Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2014 Sep 8.
Published in final edited form as: Arch Neurol. 2010 Aug;67(8):923–930. doi: 10.1001/archneurol.2010.161

Natalizumab and Progressive Multifocal Leukoencephalopathy: What are the causal factors? Can it be avoided?

Clemens Warnke 1, Til Menge 2, Hans-Peter Hartung 3, Michael K Racke 4, Petra D Cravens 5, Jeffrey L Bennett 6, Elliot M Frohman 7, Benjamin M Greenberg 8, Scott S Zamvil 9, Ralf Gold 10, Bernhard Hemmer 11, Bernd C Kieseier 12, Olaf Stüve 13
PMCID: PMC4157908  NIHMSID: NIHMS622605  PMID: 20697042

Abstract

Natalizumab (Tysabri®) was the first monoclonal antibody approved for the treatment of relapsing forms of multiple sclerosis (MS). After its initial approval, three patients undergoing natalizumab therapy in combination with other immunoregulatory and immunosuppressive agents were diagnosed with progressive multifocal leukoencephalopathy (PML). The agent was later re-approved, and its use restricted to monotherapy in patients with relapsing forms of MS. Over the past year, five additional cases of PML were reported in MS patients receiving natalizumab monotherapy. Thus, there is currently no convincing evidence that natalizumab-associated PML is restricted to combination therapy with other disease modifying or immunosuppressive agents.

The initial section of this review focuses on the scientific rationale for natalizumab in MS treatment. In the second part, our understanding of PML will be outlined. Thirdly, recent results on altered immune surveillance under natalizumab treatment are reviewed. In the forth section, the link of viral reactivation and very late activation antigen 4 (VLA-4) antagonism will be discussed. Finally, this review will address the potential impact of our current knowledge on the use of natalizumab in clinical practice.

The natalizumab experience

Multiple Sclerosis is an inflammatory demyelinating disorder of the central nervous system (CNS) and one of the most common causes of sustained neurological disability in young adults.1 The presence of leukocytes in cerebral perivascular spaces (CPVS) in areas of disease activity is one of the pathological hallmarks.24 An absolute requirement for the influx of leukocytes from the peripheral blood into the CNS is their expression of adhesion molecules, called integrins, and their interaction with counter-receptors from the immunoglobulin supergene family proteins on endothelial cells. Alpha(α)4-beta(β)1-Integrin (very late activation antigen 4 (VLA-4)) is one of the four main integrins required for the firm arrest of leukocytes following their rolling adhesion.5

Natalizumab (Tysabri®) is a recombinant humanized monoclonal IgG4-antibody that binds, amongst others, to the α-4-subunit of the α4β1 integrin, and interferes with the α-4 mediated binding to its natural ligands of the extra cellular matrix and endothelial lining, vascular cell adhesion molecule-1 (VCAM1) and fibronectin (FN).67 Although inhibition of leukocyte migration and extravasation is believed to be the leading mode of action of natalizumab, additional mechanisms might modulate the therapeutic and adverse effects of this antibody. Lindberg et al. recently showed that natalizumab has a direct effect on gene expression relevant for function and differentiation of T-lymphocytes, B-lymphocytes, neutrophils and erythrocytes.8

In vivo, antibodies against VLA-4 interfere with the binding of leukocytes to cerebral blood vessels, and effectively prevent signs of experimental autoimmune encephalomyelitis (EAE), an animal model of MS.9 Although natalizumab is highly immunogenic in mice, it reduces the influx of T cells and monocytes into the CNS and substantially ameliorates the clinical course of EAE.10

The efficacy of natalizumab in EAE led to clinical trials for the treatment of Multiple Sclerosis (MS). Following very promising results in phase II studies,1113 phase III clinical trials were performed that compared natalizumab alone versus placebo (AFFIRM trial14), and natalizumab plus interferon beta-1a (IFNβ-1a) versus placebo plus IFNβ-1a (SENTINEL trial15). Both studies showed significant advantage for the natalizumab-treated groups with respect to the primary clinical endpoints. In the AFFIRM monotherapy trial, natalizumab reduced the rate of clinical relapses at one year by 68 percent and the risk of sustained progression of disability by 42 percent over two years. Post hoc analysis of the AFFIRM trial showed disease remission defined as no activity on clinical (no relapses and no sustained disability progression) and radiological measures in 37% of the natalizumab group compared to 7% of the placebo group.16

In November 24, 2004, the Food and Drug Administration (FDA) approved natalizumab for the treatment of relapsing forms of MS. On February 28, 2005, the manufacturers of natalizumab announced the voluntary withdrawal from the market after two MS patients in the SENTINEL trial (combination therapy with IFNβ-1a) and one patient with Crohn’s disease were diagnosed with progressive multifocal leukoencephalopathy (PML).1719

In the summer 2006, natalizumab was re-approved in the United States (US), and approved in the European Union (EU) as monotherapy for the treatment of relapsing forms of MS. In the US, recommendations were made to limit the use of natalizumab to highly active (more than 2 severe relapses per year) relapsing remitting (RR)-MS and for patients not responding to or tolerating first line treatment (IFNβ-1a, IFNβ-1b, glatiramer acetate). This restricted approval was the result of a risk-benefit analysis. The initial diagnosis of PML in two patients from the combination therapy trial1718 led to the restricted approval and to risk minimization plans (TYSABRI Outreach: Unified Commitment to Health (TOUCH), TYSABRI Global Observation Program In Safety (TYGRIS), Crohn's Disease - Investigating Natalizumab through Further Observational Research and Monitoring (CD INFORM)). Systematic review of all patients treated in those studies (more than 3700 patients) showed no additional cases, therefore, calculated risk of PML was 1:1000 (95% CI) after an average treatment time of about 18 months.20 As there is no increased risk of JC-polyomavirus-replication and spread throughout the CNS in patients with MS per se,21 there is little doubt that treatment with natalizumab is linked to these cases of PML.

According to information provided by Biogen Idec, about 53,000 patients had been treated with natalizumab by December 2008 (including patients enrolled in clinical trials). Of those patients, 20,000 have received at least one year of natalizumab therapy, approximately 10,700 patients have been on therapy for 18 months and 4,300 patients have received natalizumab for at least 24 months.22

In the past 12 months, five new cases of PML have been reported, bringing up the total number of PML under natalizumab therapy to eight thus far.2223 Because of the restricted use of natalizumab, all of those new patients had been treated with monotherapy. PML was observed after 8, 12, 14, 14, 17, 26, 28 and 37 infusions. These new cases remind us of the necessity to further investigate mechanisms that predispose certain patients to an increased risk of developing PML under natalizumab therapy.

