stimuli, C/EBP is rapidly down-regulated and C/EBP and  are up-regulated. Since the induction of C/EBP and  mRNA can be observed relatively late after the inflammatory stimuli, C/EBP has been suggested to be responsible for the maintenance of the inflammatory response and not the induction of it. In a model pro- posed by Valeria Poli, pre-existing inactive factors like Stat3 (signal transducer and activators of transcription 3) and NF-B would be transiently activated first and induce transcription of pro-inflammatory molecules as well as C/EBP and , which would then maintain the activated status of the cells [100]. Pre-existing C/EBP would also play a small part in the „early‟ response by forming C/EBP homodimers or C/EBP:NF-B heterodimers. The relevance of this model for a neuroinflammatory response as well, have been supported by in vitro studies in glial cells showing decreased levels of C/EBP [97] and in- creased levels of C/EBP and  [99, 101] in response to inflammatory stimuli. The importance of a tight control of the transcriptional activity of C/EBPs is shown by the numerous ways/levels of regulation: on ex- pression level, on translational level; depending on which isoform that is favourable [102], by posttranslational modifications; phosphorylation, acetylation, sumotylation [103-105] and protein-protein interactions; with co-repressors/activators or dimer partner [106, 107]. However, as for NF-B, C/EBP,  and  activation in neurons seem to be correlated with beneficial effects, important for normal brain function. They have been implicated in processes like synthesis of trophic factors, response to trophic factors [108] as well as learning and memory [109-111].

16 C/EBP in AD In comparison to NF-B, there are not many studies that have investi- gated the role of C/EBP in AD. However, there are many parameters that indicate an important role of C/EBP in disease progression. First, many inflammatory molecules showed to be increased in AD brains have C/EBP recognition sequences in their promoters. Second, C/EBP itself is up-regulated by many cytokines [112], thereby a vicious cycle is easily created. Thirdly, C/EBP is able to dimerize with NF-B and thereby affect the cytokine production. Fourthly and most important, increased levels of C/EBP and  have actually been detected within AD brains. In a nice study by Li and colleagues it was shown that C/EBP levels was significantly increased in neocortex and limbic cortex in AD brains [113]. Interestingly, immunohistochemistry of the tissues showed that C/EBP was exclusively located within astrocytes immediately surrounding the A plaque core. This study was followed up by a study by Ko and colleagues where they studied the contribu- tion of up-regulated astrocytic C/EBP to AD progression [114]. Their results from in vitro experiments showed that increased C/EBP ex- pression within astrocytes resulted in impaired phagocytosis of dead neurons by macrophages. The responsible factor, released from astro- cytes was suggested to be pentraxin-3. Moreover, C/EBP has been proposed to be the missing link be- tween the hyper-inflammatory state observed in elderly following and the increased risk of developing AD. In a study analysing the C/EBP response after controlled cortical impact injury in adult and aged mice, aged mice showed a 12-fold increased induction in C/EBP mRNA compared to younger mice [115]. Further- more, C/EBP protein levels in aged mice following brain injury were higher at all time points studied. Finally, genome-wide expression profiles of AD hippocampus showed an approximately 4-fold increase in mRNA levels of C/EBP compared to age matched controls [78, 116].

17

Figure 3. Model of NF-B and C/EBP activation. Pre-existing inactive factors like Stat3 and NF-κB is transiently activated first in the ‘early’ inflammatory response. Upon activation, they induce transcription of inflammatory molecules and TFs in- volved in the ‘late’ inflammatory response, like C/EBPβ and δ, which then maintain the activated status of the cells. Pre-existing C/EBPβ also play a small part in the ‘early’ response, by initiating transcription after it has become phosphorylated. This model was originally suggested for induction of acute phase genes in the liver, but its relevance in the CNS has been supported by several in vitro and in vivo studies. The figure is a modified version from [100].

18 Cytokines - effector molecules in AD Cytokines are small, soluble signalling molecules strongly associated with the immune response even though they have been shown to have important functions also during normal conditions. They are often divided into anti-inflammatory cytokines (like IL-4 and IL-10) and pro-inflammatory cytokines (like IL-1, TNF and IL-6). However, most cytokines are multifunctional and it is the current circumstances that determine the outcome. Furthermore, they have been shown to have redundant functions since one cytokine often can be knocked down without pronounced physiological effects. In AD, virtually all pro-inflammatory cytokines studied have been shown to be up-regulated compared to non-demented individuals [6]. Here follows a brief presentation of two of them, IL-1 and IL-6. (IL- 1 was used in paper I-III to mimic an on-going inflammation and IL- 6 was used as a read-out in paper I).

Interleukin-1 The cytokine IL-1 is a well-known contributor to the pro- inflammatory response both in the periphery and in the CNS. Howev- er, IL-1 is also expressed in low levels within several different cell types during normal conditions and has been proposed to be involved in long-term potentiation (LTP) [117, 118]. Nevertheless, in AD brains, the increased levels of IL-1 are strongly associated with neurotoxic pro-inflammatory properties. IL-1, IL-1 and the IL-1 receptor antagonist (IL-1ra) are three members of the IL-1 family. They are encoded by three separate genes, which are located in a cluster on chromosome 2. IL-1 and IL- 1 are quite similar in terms of tertiary structure, regulatory factors and receptor affinity. However, in contrast to IL-1, the precursor protein of IL-1 has to be cleaved by IL-1 converting enzyme (caspase-1) in order to be active. Furthermore, IL-1 is secreted, working as an intercellular messenger whereas IL-1 is mostly found intracellularly. The third family member, IL-1ra, does not activate signalling pathways upon receptor binding and is thereby working as an antagonist [119]. There are two IL-1 receptors, IL-1RI and IL-1RII, which also exists in soluble forms, sIL-1RI and sIL-1RII. However, intracellular signalling is only occurring after IL-1 binding to IL-1RI since the other receptor lack the capacity to interact with the necessary adaptor proteins. The IL-1 signalling is further illustrated in figure 1 on page 7. In AD brains, microglial cells located in close proximity to A de- posits seem to be the main responsible cell type for the increased ex-

19 pression of IL-1 [120]. Interestingly, up-regulated levels of IL-1 have been linked to many of the pathological responses associated with AD: excessive glial activation and expression of inflammatory mole- cules, invasion of immune cells into the CNS, up-regulation of APP synthesis and processing and tau phosphorylation [121]. In support of an important role of IL-1 in AD are studies indicating that polymor- phisms in the IL-1A and IL-1B genes increase the risk of developing late-onset AD [122-124].

Interleukin-6 IL-6 was first discovered in 1986 by Hirano and colleagues as a factor released from lymphocytes [125]. Within the CNS, IL-6 has been de- tected in neurons, microglia and astrocytes. Among the most im- portant TFs regulating IL-6 expression are NF-B and C/EBP [126]. Astrocytes are both the main responders and producers of IL-6, which together with IL-1 and TNF is a major regulator of neuroin- flammation. IL-6 has been shown to influence proliferation, release of cytotoxic molecules and chemotaxis of astrocytes. However, during certain circumstances IL-6 is working as a neurotrophic factor with important functions in neuronal survival, differentiation and regenera- tion (reviewed in [127]). IL-6 belongs to a cytokine family that shares a common signal transducer, glycoprotein 130. The signalling is quite complex and sev- eral different parameters, such as interactions with other cytokines, concentration, brain region, stage of neuroinflammation and cellular source have been shown to determine the outcome. Interestingly, in addition to the membrane-bound receptor, IL-6 also has an agonistic soluble receptor enabling cells lacking the IL-6 receptor to respond to this cytokine [127]. The importance of IL-6 in AD is indicated by increased expression early on in disease progression, both in AD patients and in AD animal models [128-130]. Due to the dual roles of IL-6, it has been hypothesized that the initial IL-6 expression AD brains is a neuroprotective re- sponse to save the tissue, but when A is not removed properly it will do more harm than good. In addition of accelerating the inflammatory response, IL-6 has been reported to influence APP expression and processing [131, 132] as well as tau hyperphophorylation [133]. At least two different polymorphisms have been found in the IL-6 gene pro- moter, -174G/C and -572C/G, but the association to AD is debated [134-136]. Furthermore, the consequence of increased IL-6 expression in normal brain aging is not fully clear (reviewed in [137]).

20 Non-steroidal anti-inflammatory drugs and AD NSAIDs (or „aspirin-like drugs‟) are a group of pharmaceuticals that work by inhibiting the enzyme COX, which converts arachidonic acid to prostaglandins. There are two COX isoforms, the constitutively expressed COX-1 and the inflammatory inducible COX-2. In micro- glial cells, E2 has been proposed to be the main COX-2 enzymatic product. Upon prostaglandin E2 binding to microglial re- ceptors, the phagocytic capabilities are decreased and neurotoxic ac- tivities increased [138]. The major breakthrough for using NSAIDs in the treatment of AD came in the year of 1990, when McGeer and colleagues published a study in the Lancet showing that patients with rheumatoid arthritis developed AD to a much lesser extent than average [139]. Since patients with rheumatoid arthritis generally receive anti-inflammatory treat- ments well before the „risk age‟ of AD, NSAIDs was suggested as a prevention treatment. However, despite beneficial effects of NSAIDs in AD from several epidemiological studies [140-143], the clinical trials have been very disappointing. For sure the last word has not been said on this subject. In 2001, the Alzheimer‟s Disease Anti-Inflammatory Prevention Trial (ADAPT) started. This prevention trial included 2520 non-demented elderly sub- jects at high risk of AD (nicely discussed in [144]). The participants were given the non-selective COX inhibitor naproxen, the selective COX-2 inhibitor celecoxib (available by prescription) or placebo, and were supposed to be followed for ten years of treatment. However, due to safety concerns the study was abruptly ended in 2004 with very discouraging results. Remarkably, after average treatment of 24 months there were no beneficial effects of NSAID treatment, the naproxen group even seemed to have a higher occurrence of AD. Amazingly, five years later, in 2009, the study was reactivated again. 2.5 years after cessation, the incidence of AD among the participants receiving naproxen was 1/3 compared to those given placebo. There- by, naproxen was suggested to increase disease progression for per- sons in the late pre-clinical phase but be protective for normal healthy elderly individuals. It should be noted that the positive effects observed after NSAID treatment are suggested to involve additional mechanisms beside the COX inhibitory activity, such as: Altering the aggregation propensity of A by induced expression of amyloid binding proteins [145], affect- ing APP processing by modifying -secretase [146] and activating the peroxisome proliferator activated receptor  [147], a transcriptional re- pressor of several inflammatory genes.

21 Neurodegeneration in AD As mentioned earlier, one of the pathological hallmarks of AD is the gradual loss of synapses and neurons in limbic and cortical areas. It has been suggested that neurodegeneration is initiated 20-40 years before the appearance of clinical symptoms [148]. However, neuro- degeneration associated with AD has been very hard to study since most tg AD mouse models do not exhibit NFTs or neuronal loss. Apoptosis has been proposed to be involved in the death of neurons based on studies in AD brain tissues, showing increased levels of cleaved caspases and accumulation of caspase cleaved substrates in affected brain areas [149]. Further evidences for a role of caspases in AD are discussed in a later section (starting on page 23). However, direct detection of apoptosis, characterized by membrane blebbing, cell shrinkage, nuclear fragmentation and chromatin condensation, is problematic since apoptotic cells are cleared relatively fast by micro- glia.

Roles of caspases in apoptotic cell death Caspases, cysteine dependent aspartate-specific proteases, are highly conserved key players in the apoptotic machinery. There are 13 mammalian caspases commonly divided into three groups based on their action: inflammatory caspases (1, 4, 5, 11, 12 and 14), apoptosis initiator caspases (2, 8, 9 and 10) and apoptosis effector/executor caspases (3, 6 and 7). The initiator caspases can be further divided into actors in the intrinsic- or extrinsic apoptotic pathway, best character- ized for caspase-9 and -8, respectively [150]. The intrinsic apoptotic pathway is initiated by cytochrome c release from mitochondria [151]. In the cytosol cytochrome c binds to Apaf-1 (apoptotic protease-activating factor 1) and triggers its oligomeriza- tion, creating a signalling platform, the apoptosome. Thereafter, caspase-9 is recruited to the complex by homotypic interaction via a CARD (caspase-recruitment domain). Caspase-9 (as well as caspase- 8) exists as an inactive monomer in unstimulated cells and gets acti- vated by dimerization and not necessarily cleavage. Due to the long intersubunit linker, the protein is very flexible and when bound to the apoptosome there is a conformational change resulting in exposure of the active site. Thereafter, caspase-9 can activate downstream effector caspases, like capase-3 and -7, by cleavage of their intersubunit linker. Upon cleavage of the effector pro-caspase dimers, there is a conforma- tional change resulting in exposure of the catalytic residues and the substrate-binding pocket [151]. In contrast to activation of initiator caspases, activation of effector caspases are irreversible. However, the

22 effector caspases cleaves the target proteins in a specific limited way with the purpose to modify it and start a signalling cascade. The pri- mary purpose is not to degrade it even if it is a very common outcome. The extrinsic apoptotic pathway is initiated by ligand binding (e.g. Fas ligand or TNF-related apoptosis-inducing ligand) to a transmem- brane receptor leading to formation of the death-inducing signalling complex (DISC) containing caspase-8 (or -10). The activation is be- lieved to occur by an induced-proximity mechanism, similar to caspa- se-9 activation. The signal is then feed forward by caspase-8 cleavage of the intersubunit of effector caspases. Other key participants in the apoptotic signalling cascade are IAPs (inhibitor of apoptosis proteins), a protein family consisting of eight members in humans [152]. They achieve their inhibitory effect by bind- ing to different caspases and thereby inactivate the enzymatic activity, change the cellular localization or block the substrate-binding pocket. Furthermore, there are also several endogenous inhibitors of IAPs (e.g. „Smac/Diablo‟) that are released upon permeabilization of mito- chondria. These proteins potentiate the signalling cascade by inhibit- ing the IAPs.

The involvement of caspases in AD Several studies have indicated an important role of caspases in AD. Increased mRNA levels of caspase-1 and -7 have been detected in post-mortem brains of patients at high risk of developing AD (charac- terized neuropathologically and by Clinical Dementia Rating). Pa- tients suffering from severe dementia showed increased mRNA levels of caspase-1, -2, -3, -5, -6, -7, -8 and -9, compared to cognitive normal age-matched controls [153]. Active caspase-6 has also been detected by immunostaining, using a neoepitope to the p20 subunit (p20Csp6), in frontal and temporal cortex in AD brains [154]. Surpris- ingly, active caspase-6 was not detected within the nucleus but se- questered into tangles and neurites suggesting a non-apoptotic role during these circumstances. A follow-up study by the same group, investigating caspase-6 in young and aged noncognitively impaired (NCI), aged MCI and aged mild, moderate, severe and very severe AD individuals showed that p20Csp6 strongly immunostained the hippo- campi at MCI and all stages of AD [155]. Additionally, some of the aged NCI also showed p20Csp6 immunostaining, which was associat- ed with age-dependent cognitive decline. In support of a neurite re- modelling role of caspase-6 in AD, a proteomic study showed cyto- skeleton or cytoskeleton-associated proteins to be the main neuronal target for caspase-6 in AD [156].

23 Aβ has also been shown to induce caspase activation in vitro [157]. A hot area of research has been the involvement of caspase activation in Aβ inhibition of LTP. In a recently published study, caspase-3 and/or caspase-7 was suggested to be involved via cleavage of Akt1 and acti- vation of glycogen synthase kinase-3β [158]. Other cellular substrates of caspases are APP [149], tau [159] and presenilins (constituents of the γ- secretase complex) [160], which upon cleavage all accelerate disease progression in different ways. However, it should be noted that several studies have also stressed the importance of microglial cells in Aβ induced neurotoxicity [161, 162] and inhibition of LTP [163, 164](this is further discussed on page X). Furthermore, non-apoptotic functions of caspases have been shown in glial cells. Burguillos et al. recently showed that caspase-3/7 had an important role in Lipopolysaccharide (LPS)-induced microglia activation and neurotoxicity [165].

