You are on page 1of 10

REVIEWS

Pathogenesis, classification and treatment


of inflammatory myopathies
Mei Zong and Ingrid E. Lundberg
Abstract | The inflammatory myopathies—collectively, myositis—are a heterogeneous group of chronic
muscle disorders that differ in response to immunosuppressive treatment. Insufficient knowledge of the
molecular pathways that drive pathogenesis (and underlie the clinical differences between subtypes) has
hindered accurate classification, which in turn has been detrimental for clinical research. Nevertheless, new
insights into pathogenesis are paving the way for improvements in diagnosis, classification and treatment.
Accumulating data suggest that both immune and nonimmune mechanisms cause muscle weakness.
Phenotyping of the T cells that accumulate in muscle tissue has identified proinflammatory, apoptosis
resistant and cytotoxic CD4+ and CD8+ CD28null populations. Several myositis-specific autoantibodies
have been identified, associated with distinct clinical phenotypes. Thus, adaptive immunity is involved
in pathogenesis, and both T and B cells are interesting targets for therapy. Furthermore, genotyping has
revealed activation of the type I interferon pathway in patients with dermatomyositis or with expression of
particular autoantibodies. Decreased release of Ca2+ from the sarcoplasmic reticulum, as a consequence
of release of proinflammatory cytokines and high mobility group protein B1, might contribute to muscle
weakness, and nonimmune mechanisms potentially include a role for endoplasmic reticulum stress,
autophagy and hypoxia. Deeper understanding, careful phenotyping of patients—and new classification
criteria—will expedite clinical research.
Zong, M. & Lundberg, I. E. Nat. Rev. Rheumatol. 7, 297–306 (2011); published online 5 April 2011; doi:10.1038/nrrheum.2011.39

Introduction Pathogenesis
The idiopathic inflammatory myopathies (IIM), col­ Muscle is the target organ of inflammation in myositis,
lectively named myositis, are characterized clinically and the distinct histopathological features that occur in
by weakness and low endurance of skeletal muscle, and muscle tissue indicate that different disease mechanisms
histo­pathologically by the presence of inflammatory cells predominate in each IIM subtype. Thus, inflammatory
in muscle tissue. On the basis of differences in clinical and infiltrates surrounding vessels are preferentially found in
histopathological findings, separate IIM subtypes have patients with dermatomyositis, and might indicate an auto­
been identified, most often classified into poly­myositis, immune reaction directed against blood vessels, whereas
dermatomyositis and sporadic inclusion body myo­ cell infiltrates predominately surround muscle fibers
sitis (sIBM).1–3 Organs other than muscle are frequently in patients with polymyositis and sIBM, and, similarly,
involved in polymyositis and dermatomyositis, but less might indicate an autoimmune reaction directed against
often in sIBM, including skin in dermato­myositis, and muscle fibers in these myositis subsets. Histopathological
lungs in both polymyositis and dermatomyositis, suggest­ features are not always distinct, how­ever, and a substantial
ing that these ‘myopathies’ are actually systemic connective proportion of patients have mild or no histopathological
tissue diseases. changes despite pronounced symptoms. This observation
Autoantibodies are frequently present in IIM, particu­ has led to the proposal that both immune and nonimmune
larly in polymyositis and dermatomyositis, and a number mechanisms are involved in causing muscle weakness.4,5
of autoantibodies specific to myositis have recently been Furthermore, the molecular pathogenesis of sIBM is the
detected, most of them in association with distinct clinical subject of debate, and is suggested to comprise a primary
phenotypes. These findings have raised a need for novel degenerative disease process accompanied by a secondary Rheumatology Unit,
classification criteria that take account of the newly dis­ immune reaction, based on the typical findings of accumu­ Department of
covered autoantibodies. Indeed, increasing the mol­ecular lations of degenerative proteins in muscle tissue in sIBM.6,7 Medicine, Karolinska
University Hospital,
detail used in classification of patient subsets is likely to Another histo­pathological feature of some myopathies is Solna, Karolinska
improve our understanding of important pathogenic muscle fiber necrosis with few or no inflammatory infil­ Institutet, SE-171 76
Stockholm, Sweden
pathways in IIM, as we discuss in this article. trates; this disease subset is often named necrotizing myo­ (M. Zong,
pathy, yet involvement of the immune system is suggested I. E. Lundberg)
by the recently detected presence of autoantibodies.8,9
Correspondence to:
Competing interests Thus, the pathogenic mechanisms that are involved in I. E. Lundberg
The authors declare no competing interests. clinical myositis are complex. In this Review we discuss ingrid.lundberg@ki.se

NATURE REVIEWS | RHEUMATOLOGY VOLUME 7  |  MAY 2011  |  297


© 2011 Macmillan Publishers Limited. All rights reserved
REVIEWS

Key points T cells have cytotoxic effects similar to natural killer cells;
they are proinflammatory and apoptosis resistant, and
■■ Myositis is a heterogeneous group of chronic inflammatory muscle disorders, the
origins of which are not yet clear and for which efficient treatment is largely lacking
might contribute to the chronicity and treatment resis­
tance seen in myositis. However, whether these cells have
■■ Recent findings suggest that both immune and nonimmune mechanisms are
involved in the pathogenesis of myositis, and that different molecular pathways
myocytotoxic effects remains to be demonstrated.
might predominate in different subsets of myositis Other T-cell subtypes have been found in IIM cell infil­
trates. Type 17 T helper cells, which have a role in the
■■ The immune mechanisms involve immune cells—T cells, B cells, dendritic cells
and macrophages—and their products, such as cytokines and antibodies pathogenesis of several autoimmune diseases, have been
identified in inflamed muscle tissue of patients with poly­
■■ Myositis-specific autoantibodies are helpful in the diagnosis of myositis, they
identify different clinical subsets of myositis, and they might be important for myositis or dermatomyositis.19,20 These cells are a major
differentiating pathogenic mechanisms between patients with myositis source of IL‑17, a proinflammatory cytokine that plays
■■ Among nonimmune pathogenic mechanisms, there seem to be roles for
an important part in rheumatoid arthritis (RA). IL‑17
endoplasmic reticulum stress, hypoxia and autophagy, contributing to the cause in combination with IL‑1, another important cytokine in
of muscle weakness IIM, can induce expression of MHC class I molecules
■■ New classification criteria are needed to identify more homogenous subsets and IL‑6 in cultured myoblasts, and might have a role in
of patients, and subsets that are likely to share molecular pathways disease mechanisms of IIM.21
■■ Studying the outcomes of targeted therapies will facilitate understanding Another T‑cell subset, FOXP3+ regulatory T (TREG) cells,
of disease mechanisms has been detected in dermatomyositis, polymyositis and
sIBM, with localization close to other mononuclear cells
surrounding or invading muscle fibers.22 TREG cells are
the potential roles of both immune and nonimmune important regulators of the immune system, and low
mechanisms in myositis, with a focus on the subtypes that numbers or dysfunction of these cells are associated
are most widely accepted to be immune-mediated, poly­ with several autoimmune disorders. TREG cell numbers in
myositis and dermatomyositis. We also discuss the insights peripheral blood correlate inversely with the severity of
into pathogenesis that might be offered by studying the disease activity in patients with polymyositis or dermato­
clinical and molecular effects of targeted therapies. myositis.23 On the basis of these observations, therapies
that block T cells seem to be an interesting prospect, but
Immune mechanisms such blockade will probably require targeting of specific
The various histopathological patterns of inflamma­ T‑cell subsets.
tion that occur in myositis are associated with different
immune cell phenotypes. Thus, endomysial infiltrates— B cells in myositis
found predominantly in polymyositis and sIBM—are B cells are present in IIM less often than T cells, and are
mainly composed of CD8+ T cells, CD4+ T cells, dendritic preferentially found in perivascular infiltrates in dermato­
cells (DCs) and macrophages, whereas the peri­vascular myositis, but plasma cells have been detected in all three
infiltrates that predominate in dermatomyositis are mainly subtypes of IIM.24 B cells and plasma cells occasionally
composed of CD4+ T cells, DCs, and macro­phages, with form germinal centers in muscle tissue of patients with
occasional B cells (Figure 1).10–12 IIM.25 The observation that these intramuscular B cells
and plasma cells have undergone clonal expansion, class-
T cells in myositis switch recombination and somatic hypermutation indi­
The precise role of T cells in IIM pathogenesis is still cates that an antigen-driven B-cell response has occurred,
not clear, but their involvement is supported by genetic but the nature of the antigen is not yet clear.26 A role for
associa­tion in myositis; for example, HLA-DRB1*0301 is B cells in the pathogenesis of polymyositis and dermato­
a risk factor for both polymyositis and dermato­myositis, myositis is supported by favorable responses to the B‑cell-
and an important role of HLA-DR molecules is to present blocking therapy rituximab, as reported in case studies.27–31
antigens to CD4+ T cells.13 Similarly, CD8+ T cells are Furthermore, involvement of B cells in the pathogenesis of
plentiful in muscle tissue of patients with poly­myositis all IIM subtypes is indicated by the frequent occurrence
and sIBM. These cells express perforin‑1 and granzyme B, of autoantibodies, as we discuss in the next section.
suggesting that they have a myocytotoxic effect, but the
co-stimulatory factors required for such an effect have Dendritic cells and macrophages in myositis
not been convincingly demonstrated in myositis muscle Both immature and mature DCs are present in lympho­
fibers.14 Previous studies have identified clonally restricted cytic infiltrates of muscle tissue of patients with poly­
T-cell populations that might be indicative of a patho­ myositis or dermatomyositis, where they might serve as
genic process driven or perpetuated by antigens.15 In this anti­gen presenting cells.32 In addition, plasmacytoid DCs
context, the observation that many muscle-­infiltrating (pDCs), a main source of type I interferon (IFN), are also
CD4+ and CD8+ T cells in patients with IIM are so-called present in muscle tissue of patients with poly­myositis
CD28null T cells, which lack the activation molecule CD28, or dermatomyositis, as well as in the skin of patients
is interesting.16,17 CD28null T cells represent chronically with dermatomyosit­is.33–35 pDCs have occasionally been
stimulated, clonally-restricted effector cells that bypass the observed in muscle tissue of patients with sIBM.33
co-stimulation by T‑lymphocyte activation antigen CD86 Infiltrating macrophages are often detected in muscle
and are instead co-stimulated by other means.18 CD28null tissue from patients with IIM, of all subtypes. They may