PML – Primary infection and places of latency

PML is an opportunistic demyelinating disease of the brain. It is caused by reactivation of JC-Virus, a dsDNA-virus belonging, together with SV-40- and BK-virus, to the family of polyomaviridae.24 PML has almost exclusively been reported in immunocompromised patients, in particular in patients with reduced cellular immunity, including patients with HIV, patients with haematological diseases, or in patients receiving immunosuppressive medication.2526 Therefore, detection of JCV-specific cytotoxic CD8+ T lymphocytes in PML patients is associated with a favourable outcome and early disease control27 and assays measuring CD4 cell immune functions are currently being investigated as monitoring tools for immunocompromised populations to estimate their risk of polyomavirus infections.28

Primary infection by JC-virus takes place in childhood and is asymptomatic. JC-virus-antibodies are seen in 50 – 85% of all adults.2931 JCV persists in the tonsillar tissue, bone marrow, the kidney and the spleen.3234 The presence of JCV-DNA has been detected in urine, different blood compartments, and in CSF.3537 JCV viruria was found as frequently in HIV-positive individuals as in control subjects, suggesting that its detection has no clinical value.38 In a recent study, viruria was seen in 25.5 % of healthy volunteers.39

JCV-DNA has been detected in peripheral blood mononuclear cells (PBMC) in HIV-positive patients with and without PML but not in HIV-negative control subjects. There appears to be no specificity for any leukocyte subtype. The presence of JCV-DNA detection correlates with low CD4+ lymphocyte counts.38 Furthermore, JCV-DNA was recently found in peripheral blood of immunosuppressed patients with Crohn’s Disease (CD),39 but not in controls. In contrast, only one of the first three confirmed cases of PML (the patient with CD) on natalizumab had consistently elevated JCV-DNA plasma levels before onset of PML symptoms. Consequently, JCV-load might correlate with immunosuppression but may have limited predictive value as a screening tool for PML.40

Currently, our knowledge about transmission of JCV infection and its replication cycle in the healthy human population is limited. The transmittable form of JCV is commonly referred to as the JCV-archetype, as it is thought that all other genotypes originate from it. The JCV-archetype is detectable in the urine.41 In contrast, the JCV-PML-type, referred to as MAD-1 (named after the University if Wisconsin at Madison, where it was characterized) appears more neurotropic, and can be isolated from brains of PML patients. This pathological variant is characterized by deletions, duplications and point mutations in a specific JCV regulatory region.

There is growing evidence on persistency of JCV in the CNS.4244 A more recent study investigated viral protein and DNA load in brains of immunocompetent non-PML patients. No viral proteins were expressed in any of these cases. Nevertheless, fragments of the viral DNA were present in various regions of normal brain. JCV DNA was found in oligodendrocytes and astrocytes, but not in neurons.42 These findings suggest that JCV has access to the brain in immunocompetent individuals. In the setting of immunosuppression, it is conceivable that either the passing of virus during JC-viremia, or resident virus persisting in normal brain may express its genome and initiate its lytic cycle in oligodendrocytes.42

The prognosis of PML is poor. Notwithstanding that mortality has substantially dropped since the introduction of highly active antiretroviral therapy (HAART) for treatment of the acquired immune deficiency syndrome (AIDS), PML still is fatal in about 50% of HIV-associated PML cases within the first three months after diagnosis.25 No pathognomonic initial symptoms of PML have been defined, which often makes an early clinical diagnosis of this disorder very challenging. Some of the classic clinical signs and symptoms of PML include rapidly progressive dementia, motor dysfunction and vision loss, which can be difficult to differentiate from MS relapses.33, 4546

Critical for the diagnosis of PML is the demonstration of virus by JCV-PCR in the CSF and surrogate disease markers by magnetic resonance imaging (MRI). Before the introduction of HAART, JCV-PCR had a sensitivity of 72–93% and a specificity of 92–100%.25 Since HAART therapy, sensitivity is reported to be lower (about 58%) most likely due to an improved immune response under antiretroviral treatment.4749 A recent study showed positive JCV-PCR results in some MS patients without any clinical or radiological evidence for PML. Thus, low numbers of JCV-copies within the CSF might also be part of the physiological replication process.50 As there are only 8 cases in natalizumab treated patients so far, sensitivity and specificity of CSF-JCV-PCR for this subpopulation is unknown, although one case of false negative PCR in one of the recently reported PML cases proved limited sensitivity and gives reason for concern.23

MRI is a sensitive detection method, but many surrogate disease markers are not specific for PML. Typically, there are hyperintense, multifocal asymmetric signal abnormalities throughout the supratentorial subcortical white matter on T2-weighted and fluid attenuated inversion recovery (FLAIR) sequences.20,51 While there is a relative absence of gadolinium uptake in PML, in cases of immune reconstitution inflammatory syndrome (IRIS), contrast enhancement is frequently observed. This syndrome is defined as a combination of clinical worsening in a PML patient despite recovery of the immune system, explained by an overwhelming inflammatory immune response.52 This syndrome was first seen in AIDS patients after initiation of HAART, but has also been reported in natalizumab treated patients.23

Restoring the immune system is the only proven intervention for PML.26,53 The downside of restoring an immune response against JCV is the risk of IRIS. There is some evidence that IRIS can be attenuated through the administration of corticosteroids.54 While immune reconstitution is realized by HAART therapy in AIDS patients, early cessation of immunosuppressive medication was associated with favourable clinical outcomes in transplant recipients.5556 Plasma exchange has been proposed as a therapeutic tool that may allow a faster restoration of immune effector function in natalizumab-treated patients.53,57 Unfortunately, it appears that the pharmacological half-life of natalizumab far exceeds its biological half-life.58 Therefore, even accelerated clearance of treatment may not have a beneficial effect in all affected individuals. Currently a randomized multicenter trial is underway for the treatment of PML with mefloquine. The primary outcome measure is the JCV-PCR product in the CSF over up to 24 weeks.59 This trial is based on positive in vitro data on efficacy of mefloquine on JCV-replication.60