Methods to determine caspase activation and cell death Since caspase activation no longer is synonymous with cell death there is a great need of improvements of commercial products for de- tection of caspase activity and neurodegeneration. Methods like the Annexin-V staining of phosphatidylserine located to the external cell membrane [166] and terminal deoxynucleotidyl transferase dUTP nick end labelling (TUNEL) for detection of DNA fragmentation [167] only shows a very late state of apoptosis, when there is no turning back. For measurements of caspase activity, western blot and immunocyto- chemistry using directed against cleaved caspases is fre- quently used. However, there are some drawbacks with these methods. Firstly, despite cleavage one cannot be sure that the caspases are really activated since they can be bound to „invisible‟ IAPs. Secondly, initia- tor caspases do not necessarily need to be cleaved in order to be ac- tive. Furthermore, there are active site affinity ligands (such as bio- tionylated-VAD-fmk) for detection of specific caspase activation at a given time point. These ligands bind irreversible to active caspases and inhibit them. Detection is carried out by streptavidin-agarose pull- down followed by western blot using antibodies directed at the caspa- se of interest. The disadvantage with this method, even though it has been proven to give accurate measurements of casase activity, is that results are not obtained in vivo. Finally, there is a pCaspase3-sensor developed by Clontech for detection of caspase-3 activity in live cells. The construct is designed as follows: a dominant NES, a caspase-3 cleavage site, yellow fluorescent protein and an NLS. The method is based on a change in the cellular localization of the fluorescent fusion protein, from the cytoplasm to the nucleus, upon activation and re- moval of the NES. However, this method does not provide a spatio-

24 temporal understanding of the apoptotic process, i.e. where in the cell the signal is initiated.

25 Aims of the thesis

The overall aim of this thesis was to contribute to an increased under- standing of inflammatory processes involved in AD. More specifical- ly, the aims of each paper are presented below:

Paper I The aim of paper I was to bring together two earlier finding from our laboratory, namely investigate if a newly developed NF-κB decoy strategy could be applied in a relevant model system (rat primary as- tro-microglial cultures) and inhibit Aβ induced up-regulation of NF- κB and subsequent expression of the cytokine IL-6.

Paper II The aim of paper II was to study the protein levels and DNA binding activity of C/EBPβ and δ in primary astro-microglial cell cultures ex- posed to Aβ. We believed that there was a missing link between up- regulated C/EBP mRNA levels detected within AD brains and knowledge about protein translation and activity.

Paper III Paper III is a follow-up study to paper II, where we wanted to further investigate the impact of Aβ conformation on C/EBP protein expres- sion. To get a more complete picture C/EBPα was also investigated and in vivo experiments were included.

Paper IV The aim of paper IV was to develop a method for detection of caspase activation in a spatio-temporal manner in live cells. We believed this was necessary in order to receive further information of caspase ac- tivity in cell signalling and neurodegeneration associated with AD.

26 Methodological considerations

In vitro and in vivo model systems

Primary astro-microglial cultures (paper I-III) To mimic AD related neuroinflammation in vitro we chose to work with rat primary astro-microglia cultures treated with IL-1β and Aβ. Cells were obtained from male Sprague Dawly rat pups less than 24 h of age, a time when glial cells proliferate rapidly. After decapitation, a piece of frontal cortex was picked and fibroblast containing meningeal membrane was removed. Homogenization of brain tissue was ob- tained by pressing the tissue through a fine-meshed nylon net in cul- ture medium. In order to get astro-microglial cultures consisting of 90- 95% astrocytes and 5-10% microglial (a population resembling the one in human brain), cells were seeded at a low density, a culture con- dition preferred by astrocytes. Between times, cultures were analysed by staining for GFAP (an astrocytic marker) and ED1 (a marker of rat microglia). Advantages of this in vitro model are that important inter- actions between astrocytes and microglia are maintained, and also, biological properties of microglial cells in vivo are sustained.

Arc-Swe transgenic mice (paper III) To study how Aβ fibrils affect C/EBPδ expression in vivo, transgenic mice carrying both the Artic and the Swedish APP mutations (tg- ArcSwe mice) were chosen. The artic mutation is a change of amino acid in position 693 in APP where a glutamic acid is replaced by a glycine (E693G). This mutation is located within the Aβ sequence and results in early intraneuronal Aβ aggregation, enhanced Aβ protofibril formation and senile plaque deposits at young ages [10, 168, 169]. The Swedish mutation is a double mutation in APP, where a lysine and a methionine, in position 670 and 671, are replaced by an asparagine and a leucine (KM670/671NL). These mutations are located at the β- secretase cleavage site and results in increased Aβ production. The mice studied in paper III were 15.5-17 months-old when sacrificed. At this time, these mice possess a plaque burden of approximately 4 % as

27 shown in previous studies by Aβ immunostaining of tissue sections followed by quantitative image analysis [169, 170]. Also, 4 months old tg-mice, with an Aβ burden below detection limit (<20 pg Aβ/mg tis- sue), were used as control for changes related to age and Aβ deposi- tion. Brains were dissected into Aβ-containing forebrain (containing cortex and hippocampus) and Aβ-sparse hindbrain (cerebellum and ).

Cell lines (paper IV) In paper IV three different cell lines of human origin were used. First HeLa cells and U2OS cells were used for characterization of the FRET sensors. HeLa cells were the first human cancer continuous cell line to be developed and originate from a very aggressive cervix can- cer tumour. Thus, HeLa cells proliferate very rapidly. The U2OS cell line was established in 1964 by Ponten and Saksela. The cells are de- rived from a differentiated sarcoma of the tibia from a fifteen years old girl suffering from osteosarcoma. U2OS cells can be handled in the same way as HeLa cells. Further on, to study caspase activity dur- ing neurodegeneration, the neuroblastoma SH-SY5Y cell line was used. SH-SY5Y cells are a third cloned subline of SK-N-SH cells, originally derived from a bone marrow biopsy of a 4 year old girl [171, 172]. SH-SY5Y cells are quite small neuroblast-like cells with usually short processes sharing many biochemical and functional properties with neuronal cells in vivo [173]. SH-SY5Y cells can be further differ- entiated into a more cholinergic phenotype using the vitamin A me- tabolite all-trans retinoic acid (RA) [174]. RA treatment results in a stop in cell proliferation and extended neurites are received. The ad- vantages of using cell lines are that they are well defined, easy to grow and the genetic and phenotypic change is very small when cultured during right conditions. However, when studying neurodegeneration one has to be very careful to draw conclusions due to their cancero- genic and immortalized state.

Cell treatments

IL-1β and LPS (paper I-III) In order to induce an inflammatory response in the astro-microglial cell cultures, IL-1β or LPS was used. IL-1β is a cytokine (see intro- duction) known to be induced in AD brains. LPS is also known to induce a strong inflammatory response, and can be regarded as a posi-

28 tive control of inflammation, not directly related to AD. LPS is an endotoxin produced in Gram-negative bacteria, where it is a part of the outer cell membrane. The structure consists of three different do- mains; a hydrophobic domain penetrating the membrane often referred to as lipid A, a core saccharide and distal O-antigen, which is polysac- charide as well. The effect on mammalian cells is achieved by binding to TLRs.

Amyloid-β (paper I-IV) It has been debated for a long time which conformation of Aβ that is most toxic. For neurons, most studies are now pointing towards the oligomeric forms, where low concentrations are known to cause syn- aptic failure and inhibition of hippocampal LTP [175, 176]. Therefore, we chose to challenge the cells with oligomeric Aβ in paper IV. Concen- trations ranging from 5 to 20 μM, may seem high, however, the aim of this study was to study neurodegenerative processes during apoptosis. As we pointed out in the end of the discussion in paper IV, in the fu- ture, it would be very interesting to use the FRET sensors with con- centrations of Aβ not leading to cell death. As compared to neurons, not nearly that many studies analysing the effect of different confor- mations of Aβ are performed on glial cells. This can be reflected in paper I-III, where different conformations of Aβ are analysed, and also, a short model peptide of full length Aβ, Aβ(25-35), is used (pa- per I and II). Aβ(25-35) is mainly working as a model peptide for fi- brillar forms of full length Aβ due to its high aggregation capacity [177]. However, Aβ(25-35) has been suggested to be produced in vivo in brains of AD patients [178]. Similar to full length Aβ, Aβ(25-35) has been shown to induce depression of LTP [179], axonal atrophy and syn- aptic loss [180] in neurons. In glial cells, Aβ(25-35) was shown to fur- ther increase cytokine/LPS induced iNOS expression by increasing the transcriptional binding activity of C/EBP, NFκB, AP-1 and CREB [181]. The choice of Aβ(25-35) concentration, 100 μM, was due to pre- vious work showing a good response in glial cells at that concentra- tion[80, 182]. However, working as a model peptide for full length Aβ one can query why the concentration needs to be so much higher for the truncated form to receive the same effects. Full length Aβ, used in paper II and III, was prepared according to previously published pro- tocols [183, 184] to mainly consist of fibrillar or oligomeric Aβ. The dif- ferent Aβ preparations were analysed by western blot and thioflavin T assay. Thioflavin T is a commonly used dye that binds to β-sheet structures in amyloid plaques. Upon binding, there is a change in the wavelength of excitation/emission, from 350/438 to 450/482. The concentration of Aβ(1-42) was 10 μM in both papers, based on previ-

29 ous reports (e. g. [185]), shown to be the optimal concentration of oli- gomer or fibrillar enriched Aβ to obtain an inflammatory response in this type of cell culture. In all papers, the reversed peptides, Aβ(35-25) and Aβ(42-1), was used as a control peptides. However, these peptides do not form aggregates to the same extent as natural Aβ. In the future, it would be very interesting to use peptides, not related to Aβ, that aggregate into oligomers or fibrils and study the outcome since aggre- gates, independent of sequence, have been proposed to share a com- mon mechanism of toxicity [186, 187].

Cell penetrating peptides (CPP) (paper I) In order to facilitate internalization of the NF-κB decoy into glial cells, which are very hard to transfect (since they easily get activated), the decoy was hybridized to PNA via a protruding end, which in turn was coupled to a CPP via a disulphide bridge. CPPs are short cationic or amphipathic peptides that have the power of entering cells while transporting a cargo. There are both naturally occurring CPPs and syn- thetic CPPs, as Transportan 10 (TP10) [188] that we used in paper I. TP10 is amphipathic and consists of 21 amino acids of which five are positively charged. TP10 is a short analogue of TP [189], a chimeric peptide consisting of amino acids from the neuropeptide galanin in the N-terminus and amino acids from the wasp venom peptide toxin, mas- toparan, in the C-terminus.

KillerRed (paper IV) The photosensitizer KillerRed plasmid [190] used to induce apoptosis locally in the cell was kindly provided by Konstantin Lukyanov. To- day, it can be purchased from Evrogen. KillerRed is a dimeric red fluorescent protein, developed from the GFP-homolog anm2CP, con- taining two tandem copies of mitochondrial localization signals to make the light-induced cell killing more effective. Upon irradiation with green light (540-580 nm), reactive oxygen species are generated and cell death is induced via the intrinsic apoptotic pathway. An ad- vantage of KillerRed, compared to other known photosensitizers (e.g. FlAsH [191] and ReAsH [192]), is that it does not require any exogenous- ly added compound in order to work.

Staurosporine (paper IV) In paper IV, staurosporine was used to induce apoptosis in U2OS and differentiated SH-SY5Y cells. Staurosporine was identified in 1977 as a naturally occurring compound, with antimicrobial properties, pro-

30 duced by Streptomyces staurosporeus [193]. It is known to induce apop- tosis by binding to the ATP binding site of a large set of different ki- nases, including protein kinase C, and inhibit their activity [194].

Studies of protein-DNA interactions

Electrophoretic mobility shift assay (paper I-III) To study NF-κB and C/EBP binding to promoter regions, electropho- retic mobility shift assay (EMSA) was used. Garner and Revazin were among the first to describe the method in 1981 [195] and now thirty years later it is a commonly used method to study protein-DNA inter- actions. The method is very sensitive, compatible to a wide range of nucleic acid sizes and structures and crude cell lysates can be used. As the name of the method indicates, it is based on the fact that protein complexes bound to DNA cause a shift in mobility in comparison to free DNA, when run on a polyacrylamide gel. In paper I-III, radio- labeled (32P) oligonucleotides corresponding to recognition sequences of NF-κB or C/EBP were mixed with nuclear extracts from primary astro-microglial cultures exposed to different stimuli. Samples were separated under native conditions using gel electrophoresis and bands were visualized using radiography. Since all members of the NF-κB and C/EBP family acts on the same promoter region, one has to do supershift experiments in order to identify which specific protein that is bound during induced conditions. The supershift corresponds to an additional shift on the gel caused by binding of an antibody to one of the expected proteins in the protein-DNA complex. However, EMSA only gives you information if a protein is bound to the DNA sequence or not, and does not provide information if transcription is stimulated or inhibited.

Streptavidin-agarose pull-down (paper III) In addition to EMSA, a streptavidin-agarose pull-down assay was used in paper III to study the interaction of C/EBPδ to the κB DNA element. The streptavidin-agarose pull-down assay is less sensitive compared to EMSA due to several washing steps. However, there is a clear advantage not having to use radiolabeled probes. Instead of la- belling the probe with 32P, biotinylated oligonucleotides are commer- cially available. In our study, the recognition sequence of NF-κB was labelled with biotin and mixed with nuclear extracts from primary astro-microglial cultures and a streptavidin-agarose bead suspension.

31 Due to the high binding affinity of biotin to streptavidin, after two hours incubation, the protein-DNA complex can be pulled-down by centrifugation. After several washing steps to get rid of unspecific binding to the agarose beads and boiling to dissociate the complex from the beads, the sample can be loaded on a polyacrylamide gel and analysed by western blot. As a control of unspecific binding to the beads, identical κB oligonucleotides lacking biotin were used.

Reverse transcriptase-polymerase chain reaction (paper I) In paper I, an NF-κB decoy was constructed in order to selectively inhibit NF-κB activity and thereby decrease cytokine expression in primary astro-microglial cultures treated with IL-1β and Aβ(25-35). As a measurement of successful inhibition, the mRNA levels of a known NF-κB target, IL-6, was measured using reverse transcriptase- polymerase chain reaction (RT-PCR). The first step in RT-PCR, after mRNA isolation, is to transcribe the mRNA into cDNA using a RNA- dependent DNA polymerase (reverse transcriptase). In this study, we chose Molony murine leukaemia virus reverse transcriptase due to its low RNAse H activity. There are different types of primers that can be used in this first step; specific primers, random primers or oligo-dT primers. We chose to work with random primers since they have been shown to maximize the number of mRNA molecules from small sam- ples. Also, we chose 18S rRNA as an internal standard thereby ex- cluding oligo-dT primers that are only compatible with mRNA. In the second step when amplifying the genes of interest, specific primers for IL-6 and 18S were used. In order to keep the 18S amplification in the same linear range as the IL-6 amplification, the 18S primers were mixed with competimers. These are identical primers modified in the 3‟ end that cannot be extended by the Taq polymerase. Finally, the PCR products can be analysed by agarose gel electrophoresis and staining of DNA using ethidium bromide. However, there are some drawbacks using this method. It demands optimizations in primer de- sign, primer:competitor ratio and insurances that samples are analysed in the exponential amplification phase in the PCR reaction. Also, the method is viewed as a semi-quantitative method since the result may vary between samples due to RNA quality, enzyme efficiency and pipetting errors.

32 Western blot (paper II-IV) Western blot (WB) is a commonly used method when studying a spe- cific protein in a given sample. It was used in paper II and III to study different members of the C/EBP family and for detection of decompo- sition products of the FRET sensor in paper IV. Also, it was used to characterize Aβ preparations in paper III and IV. All experiments throughout this thesis were performed during denaturating conditions i. e. proteins were separated according to their molecular weight using sodium dodecyl sulphate polyacrylamide gel electrophoresis. After gel electrophoresis, proteins were transferred to polyvinylidene difluoride membranes to make them accessible to primary- and secondary anti- bodies, necessary for detection. The primary antibody is raised against the protein of interest whereas the secondary antibody is directed at the species specific heavy chain of the primary antibody. In order to quantitatively measure the protein levels, horseradish peroxidase was linked to the secondary antibody. Thereby enabling protein detection by enhanced chemiluminescence. Disadvantages with the method are time consuming optimizations of antibody concentrations, in some cases poor protein specificity of the primary antibody and differences between antibody batches.