298  |  MAY 2011  |  VOLUME 7  www.nature.com/nrrheum


© 2011 Macmillan Publishers Limited. All rights reserved
REVIEWS

a b c di
ei
Susceptibility Lymph nodes Circulation
genes
Muscle
fiber

Polymyositis
dii
eii

Environmental Muscle
factors fiber

Dermatomyositis

CD4+ T cell pDC Virus Vessel Perforin and granzyme

CD8+ T cell mDC Bacteria Type I IFN

B cell Macrophage Autoantibody Cytokines

Figure 1 | Potential immune mechanisms in myositis pathogenesis. a | Immune activators, for example, respiratory
infection, trigger disease in myositis susceptibility gene carriers. Other environmental factors might contribute. b | Innate
and adaptive responses are activated in lymph nodes; macrophages secrete cytokines; mDCs present antigens to T cells;
activated T cells assist B-cell activation and proliferation; some B cells differentiate into autoantibody-secreting plasma
cells. c | Activated immune cells enter general circulation. Upon encountering specific antigens, immune cells penetrate
endothelia, into local muscle. di | In polymyositis (and perhaps sIBM), CD8+ T cells expressing perforin-1 and granzyme
invade non-necrotic muscle fibers expressing MHC class I molecules (continuous MHC expression might be induced by
pDC-derived type I IFN); mDCs activate T cells, and macrophage-derived cytokines control T-cell homeostasis, activation
and proliferation. dii | In dermatomyositis, infiltrates—composed of CD4+ T cells, B cells, DCs and macrophages—usually
invade perivascular and perimysial muscle tissue, and might organize into structures resembling lymph node germinal
centers. ei | Immunohistochemistry of muscle tissue from a patient with polymyositis, showing CD8 + T cells localized to
endomysium. eii | Perivascular localization of CD4+ T cells in muscle tissue of a patient with dermatomyositis, revealed by
immunohistochemistry. Abbreviations: IFN, interferon; mDC, myeloid dendritic cell; pDC, plasmacytoid dendritic cell; sIBM,
sporadic inclusion body myositis.

act by presenting antigens to T cells, by invading and juvenile dermatomyositis, and with cancer in adult
clearing up necrotic muscle fibers, and/or most impor­ dermatomyositis.48 The antigen is not yet known, but
tantly, they can be a source of a multitude of cytokines has been suggested to be transcriptional intermedi­
and chemokines.36–39 ary factor γ (TIF1γ; also known as E3 ubiquitin-protein
ligase TRIM33).49 The only reliable detection method for
Autoantibodies in myositis anti‑155/140 is radioactive immunoprecipitation, which
Myositis-specific autoantibodies (MSAs) are strongly is unsuitable for routine analyses.
associated with distinct clinical profiles. The most fre­ Another MSA, anti-CADM‑140, is associated with
quent MSAs are the anti-Jo‑1 autoantibodies, which are a subset of patients with dermatomyositis who have
directed against histidyl-tRNA synthetase and are present cutaneous lesions in the absence of muscle weakness, a
in 20–25% of patients with polymyositis or dermato­ presentation known as clinically amyopathic dermato­
myositis. Their levels correlate with disease activity, and myositis (CADM). These autoantibodies have so far only
they are associated with a distinct clinical phenotype, been reported in Asian patients.50 Patients with anti-
antisynthetase syndrome (Box 1).40,41 Altogether, eight CADM‑140 sometimes develop life-threatening acute
types of antisynthetase autoantibodies have been iden­ progressive interstitial lung disease (ILD).50 The antigen
tified (Table 1),42–46 and MSAs directed against other has been identified as an RNA helicase called melanoma
autoantigens also occur. For example, antibodies to Mi‑2 differentiation-associated protein 5 (MDA‑5; also known
autoantigens (also known as chromodomain-helicase- as interferon-induced helicase C domain-containing
DNA-binding proteins) are preferentially present in protein 1), and encoded by IFIH1. Finally, a novel auto­
patients with dermatomyositis, and are usually associated antibody, recognizing 200 kD and 100 kD proteins, has
with a good response to treatment. The Mi‑2 auto­antigen been detected in patients with necrotizing myopathy who
is expressed in regenerating muscle fibers, and might were previously considered to be auto­antibody negative.8
have a modulatory role during myofiber regeneration.47 The MSAs can be helpful in the diagnosis of IIM. The
Anti-signal recognition particle (SRP) autoantibodies are specific clinical phenotypes associated with the various
associat­ed with necrotizing myopathy (Table 1). MSAs might also indicate that MSAs and/or their anti­
Other MSAs have been identified recently (Table 1). gen targets have a role in the pathogenesis of myo­sitis.
These include anti‑155/140, which is associated with Although histidyl-tRNA synthetase is ubiquitously