Impaired immune surveillance in Natalizumab-treated MS patients

Reactivation of a latent infection in natalizumab treated patients is likely linked to its therapeutic principal of action. Natalizumab was specifically designed to reduce trafficking of lymphocytes into peripheral tissues; therefore, it was postulated that treatment with natalizumab results in reduced immune surveillance of the CNS. Indeed, cerebrospinal fluid (CSF) from natalizumab-treated patients with MS contained significantly fewer white blood cells (WBC), CD4+ T cells, CD8+ T cells, CD19+ B cells and CD 138+ plasma cells compared to MS patients not treated with natalizumab.58 Furthermore, the CD4/CD8-ratio in the CSF of natalizumab-treated patients was significantly reduced to levels comparable to those of HIV-patients.61 In contrast to HIV-patients, there was no reversed CD4/CD8-ratio in peripheral blood. It was also shown, that CD4+ cells express significantly less unbound α-4-integrin pre and post natalizumab therapy compared to CD8+ cells.61 Therefore, an absolute threshold of unbound VLA-4 may be required for migration across the endothelial barrier, perhaps partially explaining reduced migration of CD4+ T cells into the CNS. Surprisingly, reduced cell counts were still detectable six months after cessation of natalizumab treatment.58 This was unexpected, as natalizumab has a biological half-life of eleven days, and its biological activity therefore can not be expected to continue beyond six weeks after cessation.62

The preferential and prolonged effect of natalizumab on CD4+ T cell number in the CNS could possibly be explained by a decrease in the number of antigen presenting cells (APC), and the expression of major histocompatibility complex (MHC) class II antigens in cerebral perivascular spaces (CPVS). Autopsy material from a patient who died of PML while on natalizumab treatment17 was investigated for MHC-expression and the number of APCs within the cerebrovascular spaces,63 and compared to anatomically-matched healthy brain tissue from MS patients not treated with natalizumab, and tissue from patients with PML not associated with natalizumab treatment. MHC II expression and the number of APCs in CPVS were significantly reduced compared to control brains. In addition, not a single CD4+ T cell was detectable in the CPVS of natalizumab treated PML patients, whereas CD8+ T cell numbers were not reduced in the CPVS as compared to controls. The latter might not be surprising, as MHC I was shown to be significantly upregulated in PML patients treated with natalizumab compared to all other controls. While it is difficult to detect JCV-responsive CD4+ T cells in patients with PML,64 JCV-specific CD8+ T cells are detectable in the peripheral blood of PML patients infected with the HIV, and their presence has been associated with a more favorable outcome.27,65,66 CD8+ T cell responses are directed toward an HLA-A*0201-restricted JCV epitope, VP1p36.67 However, Initiation and perpetuation of antigen-specific CD8+ T cell responses are likely to require the help of CD4+ T cells in the form of cytokines and other inflammatory mediators.68

VLA-4 Antagonism, viral reactivation and malignancy

There are two possible mechanisms of JCV reactivation discussed in the literature. Either the persisting virus within the CNS or passing virus during JC-viremia is responsible for JCV reactivation in the setting of immunosuppression or impaired immunosurveillance. Interestingly, CD34+ cells are susceptible to JCV infection,69 and JCV-DNA is detectable in peripheral lymphocytes from patients with and without PML.35,38,70 In addition, CD34+ haematopoietic precursor cells express high levels of VLA-4 on their surface7173 and CD34+ bone marrow cells have, in comparison to circulating cells in the peripheral blood, a higher VLA-4 expression rate and VLA-4-avidity.74 Bonig and co-workers and our group recently showed that natalizumab mobilizes CD34+ haematopoietic progenitor cells.7576 Binding of natalizumab to α4-integrin may block the VLA-4 mediated interaction of CD34+ cells of the bone marrow with its ligands in the extra cellular matrix, e.g. VCAM1, and may lead to mobilization of CD34+ cells to the peripheral blood. Antibody mediated blockage of CD34+ cell-homing into the bone marrow could play an additional role.7778 Mobilization of haematopoietic progenitor cells by monoclonal antibodies against VLA-4 in primates and mice was also demonstrated.7981

As outlined above, natalizumab upregulates transcription factors important for differentiation of B lymphocytes.8 Thus, during natalizumab induced B cell differentiation, JCV-infected bone marrow cells might be activated leading to JC-viremia and PML as a consequence of natalizumab therapy.8284 In this context, rearrangement of archetype-JCV to PML-type to the MAD-1 genotype could occur.41,85 This hypothesis could also help to explain PML cases under treatment with monoclonal antibodies against CD20 and CD52. After initial depletion, reconstitution of the B cell line might cause JC-viremia and PML.83 However, so far there is no conclusive evidence for an increased incidence of JC-viremia in natalizumab treatment. Furthermore, given the fact that natalizumab significantly reduces the extravasation of cells that express VLA-4, the presence of CD34+ cells in peripheral tissues under natalizumab therapy needs to be verified.

Thus far, there are only few reports on the reactivation of CNS latent virus other than JC virus in natalizumab-treated patients. Human herpes virus (HHV)-6 is a pleiotropic β-herpes virus commonly reactivated in the setting of acute and prolonged immunosuppression.86,87 HHV-6 has also been suggested to be involved in pathogenesis of MS,88,89 and it has also been associated with PML pathogenesis.90 Elevated serum HHV6 IgG levels and HHV6A DNA in the CSF of a subset of patients treated with natalizumab were recently reported. Also, in vitro superinfection of JC-virus infected glial cells with HHV-6 increased JCV-expression.91 Interestingly, HHV-6 has also been detected in CD34+ haematopoietic progenitor cells.92,93

JCV is probably not the only latent virus reactivated in natalizumab-treated patients. Mobilization of virus infected bone marrow cells might be a natalizumab-associated, but not a virus specific side effect. Altered immune surveillance combined with potential latent viral reactivation could also enhance the risk of malignancy. Thus far, there have only been three reported cases of melanoma in natalizumab-treated patients, which may present association by chance.15, 94

Alternative Treatment Paradigms

Existing data indicate that natalizumab is immunosuppressive, and that these properties may be a contributing factor in the susceptibility to CNS infections. Also, published reports suggest that there is a dose-duration effect on the risk of developing an infectious complication in some patients. In addition, it has not been shown that prolonged continuous therapy with natalizumab is required to ensure its efficacy. Therefore, alternative treatment paradigms appear feasible: (1) Based on published observations we know that natalizumab has an immediate effect on the number and composition of leukocytes in the CSF.58 (2) It is also known that the effect of natalizumab on leukocyte numbers in the CSF after cessation of treatment persists for at least 6 months,58 and (3) that cell numbers normalize 14 months following the discontinuation of therapy.95 In addition, it was demonstrated that the patients are clinically stable on first-line disease modifying therapies (DMTs) during the 14 months period after cessation of natalizumab.95 While an increase in T2 lesion load on MRI 15 months after cessation of treatment has been shown in one study,96 there was no change of surrogate disease markers on magnetic resonance imaging (MRI) in another.95 Thus, it may be necessary to limit the use of natalizumab for a certain period of time, followed by a treatment holiday during which patients are treated with one of the conventional disease modifying therapies. In patients with very aggressive disease natalizumab may only be used as an induction therapy. Such treatment algorithms remain speculative at present and controlled clinical trials must be performed to shed further light on these clinically relevant questions.