Laser scanning confocal microscopy (paper I and IV) Laser scanning confocal microscopy (LSCM) has several advantages compared to conventional fluorescence microscopy, which has made it an important tool in cell biology. The two most important features of LSCM are: 1. Exclusion of background information. In front of the detector there is a „pinhole‟ that eliminates out-of-focus light, i.e. fluo- rescence emission from regions outside the focus plane. Thereby, LSCM can generate images with higher resolution at the cost of emitted light intensity. 2. The source of excitation. LSCM uses a powerful laser beam to illuminate the specimen at a finely focused spot with the opti- mized wavelength and excitation intensity. The image is re- ceived by scanning a defined area where after the information is sent to a detector. In paper I we used LSCM to study the internalization of the TP10- PNA-NF-κB decoy. As discussed in Results and discussion, there was a punctate pattern of the decoys, indicating entrapment inside endo- somes, even though outcome measurement of IL-6 expression showed an effect of the decoy. This may be explained by insensitivity of the

33 system, leaving weak signals in cytoplasm or nucleus undetectable. In paper IV we used LSCM mostly because of the quantitative acceptor bleaching experiments and for light induced cell killing by KillerRed. However, studies using the newly developed anchored FRET sensors do not require LSCM. Fluorescence resonance energy transfer (FRET) Our sensors for caspase activation are based on FRET, fluorescence resonance energy transfer (or Förster resonance energy transfer). FRET is commonly used when studying different kinds of protein interactions and is based on non-radiative transfer between two fluor- ophores. The requirements for FRET to work are that the FRET pair fluorophores need to be in close proximity, less than 10 nm apart and that the „donor‟ emission spectrum and the „acceptor‟ absorption spec- trum overlap. In our experimental setup, we detect loss of FRET. Simplified, we excite ECFP (enhanced cyan fluorescent protein), which transfers energy to EYFP (enhanced yellow fluorescent pro- tein), which thereby emits yellow light. When caspases are activated, the sequence between ECFP and EYFP is cleaved resulting in ECFP emitting cyan light since it is now diffusible in the cytoplasm and no longer capable to transfer energy to EYFP. The level of FRET was calculated ratio-metrically by dividing the emission intensity in the EYFP channel (515-575 nm) by the emission intensity in the ECFP channel (455-505 nm).

Figure 4. Schematic illustration of FRET. When the FRET pair fluorophores, con- sisting of the ‘acceptor’ EYFP (white) and the ‘donor’ ECFP’ (grey), are in close proximity, energy is transferred from ECFP to EYFP that emit yellow light upon excitation of ECFP. When the distance becomes bigger than 10 nm, e.g. after caspa- se cleavage of the anchored FRET sensors, energy can no longer be transferred and ECFP emits cyan light upon excitation.

34 Results and discussion

Paper I: Targeting cytokine expression in rat primary glial cells by cellular delivery of an NF-κB decoy. In paper I we wanted to connect two earlier interesting findings from our laboratory [196, 197], namely study if the TP10-PNA(9)-NF-κB de- coy could be internalized into primary astro-microglial cell cultures and inhibit the previously seen additive effect of Aβ(25-35) on IL-1β. A great challenge working with glial cells is the ability to modulate cell function without activating the cells. Thereby, in addition to study a new future therapeutic approach targeting NF-κB activation we also wanted to evaluate the usage of a CPP as a delivering agent in this model system. As reported earlier, co-stimulation of cells with Aβ(25-35) and IL- 1β resulted in increased NF-κB activation and IL-6 expression, as compared to either treatment alone. Here we show that 1 h pre- treatment of primary astro-microglial cells with the TP10-PNA-NF- κB decoy before Aβ(25-35) and IL-1β exposure resulted in 80% and 50% inhibition of the NF-κB binding activity and IL-6 mRNA expres- sion, respectively. Thus, the effect of the decoy complex were more pronounced in the primary astro-microglial cells compared to the IL- 1β treated rat Rinm5F cells tested earlier, where 1 h pre-treatment with the decoy complex resulted in a down-regulation of 63% and 20%, respectively. This more efficient inhibition of NF-κB binding activity observed in glial cells might have several different explana- tions. Firstly, it might indicate a more central role for NF-κB in cells challenged with both Aβ and IL-1β. Secondly, it might reflect a dif- ference in NF-κB dimer subset between the two cell cultures since it has been shown that depending on the NF-κB subunits, different DNA sequences are preferred [198]. The κB binding factor of the glial TF complexes might have a higher binding affinity towards the decoy. Thirdly, in the present study we used TP10 as delivering agent where- as the earlier study primarily used TP. Although both CPPs have been reported to use the same up-take mechanisms, caveolae- and clathrin mediated endocytosis [199, 200], cargos are known to influence CPPs differently.

35 By labelling the decoy and plasma membrane fluorescently we could study the uptake using LSCM. The uptake studies showed that the TP10-PNA-NF-κB decoy was successfully internalized within cells after 1 h at the low concentration of 0.5 μM. Moreover, cells that had taken up the decoy showed no signs of activation judged by the morphology. In contrast, cells treated with the decoy alone or decoy together with TP10 but without the PNA adaptor sequence showed undetectable intracellular levels. In our primary astro-microglial cell cultures, 90-95% of the cells are astrocytes and this cell type is shown in the figure representing the decoy uptake (Fig. 5 in paper I). However, we also observed that the decoy seemed to be even more efficiently taken up by the microglia (data not shown). A possible explanation for this could be that the decoy complex was phagocytised by the microglia. Thereby, the fluo- rescent material detected within these cells might represent degraded and no longer functional decoys. Further investigations are needed to clarify this matter. Also, the fluorescently labelled decoy showed a punctate pattern in- side the cells, which would indicate an entrapment inside vesicles most likely endosomes. However, since the functional consequence of this strategy was reduced NF-κB activity and down-regulated IL-6 expression, some endosomal escape of the decoy clearly must have occurred. In conclusion, we present a new decoy strategy for NF-κB inhibi- tion in glial cells. By coupling the decoy to a transport peptide inter- nalization of the decoy was clearly improved and NF-κB activation and IL-6 expression was significantly decreased. This discovery is of great value since the plasma membrane barrier earlier has required high decoy concentrations and long incubations times. Furthermore, it opens up the possibility to use CPPs as delivering agents in general in this in vitro model system.

Future perspectives There are several inflammatory diseases like , ather- osclerosis and rheumatoid arthritis that are associated with increased and prolonged activation of NF-B. Thereby, several therapeutic strategies have been developed in order to block NF-B activity. Some examples are IKK--dominant-negative gene therapy, NF-B decoy oligonucleotides, antioxidants, curcumin and proteasome in- hibitors [83, 201]. However, since NF-B (and C/EBP) has so diverse effects in neurons and glial cells in the CNS, a drug that generally inhibits this TF could have detrimental effects. Before this approach of CPP-TF decoy could be used in a therapeutic purpose there are at least two questions that need to be addressed. 1) Is it a good idea to

36 inhibit NF-κB in the CNS? 2) If it is a good idea, could this method be improved to overcome obvious side effects? To inhibit glial NF-κB partially or transiently out of an inflammato- ry perspective would most likely slow disease progression and thereby it would be a good idea. Then one needs to direct the decoy only to- wards glial cell, astrocytes and/or microglia. One way of doing this is to use homing peptides, i. e. CPPs that are directed towards a certain cell type. However, the majority of homing peptides reported so far are directed toward cancer targets. Instead, if the difference in DNA binding affinity of NF-κB in neurons and glial cells could be used in one‟s advantage one could still use a regular CPP to only inhibit glial NF-κB. Still, one should consider choosing another CPP than TP10 since several new CPPs have been developed since the time of publi- cation, that show higher efficiency and less cytotoxicity (e.g. M918 [202]). Also, a lot of effort is made in the CPP field to increase the en- dosomal escape of the CPP-cargo [203, 203-205], thereby making it possi- ble to obtain the same effects with lower concentrations of CPP-cargo and thus avoiding unpleasant side effects. Furthermore, most certainly the TP10-PNA-NF-κB decoy would have a hard time passing the BBB. Finally, one has to remember that the experiments performed so far reflect an approach for disease prevention, where the TP10-PNA-NF- κB decoy have been added to cells before Aβ and IL-1β treatment. Of great interest is to find out how effective this approach is when cells already have been challenged with AD related pathogens.

Paper II: Alzheimer Aβ peptides block the activation of C/EBPβ and C/EBPδ in glial cells. In paper II we turned our attention towards C/EBP, another TF in- volved in AD related inflammation and IL-6 expression. As men- tioned earlier, members of the C/EBP family have been shown to be up-regulated in AD both at mRNA [78, 116, 206] and protein level [113]. Also, during certain circumstances C/EBP is known to form heterodi- mers with NF-κB [207]. Thereby, we wanted to study C/EBP expression and activity in similar experimental conditions as earlier NF-κB stud- ies. C/EBPα, β and δ have all been shown to be expressed in brain. However, in this study we chose to focus on C/EBPβ and δ. We used both full length Aβ(1-42) and the model peptide Aβ(25-35), and the response was analyzed by WB and EMSA. Due to the previous reports, we got very surprised when our results instead of showing up-regulated levels of in response to Aβ and IL-

37 1β/LPS, showed down-regulated protein levels and activity in primary astro-microglial cells. Another remarkable finding was that Aβ(25-35) by it self had no effect on C/EBP binding activity or protein level at this time point. Also, same effects were obtained using different con- formations of Aβ peptides, Aβ(1-42) highly enriched in oligomers and Aβ(25-35) mainly forming β-sheet structures (shown by congo red staining). These unexpected results may be attributed to acute effects of Aβ and IL-1β/LPS, since previous publications reporting opposite effects are derived from post mortem AD brains tissues. Hence, there is a big difference in exposure time of inflammatory agents and Aβ. This hy- pothesis is further supported by Blalock et al., showing increased lev- els of C/EBPδ mRNA only in patients with severe AD [208]. In addi- tion, further experiments are needed in order to clarify if the Aβ/IL- 1β-induced C/EBPβ and δ down-regulation is due to decreased protein synthesis or increased protein degradation. Since Aβ(25-35) co-treatment with IL-1β or LPS showed similar results, it is tempting to suggest that the effect of Aβ most likely occur downstream of MyD88, where IL-1β and LPS signalling pathways are alike (se figure 1 on page 7). Finally, when comparing results from EMSA and WB after IL-1β or LPS treatment, DNA binding activity did not fully correspond to protein levels. This indicates that EMSA is a more sensitive method than WB or that post-translational modifications of C/EBP resulted in improved binding capacity. A third possibility is that C/EBPβ and δ formed heterodimers with members of other TF families but still bound to the C/EBP recognition sequence. In conclusion, contrary to what we expected, C/EBPβ and δ levels and DNA binding activity were decreased in primary astro-microglial cell cultures treated with Aβ together with IL-1β. We propose that our conflicting results compared to existing literature are due to acute ef- fects of Aβ but requires a more comprehensive investigation.

Paper III: The CCAAT/enhancer binding protein (C/EBP) δ is differently regulated by fibrillar and oligomeric forms of the Alzheimer amyloid-β peptide. Due to surprising results in paper II, we decided to study the impact of Aβ(1-42) conformations more closely in paper III, and also investigate more long term effects of Aβ on C/EBP, using aged AD tg-ArcSwe mice. First, we focused on C/EBPα and δ in rat primary astro- microglial cell cultures treated with oligomeric or fibrillar enriched

38 Aβ(1-42), alone or together with IL-1β. LPS was used as an additional independent inflammatory inducer. The results showed that the pro- tein level of C/EBPα was only altered in response to LPS. This is in accordance with a study by Ejarque-Ortiz et al. showing that C/EBPα is mainly regulated through TLRs [97]. However, it should be noted that there was a trend towards C/EBPα down-regulation when cells were exposed to fibril-enriched Aβ and IL-1β, but this effect was not statistically significant. The results on C/EBPδ supported previous findings, namely that fi- brillar forms of Aβ had no effect on its own, but when co-treated with IL-1β significantly down-regulated the C/EBPδ levels. In this study, we showed it using full length Aβ(1-42) and the DNA binding activity was analysed as well, and shown to be down-regulated. In contrast, oligomer enriched Aβ(1-42) did not affect IL-1β induced C/EBPδ pro- tein levels, only C/EBPδ binding activity to the C/EBP element, as shown in paper II. Since C/EBPδ and NF-κB may form hetromeric complexes, we investigated if C/EBPδ during these circumstances changed preferable DNA binding site to the κB response element. In- deed, we found that IL-1β treatment increased C/EBPδ binding to the κB site. However, this effect was completely blocked when cells were co-treated with IL-1β and oligomeric or fibrillar Aβ. Whereas the ef- fect of fibrillar Aβ is well correlated with reduced protein levels, oli- gomeric Aβ reduced the κB binding to the same extent despite un- changed protein levels. Since NF-κB/C/EBP complexes bound to κB containing promoters are believed to inhibit transcription [209] this ef- fect could have deleterious effects in AD related neuroinflammation since it could result in a dysfunctional negative feed-back mechanism regulating NF-κB activity. The results from the in vitro studies were later on confirmed in vivo when we analysed protein levels of C/EBPα, β and δ in aged tg- ArcSwe mice. These mice are characterized by high levels of fibrillar Aβ deposits and therefore in our opinion serve as a good model for studies of long term effects of fibrillar Aβ. In these mice, C/EBPα was shown to be down-regulated indirectly by Aβ since a similar reduction in protein levels were observed in both forebrain and hindbrain of aged tg-ArcSwe mice compared to non-tg mice. In contrast to C/EBPα and in accordance with the in vitro studies, C/EBPδ was selectively down-regulated in plaque containing forebrains of aged tg-ArcSwe mice. Despite the reduced levels of C/EBPδ, there was support for an Aβ-induced inflammatory response in aged tgArcSwe mice based on decreased C/EBPα levels and increased C/EBPβ levels in forebrain, which is expected when an inflammatory response is induced.

39 In the beginning we were also interested in gender differences. However, rather early on, we concluded that gender did not affect C/EBP protein levels in these mice. Even though our result may seem contradictory to the study by Li et al. [113] the differences might be explained. The Aβ plaque burden in tg-ArcSwe mice is most probably of a more homogenous fibrillar na- ture compared to how it might look in the human AD brain. The re- sults presented here may correspond to a very late phase in the disease characterized by high Aβ plaque burden. Nevertheless, since Aβ oli- gomers seem to primary affect C/EBPδ function it cannot be ruled out that DNA binding and transcriptional activity is reduced also in hu- man AD brains. We speculate, with support from earlier studies show- ing different effects of oligomeric and fibrillar Aβ [210, 211], that C/EBPδ could be viewed as a marker for classical cytotoxic glial cells expressing TNFα and that fibrillar Aβ blocks induction of this pheno- type. Interestingly, in a very recently published paper where they stud- ied viability, cytokine expression profile and phagocytic function of microglia in vitro, oligomeric Aβ was shown to be much more cyto- toxic than fibrillar Aβ [212]. Oligomeric Aβ was shown to produce higher levels of TNF, nitric oxide and prostaglandin E2 and to su- press the microglial phagocytic function by down-regulation of Aβ- binding receptors. Moreover, this dysfunctional form of activated mi- croglia was shown to depend on NF-B mediated inflammation. In conclusion, our results indicate that different aggregation states of Aβ regulate C/EBPδ differently during inflammatory conditions. Furthermore, Aβ seems to cause an imbalance between NF-κB and C/EBPδ resulting in abnormal response to inflammatory stimuli. Future perspectives In AD brains, C/EBPδ is primary expressed in astrocytes and associat- ed with neuroinflammation [113]. However, C/EBPs are also expressed in neurons and proposed to be involved in neuronal development [213, 214], learning and memory formation [111, 215]. Changes in C/EBPδ due to both age and Aβ plaque burden (as we also observed in this study) may be attributed to altered expression in different cell types. In order to verify this, immunofluorescence studies of C/EBPs in tissue slices need to be performed. Moreover, it would be interesting to do out- come measurement for the disturbed NF-κB/C/EBP pathway and study the production of inflammatory mediators like IL-6 or TNF-α. Preliminary results could be obtained using primary astro-microglial cultures exposed to IL-1β and oligomeric or fibrillar Aβ, where C/EBPδ has been knocked-down using small interfering RNA. Also, we have preliminary data showing a significant inverse correlation between total Aβ and C/EBPδ expression (r=0.69; p=0.039; n=10) that

40 would be interesting to follow-up, including more mice with varying plaque burden. So far we have only analysed brain tissues with very high Aβ burden and non-transgenic mice devoid of extracellular Aβ deposition. Lastly, it would be very interesting to compare the results obtained in these mice with results from APPE693Δ-tg mice, which is a mouse model of Aβ oligomers [216].