NATURE REVIEWS | RHEUMATOLOGY VOLUME 7  |  MAY 2011  |  299


© 2011 Macmillan Publishers Limited. All rights reserved
REVIEWS

Box 1 | Clinical features of antisynthetase syndrome of myositis, which is further supported by the beneficial
therapeutic effects of anakinra (a recombinant IL‑1R
Myositis
antagonist) in patients with myositis, as discussed in
Polymyositis or dermatomyositis
further detail in the section on treatment.
Polyarthritis TNF can have a direct impact on muscle function,
Non-destructive polyarthritis of finger joints, wrists, although its importance in IIM is not well defined.58 TNF
elbows or knees expression is low in muscle tissue of patients with IIM, and
Fever TNF-blocking therapy has generated inconsistent results
Present in 80% of patients (further discussed in the section on treatment).59 Some
Mechanic’s hands evidence exists of a role for B‑cell-activating factor (BAFF;
Thick cracked skin over the tips and sides of the fingers
also known as TNF ligand superfamily member 13B),
which enhances the survival of immature and mature
Interstitial lung disease
B cells, and is involved in autoantibody production as
Present in over 70% of patients
well as in T‑cell activation and differen­t iation. High
Raynaud phenomenon serum levels of BAFF were found in patients with IIM,
Present in 60% of patients particular­ly in patients with anti-Jo‑1 autoantibodies.60
Accumulating evidence suggests a role for type I IFN
in IIM, on the basis of IFN-inducible gene expression in
expressed in cells, its expression is highest in organs tar­ muscle tissue and peripheral blood, which is particularly
geted by antisynthetase syndrome, and in regenerating striking in dermatomyositis.34,61 Complexes containing
muscle fibers.51 RNA and anti-Jo‑1, or autoantibodies to the SSA sub­
A pathogenic role for histidyl-tRNA synthetase is sup­ units Ro52 and Ro60, can act as inducers of endogenous
ported by its effect in C57BL/6 and NOD congenic mice.52 type I IFN, and activate IFN production in pDCs, similar
Immunization with the autoantibody is able to induce to mechanisms reported in SLE.33,54 Thus, patients with
myositis, ILD and anti-Jo‑1 antibodies, all features of anti-Jo‑1 or anti-Ro52 autoantibodies might share activa­
human antisynthetase syndrome, in the mice.52 Notably, tion of the type I IFN pathway regardless of whether they
the muscle inflammation is independent of T or B cells, are diag­nosed with polymyositis or dermatomyositis.
as the NOD congenic mice lack these cell types. This IFNs are major inducers of expression of MHC class I and
observed independence suggests that histidyl-tRNA class II antigens, elevated levels of which are present in
synthetase can trigger both innate and adaptive immune muscle fibers of patients with IIM (Figure 2). Expression of
responses, and lead to muscle inflammation.53 Further MHC class I molecules has been observed to correlate with
support for involvement of the anti-Jo‑1 anti­bodies in the the presence of pDCs positive for BDCA‑2 (blood den­
disease mechanisms is provided by the finding that sera dritic cell antigen 2, also known as C‑type lectin domain
containing anti-Jo‑1 autoantibodies together with nucleic family 4 member C) in IIM muscle tissue.33
acid can activate the type I IFN system,33 similar to reports IL‑6 is a proinflammatory cytokine that is critically
of the effect of antibodies to the SSA auto­antigen and U1 involved in immune responses. Occasional IL‑6+ cells have
ribonucleoprotein complex in systemic lupus erythemato­ been reported in muscle tissue from patients with poly­
sus (SLE).54 Sera from patients with myositis positive for myositis or dermatomyositis,57 and cultured muscle fibers
anti-Jo‑1 antibodies have also been shown to induce were found to release IL‑6.62 Furthermore, in a myositis
increased expression of intercellular adhesion molecule 1 mouse model, clinical outcome was improved by reducing
in human lung endothelial cells, in comparison with other levels of IL‑6.63
sera, further supporting a potential role for MSAs in the Elevated serum levels of IL‑15, as well as expression of
disease mechanisms of myositis.55 IL‑15 in muscle tissue, have been reported in patients with
polymyositis or dermatomyositis.64,65 IL‑15 is an important
Cytokines in myositis activator of T cells, and it can downregulate the expres­
Cytokines are generated by immune cells, endo­thelial cells sion of CD28 on T-cell membranes, producing CD28null
and muscle cells, and can contribute to a pro­inflammatory T cells—identified as a dominant T‑cell pheno­type in
environment in the muscle tissue of patients with myo­ muscle tissue of patients with myositis.16,17,66 These obser­
sitis. IL‑1, tumor necrosis factor (TNF) and IFNs are the vations have raised an interest in the pathogenic role of
most frequently found cyto­kines in muscle tissue. Roles IL‑15 in IIM. Importantly, IL‑15 is also involved in muscle
for other cytokines, such as IL‑6, IL‑15 and the alarmin fiber regeneration, in a mechanism yet to be elucidated.67
high mobility group protein B1 (HMGB1), have however, HMGB1 is a DNA binding protein that is regarded
also been proposed, as we discuss in this section. as an alarmin, as it is released into the extranuclear and
Overexpression of IL‑1α and IL‑1β has been demon­ extra­cellular spaces when cells are subjected to various
strated in muscle tissue from patients with polymyositis, types of stress, such as hypoxia. Implicated in the patho­
dermatomyositis or sIBM.56,57 Similarly, IL‑1 receptors genesis of rheumatic diseases,68 HMGB1 expression has
(IL-1R) have been observed in muscle fibers from patients been identified in synovial tissue of patients with RA,
with polymyositis or dermatomyositis, and co-localization and in joints of arthritis mouse models, where HMGB1-
between IL‑1R and their respective ligands was noted.56,57 neutralizing therapy can attenuate joint inflammation and
These findings suggest a role for IL‑1 in the pathogenesis damage.69,70 Expression of HMGB1 has also been observed

300  |  MAY 2011  |  VOLUME 7  www.nature.com/nrrheum


© 2011 Macmillan Publishers Limited. All rights reserved
REVIEWS

Table 1 | Myositis-specific autoantibodies


Autoantibody name Antigen Clinical association Frequency in IIM
Antibodies to aminoacyl-tRNA synthetases
Anti-Jo‑1 Histidyl-tRNA synthetase Antisynthetase syndrome 20–25%
Anti-PL‑7 Threonyl-tRNA synthetase Antisynthetase syndrome 5–10%
Anti-PL‑12 Alanyl-tRNA synthetase Antisynthetase syndrome <5%
Anti-Ej Glycyl-tRNA synthetase Antisynthetase syndrome 5–10%
Anti-OJ Isoleucyl-tRNA synthetase Antisynthetase syndrome <5%
Anti-KS Asparaginyl-tRNA synthetase Antisynthetase syndrome Rare
Anti-Ha Tyrosyl-tRNA synthetase Antisynthetase syndrome One patient
Anti-Zo Phenylalanyl-tRNA synthetase Antisynthetase syndrome One patient
Other myositis-specific autoantibodies
Anti‑P155/140 TIF1γ (putative) Juvenile dermatomyositis, 13–30%
cancer in adult dermatomyositis
Anti-CADM‑140 MDA‑5 CADM 50% of CADM
Anti-Mi‑2 Mi‑2α and Mi‑2β Dermatomyositis 8% of IIM generally, 15–20% of
dermatomyositis
Anti-SRP SRP Necrotizing myopathy 4–6% of IIM
Anti‑200/100 200 kD and 100 kD proteins Necrotizing myopathy 60% of necrotizing myopathy
Abbreviations: CADM, clinically amyopathic dermatomyositis; IIM, idiopathic inflammatory myopathies; Mi‑2, Mi‑2 autoantigens; MDA‑5, melanoma
differentiation-associated protein 5; SRP, signal recognition particle; TIF1γ, transcriptional intermediary factor γ.

