New developments in understanding and treating ... - CiteSeerX

7 downloads 0 Views 218KB Size Report
1 Cecilie Vogt Clinic for Neurology in the HELIOS Clinic Berlin-Buch, ... 10117 Berlin, Germany ...... Evangelou N, Esiri MM, Smith S, Palace J, Matthews PM.
Repository of the Max Delbrück Center for Molecular Medicine (MDC) Berlin (Germany) http://edoc.mdc-berlin.de/9668/

New developments in understanding and treating neuroinflammation C. Infante-Duarte, S. Waiczies, J. Wuerfel, and F. Zipp

Published in final edited form as:

Journal of Molecular Medicine. 2008 ; 86 (9): 975-985 | doi: 10.1007/s00109-007-0292-0 Springer (Germany) ►

Final Draft

New developments in understanding and treating neuroinflammation C. Infante-Duarte1,2, S. Waiczies1,2, J. Wuerfel1,2, and F. Zipp1,2 1

Cecilie Vogt Clinic for Neurology in the HELIOS Clinic Berlin-Buch, Charité–Universitaetsmedizin Berlin, Charitéplatz 1, 10117 Berlin, Germany 2 Max Delbrueck Center for Molecular Medicine, Berlin, Germany

ABSTRACT | We are currently witnesses to and authors of a paradigm shift in neuropathology. While classical acute and chronic neuroinflammatory diseases such as meningitis or multiple sclerosis (MS) present aspects of neurodegeneration, the disease course of progressive degenerative disorders such as Alzheimer’s disease (AD), Parkinson’s disease (PD), or stroke-mediated neuronal deficit are demonstrably affected by inflammation. These insights have immediate consequences both for research methods and for the development of novel, more efficient therapies for these diseases. In this review, we analyze the inflammatory and degenerative pathological mechanisms in the brain with particular emphasis on the classical chronic inflammatory disease MS. We demonstrate that the latest pathological considerations not only require the application of advanced research technologies to investigate new pathomechanistic pathways, but also affect the investigation, development, and monitoring of novel potential therapeutic tools. KEYWORDS | Multiple Sclerosis; Neuroinflammation; Neurodegeneration; T Cell.

Inflammation vs degeneration Traditionally, disorders of the central nervous system (CNS) have been categorized as inflammatory and noninflammatory. While multiple sclerosis (MS) represented the classical example of a chronic inflammatory disease, other CNS disorders such as Parkinson’s disease (PD) and Alzheimer’s disease (AD) were conventionally considered noninflammatory neurodegenerative processes. However, this classical paradigm has been reexamined [1], partly due to new findings, partly due to a reinterpretation of century-old data in light of new methodologies, and also due to a revision of the notion of neuroinflammation itself. For a very long time, neuroinflammation was considered to be a process accompanied by a huge invasion of the CNS by blood leukocytes [2]. According to this interpretation, CNS inflammation occurred exclusively in infectious processes of the CNS or under pathological conditions that promote blood–brain barrier (BBB) alterations and massive leukocyte infiltration, such as in MS. By contrast, the recent neuropathological concept has rediscovered longstanding investigations and redefines the presence of reactive microglia as a primary form of localized inflammation, which is a pathological hallmark of, among others, AD [3,4] and PD [5-7]. Also, supporting the inflammatory aspect of both disorders, it has been suggested that certain nonsteroidal antiinflammatory drugs (NAID) reduce disease risk [8-10]. However, recent studies show either no general effects of NAID in PD or AD [11,12] or a sex-related beneficial effect only in PD [13]. Nevertheless, it is becoming widely acknowledged that brain inflammation (and thus glia activation) is involved in the pathology of neurodegenerative disorders and should be therapeutically targeted [14]. Stroke is another example in which inflammation is now thought to exacerbate the tissue injury caused by the infarction. The interaction between the immune and nervous systems is, in the case of stroke, far from clear. It can involve many aspects, including protection, damage, or even neuroregeneration that make the development of therapeutic strategies particularly complex [15]. Similarly, the neurodegenerative aspects of acute and chronic inflammatory CNS disorders such as meningitis MDC Repository | http://edoc.mdc-berlin.de/9668/

and MS were in part disregarded for a long time. The impact and mechanisms of neuronal damage in meningitis have been extensively investigated in recent years [16,17]. By contrast, in MS, a disease that currently represents our primary research focus, evidence for neuronal damage was already provided more than a century ago [18]. The redefinition of MS pathology has immediate implications: it has now become evident that reliable monitoring of the disease course and accurate treatment strategies can only be guaranteed if both inflammatory and neurodegenerative aspects are considered and targeted [19].

The immune system and the CNS In the last few decades, it has become evident that the immune and nervous systems cannot be considered as two compartments, separated by a barrier, and not communicating. The interaction of immune cells and nervous cells is part of the general host defense and is essential in protecting the CNS from chronic infection and pathogen-induced damage. The normal CNS is continuously patrolled by activated, and possibly also inactivated, lymphocytes that, under “healthy” conditions, do not lead to inflammation or alter BBB integrity [20,21]. However, when those lymphocytes, in the context of a local infection or autoimmune process [22], reencounter their specific antigens in the CNS, which are probably presented by perivascular antigen-presenting cells [22,23], they may initiate a classical (auto)inflammatory response that promotes BBB disruption and the invasion of high numbers of activated leukocytes into CNS parenchyma [2]—both key manifestations of the earlier disease stage in MS and its animal model, experimental autoimmune encephalomyelitis (EAE) [24]. In this context, recent findings by our group and others prove T cells in the CNS to have highly dynamic properties [25,26]. MS has so far been considered as a demyelinating disorder of the CNS that principally affects young adults and is twice or in some countries more than three times more common in females than in males [27]. However, the etiology of MS is still unclear. Epidemiological data suggest that while genetic factors appear to determine 1

Infante-Duarte C et al. disease susceptibility [28], environmental factors, such as exposure to sunlight or to certain infectious agents in early life, determine disease development [29,30]. MS is thought to be a T cell-mediated autoimmune disease [31]. The role of inflammatory autoreactive CD4-positive T helper (Th) cells, producing cytokines such as IFNgamma and IL-17, has been extensively proven in models of EAE [32-35]. In patients, evidence for the involvement of T cells and proinflammatory cytokines in disease pathogenesis has come mostly from studies of lesions [36-39]. Furthermore, the “accidental” T cell response and consequent disease exacerbation observed in patients treated with an altered peptide ligand confirmed the active participation of T cells in human disease development [40]. According to the autoimmune theory of MS, it is believed that myelin-specific T cells are responsible for the initiation, coordination, and perhaps perpetuation of the autoimmune reaction, leading to oligodendrocyte damage and demyelination [18]. Although the concept of autoimmunity as a primary trigger in lesion formation is still a matter of discussion, it is now widely accepted that immune-mediated inflammation contributes to MS pathogenesis. This fact has recently been corroborated by the promising therapeutic effect in MS of two novel drugs: firstly, the alpha4 integrin antagonist natalizumab, which inhibits the migration of activated lymphocytes and monocytes into the CNS parenchyma [41-43]; and secondly, the sphingosine 1-phosphate (S1P) receptor modulator fingolimod, which captures lymphocytes within the lymph nodes and inhibits recirculation [44].