Figure 5. Summarizing figure of paper II and III. We show that different confor- mations of the Aβ peptide have different effects on the TF C/EBPδ in an inflammato- ry milieu. Oligomeric Aβ (oAβ) do not affect IL-1β induced protein expression but DNA binding activity to both the κB and C/EBP site. In contrast, fibrillar Aβ (fAβ) exerts an effect by primarily inhibiting IL-1β induced C/EBPδ expression and there- by subsequent DNA binding activity. We suggest that this Aβ induced dysregulation of C/EBPδ have a negative influence on the NF-κB inhibitory feedback mechanism resulting in a prolonged NF-κB activation. (The effect of IL-1β and Aβ co-treatment on NF-κB activation has been reported in [196]).

Paper IV: Anchored FRET sensors detect local caspase activation prior to neuronal degeneration. Lately, the view of caspase activation has started to change. In recent years, several reports have been published showing caspases to be activated in different cellular events not leading to cell death. In neu- rons, caspase-6 has been shown to be locally activated in response to trophic-factor deprivation [217] and also in neurites in AD brains [218]. Moreover, caspase-3 was recently shown to be locally activated, via caspase-9 and the intrinsic apoptotic pathway, in spines of Tg2576- APPSwe mice. This caspase activation occurred at the same time as the onset of memory deficits [219]. Also, this study indicated an im-

41 portant role of locally activated caspase-3 in memory during normal conditions. Thus, there is a strong need for new techniques that ena- bles studies of how a locally induced signal can spread in vivo. In paper IV, we developed a new set of anchored FRET sensors for detection of caspase-3, -6 or -9 activity in live cells. Our hypothesis was that by anchoring them to the cytoskeleton, they would be less dynamic compared to other existing FRET sensors. Also, by directing them towards axons/neurites via tau, neurodegeneration could more easily be studied. cDNA where constructed to contain ECFP and EYFP separated by an engineered caspase cleavage site (DEVD:caspase-3, VEID:caspase-6, LEHD:caspase-9 and LEVA: control sequence) and full length tau. The FRET sensors were charac- terized in HeLa cells, showing correct localization to microtubules in mitotic and interphase cells and undetectable levels of toxicity when transiently overexpressed. This is in accordance with a previous in vivo study showing that neither tau deletion nor overexpression had any effect on axonal transport in mouse retinal ganglion cells [220]. Moreover, quantitative acceptor bleaching experiments in differentiat- ed SH-SY5Y cells showed all sensors to have a high FRET efficiency. Initially, staurosporine was used in order to induce apoptosis. Stau- roporine treatment of U2OS cells showed the engineered caspase cleavage motif to be favourable before cleavage of other cellular sub- strates since EYFP maintained its localization to microtubule (in con- trast to ECFP) after loss of FRET. Furthermore, loss of FRET could be detected before any morphological changes occurred. To ensure that cleavage of FRET sensors was due to activation of caspases, WB was performed on staurosporine treated SH-SY5Y cells expressing the un-cleavable control sensor or sensors for caspase-3. Caspase-3-sensor expressing cells showed three specific bands corresponding to un- cleaved construct, EYFP-Tau and ECFP, respectively. In presence of caspase inhibitors and in cells expressing the control sensor, bands corresponding to caspase dependent cleavage were absent. Only the band corresponding to un-cleaved construct could be visualized. In order to investigate if the FRET sensors could really pick up a locally induced apoptotic signal, differentiated SH-SY5Y cells were transiently transfected with the sensors and exposed to staurosporine followed by FRET intensity time-lapse imaging. The results showed no difference in spatio-temporal activation of caspase-3 or -9. Howev- er, cells expressing sensors for caspase-3 showed a drop in FRET in- tensity before neurite retraction and the time course was slower as compared to cells expressing sensors for caspase-9. These findings are consistent with earlier studies showing the intrinsic apoptotic signal- ling pathway to be involved in staurosporine induced cell death [221, 222]. Cells expressing FRET sensors for caspase-6 on the other hand,

42 showed a statistically significant drop in FRET intensity in cell bodies before a decrease in neurites could be detected. Thereby the FRET sensors were proven to detect local caspase activation in a reproduci- ble way. Cells expressing the control sensor showed maintained FRET intensities despite obvious apoptotic morphological changes. Co-expression of sensors for caspase-3 and the photosensitizer KillerRed showed that an apoptotic signal could be followed when initiated in one of the neurites, as well. When apoptosis was induced in a neurite, the signal was quickly transferred to the soma but not to other neurites. They were retracted too, but without a change in FRET intensity. Finally, we analysed the effects of oligomeric Aβ on caspase acti- vation and neurodegeneration. It has been shown in vivo that Aβ depo- sitions may lead to local axonal and dendritic abnormalities and ulti- mately neurite breakage [223]. Whether caspases are involved in this process or not is debated [224, 225]. Our results showed an overall activa- tion of caspase-3 and -6 and no activity of caspase-9, when oligomeric Aβ was extracellularly added to differentiated SH-SY5Y cells. There- by, neurites and soma seemed to be equally affected by Aβ treatment and cell death was induced independent of the intrinsic apoptotic pathway. In conclusion, we have developed a new set of anchored FRET sen- sors for detection of locally activated caspase-3, -6 and -9. In vitro, the sensors were demonstrated non-toxic and they were able to detect caspase activation far before any morphological changes of the cells occurred. These sensors open up the possibility to study where in the cell an apoptotic signal is initiated and also if a specific part of the cell is more vulnerable to a certain stimuli. We believe that these sensors will help us and others to get a clearer picture of neurodegeneration associated with AD. Future perspectives To receive information about which caspases that are activated after a certain stimuli in a chosen cell, one would have to construct caspase sensors with a different FRET pair fluorophores. When co-expressed in cells, information about several caspases could be received at the same time. Also, it would be of great interest to develop FRET sensors with a different localization, like synapses for instance. In this study the aim was to study caspase activation during cell death. In the future it would be interesting to study caspase activation during conditions not leading to cell death. However, preliminary data using an Aβ concentration of 2.5 μM gave similar results. Moreover, it would be very exciting to test the FRET sensors in a more complex and natural environment like in slice physiology experiments. Moreo-

43 ver, since another family of proteases, namely calpains, has been pro- posed to be involved in AD related neurodegeneration [226, 227] it would be interesting to co-express our anchored caspase sensors with calpain sensors resembling those developed by Edelman‟s group [228]. Since these proteases are known to have a cross-talk, co-expression could give a clearer picture of neurodegeneration in AD. Finally, it would be interesting to use our new anchored FRET sen- sors in co-cultures of neurons and glial cells challenged with Aβ to get a more complete picture of the involvement of glial cells in neuro- degeneration.

Figure 6. Summarizing figure of paper IV. A new set of anchored FRET sensors were developed in order to detect local caspase activation. Caspase activity was measured as loss of FRET, light grey → dark grey. Differentiated SH-SY5Y cells were transiently transfected with the sensors and exposed to different apoptotic stimuli. Results are shown in the figure.

44 Concluding remarks

AD is a very complex disorder where chronic neuroinflammation is one of the pathological hallmarks. If neuroinflammation contributes to the development of the disease, via „inflammaging‟, or if it is a driving force activated by increased deposition of Aβ is not clear. Perhaps both are true. However, a certain degree of neuroinflammation is ben- eficial since active microglia and reactive astrocytes have the capacity to degrade Aβ. Thereby it is of great importance to define the signal- ling pathways implicated in the advantageous and disadvantageous effects of glial cells in order to develop new effective anti- inflammatory treatments for AD patients. If targeting a TF using de- coy strategies, we showed in paper I that this approach can be greatly improved by non-covalent coupling to a CPP. Using this approach, both decoy concentration and incubation times can be shortened. Also, it highlights the possibility of using CPPs as future transfection rea- gents for glial cells that are otherwise hard to transfect. The inflammatory response may also change during disease pro- gression, which has to be taken into account for the treatment starting point. Depending on the ratio between oligomers:fibrils within the AD brain, TFs can be regulated differently. This was shown in paper II and III, where C/EBP family members responded differently upon stimulation with oligomer or fibril enriched Aβ preparations in an in- flammatory milieu. Based on the obtained results (especially in paper III), we propose that C/EBPδ could be viewed as a marker for classi- cal cytotoxic glial cells and altered expression could be a contributing factor to the prolonged activation of NF-κB in AD. Finally, in hope to increase the understanding about neurodegenera- tion associated with AD, new anchored FRET sensors that enable de- tection of local caspase activation in live neuronal cells, were devel- oped in paper IV. Furthermore, these types of sensors are expected to be very useful in studies on glial cell-mediated effects on neuronal survival and death.

45 Populärvetenskaplig sammanfattning på svenska

Introduktion Alzheimers sjukdom är den vanligaste formen av demens bland äldre människor och antalet drabbade ökar konstant eftersom vi lever allt längre. Denna sjukdom, som har dödlig utgång, kommer ofta smygande med symptom som humörsvängningar, sömnsvårigheter och förvirring. När sjukdomen fortskrider och minnet försämras får den drabbade allt svårare att klara sig på egen hand och är i slutskedet helt oförmögen att utföra de enklaste göromål utan assistans. I dagsläget finns inget botemedel och fokus ligger på att behandla symptomen. En stor anledning till detta är att man inte genom enkel provtagning kan upptäcka sjukdomen i ett tidigt skede hos patienter. Detta leder till att eventuellt fungerande behandlingar påbörjas för sent och man kan därför inte dra någon slutsats om behandlingen är verksam eller inte. Inne i hjärnan kännetecknas sjukdomen av ansamlingar av plack utanför cellerna bestående av amyloid-β (Aβ), ansamlingar inne i cellerna bestående av ett protein kallat tau, massiv nervcellsdöd och kronisk neuroinflammation. Betydelsen av en pågående neuroinflammation i samband med Alzheimers sjukdom har länge varit omdebatterad eftersom den har visat sig ha både positiva och negativa effekter. En väletablerad hypotes är dock att neuroinflammation under ett tidigt skede av sjukdomsförloppet är skyddande men får förödande konsekvenser när den blir kronisk. Att behandla kronisk neuroinflammationen skulle därför vara av stort medicinskt värde. Detta stöds av att personer som lever med ledgångsreumatism och därför systematiskt behandlas med antiinflammatoriska läkemedel utvecklar Alzheimers sjukdom i betydligt mindre omfattning än genomsnittet. Målsättningar Artikel I-III I de tre första studierna var målsättningen att öka förståelsen kring neuroinflammation i samband med Alzheimers sjukdom. Vi valde att studera två olika transkriptionsfaktorer, NF-κB och C/EBP. De är

46 proteiner som bestämmer vilka gener som ska uttryckas i varje cell och ansvarar för produktion av många pro-inflammatoriska molekyler. I försök att efterlikna en kronisk inflammation i en Alzheimersjuk hjärna odlade vi immunceller (mikroglia och astrocyter) från råtthjärnor och behandlade dessa med Aβ och en inflammatorisk faktor. Vi studerade förändring av C/EBP-nivåer samt dess förmåga att binda in till DNA och därmed reglera genuttryck. Vissa av dessa studier kompletterades med studier från Alzheimersjuka transgena möss och vi testade även en ny metod för att blockera aktiviteten av NF-κB.

Artikel IV Målsättningen i den fjärde studien var att utveckla en metod för att möjliggöra studier av degenerering och död av nervceller i realtid. Innan fanns endast metoder som visade om nervceller var på väg på att dö eller inte. Ingen metod kunde visa var dödsprocessen initierades, vilket skulle vara av stor betydelse för att öka förståelsen av neurodegenerering i olika sjukdomar generellt, och Alzheimers sjukdom speciellt. Resultat Tidigare studier har föreslagit att transkriptionsfaktorn NF-κB aldrig ”stängs av” under Alzheimersliknande förhållanden. Det medför en väldigt hög produktion av pro-inflammatoriska molekyler. I artikel I presenterar vi en effektiv metod att blockera NF-κB och därmed minska den skadliga överproduktionen av de pro-inflammatoriska molekylerna. I artikel II och III visar vi att C/EBP regleras annorlunda under inflammatoriska förhållanden i närvaro av Aβ. Dessutom visar vi att olika former av Aβ, som kan symbolisera olika stadier i sjukdomsförloppet, har olika påverkan på C/EBP och eventuellt efterföljande genuttryck. I artikel III framställer vi även en hypotes om hur förändring av C/EBP påverkar NF-κBs avstängsningmekanism. Slutligen, i artikel IV presenterar vi en ny modell för att studera neurodegenerering i realtid. Genom att koppla en sensor till cellens skelett och som slår om från gult till blått när dödsprocessen initieras kan man i mikroskop enkelt följa var cellen börjar dö. Denna metod kommer att vara till stor användning för att förstå hur neuroinflammationen påverkar neurodegenereringen.

47 Acknowledgements

’Att känna tacksamhet och inte uttrycka det är som att slå in en present i ett paket men utan att ge bort det.’ William Arthur Ward

Därför vill jag självklart passa på att tacka personer runt omkring mig som på olika sätt bidragit till att det blev en avhandling. Att doktorera och disputera är långt ifrån en enmansshow.

Kerstin, tusen tack för den här tiden! Du har varit en otroligt inspirerande och entusiasmerande handledare. Jag uppskattar verkligen den frihet du gett mig och att du aldrig varit för upptagen för att lyssna. Einar, tack för ett bra år på Södertörn och gott samarbete under caspaseprojektet. Jag beundrar din förmåga att tänka utanför den berömda lådan. Tack Anna-Lena för all positiv energi du sprider runt omkring dig. Tack Anna för att man alltid kan vända sig till dig vad det än gäller. Tack Lars för lärorikt och roligt samarbete i artikel III. Tack Anders för givande föreläsningar. Tack Ülo för ett gott samarbete i artikel I. Tack Siv, Marie-Louise, Birgitta, Ulla och Sylvia för all hjälp med administrativa ärenden och alla våra trevliga pratstunder i köket.

Ett STORT tack även till alla doktorander, post-docs, forskare och ex-jobbare som gör Institutionen för Neurokemi till en unik och rolig arbetsplats. Jag vill rikta ett extra tack till några speciella: Tack Malin S! Det har varit otroligt roligt att jobba tillsammans med dig. Du är fantastisk på alla sätt och vis. Varje gång jag flyger tänker jag på dig och din päronlikör, det är pensionärsvarning på den! ;) Sofia, tack för att du är en fin vän och för strålande resesällskap i Prag. Linda T, tack för gott samarbete, men framför allt för din omtänksamhet och för att jag fick vara en del av er familj i Florens. Tack Kristin, din höga ambitionsnivå har inspirerat mig. Tack Tom för att du kom in som en frisk fläkt i caspaseprojektet, det var välbehövt. Tack Niina för trevliga pratstunder på kontoret och för alla roliga mail... Tack Linda A, Linda F, Elena och Ylva för alla givande samtal ute på labbet och i lunchrummet. Tack Nina för att du är så underbart rolig och snäll. Tack för att du är mitt privata 118100, dit jag kan ringa och fråga saker i tid och otid. När och var ses vi för vår utekväll?! Tack Marie för att du är en fin vän och för roliga (om än hysteriska) shoppingrundor. Tack Ricardo för roligt samarbete och intressanta kommentarer. Tack Jessica för att du alltid lyser upp tillvaron med roliga berättelser. Tack Tina

48 och Ulla, och resten av syjuntan, för alla skratt och mysiga kvällar med god mat. Tack Lotta och Robert för en trevlig tid på Södertörn. Tack Santhosh för alla artiklar. Tack Johanna och Caroline för allt roligt vi haft på och utanför jobbet. Av någon konstig anledning blev jag faktiskt mer effektiv när ni slutade... Men, jag saknar er och nu tycker jag att det är dax för er att flytta tillbaka till Stockholm!