Conclusions

Currently, natalizumab is only recommended as monotherapy in patients with MS not responding to first line treatment or treatment naïve patients with high clinical disease activity, or in those not capable of tolerating conventional therapy. This restricted approval originated from the observation of patients who developed progressive multifocal leukoencephalopathy (PML) under natalizumab therapy in combination therapy with IFNβ-1a in the context of clinical studies.15 Recently, five more cases of PML in patients with MS who had received natalizumab in monotherapy were reported.2223 Thus, there is currently no convincing evidence that monotherapy is safer in this regard than combination therapy with disease modifying agents. PML is not a natalizumab-specific side effect, it has been diagnosed in the context of many other immunoactive drugs as well. Clearly, further studies are warranted to understand the immunological effects of natalizumab besides blocking cell migration across the BBB.

As there is currently no proven treatment for patients suffering from PML under natalizumab treatment other than accelerated clearance of therapy,57 the early establishment of a diagnosis is crucial. Upon the re-approval of natalizumab, each country initiated a risk management program to closely monitor patients at risk. Kappos et al. developed a three step diagnostic algorithm for natalizumab-treated patients with new or worsening neurological signs and symptoms. Early suspension of natalizumab treatment and strategies for clinical, MRI, and laboratory assessments were proposed.40

As outlined above, impaired immune surveillance and viral reactivation caused by treatment with natalizumab are yet not fully understood. However, there is growing suspicion that long-term treatment with natalizumab may put patients at risk of PML and presumably other infections, as well as neoplastic growth.

These key findings suggest long-term effects on CNS intrinsic immune system and viral reactivation. These observations may provide a scientific rationale for alternative treatment algorithms for natalizumab, which need to be evaluated in controlled clinical trials. The major challenge in the near future will be to identify biomarkers associated with an elevated risk of viral reactivation.

Contributor Information

Dr. Clemens Warnke, Department of Neurology, Heinrich-Heine Universität, Düsseldorf, Germany.

Dr. Til Menge, Department of Neurology, Heinrich-Heine Universität, Düsseldorf, Germany.

Dr. Hans-Peter Hartung, Department of Neurology, Heinrich-Heine Universität, Düsseldorf, Germany.

Dr. Michael K. Racke, Department of Neurology, The Ohio State University, Clumbus, OH.

Dr. Petra D. Cravens, Department of Neurology, University of Texas Southwestern Medical Center at Dallas, Dallas, TX.

Dr. Jeffrey L. Bennett, Department of Neurology, University of Colorado, Denver, CO.

Dr. Elliot M. Frohman, Department of Neurology, University of Texas Southwestern Medical Center at Dallas, Dallas, TX.

Dr. Benjamin M. Greenberg, Department of Neurology, University of Texas Southwestern Medical Center at Dallas, Dallas, TX.

Dr. Scott S. Zamvil, Department of Neurology, University of California, San Francisco, CA.

Dr. Ralf Gold, Department of Neurology, Ruhr University Bochum, Germany.

Dr. Bernhard Hemmer, Department of Neurology, Technische Universität München, Klinik Rechts der Isar, Munich, Germany.

Dr. Bernd C. Kieseier, Department of Neurology, Heinrich-Heine Universität, Düsseldorf, Germany.

Dr. Olaf Stüve, Department of Neurology, University of Texas Southwestern Medical Center at Dallas, Dallas, TX, Division of Neurology, Dallas VA Medical Center, Dallas, TX.

