You are on page 1of 13

Review

Neuroprotection, regeneration and


immunomodulation: broadening
the therapeutic repertoire in
multiple sclerosis
Orhan Aktas, Bernd Kieseier and Hans-Peter Hartung
Department of Neurology, Heinrich-Heine-Universitat Dusseldorf, 40225 Dusseldorf, Germany

Multiple sclerosis (MS), an incurable but manageable


disorder, is characterized by chronic inflammatory
demyelination and neurodegeneration in the central
nervous system. Although the primary cause of this
often devastating disease remains elusive, major therapeutic advances have occurred during the past two
decades. Here, we present a review of current immunomodulatory treatments and outline upcoming therapy
approaches, including biologics and oral alternatives
that might have equivalent or superior efficacy and/or
enhanced tolerability compared with available treatments, and discuss the scientific rationale and expected
benefits and risks for these compounds. We also speculate about alternatives beyond immune-directed
approaches, review novel insights into the neurobiological consequences of sustained brain inflammation and
evaluate future perspectives for neuroprotective and
neuroregenerative treatment strategies for MS.
Treatment of multiple sclerosis: taking the challenge
Multiple sclerosis (MS) is the most common chronic inflammatory disease of the central nervous system (CNS), affecting 0.1% of the general population in Western countries. As
the leading cause of neurological disability in young adults,
MS represents a significant public health issue, with a high
socio-economic burden and a crucial impact on quality of
life. At diagnosis approximately 85% of patients have a
relapsing remitting disease course (RR-MS), characterized
by unpredictable and recurrent episodes of neurological
deficits, such as loss of vision or paralysis, and followed by
periods of remission. Remission is not always complete,
and usually after several years more than two-thirds of
these patients switch to an insidious secondary progressive
course (SP-MS). For 15% of patients, MS is progressive
from the very beginning (primary progressive, or PP-MS).
Since the first concise description of MS provided by JeanMartin Charcot in 1868, this heterogeneous clinical picture
has been linked to typical neuropathological changes, including the presence of multifocal inflammatory, largely
white matter brain and spinal cord lesions, indicating the
fundamental role of immunological processes for disease
pathogenesis (Figure 1).
Corresponding authors: Aktas, O. (orhan.aktas@uni-duesseldorf.de); Hartung,
H.-P. (hans-peter.hartung@uni-duesseldorf.de)

140

The approval of immunomodulatory injection therapies


in the early 1990 s heralded a new era and now MS therapy
research is the most active therapeutic field in neurology.
Here, we take the continuing series of clinical trials as an
opportunity to review the upcoming treatment strategies,
ranging from unspecific but oral immunomodulatory drugs
to specific interventions by monoclonal antibodies (mAbs).
We focus on the underlying immunological and neurobiological modes of action, link them to specific disease patterns and point to unexpected side effects originating from
the peculiar immunological characteristics of the CNS. The
observed superior efficacy of selective immunosuppression
by mAbs holds promise for future MS therapy, particularly
in early phases of the disease, when modulation of the
immune system might provide indirect neuroprotection. At
later particularly at progressive stages, even state-ofthe-art selective immunosuppressants such as alemtuzumab or rituximab have failed in MS. Therefore, we review
recently elucidated molecular insights into the diffuse and
neurodegenerative nature of MS, including novel concepts
such as disturbances of channel homeostasis and energy
balance, and work up the resulting options regarding
neuroprotective and regenerative therapies.
Current therapies hopes, experiences and facts
Currently, glatiramer acetate (GA) and the interferon
(IFN)-b variants are established as disease-modifying
immunomodulatory treatments for RR-MS (Table 1).
Although the net clinical gain of these drugs for patients
is considered to be only modest, they show a favorable risk
benefit profile and are used around the world since their
introduction [1]. Although used for relapsing forms of MS,
neither the IFN-bs nor GA have demonstrable significant
effect in relapse-free progressive disease states [2]. Moreover, despite the evidence for efficacy in clinical trials,
interindividual treatment response to these agents is
heterogeneous. A substantial proportion of patients continue to experience relapses or clinical progression with
such ongoing immunomodulation, and escalation therapy
with the nonspecific immunosuppressant mitoxantrone is
warranted [3,4]. However, owing to its potential for cumulative cardiotoxicity and the increased occurrence of
acute leukemia, mitoxantrone is considered a second-line
therapy only, restricted to non-responders to first-line

0166-2236/$ see front matter 2009 Elsevier Ltd. All rights reserved. doi:10.1016/j.tins.2009.12.002 Available online 4 January 2010

Review

Trends in Neurosciences

Vol.33 No.3

Figure 1. Concepts on lesion formation in MS. Neuropathological analyses have revealed four different patterns of tissue damage (a), claimed to be unique for individual
subjects and thus indicating interindividual heterogeneity in pathogenesis [119]. There is no doubt that antibodies against myelin contribute to autoimmune demyelination
as amplifiers of the encephalitogenic autoimmune attack and enhancer of disease pathology. Moreover, a possible pathogenic role for antibody-mediated CNS damage is
now obvious in neuromyelitis optica, or Devics syndrome, an MS-related but possibly distinct disease entity characterized by the presence of antibodies against the
astrocytic aquaporin-4 water channel and massive complement depositions in affected brains [120124]. Regarding classical MS, a recent study of the pathological events
preceding myelin phagocytosis in nascent MS lesions raised the possibility that primarily non-inflammatory oligodendrocyte cell death might initiate local macrophage
scavenger activity (b), resulting in subsequent amplification of the inflammatory response [125]. It remains disconcerted whether pathological heterogeneity in MS is
largely a result of evolution of lesional pathology (b), rather than pathogenic heterogeneity (a) [126]. One has to keep in mind that in MS and its animal model, EAE, axonal
damage and neuronal death occurs even outside focal inflammatory lesions, and in earliest disease stages [55,56], even in the spinal cord [127,128].

treatments or to highly active relapsing cases. The same


holds true for natalizumab, the first antibody-based
therapy licensed for relapsing MS in 2004, with a complex
riskbenefit ratio (Box 1).
Upcoming therapies: from peril to progress
Despite intensive efforts, it has not been possible to
identify any single etiological agent or event serving as
the critical and unique starting point for the heterogeneous
inflammatory infiltration of the CNS and neurodegeneration processes that underlie MS. Moreover, recent reports
on widespread neurobiological changes that might be independent of an inflammatory attack have ignited the debate
whether MS is primarily inflammatory in nature or
represents a neurodegenerative disease [5]. Irrespective
of these intriguing novel insights, the still-widely accepted

concept of MS pathogenesis, based and continually


explored in its animal models (such as experimental autoimmune encephalomyelitis, EAE [68]), suggests an autoimmune attack against the myelin sheath as the core
process, providing the rationale for currently evaluated
therapeutic approaches (Figure 2).
Nonspecific immunosuppression and
immunomodulation: perspectives for oral therapies
All first-line therapies for MS are available as parenteral
formulations only, which results in issues of compliance
and injection-related side effects that could be avoided with
oral agents. Currently, several are being explored, including nonselective immunosuppressive and immunomodulatory drugs (Table 2). Agents such as cladribine [9,10]
and teriflunomide [11] are antiproliferative compounds
141

Review

Trends in Neurosciences Vol.33 No.3

Table 1. Approved MS therapies


Therapy
Interferon
(IFN)-b1a and -b1b

Glatiramer acetate
(GA; copolymer-1)

Approved/used for
 CIS

Clinical effects
 Reduces relapse rate

 Relapsing forms
of MS
 SP-MS (IFN-b1b)

 Modifies disease
course (CIS)

 CIS

 Reduces relapse rate

 Standardized, randomized mixture of synthetic polypeptides


consisting of L-glutamic acid, L-lysine, L-alanine and L-tyrosine
with a defined molar residue ratio of 0.14:0.34:0.43:0.09
 Limits T lymphocytes by downregulating migration and
proliferation and by induction of anergy and apoptosis
 Induces an immunomodulatory (and possibly reparative)
T lymphocyte phenotype by shift of GA-reactive T cells from
Th1 towards Th2 during treatment and secretion of BDNF
by GA-reactive T cells
 Promotes development of anti-inflammatory type II monocytes

 RR-MS

Suggested mode of action (selection, see Ref. [1] for details)


 Endogenous immunomodulatory cytokine typically
induced during viral infections
 Limits general T lymphocyte activity by suppressing
proliferation and migration
 Targets encephalitogenic Th17 T lymphocyte population
 Downregulates antigen presentation by MHC class II molecules
 Inhibits immune cell migration by increasing the concentration of
soluble adhesion molecules (sICAM-1, sVCAM-1) and decreasing the
concentration of surface-expressed adhesion molecules (VLA-4)
 Stabilizes the bloodbrain barrier by reducing the expression
of matrix metalloproteinase 9
 Has a complex impact on the cytokine profile