in inflammatory infiltrating cells, endothelial cells and


muscle fibers of patients with polymyositis or dermato­
myositis.71 Furthermore, HMGB1 can induce reversible
upregulation of MHC class I molecules in differentiated,
healthy muscle fibers, and an irreversible decrease in Ca2+
release from the sarcoplasmic reticulum during induction
*
of fatigue in intact single fibers.72 These data indicate that
HMGB1 might be involved in the mechanisms of muscle
weakness in IIM. Moreover, HMGB1 might be impor­
tant in the initiation phase of myositis, as its extranuclear * *
expression occurs in endothelial cells and muscle fibers
of patients with early myositis without detectable inflam­
matory infiltrates in muscle.71 Further work is required,
however, to determine the role of HMGB1 in IIM, as well
as to understand its effects on calcium homeostasis in
skeletal muscle.
Figure 2 | Localization of MHC class I antigens in muscle
Nonimmune mechanisms tissue of a patient with myositis. Note the MHC class I
expression (brown staining) both on the membrane (arrow)
The potential involvement of nonimmune mechanisms
and in the cytoplasm (stars) of muscle fibers, as revealed by
in the pathogenesis of myositis is increasingly recognized. immunohistochemistry. Inset in upper right corner shows a
Such mechanisms include hypoxia, endoplasmic reticulum negative control stain using an isotype control from the same
(ER) stress and autophagy, all of which we discuss here. species as the antibody to MHC class I, in the same patient.
Although we consider these three aspects separately, they
are actually closely interlinked (Figure 3); for example, All of these findings have suggested a role for hypoxia in
hypoxia might induce ER stress, as discussed below. the pathogenesis of myositis. Further support is provided
by the low percentage of type I, oxygen-dependent, muscle
Hypoxia fibers in patients with chronic polymyositis or dermato­
Why consider hypoxia in myositis? The question was raised myositis, together with improved muscle endurance and
by several observations: firstly, substantial muscle weak­ an increased percentage of type I fibers after a 12‑week
ness can occur despite little or no apparent muscle inflam­ submaximal home exercise program.78 Moreover, after a
mation;73 secondly, capillaries are lost in IIM tissues;74–75 7‑week intensive resistance training program, the levels of
and finally, overexpression of hypoxia-inducible factor 1‑α 265 gene transcripts were substantially modified in muscle
(HIF‑1-α) and its downstream gene products, vascular tissue of patients with myositis, and seven of them showed
endothelial growth factor and erythro­poietin, and their a shift towards altered expression of enzymes involved
receptors, are all detectable in muscle tissues in IIM.76,77 with oxidative metabolism.79

NATURE REVIEWS | RHEUMATOLOGY VOLUME 7  |  MAY 2011  |  301


© 2011 Macmillan Publishers Limited. All rights reserved
REVIEWS

Myositis cyclic AMP-dependent transcription factor ATF6α


pathways). 82,83 Failure of these mechanisms leads to
cell apoptosis.
HMGB1 or other MHC-1 Muscle fatigue In skeletal muscle, several proteins that bind and/or
proinflammatory cytokines regulate calcium exist in the sarcoplasmic reticulum (SR),
a specialized form of ER. These proteins have central roles
in muscle calcium homeostasis, and improper modifica­
tions of these proteins resulting from stress will definitely
disrupt muscle functions. Thus, skeletal muscle is more
sensitive to, and more easily affected by, ER stress than
other tissue types. Upregulation of the UPR and NFκB
Hypoxia ER stress Muscle apoptosis pathways has been observed in muscle tissue from
and/or impaired autophagy
patients with IIM.5 Expression of E3 ubiquitin-protein
ligase RNF5, an ER-associated protein involved in the
Disturbed microcirculation, Continuous expression Accumulation of recognition and processing of misfolded proteins, was
trauma, infections, aging of cytokines abnormal proteins
and/or proteins elevated in patients with sIBM.84
A potential inducer of ER stress in myositis is the
Figure 3 | Potential nonimmune mechanisms in myositis pathogenesis. It is
thought that nonimmune mechanisms might affect the performance of muscle
upregula­tion of MHC class I that occurs in muscle fibers.4,5
fibers and weaken muscle in the absence of inflammatory cell infiltrates. Hypoxia, Normally, MHC class I expression is down­regulated when
potentially induced by capillary loss, can induce expression and extracellular muscle fibers undergo differentiation from myoblasts to
release of HMGB1 in endothelial cells and muscle fibers. HMGB1 might induce myotubes, and its level should remain low in differen­tiated
expression of MHC class I molecules, which in turn might cause ER stress, which myofibers, but it is overexpressed in myofibers of 75% of
can lead to expression of NFκB and perpetuation of inflammation. MHC class I patients with myositis.85 Furthermore, the UPR and NFκB
molecules might also reduce Ca2+ release from the sarcoplasmic reticulum, pathways are upregulated in a transgenic mouse model
inducing muscle fatigue. Substantial accumulation of abnormal proteins in
of myositis, with upregulation of MHC class I molecules
muscle fibers might induce ER stress and/or autophagy. If these unwanted
proteins cannot be cleared in an appropriate period then muscle apoptosis or in the muscle fibers occurring to a similar degree to that
autophagic cell death will occur, contributing to muscle fatigue. Abbreviations: observed in patients with myositis. 86 Other possible
ER, endoplasmic reticulum; HMGB1, high mobility group protein B1; NFκB, in­ducers of ER stress in myositis include viral infection and
nuclear factor κB. hypoxia. ER stress can result in NFκB activation, which
promotes transcription of cytokine genes and induces a
Hypoxia might induce muscle weakness by a number self-sustaining inflammatory response—this process can
of pathways. It can cause metabolic alterations in muscle, further contribute to muscle damage.5 Finally, ER stress
such as reduced levels of ATP and phosphocreatine, both might contribute to muscle dysfunction by inducing
of which have been reported in muscle tissue of patients muscle fiber apoptosis or autophagic cell death.87
with polymyositis or dermatomyositis.80 It might also
induce production of proinflammatory cytokines and Autophagy
adhesion molecules, such as IL‑1α, IL‑1β and transform­ Autophagy removes unwanted or damaged proteins from
ing growth factor β, in muscle tissue. Colocalization of cells; during this process, autophagosomes form to carry
HMGB1 and hypoxia can be observed in the lesions proteins destined for degradation to the lysosomes. If this
of mice with collagen-induced arthritis, and hypoxia- autophagosomal–lysosomal system cannot work nor­
induced extracellular release of HMGB1 occurs in vitro.81 mally, autophagic cell death will take place.88 Involvement
As HMGB1 can induce muscle fatigue, and is expressed of autophagy in myositis was proposed after the obser­
in muscle fibers of patients with myositis without muscle vation that muscle fiber death happens despite over­
inflammation,71,72 hypoxia might induce translocation of expression of anti-apoptotic molecules in patients with
HMGB1 from nuclei to cytoplasm of muscle fibers, and myositis.89 Autophagy seems to be particularly rele­vant for
thereby affect the fiber contractility; this hypothesis is sIBM, as abnormal intracellular accumulation of multi­
worth investigating further. protein inclusions in muscle fibers is a typical feature of
sIBM muscle. The origin and formation of these protein
ER stress inclusions are not well understood, but if they cannot be
ER stress is generated when unfolded or misfolded pro­ cleared within a proper period, they might become a detri­
teins accumulate within the organelle. Such accumula­ mental factor and influence normal muscle function. In
tion elicits the unfolded protein response (UPR) and support of this hypothesis is the finding that accumula­
the ER overload response (nuclear factor κB [NFκB] tion of amyloid causes a reduction in muscular peak force
pathway). The UPR has two primary functions: halting and lower calcium peak amplitude in mouse models of
protein translation through the PERK (pancreatic eIF2‑α sIBM.90 Moreover, autophagy-related protein LC3 B was
kinase, also known as eukaryotic translation initiation found in muscle fibers of patients with sIBM, but not in
factor 2‑α kinase 3) pathway, and activating signaling muscle biopsy samples of patients with polymyositis.91
pathways that lead to the production of molecular chap­ Disturbance of the autophagosomal–lysosomal system
erones involved in protein folding (through the serine/ in sIBM was further supported by the observation of
threonine-protein kinase/endoribonuclease IRE1 and decreased lysosomal proteolytic activity in muscle tissue

