Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Clinical Immunology 239 (2022) 109031

Contents lists available at ScienceDirect

Clinical Immunology
journal homepage: www.elsevier.com/locate/yclim

T cell dysregulation in SLE


Klaus Tenbrock a, *, Thomas Rauen b
a
Dept of Pediatrics, Pediatric Rheumatology, RWTH Aachen University, Aachen, Germany
b
Dept of Medicine, Division of Nephrology and Clinical Immunology, RWTH Aachen University, Aachen, Germany

A R T I C L E I N F O A B S T R A C T

Keywords: T helper cells aid B cells with the production of antibodies and thus play a central role in disease development of
Lupus systemic lupus erythematosus (SLE). Numerous T helper cell abnormalities have been described in SLE patients
CD4+ T cells that contribute to disease pathophysiology and provide suitable targets for therapeutic intervention. In addition,
T follicular helper cells
T effector cell also play a less well-defined role in SLE. This review focusses on underlying molecular mechanisms
Tregs
Foxp3
of different T cell alterations in SLE.
CD8+ T cells

1. Introduction and autoimmunity. This relates to the fact that, in addition to being a
necessary growth factor for T helper and cytotoxic T cells, IL-2 is also a
B cells and autoantibody production are key drivers in the patho­ non-redundant cytokine for regulatory T cells (Tregs), which are the
genesis of systemic lupus erythematosus (SLE). The most recent disease gatekeepers of immune tolerance [6]. Regulatory T cells (Tregs) fail to
classification system presupposes the presence of anti-nuclear anti­ produce IL-2 and are thereby completely dependent on external IL-2 [7].
bodies to establish the diagnosis of SLE [1]. Yet, B cells per se are Decreased IL-2 production in T cells was identified in the clinical
ineffective without appropriate support of T cells providing necessary context of enhanced viral infection rates in SLE patients suggesting an
costimulatory signals to them. The importance of these pathways in impaired T cellular cytotoxicity [5]. This was also linked to altered
promoting autoantibody production have been demonstrated in genet­ production of other cytokines including increased expression of IL-6 and
ically modified lupus-prone mice or by the use of blocking antibodies IL-10 [8]. Both contribute to non-cognate help to B cells, whereas the
against different costimulatory molecules like inducible costimulatory decreased IL-2 expression accounts for defective production of cytotoxic
ligand (ICOS-L) and CD40 ligand [2–4]. Thus, T helper cells are of cells or defective activation-induced cell death of various T-cell subsets
utmost importance for SLE disease development and progression. including those with autoreactive specificity. First, it was unclear
Numerous T cell abnormalities including alterations within cytokine whether defective IL-2 production might be taken advantage of in
production, metabolism and epigenetic modifications have been therapeutic interventions. In the meantime, a number of studies have
described that influence T helper cell function. Moreover, effector T cell provided proof of evidence that a low-dose IL-2 therapy is beneficial in
and regulatory T cell functions are also altered in SLE and will be dis­ refractory SLE patients [9–12]. It has moreover become clear that the
cussed in this review. (See Fig. 1.) mode of administration and half-life of therapeutic IL-2 affect clinical
effectiveness. This is related to the effect of IL-2 on T cells as well as on
increased selectivity for Tregs, who express a high affinity IL-2 receptor
1.1. The role and regulation of IL-2 (and of other cytokines) (i.e. CD25). Enhanced half-life of IL-2 can be achieved by conjugating
the cytokine to poly(ethylene glycol). Pegylated IL-2 has a half live of 14
T helper cells exert their function particularly by expression of cy­ days and was efficient in SLE mice models as well as in cutaneous lupus
tokines. One of the most important cytokines is interleukin (IL)-2, which in humans and results in expansion of the peripheral Tregs compartment
acts as T cell growth and proliferation factor. T cell dysfunction in SLE [13].
was first described in the 80s of the last century demonstrating that SLE
T cells exhibit a decreased IL-2 production [5]. Defects in IL-2 produc­
tion and its receptor expression were linked to severe T cell dysfunction

* Corresponding author.
E-mail address: ktenbrock@ukaachen.de (K. Tenbrock).

https://doi.org/10.1016/j.clim.2022.109031
Received 1 February 2022; Received in revised form 2 May 2022; Accepted 2 May 2022
Available online 6 May 2022
1521-6616/© 2022 Published by Elsevier Inc.
K. Tenbrock and T. Rauen Clinical Immunology 239 (2022) 109031

1.2. Molecular basis for defective IL-2 production inhibits promoter activity and thereby mRNA and protein expression. In
addition to this indirect effect on IL2 gene expression CREMα exerts
The decreased expression of IL-2 is the result of different patho­ direct effects by binding to a specific binding site within the IL2 pro­
physiological defects described in SLE T cells. Beginning with signal moter. This binding site is usually occupied by the transcriptional acti­
transduction from the cellular surface it was shown that SLE T cells vator CREB, however decreased CREB expression results in a CREM/
display an altered Ca2+ signaling which was related to decreased CREB imbalance favoring CREM binding and thus transcriptional
expression of the T cell receptor (TCR) zeta chain and rewiring of the repression of IL2 [19–21]. Moreover, CREM recruits the histone deace­
TCR complex [14]. This is accompanied by an altered lipid raft forma­ tylase 1 (HDAC1) to the promoter which results in the deaceylation of
tion on the surface of SLE T cells at the immunological synapse resulting histone H4, a chromatin closure and limited accessibility to the pro­
in enhanced signaling. In detail, the T cell receptor (TCR) zeta chain is moter also by other TFs [22]. Furthermore, CREMα mediates IL2
replaced by the FcRgamma chain which is usually an integral part of the silencing in T lymphocytes from SLE patients through a gene-wide his­
B cell receptor complex and not readily expressed in T cells. FcRgamma tone deacetylase 1-directed deacetylation of histone H3K18 and DNA
forms a complex with tyrosine kinase Syk which in turn is capable to methyltransferase 3a-directed cytosine phosphate guanosine (CpG)-
signal 100 times more effectively than the TCR zeta-ZAP70 complex. It DNA hypermethylation [23,24]. Consequently, CREM downregulation
has also been demonstrated that forced expression of the TCR zeta chain using an anti-sense CREM plasmid restores IL-2 expression in SLE T cells
in SLE T cells restores IL-2 production. This receptor/kinase rearrange­ [20].
ment explains the overexcitability of SLE T cells. A decreased expression The expression of CREM itself is regulated by different promotors.
of the zeta chain in SLE T cells occurs already on the mRNA level sug­ CREM isoforms consist of different splice variants containing tran­
gesting a dysregulation already at the transcriptional level. Two major scriptional activators and repressors depending on the presence of a
transcription factors (TF) have been identified that are involved in TCR trans-activator domain that may bind to transcriptional cofactors CREB-
zeta promoter dysregulation in SLE T cells. One is the ETS1 transcription binding protein (CBP) or p300. CREM promotor 1 activity is regulated
factor E74-like factor 1 (ELF1) which regularly undergoes post­ by protein phosphatase (PP)2A-dependent specificity protein 1 (SP1)
translational modifications affecting its activity. ELF1 becomes phos­ phosphorylation and binding, which is found to be enhanced in SLE T
phorylated and O-Nacetylglucosamine (O-GlcNAc) glycosylated. The cells in combination with enhanced activating protein 1 (AP1) activity
latter process is disturbed in SLE T cells and prevents ELF1 binding to the that regulates an additional intronic promoter P2 [25,26]. PP2A
zeta promotor and consecutive activation [15–17]. The second TF with downregulation using siRNA results in enhanced binding of pCREB to
major effects on IL2 transcription is the cAMP responsive element the IL-2 promoter and upregulation of IL-2. Vice versa, a transgenic
modulator (CREM). CREM is highly expressed in SLE T cells either in the mouse that overexpresses the PP2Ac subunit in T cells displays a
α- or ICER-isoform and accounts for numerous transcriptional, epige­ heightened susceptibility to immune-mediated glomerulonephritis by
netic and metabolic abnormalities in SLE T cells. CREMα has a direct an IL-17 dependent mechanism while PP2A knock out mice show
effect by binding to the TCR zeta promotor [18]. Subsequently, it decreased IL-17 expression and are protected in a model of experimental

Fig. 1. Schematic representation of transcription factors involved in differential cytokine expression in SLE T-cells.
AP1, Activated protein 1; CaMKII, Calmodulin Kinase II; CaMKIV, Calmodulin Kinase IV; CREB, cAMP response element binding protein; CREM, cAMP response
element binding modulator; DNMT3A, DNA methyle transferase 3A, DUSP4, Dual sepcific phosphatase 4; ELF1, ETS1 transcription factor E74-like factor 1; IL-2,
Interleukin 2; IL-17, Interleukin 17; IL-21, Interleukin 21; IRF4, Interferon regulated factor 4; mTOR, Mammalian target of rapamycin; NFAT, Nuclear factor of
activated T cells; PP2A, Protein Phosphatase 2A; STAT4, Signal transducer and activator of transcription 4; TCR, T cell receptor.