Mitoxantrone

 RR-MS
 SP-MS

 Reduces relapse rate


 Attenuates (secondary)
disease progression

 Anthracenedione chemotherapeutic agent


 Eliminates and deactivates immune cells, particularly
monocytes and macrophages
 Inhibits T lymphocyte proliferation and migration
 Inhibits B lymphocyte activation

Natalizumab

 RR-MS
 SP-MS

 Reduces relapse rate


 Attenuates disease
progression

 Humanized IgG4 mAb against a4 subunit of a4b1 integrin


 Blocks T lymphocyte entry into the CNS

BDNF=brain-derived neurotrophic factor; CIS=clinically isolated syndrome; CNS=central nervous system; IgG=immunoglobulin G; MHC=major histocompatibility complex;
RR-MS=relapsing remitting multiple sclerosis; sICAM=soluble intercellular adhesion molecule; SP-MS=secondary progressive multiple sclerosis; sVCAM=vascular cell
adhesion molecule; VLA=very late antigen.

which interfere with metabolic or signalling pathways of


activated immune cells, and thus act as immunosuppressive therapies. Recently announced data from a Phase III
trial of cladribine indicate that even a few short annual
treatment courses reduce clinical and imaging-based disease activity in RR-MS [9], resulting in recent filings for
registration of this oral agent. A more specific immunomodulatory mode of action has been proposed for two agents
currently in advanced clinical trials, dimethylfumarate
(BG00012), a substance belonging to the family of fumaric
acid esters, and laquinimod. Fumaric acid esters were
effective in RR-MS, probably as a result of the activation
of the neuroprotective nuclear factor E2-related factor 2
transcription pathway [12], and promising data with
BG00012 encouraged ongoing Phase III trials [13]. Laquinimod showed efficacy on magnetic resonance imaging
(MRI)-based measures of disease evolution in RR-MS
[14] and could offer an improved riskbenefit ratio compared with its parent substance linomide for which development was stopped because of serious cardiovascular
toxicity.
Selective immunosuppression: targeting novel
molecular switches
Raising the bar of clinical efficacy by immune-directed MS
treatments might require more selective intervention strategies. Antigen-specific approaches to treat MS are based on
the assumption that unwanted autoreactive T cell clones
drive disease progression by orchestrating an immunological attack against myelin antigens [15]. Therefore, the
142

modulation of such deleterious T lymphocytes constitutes


a key strategy for therapies that might interfere with the
immunological synapse [16], consisting of an antigen (e.g.
a peptide from a myelin-derived protein) embedded within a
major histocompatibility complex (MHC) molecule on the
surface of an antigen-presenting cell (e.g. macrophage),
which is recognized by a specific receptor on a T lymphocyte
(T cell receptor). The major challenge of such antigenspecific approaches is that the target antigen(s) of the
deviated immune response in MS have not been identified
with certainty and that an initially focused immune
response broadens as the disease evolves with spreading
of epitopes and a consequently diversified immunological
attack [17].
Biologics such as recombinant antibodies might provide
additional selective treatment strategies for MS (Table 2)
[18]. A possible candidate is alemtuzumab, which causes
sustained depletion of T lymphocytes from the circulation
upon binding to CD52, a glycosylphosphatidylinositolanchored 2128 kDa protein with unknown function on
the surface of mature lymphocytes, monocytes, macrophage and some dendritic cells [19]. In earlier studies,
alemtuzumab suppressed clinical and MRI signs of inflammation in active SP-MS patients [20] but failed to beneficially impact on disability progression, indicating that once
established neurodegeneration might be uncoupled from
extracerebral inflammatory processes. Alemtuzumab is
now being tested in early stages of RR-MS where it has
shown superior clinical efficacy compared with standard
IFN-b therapy [21]. However, the observed rare but severe

Review

Trends in Neurosciences

Vol.33 No.3

Box 1. Biologics as selective, powerful and challenging MS therapies: lessons from natalizumab
This humanized antibody targeting the a4 integrin subunit blocks
migration of activated immune cells into the CNS and reduced
disease progression and relapse rate in two phase III trials in RR-MS
[99,100]. Progressive multifocal leukoencephalopathy (PML), a lifethreatening opportunistic infection of oligodendrocytes, has emerged
as a rare, serious adverse event and could limit the broader use of
natalizumab. PML is usually confined to immunocompromised
individuals, e.g. HIV patients, and is caused by the reactivation of a
latent infection with the ubiquitous JC DNA polyomavirus (JCV).
CD8+ cytotoxic T cells appear crucial to recognize and destroy JCVinfected cells in the brain. The risk of developing PML on natalizumab
treatment appears to increase with its duration. Most cases occurred
after 24 months of exposure [101,102]. PML has also been reported in
non-MS patients on treatment with rituximab, alemtuzumab and
efalizumab, as well as for conventional immunosuppressants [102
104]. Profound immunosuppression increases replication of JCV in
the kidney and promotes virus entrance in the blood [105]. An obvious

explanation invokes compromised immunosurveillance of the brain.


Natalizumab was found to induce a sustained decrease in the CD4/CD8
ratio in the cerebrospinal fluid [106109]. Additional reasons for the
emergence of PML might be unique to the mode of action of
natalizumab. Natalizumab has been shown to elevate levels of
circulating B lineage cells and mobilize VLA-4+ CD34+ hematopoietic
stem cells from the bone marrow, all potential JCV carriers, and to
promote transcription factors required for JCV replication [110].
Moreover, in light of the substantial discrepancy between high viral
prevalence and low disease incidence, one has to consider virus- or
host-specific factors such as specific virus capsid protein phenotypes,
and not only the immune status, as relevant for PML manifestation
[111]. For the lymphocyte-depleting monoclonal antibody alemtuzumab, an increased frequency of B cell-related autoimmune thyroid
disorders and idiopathic thrombocytopenic purpura was reported,
probably originating from a B cell pool skewed towards immature
phenotypes after alemtuzumab treatment [21,112] (Figure I).

Figure I. The polyoma virus JC (JCV) resides in latent form in the bone marrow, kidnesy and lymphoid tissue but is usually not detected in the blood of
immunocompetent individuals. Rearrangement of the regulatory region sequence of the virus is required for reactivation of JCV to result in progressive multifocal
leukoencephalopathy (PML). This is usually counteracted by a CD8 T lymphocyte mediated immune response. If this protective response is compromised, JCV may
infect oligodendrocytes, proliferate and induce productive lytic infection with subsequent PML. Natalizumab may flush B cell precursors, potential harbors of JCV, from
the bone marrow into the blood and with few cells allowed access invade the CNS en passant. Natalizumab may promote PML conceivably also through activation of
transcription factors necessary for viral replication.

side effects (Box 1) call for further strategies preserving the


impressive clinical efficacy but minimizing the risk. One
alternative with a so-far favorable risk profile could be
daclizumab, a mAb to CD25, the a-subunit of the interleukin-2-receptor (IL-2R) on T cells. In MS, first-therapy
studies showed a significant suppression of newly occurring inflammatory MRI lesions and clinical activity [22,23].
Interestingly, this monoclonal antibody appears not to
suppress T cell proliferation mediated through the highaffinity receptor (a, b, g IL-2R subunits) but is associated
with intermediate affinity IL-2R-mediated expansion of a
subset of natural killer (NK) cells that express the marker
CD56 at high levels. This subset has immunoregulatory
properties.
Other biologics aim at targeting the homeostasis of B
lymphocytes that have more recently been recognized to

play a crucial role in the perpetuation of ongoing encephalitogenic immune responses through antibody synthesis or
by antigen presentation and subsequent T cell proliferation [24]. Anti-myelin antibodies amplify autoimmune
demyelination [25], and recent observations have raised
the question of whether antibody responses against
primary neuronal structures such as neurofilaments also
contribute to autoimmune neurodegeneration [26,27]. An
alternative hypothesis in progressive MS could be occurrence of compartmentalized sustained B cell activation
that is independent of systemic immunity [28]. These
aggregate data provide a strong rationale for B celldirected immunotherapies, such as the anti-CD20 mAb
rituximab. Indeed, rituximab-induced B cell depletion
has recently been demonstrated to confer therapeutic
benefit in MS, not only in severe MS cases but also in
143