References

  • 1.Compston A, Coles A. Multiple sclerosis. Lancet. 2002;359(9313):1221–31. doi: 10.1016/S0140-6736(02)08220-X. [DOI] [PubMed] [Google Scholar]
  • 2.Martin R, McFarland HF, McFarlin DE. Immunological aspects of demyelinating diseases. Annu Rev Immunol. 1992;10:153–87. doi: 10.1146/annurev.iy.10.040192.001101. [DOI] [PubMed] [Google Scholar]
  • 3.Lucchinetti C, Brück W, Parisi J, et al. Heterogeneity of multiple sclerosis lesions: implications for the pathogenesis of demyelination. Ann Neurol. 2000;47(6):707–17. doi: 10.1002/1531-8249(200006)47:6<707::aid-ana3>3.0.co;2-q. [DOI] [PubMed] [Google Scholar]
  • 4.Lassmann H, Brück W, Lucchinetti CF. The immunopathology of multiple sclerosis: an overview. Brain Pathol. 2007;17(2):210–8. doi: 10.1111/j.1750-3639.2007.00064.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Luster AD, Alon R, von Andrian UH. Immune cell migration in inflammation: present and future therapeutic targets. Nat Immunol. 2005;6(12):1182–90. doi: 10.1038/ni1275. [DOI] [PubMed] [Google Scholar]
  • 6.Ulbrich H, Eriksson EE, Lindbom L. Leukocyte and endothelial cell adhesion molecules as targets for therapeutic interventions in inflammatory disease. Trends Pharmacol Sci. 2003;24(12):640–7. doi: 10.1016/j.tips.2003.10.004. [DOI] [PubMed] [Google Scholar]
  • 7.von Andrian UH, Engelhardt B. Alpha4 integrins as therapeutic targets in autoimmune disease. N Engl J Med. 2003;348(1):68–72. doi: 10.1056/NEJMe020157. [DOI] [PubMed] [Google Scholar]
  • 8.Lindberg RLP, Achtnichts L, Hoffmann F, Kuhle J, Kappos L. Natalizumab alters transcriptional expression profiles of blood cell subpopulations of multiple sclerosis patients. J Neuroimmunol. 2008;194(1–2):153–64. doi: 10.1016/j.jneuroim.2007.11.007. [DOI] [PubMed] [Google Scholar]
  • 9.Yednock TA, Cannon C, Fritz LC, et al. Prevention of experimental autoimmune encephalomyelitis by antibodies against alpha 4 beta 1 integrin. Nature. 1992;356(6364):63–6. doi: 10.1038/356063a0. [DOI] [PubMed] [Google Scholar]
  • 10.Kanwar JR, Harrison JE, Wang D, et al. Beta7 integrins contribute to demyelinating disease of the central nervous system. J Neuroimmunol. 2000;103(2):146–52. doi: 10.1016/s0165-5728(99)00245-3. [DOI] [PubMed] [Google Scholar]
  • 11.Miller DH, Khan OA, Sheremata WA, et al. A controlled trial of natalizumab for relapsing multiple sclerosis. N Engl J Med. 2003;348(1):15–23. doi: 10.1056/NEJMoa020696. [DOI] [PubMed] [Google Scholar]
  • 12.O'Connor PW, Goodman A, Willmer-Hulme AJ, et al. Randomized multicenter trial of natalizumab in acute MS relapses: clinical and MRI effects. Neurology. 2004;62(11):2038–43. doi: 10.1212/01.wnl.0000128136.79044.d6. [DOI] [PubMed] [Google Scholar]
  • 13.Tubridy N, Behan PO, Capildeo R, et al. The effect of anti-alpha4 integrin antibody on brain lesion activity in MS. The UK Antegren Study Group. Neurology. 1999;53(3):466–72. doi: 10.1212/wnl.53.3.466. [DOI] [PubMed] [Google Scholar]
  • 14.Polman CH, O'Connor PW, Havrdova E, et al. A randomized, placebo-controlled trial of natalizumab for relapsing multiple sclerosis. N Engl J Med. 2006;354(9):899–910. doi: 10.1056/NEJMoa044397. [DOI] [PubMed] [Google Scholar]
  • 15.Rudick RA, Stuart WH, Calabresi PA, et al. Natalizumab plus interferon beta-1a for relapsing multiple sclerosis. N Engl J Med. 2006;354(9):911–23. doi: 10.1056/NEJMoa044396. [DOI] [PubMed] [Google Scholar]
  • 16.Havrdova E, Galetta S, Hutchinson M, et al. Effect of natalizumab on clinical and radiological disease activity in multiple sclerosis: a retrospective analysis of the Natalizumab Safety and Efficacy in Relapsing-Remitting Multiple Sclerosis (AFFIRM) study. Lancet Neurol. 2009;8(3):254–60. doi: 10.1016/S1474-4422(09)70021-3. [DOI] [PubMed] [Google Scholar]
  • 17.Kleinschmidt-DeMasters BK, Tyler KL. Progressive multifocal leukoencephalopathy complicating treatment with natalizumab and interferon beta-1a for multiple sclerosis. N Engl J Med. 2005;353(4):369–74. doi: 10.1056/NEJMoa051782. [DOI] [PubMed] [Google Scholar]
  • 18.Langer-Gould A, Atlas SW, Green AJ, Bollen AW, Pelletier D. Progressive multifocal leukoencephalopathy in a patient treated with natalizumab. N Engl J Med. 2005;353(4):375–81. doi: 10.1056/NEJMoa051847. [DOI] [PubMed] [Google Scholar]
  • 19.Van Assche G, Van Ranst M, Sciot R, et al. Progressive multifocal leukoencephalopathy after natalizumab therapy for Crohn's disease. N Engl J Med. 2005;353(4):362–8. doi: 10.1056/NEJMoa051586. [DOI] [PubMed] [Google Scholar]
  • 20.Yousry TA, Major EO, Ryschkewitsch C, et al. Evaluation of patients treated with natalizumab for progressive multifocal leukoencephalopathy. N Engl J Med. 2006;354(9):924–33. doi: 10.1056/NEJMoa054693. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Khalili K, White MK. Human demyelinating disease and the polyomavirus JCV. Mult Scler. 2006;12(2):133–42. doi: 10.1191/135248506ms1264oa. [DOI] [PubMed] [Google Scholar]
  • 22. [Accessed March 17, 2009];D96A171A-3815-4766-940D-BE6E2DAD8F0A_biibTysabri13Mar09.pdf (application/pdf Object) Available at: http://library.corporate-ir.net/library/14/148/148682/items/328677/D96A171A-3815-4766-940D-BE6E2DAD8F0A_biibTysabri13Mar09.pdf.
  • 23.Hartung H. New cases of progressive multifocal leukoencephalopathy after treatment with natalizumab. Lancet Neurol. 2009;8(1):28–31. doi: 10.1016/S1474-4422(08)70281-3. [DOI] [PubMed] [Google Scholar]
  • 24.Frisque RJ, Bream GL, Cannella MT. Human polyomavirus JC virus genome. J Virol. 1984;51(2):458–69. doi: 10.1128/jvi.51.2.458-469.1984. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Koralnik IJ. New insights into progressive multifocal leukoencephalopathy. Curr Opin Neurol. 2004;17(3):365–70. doi: 10.1097/00019052-200406000-00019. [DOI] [PubMed] [Google Scholar]
  • 26.Koralnik IJ. Progressive multifocal leukoencephalopathy revisited: Has the disease outgrown its name? Ann Neurol. 2006;60(2):162–73. doi: 10.1002/ana.20933. [DOI] [PubMed] [Google Scholar]