Inflammatory neurodegeneration The most representative forms of MS are: relapsing– remitting MS (RRMS) in which acute attacks are followed by complete or partial recovery and primary progressive MS (PPMS), which is characterized by disease progression from onset. More than 80% of patients show a relapsing–remitting course at the beginning of the disease, which, in the majority of the cases, converts to a progressive disease course—secondary progressive MS (SPMS)—after 10–25 years [45]. Although inflammation, demyelination, and damage to oligodendrocytes may account for the relapsing course, they fail to explain the clinical features of the progressive phase that are more similar to those observed in neurodegenerative diseases [18]. Thus, the neglected feature of neuronal injury in MS has gradually gained credence during the past few years. However, neurodegeneration, manifested in the form of axonal transection and neuronal loss, is not simply a hallmark of the progressive phase of MS. Axonal transection has been described in the early disease phase and has been shown to correlate with inflammation and to contribute to disability [46-51]. On the other hand, neuronal apoptosis in the cerebral cortex has also been described early in the course of the disease [52-54] and has been shown to correlate with cognitive deficits [55]. Although the extent of gray matter pathology—which includes loss of neurons and demyelination, axonal transection, and dendritic transaction—was suggested to account for clinical disability, especially in SPMS patients [56], its contribution to the disease is still largely underestimated. If MS is considered to be caused by demyelination and neuronal damage, the obvious question emerges as to the cause of neurodegeneration. Axonal injury has been linked to demyelination (Wallerian degeneration) [57]. However, numerous other reports show evidence for additional demyelination-independent MDC Repository | http://edoc.mdc-berlin.de/9668/

injury of axons [51,58,59] that seems to generate more damage than demyelination alone. Both reversible demyelination and irreversible axonal loss are occurrences associated with the earlier disease course. Therefore, it is essential to develop therapies that not only aim at preventing demyelination or inducing remyelination, but also at preventing axonal damage (Fig.1). So far, it has been reported that axonal loss might be triggered by deprivation of oligodendrocyte-derived trophic factors [60], by nitric oxide (NO)-mediated disruption of axonal conduction [61], or by deficient mitochondria that are unable to deliver the increased amount of energy demanded in chronically demyelinating axons [62]. This brings us to the question of whether axonal pathology is responsible for the loss of neurons observed in MS. It has been shown that damage of axons might promote neuronal death (retrograde degeneration). However, the fact that gray matter pathology was observed early in the disease, in the absence of white matter involvement, points to the occurrence of primary neuronal loss independently of demyelination or axonal damage as well [63] (Fig.1). One of the main goals of our ongoing research is to elucidate the mechanisms involved in neuronal apoptosis. Using animal models, we recently demonstrated that CD4-positive T helper cells are able to cover long distances in the brain to interact with neurons and to induce their death [25]. Because neurons do not express MHC class II molecules and neuronal damage has been reported to be antigen-independent [25,64], we believe that CD4-positive T cells might damage neurons in a bystander fashion [1,18]. In addition, our current data show that non-CNS specific CD4-positive T cells have the capacity not only to enter the CNS but also to promote alterations in the permeability of cerebral blood vessels without promoting any neuronal pathology [65]. Based on these data, we speculate that in a compromised brain, such as in MS [66], perturbations of the BBB caused by nonspecific effector T cells may have further pathological consequences. Currently, we are investigating the role of those unspecific trigger mechanisms, which might support damage in the inflamed brain in vivo. Furthermore, we demonstrated that, in mouse, the tumor necrosis factor (TNF)-related apoptosis-inducing ligand (TRAIL), a death ligand secreted by or expressed on T cells, is a specific death-inducing effector molecule in the brain, which triggers bystander apoptosis of neurons in the inflamed CNS in vivo and promotes EAE [67]. Apart from CD4positive T cells, CD8-positive cytotoxic T cells potentially cause neuronal death in the CNS. In contrast to MHC class II, the expression of MHC class I seems to be upregulated on neurons under inflammatory conditions. Therefore, an antigen-dependent fatal interaction of CD8positive cells and the target neurons is conceivable [68]. Independently of the killing activity of T lymphocytes, it has been suggested that activated microglia detected close to neurons may be involved in neuronal death [53]. In this context, we have recently defined a possible effector molecule, 7-ketocholesterol (7-KC), which might provide a link between demyelination, activation of microglia, and subsequent neuronal loss. 7-KC is a breakdown product of the oxidation of myelin lipids. On the one hand, we showed that 7-KC is present in the CSF and CNS of MS patients and correlates with disability in EAE, and on the other hand, we demonstrated that 7-KC is able to activate microglia and to indirectly promote microglia-mediated neuronal death [69].

2

Infante-Duarte C et al. Future studies will elucidate which processes are fundamental to the CNS, and thus shared by the MS pathology and other diseases such as stroke and classical neurodegeneration [1]. There is already evidence that the death ligand TRAIL [70], inflammatory cytokines such as interleukin-1 (IL-1) [71], and free radicals such as superoxide and NO [72] play a central role in different CNS pathologies. NO is produced in increased amounts in damaged mitochondria and, at the same time, impairs mitochondrial function [73]. NO, detected in brains of patients with neurodegeneration and MS, has been suggested to contribute to ongoing neuronal and oligodendroglial death, as it inhibits mitochondrial respiration, probably leading to intracellular 2+ accumulation of Ca [62].

Newly approved and experimental therapies in MS We, like many other research groups, are focusing our efforts on establishing the optimal strategy for treating MS patients. Our current understanding of the pathogenesis outlined above in which neuroinflammation and neurodegeneration may occur in parallel and in series indicate that therapies should ideally target both processes simultaneously (Fig.1). Drugs currently available for the treatment of MS reduce the risk of relapse, but only slightly prolong the progression of disability. The main treatment strategies tackled so far primarily involve measures to prevent damage from occurring by local inflammation. Clinical approaches initially involved the application of drugs, which unspecifically suppress the immune system (immunosuppressants) such as the glucocorticoids and the antineoplastic agent mitoxanthrone, both of which reduce the progression of disease but are associated with harmful side effects. Eventually, drugs that modulate the immune system (immunomodulators) in a pleiotropic manner, including the interferon beta (IFN-beta) group and glatiramer acetate (GA, a short synthetic peptide mix mimicking MBP), were also shown to subdue local inflammation and reduce the relapse rate in MS. These disease-modifying agents (DMA) have been approved during the past 10 years for the treatment of MS [74]. Other promising drugs acting similarly and, possibly, more efficiently (such as statins or flavanoids, which will be discussed further on) are still awaiting clinical trials for efficacy and safety testing. To enhance the efficacy of currently approved DMA (that have undergone thorough postmarketing surveillance), it may also be beneficial to apply add-on treatments to base DMA.

Newly approved and forthcoming therapies Neuroinflammation in MS is dominated by the infiltration of activated autoreactive T cells into the CNS. Indeed, most of the new pharmacological agents currently being introduced as therapy options for MS, such as natalizumab and fingolimod (FTY720), specifically target T cell migration. Fingolimod is a derivative of myriocin (a constituent of the fungus Isaria sinclairii) and mimics the activity of the phospholipid S1P [75]. The beneficial effect of the oral administration of fingolimod to MS patients was recently shown in a phase II clinical trial [44]. The salutary MDC Repository | http://edoc.mdc-berlin.de/9668/