Ett extra varmt tack till alla er som tog er tid att korrekturläsa avhandlingen: Kerstin, Einar, Anna-Lena, Nina, Linda T, Ricardo och Ellinor. Tack Erik för att hjälpen med bilderna.

Dessutom ett stort tack till alla kära vänner utanför jobbet. Tack Malin och Helena för att ni förstår mig så bra och för att ni alltid ställer upp vad det än gäller. Tack till Oskar, Stefan, Ellinor & Dante, Ingrid & Jocke, Mia & Jocke, Marie & Jonas, Tjodolf & Elin, Martin & Elin och Daniel & Anki för semestrar, fester, middagar, biokvällar, spelkvällar, resor och promenader. Tack för att ni lyssnat och i alla fall låtsats vara intresserade av mina forskningsrön…

Tack Sigfrid och Elsy för gott fika, trevliga och intressanta diskussioner varje gång vi kommer till Baskeri.

Tack Laila och Bengt-Åke för att ni är så otroligt snälla och alltid ställer upp. Er hjälp med barnvakt och avlastning har varit ovärderlig. Tack Kristina och Staffan för att jag och familjen fått komma förbi på middagar, lek och bus, vilket gett mig många behövliga pauser från arbetet med avhandlingen.

Mamma Ulla och pappa Jan-Erik, tack för att ni alltid har stöttat och uppmuntrat mig i allt från att bli cirkusakrobat till att disputera i neurokemi. Ni har alltid fått mig att vilja göra mitt bästa. Tack Victoria, världens bästa storasyster, för din omtänksamhet och för att du alltid har tid för mig. Att träffa dig och Peder är alltid roligt och får mig att längta hem till Norrland igen.

Slutligen, tack Carl och Victor för att ni är världens bästa och mest underbara barn. Tack för att ni får mig att tänka på allt annat än jobb. Ni är min största bedrift och de mest lyckade resultaten av alla mina experiment… Erik, ord kan inte beskriva hur mycket du betyder för mig ♥.

TACK!

49 Referenes

[1] C.P. Ferri, M. Prince, C. Brayne, H. Brodaty, L. Fratiglioni, M. Ganguli, K. Hall, K. Hasegawa, H. Hendrie, Y. Huang, A. Jorm, C. Mathers, P.R. Menezes, E. Rimmer, M. Scazufca, Alzheimer's Disease International. Global prevalence of dementia: A delphi consensus study, Lancet 366 (2005) 2112-2117.

[2] M.F. Folstein, S.E. Folstein, P.R. McHugh. "Mini-mental state". A practical method for grading the cognitive state of patients for the clinician, J. Psychiatr. Res. 12 (1975) 189-198.

[3] H.C. Huang, Z.F. Jiang. Accumulated amyloid-beta peptide and hyperphosphorylated : Relationship and links in alzheimer's disease, J. Alzheimers Dis. 16 (2009) 15-27.

[4] M. Goedert, M.G. Spillantini. A century of alzheimer's disease, Science 314 (2006) 777-781.

[5] H. Braak, E. Braak. Neuropathological stageing of alzheimer-related changes, Acta Neuropathol. 82 (1991) 239-259.

[6] H. Akiyama, S. Barger, S. Barnum, B. Bradt, J. Bauer, G.M. Cole, N.R. Cooper, P. Eikelenboom, M. Emmerling, B.L. Fiebich, C.E. Finch, S. Frautschy, W.S. Griffin, H. Hampel, M. Hull, G. Landreth, L. Lue, R. Mrak, I.R. Mackenzie, P.L. McGeer, M.K. O'Banion, J. Pachter, G. Pasinetti, C. Plata-Salaman, J. Rogers, R. Rydel, Y. Shen, W. Streit, R. Strohmeyer, I. Tooyoma, F.L. Van Muiswinkel, R. Veerhuis, D. Walker, S. Webster, B. Wegrzyniak, G. Wenk, T. Wyss-Coray. In- flammation and alzheimer's disease, Neurobiol. Aging 21 (2000) 383-421.

[7] J.A. Hardy, G.A. Higgins. Alzheimer's disease: The amyloid cascade hypothesis, Science 256 (1992) 184-185.

[8] D.J. Selkoe. The molecular pathology of alzheimer's disease, Neuron 6 (1991) 487-498.

[9] C. Haass, C.A. Lemere, A. Capell, M. Citron, P. Seubert, D. Schenk, L. Lannfelt, D.J. Selkoe. The swedish mutation causes early-onset alzheimer's disease by beta-secretase cleavage within the se- cretory pathway, Nat. Med. 1 (1995) 1291-1296.

[10] C. Nilsberth, A. Westlind-Danielsson, C.B. Eckman, M.M. Condron, K. Axelman, C. Forsell, C. Stenh, J. Luthman, D.B. Teplow, S.G. Younkin, J. Naslund, L. Lannfelt. The 'arctic' APP mutation (E693G) causes alzheimer's disease by enhanced abeta protofibril formation, Nat. Neurosci. 4 (2001) 887-893.

[11] C. Hooper, R. Killick, S. Lovestone. The GSK3 hypothesis of alzheimer's disease, J. Neurochem. 104 (2008) 1433-1439.

50 [12] A.R. Simard, D. Soulet, G. Gowing, J.P. Julien, S. Rivest. Bone marrow-derived microglia play a critical role in restricting senile plaque formation in alzheimer's disease, Neuron 49 (2006) 489- 502.

[13] M. Koistinaho, S. Lin, X. Wu, M. Esterman, D. Koger, J. Hanson, R. Higgs, F. Liu, S. Malkani, K.R. Bales, S.M. Paul. Apolipoprotein E promotes astrocyte colocalization and degradation of deposited amyloid-beta peptides, Nat. Med. 10 (2004) 719-726.

[14] K.L. Richard, M. Filali, P. Prefontaine, S. Rivest. Toll-like receptor 2 acts as a natural innate im- mune receptor to clear 1-42 and delay the cognitive decline in a mouse model of alz- heimer's disease, J. Neurosci. 28 (2008) 5784-5793.

[15] C.R. Abraham, D.J. Selkoe, H. Potter, D.L. Price, L.C. Cork. Alpha 1-antichymotrypsin is present together with the beta-protein in monkey brain amyloid deposits, Neuroscience 32 (1989) 715-720.

[16] P. Eikelenboom, F.C. Stam. Immunoglobulins and complement factors in senile plaques. an immun- operoxidase study, Acta Neuropathol. 57 (1982) 239-242.

[17] R. Veerhuis, R.S. Boshuizen, A. Familian. Amyloid associated proteins in alzheimer's and prion disease, Curr. Drug Targets CNS Neurol. Disord. 4 (2005) 235-248.

[18] P. Eikelenboom, R. Veerhuis, E. van Exel, J.J. Hoozemans, A.J. Rozemuller, W.A. van Gool. The early involvement of the innate immunity in the pathogenesis of late-onset alzheimer's disease: Neuropathological, epidemiological and genetic evidence, Curr. Alzheimer Res. 8 (2011) 142-150.

[19] D. Goldgaber, H.W. Harris, T. Hla, T. Maciag, R.J. Donnelly, J.S. Jacobsen, M.P. Vitek, D.C. Gaj- dusek. Interleukin 1 regulates synthesis of amyloid beta-protein precursor mRNA in human endo- thelial cells, Proc. Natl. Acad. Sci. U. S. A. 86 (1989) 7606-7610.

[20] J.D. Buxbaum, M. Oishi, H.I. Chen, R. Pinkas-Kramarski, E.A. Jaffe, S.E. Gandy, P. Greengard. Cholinergic agonists and interleukin 1 regulate processing and secretion of the alzheimer beta/A4 amyloid protein precursor, Proc. Natl. Acad. Sci. U. S. A. 89 (1992) 10075-10078.

[21] P. Edison, H.A. Archer, A. Gerhard, R. Hinz, N. Pavese, F.E. Turkheimer, A. Hammers, Y.F. Tai, N. Fox, A. Kennedy, M. Rossor, D.J. Brooks. Microglia, amyloid, and cognition in alzheimer's dis- ease: An [11C](R)PK11195-PET and [11C]PIB-PET study, Neurobiol. Dis. 32 (2008) 412-419.

[22] J. Koenigsknecht-Talboo, G.E. Landreth. Microglial phagocytosis induced by fibrillar beta-amyloid and IgGs are differentially regulated by proinflammatory cytokines, J. Neurosci. 25 (2005) 8240- 8249.

[23] M. Gatz, C.A. Reynolds, L. Fratiglioni, B. Johansson, J.A. Mortimer, S. Berg, A. Fiske, N.L. Pedersen. Role of genes and environments for explaining alzheimer disease, Arch. Gen. Psychiatry 63 (2006) 168-174.

[24] E. van Exel, P. Eikelenboom, H. Comijs, M. Frolich, J.H. Smit, M.L. Stek, P. Scheltens, J.E. Eef- sting, R.G. Westendorp. Vascular factors and markers of inflammation in offspring with a parental history of late-onset alzheimer disease, Arch. Gen. Psychiatry 66 (2009) 1263-1270.

[25] M. Ek, D. Engblom, S. Saha, A. Blomqvist, P.J. Jakobsson, A. Ericsson-Dahlstrand. Inflammatory response: Pathway across the blood-brain barrier, Nature 410 (2001) 430-431.

[26] C.T. Kodl, E.R. Seaquist. Cognitive dysfunction and diabetes mellitus, Endocr. Rev. 29 (2008) 494- 511.

51 [27] K. Yaffe, A. Kanaya, K. Lindquist, E.M. Simonsick, T. Harris, R.I. Shorr, F.A. Tylavsky, A.B. Newman. The metabolic syndrome, inflammation, and risk of cognitive decline, JAMA 292 (2004) 2237-2242.

[28] W.Y. Chan, S. Kohsaka, P. Rezaie. The origin and cell lineage of microglia: New concepts, Brain Res. Rev. 53 (2007) 344-354.

[29] R.M. Ransohoff, V.H. Perry. Microglial physiology: Unique stimuli, specialized responses, Annu. Rev. Immunol. 27 (2009) 119-145.

[30] W.J. Streit, N.W. Sammons, A.J. Kuhns, D.L. Sparks. Dystrophic microglia in the aging human brain, Glia 45 (2004) 208-212.

[31] V.H. Perry, J.A. Nicoll, C. Holmes. Microglia in neurodegenerative disease, Nat. Rev. Neurol. 6 (2010) 193-201.

[32] M. Meyer-Luehmann, T.L. Spires-Jones, C. Prada, M. Garcia-Alloza, A. de Calignon, A. Rozkalne, J. Koenigsknecht-Talboo, D.M. Holtzman, B.J. Bacskai, B.T. Hyman. Rapid appearance and local toxicity of amyloid-beta plaques in a mouse model of alzheimer's disease, Nature 451 (2008) 720- 724.

[33] W.J. Streit, K.R. Miller, K.O. Lopes, E. Njie. Microglial degeneration in the aging brain--bad news for neurons? Front. Biosci. 13 (2008) 3423-3438.

[34] A.H. de Haas, H.W. Boddeke, K. Biber. Region-specific expression of immunoregulatory proteins on microglia in the healthy CNS, Glia 56 (2008) 888-894.

[35] A. Nimmerjahn, F. Kirchhoff, F. Helmchen. Resting microglial cells are highly dynamic surveillants of brain parenchyma in vivo, Science 308 (2005) 1314-1318.

[36] P. Rezaie, G. Trillo-Pazos, J. Greenwood, I.P. Everall, D.K. Male. Motility and ramification of human fetal microglia in culture: An investigation using time-lapse video microscopy and image analysis, Exp. Cell Res. 274 (2002) 68-82.

[37] D.G. Walker, J.E. Dalsing-Hernandez, N.A. Campbell, L.F. Lue. Decreased expression of CD200 and CD200 receptor in alzheimer's disease: A potential mechanism leading to chronic inflamma- tion, Exp. Neurol. 215 (2009) 5-19.

[38] A. Okello, P. Edison, H.A. Archer, F.E. Turkheimer, J. Kennedy, R. Bullock, Z. Walker, A. Kenne- dy, N. Fox, M. Rossor, D.J. Brooks. Microglial activation and amyloid deposition in mild cogni- tive impairment: A PET study, Neurology 72 (2009) 56-62.

[39] A. Salminen, J. Ojala, A. Kauppinen, K. Kaarniranta, T. Suuronen. Inflammation in alzheimer's disease: Amyloid-beta oligomers trigger innate immunity defence via pattern recognition recep- tors, Prog. Neurobiol. 87 (2009) 181-194.

[40] G.E. Landreth, E.G. Reed-Geaghan. Toll-like receptors in alzheimer's disease, Curr. Top. Microbiol. Immunol. 336 (2009) 137-153.

[41] J.C. Kagan, T. Su, T. Horng, A. Chow, S. Akira, R. Medzhitov. TRAM couples endocytosis of toll- like receptor 4 to the induction of interferon-beta, Nat. Immunol. 9 (2008) 361-368.

52 [42] E.G. Reed-Geaghan, J.C. Savage, A.G. Hise, G.E. Landreth. CD14 and toll-like receptors 2 and 4 are required for fibrillar A{beta}-stimulated microglial activation, J. Neurosci. 29 (2009) 11982- 11992.

[43] K. Fassbender, S. Walter, S. Kuhl, R. Landmann, K. Ishii, T. Bertsch, A.K. Stalder, F. Muehlhauser, Y. Liu, A.J. Ulmer, S. Rivest, A. Lentschat, E. Gulbins, M. Jucker, M. Staufenbiel, K. Brechtel, J. Walter, G. Multhaup, B. Penke, Y. Adachi, T. Hartmann, K. Beyreuther. The LPS receptor (CD14) links innate immunity with alzheimer's disease, FASEB J. 18 (2004) 203-205.

[44] M. Letiembre, Y. Liu, S. Walter, W. Hao, T. Pfander, A. Wrede, W. Schulz-Schaeffer, K. Fass- bender. Screening of innate immune receptors in neurodegenerative diseases: A similar pattern, Neurobiol. Aging 30 (2009) 759-768.

[45] Y. Liu, S. Walter, M. Stagi, D. Cherny, M. Letiembre, W. Schulz-Schaeffer, H. Heine, B. Penke, H. Neumann, K. Fassbender. LPS receptor (CD14): A receptor for phagocytosis of alzheimer's amy- loid peptide, Brain 128 (2005) 1778-1789.

[46] S. Walter, M. Letiembre, Y. Liu, H. Heine, B. Penke, W. Hao, B. Bode, N. Manietta, J. Walter, W. Schulz-Schuffer, K. Fassbender. Role of the toll-like receptor 4 in neuroinflammation in alzhei- mer's disease, Cell. Physiol. Biochem. 20 (2007) 947-956.

[47] J.J. Jin, H.D. Kim, J.A. Maxwell, L. Li, K. Fukuchi. Toll-like receptor 4-dependent upregulation of cytokines in a transgenic mouse model of alzheimer's disease, J. Neuroinflammation 5 (2008) 23.

[48] P. Minoretti, C. Gazzaruso, C.D. Vito, E. Emanuele, M. Bianchi, E. Coen, M. Reino, D. Geroldi. Effect of the functional toll-like receptor 4 Asp299Gly polymorphism on susceptibility to late- onset alzheimer's disease, Neurosci. Lett. 391 (2006) 147-149.

[49] S.E. Hickman, E.K. Allison, J. El Khoury. Microglial dysfunction and defective beta-amyloid clear- ance pathways in aging alzheimer's disease mice, J. Neurosci. 28 (2008) 8354-8360.

[50] I. Allaman, M. Belanger, P.J. Magistretti. Astrocyte-neuron metabolic relationships: For better and for worse, Trends Neurosci. 34 (2011) 76-87.

[51] A. Volterra, J. Meldolesi. Astrocytes, from brain glue to communication elements: The revolution continues, Nat. Rev. Neurosci. 6 (2005) 626-640.

[52] W. Liu, Y. Tang, J. Feng. Cross talk between activation of microglia and astrocytes in pathological conditions in the central nervous system, Life Sci. 89 (2011) 141-146.

[53] M.V. Sofroniew, H.V. Vinters. Astrocytes: Biology and pathology, Acta Neuropathol. 119 (2010) 7- 35.

[54] B.A. Barres. The mystery and magic of glia: A perspective on their roles in health and disease, Neuron 60 (2008) 430-440.

