Skip to main content
Access keys NCBI Homepage MyNCBI Homepage Main Content Main Navigation
Nat Rev Microbiol. 2023; 21(1): 51–64.
Published online 2022 Aug 5. doi: 10.1038/s41579-022-00770-5
PMCID: PMC9362539
PMID: 35931816

Epstein–Barr virus and multiple sclerosis

Abstract

Epstein–Barr virus (EBV) is a ubiquitous human lymphotropic herpesvirus with a well-established causal role in several cancers. Recent studies have provided compelling epidemiological and mechanistic evidence for a causal role of EBV in multiple sclerosis (MS). MS is the most prevalent chronic inflammatory and neurodegenerative disease of the central nervous system and is thought to be triggered in genetically predisposed individuals by an infectious agent, with EBV as the lead candidate. How a ubiquitous virus that typically leads to benign latent infections can promote cancer and autoimmune disease in at-risk populations is not fully understood. Here we review the evidence that EBV is a causal agent for MS and how various risk factors may affect EBV infection and immune control. We focus on EBV contributing to MS through reprogramming of latently infected B lymphocytes and the chronic presentation of viral antigens as a potential source of autoreactivity through molecular mimicry. We consider how knowledge of EBV-associated cancers may be instructive for understanding the role of EBV in MS and discuss the potential for therapies that target EBV to treat MS.

Subject terms: Virus-host interactions, Viral pathogenesis, Neurological disorders

Epstein–Barr virus infects most of the human population and, depending on other risk factors, contributes to the development of multiple sclerosis. In this Review, Soldan and Lieberman discuss supporting evidence and potential mechanisms that link Epstein–Barr virus to multiple sclerosis.

Introduction

Epstein–Barr virus (EBV) was the first human tumour virus identified, after its discovery in tumour cells of paediatric Burkitt lymphoma1,2. We now know that EBV is ubiquitous, establishing lifelong infection in more than 90% of adults worldwide3,4. Despite its typically subclinical persistence, EBV is consistently detected in numerous cancers, including nasopharyngeal carcinoma, subtypes of Hodgkin and non-Hodgkin lymphomas, a subtype of gastric carcinomas (EBV-associated gastric carcinoma), natural killer (NK)/T cell lymphomas and leiomyosarcomas. In addition, EBV has a profound effect on the immune system, and is the most common causal agent of infectious mononucleosis5 as well as fatal lymphoproliferative disorders in various immunosuppressive conditions6. Increasingly, it is appreciated that EBV is also a major risk factor for several autoimmune disorders, notably multiple sclerosis (MS)7,8.

MS is the most prevalent chronic inflammatory and neurodegenerative disease of the central nervous system (CNS). Approximately 2.8 million (35.9/100,000) people have MS worldwide9. MS incidence is also increasing in developing countries9 and among children10. The neurological signs and symptoms of MS include impaired motor function; visual symptoms; fatigue; eye movement disorders; bladder symptoms; sensory symptoms; sexual dysfunction; ataxia; deafness; spasticity; dementia; and cognitive impairment11. The clinical progression of MS is variable and unpredictable, with three distinct clinical courses: relapsing–remitting MS (RRMS), (2) secondary progressive MS (SPMS) and (3) primary progressive MS (PPMS)12,13. In addition, clinically isolated syndrome often progresses to MS, especially when symptoms are accompanied by CNS lesions.

The aetiology of MS is complex and multifactorial, involving the interplay of known genetic susceptibility factors, predominantly in genes directing the immune system, and environmental factors, including infectious agents, lack of sun exposure and vitamin D, smoking and obesity14. Infectious agents were first suspected in the aetiology of MS soon after its classification as a discrete clinical entity in the late 1800s15. The heterogeneity and the evolution of the disease throughout a patient’s lifetime and within the MS lesion itself have further obscured the identification of a single infectious agent as a consistent disease trigger. Nevertheless, epidemiological, serological and virological evidence has accumulated to support the role of EBV in the aetiology of MS, with recent large population-based studies demonstrating that EBV infection is likely a prerequisite for disease (reviewed in refs.7,1618; Table 1). In the most definitive epidemiology study on viruses and MS to date, more than ten million US army personnel were followed up over 20 years, and a 32-fold increased risk of MS diagnosis was shown in individuals who converted to EBV seropositivity compared with those who remained seronegative; this is the largest and most comprehensive study strongly suggesting that EBV infection is required for subsequent development of MS18,19. However, determining the precise mechanisms for EBV in the development of MS remains challenging because the virus is not always found in MS lesions. Related issues arise in the analysis of some EBV-associated cancers, in which EBV is present in only a subtype or subpopulation of cancer cells and oncogenesis depends on additional mutations or environmental cofactors20. Because most EBV infections do not cause disease, understanding the role of cofactors and aberrations in the normal infection process is key. One common cofactor in EBV disease is the disruption of normal immune control of EBV infection. Furthermore, as EBV infects and transforms B cells, we consider the intrinsic relationship of EBV with its host cells as a potential source of immune dysfunction.

Table 1

Selected studies providing evidence for a role of EBV in MS

EvidenceResultStudyRefs.
EpidemiologicalLow rates of MS in areas with more childhood infectionsReview230
Increased risk of MS with a history of infectious mononucleosisReview231
Increased risk of MS with EBV seroconversionHuman serum18
Decreased risk of MS in seronegative individualsHuman serum7,18
ImmunologicalIncreased levels of EBV-specific antibodies in MSReview121,122
MS-risk alleles enriched for transcription control by EBNA2Computational GWAS176,178
Deficient cytotoxic T lymphocyte control of EBV in MSMS CD8+ T cells216
EBV-reactive OCBsMS CSF85
Molecular mimicry between EBNA1 and CNS antigensMS B cells17,116
VirologicalIncreased shedding of EBV in saliva of paediatric patients with MSPaediatric MS55
EBV BZLF1 in MS lesionsMS brain166
Prosurvival influence of EBV latency genes on memory B cellsIn vitro232
EBV loads correlate with T-bet+CXCR3+ memory cells and IFNγ productionMS B cells171

CNS, central nervous system; CSF, cerebrospinal fluid; EBV, Epstein–Barr virus; GWAS, genome-wide association study; IFNγ, interferon-γ; MS, multiple sclerosis; OCBs, oligoclonal bands.

EBV biology and life cycle

EBV (human herpesvirus 4) is one of eight known human herpesviruses, with a large (173-kb) double-stranded DNA genome with approximately 100 protein-coding genes and numerous non-coding RNAs and microRNAs (miRNAs)21,22. Like all herpesviruses, EBV has both a productive (lytic) cycle and a non-productive (latent) phase. EBV establishes long-term latent infection of B lymphocytes and productive infection in the oral mucosal epithelium23. EBV DNA is packaged as a linear genome in the infectious viral particle, but persists in the nucleus of latently infected cells as a closed, circular chromatinized genome, referred to as an ‘episome’. Although only two distinct EBV strains have been delineated, the impact of genetic variation on the pathobiology of EBV infection is poorly understood22,24. Specific strains of EBV may be associated with MS, but conclusive MS genotypes have not been identified2527.

EBV is typically acquired through oral secretory transmission before the age of 5–8 years in resource-poor regions, whereas in resource-rich environments, infection is frequently delayed until adolescence or young adulthood5,28,29. During primary infection, the virus enters squamous epithelial cells and replicates within them, subsequently crossing the mucosal epithelial cell barrier via transcytosis and infecting local infiltrating B lymphocytes of Waldeyer’s tonsillar ring30. EBV infection of naive B lymphocytes initiates a developmental process and reprogramming similar to the germinal centre (GC) that results in long-lived memory B cells harbouring EBV episomes23,31. Lifelong persistence occurs through the establishment of latent reservoirs in these cells and periodic reactivation primarily in the oropharynx32, but other sites of EBV persistence, such as the gut mucosa or meninges, have been reported but not extensively characterized33,34.

EBV can enter various cell types through different mechanisms. The viral proteins gp350/gp220 and gp42 are required for EBV entry into B lymphocytes. Engagement of gp350 with CD21 (also known as complement receptor type 2) is followed by endocytosis of the virus into a low-pH component, where fusion is facilitated by the virus’s core fusion machinery (gB, gH and gL)35. gp42 subsequently binds to gH and interacts with HLA class II, which functions as a co-receptor35. Entry into epithelial cells occurs via a CD21-independent pathway. The EBV protein BMRF2 interacts with β1 integrin to trigger fusion and subsequent interactions between EBV gH/gL and αVβ6/8 integrins, thereby mediating endothelial cell fusion and entry36,37. More recently, ephrin receptor A2 (EphA2) was identified as an important entry factor for EBV in epithelial cells38. EphA2 genetic knockouts and inhibitors reduce infection of endothelial cells, and EphA2 interacts with gH/gL and gB38. In addition to B cells and epithelial cells, EBV can infect T cells, smooth muscle cells and NK cells, where the mechanism of entry is unclear39. EBV infection of T cells and NK cells is thought to be a rare event that can lead to the development of highly aggressive NK/T cell lymphomas and chronic active EBV infection40. EBV can also infect neuroblastoma cell lines and primary fetal astrocytes in vitro, although latent infection of neurons has not been unequivocally demonstrated in clinical specimens41,42.

EBV latent infection and B cell reprogramming

EBV infection efficiently reprogrammes naive B cells towards a developmental path recapitulating GC reaction, clonal expansion and differentiation towards a memory B cell phenotype23,31. These developmental stages correspond to different viral gene programmes termed ‘latency types’. During the hyperproliferative phase, EBV adopts a type III latency in which most latency-associated genes (EBNA1, EBNA2, EBNA3A, EBNA3B, EBNA3C, EBNA-LP, LMP1, LMP2 and multiple non-coding RNAs) are expressed43. Five latent genes (EBNA1, EBNA2, EBNA3A, EBNA3C and LMP1) are required for efficient B cell immortalization in vitro44. Different degrees of transcriptional silencing result in latency types II, I and 0, in which few or no viral genes are expressed. Importantly, all EBV-related cancers are associated with latent infection, and the different latency types correlate with different EBV-associated malignancies45. However, there can be considerable variation in viral gene expression among tumour cells and stages, including sporadic and abortive lytic reactivation. At present, it is unclear whether any specific latency type or lytic infection is associated with MS pathogenesis.

Viral reactivation and lytic gene expression

EBV lytic cycle reactivation occurs in healthy individuals, and is required for transmission and potentially for replenishing the latent reservoir. However, aberrant lytic activity is associated with several diseases, including oral hairy leukoplakia46 and chronic active EBV infection. EBV reactivation occurs through regulated stages with immediate-early genes controlling the expression of late genes and viral DNA replication, followed by virus assembly and egress22. Numerous cell signalling pathways can trigger the switch to lytic infection, depending on the host cell type. Many of these pathways are related to immune cell signalling, such as activation of B cell receptor (BCR) signalling with anti-immunoglobulin or activation of protein kinase C by phorbol esters47,48. In latently infected memory B cells, the switch requires two EBV-encoded transcription factors, BZLF1 (also known as ZTA, ZEBRA and Z) and BRLF1 (also known as RTA and R), which coordinately activate many of the EBV lytic genes22. Although EBV-related malignancies are associated with specific latency types, lytic gene expression has been shown in some tumour cells, and serology studies suggest that lytic antigen immunity precedes EBV-associated malignancies, particularly nasopharyngeal carcinoma, non-Hodgkin lymphoma and post-transplant lymphoproliferative disease49,50. In addition, highly sensitive genome-wide RNA sequencing methods have demonstrated expression of a subset of lytic genes in EBV-positive tumour cells51, suggesting that aberrant lytic gene expression and abortive lytic replication may occur more frequently in EBV-associated cancers and autoimmune disorders52. Defects in the control of EBV lytic reactivation have been suggested for MS, but the findings remain controversial5355.

Immune control of EBV

Immune responses to EBV infection differ widely and are influenced by genetics, the environment and age56 (Fig. 1). While primary infection before the age of 5 years is often asymptomatic, primary infection in adolescence can result in infectious mononucleosis. During mononucleosis, CD8+ T cells and NK cells rapidly expand in number. Most individuals maintain lifelong, effective immune control of the virus, where reactivation occasionally occurs but is quickly suppressed. This effective immune control is dominated by CD8+ T cells that target latently infected cells and early lytic replication57. Various EBV latency proteins, including EBNA2, EBNA3A, EBNA3B, EBNA3C and LMP2, are recognized by CD8+ T cells through major histocompatibility complex (MHC) class I presentation58. By contrast, immune response to EBNA1 peptides is mediated predominantly by T helper 1 (TH1)-polarized CD4+ T cells59,60. In addition, NK cells play a supportive role in controlling primary and lytic infection, while NK T cells and γδ T cells restrict latency types I and II (refs.5,6163). These cytotoxic lymphocytes also successfully restrict EBV in preclinical models, indicating that the cytolytic arm of the immune system must be engaged for efficient control of EBV infection45,6466. Study of primary genetic immunodeficiencies that are associated with an increased risk of EBV-associated disease has identified key immunoregulatory factors for controlling infection, such as the co-stimulatory proteins CD27, SLAM family members, magnesium transporter and the co-inhibitory CTLA-4 receptor67,68.

An external file that holds a picture, illustration, etc.
Object name is 41579_2022_770_Fig1_HTML.jpg
The maturation of the immune system, EBV infection and the development of MS.

The consequences of Epstein–Barr virus (EBV) infection are influenced by the age and genetic background of an individual. The risk of both infectious mononucleosis and multiple sclerosis (MS) increases when primary EBV infection occurs after the age of 10 years, when thymic negative selection of autoreactive T cells slows and T helper 1 (TH1) cell-mediated responses approach their peak. Most individuals receive a diagnosis of MS between the ages of 20 years and 50 years, years after EBV exposure. EBV infection increases the survival of memory B cells and causes lasting changes in the host cytokine response. There are many gaps in our understanding of how the maturation of the immune system triggers an evolving process of EBV-driven autoimmune reactivity leading to the development of MS. CMV, cytomegalovirus.

