ArticlePDF AvailableLiterature Review

Intracellular B Lymphocyte Signalling and the Regulation of Humoral Immunity and Autoimmunity

Authors:

Abstract and Figures

B lymphocytes are critical for effective immunity; they produce antibodies and cytokines, present antigens to T lymphocytes and regulate immune responses. However, because of the inherent randomness in the process of generating their vast repertoire of antigen-specific receptors, B cells can also cause diseases through recognizing and reacting to self. Therefore, B lymphocyte selection and responses require tight regulation at multiple levels and at all stages of their development and activation to avoid diseases. Indeed, newly generated B lymphocytes undergo rigorous tolerance mechanisms in the bone marrow and, subsequently, in the periphery after their migration. Furthermore, activation of mature B cells is regulated through controlled expression of co-stimulatory receptors and intracellular signalling thresholds. All these regulatory events determine whether and how B lymphocytes respond to antigens, by undergoing apoptosis or proliferation. However, defects that alter regulated co-stimulatory receptor expression or intracellular signalling thresholds can lead to diseases. For example, autoimmune diseases can result from altered regulation of B cell responses leading to the emergence of high-affinity autoreactive B cells, autoantibody production and tissue damage. The exact cause(s) of defective B cell responses in autoimmune diseases remains unknown. However, there is evidence that defects or mutations in genes that encode individual intracellular signalling proteins lead to autoimmune diseases, thus confirming that defects in intracellular pathways mediate autoimmune diseases. This review provides a synopsis of current knowledge of signalling proteins and pathways that regulate B lymphocyte responses and how defects in these could promote autoimmune diseases. Most of the evidence comes from studies of mouse models of disease and from genetically engineered mice. Some, however, also come from studying B lymphocytes from patients and from genome-wide association studies. Defining proteins and signalling pathways that underpin atypical B cell response in diseases will help in understanding disease mechanisms and provide new therapeutic avenues for precision therapy.
Pathways of B cell development and differentiation. B cells are generated from haematopoietic progenitor cells in the bone marrow. This process involves the expression of B lineage cell-specific proteins and the rearrangement of mini antibody V(D)J genes to generate the BCR repertoire. During the pro-B cell stage, antibody heavy chains are first generated by randomly rearranging and combining V, D and J mini genes. Pre-B cells express the pre-B cell antigen receptor (BCR) on the cell surface with the fully arranged heavy chain associated with the surrogate light chain (red). At later stages, light chain V and J mini genes are rearranged and a complete BCR is expressed in association with the Ig-α and Ig-β (green) subunits of the BCR complex. Immature B cells then undergo tolerance mechanisms with B cells recognizing self-protein undergoing light chain editing, apoptosis or functional inactivation (anergy). Surviving immature B cells then exit the bone marrow and migrate to secondary lymphoid organs where they develop into transitional (T) B cells. Transitional B cells can be subdivided into a number of developmental subsets. These include T1 B cells that express a high level of IgM and T2 B cells that express both IgM and IgD. These B cells undergo a range of tolerance checkpoint and cells that recognize self-antigens with high affinity are deleted. Cells with intermediate/low affinity to self-antigens and those that do not recognize self survive and circulate for about 3 weeks to survey the body for their target antigens. Transitional B cells develop into either marginal zone (MZ) B cells or follicular B cells. MZ B cells sample antigens and those that recognize antigens expand independently of T cell help. For their expansion, MZ B cells require TLR signalling to into short-lived plasma cells that produce antibodies with limited avidity for their target antigens. Follicular B cells are activated when they encounter their target antigens in the presence T cell help. Activated follicular B cells then migrate to B cell follicles and initiate somatic maturation in germinal centres. During this process, the cells proliferate, acquire somatic mutations, produce antibodies with higher avidity and class switch to IgG. Antigen-specific mature B cells then leave germinal centres and differentiate into either plasma cells or memory B cells. Plasma cells can either remain secondary lymphoid organs or travel to bone marrow to produce antibodies. B1 cells comprise a distinct subset of B cells that develop in the bone marrow and migrate to the periphery (peritoneal and pleural cavities in mice). B1 cells produce polyreactive IgM antibodies and partake in providing a first line of immunity against pathogens
… 
This content is subject to copyright. Terms and conditions apply.
Intracellular B Lymphocyte Signalling and the Regulation
of Humoral Immunity and Autoimmunity
Taher E. Taher
1
&Jonas Bystrom
1
&Voon H. Ong
2
&David A. Isenberg
3
&
Yves Renaudineau
4
&David J. Abraham
2
&Rizgar A. Mageed
1
Published online: 29 April 2017
#The Author(s) 2017. This article is an open access publication
Abstract B lymphocytes are critical for effective immunity;
they produce antibodies and cytokines, present antigens to T
lymphocytes and regulate immune responses. However, be-
cause of the inherent randomness in the process of generating
their vast repertoire of antigen-specific receptors, B cells can
also cause diseases through recognizing and reacting to self.
Therefore, B lymphocyte selection and responses require tight
regulation at multiple levels and at all stages of their develop-
ment and activation to avoid diseases. Indeed, newly generat-
ed B lymphocytes undergo rigorous tolerance mechanisms in
the bone marrow and, subsequently, in the periphery after their
migration. Furthermore, activation of mature B cells is regu-
lated through controlled expression of co-stimulatory recep-
tors and intracellular signalling thresholds. All these regulato-
ry events determine whether and how B lymphocytes respond
to antigens, by undergoing apoptosis or proliferation.
However, defects that alter regulated co-stimulatory receptor
expression or intracellular signalling thresholds can lead to
diseases. For example, autoimmune diseases can result from
altered regulation of B cell responsesleading to the emergence
of high-affinity autoreactive B cells, autoantibody production
and tissue damage. The exact cause(s) of defective B cell
responses in autoimmune diseases remains unknown.
However, there is evidence that defects or mutations in genes
that encode individual intracellular signalling proteins lead to
autoimmune diseases, thus confirming that defects in intracel-
lular pathways mediate autoimmune diseases. This review
provides a synopsis of current knowledge of signalling pro-
teins and pathways that regulate B lymphocyte responses and
how defects in these could promote autoimmune diseases.
Most of the evidence comes from studies of mouse models
of disease and from genetically engineered mice. Some, how-
ever, also come from studying B lymphocytes from patients
and from genome-wide association studies. Defining proteins
and signalling pathways that underpin atypical B cell response
in diseases will help in understanding disease mechanisms and
provide new therapeutic avenues for precision therapy.
Keywords Blymphocytes .Intracellular signalling .
Autoimmune diseases
Introduction
Autoimmune diseases are pathological conditions in which
defects in immunological tolerance to self lead to the initiation
of effector immunity to self, chronic inflammation and tissue
and organ damage. These diseases affect about 510% of
human populations worldwide and cause significant degrees
of morbidity and early death [1]. The cause of most autoim-
mune diseases remains largely unknown. However, suscepti-
bility to develop these diseases is associated with a combina-
tion of genetic, environmental and hormonal factors [2]. These
factors combine to cause defects in the survival and selection
of self-reactive T and B lymphocytes. Evidence from the last
50 years of research indicates that T lymphocytes initiate au-
toimmune responses in conjunction with, or following
*Rizgar A. Mageed
r.a.mageed@qmul.ac.uk
1
Centre for Experimental Medicine and Rheumatology, William
Harvey Research Institute, Queen Mary University of London,
Charterhouse Square, London EC1M 6BQ, UK
2
Centre for Rheumatology and Connective Tissue Diseases, Royal
Free Hospital, University College London, London, UK
3
Centre for Rheumatology, University College London, London, UK
4
Immunology Laboratory, University of Brest Medical School,
Brest, France
Clinic Rev Allerg Immunol (2017) 53:237264
DOI 10.1007/s12016-017-8609-4
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
incitement by, B lymphocytes. In addition to autoantibody
production, there is compelling evidence that B lymphocytes
also contribute to the development of the autoimmune dis-
eases through mechanisms such as autoantigen presentation
to activate autoreactive T cells and/or promote their polariza-
tion to produce disease-promoting/perpetuating cytokines. In
this respect, it is perhaps revealing that a significant proportion
of genetic susceptibility risk factors to develop autoimmune
diseases corresponds with defects in the regulation of B cell
responses, intracellular signalling and tolerance induction.
These observations highlight changing perceptions about the
role played by B cells in autoimmune diseases (Fig. 1;
Tab les 1and 2). The importance of these roles is supported
by the therapeutic benefit gained from depleting B cells in
patients with a range of autoimmune diseases. For example,
patients with diseases such rheumatoid arthritis (RA) [71],
type 1 diabetes (T1D) [72], anti-neutrophil cytoplasmic
antibody (ANCA) vasculitis [73], multiple sclerosis (MS)
[74], systemic sclerosis (SSc) [75,76], primary Sjögrenssyn-
drome [7780] and systemic lupus erythematosus (SLE)
[8183] benefit from therapeutic depletion of B cells. Of note
in this respect is that B cell-depleting therapy has a clinical
benefit without significantly affecting autoantibody levels,
suggesting that, perhaps, other B cell functions, including an-
tigen presentation and cytokine production could be critical
aspects of B cell involvement in the pathogenesis of the auto-
immune diseases.
The need for, and the ability to generate, a vast B cell
repertoire to combat a universe of pathogens requires
tolerance checkpoints and exquisite fine-tuning of B cell
receptor (BCR) signalling to limit the emergence of path-
ogenic autoreactive B cells. Highly coordinated and inte-
grated intracellular signalling transduced through the
BCR and other co-stimulatory receptors, including innate
Fig. 1 Signalling molecules and pathways in regulating B cell selection
and responses. The diagram illustrates major signalling proteins/
pathways involved in B cell physiology and whose regulation has been
reported to be altered/defective in B cells in autoimmune disease. Proteins
indicated in yellow are kinases, red for phosphatases, pink for proteins
involved in ubiquitination, black for transcription factors and brown for
adaptor proteins. Arrows indicate proteins that promote positive
signalling, while blunt-ended lines indicate the protein negatively
regulate signalling. Minus signs (encircled) indicate proteins/signalling
pathways are reduced in mice and/or patients with autoimmune diseases
or that reduction by genetic engineering promotes B lymphocyte
hyperactivity and autoimmune disease. Positive signs (encircled)
indicate enhanced activity of the proteins/signalling pathways in B
lymphocytes from patients with autoimmune disease, mouse models or
that their genetic manipulation promotes autoimmunity
238 Clinic Rev Allerg Immunol (2017) 53:237264
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Tab le 1 Reported defects in signalling molecules and pathways and their impact on B cell responses and association with diseases
Signalling
molecule
Encoding
gene
Effect on B cell response and disease in animal models Association with human diseases
Lyn LYN B cell hyperactivity causing lupus-like disease in gene-deficient mice [3] Reduced cellular expression leading to IgG autoantibody and cytokine production [4]
SHP-1 PTPN6 Selective deficiency in B cells promotes systemic autoimmune disease [5] Reduced cellular expression and SNP association with SLE [6,7]
LYP PTPN22 Expression of the R619W variant in B cells causes systemic autoimmunity [8] The R619W is a risk allele in several systemic autoimmune diseases [9,10]
CD45 CD45 Mutation in the inhibitory wedge causes autoantibody production leading to
severe glomerulonephritis [11]
Decreased expression, increased translocation signalling domains and altered isoform
expression associated with SLE [12]
BTK BTK Over expression increases plasma cell numbers, spontaneous germinal centre
formation, autoantibody production and lupus-like disease [13,14]
Gene defect causes X-linked agammaglobulinemia, reduced B cell numbers and deficiency in
all immunoglobulin isotypes [15]. Increased phosphorylation in SLE B cells [16]
CD22 CD22 Deficiency causes autoantibody production and lupus-like disease [17,18] Splicing defect causes expression of a truncated CD22 expression and increased leukemic B
cell precursors [19]
CD19 CD19 Altered expression correlates with autoimmune diseases [20,21] Increased expression in patients with systemic sclerosis; polymorphism is associated with
susceptibility to SLE [22,23]
FCγRIIB FCγRIIB Deficiency causes SLE-like autoimmune disease and renders non-permissive
H2B mouse strain susceptible to collagen-induced arthritis (CIA) [24,25]
Decreased expression in SLE [26]
SHIP-1 INPP5D B cell-specific deficiency causes lupus-like disease [27] Hypophosphorylated in SLE patients [28]
PTEN PTEN B cell-specific deficiency causes hyperresponsiveness and anti-ssDNA autoanti-
body production [29,30]
Decreased expression in SLE patients [31]
PTP1B PTPN1 B cell-specific deficiency causes systemic autoimmunity in aged mice [32] Reduced expression in RA patients [32]
Act1 TRAF3IP2 Deficient mice develop Sjögrens syndrome-like disease [33] Susceptibility gene in psoriatic arthritis and SLE and SNP associated with RA [3436]
A20 TNFAIP3 B cell-specific deficiency causes systemic autoimmunity [3739] SNPs associated with SLE and RA [40,41]
Cbl CBL B cell-specific deficiency of c-Cbl and Cbl-b causes systemic autoimmunity [42] SNP associated with SLE and type 1 diabetes [43,44]
WASP WA S B cell-specific deficiency causes systemic autoimmune disease [45] About 40% of Wiskott-Aldrich syndrome patients develop autoimmunity [46]
The table summarizes data on reported clinical and immunological phenotypes in engineered mice gene deficient/mutated for signalling proteins. The table also provides some of the reported data on defects
in the expression or function of the corresponding protein in patients
Clinic Rev Allerg Immunol (2017) 53:237264 239
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
pattern recognition receptors such as Toll-like receptors
(TLRs), costimulatory/inhibitory molecules and cytokine
receptors, are essential for regulating the outcome of
BCR engagement by antigens. The available evidence
indicates that minimal alterations in established thresh-
olds of activating or inhibiting intracellular signalling
can lead to a breakdown of immunological tolerance.
This review provides a synopsis of current knowledge
of signalling molecules and pathways involved in medi-
ating and regulating B cell responses and how changes
could lead to aggressive self-reactivity and autoimmune
diseases.
Tabl e 2 Polymorphisms and mutations in genes encoding co-receptors, signalling proteins, transcription factors and cytokines/chemokines that are
associated with human diseases
Gene Chromosome Disease association Protein Function in B cells Reference
PTPN22 1p13.2 RA, SLE, GT, T1D LYP Lymphocyte-specific tyrosine phosphatase
a
[47]
NCF2 1q25 SLE p67phox Subcomponent of NADPH oxidase, ROS generation
a
[48]
IL10 1q31-q32 SLE, UC, T1D IL-10 Anti-inflammatory cytokine
a
[49]
PTPRC 1q31.3-q32.1 SLE, RA, MS, T1D CD45 Membrane protein tyrosine kinases [50]
FCGR2A 1q23.2 SLE, RA IGFR2 Low-affinity IgG FC receptor
a
[51,52]
RASGRP3 2p25.1-p24.1 SLE GRP3 Signalling downstream of the BCR
a
[53]
BANK1 4q24 SLE, SSc, RA BANK1 Scaffold protein involved in BCR signalling [54]
IL21 4q27 SLE, PSO, CEL IL-21 Cytokine, class switch recombination, plasma cell
differentiation
a
[55]
BACH2 6q15 SLE, AS, ATD, CEL,
CD, MS, T1D, IBD, PSC
BACH2 Negative regulator of transcription
a
[50]
PRDM1-ATG5 6q21 SLE, RA, CD Blimp1 Differentiation and development of plasma cells
a
[49]
IKZF1 7p12.2 SLE, CD Ikaros TF, differentiation, development, self-tolerance
a
[53]
BLK 8p23-p22 SLE, SS, RA, SSc, pAPS BLK Tyrosine kinase, BCR signalling, development [56]
LYN 8q12 SLE Lyn Tyrosine protein kinase, BCR signalling [4,57]
CCL21 9q13.3 RA CCL21 Chemokine, germinal centre formation [58]
ETS1 11q23.3 SLE Ets1 TF, negative regulator of differentiation
a
[59]
CXCR5 11q23.3 SS CXCR5 Chemokine receptor, migration to B cell follicles
a
[60]
SLC15A4 12q24.32 SLE PTR4 Proton-coupled amino-acid transporter located in
endolysosomes, autoantibody production
a
[53,61]
ELF1 13q13 SLE Elf1 TF, binding the IgH enhancer
a
[62]
CSK 15q24.1 SLE Csk Increases BCR-mediated activation of mature B cells
a
[63]
ITGAM 16p11.2 SLE CD11B Regulation of BCR signalling
a
[64]
IRF8 16q24.1 SLE IRF8 TF, cell development
a
[65]
IKZF3 17q21 SLE Aiolos TF, downregulation of the pre-BCR
a
[65]
CD40 20q13.12 RA CD40 Co-stimulatory molecule, promotes antibody production [58]
IKBKE 1q32.1 SLE IKKI Phosphorylates IκBα
a
[50]
TNIP1 5q33.1 SLE, SS, PS NAF1 TNFAIP3 interacting protein
a
[49,60]
TNFAIP3 6q23 SLE, SS, RA, T1D
UC, CEL, PSO
A20 Ubiquitination and negative signalling regulator ubiquitin
editing enzyme
a
[58,60]
PRKCB 16p11.2 SLE PRKCB1 Member of the PKC family, BCR-dependent NF-κB
activation
a
[66]
UBE2L3 22q11.21 SLE, CD, RA, CEL UBE2L3 Ubiquinase, NFkB activation, plasmablast and plasma cell
development
a
[67]
IRAK1/MECP2 Xq28 SLE, RA Irak1 TACI-dependent Ig class switching via MyD88
a
[68,69]
REL 2p16.1 RA Rel Survival and proliferation
a
[70]
TRAF1 9q33.1 RA Traf1 CD40 and TLR signalling
a
[58]
The table lists polymorphic risk loci associated with the development of autoimmune diseases. The data are generated in GWAS and genes cited include
those that encode proteins with known functions in B lymphocytes
RA rheumatoid arthritis, SLE systemic lupus erythematosus, GT Graves thyroiditis, T1D type 1 diabetes, CEL coeliac disease, MS multiple sclerosis, CD
Crohns disease, PSO psoriasis, UC ulcerative colitis, AS ankylosing spondylitis, ATD autoimmune thyroid disease, JIA juvenile idiopathic arthritis, AA
alopecia areata, IBD inflammatory bowel disease, PSC primary sclerosing cholangitis, SS Sjögrens syndrome, SSc systemic sclerosis, TF transcription
factor, BCR B cell receptor
a
Not specific for B cells
240 Clinic Rev Allerg Immunol (2017) 53:237264
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Signals Controlling B Cell Development
and Functions
The BCR repertoire for antigens is vast, generated through
random recombination of germline V(D)J mini genes, to pro-
vide broad immunity against pathogens. However, an intrinsic
feature of generating this vast repertoire is the randomness
with which germline V(D)J mini genes are recombined. This
leads, in up to 80% of newly generated B cells, to the
generation of BCRs that recognize self (Fig. 2). There is,
therefore, a necessity for emerging B cells to undergo toler-
ance in the bone marrow and also subsequently in the periph-
ery for B cells that escape bone marrow tolerance or those that
emerge as a result of mutations in secondary lymphoid organs.
Newly generated B cells first encounter self-antigens in the
bone marrow, and their elimination or survival depends to a
great extent on the strength with which their BCRs bind self-
antigens and strength of the resulting intracellular signalling.
Fig. 2 Pathways of B cell development and differentiation. B cells are
generated from haematopoietic progenitor cells in the bone marrow. This
process involves the expression of B lineage cell-specific proteins and the
rearrangement of mini antibody V(D)J genes to generate the BCR
repertoire. During the pro-B cell stage, antibody heavy chains are first
generated by randomly rearranging and combining V, D and J mini genes.
Pre-B cells express the pre-B cell antigen receptor (BCR) on the cell
surface with the fully arranged heavy chain associated with the
surrogate light chain (red). At later stages, light chain V and J mini
genes are rearranged and a complete BCR is expressed in association
with the Ig-αand Ig-β(green) subunits of the BCR complex. Immature
B cells then undergo tolerance mechanisms with B cells recognizing
self-protein undergoing light chain editing, apoptosis or functional
inactivation (anergy). Surviving immature B cells then exit the bone
marrow and migrate to secondary lymphoid organs where they develop
into transitional (T) B cells. Transitional B cells can be subdivided into a
number of developmentalsubsets. These include T1 B cells that express a
high level of IgM and T2 B cells that express both IgM and IgD. These B
cells undergo a range of tolerance checkpoint and cells that recognize
self-antigens with high affinity are deleted. Cells with intermediate/low
affinity to self-antigens and those that do not recognize self survive and
circulate for about 3 weeks to survey the body for their target antigens.
Transitional B cells develop into either marginal zone (MZ) B cells or
follicular B cells. MZ B cells sample antigens and those that recognize
antigens expand independently of T cell help. For their expansion, MZ B
cells require TLR signalling to into short-lived plasma cells that produce
antibodies with limited avidity for their target antigens. Follicular B cells
are activated when they encounter their target antigens in the presence T
cell help. Activated follicular B cells then migrate to B cell follicles and
initiate somatic maturation in germinal centres. During this process, the
cells proliferate, acquire somatic mutations, produce antibodies with
higher avidity and class switch to IgG. Antigen-specific mature B cells
then leave germinal centres and differentiate into either plasma cells or
memory B cells. Plasma cells can either remain secondary lymphoid
organs or travel to bone marrow to produce antibodies. B1 cells comprise
a distinct subset of B cells that develop in the bone marrow and migrate to
the periphery (peritoneal and pleural cavities in mice). B1 cells produce
polyreactive IgM antibodies and partake in providing a first line of
immunity against pathogens
Clinic Rev Allerg Immunol (2017) 53:237264 241
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Thus, recognition of self with high affinity initiates strong in-
tracellular signalling, and as a result, B cells undergo receptor
editing to replace nascent light chains that endow self-reactivity
or, if this fails, apoptosis or anergy [8487]. To facilitate B cell
tolerance yet provide effective B cell immunity, intracellular
signalling is regulated by highly refined thresholds. These
thresholds regulate the magnitude and duration of intracellular
B cell signalling and the outcome of BCR engagement by
antigens. Subsequent to undergoing tolerance in the bone mar-
row, immature B cells migrate to the periphery where they will
need tonic signalling for their transition to full maturity in read-
iness to respond to antigens. Signalling thresholds are also set
for co-stimulatory receptors that modulate B cell responses
following antigen recognition. These co-stimulatory receptors
include CD40, TLR, BAFF and receptors for a range of other
cytokines. Most of these signals involve the activation of phos-
phatidylinositol 3 kinase (PI3K) and the canonical pathway of
NF-κB activation [88,89].
BCR-Mediated Signalling
Proximal BCR Signalling Initiates B Cell Development
and Responses
The BCR associates with a heterodimer of signalling proteins,
Ig-αand Ig-β(also known as CD79a and CD79b), to form the
BCR complex. The cytoplasmic domains of both Ig-αand
Ig-βhave immunoreceptor tyrosine-based activation motifs
(ITAMs). ITAMs initiate signalling when their tyrosine resi-
dues are phosphorylated following BCR engagement and
translocation to lipid raft signalling domains that contain the
Src family tyrosine kinase Lyn. Downstream, activation of
ITAMs generates a docking site for spleen tyrosine kinase
(Syk) recruitment and phosphorylation [89,90]. The activa-
tion of Syk is fundamental for initiating signalling cascades
leading to lymphocyte activation [9092]. Substrates of acti-
vated tyrosine kinases are adaptor molecules that, in turn,
recruit other kinases to the BCR complex. B cell linker protein
(BLNK) is a Syk substrate with nine tyrosine residues that are
rapidly phosphorylated following engagement of the BCR
[92,93]. Phosphorylated BLNK is recruited to the plasma
membrane and this requires its association with CIN85. The
BLNK-CIN85 complex then coordinates the recruitment of
growth factor receptor-bound protein 2 (Grb2) and
phosphoinositide phospholipase C gamma (PLCγ)[94], a
process essential for B cell development and responses.
Defective Regulation of BCR-Mediated Signalling Leads
to Aberrant B Cell Responses and Autoimmune Diseases
Src Family Tyrosine Kinase Lyn Lyn is a key dual activity
kinase. It initiates BCR-mediated signalling by phosphorylat-
ing Ig-α/Ig-βITAMs but then regulates this signalling by
phosphorylating immunoreceptor tyrosine-based inhibition
motifs (ITIMs) in CD5, CD22 and FcγRIIB [95]. Thus,
Lyn-deficient mice develop spontaneous lupus-like autoim-
mune disease, splenomegaly and glomerulonephritis and pro-
duce anti-dsDNA autoantibodies [96,97]. In addition, BCR-
mediated calcium (Ca
2+
) influx is enhanced in B lymphocytes
in Lyn
/
, mice and there is accelerated class switching of anti-
dsDNA and anti-RNA autoantibodies [97]. Interestingly,
however, deletion of myeloid differentiation primary response
gene 88 (MyD88) in Lyn
/
mice, both globally or selectively
in B lymphocytes, suppresses B cell activation and class
switching of autoantibodies and ameliorates lupus disease
[98]. This finding suggests that aberrant B cell responses in
Lyn
/
mice are likely to be influenced not only by BCR-
mediated signalling bust also by signalling through TLRs.
In humans, there is evidence for reduced Lyn expression in
B lymphocytes from patients with SLE and that this reduction
impacts B cell responses. For example, B cells from Lyn-
insufficient SLE patients produce IgG autoantibodies to
dsDNA and disease-promoting cytokines in vitro [4]. The
association between Lyn insufficiency and SLE is supported
by genetic studies. Thus, single nucleotide polymorphism
(SNP) analyses and genome-wide association studies have
revealed that polymorphisms in LYN, as well as other Src
family tyrosine kinases including BLK, are risk factors for
susceptibility to SLE [56,99].