Review

Trends in Neurosciences Vol.33 No.3

Figure 2. Therapeutic approaches to neuroimmunomodulation. The neuropathology of MS links the heterogeneous clinical presentation to a characteristic pattern,
reflected by multifocal inflammatory brain and spinal cord white matter lesions of different age. Despite intensive efforts, no infectious agent could be identified to date. The
presumed autoimmune hypothesis was ignited by the animal model EAE. Here, immunization with myelin antigens or the transfer of activated myelin-specific T
lymphocytes results in MS-like CNS inflammation [129]. As one might expect from any type of model, EAE has several drawbacks limiting the transfer of novel findings to
MS [8]. Nevertheless, the different EAE variants have provided useful insights for the evaluation of possible therapeutic targets. (a) According to current immunological
concepts, preexisting self-reactive T lymphocytes, specific for CNS autoantigens such as myelin basic protein, escape immunoregulatory mechanisms, persist and undergo
expansion in lymphatic tissues outside the brain such as the spleen or lymph nodes. Activation enables these lymphocytes to approach the bloodbrain barrier, following a
gradient of chemokines, and to interact with adhesion molecules expressed on brain endothelial cells. Following penetration of the bloodbrain barrier, these lymphocytes
become reactivated by local antigen-presenting cells (perivascular macrophages or dendritic cells) within the perivascular space, and finally invade the CNS parenchyma.
They now create a proinflammatory environment and recruit further immunocompetent cells such as B cells and antibody-secreting plasma cells, as well as macrophages
or microglia. The concept of a myelin-specific immune attack is also compatible with the recently recognized neurodegenerative features of MS [98]. Apparently, neuronal
damage can occur as collateral damage during autoimmune demyelination [50,130]. This phenomenon might be explained by unexpected crossreactivity between myelin
and neuronal antigens resulting in cumulative autoimmune responses [131,132]. (b) The putative modes of action for the new MS drug candidates are heterogeneous and
comprise the interaction with various cellular and molecular counterparts within the immune and the nervous system. T-shaped lines in red indicate blocking of target
structures or pathways, and red arrows indicate possible therapeutic interactions. APC, antigen-presenting cell (such as macrophage or dendritic cell); OG, oligodendrocyte;
Pc, plasma cell; S1P-R, sphingosine 1 phosphate-receptor; Th, T helper cell; VLA-4, very late antigen-4.

more mild RR-MS [29]. A subset of patients with PP-MS


that show signs of inflammatory activity also appear to
receive benefit from this monoclonal [30]. Based on the
encouraging proof-of-principle experience with rituximab,
depleting antibodies such as ocrelizumab and ofatumumab
that recognize the same CD20 antigen are currently being
evaluated.
144

Targeting immune cell migration (and more):


unexpected endogenous pathways
Additional selective compounds explore alternative pathways to modulate immune cell mobility. The heterogeneous interactions of invading immune cells with the
bloodbrain barrier [31], a complex structure composed of
heterogeneous subcompartments [32], as well as the sub-

Review

Trends in Neurosciences

Vol.33 No.3

Table 2. Compounds currently tested in phase II/III trials for MS therapy (selection of promising and innovative approaches)
Compound
Leukocyte depletion
Alemtuzumab

Daclizumab

Mode of action

Route of Indication currently


application tested [Phase]

 Humanized IgG1 mAb to CD52, a surface antigen of unknown i.v.


function present on lymphocytes, monocytes and dendritic cells
 Depletes T and, to a lesser extent, B lymphocytes
 T cell depletion is sustained, B cell depletion is transient
 Increases levels of BAFF and regulatory T lymphocytes (Treg)
i.v./s.c.
 Humanized IgG1 mAb to CD25, the a-subunit of
the IL-2 receptor of T lymphocytes
 Increases number of immunoregulatory CD56bright NK cells
 Decreases number of CD8+ T lymphocytes

B lymphocyte depletion and inhibition


Ofatumumab (HuMax-CD20)  Human IgG1 mAb to CD20
 Depletes mainly CD20+ B lymphocytes (not plasma cells)
 Humanized IgG1 mAb to CD20
Ocrelizumab
 Depletes mainly CD20+ B lymphocytes (not plasma cells)
 Chimerical IgG1 mAb to CD20
Rituximab
 Depletes mainly CD20+ B lymphocytes (not plasma cells)
General immunosuppressive
 Synthetic adenosine deaminase-resistant
Cladribine
nucleoside analog with selective lymphocyte activity
(2-chlorodeoxyadenosine)
 Disrupts DNA synthesis and repair in lymphocytes
 Impedes migration of T cells and monocytes
 Modulates chemokine and cytokine levels
 Antimetabolite related to leflunomide used for
Teriflunomide
rheumatoid arthritis
 Inhibits nucleotide synthesis
Broader/unknown mechanism of action
 Immunomodulator related to linomide with unknown
Laquinimod
molecular target
 Induces immunological Th1Th2 shift
Dimethyl fumarate (BG00012)  Activates transcription factor NF-E2-related factor (Nrf) 2
 Upregulates anti-oxidant response elements
 Related to other fumaric acids already used for psoriasis
Immune cell trafficking
Fingolimod (FTY720)

BAF312
Firategrast (SB683699)
Neuroprotective
Amiloride
Lamotrigine
Topiramate

 Modulates activation of G-protein-coupled


sphingosine 1-phosphate (S1P) receptors 1, 35: mediates
persistent signalling via internalized S1P1 receptors
 Influences migration by trapping lymphocytes in
lymph nodes and thus reducing their recirculation to
peripheral inflammatory tissues
 Differentially retains effector memory cells or
proinflammatory CD4 Th17 cells
 Might diminish astrogliosis by modulating S1P1 and
3 receptors on astrocytes
 Might promote remyelination by acting on
oligodendrocyte S1P5 receptors
 Modulates activation of S1P receptors 1 and 5
 Blocks a4 integrin subunit (small molecular drug)







Blocks acid-sensing ion channel-1 (ASIC1)


Blocks voltage-sensitive Na+ channels
Inhibits excitatory neurotransmission
Blocks voltage-sensitive Na+ channels
Enhances GABA-activated chloride channels
Modulates kainate and AMPA receptors

Ref.

RR-MS [III]

[1921,113,114]

RR-MS [III]

[22,23,115]

i.v.

RR-MS [I/II]

[116]

i.v.

RR-MS [II]

[117]

i.v.

RR-MS [II]

[29,30]

Oral

RR-MS [III]

[9,10]

Oral

RR-MS [II]

[11]

Oral

RR-MS [III]

[14]

Oral

RR-MS [III]

[12,13]

Oral

RR-MS [III]

[3440]

PP-MS [III]

Oral
Oral

RR-MS [II]
RR-MS [II]

Oral
Oral
Oral

Planned: RR-MS [II] [74]


SP-MS [II ambiguous] [118]
RR-MS [II]
[61]

AMPA=a-amino-3-hydroxyl-5-methyl-4-isoxazole-propionate; ASIC1=acid-sensing ion channel 1; BAFF=B cell activating factor; GABA=g-aminobutyric acid; IFN=interferon;
MRI=magnetic resonance imaging; MBP=myelin basic protein; s.c.=subcutaneous; i.v.=intravenous; HLA=human leukocyte antigen; RR-MS=relapsing-remitting MS; SPMS=secondary progressive MS; PP-MS=primary progressive MS.

sequent impact of chemokines [33], low-molecular weight


cytokines that guide immune cells from primary to secondary immune organs to target issue, offer additional
options to restore dysregulated immune cell trafficking.
Upon phosphorylation, the sphingosine 1-phosphate (S1P)
receptor modulator fingolimod (FTY720) mediates persistent signalling via internalized S1P1 receptors [34,35]. As a

consequence, lymphocytes are trapped in lymph nodes and


prevented from recirculation (lymphocyte egress) [36].
Another S1P receptor subtype, S1P5 is expressed on oligodendrocytes and neurons and might exert independent
reparative effects (see below) [37]. Notably, fingolimod can
cross the bloodbrain barrier. Encouraging Phase II study
data showed that blockade of lymphocyte egress decreases
145

Review
the manifestation of inflammatory lesions in RR-MS
[38,39] and prompted the initiation of advanced clinical
studies, including more specific S1P modulators such as
BAF312 (Table 2). Meanwhile, positive results of a head-tohead comparison of fingolimod with standard IFN-b
therapy have been reported [40]. However, a suspicious
clustering of cases with malignant skin cancers and viral
infections was also observed [38], among them two cases of
lethal herpes simplex and varicella encephalitis as well as
a case with focal hemorrhagic encephalitis of unknown
origin [41]. Recently, the kininkallikrein system exerting
a variety of immunological functions [42] has been identified as an additional and hitherto unknown pathway
controlling encephalitogenic immune cell migration. The
kinin receptor B1 critically regulates the homing of encephalitogenic T cells into the CNS and targeting B1 might
attenuate disease in MS animal models [43]. Blunting
activity of the bradykinin-cleaving angiotensin converting
enzyme has been reported to induce potent regulatory T
cells and also to modulate encephalitogenic T cell autoimmunity [44,45], justifying clinical studies in MS. Taken