EBV deregulation of immune control

Despite a robust immune response to primary infection, EBV establishes a long-term latent infection in B lymphocytes, through a combination of viral reprogramming of B lymphocytes and disarming many innate and adaptive immune responses. EBV encodes numerous proteins that modulate the immune response. Some of these are expressed during the lytic or prelatent phase, while others are more consistently expressed during the latent infection. For example, EBNA1 can induce CXCL12 to recruit regulatory T cells69 and suppress NK cell responses by downregulating NKG2D ligands70. EBNA2 transcriptionally activates numerous genes involved in immune regulation, such as those encoding tumour necrosis factor (TNF)71, lymphotoxin-α72, IL-18R73 and PDL1 (refs.74,75). EBNA2 also suppresses interferon responses76 and HLA class II gene expression77. Virally encoded IL-10 (also known as BCRF1) suppresses pro-inflammatory cytokine secretion, such as secretion of IL-2 and interferon-γ (IFNγ), while viral BNLF2a inhibits the transporter associated with antigen processing (TAP)78. Multiple viral miRNAs target type I interferon pathways, such as IRF9, JAK1, JAK2 and RIG-I (ref.79). Functionally, EBV miRNAs suppress CD8+ T cell response and are required for the establishment of latent infection in humanized mice80. Thus, EBV encodes numerous genes that deregulate innate and adaptive immunity, and it is not yet clear which, if any, of these pathways are most involved in the pathobiology of MS.

Pathobiology of MS

The pathobiology of MS is notable for several immune abnormalities, which have been described extensively elsewhere. Briefly, oligoclonal bands in the cerebrospinal fluid (CSF) and elevated IgG concentrations in the CNS are hallmarks of MS, and can be used for diagnosis. Notably, oligoclonal bands are found in several neuroinflammatory disorders, and are typically directed against the pathogen implicated in the disease. By contrast, the oligoclonal bands in MS are reactive against multiple antigens, including viral antigens, bacterial antigens and self-antigens8186. Several studies have provided evidence of EBV infection or elevated immune responses to EBV within the CNS, while others have not replicated these findings. The presence of EBV-reactive and human herpesvirus 6-reactive oligoclonal bands and antibody reactivity to EBNA1 and EBNA2 epitopes have been reported in MS CSF85,87. In addition, cytotoxic T lymphocytes (CTLs) reactive to EBV lytic proteins have been detected in the CSF of patients with MS88. The presence of serum antibodies to EBNA1 has been correlated with elevated intrathecal IgG levels in patients with early MS, suggesting a role for EBV at the onset of MS symptoms.

Cytokine production is highly perturbed in MS, with a characteristic upregulation of several pro-inflammatory cytokines, including IL-12, TNF, IFNγ, lymphotoxin-α and osteopontin89. Before disease relapse, IL-10 secretion is downregulated and both IL-10 and TGFβ levels increase with disease remission89. Inflammatory B cells secreting higher levels of IL-10 and GM-CSF have also been identified in peripheral blood from patients with MS9092. The effects of immunomodulatory therapies in MS further underscore the role of the immune control. For example, IFNβ (type I interferon) treatment is therapeutic, while IFNγ (type II interferon) treatment exacerbates disease progression93, and functional studies have confirmed that the IFNα/β pathway is downregulated in the peripheral blood mononuclear cells of untreated patients with MS94. More recently, the important role of B cells in MS pathogenesis was revealed by the success of B cell depletion therapy targeting B cells, including anti-CD20 (see later)90.

Ultimately, the immune abnormalities in MS are associated with the development of focal demyelinating lesions (also known as plaques) in CNS white and grey matter and can be visualized by MRI. These lesions differ in size, distribution and cellular composition95. The neuropathological findings suggest that within the active lesion, inflammatory T cells, B cells, plasma cells, activated microglia and macrophages are associated with oligodendrocyte loss, demyelination and astrocyte activation as the lesion forms around veins and venules, expands into normal-appearing white matter and leads to the formation of gliotic scars9699. Within the active lesion, macrophages contain both early and late myelin degradation products. Inflammation is greatest in active lesions, but is also observed in other stages of MS plaques. Interestingly, relatively little inflammation is observed in the initial stages of white matter lesions, leading to debate as to whether lesions are initiated by a neurodegenerative or an inflammatory process and raising the possibility that initial tissue injury is initiated by lymphocyte-derived soluble factors that induce damage directly or via activation of microglia97. Inactive MS lesions are hypocellular, with loss of oligodendrocytes and myelin, astrocytosis, fewer myelin degradation products within macrophages and loss of axonal density96. In addition to demyelination, axonal loss occurs in both white matter and grey matter and, over time, there is atrophy of the brain100. Remyelination may occur as new oligodendrocytes regenerate; the extent of remyelination depends on many factors, including the location in the brain. Circumstances that determine whether inflammation within a lesion resolves and remyelinates or if it ‘smoulders’ are incompletely understood.

Immune cell composition within the lesion differs with respect to the type of MS and the stage of the lesion. T cell and B cell infiltration is greatest in active lesions in patients with RRMS. CD8+ T cells consistently outnumber CD4+ T cells in all sites of the MS lesion except for the perivascular and meningeal cuffs, where CD4+ T cells, CD20+ B cells and plasma cells predominate98,99, suggesting that CD8+ T cells are more important effectors in the immunopathogenesis of MS than previously appreciated. Notably, fewer brain lesions, fewer inflammatory cells and more spinal cord lesions are found in PPMS than in RRMS.

Immunogenetics of MS

The genetic contribution to MS susceptibility is complex and is extensively reviewed elsewhere101,102. The strongest genetic risk factor for MS is a specific haplotype of the highly polymorphic MHC103. Specifically, an increased risk of MS exists in individuals with the MHC class II alleles HLA-DR2 and HLA-DQw1 (ref.104), with the primary risk allele being HLA-DRB1*1501. Genome-wide studies have identified more than 200 MS-associated loci across the human genome, and approximately 30 are associated with the MHC locus105109. Most of these loci have well-ascribed functions in the immune system, while some are associated with myelin structure or mitochondrial function110113. Importantly, these studies also reveal shared genetic risk factors with other autoimmune conditions.

How does EBV increase the risk of MS?

MS has a complex aetiology, with multiple causative factors that can be further defined as either drivers or triggers114. EBV is a trigger (that is, it must be acquired before the onset of disease); however, its role as putative ‘driver’ of disease progression is poorly defined. Ongoing clinical studies using antivirals, vaccines and cell-based approaches targeting EBV in patients with MS (discussed later) are likely to elucidate the role of EBV as a driver of disease activity. The risk of MS increases approximately 32-fold with EBV infection, and more with symptomatic to severe infectious mononucleosis and HLA-DR2b (HLA-DRB1*1501b and HLA-DRA1*0101a)18. How these genetic and environmental factors compound risk in MS is not fully understood, and there remain many plausible mechanisms. Determining which of these are the most frequent drivers and how best to therapeutically intervene remain challenges. In this section, we discuss the evidence for EBV as a trigger and/or a driver in MS pathogenesis, and highlight critical questions that may elucidate the role of EBV as a trigger and potential driver of MS (Box 1).

Box 1 Critical questions and knowledge gaps

  • How do developmental changes in the human immune system impact the long-term control of Epstein–Barr virus (EBV) with respect to T cell responsiveness and latent B cell reservoir? And how does this inform our understanding of the timing of EBV infection and its subsequent lifetime latency and immune control?
  • What, if any, are the pathogenic roles of EBV in the central nervous system (CNS)? Do CNS-infiltrating immune cells harbour EBV or EBV-reactive immune cells, especially in multiple sclerosis (MS)? What are the specific dynamics of EBV infection in the CNS? Which cells are involved, and does this differ in patients with MS compared with healthy controls?
  • How does EBV reprogramming of B cells contribute to MS risk? Does EBV alter B cell antigen presentation and T cell miscommunication to drive autoimmunity? Does EBV rescue autoreactive B cells?
  • How do MS-risk alleles compound the effects of EBV latent infection in B cells? Is enhanced EBNA2 binding at MS-risk alleles sufficient to drive B cell autoimmunity?
  • How do EBV infection and the HLA-DR15 allele compound the risk of MS? Is there an altered presentation of EBV antigens or EBV-induced factors in this HLA haplotype?
  • Which EBV factors are most consistently associated with MS pathogenesis, and can this inform more selective drug design and immunotherapies?
  • How do effective MS therapies (for example, CD20 depletion, cladribine and CD52 depletion) affect EBV-positive cells, EBV infection cycle, the frequency of EBV-positive cells and EBV loads? Does deficient cytotoxic T lymphocyte control in MS lead to EBV reactivation and increased EBV antibody responses and CNS inflammation?

Molecular mimicry and mistaken self

Latent and persistent infection is a chronic source of viral antigenic stimulation. Several EBV antigens are the target of cross-reactive autoantibodies found in MS. This cross-reactivity between self-antigens and EBV antigens involves both cellular and humoral immune responses. Early studies found that patient-derived T cells autoreactive to myelin basic protein (MBP) were also cross-reactive to a wide range of viral peptides, including peptides from EBV115. Subsequent studies identified MBP-reactive T cells in patients with MS that cross-react with EBNA1 (ref.116). Similar cross-reactivities with EBNA1 were found for T cells autoreactive to anoctamin 2 (ref.117), α-crystallin B chain (CRYAB)88,118 and most recently glial cell adhesion molecule17. Mimicry has been reported for the lytic proteins BHRF1 and BPLF1 (ref.119). Peptides from these viral lytic proteins were found bound to the HLA-DR15 haplotype and were cross-reactive with the self-protein RASGRP2 as a target autoantigen, which is expressed in the brain and B cells and is targeted by brain-homing, autoreactive CD4+ T cells119.

Autoreactive antibodies in MS also cross-react with viral proteins, especially EBNA1 (ref.120). Higher levels of antibodies to EBNA1 are typically observed in both serum and CSF of patients with MS121,122. Elevated titres of antibodies to EBNA1 were found to have a genetic component beyond just HLA type123, and high titres of antibodies to EBNA1 are associated with an increased risk of MS124. Many of these EBNA1-specific antibodies are polyreactive, and it is not clear which antigen initiates the immunogenicity. In addition to viral mimicry, virus infection in peripheral tissue induces cellular stress proteins, such as CRYAB, that can mimic CNS tissues and elicit an autoimmune reaction to proteins such as myelin125. Interestingly, CRYAB-specific antibodies from patients with MS cross-react with EBNA1 (ref.126). Despite these correlations, the pathogenic role of autoreactive and EBV-cross-reactive antibodies in MS is not well established.

Why then do so many self-reactive immune responses in MS cross-react with EBV peptides and EBNA1 in particular? Peptide library analyses have identified several domains of EBNA1 that are recognized by autoreactive immune responses (Fig. 2). EBNA1 amino acids 391–410 peptide mimics CRYAB amino acids 1–15 with an overlapping sequence of RRPFF126. A similar domain of EBNA1 (amino acids 386–405) mimics glial cell adhesion molecule17. In the case of glial cell adhesion molecule, post-translational modification of the host protein increased cross-reactivity, providing a mechanism for epitope evolution and spreading in response to environmental signals. Some reactivity to EBNA1 was associated with germ line, unmutated BCR, suggesting that early antibodies have innate affinity for a region of EBNA1 (ref.17). Other studies have pointed to the glycine-rich regions of EBNA1, which generate repetitive, low-complexity peptides127. Autoreactive antibodies also react with peptides derived from the exposed surface of the EBNA1 DNA-binding domain, but not the DNA-binding interface itself, suggesting that the intact EBNA1–DNA complex is an important antigenic stimulus128. Paradoxically, EBNA1 also has immune-evasive properties. The internal Gly-Ala repeats (amino acids 90–303) suppress HLA presentation through multiple mechanisms, including inhibition of peptide processing129131, suppression of autophagy132 and translational suppression133,134 owing to the mRNA structure134136. How these activities are related to the high exposure of EBNA1 in autoimmune disease and what aspects of EBNA1 peptide presentation differ in patients with MS are unclear.

An external file that holds a picture, illustration, etc.
Object name is 41579_2022_770_Fig2_HTML.jpg
EBNA1 sequences and their potential role in molecular mimicry and autoreactivity.

Computational model of EBNA1 full-length protein233 indicating the most frequent peptide epitopes associated with multiple sclerosis autoimmunity. DNA is shown as solid, protein as ribbon. Epitopes are highlighted in magenta and boxed. A partial list of EBNA1 peptides and their autoimmune properties, including immune response type and cellular protein mimic. aa, amino acids; ANO2, anoctamin 2; β-SYN, β-synuclein; CRYAB, α-crystallin B chain; DBD, DNA-binding domain; (GA)n, glycine-alanine repeats; glialCAM, glial cell adhesion molecule; MBP, myelin basic protein; NTD, amino-terminal domain.

