Next Article in Journal
Maytenus octogona Superoxide Scavenging and Anti-Inflammatory Caspase-1 Inhibition Study Using Cyclic Voltammetry and Computational Docking Techniques
Next Article in Special Issue
Changes in the Mitochondria in the Aging Process—Can α-Tocopherol Affect Them?
Previous Article in Journal
The Presence of T Allele (rs35705950) of the MUC5B Gene Predicts Lower Baseline Forced Vital Capacity and Its Subsequent Decline in Patients with Hypersensitivity Pneumonitis
Previous Article in Special Issue
Naturally-Occurring Tyrosinase Inhibitors Classified by Enzyme Kinetics and Copper Chelation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Endosome Traffic Modulates Pro-Inflammatory Signal Transduction in CD4+ T Cells—Implications for the Pathogenesis of Systemic Lupus Erythematosus

1
Department of Medicine, Norton College of Medicine, State University of New York, Upstate Medical University, Syracuse, NY 13210, USA
2
Department of Biochemistry and Molecular Biology, Norton College of Medicine, State University of New York, Upstate Medical University, Syracuse, NY 13210, USA
3
Department of Microbiology and Immunology, Norton College of Medicine, State University of New York, Upstate Medical University, Syracuse, NY 13210, USA
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(13), 10749; https://doi.org/10.3390/ijms241310749
Submission received: 5 May 2023 / Revised: 10 June 2023 / Accepted: 26 June 2023 / Published: 28 June 2023
(This article belongs to the Special Issue Latest Review Papers in Molecular and Cellular Biology 2023)

Abstract

:
Endocytic recycling regulates the cell surface receptor composition of the plasma membrane. The surface expression levels of the T cell receptor (TCR), in concert with signal transducing co-receptors, regulate T cell responses, such as proliferation, differentiation, and cytokine production. Altered TCR expression contributes to pro-inflammatory skewing, which is a hallmark of autoimmune diseases, such as systemic lupus erythematosus (SLE), defined by a reduced function of regulatory T cells (Tregs) and the expansion of CD4+ helper T (Th) cells. The ensuing secretion of inflammatory cytokines, such as interferon-γ and interleukin (IL)-4, IL-17, IL-21, and IL-23, trigger autoantibody production and tissue infiltration by cells of the adaptive and innate immune system that induce organ damage. Endocytic recycling influences immunological synapse formation by CD4+ T lymphocytes, signal transduction from crosslinked surface receptors through recruitment of adaptor molecules, intracellular traffic of organelles, and the generation of metabolites to support growth, cytokine production, and epigenetic control of DNA replication and gene expression in the cell nucleus. This review will delineate checkpoints of endosome traffic that can be targeted for therapeutic interventions in autoimmune and other disease conditions.

1. Introduction

The Rab GTPase subfamily, the largest subfamily of the Ras family of GTPases, is composed of more than 60 family members in humans, and they control different steps of receptor trafficking [1]. Alterations in Rab proteins are connected to a multitude of diseases, such as cancer [2,3,4,5], neurodegenerative diseases [6,7,8], and immune disorders [9]. Receptor recycling and endocytic trafficking control the surface expression levels of T cell receptors (TCR) [10,11]. Altered TCR expression levels lead to dysregulated TCR signaling, contributing to proinflammatory skewing, a critical agent of abnormal T cell activation in autoimmune diseases, including rheumatoid arthritis and systemic lupus erythematosus (SLE) [12].
In contrast to pro-inflammatory cytokines, interleukin 2 (IL-2) plays a unique role as an anti-inflammatory cytokine given its fundamental role in regulatory T cell (Treg) differentiation [13]. SLE patients show reduced numbers and impaired function of Tregs [14,15]. Tregs are important CD4+ T cells that are identified by the expression of the transcription factor, FoxP3, and the inhibition of autoreactive T cells [16]. Tregs function via direct cell-to-cell contact or by the production of immunosuppressive cytokines, such as transforming growth factor β (TGFβ) and IL-10 [17]. TGFβ and IL-6 are required for the development of helper T 17 (Th17) cells, and IL-6 is increased in SLE [18]. Th17 cells are identified by the specific transcription factor, retinoic-acid-receptor-related orphan nuclear receptor γ (RORγt), and the production of IL-17 [19,20]. The IL-17 signature plays an important role in effector-cell-mediated tissue damage by recruiting other pro-inflammatory cells [21]. Follicular helper T (Tfh) cells are CD4+ T cells that maintain the germinal centers that derive autoreactive antibodies secreting plasma cells through the secretion of IL-4 and IL-21 [22]. Tfh cells are abnormally expanded in SLE patients and lupus-prone mice [23].
The expansion of Th17 cells and the increased production of IL-17A correlate with disease activity in SLE patients [24]. Similarly, decreased IL-2 production promotes the imbalance of Th17/Treg, which contributes to organ inflammation and damage in SLE patients [25]. Understanding the underlying defects in SLE patients’ CD4+ T cells that lead to proinflammatory skewing is important in understanding the disease’s pathophysiology. The following sections describe the disturbed balance of proinflammatory and anti-inflammatory cells in SLE and how they are regulated by their altered receptor endosomal recycling, resulting metabolic abnormalities and downstream signaling.