mechanisms seem to be the sequestration of newly generated autoreactive T cells in the medullary sinuses of lymph nodes [76] and the ligation to S1P1 on endothelial cells, which results in Rac-dependent tightening of endothelial-cell junctions [77]. In addition, fingolimod binds with highest affinity to S1P5 receptors [75], which, in the CNS, are exclusively expressed on oligodendrocytes. The S1P5 receptor is involved in oligodendrocyte survival [78] and in the inhibition of the motility of oligodendrocyte precursors, creating a possible setback in the remyelination process [79]. S1P is not only involved in lymphocyte trafficking and oligodendrocyte mobilization but also in other physiological events, such as vascular tone, heart rate regulation, and contraction of airway smooth muscle cells [80]. Thus, apart from the observed lymphopenia in patients on fingolimod [44], a higher incidence of adverse events in patients receiving the agonist is to be expected [44]. Notwithstanding these uncertainties, the results of the preliminary study by Kappos et al. provide us with a strong incentive for long-term follow-up trials on a large scale. Indeed, this treatment strategy is currently being checked in a multicentered placebo-controlled phase III clinical trial in relapsing–remitting MS. Natalizumab, a humanized recombinant monoclonal antibody (mAb) that binds to the alpha4 chain of the alpha4beta1(VLA-4) and alpha4beta7 integrins, is a selective adhesion molecule inhibitor that has already shown impressive results in RRMS patients by dramatically reducing infiltration lymphocytes in the CSF [41,43,81]. Recently, natalizumab has also been shown to robustly prevent new lesion formation as shown by MRI analysis [82]. Apart from natalizumab, various new mAb therapies have been developed to deplete harmful immune cell populations in the periphery. These therapies include alemtuzumab, targeting CD52, which is found on mononuclear cells, and rituximab, targeting CD20 and thereby specifically depleting premature and mature B cells. Alemtuzumab (Campath-1H) was originally tested in SPMS in a phase I clinical trial, and although results were initially encouraging [83, 84], in the long-term, only a minority of patients benefited. Alemtuzumab did, however, markedly reduce the relapse rate and EDSS score over a period of 2 years in RRMS patients [85] and a phase II study in RRMS showed dose-dependent reduction in the risk for relapse and for progression of disability (reduction of 87% and 66%, respectively, at the highest dose) compared with the control group treated with Rebif (data presented by Coles et al. at the 59th Annual Meeting of the American Academy of Neurology, 2007, Abstract S12.004). Similarly, preliminary results from an ongoing phase II/III clinical trial in PPMS with rituximab only revealed a temporary suppression of B cell activation [86]. Rituximab was, however, shown to be beneficial in neuromyelitis optica [87], which in view of recent data on the occurrence of aquaporin-4 antibodies, can be considered a distinct disease entity [88]. As yet unpublished results from a phase II trial with RRMS patients indicate that rituximab reduces MRI lesion load by 90% and relapse rate by 58% (data presented by Hauser et al. at the 59th Annual Meeting of the American Academy of Neurology, 2007, Abstract S12.003). Overall, selective therapies including the mAb therapies represent a new stage in MS therapy. However, even these promising therapies may have their setbacks. For instance, the rare appearance of progressive multifocal leukoencephalopathy (PML) after natalizumab [89] and 3

Infante-Duarte C et al. Graves’ disease or idiopathic thrombocytopenic purpura after alemtuzumab [90] cautions us to apply each therapeutic agent, especially such selective agents with extreme care and consideration to each individual patient [91]. One further difficulty with drugs that selectively target one molecule is a situation of redundancy where the organism will find a means of compensating for deleted targeted protein or cell. This might be one reason why the CCR1 blockade did not show any significant beneficial effect in a phase II clinical trial in patients with RRMS [92], although on the basis of data obtained in animal models [93-95], this therapy had been considered very promising. Therefore, there is an emerging school of thought which argues that complex and heterogeneous diseases such as MS would better targeted with “dirty” promiscuous drugs rather than selective therapies [96].This leaves us with the still open question as to whether we should further characterize the partially effective pleiotropic drugs, identifying their “pitfalls” in the process, or whether we should concentrate on identifying the many targets significant in the pathology of complex diseases such as MS to design new “multiple ligands.”

Experimental therapies Apart from the development of selective therapies that target one or more molecules, another therapeutic goal in patient health care is to identify therapeutic agents that may be administered via a noninvasive, i.e., oral, route to enhance patient compliance. Examples range from pilot trials to phase III studies.

Flavonoids: anti-inflammatory and neuroprotective Flavonoids, including EGCG (epigallocatechin-3-gallate), are polyphenolic plant metabolites that have been shown to have versatile biological activities, including antimicrobial [97] and antitumor properties [98-100]. While it is clear that they promote apoptosis and cell cycle arrest in transformed cells [101-104], polyphenols such as EGCG have also been shown to protect nontransformed cells from apoptosis [105,106]. Indeed, this substance, which exhibits several mechanisms of action, has been shown to have protective effects in autoimmune animal models. It reduced disease severity in collagen-induced arthritis by significantly lowering the levels of inflammatory cytokines and mediators [107]. We reported that EGCG is capable of targeting both inflammation and direct neuronal injury in EAE: the antioxidative properties of EGCG directly protected living brain tissue against neuronal injury by N-methyl-D-aspartate (NMDA) or the death ligand TRAIL, and blocked the formation of neurotoxic reactive oxygen species in neurons [108]. Neuroprotective effects by polyphenols have also been reported in neurodegenerative diseases such as PD [109,110] and AD [111] or transient global ischemia [112]. Our recent findings [108], and work by Hendriks et al. [113], suggest that flavonoids may be therapeutic candidates for the treatment of MS due to their versatility in combining antiinflammatory and neuroprotective properties. Phase II trials, including one conducted by us, are scheduled to begin shortly.

MDC Repository | http://edoc.mdc-berlin.de/9668/

HMG CoA reductase inhibitors: beyond lipid lowering The β-3-hydroxy-3-methylglutaryl (HMG)-CoA reductase inhibitors (HMGCRI), known as statins, have been used for several years as the gold standard for lipid-lowering therapy. These drugs were initially designed to lower cholesterol levels by specifically inhibiting the catalytic conversion of HMG-CoA to L-mevalonate, which is a precursor to cholesterol and also to intermediary lipid donors such as farnesyl pyrophosphate (Fpp) or geranylgeranyl pyrophosphate (GGpp). These isoprenoids are important for the posttranslational modification or isoprenylation of GTP-binding proteins of the Ras superfamily [114]. These include members such as the prototype Ras and Rap and also the Ras homologue (Rho) family members, including Rho, Rac, and Cdc42, all of which are important for signaling in inflammatory processes. Indeed, the outcome of manipulating cholesterol and isoprenoid synthesis with HMGCRI, and thereby intracellular signaling, has been studied in brain pathology, including neuroinflammation [115]. Originally, various groups, including ours, showed that the HMGCRI atorvastatin is therapeutically effective in EAE [116,117], the salutary effects probably being due either to regulatory mechanisms preventing the expansion of harmful autoreactive cells already outside the CNS compartment [118] or to the blocking of leukocytes from leaving the BBB vasculature into the CNS [116,119]. Evidence to support the latter scenario is that HMGCRI reduce levels of chemokine receptors, T cell adhesion molecules, and matrix metalloproteinase (MMP)-9 by peripheral immune cells derived from MS patients [120] and inhibit the migration of these cells through BBB-derived endothelial cells [121]. Furthermore, for statins, which are bioavailable in the brain, the antioxidative properties of these drugs [122] might hold the key to preventing cholesterol oxidation and formation of harmful oxysterol that induces microglia-mediated neuronal damage [69], thus providing further rationale for the attenuating neuroinflammatory processes in the EAE. The beneficial effect of HMGCRI in the EAE, which readily diffuse into the CNS compartment, has also been linked to the enhanced survival of oligodendrocyte progenitor cells and their differentiation to remyelinating oligodendrocytes in vivo [123]. Thus, HMGCRI represent another versatile tool, which might target both neuroinflammation and direct neuronal damage simultaneously. This has already led to positive outcomes in MS treatment in two small open-label trials, which were not free of criticism due to their size and design [124,125]. These promising data are currently undergoing evaluation in larger-scale controlled clinical trials, including our own.

New techniques inflammation

to

monitor

neuro-

Our understanding of the interactions between immune cells and cells of the CNS has been considerably improved in the last few years. However, the data obtained so far have resulted from in vitro models or from the ex vivo evaluation of CNS tissue, which can lead to more conflict and controversy than clarity regarding the in vivo situation. Nevertheless, the new era of methodological approaches to molecular imaging has the potential to elucidate the dynamics of the pathological processes involved. Molecular imaging is a noninvasive methodology that combines molecular and cell biology tools for the in vivo imaging of biological processes such 4

