You are on page 1of 27

Neuron

Review

Multiple Sclerosis: Mechanisms and Immunotherapy


Clare Baecher-Allan,1 Belinda J. Kaskow,1 and Howard L. Weiner1,*
1Department of Neurology, Ann Romney Center for Neurologic Diseases, Partners MS Center, Evergrande Center for Immunologic Diseases,

Brigham and Women’s Hospital, Harvard Medical School, Boston, MA 02115, USA
*Correspondence: hweiner@rics.bwh.harvard.edu
https://doi.org/10.1016/j.neuron.2018.01.021

Multiple sclerosis (MS) is an autoimmune disease triggered by environmental factors that act on a genet-
ically susceptible host. It features three clinical stages: a pre-clinical stage detectable only by MRI; a re-
lapsing-remitting (RRMS) stage characterized by episodes of neurologic dysfunction followed by resolu-
tion; and a progressive stage, which usually evolves from the relapsing stage. Advances in our
understanding of the immune mechanisms that contribute to MS have led to more than ten FDA-approved
immunotherapeutic drugs that target effector T cells, regulatory cells, B cells, and cell trafficking into the
nervous system. However, most drugs for relapsing MS are not effective in treating progressive disease.
Progressive MS features a compartmentalized immune response in the central nervous system, involving
microglia cells and astrocytes, as well as immune-independent processes that drive axonal dysfunction.
Major challenges for MS research involve understanding the mechanisms of disease progression, devel-
oping treatment for progressive MS, and determining the degree to which progressive disease can be
prevented by early treatment. Key priorities for MS research include developing biomarkers, personalized
medicine and advanced imaging, and a better understanding of the microbiome. With a better under-
standing of the genetic and epidemiological aspects of this disease, approaches to prevent MS are
now also being considered.

Introduction neurologic dysfunction progressively worsens, affecting, in


Multiple sclerosis (MS) is a chronic inflammatory disease of particular, a patient’s gait (Figure 1). In some patients, MS
the central nervous system (CNS) of presumed autoimmune can be progressive from onset and is termed primary progres-
etiology. Our understanding of the immune processes that sive disease. It is not known the degree to which primary pro-
contribute to MS has led to FDA-approved, disease-modifying gressive MS (PPMS) represents a unique form of MS or simply
drugs that are effective in relapsing forms of MS. However, few secondary progressive MS (SPMS), in which the relapses were
treatments are effective for progressive forms of the disease. not apparent clinically. Trials in the past 10 years demonstrate
It is now accepted that MS is a heterogeneous, multifactorial that the characteristics and behavior of RRMS cohorts in clin-
disease influenced by both genetic and environmental factors ical trials have changed and are characterized by lower clinical
(Dendrou et al., 2015). The clinical picture of MS is also better disease activity. This might possibly reflect a change in the nat-
defined and continues to evolve in the current treatment era. In ural history of MS (Uitdehaag et al., 2011). In addition, more
addition, magnetic resonance imaging (MRI) has provided a ba- recent natural history studies report longer times to disability
sis for understanding different stages of MS and for developing milestones, with the median times to requiring a cane to
treatments. Finally, with advances in our understanding of both walk, ranging from 15 to 32 years (Tremlett et al., 2010). With
immune and neurobiological processes, new treatment ap- MRI imaging and our better understanding of disease mecha-
proaches are being developed. The aim of this Review is to nisms, the clinical classification of MS has evolved to include
provide a detailed description of the pathogenesis of MS and the concept of ‘‘active’’ disease, which is associated with
the current directions in research and treatment of the disease. both the relapsing and the progressive forms of the disease.
A summary of our current understanding of MS is given in This classification reflects the inflammatory component of
Table 1. MS, which can be visualized by MRI using gadolinium
enhancement (Lublin et al., 2014).
MS Pathology and Risk Factors The MS Pathological Process
MS broadly comprises three stages: (1) a pre-clinical stage, in The MS pathological process includes breakdown of the blood-
which a combination of genetic and environmental factors brain barrier, multifocal inflammation, demyelination, oligoden-
trigger the disease; (2) a relapsing-remitting (RRMS) clinical drocyte loss, reactive gliosis, and axonal degeneration (Trapp
stage, which features discrete, self-limited episodes of neuro- and Nave, 2008). Four neuropathological patterns have been
logic dysfunction, such as optic neuritis, sensory disturbances, described for MS based on the presence or absence of comple-
or disturbances in motor and cerebellar function (this inflamma- ment and immunoglobulins, apoptotic nuclei, and/or preferential
tory stage may begin sub-clinically and be evident solely by loss of myelin protein (Lucchinetti et al., 2000). This heterogene-
MRI imaging); and (3) a progressive clinical stage during which ity in lesion pattern is observed between patients, but not within

742 Neuron 97, February 21, 2018 ª 2018 Elsevier Inc.


Neuron

Review

Table 1. Mechanisms of Multiple Sclerosis


Feature Disease Mechanism
Autoimmunity MS is a cell-mediated autoimmune disease directed against CNS myelin antigens that involves both CD4+ and
CD8+ cells. Autoantibodies may play a secondary or enhancing role. Autoreactive T cells against myelin components
exist in normal individuals and, in these individuals, do not cause disease and may even have brain-protecting
properties. MS is induced when pathogenic Th17- and Th1-type and CD8 myelin autoreactive T cells are induced.
Infection Infectious agents play a crucial role in inducing myelin-reactive pathogenic T cells. Potential mechanisms include
cross reactivity with CNS myelin antigens, triggering an already expanded autoreactive immune repertoire or a
self-limited infection of the brain that releases myelin antigens. MS is not caused by a persistent viral infection
of the brain or transmissible agent.
Genetics Both MHC and non-MHC genes are risk factors for development of MS. MHC genes determine immune repertoire
whereas non-MHC genes determine regulatory and tolerance mechanisms in MS, both of which are defective.
Environment Environmental factors can increase the risk for both the development of MS and the disease course and include
lowered vitamin D, lowered UV radiation exposure, cigarette smoking, obesity, and EBV exposure.
B cells B cells play a central role in MS. Like T cells, there are pro- and anti-inflammatory B cell subsets. In relapsing MS,
B cells serve as primary antigen-presenting cells that drive pathogenic T cells. In progressive MS, B cells enhance
the compartmentalized CNS responses through lymphoid follicles and secreted factors.
Microbiome The microbiome regulates T cell function throughout the body and contains both protective and pathogenic
microbial components, which play an important role in MS by establishing immune set points and by secreting
metabolites.
Relapsing MS Relapsing MS is driven by immune cells that migrate into the CNS. Multiple treatments have been shown to
effectively treat relapsing MS (decrease relapses and new MRI lesions) and act on the following common
pathways: decrease number and/or function of effector cells, increase number and/or function of regulatory cells,
and prevent trafficking of cells to the CNS.
Progressive MS Progressive MS mechanisms include those that are immune dependent and immune independent. In
immune-dependent forms, an innate immune response is established in the brain that involves microglia,
macrophages, B cells, and lymphoid follicles. There may also be chronic activation of peripheral T cells and
innate cells. In immune-independent forms, mitochondrial injury, oxidative stress, and ion imbalance occur.
Current therapy does not effectively target these two processes.
Autoantigen The inciting autoantigen in MS is unknown. However, when MS is diagnosed, there would be no single autoantigen
to target as there is spreading of reactivity to other organ-specific antigens as occurs in type 1 diabetes. Thus,
antigen-specific therapy would have to employ bystander suppression or be given as a preventative strategy in at
risk subjects.
Therapy MS is a heterogeneous disease. There will be responders and non-responders to each ‘‘effective’’ therapy.
The earlier in the disease course that treatment is initiated, the more likely it is to be effective. Effective treatment
will require pulse or continuous therapy and, ultimately, combination therapy. The identification of immune and
MRI biomarkers will be the cornerstone of immunotherapy of MS and the achievement of no evidence of disease
activity (NEDA).

patients. Active lesions are common in relapsing MS patients abundant in progressive stages of MS, and microglial activation
but become rare during progressive MS (Trapp et al., 1999). is more prominent than peripheral immune cell infiltration (Lass-
Transected axons in acute MS lesions are abundant and occur mann, 2012). It has been suggested that meningeal infiltrates
approximately 12 times more frequently than in chronic lesions activate microglia and that brains with increased meningeal infil-
(Kornek and Lassmann, 1999). Thus, axonal loss occurs at tration have increased B-cell-like follicles (Magliozzi et al., 2010).
disease onset and is a continuous process. The conversion of Both macrophages and microglia accumulate at sites of
RRMS to SPMS is therefore thought to occur when the brain ex- active demyelination and neurodegeneration in MS. Using newly
hausts its capacity to compensate for further axonal loss (Nave defined cell surface markers, it has been possible to distinguish
and Trapp, 2008). infiltrating macrophages from resident microglia (Butovsky et al.,
Although MS has historically been considered a demyelinating 2014). These markers have also shown that microglial activation
disease of the CNS and white matter, in recent years, demyelin- is more pronounced in the normal-appearing white matter of
ation of the cortical and deep gray matter have been recognized MS patients as compared to controls and increases with disease
and may exceed white matter demyelination (Kutzelnigg et al., duration. Furthermore, as measured by the P2RY12 marker,
2005). Cortical atrophy has been associated with disease pro- there is a loss of the homeostatic microglial phenotype in MS,
gression and with gray matter atrophy, and it is among the which reappears in inactive lesions (Zrzavy et al., 2017).
best predictors of neurological disability in MS (Fisniku et al., Genetic Components
2008). Inflammatory cortical demyelination has also been identi- MS has a genetic component as evidenced by first-degree rel-
fied in early MS (Lucchinetti et al., 2011). Subpial lesions are atives having a higher risk for MS (2%–5% risk), by a high

Neuron 97, February 21, 2018 743


Neuron

Review

Figure 1. The Stages of Multiple Sclerosis


Multiple sclerosis is thought to begin before clinical symptoms are evident, and it can be discovered incidentally on MRI as a radiologically isolated
syndrome (RIS). It then usually manifests as a clinically isolated syndrome (CIS), which is followed by a relapsing-remitting stage, characterized by discrete
episodes of neurologic dysfunction with remission. The progressive stage involves steadily worsening disability and usually evolves from the relapsing-
remitting stage, although some patients may have progressive disease from onset (primary progressive MS). MRI correlates of disease include gadolinium-
enhancing lesions, which represent the breakdown of the blood-brain barrier and the movement of cells into the CNS, accumulation of T2 burden of
disease, and a decrease in brain volume as measured by atrophy. Many treatments impact on relapsing forms of MS. Few treatment options are available
for progressive MS.

disease concordance rate in monozygotic (MZ) twins (25%), bined OR 5. Patients with DRB1*15:01 develop more lesions
and by a difference in disease prevalence among different (T2 foci) on MRI and a decrease in brain volume (Okuda et al.,
ancestral groups (Hollenbach and Oksenberg, 2015; Sawcer 2009), whereas those with B*44 have reduced T2 lesion load
et al., 2014). MS is rare in black Africans although it occurs in Af- and preserved brain volume (Healy et al., 2010). Virtually all
rican Americans (Hollenbach and Oksenberg, 2015). Genetics MS-associated genes regulate either innate or adaptive immu-
contributes to geographic variations in MS but cannot explain nity. It is likely that HLA variants are related to disease speci-
differences in risk among people of common ancestry, who ficity, whereas non-HLA variants influence broad aspects of im-
migrate to areas of high or low MS prevalence (as discussed mune activation and tolerance. Although some of the genetic
below). There is no evidence of Mendelian inheritance. The associations linked to MS are found in other autoimmune dis-
HLA-DRB1*15:01 allele in the class II major histocompatibility eases, few are shared with other neurodegenerative diseases,
complex (MHC) (odds ratio [OR] 3) is the earliest, and most such as Alzheimer’s and Parkinson’s diseases. Finally, although
dominant, risk factor identified in MS (Sawcer et al., 2005). Sub- the clinical features of the disease may have a genetic compo-
sequent genome-wide association studies (GWASs) have found nent in some families with MS (Hensiek et al., 2007) and
that the interleukin-2 receptor alpha gene (IL2RA) and inter- although the increased disease severity observed in African
leukin-7 receptor alpha gene (IL7R) are also inheritable risk fac- Americans may have a genetic component, disease severity
tors (Hafler et al., 2007). These studies have identified a total of and disease type do not appear to have a strong genetic
110 polymorphisms in 103 discreet loci outside the MHC that component. Even in the case of affected MZ twins, one twin
are associated with susceptibility (Sawcer et al., 2014). The can have mild relapsing disease while the other has disabling
class I variants HLA-A*02 (OR 0.6) and HLA B*44 are associ- progressive disease. It thus appears that the primary genetic ef-
ated with protection from disease (Sawcer et al., 2014). The fect in MS relates to disease susceptibility rather than to dis-
presence of DRB1*15:01 and absence of HLA-A*02 has a com- ease course or severity.

744 Neuron 97, February 21, 2018


Neuron

Review

Environmental Factors of MS in individuals who have infectious mononucleosis or


It is now recognized that environmental and lifestyle factors elevated antibody titers against EBV nuclear antigen 1
interact with genetics to influence both MS susceptibility and (EBNA1). There is also a higher prevalence of EBV infection in in-
the course of the disease. Importantly, some environmental fac- dividuals with MS relative to controls (Plavina et al., 2014). As
tors are modifiable. The earliest clues to environmental influ- with other environmental factors, only EBV infection during
ences came from studies into the effect of latitude and migration. adolescence or later increases the risk of MS. It is not clear
It has long been recognized that there is a latitudinal gradient in how EBV infection could contribute to MS. It could have a spe-
MS; it is more prevalent the further one is from the equator (Simp- cific effect, through a mechanism such as molecular mimicry,
son et al., 2011). Furthermore, migration studies have shown or general immune effects on B cells or on other immune regula-
that migrants who move from a high-risk to a low-risk country tory elements, or it could be a secondary phenomenon. Of note,
before adolescence have a decreased disease rate, whereas EBV infection is also reported to be a risk factor for SLE (Harley
those that move from a low-risk to a high-risk country acquire and James, 2006).
a higher risk (Gale and Martyn, 1995; Ahlgren et al., 2012). This No specific diet is associated with an increased risk of MS.
strongly suggests that exposure to an environmental factor in The role of dietary factors appears to be complex and related
the first two decades of life can influence disease risk. Further- to the influence of multiple dietary components on immune
more, the incidence of MS appears to be increasing in women regulation, including vitamins A and D, salt, omega-3-unsatu-
in certain areas of the world that were previously areas of low rated fatty acid, and polyphenol. Recent reports have sug-
MS prevalence. gested that salt modulates the differentiation of human and
Vitamin D and ultraviolet (UV) radiation are two environmental mouse Th17 cells (Wu et al., 2013). A more aggressive course
factors known to be associated with MS susceptibility that might of experimental autoimmune encephalomyelitis (EAE), which
also help to explain the latitude effect. Increased vitamin D levels, serves as the animal model for MS, was observed in mice
especially before the age of 20, are associated with a reduced fed a high-sodium diet (Wu et al., 2013). In addition, in a small,
risk of MS in later life (Munger et al., 2006). UV radiation also pro- 2-year observational study of 122 patients with MS, a higher
tects against MS, probably both through its effects on vitamin D sodium intake was associated with increased MS disease ac-
and through independent effects on the immune system (Lucas tivity, as assessed clinically and radiologically (Farez et al.,
et al., 2015). Vitamin D has broad effects on the immune sys- 2015). However, a 5-year longitudinal study (of 465 MS pa-
tem, including the suppression of both B cell and T cell prolifer- tients) found no association between salt intake and either
ation, skewing T cells away from inflammatory responses and the rate of CIS conversion to MS or the severity of disease (Fitz-
toward regulatory T cell (Treg) responses, and promoting tolero- gerald et al., 2017), and it was not associated with a reduced
genic monocyte and dendritic cell phenotypes (Aranow, 2011). time to relapse in pediatric MS (Nourbakhsh et al., 2016).
Vitamin D is not only linked to MS, but also to other autoimmune With regard to whether pregnancy affects the risk for MS, it
diseases, including rheumatoid arthritis (RA), type 1 diabetes has been found that giving birth, or having a terminated preg-
(T1D), and systemic lupus erythematous (SLE). It is now common nancy, protects women from developing MS over the next 5
practice to provide vitamin D supplementation to all MS patients years (Hedström et al., 2014).
(Stein et al., 2011). Of note, there is also evidence that people The Role of the Microbiome
with lower levels of vitamin D have increased infections (Kalluri The human microbiome encompasses trillions of organisms that
et al., 2017). inhabit the gut and that shape the gut-associated lymphoid tis-
Cigarette smoking is another well-recognized environmental sue. It is now known that the gut microbiome contributes to
factor that is linked to both MS risk and disease activity (Wing- health and disease, influencing infection, atherosclerosis, can-
erchuk, 2012). Smoking appears to mediate its effect by driving cer, autoimmunity, and inflammatory and metabolic disease
T-helper 17 (Th17)-type responses in the lung. Oral tobacco (Lynch and Pedersen, 2016). Recent studies have shown that
doesn’t have this effect (Hedström et al., 2011). As with other the gut microbiome participates in immunoregulatory pathways
risk factors, smoking appears to influence susceptibility to that can both contribute to, and protect individuals from, dis-
MS in conjunction with genetic factors and with other environ- ease. In a spontaneous, relapsing EAE model, animals raised un-
mental factors. As the environment is being more systemati- der germ-free conditions were resistant to disease, while those
cally investigated, other factors are being identified that in- exposed to commensal bacteria developed EAE (Berer et al.,
crease disease risk, including adolescent obesity, exposure 2011). It has also been shown that segmented filamentous bac-
to organic solvents, and night shift work. By contrast, oral to- teria in the gut can induce Th17 cells (Ivanov et al., 2009) and that
bacco use, cytomegalovirus virus infection, alcohol, and coffee pro-inflammatory T cell responses to gut microbiota promote
consumption are associated with decreased risk (Olsson et al., EAE (Lee et al., 2011). Conversely, Tregs can be induced in the
2017). All of these factors most likely impact on the immune gut by commensal gut bacteria (Round and Mazmanian, 2010),
system. and in the EAE model, polysaccharide A from the human
Although it has long been suspected that an infectious agent is commensal Bacteroides fragilis protects against EAE (Ochoa-
at the heart of MS, no MS infectious agent has been identified. Repáraz et al., 2010). Thus, the gut microbiome has both
Indeed, the first drug approved for treatment of MS, beta inter- disease-inducing and disease-protecting influences on CNS
feron, was tested based on the hypothesis that MS was due to autoimmunity. Moreover, it has been recently shown that trypto-
a viral infection. Nonetheless, there is a link between Epstein- phan-derived aryl hydrocarbon receptor (AHR) ligands produced
Barr virus (EBV) infection and MS as there is an increased risk by intestinal commensals can reach the CNS to modulate