[55] K.V. Kuchibhotla, C.R. Lattarulo, B.T. Hyman, B.J. Bacskai. Synchronous hyperactivity and inter- cellular calcium waves in astrocytes in alzheimer mice, Science 323 (2009) 1211-1215.

[56] A.Y. Abramov, L. Canevari, M.R. Duchen. Changes in intracellular calcium and glutathione in astrocytes as the primary mechanism of amyloid neurotoxicity, J. Neurosci. 23 (2003) 5088-5095.

53 [57] C.J. Garwood, A.M. Pooler, J. Atherton, D.P. Hanger, W. Noble. Astrocytes are important mediators of abeta-induced neurotoxicity and tau phosphorylation in primary culture, Cell. Death Dis. 2 (2011) e167.

[58] S. Fuller, M. Steele, G. Munch. Activated astroglia during chronic inflammation in alzheimer's disease--do they neglect their neurosupportive roles? Mutat. Res. 690 (2010) 40-49.

[59] S.A. Scott, S.A. Johnson, C. Zarow, L.S. Perlmutter. Inability to detect beta-amyloid protein precur- sor mRNA in alzheimer plaque-associated microglia, Exp. Neurol. 121 (1993) 113-118.

[60] M. Hartlage-Rubsamen, U. Zeitschel, J. Apelt, U. Gartner, H. Franke, T. Stahl, A. Gunther, R. Schliebs, M. Penkowa, V. Bigl, S. Rossner. Astrocytic expression of the alzheimer's disease beta- secretase (BACE1) is stimulus-dependent, Glia 41 (2003) 169-179.

[61] S. Rossner, J. Apelt, R. Schliebs, J.R. Perez-Polo, V. Bigl. Neuronal and glial beta-secretase (BACE) protein expression in transgenic Tg2576 mice with amyloid plaque pathology, J. Neurosci. Res. 64 (2001) 437-446.

[62] W.C. Huang, F.C. Yen, F.S. Shie, C.M. Pan, Y.J. Shiao, C.N. Yang, F.L. Huang, Y.J. Sung, H.J. Tsay. TGF-beta1 blockade of microglial chemotaxis toward abeta aggregates involves SMAD sig- naling and down-regulation of CCL5, J. Neuroinflammation 7 (2010) 28.

[63] C.K. Glass, K. Saijo, B. Winner, M.C. Marchetto, F.H. Gage. Mechanisms underlying inflammation in neurodegeneration, Cell 140 (2010) 918-934.

[64] C. Franceschi, M. Bonafe, S. Valensin, F. Olivieri, M. De Luca, E. Ottaviani, G. De Benedictis. Inflamm-aging. an evolutionary perspective on immunosenescence, Ann. N. Y. Acad. Sci. 908 (2000) 244-254.

[65] U. Fagiolo, A. Cossarizza, E. Scala, E. Fanales-Belasio, C. Ortolani, E. Cozzi, D. Monti, C. Frances- chi, R. Paganelli. Increased cytokine production in mononuclear cells of healthy elderly people, Eur. J. Immunol. 23 (1993) 2375-2378.

[66] B. Giunta, F. Fernandez, W.V. Nikolic, D. Obregon, E. Rrapo, T. Town, J. Tan. Inflammaging as a prodrome to alzheimer's disease, J. Neuroinflammation 5 (2008) 51.

[67] L.F. Lue, L. Brachova, W.H. Civin, J. Rogers. Inflammation, A beta deposition, and formation as correlates of alzheimer's disease neurodegeneration, J. Neuropathol. Exp. Neu- rol. 55 (1996) 1083-1088.

[68] D. Lio, L. Scola, A. Crivello, G. Colonna-Romano, G. Candore, M. Bonafe, L. Cavallone, C. Fran- ceschi, C. Caruso. Gender-specific association between -1082 IL-10 promoter polymorphism and longevity, Genes Immun. 3 (2002) 30-33.

[69] D. Lio, F. Licastro, L. Scola, M. Chiappelli, L.M. Grimaldi, A. Crivello, G. Colonna-Romano, G. Candore, C. Franceschi, C. Caruso. Interleukin-10 promoter polymorphism in sporadic alzheimer's disease, Genes Immun. 4 (2003) 234-238.

[70] A. Salminen, J. Ojala, K. Kaarniranta, A. Haapasalo, M. Hiltunen, H. Soininen. Astrocytes in the aging brain express characteristics of senescence-associated secretory phenotype, Eur. J. Neurosci. 34 (2011) 3-11.

[71] A. Freund, A.V. Orjalo, P.Y. Desprez, J. Campisi. Inflammatory networks during cellular senes- cence: Causes and consequences, Trends Mol. Med. 16 (2010) 238-246.

54 [72] R. von Bernhardi, J.E. Tichauer, J. Eugenin. Aging-dependent changes of microglial cells and their relevance for neurodegenerative disorders, J. Neurochem. 112 (2010) 1099-1114.

[73] A.M. Floden, C.K. Combs. Beta-amyloid stimulates murine postnatal and adult microglia cultures in a unique manner, J. Neurosci. 26 (2006) 4644-4648.

[74] G. Orphanides, D. Reinberg. A unified theory of gene expression, Cell 108 (2002) 439-451.

[75] R. Tupler, G. Perini, M.R. Green. Expressing the human genome, Nature 409 (2001) 832-833.

[76] D.G. Edmondson, S.Y. Roth. Chromatin and transcription, FASEB J. 10 (1996) 1173-1182.

[77] B. Kaltschmidt, M. Uherek, B. Volk, P.A. Baeuerle, C. Kaltschmidt. Transcription factor NF-kappaB is activated in primary neurons by amyloid beta peptides and in neurons surrounding early plaques from patients with alzheimer disease, Proc. Natl. Acad. Sci. U. S. A. 94 (1997) 2642-2647.

[78] W.J. Lukiw. Gene expression profiling in fetal, aged, and alzheimer hippocampus: A continuum of stress-related signaling, Neurochem. Res. 29 (2004) 1287-1297.

[79] K. Terai, A. Matsuo, E.G. McGeer, P.L. McGeer. Enhancement of immunoreactivity for NF-kappa B in human cerebral infarctions, Brain Res. 739 (1996) 343-349.

[80] M. Samuelsson, L. Fisher, K. Iverfeldt. Beta-amyloid and interleukin-1beta induce persistent NF- kappaB activation in rat primary glial cells, Int. J. Mol. Med. 16 (2005) 449-453.

[81] C. Bonaiuto, P.P. McDonald, F. Rossi, M.A. Cassatella. Activation of nuclear factor-kappa B by beta-amyloid peptides and interferon-gamma in murine microglia, J. Neuroimmunol. 77 (1997) 51- 56.

[82] Y. Du, X. Chen, X. Wei, K.R. Bales, D.T. Berg, S.M. Paul, M.R. Farlow, B. Maloney, Y.W. Ge, D.K. Lahiri. NF-(kappa)B mediates amyloid beta peptide-stimulated activity of the human apolipoprotein E gene promoter in human astroglial cells, Brain Res. Mol. Brain Res. 136 (2005) 177-188.

[83] P.P. Tak, G.S. Firestein. NF-kappaB: A key role in inflammatory diseases, J. Clin. Invest. 107 (2001) 7-11.

[84] R. Sen, D. Baltimore. Multiple nuclear factors interact with the immunoglobulin enhancer sequences, Cell 46 (1986) 705-716.

[85] Q. Li, I.M. Verma. NF-kappaB regulation in the immune system, Nat. Rev. Immunol. 2 (2002) 725- 734.

[86] F. Arenzana-Seisdedos, P. Turpin, M. Rodriguez, D. Thomas, R.T. Hay, J.L. Virelizier, C. Darge- mont. Nuclear localization of I kappa B alpha promotes active transport of NF-kappa B from the nucleus to the cytoplasm, J. Cell. Sci. 110 ( Pt 3) (1997) 369-378.

[87] M.P. Mattson, M.K. Meffert. Roles for NF-kappaB in nerve cell survival, plasticity, and disease, Cell Death Differ. 13 (2006) 852-860.

[88] M.K. Meffert, D. Baltimore. Physiological functions for brain NF-kappaB, Trends Neurosci. 28 (2005) 37-43.

55 [89] D.R. Kaplan, F.D. Miller. Neurotrophin signal transduction in the nervous system, Curr. Opin. Neurobiol. 10 (2000) 381-391.

[90] J.R. Bethea, M. Castro, R.W. Keane, T.T. Lee, W.D. Dietrich, R.P. Yezierski. Traumatic injury induces nuclear factor-kappaB activation, J. Neurosci. 18 (1998) 3251-3260.

[91] P.F. Johnson, W.H. Landschulz, B.J. Graves, S.L. McKnight. Identification of a rat liver nuclear protein that binds to the enhancer core element of three animal viruses, Genes Dev. 1 (1987) 133- 146.

[92] S.C. Williams, N.D. Angerer, P.F. Johnson. C/EBP proteins contain nuclear localization signals imbedded in their basic regions, Gene Expr. 6 (1997) 371-385.

[93] D.P. Ramji, P. Foka. CCAAT/enhancer-binding proteins: Structure, function and regulation, Bio- chem. J. 365 (2002) 561-575.

[94] T. Alam, M.R. An, J. Papaconstantinou. Differential expression of three C/EBP isoforms in multiple tissues during the acute phase response, J. Biol. Chem. 267 (1992) 5021-5024.

[95] P. Descombes, U. Schibler. A liver-enriched transcriptional activator protein, LAP, and a transcrip- tional inhibitory protein, LIP, are translated from the same mRNA, Cell 67 (1991) 569-579.

[96] A.L. Welm, N.A. Timchenko, G.J. Darlington. C/EBPalpha regulates generation of C/EBPbeta isoforms through activation of specific proteolytic cleavage, Mol. Cell. Biol. 19 (1999) 1695-1704.

[97] A. Ejarque-Ortiz, J.M. Tusell, J. Serratosa, J. Saura. CCAAT/enhancer binding protein-alpha is down-regulated by toll-like receptor agonists in microglial cells, J. Neurosci. Res. 85 (2007) 985- 993.

[98] A. Ejarque-Ortiz, M.G. Medina, J.M. Tusell, A.P. Perez-Gonzalez, J. Serratosa, J. Saura. Upregula- tion of CCAAT/enhancer binding protein beta in activated astrocytes and microglia, Glia 55 (2007) 178-188.

[99] A. Ejarque-Ortiz, N. Gresa-Arribas, M. Straccia, P. Mancera, C. Sola, J.M. Tusell, J. Serratosa, J. Saura. CCAAT/enhancer binding protein delta in microglial activation, J. Neurosci. Res. 88 (2010) 1113-1123.

[100] V. Poli. The role of C/EBP isoforms in the control of inflammatory and native immunity functions, J. Biol. Chem. 273 (1998) 29279-29282.

[101] A. Ejarque-Ortiz, M.G. Medina, J.M. Tusell, A.P. Perez-Gonzalez, J. Serratosa, J. Saura. Upregula- tion of CCAAT/enhancer binding protein beta in activated astrocytes and microglia, Glia 55 (2007) 178-188.

[102] C.F. Calkhoven, C. Muller, A. Leutz. Translational control of C/EBPalpha and C/EBPbeta isoform expression, Genes Dev. 14 (2000) 1920-1932.

[103] A. Ray, B.K. Ray. Serum amyloid A gene expression under acute-phase conditions involves partic- ipation of inducible C/EBP-beta and C/EBP-delta and their activation by phosphorylation, Mol. Cell. Biol. 14 (1994) 4324-4332.

[104] T.I. Cesena, J.R. Cardinaux, R. Kwok, J. Schwartz. CCAAT/enhancer-binding protein (C/EBP) beta is acetylated at multiple lysines: Acetylation of C/EBPbeta at lysine 39 modulates its ability to activate transcription, J. Biol. Chem. 282 (2007) 956-967.

56 [105] J. Kim, C.A. Cantwell, P.F. Johnson, C.M. Pfarr, S.C. Williams. Transcriptional activity of CCAAT/enhancer-binding proteins is controlled by a conserved inhibitory domain that is a target for sumoylation, J. Biol. Chem. 277 (2002) 38037-38044.

[106] H. Wang, B. Larris, T.H. Peiris, L. Zhang, J. Le Lay, Y. Gao, L.E. Greenbaum. C/EBPbeta acti- vates E2F-regulated genes in vivo via recruitment of the coactivator CREB-binding protein/P300, J. Biol. Chem. 282 (2007) 24679-24688.

[107] B. Stein, P.C. Cogswell, A.S. Baldwin Jr. Functional and physical associations between NF-kappa B and C/EBP family members: A rel domain-bZIP interaction, Mol. Cell. Biol. 13 (1993) 3964- 3974.

[108] N. Kfoury, G. Kapatos. Identification of neuronal target genes for CCAAT/enhancer binding pro- teins, Mol. Cell. Neurosci. 40 (2009) 313-327.

[109] K. Yukawa, T. Tanaka, S. Tsuji, S. Akira. Expressions of CCAAT/Enhancer-binding proteins beta and delta and their activities are intensified by cAMP signaling as well as Ca2+/calmodulin kinas- es activation in hippocampal neurons, J. Biol. Chem. 273 (1998) 31345-31351.

[110] E. Sterneck, R. Paylor, V. Jackson-Lewis, M. Libbey, S. Przedborski, L. Tessarollo, J.N. Crawley, P.F. Johnson. Selectively enhanced contextual fear conditioning in mice lacking the transcriptional regulator CCAAT/enhancer binding protein delta, Proc. Natl. Acad. Sci. U. S. A. 95 (1998) 10908- 10913.

[111] S.M. Taubenfeld, K.A. Wiig, B. Monti, B. Dolan, G. Pollonini, C.M. Alberini. Fornix-dependent induction of hippocampal CCAAT enhancer-binding protein [beta] and [delta] co-localizes with phosphorylated cAMP response element-binding protein and accompanies long-term memory con- solidation, J. Neurosci. 21 (2001) 84-91.

[112] J.R. Cardinaux, I. Allaman, P.J. Magistretti. Pro-inflammatory cytokines induce the transcription factors C/EBPbeta and C/EBPdelta in astrocytes, Glia 29 (2000) 91-97.

[113] R. Li, R. Strohmeyer, Z. Liang, L.F. Lue, J. Rogers. CCAAT/enhancer binding protein delta (C/EBPdelta) expression and elevation in alzheimer's disease, Neurobiol. Aging 25 (2004) 991- 999.

[114] C.Y. Ko, L.H. Chang, Y.C. Lee, E. Sterneck, C.P. Cheng, S.H. Chen, A.M. Huang, J.T. Tseng, J.M. Wang. CCAAT/enhancer binding protein delta (CEBPD) elevating PTX3 expression inhibits mac- rophage-mediated phagocytosis of dying neuron cells, Neurobiol. Aging (2010) .

[115] R. Sandhir, N.E. Berman. Age-dependent response of CCAAT/enhancer binding proteins following traumatic brain injury in mice, Neurochem. Int. 56 (2010) 188-193.

[116] V. Colangelo, J. Schurr, M.J. Ball, R.P. Pelaez, N.G. Bazan, W.J. Lukiw. Gene expression profiling of 12633 genes in alzheimer hippocampal CA1: Transcription and neurotrophic factor down- regulation and up-regulation of apoptotic and pro-inflammatory signaling, J. Neurosci. Res. 70 (2002) 462-473.

[117] A. Avital, I. Goshen, A. Kamsler, M. Segal, K. Iverfeldt, G. Richter-Levin, R. Yirmiya. Impaired interleukin-1 signaling is associated with deficits in hippocampal memory processes and neural plasticity, Hippocampus 13 (2003) 826-834.

[118] F.M. Ross, S.M. Allan, N.J. Rothwell, A. Verkhratsky. A dual role for interleukin-1 in LTP in mouse hippocampal slices, J. Neuroimmunol. 144 (2003) 61-67.

57 [119] R.E. Mrak, W.S. Griffin. Interleukin-1, neuroinflammation, and alzheimer's disease, Neurobiol. Aging 22 (2001) 903-908.

[120] W.S. Griffin, L.C. Stanley, C. Ling, L. White, V. MacLeod, L.J. Perrot, C.L. White 3rd, C. Araoz. Brain interleukin 1 and S-100 immunoreactivity are elevated in down syndrome and alzheimer disease, Proc. Natl. Acad. Sci. U. S. A. 86 (1989) 7611-7615.