Infante-Duarte C et al. as location and degree of gene and protein expression, signal transduction pathways, cell migration patterns, and cellular interactions at the site of pathology in the CNS. Molecular imaging technologies, such as MRI, positron emission tomography (PET), multiphoton microscopy (MPM), bioluminescence, or fluorescence, are applied broadly across the life sciences, from stem cell transplantation research, through cancer investigation, to the treatment of CNS diseases. As far as we are concerned, the necessity to develop these techniques is of utmost importance because the brain is not as accessible as other living organs. Multiphoton microscopy is a relatively novel imaging technique. It is based on the probability that two or more low-energy photons are absorbed quasisimultaneously by a fluorophore, exciting it to the same level as one high energy photon. Multiphoton excitation was first theoretically predicted by the German physicist Maria Göppert-Mayer more than 75 years ago [126], but its application in microscopy in the biological sciences would not come until much later, and was first described by Denk et al. in 1990 [127]. The principal advantages of using two or more photons vs one are, on the one hand, the reduced phototoxicity caused in the specimen by lowenergy wavelengths (especially important in living tissue) and, on the other hand, the fact that low energy penetrates deeper into the tissue. In addition, multiphoton excitation occurs only at the focal plane and does not affect the surrounding tissue, instead limiting photobleaching and damage to the imaged area [128]. As mentioned above, we have used MPM to investigate the activity of T cells in living brain tissue [25] and to determine the behavior of nonneural specific T cells within the CNS in vivo [65]. Instances of the use of live imaging in neurology so far include the investigation of neuronal activity [129], even in freely moving animals [130], and of microglia function [131] and the monitoring of T cell behaviors in neuroinflammation [26]. Intravital MPM can be expected to provide an answer to our many questions regarding the role of different T cell types during CNS inflammation, the mechanisms driving T cell trafficking within the CNS, and the ways in which T cells interact with CNS cells. Another approach, which may be transferred to the clinic more rapidly, is MR cellular and molecular imaging. MRI is a technique providing high spatial, temporal, and spectral information on anatomy, metabolism, and tissue function in vivo. Due to its noninvasive nature and the fact that it is virtually free of side effects, MRI has already become one of the most important and most frequently used imaging tools in medicine. Today, high magnetic field strengths and dedicated rodent scanners allow for sufficient spatial resolution in vivo, even in murine animal models, such as EAE mice. Novel MRI diffusion imaging techniques aim at differentiating between axonal damage and neurodegeneration [132]. Recent advances in MR sequences have been applied to enhance the image resolution or signal to noise ratio [133]. Using new contrast agents, we recently showed very early BBB breakdown preceding the clinical onset of disease in EAE, thereby mimicking multiple sclerosis [134]. One of these agents, Gadofluorine M, was also successfully used to detect peripheral nerve degeneration [135]. Another class of contrast agents, superparamagnetic iron oxide particles, is now widely employed in labeling phagocytic cells in MRI [136]. A novel very small iron oxide particle (VSOP) currently under our investigation is being used to MDC Repository | http://edoc.mdc-berlin.de/9668/

label various cell types, such as macrophages or T cells ex vivo, which are subsequently transferred to an animal and can be traced longitudinally in vivo. It is interesting to note that VSOP also binds specifically to proteoglycane structures, which appear to be extensively deposited in active inflammatory lesions, providing a unique tool for the molecular targeting of extracellular matrix alterations and basal membrane damage. By applying MR spectroscopy, magnetization transfer imaging or diffusion tensor techniques, even occult tissue damage and reduced nerve fiber integrity can be quantified in vivo [133,137]. In the future, novel MR contrast techniques applying hyperpolarized nuclei such as C-13 will be able to provide metabolic information regarding, e.g., glykogen or pyrovate metabolism [138]. The latest approach to developing multimodal imaging devices, such as integrated MRI–PET scanners or combined immunofluorescence and MR contrast markers, will help us to bridge the gap between the macroscopic, microscopic, and molecular levels, thus completing our pathophysiological understanding by supplying essential information, which has so far eluded us.

Postscript The Cecilie Vogt Clinic is a center for translational neurology. One of its focuses is the investigation of inflammatory disorders of the brain such as MS. This disease is a major cause of severe disability in young adults in the Western world, and consequently has a serious socioeconomic impact. Our clinic aims not only to clarify disease pathophysiology using state-of-the-art technology, but also to apply this knowledge to the development of effective, safer, more efficient therapies, and reliable markers for diagnosis, prognosis, and therapy response. We take basic findings from the bench and apply them directly to the patient’s care. This allows for an optimal coordination of theoretical and clinical knowledge, leading to new concepts in neurological treatment and providing the opportunity for patients to take part in clinical studies with novel therapeutic tactics.

Acknowledgements We gratefully acknowledge the funding from the Institut für MS-Forschung (IFMS) Göttingen/Gemeinnützige HertieStiftung (GHA), the Deutsche Forschungsgemeinschaft (DFG), the Bundesministerium für Bildung und Forschung (BMBF), and the Charité–Universitätsmedizin Berlin. We thank Andrew Mason for reading the manuscript as a native English-speaker, and Volker Siffrin and Ulf Schulze-Topphoff for the helpful discussions.

Corresponding Author Frauke Zipp, [email protected]

5

Infante-Duarte C et al.

References 1.

Zipp F, Aktas O (2006) The brain as a target of inflammation: common pathways link inflammatory and neurodegenerative diseases. Trends Neurosci 29:518–527.

2.

Hickey WF (1991) Migration of hematogenous cells through the blood–brain barrier and the initiation of CNS inflammation. Brain Pathol 1:97–105.

3.

McGeer PL, Itagaki S, Tago H, McGeer EG (1987) Reactive microglia in patients with senile dementia of the Alzheimer type are positive for the histocompatibility glycoprotein HLADR. Neurosci Lett 79:195–200.

4.

Rogers J, Luber-Narod J, Styren SD, Civin WH (1988) Expression of immune system-associated antigens by cells of the human central nervous system: relationship to the pathology of Alzheimer’s disease. Neurobiol Aging 9:339– 349.

5.

McGeer PL, Itagaki S, Boyes BE, McGeer EG (1988) Reactive microglia are positive for HLA-DR in the substantia nigra of Parkinson’s and Alzheimer’s disease brains. Neurology 38:1285–1291.

6.

Langston JW, Forno LS, Tetrud J, Reeves AG, Kaplan JA, Karluk D (1999) Evidence of active nerve cell degeneration in the substantia nigra of humans years after 1-methyl-4phenyl-1,2,3,6-tetrahydropyridine exposure. Ann Neurol 46:598–605.

7.

Imamura K, Hishikawa N, Sawada M, Nagatsu T, Yoshida M, Hashizume Y (2003) Distribution of major histocompatibility complex class II-positive microglia and cytokine profile of Parkinson’s disease brains. Acta Neuropathol (Berl) 106:518–526.

8.

McGeer PL, Rogers J, McGeer EG (2006) Inflammation, anti-inflammatory agents and Alzheimer disease: the last 12 years. J Alzheimers Dis 9:271–276.

9.

Chen H, Jacobs E, Schwarzschild MA, McCullough ML, Calle EE, Thun MJ, Ascherio A (2005) Nonsteroidal antiinflammatory drug use and the risk for Parkinson’s disease. Ann Neurol 58:963–967.

immune response in bacterial meningitis. J Clin Invest 117:2004–2013. 18. Aktas O, Ullrich O, Infante-Duarte C, Nitsch R, Zipp F (2007) Neuronal damage in brain inflammation. Arch Neurol 64:185–189. 19. Aktas O, Waiczies S, Zipp F (2007) Neurodegeneration in autoimmune demyelination: recent mechanistic insights reveal novel therapeutic targets. J Neuroimmunol 184:17– 26. 20. Hickey WF (2001) Basic principles of immunological surveillance of the normal central nervous system. Glia 36:118–124. 21. Brabb T, von Dassow P, Ordonez N, Schnabel B, Duke B, Goverman J (2000) In situ tolerance within the central nervous system as a mechanism for preventing autoimmunity. J Exp Med 192:871–880. 22. Hickey WF, Kimura H (1988) Perivascular microglial cells of the CNS are bone marrow-derived and present antigen in vivo. Science 239:290–292. 23. Becher B, Bechmann I, Greter M (2006) Antigen presentation in autoimmunity and CNS inflammation: how T lymphocytes recognize the brain. J Mol Med 84:532–543. 24. Kermode AG, Thompson AJ, Tofts P, MacManus DG, Kendall BE, Kingsley DP, Moseley IF, Rudge P, McDonald WI (1990) Breakdown of the blood–brain barrier precedes symptoms and other MRI signs of new lesions in multiple sclerosis. Pathogenetic and clinical implications. Brain 113(Pt 5):1477–1489. 25. Nitsch R, Pohl EE, Smorodchenko A, Infante-Duarte C, Aktas O, Zipp F (2004) Direct impact of T cells on neurons revealed by two-photon microscopy in living brain tissue. J Neurosci 24:2458–2464. 26. Kawakami N, Nagerl UV, Odoardi H, Flugel A (2005) Live imaging and autoantigen recognition autoimmune encephalomyelitis 201:1805–1814.