2. Endosomal Trafficking and Recycling Pathways

Cell surface expression of membrane proteins can be modified through changes in vesicular transport that arise from changes in trafficking or the behavior of entire organelles [26]. In response to signaling mechanisms, membrane trafficking changes to increase or decrease the surface expression of proteins [26]. Surface receptor expression can be decreased when stimulated by their ligands, while, for example, in response to insulin, the glucose transporter surface expression can increase [26].
Transmembrane proteins are endocytosed, a process where cargoes (receptors and their bound ligands) are transported in membrane-bound vesicles from the cell surface to inside the cell. There are two main endocytosis pathways, clathrin-independent endocytosis (CIE) and the more well-characterized clathrin-mediated endocytosis (CME) [26,27]. Clathrin is a protein that plays a major role in the formation of clathrin-coated vesicles, which make up to 95% of endocytic vesicles [28]. Once internalized into the cell, regardless of the endocytic pathway, cargoes merge into the common endosomal network: an interconnected “highway” system that controls the trafficking and transfer of cargoes between organelles. The endosomal network collects, sorts, and sends cargoes to their final destinations in the cell [29].
Rab GTPases are the “master regulators” of intracellular trafficking. Rab proteins are localized in distinct compartments, where they recruit effector proteins to regulate the transport between organelles [1,30]. Rab GTPases function as “molecular switches” and cycle between the inactive GDP-bound and active GTP-bound states. Guanine nucleotide exchange factors (GEFs) catalyze the dissociation of GDP from a GTPase for GTP to replace GDP, and GTPase-activating proteins (GAPs) catalyze the hydrolysis of the third phosphate of GTP to create GDP [31]. Since cells internalize their receptors about one to five times an hour [32], regulating Rab GTPase activity is important in retaining vesicle trafficking.
Internalized cargoes converge into a common early endosome, where they are sorted for subsequent transport to different parts of the cell. The acidic environment (pH~6.5) of the early endosome causes ligands to be released from the receptors. Most of these ligands are sorted into the late endosomes and then into the lysosomes for degradation [33]. The receptors themselves can be degraded in the lysosome, but can also have different fates, such as being transported to the trans-Golgi network (TGN) or the plasma membrane for reuse, also known as recycling [34]. While the mechanisms that mediate the vesicular transport along the lysosomal degradative pathway have been well characterized, the regulation of sorting and recycling cargoes is not fully understood.
The Rab proteins are localized to specific endosomes but are segregated to distinct regions of the membrane, defining specialized functional membrane domains (Figure 1) [35]. There are two pathways for recycling back to the plasma membrane: (1) the rapid-recycling pathway, where recycling occurs directly from the early endosome, and (2) the slow-recycling pathway, where recycling occurs indirectly via the endosomal recycling compartment (ERC), a subpopulation of recycling endosomes [35,36]. In the rapid-recycling pathway, receptors are internalized and delivered to the early endosomes that are marked by Rab5, then recycled back onto the cell surface via the recycling pathway under the control of Rab4. Alternatively, during the slow-recycling pathway, cargo proteins are transported from the early endosome to the ERC, then back onto the cell membrane. Rab11 is localized to the ERC and TGN and is important in the slow-recycling pathway [35]. In other parts of intracellular trafficking, Rab9 is localized to the late endosome and TGN [37]. Rab7 is found in late endosomes and lysosomes, and Rab5 and Rab7 are transition cargoes from early endosomes to late endosomes [38].
The composition of the plasma membrane is controlled by the balance between endocytic uptake and recycling, contributing to physiological functions such as nutrient uptake and signal transduction [27,34]. Cell surface proteins that are recycled include proteins involved in nutrient uptake, such as the transferrin receptor (CD71) (for iron), glucose transporters (GLUT), and lipoprotein receptors (for cholesterol); cell adhesion molecules (integrins and cadherins); and cell signaling proteins (i.e., ErbB family proteins) and G protein-coupled receptors (i.e., chemokine receptors) [26,33,39,40,41,42].
Rab5 is required for the formation and function of the early endosome [43,44]. Every transport step requires the GTP-bound, activated Rab GTPases to bind to effector proteins [1]. For Rab5, upon activation by its effector proteins, the Rabaptin-5/Rabex-5 complex [45], it recruits phosphoinositol-3 kinases (PI3K) [46], including hVPS34 [47], subsequently generating phosphoinositol-3-phosphate (PI3P). An environment of PI3P and Rab5 is needed for the docking protein, early endosomal antigen 1 (EEA1), to bind to the endosomal membrane [48,49,50,51]. Rab5, together with EEA1, regulates the fusion between primary endocytic vesicles and sorting endosomes [52,53,54]. Rab35 colocalizes with CD71 on the plasma membrane and vesicles, regulating the rapid recycling of CD71 [55].
Rab4, localized to the early endosomes, recycles glycosphingolipids from early endosomes through rapid recycling [56], and its dominant-negative form inhibits rapid recycling [57]. Rab4 transfers cargoes into either recycling or degradative pathways [58,59]. The HTLV-1 related endogenous retroviral sequence (HRES-1)/Rab4, also designated as Rab4A, is overexpressed in SLE patients’ T cells [10], and it regulates the expression of CD71 [60,61], CD4 [61], and TCRζ [10]. When Rab4A is overexpressed in T cells, endocytic recycling of TCRζ and CD4 is inhibited and targeted for lysosomal degradation [10,61]. The loss of TCRζ and CD4 in SLE T cells changes their downstream TCR signaling pathways [10]. The composite description of endocytic recycling may vary between immune cell types and may affect their functions through downstream pathways through changes in metabolites.