Trends in Neurosciences Vol.33 No.3

together, these data indicate that although efficient blockade of selected components of the immune system might
produce the desired clinical impact, it might also have
uncommon serious adverse effects that clearly mandate
vigorous monitoring programs and a continual assessment
of the riskbenefit ratio.
The horizon of future therapies: perspectives for
neurobiological approaches
Current immune-based therapies and prospects suggest a
narrow window of therapeutic opportunity with such
agents and point to the need to also focus on the challenging and neurodegenerative processes observed in MS.
Considerable advances have been achieved regarding
our understanding of the multifaceted damage pattern
to neurons and glia in the CNS in MS. Starting with the
rediscovery of axons as immediate rather than subsequent victims of the inflammatory process [46,47],
the timing and the precise molecular mechanisms responsible for these irreversible pathological changes are being
intensively studied (Figure 3) [15]. Axonal transection

Figure 3. Neuronal dysfunction, destruction and failed regeneration in the course of chronic inflammation. (a) Recent studies indicate that even subtle immunological
alterations should be the focus of significant attention for neurodegenerative processes: apparently, immune mechanisms observed in the course of mild and chronic
inflammation crucially contribute to neuronal dysfunction and promote CNS degeneration in the course of MS. Magnetic resonance spectroscopy studies revealed
remarkable changes within the so-called NAWM (normal-appearing white matter) at disease onset: significantly reduced levels of NAA, a marker for neuronal integrity,
indicates profound and widespread neuronal dysfunction which cannot be explained by the anatomical localization of the inflammatory lesions within the white matter [67].
Classical neuropathological studies have described dendritic pathology within the cortex of MS patients. Indeed, MS patients might exhibit a relevant cognitive decline
from the very beginning [133]. Moreover, during acute inflammatory relapses, MS patients show reversible neurological deficits. The mechanisms responsible for this
disperse dysfunction represent a relevant therapeutic target. (b) Novel insights suggest a complex damage scenario comprising the impact of inflammatory tissue
alterations such as reactive oxygen species (ROS), nitric oxide (NO), tissue acidosis, myelin-specific antibodies or death ligands, on the myelin sheath, the corresponding
oligodendrocyte and the axon itself. The disturbed ion homeostasis is reflected by an increased Ca2+ influx, leading to the harmful activation of Ca2+-dependent proteases
such as calpain, and a persisting increase in intra-axonal Na+ concentrations, leading to limited nerve conduction. These general alterations are associated with a prominent
failure of mitochondria which are not able to meet the energy needs, and with a loss of essential neurotrophic support as a result of demyelination. (c) Local regeneration
within a typical inflammatory MS plaque upon immune cell-mediated damage depends on the preservation of local progenitor cells and their capacity to migrate and
differentiate to, e.g. remyelinating oligodendrocytes (OG). The inflamed tissue itself is characterized by metabolic changes which, in turn, inhibit the activation of
endogenous repair processes. The presence of a strong astroglial scar inhibiting spontaneous tissue restoration has also to be considered. If remyelination takes place, the
resulting myelin sheaths are typically thinner and shorter, compared with developmental myelination. Importantly, the myelin sheath also provides trophic support for the
axon and decreases the neuronal vulnerability towards further inflammatory insults. Thus, neurobiological efforts to enhance endogenous repair processes including
remyelination and neuronal restoration represent an important approach beyond pure immunomodulatory or suppressive regimens.

146

Review
within acute lesions was associated with the local accumulation of macrophages, microglia as well as T cells. Several
mechanistic studies have demonstrated the capacity of T
cells to mediate collateral neuronal damage upon activation in vitro and in vivo [4854]. Moreover, neuronal cell
death and synaptic damage was found in the human cortex
[55], and linked to cortical demyelination which had been
neglected so far in studies that have used standard histochemistry methods [56]. Remarkably, these cortical
lesions lack lymphocytic infiltrates and are characterized
by activated microglia. Quantitative studies of the socalled normal appearing white matter (NAWM), devoid
of macroscopic inflammatory lesions, demonstrated a substantial reduction of axonal density [15].
In addition to axonal damage, either immediate or
subsequent to acute and massive inflammatory infiltration, chronic neurodegeneration continuously proceeds
(Figure 3a) and represents the major clinical determinant
of accumulating and irreversible disability in progressive
disease phases [57]. This insidious process is mainly driven
by minute inflammation, possibly sustained by innate
immune mechanisms [58], which might contribute to the
widespread presence of chronically activated microglia in
the MS brain even outside focal lesions. Persisting inflammation detected in various MS stages (both relapsing and
progressive) [59] challenges basic neuronal homeostasis
and particularly affects demyelinated axons which lack
myelin-derived support and might thus be prone to
changes of ion concentrations or inflammatory secretions
that might otherwise not be harmful [60]. The latter comprise a plethora of mediators, including proteolytic
enzymes such as matrix metalloproteases, cytokines, oxidative products and free radicals as integral components of
emerging concepts of neurodegeneration (Figure 3b).
Failed energy metabolism: ion channel homeostasis as a
potential therapeutic target
In the center of these concepts are reactive oxygen and
nitrogen species such as nitric oxide (NO) which are found
in raised concentrations in MS and which could directly
contribute to neurodegeneration. According to this concept
[61], NO inhibits mitochondrial respiration and leads to
subsequent energy failure within the axon [62]. The resulting ATP deficiency results in a gradual loss of function of
the Na+/K+ ATPase and a collapse of transmembrane ionic
gradients. This process is amplified by simultaneous perhaps immune-related demyelination. The acute loss of a
myelin segment leads to a conduction block which can be
mitigated by insertion of Na+ v1.6 channels along the
denuded sodium channel poor axolemma, resulting in
inefficient non-saltatory conduction along the injured segment [63]. This in turn results in a persisting Na+ influx
and, consequently, membrane depolarization. In response
to the unforeseen ionic imbalance, the energy-independent
Na+/Ca2+ exchanger starts to reverse its operation by
exporting Na+ and importing Ca2+ into the intra-axonal
compartment, thus leading to excessive intracellular Ca2+
concentrations. The axonal Ca2+ overload might be further
boosted by an independent, glutamate-dependent injury
mechanism. Neuropathological studies revealed that
macrophages and microglia found in close vicinity to dys-

Trends in Neurosciences

Vol.33 No.3

trophic axons highly express glutaminase, the major intracellular source of glutamate synthesis. Obviously,
subsequent overshooting activation of excitatory glutamate receptors might also contribute to increased intracellular Ca2+ levels [64]. Dutta et al. [65], studying lesions
of progressive MS patients, observed in the majority of
demyelinated axons direct fingerprints of Ca2+-dependent
protease activity such as fragmented neurofilaments and
depolymerized microtubules [65]. In the same study, transcript profiling revealed a decrease of mRNA levels of
mitochondrial genes, indicating an impaired activity of
mitochondrial enzyme complexes of the electron transport
chain and supporting the hypothesis that a mismatch
between energy demand and ATP availability drives
degeneration. These findings were recently confirmed
[66] and suggest damage mechanisms beyond direct
immune cell-mediated axonal and neuronal destruction.
Moreover, such subtle alterations might explain diffuse
reductions for N-acetylaspartate (NAA) observed in the
NAWM of RR-MS patients, sometimes even very early in
the disease course [67,68].
Viewed in this light, certain aspects of the neurodegenerative process might allow MS to be legitimately classified
as a channelopathy, mainly caused by energy failure, and
therapies limiting ongoing Na+ and reactive Ca2+ entry
into the axon might provide a novel neuroprotective
approach [61]. This goal could be achieved by sodium
channel-blocking compounds such as phenytoin, a clinically validated anticonvulsant which has shown therapeutic efficacy in EAE [69]. In addition to a direct effect
on axons, phenytoin also interferes with inflammatory
effector mechanisms, as it strongly reduces immunological
effector functions of activated Na+ v1.5 channel-expressing
microglia [69,70]. However, it is noteworthy that a recent
report indicates abrupt withdrawal of phenytoin results in
immediate exacerbation and even fatalities in EAE [71].
Although the precise mechanisms underlying this rebound
effect have not been elucidated yet, a possible contribution
of the so-far neglected Na+ v1.5 and v1.6 channels on
macrophages has been suggested [61], and a clinical Phase
II study in PP-MS was halted. For lamotrigine, another
sodium channel blocker, the results of a placebo-controlled
therapy trial in SP-MS were recently announced. Although
the primary endpoint, a beneficial impact on brain tissue
volume loss, was missed, patients taking lamotrigine significantly improved in walking speed, although the study
was not designed to measure such an effect as a major
outcome. Results from a placebo-controlled therapy trial in
RR-MS with topiramate, an anticonvulsant with a complex
mechanism of action [72], are pending.
A similarly complex regulation has been recently
described for TWIK-related acid-sensitive potassium
(TASK) channels which allow the regulated efflux of potassium ions. Although inflammatory plaques of human MS
patients displayed profoundly lowered expression of both
TASK isoforms 1 and 3, the same channels were also found
to influence T cell effector function in EAE [73], hampering
the transfer of these findings into clinical scenarios. Similarly, the search for further molecular candidates resulted
in the identification of the proton-gated acid-sensing ion
channel 1 (ASIC1) as a potential therapeutic target [74].
147