Neuron 97, February 21, 2018 745


Neuron

Review

astrocyte function and suppress inflammation and neurodegen- genic microorganisms, or fecal replacement therapy could serve
eration (Rothhammer et al., 2016). Thus, the gut microbiome as a physiological, non-toxic form of therapy and might also aid
participates in immunoregulatory pathways to modulate CNS disease prevention.
autoimmunity.
Studies have also investigated the gut microbiome of MS The Adaptive Immune System in MS
patients. We have found the bacterial species Methanobrevi- The success of clinical trials in MS that target immune mole-
bacter and Akkermansia to be increased and the bacteria Bu- cules or specific cell types both reinforces the idea that the im-
tyricimonas to be decreased in patients with MS relative to mune system plays a central role in MS and provides strong ev-
control subjects (Jangi et al., 2016). Patients on disease-modi- idence for the involvement of particular immune pathways in
fying treatment also show increased abundances of Prevotella the pathogenesis of the disease. These data have suggested
and Sutterella and decreased levels of Sarcina compared to that B cells and different effector T cell populations of the adap-
untreated MS patients. Immune profiling of peripheral blood tive immune system, together with natural killer (NK) and micro-
from these subjects demonstrated an association between glial cells of the innate immune system, play unique roles in
gene expression in T cells and monocytes with bacterial spe- contributing to the disease. In addition, not only have the in-
cies in the gut. Other MS studies have found that perturba- flammatory lesions within the CNS white matter been reported
tions in the gut microbiota were associated with parallel pre- to contain CD4 and CD8 T cells, but also the meninges in pro-
dicted enrichment of metabolic pathways associated with gressive MS to contain ectopic germinal centers that include B
neurodegeneration in recent-onset pediatric MS (Tremlett cells and other immune populations. Studies into how these
et al., 2016), exhibited dysbiosis with depletion of major buty- effector or regulatory cell types function in healthy donors, as
rate producers belonging to Clostridial clusters XIVa and IV compared to in patients with MS, shed light on how their func-
clusters (Miyake et al., 2015), as well as changes in gut micro- tional dysregulation might contribute to disease. As such, we
biota that are associated with glatiramer acetate and vitamin D discuss below our current understanding of the role that
supplementation (Cantarel et al., 2015), and a distinct microbi- distinct populations of effector T cells, regulatory cells, and B
al community profile that is consistent with gut microbial dys- cells might play in MS.
biosis (Chen et al., 2016). Recently, it was also shown that a Effector T Cells in MS
high frequency of intestinal Th17 cells correlates with microbi- MS is believed to be caused by the activation of peripheral au-
al alterations and disease activity in MS (Cosorich et al., 2017), toreactive effector CD4 T cells that migrate into the CNS and
that gut bacteria from MS individuals modulate human T cells initiate the disease process (Figure 2). This has been referred
and worsen EAE when transplanted to germ-free mice (Baran- to as the ‘‘outside-in hypothesis’’ of MS as opposed to the ‘‘in-
zini et al., 2017), that gut microbiota from unaffected twins side-out hypothesis, in which MS begins with an initial event in
promoted EAE and was associated with less IL-10 production the CNS (Malpass, 2012). Once in the CNS, autoreactive
in recipient mice (Berer et al., 2017), and that the human gut- effector CD4 T cells are locally reactivated by antigen-present-
derived commensal Prevotella suppresses EAE in an HLA ing cells (APCs) and recruit additional T cells and macrophages
class II transgenic mouse model (Mangalam et al., 2017). to establish the inflammatory lesion. MS lesions contain CD8
Although these reports do not establish causation, they sug- T cells, which are mostly found at the edges of lesions, and
gest that the microbiome plays an important role in MS and CD4 T cells, which are found deep in the lesions (Traugott
provide new avenues for the investigation of the microbiome et al., 1983; Hauser et al., 1986). These cells cause myelin
and MS. loss, oligodendrocyte destruction, and axonal damage, leading
How could the microbiome be related to MS? One possibility to neurologic dysfunction. In parallel, immune-modulatory net-
is that molecular mimicry exists between a gut organism and works are triggered to limit inflammation and to initiate repair,
structures in the CNS. This has been observed for Staphylo- which results in at least partial remyelination and is associated
coccus and Syndenham’s chorea. Also in neuromyelitis optica, with clinical remission. Nonetheless, without treatment, further
the immunodominant epitope of aquaporin-4 cross reacts with attacks that ultimately lead to the progressive form of the dis-
a homologous sequence found in the adenosine triphosphate- ease occur. Our current understanding of how effector CD4
binding cassette transporter of Clostridium perfringens (Cree and CD8 T cells promote the disease process is dis-
et al., 2016). Thus, it is possible that the gut holds the key to cussed below.
the long-sought after infectious agent of MS. Another possibility Effector CD4 T Cells
is that gut dysbiosis associated with diet and the environment CD4 T cells are found deep within CNS lesions and in the cere-
could affect immune regulation, promoting a less tolerant state brospinal fluid (CSF) of patients with MS (Traugott et al., 1983).
to CNS autoantigens. The hygiene hypothesis suggests that a Evidence that myelin-reactive T cells participate in the inflamma-
lack of early childhood exposure to infectious agents, symbiotic tory demyelination in MS includes a trial in which MS patients
microorganisms (such as the gut microbiota or probiotics), and were treated with an altered peptide ligand of myelin basic pro-
parasites increases an individual’s susceptibility to allergic and tein (MBP) (Martin et al., 2000). Although this treatment shifted
autoimmune diseases by suppressing the natural development the MBP response from pathogenic to tolerogenic in mouse
of the immune system and causing defects in the establishment models, in humans, treatment resulted in disease exacerbation,
of immune tolerance. A third possibility is that current immuno- as evidenced by an increase in contrast-enhancing lesions
modulatory therapy might partly act through the gut. The and up to a 2,000-fold expansion of circulating MBP-specific
modulation of the gut microbiome through probiotics, tolero- CD4 T cells. The strongest genetic risk factor for development

746 Neuron 97, February 21, 2018


Neuron

Review

Figure 2. Effector T Cells in Multiple Sclerosis


Peripherally activated T cells cross the blood-brain barrier (BBB) into the central nervous system (CNS), where they are re-activated and secrete cytokines to exert
their effector functions. T helper (Th)-1 cells produce their lineage-defining cytokine, IFNg, as well as TNFa, while Th17 cells secrete their defining cytokine IL-17,
as well as IL-21 and IL-22, and can also express IFNg, which contributes to their pathogenicity. CD8+ effector T cells (Teffs) can also be a source of IL-17 and
IFNg. This cytokine secretion leads to the activation of CNS-resident immune cells (such as microglia, astrocytes, and macrophages), as well as to the production
of cytokines, increased antigen-presenting cell (APC) function, and the enhanced production of reactive oxygen species (ROS) and reactive nitrogen species
(RNS). CD8 T cells can also release cytolytic granules causing axonal dissection. Effector T cells can be regulated in the periphery or in the CNS by FoxP3+
regulatory T cells (Tregs) and by Tr1 cells, CD8 Tregs, natural killer cells (NK cells), and regulatory B (Breg) cells.

of MS is the MHC class II locus DR2 (DRB*1501/DQ6). This Interferon gamma (IFNg)- and interleukin-17 (IL-17)-
class II allele itself regulates the activation of CD4 T cells. This secreting CD4 T cells are believed to be the pathogenic initia-
is because proteins encoded by each MHC allele contain charge tors of MS. In the late 1980s, CD4 T cells were first separated
and structural features that restrict the type of peptide that into the functionally distinct IFNg-producing Th1 cells, which
can be bound and presented to activate T cell receptor (TCR) clear extracellular pathogens, and the IL-4-producing Th2
molecules on mature peripheral T cells (Hemmer et al., 2000). cells, which mediate allergic responses (Tada et al., 1978).
MHC class II proteins can also affect thymic selection to shape Subsequent studies then identified CD4+ Th17 cells, which
the TCR repertoire and the self-tolerance of thymic emigrant play a central role in autoimmunity (Bettelli et al., 2006). Prior
T cells (Klein et al., 2014). Mice that express both MS-associated to the identification of IL-17 cells, whose secretion is stabilized
HLA-DR and MS-patient-derived TCRs develop spontaneous by IL-23, IFNg-producing Th1 cells induced by IL-12 were
EAE (Madsen et al., 1999). Some have argued that these obser- believed to be critical mediators of EAE because EAE did
vations contradict a trial in which MS patients underwent CD4 not occur in the absence of the IL-12 p40 subunit. However,
T cell depletion to no clinical benefit (van Oosten et al., 1997). this interpretation was altered upon finding that p40 is also a
However, it is important to note that anti-CD4 therapy would component of IL-23 and that genetic deletion of IFNg actually
also have depleted Treg and Tr1 CD4+ regulatory populations enhanced the severity of EAE (Becher et al., 2002). Yet, the
that were unknown at the time but are now known to limit the role of IL-23 in potentiating MS is less well supported. This
activity of autoimmune T cells. Furthermore, depleting CD4 is because a trial of the drug ustekinumab, which blocks the
T cells may not ameliorate disease as it would not affect CD8 common p40 subunit in IL-12/IL-23, failed to show a clinical
T cells, which make up the majority of CNS-resident T cells in effect in multiple sclerosis (Segal et al., 2008). Thus, although
patients and may play a crucial role in the disease once it is IL-17 is believed to a central mediator of autoimmunity in EAE,
initiated by CD4 T cells. with IFNg possibly being protective in mice (Berghmans et al.,

Neuron 97, February 21, 2018 747


Neuron

Review

2006; Furlan et al., 2001), the increased production of either infiltrate and at the edge of CNS lesions (Traugott et al., 1983).
IFNg or IL-17 is associated with human MS (Lock et al., We found that perivascular cuffs at the edge of active demy-
2002). A pathogenic role for IFNg in MS is supported by a trial elinating plaques in progressive MS contain up to 50 times
in 1987 in which the administration of IFNg to MS patients more CD8+ than CD4+ T cells (Hauser et al., 1986), while the
exacerbated their disease (Panitch et al., 1987a). A role for CD4+:CD8+ ratio in the CSF ranges from 3:1 to 6:1 and in
IL-17 is supported by a double blind trial in which the admin- the peripheral blood is 2:1 in MS patients (Friese and Fugger,
istration of an anti-IL-17 compound to MS patients reduced 2005). CD8 T cells are also often present in disease progres-
lesion formation, as assessed by MRI (Havrdová et al., sion-associated cortical plaques (Losy, 2013). A portion of
2016). Thus, both IL-17 and IFNg appear to play a role in the CD8 T cells isolated from the CNS of patients with MS exhibit
human disease, which is consistent with adoptive transfer evidence of oligoclonal expansion, indicating that these oligo-
EAE experiments that have shown that both Th1 and Th17 clonal cells have been amplified via antigen-specific re-
cells can independently induce CNS autoimmunity with sponses (Jacobsen et al., 2002). These oligoclonally
distinct patterns of EAE histology (Luger et al., 2008; Lee expanded CD8 T cells were shown to exist not only in the
and Goverman, 2013). CNS, but also in the CSF and blood in a small number of
IL-17-producing CD4 T cells can be pathogenic or non-path- tested MS patients. Strikingly, it was also shown that some
ogenic (Gaublomme et al., 2015). In contrast to non-encephali- of these expanded CD8 T cells persisted in the blood and
togenic Th17 cells, murine pathogenic Th17 cells express IL-21, CSF for several years.
which regulates B cell activation and plasma cell generation; CD8 T cells recognize peptides of endogenous intracellular
IL-22, which augments inflammation via STAT3 activation; proteins presented in the context of MHC class I molecules
and pro-inflammatory TNFa. They also exhibit decreased and kill cells via a cell-contact-mediated process involving
expression of the anti-inflammatory cytokine IL-10. Importantly, the activities of granzyme A (GzmA) and granzyme B (GzmB)
encephalitogenic IL-17-secreting T cells are also believed to (Barry and Bleackley, 2002). While MHC class I is ubiquitously
co-produce IFNg and IL-17, which may function early in disease and constitutively expressed on all cells, oligodendrocytes,
to activate resident microglia (Murphy et al., 2010). Pure popu- astrocytes, and neurons in patients with disease activity
lations of Th1/17 cells display gene signatures that show exhibit enhanced MHC class I and class II expression (Gobin
marked similarity to mouse pathogenic Th17 cells as opposed et al., 2001). The identity of pathogenic class I-restricted anti-
to nonpathogenic Th17 cells (Lee et al., 2012b). In MS-pa- gens remains elusive; however, antigen-presenting microglial
tient-derived Th1/17 cells, our group found IL-10 expression cells have the capacity to cross present exogenous antigens
to be reduced and CXCR3 expression to be elevated, while on their MHC class I molecules, potentially leading to the
Th17 cells expressed higher levels of IL-10 in stable versus high frequency of myelin-reactive CD8 T cells that has been
active patients (Hu et al., 2017). Similarly, libraries generated reported in patients with MS (Zang et al., 2004). These
from patient- and healthy donor-derived CCR6+ myelin reactive myelin-reactive CD8 T cells secrete IFNg and kill cells that ex-
T cells demonstrated that patient cells expressed more IFNg, press endogenously produced myelin. The cytotoxic function
IL-17, and GM-CSF and less IL-10 than did the cells of healthy of effector CD8 T cells might play a central role in axonal dam-
donors (Cao et al., 2015). While these data strongly implicate age, as their intracellular lytic granules are directionally posi-
IFNg and IL-17 as being pathological in MS, it is important tioned toward nearby axons in immunohistochemical analysis
to note that IL-17 can also be secreted by non-T cells, such of postmortem CNS tissue slices. Axonal injury has been
as NK cells and astrocytes (Cua and Tato, 2010; Tzartos correlated with the presence of lesional CD8 T cells in close
et al., 2008). apposition to neurons (Melzer et al., 2009), and enhanced
How IFNg and IL-17 initiate or augment disease is not levels of GzmA and GzmB have been found in the CSF of pa-
completely understood. IFNg increases the expression of HLA tients with active disease flares as compared to those in
class I and class II molecules (Keskinen et al., 1997) and en- remission (Malmeström et al., 2008). Pathogenic CD8 T cells
hances the antigen-presenting capabilities of macrophages, might also contribute to pathology by secreting IFNg and
astrocytes, and microglia (Kato and Suzumura, 2003; Mouzaki IL-17 (Huber et al., 2013). These IFNg-, IL-17-, and GzmB-pro-
et al., 2015). It is less clear how IL-17 mediates autoimmunity, ducing effector CD8 T cells also potentially undergo enhanced
as responses to IL-17 are pleiotropic and the receptors for endothelial transmigration in a blood-brain barrier model with
IL-17 are ubiquitously expressed, with IL-17-induced re- human cells and in mouse models (Larochelle et al., 2015;
sponses differing depending upon cell type (Zhang et al., Duan et al., 2013). Thus, CD8 T cells may not only cause oligo-
2013; Gu et al., 2013). However, in addition to inducing the dendrocyte death and neuronal damage, but they may also
release of a host of pro-inflammatory cytokines, it is believed potentiate IFNg- and IL-17-mediated pathology once inside
that IL-17 plays a crucial role in the breakdown of the blood- the CNS.
brain barrier by inducing reactive oxygen species within endo- Distinct populations of CD8 T cells may play unique roles in
thelial cells (Huppert et al., 2010). MS. A novel subset of CD8 T cells that are defined by the sur-
Effector CD8 T Cells face expression of CD161 is reportedly elevated in the circula-
Although much research has focused on the encephalitogenic tion of patients with MS and contains all IL-17-producing CD8
role of CD4 T cells in MS, it appears that CD8 T cells also play T cells (Annibali et al., 2011). Within this population reside the
a significant role in human MS. In MS, in contrast to EAE, CD8 potentially functionally distinct, mucosal-associated-invariant
T cells make up the majority of T cells in the CNS perivascular CD8 T cells (MAIT cells) that exhibit restricted TCR usage,

748 Neuron 97, February 21, 2018


Neuron

Review

expressing TCR Va2 (Billerbeck et al., 2010). In contrast, reports suppression (Schadenberg et al., 2011), as high numbers of
have suggested that populations of regulatory CD8 T cells non-suppressing Tregs have been found to co-exist with pro-
might also play a role in suppressing the immune response inflammatory T cells in autoimmune target tissues at peak dis-
and disease activity. These regulatory CD8 T cell subsets ease in both EAE in the CNS and in RA in the synovium (O’Connor
have been identified via a unique expression of molecules, et al., 2007; Cao et al., 2003).
including CD122 (IL-2Rb chain) (Lu et al., 2016), CD28 (Vudda- Tregs from MS patients have a reduced suppressive capacity
malay and van Meerwijk, 2017), CD103 (Uss et al., 2006), or as measured by the co-culture of MS-derived Tregs and
HLA-G (Lozano et al., 2009), and have been proposed to func- effector T cells (Teffs) (Viglietta et al., 2004; Haas et al., 2005;
tion via distinct mechanisms. In summary, it appears that Costantino et al., 2008). Most studies, however, have not found
different subsets of CD8 T cells play a pathogenic or beneficial a difference in the frequency of Tregs in the blood, suggesting
role in MS. a qualitative versus a quantitative abnormality. Although
Regulatory Cells in MS increased Treg function has been proposed as a therapy for
Subsets of T cells that modulate immune activation and control MS, such approaches do not address whether the Treg defect
the development of autoimmunity have been identified (Sakagu- observed in MS is related to the resistance of effector cells to
chi et al., 2001). At least two subsets of CD4+ regulatory T cells Treg suppression or to an inherent defect in the Tregs them-
have been identified and studied in the context of MS. Tregs are selves. Recent studies of Tregs and Teffs isolated from patients
a subset of regulatory T cells that express the transcription factor with a number of autoimmune diseases, including MS, have
Forkhead box protein 3 (FoxP3), as well as a myriad of inhibitory demonstrated that patient-derived Teff are, in fact, resistant
immune checkpoint surface molecules that contribute to their to Treg-mediated suppression (Schneider et al., 2013; Bhela
capacity to suppress T cell proliferation in vitro via a cell-con- et al., 2015). Thus, published studies suggest that in MS there
tact-mediated mechanism (Lu and Rudensky, 2009). A second exists both a defect in Tregs and a resistance to Treg suppres-
type of CD4+ regulatory T cell is the Tr1 regulatory CD4+ T cell sion by Teff cells.
that inhibits cell proliferation primarily via the secretion of IL-10 Inflammatory molecules expressed in autoimmune diseases
(Gregori et al., 2012). Each of these Tregs exhibit unique charac- can negatively affect Treg suppression. The pro-inflammatory
teristics during autoimmune neuro-inflammation, and both are cytokine IL-6, which exacerbates EAE, can alter CD4 T cell dif-
believed to be important in MS. NK cells also have regulatory ferentiation and promote the generation of Th17 cells at the
properties and suppress immune responses via their capacity expense of Tregs (Kimura and Kishimoto, 2010; Becher and
to kill activated, and potentially pathogenic, CD4 T cells. Daclizu- Segal, 2011). IL-6 can also directly inhibit Treg suppression
mab has been shown to enhance CD56hi NK regulatory activity in co-culture experiments via its ability to augment the secre-
and to be effective in relapsing forms of MS (Bielekova et al., tion of GzmB, which contributes to Teff cell resistance to Treg
2006). CD8 regulatory T cells may also exist and are dis- suppression (Bhela et al., 2015). Thus, while clinical trials to
cussed above. target IL-6 in MS have been proposed, therapeutically block-
CD4 FoxP3+ Tregs ing IL-6R (for example, with the compound tocilizumab) has
FoxP3+ Tregs, which make up less than 4% of circulating CD4 been found to induce MS-like lesions in RA patients (Coma-
T cells, are referred to as ‘‘professional’’ suppressor cells as bella, 2016). Similarly to IL-6, the pro-inflammatory cytokine
they exert a dominant inhibition of the activation of other cell TNFa inhibits Treg suppression, but unlike IL-6, TNFa
types via a process that requires cell contact (Baecher-Allan concomitantly induces Tregs to undergo proliferation (Valencia
et al., 2001; Schmidt et al., 2012). In contrast to the non-self- et al., 2006). Surprisingly, anti-TNFa therapy worsened MS
reactivity of effector T cell TCRs, the TCR repertoire of Tregs is and unmasked MS in arthritis patients being treated with
selected to recognize self-antigens (Pacholczyk and Kern, anti-TNFa (Kaltsonoudis et al., 2014) even though blocking
2008; Lee et al., 2012a). Accordingly, Tregs are activated by TNFa reduces disease in patients with psoriasis and RA (van
self-antigens. FoxP3 is absolutely required for their suppressive Oosten et al., 1996).
activity as the loss of a functional, X-linked FoxP3 gene in male CD4+ Tr1 Regulatory Cells
children born with immunodysregulation polyendocrinopathy Tr1-type regulatory T cells were first described in the mouse
enteropathy X-linked (IPEX) syndrome results in a lethal multi- model of colitis and were shown to produce IL-10 (Groux
organ autoimmunity (d’Hennezel et al., 2009). The direct role et al., 1997). They were subsequently identified in patients
that Tregs play in blocking autoimmunity has been demonstrated with severe combined immunodeficiency (SCID), who ex-
in mouse strains generated with either mutations or deletions hibited chimerism having undergone HLA-mismatched he-
of FoxP3 (Sakaguchi et al., 1985). Treg function has been shown matopoietic stem cell transplantation (Gregori et al., 2012).
to directly affect the ability to induce EAE in the mouse, as the These Tr1 cells secreted high amounts of IL-10 and killed
co-transfer of Tregs with pathogenic CD4+CD25– cells prevents myeloid APCs via a GzmB-mediated mechanism. The
EAE (Read et al., 2000). MBP-reactive, disease-ameliorating absence of specific surface antigens or transcription factors
Tregs have been identified in the mouse (Stephens et al., unique to Tr1 cells means that they are primarily identified
2009), while in humans the existence of MBP-reactive natural operationally as CD4 T cells that exhibit an IL-10high IL-2low
Tregs is unclear since suppressive, FoxP3-expressing human IFNg+ cytokine secretion profile. However, it has recently
MBP-reactive clones are suggested to have arisen from non- been proposed that Tr1 cells co-express CD49b and LAG3
Treg populations (Hong et al., 2009). Several studies have (Gagliani et al., 2013). Cellular therapy with human Tr1-type
demonstrated that inflammation can adversely affect Treg cells enriched by the in vitro co-culture of CD4 T cells with