302  |  MAY 2011  |  VOLUME 7  www.nature.com/nrrheum


© 2011 Macmillan Publishers Limited. All rights reserved
REVIEWS

of patients with the disease. Interestingly, ER stress might are cyclosporin A, which has been found to be as effec­
induce lysosomal dysfunction in sIBM.92 tive as methotrexate,100 or mycophenylate mofetil, use of
which is supported by case reports.101–103 One of the few
Classification placebo-controlled trials to report beneficial outcomes
Currently, the most used classification criteria for IIM in treatment-resistant dermatomyositis used high-dose
(known as the Bohan and Peter 93 and Tanimoto94 crit­eria) intravenous immunoglobulin in combination with con­
are based on clinical, histopathological and neuro­physio­ ventional immunosuppressive agents for 3 months.104
logical findings, in combination with elevated serum levels How­ever, an open-label study that included treatment-
of muscle enzymes, but they do not incorpo­rate more resistant dermato­myositis, polymyositis and sIBM
recent mechanistic data that might help to distin­guish could not confirm this effect in any of the three myositis
subsets of patients with myositis in a clinically meaning­ subsets, in terms either of clinical outcome measures or of
ful fashion.93,94 In a more recent set of criteria, published in­flammatory changes in muscle tissue.
in 2004,95 MSAs and MRI of muscle tissues were included. A particular therapeutic challenge is presented by
Each of these criteria sets has limitations, how­ever; for patients with myositis and ILD, who have high morbidity
example, in the Bohan and Peter criteria the muscle and mortality. In this patient subset there are even fewer
biopsy criterion is not specified in detail, and does not evidence-based treatment recommendations than for
exclude other forms of myopathy such as muscular dys­ myositis generally. Several case reports suggest a bene­
trophy. Furthermore, the Bohan and Peter criteria, which, ficial effect of cyclophosphamide or tacrolimus on ILD.
with a publication date of 1975,93 are the oldest, might An effect of tacrolimus on both pulmonary and muscular
also misclassify patients with sIBM as having polymyo­ function was suggested from an open-label retrospective
sitis, because sIBM had not been described when these study of patients with antisynthetase syndrome,106 but this
criteria were devised. Indeed, several discoveries that have observation needs to be followed up by a controlled trial.
occurred since the 1980s could affect the methods used to
subtype myo­sitis, including assays for the MSAs discussed Future prospects
above and listed in Table 1, immunohistochemical charac­ The new biologic agents that are in use or under develop­
terization of muscle biopsies—including expression of ment for the treatment of rheumatic diseases permit
MHC class I molecules in muscle fibers—and sub­typing targeted therapy, and are thereby helpful for elucidating
of invading inflammatory cells.1,96,97 Similarly, MRI can key molecular pathways, as has been the case for TNF
be used to visualize muscle inflammation and might be bloc­kade in RA. Nevertheless, extensive experience of
helpful in the classification of IIM. To solve the unmet treatment with targeted therapies using biologic agents is
need for updated and validated classification cri­teria, lacking in myositis.
an inter­national, multidisciplinary, data-driven project The B‑cell-targeting agent rituximab is the biologic
is ongoing, within the framework of the International therapy for which most beneficial reports exist, provid­
Myositis Assessment and Clinical Research (IMACS) ing support for a role for B cells in at least a subset of
program,98 within the National Institute of Environmental patients with polymyositis or dermatomyositis.28–31,107 An
Health Sciences of the USA. international multicenter trial of rituximab in myositis
has recently concluded, but no data have been reported
Treatment to date.
Current practice Beneficial effects of IL‑1 blockade with another bio­
As we lack thorough understanding of the disease mecha­ logic agent, anakinra, were reported in one patient with
nisms that underlie myositis, traditional therapies are myo­sitis with anti-Jo‑1 autoantibodies.108 Further­more,
largely nonspecific, and are associated with the risks of in a pilot study that included 15 patients with refractory
generalized immunosuppression. There have only been myositis treated with anakinra, clinical outcome was
a few controlled trials in polymyositis and dermato­ improved in half of the patients,109 but these observations
myositis, and treatment recommendations are therefore need to be confirmed in a controlled trial. The results of
based mostly on clinical experience and open-label trials, TNF bloc­kade in myositis are conflicting. In one open-
as has been recently summarized.99 The introduction of label study of infliximab treatment, some patients experi­
glucocorticoids in the 1950s had a major impact on mor­ enced disease flares, and induction of type I IFN activity
bidity and mortality, and these drugs still form the basis of was recorded. These data argue against a key role for
treatment. High doses of glucocorticoids, 0.75–1.00 mg/kg TNF in chronic, treatment-resistant cases.110 In sIBM, the
per day, are often required for several weeks. Many effects of pharmacological treatment have been limited,
patients respond with improved function, but few recover and to date there is no evidence to suggest any benefit of
their former physical performance, and side effects are im­munosuppressive treatment in sIBM.
common. Many experts therefore recom­mend combina­ Thus, there is a clear need for new therapies, not only
tion therapy with another immuno­suppressive agent to for sIBM but also for polymyositis and dermatomyositis.
improve response and reduce the need for steroids. The Successful development of such therapies will be depen­
most frequently used first-line agents are methotrexate dent on improved knowledge of the molecular pathways
at 15–25 mg/week or azathioprine at 2 mg/kg per day. If that operate in these diseases. Current therapies are
these drugs are not tolerated, or have insufficient effect, directed against immune-mediated muscle dysfunction,
the most frequently recommended second-line agents but newer therapies will need to incorporate advances

NATURE REVIEWS | RHEUMATOLOGY VOLUME 7  |  MAY 2011  |  303


© 2011 Macmillan Publishers Limited. All rights reserved
REVIEWS

in understanding of nonimmune mechanisms (such as patients with IIM is autoantibody profiling, but this
ER stress and dysfunctional autophagy). In this context, approach needs to be tested in future trials. Studying the
recent data on the beneficial effects of adding physical outcomes of targeted therapies will facilitate understand­
exercise to conventional immunosuppressive treatment ing of disease mechanisms and improvement in classifi­
are encouraging.111–113 In a pilot study, resistance exercise cation criteria, which will, in turn, increase our ability to
led to improved clinical performance, and the expres­ provide patients with properly targeted therapy.
sion of genes involved in inflammation and fibrosis was
reduced in muscle tissue.79 This observation might indi­
Review criteria
cate that exercise in established polymyositis or dermato­
myositis has a beneficial effect on molecular expression in This Review was predominately based on articles
the target organ of myositis, but it requires replication published in the past 3–5 years, and found in PubMed
in larger studies. using the search terms “myositis AND (pathogenesis OR
classification OR treatment)”, “myositis AND immune
cells”, “myositis AND (T cell OR B cell OR macrophage OR
Conclusion dendritic cell)”, “myositis AND autoantibody”, “myositis
In summary, recent advances suggest that different AND (cytokines OR chemokines)”, “myositis AND (HMGB1
molecu­lar pathways predominate in different subsets of OR IL‑6 OR IL‑15 OR BAFF)”, “myositis AND hypoxia”,
myositis, and that targeted therapies consequently vary “myositis AND (ER OR endoplasmic reticulum)”, “myositis
in efficacy among patients with IIM. Our knowledge of AND autophagy”. Older publications found using “myositis
disease mechanisms needs to improve, which will allow AND (pathogenesis OR classification OR treatment)” were
us to develop validated classification criteria and reliable included. Some additional papers were selected from
reference lists of papers identified in PubMed.
prognostic biomarkers. One promising way to subtype