2.1. Endocytic Regulation of Antigen Presentation to CD4+ T Cells

CD4+ T cells are activated through the engagement of the TCR with antigens presented by major histocompatibility complex (MHC)-II molecules [62]. The TCR’s affinity to the self or the agonist peptide-MHC II complex is important in the fate of CD4+ T cells in that it determines whether CD4+ T cells are differentiated into naïve CD4+ T cells or Tregs [63]. Recent studies addressed the role of Rab5, Rab7, Rab9, and Rab11 GTPases in MHC-II-dependent antigen presentation by dendritic cells [64]. MHC-II–peptide complexes are associated with each of these Rab GTPases during dendritic cell maturation, while MHC-II complexes sequentially traffic from Rab5+, Rab7+, and Rab9+ early endosomes to the cell surface via Rab11+ endosomes [64]. MHC-II can also be secreted from dendritic cells in exosomes [65]. Both Rab4A [66] and Rab4B are involved in MHC-II-mediated antigen presentation by B cells [67].
Exosomes are extracellular vesicles that are released from the MVB when it fuses with the plasma membrane [68]. Endosomal sorting complexes required for transport (ESCRT) proteins [69], and Rab27a and Rab27b GTPases play an important role in their biogenesis [70]. Exosomes can contain nucleic acids, lipids, metabolites, cytosolic proteins, and cell surface molecules [71]. Exosomes from antigen presenting cells (APCs) play an important role in presenting extracellular antigen to CD4+ T cells to promote T cell immunity during infection [72]. T- and B cell-derived exosomes have been found to be involved in SLE, for example, miR-451a expression in CD4+ T cell- and B cell-derived exosomes is downregulated in SLE patients with active disease and correlates with renal damage [73].