Neuron 97, February 21, 2018 749


Neuron

Review

IL-10-generated dendritic cells (DC-10) is being tested mixed results, which might be due to the marked differences
in allogeneic hematopoietic stem cell transplantation and, if between human and mouse NK cells (Shi and Van Kaer, 2006).
successful, could be applied to MS. B Cell Populations in MS
Although not as extensively studied as FoxP3+ Tregs, Tr1 cells Although MS is considered to be a T-cell-mediated disease, the
from patients with MS have also been found to be dysfunctional. dramatic effects produced by anti-CD20 therapy (rituximab and
By stimulating CD4 T cells through CD46, which strongly induces ocrelizumab) in MS demonstrate a central role for B cells in its
IL-10, it was found that MS CD4 T cells express less IL-10 than pathogenesis (Greenfield and Hauser, 2017). One of the classic
those from healthy individuals. (Astier et al., 2006). Numerous findings in MS is the presence of oligoclonal bands (OCBs) in
studies have demonstrated the importance of IL-10 in amelio- the CSF, which is observed in 95% of patients and which arise
rating murine EAE (Mayo et al., 2016) and the effects of adminis- from clonally expanded Ig-secreting cells (Heidelberger et al.,
tering Tr1-like cells induced by dexamethasone and vitamin D3 1942; Schirmer et al., 2014). Although the specificity of most
(Kleinewietfeld and Hafler, 2014). In MS, the capacity of CD4 OCBs remains unclear, CSF OCB antibodies from four MS pa-
T cells to secrete IL-10 is associated with decreased disease ac- tients exhibited specificity for a number of ubiquitous intracellular
tivity as the expression of IL-10 and CD46 is enhanced in patients proteins that are released as debris during tissue destruction
who respond to IFNb therapy compared to cells from patients (Bra€ndle et al., 2016). Although CSF antibodies that bind myelin
who did not respond (Chiarini et al., 2012). Non-pathogenic lipids have been associated with aggressive MS (Villar et al.,
Th17 cells have also been reported to secrete increasing 2005), similar anti-lipid antibodies exist in SLE and Grave’s dis-
amounts of Il-10 (Lee et al., 2012b). eases (Endo et al., 1984). Nonetheless, a number of autoantibody
Regulatory Natural Killer Cells targets have been identified in MS, including MOG in pediatric
NK cells are innate immune cells involved in early host defense and NMOSD phenotypes, KIR4.1, and MBP (discussed below
against infection and tumor transformation. NK cells secrete in Biomarkers).
both pro-inflammatory (IFNg, TNFa) and anti-inflammatory Clonally expanded B cells are found in the brain paren-
(IL-4, IL-10) cytokines and have been suggested to play dual chyma, meninges, and CSF of MS patients (Lovato et al.,
roles in disease (Schleinitz et al., 2010). It is unclear whether 2011) and are present in the CNS at greater frequency earlier
the specific types of NK cells, two of which are commonly in the disease (Henderson et al., 2009). Increased B cell fre-
defined as CD56dimCD16hi or CD56bright, play unique roles in quency in the CSF correlates with a rapider disease progres-
autoimmunity (Caruana et al., 2017; Laroni et al., 2016). Although sion (Cepok et al., 2001). Beyond their potential ability to pro-
the CD56dimCD16hi cells that represent 90% of NK cells in duce autoantibodies, B cells in the CNS could act in MS by
peripheral blood are immediately cytotoxic, they are found in secreting chemokines/cytokines and by presenting antigen
much lower frequency in tissue. In contrast, the CD56bright cells, to T cells (Figure 3). Strengthening the hypothesis that B cell
which primarily secrete cytokines and acquire cytotoxic activity antigen presentation might be central to disease pathology,
over time, predominate in tissues. the capacity to induce disease in the EAE mouse model was
Most lines of evidence suggest that NK cells play an immuno- shown to require B cell expression of MHC class II (Molnarfi
regulatory role in MS even though NK cells have been found et al., 2013). B cells can cross the blood-brain barrier and
in the demyelinating lesions of MS patients (Traugott, 1985). become long-term CNS residents in lymph-node-like follicles
CD56bright NK cells are increased by immunomodulatory and found in the meninges, which are often adjacent to cortical le-
immunosuppressant therapies, increases in NK frequency sions (Selter et al., 2013; Cross et al., 2006; Serafini et al.,
correlate with treatment response (Bielekova et al., 2006), 2004). These cellular aggregates, often referred to as ectopic
reduced NK frequency has been associated with relapse (Mor- lymphoid follicles, were identified within the meninges in
andi et al., 2008), and in vitro NK functional activity increases two-thirds of SPMS brain specimens. The existence of these
at times of remission (Infante-Duarte et al., 2005). Daclizumab, follicle-like structures suggests that B cells undergo intra-
an FDA-approved drug that reduces MS relapses, is a human- thecal expansion and differentiation to plasmablasts and
ized antibody against the IL-2 receptor that expands CD56bright plasma cells in the CNS itself (Magliozzi et al., 2007; Serafini
NK cells (Bielekova et al., 2006). A similar increase in CD56bright et al., 2004).
NK cell frequency has also been described in MS patients A central challenge in MS is to understand why depleting
treated with IFNb (Saraste et al., 2007). B cells with anti-CD20 has such a dramatic effect in abolishing
CD56bright NK cells from untreated MS patients have reduced MS relapses and a modest effect on progression in primary
ability to inhibit the proliferation of autologous activated T cells, progressive MS (Castillo-Trivino et al., 2013). CD20 is ex-
which might relate both to the dysfunction of CD56bright NK cells pressed on most B cells and is found on maturing B cells
and to the finding that MS CD4 T cells might be less sensitive just before the cell becomes a terminally differentiated plasma
to NK cell regulation (Laroni et al., 2016). This apparent ‘‘NK cell (Leandro, 2013). Anti-CD20 therapy in MS patients results
resistance’’ by patient-derived T cells was suggested to involve in an 88% reduction in gadolinium-enhancing lesions (Nai-
enhanced T cell expression of the NK-inhibitory ligand HLA-E or smith et al., 2010; Bourdette and Yadav, 2008) within 2 months
reduced expression of CD155 on patient-derived T cells (Gross in association with the immediate loss of B cells from the cir-
et al., 2016). Daclizumab treatment increases expression of culation and from the CSF by 20 weeks. Anti-CD20 also re-
CD155 on CD4 T cells and enhances in vitro suppression of duces the number of T cells in the blood and CSF by 20%
autologous activated T cells by NK cells (Gross et al., 2016). and 50%, respectively (Naismith et al., 2010), and reduces
Mouse studies of NK cells in CNS autoimmunity have given the capacity of the remaining T cells to secrete IL-17 and

750 Neuron 97, February 21, 2018


Neuron

Review

Figure 3. The Pathogenic Role of B Cells in Multiple Sclerosis


B cells present antigens in the context of MHC class II molecules to activate CD4 T cells, contributing to an inflammatory environment. Regulatory B (Breg) cells
can inhibit effector T (Teff) cells via IL-10, IL-35, TGF-b, or PD-L1 and can be dysfunctional in MS. B cells can produce autoantibodies in the periphery and CNS.
B cells can cross the blood-brain barrier and form follicle-like ectopic germinal centers in the CNS, which becomes compartmentalized and functions
independently of the periphery. B cells present in these follicles can undergo expansion and differentiation into antibody-secreting plasma cells within the CNS
itself. B cells also secrete the pro-inflammatory cytokines TNFa, IL-6, and GM-CSF, which promote CNS inflammation.

IFNg (Baker et al., 2017). Plasma cells that produce OCBs do maturation and represent the stage at which negative selection
not express CD20. Thus, the rapid effect on disease relapses against autoreactive B cells is believed to occur (Palanichamy
is not accompanied by changes in antibody levels or OCBs. In et al., 2009). Transitional B cells are expanded in human immu-
contrast, treatment with natalizumab does lead to a reduction nodeficiencies (Cuss et al., 2006). After an early transitional
in OCBs over time (Mancuso et al., 2014). Of note, another B- state, B cells undergo maturation from the mature naive to the
cell-targeting drug, atacicept, which blocks memory B cells pre-germinal center non-switched memory, switched memory
and plasmablasts from using the TACI growth factor, exacer- plasmablast, and, ultimately, to the antibody-secreting plasma
bated MS, suggesting that it negatively affects a disease- cell. The B cell becomes more proliferative and exhibits
ameliorating B cell subset (Bible, 2014). Thus, it appears that enhanced survival as it matures (Palanichamy et al., 2009).
the rapid onset of the beneficial effect of anti-CD20 relates CD19, which modulates the signaling threshold for B cell activa-
to deleting a pro-inflammatory B cell subset that drives tion, is expressed on all circulating B cell lineage cells from the
T cell activation via antigen presentation or cytokine secretion. earliest B lineage cells coming from the bone marrow through
The longer-term benefit of such treatment might arise from re- to the plasmablast. CD20 is a surface protein whose function
establishing the activity of inhibitory regulatory B cell popula- in unknown and is expressed at the pre-B cell stage through to
tions (Barun and Bar-Or, 2012). the development of mature naive and memory B cells but is
Functionally Distinct B Cell Populations in MS lost when the cell develops into a plasma cell. As a result,
The striking disease-ameliorating effect of anti-CD20 has CD20 depletion targets all the mature naive and memory mature
focused interest on understanding the immune-regulating prop- B cell populations but leaves the immature and plasma cell com-
erties of human B cell populations. B cell lymphopoiesis begins partments untouched. Thus, although memory B cells, short-
in the bone marrow when B cell precursors undergo antigen- lived plasmablasts, and long-lived plasma cells are increased
independent differentiation into transitional B cells that migrate in the CNS in MS (Michel et al., 2015) and correlate with MRI le-
from the marrow into the blood. These transitional B cells (CD38hi sions (Kuenz et al., 2008), anti-CD20 therapy may only deplete
CD24hiIgD+CD27 CD10+/ ) are then subject to antigen-driven the memory B cell population.

Neuron 97, February 21, 2018 751


Neuron

Review

Dysfunctional Pro-inflammatory B Cells in MS by their production of IL-10; to date, no specific surface markers
B cells are heterogeneous in their production of pro-inflamma- or transcription factors have been identified that define a Breg
tory (IL-6, TNFa, LTa, and GM-CSF) and anti-inflammatory (Mauri and Menon, 2015). Bregs have also been reported to
(IL-10 and IL-35) cytokines. Both STAT5 and STAT6 can induce suppress via IL-35, TGFb, PD-L1, and GzmB (Ray and Dittel,
B cell GM-CSF production. B cells from MS patients secrete 2017; Molnarfi et al., 2017; Yang et al., 2013). Thus, they may
more GM-CSF and less IL-10 than do the B cells of healthy sub- be a heterogeneous population at different stages of differentia-
jects, and B-cell-derived GM-CSF enhances myeloid secretion tion. Mechanisms of Breg suppression include IL-10-mediated
of the T-cell-activating cytokines IL-12 and IL-6, which play inhibition of dendritic cell priming, the induction of Tr1, and
a role in directing Th1 and Th17 development (Li et al., 2015). induced Tregs (iTregs) via IL-10, TGFb, or PD-L1 and the killing
Strikingly, in MS patients treated with anti-CD20, reconstituting of T cells via GzmB cytotoxicity. In contrast to naive B cells,
B cells were found to produce reduced amounts of GM-CSF B10-type Bregs reside in the memory CD19+CD24hiCD27+
and increased amounts of IL-10. As B cell production of GM- B cell population and are identified by their ability to produce
CSF and IL-10 is mutually exclusive, it is unclear whether the high amounts of IL-10 in response to TLR4/9 stimulation (Iwata
same B cell is differentially regulated to secrete either pro- or et al., 2011). Breg activity is also present in immature transitional
anti-inflammatory cytokines or whether these cytokines are pro- CD27 CD19+CD24hiCD38hi B cells and CD27+CD24 CD38hi
duced by distinct subsets within the mature B cell population. plasmablasts, which can utilize IL-10 and PD-L1 to induce CD4
Regardless, anti-CD20 treatment results in a long-lasting benefit T cells to become Tr1 or iTregs, which are the suppressive
that appears to correlate with the absence of memory B cells and FoxP3+ cells that are generated in the periphery after thymic se-
with reduced B cell secretion of pro-inflammatory cytokines lection (Blair et al., 2010; Flores-Borja et al., 2013; Siewe et al.,
(Hauser et al., 2013). 2014). In contrast, the Breg cells that express CD19+CD38+
Changes in mature B cell number and activity correlate with CD1d+IgM+CD147+, which are induced by IL-21, produce both
disease activity and treatment response. Unlike B cells in pa- IL-10 and GzmB and suppress by killing (Lindner et al., 2013).
tients with RA or T1D, the deletion of autoreactive B cells at Using a functional definition to identify Bregs makes it difficult
the development stage of the transitional B cell is not aberrant to determine whether disease-associated differences in Breg
in MS. Yet, despite an apparently normal response to checkpoint activity are due to defective Breg function or to Breg paucity.
deletion, mature naive B cells exhibit abnormal activation and Furthermore, being restrained by a functional definition makes
proliferation in MS (Kinnunen et al., 2013). Furthermore, memory it difficult to determine whether Breg subsets have a fixed or a
B cells from patients with MS show enhanced expression of transient suppressive function.
CD80 and CD86 (Bar-Or et al., 2001). The unsuccessful trial of Changes in Breg frequency and activity have been reported
atacicept, designed to block BAFF growth factor activation in disease. Bregs from MS patients who had a concurrent
of mature B cells, actually enhanced memory B cell activation Helminth infection exhibited increased IL-10 production and
and exacerbated CNS inflammation (Sergott et al., 2015b). a greater capacity to suppress T cells in vitro (Correale
This exacerbation appeared to involve the unexpected induction et al., 2008). The frequency of IL-10-secreting B cells are
of an IL-15-mediated increase in memory B cell activity by ataci- decreased in patients with RA (Flores-Borja et al., 2013)
cept (Ma et al., 2014). Enhanced B cell activation might also have and Grave’s disease (Zha et al., 2012), while Bregs in patients
been responsible for the development of MS following anti-TNFa with SLE exhibit an impaired secretion of IL-10 in response to
therapy in patients with RA or Crohn’s disease as treatment re- CD40 activation (Blair et al., 2010). However, it is not clear
sulted in increased numbers of memory B cells (Roll et al., whether the frequency of IL-10-secreting Bregs is altered in
2012; Souto-Carneiro et al., 2009). Also, IL-21, which is pro- MS (Michel et al., 2014a; Habib et al., 2015). The reset of
duced by pathogenic Th17 cells, might enhance the proliferation Breg function could be one way in which anti-CD20 therapy
and survival of memory B cells (Gharibi et al., 2016). Thus, in MS, in SLE patients restored the expression of CD1d on
memory B cells exhibit enhanced pro-inflammatory properties CD19+CD24hiCD38hi Bregs, along with restoring other im-
and have the capacity to drive T cell activation. mune populations (such as invariant NK T cell activity)
In addition to affecting mature B cells, anti-CD20 treatments (Bosma et al., 2012). In MS, one way to explain the rapid ef-
might affect other immune system populations. Unique CD20- fect of CD20 therapy is that it quickly depletes the pro-in-
transitional B cell subsets are expanded relative to other popula- flammatory cytokine-producing memory B cells that are
tions after anti-CD20 therapy and may be involved in restoring increased in MS (von Bu €dingen et al., 2015). In this regard,
immune homeostasis (Palanichamy et al., 2009). An ignored the long-term benefit of Alemtuzumab (an anti-CD52) therapy
area related to anti-CD20 efficacy in MS is a small compartment for MS patients might be due to the long-term reversal of
of T cells that express CD20. These CD20+ T cells are highly pro- naive versus memory B cell ratios (Baker et al., 2017).
apoptotic, contain both CD4 and CD8 subpopulations, and Thus, the long-term benefit of anti-CD20 in MS may be
show increased production of IL-17, IFNg, and TNFa (Schuh related to both a reduction in pro-inflammatory memory B
et al., 2016). The depletion of these CD20+ T cells by anti- cells and the restoration of Bregs.
CD20 therapy would be beneficial were they to play a pathogenic As more has been learned about both MS and the complex
role in MS (Palanichamy et al., 2014). immune networks that drive autoimmunity, pivotal roles for the
Regulatory Anti-inflammatory B Cells in MS innate immune system, and for B cells, endothelial function,
Subsets of B cells that suppress immune responses are referred and different T cells subpopulations, have been identified, estab-
to as regulatory B cells or Bregs. Bregs are primarily identified lishing them as key contributors to MS immune dysregulation.

752 Neuron 97, February 21, 2018


Neuron

Review
Figure 4. Mechanisms in Progressive
Multiple Sclerosis
CNS resident immune cells, including microglia
and astrocytes, as well as B cells potentially from
ectopic germinal centers, can drive the disease
processes in progressive MS that contribute
to neurodegeneration. The immune-dependent
components of disease might activate various
disease processes that then become self-main-
taining and immune-independent deleterious
mechanisms in the neuron. Mitochondrial injury
occurs as a result of mitochondrial DNA
(mtDNA) mutation and reduced respiratory chain
activity, leading to energy deficiency, owing to
the enhanced production of ROS/RNS. Iron is
released from regions of active demyelination
and contributes to oxidative stress, while gluta-
mate excitotoxicity causes a massive influx of
calcium into neurons, resulting in ionic imbalance.
These immune-dependent and -independent
mechanisms all lead to axonal atrophy and neu-
rodegeneration. ROS, reactive oxygen species;
RNS, reactive nitrogen species; Fe3+, iron; Ca2+,
calcium.

Progressive MS of meningeal inflammation and in the formation of ectopic


The pathologic features associated with progressive MS germinal centers and follicle-like structures, which are en-
include brain atrophy, which is related to axonal loss, cortical riched in B cells, plasma cells, and follicular dendritic cells
demyelination, microglia activation, and failure of remyelination. (Serafini et al., 2004). Meningeal inflammation appears to be
The physical decline and disability experienced by MS patients an important cause of cortical demyelination, which might be
are generally related to this form of the illness. Progressive MS induced by soluble factors that diffuse from the meninges. Mi-
can develop from the relapsing form of MS or can be progres- croglia are resident innate immune cells of the brain and have
sive from the outset (primary progressive). From MRI imaging a unique molecular signature (Butovsky et al., 2014). They are
studies, it is clear that, in most instances, primary progressive activated in MS and can secrete pro-inflammatory cytokines,
MS is simply the secondary progressive form, in which the re- reactive oxygen intermediates, proteinase, and complement
lapsing stage involved sub-clinical relapses. However, in some proteins. PET imaging demonstrates diffuse microglial inflam-
instances, the disease begins as primary progressive MS, rep- mation in progressive MS, including in normal-appearing white
resenting a unique subtype of the disease, although the pro- matter (Rissanen et al., 2014). Visualization of leptomeningeal
cesses that drive both primary and secondary progressive enhancements at T7 may also be useful (Zurawski et al.,
MS are likely to be the same. In general terms, relapsing dis- 2017). Microglia also have neuroprotective phenotypes, such
ease involves the movement of immune cells from the periph- as upregulating indoleamine 2,3-dioxygenase, which modu-
ery into the CNS, whereas progressive disease involves the lates tryptophan metabolism and promotes an anti-inflamma-
development of a compartmentalized pathological process in tory microenvironment (Kwidzinski et al., 2005). However, mi-
the brain. This is evident by the decrease in gadolinium croglia lose their normal homeostatic phenotype in MS (Zrzavy
enhancement on MRI of the brain of progressive MS patients, et al., 2017). Astrocytes are involved in multiple aspects of
indicating a reduced breakdown of the blood-brain barrier CNS function, including synapse maintenance, neurotransmis-
and less movement of cells into the brain. Importantly, few sion, phagocytosis, and blood-brain barrier formation. A
drugs that are effective in relapsing forms of MS are effective neurotoxic subpopulation of astrocytes (termed A1 astrocytes)
in progressive disease as most of these drugs target the adap- has been described (Liddelow et al., 2017), and the primary
tive peripheral immune system as opposed to a compartmen- driver of A1 astrocytes are microglia. Astrocytes promote
talized CNS immune response. Unlike relapsing MS, which is MS by producing neurotoxic molecules, such as NO and
primarily immune system driven, progressive MS involves TNFa, and by recruiting neurotoxic inflammatory monocytes
both immune-dependent and -independent mechanisms, to the CNS. Of note, in the acute EAE model of MS, the dele-
although the immune-independent mechanisms are triggered tion of astrocytes worsens disease whereas astrocyte deletion
by the immune system (Figure 4). in the progressive EAE model improves disease (Mayo et al.,
Immune-Dependent and -Independent Mechanisms of 2014). Changes in the innate immune system have also been
Progression identified in the peripheral blood of patients with progressive
Whereas the peripheral adaptive immune system (T cells) MS, who exhibit enhanced IFNg production by non-T cells
drives relapsing MS, the innate immune system (microglia (Balashov et al., 1997) as well as increased numbers of
and astrocytes) with B cells is the primary driver of progressive blood-derived myeloid dendritic cells that express CD80 and
MS. B lymphocytes are prominent, particularly in the context secrete IL-12 and TNFa (Karni et al., 2006). Once CNS