2
K. Tenbrock and T. Rauen Clinical Immunology 239 (2022) 109031

autoimmune encephalitis [27–30]. molecular mimicry between bacterial epitopes and autoantigens ac­
Moreover, CREM is regulated by The calcium modulin kinase IV counts for abnormal CD4+ T cell differentiation including enhanced
(CamKIV) which is activated in SLE T cells and can be activated in T cells expression of IL-17. It was shown in lupus-prone mice that supplemen­
of healthy controls in the presence of SLE sera containing IgG complexes tation with Lactobacillus acidophilus reduced IL-17 and DN T cell
that specifically activate the TCR in SLE. Thus, inhibition of CamKIV expression during treatment with tacrolimus while it enhanced expres­
prevents CREM activation and rescues IL-2 expression as well as kidney sion of IL-10 [45]. It is tempting to speculate that metabolites provided
damage [31,32]. Beyond IL-2, CREM also regulates expression of other by altered gut microbiota influence T cell function and it was shown that
cytokines like IL-17 and IL-21, which both exert proinflammatory glycolysis is enhanced in Th17 cells. Again, CREM/ICER regulate
functions and are involved in organ damage [33,34]. expression of pyruvate dehydrogenase (PDH), which in its active
dephosphorylated form advances energy supply by using oxidative
1.3. Regulation of IL-17 expression phosphorylation. CREM/ICER thereby bind and suppress the Pdp2 pro­
moter, which regulates expression of PDH [46] resulting in an enhanced
IL-17-producing T cells particularly contribute to kidney pathology glycolysis and IL-17 production. Moreover, another metabolic alteration
in SLE. Migration of these proinflammatory cells is partially mediated by is enhanced glutaminolysis. Th17 induction depends on glutaminolysis
cleaved CD95 ligand, which induces S1P receptor 3 in Th17 cells that and upregulation of glutaminase 1 (Gls1) which is one of the key en­
induces migration into the inflamed kidney and the promotion of organ zymes to mediate the metabolic breakdown of glutamine. The Gls1
damage [35]. IL-17A and IL-17F thereby both contribute to tissue promoter contains a CRE site and its function depends on CREM/ICER
inflammation by induction of different chemokines (e.g. MCP-1), which binding. CREM-deficient murine T cells show reduced activity of
recruit additional monocytes and neutrophils to the kidney. The isoform oxidative phosphorylation rates in the presence of glutamine and
CREMα influences transcriptional control of both IL17A and IL17F [33]. reduced Gls1 expression [46]. Vice versa, glutaminase 1 inhibition
The human IL17A and IL17F genes are located in close proximity on improved autoimmune pathology in MRL.lpr mice, and suppressed Th17
chromosome 6 (within a range of about 85 kb). CREMα binds to a cAMP- differentiation of T cells from patients with SLE but not in those from
responsive element, CRE (− 111/− 104), within the proximal human healthy donors [47].
IL17A promoter and increases its activity. Increased CREMα recruitment Additional metabolic alterations have been unraveled in another
to the IL17A and the IL17F promoters has been documented in T cells murine lupus model demonstrating abundance of the metabolite tryp­
from SLE patients [33]. This recruitment induces epigenetic remodeling tophane which is produced by altered gut microbiota [43]. Pathophy­
of the IL17A and F gene loci. In contrast to the effects observed at the IL2 siologically, it is suggested that the tryptophane metabolite kynurenine
gene locus, CREMα binding results in lower abundance of histone binds AhR and activates mechanistic target of rapamycin (mTOR) C1 in
deactetylase HDAC1 and DNA methyltransferase DNMT3a in the vicin­ human T cells, which provides a rationale for treatment with the mTOR
ity to the IL17A promoter resulting in an overall increased histone H3 inhibitors rapamycine and sirolimus. This is particularly interesting
acetylation and reduced histone methylation as well as CpG-DNA since mTOR regulates cellular growth and energy utilization. CD4+ T
hypomethylation of the IL17 locus [24]. Moreover, CREMα induces cells from SLE patients display enhanced mTOR activation [48], which
dual specificity protein phosphatase (DUSP) 4 in effector CD4+ T cells enhances glycolysis and fatty acid synthesis, 2 major metabolic path­
through corecruitment of p300 [36]. DUSP4 induces IL-17A while ways that favor Th17 differentiation [49,50] resulting in a Th17/Treg
limiting IL-2 expression, which is also found in T cells jSLE patients. In imbalance in SLE patients.
line with this, a T cells specific transgenic overexpression of CREMα in
mice also enhances IL-17 levels, while CREM− /− mice in which the DNA 1.4. Regulation of IL-21 expression
binding domain has been disrupted and thus all CREM isoforms are
missing including ICER, IL-17 production is impaired and CREM- IL-21 is a pleiotropic cytokine primarily expressed by T follicular
deficient B6.lpr mice are protected from developing autoimmunity helper cells (TfH cells). TfH cells are located in the germinal centers and
[37]. While IL-17A expression is enhanced, the expression of IL17F is provide IL-21-dependent help for B cell differentiation in order to pro­
downredulated by binding of CREMα to a CRE site within the IL-17F duce (auto)antibodies. This cytokine is indeed of utmost importance for
promoter. It is suggested that an imbalance between IL-17A and IL- the generation of autoreactive B cells as demonstrated in experiments in
17F levels and impaired assembly of IL-17A/IL-17F heterodimers, which the IL-21R signaling pathway was blocked in mice by genetic
which have reduced proinflammatory activity as compared with IL-17A modification [51] or the use of antibodies. In addition to IL-21, TfH cells
homodimers aggravates lupus pathology [38]. As already discussed also provide contact-dependent T cell help using molecules like CD40L-
above, transcription of CREM itself is dependent on CamKIV activation CD40 [52], inducible T cell co-stimulator (ICOS)- ICOSL signaling and
[31]. As a proof of principle, pharmacologic inhibition of CamKIV by the signaling lymphocyte activation molecule-associated protein (SAP) [2].
inhibitor KN93 or genetic inhibition of CamKIV decreases frequency of The first one has been evaluated in clinical trials, yet not proven to be
IL-17-producing cells and efficiently decreases lupus-like disease in promising due to enhanced thromboembolic complications, most likely
murine models [32,34]. In addition to CamKIV, Rock2, one of two Rho since CD40 is also expressed on thrombocytes [4]. However, mice
kinases has an increased activity in SLE PBMCs and blockade of this models demonstrated decreased autoantibody production after use of
kinase reverts enhanced IL-17 as well as IL-21 expression. Moreover, blocking antibodies against ICOSL [53] and CD40L [52].
Th17 cells express TLR2, which upon activation increases histone With regard to IL-21, the downregulation of the CD3 zeta chain in­
acetylation and reduces methylation of the IL17A and IL17F promoter duces an altered lipid raft formation at the T cellular receptor complex.
region [50]. Moreover, apart from CD4 T cells also CD4-CD8- double This affects signal transduction of the immunological synapse in T cells,
negative (DN) T cells infiltrate the kidney and contribute to tissue which results in a lower excitation threshold and enhanced Ca2+
damage by production of IL-17 [39]. Interestingly, CREM binding to the signaling [14]. Mice with a transgenic overexpression of CREMα spe­
promoter of the transmembrane glycoprotein CD8 contributes to the cifically in T cells display decreased IL-2 expression, but strikingly, and
generation of pathogenic DN T cells. With that respect, we were able to in analogy to human SLE also enhanced Ca2+ signaling in T cells. Ca2+ is
show that CREMα overexpression in a Fas (CD95)− /− lupus mouse model central for IL-21 transcription, since it activates NFAT, which is the most
strongly accelerates lymphadenopathy and splenomegaly, which was important transcription factor for activation of the IL-21 promoter [54].
related to a massive expansion of DN T cells [40]. In addition, a CRE-halfside was identified within the IL21 promoter,
Recent papers suggest that gut microbiota dysbiosis is involved in which is crucial for the whole promoter activity and is able to bind
SLE pathogenesis [41–44]. It is currently unclear whether this is caus­ CREMα. Interestingly, in analogy to the murine and human IL17 pro­
ative or merely an epiphenomenon. It has been suggested that the moter CREMα displays activator function for IL-21. Autoimmune-prone