2.2. Contribution of Endosomal Traffic to T Cell Synapse Formation

When a T cell recognizes an antigen presented by an APC, the TCR on the T cell organizes an immune synapse, involving the TCR/CD3 complex and co-receptor CD4 [74]. The TCR is a heterodimer, commonly consisting of the α and β chains, which recognizes antigens presented by APCs on their MHC-I or -II [75]. The TCR is assembled with a complex of CD3 proteins with the subunits δ, ε, γ, and ζ [76]. The co-receptor, CD4, also binds to MHC-II [77].
When the TCR engages with a peptide–MHC II complex presented by an APC, signaling molecules are recruited via endocytic traffic to the site of interaction, which is termed the immunological synapse (IS). The IS is also called the supramolecular activation complex (SMAC), which is comprised of concentric structures, designated as central, peripheral, and distal, or c-SMAC, p-SMAC, and d-SMAC, respectively [78]. The c-SMAC contains the TCR, CD3, CD4, and CD28; the p-SMAC contains adhesion molecules, lymphocyte function-associated antigen 1 (LFA-1) and intracellular adhesion molecule (ICAM) molecules, and signal transducers such as CD71 [79]. The movement of CD4 and CD71 in and out of the IS is triggered by the activation of protein kinase C (PKC) [80] and mediated via endosome traffic by Rab4A [61]. Signal transduction from the CD3ζ chain depends on the phosphorylation of tyrosine residues by the lymphocyte-specific protein tyrosine kinase (Lck), which is recycled via endocytic traffic controlled by Rab11 and Rab11 family interacting protein-3 [81]. CD3ζ has three immunoreceptor tyrosine-based activation motif (ITAM) domains. Phosphorylated ITAM domains recruit ζ-associated protein kinase 70 (ZAP-70), which becomes phosphorylated and activated by Lck [82]. Activated ZAP-70 phosphorylates the adaptor proteins, LAT, and lymphocyte cytosolic protein 2 (LCP2 or SLP-76), which then bind to and activate phospholipase cCγ (PLC-γ). Activated PLC-γ hydrolyzes phosphatidylinositol-4,5-biphosphate (PIP2), producing inositol 1,4,5-triphosphate and diacylglycerol. This results in calcium flux, and PKC becomes activated. Then, Ras guanine releasing protein 1 (RasGRP1) is recruited, and the Ras-mitogen-activated protein kinase pathway is activated [83,84] (Figure 2).
Moreover, endosome traffic delivers other essential cargoes for the formation of a functional IS on CD4+ T cells, such as the linker for the activation of T cells (LAT1), the guanine exchange factor Vav, and several actin polymerization regulatory proteins, like the Wiscott–Aldrich syndrome protein (WASp), the WASp-interacting protein, WAVE-2, and coronin-1 [85]. Following activation, CD3, CD4, and CD71 are internalized by Rab5+ endosomes to be redirected to the cell surface in Rab4+ or Rab11+ early endosomes [86]. When the TCR engages self-antigens, death receptors, such as CD95/Fas/Apo1, become activated and trigger a wave of endosomal traffic to the mitochondria, thus initiating apoptosis [87]. The retrograde endosomal traffic of LAT1 between the IS and the TGN is controlled by Rab6 and Syntaxin-16 [88]. In turn, the association with sorting nexin 27 (SNX27) promotes endosome rerouting to the IS by preventing its traffic to lysosomes [89]. Notably, SNX27+ sorting endosomes are associated with Rab4 [90]. Future studies may be aimed at the role of SNX27 in directing Rab4-mediated endosome traffic between the cell surface and lysosomes.

3. Trafficked Receptors Impact Metabolic Abnormalities in T Cells

In autoimmune diseases, such as SLE, changes in the expression of several surface proteins are linked to the functional defects in T cells [91]. Surface receptors on T cells promote signaling cascades that determine the cell’s fate, cytokine production, and differentiation. When T cells differentiate and execute effector functions, they undergo metabolic reprogramming [92]. In SLE, T cell metabolism is dysregulated in multiple ways [93]. Therefore, it is important to link how the expression levels of surface proteins that are regulated by endosomal recycling may lead to metabolic abnormalities found in SLE.