Neuron 97, February 21, 2018 753


754 Neuron 97, February 21, 2018

Table 2. Treatment for Multiple Sclerosisa


1. FDA-approved therapies
Drug (Date of Approval) Target Mechanism of Action Efficacy Adverse Events References
Commercial Name
IFNb (1993) (IFNb-1b Binds to type I IFN Inhibits T cell division, matrix Reduction in annualized Post-injection flu-like Bermel et al., 2013;
Betaseron, Extavia) receptor on human metallo-proteinase, BBB relapse rate in RRMS. Delays symptoms, liver toxicity, and Schubert et al., 2015;
(IFNb-1a Avonex, cells migration, and conversion to clinically depression. Neutralizing Abs Goodkin, 1996; Kappos
Rebif, Plegridy) proinflammatory cytokines. definite MS in CIS. associated with reduced et al., 2016
Induces Tregs, suppressive efficacy
transitional CD19+CD24++
CD38++ B cells
Glatiramer Acetate Random polymers of Competes with peptide Reduction in annualized Post-injection idiosyncratic Arnon and Aharoni, 2004;
(1997) Copaxone, glutamic acid, lysine, binding to MHC, increases relapse rate in RRMS. reactions, lymphadenopathy Racke et al., 2010; Simpson
Glatopa alanine, and tyrosine IL- 10, IL-4, TGFb, and CD8 et al., 2002; Tennakoon et al.,
to bind to MHC Tregs, induces type II 2006; Weber et al., 2007
monocytes that cause a Th1
to Th2 shift
Mitoxantrone (2000) Interferes with DNA Causes nucleotide Reduction in annualized Bone marrow suppression, Martinelli Boneschi et al.,
Novantrone repair crosslinking and DNA strand relapse rate and disease cardiomyopathy, leukemia 2013; Vollmer et al., 2010
breaks. Inhibits lymphocyte progression in clinically
and monocyte migration, B worsening RRMS
cell function, and secretion of and SPMS.
TNFa, IL-2 and IFNg
Natalizumab (2004) CD49d, the a4 Blocks B and T cell migration Reduction in annualized Reactivation of the JC virus Polman et al., 2006;
Tysabri subunit of VLA4 into the CNS. Blocks VLA4 relapse rate in RRMS. in the CNS may occur in Sellebjerg et al., 2016
integrin (humanized binding to VCAM-1 and some patients (PML),
mAb) fibronectin infusion reactions
Fingolimod (2010) S1PR (Sphinosine-1- Sequesters lymphocytes in Reduction in annualized Bradycardia, infection, Willis and Cohen, 2013
Gilenya phosphate receptor) lymph nodes by inhibiting relapse rate in RRMS. macular edema,
lymphocyte egress. Inhibits lymphopenia rare PML cases
the migration of dendritic
cells to secondary lymphoid
organs
Teriflunomide (2012) Inhibits dihydroorotate Inhibits pyrimidine synthesis. Reduction in annualized Hair thinning, liver toxicity, O’Connor et al., 2011;
Aubagio dehydrogenase Inhibits secretion of relapse rate in RRMS. teratogenesis Tallantyre et al., 2008
proinflammatory cytokines

Review
and T cell activation
(Continued on next page)

Neuron
Review
Neuron
Table 2. Continued
Dimethyl fumarate Nrf2 pathway Activates the Nrf2 Reduction in annualized GI side effects, flushing, Kappos et al., 2008
(2013) Tecfidera transcriptional pathway relapse rate in RRMS. lymphopenia, and rare PML
Alemtuzumab (2014) CD52 on T and B cells Depletes B and T cells via Reduction in annualized Autoimmune diseases Coles et al., 2008; Havrdová
Lemtrada (humanized mAb) ADCC and complement relapse rate in RRMS. including thyroid, immune et al., 2015
thrombocytopenia purpura,
and glomerulonephritis
Daclizumab (2016) CD25, anti-IL2R Prevents IL-2 signaling Reduction in annualized Liver toxicity, skin reactions Giovannoni et al., 2014;
Zinbryta (humanized mAb) through the high affinity relapse rate in RRMS. Gross et al., 2016; Kappos
IL-2R. Augments CD56+ NK et al., 2015
cell activity
Ocrelizumab (2017) CD20+ B cells Depletes CD20+ B cells. Decreased annualized Infusion reactions, risk of Hauser et al., 2017;
Ocrevus (humanized mAb) Reduces pathogenic B cell relapse rate in RRMS. breast cancer, low risk Montalban et al., 2017;
antigen presentation Reduced disease of PML Sørensen and
progression in PPMS. Blinkenberg, 2016
Cladribine (2017) Adenosine deaminase Depletes immune cells by Reduction in annualized Infection, lymphopenia Giovannoni et al., 2010;
(EMA/Canadian inducing lymphocyte relapse rate in RRMS. Leist et al., 2014
approval) apoptosis. Sustained
reduction in CD4 and CD8 T
cells and transient reduction
in B cells
2. Off-label immune therapy
Drug Target Mechanism of Action Efficacy Adverse Events References
Bone Marrow Immune reconstitution Global immune suppression Suppresses relapses and Transplant-related morbidity Burt et al., 2015; Fassas
Transplantation may slow disease and secondary malignancy et al., 1997; Daikeler
progression. et al., 2012
Cyclophosphamide Alkylating metabolite Immune suppression and Suppresses relapses and Alopecia, myelosuppression, Comabella et al., 1998;
immune deviation may slow disease cancer, sterility Stankiewicz et al., 2013
progression.
Azathioprine; General immune Inhibits T and B cell activity, Suppresses relapses in GI and hepatic toxicity; GI Vollmer et al., 2010; Gray
Methotrexate; suppressants proliferation, and BBB uncontrolled trials. toxicity, myelosuppression, et al., 2004; Michel
Neuron 97, February 21, 2018 755

Mycophenolate migration. Inhibits de novo GI toxicity, pneumonitis; et al., 2014b


mofetil purine synthesis; leukopenia, anemia,
dihydrofolic acid reductase; headaches and diarrhea
and inosine monophosphate
dehydrogenase
(Continued on next page)
756 Neuron 97, February 21, 2018

Table 2. Continued
Intravenous steroids Anti-inflammatory Dampens inflammatory Accelerates relapse recovery Sleep problems, mood Sloka and Stefanelli, 2005;
cytokine cascade, inhibits (3–5 day course). disturbance, weight gain Sellebjerg et al., 1998
the activation of T cells, limits
lymphocyte extravasation
into CNS induces apoptosis
of activated immune cells
Intravenous Immune modulation Idiotypic regulation of B May be useful in RRMS and Hypercoagulability, renal Katz et al., 2006; Sørensen et
Immunoglobulin cells, reduces serum pregnancy. failure al., 2002
autoantibody and
inflammatory cytokines
Rituximab Rituxan CD20+ B cells Depletes CD20+ B cells. Reduces relapse rate Infusion reactions Castillo-Trivino et al., 2013;
(Chimeric Ab) Reduces pathogenic B cell in RRMS. Hauser et al., 2008
antigen presentation
Plasmapheresis Removal of serum Immune modulation Accelerates relapse recovery Hypotension and Weiner et al., 1989;
factors and limits disability accrual. hypocoagulability Weinshenker, 2001
3. Clinical trials
Drug Target Mechanism of Action Efficacy Adverse Events References
Siponimod; Ozanimod S1PR (Sphinosine-1- Sequesters lymphocytes in Reduction of annualized Bradycardia, lymphopenia Kappos et al., 2016, 2017;
phosphate receptor) Lymph nodes by inhibiting relapse rate in RRMS. Selmaj et al., 2013; Cohen
egress. May have direct Reduces risk of disability et al., 2016
antiinflammatory effects progression in SPMS.
in CNS
Laquinimod Derivative of linomide Induces Tregs and secretion Reduction of annualized Liver toxicity Comi et al., 2012; Polman
(quinoline- of TGFb and Th2 cytokines. relapse rate in RRMS. et al., 2005
3-carboxamide) Neuro-protective effect:
increased serum BDNF, and
reduced excitatory currents
High-dose biotin Vitamin B7 An essential co-factor for Stabilization of disease Infrequent diarrhea, renders Sedel et al., 2015; Tourbah
fatty acid synthesis and progression and biotinylated assays et al., 2016
energy production improvement of disability in inaccurate
30% of patients.
Secukinumab Anti-IL-17A Blocks the activity of IL-17 Reduction of new Infection Havrdová et al.; 2016
gadolinium-enhancing T1
lesions.
4. Treatments that worsened MS
Drug (Date of Trial) Target Mechanism of Action Comments References

Review
IFNg (1987) IFNg receptor Increased monocyte MHC Induced disease exacerbation Panitch et al., 1987b
class II expression in 7/15 patients.
(Continued on next page)

Neuron
Review
Neuron
Table 2. Continued
TNFa (1996) TNFa blockade Reduction of TNFa activity Increased MRI and disease activity in an open label study. A Arnason et al., 1999; van
Lenercept/Infliximab (sTNFR-Ig) multi-center trial of Lenercept was halted due to increased Oosten et al., 1996
relapses and severity of exacerbations.
Atacicept (2014) BLyS/April receptor Depletes mature B cells and Exacerbated disease activity in RRMS (beneficial in RA Kappos et al., 2014; Sergott
blockade plasma cells and SLE). et al., 2015a
Tocilizumab (2016) IL-6R (humanized Blocks IL6R In RA patients, blocking IL-6R induced MS-like lesions. Beauchemin and
mAb) Carruthers, 2016;
Comabella, 2016
5. Antigen-specific therapies for MS
Drug (Date of Trial) Target Mechanism of Action Comments References
Oral Myelin (1993) Ag-specific cells Mucosal induction of No side effects. No detectable benefit. Weiner et al., 1993
regulatory cells, tolerization
of antigen-specific cells
APL for MBP (2000) Antigen peptide of Block T cell responses by Worsened disease by MRI in some patients; increased MBP Bielekova et al., 2000
MBP with modifications acting as TCR antagonist or reactive T cells in circulation.
made to TCR contact by inducing regulatory T cell
sites populations
BHT-3009 (2008) DNA vaccine Immune tolerance to myelin Reduction of gadolinium enhancement in double blind trial. Garren et al., 2008
encoding myelin basic basic protein
protein
Transdermal Myelin MOG, MBP, and PLP Immune tolerance to myelin Reduction in antigen-specific T cell responses. Walczak et al., 2013
Peptides (2013) peptides antigens
ETIMS antigen Autologous PBMCs Induction of antigen-specific No new lesions in low-disease patients, new lesions in highly Lutterotti et al., 2013
specific therapy coupled to peptides tolerance. Decreased active patients.
(2013) antigen-specific T cell
responses
ATX-MS-1467 (2018) Antigen-specific Cocktail of four myelin First trial (6 patients): no adverse events, unchanged EDSS Streeter et al., 2015;
pathogenic T cells antigen peptides to score. Second trial (43 patients): effect on gadolinium Chataway et al., 2018
induce tolerance, given enhancement measured at 16 weeks.
intradermally
Neuron 97, February 21, 2018 757

a
see MScenter.partners.org for updates to table
Neuron

Review

damage has occurred, MS progression is predominantly phase II placebo-controlled trial showing a reduction in the
driven by immune-independent mechanisms, which also partly annualized whole-brain atrophy rate (Chataway et al., 2014).
explains why immunomodulatory therapy has generally been Other approaches include re-myelination strategies (Franklin,
unsuccessful in treating progressive MS. These immune-inde- 2017; Stangel et al., 2017), inhibition of ion channels, and tar-
pendent mechanisms include mitochondrial injury, iron accu- geting mitochondria (Correale et al., 2017). To that end, Lipoic
mulation leading to oxidative stress, ion channel dysfunction, acid, an antioxidant, reduced annualized changes in brain vol-
glutamate excitotoxicity, excess intra-axonal calcium, tissue ume in SPMS patients (Spain et al., 2017), while Clemastine
hypoxia, and defects in axonal transport (Mahad et al., 2015; fumerate, a remyelination drug, showed efficacy in RRMS sub-
Correale et al., 2017). jects (Green et al., 2017). The effective treatment of progres-
Treatment of Progressive MS sive forms of MS will depend on the early treatment of inflam-
Most trials of disease-modifying therapy for progressive MS mation to prevent compartmentalized inflammation developing
have not succeeded, and we refer readers to recent reviews in the CNS, treatment that targets innate immune processes in
for more information on these trials (Comi, 2013; Correale the brain, and non-immune approaches that target neurode-
et al., 2017). The reason for the failure of these trials has generation and axonal dysfunction. Progress in meeting this
become clearer as the different pathophysiologic mechanisms challenge relies on the development of biomarkers and of
of progressive versus relapsing disease have been defined. new imaging approaches that can, for example, image gray
The partial success of immunomodulatory therapy for progres- matter and microglial activation. MRI imaging has been crucial
sive MS has generally been due to the treatment of cohorts in the development of the many treatments now available for
who were in the early stages of progressive disease and still relapsing forms of MS.
had a component of inflammation (Hohol et al., 1999). It is
noteworthy that anti-CD20 treatment with ocrelizumab did Treatment for Relapsing MS
succeed in a phase III clinical trial of patients with primary pro- The remarkable advances made in the treatment of MS are
gressive MS that were diagnosed with primary progressive shown in Table 2, with more than ten drugs approved by the
multiple sclerosis and had disease for either less than 15 years FDA for treating this disease. All of these drugs target the im-
with an Expanded Disability Status Scale (EDSS) score of mune system and impact the immune networks shown in Fig-
more than 5.0 or less than 10 years with an EDSS score of ures 2 and 3, and virtually all of them impact the inflammatory
up to 5.0 while excluding patients with relapsing-remitting, relapsing stage of the disease. Treatment for progressive
secondary progressive, or progressive relapsing disease forms of MS constitute a major unmet need. Broadly consid-
(Montalban et al., 2017). As a result, ocrelizumab has been ered, these immunomodulatory drugs have the following
approved to treat primary progressive MS by the FDA. Siponi- modes of action: decrease disease-inducing Th1/Th17 cells,
mod, a selective S1P inhibitor, has also reported positive re- induce Treg cells, affect the trafficking of cells to the nervous
sults in a phase III trial of progressive MS even though another system, and affect B cells. The availability of so many treat-
S1P inhibitor (fingolimod) did not succeed (Smith and Cohen, ments for relapsing forms of MS raises two important consid-
2016) The mechanism(s) by which these treatments ameliorate erations in deciding how to use them in patients: what is the
progressive MS is not well understood but may relate to direct relative efficacy of one drug versus another, and what are
or indirect effects on the innate immune response in the CNS. the side effects associated with treatment? Although head-
A trial of intrathecal anti-CD20 has been proposed to better to-head comparative studies are not available for all drugs,
target B cells in the CNS. among FDA-approved drugs, alemtuzumab, natalizumab,
Studies in animal models of progressive MS have identified and ocrelizumab appear to have the strongest anti-inflamma-
unique pathways by which the CNS innate immune system tory effects as measured by reducing relapses and affecting
may be targeted. Sphingosine 1-phosphate receptor modula- inflammation (as assessed by MRI, using gadolinium enhance-
tion suppresses pathogenic astrocyte activation and progres- ment) (Tramacere et al., 2015). Among oral medications, fingo-
sive CNS inflammation (Rothhammer et al., 2017) as does limod appears to have stronger anti-inflammatory effects than
modulation of astrocyte activity by dietary tryptophan (Roth- either teriflunomide or dimethylfumarate. Injectable medica-
hammer et al., 2016). Laquinomid, which affects brain atrophy tions, such as interferons and glatiramer acetate, appear to
and progression in MS but didn’t affect relapses, might act by be comparable, with higher doses of interferon providing a
inducing an effect on astrocytes. Nasally delivered anti-CD3 in- better treatment outcome than lower doses. All FDA-approved
duces IL-10-secreting Tregs that migrate to the brain, drugs have an effect on brain atrophy as measured by MRI,
decrease both microglial and astrocyte activation, and slow although it is difficult to compare drugs given the different
disease progression (Mayo et al., 2016). A carbon-based technologies used to quantify brain atrophy. Clinicians gener-
fullerene linked to an NMDA receptor with anti-excitotoxic ally base treatment decisions for an individual patient on the
properties slows disease progression and prevents axonal patient’s relapse history and the number of inflammatory
damage in a progressive animal model (Basso et al., 2008). A events and new T2 lesions on MRI. In the future, measures
unique approach that has shown promise in a double-blind of atrophy linked to progression might also become a basis
study of severe progressive MS is high-dose biotin (Tourbah for treatment decisions.
et al., 2016), which might reduce axonal hypoxia through There are various side effects associated with MS immuno-
enhanced energy production and also promote axonal remye- therapy (see Table 2), and these side effects can often determine
lination. High-dose statins (simvastatin) were also tested in a which drug is used to treat MS once the risks and benefits

758 Neuron 97, February 21, 2018


Neuron

Review

are considered. One of the most concerning side effects is pro- Despite the advances in treatment options and an emerging
gressive multifocal leukoencephalopathy (PML), which is partic- recognition of disease heterogeneity, to date, no single biochem-
ularly associated with treatment with natalizumab. Risk factors ical or immune marker is able to accurately reflect disease path-
for developing natalizumab-associated PML include prior immu- ogenesis, segregate the different disease phenotypes, serve as a
nosuppressive therapy, the presence of antibodies to John monitor of disease activity, or predict the future course of dis-
Cunningham virus (JCV), and treatment for more than 2 years. ease. It is beyond the scope of this article to provide a detailed
The incidence of PML is 13 per 1,000 MS patients treated with discussion of the plethora of biomarkers under study; instead,
all three risk factors. Cases of PML have also been reported we refer readers to recent comprehensive reviews on this topic
following treatment with fingolimod and dimethylfumarate. An (El Ayoubi and Khoury, 2017; Comabella and Montalban, 2014).
important conceptual point to consider is whether early treat- One particular biomarker of note, serum neurofilament levels (a
ment with drugs that have the strongest anti-inflammatory prop- biomarker of neuroaxonal damage), appears to be associated
erties is the best approach to preventing later disability. There is with the progression of brain atrophy and disability in early MS
no consensus on the best sequence of treatment or on the (Kuhle et al., 2017), although elevated serum neurofilament levels
optimal combination therapy. Because it is considered unethical are not specific to MS. Serum antigen-array patterns (Quintana
not to treat relapsing MS, the testing of new drugs against a pla- et al., 2008) and microRNAs (Gandhi et al., 2013) can also distin-
cebo is problematical. guish different subtypes of MS, and cellular transcription signa-
There have also been immunomodulatory treatments that tures can identify patients with different levels of disease activity
have made MS worse (Table 2). Some of these effects, such (Ottoboni et al., 2012). Nevertheless, studies of biomarkers suffer
as the increased number of attacks following IFNg, which is from a lack of independent validation. For example, reports that
known to increase class II expression and to stimulate the anti-myelin antibodies predict conversion from CIS to RRMS
innate immune system, were not unexpected given our current (Berger et al., 2003) and that the potassium channel KIR4.1 is
knowledge. Less easy to explain is the worsening of MS an immune target in MS (Srivastava et al., 2012) could not be
following treatment with anti-TNFa and atacicept. Anti-TNFa replicated. Thus, the implementation of biomarkers for MS
is an effective therapy for RA, and yet some RA patients requires vigorous clinical and multicenter validation.
treated with anti-TNFa develop MS. Worsening of MS with No Evidence of Disease Activity
the anti-B cell drug Ataticept illustrates the complex role of The emergence of ‘‘no evidence of disease activity’’ (NEDA) as a
B cells in MS. concept has its roots in rheumatology, in which disease-free
If MS is an antigen-specific, cell-mediated autoimmune dis- status is the goal for clinical trials and for patient care. In MS,
ease, the most specific and least toxic therapy for MS would the NEDA composite measure consists of the following fea-
be an antigen-specific approach, which has been attempted in tures: (1) no new or enlarging T2-weighted lesions, (2) no new
MS with limited success. One of the difficulties of this approach gadolinium-enhancing lesions, (3) no relapses, and (4) no
is the development of reactivity to multiple CNS antigens as the confirmed worsening of EDSS scores (Bevan and Cree, 2014).
disease progresses and the lack of a clear inciting antigen for Clinical trials have begun to report NEDA as a secondary
which there is common reactivity in MS, such as reactivity to outcome measure. The percentage of patients achieving
AQP4 in neuromyelitis optica. NEDA is usually measured in the first 2 years and varies among
Biomarkers studies from 28% to 42%. These numbers may be very different
Biomarkers are beginning to play an increasingly important role in a real-world setting compared with those obtained in clinical
in MS with a better understanding of its disease mechanisms trials. We recently explored the percentage of patients
and with advances in technology. The most commonly used achieving NEDA status in a real-world setting (Rotstein et al.,
biomarker is MRI imaging, which is primarily used as a 2015), and although 46.0% of patients in this study demon-
biomarker of disease activity. MRI serves as a measure of the strated NEDA after 1 year of observation during therapy, only
inflammatory component of MS as evidenced by gadolinium- 7.9% of patients maintained durable NEDA after 7 years. This
enhancing lesions and the appearance of new T2 lesions. Given demonstrates that NEDA is difficult to sustain in a real-world
that there are multiple drugs for the treatment of the relapsing cohort of MS patients.
and inflammatory components of MS, MRI imaging is routinely
used to help adjust therapy to minimize the appearance of new Conclusion
lesions. There are no analogous MRI markers for progressive Enormous progress has been made in our understanding of MS.
MS, although physicians may choose to treat this form of MS This understanding has led to treatments for MS, especially in the
with more aggressive forms of therapy depending on the de- relapsing and inflammatory stages. Major advances in MS are
gree of MRI disease activity, including brain atrophy and spinal linked to the identification of abnormal immune responses in
cord involvement. Another clinically useful biomarker is the the disease and how to measure and modify them. It is now clear
measurement of JC virus antibody titers, the presence of which that MS is primarily an autoimmune disease and not a persistent
increases the risk of PML. JV virus titers are measured in all pa- infection even though an infectious agent may trigger the disease
tients being treated with natalizumab and treatment with this or be associated with the disease. All genetic studies point to the
drug is generally limited if a patient’s titers are high. The pres- immune system as being the primary cause of MS pathology, with
ence of elevated IgG in the CSF has classically been used as a a complex interaction occurring between genes and environ-
diagnostic biomarker but has now largely been replaced ment. MRI imaging provides an important opportunity to monitor
by MRI. the disease. There have been surprises along the way, including