3
K. Tenbrock and T. Rauen Clinical Immunology 239 (2022) 109031

Fas (CD95)− /− mice with a transgenic overexpression of CREM in T cells Notch-3 axis in SLE patients [69]. Y-box binding protein-1 (YB-1) is a
show enhanced anti-dsDNA autoantibody and enhanced total IgG pro­ pleiotropic protein that acts as transcriptional or translational regulator,
duction [55,56] which could be related to this fact. but may also be actively secreted by inflammatory cells such as activated
In addition to CREM, other factors influence IL-21 expression in TfH monocytes [70]. In sera from active SLE patients and kidneys with lupus
cells in SLE as well including signal transducer and activator of tran­ nephritis, we found increased levels of YB-1 with a specific post­
scription factor STAT4. Levels of pSTAT4 expression in circulating Tfh- translational modification, i.e. guanidinylated YB-1 (YB-1-G). Notably,
like cells from SLE patients were significantly increased as compared to YB-1-G can bind to the extracellular domains of transmembrane Notch-3
healthy controls [57]. Maintenance of Tfh cell cytokine synthesis receptors and thereby activate its intracellular signaling processes that
including IL-21 was driven by type I IFNs, which activate pSTAT4 eventually result in enhanced IL-10 production. Along these lines, lupus-
signaling. In addition to follicular T helper cells, which are characterized prone mice with constitutional Notch-3 depletion showed an aggravated
by CXCR5 expression and provide B cell help in germinal centers, PD- lupus phenotype with significantly increased mortality, enlarged
1hiCXCR5-CD4+ T cells from SLE patients are T peripheral helper (Tph) lymphoid organs and fewer Tregs, however reduced IL-10 amounts and
cells, a CXCR5- T cell population that also stimulates B cell responses via aggravated nephritis [69].
IL-21 in the extrafollicular department. This population is critically Specific blockade of IL-10 prevents B cell responses in a murine
dependent on the transcription factor c-MAF and a deletion of MAF model of SLE [8]. In a recent publication extrafollicular, IL-10-
inhibited both, IL21 and IL10 mRNA expression in human CD4+ T cells producing CCR6+ helper T cells were identified which are found high­
and reduced secretion of IL-21 [58]. ly abundant in lymph nodes of SLE patients and colocalize with B cells at
Beyond Tfh cells, there is evidence of extrafollicular germinal center the margins of follicles. These cells can also be identified in lymph nodes
existence in patients with SLE. B cells and plasma cells within these and circulation of autoimmune prone mice [71]. While it is clear that IL-
extrafollicular germinal centers depend on the help of extrafollicular 10 exerts anti-inflammatory properties, it is also an important cytokine
CD4+ T cells. As in TfH cells, extrafollicular CD4+ T cells express Bcl-6 for the B cell response against e.g. plasmodium infections [72] and
as a master transcriptional regulator and their antibody response de­ interestingly, patients with Plasmodium falciparum malaria display
pends on IL-21. While these extrafollicularly enriched circulating CD4+ increased levels of circulating anti-DNA antibodies and immune com­
T cells express the lymph node homing receptor CXCR5+ and have been plexes containing parasite DNA [73].
documented in lupus nephritic kidneys in close contact with B cells,
there is evidence for a CXCR5-negative CD4+ T helper cell subset named 1.6. Tregs in SLE
peripheral (TpH) cells. These cells have also been identified in other
rheumatic diseases including juvenile idiopathic arthritis [59]. Since Regulatory T cells (Tregs) are gatekeepers of immunological toler­
these cells also produce IL-21 and provide B cell help [58,60] it is now ance, characterized by the master transcriptional regulator Foxp3 [74]
quite accepted that there are different populations of IL-21-producing T and have been extensively studied with respect to numbers and partially
helper cells that provide B cell help in different niches including the function in human SLE. However, the results are mixed, which is related
blood and which may transmigrate to inflamed tissue. to different markers used for staining and isolation of these cells. Foxp3
It has been recently published that function and expression of TFH can naturally not be used for sorting and functional analysis, since it is
cells is metabolically controlled by phosphatidylethanolamine [61]. an intracellular marker. Therefore, some studies report reduced
Interestingly, lipid profiles of SLE patients are different from patients numbers or impaired function of circulating Treg cells in SLE patients
with rheumatoid arthritis (RA) and healthy controls. UPLC-MS/MS- [75,76], while other groups observed no abnormalities [77,78]. Indeed,
based plasma lipidomics reveal strikingly enhanced levels of phospha­ some studies report increased levels of Treg cells in SLE as compared to
tidylethanolamine in accordance to disease activity levels as measured those in healthy controls [79,80] or a resistance of lupus effector T cells
by SLEDAI [62]. Moreover, anti-phosphatidylethanolamine antibodies to Treg suppression instead of functional defects of SLE Treg cells [81].
play a key role in the diagnosis of the antiphospholipid syndrome. In The expression of Foxp3 and thus the functional capacities of Tregs
addition, mechanistic target of rapamycin (mTOR) is enhanced in SLE critically dependent on the methylation status within the so called
CD4+ T cells and IL-21 itself stimulates mTORC1 and mTORC2. IL-21 TSDR. There is currently no evidence that this methylation status is
thereby impairs autophagy, differentiation, and function of Treg cells impaired in SLE Tregs suggesting that there is no primary defect in Tregs
[63]. in SLE. Nevertheless, different cytokines within the inflamed local
milieu might influence Treg function. As mentioned above, IL-21 affects
1.5. Regulation of IL-10 expression mTOR activity and mTOR is one of the important energy sensors in
living cells including Tregs [63].
IL-10 is regarded as an antiinflammatory cytokine, which is one of One critical point with regard to Treg expansion is indeed the
the key regulatory molecules mainly provided by Tregs, but also by type external supplementation with IL-2, since Tregs are not able to produce
1 regulatory cells (Tr1), which are primarily characterized by IL-10 IL-2 by themselves. This again suggest a rather secondary effect in Tregs
production. The switch towards a Tr1 phenotype can be mediated by exemplified by the fact that low dose IL-2 treatment enhances expression
T cell activation via complement regulator CD46, again a receptor of Tregs in human SLE patients as well as in murine lupus models.
dependent on IL-2 and expression of CREM/ICER [64]. IL-10 inhibits
upregulation of MHCII on antigen-presenting cells and prevents the 1.7. CD8 T cells in SLE
production of cytokines by inhibition of glycolysis and promotion of
oxidative phosphorylation. Furthermore, IL-10 suppresses mammalian CD8 effector T cells are important for the clearance of virus-infected
target of rapamycin (mTOR) activity in antigen-presenting cells [65]. cells through production of perforins and granzymes. Several functional
Despite its tolerogenic effects on T cells, IL-10 is a well-known B cell deficiencies in CD8 T cells obtained from SLE patients have been noted
growth and differentiation factor. It enhances survival, proliferation, like downregulation of SLAM4 [82], which might correlate with poor
isotype switching, and differentiation of human B cells to plasma cells control of Epstein-Barr virus infection [83,84] and predisposition to
and has thus been associated with SLE. infection. With that respect, low dose IL-2 therapy enhances viral
Increased levels of IL-10 were found in sera and tissues from SLE clearance and control of viral infections in SLE patients as well as in
patients [59,66] and have been linked to reduced DNA methylation of lupus mouse models infected with influenza A [85], however also ex­
the IL10 gene [67]. Furthermore, blockade of IL-10 relieved some of the pands IL-13- and interferon-γ-producing CD8 + T cells via the STAT6-
cellular features in vitro [68]. We were able to demonstrate that IL-10 GATA-3 axis in SLE patients [86]. These cells can be pathogenic under
production is also regulated at the extracellular level via the YB-1/ certain circumstances. This is particularly important since it has not