Rescue of autoreactive and inflammatory B cells

EBV is highly efficient at immortalizing naive and resting B cells (Fig. 3). However, it is not fully established which B cell subtypes may or may not be susceptible to EBV. EBV immortalization of a ‘forbidden’ autoreactive B cell clone has been proposed as a potential mechanism triggering MS137. EBV transformation could bypass the normal process of elimination of autoreactive B cells, although most of this selection occurs in the bone marrow at early stages of B cell development138. Similar mechanisms of immune evasion are proposed for EBV-associated cancers139 (Box 2). EBV immortalization bypasses many of the requirements for T cell help through the virally encoded CD40-like receptor LMP1 and BCR-like receptor LMP2 (refs.23,31). Their combined expression is sufficient to drive lymphomagenesis in transgenic mice140, and it is likely that these ligand-independent receptors provide signals that can rescue autoreactive B cells. EBV-infected B cells also express mature BCR and IgG without necessarily passing through GC selection, further enabling the survival of B cell clones reactive to self 141. EBV-infected B cells alter T cell interactions mediated by CD70–CD27 and OX40L that disable T cell control and enable B cell lymphomagenesis142,143. Whether these forbidden B cells are antigen-presenting cells or antibody-producing cells is not yet known. However, recent B cell depletion studies suggest that B cell subtypes, and not plasma cell numbers or overall circulating antibody levels, best correlate with CNS pathogenesis in patients with MS90.

An external file that holds a picture, illustration, etc.
Object name is 41579_2022_770_Fig3_HTML.jpg
EBV latency drives B cell survival of inflammatory B cells.

Epstein–Barr virus (EBV) infection promotes the proliferation and survival of memory B cells that may alter T cell control of EBV infection and autoimmune reactive B cells and T cells. EBV LMP1 and LMP2 function as CD40-like and B cell receptor (BCR)-like mimics to bypass T cell-dependent germinal centre reactions. EBNA1 and EBNA2 drive gene regulatory changes that may affect preferentially multiple sclerosis (MS)-risk alleles. EBV-induced viral and cellular factors may promote inflammation, driving autoreactivity through direct interaction with T cells or natural killer (NK) cells, as well as through soluble factors, including exosomes. EBNA1 is frequently processed as an antigenic epitope that can stimulate autoreactive B cell and T cell development. Although EBV-positive cells are shown as antigen-presenting cells, it is not known whether they actually present EBNA1 peptides or whether these are presented by uninfected antigen-presenting cells, including uninfected dendritic cells and macrophages that captured infected cell debris. OPN, osteopontin; TNF, tumour necrosis factor.

Box 2 Common themes of EBV-associated cancers and MS

There are several common features of Epstein–Barr virus (EBV) infection as an aetiological agent in both cancer and multiple sclerosis (MS). Most EBV-associated cancers result from EBV prolonging the survival of a cell that acquires additional oncogenic mutations or epigenetic changes that drive cancer cell evolution. Cancer may also arise from EBV entering a cell with precancerous mutations that may enable EBV to establish an oncogenic infection, such as a type II latency in an epithelial cell. It is also possible that EBV acquires mutations and induces epigenetic changes in the host cell that drive oncogenesis. These rare events amount to a significant incidence of cancer cases owing to the high prevalence and persistence of EBV. Similar types of aberrations may need to be considered for MS. Does EBV infect a rare ‘forbidden’ B cell? If so, what are the B cells that are infected in patients with MS, and how may these differ from non-pathogenic EBV-positive B cells that do not drive MS? Could EBV have acquired rare mutations or polymorphisms that drive MS? Because EBV is so ubiquitous and because it is usually acquired early in life, the question of how the virus may be tolerated as ‘self’ versus chronically rejected as ‘non-self’ may depend on the age at primary infection. Antigens acquired before a certain stage of immune development and presented in the appropriate HLA context may be considered self-antigens and acquire tolerance. Similarly, foreign antigens that mimic self-antigens may escape immune recognition by posing as self or exhausting T cells. EBV modulation of many B cell immunoregulatory genes is also likely to contribute to pathogenesis in both cancer and MS. Indeed, similarly to MS, infectious mononucleosis in adolescence increases the risk of developing Hodgkin lymphoma (an approximately fourfold increase). In Hodgkin lymphoma, EBV rescues defective germinal centre B cells from apoptosis and initiates early events in lymphomagenesis by altering normal B cell gene expression programmes139. Therefore, it is possible that an analogous EBV-mediated rescue of autoreactive B cells or other B cell subsets may set the stage for the development of MS.

EBV infiltrating the CNS

EBV-infected B cells migrate to the CNS, where they may have altered immune reactivities and are associated with EBV-associated diseases, including primary CNS lymphoma (Fig. 4). EBV-positive B cells and plasma cells have been identified after death in MS lesions in the CNS of patients with MS, but not in controls144146. EBV gene expression was a variable mixture of both latent transcripts (EBV-encoded small RNAs (EBERs), EBNA3A, LMP2A and LMP2B) and lytic transcripts (BZLF1 and gp350) in these brain lesions33,144,146148. In situ hybridization experiments revealed a significant number of EBER+ B cells and a small number of BZLF1+ cells, although some EBV-positive B cells were also found in the brains of controls148. EBV-infected plasma cells in the CNS have been found synapsed with cytotoxic CD8+ T cells, suggesting a local inflammatory interaction initiated by EBV-positive B cells in the CNS149. There is evidence that EBV establishes an extralymphatic viral sanctuary in the CNS150, especially in vulnerable individuals during infectious mononucleosis, in which approximately 50% of memory B cells can be EBV positive151. However, several studies failed to find evidence of  EBV-positive B cells in the CSF of patients with MS or MS lesions in the CNS152157. Some of these conflicting findings may be due to technical challenges of detecting transient EBV gene expression in migratory B cells in the CNS of patients with MS and post-mortem samples158.

An external file that holds a picture, illustration, etc.
Object name is 41579_2022_770_Fig4_HTML.jpg
Mechanisms of EBV-mediated inflammatory cascades in periphery and CNS.

Epstein–Barr virus (EBV) may drive inflammatory events in both the periphery and the central nervous system (CNS), leading to the development of the multiple sclerosis lesion in the CNS. EBV immune-evasive features and risk-associated immune deficiencies promote EBV inflammatory cascades in the periphery. CNS pathogenesis may be initiated through multiple mechanisms, including natural and EBV-driven CNS trafficking of autoreactive B cells and T cells, molecular mimicry driven by chronic EBV infection, EBV-driven inflammatory cytokines and exosomes, and aberrant EBV lytic infection and tropisms owing to deficient immune control. CTL, cytotoxic T lymphocyte; glialCAM, glial cell adhesion molecule; MBP, myelin basic protein; OPN, osteopontin.

Deficient CTL control of EBV infection

T cell control of EBV infection is required for homeostatic viral persistence, and immune dysregulation is observed in all EBV-associated disease. In healthy carriers of  latent EBV infection (more than 90% of the adult population), nearly 1% of all T cells are reactive to EBV latent or lytic antigens159,160. Immune response to EBV is frequently skewed in patients with MS. Higher titres of EBNA1-reactive IgG are found several years before the onset of MS symptoms and correlate with MS risk8,161,162. EBNA1-specific T cell frequencies increase and specificities broaden in MS. CD4+ T cells show TH1 polarization and CD8+ T cell responses correlate with disease activity163166. MS progression correlates with a decreased functionality of EBV-specific CD4+ T cells and CD8+ T cells, as measured by IFNγ production and cytotoxic activity167169. T cell exhaustion may partly account for the failure to control chronic EBV infection. Developmental changes in the immune system are also critical for control of EBV infection. Childhood experience (time of exposure to EBV and geographical risk) indicate that immune system maturation, exposure and education are important components of MS aetiology8. Poorly defined, idiosyncratic CTL deficiencies may also enable EBV-positive B cells to proliferate, migrate to the CNS and express inflammatory viral and host factors169.

EBV-associated inflammation

Both B cells and T cells from patients with MS have atypical inflammatory features. Patients with MS with high EBV loads have T-bet+CXCR3+ memory B cells induced by IFNγ and TLR9 signals and EBV-reactive CTLs autoreactive to neuronal tissue170. EBV load correlated with the early emergence of CXCR3+ class-switched memory B cells, GC-like B cell development and trafficking of these cells to the CNS in mice170. These CXCR3+ B cells had enhanced ability to secrete anti-EBNA1 IgG171. It is important to note that EBV-negative B cells from patients with MS also have inflammatory features, and memory B cell subsets, in particular, were found to secrete higher levels of GM-CSF in patients with MS relative to healthy controls92. EBV-infected B cells produce high levels of inflammatory cytokines and exosomes that contain inflammatory components, including small viral nucleic acids, such as EBERs and miRNAs172. Exosomes containing EBERs with 5′-triphosphate pathogen-associated molecular patterns stimulated dendritic cell antiviral inflammatory activity, similar to systemic lupus erythematosus172. EBV miRNAs, which can be transported in exosomes, can target MS risk-associated genes, such as ZC3HAV1 regulating interferon response173. Exosomes may cross the blood–brain barrier, and are endocytosed by brain microvascular endothelial cells174. Therefore, it is possible that EBV-positive B cells in the periphery produce exosomes that cross into the CNS and/or that EBV-positive B cells in the CNS are a source of these inflammatory exosomes (Fig. 4). EBV-positive B cells may also induce autoreactive T cells through modification of their antigen presentation8,90.

Deregulation of B cell gene expression and autoimmune control

EBV encodes several transcriptional regulators and signalling molecules that reprogramme B cell gene networks implicated in cancer and autoimmunity. The latency nuclear regulatory factor EBNA2 is essential for EBV immortalization and drives B cell proliferation175. EBNA2 interacts with several host transcription factors, and studies involving chromatin immunoprecipitation followed by sequencing revealed that EBNA2 binds to almost half of the risk alleles for seven autoimmune disorders176. Genome-wide chromatin accessibility (assay for transposase-accessible chromatin using sequencing) and DNA looping (HiC) further demonstrated the role of EBNA2 in altering chromatin structure at many autoimmune genetic risk alleles177. Risk alleles were enriched for EBNA2 binding relative to non-risk alleles, as demonstrated for a few specific examples, such as ZMIZ1 (ref.177).

Genome-wide linkage studies have further implicated EBV as a regulator of MS-risk alleles178. Expression quantitative trait locus analysis found that genes located near MS-risk SNPs were linked with EBV type III latency. These genes include BATF, IRF5, IRF7 and STAT genes. In a related study, EBNA2 bound preferentially to five of six MS-risk alleles, relative to non-risk alleles, and a peptide inhibitor that disrupts EBNA2 interaction with the cellular transcription factor RBPJ altered high-risk allele expression71. Thus, MS-risk alleles could increase the efficiency of EBNA2 to promote B cell survival and immortalization71. EBNA2 targets also overlap with those of vitamin D receptor, which is another risk factor for MS179. Furthermore, polymorphisms in EBNA2 correlate with MS risk, suggesting that the virus strain may also be a risk factor180. The precise mechanism of gene deregulation in MS may be further nuanced and influenced by epigenetic control. DNA methylation and genomic imprinting of alleles associated with MS have been implicated in MS181,182. For example, HLA-DRB*1501 is hypomethylated and expressed at high levels in antigen-presenting cells in patients with MS183,184. Alternative splicing has been seen in MS B cells, and may be related to EBV transcriptional reprogramming185.

EBV encodes several other transcription regulatory factors that can influence B cell biology. The EBV lytic activator BZLF1 is a potent transcriptional regulator of numerous viral and cellular genes. BZLF1 expression has been identified in plasma B cells in post-mortem brain samples from patients with MS and has been associated with reactive cytotoxic CD8+ T cell infiltration166. EBV-induced G protein-coupled receptor 2 (EBI2; also known as GPR183) is a G protein receptor for dihydroxycholesterol, which is overexpressed in MS lesions and involved in migration of CD4+ T cells186.

EBV genomes are also regulated by epigenetic modification, especially DNA methylation, which can impact viral gene expression and latency type187,188. Epigenetic control of EBV is an important component of EBV cancer aetiology, but its role in autoimmune disease is not well described. Studies in MS patients and animal models have identified gene variants, miRNAs and viral co-factors that exert epigenetic control to increase inflammation, immune cell differentiation and myelin breakdown189. Epigenetic modification of genes that promote neuroinvasion of EBV-positive B cells, including the genes encoding osteopontin and CXCR4, has been described in some experimental models, suggesting that EBV may affect epigenetic mechanisms driving MS190.

EBV interactions with HLA

HLA alleles have different binding affinities and specificities for antigenic peptides that impact T cell immunogenicity and functionality191. Antigenic peptides derived from MBP have been identified from B cells from patients with MS and correlate with higher levels of MBP-specific T cells in patients with MS than in controls81. Related studies implicate variant peptide binding of the high-risk HLA-DRB1*15 allele in the presentation of various autoreactive peptides. Some of these peptides may be derived from EBV proteins, providing a potential mechanism to explain the combined risk of EBV infection and HLA-DRB1*15. For example, humanized mice reconstituted with HLA-DR15 had elevated CD8+ T cell responses and CD4+ T cells cross-reacting with MBP192. In addition, some studies have found that HLA-DR15 and HLA-DRB*07 patients with MS have higher EBV viral loads, whereas HLA-A*02 individuals have lower viral loads, suggesting that class I and class II MHC molecules modulate EBV latency control193,194. However, other studies did not find increased EBV viral loads in MS or changes that immediately precede or coincide with relapses195197. Nevertheless, HLA-A*02 correlates with a decreased risk of MS (reviewed in ref.198). Alternatively, but perhaps related, MS-associated risk alleles, including HLA-DRB5, are also correlated with differentially regulated gene expression199,200. Higher expression levels of HLA alleles may also affect peptide selection and presentation that contribute to peptide mimicry and autoreactivity. Another intriguing finding is that the HLA-DR15 allele can serve as a co-receptor for EBV entry into B cells, raising the possibility that viral entry pathways may also contribute to MS risk201.