Neuron 97, February 21, 2018 759


Neuron

Review

the lack of a beneficial effect from anti-TNFa therapy and the dra- Balashov, K.E., Smith, D.R., Khoury, S.J., Hafler, D.A., and Weiner, H.L. (1997).
Increased interleukin 12 production in progressive multiple sclerosis: induction
matic effect of anti-CD20 therapy. Precision medicine has by activated CD4+ T cells via CD40 ligand. Proc. Natl. Acad. Sci. USA 94,
evolved as a major theme in medicine and is becoming a major 599–603.
goal in MS (Zhao et al., 2017). The current key unanswered ques-
Bar-Or, A., Oliveira, E.M., Anderson, D.E., Krieger, J.I., Duddy, M., O’Connor,
tion in MS is how to treat and better understand the progressive K.C., and Hafler, D.A. (2001). Immunological memory: contribution of memory
form of this disease. This will require new approaches to target B cells expressing costimulatory molecules in the resting state. J. Immunol.
167, 5669–5677.
compartmentalized processes in the CNS.
Finally, if MS is triggered by the environment in susceptible Baranzini, S.E., Cekanaviciute, E., Debelius, J., Singh, S., Runia, T., Yoo, B.,
individuals, it might be possible to prevent MS. Some believe Crabtree-Hartman, E., Bove, R., Gelfand, J., and Jia, S. (2017). The MS-asso-
ciated gut microbiome. Mult. Scler. J. 23, 100.
vaccination against EBV or treating children at high risk for
development of MS with vitamin D could be effective initial ap- Barry, M., and Bleackley, R.C. (2002). Cytotoxic T lymphocytes: all roads lead
to death. Nat. Rev. Immunol. 2, 401–409.
proaches. Prevention would be the ultimate cure for MS and
would depend on a treatment that modulates the immune sys- Barun, B., and Bar-Or, A. (2012). Treatment of multiple sclerosis with anti-
tem in childhood or adolescence. Such a vaccine could initially CD20 antibodies. Clin. Immunol. 142, 31–37.
be given to higher-risk individuals, such as children whose par- Basso, A.S., Frenkel, D., Quintana, F.J., Costa-Pinto, F.A., Petrovic-Stojkovic,
ents have MS and where there is a strong family history. Modu- S., Puckett, L., Monsonego, A., Bar-Shir, A., Engel, Y., Gozin, M., and Weiner,
H.L. (2008). Reversal of axonal loss and disability in a mouse model of progres-
lation of the gut microbiome appears as an attractive approach sive multiple sclerosis. J. Clin. Invest. 118, 1532–1543.
as it may have no negative effects. In addition, our understanding
of MS could change with new discoveries. The most impactful Beauchemin, P., and Carruthers, R. (2016). MS arising during Tocilizumab
therapy for rheumatoid arthritis. Mult. Scler. 22, 254–256.
would be the identification of an infectious agent truly linked
to the disease. Over the last century, MS has moved from an un- Becher, B., and Segal, B.M. (2011). T(H)17 cytokines in autoimmune neuro-
inflammation. Curr. Opin. Immunol. 23, 707–712.
known, untreatable disease to one with many treatment options
and many unexplored roads ahead. Becher, B., Durell, B.G., and Noelle, R.J. (2002). Experimental autoimmune
encephalitis and inflammation in the absence of interleukin-12. J. Clin. Invest.
110, 493–497.
ACKNOWLEDGMENTS
Berer, K., Mues, M., Koutrolos, M., Rasbi, Z.A., Boziki, M., Johner, C., Wekerle,
H., and Krishnamoorthy, G. (2011). Commensal microbiota and myelin autoan-
Supported by grants from the National Multiple Sclerosis Society to H.L.W. tigen cooperate to trigger autoimmune demyelination. Nature 479, 538–541.
(RR 205-A-13, RG-1507-05029. NMSS ICT 0010) and to C.B.-A. (RG-1506-
04836), by NIH grants to H.L.W. (R01 NS087226, UH3TR000890), by the Ever- Berer, K., Gerdes, L.A., Cekanaviciute, E., Jia, X., Xiao, L., Xia, Z., Liu, C., Klotz,
grande Center for Immunologic Diseases, and by the Nancy Davis Foundation. L., Stauffer, U., Baranzini, S.E., et al. (2017). Gut microbiota from multiple scle-
We thank Dr. Tanuja Chitnis, Dr. Vijay Kuchroo, and Dr. Rohit Bakshi for their rosis patients enables spontaneous autoimmune encephalomyelitis in mice.
discussions on the manuscript. Proc. Natl. Acad. Sci. U.S.A. 114, 10719–10724.

Berger, T., Rubner, P., Schautzer, F., Egg, R., Ulmer, H., Mayringer, I., Dilitz, E.,
REFERENCES Deisenhammer, F., and Reindl, M. (2003). Antimyelin antibodies as a predictor
of clinically definite multiple sclerosis after a first demyelinating event. N. Engl.
J. Med. 349, 139–145.
Ahlgren, C., Odén, A., and Lycke, J. (2012). A nationwide survey of the preva-
lence of multiple sclerosis in immigrant populations of Sweden. Mult. Scler. 18, Berghmans, N., Dillen, C., and Heremans, H. (2006). Exogenous IL-12
1099–1107. suppresses experimental autoimmune encephalomyelitis (EAE) by tuning
IL-10 and IL-5 levels in an IFN-gamma-dependent way. J. Neuroimmunol.
Annibali, V., Ristori, G., Angelini, D.F., Serafini, B., Mechelli, R., Cannoni, S., 176, 63–75.
Romano, S., Paolillo, A., Abderrahim, H., Diamantini, A., et al. (2011).
CD161(high)CD8+T cells bear pathogenetic potential in multiple sclerosis. Bermel, R.A., You, X., Foulds, P., Hyde, R., Simon, J.H., Fisher, E., and Rudick,
Brain 134, 542–554. R.A. (2013). Predictors of long-term outcome in multiple sclerosis patients
treated with interferon beta. Ann. Neurol. 73, 95–103.
Aranow, C. (2011). Vitamin D and the immune system. J. Investig. Med. 59,
881–886. Bettelli, E., Carrier, Y., Gao, W., Korn, T., Strom, T.B., Oukka, M., Weiner, H.L.,
and Kuchroo, V.K. (2006). Reciprocal developmental pathways for the gener-
Arnason, B., Jacobs, G., Hanlon, M., Clay, B., Noronha, A., Auty, A., Davis, B., ation of pathogenic effector TH17 and regulatory T cells. Nature 441, 235–238.
Nath, A., Bouchard, J., and Belanger, C.; The Lenercept Multiple Sclerosis
Study Group and The University of British Columbia MS/MRI Analysis Group Bevan, C.J., and Cree, B.A. (2014). Disease activity free status: a new end
(1999). TNF neutralization in MS: results of a randomized, placebo-controlled point for a new era in multiple sclerosis clinical research? JAMA Neurol. 71,
multicenter study. Neurology 53, 457–465. 269–270.

Arnon, R., and Aharoni, R. (2004). Mechanism of action of glatiramer acetate in Bhela, S., Kempsell, C., Manohar, M., Dominguez-Villar, M., Griffin, R., Bhatt,
multiple sclerosis and its potential for the development of new applications. P., Kivisakk-Webb, P., Fuhlbrigge, R., Kupper, T., Weiner, H., and Baecher-
Proc. Natl. Acad. Sci. USA 101 (Suppl 2 ), 14593–14598. Allan, C. (2015). Nonapoptotic and extracellular activity of granzyme B
mediates resistance to regulatory T cell (Treg) suppression by HLA-DR-
Astier, A.L., Meiffren, G., Freeman, S., and Hafler, D.A. (2006). Alterations CD25hiCD127lo Tregs in multiple sclerosis and in response to IL-6.
in CD46-mediated Tr1 regulatory T cells in patients with multiple sclerosis. J. Immunol. 194, 2180–2189.
J. Clin. Invest. 116, 3252–3257.
Bible, E. (2014). Multiple sclerosis: atacicept increases relapse rates in multiple
Baecher-Allan, C., Brown, J.A., Freeman, G.J., and Hafler, D.A. (2001). sclerosis. Nat. Rev. Neurol. 10, 182.
CD4+CD25high regulatory cells in human peripheral blood. J. Immunol. 167,
1245–1253. Bielekova, B., Goodwin, B., Richert, N., Cortese, I., Kondo, T., Afshar, G.,
Gran, B., Eaton, J., Antel, J., Frank, J.A., et al. (2000). Encephalitogenic
Baker, D., Marta, M., Pryce, G., Giovannoni, G., and Schmierer, K. (2017). potential of the myelin basic protein peptide (amino acids 83-99) in multiple
Memory B cells are major targets for effective immunotherapy in relapsing sclerosis: results of a phase II clinical trial with an altered peptide ligand.
multiple sclerosis. EBioMedicine 16, 41–50. Nat. Med. 6, 1167–1175.

760 Neuron 97, February 21, 2018


Neuron

Review
Bielekova, B., Catalfamo, M., Reichert-Scrivner, S., Packer, A., Cerna, M., patients have a distinct gut microbiota compared to healthy controls. Sci.
Waldmann, T.A., McFarland, H., Henkart, P.A., and Martin, R. (2006). Rep. 6, 28484.
Regulatory CD56(bright) natural killer cells mediate immunomodulatory effects
of IL-2Ralpha-targeted therapy (daclizumab) in multiple sclerosis. Proc. Natl. Chiarini, M., Serana, F., Zanotti, C., Capra, R., Rasia, S., Rottoli, M., Rovaris,
Acad. Sci. USA 103, 5941–5946. M., Caputo, D., Cavaletti, G., Frigo, M., et al. (2012). Modulation of the central
memory and Tr1-like regulatory T cells in multiple sclerosis patients responsive
Billerbeck, E., Kang, Y.H., Walker, L., Lockstone, H., Grafmueller, S., Fleming, to interferon-beta therapy. Mult. Scler. 18, 788–798.
V., Flint, J., Willberg, C.B., Bengsch, B., Seigel, B., et al. (2010). Analysis of
CD161 expression on human CD8+ T cells defines a distinct functional subset Cohen, J.A., Arnold, D.L., Comi, G., Bar-Or, A., Gujrathi, S., Hartung, J.P., Cra-
with tissue-homing properties. Proc. Natl. Acad. Sci. USA 107, 3006–3011. vets, M., Olson, A., Frohna, P.A., and Selmaj, K.W. (2016). Safety and efficacy
of the selective sphingosine 1-phosphate receptor modulator ozanimod in re-
Blair, P.A., Noreña, L.Y., Flores-Borja, F., Rawlings, D.J., Isenberg, D.A., lapsing multiple sclerosis (RADIANCE): a randomised, placebo-controlled,
Ehrenstein, M.R., and Mauri, C. (2010). CD19(+)CD24(hi)CD38(hi) B cells phase 2 trial. Lancet Neurol 15, 373–381.
exhibit regulatory capacity in healthy individuals but are functionally impaired
in systemic Lupus Erythematosus patients. Immunity 32, 129–140. Coles, A.J., Compston, D.A., Selmaj, K.W., Lake, S.L., Moran, S., Margolin,
D.H., Norris, K., and Tandon, P.K.; CAMMS223 Trial Investigators (2008).
Bosma, A., Abdel-Gadir, A., Isenberg, D.A., Jury, E.C., and Mauri, C. (2012). Alemtuzumab vs. interferon beta-1a in early multiple sclerosis. N. Engl. J.
Lipid-antigen presentation by CD1d(+) B cells is essential for the maintenance Med. 359, 1786–1801.
of invariant natural killer T cells. Immunity 36, 477–490.
Comabella, M. (2016). Tocilizumab and multiple sclerosis: a causal relation-
Bourdette, D., and Yadav, V. (2008). B-cell depletion with rituximab in relaps- ship? Clinical Commentary on the case report entitled–MS arising during
ing-remitting multiple sclerosis. Curr. Neurol. Neurosci. Rep. 8, 417–418. Tocilizumab therapy for rheumatoid arthritis. Mult. Scler. 22, 257–258.

Bra€ndle, S.M., Obermeier, B., Senel, M., Bruder, J., Mentele, R., Khademi, M., Comabella, M., Balashov, K., Issazadeh, S., Smith, D., Weiner, H.L., and
Olsson, T., Tumani, H., Kristoferitsch, W., Lottspeich, F., et al. (2016). Distinct Khoury, S.J. (1998). Elevated interleukin-12 in progressive multiple sclerosis
oligoclonal band antibodies in multiple sclerosis recognize ubiquitous self-pro- correlates with disease activity and is normalized by pulse cyclophosphamide
teins. Proc. Natl. Acad. Sci. USA 113, 7864–7869. therapy. J. Clin. Invest. 102, 671–678.

Burt, R.K., Balabanov, R., Han, X., Sharrack, B., Morgan, A., Quigley, K., Comabella, M., and Montalban, X. (2014). Body fluid biomarkers in multiple
Yaung, K., Helenowski, I.B., Jovanovic, B., Spahovic, D., et al. (2015). Associ- sclerosis. Lancet Neurol. 13, 113–126.
ation of nonmyeloablative hematopoietic stem cell transplantation with neuro-
logical disability in patients with relapsing-remitting multiple sclerosis. JAMA Comi, G. (2013). Disease-modifying treatments for progressive multiple scle-
313, 275–284. rosis. Mult. Scler. 19, 1428–1436.

Butovsky, O., Jedrychowski, M.P., Moore, C.S., Cialic, R., Lanser, A.J., Comi, G., Jeffery, D., Kappos, L., Montalban, X., Boyko, A., Rocca, M.A., and
Gabriely, G., Koeglsperger, T., Dake, B., Wu, P.M., Doykan, C.E., et al. Filippi, M. (2012). Placebo-controlled trial of oral laquinimod for multiple scle-
(2014). Identification of a unique TGF-b-dependent molecular and functional rosis. N. Engl. J. Med. 366, 1000–1009.
signature in microglia. Nat. Neurosci. 17, 131–143.
Correale, J., Farez, M., and Razzitte, G. (2008). Helminth infections associated
Cantarel, B.L., Waubant, E., Chehoud, C., Kuczynski, J., DeSantis, T.Z., with multiple sclerosis induce regulatory B cells. Ann. Neurol. 64, 187–199.
Warrington, J., Venkatesan, A., Fraser, C.M., and Mowry, E.M. (2015). Gut
microbiota in multiple sclerosis: possible influence of immunomodulators. Correale, J., Gaitán, M.I., Ysrraelit, M.C., and Fiol, M.P. (2017). Progressive
J. Investig. Med. 63, 729–734. multiple sclerosis: from pathogenic mechanisms to treatment. Brain 140,
527–546.
Cao, D., Malmström, V., Baecher-Allan, C., Hafler, D., Klareskog, L., and
Trollmo, C. (2003). Isolation and functional characterization of regulatory Cosorich, I., Dalla-Costa, G., Sorini, C., Ferrarese, R., Messina, M.J., Dolpady,
CD25brightCD4+ T cells from the target organ of patients with rheumatoid J., Radice, E., Mariani, A., Testoni, P.A., Canducci, F., et al. (2017). High
arthritis. Eur. J. Immunol. 33, 215–223. frequency of intestinal TH17 cells correlates with microbiota alterations and
disease activity in multiple sclerosis. Sci. Adv. 3, e1700492.
Cao, Y., Goods, B.A., Raddassi, K., Nepom, G.T., Kwok, W.W., Love, J.C., and
Hafler, D.A. (2015). Functional inflammatory profiles distinguish myelin-reac- Costantino, C.M., Baecher-Allan, C.M., and Hafler, D.A. (2008). Human regu-
tive T cells from patients with multiple sclerosis. Sci. Transl. Med. 7, 287ra74. latory T cells and autoimmunity. Eur. J. Immunol. 38, 921–924.

Caruana, P., Lemmert, K., Ribbons, K., Lea, R., and Lechner-Scott, J. (2017). Cree, B.A., Spencer, C.M., Varrin-Doyer, M., Baranzini, S.E., and Zamvil, S.S.
Natural killer cell subpopulations are associated with MRI activity in a relaps- (2016). Gut microbiome analysis in neuromyelitis optica reveals overabun-
ing-remitting multiple sclerosis patient cohort from Australia. Mult. Scler. 23, dance of Clostridium perfringens. Ann. Neurol. 80, 443–447.
1479–1487.
Cross, A.H., Stark, J.L., Lauber, J., Ramsbottom, M.J., and Lyons, J.A. (2006).
Castillo-Trivino, T., Braithwaite, D., Bacchetti, P., and Waubant, E. (2013). Rituximab reduces B cells and T cells in cerebrospinal fluid of multiple scle-
Rituximab in relapsing and progressive forms of multiple sclerosis: a system- rosis patients. J. Neuroimmunol. 180, 63–70.
atic review. PLoS ONE 8, e66308.
Cua, D.J., and Tato, C.M. (2010). Innate IL-17-producing cells: the sentinels of
Cepok, S., Jacobsen, M., Schock, S., Omer, B., Jaekel, S., Böddeker, I., the immune system. Nat. Rev. Immunol. 10, 479–489.
Oertel, W.H., Sommer, N., and Hemmer, B. (2001). Patterns of cerebrospinal
fluid pathology correlate with disease progression in multiple sclerosis. Brain Cuss, A.K., Avery, D.T., Cannons, J.L., Yu, L.J., Nichols, K.E., Shaw, P.J., and
124, 2169–2176. Tangye, S.G. (2006). Expansion of functionally immature transitional B cells
is associated with human-immunodeficient states characterized by impaired
Chataway, J., Martin, K., Barrell, K., Sharrack, B., Stolt, P., and Wraith, D. humoral immunity. J. Immunol. 176, 1506–1516.
(2018). Effects of ATX-MS-1467 immunotherapy over 16 weeks in
relapsing multiple sclerosis. Neurology. https://doi.org/10.1212/WNL. d’Hennezel, E., Ben-Shoshan, M., Ochs, H.D., Torgerson, T.R., Russell, L.J.,
0000000000005118. Lejtenyi, C., Noya, F.J., Jabado, N., Mazer, B., and Piccirillo, C.A. (2009).
FOXP3 forkhead domain mutation and regulatory T cells in the IPEX syndrome.
Chataway, J., Schuerer, N., Alsanousi, A., Chan, D., Macmanus, D., Hunter, K., N. Engl. J. Med. 361, 1710–1713.
Anderson, V., Bangham, C.R., Clegg, S., Nielsen, C., et al. (2014). Effect of
high-dose simvastatin on brain atrophy and disability in secondary progressive Daikeler, T., Tichelli, A., and Passweg, J. (2012). Complications of autologous
multiple sclerosis (MS-STAT): a randomised, placebo-controlled, phase 2 trial. hematopoietic stem cell transplantation for patients with autoimmune dis-
Lancet 383, 2213–2221. eases. Pediatr. Res. 71, 439–444.

Chen, J., Chia, N., Kalari, K.R., Yao, J.Z., Novotna, M., Soldan, M.M., Luckey, Dendrou, C.A., Fugger, L., and Friese, M.A. (2015). Immunopathology of mul-
D.H., Marietta, E.V., Jeraldo, P.R., Chen, X., et al. (2016). Multiple sclerosis tiple sclerosis. Nat. Rev. Immunol. 15, 545–558.