CD45 Tyrosine Phosphatase CD45 is a membrane protein
tyrosine phosphatase that positively regulates Lyn activation
by dephosphorylating a tyrosine residue at position 507
(Y-507). This causes a conformational change that exposes
the catalytic domain of Lyn and promotes autophosphoryla-
tion of the positive regulatory tyrosine at position 396 (Y-396)
[100]. In addition to Lyn, CD45 regulates the activation of
other kinases, such as Janus kinases (JAKs) and, thus, influ-
ences cytokine signalling [101], Src kinases involved in cell
adhesion [102], TLR signalling [103] and apoptosis [104].
Dysregulation of CD45, therefore, can affect multiple B lym-
phocyte functions leading to autoimmune-like diseases, but
the precise impact of changes of each of the multiple pathways
that CD45 regulates in promoting autoimmune disease re-
mains unclear. In genetically engineered mice, a single nucle-
otide replacement in the dimerization wedge of the CD45
molecule was shown to lead to autoantibody production and
the development of lupus-like disease [11]. However, it is
established that CD45 also influences apoptosis and defects
in its expression have been shown to promote lupus disease in
Fas ligand-mutant (Fasl
gld/gld
) mice. In this setting, reduced
CD45 expression enhanced B lymphocyte hyperactivity and
auto-Ab production [105]. Furthermore, defects in CD45 reg-
ulation has been shown to affect B lymphocyte tolerance. For
example, CD45
/
mice and CD45
/
B cell lines show re-
duced CD22 activation, SHP-1 recruitment, increased Syk
242 Clinic Rev Allerg Immunol (2017) 53:237264
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
activation [106]andCa
2+
influx [107]. Of note, however, is
that loss of function mutations in sialic acid acetyl esterase
(SIAE), which is required for the inhibitory function of
CD22, has been shown to also create a significant risk for
developing RA, T1D and SLE [108]. In hen egg lysozyme
(HEL) transgenic mice, HEL induced tolerance in mature
CD45
+/+
B lymphocytes but led to the activation and accumu-
lation of long-lived CD45
/
HEL-reactive B lymphocytes
[109].
Studies of B lymphocytes in patients with SLE in our lab-
oratory revealed that Lyn insufficiency was associated with
increased CD45 translocation to lipid raft signalling domains
and, ultimately, to reduced cellular expression of this phospha-
tase [12]. The noted increase in the translocation of CD45 to
lipid raft signalling domains is likely to be relevant to reduced
Lyn expression since CD45 promotes Lyn activation, ulti-
mately its degradation in the proteasome [4].
Downstream Kinases Defects in the regulation of BCR-
associated signalling molecules downstream of Lyn have also
been reported and shown to promote aberrant B cell responses
and autoimmune diseases. For example, defective regulation
of the adaptor protein B cell adaptor protein with ankyrin
repeats 1 (BANK1) which initiates BCR-mediated Ca
2+
sig-
nalling after Lyn-mediated phosphorylation of inositol 1,4,5-
trisphosphate receptor (IP3R) causes autoimmune-like disease
in mice. Thus, while BCR-mediated Ca
2+
influx was shown to
be normal in Bank1
/
primary B cells, this deficiency led to
enhanced CD40-mediated proliferation, survival, increased
Akt activation, and enhanced T-dependent antibody produc-
tion and formation of germinal centres [110]. In humans, ge-
netic studies have revealed that two variants of BANK1,
R61H and A383T, are strongly associated with susceptibility
to SLE [54]. The molecular basis for this association, howev-
er, remains to be determined. Nevertheless, the increase in
CD40-mediated Akt activation in Bank1
/
B cells suggests
that the allelic variants may promote autoimmunity through
affecting cognate B-T cell interactions.
More recent studies of B cells from patients with SLE car-
ried out in our laboratory revealed that the extent of defects in
intercellular signalling is more complex and extensive than
previously thought with each of the many identified defects
likely to impact different B cell responses and clinical symp-
toms differently [28]. For example, these studies identified
defective regulation of PI3K, MAPK, cyclin-dependent ki-
nase1 (CDK1) and PKC to varying degrees in B cells from
patients with SLE compared with matched healthy controls.
These studies also revealed that the activity of Rho, a serine/
threonine kinase involved in cell motility, was reduced in B
cells from patients with SLE. Although as stated above, the
relevance of many of these defects remains to be determined,
it is likely that reduced activity of Rho can lead to defective
migration of B lymphocytes. In addition to the above defects,
reduced activity of the cell cycle kinase ATR was noted in the
SLE patients. ATR is involved in activating the DNA damage
response pathway, which leads either to cell cycle arrest or
apoptosis and is, therefore, a key checkpoint in regulating cell
responses to DNA damage.
Protein Tyrosine Phosphatases In addition to kinases that
positively regulate BCR-mediated signalling, defects in phos-
phatases that control the activation of kinases downstream of
Lyn have also been reported to be involved in promoting
aberrant B cell responses in autoimmune diseases. For exam-
ple, defects in LYP tyrosine phosphatase, which is encoded by
the protein tyrosine phosphatase non-receptor 22 (PTPN22),
were shown to be sufficient to promote systemic autoimmu-
nity. GWAS studies also revealed that a SNP in PTPN22,
1858C/T that resulted in R620W amino acid substitution is
associated with increased risk of SLE, TID and RA
[111113]. Interestingly, expression of the R619W LYP vari-
ant in B cells alone was shown to be sufficient to develop
splenomegaly, spontaneous germinal centre formation, glo-
merulonephritis and anti-dsDNA autoantibody production [8].
Calcium and Diacylglycerol Signalling
In Transcriptional Activation and B Cell Survival
The recruitment of PLCγto the BCR signalling complex fol-
lowing engagement by antigens initiates phosphatidylinositol
4,5 biphosphate (PIP2) hydrolysis leading to the generation of
inositol 1,4,5-triphosphate (IP3) and diacylglycerol (DAG)
[94,114]. DAG binds to the cysteine-rich domain of Ras/
Rap guanyl-releasing protein and this activates rat sarcoma
(Ras) and ras-related protein (Rap) GTPases, the serine/
threonine protein kinase C (PKC) and protein kinase D
(PKD). IP3, in contrast, binds to IP3 receptors on the endo-
plasmic reticulum (ER) to release Ca
2+
from its stores and,
thus, increase cytosolic Ca
2+
concentration [114]. The deple-
tion of Ca
2+
stores in the ER is sensed by stromal interaction
molecules 1 and 2 (STIM1 and STIM2). As a result, these
proteins relocate to the ER-plasma membrane junction where
they bind to the Ca
2+
-release activated channel (CRAC) pro-
tein Orail and/or canonical transient receptor potential 1
(TRPC1) channels allowing extracellular Ca
2+
entry to in-
crease its intracellular level. As a consequence, the sustained
increase in intracellular Ca
2+
triggers the activation of the
Ca
2+
/calmodulin-dependent protein kinase kinases
(CaMKKs), serine threonine kinases involved in the regula-
tion of important cellular processes such as survival and cyto-
skeletal reorganization [115]. The BCR-induced increase in
intracellular Ca
2+
levels also activates calcineurin (also known
as protein phosphatase 2B, PP2B), a protein phosphatase that
controls intracellular localization of nuclear factor of activated
T cells (NFAT) family of transcription factors [116,117]. In
Clinic Rev Allerg Immunol (2017) 53:237264 243
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
resting B cells, NFATs are constitutively phosphorylated by
casein kinase 1 and glycogen synthase kinase-3 (GSK3) and
are sequestered in the cytosol as a result of binding to 14-3-3
proteins. The BCR-induced activation of calcineurin leads to
dephosphorylation of NFATs, thus permitting their transloca-
tion to the nucleus. In the nucleus, NFATs form complexes
with other transcription factors to regulate the transcription of
target genes including IL-2, IL-4, IL-10, tumour-necrosis fac-
tor alpha (TNFα) and interferon gamma (IFNγ)[117,118].
Defective Regulation of Ca
2+
Signalling Promotes Defective
B Cell Tolerance
Numerous studies have examined molecular mechanisms
leading to aberrant Ca
2+
signalling in B lymphocytes with
special emphasis on how BCR and co-receptors CD19 and
CD21 mediate PLCγ2-IP3-Ca
2+
signalling. CD21-mediated
Ca
2+
signalling plays an important role in breaching B lym-
phocyte tolerance leading to autoantibody production [119].
Furthermore, changes in the regulation of Ca
2+
signalling are
recognized as a major signalling event that contributes to the
loss of B lymphocyte tolerance. Noteworthy, and perhaps par-
adoxically, is that an elevation in the baseline level of Ca
2+
and reduced BCR-mediated elevation have been noted in
tolerized/anergic B cells [84,120]. Elevated baseline level of
Ca
2+
is likely to be due to persistent but low level engagement
of the BCR by self-antigens and a recognized characteristic of
anergic B cells in experimental models [84,121]. In response
to antigen engagement, naïve B cells show a rapid increase in
intracellular Ca
2+
followed by a drop to reach a plateau within
minutes. This plateau is similar to the basal level seen in
anergic B cells and continues as long as the BCR is engaged
by the antigen. Thus, anergic B cells represent the physiolog-
ical equivalent of chronically antigen-stimulated naïve B cells.
Modulating BCR-mediated mechanisms of Ca
2+
signalling
could, therefore, provide a potential therapeutic approach for
treating autoimmune diseases. Indeed, treatment of B cells
with 1,4-benzodiazepine Bz-423 which increases sensitivity
to BCR engagement by causing sustained high level of Ca
2+
promotes apoptosis [122]. Since hyperactivation and altered
Ca
2+
signalling are distinguishing features of autoreactive B
cells, treatment with Bz-423 has been suggested to be a useful
approach for eliminating autoreactive B cells in autoimmune
diseases.
Phosphatidylinositol 3 Kinase Signalling
PI3K Signalling in B Cell Development, Survival
and Activation
PI3K signalling is important for B cell development, survival
and activation. PI3Ks represent a family of lipid and protein
kinases that function mainly through phosphorylation of
phosphoinositide [123,124]. Based on molecular structure
and functions, PI3Ks are divided into four classes: I, II, III
and IV. Members of class I PI3Ks are the ones whose altered
activation is implicated in autoimmunity and inflammation.
This class of PI3Ks is subdivided into two distinct subgroups,
IA and IB. In mammals, the IA subgroup includes three mem-
bers: PI3Kα,PI3Kβand PI3Kδ[125]. All three kinase mem-
bers of the IA subgroup are heterodimers consisting of p110
catalytic subunits (p110α,p110βand p110δ)andaregulatory
subunit, usually referred to as p85 [125]. Subgroup IB, in
contrast, consists of one catalytic subunit, p110γ,associated
with either a p101 or p84 regulatory subunit [123]. These
different PI3Ks function in different signalling pathways in
lymphocytes with p110δexpression been restricted to
haematopoietic cells. Upon receptor activation, PI3Ks phos-
phorylate PIP2 leading to the production of PIP3 [126]. The
production of PIP3 requires recruitment of PI3Ks to the plas-
ma membrane either through binding of the SH2 domain of
their regulatory units to the phosphorylated tyrosine residues
in receptor signalling complex domains and adaptors, or
through direct recruitment by Ras. Signalling through PI3K
is negatively regulated by the lipid phosphatase, SH2 domain-
containing inositol phosphatase (SHIP) and phosphatase and
tensin homologue deleted on chromosome ten (PTEN) [127].
Upon engagement of the BCR, CD19 recruits PI3K to the
plasma membrane through binding of p85 to its tyrosine-
phosphorylated cytoplasmic domain. In B lymphocytes, the
B cell activating factor (BAFF) and low basal signalling by
un-engaged BCR maintains low PIP3 levels [87]. The level of
PIP3 increases dramatically following BCR engagement by
antigens and co-stimulation through CD19, IL-4 receptor
and/or TLRs. The recruitment and binding of the key down-
stream target of PI3K, Akt [also known as protein kinase B
(PKB)], to the PIP3 through its pleckstrin homology (PH)
domain causes conformational changes to Akt and, as a result,
permits PIP3-dependent kinase 1 (PDK1)-mediated phos-
phorylation of Akt at threonine 304 within its catalytic do-
main. PDK1 has a PH domain that binds PIP3 and promotes
its translocation to the plasma membrane to co-localize with
Akt [128]. Once activated, Akt phosphorylates important
downstream targets including Rheb GAP TSC2, FOX1/3
and Fox4A. Akt-induced phosphorylation of Ras homologue
enriched in brain (Rheb) GAP TSC2 that leads to the accumu-
lation of Rheb-GTP complex results in the activation of mam-
malian target of rapamycin complex 1 (mTORc1) [129]. The
Faxo family of transcription factors is active and located in the
nucleus in resting cells; however, when phosphorylated by
Akt, they translocate to the cytosol where their transcriptional
activities are terminated. Akt is, therefore, important for me-
tabolism and cell survival in peripheral B lymphocytes [87].
Additionally, Akt/Foxo pathway plays a critical role in regu-
lating the expression of recombinase activating genes (RAGs)
that are responsible for antigen-receptor rearrangement in B
244 Clinic Rev Allerg Immunol (2017) 53:237264
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
cells [130]. When Akt is inactive in quiescent B lymphocytes,
Foxo1, Foxo3 and Foxo4A drive transcription of genes
encoding IL-7, an essential homeostatic cytokine for lympho-
cytes, as well as for Kruppel-like factor 2 (KLF2) transcription
factor [131]. KLF2 directly regulates the expression of adhe-
sion molecules and chemokine receptors responsible for con-
trolling B lymphocyte entry into and exit from secondary
lymphoid organs.
Dysregulated PI3K Signalling Alters Normal B Cell
Development and Differentiation to Plasma Cells
in Autoimmune Diseases
The involvement of aberrant PI3K signalling in the pathogen-
esis of autoimmune diseases is intriguing as all leukocytes
express all members of class I PI3Ks. However, evidence for
the involvement of defective regulation of PI3K signalling has
mainly emerged from studying pathways involving PI3Kγ
and PI3Kδas these two class I PI3Ks are exclusively
expressed in immune cells. In contrast to PI3Kαand PI3Kβ
where ablation of their genes leads to embryonic lethality
[132], PI3Kγ- and PI3Kδ-deficient mice are viable but are
immunodeficient [133137]. Furthermore, enhanced activity
of either PI3Kγor PI3Kδhas been implicated in promoting
autoimmunity [138140]. In murine models of lupus and in
SLE patients, the activity of PI3K is increased [141]. The
exact cause(s) and impact of enhanced PI3K activity on SLE
and autoimmune diseases in general remains to be determined.
However, PI3K promotes B cell survival and the generation of
short-lived plasma cells and suppresses class switch recombi-
nation through activating Akt, which, in turn, represses Foxo
transcription factors [142,143]. Of note, is that PI3Kδis the
main PI3K family member that is involved in regulating B
lymphocyte responses. Thus, mice lacking PI3Kδshow re-
duced development of pro-B to pre-B cells in the bone marrow
and impaired responses of mature B cells [136,144].
Additionally, PI3Kδis involved in regulating marginal zone
(MZ) and B-1 B cell responses including antibody production
[145]. Interestingly, B cell development in PI3Kγ
/
δ
/
mice
is similar to PI3Kδ
/
mice, whereas no defects are seen in B
cell development in PI3Kγ
/
mice [146]. These observations
indicate that PI3Kγdoes not play a notable role in B cell
development. Indeed, genetically engineered mice expressing
a catalytically inactive PI3Kδmanifest impaired BCR signal-
ling and reduced IgM and IgG antibody production [144].
Similarly, heterozygous deletion of PI3Kδdiminishes autoan-
tibody production, ameliorates nephritis and improves surviv-
al in Lyn-deficient mice that develop lupus-like disease. In
contrast, mice expressing constitutively active PI3Kδshow a
reduced ability to eliminate autoreactive B lymphocytes [140].
In addition to its involvement in regulating BCR-mediated
signalling, PI3Kδis involved in mediating inflammation trig-
gered by the engagement of TLRs [147]. These observations
suggest that targeting of PI3Kδcould be an attractive thera-
peutic option for treating patients with autoimmune diseases
and chronic inflammation [148150]. Of note in this respect is
that studies using mouse models of lupus have shown that
inhibiting PI3K blocked glomerulonephritisand extended sur-
vival [139].
In addition to direct evidence for the role of dysregulated
PI3K signalling in promoting autoimmune diseases in mice,
there is indirect evidence for its involvement in promoting
disease in patients. For example, there is evidence for de-
creased expression of PTEN, a lipid phosphatase that nega-
tively regulates PI3K signalling in B cell subsets, except in
memory B cells, in patients with SLE [31]. Furthermore, the
level of PTEN in B cells from patients with SLE is inversely
related to disease activity. Decreased levels of PTEN also
concur with the upregulation of microRNA (miR-7) that
downregulates PTEN expression. These findings suggest that
defective miR-7 regulation of PTEN could contribute to B cell
hyperresponsiveness in SLE [31]. Functional screening of a
microRNA library also revealed that another miR, miR-148a,
is a potent regulator of B cell tolerance [151]. Furthermore,
increased expression levels of miR-148a were reported in pa-
tients with lupus and also in lupus-prone mice [151]. Elevated
miR-148a levels impair B cell tolerance through enhancing
the survival of immature B cells following BCR engagement
by self-antigens [151]. Molecular studies revealed that miR-
148a functions by suppressing the expression of Gadd45α,
PTEN and the pro-apoptotic protein Bim. Furthermore, in-
creased expression of miR-148a leads to lethal autoimmune
disease in a mouse model of lupus [151]. Using adoptive
transfer of anergic B cells, a recent study revealed that contin-
uous signalling through the inhibitory molecules SHP-1 and
SHIP-1 was required to maintain B cell anergy. Furthermore,
reducing signalling through either of these two signalling
pathways leads to rapid B cell activation, proliferation and
the generation of short-lived plasma cells [152].
Ubiquitination-Regulated Signalling
Ubiquitination Regulation of BCR-Mediated Signalling
and Antigen Processing
Ubiquitination is an important posttranslational modification
process that regulates signal transduction through covalent
attachment of ubiquitin (Ub) moieties, a 76-amino acid pep-
tide, to targeted proteins. The process involves at least three
enzymes, Ub-activating enzyme (E1) that activates Ub, Ub-
conjugating enzyme (E2) and Ub ligase (E3). E3 enzymes,
such as Cbl, catalyse ligation of the C-terminal residue of
Ub to a lysine residue on the target protein [153]. Lysine
residues K6, K11, K27, K29, K33, K48 and K63 of Ub can
potentially form seven different types of linkages in branched
poly Ub chains, whereas a linear form of the Ub chain can be
Clinic Rev Allerg Immunol (2017) 53:237264 245
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
formed when only one lysine in each Ub in involved in link-
age formation [153,154]. Mono ubiquitination promotes
endocytic trafficking and DNA repair and the K48-linked
Ub moieties tag proteins for degradation via the proteasomal
system. In contrast, K63-linked and linearly linked Ub chains
provide docking sites for downstream effectors and promote
protein-protein interactions and signalling [155].
Ubiquitination can also be regulated through deubiquitinating
enzymes, proteases that remove mono-ubiquitins and poly-
ubiquitins from proteins. In this regard, A20 acts as a
deubiquitinating as well as an ubiquitin-editing enzyme.
A20 inhibits the activation of NF-κB. It also restricts apopto-
sis induced by TNFα[156]. The following section will review
data on two key effector enzymes involved in the
ubiquitination cycle, Cbl and A20, since there is an abundance
of evidence for their involvement in autoimmunity.
In mammals, the Casitas B lineage lymphoma (Cbl) family
of proteins has three members: c-Cbl, Cbl-b, and Cbl-3. c-Cbl
and Cbl-b are expressed in B cells [157] and function as prom-
inent substrates for tyrosine phosphorylation and regulators of
the threshold of signalling [157160]. c-Cbl effectively in-
hibits B cell responses through downregulating Syk kinase
[161]. c-Cbl and Cbl-b interact with several BCR-associated
signalling molecules such as PLCγ2, BLNK, PI3 kinase, Lyn,
Va v a n d S y k [ 42,162,163]. Subsequent to binding to ITAMs,
Syk is phosphorylated on tyrosine 323 and this creates a bind-
ing site for c-Cbl [164]. c-Cbl recruits components of the
ubiquitin conjugation pathway and acts as an ubiquitin ligase
[165]. Binding of c-Cbl results in Syk ubiquitination and
downregulation of BCR signalling [164]. Apart from regulat-
ing BCR signalling, c-Cbl mediates BCR ubiquitination, a
process crucial for facilitating antigen processing and presen-
tation by B cells through the internalization of antigen-BCR
complexes and guiding them to multi-vesicular body-like
MIIC. In these multi-vesicular body-like MIICs, antigen-
BCR complexes are processed into peptides and loaded onto
MHC class II for presentation to T cells [166169]. The re-
cruitment of Cbl-b to clustered BCRs is also required for the
entry of endocytosed BCRs into late endosomes. Recruitment
of Cbl-b is also required for the entry of TLR9 into endosomes
as has been noted after in vitro activation of TLR9 by BCR-
captured antigens [170].
In contrast to Cbl, A20 is a widely expressed cytoplasmic
protein that inhibits NF-κB activation and signalling down-
stream of interleukin-1 receptor (IL-1R), TNF receptor 1
(TNFR1), CD40 as well as signalling through innate-type re-
ceptors such as TLRs and NOD-like receptors (NLRs)
[171174]. In addition, A20 promotes cell survival through
which it can regulate immune responses [174]. By
destabilizing E2 enzymes, A20 can disrupt the interaction be-
tween E2 and E3 and, therefore, restrict ubiquitination of tar-
get proteins [175]. To achieve its critical biochemical func-
tions, A20 interacts with key effectors including the receptor
interacting kinase-1 (RIPK1), a key player in inflammation
and cell death, E2, E3, ABIN-1 (ubiquitin sensors) and
NEMO/IKKγ,akeyplayerinNF-κBsignalling[
176181].
Additionally, A20 binds directly to ubiquitin chains [177,179]
and modifies ubiquitinated protein substrates in multiple
ways. For example, A20 cleaves poly-ubiquitin chains, there-
by, exhibiting a deubiquitinating activity. In addition, A20
works with E1 and E2 proteins to build ubiquitin chains, thus
displaying E3-like activity [171,182]. Through its Ub-editing
functions, A20 also removes K63-linked poly-ubiquitin
chains from substrates and builds K48-linked ubiquitin chains
[182].
Altered Ubiquitination in Defective in B Cell Tolerance
Inappropriate ubiquitination has been associated with the de-
velopment of autoimmune diseases. A large body of evidence
implicates defects in the level and regulation of Cbl and A20
in the pathogenesis of autoimmune diseases. Thus, Cbl-b-
deficient mice develop autoimmune diseases and highlight a
connection between Cbl-b-mediated protein degradation and
the regulation of BCR signalling thresholds [158]. These mice
produce high levels of autoantibodies to double-stranded
DNA and develop signs of spontaneous lupus-like disease
[158]. Another study revealed that Cbl-b-deficient mice had
an enhanced susceptibility to develop experimental autoim-
mune encephalitis (EAE) [183]. B cells from Cbl-b-deficient
mice showed an enhanced ability to proliferate in response to
BCR and CD40 engagements [158]. The lowering of BCR
thresholds caused by the loss of Cbl-b correlated with in-
creased susceptibility to develop autoimmune disease.
Many signalling proteins associate with Cbl-b, including
PLCγ, PI3K, Syk and the adaptor proteins Slp-76 and Vav.
However, Cbl-b-deficient cells have a selective enhancement
of Vav phosphorylation, indicating that Cbl-b is a negative
regulator of Vav phosphorylation. Vav is a key guanine nucle-
otide exchange factor for the Rho family of GTP-binding pro-
teins [184], and mice with B cell-specific ablation of c-Cbl
and Cbl-b manifest lupus-like disease and have a significant
increase in MZ and B1 B cell numbers [184]. Interestingly,
however, c-Cbl/Cbl-b-deficient B cells were not hyperrespon-
sive to BCR engagement, did not proliferate extensively nor
produced antibodies but tolerance induction was impaired
[42]. Apart from attenuated BLNK phosphorylation, these
mutant B cells showed enhanced phosphorylation of BCR-
proximal signalling proteins including Syk, PLCγ-2 and Vav
and increased Ca
2+
mobilization. These results, therefore, in-
dicate that Cbl proteins regulate B cell tolerance possibly
through fine-tuning of BCR-mediated signalling thresholds
[42].
In contrast to Cbl proteins, as cited earlier, A20 is expressed
in all cell types and regulates the canonical pathway of NF-κB
activation and promote cell survival. The regulation of these
246 Clinic Rev Allerg Immunol (2017) 53:237264
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
signals by A20 is important for preventing autoimmune dis-
eases and defects could lead to autoimmune inflammatory
diseases. For example, A20-deficient mice were shown to
develop multi-organ inflammation and perinatal lethality
which prevented detailed studies of A20 functions in adult
mice [173]. However, mice lacking A20 expression specifi-
cally in B cells provided better insights into how A20 regu-
lates B cell development and functions. These mice spontane-
ously developed a lupus-like disease characterized by in-
creased plasma cell and germinal centre B cell numbers, ele-
vated levels of IgM and IgG autoantibodies and immunoglob-
ulin deposits in the kidney [3739]. The increase in germinal
centre B cell numbers could be due to resistance to FAS-
mediated apoptosis [39] and/or enhanced expression of
NF-κB-dependent anti-apoptotic proteins including Bcl-X.
Of note, however, is that these mice did not develop renal
failure but severe nephritis when lupus-prone mice were used.
Furthermore, heterozygous mice which expressed reduced
levels of A20 specifically in B cells manifested increased
numbers of germinal centre B cells and produced autoanti-
bodies [38]. In addition to enhanced BCR-mediated signal-
ling, A20-deficient mice were hyperresponsive to TLR and
CD40 engagement. Furthermore, when stimulated, A20-
deficient B cells produced higher levels of IL-6 compared
with wild-type B cells [39]. Enhanced IL-6 production in
A20-deficient B cells may account for the moderate increase
in T cell numbers in mice lacking A20 expression in B cells
[39].