Review
ASIC1 contributes to axonal dysfunction and degeneration
in inflammatory lesions by augmenting neuronal Na+ and
Ca2+ influxes during tissue acidosis, and experimental
ASIC1 blockade by amiloride conferred axonal integrity
in the EAE model. Again, an old drug, already approved for
treatment of arterial hypertension, offers new hope for
primary neuroprotection and the basis for a future therapeutic trial. One has to keep in mind that, similar to the
other targets discussed above, ASIC1 is also expressed on
immune cells, thus the safety and efficacy of such channel
blockers for neuroprotection in MS remains uncertain.
Failed neuroregeneration: how to enhance endogenous
regeneration?
Experimental data suggest that in MS, deleterious neurodegenerative processes commence early and are already
established in the brain at the time of the first clinical
manifestation. Therefore, in addition to studying immunedirected prophylactic or neuroprotective strategies, there
is also an urgent need to discover regenerative approaches,
with the goal to restore and repair the injured tissue ab
initio (Figure 3c). In light of demyelination as a predominant neuropathological feature in MS, and the particular
vulnerability of denuded axons towards otherwise mild
inflammatory alterations, restoration of myelin sheath
integrity is a crucial issue [75]. Even within an acute
inflammatory brain lesion, spontaneous remyelination is
possible and regularly observed [76], yet a significant
feature of the disease is remyelination failure, probably
as a result of a differentiation block of oligodendroglial
progenitor cells (OPCs) [77]. Thus, regenerative treatment
approaches have aimed at restoring the myelin sheath by
transplanting exogenous myelin-forming cells [75]. However, the success of such cell replacement therapies is
clearly hampered by the multifocal (and sometimes even
diffuse) manifestation of MS lesions, hence requiring challenging neurosurgical procedures, and posing an immunological threat as a result of tissue incompatibility between
the donors and recipients immune systems. There might
be little benefit to be gained by delivering remyelinating
cells into inflammatory MS lesions that already contain
abundant progenitor cell populations with the capacity to
generate new oligodendrocytes [78]. Remyelination failure
is substantially caused by the hostile proinflammatory
environment that typically initiates and sustains demyelination [75] and to which, if not controlled, any new tissue
growth would be subject. Moreover, similar to other regenerative processes in the body, the overall remyelination
capacity in the CNS declines with age [79].
Regarding the underlying molecular mechanisms of
observed natural but limited remyelination, a recent study
revealed the substantial role of the leucine rich repeat and
Ig domain containing-1 (LINGO-1) which is exclusively
expressed on neurons and negatively regulates myelination. In EAE, systemic treatment with an antagonistic
LINGO-1 antibody leads to functional recovery and
increased axonal integrity without affecting the encephalitogenic immune response [80], justifying early clinical
studies in humans. Another possible strategy to enhance
endogenous remyelination is the identification of intracellular pathways repressing the final differentiation of
148

Trends in Neurosciences Vol.33 No.3

progenitors. As recently reported, differentiation of oligodendrocyte progenitors during remyelination (upon toxic
experimental demyelination) is regulated by histone deacetylases, offering novel epigenetic clues for myelin repair
strategies [79]. Regarding the underlying molecular mechanisms, the cyclin-dependent kinase inhibitor p57kip2 as
well as the Notch pathway might represent potent candidates, as they are responsible for the regulation of oligodendroglial differentiation and, at the same time,
modulated in the course of autoimmune neuroinflammation [8183]. Perhaps surprisingly, the above discussed
lymphocyte egress blocker, fingolimod, has recently been
shown to have a favorable impact on OPC homeostasis and
remyelination processes by modulating the S1P5 receptor
in vitro [37] and could thus uniquely combine immunomodulatory with neuroreparative properties.
Neuronal restoration
Many pathological conditions in the brain are characterized
by insufficient replacement of damaged neurons and poor
rewiring of axons, thus accounting for fixed and irreversible
clinical deficits. Several lines of evidence indicate that functional repair based on axonal outgrowth is hampered by
abundantly expressed myelin-associated inhibitory factors
in the CNS. Among them, the myelin component Nogo-A has
already been the focus of considerable attention, as it interacts with the neuronal Nogo receptor complex and limits
axonal sprouting in a contact-dependent manner. In experimental models of spinal cord injury, significant functional
recovery was achieved upon antibody-mediated blockade of
Nogo interactions [84]. These findings encouraged similar
studies in animal models of MS, yielding conflicting results
in independent studies. Indeed, Nogo has multifaceted roles
in the course of autoimmune demyelination and also influences the encephalitogenic immune response [85,86]. Thus,
prior to considering Nogo modulation as a future regenerative approach in MS, many issues remain to be elucidated to
dissect pure neurobiological from rather immunological
Nogo functions [86].
Very little is known about the contribution of other
inhibitory structures such as the prominent scarring in
MS lesions. The extracellular matrix within these scars
mainly consists of collagen fibers which themselves exert
a repulsive effect on sprouting axons [87]. Moreover, regarding neuronal regeneration in MS lesions, recent animal
experimental and human studies indicate that inflammation might trigger the proliferation of neural progenitors
both in the subventricular zone [88] as well as in chronic
lesions [89]. However, the final differentiation into mature
and integrated neurons obviously fails in MS and EAE
[89,90]. The cellular composition of chronic lesions is typically dominated by the abundant presence of activated
astrocytes, a uniform and very early reaction of the CNS
tissue in the course of autoimmune demyelination [91,92].
This so-called reactive astrogliosis [93] might be relevant for
tissue stabilization in the acute injury phase but seems to be
responsible for inadequate neuronal regeneration later on
[94]. Recent reports suggest a cardinal role for basal metabolic changes of the CNS environment. Obviously, subtle
modulation of the redox equilibrium found even in mild
inflammation impairs the proliferation of neural or more

Review
restricted glial stem cells [9597] and lead to the generation
of astrocytes at the expense of neurons [97]. In the latter
study, the altered differentiation pattern was mediated by
the histone deacetylase silent information regulator 1
(Sirt1), which serves as a sensor for the redox potential in
neural stem cells and directs their differentiation via the
modulation of the basic helixloophelix transcription factor
hairy/enchancer of split 1 (Hes1) and the subsequent inhibition of proneuronal mammalian achaete-scute homolog 1
(Mash1). Another recent study implicated the related sonic
hedgehog (Shh) pathway [90]. Apparently, proinflammatory
IFN-g secreted during autoimmune demyelination inhibits
Shh-mediated differentiation of neural progenitors and
results in the disruption of the neurogenic rostral migratory
stream. As mild and sustained inflammatory processes
occur not only in MS but also in several other neurological
diseases, such as stroke, spinal cord injury or primary
neurodegenerative diseases [98], the characterization of
these regulatory circuits could help to reinforce the endogenous restoration programs of the brain.
Concluding remarks and outlook
With the successful introduction of immunomodulatory
therapies, we have witnessed major therapeutic advances
in MS during the past two decades. The dominant inflammatory pathology of the disease focuses upcoming therapies on aberrant immunological processes. Drug
development will optimistically shortly offer oral alternatives to current parenteral agents. Experimental
approaches that bridge immunological and neurobiological
aspects of the disease are underway and will provide a
better understanding of the molecular steps leading to
inflammation-mediated axonal dysfunction and failure of
remyelination (Box 2). Targeting these processes, which
arise from the interface of immune response and brain
Box 2. Outstanding questions
 What are the molecular pathways of chronic neuronal dysfunction
in the course of mild but persisting CNS inflammation?
 What are appropriate endpoints and biomarkers for therapeutic
studies of neuroprotective, remyelinating and repair-promoting
therapies?
 Is neurodegeneration always driven or conditioned by inflammation?
 Which role does innate immunity play in the neurodegenerative
stages of MS?
 What are the determinants of lesion localization within the CNS
white and gray matter?
 How can we target the astroglial network for treating inflammatory neurodegeneration?
 How can we better identify immunological and neurobiological
subgroups of MS patients?
 How can we identify responders to immune-based therapies?
 How can we selectively modulate the activated immune system
without impairing beneficial immune reactions?
 Could resident CNS immunity be modulated by deactivating
microglia?
 How can we target sequestered compartmentalized immune
responses in the brain (drugs need to penetrate the bloodbrain
barrier)?
 How can we foster endogenous repair mechanisms to promote
remyelination, axonal survival and outgrowth?
 Is there prospect for cell-based therapies as replacement or
delivery vehicles?