Opportunities for therapeutic intervention

Existing immunomodulatory therapies and their potential effect on EBV

The effectiveness of immunosuppressive and anti-inflammatory therapies in MS supports the autoimmune component in disease pathogenesis. Corticosteroids effectively treat MS flares202, but are too immunosuppressive for long-term use. Several immunosuppressive and chemotherapeutic drugs that dramatically decrease the levels of circulating immune cells, including cyclophosphamide, cladribine, mitoxantrone, methotrexate and teriflunomide, have been used with variable success203,204. It is now appreciated that B cells play an essential role in MS pathogenesis, on the basis of the success of CD20-specific depletion. Monoclonal antibodies to the B cell antigen CD20 (ocrelizumab and ofatumumab) reduce MS relapse and lesion formation, while a monoclonal antibody (anti-IL-12 p40 and anti-IL-23 p40, ustekinumab) that targets both TH1 cells and TH17 cells did not show similar efficacy205207. Importantly, additional therapeutics that broadly target B cells, including anti-CD52 monoclonal antibody and cladribine act as B cell-depleting drugs and are of therapeutic use in MS. By contrast, treatments that target naive and plasma B cells (for example, atacicept) or boost memory B cells (for example, infliximab) further aggravate MS via TNF blockade208. The effects of these treatments on EBV load is not yet known. Interestingly, teriflunomide has been shown to reduce both EBV-induced lymphoproliferation and lytic viral replication209.

EBV-specific CTL therapy

Cell-based immunotherapies, including EBV-specific CTL lines, have proven successful in the treatment of post-transplantation lymphoproliferative disorder and EBV-associated lymphomas and nasopharyngeal carcinoma, with low rates of graft-versus-host disease210212. Therefore, the use of autologous T cell therapy has been expanded to clinical trials in MS213216. These therapies attempt to compensate for deficient CTL control of EBV-infected B cells. Phase I trials using ATA188, an allogenic T cell therapy using T cells from healthy donors, have been initiated to evaluate allogenic EBV CTL therapy in PPMS and SPMS (NCT03283826), and the first clinical episode highly suggestive of MS (NCT02912897). Initial reports have demonstrated increased circulation of LMP-reactive and lymphoblastoid cell line (LCL)-reactive effector CD8+ memory cell populations. Notably, patients with PPMS have reported clinical improvement after autologous EBV-specific T cell therapy targeting EBNA1, LMP1 and LMP2A214. Early results suggest that ATA188 is safe and well tolerated, with a decrease in Expanded Disability Status Scale (EDSS) score217.

Antivirals, vaccines and their potential to target EBV in MS pathogenesis

Specific antivirals for treating EBV infection have not, to date, been approved for treatment of MS. Moreover, several clinical trials testing the efficacy of antivirals, specifically those with broad antiherpesvirus activity, including acyclovir and valacyclovir, did not demonstrate a clear benefit in MS218220. IFNβ, a cytokine with broad antiviral, antiproliferative and anti-inflammatory effects, was the first immunomodulatory therapy to successfully modify the disease course of MS, and is still one of the most frequently used therapeutic options for MS; it is considered a first-line therapy with modest efficacy in controlling ongoing disease221. The exact mode of action of IFNβ in MS is only partly understood. IFNβ has potent antiviral activity and is known to counteract many immunomodulatory actions of EBV222,223. Antiviral nucleoside analogues may also be effective for treating EBV infection in MS. Recent studies have shown that the non-cyclic nucleoside analogue tenofovir alafenamide (TAF), which was developed as a specific inhibitor for the HIV and hepatitis B virus reverse transcriptases and is frequently used in HIV pre-exposure prophylaxis, also inhibits the EBV DNA polymerase224. Notably, TAF was twice as potent as ganciclovir in direct inhibition of EBV replication and DNA polymerase activity225. In addition, anecdotal reports and case studies have suggested that there may be a clinical benefit and decreased relapses in patients with RRMS receiving TAF regimens226. A clinical trial (NCT04880577) has been initiated to test the ability of TAF as an add-on therapy to ocrelizumab to reduce symptoms and promote neuroprotection in RRMS.

There are distinct challenges for EBV vaccine development. Sterilizing immunity to EBV may not be possible given the efficiency of EBV transmission and persistence, and merely delaying the time of infection is undesirable, because it increases the risk of mononucleosis and MS. Furthermore, identification of the most appropriate viral antigens is complex for both prevention of infection and treatment of existing disease. Vaccine approaches to block early events in EBV primary infection would require neutralizing antibodies that target components of viral entry proteins (including gp350, gp42, gH, gL and gB). Therapeutic vaccines for various EBV-associated cancers or autoimmune disease may need to target multiple viral proteins, as both latent and lytic viral genes have been implicated in disease pathogenesis. In addition to careful consideration of the viral antigens included in the vaccine, a successful vaccine strategy for EBV must stimulate both the humoral arm and the cell-mediated arm of the adaptive immune system and induce production of effector and long-lived memory cells. The development of a vaccine targeted at preventing the development of mononucleosis in EBV-seronegative children could potentially reduce the likelihood that these individuals will later develop MS. A small trial examining the efficacy of vaccination with the HLA-B*0801-restricted CD8+ T cell epitope FLRGRAYGL demonstrated a reduced likelihood of developing mononucleosis in those children who seroconverted227. Similarly, vaccination of EBV-seronegative young adults with a recombinant gp350 subunit vaccine prevented the development of mononucleosis, although it did not decrease rates of asymptomatic EBV infection228. Phase I/II trials demonstrated that this gp350 subunit vaccine was well tolerated and immunogenic, inducing robust gp350 antibody responses as well as EBV-neutralizing antibody responses229. Following the success of its severe acute respiratory syndrome coronavirus 2 vaccine, Moderna launched a vaccine trial (NCT05164094) using mRNA encoding EBV gp350, gB, gH/gL and gp42 in seronegative 18–30-year-old adults. Further studies are required to determine whether this approach or other vaccine approaches could ultimately decrease the likelihood of developing mononucleosis and, subsequently, MS.

Conclusions

Despite years of controversy, the role of EBV infection and seropositivity as essential co-factors for most forms of MS may now be settled. As the severity of EBV primary infection strongly correlates with the development of MS many years later, it is likely that MS depends on the initial immune response to EBV infection. Failure to control this primary infection may lead to colonization of resident memory B cell and T cell follicles in CNS accessible sites, such as tertiary lymphoid structures, that are uniquely prone to inducing immune pathology in the CNS. The time of infection likely contributes to immune system elimination of viral, autoreactive T cells and antibodies that target CNS components. These events must be further exacerbated by numerous genetic risk alleles, especially HLA-DRB1*1501, that may compound the effects of EBV infection through aberrant presentation of autoreactive peptides. Other alleles can cooperate with EBV transcription regulatory factors, such as EBNA2, through altered binding specificity and gene programmes promoting inflammatory B cell proliferation. Whether there are any special features of EBV antigens, such as EBNA1, that induce high rates of polyreactivity and self-mimicry needs to be further investigated. Among the most pressing questions is whether EBV-infected cells or viral products act within the CNS or indirectly through inflammatory events in the periphery. Ultimately, how autoreactive immune cells and antibodies form and accumulate in the CNS remain high-priority questions. Knowing that EBV is a likely driver of inflammatory autoimmune disease provides a target for future therapies.

Glossary

Burkitt lymphomaAn aggressive form of non-Hodgkin lymphoma endemic to sub-Saharan Africa, where it is associated with Epstein–Barr virus infection.
Relapsing–remitting MS(RRMS). A form of multiple sclerosis (MS) where disease exacerbations are interspersed with periods of disease inactivity.
Secondary progressive MS(SPMS). A form of multiple sclerosis (MS) that follows relapsing–remitting MS where progressive disability accumulates with or without discernible relapse.
Primary progressive MS(PPMS). A form of multiple sclerosis (MS) that lacks distinct periods of disease exacerbations.
Clinically isolated syndromeAn initial episode of neurological symptoms associated with inflammation and demyelination with symptoms characteristic of multiple sclerosis that frequently, although not always, progresses to multiple sclerosis.
Waldeyer’s tonsillar ringA ring of lymphoid tissue surrounding the nasopharynx and oropharynx that includes the tonsils and adenoids.
Germinal centre(GC). An area within lymph nodes and other secondary lymphoid organs, including the spleen, where T cell-dependent B cell activation, differentiation and proliferation occur. Germinal centres are concentrated areas of B cell somatic mutation and selection.
Chronic active EBV infectionA rare condition marked by poor control of Epstein–Barr virus (EBV) infection, resulting in high EBV plasma viral loads and systemic infiltration by EBV-positive B cells or EBV-positive T cells.
Oral hairy leukoplakiaA white lesion on the tongue with a ‘hairy’ appearance that is caused by Epstein–Barr virus lytic infection and that can occur in immunocompromised individuals, especially those with HIV/AIDS.
Infectious mononucleosisA self-limiting disorder characterized by fever, extreme fatigue, sore throat and highly swollen lymph nodes; most frequently caused by immune response to primary Epstein–Barr virus infection, although milder forms are associated with cytomegalovirus infection.
Graft-versus-host diseaseA condition in which the donor’s immune system (the graft) rejects the recipient (the host) as non-self.

Author contributions

The authors contributed equally to all aspects of the article.

Peer review

Peer review information

Nature Reviews Microbiology thanks Christian Münz and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Competing interests

P.M.L. founded and is an adviser to Vironika LLC. P.M.L. is named on a patent for inhibitors of EBNA1. S.S.S. declares no competing interests.