In humans, GWAS and SNP analyses of TNFAIP3, the
gene encoding A20, revealed a potential role for A20 in sus-
ceptibility to autoimmune diseases in humans (Table 2)[185].
Subsequent studies confirmed an association with a number of
autoimmune diseases including SLE [186], RA [187], psoria-
sis [188], T1D [189,190] and SSc [191,192]. Since mice
expressing low levels of A20 develop spontaneous inflamma-
tion and autoimmune diseases [3739], TNFAIP3 SNPs might
affect its function or expression. Indeed, reduced A20 func-
tions in patients with SLE were associated with a SNP in the
coding region of TNFAIP3 that caused a substitution in residue
127 from phenylalanine to cysteine. In contrast, reduced A20
level was associated with a SNP at the 3enhancer region of
TNFAIP3 [193]. Additionally, it was suggested that SNPs lo-
cated outside of the coding regions of TNFAIP3 may confer
susceptibility to diseases by reducing A20 expression [194,
195]. Polymorphisms could also have prognostic and thera-
peutic values. Thus, TNFAIP3 polymorphisms and altered
A20 expression levels were associated with therapeutic re-
sponses to RA patient treated with anti-TNFαagents [196].
The association of TNFAIP3 polymorphisms with lymphoma
in patients with Sjögrens also highlights the potential role of
A20 in regulating B cell hyperactivity and malignant transfor-
mation leading to lymphomagenesis [197]. Moreover, the
presence of certain TNFAIP3 SNPs was associated with the
risk of severe renal or haematological complications in pa-
tients with SLE [193].
The NF-κB-associated signalling cascade is regulated by
an E2 enzyme, UBE2L3 (also called UBCH7). UBE2L3 par-
ticipates in the ubiquitination of p53, c-Fos and the NF-κB
precursor p105, and defects are associated with increased sus-
ceptibility to many autoimmune diseases including RA and
SLE [198,199]. A single haplotype spanning UBE2L3,
rs140490, was associated with increased UBE2L3 expression
in B cells and aligned across multiple autoimmune diseases.
Additionally, the UBE2L3 risk allele correlated with increased
numbers of plasmablasts and plasma cells in patients with
SLE suggesting a role for UBE2L3 in plasmablast and
plasmacyte development [67,200].
Innate Immune Receptor-Mediated Signalling
Innate Immune Receptor-Mediated Signalling and B Cell
Tole ra n ce
Innate immune receptors, such as TLRs, are pattern recogni-
tion molecules that bind conserved pathogen-associated mo-
lecular patterns (PAMPs) on pathogens. Ten TLRs are
expressed in human cells, whereas in murine cells, there are
13 such receptors. Naïve human B cells express TLR1, 2, 3, 4,
6, 7 and 9, while plasma cells only express TLR3 and 4 [201].
TLR7 and TLR9 are known to be able to directly influence B
cell tolerance. Engagement of TLR2 and TLR4, in contrast,
has been implicated in promoting autoimmune diseases in
mice although there is no direct evidence to support how, or
indeed if, these two receptors modulate B cell. TLR7 and
TLR9 are intracellular receptors that bind their ligands in
endosomes. TLR7 binds ssRNA while TLR9 binds CpG
DNA in viruses and bacteria. Interestingly, these receptors
can be stimulated in self-reactive B cells by RNA and/or
DNA-containing immune complexes. The two receptors di-
merize upon ligand binding and recruit the adaptor protein
myeloid differentiation primary response gene 88 (MyD88).
The IL-1 receptor-associated kinase 4 (IRAK4) binds to
MyD88 and activates IRAK1 and IRAK2. The resulting sig-
nalling complexes initiate the activation of NF-κB, MAPK
and IFN-regulatory factor 1 (IRF1) and IRF5 signalling path-
ways and regulate the production of pro-inflammatory cyto-
kine [202205].
Dysregulated Innate Immune Receptor-Mediated Signalling
Promotes B Cell Autoreactivity
A key feature of immunological abnormality in patients with
autoimmune diseases is the production of autoantibodies, such
as autoantibodies with specificity for nuclear antigens includ-
ing DNA and proteins. There is evidence that crosstalk be-
tween signalling mediated by the BCR and TLRs could play
Clinic Rev Allerg Immunol (2017) 53:237264 247
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
an important role in the loss of B cell tolerance to these anti-
gens [206,207]. For example, BCR engagement by nucleic
acid associated with self-antigens facilitates trafficking to
endosomal compartments where TLRs reside leading to their
engagement and B cell activation [208,209]. In this respect,
both TLR7 and TLR9 which initiate MYD88-dependent sig-
nalling pathways have been implicated in the pathogenesis of
animal models of lupus and the production of anti-nuclear
autoantibodies. Indeed, deletion of the TLR7 gene in lupus
mice suppresses the production of autoantibodies to RNA-
associated proteins and ameliorate systemic autoimmunity.
Paradoxically, however, deletion of the TLR9 gene abolishes
anti-dsDNA and anti-chromatin autoantibody production but
exacerbates clinical symptoms [210,211]. Since both TLR7
and TLR9 are expressed in B cells and myeloid cells, it is
unclear whether the phenotype seen in these mice could be
attributed to the effect of the two receptors on myeloid and/or
B cells. However, deletion of TLR7 in Wiskott-Aldrich syn-
drome protein (WASp) in mice inhibited systemic autoimmu-
nity, whereas deletion of TLR9 promoted systemic autoimmu-
nity which recapitulates the phenotype seen in TLR7/9-defi-
cient lupus mice [212214]. WASp is expressed in
haematopoietic cells and is implicated in BCR- and TLR-
mediated signalling. Mutations in the WASp gene in humans
cause Wiskott-Aldrich syndrome, an X-linked recessive dis-
ease characterized by primary immunodeficiency and high
levels of autoantibodies [215]. In contrast to attenuating T cell
receptor (TCR)-mediated signalling, WASp-deficient B cells
are hyperresponsive to both BCR and TLR engagement lead-
ing to enhanced signalling adequate to mediate autoimmune
disease even in the autoimmune-resistant B6 mouse [216]. In
addition, WASp-deficient B cells are capable of activating
wild-type CD4
+
T cells and inducing spontaneous germinal
centre formation, glomerulonephritis and the production of
class-switched autoantibodies in mixed bone marrow chi-
meras in mice [216]. These effects were all MyD88-
dependent since deletion of MyD88 in B cells abrogated T
cell activation and spontaneous germinal centre formation
[98,217219]. The pivotal role of TLR7 signalling in the
pathogenesis of lupus was confirmed in several mouse models
with the Y-chromosome-linked genomic-modifier Yaa in
which there is duplication of the Tlr7 gene [220]. In Yaa
mouse models, duplication of the Tlr7 gene was reported to
be the sole requirement for accelerated autoimmunity and that
reduction of Tlr7 gene dosage abolished the autoimmune phe-
notype.Furthermore,inTLR7 transgenic mice, B cells prefer-
entially homed to spontaneous germinal centres in competi-
tive chimeras suggesting a key role for TLR7-expressing B
cells in driving the formation of autoreactive germinal centres
[221]. Of note, overexpression of soluble RNAase ameliorat-
ed autoimmunity in TLR7-transgenic mice suggesting an im-
portant role for RNA in the pathogenesis of disease in these
mice [222]. In genetic studies in humans, SNPs within Tlr7
and polymorphisms in genes encoding proteins and transcrip-
tion factors downstream of TLR signalling, including
TNFAIP3, TNIP1 and IRF5, associate with susceptibility to
SLE [53,223226]. In addition, variants of SLC15A4, a his-
tidine transporter involved in lysosomal TLR signalling, also
associate with susceptibility to SLE. Furthermore, deletion of
SLC15A4inBcellslimitsautoimmunityinmurinemodelsof
the disease [61]. Noteworthy in this respect is that humans
deficient in either IRAK4 or MYD88, downstream effectors
of TLR signalling, show increased autoreactivity within the
naïve B cell compartment suggesting a pivotal role for TLR
signalling in regulating tolerance in B cells [227,228].
Co-stimulatory Receptor-Mediated Signalling
Co-stimulatory Receptor Signalling and B Lymphocyte
Responses
The outcome of BCR engagement is influenced by signalling
generated through a number of co-stimulatory receptors in-
cluding CD5, CD19, CD21, CD22, CD40, CD45, CD72 and
FcγRIIB. Signalling through these molecules upregulate and/
or downregulate BCR-mediated signalling to fine-tune B cell
responses. Any imbalance, or dysregulation, in signalling me-
diated through these co-receptors can either mediate autoim-
mune responses, or limit the ability of the immune system to
mount an effective humoral response.
One of the key co-receptors involved in modulating BCR-
mediated signalling is CD19. The cytoplasmic domain of
CD19 has nine tyrosine residues which, when phosphorylat-
ed, act as docking sites for SH2-containing adaptors and ki-
nases including PI3Ks, Vav-family guanosine exchange fac-
tors (GEFs) and growth factor receptor-bound protein 2
(Grb2). The engagement of CD40 by its ligand, CD40L, in
contrast, initiates signalling through TNFR-associated factors
(TRAFs) leading to the activation of downstream signalling
pathways including MAPKs and NF-κB.
The activation of BCR-mediated signalling is also regulat-
ed by protein tyrosine phosphatases (PTPs), some of which,
such as CD45, play dual positive and negative roles as cited
earlier. Cytoplasmic phosphatases are recruited to the BCR
complex through ITIM-containing co-receptors, such as
CD5, CD22 and the low-affinity Fcγreceptors, specifically
FcγRIIB [229]. Altered expression and/or activation either of
kinases or phosphatases can lead to defective BCR-mediated
signalling which, in turn, alters B-lymphocyte responses
[230232].
Co-ligation of the BCR and the FcγRIIB by antigen-
antibody complexes leads to tyrosine phosphorylation of
ITIMs [233], which, in turn, recruit SHIP, a lipid phosphatase
with specificity for 5-phosphate of PIP3 [234]throughSH2-
domain-mediated binding. SHIP dephosphorylates PIP3 to
produce PI(3,4)P2 and, thus, diminish BCR-mediated
248 Clinic Rev Allerg Immunol (2017) 53:237264
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
elevation of PIP3. B lymphocytes also express the siglec fam-
ily member CD22, an ITIM-containing receptor which inter-
acts with ligands carrying a 26-linked sialic acids [235].
CD22 modulates BCR signalling threshold and inhibits sig-
nalling by recruiting SHP-1, a tyrosine phosphatase.
CD72 is constitutively expressed on B cells at all stages of
their development except on plasma cells. CD72 negatively
regulates BCR-induced signalling by recruiting SHP-1
through its cytoplasmic ITIM motif [236]. CD72 plays an
essential regulatory role in modulating BCR-mediated signal-
ling in autoreactive B cells [237]. In anergic B cells, CD72
downregulates BCR-mediated signalling by limiting antigen-
induced Ca
2+
influx and the activation of NFATc1, NF-κB,
MAPK and Akt. Noteworthy, CD72 associates with SHP-1
and Cbl-b, suggesting a role for SHP-1 and Cbl-b in CD72-
mediated inhibitory effects on BCR-mediated signalling in
anergic B cells. CD100, a ligand for CD72, can turn off the
negative effect of CD72 by inhibiting the phosphorylation of
CD72 and, consequently, disrupting the interaction between
SHP-1 and CD72 [238,239].
Dysregulated Co-stimulatory Receptor-Mediated Signalling
Promotes Autoreactive B Cell Expansion and Autoantibody
Production
Cognate interactions between B and T cells involving co-
stimulatory receptors such as CD40 and CD40L are critical
for thymus-dependent humoral immunity. Ligation of CD40
induces B cell proliferation, class switching and somatic mu-
tations. In lupus disease, loss or blockade of cognate B-T cell
interactions involving CD40-CD40L ameliorates disease and
prolongs survival in the NZB/NZW F1 and MRL-lpr sponta-
neous models of lupus [240242]. In addition, the use of ag-
onist anti-CD40 antibodies inhibits apoptosis of rheumatoid
factor (RF) precursor B cells in arthritic mice, while blockade
of CD40-CD40L abolishes RF production in transgenic mice
[243,244]. Furthermore, cd40l gene-deficient mice or treat-
ment of neonatal NOD mice with anti-CD40L antibodies sup-
presses autoimmune diabetes [245248]. Moreover, treatment
of EAE mice with anti-CD40L antibody improved disease
[249,250]. Preclinical assessment of anti-CD40 antibody in
a model of multiple sclerosis (MS) in monkeys provided ad-
ditional support for the importance of CD40-CD40L interac-
tion in autoimmune diseases [251253]. However, clinical
trials of anti-CD40L in patients with lupus had mixed out-
comes, and in addition, some patients developed thromboem-
bolism [254256].
As cited above, in addition to CD40, defective regulation
of engagement or signalling through other co-stimulatory re-
ceptors such as CD5, CD22 and FcγRIIB can also promote
autoimmune diseases. CD5 and CD22 negatively regulate
BCR-mediated signalling through ITIMs in their intracellular
domains and PTPs. The PTPs can have dual inhibitory and
activating effects. In autoimmune diseases, there is substan-
tive evidence that defects in the regulation of co-stimulatory
receptors and associated PTPs promote lupus disease, both in
animal models and in patients. For example, there is evidence
for altered expression of CD22 and SHP-1 in patients with
SLE [4,6,257]. In genetically engineered mice, deletion of
cd22,FcγRIIB or PTPN6, which encodes SHP-1, leads to B
lymphocyte hyperactivity, auto-Ab production and lupus-like
disease [5,96,258]. However, it remains unclear whether
defects in the regulation of these co-stimulatory receptors
and associated PTPs in patients are inherent or result from
the disease process or, indeed, if they have a causal relation-
ship with the disease.
In addition to the role of dysregulated kinases and phos-
phatases that regulate proximal BCR signalling, defects in co-
stimulatory receptors that regulate downstream signalling
have been associated with the development of autoimmune
diseases. For example, dysregulation of CD72 has been
shown to promote autoimmune diseases. In anergic B cells,
CD72 constitutively regulates BCR-mediated signalling and
limits proliferation and survival through suppressing cyclin
D2 expression and Rb phosphorylation, key regulators of the
cell cycle. Indeed, CD72-deficient mice spontaneously pro-
duce autoantibodies and develop lupus-like disease.
Furthermore, CD72-deficient B cells proliferate and survive
when their BCRs are engaged by self-antigens. The prolifera-
tive response of anergic B cells to BCR engagement by self-
antigens results in the loss of immunological self-tolerance,
upregulation of cyclin D2 and Bcl-xL, proliferation and sur-
vival of autoreactive B cells [259]. In contrast to anergic B
cells where calcineurin/NFAT and NF-κBsignallingpathways
are defective [260,261], self-antigen binding to anergic
CD72
/
B cells leads to the activation of both calcineurin/
NFAT and NF-κB[259]. Both calcineurin/NFAT and NF-κB
are required for the induction of cyclin D2 [262,263] and
activation of MAPK and Akt, key regulators of cell cycle
and survival.
Cytokine-Mediated Signalling
Cytokine Signalling in B Cell Differentiation
Dynamic regulation of cytokine production and cytokine re-
ceptor expression is required for B cell development, differ-
entiation and efficient immune responses. Cytokines are in-
volved in cellular communications and signalling and initiate
a wide range of effects including cell differentiation, prolifer-
ation and regulation. Cytokines involved in B cell differenti-
ation and responses include interferons (IFNs), interleukins
(ILs) and members of the TNF family of ligands and recep-
tors. Almost 40 cytokine receptors are known to initiate intra-
cellular signalling mostly through JAKs and signal transduc-
ers and activators of transcription (STATs) [264,265]. In
Clinic Rev Allerg Immunol (2017) 53:237264 249
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
addition to activating JAKs and STATs, cytokines can also
initiate other signalling pathways such as activating Ras and
PI3Ks [266,267]. The binding of cytokines to their receptors
induces dimerization or polymerization of the receptors and
this activates associated JAKs. Activated JAKs induce phos-
phorylation and homo- and hetero-dimerization of STATs.
Dimerized STATs translocate to the nucleus where they induce
transcription of their target genes.There are four JAKs (JAK1,
JAK2, JAK3 and TYK2) and seven known STATs (STAT1,
STAT2, STAT3, STAT4, STAT5a, STAT5b and STAT6).
Receptors for type 1 IFN signal via JAK1 and Tyk2, whereas
receptors for IL-12 and IL-23 signal through JAK2 and Tyk2.
Receptors for IL-2, IL-4, IL-7, IL-9, IL-15 and IL-23 signal
through JAK1 and JAK3, whereas IFNγreceptor signals via
JAK1 and JAK2 [268].
BAFF, which is a member of the TNF family of ligands and
receptors, is crucial for B cell survival and development [269,
270]. The cytokine can bind to three different receptors:
BAFF-R, B cell maturation antigen (BCMA) and transmem-
brane activator and Ca
2+
modulator and cyclophilin ligand
interactor (TACI) [271,272]. Binding of BAFF to BAFF-R
plays a predominant role in B cell maturation and survival,
and mice deficient in either BAFF or BAFF-R exhibit a B cell
developmental block at the T2 stage of transitional B cell
maturation. In contrast, B cells in BCMA- or TACI-deficient
mice develop normally into mature cells [273]. BAFF-R is
linked to TRAFs and signals through both the canonical and
alternative NF-κB pathways as well as through MAPK and
PI3K pathways [274,275]. Activation of the NF-κBpathway
by engagement of the BAFF-R rescues transitional B cells
receiving BCR signals from apoptosis in response to engage-
ment by self-antigens, possibly through increased transcrip-
tion of anti-apoptotic proteins and posttranscriptional modifi-
cations of pro-apoptotic proteins [276278]. Interestingly, re-
cent studies have revealed that there is crosstalk between
BAFF-R and BCR signals that can induce survival signals
independent of NF-κB activation [279,280]. Thus, BAFF
promotes rapid phosphorylation of proximal BCR signalling
involving Ig-αand Syk. Deletion of Syk impairs survival and
renders B cells non-responsive to BAFF. Survival in this set-
ting can partially be restored by ectopic activation of MAPK
or PI3K suggesting that BAFF-R signalling is likely to facil-
itate BCR-induced survival through activation of MAPK and
PI3K.Inthisrespect,BCRandCD19havebeenimplicatedin
regulating BAFF-R levels [86,273].
Altered Cytokine Prof iles Impacts Intracellular B Cell
Signalling and Responses
Studies over the last few years have suggested that B cells can
be subdivided into effector subsets based on the profile of
cytokines they produce. Thus, B cells have been subdivided,
in a manner akin to the subdivision of Th1 and Th2 cells, into
B effector 1 cells that produce IFNγand IL-12 and B effector
2 cells that produce IL-2, IL-4 and IL-6. More recent studies
identified another effector B cell subset, B-regulatory cells
(Bregs), characterized by their ability to produce IL-10,
TGFβand IL-35 and with immunosuppressive functions
[281]. However, the available evidence indicates that, in con-
trast to effector T cell subsets, effector B cell subsets do not
fulfil requirements of classic immune lineages such as defin-
ing transcription factors and may also exhibit plasticity de-
pending on their microenvironmental settings. Nevertheless,
the evidence provides support for a differential profile of cy-
tokine production in B cells in different pathophysiological
conditions. For example, altered profiles of cytokine produc-
tion have been implicated in aberrant B cell responses in au-
toimmune diseases. Furthermore, in addition to the role of
cognate T-B cell interactions in diverging cytokine production
in B cells, TLR co-engagement with CD40 has been shown to
synergize in promoting IL-10 and IL-35 production
[282284]. Both cytokines have important regulatory func-
tions including limiting the generation of autoreactive germi-
nal centres [284]. Interestingly, IL-10 can have dual effects on
autoimmune diseases: acting as a B cell stimulator and also
suppressor of T cell activations [285]. In this respect, IL-10
impacts autoimmune disease pathology differently depending
on which cell and/or mechanism drives a disease. For exam-
ple, while the transfer of IL-10-producing Bregs drives Treg
cell expansion and modulates arthritis in mice, treatment of
SLE patients with IL-10-specific monoclonal antibodies ame-
liorates disease [286,287]. This outcome is consistent with
evidence showing that high levels of IL-10 correlate with lu-
pus disease activity in patients [288]. Interestingly, however,
there is also evidence for defective signalling that regulate IL-
10 production by B cells in patients with SLE [281,282].
Similar to IL-10, IL-35 production by B cells has been shown
to have immune regulatory functions and essential for recov-
ery from EAE in mice [283].
In addition to the differentiation of B cells to distinct effec-
tor subsets with different cytokine profiles that impact auto-
immune diseases differently, altered regulation of cytokine
signalling in B cells can promote or enhance autoimmune
disease pathology. For example, lupus is associated with high
levels of IFNαand INFγproduction with both altering B
lymphocyte responses and autoantibody isotype production
[289291]. IFNαlowers BCR activation thresholds and pro-
motes B cell differentiation through activating IRF5 transcrip-
tion factor [292]. Indeed, polymorphism in the IRF5 gene has
been associated with susceptibility to SLE [7]. Interestingly,
excess production of IFNαcan be induced by immune com-
plexes suggesting a positive feedback circuit between IFNα
and autoreactive B cells in lupus [290]. High levels of IFNγ,
in contrast, enhance the production of complement-fixing IgG
subclass of autoantibodies in lupus mice and promote lupus
disease in patients with RA when treated with the cytokine
250 Clinic Rev Allerg Immunol (2017) 53:237264
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Tab l e 3 Studies, clinical trials and approved therapeutic targeting of cytokines and signalling pathways in B cells for treating autoimmune diseases
Targeted signalling
molecule/pathway
Agent used Structure of agent Biological effects on B lymphocytes Disease/status Reference
JAK1/JAK2/JAK3
and to a lesser
extent TYK2
Tofacitinib Chemical inhibitor Inhibits the JAK/STAT pathway and blocks cytokine
signalling
Approved for treating RA in many countries but
not yet in the EU
[302]
JAK1/JAK2 Baricitinib Chemical inhibitor Inhibits the JAK/STAT pathway and blocks cytokine
signalling
RA in phase III clinical trials [302]
JAK3 Decernotinib
(VX-509)
Chemical inhibitor Inhibits the JAK/STAT pathway and blocks cytokine
signalling
RA in phase II clinical trials [302]
Pan-JAK Peficitinib (ASP015
K)
Chemical inhibitor Inhibits the JAK/STAT pathway and blocks cytokine
signalling
RA in phase II clinical trials [302]
JAK1 Filgotinib
(GLPG-0634)
Chemical inhibitor Inhibits the JAK/STAT pathway and blocks cytokine
signalling
RA in phase II clinical trials [302]
JAK1 ABT-494 Chemical inhibitor Inhibits the JAK/STAT pathway and blocks cytokine
signalling
RA in phase II clinical trials [302]
JAK1/JAK2 INCB039110 Chemical inhibitor Inhibits the JAK/STAT pathway and blocks cytokine
signalling
RA in phase II clinical trials [302]
JAK/SYK R333 Chemical inhibitor Inhibits the JAK/STAT pathway and blocks cytokine
signalling
Discoid lupus in phase II clinical trials [303]
JAK1 GSK2586184 Chemical inhibitor Inhibits the JAK/STAT pathway and blocks cytokine
signalling
SLE in phase II clinical trials [303]
JAK1 GLG0778 Chemical inhibitor Inhibits the JAK/STAT pathway and blocks cytokine
signalling
SLE in phase II clinical trials [303]
SYK Fostamatinib Chemical inhibitor Inhibits SYK and blocks BCR and FcγR signalling Clinical trials concluded that it is effective in
treating RA; however, its clinical application
is precluded due to unexpected side effects
[304]
BLyS (BAFF) Atacicept Recombinant fusion
protein (TACI-Ig)
Blocks BLyS/APRIL binding and reduces survival and
the number of some B cell subsets
Reduced B cell and plasma cell numbers and
SLE disease activity but phase II/III trial
stopped due to low blood Ab levels and
pneumonia
[305]
Belimumab Fully human monoclonal
Ab (mAb)
Inhibits BLyS binding to membrane receptors;
promotes apoptosis of B lymphocytes
Approved for treating SLE. However, patients
with active lupus nephritis are excluded.
Use for active lupus nephritis at phase III clinical
trials. Sjögrens syndrome phase III clinical
trials
[305]
Briobacept (BR3-Fc) Recombinant fusion
protein
Inhibits BLyS binding to its receptor and promotes
apoptosis
SLE clinical trials did not show sufficient
efficacy
[305]
Blisibimod
(AMG-623)
Peptide-Fc fusion protein
with 4 BLyS binding
domains
Inhibits BLyS binding to its receptors and promotes
apoptosis
SLE clinical trial is in phase III [305]
IL-6 Sirukumab Fully human mAb Reduces B lymphocyte proliferation and differentiation Clinical trials concluded its effectiveness in
inhibiting progression of joint damage and
improved signs and symptoms of disease in
RA
[306]
IL-6R Tocilizumab Humanized mAb Blocks B lymphocyte differentiation and reduces Ab
production
Clinical trials concluded its effectiveness as a
therapy for treating early RA
[307]
IFNαRontalizumab Humanized mAb Inhibits B lymphocyte activation and Ab production Clinical trials concluded its effectiveness in
treating SLE
[307]
Clinic Rev Allerg Immunol (2017) 53:237264 251
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
[293]. Further evidence for the harmful role of excess IFNγ
production in promoting B cell abnormalities and lupus dis-
eases comes from studies in which deletion of IFNγreceptor
in B cells abrogates spontaneous germinal centre formation,
class switching of autoantibodies and nephritis in lupus-prone
mice [294].
Studies in humans and mice have also revealed an impor-
tant role for excess IL-21 production in promoting B cells to
differentiate to plasma cells [295]. Thus, IL-21
/
mice have a
diminished ability to produce IgG1 in response to immuniza-
tion, whereas transgenic mice with enhanced expression of IL-
21 develop hypergammaglobulinemia [295]. In contrast, IL-
21 blockade successfully ameliorates lupus symptoms in
lupus-prone MRL mice. Furthermore, knocking out il-21r
gene suppresses lupus manifestations in the BXSB Yaa mouse
model [295].