Trends in Neurosciences

Vol.33 No.3

homeostasis, might offer novel therapeutic avenues to


control autoimmune inflammation and to prevent neurodegeneration in this still incurable disorder.
Acknowledgements
The authors wish to thank Dr. Stephen C. Reingold, New York, for helpful
discussions. Supported by Forschungskomission, Medical Faculty, HHU,
the Deutsche Forschungsgemeinschaft (SFB-Transregio 43), the
Heinrich- und Erna-Schaufler-Stiftung, and KKNMS (German
Competence Network on Multiple Sclerosis).

References
1 Menge, T. et al. (2008) Disease-modifying agents for multiple sclerosis:
recent advances and future prospects. Drugs 68, 24452468
2 Hartung, H.P. and Aktas, O. (2009) Bleak prospects for primary
progressive multiple sclerosis therapy: downs and downs, but a
glimmer of hope. Ann. Neurol. 66, 429432
3 Rudick, R.A. and Polman, C.H. (2009) Current approaches to the
identification and management of breakthrough disease in patients
with multiple sclerosis. Lancet Neurol. 8, 545559
4 Hartung, H.P. et al. (2002) Mitoxantrone in progressive multiple
sclerosis: a placebo-controlled, double-blind, randomised, multicentre
trial. Lancet 360, 20182025
5 Trapp, B.D. and Nave, K.A. (2008) Multiple sclerosis: an immune or
neurodegenerative disorder? Annu. Rev. Neurosci. 31, 247269
6 Steinman, L. (2003) Optic neuritis, a new variant of experimental
encephalomyelitis, a durable model for all seasons, now in its
seventieth year. J. Exp. Med. 197, 10651071
7 Gold, R. et al. (2006) Understanding pathogenesis and therapy of
multiple sclerosis via animal models: 70 years of merits and culprits in
experimental autoimmune encephalomyelitis research. Brain 129,
19531971
8 Ransohoff, R.M. (2006) EAE: pitfalls outweigh virtues of screening
potential treatments for multiple sclerosis. Trends Immunol. 27, 167
168
9 Giovannoni, G. (2009) Results from the CLARITY study: a phase III,
randomized, double-blind study to evaluate the safety and efficacy of
oral cladribine in relapsing-remitting multiple sclerosis (RRMS).
Annual Meeting of the American Academy of Neurology, Seattle,
WA, USA. Neurology Abstract LBS.001
10 Hartung, H.P. et al. Development of oral cladribine for the treatment
of multiple sclerosis. J. Neurol.(in press), Epub (Nov 18)
11 Merrill, J.E. et al. (2009) Teriflunomide reduces behavioral,
electrophysiological, and histopathological deficits in the Dark
Agouti at model of experimental autoimmune encephalomyelitis. J.
Neurol. 256, 89103
12 Nguyen, T. et al. (2009) The Nrf2-antioxidant response element
signaling pathway and its activation by oxidative stress. J. Biol.
Chem. 284, 1329113295
13 Kappos, L. et al. (2008) Efficacy and safety of oral fumarate in patients
with relapsing-remitting multiple sclerosis: a multicentre,
randomised, double-blind, placebo-controlled phase IIb study.
Lancet 372, 14631472
14 Comi, G. et al. (2008) Effect of laquinimod on MRI-monitored disease
activity in patients with relapsing-remitting multiple sclerosis: a
multicentre, randomised, double-blind, placebo-controlled phase IIb
study. Lancet 371, 20852092
15 Aktas, O. et al. (2007) Neurodegeneration in autoimmune
demyelination: recent mechanistic insights reveal novel therapeutic
targets. J. Neuroimmunol. 184, 1726
16 Dustin, M.L. (2009) The cellular context of T cell signaling. Immunity
30, 482492
17 Sospedra, M. and Martin, R. (2005) Immunology of multiple sclerosis.
Annu. Rev. Immunol. 23, 683747
18 Lutterotti, A. and Martin, R. (2008) Getting specific: monoclonal
antibodies in multiple sclerosis. Lancet Neurol. 7, 538547
19 Waldmann, H. and Hale, G. (2005) CAMPATH: from concept to clinic.
Philos. Trans. R. Soc. Lond. B Biol. Sci. 360, 17071711
20 Coles, A.J. et al. (1999) Monoclonal antibody treatment exposes three
mechanisms underlying the clinical course of multiple sclerosis. Ann.
Neurol. 46, 296304
21 Coles, A.J. et al. (2008) Alemtuzumab vs. interferon beta-1a in early
multiple sclerosis. N. Engl. J. Med. 359, 17861801
149

Review
22 Rose, J.W. et al. (2007) Daclizumab phase II trial in relapsing and
remitting multiple sclerosis: MRI and clinical results. Neurology 69,
785789
23 Montalban, X. et al. (2007) Preliminary CHOICE results: a phase 2,
randomised, placebo-controlled multicentre study of subcutaneous
daclizumab in patients with active, relapsing forms of multiple
sclerosis on interferon beta. 23rd Congress of the European
Committee for Treatment and Research in Multiple Sclerosis
(ECTRIMS), Prague, Czech Republic. Mult. Scler. 13, S18
24 Racke, M.K. (2008) The role of B cells in multiple sclerosis: rationale
for B-cell-targeted therapies. Curr. Opin. Neurol. 21 (Suppl. 1), S9
S18
25 Linington, C. et al. (1988) Augmentation of demyelination in rat acute
allergic encephalomyelitis by circulating mouse monoclonal
antibodies directed against a myelin/oligodendrocyte glycoprotein.
Am. J. Pathol. 130, 443454
26 Mathey, E.K. et al. (2007) Neurofascin as a novel target for
autoantibody-mediated axonal injury. J. Exp. Med. 204, 23632372
27 Huizinga, R. et al. (2007) Immunization with neurofilament light
protein induces spastic paresis and axonal degeneration in Biozzi
ABH mice. J. Neuropathol. Exp. Neurol. 66, 295304
28 Franciotta, D. et al. (2008) B cells and multiple sclerosis. Lancet
Neurol. 7, 852858
29 Hauser, S.L. et al. (2008) B-cell depletion with rituximab in relapsingremitting multiple sclerosis. N. Engl. J. Med. 358, 676688
30 Hawker, K. et al. (2009) Rituximab in patients with primary
progressive multiple sclerosis: results of a randomized double-blind
placebo-controlled multicenter trial. Ann. Neurol. 66, 460471
31 Bartholomaus, I. et al. (2009) Effector T cell interactions with
meningeal vascular structures in nascent autoimmune CNS
lesions. Nature 462, 9498
32 Bechmann, I. et al. (2007) What is the bloodbrain barrier (not)?
Trends Immunol. 28, 511
33 Charo, I.F. and Ransohoff, R.M. (2006) The many roles of chemokines
and chemokine receptors in inflammation. N. Engl. J. Med. 354, 610
621
34 Mullershausen, F. et al. (2009) Persistent signaling induced by
FTY720-phosphate is mediated by internalized S1P1 receptors.
Nat. Chem. Biol. 5, 428434
35 Brinkmann, V. (2009) FTY720 (fingolimod) in multiple sclerosis:
therapeutic effects in the immune and the central nervous system.
Br. J. Pharmacol. 158, 11731182
36 Brinkmann, V. (2007) Sphingosine 1-phosphate receptors in health
and disease: mechanistic insights from gene deletion studies and
reverse pharmacology. Pharmacol. Ther. 115, 84105
37 Miron, V.E. et al. (2008) FTY720 modulates human oligodendrocyte
progenitor process extension and survival. Ann. Neurol. 63, 6171
38 Kappos, L. et al. (2006) Oral fingolimod (FTY720) for relapsing
multiple sclerosis. N. Engl. J. Med. 355, 11241140
39 OConnor, P. et al. (2009) Oral fingolimod (FTY720) in multiple
sclerosis: two-year results of a phase II extension study. Neurology
72, 7379
40 Cohen, J. et al. (2010) Oral Fingolimod vs intramuscular interferon in
relapsing multiple sclerosis, New Engl J Med, E-pub (Jan 20)
41 Leypoldt, F. et al. (2009) Hemorrhaging focal encephalitis under
fingolimod (FTY720) treatment: a case report. Neurology 72, 1022
1024
42 Schulze-Topphoff, U. et al. (2008) Roles of the kallikrein/kinin system
in the adaptive immune system. Int. Immunopharmacol. 8, 155160
43 Schulze-Topphoff, U. et al. (2009) Activation of kinin receptor B1
limits encephalitogenic T lymphocyte recruitment to the central
nervous system. Nat. Med. 15, 788793
44 Platten, M. et al. (2009) Blocking angiotensin-converting enzyme
induces potent regulatory T cells and modulates TH1- and TH17mediated autoimmunity. Proc. Natl. Acad. Sci. U. S. A. 106, 14948
14953
45 Stegbauer, J. et al. (2009) Role of the reninangiotensin system in
autoimmune inflammation of the central nervous system. Proc. Natl.
Acad. Sci. U. S. A. 106, 1494214947
46 Ferguson, B. et al. (1997) Axonal damage in acute multiple sclerosis
lesions. Brain 120, 393399
47 Trapp, B.D. et al. (1998) Axonal transection in the lesions of multiple
sclerosis. N. Engl. J. Med. 338, 278285