  • 27.Du Pasquier RA, Kuroda MJ, Zheng Y, et al. A prospective study demonstrates an association between JC virus-specific cytotoxic T lymphocytes and the early control of progressive multifocal leukoencephalopathy. Brain. 2004;127(Pt 9):1970–8. doi: 10.1093/brain/awh215. [DOI] [PubMed] [Google Scholar]
  • 28.Batal I, Zeevi A, Heider A, et al. Measurements of global cell-mediated immunity in renal transplant recipients with BK virus reactivation. Am J Clin Pathol. 2008;129(4):587–91. doi: 10.1309/23YGPB1E758ECCFP. [DOI] [PubMed] [Google Scholar]
  • 29.Knowles WA, Pipkin P, Andrews N, et al. Population-based study of antibody to the human polyomaviruses BKV and JCV and the simian polyomavirus SV40. J Med Virol. 2003;71(1):115–23. doi: 10.1002/jmv.10450. [DOI] [PubMed] [Google Scholar]
  • 30.Padgett BL, Walker DL. Prevalence of antibodies in human sera against JC virus, an isolate from a case of progressive multifocal leukoencephalopathy. J Infect Dis. 1973;127(4):467–70. doi: 10.1093/infdis/127.4.467. [DOI] [PubMed] [Google Scholar]
  • 31.Stolt A, Sasnauskas K, Koskela P, Lehtinen M, Dillner J. Seroepidemiology of the human polyomaviruses. J Gen Virol. 2003;84(Pt 6):1499–504. doi: 10.1099/vir.0.18842-0. [DOI] [PubMed] [Google Scholar]
  • 32.Ferrante P, Caldarelli-Stefano R, Omodeo-Zorini E, et al. Comprehensive investigation of the presence of JC virus in AIDS patients with and without progressive multifocal leukoencephalopathy. J Med Virol. 1997;52(3):235–42. doi: 10.1002/(sici)1096-9071(199707)52:3<235::aid-jmv1>3.0.co;2-3. [DOI] [PubMed] [Google Scholar]
  • 33.Major EO, Amemiya K, Tornatore CS, Houff SA, Berger JR. Pathogenesis and molecular biology of progressive multifocal leukoencephalopathy, the JC virus-induced demyelinating disease of the human brain. Clin Microbiol Rev. 1992;5(1):49–73. doi: 10.1128/cmr.5.1.49. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Sabath BF, Major EO. Traffic of JC virus from sites of initial infection to the brain: the path to progressive multifocal leukoencephalopathy. J Infect Dis. 2002;186 (Suppl 2):S180–6. doi: 10.1086/344280. [DOI] [PubMed] [Google Scholar]
  • 35.Azzi A, De Santis R, Ciappi S, et al. Human polyomaviruses DNA detection in peripheral blood leukocytes from immunocompetent and immunocompromised individuals. J Neurovirol. 1996;2(6):411–6. doi: 10.3109/13550289609146907. [DOI] [PubMed] [Google Scholar]
  • 36.Kitamura T, Sugimoto C, Kato A, et al. Persistent JC virus (JCV) infection is demonstrated by continuous shedding of the same JCV strains. J Clin Microbiol. 1997;35(5):1255–7. doi: 10.1128/jcm.35.5.1255-1257.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Ling PD, Lednicky JA, Keitel WA, et al. The dynamics of herpesvirus and polyomavirus reactivation and shedding in healthy adults: a 14-month longitudinal study. J Infect Dis. 2003;187(10):1571–80. doi: 10.1086/374739. [DOI] [PubMed] [Google Scholar]
  • 38.Koralnik IJ, Schmitz JE, Lifton MA, Forman MA, Letvin NL. Detection of JC virus DNA in peripheral blood cell subpopulations of HIV-1-infected individuals. J Neurovirol. 1999;5(4):430–5. doi: 10.3109/13550289909029484. [DOI] [PubMed] [Google Scholar]
  • 39.Verbeeck J, Van Assche G, Ryding J, et al. JC viral loads in patients with Crohn's disease treated with immunosuppression: can we screen for elevated risk of progressive multifocal leukoencephalopathy? Gut. 2008;57(10):1393–7. doi: 10.1136/gut.2007.145698. [DOI] [PubMed] [Google Scholar]
  • 40.Kappos L, Bates D, Hartung H, et al. Natalizumab treatment for multiple sclerosis: recommendations for patient selection and monitoring. Lancet Neurol. 2007;6(5):431–41. doi: 10.1016/S1474-4422(07)70078-9. [DOI] [PubMed] [Google Scholar]
  • 41.Imperiale M, Major EO. Fields Virology. 5. 2007. Polyomaviruses; pp. 2263–2298. [Google Scholar]
  • 42.Perez-Liz G, Del Valle L, Gentilella A, Croul S, Khalili K. Detection of JC virus DNA fragments but not proteins in normal brain tissue. Ann Neurol. 2008;64(4):379–87. doi: 10.1002/ana.21443. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Rollison DEM, Utaipat U, Ryschkewitsch C, et al. Investigation of human brain tumors for the presence of polyomavirus genome sequences by two independent laboratories. Int J Cancer. 2005;113(5):769–74. doi: 10.1002/ijc.20641. [DOI] [PubMed] [Google Scholar]
  • 44.Vago L, Cinque P, Sala E, et al. JCV-DNA and BKV-DNA in the CNS tissue and CSF of AIDS patients and normal subjects. Study of 41 cases and review of the literature. J Acquir Immune Defic Syndr Hum Retrovirol. 1996;12(2):139–46. doi: 10.1097/00042560-199606010-00006. [DOI] [PubMed] [Google Scholar]
  • 45.Astrom KE, Mancall EL, Richardson EP. Progressive multifocal leuko-encephalopathy; a hitherto unrecognized complication of chronic lymphatic leukaemia and Hodgkin's disease. Brain. 1958;81(1):93–111. doi: 10.1093/brain/81.1.93. [DOI] [PubMed] [Google Scholar]
  • 46.Richardson EP. Progressive multifocal leukoencephalopathy. N Engl J Med. 1961;265:815–23. doi: 10.1056/NEJM196110262651701. [DOI] [PubMed] [Google Scholar]
  • 47.Falcó V, Olmo M, del Saz SV, et al. Influence of HAART on the clinical course of HIV-1-infected patients with progressive multifocal leukoencephalopathy: results of an observational multicenter study. J Acquir Immune Defic Syndr. 2008;49(1):26–31. doi: 10.1097/QAI.0b013e31817bec64. [DOI] [PubMed] [Google Scholar]
  • 48.Marzocchetti A, Di Giambenedetto S, Cingolani A, et al. Reduced rate of diagnostic positive detection of JC virus DNA in cerebrospinal fluid in cases of suspected progressive multifocal leukoencephalopathy in the era of potent antiretroviral therapy. J Clin Microbiol. 2005;43(8):4175–7. doi: 10.1128/JCM.43.8.4175-4177.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Marzocchetti A, Sanguinetti M, Giambenedetto SD, et al. Characterization of JC virus in cerebrospinal fluid from HIV-1 infected patients with progressive multifocal leukoencephalopathy: insights into viral pathogenesis and disease prognosis. J Neurovirol. 2007;13(4):338–46. doi: 10.1080/13550280701381324. [DOI] [PubMed] [Google Scholar]