F, Bonhoeffer T, Wekerle of effector cell trafficking within the unfolding lesion. J Exp Med

10. Chen H, Zhang SM, Hernan MA, Schwarzschild MA, Willett WC, Colditz GA, Speizer FE, Ascherio A (2003) Nonsteroidal anti-inflammatory drugs and the risk of Parkinson disease. Arch Neurol 60:1059–1064.

27. Orton SM, Herrera BM, Yee IM, Valdar W, Ramagopalan SV, Sadovnick AD, Ebers GC (2006) Sex ratio of multiple sclerosis in Canada: a longitudinal study. Lancet Neurol 5:932–936.

11. Hancock DB, Martin ER, Stajich JM, Jewett R, Stacy MA, Scott BL, Vance JM, Scott WK (2007) Smoking, caffeine, and nonsteroidal anti-inflammatory drugs in families with Parkinson disease. Arch Neurol 64:576–580.

28. Kantarci O, Wingerchuk D (2006) Epidemiology and natural history of multiple sclerosis: new insights. Curr Opin Neurol 19:248–254.

12. Group AR, Lyketsos CG, Breitner JC, Green RC, Martin BK, Meinert C, Piantadosi S, Sabbagh M (2007) Naproxen and celecoxib do not prevent AD in early results from a randomized controlled trial. Neurology 68:1800–1808. 13. Hernan MA, Logroscino G, Garcia Rodriguez LA (2006) Nonsteroidal anti-inflammatory drugs and the incidence of Parkinson disease. Neurology 66:1097–1099.

29. Marrie RA (2004) Environmental risk factors in multiple sclerosis aetiology. Lancet Neurol 3:709–718. 30. Giovannoni G, Cutter GR, Lunemann J, Martin R, Munz C, Sriram S, Steiner I, Hammerschlag MR, Gaydos CA (2006) Infectious causes of multiple sclerosis. Lancet Neurol 5:887–894. 31. Sospedra M, Martin R (2005) Immunology of multiple sclerosis. Annu Rev Immunol 23:683–747.

14. Block ML, Zecca L, Hong JS (2007) Microglia-mediated neurotoxicity: uncovering the molecular mechanisms. Nat Rev Neurosci 8:57–69.

32. Zamvil SS, Steinman L (1990) The T lymphocyte in experimental allergic encephalomyelitis. Annu Rev Immunol 8:579–621.

15. Dirnagl U, Klehmet J, Braun JS, Harms H, Meisel C, Ziemssen T, Prass K, Meisel A (2007) Stroke-induced immunodepression: experimental evidence and clinical relevance. Stroke 38:770–773.

33. Langrish CL, Chen Y, Blumenschein WM, Mattson J, Basham B, Sedgwick JD, McClanahan T, Kastelein RA, Cua DJ (2005) IL-23 drives a pathogenic T cell population that induces autoimmune inflammation. J Exp Med 201:233– 240.

16. Weber JR, Tuomanen EI (2007) Cellular damage in bacterial meningitis: an interplay of bacterial and host driven toxicity. J Neuroimmunol 184:45–52. 17. Hoffmann O, Priller J, Prozorovski T, Schulze-Topphoff U, Baeva N, Lunemann JD, Aktas O, Mahrhofer C, Stricker S, Zipp F, Weber JR (2007) TRAIL limits excessive host

MDC Repository | http://edoc.mdc-berlin.de/9668/

34. Ivanov II, McKenzie BS, Zhou L, Tadokoro CE, Lepelley A, Lafaille JJ, Cua DJ, Littman DR (2006) The orphan nuclear receptor RORgammat directs the differentiation program of proinflammatory IL-17+ T helper cells. Cell 126:1121–1133.

6

Infante-Duarte C et al. 35. Hofstetter HH, Toyka KV, Tary-Lehmann M, Lehmann PV (2007) Kinetics and organ distribution of IL-17-producing CD4 cells in proteolipid protein 139–151 peptide-induced experimental autoimmune encephalomyelitis of SJL mice. J Immunol 178:1372–1378.

50. Evangelou N, Esiri MM, Smith S, Palace J, Matthews PM (2000) Quantitative pathological evidence for axonal loss in normal appearing white matter in multiple sclerosis. Ann Neurol 47:391–395.

36. Nyland H, Mork S, Matre R (1982) In-situ characterization of mononuclear cell infiltrates in lesions of multiple sclerosis. Neuropathol Appl Neurobiol 8:403–411.

51. Kuhlmann T, Lingfeld G, Bitsch A, Schuchardt J, Bruck W (2002) Acute axonal damage in multiple sclerosis is most extensive in early disease stages and decreases over time. Brain 125:2202–2212.

37. Traugott U, Reinherz EL, Raine CS (1983) Multiple sclerosis: distribution of T cell subsets within active chronic lesions. Science 219:308–310.

52. Kidd D, Barkhof F, McConnell R, Algra PR, Allen IV, Revesz T (1999) Cortical lesions in multiple sclerosis. Brain 122(Pt 1):17–26.

38. Hauser SL, Bhan AK, Gilles F, Kemp M, Kerr C, Weiner HL (1986) Immunohistochemical analysis of the cellular infiltrate in multiple sclerosis lesions. Ann Neurol 19:578– 587.

53. Peterson JW, Bo L, Mork S, Chang A, Trapp BD (2001) Transected neurites, apoptotic neurons, and reduced inflammation in cortical multiple sclerosis lesions. Ann Neurol 50:389–400.

39. Lock C, Hermans G, Pedotti R, Brendolan A, Schadt E, Garren H, Langer-Gould A, Strober S, Cannella B, Allard J, Klonowski P, Austin A, Lad N, Kaminski N, Galli SJ, Oksenberg JR, Raine CS, Heller R, Steinman L (2002) Gene-microarray analysis of multiple sclerosis lesions yields new targets validated in autoimmune encephalomyelitis. Nat Med 8:500–508.

54. Sailer M, Fischl B, Salat D, Tempelmann C, Schonfeld MA, Busa E, Bodammer N, Heinze HJ, Dale A (2003) Focal thinning of the cerebral cortex in multiple sclerosis. Brain 126:1734–1744.

40. Bielekova B, Goodwin B, Richert N, Cortese I, Kondo T, Afshar G, Gran B, Eaton J, Antel J, Frank JA, McFarland HF, Martin R (2000) Encephalitogenic potential of the myelin basic protein peptide (amino acids 83–99) in multiple sclerosis: results of a phase II clinical trial with an altered peptide ligand. Nat Med 6:1167–1175. 41. Miller DH, Khan OA, Sheremata WA, Blumhardt LD, Rice GP, Libonati MA, Willmer-Hulme AJ, Dalton CM, Miszkiel KA, O’Connor PW (2003) A controlled trial of natalizumab for relapsing multiple sclerosis. N Engl J Med 348:15–23.

55. Mathiesen HK, Jonsson A, Tscherning T, Hanson LG, Andresen J, Blinkenberg M, Paulson OB, Sorensen PS (2006) Correlation of global N-acetyl aspartate with cognitive impairment in multiple sclerosis. Arch Neurol 63:533–536. 56. Vercellino M, Plano F, Votta B, Mutani R, Giordana MT, Cavalla P (2005) Grey matter pathology in multiple sclerosis. J Neuropathol Exp Neurol 64:1101–1107. 57. Evangelou N, Konz D, Esiri MM, Smith S, Palace J, Matthews PM (2000) Regional axonal loss in the corpus callosum correlates with cerebral white matter lesion volume and distribution in multiple sclerosis. Brain 123(Pt 9):1845– 1849.