Footnotes

Publisher’s note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

References

1. Young LS, Rickinson AB. Epstein-Barr virus: 40 years on. Nat. Rev. Cancer. 2004;4:757–768. doi: 10.1038/nrc1452. [PubMed] [CrossRef] [Google Scholar]
2. Young LS, Yap LF, Murray PG. Epstein-Barr virus: more than 50 years old and still providing surprises. Nat. Rev. Cancer. 2016;16:789–802. doi: 10.1038/nrc.2016.92. [PubMed] [CrossRef] [Google Scholar]
3. Wong Y, Meehan MT, Burrows SR, Doolan DL, Miles JJ. Estimating the global burden of Epstein-Barr virus-related cancers. J. Cancer Res. Clin. Oncol. 2021 doi: 10.1007/s00432-021-03824-y. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
4. Shannon-Lowe C, Rickinson A. The global landscape of EBV-associated tumors. Front. Oncol. 2019;9:713. doi: 10.3389/fonc.2019.00713. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
5. Dunmire SK, Verghese PS, Balfour HH., Jr Primary Epstein-Barr virus infection. J. Clin. Virol. 2018;102:84–92. doi: 10.1016/j.jcv.2018.03.001. [PubMed] [CrossRef] [Google Scholar]
6. Fournier B, Latour S. Immunity to EBV as revealed by immunedeficiencies. Curr. Opin. Immunol. 2021;72:107–115. doi: 10.1016/j.coi.2021.04.003. [PubMed] [CrossRef] [Google Scholar]
7. Ascherio A, Munger KL. Epidemiology of multiple sclerosis: from risk factors to prevention-an update. Semin. Neurol. 2016;36:103–114. doi: 10.1055/s-0036-1579693. [PubMed] [CrossRef] [Google Scholar]
8. Laderach F, Munz C. Epstein Barr virus exploits genetic susceptibility to increase multiple sclerosis risk. Microorganisms. 2021 doi: 10.3390/microorganisms9112191. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
9. Walton C, et al. Rising prevalence of multiple sclerosis worldwide: insights from the Atlas of MS, third edition. Mult. Scler. 2020;26:1816–1821. doi: 10.1177/1352458520970841. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
10. Alroughani R, Boyko A. Pediatric multiple sclerosis: a review. BMC Neurol. 2018;18:27. doi: 10.1186/s12883-018-1026-3. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
11. Rodgers MM, et al. Gait characteristics of individuals with multiple sclerosis before and after a 6-month aerobic training program. J. Rehabil. Res. Dev. 1999;36:183–188. [PubMed] [Google Scholar]
12. Confavreux C, Vukusic S. The clinical course of multiple sclerosis. Handb. Clin. Neurol. 2014;122:343–369. doi: 10.1016/B978-0-444-52001-2.00014-5. [PubMed] [CrossRef] [Google Scholar]
13. Thompson AJ, et al. Diagnosis of multiple sclerosis: 2017 revisions of the McDonald criteria. Lancet Neurol. 2018;17:162–173. doi: 10.1016/S1474-4422(17)30470-2. [PubMed] [CrossRef] [Google Scholar]
14. Brodin P, et al. Variation in the human immune system is largely driven by non-heritable influences. Cell. 2015;160:37–47. doi: 10.1016/j.cell.2014.12.020. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
15. Soldan, S. S. & Jacobson, S. in Neurotropic Viral Infections (ed. Reiss, C.) 175–220 (Springer, 2016).
16. Ruprecht K. The role of Epstein-Barr virus in the etiology of multiple sclerosis: a current review. Expert Rev. Clin. Immunol. 2020;16:1143–1157. doi: 10.1080/1744666X.2021.1847642. [PubMed] [CrossRef] [Google Scholar]
17. Lanz TV, et al. Clonally expanded B cells in multiple sclerosis bind EBV EBNA1 and GlialCAM. Nature. 2022;603:321–327. doi: 10.1038/s41586-022-04432-7. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
18. Bjornevik K, et al. Longitudinal analysis reveals high prevalence of Epstein-Barr virus associated with multiple sclerosis. Science. 2022;375:296–301. doi: 10.1126/science.abj8222. [PubMed] [CrossRef] [Google Scholar]
19. Bar-Or A, Banwell B, Berger JR, Lieberman PM. Guilty by association: Epstein-Barr virus in multiple sclerosis. Nat. Med. 2022;28:904–906. doi: 10.1038/s41591-022-01823-1. [PubMed] [CrossRef] [Google Scholar]
20. Cancer Genome Atlas Research Network Comprehensive molecular characterization of gastric adenocarcinoma. Nature. 2014;513:202–209. doi: 10.1038/nature13480. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
21. Baer R, et al. DNA sequence and expression of the B95-8 Epstein-Barr virus genome. Nature. 1984;310:207–211. doi: 10.1038/310207a0. [PubMed] [CrossRef] [Google Scholar]
22. Kanda T, Yajima M, Ikuta K. Epstein-Barr virus strain variation and cancer. Cancer Sci. 2019;110:1132–1139. doi: 10.1111/cas.13954. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
23. Thorley-Lawson DA. EBV persistence–introducing the virus. Curr. Top. Microbiol. Immunol. 2015;390:151–209. [PMC free article] [PubMed] [Google Scholar]
24. Farrell PJ. Epstein-Barr virus strain variation. Curr. Top. Microbiol. Immunol. 2015;390:45–69. [PubMed] [Google Scholar]
25. Santpere G, et al. Genome-wide analysis of wild-type Epstein-Barr virus genomes derived from healthy individuals of the 1000 Genomes Project. Genome Biol. Evol. 2014;6:846–860. doi: 10.1093/gbe/evu054. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
26. Lay ML, et al. Epstein-Barr virus genotypes and strains in central nervous system demyelinating disease and Epstein-Barr virus-related illnesses in Australia. Intervirology. 2012;55:372–379. doi: 10.1159/000334693. [PubMed] [CrossRef] [Google Scholar]
27. Brennan RM, et al. Strains of Epstein-Barr virus infecting multiple sclerosis patients. Mult. Scler. 2010;16:643–651. doi: 10.1177/1352458510364537. [PubMed] [CrossRef] [Google Scholar]
28. de-Thé, G. et al. Sero-epidemiology of the Epstein-Barr virus: preliminary analysis of an international study- a review 3–16 (IARC Science Publications, 1975). [PubMed]
29. Balfour HH, Jr, et al. Age-specific prevalence of Epstein-Barr virus infection among individuals aged 6-19 years in the United States and factors affecting its acquisition. J. Infect. Dis. 2013;208:1286–1293. doi: 10.1093/infdis/jit321. [PubMed] [CrossRef] [Google Scholar]
30. Chandran, B. & Hutt-Fletcher, L. in Human Herpesviruses: Biology, Therapy, and Immunoprophylaxis (eds Arvin, A. et al.) (Cambridge Univ. Press, 2007). [PubMed]
31. Thorley-Lawson DA. Epstein-Barr virus: exploiting the immune system. Nat. Rev. Immunol. 2001;1:75–82. doi: 10.1038/35095584. [PubMed] [CrossRef] [Google Scholar]
32. Thompson MP, Kurzrock R. Epstein-Barr virus and cancer. Clin. Cancer Res. 2004;10:803–821. doi: 10.1158/1078-0432.CCR-0670-3. [PubMed] [CrossRef] [Google Scholar]
33. Hassani A, Corboy JR, Al-Salam S, Khan G. Epstein-Barr virus is present in the brain of most cases of multiple sclerosis and may engage more than just B cells. PLoS ONE. 2018;13:e0192109. doi: 10.1371/journal.pone.0192109. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
34. Gianella S, et al. Effect of cytomegalovirus and Epstein-Barr virus replication on intestinal mucosal gene expression and microbiome composition of HIV-infected and uninfected individuals. AIDS. 2017;31:2059–2067. doi: 10.1097/QAD.0000000000001579. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
35. Speck P, Haan KM, Longnecker R. Epstein-Barr virus entry into cells. Virology. 2000;277:1–5. doi: 10.1006/viro.2000.0624. [PubMed] [CrossRef] [Google Scholar]
36. Xiao J, Palefsky JM, Herrera R, Tugizov SM. Characterization of the Epstein-Barr virus glycoprotein BMRF-2. Virology. 2007;359:382–396. doi: 10.1016/j.virol.2006.09.047. [PubMed] [CrossRef] [Google Scholar]
37. Xiao J, Palefsky JM, Herrera R, Berline J, Tugizov SM. EBV BMRF-2 facilitates cell-to-cell spread of virus within polarized oral epithelial cells. Virology. 2009;388:335–343. doi: 10.1016/j.virol.2009.03.030. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
38. Zhang H, et al. Ephrin receptor A2 is an epithelial cell receptor for Epstein-Barr virus entry. Nat. Microbiol. 2018;3:1–8. [PubMed] [Google Scholar]
39. Stubbins RJ, et al. Epstein-Barr virus associated smooth muscle tumors in solid organ transplant recipients: incidence over 31 years at a single institution and review of the literature. Transpl. Infect. Dis. 2019;21:e13010. doi: 10.1111/tid.13010. [PubMed] [CrossRef] [Google Scholar]
40. Kimura H, Cohen JI. Chronic active Epstein-Barr virus disease. Front. Immunol. 2017;8:1867. doi: 10.3389/fimmu.2017.01867. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
41. Jha HC, et al. Gammaherpesvirus infection of human neuronal cells. mBio. 2015;6:e01844–e01815. doi: 10.1128/mBio.01844-15. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
42. Menet A, et al. Epstein-Barr virus infection of human astrocyte cell lines. J. Virol. 1999;73:7722–7733. doi: 10.1128/JVI.73.9.7722-7733.1999. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
43. Kanda T. EBV-encoded latent genes. Adv. Exp. Med. Biol. 2018;1045:377–394. doi: 10.1007/978-981-10-7230-7_17. [PubMed] [CrossRef] [Google Scholar]
44. Kieff, E. & Rickinson, A. B. in Fields Virology (eds Knipe, D. M. & Howley, P. M.) 2603–2654 (Lippincott Williams and Wilkins, 2007).
45. Shinozaki-Ushiku A, Kunita A, Fukayama M. Update on Epstein-Barr virus and gastric cancer (review) Int. J. Oncol. 2015;46:1421–1434. doi: 10.3892/ijo.2015.2856. [PubMed] [CrossRef] [Google Scholar]
46. Greenspan JS, Greenspan D, Webster-Cyriaque J. Hairy leukoplakia; lessons learned: 30-plus years. Oral Dis. 2016;22:120–127. doi: 10.1111/odi.12393. [PubMed] [CrossRef] [Google Scholar]
47. Murata T, et al. Molecular basis of Epstein-Barr virus latency establishment and lytic reactivation. Viruses. 2021 doi: 10.3390/v13122344. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
48. McKenzie J, El-Guindy A. Epstein-Barr virus lytic cycle reactivation. Curr. Top. Microbiol. Immunol. 2015;391:237–261. [PubMed] [Google Scholar]
49. Chan CK, et al. Epstein-Barr virus antibody patterns preceding the diagnosis of nasopharyngeal carcinoma. Cancer Causes Control. 1991;2:125–131. doi: 10.1007/BF00053132. [PubMed] [CrossRef] [Google Scholar]
50. Mueller N, et al. Epstein-Barr virus antibody patterns preceding the diagnosis of non-Hodgkin’s lymphoma. Int. J. Cancer. 1991;49:387–393. doi: 10.1002/ijc.2910490313. [PubMed] [CrossRef] [Google Scholar]
51. Lu F, et al. Defective Epstein-Barr virus genomes and atypical viral gene expression in B-cell lines derived from multiple myeloma patients. J. Virol. 2021;95:e0008821. doi: 10.1128/JVI.00088-21. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
52. Rosemarie Q, Sugden B. Epstein-Barr virus: how its lytic phase contributes to oncogenesis. Microorganisms. 2020 doi: 10.3390/microorganisms8111824. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
53. Maple PAC, Gran B, Tanasescu R, Pritchard DI, Constantinescu CS. An absence of Epstein-Barr virus reactivation and associations with disease activity in people with multiple sclerosis undergoing therapeutic hookworm vaccination. Vaccines. 2020 doi: 10.3390/vaccines8030487. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
54. Torkildsen O, Nyland H, Myrmel H, Myhr KM. Epstein-Barr virus reactivation and multiple sclerosis. Eur. J. Neurol. 2008;15:106–108. [PubMed] [Google Scholar]
55. Yea C, et al. Epstein-Barr virus in oral shedding of children with multiple sclerosis. Neurology. 2013;81:1392–1399. doi: 10.1212/WNL.0b013e3182a841e4. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
56. Soldan SS, Lieberman PM. Epstein-Barr virus infection in the development of neurological disorders. Drug Discov. Today Dis. Model. 2020;32:35–52. doi: 10.1016/j.ddmod.2020.01.001. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
57. Munz C. Latency and lytic replication in Epstein-Barr virus-associated oncogenesis. Nat. Rev. Microbiol. 2019;17:691–700. doi: 10.1038/s41579-019-0249-7. [PubMed] [CrossRef] [Google Scholar]
58. Leen A, et al. Differential immunogenicity of Epstein-Barr virus latent-cycle proteins for human CD4+ T-helper 1 responses. J. Virol. 2001;75:8649–8659. doi: 10.1128/JVI.75.18.8649-8659.2001. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
59. Bickham K, et al. EBNA1-specific CD4+ T cells in healthy carriers of Epstein-Barr virus are primarily Th1 in function. J. Clin. Invest. 2001;107:121–130. doi: 10.1172/JCI10209. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
60. Munz C, et al. Human CD4+ T lymphocytes consistently respond to the latent Epstein-Barr virus nuclear antigen EBNA1. J. Exp. Med. 2000;191:1649–1660. doi: 10.1084/jem.191.10.1649. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
61. Azzi T, et al. Role for early-differentiated natural killer cells in infectious mononucleosis. Blood. 2014;124:2533–2543. doi: 10.1182/blood-2014-01-553024. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
62. Dunmire SK, Grimm JM, Schmeling DO, Balfour HH, Jr, Hogquist KA. The incubation period of primary Epstein-Barr virus infection: viral dynamics and immunologic events. PLoS Pathog. 2015;11:e1005286. doi: 10.1371/journal.ppat.1005286. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
63. Williams H, et al. The immune response to primary EBV infection: a role for natural killer cells. Br. J. Haematol. 2005;129:266–274. doi: 10.1111/j.1365-2141.2005.05452.x. [PubMed] [CrossRef] [Google Scholar]
64. Strowig T, et al. Priming of protective T cell responses against virus-induced tumors in mice with human immune system components. J. Exp. Med. 2009;206:1423–1434. doi: 10.1084/jem.20081720. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
65. Chijioke O, et al. Human natural killer cells prevent infectious mononucleosis features by targeting lytic Epstein-Barr virus infection. Cell Rep. 2013;5:1489–1498. doi: 10.1016/j.celrep.2013.11.041. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