  • 50.Iacobaeus E, Ryschkewitsch C, Gravell M, et al. Analysis of cerebrospinal fluid and cerebrospinal fluid cells from patients with multiple sclerosis for detection of JC virus DNA. Mult Scler. 2009;15(1):28–35. doi: 10.1177/1352458508096870. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Whiteman ML, Post MJ, Berger JR, et al. Progressive multifocal leukoencephalopathy in 47 HIV-seropositive patients: neuroimaging with clinical and pathologic correlation. Radiology. 1993;187(1):233–40. doi: 10.1148/radiology.187.1.8451420. [DOI] [PubMed] [Google Scholar]
  • 52.Tan K, Roda R, Ostrow L, McArthur J, Nath A. PML-IRIS in patients with HIV infection. Clinical manifestations and treatment with steroids. [Accessed February 27, 2009];Neurology. 2009 doi: 10.1212/01.wnl.0000343510.08643.74. Available at: http://www.ncbi.nlm.nih.gov/pubmed/19129505. [DOI] [PMC free article] [PubMed]
  • 53.Stüve O, Marra CM, Cravens PD, et al. Potential risk of progressive multifocal leukoencephalopathy with natalizumab therapy: possible interventions. Arch Neurol. 2007;64(2):169–76. doi: 10.1001/archneur.64.2.169. [DOI] [PubMed] [Google Scholar]
  • 54.Berger JR. Steroids for PML-IRIS. A double-edged sword? [Accessed February 27, 2009];Neurology. 2009 doi: 10.1212/01.wnl.0000343735.44983.5e. Available at: http://www.ncbi.nlm.nih.gov/pubmed/19144930. [DOI] [PubMed]
  • 55.Crowder CD, Gyure KA, Drachenberg CB, et al. Successful outcome of progressive multifocal leukoencephalopathy in a renal transplant patient. Am J Transplant. 2005;5(5):1151–8. doi: 10.1111/j.1600-6143.2005.00800.x. [DOI] [PubMed] [Google Scholar]
  • 56.Shitrit D, Lev N, Bar-Gil-Shitrit A, Kramer MR. Progressive multifocal leukoencephalopathy in transplant recipients. Transpl Int. 2005;17(11):658–65. doi: 10.1007/s00147-004-0779-3. [DOI] [PubMed] [Google Scholar]
  • 57.Khatri BO, Man S, Giovannoni G, et al. Effect of plasma exchange in accelerating natalizumab clearance and restoring leukocyte function. Neurology. 2009;72(5):402–9. doi: 10.1212/01.wnl.0000341766.59028.9d. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Stüve O, Marra CM, Jerome KR, et al. Immune surveillance in multiple sclerosis patients treated with natalizumab. Ann Neurol. 2006;59(5):743–7. doi: 10.1002/ana.20858. [DOI] [PubMed] [Google Scholar]
  • 59. [Accessed March 17, 2009];Study to Explore the Effect of Mefloquine in Subjects With Progressive Multifocal Leukoencephalopathy (PML) - Full Text View - ClinicalTrials.gov. Available at: http://clinicaltrials.gov/ct2/show/NCT00746941?term=pml+mefloquine&rank=1.
  • 60.Brickelmaier M, Lugovskoy A, Kartikeyan R, et al. Identification and charaterization of mefloquine efficacy against JC virus in vitro. [Accessed March 17, 2009];Antimicrob Agents Chemother. 2009 doi: 10.1128/AAC.01614-08. Available at: http://www.ncbi.nlm.nih.gov/pubmed/19258267. [DOI] [PMC free article] [PubMed]
  • 61.Stüve O, Marra CM, Bar-Or A, et al. Altered CD4+/CD8+ T-cell ratios in cerebrospinal fluid of natalizumab-treated patients with multiple sclerosis. Arch Neurol. 2006;63(10):1383–7. doi: 10.1001/archneur.63.10.1383. [DOI] [PubMed] [Google Scholar]
  • 62.Sheremata WA, Vollmer TL, Stone LA, Willmer-Hulme AJ, Koller M. A safety and pharmacokinetic study of intravenous natalizumab in patients with MS. Neurology. 1999;52(5):1072–4. doi: 10.1212/wnl.52.5.1072. [DOI] [PubMed] [Google Scholar]
  • 63.Martin MDP, Cravens PD, Winger R, et al. Decrease in the Numbers of Dendritic Cells and CD4+ T Cells in Cerebral Perivascular Spaces Due to Natalizumab. Arch Neurol. 2008;65(12):1596–603. doi: 10.1001/archneur.65.12.noc80051. [DOI] [PubMed] [Google Scholar]
  • 64.Gasnault J, Kahraman M, de Goër de Herve MG, et al. Critical role of JC virus-specific CD4 T-cell responses in preventing progressive multifocal leukoencephalopathy. AIDS. 2003;17(10):1443–9. doi: 10.1097/00002030-200307040-00004. [DOI] [PubMed] [Google Scholar]
  • 65.Du Pasquier RA, Clark KW, Smith PS, et al. JCV-specific cellular immune response correlates with a favorable clinical outcome in HIV-infected individuals with progressive multifocal leukoencephalopathy. J Neurovirol. 2001;7(4):318–22. doi: 10.1080/13550280152537175. [DOI] [PubMed] [Google Scholar]
  • 66.Koralnik IJ, Du Pasquier RA, Kuroda MJ, et al. Association of prolonged survival in HLA-A2+ progressive multifocal leukoencephalopathy patients with a CTL response specific for a commonly recognized JC virus epitope. J Immunol. 2002;168(1):499–504. doi: 10.4049/jimmunol.168.1.499. [DOI] [PubMed] [Google Scholar]
  • 67.Du Pasquier RA, Kuroda MJ, Schmitz JE, et al. Low frequency of cytotoxic T lymphocytes against the novel HLA-A*0201-restricted JC virus epitope VP1(p36) in patients with proven or possible progressive multifocal leukoencephalopathy. J Virol. 2003;77(22):11918–26. doi: 10.1128/JVI.77.22.11918-11926.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Bennett SR, Carbone FR, Karamalis F, Miller JF, Heath WR. Induction of a CD8+ cytotoxic T lymphocyte response by cross-priming requires cognate CD4+ T cell help. J Exp Med. 1997;186(1):65–70. doi: 10.1084/jem.186.1.65. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Monaco MC, Atwood WJ, Gravell M, Tornatore CS, Major EO. JC virus infection of hematopoietic progenitor cells, primary B lymphocytes, and tonsillar stromal cells: implications for viral latency. J Virol. 1996;70(10):7004–12. doi: 10.1128/jvi.70.10.7004-7012.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Tornatore C, Berger JR, Houff SA, et al. Detection of JC virus DNA in peripheral lymphocytes from patients with and without progressive multifocal leukoencephalopathy. Ann Neurol. 1992;31(4):454–62. doi: 10.1002/ana.410310426. [DOI] [PubMed] [Google Scholar]
  • 71.Bellucci R, De Propris MS, Buccisano F, et al. Modulation of VLA-4 and L-selectin expression on normal CD34+ cells during mobilization with G-CSF. Bone Marrow Transplant. 1999;23(1):1–8. doi: 10.1038/sj.bmt.1701522. [DOI] [PubMed] [Google Scholar]