150

Trends in Neurosciences Vol.33 No.3


48 Diestel, A. et al. (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, 17291740
49 Giuliani, F. et al. (2003) Vulnerability of human neurons to T cellmediated cytotoxicity. J. Immunol. 171, 368379
50 Aktas, O. et al. (2005) Neuronal damage in autoimmune
neuroinflammation mediated by the death ligand TRAIL. Neuron
46, 421432
51 Nitsch, R. et al. (2004) Direct impact of T cells on neurons revealed by
two-photon microscopy in living brain tissue. J. Neurosci. 24, 2458
2464
52 Sobottka, B. et al. (2009) Collateral bystander damage by myelindirected CD8+ T cells causes axonal loss. Am. J. Pathol. 175, 1160
1166
53 Gobel, K. et al. Collateral neuronal apoptosis in CNS gray matter
during an oligodendrocyte-directed CD8(+) T cell attack. Glia. (in
press), Epub (Sept 24)
54 Friese, M.A. and Fugger, L. (2009) Pathogenic CD8(+) T cells in
multiple sclerosis. Ann. Neurol. 66, 132141
55 Peterson, J.W. et al. (2001) Transected neurites, apoptotic neurons,
and reduced inflammation in cortical multiple sclerosis lesions. Ann.
Neurol. 50, 389400
56 Kutzelnigg, A. et al. (2005) Cortical demyelination and diffuse white
matter injury in multiple sclerosis. Brain 128, 27052712
57 Frohman, E.M. et al. (2005) Characterizing the mechanisms of
progression in multiple sclerosis: evidence and new hypotheses for
future directions. Arch Neurol. 62, 13451356
58 Weiner, H.L. (2009) The challenge of multiple sclerosis: how do we
cure a chronic heterogeneous disease? Ann. Neurol. 65, 239248
59 Frischer, J.M. et al. (2009) The relation between inflammation and
neurodegeneration in multiple sclerosis brains. Brain 132, 1175
1189
60 Nave, K.A. and Trapp, B.D. (2008) Axon-glial signaling and the glial
support of axon function. Annu. Rev. Neurosci. 31, 535561
61 Waxman, S.G. (2008) Mechanisms of disease: sodium channels and
neuroprotection in multiple sclerosis current status. Nat. Clin.
Pract. Neurol. 4, 159169
62 Kapoor, R. et al. (2003) Blockers of sodium and calcium entry protect
axons from nitric oxide-mediated degeneration. Ann. Neurol. 53, 174
180
63 Craner, M.J. et al. (2004) Molecular changes in neurons in multiple
sclerosis: altered axonal expression of Nav1.2 and Nav1.6 sodium
channels and Na+/Ca2+ exchanger. Proc. Natl. Acad. Sci. U. S. A. 101,
81688173
64 Werner, P. et al. (2001) Multiple sclerosis: altered glutamate
homeostasis in lesions correlates with oligodendrocyte and axonal
damage. Ann. Neurol. 50, 169180
65 Dutta, R. et al. (2006) Mitochondrial dysfunction as a cause of axonal
degeneration in multiple sclerosis patients. Ann. Neurol. 59, 478489
66 Mahad, D.J. et al. (2009) Mitochondrial changes within axons in
multiple sclerosis. Brain 132, 11611174
67 Filippi, M. et al. (2003) Evidence for widespread axonal damage at the
earliest clinical stage of multiple sclerosis. Brain 126, 433437
68 Cader, S. et al. (2007) Discordant white matter N-acetylaspartate and
diffusion MRI measures suggest that chronic metabolic dysfunction
contributes to axonal pathology in multiple sclerosis. Neuroimage 36,
1927
69 Craner, M.J. et al. (2005) Sodium channels contribute to microglia/
macrophage activation and function in EAE and MS. Glia 49, 220229
70 Black, J.A. et al. (2009) Sodium channel activity modulates multiple
functions in microglia. Glia 57, 10721081
71 Black, J.A. et al. (2007) Exacerbation of experimental autoimmune
encephalomyelitis after withdrawal of phenytoin and carbamazepine.
Ann. Neurol. 62, 2133
72 Rogawski, M.A. and Loscher, W. (2004) The neurobiology of
antiepileptic drugs. Nat. Rev. Neurosci. 5, 553564
73 Meuth, S.G. et al. (2008) TWIK-related acid-sensitive K+ channel 1
(TASK1) and TASK3 critically influence T lymphocyte effector
functions. J. Biol. Chem. 283, 1455914570
74 Friese, M.A. et al. (2007) Acid-sensing ion channel-1 contributes to
axonal degeneration in autoimmune inflammation of the central
nervous system. Nat. Med. 13, 14831489