Overproduction of other cytokines, such as BAFF, a cyto-
kine crucial for peripheral B cell development, has also been
implicated in promoting autoimmune diseases [269,270]. As
cited earlier, BAFF binds to BAFF-R, BCMA and TACI.
BAFF-R plays a key role in B cell maturation and survival
and excess BAFF production is noted in many autoimmune
diseases including SLE, RA and MS [296]. Indeed, transgenic
mice expressing a high level of BAFF (BAFF-Tg) display B
cell hyperplasia and develop lupus-like disease [297]. Of note
in this respect is that high levels of BAFF rescue low-affinity
self-reactive transitional B cells from negative selection at
tolerance checkpoints and allows them to become mature B
cells [298,299]. These observations suggest that signalling
through BAFF-R synergizes with BCR-mediated signalling
in autoreactive B cells to override tolerance in these cells
and permit the generation of pathogenic autoreactive B cells.
Interestingly, MyD88, essential for TLR signalling, is crucial
for autoantibody production in BAFF-Tg mice, suggesting
that there is an interplay between BAFF-R and TLR signalling
in promoting autoimmune diseases [297].
In addition to overt changes in cytokine production and
signalling abnormalities in B cells, subtle changes in the reg-
ulation of signalling pathways activated by cytokine binding
can also promote humoral autoimmunity. For example, het-
erozygous mutations in the stat3 gene leading to replacements
of key amino acids, such as K392R, M394T and K658N, and
enhanced STAT3 binding, or dimerization are associated with
multi-organ autoimmunity, lymphoproliferation and
hypogammaglobulinemia with terminal B cell maturational
arrest [300]. Similarly, SNPs in stat3 and stat4 genes have
been associated with autoimmune thyroid diseases [301].
However, the mechanism involved in promoting altered B cell
responses inindividuals with these polymorphisms and how B
cells promote disease are yet to be determined.
It is noteworthy that several JAK inhibitors have been re-
cently developed as new therapies for treating patients with
inflammatory autoimmune diseases such as RA. Thus,
Tab l e 3 (continued)
Targeted signalling
molecule/pathway
Agent used Structure of agent Biological effects on B lymphocytes Disease/status Reference
Sifalimumab Fully human mAb Inhibits B lymphocyte activation and Ab production Clinical trials concluded its effectiveness in
treating SLE
[308]
TLR4 NI-0101 Humanized mAb Inhibits signalling through TLR4 In phase I clinical trial for treating RA [309]
TLR7/8/9 Chloroquine Chemical TLR7/8/9
antagonist
Reduces endosomal acidification and inhibits signalling
through TLR
A mainstay therapy for SLE [308]
TLR7/8/9 IMO-8400 Chemical TLR7/8/9
antagonist
Inhibits signalling through TLR7/8/9 In phase I clinical trials for SLE [308]
TLR7-RLR9 IMO-3100 Chemical TLR7/9 antag-
onist
Inhibits signalling through TLR7/9 SLE clinical trial is in phase I [308]
The table summarizes available information on the use of therapeutic agents to target signalling pathways in B lymphocytes in clinical trials and in practice
252 Clinic Rev Allerg Immunol (2017) 53:237264
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
tofacitinib that inhibits JAK1, JAK2, JAK3 and, to a lesser
extent, TYK2 is used in the clinic for treating RA patients in
many countries (Table 3). JAKs are a family of non-receptor
tyrosine kinases that are critical for signalling through cyto-
kine receptors. At present, a number of other JAK inhibitors
are in clinical trials for treating patients with diseases includ-
ing RA and SLE (Table 3).
Conclusions
It is now established that B cells play a key role in initi-
ating and driving pathogenesis in many autoimmune dis-
eases including SLE, RA, MS and SSc beyond their role
in producing autoantibodies. Pathogenic roles played by B
cells depend on tolerance status of the cells, which co-
receptors are co-engaged with the BCR, the cytokine mi-
lieu and, ultimately, the nature and extent of intracellular
signalling generated within B cells. The outcome of BCR
engagement by antigens, self or exogenous, is determined
by optimal levels or engagement and activation of kinases
and phosphatases that regulate the strength and duration
of intracellular signalling. Therefore, abnormal or incon-
gruous engagement/regulation of signalling proteins, co-
receptors or cytokines/TLRs can override B cell tolerance
and lead to autoimmune disease development. Indeed, the
available evidence indicates that regulated and coordinat-
ed engagement of co-receptors and signalling pathways
ensures immune response specificity and efficiency and
prevents the development of autoimmune diseases. This
article has provided an overview of some of the key B cell
signalling pathways and how defects in these could im-
pact pathophysiology (Fig. 1). The available functional
evidence on how changes in the level or function of sig-
nalling proteins and pathways impact B cell responses
provides molecular mechanisms for how GWAS and
SNP association with diseases such as SLE, RA, MS
and TID could predispose to these diseases (Table 2).
Noteworthy is that even in healthy individuals,
autoreactive B cells, perhaps with low affinity for self-
antigens, persist in the naïve B cell repertoire, yet these
fail to become high-affinity pathogenic B cells. Thus, de-
fining the molecular mechanisms that constrain the acti-
vation and/or regulation of self-reactive B cells will help
understanding disease mechanisms and could also pave
the way for new therapeutic strategies in precision medi-
cine. Although B cell depletion therapy has proved to be
highly effective in treating a number of autoimmune dis-
eases including RA, MS and SSc, total ablation of B cells
carries its own risks including life-threatening infections.
In this respect, some recent studies have revealed that
treatment of lupus mice with inhibitors of Brutonstyro-
sine kinase (BTK) can ameliorate disease [310,311].
Thus, targeting dysregulated signalling effectors associat-
ed with proximal or downstream of BCR, CD40, TLR or
cytokine receptors could prove an effective therapeutic
strategy. However, despite substantial progress in the past
few years in defining the role of altered B cell signalling
in autoimmune diseases many questions and challenges
remain.
Take-Home Points
B lymphocytes are essential for effective immunity to
pathogens, but they can also cause diseases through pro-
ducing autoantibodies, disease-promoting cytokines and
presenting antigens to autoreactive T lymphocytes.
The potential of B lymphocyte to cause diseases is
prevented by tolerance and by a tight control of intracel-
lular signalling pathways.
Defects in intracellular signalling, however, can occur
and these lead to autoimmune, lymphoproliferative or
immunodeficient diseases.
Defects in proximal BCR signalling due to reduced Lyn
and/or CD45 levels promote autoreactive B cell activa-
tion, class switching to IgG autoantibodies and disease.
Defects in downstream BCR signalling, e.g. Ca
+2
or di-
acylglycerol pathways, lead to increased activation of
NFAT transcription factor, tolerance defects, B lympho-
cyte hyperactivity and autoimmune diseases. Defects in
PI3K lead to apoptosis defects and the generation of
short-lived autoreactive plasma cells.
Defects in ubiquitination, which regulates signalling in B
lymphocytes through controlling protein levels and tran-
scription factors, enhance activation leading to enhanced
B lymphocyte proliferation to BCR and CD40 engage-
ments and autoantigen presentation to T lymphocytes and
lead to autoimmune diseases.
Defects in innate immune receptor signalling, such as
TLRs, exaggerate defects in BCR signalling and enhance
autoreactive B cell expansion and responses.
Dysregulated expression of co-stimulatory receptors,
stimulatory or inhibitory, including CD5, CD19, CD21,
CD22, CD40, CD45, CD72 and FcγRIIB, leads to B
lymphocyte hyperactivity, auto-Ab production and auto-
immune diseases.
Altered cytokine production, cytokine receptor expres-
sion and/or cytokine signalling help rescue low-affinity
self-reactive B lymphocytes, their differentiation and IgG
isotype autoantibody production.
Defects in the level and activation of JAKs and STATs
enhance B lymphocyte differentiation leading to multi-
organ autoimmunity, lymphoproliferation or
hypogammaglobulinemia.
Clinic Rev Allerg Immunol (2017) 53:237264 253
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Acknowledgments Grant UC5 from the Raynauds and Scleroderma
Association (new name: Scleroderma research-UK) awarded to Rizgar
Mageed and David Abraham.
Compliance with Ethical Standards
Conflict of Interest The authors declare that they have no conflict of
interest.
Funding Dr. Taher E Taher was funded by grant UC5 from
Scleroderma and Raynauds-UK (SR-UK) awarded to Professors Rizgar
Mageed and David Abraham.
Ethical Approval This articledoes not contain any studies with human
participants or animals performed by any of the authors.
Informed Consent Informed consent was obtained from all individual
participants included in the study.
References
1. Yu X (2015) Autoimmune diseases: what have we learned from
mice? Curr Pharm Des 21(18):23062307. doi:10.1038/mp.2016.
165
2. Konsta OD, Le Dantec C, Brooks WH, Renaudineau Y (2015)
Genetics and epigenetics of autoimmune diseases. eLS 19;
Brooks Wh, J Autoimmunity, 2010
3. Hibbs ML, Tarlinton DM, Armes J, Grail D, Hodgson G, Maglitto
R, Stacker SA, Dunn AR (1995) Multiple defects in the immune
system of Lyn-deficient mice, culminating in autoimmune disease.
Cell 83(2):301311
4. Flores-Borja F, Kabouridis PS, Jury EC, Isenberg DA, Mageed
RA (2005) Decreased Lyn expression and translocation to lipid
raft signaling domains in B lymphocytes from patients with sys-
temic lupus erythematosus. Arthritis Rheum 52(12):39553965.
doi:10.1002/art.21416
5. Mary C, Laporte C, Parzy D, Santiago ML, Stefani F, Lajaunias F,
Parkhouse RM, OKeefe TL, Neuberger MS, Izui S et al (2000)
Dysregulated expression of the Cd22 gene as a result of a short
interspersed nucleotide element insertion in Cd22a lupus-prone
mice. J Immunol 165(6):29872996
6. Liossis SN, Solomou EE, Dimopoulos MA, Panayiotidis P,
Mavrikakis MM, Sfikakis PP (2001) B-cell kinase lyn deficiency
in patients with systemic lupus erythematosus. J Investig Med
49(2):157165. doi:10.2310/6650.2001.34042
7. Alarcón-Riquelme ME, Ziegler JT, Molineros J, Howard TD,
Moreno-Estrada A, Sánchez-Rodríguez E, Ainsworth HC, Ortiz-
Tello P, Comeau ME, Rasmussen A et al (2016) Genome-wide
association study in an Amerindian ancestry population reveals
novel systemic lupus erythematosus risk loci and the role of
European admixture. Arthritis Rheumatol 68(4):932943. doi:
10.1002/art.39504
8. Dai X, James RG, Habib T, Singh S, Jackson S, Khim S, Moon
RT, Liggitt D, Wolf-Yadlin A, Buckner JH et al (2013) A disease-
associated PTPN22 variant promotes systemic autoimmunity in
murine models. J Clin Invest 123(5):20242036. doi:10.1172/
JCI66963
9. Miyagawa H, Yamai M, Sakaguchi D, Kiyohara C, Tsukamoto H,
Kimoto Y, Nakamura T, Lee JH, Tsai CY, Chiang BL et al (2008)
Association of polymorphisms in complement component C3
gene with susceptibility to systemic lupus erythematosus.
Rheumatol 47(2):158164. doi:10.1093/rheumatology/kem321
10. Vang T, Miletic AV, Bottini N, Mustelin T (2007) Protein tyrosine
phosphatase PTPN22 in human autoimmunity. Autoimmunity
40(6):453461. doi:10.1080/08916930701464897
11. Hermiston ML, Tan AL, Gupta VA, Majeti R, Weiss A (2005) The
juxtamembrane wedge negatively regulates CD45 function in B
cells. Immunity 23(6):635647. doi:10.1016/j.immuni.2005.11.001
12. Flores-Borja F, Kabouridis PS, Jury EC, Isenberg DA, Mageed
RA (2007) Altered lipid raft-associated proximal signaling and
translocation of CD45 tyrosine phosphatase in B lymphocytes
from patients with systemic lupus erythematosus. Arthritis
Rheum 56(1):291302. doi:10.1002/art.22309
13. Halcomb KE, Musuka S, Gutierrez T, Wright HL, Satterthwaite
AB (2008) Btk regulates localization, in vivo activation, and class
switching of anti-DNA B cells. Mol Immunol 46(2):233241. doi:
10.1016/j.molimm.2008.08.278
14. Kil LP, de Bruijn MJ, van Nimwegen M, Corneth OB, van
Hamburg JP, Dingjan GM, Thaiss F, Rimmelzwaan GF, Elewaut
D, Delsing D et al (2012) Btk levels set the threshold for B-cell
activation and negative selection of autoreactive B cells in mice.
Blood 119(16):37443756. doi:10.1182/blood-2011-12-397919
15. Maas A, Hendriks RW (2001) Role of Brutons tyrosine kinase in
B cell development. Dev Immunol 8(34):171181
16. Honigberg LA, Smith AM, Sirisawad M, Verner E, Loury D,
Chang B, Li S, Pan Z, Thamm DH, Miller RA et al (2010) The
Bruton tyrosine kinase inhibitor PCI-32765 blocks B-cell activa-
tion and is efficacious in models of autoimmune disease and B-cell
malignancy. Proc Natl Acad Sci U S A 107(29):1307513080.
doi:10.1073/pnas.1004594107
17. OKeefe TL, Williams GT, Batista FD, Neuberger MS (1999)
Deficiency in CD22, a B cell-specific inhibitory receptor, is suffi-
cient to predispose to development of high affinity autoantibodies.
J Exp Med 189(8):13071313
18. OKeefe TL, Williams GT, Davies SL, Neuberger MS (1996)
Hyperresponsive B cells in CD22-deficient mice. Science
274(5288):798801
19. Uckun FM, Goodman P, Ma H, Dibirdik I, Qazi S (2010) CD22
EXON 12 deletion as a pathogenic mechanism of human B-
precursor leukemia. Proc Natl Acad Sci U S A 107(39):16852
16857. doi:10.1073/pnas.1007896107
20. Saito E, Fujimoto M, Hasegawa M, Komura K, Hamaguchi Y,
Kaburagi Y, Nagaoka T, Takehara K, Tedder TF, Sato S (2002)
CD19-dependent B lymphocyte signaling thresholds influence
skin fibrosis and autoimmunity in the tight-skin mouse. J Clin
Invest 109(11):14531462. doi:10.1172/JCI15078
21. Sato S, Hasegawa M, Fujimoto M, Tedder TF, Takehara K (2000)
Quantitative genetic variation in CD19 expression correlates with
autoimmunity. J Immunol 165(11):66356643
22. Kuroki K, Tsuchiya N, Tsao BP, Grossman JM, Fukazawa T,
Hagiwara K, Kano H, Takazoe M, Iwata T, Hashimoto H et al
(2002) Polymorphisms of human CD19 gene: possible association
with susceptibility to systemic lupus erythematosus in Japanese.
Genes Immun 3(Suppl 1):S21S30. doi:10.1038/sj.gene.6363906
23. Yoshizaki A (2016) B lymphocytes in systemic sclerosis: abnor-
malities and therapeutic targets. J Dermatol 43(1):3945. doi:10.
1111/1346-8138.13184
24. Bolland S, Ravetch JV (2000) Spontaneous autoimmune disease
in Fc(gamma)RIIB-deficient mice results from strain-specific
epistasis. Immunity 13(2):277285
254 Clinic Rev Allerg Immunol (2017) 53:237264
Open Access This article is distributed under the terms of the Creative
Commons Attribution 4.0 International License (http://
creativecommons.org/licenses/by/4.0/), which permits unrestricted
use, distribution, and reproduction in any medium, provided you give
appropriate credit to the original author(s) and the source, provide a link
to the Creative Commons license, and indicate if changes were made.
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
25. Yuasa T, Kubo S, Yoshino T, Ujike A, Matsumura K, Ono M,
Ravetch JV, Takai T (1999) Deletion of fcgamma receptor IIB
renders H-2(b) mice susceptible to collagen-induced arthritis. J
Exp Med 189(1):187194
26. Mackay M, Stanevsky A, Wang T, Aranow C, Li M, Koenig S,
Ravetch JV, Diamond B (2006) Selective dysregulation of the
FcgammaIIB receptor on memory B cells in SLE. J Exp Med
203(9):21572164. doi:10.1084/jem.20051503
27. Maxwell MJ, Duan M, Armes JE, Anderson GP, Tarlinton DM,
Hibbs ML (2011) Genetic segregation of inflammatory lung dis-
ease and autoimmune disease severity in SHIP-1/mice. J
Immunol 186(12):71647175. doi:10.4049/jimmunol.1004185
28. Taher TE, Parikh K, Flores-Borja F, Mletzko S, Isenberg DA,
Peppelenbosch MP, Mageed RA (2010) Protein phosphorylation
and kinome profiling reveal altered regulation of multiple signal-
ing pathways in B lymphocytes from patients with systemic lupus
erythematosus. Arthritis Rheum 62(8):24122423. doi:10.1002/
art.27505
29. Anzelon AN, Wu H, Rickert RC (2003) Pten inactivation alters
peripheral B lymphocyte fate and reconstitutes CD19 function.
Nat Immunol 4(3):287294. doi:10.1038/ni892
30. Suzuki A, Kaisho T, Ohishi M, Tsukio-Yamaguchi M, Tsubata T,
Koni PA, Sasaki T, Mak TW, Nakano T (2003) Critical roles of
Pten in B cell homeostasis and immunoglobulin class switch re-
combination. J Exp Med 197(5):657667
31. Wu XN, Ye YX, Niu JW, Li Y, Li X, You X, Chen H, Zhao LD,
Zeng XF, Zhang FC et al (2014) Defective PTEN regulation con-
tributes to B cell hyperresponsiveness in systemic lupus erythe-
matosus. Sci Transl Med 6(2446):246ra299. doi:10.1126/
scitranslmed.3009131
32. Medgyesi D, Hobeika E, Biesen R, Kollert F, Taddeo A, Voll RE,
Hiepe F, Reth M (2014) The protein tyrosine phosphatase PTP1B
is a negative regulator of CD40 and BAFF-R signaling and con-
trols B cell autoimmunity. J Exp Med 211(3):427440. doi:10.
1084/jem.20131196
33. Qian Y, Giltiay N, Xiao J, Wang Y, Tian J, Han S, Scott M, Carter
R, Jorgensen TN, Li X (2008) Deficiency of Act1, a critical mod-
ulator of B cell function, leads to development of Sjogrenssyn-
drome. Eur J Immunol 38(8):22192228. doi:10.1002/eji.
200738113
34. Huffmeier U, Uebe S, Ekici AB, Bowes J, Giardina E,
Korendowych E, Juneblad K, Apel M, McManus R, Ho P et al
(2010) Common variants at TRAF3IP2 are associated with sus-
ceptibility to psoriatic arthritis and psoriasis. Nat Genet 42(11):
996999. doi:10.1038/ng.688
35. Perricone C, Ciccacci C, Ceccarelli F, Di Fusco D, Spinelli FR,
Cipriano E, Novelli G, Valesini G, Conti F, Borgiani P (2013)
TRAF3IP2 gene and systemic lupus erythematosus: association
with disease susceptibility and pericarditis development.
Immunogenetics 65(10):703709. doi:10.1007/s00251-013-
0717-6
36. Potter C, Eyre S, Cope A, Worthington J, Barton A (2007)
Investigation of association between the TRAF family genes and
RA susceptibility. Ann Rheum Dis 66(10):13221326. doi:10.
1136/ard.2006.065706
37. Chu Y, Vahl JC, Kumar D, Heger K, Bertossi A, Wojtowicz E,
Soberon V, Schenten D, Mack B, Reutelshofer M et al (2011) B
cells lacking the tumor suppressor TNFAIP3/A20 display im-
paired differentiation and hyperactivation and cause inflammation
and autoimmunity in aged mice. Blood 117(7):22272236. doi:
10.1182/blood-2010-09-306019
38. Hovelmeyer N, Reissig S, Xuan NT, Adams-Quack P, Lukas D,
Nikolaev A, Schluter D, Waisman A (2011) A20 deficiency in B
cells enhances B-cell proliferation and results in the development
of autoantibodies. Eur J Immunol 41(3):595601. doi:10.1002/eji.
201041313
39. Tavares RM, Turer EE, Liu CL, Advincula R, Scapini P, Rhee L,
Barrera J, Lowell CA, Utz PJ et al (2010) The ubiquitin modifying
enzyme A20 restricts B cell survival and prevents autoimmunity.
Immunity 33(2):181191. doi:10.1016/j.immuni.2010.07.017
40. Musone SL, Taylor KE, Lu TT, Nititham J, Ferreira RC, Ortmann
W, Shifrin N, Petri MA, Kamboh MI, Manzi S et al (2008)
Multiple polymorphisms in the TNFAIP3 region are independent-
ly associated with systemic lupus erythematosus. Nat Genet 40(9):
10621064. doi:10.1038/ng.202
41. Plenge RM, Cotsapas C,Davies L, Price AL, de Bakker PI, Maller
J, Peer I, Burtt NP, Blumenstiel B, DeFelice M et al (2007) Two
independent alleles at 6q23 associated with risk of rheumatoid
arthritis. Nat Genet 39(12):14771482. doi:10.1038/ng.2007.27
42. Kitaura Y, Jang IK, Wang Y, Han YC, Inazu T, Cadera EJ,
Schlissel M, Hardy RR, Gu H (2007) Control of the B cell-
intrinsic tolerance programs by ubiquitin ligases Cbl and Cbl-b.
Immunity 26(5):567578. doi:10.1016/j.immuni.2007.03.015
43. Bergholdt R, Taxvig C, Eising S, Nerup J, Pociot F (2005) CBLB
variants in type 1 diabetes and their genetic interaction with
CTLA4. J Leukoc Biol 77(4):579585. doi:10.1189/jlb.0904524
44. Gomez-Martin D, Ibarra-Sanchez M, Romo-Tena J, Cruz-Ruiz J,
Esparza-Lopez J, Galindo-Campos M, Diaz-Zamudio M, Alcocer-
Varela J (2013) Casitas B lineage lymphoma b is a key regulator of
peripheral tolerance in systemic lupus erythematosus. Arthritis
Rheum 65(4):10321042. doi:10.1002/art.3783
45. Recher M, Burns SO, de la Fuente MA, Volpi S, Dahlberg C,
Walter JE, Moffitt K, Mathew D, Honke N, Lang PA et al
(2012) B cell-intrinsic deficiency of the Wiskott-Aldrich syn-
drome protein (WASp) causes severe abnormalities of the periph-
eral B-cell compartment in mice. Blood 119(12):28192828
46. Schurman SH, Candotti F (2003) Autoimmunity in Wiskott-
Aldrich syndrome. Curr Opin Rheumatol 15(4):446453
47. Rieck M, Arechiga A, Onengut-Gumuscu S, Greenbaum C,
Concannon P, Buckner JH (2007) Genetic variation in PTPN22
corresponds to altered function of T and B lymphocytes. J
Immunol 179(7):47044710
48. Jacob CO, Eisenstein M, Dinauer MC, Ming W, Liu Q, John S,
Quismorio FP Jr, Reiff A, Myones BL, Kaufman KM et al (2012)
Lupus-associated causal mutation in neutrophil cytosolic factor 2
(NCF2) brings unique insights to the structure and function of
NADPH oxidase. Proc Natl Acad Sci U S A 109(2):E59E67.
doi:10.1073/pnas.1113251108
49. Gateva V, Sandling JK, Hom G, Taylor KE, Chung SA, Sun X,
Ortmann W, Kosoy R, Ferreira RC, Nordmark G et al (2009) A
large-scale replication study identifies TNIP1, PRDM1, JAZF1,
UHRF1BP1 and IL10 as risk loci for systemic lupus erythemato-
sus. Nat Genet 41(11):12281233. doi:10.1038/ng.468
50. Morris DL, Sheng Y, Zhang Y, Wang YF, Zhu Z, Tombleson P,
Chen L, Cunninghame Graham DS, Bentham J, Roberts AL et al
(2016) Genome-wide association meta-analysis in Chinese and
European individuals identifies ten new loci associated with sys-
temic lupus erythematosus. Nat Genet 48(8):940946. doi:10.
1038/ng.3603
51. Harley JB, Alarcon-Riquelme ME, Criswell LA, Jacob CO,
Kimberly RP, Moser KL, Tsao BP, Vyse TJ, Langefeld CD,
Nath SK et al (2008) Genome-wide association scan in women
with systemic lupus erythematosus identifies susceptibility vari-
ants in ITGAM, PXK, KIAA1542 and other loci. Nat Genet
40(2):204210. doi:10.1038/ng.81
52. Raychaudhuri S, Thomson BP, Remmers EF, Eyre S, Hinks A,
Guiducci C, Catanese JJ, Xie G, Stahl EA, Chen R et al (2009)
Genetic variants at CD28, PRDM1 and CD2/CD58 are associated
with rheumatoid arthritis risk. Nat Genet 41(12):13131318. doi:
10.1038/ng.479
53. Han JW, Zheng HF, Cui Y, Sun LD, Ye DQ, Hu Z, Xu JH, Cai
ZM, Huang W, Zhao GP et al (2009) Genome-wide association
Clinic Rev Allerg Immunol (2017) 53:237264 255
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
study in a Chinese Han population identifies nine new suscepti-
bility loci for systemic lupus erythematosus. Nat Genet 41(11):
12341237. doi:10.1038/ng.472
54. Kozyrev SV, Abelson AK, Wojcik J, Zaghlool A, Linga Reddy
MV, Sanchez E, Gunnarsson I, Svenungsson E, Sturfelt G, Jonsen
A et al (2008) Functional variants in the B-cell gene BANK1 are
associated with systemic lupus erythematosus. Nat Genet 2008
40(2):211216. doi:10.1038/ng.79
55. Sawalha AH, Kaufman KM, Kelly JA, Adler AJ, Aberle T,
Kilpatrick J, Wakeland EK, Li QZ, Wandstrat AE, Karp DR
et al (2008) Genetic association of interleukin-21 polymorphisms
with systemic lupus erythematosus. Ann Rheum Dis 67(4):458
461. doi:10.1136/ard.2007.075424
56. Hom G, Graham RR, Modrek B, Taylor KE, Ortmann W, Garnier
S, Lee AT, Chung SA, Ferreira RC, Pant PV et al (2008)
Association of systemic lupus erythematosus with C8orf13-BLK
and ITGAM-ITGAX. N Engl J Med 358(9):900909. doi:10.