  • 72.Steen R, Tjønnfjord GE, Gunnes Grøseth LA, Heldal D, Egeland T. Efflux of CD34+ cells from bone marrow to peripheral blood is selective in steady-state hematopoiesis and during G-CSF administration. J Hematother. 1997;6(6):563–73. doi: 10.1089/scd.1.1997.6.563. [DOI] [PubMed] [Google Scholar]
  • 73.Yamaguchi M, Ikebuchi K, Hirayama F, et al. Different adhesive characteristics and VLA-4 expression of CD34(+) progenitors in G0/G1 versus S+G2/M phases of the cell cycle. Blood. 1998;92(3):842–8. [PubMed] [Google Scholar]
  • 74.Lichterfeld M, Martin S, Burkly L, Haas R, Kronenwett R. Mobilization of CD34+ haematopoietic stem cells is associated with a functional inactivation of the integrin very late antigen 4. Br J Haematol. 2000;110(1):71–81. doi: 10.1046/j.1365-2141.2000.02130.x. [DOI] [PubMed] [Google Scholar]
  • 75.Bonig H, Wundes A, Chang K, Lucas S, Papayannopoulou T. Increased numbers of circulating hematopoietic stem/progenitor cells are chronically maintained in patients treated with the CD49d blocking antibody natalizumab. Blood. 2008;111(7):3439–41. doi: 10.1182/blood-2007-09-112052. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Zohren F, Toutzaris D, Klärner V, et al. The monoclonal anti-VLA-4 antibody natalizumab mobilizes CD34+ hematopoietic progenitor cells in humans. Blood. 2008;111(7):3893–5. doi: 10.1182/blood-2007-10-120329. [DOI] [PubMed] [Google Scholar]
  • 77.Fabbri M, Bianchi E, Fumagalli L, Pardi R. Regulation of lymphocyte traffic by adhesion molecules. Inflamm Res. 1999;48(5):239–46. doi: 10.1007/s000110050454. [DOI] [PubMed] [Google Scholar]
  • 78.Papayannopoulou T. Mechanisms of stem-/progenitor-cell mobilization: the anti-VLA-4 paradigm. Semin Hematol. 2000;37(1 Suppl 2):11–8. doi: 10.1016/s0037-1963(00)90084-2. [DOI] [PubMed] [Google Scholar]
  • 79.Christ O, Kronenwett R, Haas R, Zöller M. Combining G-CSF with a blockade of adhesion strongly improves the reconstitutive capacity of mobilized hematopoietic progenitor cells. Exp Hematol. 2001;29(3):380–90. doi: 10.1016/s0301-472x(00)00674-3. [DOI] [PubMed] [Google Scholar]
  • 80.Craddock CF, Nakamoto B, Andrews RG, Priestley GV, Papayannopoulou T. Antibodies to VLA4 integrin mobilize long-term repopulating cells and augment cytokine-induced mobilization in primates and mice. Blood. 1997;90(12):4779–88. [PubMed] [Google Scholar]
  • 81.Papayannopoulou T, Nakamoto B. Peripheralization of hemopoietic progenitors in primates treated with anti-VLA4 integrin. Proc Natl Acad Sci U S A. 1993;90(20):9374–8. doi: 10.1073/pnas.90.20.9374. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Bennett JL. Natalizumab and progressive multifocal leukoencephalopathy: migrating towards safe adhesion molecule therapy in multiple sclerosis. Neurol Res. 2006;28(3):291–8. doi: 10.1179/016164106X98189. [DOI] [PubMed] [Google Scholar]
  • 83.Houff SA, Berger J, Major EO. Response to Lindberg et al. Natalizumab alters transcriptional expression profiles of blood cell subpopulations of multiple sclerosis patients. J Neuroimmunol. 2008;199(1–2):160–161. doi: 10.1016/j.jneuroim.2008.05.007. [DOI] [PubMed] [Google Scholar]
  • 84.Ransohoff RM. Natalizumab and PML. Nat Neurosci. 2005;8(10):1275. doi: 10.1038/nn1005-1275. [DOI] [PubMed] [Google Scholar]
  • 85.Del Valle L, White MK, Khalili K. Potential mechanisms of the human polyomavirus JC in neural oncogenesis. J Neuropathol Exp Neurol. 2008;67(8):729–40. doi: 10.1097/NEN.0b013e318180e631. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Ahlqvist J, Fotheringham J, Akhyani N, et al. Differential tropism of human herpesvirus 6 (HHV-6) variants and induction of latency by HHV-6A in oligodendrocytes. J Neurovirol. 2005;11(4):384–94. doi: 10.1080/13550280591002379. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Donati D, Martinelli E, Cassiani-Ingoni R, et al. Variant-specific tropism of human herpesvirus 6 in human astrocytes. J Virol. 2005;79(15):9439–48. doi: 10.1128/JVI.79.15.9439-9448.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Fotheringham J, Jacobson S. Human herpesvirus 6 and multiple sclerosis: potential mechanisms for virus-induced disease. Herpes. 2005;12(1):4–9. [PubMed] [Google Scholar]
  • 89.Gardell JL, Dazin P, Islar J, et al. Apoptotic effects of Human Herpesvirus-6A on glia and neurons as potential triggers for central nervous system autoimmunity. J Clin Virol. 2006;37 (Suppl 1):S11–6. doi: 10.1016/S1386-6532(06)70005-1. [DOI] [PubMed] [Google Scholar]
  • 90.Mock DJ, Powers JM, Goodman AD, et al. Association of human herpesvirus 6 with the demyelinative lesions of progressive multifocal leukoencephalopathy. J Neurovirol. 1999;5(4):363–73. doi: 10.3109/13550289909029477. [DOI] [PubMed] [Google Scholar]
  • 91.Yao K, Gagnon S, Akhyani N, et al. Reactivation of human herpesvirus-6 in natalizumab treated multiple sclerosis patients. PLoS ONE. 2008;3(4):e2028. doi: 10.1371/journal.pone.0002028. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Andre-Garnier E, Milpied N, Boutolleau D, et al. Reactivation of human herpesvirus 6 during ex vivo expansion of circulating CD34+ haematopoietic stem cells. J Gen Virol. 2004;85(Pt 11):3333–6. doi: 10.1099/vir.0.80319-0. [DOI] [PubMed] [Google Scholar]
  • 93.Luppi M, Barozzi P, Morris C, et al. Human herpesvirus 6 latently infects early bone marrow progenitors in vivo. J Virol. 1999;73(1):754–9. doi: 10.1128/jvi.73.1.754-759.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Mullen JT, Vartanian TK, Atkins MB. Melanoma complicating treatment with natalizumab for multiple sclerosis. N Engl J Med. 2008;358(6):647–8. doi: 10.1056/NEJMc0706103. [DOI] [PubMed] [Google Scholar]
  • 95.Stüve O, Cravens PD, Frohman EM, et al. Immunologic, clinical, and radiologic status 14 months after cessation of natalizumab therapy. Neurology. 2009;72(5):396–401. doi: 10.1212/01.wnl.0000327341.89587.76. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Vellinga MM, Castelijns JA, Barkhof F, Uitdehaag BMJ, Polman CH. Postwithdrawal rebound increase in T2 lesional activity in natalizumab-treated MS patients. Neurology. 2008;70(13 Pt 2):1150–1. doi: 10.1212/01.wnl.0000265393.03231.e5. [DOI] [PubMed] [Google Scholar]

RESOURCES