Review
75 Franklin, R.J. and Ffrench-Constant, C. (2008) Remyelination in the
CNS: from biology to therapy. Nat. Rev. Neurosci. 9, 839855
76 Patani, R. et al. (2007) Remyelination can be extensive in multiple
sclerosis despite a long disease course. Neuropathol. Appl. Neurobiol.
33, 277287
77 Kuhlmann, T. et al. (2008) Differentiation block of oligodendroglial
progenitor cells as a cause for remyelination failure in chronic
multiple sclerosis. Brain 131, 17491758
78 Chang, A. et al. (2002) Premyelinating oligodendrocytes in chronic
lesions of multiple sclerosis. N. Engl. J. Med. 346, 165173
79 Shen, S. et al. (2008) Age-dependent epigenetic control of
differentiation inhibitors is critical for remyelination efficiency.
Nat. Neurosci. 11, 10241034
80 Mi, S. et al. (2007) LINGO-1 antagonist promotes spinal cord
remyelination and axonal integrity in MOG-induced experimental
autoimmune encephalomyelitis. Nat. Med. 13, 12281233
81 John, G.R. et al. (2002) Multiple sclerosis: re-expression of a
developmental pathway that restricts oligodendrocyte maturation.
Nat. Med. 8, 11151121
82 Jurynczyk, M. et al. (2005) Inhibition of Notch signaling enhances
tissue repair in an animal model of multiple sclerosis. J.
Neuroimmunol. 170, 310
83 Kremer, D. et al. (2009) p57kip2 is dynamically regulated in
experimental autoimmune encephalomyelitis and interferes with
oligodendroglial maturation. Proc. Natl. Acad. Sci. U. S. A. 106,
90879092
84 Schnell, L. and Schwab, M.E. (1990) Axonal regeneration in the rat
spinal cord produced by an antibody against myelin-associated
neurite growth inhibitors. Nature 343, 269272
85 Karnezis, T. et al. (2004) The neurite outgrowth inhibitor Nogo A is
involved in autoimmune-mediated demyelination. Nat. Neurosci. 7,
736744
86 Fontoura, P. and Steinman, L. (2006) Nogo in multiple sclerosis:
growing roles of a growth inhibitor. J. Neurol. Sci. 245, 201210
87 Hermanns, S. et al. (2006) The collagenous wound healing scar in the
injured central nervous system inhibits axonal regeneration. Adv.
Exp. Med. Biol. 557, 177190
88 Nait-Oumesmar, B. et al. (2007) Activation of the subventricular zone
in multiple sclerosis: evidence for early glial progenitors. Proc. Natl.
Acad. Sci. U. S. A. 104, 46944699
89 Chang, A. et al. (2008) Neurogenesis in the chronic lesions of multiple
sclerosis. Brain 131, 23662375
90 Wang, Y. et al. (2008) Paradoxical dysregulation of the neural stem
cell pathway sonic hedgehog-Gli1 in autoimmune encephalomyelitis
and multiple sclerosis. Ann. Neurol. 64, 417427
91 Wang, D. et al. (2005) Astrocyte-associated axonal damage in preonset stages of experimental autoimmune encephalomyelitis. Glia 51,
235240
92 Luo, J. et al. (2007) Glia-dependent TGF-beta signaling, acting
independently of the TH17 pathway, is critical for initiation of
murine autoimmune encephalomyelitis. J. Clin. Invest. 117, 3306
3315
93 Ridet, J.L. et al. (1997) Reactive astrocytes: cellular and molecular
cues to biological function. Trends Neurosci. 20, 570577
94 Sofroniew, M.V. (2005) Reactive astrocytes in neural repair and
protection. Neuroscientist 11, 400407
95 Studer, L. et al. (2000) Enhanced proliferation, survival, and
dopaminergic differentiation of CNS precursors in lowered oxygen.
J. Neurosci. 20, 73777383
96 Smith, J. et al. (2000) Redox state is a central modulator of the balance
between self-renewal and differentiation in a dividing glial precursor
cell. Proc. Natl. Acad. Sci. U. S. A. 97, 1003210037
97 Prozorovski, T. et al. (2008) Sirt1 contributes critically to the
redox-dependent fate of neural progenitors. Nat. Cell Biol. 10, 385
394
98 Zipp, F. and Aktas, O. (2006) The brain as a target of inflammation:
common pathways link inflammatory and neurodegenerative
diseases. Trends Neurosci. 29, 518527
99 Polman, C.H. et al. (2006) A randomized, placebo-controlled trial of
natalizumab for relapsing multiple sclerosis. N. Engl. J. Med. 354,
899910
100 Rudick, R.A. et al. (2006) Natalizumab plus interferon beta-1a for
relapsing multiple sclerosis. N. Engl. J. Med. 354, 911923

Trends in Neurosciences

Vol.33 No.3

101 Yousry, T.A. et al. (2006) Evaluation of patients treated with


natalizumab for progressive multifocal leukoencephalopathy. N.
Engl. J. Med. 354, 924933
102 Hartung, H.P. (2009) New cases of progressive multifocal
leukoencephalopathy after treatment with natalizumab. Lancet
Neurol. 8, 2831
103 Hemmer, B. et al. (2006) Central nervous system infections a
potential complication of systemic immunotherapy. Curr. Opin.
Neurol. 19, 271276
104 Berger, J.R. and Houff, S. (2009) Opportunistic infections and other
risks with newer multiple sclerosis therapies. Ann. Neurol. 65, 367
377
105 Verbeeck, J. et al. (2008) JC viral loads in patients with Crohns
disease treated with immunosuppression: can we screen for elevated
risk of progressive multifocal leukoencephalopathy? Gut 57, 1393
1397
106 Stuve, O. et al. (2006) Altered CD4+/CD8+ T-cell ratios in
cerebrospinal fluid of natalizumab-treated patients with multiple
sclerosis. Arch. Neurol. 63, 13831387
107 Ransohoff, R.M. (2007) Thinking without thinking about
natalizumab and PML. J. Neurol. Sci. 259, 5052
108 Krumbholz, M. et al. (2008) Natalizumab disproportionately
increases circulating pre-B and B cells in multiple sclerosis.
Neurology 71, 13501354
109 Zohren, F. et al. (2008) The monoclonal anti-VLA-4 antibody
natalizumab mobilizes CD34+ hematopoietic progenitor cells in
humans. Blood 111, 38933895
110 Lindberg, R.L. et al. (2008) Natalizumab alters transcriptional
expression profiles of blood cell subpopulations of multiple sclerosis
patients. J. Neuroimmunol. 194, 153164
111 Sunyaev, S.R. et al. (2009) Adaptive mutations in the JC virus protein
capsid
are
associated
with
progressive
multifocal
leukoencephalopathy (PML). PLoS Genet. 5, e1000368
112 Thompson, S.A. et al. B-cell reconstitution and BAFF after
alemtuzumab (Campath-1H) treatment of multiple sclerosis. J.
Clin. Immunol. (in press), Epub (Sep 10)
113 Bloom, D. et al. (2009) BAFF is increased in renal transplant patients
following treatment with alemtuzumab. Am. J. Transplant. 9, 1835
1845
114 Cox, A.L. et al. (2005) Lymphocyte homeostasis following therapeutic
lymphocyte depletion in multiple sclerosis. Eur. J. Immunol. 35,
33323342
115 Bielekova, B. et al. (2006) Regulatory CD56 (bright) natural killer
cells mediate immunomodulatory effects of IL-2Ra-targeted therapy
(daclizumab) in multiple sclerosis. Proc. Natl. Acad. Sci. U. S. A. 103,
59415946
116 Castillo, J. et al. (2009) Ofatumumab, a second-generation anti-CD20
monoclonal antibody, for the treatment of lymphoproliferative and
autoimmune disorders. Expert Opin. Investig. Drugs 18, 491500
117 Hutas, G. (2008) Ocrelizumab, a humanized monoclonal antibody
against CD20 for inflammatory disorders and B-cell malignancies.
Curr. Opin. Investig. Drugs 9, 12061215
118 Bechtold, D.A. et al. (2006) Axonal protection achieved in a model of
multiple sclerosis using lamotrigine. J. Neurol. 253, 15421551
119 Lucchinetti, C.F. et al. (1996) Distinct patterns of multiple sclerosis
pathology indicates heterogeneity in pathogenesis. Brain Pathol. 6,
259274
120 Lucchinetti, C.F. et al. (2002) A role for humoral mechanisms in
the pathogenesis of Devics neuromyelitis optica. Brain 125, 1450
1461
121 Lennon, V.A. et al. (2005) IgG marker of optic-spinal multiple sclerosis
binds to the aquaporin-4 water channel. J. Exp. Med. 202, 473477
122 Tait, M.J. et al. (2008) Water movements in the brain: role of
aquaporins. Trends Neurosci. 31, 3743
123 Aktas, O. and Hartung, H.P. (2009) Neuromyelitis and more: the
unfolding spectrum of aquaporin 4-related neurological diseases. J.
Neurol. 256, 19061908
124 Bradl, M. et al. (2009) Neuromyelitis optica: pathogenicity of patient
immunoglobulin in vivo. Ann. Neurol. 66, 630643
125 Barnett, M.H. et al. (2009) Immunoglobulins and complement in
postmortem multiple sclerosis tissue. Ann. Neurol. 65, 3246
126 Breij, E.C. et al. (2008) Homogeneity of active demyelinating lesions
in established multiple sclerosis. Ann. Neurol. 63, 1625

151

Review
127 Vogt, J. et al. (2009) Lower motor neuron loss in multiple sclerosis and
experimental autoimmune encephalomyelitis. Ann. Neurol. 66, 310
322
128 Schirmer, L. et al. (2009) Substantial early, but nonprogressive
neuronal loss in multiple sclerosis (ms) spinal cord. Ann. Neurol.
66, 698704
129 Steinman, L. and Zamvil, S.S. (2005) Virtues and pitfalls of EAE for
the development of therapies for multiple sclerosis. Trends Immunol.
26, 565571
130 Kornek, B. et al. (2000) Multiple sclerosis and chronic autoimmune
encephalomyelitis: a comparative quantitative study of axonal injury

152

Trends in Neurosciences Vol.33 No.3


in active, inactive, and remyelinated lesions. Am. J. Pathol. 157, 267
276
131 Derfuss, T. et al. (2009) Contactin-2/TAG-1-directed autoimmunity is
identified in multiple sclerosis patients and mediates gray matter
pathology in animals. Proc. Natl. Acad. Sci. U. S. A. 106, 83027
132 Krishnamoorthy, G. et al. (2009) Myelin-specific T cells also recognize
neuronal autoantigen in a transgenic mouse model of multiple
sclerosis. Nat. Med. 15, 626632
133 Morgen, K. et al. (2006) Evidence for a direct association between
cortical atrophy and cognitive impairment in relapsing-remitting MS.
Neuroimage 30, 891898

You might also like