1. Dalakas, M. C. Polymyositis, dermatomyositis proliferation inducing ligand. Rheumatology 23. Banica, L. et al. Quantification and molecular
and inclusion-body myositis. N. Engl. J. Med. 325, (Oxford) 49, 1867–1877 (2010). characterization of regulatory T cells in
1487–1498 (1991). 13. Englund, P., Lindroos, E., Nennesmo, I., connective tissue diseases. Autoimmunity 42,
2. Dalakas, M. C. & Hohlfeld, R. Polymyositis and Klareskog, L. & Lundberg, I. E. Skeletal muscle 41–49 (2009).
dermatomyositis. Lancet 362, 971–982 (2003). fibers express major histocompatibility complex 24. Greenberg, S. A. et al. Plasma cells in muscle in
3. Plotz, P. H. et al. Current concepts in the class II antigens independently of inflammatory inclusion body myositis and polymyositis.
idiopathic inflammatory myopathies: infiltrates in inflammatory myopathies. Am. J. Neurology 65, 1782–1787 (2005).
polymyositis, dermatomyositis, and related Pathol. 159, 1263–1273 (2001). 25. De Bleecker, J. L., Engel, A. G. & Butcher, E. C.
disorders. Ann. Intern. Med. 111, 143–157 14. Behrens, L. et al. Human muscle cells express a Peripheral lymphoid tissue-like adhesion
(1989). functional costimulatory molecule distinct from molecule expression in nodular infiltrates in
4. Li, C. K. et al. Overexpression of MHC class I B7.1 (CD80) and B7.2 (CD86) in vitro and in inflammatory myopathies. Neuromuscul. Disord.
heavy chain protein in young skeletal muscle inflammatory lesions. J. Immunol. 161, 6, 255–260 (1996).
leads to severe myositis: implications for 5943–5951 (1998). 26. Bradshaw, E. M. et al. A local antigen-driven
juvenile myositis. Am. J. Pathol. 175, 1030–1040 15. Hofbauer, M. et al. Clonal tracking of humoral response is present in the inflammatory
(2009). autoaggressive T cells in polymyositis by myopathies. J. Immunol. 178, 547–556 (2007).
5. Nagaraju, K. et al. Activation of the endoplasmic combining laser microdissection, single-cell PCR, 27. Levine, T. D. Rituximab in the treatment of
reticulum stress response in autoimmune and CDR3-spectratype analysis. Proc. Natl Acad. dermatomyositis: an open-label pilot study.
myositis: potential role in muscle fiber damage Sci. USA 100, 4090–4095 (2003). Arthritis Rheum. 52, 601–607 (2005).
and dysfunction. Arthritis Rheum. 52, 16. Fasth, A. E. et al. T cell infiltrates in the muscles 28. Majmudar, S., Hall, H. A. & Zimmermann, B.
1824–1835 (2005). of patients with dermatomyositis and Treatment of adult inflammatory myositis with
6. Askanas, V. & Engel, W. K. Inclusion-body polymyositis are dominated by CD28null T cells. rituximab: an emerging therapy for refractory
myositis: a myodegenerative conformational J. Immunol. 183, 4792–4799 (2009). patients. J. Clin. Rheumatol. 15, 338–340
disorder associated with Aβ, protein misfolding, 17. Pandya, J. M. et al. Expanded TCR-Vβ restricted (2009).
and proteasome inhibition. Neurology 66, T cells from sporadic inclusion body myositis 29. Rios Fernandez, R., Callejas Rubio, J. L.,
S39–S48 (2006). patients are proinflammatory and cytotoxic Sanchez Cano, D., Saez Moreno, J. A. &
7. Greenberg, S. A. How citation distortions create CD28(null) T cells. Arthritis Rheum. 62, Ortego Centeno, N. Rituximab in the treatment
unfounded authority: analysis of a citation 3457–3466 (2010). of dermatomyositis and other inflammatory
network. BMJ 339, b2680 (2009). 18. Fasth, A. E., Bjorkstrom, N. K., Anthoni, M., myopathies. A report of 4 cases and review of
8. Christopher-Stine, L. et al. A novel autoantibody Malmberg, K. J. & Malmstrom, V. Activating the literature. Clin. Exp. Rheumatol. 27,
recognizing 200‑kd and 100‑kd proteins is NK‑cell receptors co-stimulate CD4(+)CD28(–) 1009–1016 (2009).
associated with an immune-mediated necrotizing T cells in patients with rheumatoid arthritis. Eur. 30. Tournadre, A. et al. Polymyositis and pemphigus
myopathy. Arthritis Rheum. 62, 2757–2766 J. Immunol. 40, 378–387 (2010). vulgaris in a patient: successful treatment with
(2010). 19. Bettelli, E., Korn, T., Oukka, M. & Kuchroo, V. K. rituximab. Joint Bone Spine 75, 728–729
9. Hengstman, G. J. et al. Anti-signal recognition Induction and effector functions of T(H)17 cells. (2008).
particle autoantibodies: marker of a necrotising Nature 453, 1051–1057 (2008). 31. Vandenbroucke, E. et al. Rituximab in life
myopathy. Ann. Rheum. Dis. 65, 1635–1638 20. Page, G. et al. Plasma cell-like morphology of threatening antisynthetase syndrome.
(2006). Th1‑cytokine‑producing cells associated with the Rheumatol. Int. 29, 1499–1502 (2009).
10. de Padilla, C. M. & Reed, A. M. Dendritic cells loss of CD3 expression. Am. J. Pathol. 164, 32. Page, G., Chevrel, G. & Miossec, P. Anatomic
and the immunopathogenesis of idiopathic 409–417 (2004). localization of immature and mature dendritic
inflammatory myopathies. Curr. Opin. Rheumatol. 21. Chevrel, G. et al. Interleukin‑17 increases the cell subsets in dermatomyositis and
20, 669–674 (2008). effects of IL‑1β on muscle cells: arguments for polymyositis: interaction with chemokines and
11. Reed, A. M. & Ernste, F. The inflammatory milieu the role of T cells in the pathogenesis of Th1 cytokine-producing cells. Arthritis Rheum.
in idiopathic inflammatory myositis. Curr. myositis. J. Neuroimmunol. 137, 125–133 50, 199–208 (2004).
Rheumatol. Rep. 11, 295–301 (2009). (2003). 33. Eloranta, M. L. et al. A possible mechanism for
12. Szodoray, P. et al. Idiopathic inflammatory 22. Waschbisch, A., Schwab, N., Ruck, T., endogenous activation of the type I interferon
myopathies, signified by distinctive peripheral Stenner, M. P. & Wiendl, H. FOXP3+ T regulatory system in myositis patients with anti‑Jo‑1 or
cytokines, chemokines and the TNF family cells in idiopathic inflammatory myopathies. anti-Ro 52/anti-Ro 60 autoantibodies. Arthritis
members B‑cell activating factor and a J. Neuroimmunol. 225, 137–142 (2010). Rheum. 56, 3112–3124 (2007).