[121] N.J. Rothwell, G.N. Luheshi. Interleukin 1 in the brain: Biology, pathology and therapeutic target, Trends Neurosci. 23 (2000) 618-625.

[122] J.A. Nicoll, R.E. Mrak, D.I. Graham, J. Stewart, G. Wilcock, S. MacGowan, M.M. Esiri, L.S. Murray, D. Dewar, S. Love, T. Moss, W.S. Griffin. Association of interleukin-1 gene polymor- phisms with alzheimer's disease, Ann. Neurol. 47 (2000) 365-368.

[123] F. Licastro, F. Veglia, M. Chiappelli, L.M. Grimaldi, E. Masliah. A polymorphism of the interleu- kin-1 beta gene at position +3953 influences progression and neuro-pathological hallmarks of alz- heimer's disease, Neurobiol. Aging 25 (2004) 1017-1022.

[124] F.L. Sciacca, C. Ferri, F. Licastro, F. Veglia, I. Biunno, A. Gavazzi, E. Calabrese, F. Martinelli Boneschi, S. Sorbi, C. Mariani, M. Franceschi, L.M. Grimaldi. Interleukin-1B polymorphism is associated with age at onset of alzheimer's disease, Neurobiol. Aging 24 (2003) 927-931.

[125] T. Hirano, K. Yasukawa, H. Harada, T. Taga, Y. Watanabe, T. Matsuda, S. Kashiwamura, K. Nakajima, K. Koyama, A. Iwamatsu. Complementary DNA for a novel human interleukin (BSF-2) that induces B lymphocytes to produce immunoglobulin, Nature 324 (1986) 73-76.

[126] U. Dendorfer, P. Oettgen, T.A. Libermann. Multiple regulatory elements in the interleukin-6 gene mediate induction by prostaglandins, cyclic AMP, and lipopolysaccharide, Mol. Cell. Biol. 14 (1994) 4443-4454.

[127] A. Spooren, K. Kolmus, G. Laureys, R. Clinckers, J. De Keyser, G. Haegeman, S. Gerlo. Interleu- kin-6, a mental cytokine, Brain Res. Rev. 67 (2011) 157-183.

[128] R. Tehranian, H. Hasanvan, K. Iverfeldt, C. Post, M. Schultzberg. Early induction of interleukin-6 mRNA in the hippocampus and cortex of APPsw transgenic mice Tg2576, Neurosci. Lett. 301 (2001) 54-58.

[129] M. Huell, S. Strauss, B. Volk, M. Berger, J. Bauer. Interleukin-6 is present in early stages of plaque formation and is restricted to the brains of alzheimer's disease patients, Acta Neuropathol. 89 (1995) 544-551.

[130] J. Bauer, S. Strauss, U. Schreiter-Gasser, U. Ganter, P. Schlegel, I. Witt, B. Yolk, M. Berger. Inter- leukin-6 and alpha-2-macroglobulin indicate an acute-phase state in alzheimer's disease cortices, FEBS Lett. 285 (1991) 111-114.

[131] R. Del Bo, N. Angeretti, E. Lucca, M.G. De Simoni, G. Forloni. Reciprocal control of inflammatory cytokines, IL-1 and IL-6, and beta-amyloid production in cultures, Neurosci. Lett. 188 (1995) 70- 74.

[132] B. Brugg, Y.L. Dubreuil, G. Huber, E.E. Wollman, N. Delhaye-Bouchaud, J. Mariani. Inflammato- ry processes induce beta-amyloid precursor protein changes in mouse brain, Proc. Natl. Acad. Sci. U. S. A. 92 (1995) 3032-3035.

58 [133] R.A. Quintanilla, D.I. Orellana, C. Gonzalez-Billault, R.B. Maccioni. Interleukin-6 induces alz- heimer-type phosphorylation of tau protein by deregulating the cdk5/p35 pathway, Exp. Cell Res. 295 (2004) 245-257.

[134] M. van Oijen, P.P. Arp, F.J. de Jong, A. Hofman, P.J. Koudstaal, A.G. Uitterlinden, M.M. Breteler. Polymorphisms in the and transforming growth factor beta1 gene and risk of demen- tia. the rotterdam study, Neurosci. Lett. 402 (2006) 113-117.

[135] M.X. He, W.L. Yang, M.M. Zhang, Y.J. Lian, H.Y. Hua, J.S. Zeng, L.R. Zhang. Association be- tween interleukin-6 gene promoter -572C/G polymorphism and the risk of sporadic alzheimer's disease, Neurol. Sci. 31 (2010) 165-168.

[136] M. Wang, J. Jia. The interleukin-6 gene -572C/G promoter polymorphism modifies alzheimer's risk in APOE epsilon 4 carriers, Neurosci. Lett. 482 (2010) 260-263.

[137] J.P. Godbout, R.W. Johnson. Interleukin-6 in the aging brain, J. Neuroimmunol. 147 (2004) 141- 144.

[138] M.T. Heneka, M.K. O'Banion, D. Terwel, M.P. Kummer. Neuroinflammatory processes in alzhei- mer's disease, J. Neural Transm. 117 (2010) 919-947.

[139] P.L. McGeer, E. McGeer, J. Rogers, J. Sibley. Anti-inflammatory drugs and alzheimer disease, Lancet 335 (1990) 1037.

[140] J.C. Anthony, J.C. Breitner, P.P. Zandi, M.R. Meyer, I. Jurasova, M.C. Norton, S.V. Stone. Re- duced prevalence of AD in users of NSAIDs and H2 receptor antagonists: The cache county study, Neurology 54 (2000) 2066-2071.

[141] J.C. Breitner, K.A. Welsh, M.J. Helms, P.C. Gaskell, B.A. Gau, A.D. Roses, M.A. Pericak-Vance, A.M. Saunders. Delayed onset of alzheimer's disease with nonsteroidal anti-inflammatory and his- tamine H2 blocking drugs, Neurobiol. Aging 16 (1995) 523-530.

[142] B.A. in t'Veld, A. Ruitenberg, A. Hofman, L.J. Launer, C.M. van Duijn, T. Stijnen, M.M. Breteler, B.H. Stricker. Nonsteroidal antiinflammatory drugs and the risk of alzheimer's disease, N. Engl. J. Med. 345 (2001) 1515-1521.

[143] J. Rogers, L.C. Kirby, S.R. Hempelman, D.L. Berry, P.L. McGeer, A.W. Kaszniak, J. Zalinski, M. Cofield, L. Mansukhani, P. Willson. Clinical trial of indomethacin in alzheimer's disease, Neurol- ogy 43 (1993) 1609-1611.

[144] K.R. Zahs, K.H. Ashe. 'Too much good news' - are alzheimer mouse models trying to tell us how to prevent, not cure, alzheimer's disease? Trends Neurosci. 33 (2010) 381-389.

[145] I. Ray, A. Chauhan, H.M. Wisniewski, J. Wegiel, K.S. Kim, V.P. Chauhan. Binding of amyloid beta-protein to intracellular brain proteins in rat and human, Neurochem. Res. 23 (1998) 1277- 1282.

[146] S. Weggen, J.L. Eriksen, P. Das, S.A. Sagi, R. Wang, C.U. Pietrzik, K.A. Findlay, T.E. Smith, M.P. Murphy, T. Bulter, D.E. Kang, N. Marquez-Sterling, T.E. Golde, E.H. Koo. A subset of NSAIDs lower amyloidogenic Abeta42 independently of cyclooxygenase activity, Nature 414 (2001) 212- 216.

59 [147] J.M. Lehmann, J.M. Lenhard, B.B. Oliver, G.M. Ringold, S.A. Kliewer. Peroxisome proliferator- activated receptors alpha and gamma are activated by indomethacin and other non-steroidal anti- inflammatory drugs, J. Biol. Chem. 272 (1997) 3406-3410.

[148] Z.S. Khachaturian. Toward a comprehensive theory of alzheimer's disease--challenges, caveats, and parameters, Ann. N. Y. Acad. Sci. 924 (2000) 184-193.

[149] F.G. Gervais, D. Xu, G.S. Robertson, J.P. Vaillancourt, Y. Zhu, J. Huang, A. LeBlanc, D. Smith, M. Rigby, M.S. Shearman, E.E. Clarke, H. Zheng, L.H. Van Der Ploeg, S.C. Ruffolo, N.A. Thorn- berry, S. Xanthoudakis, R.J. Zamboni, S. Roy, D.W. Nicholson. Involvement of caspases in prote- olytic cleavage of alzheimer's amyloid-beta precursor protein and amyloidogenic A beta peptide formation, Cell 97 (1999) 395-406.

[150] E.M. Ribe, E. Serrano-Saiz, N. Akpan, C.M. Troy. Mechanisms of neuronal death in disease: Defining the models and the players, Biochem. J. 415 (2008) 165-182.

[151] S.J. Riedl, G.S. Salvesen. The apoptosome: Signalling platform of cell death, Nat. Rev. Mol. Cell Biol. 8 (2007) 405-413.

[152] S.M. Srinivasula, J.D. Ashwell. IAPs: What's in a name? Mol. Cell 30 (2008) 123-135.

[153] P.N. Pompl, S. Yemul, Z. Xiang, L. Ho, V. Haroutunian, D. Purohit, R. Mohs, G.M. Pasinetti. Caspase gene expression in the brain as a function of the clinical progression of alzheimer disease, Arch. Neurol. 60 (2003) 369-376.

[154] H. Guo, S. Albrecht, M. Bourdeau, T. Petzke, C. Bergeron, A.C. LeBlanc. Active caspase-6 and caspase-6-cleaved tau in neuropil threads, neuritic plaques, and neurofibrillary tangles of alzhei- mer's disease, Am. J. Pathol. 165 (2004) 523-531.

[155] S. Albrecht, M. Bourdeau, D. Bennett, E.J. Mufson, M. Bhattacharjee, A.C. LeBlanc. Activation of caspase-6 in aging and mild cognitive impairment, Am. J. Pathol. 170 (2007) 1200-1209.

[156] G. Klaiman, T.L. Petzke, J. Hammond, A.C. Leblanc. Targets of caspase-6 activity in human neurons and alzheimer disease, Mol. Cell. Proteomics 7 (2008) 1541-1555.

[157] J.W. Allen, B.A. Eldadah, X. Huang, S.M. Knoblach, A.I. Faden. Multiple caspases are involved in beta-amyloid-induced neuronal apoptosis, J. Neurosci. Res. 65 (2001) 45-53.

[158] J. Jo, D.J. Whitcomb, K.M. Olsen, T.L. Kerrigan, S.C. Lo, G. Bru-Mercier, B. Dickinson, S. Scul- lion, M. Sheng, G. Collingridge, K. Cho. Abeta(1-42) inhibition of LTP is mediated by a signaling pathway involving caspase-3, Akt1 and GSK-3beta, Nat. Neurosci. 14 (2011) 545-547.

[159] R.A. Rissman, W.W. Poon, M. Blurton-Jones, S. Oddo, R. Torp, M.P. Vitek, F.M. LaFerla, T.T. Rohn, C.W. Cotman. Caspase-cleavage of tau is an early event in alzheimer disease tangle pathol- ogy, J. Clin. Invest. 114 (2004) 121-130.

[160] T.W. Kim, W.H. Pettingell, Y.K. Jung, D.M. Kovacs, R.E. Tanzi. Alternative cleavage of alz- heimer-associated presenilins during apoptosis by a caspase-3 family protease, Science 277 (1997) 373-376.

[161] A.E. Roher, M.O. Chaney, Y.M. Kuo, S.D. Webster, W.B. Stine, L.J. Haverkamp, A.S. Woods, R.J. Cotter, J.M. Tuohy, G.A. Krafft, B.S. Bonnell, M.R. Emmerling. Morphology and toxicity of abeta-(1-42) dimer derived from neuritic and vascular amyloid deposits of alzheimer's disease, J. Biol. Chem. 271 (1996) 20631-20635.

60 [162] X.D. Pan, X.C. Chen, Y.G. Zhu, L.M. Chen, J. Zhang, T.W. Huang, Q.Y. Ye, H.P. Huang. Trip- chlorolide protects neuronal cells from microglia-mediated beta-amyloid neurotoxicity through in- hibiting NF-kappaB and JNK signaling, Glia 57 (2009) 1227-1238.

[163] Q. Wang, M.J. Rowan, R. Anwyl. Beta-amyloid-mediated inhibition of NMDA receptor-dependent long-term potentiation induction involves activation of microglia and stimulation of inducible ni- tric oxide synthase and superoxide, J. Neurosci. 24 (2004) 6049-6056.

[164] A.M. Minogue, A.W. Schmid, M.P. Fogarty, A.C. Moore, V.A. Campbell, C.E. Herron, M.A. Lynch. Activation of the c-jun N-terminal kinase signaling cascade mediates the effect of amyloid- beta on long term potentiation and cell death in hippocampus: A role for interleukin-1beta? J. Biol. Chem. 278 (2003) 27971-27980.

[165] M.A. Burguillos, T. Deierborg, E. Kavanagh, A. Persson, N. Hajji, A. Garcia-Quintanilla, J. Cano, P. Brundin, E. Englund, J.L. Venero, B. Joseph. Caspase signalling controls microglia activation and neurotoxicity, Nature 472 (2011) 319-324.

[166] I. Vermes, C. Haanen, H. Steffens-Nakken, C. Reutelingsperger. A novel assay for apoptosis. flow cytometric detection of phosphatidylserine expression on early apoptotic cells using fluorescein labelled annexin V, J. Immunol. Methods 184 (1995) 39-51.

[167] Y. Gavrieli, Y. Sherman, S.A. Ben-Sasson. Identification of programmed cell death in situ via specific labeling of nuclear DNA fragmentation, J. Cell Biol. 119 (1992) 493-501.

[168] A. Lord, H. Kalimo, C. Eckman, X.Q. Zhang, L. Lannfelt, L.N. Nilsson. The arctic alzheimer mutation facilitates early intraneuronal abeta aggregation and senile plaque formation in transgenic mice, Neurobiol. Aging 27 (2006) 67-77.

[169] O. Philipson, P. Hammarstrom, K.P. Nilsson, E. Portelius, T. Olofsson, M. Ingelsson, B.T. Hyman, K. Blennow, L. Lannfelt, H. Kalimo, L.N. Nilsson. A highly insoluble state of abeta similar to that of alzheimer's disease brain is found in arctic APP transgenic mice, Neurobiol. Aging 30 (2009) 1393-1405.

[170] A. Codita, A. Gumucio, L. Lannfelt, P. Gellerfors, B. Winblad, A.H. Mohammed, L.N. Nilsson. Impaired behavior of female tg-ArcSwe APP mice in the IntelliCage: A longitudinal study, Behav. Brain Res. 215 (2010) 83-94.

[171] J.L. Biedler, L. Helson, B.A. Spengler. Morphology and growth, tumorigenicity, and cytogenetics of human neuroblastoma cells in continuous culture, Cancer Res. 33 (1973) 2643-2652.

[172] J.L. Biedler, S. Roffler-Tarlov, M. Schachner, L.S. Freedman. Multiple neurotransmitter synthesis by human neuroblastoma cell lines and clones, Cancer Res. 38 (1978) 3751-3757.

[173] V. Ciccarone, B.A. Spengler, M.B. Meyers, J.L. Biedler, R.A. Ross. Phenotypic diversification in human neuroblastoma cells: Expression of distinct neural crest lineages, Cancer Res. 49 (1989) 219-225.

[174] S. Pahlman, A.I. Ruusala, L. Abrahamsson, M.E. Mattsson, T. Esscher. Retinoic acid-induced differentiation of cultured human neuroblastoma cells: A comparison with phorbolester-induced differentiation, Cell Differ. 14 (1984) 135-144.

[175] D.M. Walsh, I. Klyubin, J.V. Fadeeva, W.K. Cullen, R. Anwyl, M.S. Wolfe, M.J. Rowan, D.J. Selkoe. Naturally secreted oligomers of amyloid beta protein potently inhibit hippocampal long- term potentiation in vivo, Nature 416 (2002) 535-539.

61 [176] C. Haass, D.J. Selkoe. Soluble protein oligomers in neurodegeneration: Lessons from the alzhei- mer's amyloid beta-peptide, Nat. Rev. Mol. Cell Biol. 8 (2007) 101-112.