1056/NEJMoa0707865
57. Castillejo-Lopez C, Delgado-Vega AM, Wojcik J, Kozyrev SV,
Thavathiru E, Wu YY, Sanchez E, Pollmann D, Lopez-Egido
JR, Fineschi S et al (2012) Genetic and physical interaction of
the B-cell systemic lupus erythematosus-associated genes
BANK1 and BLK. Ann Rheum Dis 71(1):136142. doi:10.
1136/annrheumdis-2011-200085
58. Raychaudhuri S, Remmers EF, Lee AT, Hackett R, Guiducci C,
Burtt NP, Gianniny L, Korman BD, Padyukov L, Kurreeman FA
et al (2008) Common variants at CD40 and other loci confer risk
of rheumatoid arthritis. Nat Genet 40(10):12161223. doi:10.
1038/ng.233
59. Yang W, Shen N, Ye DQ, Liu Q, Zhang Y, Qian XX, Hirankarn N,
Ying D, Pan HF, Mok CC et al (2010) Genome-wide association
study in Asian populations identifies variants in ETS1 and
WDFY4 associated with systemic lupus erythematosus. PLoS
Genet 6(2):e1000841. doi:10.1371/journal.pgen.1000841
60. Lessard CJ, Li H, Adrianto I, Ice JA, Rasmussen A, Grundahl
KM, Kelly JA, Dozmorov MG, Miceli-Richard C, Bowman S
et al (2013) Variants at multiple loci implicated in both innate
and adaptive immune responses are associated with Sjogrenssyn-
drome. Nat Genet 45(11):12841292. doi:10.1038/ng.2792
61. Kobayashi T, Shimabukuro-Demoto S, Yoshida-Sugitani R,
Furuyama-Tanaka K, Karyu H, Sugiura Y, Shimizu Y, Hosaka
T, Goto M, Kato N et al (2014) The histidine transporter
SLC15A4 coordinates mTOR-dependent inflammatory responses
and pathogenic antibody production. Immunity 41(3):375388.
doi:10.1016/j.immuni.2014.08.011
62. Yang J, Yang W, Hirankarn N, Ye DQ, Zhang Y, Pan HF, Mok CC,
Chan TM, Wong RW, Mok MY et al (2011) ELF1 is associated
with systemic lupus erythematosus in Asian populations. Hum
Mol Genet 20(3):601607. doi:10.1093/hmg/ddq474
63. Manjarrez-Orduno N, Marasco E, Chung SA, Katz MS, Kiridly
JF, Simpfendorfer KR, Freudenberg J, Ballard DH, Nashi E,
Hopkins TJ et al (2012) CSK regulatory polymorphism is associ-
ated with systemic lupus erythematosus and influences B-cell sig-
naling and activation. Nat Genet 44(11):12271230. doi:10.1038/
ng.2439
64. Ding C, Ma Y, Chen X, Liu M, Cai Y, Hu X, Xiang D, Nath S,
Zhang HG, Ye H et al (2013) Integrin CD11b negatively regulates
BCR signalling to maintain autoreactive B cell tolerance. Nat
Commun 4:2813. doi:10.1038/ncomms3813
65. Lessard CJ, Adrianto I, Ice JA, Wiley GB, Kelly JA, Glenn SB,
Adler AJ, Li H, Rasmussen A, Williams AH et al (2012)
Identification of IRF8, TMEM39A, and IKZF3-ZPBP2 as suscep-
tibility loci for systemic lupus erythematosus in a large-scale mul-
tiracial replication study. Am J Hum Genet 90(4):648660. doi:10.
1016/j.ajhg.2012.02.023
66. Sheng YJ, Gao JP, Li J, Han JW, Xu Q, Hu WL, Pan TM, Cheng
YL, Yu ZY, Ni C et al (2011) Follow-up study identifies two novel
susceptibility loci PRKCB and 8p11.21 for systemic lupus erythe-
matosus. Rheumatology (Oxford) 50(4):682688. doi:10.1093/
rheumatology/keq313
67. Lewis MJ, Vyse S, Shields AM, Boeltz S, Gordon PA, Spector
TD, Lehner PJ, Walczak H, Vyse TJ (2015) UBE2L3 polymor-
phism amplifies NF-kappaB activation and promotes plasma cell
development, linking linear ubiquitination to multiple autoim-
mune diseases. Am J Hum Genet 96(2):221234. doi:10.1016/j.
ajhg.2014.12.024
68. Kaufman KM, Zhao J, Kelly JA, Hughes T, Adler A, Sanchez E,
Ojwang JO, Langefeld CD, Ziegler JT, Williams AH et al (2013)
Fine mapping of Xq28: both MECP2 and IRAK1 contribute to
risk for systemic lupus erythematosus in multiple ancestral groups.
Ann Rheum Dis 72(3):437444. doi:10.1136/annrheumdis-2012-
201851
69. HeB,SantamariaR,XuW,ColsM,ChenK,PugaI,ShanM,
Xiong H, Bussel JB, Chiu A et al (2010) The transmembrane
activator TACI triggers immunoglobulin class switching by acti-
vating B cells through the adaptor MyD88. Nat Immunol 11(9):
836845. doi:10.1038/ni.1914
70. Gregersen PK, AmosCI, Lee AT, Lu Y, Remmers EF, Kastner DL,
Seldin MF, Criswell LA, Plenge RM, Holers VM et al (2009)
REL, encoding a member of the NF-kappaB family of transcrip-
tion factors, is a newly defined risk locus for rheumatoid arthritis.
Nat Genet 41(7):820823. doi:10.1038/ng.395
71. Edwards JC, Szczepanski L, Szechinski J, Filipowicz-Sosnowska
A, Emery P, Close DR, Stevens RM, Shaw T (2004) Efficacy of
B-cell-targeted therapy with rituximab in patients with rheumatoid
arthritis. N Engl J Med 350(25):25722581. doi:10.1056/
NEJMoa032534
72. Pescovitz MD, Greenbaum CJ, Krause-Steinrauf H, Becker DJ,
Gitelman SE, Goland R, Gottlieb PA, Marks JB, McGee PF,
Moran AM et al (2009) Rituximab, B-lymphocyte depletion,
and preservation of beta-cell function. N Engl J Med 361(22):
21432152. doi:10.1056/NEJMoa0904452
73. Stone JH, Merkel PA, Spiera R, Seo P, Langford CA, Hoffman
GS, Kallenberg CG, St Clair EW, Turkiewicz A, Tchao NK (2010)
Rituximab versus cyclophosphamide for ANCA-associated vas-
culitis. N Engl J Med 363(3):221232. doi:10.1056/
NEJMoa0909905
74. Hauser SL, Waubant E, Arnold DL, Vollmer T, Antel J, Fox RJ,
Bar-Or A, Panzara M, Sarkar N, Agarwal S, HERMES Trial
Group et al (2008) B-cell depletion with rituximab in relapsing-
remitting multiple sclerosis. N Engl J Med 358(7):676688. doi:
10.1056/NEJMoa0706383
75. Daoussis D, Andonopoulos AP (2011) Rituximab in the treatment
of systemic sclerosis-associated interstitial lung disease: comment
on the article by Yoo. Rheumatol Int 31(6):841842. doi:10.1007/
s00296-010-1485-3
76. Jordan S, Distler JH, Maurer B, Huscher D, van Laar JM, Allanore
Y, Distler O, group ERs (2015) Effects and safety of rituximab in
systemic sclerosis: an analysis from the European Scleroderma
Trial and Research (EUSTAR) group. Ann Rheum Dis 74(6):
11881194. doi:10.1136/annrheumdis-2013-204522
77. Devauchelle-Pensec V, Morvan J, Rat AC, Jousse-Joulin S,
Pennec Y, Pers JO, Jamin C, Renaudineau Y, Quintin-Roué I,
Cochener B et al (2011) Effects of rituximab therapy on quality
of life in patients with primary Sjögrens syndrome. Clin Exp
Rheumatol 29(1):612
78. Jousse-Joulin S, Devauchelle-Pensec V, Morvan J, Guias B,
Pennec Y, Pers JO, Daridon C, Jamin C, Renaudineau Y, Roué
IQ et al (2007) Ultrasound assessment of salivary glands in pa-
tients with primary Sjögrens syndrome treated with rituximab:
256 Clinic Rev Allerg Immunol (2017) 53:237264
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
quantitative and Doppler waveform analysis. Biologics 1(3):311
319
79. Pers JO, Devauchelle V, Daridon C, Bendaoud B, Le Berre R,
Bordron A, Hutin P, Renaudineau Y, Dueymes M, Loisel S et al
(2007) BAFF-modulated repopulation of B lymphocytes in the
blood and salivary glands of rituximab-treated patients with
Sjögrens syndrome. Arthritis Rheum 56(5):14641477. doi:10.
1002/art.22603
80. Devauchelle-Pensec V, Pennec Y, Morvan J, Pers JO, Daridon C,
Jousse-Joulin S, Roudaut A, Jamin C, Renaudineau Y, Roué IQ
et al (2007) Improvement of Sjögrens syndrome after two infu-
sions of rituximab (anti-CD20). Arthritis Rheum 57(2):310317.
doi:10.1002/art.22536
81. Taher TE, Muhammad HA, Bariller E, Flores-Borja F,
Renaudineau Y, Isenberg DA, Mageed RA (2013) B-
lymphocyte signalling abnormalities and lupus immunopatholo-
gy. Int Rev Immunol 32(4):428444. doi:10.3109/08830185.
2013.788648
82. Isenberg DA (2012) Rituximabit was the best of times, it was
the worst of times. Autoimmun Rev 11(11):790791. doi:10.
1016/j.autrev.2012.02.005
83. Seret G, Le Meur Y, Renaudineau Y, Youinou P (2012) Mesangial
cell-specific antibodies are central to the pathogenesis of lupus
nephritis. Clin Dev Immunol 2012:579670. doi:10.1155/2012/
579670
84. Cancro MP (2009) Signalling crosstalk in B cells: managing worth
and need. Nat Rev Immunol 9(9):657661. doi:10.1038/nri2621
85. Cambier JC, Gauld SB, Merrell KT, Vilen BJ (2007) B-cell aner-
gy: from transgenic models to naturally occurring anergic B cells?
Nat Rev Immunol 7(8):633643. doi:10.1038/nri2133
86. Hartley SB, Crosbie J, Brink R, Kantor AB, Basten A, Goodnow
CC (1991) Elimination from peripheral lymphoid tissues of self-
reactive B lymphocytes recognizing membrane-bound antigens.
Nature 353(6346):765769. doi:10.1038/353765a0
87. Nemazee D (2006) Receptor editing in lymphocyte development
and central tolerance. Nat Rev Immunol 6(10):728740. doi:10.
1038/nri1939
88. Srinivasan L, Sasaki Y, Calado DP, Zhang B, Paik JH, DePinho
RA, Kutok JL, Kearney JF, Otipoby KL, Rajewsky K (2009) PI3
kinase signals BCR-dependent mature B cell survival. Cell
139(3):573586. doi:10.1016/j.cell.2009.08.041
89. Wang LD, Clark MR (2003) B-cell antigen-receptor signalling in
lymphocyte development. Immunol 110(4):411420
90. Mocsai A, Ruland J, Tybulewicz VL (2010) The SYK tyrosine
kinase: a crucial player in diverse biological functions. Nat Rev
Immunol 10(6):387402. doi:10.1038/nri2765
91. Koretzky GA, Abtahian F, Silverman MA (2006) SLP76 and
SLP65: complex regulation of signalling in lymphocytes and be-
yond. Nat Rev Immunol 6(1):6778. doi:10.1038/nri1750
92. Kurosaki T, Hikida M (2009) Tyrosine kinases and their substrates
in B lymphocytes. Immunol Rev 228(1):132148. doi:10.1111/j .
1600-065X.2008.00748.x
93. Oellerich T, Bremes V, Neumann K, Bohnenberger H, Dittmann
K, Hsiao HH, Engelke M, Schnyder T, Batista FD, Urlaub H et al
(2011) The B-cell antigen receptor signals through a preformed
transducer module of SLP65 and CIN85. EMBO J 30(17):3620
3634. doi:10.1038/emboj.2011.251
94. Oh-hora M, Rao A (2008) Calcium signaling in lymphocytes.
Curr Opin Immunol 20(3):250258. doi:10.1016/j.coi.2008.04.
004
95. Gross AJ, Lyandres JR, Panigrahi AK, Prak ET, DeFranco AL
(2009) Developmental acquisition of the Lyn-CD22-SHP-1 inhib-
itory pathway promotes B cell tolerance. J Immunol182(9):5382
5392. doi:10.4049/jimmunol.0803941
96. Cornall RJ, Cyster JG, Hibbs ML, Dunn AR, Otipoby KL, Clark
EA, Goodnow CC (1998) Polygenic autoimmune traits: Lyn,
CD22, and SHP-1 are limiting elements of a biochemical pathway
regulating BCR signaling and selection. Immunity 8(4):497508
97. Lamagna C, Hu Y, DeFranco AL, Lowell CA (2014) B cell-
specific loss of Lyn kinase leads to autoimmunity. J Immunol
192(3):919928. doi:10.4049/jimmunol.1301979
98. Hua Z, Gross AJ, Lamagna C, Ramos-Hernandez N, Scapini P, Ji
M, Shao H, Lowell CA, Hou B, DeFranco AL (2014)
Requirement for MyD88 signaling in B cells and dendritic cells
for germinal center anti-nuclear antibody production in Lyn-
deficient mice. J Immunol 192(3):875885. doi:10.4049/
jimmunol.1300683
99. Lu R, Vidal GS, Kelly JA, Delgado-Vega AM, Howard XK,
Macwana SR, Dominguez N, Klein W, Burrell C, Harley IT
et al (2009) Genetic associations of LYN with systemic lupus
erythematosus. Genes Immun 10(5):397403. doi:10.1038/gene.
2009.19
100. Hermiston ML, Xu Z, Weiss A (2003) CD45: a critical regulator of
signaling thresholds in immune cells. Ann Rev Immunol 21:107
137. doi:10.1146/annurev.immunol.21.120601.140946
101. Irie-Sasaki J, Sasaki T, Matsumoto W, Opavsky A, Cheng M,
Welstead G, Griffiths E, Krawczyk C, Richardson CD, Aitken K
et al (2001) CD45 is a JAK phosphatase and negatively regulates
cytokine receptor signalling. Nature 409(6818):349354. doi:10.
1038/35053086
102. Roach T, Slater S, Koval M, White L, Cahir McFarland ED,
Okumura M, Thomas M, Brown E (1997) CD45 regulates Src
family member kinase activity associated with macrophage
integrin-mediated adhesion. Curr Biol 7(6):408417
103. Piercy J, Petrova S, Tchilian EZ, Beverley PC (2006) CD45 neg-
atively regulates tumour necrosis factor and interleukin-6 produc-
tion in dendritic cells. Immunology 118(2):250256. doi:10.1111/
j.1365-2567.2006.02363.x
104. Gregori S, Mangia P, Bacchetta R, Tresoldi E, Kolbinger F,
Traversari C, Carballido JM, de Vries JE, Korthauer U,
Roncarolo MG (2005) An anti-CD45RO/RB monoclonal anti-
body modulates T cell responses via induction of apoptosis and
generation of regulatory T cells. J Exp Med 201(8):12931305.
doi:10.1084/jem.20040912
105. Brooks WP, Lynes MA (2001) Effects of hemizygous CD45 ex-
pression in the autoimmune Fasl(gld/gld) syndrome. Cell
Immunol 212(1):2434. doi:10.1006/cimm.2001.1845
106. Greer SF, Justement LB (1999) CD45 regulates tyrosine phos-
phorylation of CD22 and its association with the protein tyrosine
phosphatase SHP-1. J Immunol 162(9):52785286
107. Benatar T, Carsetti R, Furlonger C, Kamalia N, Mak T, Paige CJ
(1996) Immunoglobulin-mediated signal transduction in B cells
from CD45-deficient mice. J Exp Med 183(1):329334
108. Surolia I, Pirnie SP, Chellappa V, Taylor KN, Cariappa A, Moya J,
Liu H, Bell DW, Driscoll DR, Diederichs S et al (2010)
Functionally defective germline variants of sialic acid
acetylesterase in autoimmunity. Nature 466(7303):243247. doi:
10.1038/nature09115
109. Cyster JG, Healy JI, Kishihara K, Mak TW, Thomas ML,
Goodnow CC (1996) Regulation of B-lymphocyte negative and
positive selection by tyrosine phosphatase CD45. Nature
381(6580):325328. doi:10.1038/381325a0
110. Aiba Y, Yamazaki T, Okada T, Gotoh K, Sanjo H, Ogata M,
Kurosaki T (2006) BANK negatively regulates Akt activation
and subsequent B cell responses. Immunity 24(3):259268. doi:
10.1016/j.immuni.2006.01.002
111. Begovich AB, Carlton VE, Honigberg LA, Schrodi SJ,
Chokkalingam AP, Alexander HC, Ardlie KG, Huang Q, Smith
AM, Spoerke JM et al (2004) A missense single-nucleotide poly-
morphism in a gene encoding a protein tyrosine phosphatase
(PTPN22) is associated with rheumatoid arthritis. Am J Hum
Genet 75(2):330337. doi:10.1086/422827
Clinic Rev Allerg Immunol (2017) 53:237264 257
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
112. Bottini N, Musumeci L, Alonso A, Rahmouni S, Nika K,
Rostamkhani M, MacMurray J, Meloni GF, Lucarelli P,
Pellecchia M et al (2004) A functional variant of lymphoid tyro-
sine phosphatase is associated with type I diabetes. Nat Genet
36(4):337338. doi:10.1038/ng1323
113. Kyogoku C, Langefeld CD, Ortmann WA, Lee A, Selby S,
Carlton VE, Chang M, Ramos P, Baechler EC, Batliwalla FM
et al (2004) Genetic association of the R620W polymorphism of
protein tyrosine phosphatase PTPN22 with human SLE. Am J
Hum Genet 75(3):504507. doi:10.1086/423790
114. Bootman MD (2012) Calcium signaling. Cold Spring Harb
Perspect Biol 4(7):a011171. doi:10.1101/cshperspect.a011171
115. Matthews SA, Cantrell DA (2009) New insights into the regula-
tion and function of serine/threonine kinases in T lymphocytes.
Immunol Rev 228(1):241252. doi:10.1111/j.1600-065X.2008.
00759.x
116. Im SH, Rao A (2004) Activation and deactivation of gene expres-
sion by Ca2+/calcineurin-NFAT-mediated signaling. Mol Cells
18(1):19
117. Muller MR, Rao A (2010) NFAT, immunity and cancer: a tran-
scription factor comes of age. Nat Rev Immunol 10(9):645656.
doi:10.1038/nri2818
118. Garaud S, Morva A, Lemoine S, Hillion S, Bordron A, Pers JO,
Berthou C, Mageed RA, Renaudineau Y, Youinou P (2011) CD5
promotes IL-10 production in chronic lymphocytic leukemia B
cells through STAT3 and NFAT2 activation. J Immunol 186(8):
48354844. doi:10.4049/jimmunol.1003050
119. Lyubchenko T, dal Porto J, Cambier JC, Holers VM (2005)
Coligation of the B cell receptor with complement receptor type
2 (CR2/CD21) using its natural ligand C3dg: activation without
engagement of an inhibitory signaling pathway. J Immunol
174(6):32643272
120. Yarkoni Y, Getahun A, Cambier JC (2010) Molecular underpin-
ning of B-cell anergy. Immunol Rev 237(1):249263. doi:10.
1111/j.1600-065X.2010.00936.x
121. Gauld SB, Benschop RJ, Merrell KT, Cambier JC (2005)
Maintenance of B cell anergy requires constant antigen receptor
occupancy and signaling. Nat Immunol 6(11):11601167. doi:10.
1038/ni1256
122. Bednarski JJ, Lyssiotis CA, Roush R, Boitano AE, Glick GD,
Opipari AW Jr (2004) A novel benzodiazepine increases the sen-
sitivity of B cells to receptor stimulation with synergistic effects on
calcium signaling and apoptosis. J Biol Chem 279(28):29615
29621. doi:10.1074/jbc.M403507200
123. Dhand R, Hara K, Hiles I, Bax B, Gout I, Panayotou G, Fry MJ,
Yonezawa K, Kasuga M, Waterfield MD (1994) PI 3-kinase:
structural and functional analysis of intersubunit interactions.
EMBO J 13(3):511521
124. Lehmann K, Muller JP, Schlott B, Skroblin P, Barz D, Norgauer J,
Wetzker R (2009) PI3Kgamma controls oxidative bursts in neu-
trophils via interactions with PKCalpha and p47phox. Biochem J
419(3):603610. doi:10.1042/BJ20081268
125. Geering B, Cutillas PR, Nock G, Gharbi SI, Vanhaesebroeck B
(2007) Class IA phosphoinositide 3-kinases are obligate p85-p110
heterodimers. Proc Natl Acad Sci U S A 104(19):78097814. doi:
10.1073/pnas.0700373104
126. Pirola L, Zvelebil MJ, Bulgarelli-Leva G, Van Obberghen E,
Waterfield MD, Wymann MP (2001) Activation loop sequences
confer substrate specificity to phosphoinositide 3-kinase alpha
(PI3Kalpha ). Functions of lipid kinase-deficient PI3Kalpha in
signaling. J Biol Chem 276(24):2154421554. doi:10.1074/jbc.
M011330200
127. Dyson JM, Fedele CG, Davies EM, Becanovic J, Mitchell CA
(2012) Phosphoinositide phosphatases: just as important as the
kinases. Subcell Biochem 58:215279. doi:10.1007/978-94-007-
3012-0_7
128. Finlay D, Cantrell D (2011)The coordination of T-cell function by
serine/threonine kinases. Cold Spring Harb Perspect Biol 3(1):
a002261. doi:10.1101/cshperspect.a002261
129. Laplante M, Sabatini DM (2012) mTOR signaling. Cold Spring
Harb Perspect Biol 4(2). doi:10.1101/cshperspect.a011593
130. Kuo TC, Schlissel MS (2009) Mechanisms controlling expression
of the RAG locus during lymphocyte development. Curr Opin
Immunol 21(2):173178. doi:10.1016/j.coi.2009.03.008
131. Hart GT, Wang X, Hogquist KA, Jameson SC (2011) Kruppel-like
factor 2 (KLF2) regulates B-cell reactivity, subset differentiation,
and trafficking molecule expression. Proc Natl Acad Sci U S A
108(2):716721. doi:10.1073/pnas.1013168108
132. Bi L, Okabe I, Bernard DJ, Nussbaum RL (2002) Early embryonic
lethality in mice deficient in the p110beta catalytic subunit of PI 3-
kinase. Mamm Genome 13(3):169172. doi:10.1007/s00335-
001-2123-x
133. Bilancio A, Okkenhaug K, Camps M, Emery JL, Ruckle T,
Rommel C, Vanhaesebroeck B (2006) Key role of the p110delta
isoform of PI3K in B-cell antigen and IL-4 receptor signaling:
comparative analysis of genetic and pharmacologic interference
with p110delta function in B cells. Blood 107(2):642650. doi:10.
1182/blood-2005-07-3041
134. Clayton E, Bardi G, Bell SE, Chantry D, Downes CP, Gray A,
Humphries LA, Rawlings D, Reynolds H, Vigorito E et al (2002)
A crucial role for the p110delta subunit of phosphatidylinositol 3-
kinase in B cell development and activation. J Exp Med 196(6):
753763
135. Hirsch E, Katanaev VL, Garlanda C, Azzolino O, Pirola L,
Silengo L, Sozzani S, Mantovani A, Altruda F, Wymann MP
(2000) Central role for G protein-coupled phosphoinositide 3-
kinase gamma in inflammation. Science 287(5455):1049
1053
136. Jou ST, Carpino N, Takahashi Y, Piekorz R, Chao JR, Carpino N,
Wang D, Ihle JN (2002) Essential, nonredundant role for the
phosphoinositide 3-kinase p110delta in signaling by the B-cell
receptor complex. Mol Cell Biol 22(24):85808591
137. Sasaki T, Irie-Sasaki J, Jones RG, Oliveira-dos-Santos AJ,
Stanford WL, Bolon B, Wakeham A, Itie A, Bouchard D,
Kozieradzki I et al (2000) Function of PI3Kgamma in thymocyte
development, T cell activation, and neutrophil migration. Science
287(5455):10401046
138. Barber DF, Bartolome A, Hernandez C, Flores JM, Fernandez-
Arias C, Rodriguez-Borlado L, Hirsch E, Wymann M,
Balomenos D, Carrera AC (2006) Class IB-phosphatidylinositol
3-kinase (PI3K) deficiency ameliorates IA-PI3K-induced system-
ic lupus but not T cell invasion. J Immunol 176(1):589593
139. Barber DF, Bartolome A, Hernandez C, Flores JM, Redondo C,
Fernandez-Arias C, Camps M, Ruckle T, Schwarz MK, Rodriguez
S et al (2005) PI3Kgamma inhibition blocks glomerulonephritis
and extends lifespan in a mouse model of systemic lupus. Nat Med
11(9 ):933935. doi:10.1038/nm1291
140. Maxwell MJ, Tsantikos E, Kong AM, Vanhaesebroeck B,
Tarlinton DM, Hibbs ML (2012) Attenuation of phosphoinositide
3-kinase delta signaling restrains autoimmune disease. J
Autoimmun 38(4):381391. doi:10.1016/j.jaut.2012.04.001
141. Suarez-Fueyo A, Barber DF, Martinez-Ara J, Zea-Mendoza AC,
Carrera AC (2011) Enhanced phosphoinositide 3-kinase delta ac-
tivity is a frequent event in systemic lupus erythematosus that
confers resistance to activation-induced T cell death. J Immunol
187(5):23762385. doi:10.4049/jimmunol.1101602
142. Omori SA, Cato MH, Anzelon-Mills A, Puri KD, Shapiro-Shelef
M, Calame K, Rickert RC (2006) Regulation of class-switch re-
combination and plasma cell differentiation by phos-
phatidylinositol 3-kinase signaling. Immunity 25(4):545557.
doi:10.1016/j.immuni.2006.08.015
258 Clinic Rev Allerg Immunol (2017) 53:237264
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
143. Omori SA, Rickert RC (2007) Phosphatidylinositol 3-kinase
(PI3K) signaling and regulation of the antibody response. Cell
Cycle 6(4):397402. doi:10.4161/cc.6.4.3837
144. Okkenhaug K, Bilancio A, Farjot G, Priddle H, Sancho S, Peskett
E, Pearce W, Meek SE, Salpekar A, Waterfield MD et al (2002)
Impaired B and T cell antigen receptor signaling in p110delta PI 3-
kinase mutant mice. Science 297(5583):10311034. doi:10.1126/
science.1073560
145. Durand CA, Hartvigsen K, Fogelstrand L, Kim S, Iritani S,
Vanhaesebroeck B, Witztum JL, Puri KD, Gold MR (2009)
Phosphoinositide 3-kinase p110 delta regulates natural antibody
production, marginal zone and B-1 B cell function, and autoanti-
body responses. J Immunol 183(9):56735684. doi:10.4049/
jimmunol.0900432
146. Beer-Hammer S, Zebedin E, von Holleben M, Alferink J, Reis B,
Dresing P, Degrandi D,Scheu S,Hirsch E, Sexl Vet al (2010) The
catalytic PI3K isoforms p110gamma and p110delta contribute to
B cell development and maintenance, transformation, and prolif-
eration. J Leukoc Biol 87(6):10831095. doi:10.1189/jlb.0809585
147. Dil N, Marshall AJ (2009) Role of phosphoinositide 3-kinase
p110 delta in TLR4- and TLR9-mediated B cell cytokine produc-
tion and differentiation. Mol Immunol 46(10):19701978. doi:10.