42. Balcer LJ, Galetta SL, Calabresi PA, Confavreux C, Giovannoni G, Havrdova E, Hutchinson M, Kappos L, Lublin FD, Miller DH, O’Connor PW, Phillips JT, Polman CH, Radue EW, Rudick RA, Stuart WH, Wajgt A, WeinstockGuttman B, Wynn DR, Lynn F, Panzara MA (2007) Natalizumab reduces visual loss in patients with relapsing multiple sclerosis. Neurology 68:1299–1304.

58. Bitsch A, Schuchardt J, Bunkowski S, Kuhlmann T, Bruck W (2000) Acute axonal injury in multiple sclerosis. Correlation with demyelination and inflammation. Brain 123(Pt 6):1174– 1183.

43. Polman CH, O’Connor PW, Havrdova E, Hutchinson M, Kappos L, Miller DH, Phillips JT, Lublin FD, Giovannoni G, Wajgt A, Toal M, Lynn F, Panzara MA, Sandrock AW (2006) A randomized, placebo-controlled trial of natalizumab for relapsing multiple sclerosis. N Engl J Med 354:899–910.

60. Griffiths I, Klugmann M, Anderson T, Yool D, Thomson C, Schwab MH, Schneider A, Zimmermann F, McCulloch M, Nadon N, Nave KA (1998) Axonal swellings and degeneration in mice lacking the major proteolipid of myelin. Science 280:1610–1613.

44. Kappos L, Antel J, Comi G, Montalban X, O’Connor P, Polman CH, Haas T, Korn AA, Karlsson G, Radue EW (2006) Oral fingolimod (FTY720) for relapsing multiple sclerosis. N Engl J Med 355:1124–1140.

61. Smith KJ, Lassmann H (2002) The role of nitric oxide in multiple sclerosis. Lancet Neurol 1:232–241.

45. O’Connor P (2002) Key issues in the diagnosis and treatment of multiple sclerosis. An overview. Neurology 59:S1–33. 46. Ferguson B, Matyszak MK, Esiri MM, Perry VH (1997) Axonal damage in acute multiple sclerosis lesions. Brain 120(Pt 3):393–399. 47. Trapp BD, Peterson J, Ransohoff RM, Rudick R, Mork S, Bo L (1998) Axonal transection in the lesions of multiple sclerosis. N Engl J Med 338:278–285. 48. Coles AJ, Wing MG, Molyneux P, Paolillo A, Davie CM, Hale G, Miller D, Waldmann H, Compston A (1999) Monoclonal antibody treatment exposes three mechanisms underlying the clinical course of multiple sclerosis. Ann Neurol 46:296–304. 49. Kornek B, Storch MK, Weissert R, Wallstroem E, Stefferl A, Olsson T, Linington C, Schmidbauer M, Lassmann H (2000) Multiple sclerosis and chronic autoimmune encephalomyelitis: a comparative quantitative study of axonal injury in active, inactive, and remyelinated lesions. Am J Pathol 157:267–276.

MDC Repository | http://edoc.mdc-berlin.de/9668/

59. DeLuca GC, Williams K, Evangelou N, Ebers GC, Esiri MM (2006) The contribution of demyelination to axonal loss in multiple sclerosis. Brain 129:1507–1516.

62. Dutta R, McDonough J, Yin X, Peterson J, Chang A, Torres T, Gudz T, Macklin WB, Lewis DA, Fox RJ, Rudick R, Mirnics K, Trapp BD (2006) Mitochondrial dysfunction as a cause of axonal degeneration in multiple sclerosis patients. Ann Neurol 59:478–489. 63. Inglese M, Ge Y, Filippi M, Falini A, Grossman RI, Gonen O (2004) Indirect evidence for early widespread gray matter involvement in relapsing–remitting multiple sclerosis. Neuroimage 21:1825–1829. 64. Giuliani F, Yong VW (2003) Immune-mediated neurodegeneration and neuroprotection in MS. Int MS J 10:122–130. 65. Smorodchenko A, Wuerfel J, Pohl EE, Vogt J, Tysiak E, Glumm R, Hendrix S, Nitsch R, Zipp F, Infante-Duarte C (2007) CNS-irrelevant T-cells enter the brain, cause blood– brain barrier disruption but no glial pathology. Eur J Neurosci 26:1387–1398. 66. Minagar A, Alexander JS (2003) Blood–brain barrier disruption in multiple sclerosis. Mult Scler 9:540–549.

7

Infante-Duarte C et al. 67. Aktas O, Smorodchenko A, Brocke S, Infante-Duarte C, Topphoff US, Vogt J, Prozorovski T, Meier S, Osmanova V, Pohl E, Bechmann I, Nitsch R, Zipp F (2005) Neuronal damage in autoimmune neuroinflammation mediated by the death ligand TRAIL. Neuron 46:421–432.

83. Paolillo A, Coles AJ, Molyneux PD, Gawne-Cain M, MacManus D, Barker GJ, Compston DA, Miller DH (1999) Quantitative MRI in patients with secondary progressive MS treated with monoclonal antibody Campath 1H. Neurology 53:751–757.

68. Neumann H, Medana IM, Bauer J, Lassmann H (2002) Cytotoxic T lymphocytes in autoimmune and degenerative CNS diseases. Trends Neurosci 25:313–319.

84. Coles AJ, Wing M, Smith S, Coraddu F, Greer S, Taylor C, Weetman A, Hale G, Chatterjee VK, Waldmann H, Compston A (1999) Pulsed monoclonal antibody treatment and autoimmune thyroid disease in multiple sclerosis. Lancet 354:1691–1695.

69. Diestel A, Aktas O, Hackel D, Hake I, Meier S, Raine CS, Nitsch R, Zipp F, Ullrich O (2003) Activation of microglial poly(ADP-ribose)-polymerase-1 by cholesterol breakdown products during neuroinflammation: a link between demyelination and neuronal damage. J Exp Med 198:1729– 1740. 70. Martin-Villalba A, Herr I, Jeremias I, Hahne M, Brandt R, Vogel J, Schenkel J, Herdegen T, Debatin KM (1999) CD95 ligand (Fas-L/APO-1L) and tumor necrosis factor-related apoptosis-inducing ligand mediate ischemia-induced apoptosis in neurons. J Neurosci 19:3809–3817. 71. Allan SM, Tyrrell PJ, Rothwell NJ (2005) Interleukin-1 and neuronal injury. Nat Rev Immunol 5:629–640. 72. Moncada S, Bolanos JP (2006) Nitric bioenergetics and neurodegeneration. J 97:1676–1689.

oxide, cell Neurochem

73. Beal MF (2005) Mitochondria take center stage in aging and neurodegeneration. Ann Neurol 58:495–505. 74. Hartung HP, Bar-Or A, Zoukos Y (2004) What do we know about the mechanism of action of disease-modifying treatments in MS? J Neurol 251(Suppl 5):v12–v29. 75. Brinkmann V, Davis MD, Heise CE, Albert R, Cottens S, Hof R, Bruns C, Prieschl E, Baumruker T, Hiestand P, Foster CA, Zollinger M, Lynch KR (2002) The immune modulator FTY720 targets sphingosine 1-phosphate receptors. J Biol Chem 277:21453–21457. 76. Massberg S, von Andrian UH (2006) Fingolimod and sphingosine-1-phosphate-modifiers of lymphocyte migration. N Engl J Med 355:1088–1091. 77. Garcia JG, Liu F, Verin AD, Birukova A, Dechert MA, Gerthoffer WT, Bamberg JR, English D (2001) Sphingosine 1-phosphate promotes endothelial cell barrier integrity by Edg-dependent cytoskeletal rearrangement. J Clin Invest 108:689–701. 78. Jaillard C, Harrison S, Stankoff B, Aigrot MS, Calver AR, Duddy G, Walsh FS, Pangalos MN, Arimura N, Kaibuchi K, Zalc B, Lubetzki C (2005) Edg8/S1P5: an oligodendroglial receptor with dual function on process retraction and cell survival. J Neurosci 25:1459–1469. 79. Novgorodov AS, El-Alwani M, Bielawski J, Obeid LM, Gudz TI (2007) Activation of sphingosine-1-phosphate receptor S1P5 inhibits oligodendrocyte progenitor migration. FASEB J 21:1503–1514. 80. Ishii I, Fukushima N, Ye X, Chun J (2004) Lysophospholipid receptors: signaling and biology. Annu Rev Biochem 73:321–354. 81. Rudick RA, Stuart WH, Calabresi PA, Confavreux C, Galetta SL, Radue EW, Lublin FD, Weinstock-Guttman B, Wynn DR, Lynn F, Panzara MA, Sandrock AW (2006) Natalizumab plus interferon beta-1a for relapsing multiple sclerosis. N Engl J Med 354:911–923. 82. Miller DH, Soon D, Fernando KT, MacManus DG, Barker GJ, Yousry TA, Fisher E, O’Connor PW, Phillips JT, Polman CH, Kappos L, Hutchinson M, Havrdova E, Lublin FD, Giovannoni G, Wajgt A, Rudick R, Lynn F, Panzara MA, Sandrock AW (2007) MRI outcomes in a placebo-controlled trial of natalizumab in relapsing MS. Neurology 68:1390– 1401.