[177] M.Q. Liao, Y.J. Tzeng, L.Y. Chang, H.B. Huang, T.H. Lin, C.L. Chyan, Y.C. Chen. The correlation between neurotoxicity, aggregative ability and secondary structure studied by sequence truncated abeta peptides, FEBS Lett. 581 (2007) 1161-1165.

[178] T. Kubo, S. Nishimura, Y. Kumagae, I. Kaneko. In vivo conversion of racemized beta-amyloid ([D- ser 26]A beta 1-40) to truncated and toxic fragments ([D-ser 26]A beta 25-35/40) and fragment presence in the brains of alzheimer's patients, J. Neurosci. Res. 70 (2002) 474-483.

[179] D.B. Freir, D.A. Costello, C.E. Herron. A beta 25-35-induced depression of long-term potentiation in area CA1 in vivo and in vitro is attenuated by verapamil, J. Neurophysiol. 89 (2003) 3061-3069.

[180] C. Tohda, N. Matsumoto, K. Zou, M.R. Meselhy, K. Komatsu. Abeta(25-35)-induced memory impairment, axonal atrophy, and synaptic loss are ameliorated by M1, A metabolite of protopanax- adiol-type saponins, Neuropsychopharmacology 29 (2004) 860-868.

[181] K. Ayasolla, M. Khan, A.K. Singh, I. Singh. Inflammatory mediator and beta-amyloid (25-35)- induced ceramide generation and iNOS expression are inhibited by vitamin E, Free Radic. Biol. Med. 37 (2004) 325-338.

[182] R. Del Bo, N. Angeretti, E. Lucca, M.G. De Simoni, G. Forloni. Reciprocal control of inflammatory cytokines, IL-1 and IL-6, and beta-amyloid production in cultures, Neurosci. Lett. 188 (1995) 70- 74.

[183] K.N. Dahlgren, A.M. Manelli, W.B. Stine Jr, L.K. Baker, G.A. Krafft, M.J. LaDu. Oligomeric and fibrillar species of amyloid-beta peptides differentially affect neuronal viability, J. Biol. Chem. 277 (2002) 32046-32053.

[184] W.B. Stine Jr, K.N. Dahlgren, G.A. Krafft, M.J. LaDu. In vitro characterization of conditions for amyloid-beta peptide oligomerization and fibrillogenesis, J. Biol. Chem. 278 (2003) 11612-11622.

[185] J.A. White, A.M. Manelli, K.H. Holmberg, L.J. Van Eldik, M.J. Ladu. Differential effects of oli- gomeric and fibrillar amyloid-beta 1-42 on astrocyte-mediated inflammation, Neurobiol. Dis. 18 (2005) 459-465.

[186] R. Kayed, E. Head, J.L. Thompson, T.M. McIntire, S.C. Milton, C.W. Cotman, C.G. Glabe. Com- mon structure of soluble amyloid oligomers implies common mechanism of pathogenesis, Science 300 (2003) 486-489.

[187] M. Bucciantini, E. Giannoni, F. Chiti, F. Baroni, L. Formigli, J. Zurdo, N. Taddei, G. Ramponi, C.M. Dobson, M. Stefani. Inherent toxicity of aggregates implies a common mechanism for pro- tein misfolding diseases, Nature 416 (2002) 507-511.

[188] U. Soomets, M. Lindgren, X. Gallet, M. Hallbrink, A. Elmquist, L. Balaspiri, M. Zorko, M. Pooga, R. Brasseur, U. Langel. Deletion analogues of transportan, Biochim. Biophys. Acta 1467 (2000) 165-176.

[189] M. Pooga, M. Hallbrink, M. Zorko, U. Langel. Cell penetration by transportan, FASEB J. 12 (1998) 67-77.

62 [190] M.E. Bulina, D.M. Chudakov, O.V. Britanova, Y.G. Yanushevich, D.B. Staroverov, T.V. Chepurnykh, E.M. Merzlyak, M.A. Shkrob, S. Lukyanov, K.A. Lukyanov. A genetically encoded photosensitizer, Nat. Biotechnol. 24 (2006) 95-99.

[191] B.A. Griffin, S.R. Adams, R.Y. Tsien. Specific covalent labeling of recombinant protein molecules inside live cells, Science 281 (1998) 269-272.

[192] S.R. Adams, R.E. Campbell, L.A. Gross, B.R. Martin, G.K. Walkup, Y. Yao, J. Llopis, R.Y. Tsien. New biarsenical ligands and tetracysteine motifs for protein labeling in vitro and in vivo: Synthesis and biological applications, J. Am. Chem. Soc. 124 (2002) 6063-6076.

[193] S. Omura, Y. Iwai, A. Hirano, A. Nakagawa, J. Awaya, H. Tsuchya, Y. Takahashi, R. Masuma. A new alkaloid AM-2282 OF streptomyces origin. taxonomy, fermentation, isolation and preliminary characterization, J. Antibiot. (Tokyo) 30 (1977) 275-282.

[194] M.W. Karaman, S. Herrgard, D.K. Treiber, P. Gallant, C.E. Atteridge, B.T. Campbell, K.W. Chan, P. Ciceri, M.I. Davis, P.T. Edeen, R. Faraoni, M. Floyd, J.P. Hunt, D.J. Lockhart, Z.V. Milanov, M.J. Morrison, G. Pallares, H.K. Patel, S. Pritchard, L.M. Wodicka, P.P. Zarrinkar. A quantitative analysis of kinase inhibitor selectivity, Nat. Biotechnol. 26 (2008) 127-132.

[195] M.M. Garner, A. Revzin. A gel electrophoresis method for quantifying the binding of proteins to specific DNA regions: Application to components of the escherichia coli lactose operon regulatory system, Nucleic Acids Res. 9 (1981) 3047-3060.

[196] L. Holmlund, V. Cortes Toro, K. Iverfeldt. Additive effects of amyloid beta fragment and interleu- kin-1beta on interleukin-6 secretion in rat primary glial cultures, Int. J. Mol. Med. 10 (2002) 245- 250.

[197] L. Fisher, U. Soomets, V. Cortes Toro, L. Chilton, Y. Jiang, U. Langel, K. Iverfeldt. Cellular deliv- ery of a double-stranded oligonucleotide NFkappaB decoy by hybridization to complementary PNA linked to a cell-penetrating peptide, Gene Ther. 11 (2004) 1264-1272.

[198] C. Kunsch, S.M. Ruben, C.A. Rosen. Selection of optimal kappa B/Rel DNA-binding motifs: Interaction of both subunits of NF-kappa B with DNA is required for transcriptional activation, Mol. Cell. Biol. 12 (1992) 4412-4421.

[199] P. Lundin, H. Johansson, P. Guterstam, T. Holm, M. Hansen, U. Langel, S. EL Andaloussi. Distinct uptake routes of cell-penetrating peptide conjugates, Bioconjug. Chem. 19 (2008) 2535-2542.

[200] P. Saalik, K. Padari, A. Niinep, A. Lorents, M. Hansen, E. Jokitalo, U. Langel, M. Pooga. Protein delivery with transportans is mediated by caveolae rather than flotillin-dependent pathways, Bio- conjug. Chem. 20 (2009) 877-887.

[201] Y. Yamamoto, R.B. Gaynor. Therapeutic potential of inhibition of the NF-kappaB pathway in the treatment of inflammation and cancer, J. Clin. Invest. 107 (2001) 135-142.

[202] S. El-Andaloussi, H.J. Johansson, T. Holm, U. Langel. A novel cell-penetrating peptide, M918, for efficient delivery of proteins and peptide nucleic acids, Mol. Ther. 15 (2007) 1820-1826.

[203] S. El Andaloussi, T. Lehto, I. Mager, K. Rosenthal-Aizman, I.I. Oprea, O.E. Simonson, H. Sork, K. Ezzat, D.M. Copolovici, K. Kurrikoff, J.R. Viola, E.M. Zaghloul, R. Sillard, H.J. Johansson, F. Said Hassane, P. Guterstam, J. Suhorutsenko, P.M. Moreno, N. Oskolkov, J. Halldin, U. Tedebark, A. Metspalu, B. Lebleu, J. Lehtio, C.I. Smith, U. Langel. Design of a peptide-based vector, Pep- Fect6, for efficient delivery of siRNA in cell culture and systemically in vivo, Nucleic Acids Res. 39 (2011) 3972-3987.

63 [204] J.S. Wadia, R.V. Stan, S.F. Dowdy. Transducible TAT-HA fusogenic peptide enhances escape of TAT-fusion proteins after lipid raft macropinocytosis, Nat. Med. 10 (2004) 310-315.

[205] M. Mae, S. El Andaloussi, P. Lundin, N. Oskolkov, H.J. Johansson, P. Guterstam, U. Langel. A stearylated CPP for delivery of splice correcting oligonucleotides using a non-covalent co- incubation strategy, J. Control. Release 134 (2009) 221-227.

[206] E.M. Blalock, J.W. Geddes, K.C. Chen, N.M. Porter, W.R. Markesbery, P.W. Landfield. Incipient alzheimer's disease: Microarray correlation analyses reveal major transcriptional and tumor sup- pressor responses, Proc. Natl. Acad. Sci. U. S. A. 101 (2004) 2173-2178.

[207] B. Stein, P.C. Cogswell, A.S. Baldwin Jr. Functional and physical associations between NF-kappa B and C/EBP family members: A rel domain-bZIP interaction, Mol. Cell. Biol. 13 (1993) 3964- 3974.

[208] E.M. Blalock, J.W. Geddes, K.C. Chen, N.M. Porter, W.R. Markesbery, P.W. Landfield. Incipient alzheimer's disease: Microarray correlation analyses reveal major transcriptional and tumor sup- pressor responses, Proc. Natl. Acad. Sci. U. S. A. 101 (2004) 2173-2178.

[209] S. Prosch, A.K. Heine, H.D. Volk, D.H. Kruger. CCAAT/enhancer-binding proteins alpha and beta negatively influence the capacity of tumor necrosis factor alpha to up-regulate the human cyto- megalovirus IE1/2 enhancer/promoter by nuclear factor kappaB during monocyte differentiation, J. Biol. Chem. 276 (2001) 40712-40720.

[210] J.A. White, A.M. Manelli, K.H. Holmberg, L.J. Van Eldik, M.J. Ladu. Differential effects of oli- gomeric and fibrillar amyloid-beta 1-42 on astrocyte-mediated inflammation, Neurobiol. Dis. 18 (2005) 459-465.

[211] S. Jimenez, D. Baglietto-Vargas, C. Caballero, I. Moreno-Gonzalez, M. Torres, R. Sanchez-Varo, D. Ruano, M. Vizuete, A. Gutierrez, J. Vitorica. Inflammatory response in the hippocampus of PS1M146L/APP751SL mouse model of alzheimer's disease: Age-dependent switch in the micro- glial phenotype from alternative to classic, J. Neurosci. 28 (2008) 11650-11661.

[212] X.D. Pan, Y.G. Zhu, N. Lin, J. Zhang, Q.Y. Ye, H.P. Huang, X.C. Chen. Microglial phagocytosis induced by fibrillar beta-amyloid is attenuated by oligomeric beta-amyloid: Implications for alz- heimer's disease, Mol. Neurodegener 6 (2011) 45.

[213] C. Menard, P. Hein, A. Paquin, A. Savelson, X.M. Yang, D. Lederfein, F. Barnabe-Heider, A.A. Mir, E. Sterneck, A.C. Peterson, P.F. Johnson, C. Vinson, F.D. Miller. An essential role for a MEK-C/EBP pathway during growth factor-regulated cortical neurogenesis, Neuron 36 (2002) 597-610.

[214] A. Paquin, F. Barnabe-Heider, R. Kageyama, F.D. Miller. CCAAT/enhancer-binding protein phos- phorylation biases cortical precursors to generate neurons rather than astrocytes in vivo, J. Neuro- sci. 25 (2005) 10747-10758.

[215] M. Merhav, S. Kuulmann-Vander, A. Elkobi, S. Jacobson-Pick, A. Karni, K. Rosenblum. Behav- ioral interference and C/EBPbeta expression in the insular-cortex reveal a prolonged time period for taste memory consolidation, Learn. Mem. 13 (2006) 571-574.

[216] T. Tomiyama, S. Matsuyama, H. Iso, T. Umeda, H. Takuma, K. Ohnishi, K. Ishibashi, R. Teraoka, N. Sakama, T. Yamashita, K. Nishitsuji, K. Ito, H. Shimada, M.P. Lambert, W.L. Klein, H. Mori. A mouse model of amyloid beta oligomers: Their contribution to synaptic alteration, abnormal tau phosphorylation, glial activation, and neuronal loss in vivo, J. Neurosci. 30 (2010) 4845-4856.

64 [217] A. Nikolaev, T. McLaughlin, D.D. O'Leary, M. Tessier-Lavigne. APP binds DR6 to trigger axon pruning and neuron death via distinct caspases, Nature 457 (2009) 981-989.

[218] H. Guo, S. Albrecht, M. Bourdeau, T. Petzke, C. Bergeron, A.C. LeBlanc. Active caspase-6 and caspase-6-cleaved tau in neuropil threads, neuritic plaques, and neurofibrillary tangles of alzhei- mer's disease, Am. J. Pathol. 165 (2004) 523-531.

[219] M. D'Amelio, V. Cavallucci, S. Middei, C. Marchetti, S. Pacioni, A. Ferri, A. Diamantini, D. De Zio, P. Carrara, L. Battistini, S. Moreno, A. Bacci, M. Ammassari-Teule, H. Marie, F. Cecconi. Caspase-3 triggers early synaptic dysfunction in a mouse model of alzheimer's disease, Nat. Neu- rosci. 14 (2011) 69-76.

[220] A. Yuan, A. Kumar, C. Peterhoff, K. Duff, R.A. Nixon. Axonal transport rates in vivo are unaffect- ed by tau deletion or overexpression in mice, J. Neurosci. 28 (2008) 1682-1687.

[221] K. Li, Y. Li, J.M. Shelton, J.A. Richardson, E. Spencer, Z.J. Chen, X. Wang, R.S. Williams. Cyto- chrome c deficiency causes embryonic lethality and attenuates stress-induced apoptosis, Cell 101 (2000) 389-399.

[222] M.C. Wei, W.X. Zong, E.H. Cheng, T. Lindsten, V. Panoutsakopoulou, A.J. Ross, K.A. Roth, G.R. MacGregor, C.B. Thompson, S.J. Korsmeyer. Proapoptotic BAX and BAK: A requisite gateway to mitochondrial dysfunction and death, Science 292 (2001) 727-730.

[223] J. Tsai, J. Grutzendler, K. Duff, W.B. Gan. Fibrillar amyloid deposition leads to local synaptic abnormalities and breakage of neuronal branches, Nat. Neurosci. 7 (2004) 1181-1183.

[224] K.J. Ivins, E.T. Bui, C.W. Cotman. Beta-amyloid induces local neurite degeneration in cultured hippocampal neurons: Evidence for neuritic apoptosis, Neurobiol. Dis. 5 (1998) 365-378.

[225] M.S. Song, L. Saavedra, E.I. de Chaves. Apoptosis is secondary to non-apoptotic axonal degenera- tion in neurons exposed to abeta in distal axons, Neurobiol. Aging 27 (2006) 1224-1238.

[226] Z. Wei, M.S. Song, D. MacTavish, J.H. Jhamandas, S. Kar. Role of calpain and caspase in beta- amyloid-induced cell death in rat primary septal cultured neurons, Neuropharmacology 54 (2008) 721-733.

[227] M. Higuchi, M. Tomioka, J. Takano, K. Shirotani, N. Iwata, H. Masumoto, M. Maki, S. Itohara, T.C. Saido. Distinct mechanistic roles of calpain and caspase activation in neurodegeneration as revealed in mice overexpressing their specific inhibitors, J. Biol. Chem. 280 (2005) 15229-15237.

[228] P.W. Vanderklish, L.A. Krushel, B.H. Holst, J.A. Gally, K.L. Crossin, G.M. Edelman. Marking synaptic activity in dendritic spines with a calpain substrate exhibiting fluorescence resonance en- ergy transfer, Proc. Natl. Acad. Sci. U. S. A. 97 (2000) 2253-2258.

65