1016/j.molimm.2009.03.010
148. Marone R, Cmiljanovic V, Giese B, Wymann MP (2008)
Targeting phosphoinositide 3-kinase: moving towards therapy.
Biochim Biophys Acta 1784(1):159185. doi:10.1016/j.bbapap.
2007.10.003
149. Rommel C, Camps M, Ji H (2007) PI3K delta and PI3K gamma:
partners in crime in inflammation in rheumatoid arthritis and be-
yond? Nat Rev Immunol 7(3):191201. doi:10.1038/nri2036
150. Camps M, Ruckle T, Ji H, Ardissone V, Rintelen F, Shaw J,
Ferrandi C, Chabert C, Gillieron C, Francon B et al (2005)
Blockade of PI3Kgamma suppressesjoint inflammation and dam-
age in mouse models of rheumatoid arthritis. Nat Med 11(9):936
943. doi:10.1038/nm1284
151. Gonzalez-Martin A, Adams BD, Lai M, Shepherd J, Salvador-
Bernaldez M, Salvador JM, Lu J, Nemazee D, Xiao C (2016)
The microRNA miR-148a functions as a critical regulator of B
cell tolerance and autoimmunity. Nat Immunol 17(4):433440.
doi:10.1038/ni.3385
152. Getahun A, Beavers NA, Larson SR, Shlomchik MJ, Cambier JC
(2016) Continuous inhibitory signaling by both SHP-1 and SHIP-
1 pathways is required to maintain unresponsiveness of anergic B
cells. J Exp Med 213(5):751769. doi:10.1084/jem.20150537
153. Hershko A, Ciechanover A (1998) The ubiquitin system. Annu
Rev Biochem 67:425479. doi:10.1146/annurev.biochem.67.1.
425
154. Chen J, Chen ZJ (2013) Regulation of NF-kappaB by
ubiquitination. Curr Opin Immunol 25(1):412. doi:10.1016/j.
coi.2012.12.005
155. Haglund K, Di Fiore PP, Dikic I (2003) Distinct monoubiquitin
signals in receptor endocytosis. Trends Biochem Sci 28(11):598
603. doi:10.1016/j.tibs.2003.09.005
156. Hitotsumatsu O, Ahmad RC, Tavares R, Wang M, Philpott D,
Turer EE, Lee BL, Shiffin N, Advincula R, Malynn BA et al
(2008) The ubiquitin-editing enzyme A20 restricts nucleotide-
binding oligomerization domain containing 2-triggered signals.
Immunity 28(3):381390. doi:10.1016/j.immuni.2008.02.002
157. Duan L, Reddi AL, Ghosh A, Dimri M, Band H (2004) The Cbl
family and other ubiquitin ligases: destructive forces in control of
antigen receptor signaling. Immunity 21(1):717. doi:10.1016/j.
immuni.2004.06.012
158. Bachmaier K, Krawczyk C, Kozieradzki I, Kong YY, Sasaki T,
Oliveira-dos-Santos A, Mariathasan S, Bouchard D, Wakeham A,
Itie A et al (2000) Negative regulation of lymphocyte activation
and autoimmunity by the molecular adaptor Cbl-b. Nature
403(6766):211216. doi:10.1038/35003228
159. Keane MM, Rivero-Lezcano OM, Mitchell JA, Robbins KC,
Lipkowitz S (1995) Cloning and characterization of cbl-b: a
SH3 binding protein with homology to the c-cbl proto-oncogene.
Oncogene 10(12):23672377
160. Lupher ML Jr, Rao N, Eck MJ, Band H (1999) The Cbl
protooncoprotein: a negative regulator of immune receptor signal
transduction. Immunol Today 20(8):375382
161. Ota Y, Beitz LO, Scharenberg AM, Donovan JA, Kinet JP,
Samelson LE (1996) Characterization of Cbl tyrosine phosphory-
lation and a Cbl-Syk complex in RBL-2H3 cells. J Exp Med
184(5):17131723
162. Yasuda T, Maeda A, Kurosaki M, Tezuka T, Hironaka K,
Yamamoto T, Kurosaki T (2000) Cbl suppresses B cell receptor-
mediated phospholipase C (PLC)-gamma2 activation by regulat-
ing B cell linker protein-PLC-gamma2 binding. J Exp Med
191(4):641650
163. Yasuda T, Tezuka T, Maeda A, Inazu T, Yamanashi Y, Gu H,
Kurosaki T, Yamamoto T (2002) Cbl-b positively regulates Btk-
mediated activation of phospholipase C-gamma2 in B cells. J Exp
Med 196(1):5163
164. Rao N, Dodge I, Band H (2002) The Cbl family of ubiquitin
ligases: critical negative regulators of tyrosine kinase signaling
in the immune system. J Leukoc Biol 71(5):753763
165. Joazeiro CA, Wing SS, Huang H, Leverson JD, Hunter T, Liu YC
(1999) The tyrosine kinase negative regulator c-Cbl as a RING-
type, E2-dependent ubiquitin-protein ligase. Science 286(5438):
309312
166. Drake L, McGovern-Brindisi EM, Drake JR (2006) BCR
ubiquitination controls BCR-mediated antigen processing and
presentation. Blood 108(13):40864093. doi:10.1182/blood-
2006-05-025338
167. Katkere B, Rosa S, Caballero A, Repasky EA, Drake JR (2010)
Physiological-range temperature changes modulate cognate anti-
gen processing and presentation mediated by lipid raft-restricted
ubiquitinated B cell receptor molecules. J Immunol 185(9):5032
5039. doi:10.4049/jimmunol.1001653
168. Katkere B, Rosa S, Drake JR (2012) The Syk-binding ubiquitin
ligase c-Cbl mediates signaling-dependent B cell receptor
ubiquitination and B cell receptor-mediated antigen processing
and presentation. J Biol Chem 287(20):1663616644. doi:10.
1074/jbc.M112.357640
169. Zhang M, Veselits M, ONeill S, Hou P, Reddi AL, Berlin I, Ikeda
M, Nash PD, Longnecker R, Band H et al (2007) Ubiquitinylation
of Ig beta dictates the endocytic fate of the B cell antigen receptor.
J Immunol 179(7):44354443
170. Veselits M, Tanaka A, Lipkowitz S, ONeill S, Sciammas R,
Finnegan A, Zhang J, Clark MR (2014) Recruitment of Cbl-b to
B cell antigen receptor couples antigen recognition to Toll-like
receptor 9 activation in late endosomes. PLoS One 9(3):e89792.
doi:10.1371/journal.pone.0089792
171. Boone DL, Turer EE, Lee EG, Ahmad RC, Wheeler MT, Tsui C,
Hurley P, Chien M, Chai S, Hitotsumatsu O et al (2004) The
ubiquitin-modifying enzyme A20 is required for termination of
Toll-like receptor responses. Nat Immunol 5(10):10521060.
doi:10.1038/ni1110
172. Heyninck K, De Valck D, Vanden Berghe W, Van Criekinge W,
Contreras R, Fiers W, Haegeman G, Beyaert R (1999) The zinc
finger protein A20 inhibits TNF-induced NF-kappaB-dependent
gene expression by interfering with an RIP- or TRAF2-mediated
transactivation signal and directly binds to a novel NF-kappaB-
inhibiting protein ABIN. J Cell Biol 145(7):14711482
173. Lee EG, Boone DL, Chai S, Libby SL, Chien M, Lodolce JP, Ma
A (2000) Failure to regulate TNF-induced NF-kappaB and cell
Clinic Rev Allerg Immunol (2017) 53:237264 259
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
death responses in A20-deficient mice. Science 289(5488):2350
2354
174. Shembade N, Ma A, Harhaj EW (2010) Inhibition of NF-kappaB
signaling by A20 through disruption of ubiquitin enzyme com-
plexes. Science 327(5969):11351139. doi:10.1126/science.
1182364
175. Beyaert R, Heyninck K, Van Huffel S (2000) A20 and A20-
binding proteins as cellular inhibitors of nuclear factor-kappa B-
dependent gene expression and apoptosis. Biochem Pharmacol
60(8):11431151
176. Bosanac I, Wertz IE, Pan B, Yu C, Kusam S, Lam C, Phu L, Phung
Q, Maurer B, Arnott D, Kirkpatrick DS et al (2010) Ubiquitin
binding to A20 ZnF4 is required for modulation of NF-kappaB
signaling. Mol Cell 40(4):548557. doi:10.1016/j.molcel.2010.
10.009
177. Lu TT, Onizawa M, Hammer GE, Turer EE, Yin Q, Damko E,
Agelidis A, Shifrin N, Advincula R, Barrera J et al (2013)
Dimerization and ubiquitin mediated recruitment of A20, a com-
plex deubiquitinating enzyme. Immunity 38(5):896905. doi:10.
1016/j.immuni.2013.03.008
178. Skaug B, Chen J, Du F, He J, Ma A, Chen ZJ (2011) Direct,
noncatalytic mechanism of IKK inhibition by A20. Mol Cell
44(4):559571. doi:10.1016/j.molcel.2011.09.015
179. Tokunaga F, Nishimasu H, Ishitani R, Goto E, Noguchi T, Mio K,
Kamei K, Ma A, Iwai K, Nureki O (2012) Specific recognition of
linear polyubiquitin by A20 zinc finger 7 is involved in NF-
kappaB regulation. EMBO J 31(19):38563870. doi:10.1038/
emboj.2012.241
180. Verhelst K, Carpentier I, Kreike M, Meloni L, Verstrepen L,
Kensche T, Dikic I, Beyaert R (2012) A20 inhibits LUBAC-
mediated NF-kappaB activation by binding linear polyubiquitin
chains via its zinc finger 7. EMBO J 31(19):38453855. doi:10.
1038/emboj.2012.240
181. Lin SC, Chung JY, Lamothe B, Rajashankar K, Lu M, Lo YC,
Lam AY, Darnay BG, Wu H (2008) Molecular basis for the unique
deubiquitinating activity of the NF-kappaB inhibitor A20. J Mol
Biol 376(2):526540. doi:10.1016/j.jmb.2007.11.092
182. Wertz IE, ORourke KM, Zhou H, Eby M, Aravind L, Seshagiri S,
Wu P, Wiesmann C, Baker R, Boone DL et al (2004) De-
ubiquitination and ubiquitin ligase domains of A20 downregulate
NF-kappaB signalling. Nature 430(7000):694699. doi:10.1038/
nature02794
183. Chiang YJ, Kole HK, Brown K, Naramura M, Fukuhara S, Hu RJ,
Jang IK, Gutkind JS, Shevach E, Gu H (2000) Cbl-b regulates the
CD28 dependence of T-cell activation. Nature 403:216(6766)
216(67220. doi:10.1038/35003235
184. Rudd CE, Schneider H (2000) Lymphocyte signaling: Cbl sets the
threshold for autoimmunity. Curr Biol 10(9):R344R347
185. Vereecke L, Beyaert R, van Loo G (2009) The ubiquitin-editing
enzyme A20 (TNFAIP3) is a central regulator of immunopathol-
ogy. Trends Immunol 30(8):383391. doi:10.1016/j.it.2009.05.
007
186. Graham RR, Cotsapas C, Davies L, Hackett R, Lessard CJ, Leon
JM, Burtt NP, Guiducci C, Parkin M, Gates C et al (2008) Genetic
variants near TNFAIP3 on 6q23 are associated with systemic lu-
pus erythematosus. Nat Genet 40(9):10591061. doi:10.1038/ng.
200
187. Hughes LB, Reynolds RJ, Brown EE, Kelley JM, Thomson B,
Conn DL, Jonas BL, Westfall AO, Padilla MA, Callahan LF et al
(2010) Most common single-nucleotide polymorphisms associat-
ed with rheumatoid arthritis in persons of European ancestry con-
fer risk of rheumatoid arthritis in African Americans. Arthritis
Rheum 62(12):35473553. doi:10.1002/art.27732
188. Nair RP, Duffin KC, Helms C, Ding J, Stuart PE, Goldgar D,
Gudjonsson JE, Li Y, Tejasvi T, Feng BJ et al (2009) Genome-
wide scan reveals association of psoriasis with IL-23 and NF-
kappaB pathways. Nat Genet 41(2):199204. doi:10.1038/ng.311
189. Eyre S, Hinks A, Bowes J, Flynn E, Martin P, Wilson AG, Morgan
AW, Emery P, Steer S, Hocking LJ, Early Arthritis Consortium
et al (2010) Overlapping genetic susceptibility variants between
three autoimmune disorders: rheumatoid arthritis, type 1 diabetes
and coeliac disease. Arthritis Res Ther 12(5):R175. doi:10.1186/
ar3139
190. Fung EY, Smyth DJ, Howson JM, Cooper JD, Walker NM,
Stevens H, Wicker LS, Todd JA (2009) Analysis of 17 autoim-
mune disease-associated variants in type 1 diabetes identifies
6q23/TNFAIP3 as a susceptibility locus. Genes Immun 10(2):
188191. doi:10.1038/gene.2008.99
191. Dieude P, Guedj M, Wipff J, Ruiz B, Riemekasten G, Matucci-
Cerinic M, Melchers I, Hachulla E, Airo P, Diot E et al (2010)
Association of the TNFAIP3 rs5029939 variant with systemic
sclerosis in the European Caucasian population. Ann Rheum Dis
69(11):19581964. doi:10.1136/ard.2009.127928
192. Koumakis E, Giraud M, Dieude P, Cohignac V, Cuomo G, Airo P,
Hachulla E, Matucci-Cerinic M, Diot E, Caramaschi P et al (2012)
Brief report: Candidate gene study in systemic sclerosis identifies
a rare and functional variant of the TNFAIP3 locus as a risk factor
for polyautoimmunity. Arthritis Rheum 64(8):27462752. doi:10.
1002/art.34490
193. Adrianto I, Wen F, Templeton A, Wiley G, King JB, Lessard CJ,
Bates JS, Hu Y, Kelly JA, Kaufman KM et al (2011) Association
of a functional variant downstream of TNFAIP3 with systemic
lupus erythematosus. Nat Genet 43(3):253258. doi:10.1038/ng.
766
194. Boonyasrisawat W, Eberle D, Bacci S, Zhang YY, Nolan D,
Gervino EV, Johnstone MT, Trischitta V, Shoelson SE, Doria A
(2007) Tag polymorphisms at the A20 (TNFAIP3) locus are asso-
ciated with lower gene expression and increased risk of coronary
artery disease in type 2 diabetes. Diabetes 56(2):499505. doi:10.
2337/db06-0946
195. Elsby LM, Orozco G, Denton J, Worthington J, Ray DW, Donn
RP (2010) Functional evaluation of TNFAIP3 (A20) in rheuma-
toid arthritis. Clin Exp Rheumatol 28(5):708714
196. Koczan D, Drynda S, Hecker M, Drynda A, Guthke R, Kekow J,
Thiesen HJ (2008) Molecular discrimination of responders and
nonresponders to anti-TNF alpha therapy in rheumatoid arthritis
by etanercept. Arthritis Res Ther 10(3):R50. doi:10.1186/ar2419
197. Nocturne G, Tarn J, Boudaoud S, Locke J, Miceli-Richard C,
Hachulla E, Dubost JJ, Bowman S, Gottenberg JE, Criswell LA
et al (2016) Germline variation of TNFAIP3 in primary Sjögrens
syndrome-associated lymphoma. Ann Rheum Dis 75(4):780783.
doi:10.1136/annrheumdis-2015-207731
198. Orozco G, Eyre S, Hinks A, Bowes J, Morgan AW, Wilson AG,
Wordsworth P, Steer S, Hocking L, Consortium U et al (2011)
Study of the common genetic background for rheumatoid arthritis
and systemic lupus erythematosus. Ann Rheum Dis 70(3):463
468. doi:10.1136/ard.2010.137174
199. Wang S, Adrianto I, Wiley GB, Lessard CJ, Kelly JA, Adler AJ,
Glenn SB, Williams AH, Ziegler JT, Comeau ME et al (2012) A
functional haplotype of UBE2L3 confers risk for systemic lupus
erythematosus. Genes Immun 13(5):380387. doi:10.1038/gene.
2012.6
200. Lewis M, Vyse S, Shields A, Boeltz S, Gordon P, Spector T,
Lehner P, Walczak H, Vyse T (2015) Effect of UBE2L3 genotype
on regulation of the linear ubiquitin chain assembly complex in
systemic lupus erythematosus. Lancet 385(Suppl 1):S9. doi:10.
1016/S0140-6736(15)60324-5
201. Browne EP (2012) Regulation of B-cell responses by Toll-like
receptors. Immunology 136(4):370379. doi:10.1111/j .1365-
2567.2012.03587.x
260 Clinic Rev Allerg Immunol (2017) 53:237264
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
202. Kawagoe T, Sato S, Matsushita K, Kato H, Matsui K, Kumagai Y,
Saitoh T, Kawai T, Takeuchi O, Akira S (2008) Sequential control
of Toll-like receptor-dependent responses by IRAK1 and IRAK2.
Nat Immunol 9(6):684691. doi:10.1038/ni.160
203. Negishi H, Fujita Y, Yanai H, Sakaguchi S, Ouyang X, Shinohara
M, Takayanagi H, Ohba Y, Taniguchi T, Honda K (2006)
Evidence for licensing of IFN-gamma-induced IFN regulatory
factor 1 transcription factor by MyD88 in Toll-like receptor-de-
pendent gene induction program. Proc Natl Acad Sci U S A
103(41):1513615141. doi:10.1073/pnas.0607181103
204. Schoenemeyer A, Barnes BJ, Mancl ME, Latz E, Goutagny N,
Pitha PM, Fitzgerald KA, Golenbock DT (2005) The interferon
regulatory factor, IRF5, is a central mediator of toll-like receptor 7
signaling. J Biol Chem 280(17):1700517012. doi:10.1074/jbc.
M412584200
205. Takaoka A, Yanai H, Kondo S, Duncan G, Negishi H, Mizutani T,
Kano S, Honda K, Ohba Y, Mak TW et al (2005) Integral role of
IRF-5 in the gene induction programme activated by Toll-like
receptors. Nature 434(7030):243249. doi:10.1038/nature03308
206. Jackson SW, Kolhatkar NS, Rawlings DJ (2015) B cells take the
front seat: dysregulated B cell signals orchestrate loss of tolerance
and autoantibody production. Curr Opin Immunol 33:7077. doi:
10.1016/j.coi.2015.01.018
207. Suurmond J, Diamond B (2015) Autoantibodies in systemic auto-
immune diseases: specificity and pathogenicity. J Clin Invest
125(6):21942202. doi:10.1172/JCI78084
208. Chaturvedi A, Dorward D, Pierce SK (2008) The B cell receptor
governs the subcellular location of Toll-like receptor 9 leading to
hyperresponses to DNA-containing antigens. Immunity 28(6):
799809. doi:10.1016/j.immuni.2008.03.019
209. Chaturvedi A, Pierce SK (2009) How location governs toll-like
receptor signaling. Traffic 10(6):621628. doi:10.1111/j.1600-
0854.2009.00899.x
210. Berland R, Fernandez L, Kari E, Han JH, Lomakin I, Akira S,
Wortis HH, Kearney JF, Ucci AA, Imanishi-Kari T (2006) Toll-
like receptor 7-dependent loss of B cell tolerance in pathogenic
autoantibody knockin mice. Immunity 25(3):429440. doi:10.
1016/j.immuni.2006.07.014
211. Christensen SR, Kashgarian M, Alexopoulou L, Flavell RA, Akira
S, Shlomchik MJ (2005) Toll-like receptor 9 controls anti-DNA
autoantibody production in murine lupus. J Exp Med 202(2):321
331. doi:10.1084/jem.20050338
212. Christensen SR, Shupe J, Nickerson K, Kashgarian M, Flavell
RA, Shlomchik MJ (2006) Toll-like receptor 7 and TLR9 dictate
autoantibody specificity and have opposing inflammatory and reg-
ulatory roles in a murine model of lupus. Immunity 25(3):417
428. doi:10.1016/j.immuni.2006.07.013
213. Lartigue A, Courville P, Auquit I, Francois A, Arnoult C, Tron F,
Gilbert D, Musette P (2006) Role of TLR9 in anti-nucleosome and
anti-DNA antibody production in lpr mutation-induced murine
lupus. J Immunol 177(2):13491354
214. Jackson SW, Scharping NE, Kolhatkar NS, Khim S, Schwartz
MA, Li QZ, Hudkins KL, Alpers CE, Liggitt D, Rawlings DJ
(2014) Opposing impact of B cell-intrinsic TLR7 and TLR9 sig-
nals on autoantibody repertoire and systemic inflammation. J
Immunol 192(10):45254532. doi:10.4049/jimmunol.1400098
215. Sullivan KE, Mullen CA, Blaese RM, Winkelstein JA (1994) A
multiinstitutional survey of the Wiskott-Aldrich syndrome. J
Pediatr 125(6 Pt 1):876885
216. Becker-Herman S, Meyer-Bahlburg A, Schwartz MA, Jackson
SW, Hudkins KL, Liu C, Sather BD, Khim S, Liggitt D, Song
W et al (2011) WASp-deficient B cells play a critical, cell-intrinsic
role in triggering autoimmunity. J Exp Med 208(10):20332042.
doi:10.1084/jem.20110200
217. Ehlers M, Fukuyama H, McGaha TL, Aderem A, Ravetch JV
(2006) TLR9/MyD88 signaling is required for class switching to
pathogenic IgG2a and 2b autoantibodies in SLE. J Exp Med
203(3):553561
218. Groom JR, Fletcher CA, Walters SN, Grey ST, Watt SV, Sweet
MJ, Smyth MJ, Mackay CR, Mackay F (2007) BAFF and MyD88
signals promote a lupus like disease independent of T cells. J Exp
Med 204(8):19591971. doi:10.1084/jem.20062567
219. Teichmann LL, Schenten D, Medzhitov R, Kashgarian M,
Shlomchik MJ (2013) Signals via the adaptor MyD88 in B cells
and DCs make distinct and synergistic contributions to immune
activation and tissue damage in lupus. Immunity 38(3):528540.
doi:10.1016/j.immuni.2012.11.017
220. Deane JA, Pisitkun P, Barrett RS, Feigenbaum L, Town T, Ward
JM, Flavell RA, Bolland S (2007) Control of toll-like receptor 7
expression is essential to restrict autoimmunity and dendritic cell
proliferation. Immunity 27(5):801810. doi:10.1016/j.immuni.
2007.09.009
221. Walsh ER, Pisitkun P, Voynova E, Deane JA, Scott BL, Caspi RR,
Bolland S (2012) Dual signaling by innate and adaptive immune
receptors is required for TLR7-induced B-cell-mediated autoim-
munity. Proc Natl Acad Sci U S A 109(40):1627616281. doi:10.
1073/pnas.1209372109
222. Sun X, Wiedeman A, Agrawal N, Teal TH, Tanaka L, Hudkins
KL, Alpers CE, Bolland S, Buechler MB, Hamerman JA et al
(2013) Increased ribonuclease expression reduces inflammation
and prolongs survival in TLR7 transgenic mice. J Immunol
190(6):25362543. doi:10.4049/jimmunol.1202689
223. Garcia-Ortiz H, Velazquez-Cruz R, Espinosa-Rosales F, Jimenez-
Morales S, Baca V, Orozco L (2010) Association of TLR7 copy
number variation with susceptibility to childhood-onset systemic
lupus erythematosus in Mexican population. Ann Rheum Dis
69(10):18611865. doi:10.1136/ard.2009.124313
224. Kawasaki A, Furukawa H, Kondo Y, Ito S, Hayashi T, Kusaoi M,
Matsumoto I, Tohma S, Takasaki Y, Hashimoto H et al (2011)
TLR7 single-nucleotide polymorphisms in the 3untranslated re-
gion and intron 2 independently contribute to systemic lupus ery-
thematosus in Japanese women: a case-control association study.