4
K. Tenbrock and T. Rauen Clinical Immunology 239 (2022) 109031

been noticed within the local inflamed tissue. Interestingly some of these Kiss, S. Hasni, P.M. Izmirly, M. Jung, G. Kumánovics, X. Mariette, I. Padjen, J.
M. Pego-Reigosa, J. Romero-Diaz, Í. Rúa-Figueroa Fernández, R. Seror, G.
CD8 effector cells show a metabolically and functionally exhausted
H. Stummvoll, Y. Tanaka, M.G. Tektonidou, C. Vasconcelos, E.M. Vital, D.
phenotype [87] while others are functionally intact and build different J. Wallace, S. Yavuz, P.L. Meroni, M.J. Fritzler, R. Naden, T. Dörner, S.
clusters of CD8 effector cells, e.g. in affected kidneys in cooperation with R. Johnson, European league against rheumatism/American College of
other cell types like NK cells. Thus, there are discrepancies between the Rheumatology classification criteria for systemic lupus erythematosus, Ann.
Rheum. Dis. 78 (2019) (2019) 1151–1159.
functional effector properties in tissue resident CD8 cells and blood [2] J.M. Odegard, B.R. Marks, L.D. DiPlacido, A.C. Poholek, D.H. Kono, C. Dong, R.
derived CD8 cells, which are not easily explained and which could be A. Flavell, J. Craft, ICOS-dependent extrafollicular helper T cells elicit IgG
affected by future treatment regimens. production via IL-21 in systemic autoimmunity, J. Exp. Med. 205 (2008)
2873–2886.
[3] B.A. Sullivan, W. Tsuji, A. Kivitz, J. Peng, G.E. Arnold, M.J. Boedigheimer,
1.8. Double Negative T cells in SLE K. Chiu, C.L. Green, A. Kaliyaperumal, C. Wang, J. Ferbas, J.B. Chung, Inducible
T-cell co-stimulator ligand (ICOSL) blockade leads to selective inhibition of anti-
KLH IgG responses in subjects with systemic lupus erythematosus, Lupus Sci.
CD4-CD8- double negative (DN) T cells are expanded in SL, which Med. 3 (2016), e000146.
most probably arise from CD8+ T cells. DN T cells are pathogenic, since [4] M. Ramanujam, J. Steffgen, S. Visvanathan, C. Mohan, J.S. Fine, C. Putterman,
Phoenix from the flames: rediscovering the role of the CD40-CD40L pathway in
they produce IL-17, infiltrate the kidney and induce tissue damage [39]. systemic lupus erythematosus and lupus nephritis, Autoimmun. Rev. 19 (2020),
It was shown that CREMα trans-represses CD8 expression. CREMα or­ 102668.
chestrates epigenetic remodeling of the CD8 cluster through the [5] C.S. Via, G.C. Tsokos, B. Bermas, M. Clerici, G.M. Shearer, T cell-antigen-
presenting cell interactions in human systemic lupus erythematosus. Evidence for
recruitment of DNMT 3a and histone methyltransferase G9a [88–90]. heterogeneous expression of multiple defects, J. Immunol. 151 (1993)
CD8 T cells lose CD8 expression and become DN only when cognate Ag is 3914–3922.
sensed as self, this means that infectious antigens are unable to induce [6] W. Liao, J.X. Lin, W.J. Leonard, IL-2 family cytokines: new insights into the
complex roles of IL-2 as a broad regulator of T helper cell differentiation, Curr.
DN T cells. In addition, a loss of marginal zone macrophages (MZM) is
Opin. Immunol. 23 (2011) 598–604.
described in patients with SLE and induces expression of DN T cells [7] J.D. Fontenot, J.P. Rasmussen, M.A. Gavin, A.Y. Rudensky, A function for
[90–92]. The loss of these macrophages impairs the tolerogenic clear­ interleukin 2 in Foxp3-expressing regulatory T cells, Nat. Immunol. 6 (2005)
ance of apoptotic cells. MZM defects generate an inflammatory milieu, 1142–1151.
[8] L. Llorente, W. Zou, Y. Levy, Y. Richaud-Patin, J. Wijdenes, J. Alcocer-Varela,
in which elevated IL-23 along with reduced TGFβ facilitate self-reactive B. Morel-Fourrier, J.C. Brouet, D. Alarcon-Segovia, P. Galanaud, D. Emilie, Role
DN T-cell activation, expansion, and survival, which are generated from of interleukin 10 in the B lymphocyte hyperactivity and autoantibody production
self-reactive CD8+ T cells. Importantly, the therapeutic effects of rapa­ of human systemic lupus erythematosus, J. Exp. Med. 181 (1995) 839–844.
[9] M. Miao, Y. Li, D. Xu, R. Zhang, J. He, X. Sun, Z. Li, Therapeutic responses and
mycin in SLE were associated with the reduction of IL-17-producing DN predictors of low-dose interleukin-2 in systemic lupus erythematosus, Clin. Exp.
T cells in a recently published trial [93]. Rheumatol. (2021), https://doi.org/10.55563/clinexprheumatol/1o6pn1. Online
ahead of print.
[10] C. von Spee-Mayer, E. Siegert, D. Abdirama, A. Rose, A. Klaus, T. Alexander,
1.9. T cell signatures as potential biomarkers P. Enghard, B. Sawitzki, F. Hiepe, A. Radbruch, G.R. Burmester, G. Riemekasten,
J.Y. Humrich, Low-dose interleukin-2 selectively corrects regulatory T cell defects
in patients with systemic lupus erythematosus, Ann. Rheum. Dis. 75 (2016)
With regard to personalized medicine approaches it becomes 1407–1415.
increasingly important to stratify patients into categories related to [11] J. He, X. Zhang, Y. Wei, X. Sun, Y. Chen, J. Deng, Y. Jin, Y. Gan, X. Hu, R. Jia,
organ involvement, disease manifestations and biomarkers that define C. Xu, Z. Hou, Y.A. Leong, L. Zhu, J. Feng, Y. An, Y. Jia, C. Li, X. Liu, H. Ye, L. Ren,
R. Li, H. Yao, Y. Li, S. Chen, X. Zhang, Y. Su, J. Guo, N. Shen, E.F. Morand, D. Yu,
response to therapy. The so called interferon signature, which compro­ Z. Li, Low-dose interleukin-2 treatment selectively modulates CD4(+) T cell
mises a RNA expression analysis of 5–15 genes is one of the biomarkers subsets in patients with systemic lupus erythematosus, Nat. Med. 22 (2016)
used but there are conflicting results regarding their predictive values 991–993.
[12] J. He, R. Zhang, M. Shao, X. Zhao, M. Miao, J. Chen, J. Liu, X. Zhang, X. Zhang,
with respect to disease activity, prednisolone usage and therapeutic
Y. Jin, Y. Wang, S. Zhang, L. Zhu, A. Jacob, R. Jia, X. You, X. Li, C. Li, Y. Zhou,
response [94–96]. This situation is even more complex with regard to T Y. Yang, H. Ye, Y. Liu, Y. Su, N. Shen, J. Alexander, J. Guo, J. Ambrus, X. Lin,
cell specific markers. Numerous markers on T cells including DNA D. Yu, X. Sun, Z. Li, Efficacy and safety of low-dose IL-2 in the treatment of
systemic lupus erythematosus: a randomised, double-blind, placebo-controlled
methylations on promoters and miRNA expression have been exploited
trial, Ann. Rheum. Dis. 79 (2020) 141–149.
in SLE cohorts and correlated with SLEDAI scores with mixed results [13] B. Zhang, J. Sun, Y. Wang, D. Ji, Y. Yuan, S. Li, Y. Sun, Y. Hou, P. Li, L. Zhao,
[97–100]. This may reflect the complexity of disease and also the F. Yu, W. Ma, B. Cheng, L. Wu, J. Hu, M. Wang, W. Song, X. Li, H. Li, Y. Fei,
different involvement of T cells in different organ manifestations. H. Chen, L. Zhang, G.C. Tsokos, D. Zhou, X. Zhang, Site-specific PEGylation of
interleukin-2 enhances immunosuppression via the sustained activation of
Nevertheless, it is highly probable that future therapeutic strategies regulatory T cells, Nat. Biomed. Eng. 5 (2021) 1288–1305.
include biomarker analysis prior to start of medication. [14] K. Tenbrock, Y.T. Juang, V.C. Kyttaris, G.C. Tsokos, Altered signal transduction in
SLE T cells, Rheumatology (Oxford) 46 (2007) 1525–1530.
[15] G.C. Tsokos, M.P. Nambiar, Y.T. Juang, Activation of the Ets transcription factor
2. Conclusions Elf-1 requires phosphorylation and glycosylation: defective expression of
activated Elf-1 is involved in the decreased TCR zeta chain gene expression in
T cells are of utmost importance in the pathogenesis of SLE and patients with systemic lupus erythematosus, Ann. N. Y. Acad. Sci. 987 (2003)
240–245.
already serve as drug targets. IL-2 is the example par excellence in which [16] Y.T. Juang, K. Tenbrock, M.P. Nambiar, M.F. Gourley, G.C. Tsokos, Defective
a clinical observation made it to the bench and back to the patient and production of functional 98-kDa form of Elf-1 is responsible for the decreased
thus resulted in a therapeutic intervention that is about to reach expression of TCR zeta-chain in patients with systemic lupus erythematosus,
J. Immunol. 169 (2002) 6048–6055.
approval by regulatory authorities. Some of the herein described T cell [17] Y.T. Juang, E.E. Solomou, B. Rellahan, G.C. Tsokos, Phosphorylation and O-
abnormalities might also be suitable targets for therapeutic in­ linked glycosylation of Elf-1 leads to its translocation to the nucleus and binding
terventions in the future. With respect to our patients, they are urgently to the promoter of the TCR zeta-chain, J. Immunol. 168 (2002) 2865–2871.
[18] K. Tenbrock, V.C. Kyttaris, M. Ahlmann, J.M. Ehrchen, M. Tolnay, H. Melkonyan,
needed. C. Mawrin, J. Roth, C. Sorg, Y.T. Juang, G.C. Tsokos, The cyclic AMP response
element modulator regulates transcription of the TCR zeta-chain, J. Immunol.
References 175 (2005) 5975–5980.
[19] E.E. Solomou, Y.T. Juang, M.F. Gourley, G.M. Kammer, G.C. Tsokos, Molecular
basis of deficient IL-2 production in T cells from patients with systemic lupus
[1] M. Aringer, K. Costenbader, D. Daikh, R. Brinks, M. Mosca, R. Ramsey-Goldman,
erythematosus, J. Immunol. 166 (2001) 4216–4222.
J.S. Smolen, D. Wofsy, D.T. Boumpas, D.L. Kamen, D. Jayne, R. Cervera,
[20] K. Tenbrock, Y.T. Juang, M.F. Gourley, M.P. Nambiar, G.C. Tsokos, Antisense
N. Costedoat-Chalumeau, B. Diamond, D.D. Gladman, B. Hahn, F. Hiepe,
cyclic adenosine 5′ -monophosphate response element modulator up-regulates IL-
S. Jacobsen, D. Khanna, K. Lerstrøm, E. Massarotti, J. McCune, G. Ruiz-Irastorza,
2 in T cells from patients with systemic lupus erythematosus, J. Immunol. 169
J. Sanchez-Guerrero, M. Schneider, M. Urowitz, G. Bertsias, B.F. Hoyer,
(2002) 4147–4152.
N. Leuchten, C. Tani, S.K. Tedeschi, Z. Touma, G. Schmajuk, B. Anic, F. Assan, T.
M. Chan, A.E. Clarke, M.K. Crow, L. Czirják, A. Doria, W. Graninger, B. Halda-