Neuron 97, February 21, 2018 761


Neuron

Review
Duan, H., Xing, S., Luo, Y., Feng, L., Gramaglia, I., Zhang, Y., Lu, D., Zeng, Q., Gobin, S.J., Montagne, L., Van Zutphen, M., Van Der Valk, P., Van Den Elsen,
Fan, K., Feng, J., et al. (2013). Targeting endothelial CD146 attenuates P.J., and De Groot, C.J. (2001). Upregulation of transcription factors control-
neuroinflammation by limiting lymphocyte extravasation to the CNS. Sci. ling MHC expression in multiple sclerosis lesions. Glia 36, 68–77.
Rep. 3, 1687.
Goodkin, D.E. (1996). Interferon beta treatment for multiple sclerosis: persist-
El Ayoubi, N.K., and Khoury, S.J. (2017). Blood biomarkers as outcome mea- ing questions. Mult. Scler. 1, 321–324.
sures in inflammatory neurologic diseases. Neurotherapeutics 14, 135–147.
Gray, O., McDonnell, G.V., and Forbes, R.B. (2004). Methotrexate for multiple
Endo, T., Scott, D.D., Stewart, S.S., Kundu, S.K., and Marcus, D.M. (1984). sclerosis. Cochrane Database Syst. Rev. (2), CD003208.
Antibodies to glycosphingolipids in patients with multiple sclerosis and SLE.
J. Immunol. 132, 1793–1797. Green, A.J., Gelfand, J.M., Cree, B.A., Bevan, C., Boscardin, W.J., Mei, F., In-
man, J., Arnow, S., Devereux, M., Abounasr, A., et al. (2017). Clemastine fuma-
Farez, M.F., Fiol, M.P., Gaitán, M.I., Quintana, F.J., and Correale, J. (2015). rate as a remyelinating therapy for multiple sclerosis (ReBUILD): a randomised,
Sodium intake is associated with increased disease activity in multiple scle- controlled, double-blind, crossover trial. Lancet 390, 2481–2489.
rosis. J. Neurol. Neurosurg. Psychiatry 86, 26–31.
Greenfield, A.L., and Hauser, S.L. (2017). B cell therapy for multiple sclerosis:
Fassas, A., Anagnostopoulos, A., Kazis, A., Kapinas, K., Sakellari, I., Kimiski-
entering an era. Ann. Neurol. Published online December 15, 2017. https://doi.
dis, V., and Tsompanakou, A. (1997). Peripheral blood stem cell transplanta-
org/10.1002/ana.25119.
tion in the treatment of progressive multiple sclerosis: first results of a pilot
study. Bone Marrow Transplant. 20, 631–638.
Gregori, S., Goudy, K.S., and Roncarolo, M.G. (2012). The cellular and molec-
ular mechanisms of immuno-suppression by human type 1 regulatory T cells.
Fisniku, L.K., Chard, D.T., Jackson, J.S., Anderson, V.M., Altmann, D.R., Mis-
Front. Immunol. 3, 30.
zkiel, K.A., Thompson, A.J., and Miller, D.H. (2008). Gray matter atrophy is
related to long-term disability in multiple sclerosis. Ann. Neurol. 64, 247–254.
Gross, C.C., Schulte-Mecklenbeck, A., Ru €nzi, A., Kuhlmann, T., Posevitz-
Fitzgerald, K.C., Munger, K.L., Hartung, H.P., Freedman, M.S., Montalbán, X., ,
Fejfár, A., Schwab, N., Schneider-Hohendorf, T., Herich, S., Held, K., Konjevic
Edan, G., Wicklein, E.M., Radue, E.W., Kappos, L., Pohl, C., and Ascherio, A.; M., et al. (2016). Impaired NK-mediated regulation of T-cell activity in multiple
BENEFIT Study Group (2017). Sodium intake and multiple sclerosis activity sclerosis is reconstituted by IL-2 receptor modulation. Proc. Natl. Acad. Sci.
and progression in BENEFIT. Ann. Neurol. 82, 20–29. USA 113, E2973–E2982.

Flores-Borja, F., Bosma, A., Ng, D., Reddy, V., Ehrenstein, M.R., Isenberg, Groux, H., O’Garra, A., Bigler, M., Rouleau, M., Antonenko, S., de Vries, J.E.,
D.A., and Mauri, C. (2013). CD19+CD24hiCD38hi B cells maintain regula- and Roncarolo, M.G. (1997). A CD4+ T-cell subset inhibits antigen-specific
tory T cells while limiting TH1 and TH17 differentiation. Sci. Transl. Med. 5, T-cell responses and prevents colitis. Nature 389, 737–742.
173ra23.
Gu, C., Wu, L., and Li, X. (2013). IL-17 family: cytokines, receptors and
Franklin, R.J. (2017). Regenerating CNS myelin—from mechanisms to exper- signaling. Cytokine 64, 477–485.
imental medicines. Nat. Rev. Neurosci. 18, 753.
Haas, J., Hug, A., Viehöver, A., Fritzsching, B., Falk, C.S., Filser, A., Vetter, T.,
Friese, M.A., and Fugger, L. (2005). Autoreactive CD8+ T cells in multiple scle- Milkova, L., Korporal, M., Fritz, B., et al. (2005). Reduced suppressive effect
rosis: a new target for therapy? Brain 128, 1747–1763. of CD4+CD25high regulatory T cells on the T cell immune response against
myelin oligodendrocyte glycoprotein in patients with multiple sclerosis. Eur.
Furlan, R., Brambilla, E., Ruffini, F., Poliani, P.L., Bergami, A., Marconi, P.C., J. Immunol. 35, 3343–3352.
Franciotta, D.M., Penna, G., Comi, G., Adorini, L., and Martino, G. (2001). Intra-
thecal delivery of IFN-gamma protects C57BL/6 mice from chronic-progres- Habib, J., Deng, J., Lava, N., Tyor, W., and Galipeau, J. (2015). Blood B cell and
sive experimental autoimmune encephalomyelitis by increasing apoptosis of regulatory subset content in multiple sclerosis patients. J. Mult. Scler. (Foster
central nervous system-infiltrating lymphocytes. J. Immunol. 167, 1821–1829. City) 2, 1000139.

Gagliani, N., Magnani, C.F., Huber, S., Gianolini, M.E., Pala, M., Licona-Limon, Hafler, D.A., Compston, A., Sawcer, S., Lander, E.S., Daly, M.J., De Jager,
P., Guo, B., Herbert, D.R., Bulfone, A., Trentini, F., et al. (2013). Coexpression P.L., de Bakker, P.I., Gabriel, S.B., Mirel, D.B., Ivinson, A.J., et al.; International
of CD49b and LAG-3 identifies human and mouse T regulatory type 1 cells. Multiple Sclerosis Genetics Consortium (2007). Risk alleles for multiple scle-
Nat. Med. 19, 739–746. rosis identified by a genomewide study. N. Engl. J. Med. 357, 851–862.
Gale, C.R., and Martyn, C.N. (1995). Migrant studies in multiple sclerosis. Prog. Harley, J.B., and James, J.A. (2006). Epstein-Barr virus infection induces lupus
Neurobiol. 47, 425–448. autoimmunity. Bull. NYU Hosp. Jt. Dis. 64, 45–50.
Gandhi, R., Healy, B., Gholipour, T., Egorova, S., Musallam, A., Hussain, M.S.,
Hauser, S.L., Bhan, A.K., Gilles, F., Kemp, M., Kerr, C., and Weiner, H.L. (1986).
Nejad, P., Patel, B., Hei, H., Khoury, S., et al. (2013). Circulating microRNAs as
Immunohistochemical analysis of the cellular infiltrate in multiple sclerosis
biomarkers for disease staging in multiple sclerosis. Ann. Neurol. 73, 729–740.
lesions. Ann. Neurol. 19, 578–587.
Garren, H., Robinson, W.H., Krasulova, E., Havrdová, E., Nadj, C., Selmaj, K.,
Hauser, S.L., Waubant, E., Arnold, D.L., Vollmer, T., Antel, J., Fox, R.J., Bar-Or,
Losy, J., Nadj, I., Radue, E.W., Kidd, B.A., et al. (2008). Phase 2 trial of a DNA
A., Panzara, M., Sarkar, N., Agarwal, S., et al.; HERMES Trial Group (2008).
vaccine encoding myelin basic protein for multiple sclerosis. Ann. Neurol. 63,
B-cell depletion with rituximab in relapsing-remitting multiple sclerosis.
611–620.
N. Engl. J. Med. 358, 676–688.
Gaublomme, J.T., Yosef, N., Lee, Y., Gertner, R.S., Yang, L.V., Wu, C., Pan-
dolfi, P.P., Mak, T., Satija, R., Shalek, A.K., et al. (2015). Single-cell genomics Hauser, S.L., Chan, J.R., and Oksenberg, J.R. (2013). Multiple sclerosis: pros-
unveils critical regulators of Th17 cell pathogenicity. Cell 163, 1400–1412. pects and promise. Ann. Neurol. 74, 317–327.

Gharibi, T., Majidi, J., Kazemi, T., Dehghanzadeh, R., Motallebnezhad, M., and Hauser, S.L., Bar-Or, A., Comi, G., Giovannoni, G., Hartung, H.P., Hemmer, B.,
Babaloo, Z. (2016). Biological effects of IL-21 on different immune cells and its Lublin, F., Montalban, X., Rammohan, K.W., Selmaj, K., et al.; OPERA I and
role in autoimmune diseases. Immunobiology 221, 357–367. OPERA II Clinical Investigators (2017). Ocrelizumab versus interferon beta-1a
in relapsing multiple sclerosis. N. Engl. J. Med. 376, 221–234.
Giovannoni, G., Comi, G., Cook, S., Rammohan, K., Rieckmann, P., Soelberg
Sørensen, P., Vermersch, P., Chang, P., Hamlett, A., Musch, B., and Green- Havrdová, E., Horakova, D., and Kovarova, I. (2015). Alemtuzumab in the treat-
berg, S.J.; CLARITY Study Group (2010). A placebo-controlled trial of oral ment of multiple sclerosis: key clinical trial results and considerations for use.
cladribine for relapsing multiple sclerosis. N. Engl. J. Med. 362, 416–426. Ther. Adv. Neurol. Disorder. 8, 31–45.

Giovannoni, G., Radue, E.W., Havrdova, E., Riester, K., Greenberg, S., Mehta, Havrdová, E., Belova, A., Goloborodko, A., Tisserant, A., Wright, A., Wall-
L., and Elkins, J. (2014). Effect of daclizumab high-yield process in patients stroem, E., Garren, H., Maguire, R.P., and Johns, D.R. (2016). Activity of secu-
with highly active relapsing-remitting multiple sclerosis. J. Neurol. 261, kinumab, an anti-IL-17A antibody, on brain lesions in RRMS: results from a
316–323. randomized, proof-of-concept study. J. Neurol. 263, 1287–1295.

762 Neuron 97, February 21, 2018


Neuron

Review
Healy, B.C., Liguori, M., Tran, D., Chitnis, T., Glanz, B., Wolfish, C., Gauthier, Kalluri, H.V., Sacha, L.M., Ingemi, A.I., Shullo, M.A., Johnson, H.J., Sood, P.,
S., Buckle, G., Houtchens, M., Stazzone, L., et al. (2010). HLA B*44: protective Tevar, A.D., Humar, A., and Venkataramanan, R. (2017). Low vitamin D expo-
effects in MS susceptibility and MRI outcome measures. Neurology 75, sure is associated with higher risk of infection in renal transplant recipients.
634–640. Clin. Transplant. 31, 31.

Hedström, A.K., Ba €a
€rnhielm, M., Olsson, T., and Alfredsson, L. (2011). Expo- Kaltsonoudis, E., Voulgari, P.V., Konitsiotis, S., and Drosos, A.A. (2014).
sure to environmental tobacco smoke is associated with increased risk for Demyelination and other neurological adverse events after anti-TNF therapy.
multiple sclerosis. Mult. Scler. 17, 788–793. Autoimmun. Rev. 13, 54–58.

Hedström, A.K., Hillert, J., Olsson, T., and Alfredsson, L. (2014). Reverse Kappos, L., Gold, R., Miller, D.H., Macmanus, D.G., Havrdova, E., Limmroth,
causality behind the association between reproductive history and MS. Mult. V., Polman, C.H., Schmierer, K., Yousry, T.A., Yang, M., et al.; BG-12
Scler. 20, 406–411. Phase IIb Study Investigators (2008). Efficacy and safety of oral fumarate in pa-
tients with relapsing-remitting multiple sclerosis: a multicentre, randomised,
Heidelberger, M., Kabat, E.A., and Mayer, M. (1942). A further study of the double-blind, placebo-controlled phase IIb study. Lancet 372, 1463–1472.
cross reaction between the specific polysaccharides of types Iii and Viii pneu-
mococci in horse antisera. J. Exp. Med. 75, 35–47. Kappos, L., Hartung, H.P., Freedman, M.S., Boyko, A., Radu €, E.W., Mikol,
D.D., Lamarine, M., Hyvert, Y., Freudensprung, U., Plitz, T., and van Beek,
Hemmer, B., Pinilla, C., Gran, B., Vergelli, M., Ling, N., Conlon, P., McFarland, J.; ATAMS Study Group (2014). Atacicept in multiple sclerosis (ATAMS): a
H.F., Houghten, R., and Martin, R. (2000). Contribution of individual amino randomised, placebo-controlled, double-blind, phase 2 trial. Lancet Neurol.
acids within MHC molecule or antigenic peptide to TCR ligand potency. 13, 353–363.
J. Immunol. 164, 861–871.
Kappos, L., Wiendl, H., Selmaj, K., Arnold, D.L., Havrdova, E., Boyko, A., Kauf-
Henderson, A.P., Barnett, M.H., Parratt, J.D., and Prineas, J.W. (2009). man, M., Rose, J., Greenberg, S., Sweetser, M., et al. (2015). Daclizumab HYP
Multiple sclerosis: distribution of inflammatory cells in newly forming lesions. versus interferon beta-1a in relapsing multiple sclerosis. N. Engl. J. Med. 373,
Ann. Neurol. 66, 739–753. 1418–1428.

Hensiek, A.E., Seaman, S.R., Barcellos, L.F., Oturai, A., Eraksoi, M., Cocco, E., Kappos, L., Bar-Or, A., Cree, B., Fox, R., Giovannoni, G., Gold, R., Vermersch,
Vecsei, L., Stewart, G., Dubois, B., Bellman-Strobl, J., et al. (2007). Familial P., Arnould, S., Sidorenko, T., Wolf, C., et al. (2016). Efficacy and safety of
effects on the clinical course of multiple sclerosis. Neurology 68, 376–383. siponimod in secondary progressive multiple sclerosis - results of the placebo
controlled, double-blind, Phase III EXPAND study. Proceedings of the 32nd
Hohol, M.J., Olek, M.J., Orav, E.J., Stazzone, L., Hafler, D.A., Khoury, S.J., Congress of the European Committee for Treatment and Research in Multiple
Dawson, D.M., and Weiner, H.L. (1999). Treatment of progressive multiple Sclerosis (ECTRIMS).
sclerosis with pulse cyclophosphamide/methylprednisolone: response to
therapy is linked to the duration of progressive disease. Mult. Scler. 5, Kappos, L., Bar-Or, A., Cree, B., Fox, R., Giovannoni, G., Gold, R., Vermersch,
403–409. P., Arnould, S., Sidorenko, T., Wolf, C., et al. (2017). Efficacy of siponimod
in secondary progressive multiple sclerosis: results of the phase 3 study
Hollenbach, J.A., and Oksenberg, J.R. (2015). The immunogenetics of multiple (CT.002). Neurology 88 (Suppl. 16 ), CT.002.
sclerosis: a comprehensive review. J. Autoimmun. 64, 13–25.
Karni, A., Abraham, M., Monsonego, A., Cai, G., Freeman, G.J., Hafler, D.,
Hong, J., Li, H., Chen, M., Zang, Y.C., Skinner, S.M., Killian, J.M., and Zhang, Khoury, S.J., and Weiner, H.L. (2006). Innate immunity in multiple sclerosis:
J.Z. (2009). Regulatory and pro-inflammatory phenotypes of myelin basic pro- myeloid dendritic cells in secondary progressive multiple sclerosis are acti-
tein-autoreactive T cells in multiple sclerosis. Int. Immunol. 21, 1329–1340. vated and drive a proinflammatory immune response. J. Immunol. 177, 4196–
4202.
Hu, D., Notarbartolo, S., Croonenborghs, T., Patel, B., Cialic, R., Yang, T.-H.,
Aschenbrenner, D., Andersson, K.M., Gattorno, M., Pham, M., et al. (2017). Kato, H., and Suzumura, A. (2003). [Cytokines in MS lesion]. Nihon Rinsho 61,
Transcriptional signature of human pro-inflammatory TH17 cells: Reduction 1428–1434.
of IL10 gene expression in multiple sclerosis. Nat. Commun. 8, 1600.
Katz, U., Kishner, I., Magalashvili, D., Shoenfeld, Y., and Achiron, A. (2006).
Huber, M., Heink, S., Pagenstecher, A., Reinhard, K., Ritter, J., Visekruna, A., Long term safety of IVIg therapy in multiple sclerosis: 10 years experience.
Guralnik, A., Bollig, N., Jeltsch, K., Heinemann, C., et al. (2013). IL-17A secre- Autoimmunity 39, 513–517.
tion by CD8+ T cells supports Th17-mediated autoimmune encephalomyelitis.
J. Clin. Invest. 123, 247–260. Keskinen, P., Ronni, T., Matikainen, S., Lehtonen, A., and Julkunen, I. (1997).
Regulation of HLA class I and II expression by interferons and influenza A virus
Huppert, J., Closhen, D., Croxford, A., White, R., Kulig, P., Pietrowski, E., in human peripheral blood mononuclear cells. Immunology 91, 421–429.
Bechmann, I., Becher, B., Luhmann, H.J., Waisman, A., and Kuhlmann, C.R.
(2010). Cellular mechanisms of IL-17-induced blood-brain barrier disruption. Kimura, A., and Kishimoto, T. (2010). IL-6: regulator of Treg/Th17 balance. Eur.
FASEB J. 24, 1023–1034. J. Immunol. 40, 1830–1835.

Infante-Duarte, C., Weber, A., Kra €tzschmar, J., Prozorovski, T., Pikol, S., Kinnunen, T., Chamberlain, N., Morbach, H., Cantaert, T., Lynch, M., Preston-
Hamann, I., Bellmann-Strobl, J., Aktas, O., Dörr, J., Wuerfel, J., et al. (2005). Hurlburt, P., Herold, K.C., Hafler, D.A., O’Connor, K.C., and Meffre, E. (2013).
Frequency of blood CX3CR1-positive natural killer cells correlates with dis- Specific peripheral B cell tolerance defects in patients with multiple sclerosis.
ease activity in multiple sclerosis patients. FASEB J. 19, 1902–1904. J. Clin. Invest. 123, 2737–2741.

Ivanov, I.I., Atarashi, K., Manel, N., Brodie, E.L., Shima, T., Karaoz, U., Wei, D., Klein, L., Kyewski, B., Allen, P.M., and Hogquist, K.A. (2014). Positive and
Goldfarb, K.C., Santee, C.A., Lynch, S.V., et al. (2009). Induction of intestinal negative selection of the T cell repertoire: what thymocytes see (and don’t
Th17 cells by segmented filamentous bacteria. Cell 139, 485–498. see). Nat. Rev. Immunol. 14, 377–391.

Iwata, Y., Matsushita, T., Horikawa, M., Dilillo, D.J., Yanaba, K., Venturi, G.M., Kleinewietfeld, M., and Hafler, D.A. (2014). Regulatory T cells in autoimmune
Szabolcs, P.M., Bernstein, S.H., Magro, C.M., Williams, A.D., et al. (2011). neuroinflammation. Immunol. Rev. 259, 231–244.
Characterization of a rare IL-10-competent B-cell subset in humans that par-
allels mouse regulatory B10 cells. Blood 117, 530–541. Kornek, B., and Lassmann, H. (1999). Axonal pathology in multiple sclerosis.
A historical note. Brain Pathol. 9, 651–656.
Jacobsen, M., Cepok, S., Quak, E., Happel, M., Gaber, R., Ziegler, A., Schock,
S., Oertel, W.H., Sommer, N., and Hemmer, B. (2002). Oligoclonal expansion Kuenz, B., Lutterotti, A., Ehling, R., Gneiss, C., Haemmerle, M., Rainer, C.,
of memory CD8+ T cells in cerebrospinal fluid from multiple sclerosis patients. Deisenhammer, F., Schocke, M., Berger, T., and Reindl, M. (2008). Cerebro-
Brain 125, 538–550. spinal fluid B cells correlate with early brain inflammation in multiple sclerosis.
PLoS ONE 3, e2559.
Jangi, S., Gandhi, R., Cox, L.M., Li, N., von Glehn, F., Yan, R., Patel, B.,
Mazzola, M.A., Liu, S., Glanz, B.L., et al. (2016). Alterations of the human gut Kuhle, J., Nourbakhsh, B., Grant, D., Morant, S., Barro, C., Yaldizli, Ö.,
microbiome in multiple sclerosis. Nat. Commun. 7, 12015. Pelletier, D., Giovannoni, G., Waubant, E., and Gnanapavan, S. (2017). Serum

Neuron 97, February 21, 2018 763


Neuron

Review
neurofilament is associated with progression of brain atrophy and disability in Lozano, J.M., González, R., Luque, J., Frias, M., Rivero, A., and Peña, J.
early MS. Neurology 88, 826–831. (2009). CD8(+)HLA-G(+) regulatory T cells are expanded in HIV-1-infected
patients. Viral Immunol. 22, 463–465.
€ck, W., Rauschka, H.,
Kutzelnigg, A., Lucchinetti, C.F., Stadelmann, C., Bru
Bergmann, M., Schmidbauer, M., Parisi, J.E., and Lassmann, H. (2005). Lu, L.F., and Rudensky, A. (2009). Molecular orchestration of differentiation
Cortical demyelination and diffuse white matter injury in multiple sclerosis. and function of regulatory T cells. Genes Dev. 23, 1270–1282.
Brain 128, 2705–2712.
Lu, W., Chen, S., Lai, C., Lai, M., Fang, H., Dao, H., Kang, J., Fan, J., Guo, W.,
Kwidzinski, E., Bunse, J., Aktas, O., Richter, D., Mutlu, L., Zipp, F., Nitsch, R., Fu, L., and Andrieu, J.M. (2016). Suppression of HIV replication by CD8(+)
and Bechmann, I. (2005). Indolamine 2,3-dioxygenase is expressed in the CNS regulatory T cells in elite controllers. Front. Immunol. 7, 134.
and down-regulates autoimmune inflammation. FASEB J. 19, 1347–1349.
Lublin, F.D., Reingold, S.C., Cohen, J.A., Cutter, G.R., Sørensen, P.S.,
Thompson, A.J., Wolinsky, J.S., Balcer, L.J., Banwell, B., Barkhof, F., et al.
Larochelle, C., Lécuyer, M.A., Alvarez, J.I., Charabati, M., Saint-Laurent, O.,
(2014). Defining the clinical course of multiple sclerosis: the 2013 revisions.
Ghannam, S., Kebir, H., Flanagan, K., Yednock, T., Duquette, P., et al.
Neurology 83, 278–286.
(2015). Melanoma cell adhesion molecule-positive CD8 T lymphocytes
mediate central nervous system inflammation. Ann. Neurol. 78, 39–53. Lucas, R.M., Byrne, S.N., Correale, J., Ilschner, S., and Hart, P.H. (2015).
Ultraviolet radiation, vitamin D and multiple sclerosis. Neurodegener. Dis.
Laroni, A., Armentani, E., Kerlero de Rosbo, N., Ivaldi, F., Marcenaro, E., Sivori, Manag. 5, 413–424.
S., Gandhi, R., Weiner, H.L., Moretta, A., Mancardi, G.L., and Uccelli, A. (2016).
Dysregulation of regulatory CD56(bright) NK cells/T cells interactions in multi- €ck, W., Parisi, J., Scheithauer, B., Rodriguez, M., and Lass-
Lucchinetti, C., Bru
ple sclerosis. J. Autoimmun. 72, 8–18. mann, H. (2000). Heterogeneity of multiple sclerosis lesions: implications for
the pathogenesis of demyelination. Ann. Neurol. 47, 707–717.
Lassmann, H. (2012). Cortical lesions in multiple sclerosis: inflammation versus
neurodegeneration. Brain 135, 2904–2905. Lucchinetti, C.F., Popescu, B.F., Bunyan, R.F., Moll, N.M., Roemer, S.F., Lass-
mann, H., Bru€ck, W., Parisi, J.E., Scheithauer, B.W., Giannini, C., et al. (2011).
Leandro, M.J. (2013). B-cell subpopulations in humans and their differential Inflammatory cortical demyelination in early multiple sclerosis. N. Engl. J. Med.
susceptibility to depletion with anti-CD20 monoclonal antibodies. Arthritis 365, 2188–2197.
Res. Ther. 15 (Suppl 1 ), S3.
Luger, D., Silver, P.B., Tang, J., Cua, D., Chen, Z., Iwakura, Y., Bowman, E.P.,
Lee, S.Y., and Goverman, J.M. (2013). The influence of T cell Ig mucin-3 Sgambellone, N.M., Chan, C.C., and Caspi, R.R. (2008). Either a Th17 or a
signaling on central nervous system autoimmune disease is determined by Th1 effector response can drive autoimmunity: conditions of disease induction
the effector function of the pathogenic T cells. J. Immunol. 190, 4991–4999. affect dominant effector category. J. Exp. Med. 205, 799–810.