Arthritis Res Ther 13(2):R41. doi:10.1186/ar3277
225. Shen N, Fu Q, Deng Y, Qian X, Zhao J, Kaufman KM, Wu YL, Yu
CY, Tang Y, Chen JY et al (2010) Sex-specific association of X-
linked Toll-like receptor 7 (TLR7) with male systemic lupus ery-
thematosus. Proc Natl Acad Sci U S A 107(36):1583815843. doi:
10.1073/pnas.1001337107
226. Konsta OD, Le Dantec C, Charras A, Brooks WH, Arleevskaya
MI, Bordron A, Renaudineau Y (2015) An in silico approach
reveals associations between genetic and epigenetic factors within
regulatory elements in B cells from primary Sjögrens syndrome
patients. Front Immunol 6:437. doi:10.3389/fimmu.2015.00437
227. Desnues B, MacedoAB, Roussel-Queval A, Bonnardel J, Henri S,
Demaria O, Alexopoulou L (2014) TLR8 on dendritic cells and
TLR9 on B cells restrain TLR7-mediated spontaneous autoimmu-
nity in C57BL/6 mice. Proc Natl Acad Sci U S A 111(4):1497
1502. doi:10.1073/pnas.1314121111
228. Isnardi I, Ng YS, Srdanovic I, Motaghedi R, Rudchenko S, von
Bernuth H, Zhang SY, Puel A, Jouanguy E, Picard C et al (2008)
IRAK-4- and MyD88-dependent pathways are essential for the
removal of developing autoreactive B cells in humans. Immunity
29(5):746757. doi:10.1016/j.immuni.2008.09.015
229. Kurosaki T, Shinohara H, Baba Y (2010) B cell signaling and fate
decision. Annu Rev Immunol 28:2155. doi:10.1146/annurev.
immunol.021908.132541
230. Mageed RA, Garaud S, Taher TE, Parikh K, Pers JO, Jamin C,
Renaudineau Y, Youinou P (2012) CD5 expression promotes mul-
tiple intracellular signaling pathways in B lymphocyte.
Autoimmun Rev 11(11):795798. doi:10.1016/j.autrev.2012.02.
007
Clinic Rev Allerg Immunol (2017) 53:237264 261
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
231. Renaudineau Y, Hillion S, Saraux A, Mageed RA, Youinou P
(2005) An alternative exon 1 of the CD5 gene regulates CD5
expression in human B lymphocytes. Blood 106(8):27812789.
doi:10.1182/blood-2005-02-0597
232. Youinou P, Renaudineau Y (2007) The paradox of CD5-
expressing B cells in systemic lupus erythematosus. Autoimmun
Rev 7(2):149154. doi:10.1016/j.autrev.2007.02.016
233. Daeron M, Lesourne R (2006) Negative signaling in Fc receptor
complexes. Adv Immunol 89:3986. doi:10.1016/S0065-
2776(05)89002-9
234. Parry RV, Harris SJ, Ward SG (2010) Fine tuning T lymphocytes:
a role for the lipid phosphatase SHIP-1. Biochim Biophys Acta
1804(3):592597. doi:10.1016/j.bbapap.2009.09.019
235. Séïté JF, Cornec D, Renaudineau Y, Youinou P, Mageed RA,
Hillion S (2010) IVIg modulates BCR signaling through CD22
and promotes apoptosis in mature human B lymphocytes. Blood
116(10):16981704. doi:10.1182/blood-2009-12-261461
236. Kumanogoh A, Watanabe C, Lee I, Wang X, Shi W, Araki H,
Hirata H, Iwahori K, Uchida J, Yasui T et al (2000)
Identification of CD72 as a lymphocyte receptor for the class IV
semaphorin CD100: a novel mechanism for regulating B cell sig-
naling. Immunity 13(5):621631
237. Wu Y, Nadler MJ, Brennan LA, Gish GD, Timms JF, Fusaki N,
Jongstra-Bilen J, Tada N, Pawson T, Wither J et al (1998) The B-
cell transmembrane protein CD72 binds to and is an in vivo sub-
strate of the protein tyrosine phosphatase SHP-1. Curr Biol 8(18):
10091017
238. Li DH, Tung JW, Tarner IH, Snow AL, Yukinari T,
Ngernmaneepothong R, Martinez OM, Parnes JR (2006) CD72
down-modulates BCR-induced signal transduction and dimin-
ishes survival in primary mature B lymphocytes. J Immunol
176(9):53215328
239. Shi W, Kumanogoh A, Watanabe C, Uchida J, Wang X, Yasui T,
Yukawa K, Ikawa M, Okabe M, Parnes JR et al (2000) The class
IV semaphorin CD100 plays nonredundant roles in the immune
system: defective B and T cell activation in CD100-deficient mice.
Immunity 13(5):633642
240. Early GS, Zhao W, Burns CM (1996) Anti-CD40 ligand antibody
treatment prevents the development of lupus-like nephritis in a
subset of New Zealand black x New Zealand white mice.
Response correlates with the absence of an anti-antibody response.
J Immunol 157(7):31593164
241. Ma J, Xu J, Madaio MP, Peng Q, Zhang J, Grewal IS, Flavell RA,
Craft J (1996) Autoimmune lpr/lpr mice deficient in CD40 ligand:
spontaneous Ig class switching with dichotomy of autoantibody
responses. J Immunol 157(2):417426
242. Wang X, Huang W, Schiffer LE, Mihara M, Akkerman A,
Hiromatsu K, Davidson A (2003) Effects of anti-CD154 treatment
on B cells in murine systemic lupus erythematosus. Arthritis
Rheum 48(2):495506. doi:10.1002/art.10929
243. Kyburz D, Corr M, Brinson DC, Von Damm A, Tighe H, Carson
DA (1999) Human rheumatoid factor production is dependent on
CD40 signaling and autoantigen. J Immunol 163(6):31163122
244. Tellander AC, Michaelsson E, Brunmark C, Andersson M (2000)
Potent adjuvant effect by anti-CD40 in collagen-induced arthritis.
Enhanced disease is accompanied by increased production of col-
lagen type-II reactive IgG2a and IFN-gamma. J Autoimmun
14(4):295302. doi:10.1006/jaut.2000.0374
245. Baker RL, Wagner DH Jr, Haskins K (2008) CD40 on NOD CD4
T cells contributes to their activation and pathogenicity. J
Autoimmun 31(4):385392. doi:10.1016/j.jaut.2008.09.001
246. Balasa B, Krahl T, Patstone G, Lee J, Tisch R, McDevitt HO,
Sarvetnick N (1997) CD40 ligand-CD40 interactions are neces-
sary for the initiation of insulitis and diabetes in nonobese diabetic
mice. J Immunol 159(9):46204627
247. Bour-Jordan H, Salomon BL, Thompson HL, Szot GL, Bernhard
MR, Bluestone JA (2004) Costimulation controls diabetes by al-
tering the balance of pathogenic and regulatory T cells. J Clin
Invest 114(7):979987. doi:10.1172/JCI20483
248. Green EA, Wong FS, Eshima K, Mora C, Flavell RA (2000)
Neonatal tumor necrosis factor alpha promotes diabetes in
nonobese diabetic mice by CD154-independent antigen presenta-
tion to CD8(+) T cells. J Exp Med 191(2):225238
249. Gerritse K, Laman JD, Noelle RJ, Aruffo A, Ledbetter JA,
Boersma WJ, Claassen E (1996) CD40-CD40 ligand interactions
in experimental allergic encephalomyelitis and multiple sclerosis.
Proc Natl Acad Sci U S A 93(6):24992504
250. Howard LM, Miga AJ, Vanderlugt CL, Dal Canto MC, Laman JD,
Noelle RJ, Miller SD (1999) Mechanisms of immunotherapeutic
intervention by anti-CD40L (CD154) antibody in an animal model
of multiple sclerosis. J Clin Invest 103(2):281290. doi:10.1172/
JCI5388
251. Boon L, Brok HP, Bauer J, Ortiz-Buijsse A, Schellekens MM,
Ramdien-Murli S, Blezer E, van Meurs M, Ceuppens J, de Boer
M et al (2001) Prevention of experimental autoimmune encepha-
lomyelitis in the common marmoset (Callithrix jacchus)usinga
chimeric antagonist monoclonal antibody against human CD40 is
associated with altered B cell responses. J Immunol 167(5):2942
2949
252. Boon L, Laman JD, Ortiz-Buijsse A, den Hartog MT, Hoffenberg
S, Liu P, Shiau F, de Boer M (2002) Preclinical assessment of anti-
CD40 Mab 5D12 in cynomolgus monkeys. Toxicology 174(1):
5365
253. Laman JD, t Hart BA, Brok H, Meurs M, Schellekens MM,
Kasran A, Boon L, Bauer J, Boer M, Ceuppens J (2002)
Protection of marmoset monkeys against EAE by treatment with
a murine antibody blocking CD40 (mu5D12). Eur J Immunol
32(8):22182228. doi:10.1002/1521-4141(200208)32:8<2218::
AID-IMMU2218>3.0.CO;2-0
254. Boumpas DT, Furie R, Manzi S, Illei GG, Wallace DJ, Balow JE,
Vaishnaw A, Group BGLNT (2003) A short course of BG9588
(anti-CD40 ligand antibody) improves serologic activity and de-
creases hematuria in patients with proliferative lupus glomerulo-
nephritis. Arthritis Rheum 48(3):719727. doi:10.1002/art.10856
255. Ford ML, Adams AB, Pearson TC (2004) Targeting co-
stimulatory pathways: transplantation and autoimmunity. Nat
Rev Nephrol 10(1):1424. doi:10.1038/nrneph.2013.183
256. Kalunian KC, Davis JC Jr, Merrill JT, Totoritis MC, Wofsy D,
Group I-LS (2002) Treatment of systemic lupus erythematosus
by inhibition of T cell costimulation with anti-CD154: a random-
ized, double-blind, placebo-controlled trial. Arthritis Rheum
46(12):32513258. doi:10.1002/art.10681
257. Zouali M, Sarmay G (2004) B lymphocyte signaling pathways in
systemic autoimmunity: implications for pathogenesis and treat-
ment. Arthritis Rheum 50(9):27302741. doi:10.1002/art.20487
258. Jiang Y, Hirose S, Sanokawa-Akakura R, Abe M, Mi X, Li N,
Miura Y, Shirai J, Zhang D, Hamano Y, Shirai T (1999)
Genetically determined aberrant down-regulation of
FcgammaRIIB1 in germinal center B cells associated with
hyper-IgG and IgG autoantibodies in murine systemic lupus ery-
thematosus. Int Immunol 11(10):16851691
259. Li DH, Winslow MM, Cao TM, Chen AH, Davis CR, Mellins ED,
Utz PJ, Crabtree GR, Parnes JR (2008) Modulation of peripheral
B cell tolerance by CD72 in a murine model. Arthritis Rheum
58(10):31923204. doi:10.1002/art.23812
260. Healy JI, Dolmetsch RE, Timmerman LA, Cyster JG, Thomas
ML, Crabtree GR, Lewis RS, Goodnow CC (1997) Different nu-
clear signals are activated by the B cell receptor during positive
versus negative signaling. Immunity 6(4):419428
261. Dolmetsch RE, Lewis RS, Goodnow CC, Healy JI (1997)
Differential activation of transcription factors induced by Ca2+
262 Clinic Rev Allerg Immunol (2017) 53:237264
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
response amplitude and duration. Nature 386(6627):855858.
doi:10.1038/386855a0
262. Chiles TC (2004) Regulation and function of cyclin D2 in B lym-
phocyte subsets. J Immunol 173(5):29012907
263. Winslow MM, Gallo EM, Neilson JR, Crabtree GR (2006) The
calcineurin phosphatase complex modulates immunogenic B cell
responses. Immunity 24(2):141152. doi:10.1016/j.immuni.2005.
12.013
264. Ghoreschi K, Laurence A, OShea JJ (2009) Janus kinases in
immune cell signaling. Immunol Rev 228(1):273287. doi:10.
1111/j.1600-065X.2008.00754.x
265. Harrison DA (2012) The Jak/STAT pathway. Cold Spring Harb
Perspect Biol 4(3). doi:10.1101/cshperspect.a011205
266. Cantrell DA (2003) GTPases and T cell activation. Immunol Rev
192:122130
267. Okkenhaug K, Fruman DA (2010) PI3Ks in lymphocyte signaling
and development. Curr Top Microbiol Immunol 346:5785. doi:
10.1007/82_2010_45
268. OShea JJ, Paul WE (2010) Mechanisms underlying lineage com-
mitment and plasticity of helper CD4+ Tcells. Science 327(5969):
10981102. doi:10.1126/science.1178334
269. Batten M, Groom J, Cachero TG, Qian F, Schneider P, Tschopp J,
Browning JL, Mackay F (2000) BAFF mediates survival of pe-
ripheral immature B lymphocytes. J Exp Med 192(10):14531466
270. Schiemann B, Gommerman JL, Vora K, Cachero TG, Shulga-
Morskaya S, Dobles M, Frew E, Scott ML (2001) An essential
role for BAFF in the normal development of B cells through a
BCMA-independent pathway. Science 293(5537):21112114.
doi:10.1126/science.1061964
271. Marsters SA, Yan M, Pitti RM, Haas PE, Dixit VM, Ashkenazi A
(2000) Interaction of the TNF homologues BLyS and APRIL with
the TNF receptor homologues BCMA and TACI. Curr Biol
10(13):785788
272. Thompson JS, Bixler SA, Qian F, Vora K, Scott ML, Cachero TG,
Hession C, Schneider P, Sizing ID, Mullen C et al (2001) BAFF-
R, a newly identified TNF receptor that specifically interacts with
BAFF. Science 293(5537):21082111. doi:10.1126/science.
1061965
273. Fairfax KA, Tsantikos E, Figgett WA, Vincent FB, Quah PS,
LePage M, Hibbs ML, Mackay F (2015) BAFF-driven autoim-
munity requires CD19 expression. J Autoimmun 62:110. doi:10.
1016/j.jaut.2015.06.001
274. Elgueta R, Benson MJ, de Vries VC, Wasiuk A, Guo Y, Noelle RJ
(2009) Molecular mechanism and function of CD40/CD40L en-
gagement in the immune system. Immunol Rev 229(1):152172.
doi:10.1111/j.1600-065X.2009.00782.x
275. Rickert RC, Jellusova J, Miletic AV (2011) Signaling by the tumor
necrosis factor receptor superfamily in B-cell biology and disease.
Immunol Rev 244(1):115133. doi:10.1111/j.1600-065X.2011.
01067.x
276. Brightbill HD, Jackman JK, Suto E, Kennedy H, Jones C 3rd,
Chalasani S, Lin Z, Tam L, Roose-Girma M, Balazs M et al
(2015) Conditional deletion of NF-kappaB-inducing kinase
(NIK) in adult mice disrupts mature B cell survival and activation.
J Immunol 195(3):953964. doi:10.4049/jimmunol.1401514
277. Mackay F, Schneider P (2009) Cracking the BAFF code. Nat Rev
Immunol 9(7):491502. doi:10.1038/nri2572
278. Otipoby KL, Sasaki Y, Schmidt-Supprian M, Patke A, Gareus R,
Pasparakis M, Tarakhovsky A, Rajewsky K (2008) BAFF acti-
vates Akt and Erk through BAFF-R in an IKK1-dependent man-
ner in primary mouse B cells. Proc Natl Acad Sci U S A 105(34):
1243512438. doi:10.1073/pnas.0805460105
279. Hobeika E, Levit-Zerdoun E, Anastasopoulou V, Pohlmeyer R,
Altmeier S, Alsadeq A, Dobenecker MW, Pelanda R, Reth M
(2015) CD19 and BAFF-R can signal to promote B-cell survival
in the absence of Syk. EMBO J 34(7):925939. doi:10.15252/
embj.201489732
280. Schweighoffer E, Vanes L, Nys J, Cantrell D, McCleary S,
Smithers N, Tybulewicz VL (2013) The BAFF receptor trans-
duces survival signals by co-opting the B cell receptor signaling
pathway. Immunity 38(3):475488. doi:10.1016/j.immuni.2012.
11.015
281. Lund FE (2008) Cytokine-producing B lymphocyteskey regu-
lators of immunity. Curr Opin Immunol 20(3):332338. doi:10.
1016/j.coi.2008.03.003
282. Dang VD, Hilgenberg E, Ries S, Shen P, Fillatreau S (2014) From
the regulatory functions of B cells tothe identification of cytokine-
producing plasma cell subsets. Curr Opin Immunol 28:7783. doi:
10.1016/j.coi.2014.02.009
283. Shen P, Roch T, Lampropoulou V, OConnor RA, Stervbo U,
Hilgenberg E, Ries S, Dang VD, Jaimes Y, Daridon C et al
(2014) IL-35-producing B cells are critical regulators of immunity
during autoimmune and infectious diseases. Nature 507(7492):
366370. doi:10.1038/nature12979
284. Sweet RA, Lee SK, Vinuesa CG (2012) Developing connections
amongst key cytokines and dysregulated germinal centers in au-
toimmunity. Curr Opin Immunol 24(6):658664. doi:10.1016/j.
coi.2012.10.003
285. Blenman KR, Duan B, Xu Z, Wan S, Atkinson MA, Flotte TR,
Croker BP, Morel L (2006) IL-10 regulation of lupus in the
NZM2410 murine model. Lab Investig 86(11):11361148. doi:
10.1038/labinvest.3700468
286. Llorente L, Richaud-Patin Y, Garcia-Padilla C, Claret E, Jakez-
Ocampo J, Cardiel MH, Alcocer-Varela J, Grangeot-Keros L,
Alarcon-Segovia D, Wijdenes J et al (2000) Clinical and biologic
effects of anti-interleukin-10 monoclonal antibody administration
in systemic lupus erythematosus. Arthritis Rheum 43(8):1790
1800. doi:10.1002/1529-0131(200008)43:8<1790::AID-
ANR15>3.0.CO;2-2
287. Watanabe R, Ishiura N, Nakashima H, Kuwano Y, Okochi H,
Tamaki K, Sato S, Tedder TF, Fujimoto M (2010) Regulatory B
cells (B10 cells) have a suppressive role in murine lupus: CD19
and B10 cell deficiency exacerbates systemic autoimmunity. J
Immunol 184(9):48014809. doi:10.4049/jimmunol.0902385
288. Capper ER, Maskill JK, Gordon C, Blakemore AI (2004)
Interleukin (IL)-10, IL-1ra and IL-12 profiles in active and quies-
cent systemic lupus erythematosus: could longitudinal studies re-
veal patient subgroups of differing pathology? Clin Exp Immunol
138(2):348356. doi:10.1111/j.1365-2249.2004.02607.x
289. Braun D, Caramalho I, Demengeot J (2002) IFN-alpha/beta en-
hances BCR-dependent B cell responses. Int Immunol 14(4):411
419
290. Graham RR, Kozyrev SV, Baechler EC, Reddy MV, Plenge RM,
Bauer JW, Ortmann WA, Koeuth T, Gonzalez Escribano MF et al
(2006) A common haplotype of interferon regulatory factor 5
(IRF5) regulates splicing and expression and is associated with
increased risk of systemic lupus erythematosus. Nat Genet 38(5):
550555. doi:10.1038/ng1782
291. Ronnblom L, Eloranta ML, Alm GV (2006) The type I interferon
system in systemic lupus erythematosus. Arthritis Rheum 54(2):
408420. doi:10.1002/art.21571
292. Nguyen KB, Watford WT, Salomon R, Hofmann SR, Pien GC,
Morinobu A, Gadina M, OShea JJ, Biron CA (2002) Critical role
for STAT4 activation by type 1 interferons in the interferon-
gamma response to viral infection. Science 297(5589):2063
2066. doi:10.1126/science.1074900
293. Graninger WB, Hassfeld W, Pesau BB, Machold KP, Zielinski
CC, Smolen JS (1991) Induction of systemic lupus erythematosus
by interferon-gamma in a patient with rheumatoid arthritis. J
Rheumatol 18(10):16211622
Clinic Rev Allerg Immunol (2017) 53:237264 263
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
294. Jackson SW, Jacobs HM, Arkatkar T, Dam EM, Scharping NE,
Kolhatkar NS, Hou B, Buckner JH, Rawlings DJ (2016) B cell
IFN-gamma receptor signaling promotes autoimmune germinal
centers via cell-intrinsic induction of BCL-6. J Exp Med 213(5):
733750. doi:10.1084/jem.20151724
295. Ettinger R, Kuchen S, Lipsky PE (2008) Interleukin 21 as a target
of intervention in autoimmune disease. Ann Rheum Dis 67(Suppl
3):iii83iii86. doi:10.1136/ard.2008.098400
296. Moisini I, Davidson A (2009) BAFF: a local and systemic target in
autoimmunediseases. Clin Exp Immunol 158(2):155163. doi:10.
1111/j.1365-2249.2009.04007.x
297. Mackay F, Woodcock SA, Lawton P, Ambrose C, Baetscher M,
Schneider P, Tschopp J, Browning JL (1999) Mice transgenic for
BAFF develop lymphocytic disorders along with autoimmune
manifestations. J Exp Med 190(11):16971710
298. Lesley R, Xu Y, Kalled SL, Hess DM, Schwab SR, Shu HB,
Cyster JG (2004) Reduced competitiveness of autoantigen-
engaged B cells due to increased dependence on BAFF.
Immunity 20(4):441453
299. Thien M, Phan TG, Gardam S, Amesbury M, Basten A, Mackay
F, Brink R (2004) Excess BAFF rescues self-reactive B cells from
peripheral deletion and allows them to enter forbidden follicular
and marginal zone niches. Immunity 20(6):785798. doi:10.1016/
j.immuni.2004.05.010
300. Haapaniemi EM, Kaustio M, Rajala HL, van Adrichem AJ,
Kainulainen L, Glumoff V, Doffinger R, Kuusanmaki H,
Heiskanen-Kosma T, Trotta L et al (2015) Autoimmunity,
hypogammaglobulinemia, lymphoproliferation, and mycobacteri-
al disease in patients with activating mutations in STAT3. Blood
125(4):639648. doi:10.1182/blood-2014-04-570101
301. Kotkowska A, Sewerynek E, Domańska D, Pastuszak-
Lewandoska D, Brzeziańska E (2015) Single nucleotide polymor-
phisms in the STAT3 gene influence AITD susceptibility, thyroid
autoantibody levels, and IL6 and IL17 secretion. Cell Mol Biol
Lett 20(1):88101. doi:10.1515/cmble-2015-0004
302. Nakayamada S, Kubo S, Iwata S, Tanaka Y (2016) Chemical JAK
inhibitors for the treatment of rheumatoid arthritis. Expert Opin
Pharmacother. doi:10.1080/14656566.2016.1241237
303. Deng GM, Kyttaris VC, Tsokos GC (2016) Targeting Syk in au-
toimmune rheumatic diseases. Front Immunol 7(7):7882. doi:10.
3389/fimmu.2016.00078
304. Schwartz DM, Bonelli M, Gadina M, OShea JJ (2016) Type I/II
cytokines, JAKs, and new strategies for treating autoimmune dis-
eases. Nat Rev Rheumatol 12(1):2536. doi:10.1038/nrrheum.
2015.167
305. Lazaro E, Scherlinger M, Truchetet ME, Chiche L, Schaeverbeke
T, Blanco P, Richez C (2016) Biotherapies in systemic lupus ery-
thematosus: new targets. Joint Bone Spine 20(16):3012330133.
doi:10.1016/j.jbspin.2016.07.004
306. Lazzerini PE, Capecchi PL, Guidelli GM, Selvi E, Acampa M,
Laghi-Pasini F (2016) Spotlight on sirukumab for the treatment of
rheumatoid arthritis: the evidence to date. Drug Des Devel Ther
26(10):30833098
307. Burmester GR, Rigby WF, van Vollenhoven RF, Kay J, Rubbert-
Roth A, Kelman A, Dimonaco S, Mitchell N (2016) Tocilizumab
in early progressive rheumatoid arthritis: FUNCTION, a
randomised controlled trial. Ann Rheum Dis 75(6):10811091.
doi:10.1136/annrheumdis-2015-207628
308. Oon S, Wilson NJ, Wicks I (2016) Targeted therapeutics in SLE:
emerging strategies to modulate the interferon pathway. Clin
Transl Immunology 5(5):e79. doi:10.1038/cti.2016.26
309. Achek A, Yesudhas D, Choi S (2016) Toll-like receptors: promis-
ing therapeutic targets for inflammatory diseases. Ach Pharm Res
39(8):10321049. doi:10.1007/s12272-016-0806-9
310. Chalmers SA, Doerner J, Bosanac T, Khalil S, Smith D, Harcken
C, Dimock J, Der E, Herlitz L, Webb D et al (2016) Therapeutic
blockade of immune complex-mediated glomerulonephritis by
highly selective inhibition of Brutons tyrosine kinase. Sci Rep
6:26164. doi:10.1038/srep26164
311. Park JK, Byun JY, Park JA, Kim YY, Lee YJ, Oh JI, Jang SY, Kim
YH, Song YW, Son J et al (2015) HM71224, a novel Brutons
tyrosine kinase inhibitor, suppresses B cell and monocyte activa-
tion and ameliorates arthritis in a mouse model: a potential drug
for rheumatoid arthritis. Arthritis Res Ther 18:91. doi:10.1186/
s13075-016-0988-z
264 Clinic Rev Allerg Immunol (2017) 53:237264
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com
... of bystander-activated CD8 + T cells is their ability to be activated independently of TCR signaling (5)(6)(7)(8)(9)(10). ...