5
K. Tenbrock and T. Rauen Clinical Immunology 239 (2022) 109031

[21] K. Tenbrock, Y.T. Juang, M. Tolnay, G.C. Tsokos, The cyclic adenosine 5′ - landscape of the gut microbiome of systemic lupus erythematosus in Japanese,
monophosphate response element modulator suppresses IL-2 production in Ann. Rheum. Dis. 80 (2021) 1575–1583.
stimulated T cells by a chromatin-dependent mechanism, J. Immunol. 170 (2003) [42] H. Wang, G. Wang, N. Banerjee, Y. Liang, X. Du, P.J. Boor, K.L. Hoffman, M.
2971–2976. F. Khan, Aberrant gut microbiome contributes to intestinal oxidative stress,
[22] K. Tenbrock, Y.T. Juang, N. Leukert, J. Roth, G.C. Tsokos, The transcriptional barrier dysfunction, inflammation and systemic autoimmune responses in MRL/
repressor cAMP response element modulator alpha interacts with histone lpr mice, Front. Immunol. 12 (2021), 651191.
deacetylase 1 to repress promoter activity, J. Immunol. 177 (2006) 6159–6164. [43] S.C. Choi, J. Brown, M. Gong, Y. Ge, M. Zadeh, W. Li, B.P. Croker, G. Michailidis,
[23] C.M. Hedrich, T. Rauen, G.C. Tsokos, cAMP-responsive element modulator T.J. Garrett, M. Mohamadzadeh, L. Morel, Gut microbiota dysbiosis and altered
(CREM)α protein signaling mediates epigenetic remodeling of the human tryptophan catabolism contribute to autoimmunity in lupus-susceptible mice, Sci.
interleukin-2 gene: implications in systemic lupus erythematosus, J. Biol. Chem. Transl. Med. 12 (2020).
286 (2011) 43429–43436. [44] M. Guo, H. Wang, S. Xu, Y. Zhuang, J. An, C. Su, Y. Xia, J. Chen, Z.Z. Xu, Q. Liu,
[24] C.M. Hedrich, J.C. Crispin, T. Rauen, C. Ioannidis, S.A. Apostolidis, M.S. Lo, V. J. Wang, Z. Dan, K. Chen, X. Luan, Z. Liu, K. Liu, F. Zhang, Y. Xia, X. Liu,
C. Kyttaris, G.C. Tsokos, cAMP response element modulator α controls IL2 and Alteration in gut microbiota is associated with dysregulation of cytokines and
IL17A expression during CD4 lineage commitment and subset distribution in glucocorticoid therapy in systemic lupus erythematosus, Gut Microbes 11 (2020)
lupus, Proc. Natl. Acad. Sci. U. S. A. 109 (2012) 16606–16611. 1758–1773.
[25] Y.T. Juang, T. Rauen, Y. Wang, K. Ichinose, K. Benedyk, K. Tenbrock, G.C. Tsokos, [45] D.S. Kim, Y. Park, J.W. Choi, S.H. Park, M.L. Cho, S.K. Kwok, Lactobacillus
Transcriptional activation of the cAMP-responsive modulator promoter in human acidophilus supplementation exerts a synergistic effect on tacrolimus efficacy by
T cells is regulated by protein phosphatase 2A-mediated dephosphorylation of SP- modulating Th17/Treg balance in lupus-prone mice via the SIGNR3 pathway,
1 and reflects disease activity in patients with systemic lupus erythematosus, Front. Immunol. 12 (2021), 696074.
J. Biol. Chem. 286 (2011) 1795–1801. [46] M. Kono, N. Yoshida, K. Maeda, G.C. Tsokos, Transcriptional factor ICER
[26] T. Rauen, K. Benedyk, Y.T. Juang, C. Kerkhoff, V.C. Kyttaris, J. Roth, G.C. Tsokos, promotes glutaminolysis and the generation of Th17 cells, Proc. Natl. Acad. Sci.
K. Tenbrock, A novel intronic cAMP response element modulator (CREM) U. S. A. 115 (2018) 2478–2483.
promoter is regulated by activator protein-1 (AP-1) and accounts for altered [47] M. Kono, N. Yoshida, K. Maeda, A. Suárez-Fueyo, V.C. Kyttaris, G.C. Tsokos,
activation-induced CREM expression in T cells from patients with systemic lupus Glutaminase 1 inhibition reduces glycolysis and ameliorates lupus-like disease in
erythematosus, J. Biol. Chem. 286 (2011) 32366–32372. MRL/lpr mice and experimental autoimmune encephalomyelitis, arthritis,
[27] Q. Xu, X. Jin, M. Zheng, D. Rohila, G. Fu, Z. Wen, J. Lou, S. Wu, R. Sloan, L. Wang, Rheumatol 71 (2019) 1869–1878.
H. Hu, X. Gao, L. Lu, Phosphatase PP2A is essential for T(H)17 differentiation, [48] H. Kato, A. Perl, Mechanistic target of rapamycin complex 1 expands Th17 and IL-
Proc. Natl. Acad. Sci. U. S. A. 116 (2019) 982–987. 4+ CD4-CD8- double-negative T cells and contracts regulatory T cells in systemic
[28] J.C. Crispín, S.A. Apostolidis, F. Rosetti, M. Keszei, N. Wang, C. Terhorst, T. lupus erythematosus, J. Immunol. 192 (2014) 4134–4144.
N. Mayadas, G.C. Tsokos, Cutting edge: protein phosphatase 2A confers [49] D. Cluxton, A. Petrasca, B. Moran, J.M. Fletcher, Differential regulation of human
susceptibility to autoimmune disease through an IL-17-dependent mechanism, Treg and Th17 cells by fatty acid synthesis and glycolysis, Front. Immunol. 10
J. Immunol. 188 (2012) 3567–3571. (2019) 115.
[29] J.C. Crispín, S.A. Apostolidis, M.I. Finnell, G.C. Tsokos, Induction of PP2A Bβ, a [50] J. Shan, H. Jin, Y. Xu, T cell metabolism: a new perspective on Th17/Treg cell
regulator of IL-2 deprivation-induced T-cell apoptosis, is deficient in systemic imbalance in systemic lupus erythematosus, Front. Immunol. 11 (2020) 1027.
lupus erythematosus, Proc. Natl. Acad. Sci. U. S. A. 108 (2011) 12443–12448. [51] J.A. Bubier, T.J. Sproule, O. Foreman, R. Spolski, D.J. Shaffer, H.C. Morse 3rd, W.
[30] C.G. Katsiari, V.C. Kyttaris, Y.T. Juang, G.C. Tsokos, Protein phosphatase 2A is a J. Leonard, D.C. Roopenian, A critical role for IL-21 receptor signaling in the
negative regulator of IL-2 production in patients with systemic lupus pathogenesis of systemic lupus erythematosus in BXSB-Yaa mice, Proc. Natl.
erythematosus, J. Clin. Invest. 115 (2005) 3193–3204. Acad. Sci. U. S. A. 106 (2009) 1518–1523.
[31] Y.T. Juang, Y. Wang, E.E. Solomou, Y. Li, C. Mawrin, K. Tenbrock, V.C. Kyttaris, [52] A. Oxenius, K.A. Campbell, C.R. Maliszewski, T. Kishimoto, H. Kikutani,
G.C. Tsokos, Systemic lupus erythematosus serum IgG increases CREM binding to H. Hengartner, R.M. Zinkernagel, M.F. Bachmann, CD40-CD40 ligand
the IL-2 promoter and suppresses IL-2 production through CaMKIV, J. Clin. interactions are critical in T-B cooperation but not for other anti-viral CD4+ T cell
Invest. 115 (2005) 996–1005. functions, J. Exp. Med. 183 (1996) 2209–2218.
[32] K. Ichinose, Y.T. Juang, J.C. Crispín, K. Kis-Toth, G.C. Tsokos, Suppression of [53] H. Ding, X. Wu, J. Wu, H. Yagita, Y. He, J. Zhang, J. Ren, W. Gao, Delivering PD-1