€rner, K.H., Stellmann, J.P., Breiden,


Lutterotti, A., Yousef, S., Sputtek, A., Stu
Lee, Y.K., Menezes, J.S., Umesaki, Y., and Mazmanian, S.K. (2011). Proinflam-
matory T-cell responses to gut microbiota promote experimental autoimmune P., Reinhardt, S., Schulze, C., Bester, M., Heesen, C., et al. (2013). Antigen-
encephalomyelitis. Proc. Natl. Acad. Sci. USA 108 (Suppl 1 ), 4615–4622. specific tolerance by autologous myelin peptide-coupled cells: a phase 1 trial
in multiple sclerosis. Sci. Transl. Med. 5, 188ra75.
Lee, H.M., Bautista, J.L., Scott-Browne, J., Mohan, J.F., and Hsieh, C.S.
Lynch, S.V., and Pedersen, O. (2016). The human intestinal microbiome in
(2012a). A broad range of self-reactivity drives thymic regulatory T cell selec-
health and disease. N. Engl. J. Med. 375, 2369–2379.
tion to limit responses to self. Immunity 37, 475–486.
Ma, N., Xiao, H., Marrero, B., Xing, C., Wang, X., Zheng, M., Han, G., Chen, G.,
Lee, Y., Awasthi, A., Yosef, N., Quintana, F.J., Xiao, S., Peters, A., Wu, C., Klei- Hou, C., Shen, B., et al. (2014). Combination of TACI-IgG and anti-IL-15 treats
newietfeld, M., Kunder, S., Hafler, D.A., et al. (2012b). Induction and molecular murine lupus by reducing mature and memory B cells. Cell. Immunol. 289,
signature of pathogenic TH17 cells. Nat. Immunol. 13, 991–999. 140–144.

Leist, T.P., Comi, G., Cree, B.A., Coyle, P.K., Freedman, M.S., Hartung, H.P., Madsen, L.S., Andersson, E.C., Jansson, L., krogsgaard, M., Andersen, C.B.,
Vermersch, P., Casset-Semanaz, F., and Scaramozza, M. (2014). Effect of oral Engberg, J., Strominger, J.L., Svejgaard, A., Hjorth, J.P., Holmdahl, R., et al.
cladribine on time to conversion to clinically definite multiple sclerosis in pa- (1999). A humanized model for multiple sclerosis using HLA-DR2 and a human
tients with a first demyelinating event (ORACLE MS): a phase 3 randomised T-cell receptor. Nat. Genet. 23, 343–347.
trial. Lancet Neurol. 13, 257–267.
Magliozzi, R., Howell, O., Vora, A., Serafini, B., Nicholas, R., Puopolo, M., Rey-
Li, R., Rezk, A., Miyazaki, Y., Hilgenberg, E., Touil, H., Shen, P., Moore, C.S., nolds, R., and Aloisi, F. (2007). Meningeal B-cell follicles in secondary progres-
Michel, L., Althekair, F., Rajasekharan, S., et al.; Canadian B cells in MS sive multiple sclerosis associate with early onset of disease and severe cortical
Team (2015). Proinflammatory GM-CSF-producing B cells in multiple sclerosis pathology. Brain 130, 1089–1104.
and B cell depletion therapy. Sci. Transl. Med. 7, 310ra166.
Magliozzi, R., Howell, O.W., Reeves, C., Roncaroli, F., Nicholas, R., Serafini,
Liddelow, S.A., Guttenplan, K.A., Clarke, L.E., Bennett, F.C., Bohlen, C.J., B., Aloisi, F., and Reynolds, R. (2010). A Gradient of neuronal loss and menin-
€nch, A.E., Chung, W.S., Peterson, T.C., et al.
Schirmer, L., Bennett, M.L., Mu geal inflammation in multiple sclerosis. Ann. Neurol. 68, 477–493.
(2017). Neurotoxic reactive astrocytes are induced by activated microglia.
Nature 541, 481–487. Mahad, D.H., Trapp, B.D., and Lassmann, H. (2015). Pathological mechanisms
in progressive multiple sclerosis. Lancet Neurol. 14, 183–193.
Lindner, S., Dahlke, K., Sontheimer, K., Hagn, M., Kaltenmeier, C., Barth, T.F.,
Malmeström, C., Lycke, J., Haghighi, S., Andersen, O., Carlsson, L., Waden-
Beyer, T., Reister, F., Fabricius, D., Lotfi, R., et al. (2013). Interleukin
vik, H., and Olsson, B. (2008). Relapses in multiple sclerosis are associated
21-induced granzyme B-expressing B cells infiltrate tumors and regulate
with increased CD8+ T-cell mediated cytotoxicity in CSF. J. Neuroimmunol.
T cells. Cancer Res. 73, 2468–2479.
196, 159–165.
Lock, C., Hermans, G., Pedotti, R., Brendolan, A., Schadt, E., Garren, H., Malpass, K. (2012). Multiple sclerosis: ‘outside-in’ demyelination in MS. Nat.
Langer-Gould, A., Strober, S., Cannella, B., Allard, J., et al. (2002). Gene- Rev. Neurol. 8, 61.
microarray analysis of multiple sclerosis lesions yields new targets validated
in autoimmune encephalomyelitis. Nat. Med. 8, 500–508. Mancuso, R., Franciotta, D., Rovaris, M., Caputo, D., Sala, A., Hernis, A.,
Agostini, S., Calvo, M., and Clerici, M. (2014). Effects of natalizumab on oligo-
Losy, J. (2013). Is MS an inflammatory or primary degenerative disease? clonal bands in the cerebrospinal fluid of multiple sclerosis patients: a longitu-
J. Neural Transm. (Vienna) 120, 1459–1462. dinal study. Mult. Scler. 20, 1900–1903.

Lovato, L., Willis, S.N., Rodig, S.J., Caron, T., Almendinger, S.E., Howell, O.W., Mangalam, A., Shahi, S.K., Luckey, D., Karau, M., Marietta, E., Luo, N.,
Reynolds, R., O’Connor, K.C., and Hafler, D.A. (2011). Related B cell clones Choung, R.S., Ju, J., Sompallae, R., Gibson-Corley, K., et al. (2017). Human
populate the meninges and parenchyma of patients with multiple sclerosis. gut-derived commensal bacteria suppress CNS inflammatory and demyelin-
Brain 134, 534–541. ating disease. Cell Rep. 20, 1269–1277.

764 Neuron 97, February 21, 2018


Neuron

Review
Martin, R., Bielekova, B., Gran, B., and McFarland, H.F. (2000). Lessons from Nave, K.A., and Trapp, B.D. (2008). Axon-glial signaling and the glial support of
studies of antigen-specific T cell responses in Multiple Sclerosis. J. Neural axon function. Annu. Rev. Neurosci. 31, 535–561.
Transm. Suppl. (60), 361–373.
Nourbakhsh, B., Graves, J., Casper, T.C., Lulu, S., Waldman, A., Belman, A.,
Martinelli Boneschi, F., Vacchi, L., Rovaris, M., Capra, R., and Comi, G. (2013). Greenberg, B., Weinstock-Guttman, B., Aaen, G., Tillema, J.M., et al.; Network
Mitoxantrone for multiple sclerosis. Cochrane Database Syst. Rev. (5), of Pediatric Multiple Sclerosis Centers (2016). Dietary salt intake and time to
CD002127. relapse in paediatric multiple sclerosis. J. Neurol. Neurosurg. Psychiatry 87,
1350–1353.
Mauri, C., and Menon, M. (2015). The expanding family of regulatory B cells.
Int. Immunol. 27, 479–486. O’Connor, R.A., Malpass, K.H., and Anderton, S.M. (2007). The inflamed cen-
tral nervous system drives the activation and rapid proliferation of Foxp3+ reg-
Mayo, L., Trauger, S.A., Blain, M., Nadeau, M., Patel, B., Alvarez, J.I., Mascan- ulatory T cells. J. Immunol. 179, 958–966.
€kk, P., Kallas, K., et al. (2014). Regulation of astro-
froni, I.D., Yeste, A., Kivisa
cyte activation by glycolipids drives chronic CNS inflammation. Nat. Med. 20, O’Connor, P., Wolinsky, J.S., Confavreux, C., Comi, G., Kappos, L., Olsson,
1147–1156. T.P., Benzerdjeb, H., Truffinet, P., Wang, L., Miller, A., et al. (2011). Random-
ized trial of oral teriflunomide for relapsing multiple sclerosis. N. Engl. J.
Mayo, L., Cunha, A.P., Madi, A., Beynon, V., Yang, Z., Alvarez, J.I., Prat, A., So- Med. 365, 1293–1303.
bel, R.A., Kobzik, L., Lassmann, H., et al. (2016). IL-10-dependent Tr1 cells
attenuate astrocyte activation and ameliorate chronic central nervous system Ochoa-Repáraz, J., Mielcarz, D.W., Wang, Y., Begum-Haque, S., Dasgupta,
inflammation. Brain 139, 1939–1957. S., Kasper, D.L., and Kasper, L.H. (2010). A polysaccharide from the human
commensal Bacteroides fragilis protects against CNS demyelinating disease.
Melzer, N., Meuth, S.G., and Wiendl, H. (2009). CD8+ T cells and neuronal Mucosal Immunol. 3, 487–495.
damage: direct and collateral mechanisms of cytotoxicity and impaired elec-
trical excitability. FASEB J. 23, 3659–3673. Okuda, D.T., Srinivasan, R., Oksenberg, J.R., Goodin, D.S., Baranzini, S.E.,
Beheshtian, A., Waubant, E., Zamvil, S.S., Leppert, D., Qualley, P., et al.
Michel, L., Chesneau, M., Manceau, P., Genty, A., Garcia, A., Salou, M., Elong (2009). Genotype-Phenotype correlations in multiple sclerosis: HLA genes in-
Ngono, A., Pallier, A., Jacq-Foucher, M., Lefrère, F., et al. (2014a). Unaltered fluence disease severity inferred by 1HMR spectroscopy and MRI measures.
regulatory B-cell frequency and function in patients with multiple sclerosis. Brain 132, 250–259.
Clin. Immunol. 155, 198–208.
Olsson, T., Barcellos, L.F., and Alfredsson, L. (2017). Interactions between
Michel, L., Vukusic, S., De Seze, J., Ducray, F., Ongagna, J.C., Lefrere, F., genetic, lifestyle and environmental risk factors for multiple sclerosis. Nat.
Jacq-Foucher, M., Confavreux, C., Wiertlewski, S., and Laplaud, D.A. Rev. Neurol. 13, 25–36.
(2014b). Mycophenolate mofetil in multiple sclerosis: a multicentre retrospec-
tive study on 344 patients. J. Neurol. Neurosurg. Psychiatry 85, 279–283. Ottoboni, L., Keenan, B.T., Tamayo, P., Kuchroo, M., Mesirov, J.P., Buckle,
G.J., Khoury, S.J., Hafler, D.A., Weiner, H.L., and De Jager, P.L. (2012). An
Michel, L., Touil, H., Pikor, N.B., Gommerman, J.L., Prat, A., and Bar-Or, A.
RNA profile identifies two subsets of multiple sclerosis patients differing in dis-
(2015). B cells in the multiple sclerosis central nervous system: trafficking
ease activity. Sci. Transl. Med. 4, 153ra131.
and contribution to CNS-compartmentalized inflammation. Front. Immunol.
6, 636.
Pacholczyk, R., and Kern, J. (2008). The T-cell receptor repertoire of regulatory
T cells. Immunology 125, 450–458.
Miyake, S., Kim, S., Suda, W., Oshima, K., Nakamura, M., Matsuoka, T.,
Chihara, N., Tomita, A., Sato, W., Kim, S.W., et al. (2015). Dysbiosis in the
Palanichamy, A., Barnard, J., Zheng, B., Owen, T., Quach, T., Wei, C., Looney,
gut microbiota of patients with multiple sclerosis, with a striking depletion of
R.J., Sanz, I., and Anolik, J.H. (2009). Novel human transitional B cell popula-
species belonging to clostridia XIVa and IV clusters. PLoS ONE 10, e0137429.
tions revealed by B cell depletion therapy. J. Immunol. 182, 5982–5993.
Molnarfi, N., Schulze-Topphoff, U., Weber, M.S., Patarroyo, J.C., Prod’homme,
T., Varrin-Doyer, M., Shetty, A., Linington, C., Slavin, A.J., Hidalgo, J., et al. Palanichamy, A., Jahn, S., Nickles, D., Derstine, M., Abounasr, A., Hauser,
€dingen, H.C. (2014). Rituximab
S.L., Baranzini, S.E., Leppert, D., and von Bu
(2013). MHC class II-dependent B cell APC function is required for induction
of CNS autoimmunity independent of myelin-specific antibodies. J. Exp. Med. efficiently depletes increased CD20-expressing T cells in multiple sclerosis
210, 2921–2937. patients. J. Immunol. 193, 580–586.

Molnarfi, N., Bjarnadóttir, K., Benkhoucha, M., Juillard, C., and Lalive, P.H. Panitch, H.S., Hirsch, R.L., Haley, A.S., and Johnson, K.P. (1987a). Exacerba-
(2017). Activation of human B cells negatively regulates TGF-b1 production. tions of multiple sclerosis in patients treated with gamma interferon. Lancet 1,
J. Neuroinflammation 14, 13. 893–895.

Montalban, X., Belachew, S., and Wolinsky, J.S. (2017). Ocrelizumab in pri- Panitch, H.S., Hirsch, R.L., Schindler, J., and Johnson, K.P. (1987b). Treat-
mary progressive and relapsing multiple sclerosis. N. Engl. J. Med. 376, 1694. ment of multiple sclerosis with gamma interferon: exacerbations associated
with activation of the immune system. Neurology 37, 1097–1102.
Morandi, B., Bramanti, P., Bonaccorsi, I., Montalto, E., Oliveri, D., Pezzino, G.,
Navarra, M., and Ferlazzo, G. (2008). Role of natural killer cells in the pathogen- Plavina, T., Subramanyam, M., Bloomgren, G., Richman, S., Pace, A., Lee,
esis and progression of multiple sclerosis. Pharmacol. Res. 57, 1–5. S., Schlain, B., Campagnolo, D., Belachew, S., and Ticho, B. (2014). Anti-
JC virus antibody levels in serum or plasma further define risk of natalizu-
Mouzaki, A., Rodi, M., Dimisianos, N., Emmanuil, A., Kalavrizioti, D., mab-associated progressive multifocal leukoencephalopathy. Ann. Neurol.
Lagoudaki, R., Grigoriadis, N.C., and Papathanasopoulos, P. (2015). Immune 76, 802–812.
parameters that distinguish multiple sclerosis patients from patients with other
neurological disorders at presentation. PLoS ONE 10, e0135434. Polman, C., Barkhof, F., Sandberg-Wollheim, M., Linde, A., Nordle, O., and
Nederman, T.; Laquinimod in Relapsing MS Study Group (2005). Treatment
Munger, K.L., Levin, L.I., Hollis, B.W., Howard, N.S., and Ascherio, A. (2006). with laquinimod reduces development of active MRI lesions in relapsing MS.
Serum 25-hydroxyvitamin D levels and risk of multiple sclerosis. JAMA 296, Neurology 64, 987–991.
2832–2838.
Polman, C.H., O’Connor, P.W., Havrdová, E., Hutchinson, M., Kappos, L.,
Murphy, A.C., Lalor, S.J., Lynch, M.A., and Mills, K.H. (2010). Infiltration of Miller, D.H., Phillips, J.T., Lublin, F.D., Giovannoni, G., Wajgt, A., et al.; AFFIRM
Th1 and Th17 cells and activation of microglia in the CNS during the course Investigators (2006). A randomized, placebo-controlled trial of natalizumab for
of experimental autoimmune encephalomyelitis. Brain Behav. Immun. 24, relapsing multiple sclerosis. N. Engl. J. Med. 354, 899–910.
641–651.
Quintana, F.J., Farez, M.F., Viglietta, V., Iglesias, A.H., Merbl, Y., Izquierdo, G.,
Naismith, R.T., Piccio, L., Lyons, J.A., Lauber, J., Tutlam, N.T., Parks, B.J., Lucas, M., Basso, A.S., Khoury, S.J., Lucchinetti, C.F., et al. (2008). Antigen
Trinkaus, K., Song, S.K., and Cross, A.H. (2010). Rituximab add-on therapy microarrays identify unique serum autoantibody signatures in clinical and
for breakthrough relapsing multiple sclerosis: a 52-week phase II trial. pathologic subtypes of multiple sclerosis. Proc. Natl. Acad. Sci. USA 105,
Neurology 74, 1860–1867. 18889–18894.

Neuron 97, February 21, 2018 765


Neuron

Review
Racke, M.K., Lovett-Racke, A.E., and Karandikar, N.J. (2010). The mechanism T cell resistance to adaptive T(regs) involves IL-6-mediated signaling. Sci.
of action of glatiramer acetate treatment in multiple sclerosis. Neurology 74 Transl. Med. 5, 170ra15.
(Suppl 1 ), S25–S30.
Schubert, R.D., Hu, Y., Kumar, G., Szeto, S., Abraham, P., Winderl, J., Guthridge,
Ray, A., and Dittel, B.N. (2017). Mechanisms of regulatory B cell function in J.M., Pardo, G., Dunn, J., Steinman, L., and Axtell, R.C. (2015). IFN-b treatment
autoimmune and inflammatory diseases beyond IL-10. J. Clin. Med. 6, 6. requires B cells for efficacy in neuroautoimmunity. J. Immunol. 194, 2110–2116.