Preprint
Full-text available
Background: Liver cirrhosis could lead to immune dysfunction. During the pathogenesis of liver cirrhosis, CD8⁺ T cells play a critical role. While CD38⁺HLA-DR⁺CD8⁺ T cells, also called bystander activation CD8⁺ T cells, had been shown to be involved in host injury, its specific contribution to liver cirrhosis had remained not unclear. The aim of this study was to understand how these CD38⁺HLA-DR⁺CD8⁺ T cells exerted a pathogenic role in liver cirrhosis. Methods: Flow cytometry was performed to detect the immunophenotype, antigen-specific T cells, cytokines secretion, and cytotoxicity related indicators of CD38⁺HLA-DR⁺CD8⁺ T cells. Transcriptome analysis was utilized to analyze the functional properties of these cells. The cytotoxicity of CD38⁺HLA-DR⁺CD8⁺ T cells was detected by cytotoxicity assay and antibody blocking assay. Results: The percentage of CD38⁺HLA-DR⁺CD8⁺ T cells in patients with liver cirrhosis significantly increased and was correlated with liver injury. These CD8⁺ T cells contained largely non-HBV specific T cells. Transcriptome analysis revealed that these CD8⁺ T cells subsets exhibited innate-like functional characteristic. In addition, these cells mainly consisted of effector memory T cells and displayed high expression levels of cytotoxicity-related cytokines, especially granzyme B and perforin. Stimulation experiments with cytokines shown that IL-15 could promote the activation and proliferation of these CD8⁺ T cells. Lastly, blocking assays indicated that CD38⁺HLA-DR⁺CD8⁺ T cells had strong cytotoxic effects in a TCR-independent manner, mediated by NKG2D. Conclusion: CD38⁺HLA-DR⁺CD8⁺ T cells were correlated with the liver injury in liver cirrhosis, and these cells exerted liver damaging effects through NKG2D in a TCR-independent manner.
... All B and T lymphocytes, along with antigen-presenting cells (APCs), cooperatively mediate this process through expressing specific receptors [83,84]. B lymphocytes are vital for humoral immunity; they secrete antibodies and cytokines, serve as APCs, and generate immunological memory [85,86]. However, the importance of these cell groups in BC is arguable. ...
Article
Full-text available
Antigen recognition and presentation are highlighted as the first steps in developing specialized antigen responses. Dendritic cells (DCs) are outstanding professional antigen-presenting cells (APCs) responsible for priming cellular immunity in pathological states, including cancer. However, the diminished or repressed function of DCs is thought to be a substantial mechanism through which tumors escape from the immune system. In this regard, DCs obtained from breast cancer (BC) patients represent a notably weakened potency to encourage specific T-cell responses. Additionally, impaired DC-T-cell cross-talk in BC facilitates the immune evade of cancer cells and is connected with tumor advancement, immune tolerance, and adverse prognosis for patients. In this review we aim to highlight the available knowledge on DC-T-cell interactions in BC aggressiveness and show its therapeutic potential in BC treatment.
... Furthermore, BCRinitiated B-cell activation is tightly regulated to control the quality, quantity, and duration of antibody responses based on the nature of antigenic challenges (11,12). Dysregulation of this process, such as prolonged or constitutive BCR activation, can lead to the activation of self-reactive B-cells and autoantibody production, as well as B-cell malignancies (13)(14)(15)(16). Thus, understanding the mechanisms underlying BCR signaling regulation is essential for developing vaccines for pathogens and therapies for autoimmune diseases and cancer. ...
Article
Full-text available
Antigen (Ag)-triggered B-cell receptor (BCR) signaling initiates antibody responses. However, prolonged or uncontrolled BCR signaling is associated with the development of self-reactive B-cells and autoimmune diseases. We previously showed that actin-mediated B-cell contraction on Ag-presenting surfaces negatively regulates BCR signaling. Non-muscle myosin II (NMII), an actin motor, is involved in B-cell development and antibody responses by mediating B-cell migration, cytokinesis, and Ag extraction from Ag-presenting cells. However, whether and how NMII regulates humoral responses through BCR signaling remains elusive. Utilizing a B-cell-specific, partial NMIIA knockout (cIIAKO) mouse model and NMII inhibitors, this study examined the role of NMII in BCR signaling. Upon BCR binding to antibody-coated planar lipid bilayers (PLB), NMIIA was recruited to the B-cell contact membrane and formed a ring-like structure during B-cell contraction. NMII recruitment depended on phosphatidylinositol 5-phosphatase (SHIP1), an inhibitory signaling molecule. NMII inhibition by cIIAKO did not affect B-cell spreading on PLB but delayed B-cell contraction and altered BCR clustering. Surface BCR “cap” formation induced by soluble stimulation was enhanced in cIIAKO B-cells. Notably, NMII inhibition by cIIAKO and inhibitors up-regulated BCR signaling in response to both surface-associated and soluble stimulation, increasing phosphorylated tyrosine, CD79a, BLNK, and Erk and decreasing phosphorylated SHIP1. While cIIAKO did not affect B-cell development, the number of germinal center B-cells was significantly increased in unimmunized cIIAKO mice, compared to control mice. While cIIAKO mice mounted similar antibody responses when compared to control mice upon immunization, the percentages of high-affinity antibodies, Ag-specific germinal center B-cells and isotype switched B-cells were significantly lower in cIIAKO mice than in control mice. Furthermore, autoantibody levels were elevated in cIIAKO mice, compared to control mice. Collectively, our results reveal that NMII exerts a B-cell-intrinsic inhibition on BCR signaling by regulating B-cell membrane contraction and surface BCR clustering, which curtails the activation of non-specific and self-reactive B-cells.
... Clinical evidence suggests that aberrant RANK signaling in B cells contributes to the induction of B cell autoimmunity and malignancy [112][113][114][115][116]. Somatically acquired mutations of RANK in which the intracellular signaling domain Lys at amino acid 240 changes to Glu (RANK K240E ) have been detected in human diffuse large B cell lymphoma specimens. ...
Article
Since the receptor activator of nuclear factor-kappa B ligand (RANKL), its cognate receptor receptor activator of nuclear factor-kappa B (RANK), and the decoy receptor osteoprotegerin (OPG) were discovered, a number of studies have uncovered the crucial role of the RANKL-RANK-OPG pathway in controlling the key aspect of bone homeostasis, the immune system, inflammation, cancer, and other systems under pathophysiological condition. These findings have expanded the understanding of the multifunctional biology of the RANKL-RANK-OPG pathway and led to the development of therapeutic potential targeting this pathway. The successful development and application of anti-RANKL antibody in treating diseases causing bone loss validates the utility of therapeutic approaches based on the modulation of this pathway. Moreover, recent studies have demonstrated the involvement of the RANKL-RANK pathway in osteoblast differentiation and bone formation, shedding light on the RANKL-RANK dual signaling in coupling bone resorption and bone formation. In this review, we will summarize the current understanding of the RANKL-RANK-OPG system in the context of the bone and the immune system as well as the impact of this pathway in disease conditions, including cancer development and metastasis.
Article
Bones can form the scaffolding of the body, support the organism, coordinate somatic movements, and control mineral homeostasis and hematopoiesis. The immune system plays immune supervisory, defensive, and regulatory roles in the organism, which mainly consists of immune organs (spleen, bone marrow, tonsils, lymph nodes, etc.), immune cells (granulocytes, platelets, lymphocytes, etc.), and immune molecules (immune factors, interferons, interleukins, tumor necrosis factors, etc.). Bone and the immune system have long been considered two distinct fields of study, and the bone marrow, as a shared microenvironment between the bone and the immune system, closely links the two. Osteoimmunology organically combines bone and the immune system, elucidates the role of the immune system in bone, and creatively emphasizes its interdisciplinary characteristics and the function of immune cells and factors in maintaining bone homeostasis, providing new perspectives for skeletal‐related field research. In recent years, bone immunology has gradually become a hot spot in the study of bone‐related diseases. As a new branch of immunology, bone immunology emphasizes that the immune system can directly or indirectly affect bones through the RANKL/RANK/OPG signaling pathway, IL family, TNF‐α, TGF‐β, and IFN‐γ. These effects are of great significance for understanding inflammatory bone loss caused by various autoimmune or infectious diseases. In addition, as an external environment that plays an important role in immunity and bone, this study pays attention to the role of exercise‐mediated bone immunity in bone reconstruction.
Article
Although various studies have been performed on the function of polymorphonuclear myeloid-derived suppressor cells (PMN-MDSCs) in RA, the results were conflicting. Here we were trying to clarify the role of PMN-MDSCs in the pathogenesis of RA and its specific mechanisms. We detected the frequencies and counts of PMN-MDSCs, TNF-α+ B cells and Ki67+ B cells in spleen and inflamed joints of collagen-induced arthritis (CIA) mice using flow cytometry. The pathological role of PMN-MDSCs was examined by anti-Ly6G neutralizing antibodies against PMN-MDSCs or adoptive transfer of PMN-MDSCs. And the modulation of PMN-MDSCs on B cells was conducted by coculture assays, RNA-Seq, RT-qPCR, and so on. The mechanism of BAFF regulating B cells was verified through western blot and flow cytometry. PMN-MDSCs accumulated in the spleen and joints of CIA mice. PMN-MDSCs depletion could alleviate the arthritis severity, which was accompanied by decreased TNF-α secretion and proliferation of B cells. And its adoptive transfer also facilitated disease progress. Furthermore, PMN-MDSCs from CIA mice had higher expression level of BAFF, which regulated TNF-α expression, proliferation and apoptosis of B cells in vitro. What's more, BAFF promoted phosphorylation of BTK/NF-κB signalling pathway. And Ibrutinib (BTK inhibitor) could reverse the effect of BAFF on TNF-α expression of B cells. Our study suggested that PMN-MDSCs enhanced disease severity of CIA and manipulated TNF-α expression, proliferation and apoptosis of B cells via BAFF, furthermore, BAFF promoted TNF-α expression through BTK/NF-κB signalling pathway, which demonstrated a novel pathogenesis of PMN-MDSCs in CIA.
Article
The pleiotropic cytokine IL-9 signals to target cells by binding to a heterodimeric receptor consisting of the unique subunit IL-9R and the common subunit γ-chain shared by multiple cytokines of the γ-chain family. In the current study, we found that the expression of IL-9R was strikingly upregulated in mouse naive follicular B cells genetically deficient in TNFR-associated factor 3 (TRAF3), a critical regulator of B cell survival and function. The highly upregulated IL-9R on Traf3-/- follicular B cells conferred responsiveness to IL-9, including IgM production and STAT3 phosphorylation. Interestingly, IL-9 significantly enhanced class switch recombination to IgG1 induced by BCR crosslinking plus IL-4 in Traf3-/- B cells, which was not observed in littermate control B cells. We further demonstrated that blocking the JAK-STAT3 signaling pathway abrogated the enhancing effect of IL-9 on class switch recombination to IgG1 induced by BCR crosslinking plus IL-4 in Traf3-/- B cells. Our study thus revealed, to our knowledge, a novel pathway that TRAF3 suppresses B cell activation and Ig isotype switching by inhibiting IL-9R-JAK-STAT3 signaling. Taken together, our findings provide (to our knowledge) new insights into the TRAF3-IL-9R axis in B cell function and have significant implications for the understanding and treatment of a variety of human diseases involving aberrant B cell activation such as autoimmune disorders.
Article
Full-text available
Immune-mediated diseases (IMDs) are chronic conditions that have an immune-mediated etiology. Clinically, these diseases appear to be unrelated, but pathogenic pathways have been shown to connect them. While inflammation is a common occurrence in the body, it may either stimulate a favorable immune response to protect against harmful signals or cause illness by damaging cells and tissues. Nanomedicine has tremendous promise for regulating inflammation and treating IMIDs. Various nanoparticles coated with nanotherapeutics have been recently fabricated for effective targeted delivery to inflammatory tissues. RNA interference (RNAi) offers a tremendous genetic approach, particularly if traditional treatments are ineffective against IMDs. In cells, several signaling pathways can be suppressed by using RNAi, which blocks the expression of particular messenger RNAs. Using this molecular approach, the undesirable effects of anti-inflammatory medications can be reduced. Still, there are many problems with using short-interfering RNAs (siRNAs) to treat IMDs, including poor localization of the siRNAs in target tissues, unstable gene expression, and quick removal from the blood. Nanotherapeutics have been widely used in designing siRNA-based carriers because of the restricted therapy options for IMIDs. In this review, we have discussed recent trends in the fabrication of siRNA nanodelivery systems, including lipid-based siRNA nanocarriers, liposomes, and cationic lipids, stable nucleic acid-lipid particles, polymeric-based siRNA nanocarriers, polyethylenimine (PEI)-based nanosystems, chitosan-based nanoformulations, inorganic material-based siRNA nanocarriers, and hybrid-based delivery systems. We have also introduced novel siRNA-based nanocarriers to control IMIDs, such as pulmonary inflammation, psoriasis, inflammatory bowel disease, ulcerative colitis, rheumatoid arthritis, etc. This study will pave the way for new avenues of research into the diagnosis and treatment of IMDs.
Article
G‐protein‐coupled receptors (GPCRs) are the largest and most diverse group of membrane receptors. They are involved in almost every physiologic process and consequently have a pivotal role in an extensive number of pathologies, including genetic, neurologic, and immune system disorders. Indeed, the vast array of GPCRs mechanisms have led to the development of a tremendous number of drug therapies and already account for about a third of marketed drugs. These receptors mediate their downstream signals primarily via G proteins. The regulators of G‐protein signaling (RGS) proteins are now in the spotlight as the critical modulatory factors of active GTP‐bound Gα subunits of heterotrimeric G proteins to fine‐tune the biologic responses driven by the GPCRs. Also, they possess noncanonical functions by multiple mechanisms, such as protein–protein interactions. Essential roles and impacts of these RGS proteins have been revealed in physiology, including hematopoiesis and immunity, and pathologies, including asthma, cancers, and neurologic disorders. This review focuses on the largest subfamily of R4 RGS proteins and provides a brief overview of their structures and G‐proteins selectivity. With particular interest, we explore and highlight, their expression in the hematopoietic system and the regulation in the engraftment of hematopoietic stem/progenitor cells (HSPCs). Distinct expression patterns of R4 RGS proteins in the hematopoietic system and their pivotal roles in stem cell trafficking pave the way for realizing new strategies for enhancing the clinical performance of hematopoietic stem cell transplantation. Finally, we discuss the exciting future trends in drug development by targeting RGS activity and expression with small molecules inhibitors and miRNA approaches. R4 RGS proteins exhibit profound impact on the trafficking of blood stem cells and represent promising drug targets for enhancing clinical stem cell transplantation.
Article
Full-text available
Metal‐based nanoparticles (NPs) are particularly important tools in tissue engineering‐, drug carrier‐, interventional therapy‐, and biobased technologies. However, their complex and varied migration and transformation pathways, as well as their continuous accumulation in closed biological systems, cause various unpredictable toxic effects that threaten human and ecosystem health. Considerable experimental and theoretical efforts have been made toward understanding these cytotoxic effects, though more research on metal‐based NPs integrated with clinical medicine is required. This review summarizes the mechanisms and evaluation methods of cytotoxicity and provides an in‐depth analysis of the typical effects generated in the nervous, immune, reproductive, and genetic systems. In addition, the challenges and opportunities are discussed to enhance future investigations on safer metal‐based NPs for practical commercial adoption. Considerable experimental and theoretical efforts have been directed toward understanding the cytotoxicity of metal‐based nanoparticles but additional efforts are still needed before metal‐based nanoparticles can be routinely used in clinical medicine field. This review summarizes the state of toxic mechanisms and evaluation methods of cytotoxicity, and gives an in‐depth analysis of typical cytotoxic effects.
Article
Full-text available
Pietro Enea Lazzerini,1 Pier Leopoldo Capecchi,1 Giacomo Maria Guidelli,1 Enrico Selvi,1 Maurizio Acampa,2 Franco Laghi-Pasini1 1Department of Medical Sciences, Surgery and Neurosciences, University of Siena, 2Stroke Unit, University Hospital of Siena, Siena, Italy Abstract: Rheumatoid arthritis (RA) is a chronic autoimmune inflammatory disease primarily affecting synovial joints and is characterized by persistent high-grade systemic inflammation. Proinflammatory cytokines, particularly interleukin-6 (IL-6), are of crucial importance in the pathogenesis of the disease, driving both joint inflammation and extra-articular comorbidities. Tocilizumab, a humanized IL-6 receptor-inhibiting monoclonal antibody, has been the first, and, to date, the only, IL-6 inhibitor approved for the treatment of RA. Many studies have demonstrated the potency and effectiveness of tocilizumab in controlling disease activity and radiological progression of RA. These successful results have encouraged the development of novel IL-6 inhibitors, among which a promising agent is sirukumab (SRK), a human anti-IL-6 monoclonal antibody currently under evaluation in Phase II/III studies in patients with RA, systemic lupus erythematosus, giant-cell arteritis, and major depressive disorder. The evidence to date indicates SRK as an effective and well-tolerated new therapeutic tool for patients with active RA, with some preliminary data suggesting a specific beneficial impact on relevant systemic complications associated with the disease, such as depression and cardiovascular disease. Conversely, although pathophysiological considerations make plausible the hypothesis that IL-6 blockade with SRK may also be beneficial in the treatment of many diseases other than RA (either autoimmune or not), available clinical data in patients with systemic lupus erythematosus do not seem to support this view, also giving rise to potentially relevant concerns about drug safety. If large Phase III clinical trials currently in progress in patients with RA confirm the efficacy and tolerability of SRK, then in the long term, this drug could, in the near future, occupy a place in the treatment of the disease, potentially also opening the doors to a more extended use of SRK in a wide range of disorders in which IL-6 plays a key pathogenic role. Keywords: sirukumab, rheumatoid arthritis, interleukin-6, tocilizumab, systemic lupus erythematosus, cardiovascular disease, interleukin-6
Article
Full-text available
The health of living organisms is constantly challenged by bacterial and viral threats. The recognition of pathogenic microorganisms by diverse receptors triggers a variety of host defense mechanisms, leading to their eradication. Toll-like receptors (TLRs), which are type I transmembrane proteins, recognize specific signatures of the invading microbes and activate a cascade of downstream signals inducing the secretion of inflammatory cytokines, chemokines, and type I interferons. The TLR response not only counteracts the pathogens but also initiates and shapes the adaptive immune response. Under normal conditions, inflammation is downregulated after the removal of the pathogen and cellular debris. However, a dysfunctional TLR-mediated response maintains a chronic inflammatory state and leads to local and systemic deleterious effects in host cells and tissues. Such inappropriate TLR response has been attributed to the development and progression of multiple diseases such as cancer, autoimmune, and inflammatory diseases. In this review, we discuss the emerging role of TLRs in the pathogenesis of inflammatory diseases and how targeting of TLRs offers a promising therapeutic strategy for the prevention and treatment of various inflammatory diseases. Additionally, we highlight a number of TLR-targeting agents that are in the developmental stage or in clinical trials.
Article
Full-text available
Systemic lupus erythematosus (SLE; OMIM 152700) is a genetically complex autoimmune disease. Genome-wide association studies (GWASs) have identified more than 50 loci as robustly associated with the disease in single ancestries, but genome-wide transancestral studies have not been conducted. We combined three GWAS data sets from Chinese (1,659 cases and 3,398 controls) and European (4,036 cases and 6,959 controls) populations. A meta-analysis of these studies showed that over half of the published SLE genetic associations are present in both populations. A replication study in Chinese (3,043 cases and 5,074 controls) and European (2,643 cases and 9,032 controls) subjects found ten previously unreported SLE loci. Our study provides further evidence that the majority of genetic risk polymorphisms for SLE are contained within the same regions across both populations. Furthermore, a comparison of risk allele frequencies and genetic risk scores suggested that the increased prevalence of SLE in non-Europeans (including Asians) has a genetic basis.
Article
Full-text available
Lupus nephritis (LN) is a potentially dangerous end organ pathology that affects upwards of 60% of lupus patients. Bruton’s tyrosine kinase (BTK) is important for B cell development, Fc receptor signaling, and macrophage polarization. In this study, we investigated the effects of a novel, highly selective and potent BTK inhibitor, BI-BTK-1, in an inducible model of LN in which mice receive nephrotoxic serum (NTS) containing anti-glomerular antibodies. Mice were treated once daily with vehicle alone or BI-BTK-1, either prophylactically or therapeutically. When compared with control treated mice, NTS-challenged mice treated prophylactically with BI-BTK-1 exhibited significantly attenuated kidney disease, which was dose dependent. BI-BTK-1 treatment resulted in decreased infiltrating IBA-1+ cells, as well as C3 deposition within the kidney. RT-PCR on whole kidney RNA and serum profiling indicated that BTK inhibition significantly decreased levels of LN-relevant inflammatory cytokines and chemokines. Renal RNA expression profiling by RNA-seq revealed that BI-BTK-1 dramatically modulated pathways related to inflammation and glomerular injury. Importantly, when administered therapeutically, BI-BTK-1 reversed established proteinuria and improved renal histopathology. Our results highlight the important role for BTK in the pathogenesis of immune complex-mediated nephritis, and BTK inhibition as a promising therapeutic target for LN.
Article
Full-text available
Systemic lupus erythematosus (SLE) is a prototypic autoimmune disease characterized by impaired immune tolerance, resulting in the generation of pathogenic autoantibodies and immune complexes. Although autoreactive B lymphocytes have been the first targets for biologic therapies in SLE, the importance of the innate immune system, and in particular, pathways involved in interferon (IFN) signaling, has emerged. There are now data supporting a central role for a plasmacytoid dendritic cell-derived type I IFN pathway in SLE, with a number of biologic therapeutics and small-molecule inhibitors undergoing clinical trials. Monoclonal antibodies targeting IFN-α have completed phase II clinical trials, and an antibody against the type I IFN receptor is entering a phase III trial. However, other IFNs, such as IFN gamma, and the more recently discovered type III IFNs, are also emerging as targets in SLE; and blockade of upstream components of the IFN signaling pathway may enable inhibition of more than one IFN subtype. In this review, we discuss the current understanding of IFNs in SLE, focusing on emerging therapies.
Article
Full-text available
Many autoreactive B cells persist in the periphery in a state of unresponsiveness called anergy. This unresponsiveness is rapidly reversible, requiring continuous BCR interaction with self-antigen and resultant regulatory signaling for its maintenance. Using adoptive transfer of anergic B cells with subsequent acute induction of gene deletion or expression, we demonstrate that the continuous activities of independent inhibitory signaling pathways involving the tyrosine phosphatase SHP-1 and the inositol phosphatase SHIP-1 are required to maintain anergy. Acute breach of anergy by compromise of either of these pathways leads to rapid cell activation, proliferation, and generation of short-lived plasma cells that reside in extrafollicular foci. Results are consistent with predicted/observed reduction in the Lyn–SHIP-1–PTEN–SHP-1 axis function in B cells from systemic lupus erythematosus patients.
Article
Introduction: Considerable advances in the treatment of rheumatoid arthritis (RA) have been made following the advent of biological disease-modifying anti-rheumatic drugs (DMARDs). However, biological DMARDs require intravenous or subcutaneous injection and some patients fail to respond to these drugs or lose their primary response. Currently, Janus kinase (JAK) inhibitors have been developed as a new class of DMARD that inhibits the non-receptor tyrosine kinase family JAK involved in intracellular signaling of various cytokines and growth factors. Area covered: Several JAK inhibitors such as tofacitinib and baricitinib are oral synthetic DMARD that inhibit JAK1, 2 and 3. Both drugs have shown feasible efficacy and tolerable safety. In this article, efficacy and adverse events from the phase III trials of JAK inhibitors are overviewed. In addition, pharmacokinetics and mechanism of action of JAK inhibitors in relevance to efficacy and adverse events are covered. Expert opinion: JAK inhibitors are novel therapies for RA that inhibit multiple cytokines and signaling pathways. Further studies are needed to determine their risk-benefit ratio and selection of the most appropriate patients for such therapy.
Article
Autism spectrum disorder (ASD) occurs in 1 in 68 births, preferentially affecting males. It encompasses a group of neurodevelopmental abnormalities characterized by impaired social interaction and communication, stereotypic behaviors and motor dysfunction. Although recent advances implicate maternal brain-reactive antibodies in a causative role in ASD, a definitive assessment of their pathogenic potential requires cloning of such antibodies. Here, we describe the isolation and characterization of monoclonal brain-reactive antibodies from blood of women with brain-reactive serology and a child with ASD. We further demonstrate that male but not female mice exposed in utero to the C6 monoclonal antibody, binding to contactin-associated protein-like 2 (Caspr2), display abnormal cortical development, decreased dendritic complexity of excitatory neurons and reduced numbers of inhibitory neurons in the hippocampus, as well as impairments in sociability, flexible learning and repetitive behavior. Anti-Caspr2 antibodies are frequent in women with brain-reactive serology and a child with ASD. Together these studies provide a methodology for obtaining monclonal brain-reactive antibodies from blood B cells, demonstrate that ASD can result from in utero exposure to maternal brain-reactive antibodies of single specificity and point toward the exciting possibility of prognostic and protective strategies.
Article
Systemic lupus erythematosus (SLE) is an autoimmune disease with a polymorphic presentation. The variability in the clinical expression and severity of SLE makes new treatments both essential and challenging to develop. Several biotherapies targeting different pathophysiological pathways have been developed over the past 15 years. The results of Phase II trials were encouraging but rarely borne out by Phase III trials. Recent data, which are discussed in detail in this review, allowed belimumab - a monoclonal antibody against BLyS (B-lymphocyte stimulator) - to become the first biotherapy approved for use in SLE. Other molecules targeting B cells include the two anti-BLyS antibodies tabalumab and blisibimod; atacicept, which targets both BLyS and APRIL (a proliferation-inducing ligand); and the monoclonal antibody to CD22 epratuzumab. The rekindling of interest in the B-cell pathway has also driven new clinical research into rituximab, a monoclonal antibody targeting CD20 with evaluations of new strategies. A new and promising approach is the use of inhibitors of the type 1 interferon (IFN) pathway, of which the most promising is anifrolumab, a monoclonal antibody targeting the type 1 IFN receptor. In this review, we discuss study findings and their clinical relevance, present the most promising targets, and analyze possible explanations to negative results, such as inappropriate patient selection and treatment response criteria or the erratic use of high-dose glucocorticoid therapy.