autoimmunity and organ pathology in lupus-prone mice upon inhibition of inhibitory signal concomitant with blocking ICOS co-stimulation suppresses
calcium/calmodulin-dependent protein kinase type IV, Arthritis Rheum. 63 lupus-like syndrome in autoimmune BXSB mice, Clin. Immunol. 118 (2006)
(2011) 523–529. 258–267.
[33] T. Rauen, C.M. Hedrich, Y.T. Juang, K. Tenbrock, G.C. Tsokos, cAMP-responsive [54] H.P. Kim, L.L. Korn, A.M. Gamero, W.J. Leonard, Calcium-dependent activation
element modulator (CREM)α protein induces interleukin 17A expression and of interleukin-21 gene expression in T cells, J. Biol. Chem. 280 (2005)
mediates epigenetic alterations at the interleukin-17A gene locus in patients with 25291–25297.
systemic lupus erythematosus, J. Biol. Chem. 286 (2011) 43437–43446. [55] R. Lippe, K. Ohl, G. Varga, T. Rauen, J.C. Crispin, Y.T. Juang, S. Kuerten, F. Tacke,
[34] T. Koga, C.M. Hedrich, M. Mizui, N. Yoshida, K. Otomo, L.A. Lieberman, M. Wolf, K. Roebrock, T. Vogl, E. Verjans, N. Honke, J. Ehrchen, D. Foell,
T. Rauen, J.C. Crispin, G.C. Tsokos, CaMK4-dependent activation of AKT/mTOR B. Skryabin, N. Wagner, G.C. Tsokos, J. Roth, K. Tenbrock, CREMalpha
and CREM-alpha underlies autoimmunity-associated Th17 imbalance, J. Clin. overexpression decreases IL-2 production, induces a T(H)17 phenotype and
Invest. 124 (2014) 2234–2245. accelerates autoimmunity, J. Mol. Cell Biol. 4 (2012) 121–123.
[35] A. Poissonnier, D. Sanséau, M. Le Gallo, M. Malleter, N. Levoin, R. Viel, L. Morere, [56] K. Ohl, A. Wiener, A. Schippers, N. Wagner, K. Tenbrock, Interleukin-2 treatment
A. Penna, P. Blanco, A. Dupuy, F. Poizeau, A. Fautrel, J. Seneschal, F. Jouan, reverses effects of cAMP-responsive element modulator alpha-over-expressing T
J. Ritz, E. Forcade, N. Rioux, C. Contin-Bordes, T. Ducret, A.M. Vacher, P. cells in autoimmune-prone mice, Clin. Exp. Immunol. 181 (2015) 76–86.
A. Barrow, R.J. Flynn, P. Vacher, P. Legembre, CD95-mediated calcium signaling [57] X. Dong, O.Q. Antao, W. Song, G.M. Sanchez, K. Zembrzuski, F. Koumpouras,
promotes T helper 17 trafficking to inflamed organs in lupus-prone mice, A. Lemenze, J. Craft, J.S. Weinstein, Type I interferon-activated STAT4 regulation
Immunity 45 (2016) 209–223. of follicular helper T cell-dependent cytokine and immunoglobulin production in
[36] S.R. Hofmann, K. Mäbert, F. Kapplusch, S. Russ, S. Northey, M.W. Beresford, G. lupus, arthritis, Rheumatol 73 (2021) 478–489.
C. Tsokos, C.M. Hedrich, cAMP response element modulator α induces dual [58] A.V. Bocharnikov, J. Keegan, V.S. Wacleche, Y. Cao, C.Y. Fonseka, G. Wang, E.
specificity protein phosphatase 4 to promote effector T cells in juvenile-onset S. Muise, K.X. Zhang, A. Arazi, G. Keras, Z.J. Li, Y. Qu, M.F. Gurish, M. Petri, J.
lupus, J. Immunol. 203 (2019) 2807–2816. P. Buyon, C. Putterman, D. Wofsy, J.A. James, J.M. Guthridge, B. Diamond, J.
[37] N. Yoshida, D. Comte, M. Mizui, K. Otomo, F. Rosetti, T.N. Mayadas, J.C. Crispín, H. Anolik, M.F. Mackey, S.E. Alves, P.A. Nigrovic, K.H. Costenbader, M.
S.J. Bradley, T. Koga, M. Kono, M.P. Karampetsou, V.C. Kyttaris, K. Tenbrock, G. B. Brenner, J.A. Lederer, D.A. Rao, PD-1hiCXCR5- T peripheral helper cells
C. Tsokos, ICER is requisite for Th17 differentiation, Nat. Commun. 7 (2016) promote B cell responses in lupus via MAF and IL-21, JCI Insight 4 (2019).
12993. [59] J. Fischer, J. Dirks, J. Klaussner, G. Haase, A. Holl-Wieden, C. Hofmann,
[38] C.M. Hedrich, T. Rauen, K. Kis-Toth, V.C. Kyttaris, G.C. Tsokos, cAMP-responsive S. Hackenberg, H. Girschick, H. Morbach, Effect of clonally expanded PD-1(high)
element modulator α (CREMα) suppresses IL-17F protein expression in T CXCR5-CD4+ peripheral T helper cells on B cell differentiation in the joints of
lymphocytes from patients with systemic lupus erythematosus (SLE), J. Biol. patients with antinuclear antibody-positive juvenile idiopathic arthritis, arthritis,
Chem. 287 (2012) 4715–4725. Rheumatol 74 (2022) 150–162.
[39] J.C. Crispín, M. Oukka, G. Bayliss, R.A. Cohen, C.A. Van Beek, I.E. Stillman, V. [60] A. Arazi, D.A. Rao, C.C. Berthier, A. Davidson, Y. Liu, P.J. Hoover, A. Chicoine, T.
C. Kyttaris, Y.T. Juang, G.C. Tsokos, Expanded double negative T cells in patients M. Eisenhaure, A.H. Jonsson, S. Li, D.J. Lieb, F. Zhang, K. Slowikowski, E.
with systemic lupus erythematosus produce IL-17 and infiltrate the kidneys, P. Browne, A. Noma, D. Sutherby, S. Steelman, D.E. Smilek, P. Tosta,
J. Immunol. 181 (2008) 8761–8766. W. Apruzzese, E. Massarotti, M. Dall’Era, M. Park, D.L. Kamen, R.A. Furie,
[40] R. Lippe, K. Ohl, G. Varga, T. Rauen, J.C. Crispin, Y.T. Juang, S. Kuerten, F. Tacke, F. Payan-Schober, W.F. Pendergraft 3rd, E.A. McInnis, J.P. Buyon, M.A. Petri,
M. Wolf, K. Roebrock, T. Vogl, E. Verjans, N. Honke, J. Ehrchen, D. Foell, C. Putterman, K.C. Kalunian, E.S. Woodle, J.A. Lederer, D.A. Hildeman,
B. Skryabin, N. Wagner, G.C. Tsokos, J. Roth, K. Tenbrock, CREMα C. Nusbaum, S. Raychaudhuri, M. Kretzler, J.H. Anolik, M.B. Brenner, D. Wofsy,
overexpression decreases IL-2 production, induces a T(H)17 phenotype and N. Hacohen, B. Diamond, The immune cell landscape in kidneys of patients with
accelerates autoimmunity, J. Mol. Cell Biol. 4 (2012) 121–123. lupus nephritis, Nat. Immunol. 20 (2019) 902–914.
[41] Y. Tomofuji, Y. Maeda, E. Oguro-Igashira, T. Kishikawa, K. Yamamoto, [61] G. Fu, C.S. Guy, N.M. Chapman, G. Palacios, J. Wei, P. Zhou, L. Long, Y.D. Wang,
K. Sonehara, D. Motooka, Y. Matsumoto, H. Matsuoka, M. Yoshimura, M. Yagita, C. Qian, Y. Dhungana, H. Huang, A. Kc, H. Shi, S. Rankin, S.A. Brown, A. Johnson,
T. Nii, S. Ohshima, S. Nakamura, H. Inohara, K. Takeda, A. Kumanogoh, R. Wakefield, C.G. Robinson, X. Liu, A. Sheyn, J. Yu, S. Jackowski, H. Chi,
Y. Okada, Metagenome-wide association study revealed disease-specific Metabolic control of T(FH) cells and humoral immunity by
phosphatidylethanolamine, Nature 595 (2021) 724–729.