66. Zumwalde NA, et al. Adoptively transferred Vγ9Vδ2 T cells show potent antitumor effects in a preclinical B cell lymphomagenesis model. JCI Insight. 2017 doi: 10.1172/jci.insight.93179. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
67. Lino CNR, Ghosh S. Epstein-Barr virus in inborn immunodeficiency-more than infection. Cancers. 2021 doi: 10.3390/cancers13194752. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
68. Cohen JI. Primary immunodeficiencies associated with EBV disease. Curr. Top. Microbiol. Immunol. 2015;390:241–265. [PMC free article] [PubMed] [Google Scholar]
69. Huo S, et al. EBV-EBNA1 constructs an immunosuppressive microenvironment for nasopharyngeal carcinoma by promoting the chemoattraction of Treg cells. J. Immunother. Cancer. 2020 doi: 10.1136/jitc-2020-001588. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
70. Westhoff Smith D, Chakravorty A, Hayes M, Hammerschmidt W, Sugden B. The Epstein-Barr virus oncogene EBNA1 suppresses natural killer cell responses and apoptosis early after infection of peripheral B cells. mBio. 2021;12:e0224321. doi: 10.1128/mBio.02243-21. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
71. Keane JT, et al. The interaction of Epstein-Barr virus encoded transcription factor EBNA2 with multiple sclerosis risk loci is dependent on the risk genotype. EBioMedicine. 2021;71:103572. doi: 10.1016/j.ebiom.2021.103572. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
72. Spender LC, et al. Cell target genes of Epstein-Barr virus transcription factor EBNA-2: induction of the p55alpha regulatory subunit of PI3-kinase and its role in survival of EREB2.5 cells. J. Gen. Virol. 2006;87:2859–2867. doi: 10.1099/vir.0.82128-0. [PubMed] [CrossRef] [Google Scholar]
73. Pages F, et al. Epstein-Barr virus nuclear antigen 2 induces interleukin-18 receptor expression in B cells. Blood. 2005;105:1632–1639. doi: 10.1182/blood-2004-08-3196. [PubMed] [CrossRef] [Google Scholar]
74. Anastasiadou E, et al. Epstein-Barr virus-encoded EBNA2 alters immune checkpoint PD-L1 expression by downregulating miR-34a in B-cell lymphomas. Leukemia. 2019;33:132–147. doi: 10.1038/s41375-018-0178-x. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
75. Yanagi Y, et al. RNAseq analysis identifies involvement of EBNA2 in PD-L1 induction during Epstein-Barr virus infection of primary B cells. Virology. 2021;557:44–54. doi: 10.1016/j.virol.2021.02.004. [PubMed] [CrossRef] [Google Scholar]
76. Kanda K, et al. The EBNA2-related resistance towards alpha interferon (IFN-alpha) in Burkitt’s lymphoma cells effects induction of IFN-induced genes but not the activation of transcription factor ISGF-3. Mol. Cell Biol. 1992;12:4930–4936. [PMC free article] [PubMed] [Google Scholar]
77. Su C, et al. EBNA2 driven enhancer switching at the CIITA-DEXI locus suppresses HLA class II gene expression during EBV infection of B-lymphocytes. PLoS Pathog. 2021;17:e1009834. doi: 10.1371/journal.ppat.1009834. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
78. Jochum S, Moosmann A, Lang S, Hammerschmidt W, Zeidler R. The EBV immunoevasins vIL-10 and BNLF2a protect newly infected B cells from immune recognition and elimination. PLoS Pathog. 2012;8:e1002704. doi: 10.1371/journal.ppat.1002704. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
79. Bouvet M, et al. Multiple viral microRNAs regulate interferon release and signaling early during infection with Epstein-Barr virus. mBio. 2021 doi: 10.1128/mBio.03440-20. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
80. Murer A, et al. MicroRNAs of Epstein-Barr virus attenuate T-cell-mediated immune control in vivo. mBio. 2019 doi: 10.1128/mBio.01941-18. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
81. Joshi N, Usuku K, Hauser SL. The T-cell response to myelin basic protein in familial multiple sclerosis: diversity of fine specificity, restricting elements, and T-cell receptor usage. Ann. Neurol. 1993;34:385–393. doi: 10.1002/ana.410340313. [PubMed] [CrossRef] [Google Scholar]
82. Martin C, et al. Absence of seven human herpesviruses, including HHV-6, by polymerase chain reaction in CSF and blood from patients with multiple sclerosis and optic neuritis. Acta Neurol. Scand. 1997;95:280–283. doi: 10.1111/j.1600-0404.1997.tb00210.x. [PubMed] [CrossRef] [Google Scholar]
83. Sindic CJ, Monteyne P, Laterre EC. The intrathecal synthesis of virus-specific oligoclonal IgG in multiple sclerosis. J. Neuroimmunol. 1994;54:75–80. doi: 10.1016/0165-5728(94)90233-X. [PubMed] [CrossRef] [Google Scholar]
84. Sriram S, et al. Chlamydia pneumoniae infection of the central nervous system in multiple sclerosis. Ann. Neurol. 1999;46:6–14. doi: 10.1002/1531-8249(199907)46:1<6::AID-ANA4>3.0.CO;2-M. [PubMed] [CrossRef] [Google Scholar]
85. Virtanen JO, Wohler J, Fenton K, Reich DS, Jacobson S. Oligoclonal bands in multiple sclerosis reactive against two herpesviruses and association with magnetic resonance imaging findings. Mult. Scler. 2014;20:27–34. doi: 10.1177/1352458513490545. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
86. Franciotta D, et al. Cerebrospinal BAFF and Epstein-Barr virus-specific oligoclonal bands in multiple sclerosis and other inflammatory demyelinating neurological diseases. J. Neuroimmunol. 2011;230:160–163. doi: 10.1016/j.jneuroim.2010.10.027. [PubMed] [CrossRef] [Google Scholar]
87. Wang Z, et al. Antibodies from multiple sclerosis brain identified Epstein-Barr virus nuclear antigen 1 & 2 epitopes which are recognized by oligoclonal bands. J. Neuroimmune Pharmacol. 2021;16:567–580. doi: 10.1007/s11481-020-09948-1. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
88. van Nierop GP, Mautner J, Mitterreiter JG, Hintzen RQ, Verjans GM. Intrathecal CD8 T-cells of multiple sclerosis patients recognize lytic Epstein-Barr virus proteins. Mult. Scler. 2016;22:279–291. doi: 10.1177/1352458515588581. [PubMed] [CrossRef] [Google Scholar]
89. Chabas D, et al. The influence of the proinflammatory cytokine, osteopontin, on autoimmune demyelinating disease. Science. 2001;294:1731–1735. doi: 10.1126/science.1062960. [PubMed] [CrossRef] [Google Scholar]
90. Cencioni MT, Mattoscio M, Magliozzi R, Bar-Or A, Muraro PA. B cells in multiple sclerosis - from targeted depletion to immune reconstitution therapies. Nat. Rev. Neurol. 2021;17:399–414. doi: 10.1038/s41582-021-00498-5. [PubMed] [CrossRef] [Google Scholar]
91. Lisak RP, et al. B cells from patients with multiple sclerosis induce cell death via apoptosis in neurons in vitro. J. Neuroimmunol. 2017;309:88–99. doi: 10.1016/j.jneuroim.2017.05.004. [PubMed] [CrossRef] [Google Scholar]
92. Li R, et al. Proinflammatory GM-CSF-producing B cells in multiple sclerosis and B cell depletion therapy. Sci. Transl. Med. 2015;7:310ra166. doi: 10.1126/scitranslmed.aab4176. [PubMed] [CrossRef] [Google Scholar]
93. Panitch HS, Hirsch RL, Schindler J, Johnson KP. Treatment of multiple sclerosis with gamma interferon: exacerbations associated with activation of the immune system. Neurology. 1987;37:1097–1102. doi: 10.1212/WNL.37.7.1097. [PubMed] [CrossRef] [Google Scholar]
94. Feng X, et al. Low expression of interferon-stimulated genes in active multiple sclerosis is linked to subnormal phosphorylation of STAT1. J. Neuroimmunol. 2002;129:205–215. doi: 10.1016/S0165-5728(02)00182-0. [PubMed] [CrossRef] [Google Scholar]
95. Lucchinetti C, et al. Heterogeneity of multiple sclerosis lesions: implications for the pathogenesis of demyelination. Ann. Neurol. 2000;47:707–717. doi: 10.1002/1531-8249(200006)47:6<707::AID-ANA3>3.0.CO;2-Q. [PubMed] [CrossRef] [Google Scholar]
96. Lassmann H. Multiple sclerosis pathology. Cold Spring Harb. Perspect. Med. 2018 doi: 10.1101/cshperspect.a028936. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
97. Barnett MH, Prineas JW. Relapsing and remitting multiple sclerosis: pathology of the newly forming lesion. Ann. Neurol. 2004;55:458–468. doi: 10.1002/ana.20016. [PubMed] [CrossRef] [Google Scholar]
98. Salou M, Nicol B, Garcia A, Laplaud DA. Involvement of CD8+ T cells in multiple sclerosis. Front. Immunol. 2015;6:604. doi: 10.3389/fimmu.2015.00604. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
99. Salou M, et al. Expanded CD8 T-cell sharing between periphery and CNS in multiple sclerosis. Ann. Clin. Transl. Neurol. 2015;2:609–622. doi: 10.1002/acn3.199. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
100. Cagol A, et al. Association of brain atrophy with disease progression independent of relapse activity in patients with relapsing multiple sclerosis. JAMA Neurol. 2022 doi: 10.1001/jamaneurol.2022.1025. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
101. Kim W, Patsopoulos NA. Genetics and functional genomics of multiple sclerosis. Semin. Immunopathol. 2022;4:63–79. doi: 10.1007/s00281-021-00907-3. [PubMed] [CrossRef] [Google Scholar]
102. Yuan S, Xiong Y, Larsson SC. An atlas on risk factors for multiple sclerosis: a Mendelian randomization study. J. Neurol. 2021;268:114–124. doi: 10.1007/s00415-020-10119-8. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
103. Jersild C, Dupont B, Fog T, Platz PJ, Svejgaard A. Histocompatibility determinants in multiple sclerosis. Transplant. Rev. 1975;22:148–163. [PubMed] [Google Scholar]
104. Cook SD. Multiple sclerosis and viruses. Mult. Scler. 1997;3:388–389. doi: 10.1177/135245859700300606. [PubMed] [CrossRef] [Google Scholar]
105. AustraliaNew Zealand Multiple Sclerosis Genetics Consortium. Genome-wide association study identifies new multiple sclerosis susceptibility loci on chromosomes 12 and 20. Nat. Genet. 2009;41:824–828. doi: 10.1038/ng.396. [PubMed] [CrossRef] [Google Scholar]
106. De Jager PL, et al. Meta-analysis of genome scans and replication identify CD6, IRF8 and TNFRSF1A as new multiple sclerosis susceptibility loci. Nat. Genet. 2009;41:776–782. doi: 10.1038/ng.401. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
107. International Multiple Sclerosis Genetics Consortium Risk alleles for multiple sclerosis identified by a genomewide study. N. Engl. J. Med. 2007;357:851–862. doi: 10.1056/NEJMoa073493. [PubMed] [CrossRef] [Google Scholar]
108. Cree BA. Multiple sclerosis genetics. Handb. Clin. Neurol. 2014;122:193–209. doi: 10.1016/B978-0-444-52001-2.00009-1. [PubMed] [CrossRef] [Google Scholar]
109. Lin, X., Deng, F. Y., Lu, X. & Lei, S. F. Susceptibility genes for multiple sclerosis identified in a gene-based genome-wide association study. J. Clin. Neurol. (2015). [PMC free article] [PubMed]
110. He B, Yang B, Lundahl J, Fredrikson S, Hillert J. The myelin basic protein gene in multiple sclerosis: identification of discrete alleles of a 1.3 kb tetranucleotide repeat sequence. Acta Neurol. Scand. 1998;97:46–51. doi: 10.1111/j.1600-0404.1998.tb00608.x. [PubMed] [CrossRef] [Google Scholar]
111. Kellar-Wood H, Robertson N, Govan GG, Compston DA, Harding AE. Leber’s hereditary optic neuropathy mitochondrial DNA mutations in multiple sclerosis. Ann. Neurol. 1994;36:109–112. doi: 10.1002/ana.410360121. [PubMed] [CrossRef] [Google Scholar]
112. Reynier P, et al. mtDNA haplogroup J: a contributing factor of optic neuritis. Eur. J. Hum. Genet. 1999;7:404–406. doi: 10.1038/sj.ejhg.5200293. [PubMed] [CrossRef] [Google Scholar]
113. Thompson RJ, et al. Analysis of polymorphisms of the 2’,3’-cyclic nucleotide-3’-phosphodiesterase gene in patients with multiple sclerosis. Mult. Scler. 1996;2:215–221. doi: 10.1177/135245859600200501. [PubMed] [CrossRef] [Google Scholar]
114. Sollid LM. Epstein-Barr virus as a driver of multiple sclerosis. Sci. Immunol. 2022;7:eabo7799. doi: 10.1126/sciimmunol.abo7799. [PubMed] [CrossRef] [Google Scholar]
115. Wucherpfennig KW, Strominger JL. Molecular mimicry in T cell-mediated autoimmunity: viral peptides activate human T cell clones specific for myelin basic protein. Cell. 1995;80:695–705. doi: 10.1016/0092-8674(95)90348-8. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
116. Lunemann JD, et al. EBNA1-specific T cells from patients with multiple sclerosis cross react with myelin antigens and co-produce IFN-gamma and IL-2. J. Exp. Med. 2008;205:1763–1773. doi: 10.1084/jem.20072397. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
117. Tengvall K, et al. Molecular mimicry between Anoctamin 2 and Epstein-Barr virus nuclear antigen 1 associates with multiple sclerosis risk. Proc. Natl Acad. Sci. USA. 2019;116:16955–16960. doi: 10.1073/pnas.1902623116. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
118. van Sechel AC, et al. EBV-induced expression and HLA-DR-restricted presentation by human B cells of alpha B-crystallin, a candidate autoantigen in multiple sclerosis. J. Immunol. 1999;162:129–135. [PubMed] [Google Scholar]
119. Jelcic I, et al. Memory B cells activate brain-homing, autoreactive CD4(+) T cells in multiple sclerosis. Cell. 2018;175:85–100 e123. doi: 10.1016/j.cell.2018.08.011. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
120. Nociti V, et al. Epstein-Barr virus antibodies in serum and cerebrospinal fluid from multiple sclerosis, chronic inflammatory demyelinating polyradiculoneuropathy and amyotrophic lateral sclerosis. J. Neuroimmunol. 2010;225:149–152. doi: 10.1016/j.jneuroim.2010.04.007. [PubMed] [CrossRef] [Google Scholar]
121. Ascherio A, Munger KL, Lunemann JD. The initiation and prevention of multiple sclerosis. Nat. Rev. Neurol. 2012;8:602–612. doi: 10.1038/nrneurol.2012.198. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
122. Lunemann JD, Ascherio A. Immune responses to EBNA1: biomarkers in MS. Neurology. 2009;73:13–14. doi: 10.1212/WNL.0b013e3181aa2a5f. [PubMed] [CrossRef] [Google Scholar]