MDC Repository | http://edoc.mdc-berlin.de/9668/

85. Coles A, Deans J, Compston A (2004) Campath-1H treatment of multiple sclerosis: lessons from the bedside for the bench. Clin Neurol Neurosurg 106:270–274. 86. Monson NL, Cravens PD, Frohman EM, Hawker K, Racke MK (2005) Effect of rituximab on the peripheral blood and cerebrospinal fluid B cells in patients with primary progressive multiple sclerosis. Arch Neurol 62:258–264. 87. Cree BA, Lamb S, Morgan K, Chen A, Waubant E, Genain C (2005) An open label study of the effects of rituximab in neuromyelitis optica. Neurology 64:1270–1272. 88. Paul F, Jarius S, Aktas O, Bluthner M, Bauer O, Appelhans H, Franciotta D, Bergamaschi R, Littleton E, Palace J, Seelig HP, Hohlfeld R, Vincent A, Zipp F (2007) Antibody to aquaporin 4 in the diagnosis of neuromyelitis optica. PLoS Med 4:e133. 89. Langer-Gould A, Steinman L (2006) What went wrong in the natalizumab trials? Lancet 367:708–710. 90. Coles AJ, Cox A, Le Page E, Jones J, Trip SA, Deans J, Seaman S, Miller DH, Hale G, Waldmann H, Compston DA (2006) The window of therapeutic opportunity in multiple sclerosis: evidence from monoclonal antibody therapy. J Neurol 253:98–108. 91. Kappos L, Bates D, Hartung HP, Havrdova E, Miller D, Polman CH, Ravnborg M, Hauser SL, Rudick RA, Weiner HL, O’Connor PW, King J, Radue EW, Yousry T, Major EO, Clifford DB (2007) Natalizumab treatment for multiple sclerosis: recommendations for patient selection and monitoring. Lancet Neurol 6:431–441. 92. Zipp F, Hartung HP, Hillert J, Schimrigk S, Trebst C, Stangel M, Infante-Duarte C, Jakobs P, Wolf C, Sandbrink R, Pohl C, Filippi M (2006) Blockade of chemokine signaling in patients with multiple sclerosis. Neurology 67:1880–1883. 93. Rottman JB, Slavin AJ, Silva R, Weiner HL, Gerard CG, Hancock WW (2000) Leukocyte recruitment during onset of experimental allergic encephalomyelitis is CCR1 dependent. Eur J Immunol 30:2372–2377. 94. Liang M, Mallari C, Rosser M, Ng HP, May K, Monahan S, Bauman JG, Islam I, Ghannam A, Buckman B, Shaw K, Wei GP, Xu W, Zhao Z, Ho E, Shen J, Oanh H, Subramanyam B, Vergona R, Taub D, Dunning L, Harvey S, Snider RM, Hesselgesser J, Morrissey MM, Perez HD (2000) Identification and characterization of a potent, selective, and orally active antagonist of the CC chemokine receptor-1. J Biol Chem 275:19000–19008. 95. Eltayeb S, Sunnemark D, Berg AL, Nordvall G, Malmberg A, Lassmann H, Wallstrom E, Olsson T, Ericsson-Dahlstrand A (2003) Effector stage CC chemokine receptor-1 selective antagonism reduces multiple sclerosis-like rat disease. J Neuroimmunol 142:75–85. 96. Frantz S (2005) Drug discovery: playing dirty. Nature 437:942–943. 97. Matsunaga K, Klein TW, Friedman H, Yamamoto Y (2001) Legionella pneumophila replication in macrophages inhibited by selective immunomodulatory effects on cytokine formation by epigallocatechin gallate, a major form of tea catechins. Infect Immun 69:3947–3953.

8

Infante-Duarte C et al. 98. Mukhtar H, Ahmad N (1999) Green tea in chemoprevention of cancer. Toxicol Sci 52:111–117.

GTPase activity and are protective in experimental allergic encephalitis. J Exp Med 200:1667–1672.

99. Suganuma M, Okabe S, Sueoka N, Sueoka E, Matsuyama S, Imai K, Nakachi K, Fujiki H (1999) Green tea and cancer chemoprevention. Mutat Res 428:339–344.

114. Waiczies S, Prozorovski T, Zipp F (2005) Modulating T cell signaling cascades by HMG-CoA reductase inhibitors. Signal Transduct 5:231–244.

100. Chung FL, Schwartz J, Herzog CR, Yang YM (2003) Tea and cancer prevention: studies in animals and humans. J Nutr 133:3268S–3274S.

115. Zipp F, Waiczies S, Aktas O, Neuhaus O, Hemmer B, Schraven B, Nitsch R, Hartung H-P (2007) Impact of HMGCoA reductase inhibition on brain pathology. Trends Pharmacol Sci 28:342–349.

101. Ahmad N, Feyes DK, Nieminen AL, Agarwal R, Mukhtar H (1997) Green tea constituent epigallocatechin-3-gallate and induction of apoptosis and cell cycle arrest in human carcinoma cells. J Natl Cancer Inst 89:1881–1886. 102. Yang GY, Liao J, Kim K, Yurkow EJ, Yang CS (1998) Inhibition of growth and induction of apoptosis in human cancer cell lines by tea polyphenols. Carcinogenesis 19:611–616. 103. Liang YC, Lin-Shiau SY, Chen CF, Lin JK (1999) Inhibition of cyclin-dependent kinases 2 and 4 activities as well as induction of Cdk inhibitors p21 and p27 during growth arrest of human breast carcinoma cells by (−)-epigallocatechin-3gallate. J Cell Biochem 75:1–12. 104. Lu YP, Lou YR, Xie JG, Peng QY, Liao J, Yang CS, Huang MT, Conney AH (2002) Topical applications of caffeine or (−)-epigallocatechin gallate (EGCG) inhibit carcinogenesis and selectively increase apoptosis in UVB-induced skin tumors in mice. Proc Natl Acad Sci USA 99:12455–12460. 105. Hsu S, Bollag WB, Lewis J, Huang Q, Singh B, Sharawy M, Yamamoto T, Schuster G (2003) Green tea polyphenols induce differentiation and proliferation in epidermal keratinocytes. J Pharmacol Exp Ther 306:29–34. 106. Yamamoto T, Hsu S, Lewis J, Wataha J, Dickinson D, Singh B, Bollag WB, Lockwood P, Ueta E, Osaki T, Schuster G (2003) Green tea polyphenol causes differential oxidative environments in tumor versus normal epithelial cells. J Pharmacol Exp Ther 307:230–236. 107. Haqqi TM, Anthony DD, Gupta S, Ahmad N, Lee MS, Kumar GK, Mukhtar H (1999) Prevention of collageninduced arthritis in mice by a polyphenolic fraction from green tea. Proc Natl Acad Sci USA 96:4524–4529. 108. Aktas O, Prozorovski T, Smorodchenko A, Savaskan NE, Lauster R, Kloetzel PM, Infante-Duarte C, Brocke S, Zipp F (2004) Green tea epigallocatechin-3-gallate mediates T cellular NF-kappa B inhibition and exerts neuroprotection in autoimmune encephalomyelitis. J Immunol 173:5794–5800. 109. Levites Y, Weinreb O, Maor G, Youdim MB, Mandel S (2001) Green tea polyphenol (−)-epigallocatechin-3-gallate prevents N-methyl-4-phenyl-1,2,3,6-tetrahydropyridineinduced dopaminergic neurodegeneration. J Neurochem 78:1073–1082. 110. Choi JY, Park CS, Kim DJ, Cho MH, Jin BK, Pie JE, Chung WG (2002) Prevention of nitric oxide-mediated 1-methyl-4phenyl-1,2,3,6-tetrahydropyridine-induced Parkinson’s disease in mice by tea phenolic epigallocatechin 3-gallate. Neurotoxicology 23:367–374. 111. Rezai-Zadeh K, Shytle D, Sun N, Mori T, Hou H, Jeanniton D, Ehrhart J, Townsend K, Zeng J, Morgan D, Hardy J, Town T, Tan J (2005) Green tea epigallocatechin-3-gallate (EGCG) modulates amyloid precursor protein cleavage and reduces cerebral amyloidosis in Alzheimer transgenic mice. J Neurosci 25:8807–8814. 112. Lee S, Suh S, Kim S (2000) Protective effects of the green tea polyphenol (−)-epigallocatechin gallate against hippocampal neuronal damage after transient global ischemia in gerbils. Neurosci Lett 287:191–194. 113. Hendriks JJ, Alblas J, van der Pol SM, van Tol EA, Dijkstra CD, de Vries HE (2004) Flavonoids influence monocytic