6
K. Tenbrock and T. Rauen Clinical Immunology 239 (2022) 109031

[62] J. Chen, C. Liu, S. Ye, R. Lu, H. Zhu, J. Xu, UPLC-MS/MS-based plasma lipidomics member 4-positive CD8+ T cells contributes to the decreased cytotoxic cell
reveal a distinctive signature in systemic lupus erythematosus patients, activity in systemic lupus erythematosus, arthritis, Rheumatol 68 (2016)
MedComm 2 (2021) (2020) 269–278. 164–173.
[63] H. Kato, A. Perl, Blockade of Treg cell differentiation and function by the [83] M. Larsen, D. Sauce, C. Deback, L. Arnaud, A. Mathian, M. Miyara, D. Boutolleau,
Interleukin-21-mechanistic target of rapamycin Axis Via suppression of C. Parizot, K. Dorgham, L. Papagno, V. Appay, Z. Amoura, G. Gorochov,
autophagy in patients with systemic lupus erythematosus, arthritis, Rheumatol 70 Exhausted cytotoxic control of Epstein-Barr virus in human lupus, PLoS Pathog. 7
(2018) 427–438. (2011), e1002328.
[64] J. Cardone, G. Le Friec, P. Vantourout, A. Roberts, A. Fuchs, I. Jackson, [84] B.R. Berner, M. Tary-Lehmann, N.L. Yonkers, A.D. Askari, P.V. Lehmann, D.
T. Suddason, G. Lord, J.P. Atkinson, A. Cope, A. Hayday, C. Kemper, Complement D. Anthony, Phenotypic and functional analysis of EBV-specific memory CD8 cells
regulator CD46 temporally regulates cytokine production by conventional and in SLE, Cell. Immunol. 235 (2005) 29–38.
unconventional T cells, Nat. Immunol. 11 (2010) 862–871. [85] P. Zhou, J. Chen, J. He, T. Zheng, J. Yunis, V. Makota, Y.O. Alexandre, F. Gong,
[65] W.K.E. Ip, N. Hoshi, D.S. Shouval, S. Snapper, R. Medzhitov, Anti-inflammatory X. Zhang, W. Xie, Y. Li, M. Shao, Y. Zhu, J.E. Sinclair, M. Miao, Y. Chen, K.
effect of IL-10 mediated by metabolic reprogramming of macrophages, Science R. Short, S.N. Mueller, X. Sun, D. Yu, Z. Li, Low-dose IL-2 therapy invigorates CD8
356 (2017) 513–519. + T cells for viral control in systemic lupus erythematosus, PLoS Pathog. 17
[66] F.A. Houssiau, C. Lefebvre, M. Vanden Berghe, M. Lambert, J.P. Devogelaer, J. (2021), e1009858.
C. Renauld, Serum interleukin 10 titers in systemic lupus erythematosus reflect [86] H. Kato, A. Perl, Double-edged sword: Interleukin-2 promotes T regulatory cell
disease activity, Lupus 4 (1995) 393–395. differentiation but also expands Interleukin-13- and interferon-γ-producing CD8(
[67] C.M. Hedrich, T. Rauen, S.A. Apostolidis, A.P. Grammatikos, N. Rodriguez +) T cells via STAT6-GATA-3 Axis in systemic lupus erythematosus, Front.
Rodriguez, C. Ioannidis, V.C. Kyttaris, J.C. Crispin, G.C. Tsokos, Stat3 promotes Immunol. 12 (2021), 635531.
IL-10 expression in lupus T cells through trans-activation and chromatin [87] J.S. Tilstra, L. Avery, A.V. Menk, R.A. Gordon, S. Smita, L.P. Kane, M. Chikina, G.
remodeling, Proc. Natl. Acad. Sci. U. S. A. 111 (2014) 13457–13462. M. Delgoffe, M.J. Shlomchik, Kidney-infiltrating T cells in murine lupus nephritis
[68] B.R. Lauwerys, N. Garot, J.C. Renauld, F.A. Houssiau, Interleukin-10 blockade are metabolically and functionally exhausted, J. Clin. Invest. 128 (2018)
corrects impaired in vitro cellular immune responses of systemic lupus 4884–4897.
erythematosus patients, Arthritis Rheum. 43 (2000) 1976–1981. [88] C.M. Hedrich, J.C. Crispín, T. Rauen, C. Ioannidis, T. Koga, N. Rodriguez
[69] D.M. Breitkopf, V. Jankowski, K. Ohl, J. Hermann, D. Hermert, K. Tenbrock, Rodriguez, S.A. Apostolidis, V.C. Kyttaris, G.C., Tsokos, cAMP responsive element
X. Liu, I.V. Martin, J. Wang, F. Groll, E. Gröne, J. Floege, T. Ostendorf, T. Rauen, modulator (CREM) α mediates chromatin remodeling of CD8 during the
U. Raffetseder, The YB-1:Notch-3 axis modulates immune cell responses and generation of CD3+ CD4- CD8- T cells, J. Biol. Chem. 289 (2014) 2361–2370.
organ damage in systemic lupus erythematosus, Kidney Int. 97 (2020) 289–303. [89] C.M. Hedrich, T. Rauen, J.C. Crispin, T. Koga, C. Ioannidis, M. Zajdel, V.
[70] B.C. Frye, S. Halfter, S. Djudjaj, P. Muehlenberg, S. Weber, U. Raffetseder, A. En- C. Kyttaris, G.C. Tsokos, cAMP-responsive element modulator α (CREMα) trans-
Nia, H. Knott, J.M. Baron, S. Dooley, J. Bernhagen, P.R. Mertens, Y-box protein-1 represses the transmembrane glycoprotein CD8 and contributes to the generation
is actively secreted through a non-classical pathway and acts as an extracellular of CD3+CD4-CD8- T cells in health and disease, J. Biol. Chem. 288 (2013)
mitogen, EMBO Rep. 10 (2009) 783–789. 31880–31887.
[71] F. Facciotti, P. Larghi, R. Bosotti, C. Vasco, N. Gagliani, C. Cordiglieri, S. Mazzara, [90] H. Li, I.E. Adamopoulos, V.R. Moulton, I.E. Stillman, Z. Herbert, J.J. Moon,
V. Ranzani, E. Rottoli, S. Curti, A. Penatti, B. Karnani, Y. Kobayashi, M. Crosti, A. Sharabi, S. Krishfield, M.G. Tsokos, G.C. Tsokos, Systemic lupus erythematosus
M. Bombaci, J.P. van Hamburg, G. Rossetti, R. Gualtierotti, M. Gerosa, S. Gatti, favors the generation of IL-17 producing double negative T cells, Nat. Commun.
S. Torretta, L. Pignataro, S.W. Tas, S. Abrignani, M. Pagani, F. Grassi, P.L. Meroni, 11 (2020) 2859.
R.A. Flavell, J. Geginat, Evidence for a pathogenic role of extrafollicular, IL-10- [91] H. Li, Y.X. Fu, Q. Wu, Y. Zhou, D.K. Crossman, P. Yang, J. Li, B. Luo, L.M. Morel,
producing CCR6(+)B helper T cells in systemic lupus erythematosus, Proc. Natl. J.H. Kabarowski, H. Yagita, C.F. Ware, H.C. Hsu, J.D. Mountz, Interferon-induced
Acad. Sci. U. S. A. 117 (2020) 7305–7316. mechanosensing defects impede apoptotic cell clearance in lupus, J. Clin. Invest.