123. Mescheriakova JY, van Nierop GP, van der Eijk AA, Kreft KL, Hintzen RQ. EBNA-1 titer gradient in families with multiple sclerosis indicates a genetic contribution. Neurol. Neuroimmunol. Neuroinflamm. 2020 doi: 10.1212/NXI.0000000000000872. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
124. Hedström AK, et al. High levels of Epstein-Barr virus nuclear antigen-1-specific antibodies and infectious mononucleosis act both independently and synergistically to increase multiple sclerosis risk. Front. Neurol. 2019;10:1368. doi: 10.3389/fneur.2019.01368. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
125. van Noort JM, Bajramovic JJ, Plomp AC, van Stipdonk MJ. Mistaken self, a novel model that links microbial infections with myelin-directed autoimmunity in multiple sclerosis. J. Neuroimmunol. 2000;105:46–57. doi: 10.1016/S0165-5728(00)00181-8. [PubMed] [CrossRef] [Google Scholar]
126. Hecker M, et al. High-density peptide microarray analysis of IgG autoantibody reactivities in serum and cerebrospinal fluid of multiple sclerosis patients. Mol. Cell Proteom. 2016;15:1360–1380. doi: 10.1074/mcp.M115.051664. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
127. Capone G, et al. Peptide matching between Epstein-Barr virus and human proteins. Pathog. Dis. 2013;69:205–212. doi: 10.1111/2049-632X.12066. [PubMed] [CrossRef] [Google Scholar]
128. Meier UC, Cipian RC, Karimi A, Ramasamy R, Middeldorp JM. Cumulative roles for Epstein-Barr virus, human endogenous retroviruses, and human herpes virus-6 in driving an inflammatory cascade underlying MS pathogenesis. Front. Immunol. 2021;12:757302. doi: 10.3389/fimmu.2021.757302. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
129. Dantuma NP, Sharipo A, Masucci MG. Avoiding proteasomal processing: the case of EBNA1. Curr. Top. Microbiol. Immunol. 2002;269:23–36. [PubMed] [Google Scholar]
130. Levitskaya J, Sharipo A, Leonchiks A, Ciechanover A, Masucci MG. Inhibition of ubiquitin/proteasome-dependent protein degradation by the Gly-Ala repeat domain of the Epstein-Barr virus nuclear antigen 1. Proc. Natl Acad. Sci. USA. 1997;94:12616–12621. doi: 10.1073/pnas.94.23.12616. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
131. Levitskaya J, et al. Inhibition of antigen processing by the internal repeat region of the Epstein-Barr virus nuclear antigen-1. Nature. 1995;375:685–688. doi: 10.1038/375685a0. [PubMed] [CrossRef] [Google Scholar]
132. Tovar Fernandez MC, et al. Substrate-specific presentation of MHC class I-restricted antigens via autophagy pathway. Cell Immunol. 2022;374:104484. doi: 10.1016/j.cellimm.2022.104484. [PubMed] [CrossRef] [Google Scholar]
133. Apcher S, Daskalogianni C, Manoury B, Fahraeus R. Epstein Barr virus-encoded EBNA1 interference with MHC class I antigen presentation reveals a close correlation between mRNA translation initiation and antigen presentation. PLoS Pathog. 2010;6:e1001151. doi: 10.1371/journal.ppat.1001151. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
134. Tellam JT, et al. mRNA Structural constraints on EBNA1 synthesis impact on in vivo antigen presentation and early priming of CD8+ T cells. PLoS Pathog. 2014;10:e1004423. doi: 10.1371/journal.ppat.1004423. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
135. Murat P, et al. G-quadruplexes regulate Epstein-Barr virus-encoded nuclear antigen 1 mRNA translation. Nat. Chem. Biol. 2014;10:358–364. doi: 10.1038/nchembio.1479. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
136. Tellam JT, Lekieffre L, Zhong J, Lynn DJ, Khanna R. Messenger RNA sequence rather than protein sequence determines the level of self-synthesis and antigen presentation of the EBV-encoded antigen, EBNA1. PLoS Pathog. 2012;8:e1003112. doi: 10.1371/journal.ppat.1003112. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
137. Pender MP. The essential role of Epstein-Barr virus in the pathogenesis of multiple sclerosis. Neuroscientist. 2011;17:351–367. doi: 10.1177/1073858410381531. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
138. Melchers F, Rolink AR. B cell tolerance–how to make it and how to break it. Curr. Top. Microbiol. Immunol. 2006;305:1–23. [PubMed] [Google Scholar]
139. Weniger MA, Kuppers R. Molecular biology of Hodgkin lymphoma. Leukemia. 2021;35:968–981. doi: 10.1038/s41375-021-01204-6. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
140. Sommermann T, et al. Functional interplay of Epstein-Barr virus oncoproteins in a mouse model of B cell lymphomagenesis. Proc. Natl Acad. Sci. USA. 2020;117:14421–14432. doi: 10.1073/pnas.1921139117. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
141. Laurence M, Benito-Leon J. Epstein-Barr virus and multiple sclerosis: updating Pender’s hypothesis. Mult. Scler. Relat. Disord. 2017;16:8–14. doi: 10.1016/j.msard.2017.05.009. [PubMed] [CrossRef] [Google Scholar]
142. Choi IK, et al. Mechanism of EBV inducing anti-tumour immunity and its therapeutic use. Nature. 2021;590:157–162. doi: 10.1038/s41586-020-03075-w. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
143. Deng Y, et al. CD27 is required for protective lytic EBV antigen-specific CD8+ T-cell expansion. Blood. 2021;137:3225–3236. doi: 10.1182/blood.2020009482. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
144. Veroni C, Serafini B, Rosicarelli B, Fagnani C, Aloisi F. Transcriptional profile and Epstein-Barr virus infection status of laser-cut immune infiltrates from the brain of patients with progressive multiple sclerosis. J. Neuroinflamm. 2018;15:18. doi: 10.1186/s12974-017-1049-5. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
145. Magliozzi R, et al. B-cell enrichment and Epstein-Barr virus infection in inflammatory cortical lesions in secondary progressive multiple sclerosis. J. Neuropathol. Exp. Neurol. 2013;72:29–41. doi: 10.1097/NEN.0b013e31827bfc62. [PubMed] [CrossRef] [Google Scholar]
146. Serafini B, et al. Epstein-Barr virus latent infection and BAFF expression in B cells in the multiple sclerosis brain: implications for viral persistence and intrathecal B-cell activation. J. Neuropathol. Exp. Neurol. 2010;69:677–693. doi: 10.1097/NEN.0b013e3181e332ec. [PubMed] [CrossRef] [Google Scholar]
147. Tzartos JS, et al. Association of innate immune activation with latent Epstein-Barr virus in active MS lesions. Neurology. 2012;78:15–23. doi: 10.1212/WNL.0b013e31823ed057. [PubMed] [CrossRef] [Google Scholar]
148. Moreno MA, et al. Molecular signature of Epstein-Barr virus infection in MS brain lesions. Neurol. Neuroimmunol. Neuroinflamm. 2018;5:e466. doi: 10.1212/NXI.0000000000000466. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
149. Serafini B, Rosicarelli B, Veroni C, Mazzola GA, Aloisi F. Epstein-Barr virus-specific CD8 T cells selectively infiltrate the brain in multiple sclerosis and interact locally with virus-infected cells: clue for a virus-driven immunopathological mechanism. J. Virol. 2019 doi: 10.1128/JVI.00980-19. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
150. Recher M, et al. Extralymphatic virus sanctuaries as a consequence of potent T-cell activation. Nat. Med. 2007;13:1316–1323. doi: 10.1038/nm1670. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
151. Hochberg D, et al. Acute infection with Epstein-Barr virus targets and overwhelms the peripheral memory B-cell compartment with resting, latently infected cells. J. Virol. 2004;78:5194–5204. doi: 10.1128/JVI.78.10.5194-5204.2004. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
152. Veroni C, et al. Immune and Epstein-Barr virus gene expression in cerebrospinal fluid and peripheral blood mononuclear cells from patients with relapsing-remitting multiple sclerosis. J. Neuroinflamm. 2015;12:132. doi: 10.1186/s12974-015-0353-1. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
153. Kiriyama T, Kataoka H, Kasai T, Nonomura A, Ueno S. Negative association of Epstein-Barr virus or herpes simplex virus-1 with tumefactive central nervous system inflammatory demyelinating disease. J. Neurovirol. 2010;16:466–471. doi: 10.1007/BF03210852. [PubMed] [CrossRef] [Google Scholar]
154. Sargsyan SA, et al. Absence of Epstein-Barr virus in the brain and CSF of patients with multiple sclerosis. Neurology. 2010;74:1127–1135. doi: 10.1212/WNL.0b013e3181d865a1. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
155. Willis SN, et al. Epstein-Barr virus infection is not a characteristic feature of multiple sclerosis brain. Brain. 2009;132:3318–3328. doi: 10.1093/brain/awp200. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
156. Peferoen LA, et al. Epstein Barr virus is not a characteristic feature in the central nervous system in established multiple sclerosis. Brain. 2010;133:e137. doi: 10.1093/brain/awp296. [PubMed] [CrossRef] [Google Scholar]
157. Torkildsen O, et al. Upregulation of immunoglobulin-related genes in cortical sections from multiple sclerosis patients. Brain Pathol. 2010;20:720–729. doi: 10.1111/j.1750-3639.2009.00343.x. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
158. Lassmann H, Niedobitek G, Aloisi F, Middeldorp JM, NeuroproMiSe EBV Working Group Epstein-Barr virus in the multiple sclerosis brain: a controversial issue–report on a focused workshop held in the Centre for Brain Research of the Medical University of Vienna, Austria. Brain. 2011;134:2772–2786. doi: 10.1093/brain/awr197. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
159. Hislop AD, Taylor GS. T-cell responses to EBV. Curr. Top. Microbiol. Immunol. 2015;391:325–353. [PubMed] [Google Scholar]
160. Hislop AD, Taylor GS, Sauce D, Rickinson AB. Cellular responses to viral infection in humans: lessons from Epstein-Barr virus. Annu. Rev. Immunol. 2007;25:587–617. doi: 10.1146/annurev.immunol.25.022106.141553. [PubMed] [CrossRef] [Google Scholar]
161. Munger KL, Levin LI, O’Reilly EJ, Falk KI, Ascherio A. Anti-Epstein-Barr virus antibodies as serological markers of multiple sclerosis: a prospective study among United States military personnel. Mult. Scler. 2011;17:1185–1193. doi: 10.1177/1352458511408991. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
162. Levin LI, et al. Temporal relationship between elevation of Epstein-Barr virus antibody titers and initial onset of neurological symptoms in multiple sclerosis. JAMA. 2005;293:2496–2500. doi: 10.1001/jama.293.20.2496. [PubMed] [CrossRef] [Google Scholar]
163. Lunemann JD, et al. Increased frequency and broadened specificity of latent EBV nuclear antigen-1-specific T cells in multiple sclerosis. Brain. 2006;129:1493–1506. doi: 10.1093/brain/awl067. [PubMed] [CrossRef] [Google Scholar]
164. International Multiple Sclerosis Genetics Consortium et al. Genetic risk and a primary role for cell-mediated immune mechanisms in multiple sclerosis. Nature. 2011;476:214–219. doi: 10.1038/nature10251. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
165. Jilek S, et al. Strong EBV-specific CD8+ T-cell response in patients with early multiple sclerosis. Brain. 2008;131:1712–1721. doi: 10.1093/brain/awn108. [PubMed] [CrossRef] [Google Scholar]
166. Angelini DF, et al. Increased CD8+ T cell response to Epstein-Barr virus lytic antigens in the active phase of multiple sclerosis. PLoS Pathog. 2013;9:e1003220. doi: 10.1371/journal.ppat.1003220. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
167. Pender MP, Csurhes PA, Pfluger CM, Burrows SR. Deficiency of CD8+ effector memory T cells is an early and persistent feature of multiple sclerosis. Mult. Scler. 2014;20:1825–1832. doi: 10.1177/1352458514536252. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
168. Pender MP, Csurhes PA, Pfluger CM, Burrows SR. Decreased CD8+ T cell response to Epstein-Barr virus infected B cells in multiple sclerosis is not due to decreased HLA class I expression on B cells or monocytes. BMC Neurol. 2011;11:95. doi: 10.1186/1471-2377-11-95. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
169. Pender MP, Csurhes PA, Pfluger CM, Burrows SR. CD8 T cell deficiency impairs control of Epstein–Barr virus and worsens with age in multiple sclerosis. J. Neurol. Neurosurg. Psychiatry. 2012;83:353–354. doi: 10.1136/jnnp-2011-300213. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
170. Veroni C, Aloisi F. The CD8 T cell-Epstein-Barr virus-B cell trialogue: a central issue in multiple sclerosis pathogenesis. Front. Immunol. 2021;12:665718. doi: 10.3389/fimmu.2021.665718. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
171. van Langelaar J, et al. The association of Epstein-Barr virus infection with CXCR3+ B-cell development in multiple sclerosis: impact of immunotherapies. Eur. J. Immunol. 2021;51:626–633. doi: 10.1002/eji.202048739. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
172. Baglio SR, et al. Sensing of latent EBV infection through exosomal transfer of 5’pppRNA. Proc. Natl Acad. Sci. USA. 2016;113:E587–E596. doi: 10.1073/pnas.1518130113. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
173. Afrasiabi A, et al. The interaction of human and Epstein-Barr virus miRNAs with multiple sclerosis risk loci. Int. J. Mol. Sci. 2021 doi: 10.3390/ijms22062927. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
174. Chen CC, et al. Elucidation of exosome migration across the blood–brain barrier model in vitro. Cell. Mol. Bioeng. 2016;9:509–529. doi: 10.1007/s12195-016-0458-3. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
175. Jiang S, et al. The Epstein-Barr virus regulome in lymphoblastoid cells. Cell Host Microbe. 2017;22:561–573.e4. doi: 10.1016/j.chom.2017.09.001. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
176. Harley JB, et al. Transcription factors operate across disease loci, with EBNA2 implicated in autoimmunity. Nat. Genet. 2018;50:699–707. doi: 10.1038/s41588-018-0102-3. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