Read, S., Malmström, V., and Powrie, F. (2000). Cytotoxic T lymphocyte-asso- Schuh, E., Berer, K., Mulazzani, M., Feil, K., Meinl, I., Lahm, H., Krane, M.,
ciated antigen 4 plays an essential role in the function of CD25(+)CD4(+) reg- Lange, R., Pfannes, K., Subklewe, M., et al. (2016). Features of human
ulatory cells that control intestinal inflammation. J. Exp. Med. 192, 295–302. CD3+CD20+ T cells. J. Immunol. 197, 1111–1117.
Rissanen, E., Tuisku, J., Rokka, J., Paavilainen, T., Parkkola, R., Rinne, J.O., and Sedel, F., Papeix, C., Bellanger, A., Touitou, V., Lebrun-Frenay, C., Galanaud,
Airas, L. (2014). In vivo detection of diffuse inflammation in secondary progressive D., Gout, O., Lyon-Caen, O., and Tourbah, A. (2015). High doses of biotin in
multiple sclerosis using PET imaging and the radioligand 11C-PK11195. J. Nucl. chronic progressive multiple sclerosis: a pilot study. Mult. Scler. Relat. Disord.
Med. 55, 939–944. 4, 159–169.
Roll, P., Muhammad, K., Schumann, M., Kleinert, S., and Tony, H.P. (2012). RF
Segal, B.M., Constantinescu, C.S., Raychaudhuri, A., Kim, L., Fidelus-Gort, R.,
positivity has substantial influence on the peripheral memory B-cell compart-
and Kasper, L.H.; Ustekinumab MS Investigators (2008). Repeated subcu-
ment and its modulation by TNF inhibition. Scand. J. Rheumatol. 41, 180–185.
taneous injections of IL12/23 p40 neutralising antibody, ustekinumab, in pa-
tients with relapsing-remitting multiple sclerosis: a phase II, double-blind, pla-
Rothhammer, V., Mascanfroni, I.D., Bunse, L., Takenaka, M.C., Kenison, J.E.,
cebo-controlled, randomised, dose-ranging study. Lancet Neurol. 7, 796–804.
Mayo, L., Chao, C.C., Patel, B., Yan, R., Blain, M., et al. (2016). Type I inter-
ferons and microbial metabolites of tryptophan modulate astrocyte activity
Sellebjerg, F., Frederiksen, J.L., Nielsen, P.M., and Olesen, J. (1998). Double-
and central nervous system inflammation via the aryl hydrocarbon receptor.
blind, randomized, placebo-controlled study of oral, high-dose methylprednis-
Nat. Med. 22, 586–597.
olone in attacks of MS. Neurology 51, 529–534.
Rothhammer, V., Kenison, J.E., Tjon, E., Takenaka, M.C., de Lima, K.A.,
Borucki, D.M., Chao, C.C., Wilz, A., Blain, M., Healy, L., et al. (2017). Sphingo- Sellebjerg, F., Cadavid, D., Steiner, D., Villar, L.M., Reynolds, R., and Mikol, D.
sine 1-phosphate receptor modulation suppresses pathogenic astrocyte acti- (2016). Exploring potential mechanisms of action of natalizumab in secondary
vation and chronic progressive CNS inflammation. Proc. Natl. Acad. Sci. USA progressive multiple sclerosis. Ther. Adv. Neurol. Disorder. 9, 31–43.
114, 2012–2017.
Selmaj, K., Li, D.K., Hartung, H.P., Hemmer, B., Kappos, L., Freedman, M.S.,
Rotstein, D.L., Healy, B.C., Malik, M.T., Chitnis, T., and Weiner, H.L. (2015). Stu€ve, O., Rieckmann, P., Montalban, X., Ziemssen, T., et al. (2013). Siponi-
Evaluation of no evidence of disease activity in a 7-year longitudinal multiple mod for patients with relapsing-remitting multiple sclerosis (BOLD): an adap-
sclerosis cohort. JAMA Neurol. 72, 152–158. tive, dose-ranging, randomised, phase 2 study. Lancet Neurol. 12, 756–767.

Round, J.L., and Mazmanian, S.K. (2010). Inducible Foxp3+ regulatory T-cell Selter, R.C., Biberacher, V., Grummel, V., Buck, D., Eienbröker, C., Oertel,
development by a commensal bacterium of the intestinal microbiota. Proc. W.H., Berthele, A., Tackenberg, B., and Hemmer, B. (2013). Natalizumab treat-
Natl. Acad. Sci. USA 107, 12204–12209. ment decreases serum IgM and IgG levels in multiple sclerosis patients. Mult.
Scler. 19, 1454–1461.
Sakaguchi, S., Fukuma, K., Kuribayashi, K., and Masuda, T. (1985). Organ-
specific autoimmune diseases induced in mice by elimination of T cell Serafini, B., Rosicarelli, B., Magliozzi, R., Stigliano, E., and Aloisi, F. (2004).
subset. I. Evidence for the active participation of T cells in natural self-toler- Detection of ectopic B-cell follicles with germinal centers in the meninges
ance; deficit of a T cell subset as a possible cause of autoimmune disease. of patients with secondary progressive multiple sclerosis. Brain Pathol. 14,
J. Exp. Med. 161, 72–87. 164–174.

Sakaguchi, S., Sakaguchi, N., Shimizu, J., Yamazaki, S., Sakihama, T., Itoh, Sergott, R.C., Bennett, J.L., Rieckmann, P., Montalban, X., Mikol, D., Freuden-
M., Kuniyasu, Y., Nomura, T., Toda, M., and Takahashi, T. (2001). Immunologic sprung, U., Plitz, T., and Van Beek, J. (2015a). ATON: results from a Phase II
tolerance maintained by CD25+ CD4+ regulatory T cells: their common role in randomized trial of the B-cell targeting agent atacicept in patients with optic
controlling autoimmunity, tumor immunity, and transplantation tolerance. Im- neuritis. J. Neurol. Sci. 351, 174–178.
munol. Rev. 182, 18–32.
Sergott, R.C., Bennett, J.L., Rieckmann, P., Montalban, X., Mikol, D., Freuden-
Saraste, M., Irjala, H., and Airas, L. (2007). Expansion of CD56Bright natural sprung, U., Plitz, T., and van Beek, J.; ATON Trial Group (2015b). ATON: results
killer cells in the peripheral blood of multiple sclerosis patients treated with from a Phase II randomized trial of the B-cell-targeting agent atacicept in pa-
interferon-beta. Neurol. Sci. 28, 121–126. tients with optic neuritis. J. Neurol. Sci. 351, 174–178.
Sawcer, S., Ban, M., Maranian, M., Yeo, T.W., Compston, A., Kirby, A., Daly, Shi, F.D., and Van Kaer, L. (2006). Reciprocal regulation between natural killer
M.J., De Jager, P.L., Walsh, E., Lander, E.S., et al.; International Multiple cells and autoreactive T cells. Nat. Rev. Immunol. 6, 751–760.
Sclerosis Genetics Consortium (2005). A high-density screen for linkage in
multiple sclerosis. Am. J. Hum. Genet. 77, 454–467. Siewe, B., Wallace, J., Rygielski, S., Stapleton, J.T., Martin, J., Deeks, S.G.,
and Landay, A. (2014). Regulatory B cells inhibit cytotoxic T lymphocyte
Sawcer, S., Franklin, R.J., and Ban, M. (2014). Multiple sclerosis genetics.
(CTL) activity and elimination of infected CD4 T cells after in vitro reactivation
Lancet Neurol. 13, 700–709.
of HIV latent reservoirs. PLoS ONE 9, e92934.
Schadenberg, A.W., Vastert, S.J., Evens, F.C., Kuis, W., van Vught, A.J.,
Simpson, D., Noble, S., and Perry, C. (2002). Glatiramer acetate: a review of its
Jansen, N.J., and Prakken, B.J. (2011). FOXP3+ CD4+ Tregs lose suppressive
use in relapsing-remitting multiple sclerosis. CNS Drugs 16, 825–850.
potential but remain anergic during transient inflammation in human. Eur. J.
Immunol. 41, 1132–1142.
Simpson, S., Jr., Blizzard, L., Otahal, P., Van der Mei, I., and Taylor, B. (2011).
Schirmer, L., Srivastava, R., and Hemmer, B. (2014). To look for a needle in a Latitude is significantly associated with the prevalence of multiple sclerosis: a
haystack: the search for autoantibodies in multiple sclerosis. Mult. Scler. 20, meta-analysis. J. Neurol. Neurosurg. Psychiatry 82, 1132–1141.
271–279.
Sloka, J.S., and Stefanelli, M. (2005). The mechanism of action of methylpred-
Schleinitz, N., Vély, F., Harlé, J.R., and Vivier, E. (2010). Natural killer cells in nisolone in the treatment of multiple sclerosis. Mult. Scler. 11, 425–432.
human autoimmune diseases. Immunology 131, 451–458.
Smith, A.L., and Cohen, J.A. (2016). Multiple sclerosis: fingolimod failure in
Schmidt, A., Oberle, N., and Krammer, P.H. (2012). Molecular mechanisms of progressive MS informs future trials. Nat. Rev. Neurol. 12, 253–254.
treg-mediated T cell suppression. Front. Immunol. 3, 51.
Sørensen, P.S., and Blinkenberg, M. (2016). The potential role for ocrelizumab
Schneider, A., Long, S.A., Cerosaletti, K., Ni, C.T., Samuels, P., Kita, M., and in the treatment of multiple sclerosis: current evidence and future prospects.
Buckner, J.H. (2013). In active relapsing-remitting multiple sclerosis, effector Ther. Adv. Neurol. Disorder. 9, 44–52.

766 Neuron 97, February 21, 2018


Neuron

Review
Sørensen, P.S., Fazekas, F., and Lee, M. (2002). Intravenous immunoglobulin Tremlett, H., Fadrosh, D.W., Faruqi, A.A., Zhu, F., Hart, J., Roalstad, S.,
G for the treatment of relapsing-remitting multiple sclerosis: a meta-analysis. Graves, J., Lynch, S., and Waubant, E.; US Network of Pediatric MS Centers
Eur. J. Neurol. 9, 557–563. (2016). Gut microbiota in early pediatric multiple sclerosis: a case-control
study. Eur. J. Neurol. 23, 1308–1321.
Souto-Carneiro, M.M., Mahadevan, V., Takada, K., Fritsch-Stork, R., Nanki, T.,
Brown, M., Fleisher, T.A., Wilson, M., Goldbach-Mansky, R., and Lipsky, P.E. Tzartos, J.S., Friese, M.A., Craner, M.J., Palace, J., Newcombe, J., Esiri, M.M.,
(2009). Alterations in peripheral blood memory B cells in patients with active and Fugger, L. (2008). Interleukin-17 production in central nervous system-
rheumatoid arthritis are dependent on the action of tumour necrosis factor. infiltrating T cells and glial cells is associated with active disease in multiple
Arthritis Res. Ther. 11, R84. sclerosis. Am. J. Pathol. 172, 146–155.

Spain, R., Powers, K., Murchison, C., Heriza, E., Winges, K., Yadav, V., Ca- Uitdehaag, B.M., Barkhof, F., Coyle, P.K., Gardner, J.D., Jeffery, D.R., and
meron, M., Kim, E., Horak, F., Simon, J., et al. (2017). Lipoic acid in secondary Mikol, D.D. (2011). The changing face of multiple sclerosis clinical trial popula-
progressive MS: a randomized controlled pilot trial. Neurol. Neuroimmunol. tions. Curr. Med. Res. Opin. 27, 1529–1537.
Neuroinflamm. 4, e374.
Uss, E., Rowshani, A.T., Hooibrink, B., Lardy, N.M., van Lier, R.A., and ten
Srivastava, R., Aslam, M., Kalluri, S.R., Schirmer, L., Buck, D., Tackenberg, B., Berge, I.J. (2006). CD103 is a marker for alloantigen-induced regulatory
Rothhammer, V., Chan, A., Gold, R., Berthele, A., et al. (2012). Potassium CD8+ T cells. J. Immunol. 177, 2775–2783.
channel KIR4.1 as an immune target in multiple sclerosis. N. Engl. J. Med.
367, 115–123. Valencia, X., Stephens, G., Goldbach-Mansky, R., Wilson, M., Shevach, E.M.,
and Lipsky, P.E. (2006). TNF downmodulates the function of human
Stangel, M., Kuhlmann, T., Matthews, P.M., and Kilpatrick, T.J. (2017). CD4+CD25hi T-regulatory cells. Blood 108, 253–261.
Achievements and obstacles of remyelinating therapies in multiple sclerosis.
Nat. Rev. Neurol. 13, 742. van Oosten, B.W., Barkhof, F., Truyen, L., Boringa, J.B., Bertelsmann, F.W.,
von Blomberg, B.M., Woody, J.N., Hartung, H.P., and Polman, C.H. (1996).
Stankiewicz, J.M., Kolb, H., Karni, A., and Weiner, H.L. (2013). Role of immu- Increased MRI activity and immune activation in two multiple sclerosis patients
nosuppressive therapy for the treatment of multiple sclerosis. Neurotherapeu- treated with the monoclonal anti-tumor necrosis factor antibody cA2.
tics 10, 77–88. Neurology 47, 1531–1534.

Stein, M.S., Liu, Y., Gray, O.M., Baker, J.E., Kolbe, S.C., Ditchfield, M.R., Egan, van Oosten, B.W., Lai, M., Hodgkinson, S., Barkhof, F., Miller, D.H., Moseley,
G.F., Mitchell, P.J., Harrison, L.C., Butzkueven, H., and Kilpatrick, T.J. (2011). I.F., Thompson, A.J., Rudge, P., McDougall, A., McLeod, J.G., et al. (1997).
A randomized trial of high-dose vitamin D2 in relapsing-remitting multiple Treatment of multiple sclerosis with the monoclonal anti-CD4 antibody cM-
sclerosis. Neurology 77, 1611–1618. T412: results of a randomized, double-blind, placebo-controlled, MR-moni-
tored phase II trial. Neurology 49, 351–357.
Stephens, L.A., Malpass, K.H., and Anderton, S.M. (2009). Curing CNS
autoimmune disease with myelin-reactive Foxp3+ Treg. Eur. J. Immunol. 39, Viglietta, V., Baecher-Allan, C., Weiner, H.L., and Hafler, D.A. (2004). Loss of
1108–1117. functional suppression by CD4+CD25+ regulatory T cells in patients with mul-
tiple sclerosis. J. Exp. Med. 199, 971–979.
Streeter, H.B., Rigden, R., Martin, K.F., Scolding, N.J., and Wraith, D.C. (2015).
Villar, L.M., Sádaba, M.C., Roldán, E., Masjuan, J., González-Porqué, P., Vil-
Preclinical development and first-in-human study of ATX-MS-1467 for immu-
larrubia, N., Espiño, M., Garcı́a-Trujillo, J.A., Bootello, A., and Alvarez-Cer-
notherapy of MS. Neurol. Neuroimmunol. Neuroinflamm. 2, e93.
meño, J.C. (2005). Intrathecal synthesis of oligoclonal IgM against myelin lipids
Tada, T., Takemori, T., Okumura, K., Nonaka, M., and Tokuhisa, T. (1978). predicts an aggressive disease course in MS. J. Clin. Invest. 115, 187–194.
Two distinct types of helper T cells involved in the secondary antibody
Vollmer, T., Stewart, T., and Baxter, N. (2010). Mitoxantrone and cytotoxic
response: independent and synergistic effects of Ia- and Ia+ helper T cells.
drugs’ mechanisms of action. Neurology 74 (Suppl 1 ), S41–S46.
J. Exp. Med. 147, 446–458.
von Bu€dingen, H.C., Palanichamy, A., Lehmann-Horn, K., Michel, B.A., and
Tallantyre, E., Evangelou, N., and Constantinescu, C.S. (2008). Spotlight on
Zamvil, S.S. (2015). Update on the autoimmune pathology of multiple
teriflunomide. Int. MS J. 15, 62–68.
sclerosis: B-cells as disease-drivers and therapeutic targets. Eur. Neurol. 73,
238–246.
Tennakoon, D.K., Mehta, R.S., Ortega, S.B., Bhoj, V., Racke, M.K., and Karan-
dikar, N.J. (2006). Therapeutic induction of regulatory, cytotoxic CD8+ T cells Vuddamalay, Y., and van Meerwijk, J.P. (2017). CD28- and CD28lowCD8+
in multiple sclerosis. J. Immunol. 176, 7119–7129. regulatory T cells: of mice and men. Front. Immunol. 8, 31.
Tourbah, A., Lebrun-Frenay, C., Edan, G., Clanet, M., Papeix, C., Vukusic, S., Walczak, A., Siger, M., Ciach, A., Szczepanik, M., and Selmaj, K. (2013).
De Sèze, J., Debouverie, M., Gout, O., Clavelou, P., et al.; MS-SPI study group Transdermal application of myelin peptides in multiple sclerosis treatment.
(2016). MD1003 (high-dose biotin) for the treatment of progressive multiple JAMA Neurology 70, 1105–1109.
sclerosis: A randomised, double-blind, placebo-controlled study. Mult. Scler.
22, 1719–1731. Weber, M.S., Prod’homme, T., Youssef, S., Dunn, S.E., Rundle, C.D., Lee, L.,
€ve, O., Sobel, R.A., Steinman, L., and Zamvil, S.S. (2007).
Patarroyo, J.C., Stu
Tramacere, I., Del Giovane, C., Salanti, G., D’Amico, R., and Filippini, G. (2015). Type II monocytes modulate T cell-mediated central nervous system auto-
Immunomodulators and immunosuppressants for relapsing-remitting multi- immune disease. Nat. Med. 13, 935–943.
ple sclerosis: a network meta-analysis. Cochrane Database Syst. Rev. (9),
CD011381. Weiner, H.L., Dau, P.C., Khatri, B.O., Petajan, J.H., Birnbaum, G., Mcquillen,
M.P., Fosburg, M.T., Feldstein, M., and Orav, E.J. (1989). Double-blind study
Trapp, B.D., and Nave, K.A. (2008). Multiple sclerosis: an immune or neurode- of true vs. sham plasma exchange in patients treated with immunosuppression
generative disorder? Annu. Rev. Neurosci. 31, 247–269. for acute attacks of multiple sclerosis. Neurology 39, 1143–1149.
Trapp, B.D., Bö, L., Mörk, S., and Chang, A. (1999). Pathogenesis of tissue Weiner, H.L., Mackin, G.A., Matsui, M., Orav, E.J., Khoury, S.J., Dawson, D.M.,
injury in MS lesions. J. Neuroimmunol. 98, 49–56. and Hafler, D.A. (1993). Double-blind pilot trial of oral tolerization with myelin
antigens in multiple sclerosis. Science 259, 1321–1324.
Traugott, U. (1985). Characterization and distribution of lymphocyte subpopu-
lations in multiple sclerosis plaques versus autoimmune demyelinating lesions. Weinshenker, B.G. (2001). Plasma exchange for severe attacks of inflamma-
Springer Semin. Immunopathol. 8, 71–95. tory demyelinating diseases of the central nervous system. J. Clin. Apher.
16, 39–42.
Traugott, U., Reinherz, E.L., and Raine, C.S. (1983). Multiple sclerosis.
Distribution of T cells, T cell subsets and Ia-positive macrophages in lesions Willis, M.A., and Cohen, J.A. (2013). Fingolimod therapy for multiple sclerosis.
of different ages. J. Neuroimmunol. 4, 201–221. Semin. Neurol. 33, 37–44.

Tremlett, H., Zhao, Y., Rieckmann, P., and Hutchinson, M. (2010). New per- Wingerchuk, D.M. (2012). Smoking: effects on multiple sclerosis susceptibility
spectives in the natural history of multiple sclerosis. Neurology 74, 2004–2015. and disease progression. Ther. Adv. Neurol. Disorder. 5, 13–22.

Neuron 97, February 21, 2018 767


Neuron

Review
Wu, C., Yosef, N., Thalhamer, T., Zhu, C., Xiao, S., Kishi, Y., Regev, A., and Zhang, B., Liu, C., Qian, W., Han, Y., Li, X., and Deng, J. (2013). Crystal struc-
Kuchroo, V.K. (2013). Induction of pathogenic TH17 cells by inducible salt- ture of IL-17 receptor B SEFIR domain. J. Immunol. 190, 2320–2326.
sensing kinase SGK1. Nature 496, 513–517.
Zhao, Y., Healy, B.C., Rotstein, D., Guttmann, C.R., Bakshi, R., Weiner, H.L.,
Yang, M., Rui, K., Wang, S., and Lu, L. (2013). Regulatory B cells in autoim- Brodley, C.E., and Chitnis, T. (2017). Exploration of machine learning
mune diseases. Cell. Mol. Immunol. 10, 122–132. techniques in predicting multiple sclerosis disease course. PLoS ONE 12,
e0174866.

Zang, Y.C., Li, S., Rivera, V.M., Hong, J., Robinson, R.R., Breitbach, W.T., Kill- Zrzavy, T., Hametner, S., Wimmer, I., Butovsky, O., Weiner, H.L., and Lass-
ian, J., and Zhang, J.Z. (2004). Increased CD8+ cytotoxic T cell responses to mann, H. (2017). Loss of ‘homeostatic’ microglia and patterns of their activa-
myelin basic protein in multiple sclerosis. J. Immunol. 172, 5120–5127. tion in active multiple sclerosis. Brain 140, 1900–1913.

Zha, B., Wang, L., Liu, X., Liu, J., Chen, Z., Xu, J., Sheng, L., Li, Y., and Chu, Y. Zurawski, J., Lassmann, H., and Bakshi, R. (2017). Use of magnetic resonance
(2012). Decrease in proportion of CD19+ CD24(hi) CD27+ B cells and impair- imaging to visualize leptomeningeal inflammation in patients with multiple
ment of their suppressive function in Graves’ disease. PLoS ONE 7, e49835. sclerosis: a review. JAMA Neurology 74, 100–109.

768 Neuron 97, February 21, 2018

You might also like