[72] F.A. Surette, J.J. Guthmiller, L. Li, A.J. Sturtz, R. Vijay, R.L. Pope, B.L. McClellan, 125 (2015) 2877–2890.
A.D. Pack, R.A. Zander, P. Shao, L.Y. Lan, D. Fernandez-Ruiz, W.R. Heath, P. [92] H. Li, Q. Wu, J. Li, P. Yang, Z. Zhu, B. Luo, H.C. Hsu, J.D. Mountz, Cutting edge:
C. Wilson, N.S. Butler, Extrafollicular CD4 T cell-derived IL-10 functions rapidly defective follicular exclusion of apoptotic antigens due to marginal zone
and transiently to support anti-plasmodium humoral immunity, PLoS Pathog. 17 macrophage defects in autoimmune BXD2 mice, J. Immunol. 190 (2013)
(2021), e1009288. 4465–4469.
[73] I.C. Hirako, C. Gallego-Marin, M.A. Ataide, W.A. Andrade, H. Gravina, B. [93] Z.W. Lai, R. Kelly, T. Winans, I. Marchena, A. Shadakshari, J. Yu, M. Dawood,
C. Rocha, R.B. de Oliveira, D.B. Pereira, J. Vinetz, B. Diamond, S. Ram, D. R. Garcia, H. Tily, L. Francis, S.V. Faraone, P.E. Phillips, A. Perl, Sirolimus in
T. Golenbock, R.T. Gazzinelli, DNA-containing Immunocomplexes promote patients with clinically active systemic lupus erythematosus resistant to, or
Inflammasome assembly and release of pyrogenic cytokines by CD14+ CD16+ intolerant of, conventional medications: a single-arm, open-label, phase 1/2 trial,
CD64high CD32low inflammatory monocytes from malaria patients, mBio 6 Lancet 391 (2018) 1186–1196.
(2015) e01605–e01615. [94] H. Enocsson, J. Wetterö, M.L. Eloranta, B. Gullstrand, C. Svanberg, M. Larsson, A.
[74] J.D. Fontenot, M.A. Gavin, A.Y. Rudensky, Foxp3 programs the development and A. Bengtsson, L. Rönnblom, C. Sjöwall, Comparison of surrogate markers of the
function of CD4+CD25+ regulatory T cells, Nat. Immunol. 4 (2003) 330–336. type I interferon response and their ability to Mirror disease activity in systemic
[75] M. Bonelli, A. Savitskaya, K. von Dalwigk, C.W. Steiner, D. Aletaha, J.S. Smolen, lupus erythematosus, Front. Immunol. 12 (2021), 688753.
C. Scheinecker, Quantitative and qualitative deficiencies of regulatory T cells in [95] L. Mai, A. Asaduzzaman, B. Noamani, P.R. Fortin, D.D. Gladman, Z. Touma, M.
patients with systemic lupus erythematosus (SLE), Int. Immunol. 20 (2008) B. Urowitz, J. Wither, The baseline interferon signature predicts disease severity
861–868. over the subsequent 5 years in systemic lupus erythematosus, Arthritis Res. Ther.
[76] C. Scheinecker, M. Bonelli, J.S. Smolen, Pathogenetic aspects of systemic lupus 23 (2021) 29.
erythematosus with an emphasis on regulatory T cells, J. Autoimmun. 35 (2010) [96] C. Wilkinson, R.B. Henderson, A.R. Jones-Leone, S.M. Flint, M. Lennon, R.
269–275. A. Levy, B. Ji, D.L. Bass, D. Roth, The role of baseline BLyS levels and type 1
[77] M. Bonelli, A. Savitskaya, C.W. Steiner, E. Rath, J.S. Smolen, C. Scheinecker, interferon-inducible gene signature status in determining belimumab response in
Phenotypic and functional analysis of CD4+ CD25- Foxp3+ T cells in patients systemic lupus erythematosus: a post hoc meta-analysis, Arthritis Res. Ther. 22
with systemic lupus erythematosus, J. Immunol. 182 (2009) 1689–1695. (2020) 102.
[78] M. Bonelli, K. von Dalwigk, A. Savitskaya, J.S. Smolen, C. Scheinecker, Foxp3 [97] S. Vordenbäumen, A. Rosenbaum, C. Gebhard, J. Raithel, A. Sokolowski,
expression in CD4+ T cells of patients with systemic lupus erythematosus: a C. Düsing, G. Chehab, J.G. Richter, R. Brinks, M. Rehli, M. Schneider,
comparative phenotypic analysis, Ann. Rheum. Dis. 67 (2008) 664–671. Associations of site-specific CD4(+)-T-cell hypomethylation within CD40-ligand
[79] Z.J. Yin, B.M. Ju, L. Zhu, N. Hu, J. Luo, M. He, X.Y. Feng, X.H. Lv, D. Pu, L. He, promotor and enhancer regions with disease activity of women with systemic
Increased CD4(+)CD25(− )Foxp3(+) T cells in Chinese systemic lupus lupus erythematosus, Lupus 30 (2021) 45–51.
erythematosus: correlate with disease activity and organ involvement, Lupus 27 [98] R.P. Singh, B.H. Hahn, D.S. Bischoff, Identification and contribution of
(2018) 2057–2068. inflammation-induced novel MicroRNA in the pathogenesis of systemic lupus
[80] T. Alexander, A. Sattler, L. Templin, S. Kohler, C. Groß, A. Meisel, B. Sawitzki, G. erythematosus, Front. Immunol. 13 (2022), 848149.
R. Burmester, R. Arnold, A. Radbruch, A. Thiel, F. Hiepe, Foxp3+ Helios+ [99] Q. Luo, J. Ye, L. Zeng, X. Li, L. Fang, B. Ju, Z. Huang, J. Li, Elevated expression of
regulatory T cells are expanded in active systemic lupus erythematosus, Ann. TIGIT on CD3(+)CD4(+) T cells correlates with disease activity in systemic lupus
Rheum. Dis. 72 (2013) 1549–1558. erythematosus, Allergy, Asthma Clin. Immunol. 13 (2017) 15.
[81] R.K. Venigalla, T. Tretter, S. Krienke, R. Max, V. Eckstein, N. Blank, C. Fiehn, A. [100] D.C. Salazar-Camarena, P. Ortíz-Lazareno, M. Marín-Rosales, A. Cruz, F. Muñoz-
D. Ho, H.M. Lorenz, Reduced CD4+,CD25- T cell sensitivity to the suppressive Valle, R. Tapia-Llanos, G. Orozco-Barocio, R. Machado-Contreras, C.A. Palafox-
function of CD4+,CD25high,CD127 − /low regulatory T cells in patients with Sánchez, BAFF-R and TACI expression on CD3+ T cells: interplay among BAFF,
active systemic lupus erythematosus, Arthritis Rheum. 58 (2008) 2120–2130. APRIL and T helper cytokines profile in systemic lupus erythematosus, Cytokine
[82] K. Kis-Toth, D. Comte, M.P. Karampetsou, V.C. Kyttaris, L. Kannan, C. Terhorst, G. 114 (2019) 115–127.
C. Tsokos, Selective loss of signaling lymphocytic activation molecule family

You might also like