MDC Repository | http://edoc.mdc-berlin.de/9668/

116. Aktas O, Waiczies S, Smorodchenko A, Dorr J, Seeger B, Prozorovski T, Sallach S, Endres M, Brocke S, Nitsch R, Zipp F (2003) Treatment of relapsing paralysis in experimental encephalomyelitis by targeting Th1 cells through atorvastatin. J Exp Med 197:725–733. 117. Youssef S, Stuve O, Patarroyo JC, Ruiz PJ, Radosevich JL, Hur EM, Bravo M, Mitchell DJ, Sobel RA, Steinman L, Zamvil SS (2002) The HMG-CoA reductase inhibitor, atorvastatin, promotes a Th2 bias and reverses paralysis in central nervous system autoimmune disease. Nature 420:78–84. 118. Waiczies S, Prozorovski T, Infante-Duarte C, Hahner A, Aktas O, Ullrich O, Zipp F (2005) Atorvastatin induces T cell anergy via phosphorylation of ERK1. J Immunol 174:5630– 5635. 119. Waiczies S, Bendix I, Prozorovski T, Ratner M, Nazarenko I, Pfueller CF, Brandt AU, Herz J, Brocke S, Ullrich O, Zipp F (2007) Geranylgeranylation but not GTP-loading determines Rho migratory-function in T cells. J Immunol 179:6024– 6032. 120. Neuhaus O, Strasser-Fuchs S, Fazekas F, Kieseier BC, Niederwieser G, Hartung HP, Archelos JJ (2002) Statins as immunomodulators: comparison with interferon-beta 1b in MS. Neurology 59:990–997. 121. Ifergan I, Wosik K, Cayrol R, Kebir H, Auger C, Bernard M, Bouthillier A, Moumdjian R, Duquette P, Prat A (2006) Statins reduce human blood–brain barrier permeability and restrict leukocyte migration: relevance to multiple sclerosis. Ann Neurol 60:45–55. 122. Suzumura K, Yasuhara M, Tanaka K, Suzuki T (1999) Protective effect of fluvastatin sodium (XU-62-320), a 3hydroxy-3-methylglutaryl coenzyme A (HMG-CoA) reductase inhibitor, on oxidative modification of human lowdensity lipoprotein in vitro. Biochem Pharmacol 57:697–703. 123. Paintlia AS, Paintlia MK, Khan M, Vollmer T, Singh AK, Singh I (2005) HMG-CoA reductase inhibitor augments survival and differentiation of oligodendrocyte progenitors in animal model of multiple sclerosis. FASEB J 19:1407–1421. 124. Vollmer T, Key L, Durkalski V, Tyor W, Corboy J, MarkovicPlese S, Preiningerova J, Rizzo M, Singh I (2004) Oral simvastatin treatment in relapsing–remitting multiple sclerosis. Lancet 363:1607–1608. 125. Sena A, Pedrosa R, Graca Morais M (2003) Therapeutic potential of lovastatin in multiple sclerosis. J Neurol 250:754–755. 126. Göppert-Mayer M (1931) Über Elementarakte mit zwei Quantensprüngen. Ann Phys 401:273–294. 127. Denk W, Strickler JH, Webb WW (1990) Two-photon laser scanning fluorescence microscopy. Science 248:73–76. 128. Williams RM, Zipfel WR, Webb WW (2001) Multiphoton microscopy in biological research. Curr Opin Chem Biol 5:603–608. 129. Kerr JN, Greenberg D, Helmchen F (2005) Imaging input and output of neocortical networks in vivo. Proc Natl Acad Sci USA 102:14063–14068.

9

Infante-Duarte C et al. 130. Helmchen F, Fee MS, Tank DW, Denk W (2001) A miniature head-mounted two-photon microscope. highresolution brain imaging in freely moving animals. Neuron 31:903–912. 131. Nimmerjahn A, Kirchhoff F, Helmchen F (2005) Resting microglial cells are highly dynamic surveillants of brain parenchyma in vivo. Science 308:1314–1318. 132. Sun SW, Liang HF, Trinkaus K, Cross AH, Armstrong RC, Song SK (2006) Noninvasive detection of cuprizone induced axonal damage and demyelination in the mouse corpus callosum. Magn Reson Med 55:302–308. 133. Boretius S, Wurfel J, Zipp F, Frahm J, Michaelis T (2007) High-field diffusion tensor imaging of mouse brain in vivo using single-shot STEAM MRI. J Neurosci Methods 161:112–117. 134. Wuerfel J, Tysiak E, Prozorovski T, Smyth M, Mueller S, Schnorr J, Taupitz M, Zipp F (2007) Mouse model mimics multiple sclerosis in the clinico-radiological paradox. Eur J Neurosci 26:190–198. 135. Bendszus M, Wessig C, Schutz A, Horn T, Kleinschnitz C, Sommer C, Misselwitz B, Stoll G (2005) Assessment of nerve degeneration by gadofluorine M-enhanced magnetic resonance imaging. Ann Neurol 57:388–395. 136. Rogers WJ, Meyer CH, Kramer CM (2006) Technology insight: in vivo cell tracking by use of MRI. Nat Clin Pract Cardiovasc Med 3:554–562. 137. Nakada T (2007) Clinical application of high and ultra highfield MRI. Brain Develop 29:325–335. 138. Taraseviciene L, Miczak A, Apirion D (1991) The gene specifying RNase E (rne) and a gene affecting mRNA stability (ams) are the same gene. Mol Microbiol 5:851–855.

MDC Repository | http://edoc.mdc-berlin.de/9668/

10

Infante-Duarte C et al.

Fig.1: Current understanding of the pathogenesis of MS and potential therapeutic targets. Neuroinflammation has been classically described as the starting point in the pathology of MS. Myelin-specific T cells are believed to orchestrate the autoimmune attack that leads to oligodendrocyte damage and demyelination. Another hallmark of MS, however, is neuronal damage involving both axonal and neuronal loss. Demyelination, axonal transection, and neuronal loss are partly mediated by inflammation but might also occur independently of inflammatory activity. This current understanding of MS pathology impacts directly on therapy development. Ideally, patients need therapies that target both the process of inflammation and the process of neurodegeneration. While state-of-the-art and experimental therapies (such as natalizumab and statins, respectively) target inflammation, other experimental compounds, such as EGCG, provide a promising basis for drugs regulating inflammation and blocking neuronal degeneration

MDC Repository | http://edoc.mdc-berlin.de/9668/

11