304  |  MAY 2011  |  VOLUME 7  www.nature.com/nrrheum


© 2011 Macmillan Publishers Limited. All rights reserved
REVIEWS

34. Greenberg, S. A. et al. Interferon α/β-mediated muscle and lung inflammation. J. Autoimmun. 71. Ulfgren, A. K. et al. Down-regulation of the
innate immune mechanisms in dermatomyositis. 29, 174–186 (2007). aberrant expression of the inflammation
Ann. Neurol. 57, 664–678 (2005). 54. Lovgren, T. et al. Induction of interferon-α by mediator high mobility group box chromosomal
35. McNiff, J. M. & Kaplan, D. H. Plasmacytoid immune complexes or liposomes containing protein 1 in muscle tissue of patients with
dendritic cells are present in cutaneous systemic lupus erythematosus autoantigen- and polymyositis and dermatomyositis treated with
dermatomyositis lesions in a pattern distinct Sjögren’s syndrome autoantigen-associated corticosteroids. Arthritis Rheum. 50,
from lupus erythematosus. J. Cutan. Pathol. 35, RNA. Arthritis Rheum. 54, 1917–1927 (2006). 1586–1594 (2004).
452–456 (2008). 55. Barbasso Helmers, S. et al. Sera from anti‑Jo‑1- 72. Grundtman, C. et al. Effects of HMGB1 on in vitro
36. Civatte, M. et al. Expression of the β chemokines positive patients with polymyositis and responses of isolated muscle fibers and
CCL3, CCL4, CCL5 and their receptors in interstitial lung disease induce expression of functional aspects in skeletal muscles of
idiopathic inflammatory myopathies. intercellular adhesion molecule 1 in human lung idiopathic inflammatory myopathies. FASEB J.
Neuropathol. Appl. Neurobiol. 31, 70–79 (2005). endothelial cells. Arthritis Rheum. 60, 24, 570–578 (2010).
37. Confalonieri, P. et al. Increased expression of 2524–2530 (2009). 73. Alexanderson, H. et al. Patients with idiopathic
β-chemokines in muscle of patients with 56. Grundtman, C. et al. Immunolocalization of inflammatory myopathies have low muscle
inflammatory myopathies. J. Neuropathol. Exp. interleukin‑1 receptors in the sarcolemma and endurance rather than low muscle strength
Neurol. 59, 164–169 (2000). nuclei of skeletal muscle in patients with [abstract 823]. Arthritis Rheum. 60 (Suppl.),
38. De Paepe, B. & De Bleecker, J. L. β-Chemokine idiopathic inflammatory myopathies. Arthritis S307 (2009).
receptor expression in idiopathic inflammatory Rheum. 56, 674–687 (2007). 74. Emslie-Smith, A. M. & Engel, A. G. Microvascular
myopathies. Muscle Nerve 31, 621–627 (2005). 57. Lundberg, I., Ulfgren, A. K., Nyberg, P., changes in early and advanced dermatomyositis:
39. Rostasy, K. M. et al. Monocyte/macrophage Andersson, U. & Klareskog, L. Cytokine a quantitative study. Ann. Neurol. 27, 343–356
differentiation in dermatomyositis and production in muscle tissue of patients with (1990).
polymyositis. Muscle Nerve 30, 225–230 (2004). idiopathic inflammatory myopathies. Arthritis 75. Estruch, R. et al. Microvascular changes in
40. Stone, K. B. et al. Anti‑Jo‑1 antibody levels Rheum. 40, 865–874 (1997). skeletal muscle in idiopathic inflammatory
correlate with disease activity in idiopathic 58. St. Pierre, B. A., Kasper, C. E. & Lindsey, A. M. myopathy. Hum. Pathol. 23, 888–895 (1992).
inflammatory myopathy. Arthritis Rheum. 56, Fatigue mechanisms in patients with cancer: 76. Grundtman, C., Tham, E., Ulfgren, A. K. &
3125–3131 (2007). effects of tumor necrosis factor and exercise on Lundberg, I. E. Vascular endothelial growth
41. Brouwer, R. et al. Autoantibody profiles in the skeletal muscle. Oncol. Nurs. Forum 19, factor is highly expressed in muscle tissue of
sera of European patients with myositis. Ann. 419–425 (1992). patients with polymyositis and patients with
Rheum. Dis. 60, 116–123 (2001). 59. Tateyama, M. et al. Expression of tumor necrosis dermatomyositis. Arthritis Rheum. 58,
42. Mahouachi, R. et al. Antisynthetase syndrome factor-α in muscles of polymyositis. J. Neurol. Sci. 3224–3238 (2008).
[French]. Tunis Med. 86, 195–196 (2008). 146, 45–51 (1997). 77. Probst-Cousin, S., Neundorfer, B. & Heuss, D.
43. Jordan Greco, A. S., Métrailler, J. C. & Dayer, E. 60. Krystufkova, O. et al. Increased serum levels of Microvasculopathic neuromuscular diseases:
The antisynthetase syndrome: a cause of B cell activating factor (BAFF) in subsets of lessons from hypoxia-inducible factors.
rapidly progressive interstitial lung disease. patients with idiopathic inflammatory Neuromuscul. Disord. 20, 192–197 (2010).
Rev. Med. Suisse 3, 2675–2676, 2679–2681 myopathies. Ann. Rheum. Dis. 68, 836–843 78. Dastmalchi, M. et al. Effect of physical training
(2007). (2009). on the proportion of slow-twitch type I muscle
44. Gomard-Mennesson, E. et al. Clinical 61. Walsh, R. J. et al. Type I interferon-inducible gene fibers, a novel nonimmune-mediated mechanism
significance of anti‑histidyl‑tRNA synthetase expression in blood is present and reflects for muscle impairment in polymyositis or
(Jo1) autoantibodies. Ann. NY Acad. Sci. 1109, disease activity in dermatomyositis and dermatomyositis. Arthritis Rheum. 57,
414–420 (2007). polymyositis. Arthritis Rheum. 56, 3784–3792 1303–1310 (2007).
45. Genth, E. Inflammatory muscle diseases: (2007). 79. Nader, G. A. et al. A longitudinal, integrated,
dermatomyositis, polymyositis, and inclusion 62. Serrano, A. L., Baeza-Raja, B., Perdiguero, E., clinical, histological and mRNA profiling study of
body myositis [German]. Internist (Berl.) 46, Jardi, M. & Munoz-Canoves, P. Interleukin‑6 is an resistance exercise in myositis. Mol. Med. 16,
1218–1232 (2005). essential regulator of satellite cell-mediated 455–464 (2010).
46. Legout, L. et al. The antisynthetase syndrome: skeletal muscle hypertrophy. Cell. Metab. 7, 80. Park, J. H. et al. Use of magnetic resonance
a subgroup of inflammatory myopathies not to 33–44 (2008). imaging and P‑31 magnetic resonance
be unrecognized [French]. Rev. Med. Interne 23, 63. Scuderi, F., Mannella, F., Marino, M., spectroscopy to detect and quantify muscle
273–282 (2002). Provenzano, C. & Bartoccioni, E. IL‑6‑deficient dysfunction in the amyopathic and myopathic
47. Mammen, A. L. et al. Expression of the mice show impaired inflammatory response in a variants of dermatomyositis. Arthritis Rheum.
dermatomyositis autoantigen Mi‑2 in model of myosin-induced experimental myositis. 38, 68–77 (1995).
regenerating muscle. Arthritis Rheum. 60, J. Neuroimmunol 176, 9–15 (2006). 81. Hamada, T. et al. Extracellular high mobility group
3784–3793 (2009). 64. Sugiura, T. et al. Increased IL‑15 production of box chromosomal protein 1 is a coupling factor
48. Kaji, K. et al. Identification of a novel autoantibody muscle cells in polymyositis and for hypoxia and inflammation in arthritis. Arthritis
reactive with 155 and 140 kDa nuclear proteins in dermatomyositis. Int. Immunol. 14, 917–924 Rheum. 58, 2675–2685 (2008).
patients with dermatomyositis: an association (2002). 82. Kim, I., Xu, W. & Reed, J. C. Cell death and
with malignancy. Rheumatology (Oxford) 46, 65. Suzuki, J. et al. Serum levels of interleukin 15 in endoplasmic reticulum stress: disease
25–28 (2007). patients with rheumatic diseases. J. Rheumatol. relevance and therapeutic opportunities. Nat.
49. Targoff, I. N. et al. A novel autoantibody to a 28, 2389–2391 (2001). Rev. Drug Discov. 7, 1013–1030 (2008).
155‑kd protein is associated with 66. Chiu, W. K., Fann, M. & Weng, N. P. Generation 83. Hosoi, T. & Ozawa, K. Endoplasmic reticulum
dermatomyositis. Arthritis Rheum. 54, and growth of CD28nullCD8+ memory T cells stress in disease: mechanisms and therapeutic
3682–3689 (2006). mediated by IL‑15 and its induced cytokines. opportunities. Clin. Sci. (Lond.) 118, 19–29
50. Sato, S. et al. RNA helicase encoded by J. Immunol. 177, 7802–7810 (2006). (2010).
melanoma differentiation-associated gene 5 is 67. Furmanczyk, P. S. & Quinn, L. S. Interleukin‑15 84. Delaunay, A. et al. The ER‑bound RING finger
a major autoantigen in patients with clinically increases myosin accretion in human skeletal protein 5 (RNF5/RMA1) causes degenerative
amyopathic dermatomyositis: association with myogenic cultures. Cell Biol. Int. 27, 845–851 myopathy in transgenic mice and is deregulated
rapidly progressive interstitial lung disease. (2003). in inclusion body myositis. PLoS ONE 3, e1609
Arthritis Rheum. 60, 2193–2200 (2009). 68. Andersson, U. & Harris, H. E. The role of HMGB1 (2008).
51. Levine, S. M. et al. Novel conformation of in the pathogenesis of rheumatic disease. 85. van der Pas, J., Hengstman, G. J., ter Laak, H. J.,
histidyl-transfer RNA synthetase in the lung: Biochim. Biophys. Acta 1799, 141–148 (2010). Borm, G. F. & van Engelen, B. G. Diagnostic
the target tissue in Jo‑1 autoantibody-associated 69. Kokkola, R. et al. Successful treatment of value of MHC class I staining in idiopathic
myositis. Arthritis Rheum. 56, 2729–2739 collagen-induced arthritis in mice and rats by inflammatory myopathies. J. Neurol. Neurosurg.
(2007). targeting extracellular high mobility group box Psychiatry 75, 136–139 (2004).
52. Soejima, M. et al. Role of innate immunity in a chromosomal protein 1 activity. Arthritis Rheum. 86. Salomonsson, S. et al. Upregulation of MHC
model of histidyl-tRNA synthetase (Jo‑1)- 48, 2052–2058 (2003). class I in transgenic mice results in reduced
mediated myositis. Arthritis Rheum. (2010). 70. Jiang, W. & Pisetsky, D. S. Mechanisms of force-generating capacity in slow-twitch muscle.
53. Katsumata, Y. et al. Species-specific immune disease: the role of high-mobility group protein 1 Muscle Nerve 39, 674–682 (2009).
responses generated by histidyl-tRNA in the pathogenesis of inflammatory arthritis. 87. Henriques-Pons, A. & Nagaraju, K. Nonimmune
synthetase immunization are associated with Nat. Clin. Pract. Rheumatol. 3, 52–58 (2007). mechanisms of muscle damage in myositis:

NATURE REVIEWS | RHEUMATOLOGY VOLUME 7  |  MAY 2011  |  305


© 2011 Macmillan Publishers Limited. All rights reserved
REVIEWS

role of the endoplasmic reticulum stress cell-mediated cytotoxicity in some diseases, and tissue of patients with inflammatory myopathies.
response and autophagy in the disease implications for the pathogenesis of the different Ann. Rheum. Dis. 66, 1276–1283 (2007).
pathogenesis. Curr. Opin. Rheumatol. 21, inflammatory myopathies. Hum. Pathol. 17, 106. Wilkes, M. R., Sereika, S. M., Fertig, N.,
581–587 (2009). 704–721 (1986). Lucas, M. R. & Oddis, C. V. Treatment of
88. Askanas, V., Engel, W. K. & Nogalska, A. 97. Arahata, K. & Engel, A. G. Monoclonal antibody antisynthetase-associated interstitial lung
Inclusion body myositis: a degenerative muscle analysis of mononuclear cells in myopathies. disease with tacrolimus. Arthritis Rheum. 52,
disease associated with intra-muscle fiber multi- III: Immunoelectron microscopy aspects of cell- 2439–2446 (2005).
protein aggregates, proteasome inhibition, mediated muscle fiber injury. Ann. Neurol. 19, 107. Haroon, M. & Devlin, J. Rituximab as a first-line
endoplasmic reticulum stress and decreased 112–125 (1986). agent for the treatment of dermatomyositis.
lysosomal degradation. Brain Pathol. 19, 98 National Institue of Environmental Health Rheumatol. Int. doi:10.1007/S00296-010-
493–506 (2009). Sciences. International Myositis Assessment and 1458-6.
89. Nagaraju, K. et al. The inhibition of apoptosis in Clinical Research (IMACS) [online], http:// 108. Furlan, A., Botsios, C., Ruffatti, A., Todesco, S. &
myositis and in normal muscle cells. J. Immunol. www.niehs.nih.gov/news/events/pastmtg/ Punzi, L. Antisynthetase syndrome with
164, 5459–5465 (2000). imacs/index.cfm (2011). refractory polyarthritis and fever successfully
90. Shtifman, A. et al. Amyloid-β protein impairs 99. Dalakas, M. C. Immunotherapy of myositis: treated with the IL‑1 receptor antagonist,
Ca(2+) release and contractility in skeletal issues, concerns and future prospects. Nat. Rev. anakinra: A case report. Joint Bone Spine 75,
muscle. Neurobiol. Aging 31, 2080–2090 Rheumatol. 6, 129–137 (2010). 366–367 (2008).
(2010). 100. Vencovsky, J. et al. Cyclosporine A versus 109. Dorph, C. et al. Anakinra in patients with
91. Temiz, P., Weihl, C. C. & Pestronk, A. methotrexate in the treatment of polymyositis refractory idiopathic inflammatory myopathies
Inflammatory myopathies with mitochondrial and dermatomyositis. Scand. J. Rheumatol. 29, [abstract 589]. Arthritis Rheum. 60 (Suppl.)
pathology and protein aggregates. J. Neurol. Sci. 95–102 (2000). S218 (2009).
278, 25–29 (2009). 101. Caramaschi, P., Volpe, A., Carletto, A., 110. Dastmalchi, M. et al. A high incidence of disease
92. Nogalska, A., D’Agostino, C., Terracciano, C., Bambara, L. M. & Biasi, D. Long-standing flares in an open pilot study of infliximab in
Engel, W. K. & Askanas, V. Impaired autophagy in refractory polymyositis responding to patients with refractory inflammatory
sporadic inclusion-body myositis and in mycophenolate mofetil: a case report and review myopathies. Ann. Rheum. Dis. 67, 1670–1677
endoplasmic reticulum stress-provoked cultured of the literature. Clin. Rheumatol. 26, (2008).
human muscle fibers. Am. J. Pathol. 177, 1795–1796 (2007). 111. Alexanderson, H. Exercise effects in patients
1377–1387 (2010). 102. Gelber, A. C., Nousari, H. C. & Wigley, F. M. with adult idiopathic inflammatory myopathies.
93. Bohan, A. & Peter, J. B. Polymyositis and Mycophenolate mofetil in the treatment of severe Curr. Opin. Rheumatol. 21, 158–163 (2009).
dermatomyositis (first of two parts). N. Engl. J. skin manifestations of dermatomyositis: a series 112. Alexanderson, H., Dastmalchi, M., Esbjornsson-
Med. 292, 344–347 (1975). of 4 cases. J. Rheumatol 27, 1542–1545 (2000). Liljedahl, M., Opava, C. H. & Lundberg, I. E.
94. Tanimoto, K. et al. Classification criteria for 103. López de la Osa, A., Sánchez Tapia, C., Benefits of intensive resistance training in
polymyositis and dermatomyositis. J. Rheumatol. Arias Díaz, M. & Terrancle de Juan, I. patients with chronic polymyositis or
22, 668–674 (1995). Antisynthetase syndrome with good response to dermatomyositis. Arthritis Rheum. 57, 768–777
95. Hoogendijk, J. E. et al. 119th ENMC international mycophenolate mofetil [Spanish]. Rev. Clin. Esp. (2007).
workshop: trial design in adult idiopathic 207, 269–270 (2007). 113. de Salles Painelli, V. et al. The possible role of
inflammatory myopathies, with the exception of 104. Dalakas, M. C. et al. A controlled trial of high- physical exercise on the treatment of idiopathic
inclusion body myositis, 10–12 October 2003, dose intravenous immune globulin infusions as inflammatory myopathies. Autoimmun. Rev. 8,
Naarden, The Netherlands. Neuromuscul. Disord. treatment for dermatomyositis. N. Engl. J. Med. 355–359 (2009).
14, 337–345 (2004). 329, 1993–2000 (1993).
96. Engel, A. G. & Arahata, K. Mononuclear cells 105. Barbasso Helmers, S. et al. Limited effects of Author contributions
in myopathies: quantitation of functionally high-dose intravenous immunoglobulin (IVIG) M. Zong and I. E. Lundberg contributed equally to all
distinct subsets, recognition of antigen-specific treatment on molecular expression in muscle aspects of the preparation of this manuscript.

306  |  MAY 2011  |  VOLUME 7  www.nature.com/nrrheum


© 2011 Macmillan Publishers Limited. All rights reserved

You might also like