177. Hong T, et al. Epstein-Barr virus nuclear antigen 2 extensively rewires the human chromatin landscape at autoimmune risk loci. Genome Res. 2021 doi: 10.1101/gr.264705.120. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
178. Afrasiabi A, et al. Evidence from genome wide association studies implicates reduced control of Epstein-Barr virus infection in multiple sclerosis susceptibility. Genome Med. 2019;11:26. doi: 10.1186/s13073-019-0640-z. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
179. Ricigliano VA, et al. EBNA2 binds to genomic intervals associated with multiple sclerosis and overlaps with vitamin D receptor occupancy. PLoS ONE. 2015;10:e0119605. doi: 10.1371/journal.pone.0119605. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
180. Mechelli R, et al. Epstein-Barr virus genetic variants are associated with multiple sclerosis. Neurology. 2015;84:1362–1368. doi: 10.1212/WNL.0000000000001420. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
181. Zhou Y, et al. Utilising multi-large omics data to elucidate biological mechanisms within multiple sclerosis genetic susceptibility loci. Mult. Scler. 2021;27:2141–2149. doi: 10.1177/13524585211004422. [PubMed] [CrossRef] [Google Scholar]
182. Ruhrmann S, Stridh P, Kular L, Jagodic M. Genomic imprinting: a missing piece of the multiple sclerosis puzzle? Int. J. Biochem. Cell Biol. 2015;67:49–57. doi: 10.1016/j.biocel.2015.05.010. [PubMed] [CrossRef] [Google Scholar]
183. Kular L, Jagodic M. Epigenetic insights into multiple sclerosis disease progression. J. Intern. Med. 2020;288:82–102. doi: 10.1111/joim.13045. [PubMed] [CrossRef] [Google Scholar]
184. Kular L, et al. DNA methylation as a mediator of HLA-DRB1*15:01 and a protective variant in multiple sclerosis. Nat. Commun. 2018;9:2397. doi: 10.1038/s41467-018-04732-5. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
185. He Y, Huang L, Tang Y, Yang Z, Han Z. Genome-wide identification and analysis of splicing QTLs in multiple sclerosis by RNA-seq data. Front. Genet. 2021;12:769804. doi: 10.3389/fgene.2021.769804. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
186. Wanke F, et al. EBI2 is highly expressed in multiple sclerosis lesions and promotes early CNS migration of encephalitogenic CD4 T cells. Cell Rep. 2017;18:1270–1284. doi: 10.1016/j.celrep.2017.01.020. [PubMed] [CrossRef] [Google Scholar]
187. Guo R, Gewurz BE. Epigenetic control of the Epstein-Barr lifecycle. Curr. Opin. Virol. 2022;52:78–88. doi: 10.1016/j.coviro.2021.11.013. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
188. Tempera I, Lieberman PM. Epigenetic regulation of EBV persistence and oncogenesis. Semin. Cancer Biol. 2014;26:22–29. doi: 10.1016/j.semcancer.2014.01.003. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
189. Kucukali CI, Kurtuncu M, Coban A, Cebi M, Tuzun E. Epigenetics of multiple sclerosis: an updated review. Neuromol. Med. 2015;17:83–96. doi: 10.1007/s12017-014-8298-6. [PubMed] [CrossRef] [Google Scholar]
190. Soldan SS, et al. Epigenetic plasticity enables CNS-trafficking of EBV-infected B lymphocytes. PLoS Pathog. 2021;17:e1009618. doi: 10.1371/journal.ppat.1009618. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
191. Greer JM, et al. Immunogenic and encephalitogenic epitope clusters of myelin proteolipid protein. J. Immunol. 1996;156:371–379. [PubMed] [Google Scholar]
192. Zdimerova H, et al. Attenuated immune control of Epstein-Barr virus in humanized mice is associated with the multiple sclerosis risk factor HLA-DR15. Eur. J. Immunol. 2021;51:64–75. doi: 10.1002/eji.202048655. [PubMed] [CrossRef] [Google Scholar]
193. Agostini S, et al. HLA alleles modulate EBV viral load in multiple sclerosis. J. Transl. Med. 2018;16:80. doi: 10.1186/s12967-018-1450-6. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
194. Wandinger K, et al. Association between clinical disease activity and Epstein-Barr virus reactivation in MS. Neurology. 2000;55:178–184. doi: 10.1212/WNL.55.2.178. [PubMed] [CrossRef] [Google Scholar]
195. Cocuzza CE, et al. Quantitative detection of Epstein-Barr virus DNA in cerebrospinal fluid and blood samples of patients with relapsing-remitting multiple sclerosis. PLoS ONE. 2014;9:e94497. doi: 10.1371/journal.pone.0094497. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
196. Lindsey JW, Hatfield LM, Crawford MP, Patel S. Quantitative PCR for Epstein-Barr virus DNA and RNA in multiple sclerosis. Mult. Scler. 2009;15:153–158. doi: 10.1177/1352458508097920. [PubMed] [CrossRef] [Google Scholar]
197. Buljevac D, et al. Epstein-Barr virus and disease activity in multiple sclerosis. J. Neurol. Neurosurg. Psychiatry. 2005;76:1377–1381. doi: 10.1136/jnnp.2004.048504. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
198. Hollenbach JA, Oksenberg JR. The immunogenetics of multiple sclerosis: a comprehensive review. J. Autoimmun. 2015;64:13–25. doi: 10.1016/j.jaut.2015.06.010. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
199. Enz LS, et al. Increased HLA-DR expression and cortical demyelination in MS links with HLA-DR15. Neurol. Neuroimmunol. Neuroinflamm. 2020 doi: 10.1212/NXI.0000000000000656. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
200. Martin R, Sospedra M, Eiermann T, Olsson T. Multiple sclerosis: doubling down on MHC. Trends Genet. 2021;37:784–797. doi: 10.1016/j.tig.2021.04.012. [PubMed] [CrossRef] [Google Scholar]
201. Menegatti J, Schub D, Schafer M, Grasser FA, Ruprecht K. HLA-DRB1*15:01 is a co-receptor for Epstein-Barr virus, linking genetic and environmental risk factors for multiple sclerosis. Eur. J. Immunol. 2021;51:2348–2350. doi: 10.1002/eji.202149179. [PubMed] [CrossRef] [Google Scholar]
202. Burnham JA, Wright RR, Dreisbach J, Murray RS. The effect of high-dose steroids on MRI gadolinium enhancement in acute demyelinating lesions. Neurology. 1991;41:1349–1354. doi: 10.1212/WNL.41.9.1349. [PubMed] [CrossRef] [Google Scholar]
203. Baker D, Marta M, Pryce G, Giovannoni G, Schmierer K. Memory B cells are major targets for effective immunotherapy in relapsing multiple sclerosis. EBioMedicine. 2017;16:41–50. doi: 10.1016/j.ebiom.2017.01.042. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
204. Ceronie B, et al. Cladribine treatment of multiple sclerosis is associated with depletion of memory B cells. J. Neurol. 2018;265:1199–1209. doi: 10.1007/s00415-018-8830-y. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
205. Hauser SL, et al. B-cell depletion with rituximab in relapsing-remitting multiple sclerosis. N. Engl. J. Med. 2008;358:676–688. doi: 10.1056/NEJMoa0706383. [PubMed] [CrossRef] [Google Scholar]
206. Kappos L, et al. Ocrelizumab in relapsing-remitting multiple sclerosis: a phase 2, randomised, placebo-controlled, multicentre trial. Lancet. 2011;378:1779–1787. doi: 10.1016/S0140-6736(11)61649-8. [PubMed] [CrossRef] [Google Scholar]
207. Segal BM, et al. Repeated subcutaneous injections of IL12/23 p40 neutralising antibody, ustekinumab, in patients with relapsing-remitting multiple sclerosis: a phase II, double-blind, placebo-controlled, randomised, dose-ranging study. Lancet Neurol. 2008;7:796–804. doi: 10.1016/S1474-4422(08)70173-X. [PubMed] [CrossRef] [Google Scholar]
208. Kappos L, et al. Atacicept in multiple sclerosis (ATAMS): a randomised, placebo-controlled, double-blind, phase 2 trial. Lancet Neurol. 2014;13:353–363. doi: 10.1016/S1474-4422(14)70028-6. [PubMed] [CrossRef] [Google Scholar]
209. Bilger A, et al. Leflunomide/teriflunomide inhibit Epstein-Barr virus (EBV)- induced lymphoproliferative disease and lytic viral replication. Oncotarget. 2017;8:44266–44280. doi: 10.18632/oncotarget.17863. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
210. Doubrovina E, et al. Adoptive immunotherapy with unselected or EBV-specific T cells for biopsy-proven EBV+ lymphomas after allogeneic hematopoietic cell transplantation. Blood. 2012;119:2644–2656. doi: 10.1182/blood-2011-08-371971. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
211. Heslop HE, et al. Long-term outcome of EBV-specific T-cell infusions to prevent or treat EBV-related lymphoproliferative disease in transplant recipients. Blood. 2010;115:925–935. doi: 10.1182/blood-2009-08-239186. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
212. Savoldo B, et al. Treatment of solid organ transplant recipients with autologous Epstein Barr virus-specific cytotoxic T lymphocytes (CTLs) Blood. 2006;108:2942–2949. doi: 10.1182/blood-2006-05-021782. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
213. Pender MP, et al. Epstein-Barr virus-specific adoptive immunotherapy for progressive multiple sclerosis. Mult. Scler. 2014;20:1541–1544. doi: 10.1177/1352458514521888. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
214. Pender MP, et al. Epstein-Barr virus-specific T cell therapy for progressive multiple sclerosis. JCI Insight. 2018 doi: 10.1172/jci.insight.124714. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
215. Pender MP, et al. Epstein-Barr virus-specific T cell therapy for progressive multiple sclerosis. JCI Insight. 2020 doi: 10.1172/jci.insight.144624. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
216. Pender MP, Csurhes PA, Burrows JM, Burrows SR. Defective T-cell control of Epstein-Barr virus infection in multiple sclerosis. Clin. Transl. Immunol. 2017;6:e126. doi: 10.1038/cti.2016.87. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
217. Bar-Or A, et al. Updated open-label extension clinical data and new magnetization transfer ratio imaging data from a phase I study of ATA188, an off-the-shelf, allogeneic Epstein-Barr virus-targeted T-cell immunotherapy for progressive multiple sclerosis [ECTRIMS 2021 poster]. Multiple Sclerosis J. 2021;27((2_suppl.)):P638. [Google Scholar]
218. Lycke J, et al. Acyclovir treatment of relapsing-remitting multiple sclerosis. A randomized, placebo-controlled, double-blind study. J. Neurol. 1996;243:214–224. doi: 10.1007/BF00868517. [PubMed] [CrossRef] [Google Scholar]
219. Bech E, et al. A randomized, double-blind, placebo-controlled MRI study of anti-herpes virus therapy in MS. Neurology. 2002;58:31–36. doi: 10.1212/WNL.58.1.31. [PubMed] [CrossRef] [Google Scholar]
220. Friedman JE, et al. A randomized clinical trial of valacyclovir in multiple sclerosis. Mult. Scler. 2005;11:286–295. doi: 10.1191/1352458505ms1185oa. [PubMed] [CrossRef] [Google Scholar]
221. Annibali V, et al. IFN-beta and multiple sclerosis: from etiology to therapy and back. Cytokine Growth Factor. Rev. 2015;26:221–228. doi: 10.1016/j.cytogfr.2014.10.010. [PubMed] [CrossRef] [Google Scholar]
222. Bentz GL, Liu R, Hahn AM, Shackelford J, Pagano JS. Epstein-Barr virus BRLF1 inhibits transcription of IRF3 and IRF7 and suppresses induction of interferon-beta. Virology. 2010;402:121–128. doi: 10.1016/j.virol.2010.03.014. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
223. Hahn AM, Huye LE, Ning S, Webster-Cyriaque J, Pagano JS. Interferon regulatory factor 7 is negatively regulated by the Epstein-Barr virus immediate-early gene, BZLF-1. J. Virol. 2005;79:10040–10052. doi: 10.1128/JVI.79.15.10040-10052.2005. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
224. De Clercq E. Potential of acyclic nucleoside phosphonates in the treatment of DNA virus and retrovirus infections. Expert Rev. Anti Infect. Ther. 2003;1:21–43. doi: 10.1586/14787210.1.1.21. [PubMed] [CrossRef] [Google Scholar]
225. Drosu NC, Edelman ER, Housman DE. Tenofovir prodrugs potently inhibit Epstein-Barr virus lytic DNA replication by targeting the viral DNA polymerase. Proc. Natl Acad. Sci. USA. 2020;117:12368–12374. doi: 10.1073/pnas.2002392117. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
226. Torkildsen O, Myhr KM, Skogen V, Steffensen LH, Bjornevik K. Tenofovir as a treatment option for multiple sclerosis. Mult. Scler. Relat. Disord. 2020;46:102569. doi: 10.1016/j.msard.2020.102569. [PubMed] [CrossRef] [Google Scholar]
227. Elliott SL, et al. Phase I trial of a CD8+ T-cell peptide epitope-based vaccine for infectious mononucleosis. J. Virol. 2008;82:1448–1457. doi: 10.1128/JVI.01409-07. [PMC free article] [PubMed] [CrossRef] [Google Scholar]
228. Sokal EM, et al. Recombinant gp350 vaccine for infectious mononucleosis: a phase 2, randomized, double-blind, placebo-controlled trial to evaluate the safety, immunogenicity, and efficacy of an Epstein-Barr virus vaccine in healthy young adults. J. Infect. Dis. 2007;196:1749–1753. doi: 10.1086/523813. [PubMed] [CrossRef] [Google Scholar]
229. Moutschen M, et al. Phase I/II studies to evaluate safety and immunogenicity of a recombinant gp350 Epstein-Barr virus vaccine healthy adults. Vaccine. 2007;25:4697–4705. doi: 10.1016/j.vaccine.2007.04.008. [PubMed] [CrossRef] [Google Scholar]
230. Bach JF. The effect of infections on susceptibility to autoimmune and allergic diseases. N. Engl. J. Med. 2002;347:911–920. doi: 10.1056/NEJMra020100. [PubMed] [CrossRef] [Google Scholar]
231. Sheik-Ali S. Infectious mononucleosis and multiple sclerosis - updated review on associated risk. Mult. Scler. Relat. Disord. 2017;14:56–59. doi: 10.1016/j.msard.2017.02.019. [PubMed] [CrossRef] [Google Scholar]
232. Dirmeier U, et al. Latent membrane protein 1 of Epstein-Barr virus coordinately regulates proliferation with control of apoptosis. Oncogene. 2005;24:1711–1717. doi: 10.1038/sj.onc.1208367. [PubMed] [CrossRef] [Google Scholar]
233. Hussain M, Gatherer D, Wilson JB. Modelling the structure of full-length Epstein-Barr virus nuclear antigen 1. Virus Genes. 2014;49:358–372. doi: 10.1007/s11262-014-1101-9. [PubMed] [CrossRef] [Google Scholar]

Articles from Nature Reviews. Microbiology are provided here courtesy of Nature Publishing Group

-