3.1. TCR and STIM1 Trafficking Affects Calcium Flux

T cell activation causes CIE of the TCR/CD3ζ subunit (TCRζ), which has decreased expression in SLE T cells [94]. Previous work in our lab has shown that the Rab4A-dependent lysosomal degradation contributes to the loss of TCRζ and CD4 in SLE T cells [10] (Figure 1).
T cells from SLE patients are marked by aberrant TCR signaling, causing hyperresponsiveness of T cells [95]. Changes in the expression of TCRζ caused by the receptor trafficking and recycling pathways have been found to be behind this phenomenon [10]. In place of the CD3ζ protein, Fc epsilon receptor I gamma chain (FcεRIγ) is substituted, which is not normally expressed in resting T cells [96]. CD3ζ and FcεRIγ are structurally homologous [97]. In SLE T cells, FcεRIγ recruits Syk instead of ZAP-70. The interaction between FcεRIγ and Syk is significantly stronger than the normal interaction between CD3ζ and ZAP-70, increasing calcium influx in these T cells [98,99,100] (Figure 2). Interestingly, increased intracellular calcium concentration in T cells can initiate exosome secretion [101]. In SLE T cells, activation accompanied by calcium influx play a role in cytokine production and inflammatory lineage development through multiple pathways.
Furthermore, TCR-induced calcium fluxing involves the trafficking and activation of calcium release-activated calcium channels (CRACs) [102]. CRAC1 is a calcium-selective ion channel protein that is encoded by the ORAI1 gene [103]. The inactivation of CRAC1 due to genetic mutations causes severe combined immunodeficiency (SCID). Stromal interaction molecule 1 (STIM1) is a transmembrane protein that is mainly localized to the endoplasmic reticulum, where it senses calcium depletion, which triggers its translocation to the plasma membrane and association with ORAI1 [104]. Of note, Rab4 and Rab5 regulate the endocytic traffic of STIM1 and thus modulate calcium flux via ORAI1 [105]. Calcium current through CRAC can be inhibited pharmacologically in CD4+ T cells by RO2959 [106]. Similar to STIM1, another calcium binding protein, CRAC regulator 2A (CRACR2A) also possesses an EF-hand, as well as Rab GTPase activity, that acts as an adaptor for dynein and transports intracellular cargoes [107]. CRACR2A is a cytosolic Ca2+ sensor that traffics STIM1 to the plasma membrane to ORAI1. Biallelic CRACR2A mutations also cause immune deficiency with autoimmune inflammatory complications [108].

3.2. Glucose Transporters and Metabolism

There are two steps to glucose metabolism: (1) the breakdown of glucose and conversion into pyruvate and (2) the TCA cycle to fuel OXPHOS in the mitochondria [109] (Figure 2). The primary metabolic pathway for Th17 cells is aerobic glycolysis, while for Tregs, it is OXPHOS through fatty acid oxidation [110,111].
Glucose transporters (GLUT) mediate glucose uptake, and for T cells, GLUT1 is the major glucose transporter [112]. IL-7 promotes glycolysis and GLUT1 trafficking [113]. In ovarian cancer cells, Rab25 regulates the recycling of GLUT1 to the cell surface [114]. Th17 cells have higher GLUT1 surface expression than Tregs [115,116]. In SLE, increased GLUT1 expression in CD4+ T cells is associated with increased activation and IL-17 production [117]. In mouse models, depleting GLUT1 reduces Th17 differentiation, while Treg differentiation remains unchanged, improving inflammatory bowel disease and GVHD [115]. Similarly, treating lupus-prone mice with CG-5, a glucose transporter inhibitor, reduced Th17 differentiation while promoting Treg differentiation, improving disease activity [118]. The glucose-dependent metabolism activates the calcium signaling pathway through generating phosphoenolpyruvate (PEP) during glycolysis [119] (Figure 2).
Glucose metabolism provides two essential metabolites for cell growth and survival via the pentose phosphate pathway (PPP): (1) ribose 5-phosphate for the synthesis of nucleotides and (2) NADPH for the synthesis of lipids and the maintenance of a reducing environment. NADPH can directly or indirectly neutralize reactive oxygen species (ROS) via regenerating reduced glutathione (GSH) from its oxidized form, glutathione disulfide (GSSG) [120]. The PPP enzyme transaldolase regulates the availability of NADPH and intracellular GSH in CD4+ T cells [121]. NADPH and GSH levels impact signal transduction via the control of the mitochondrial transmembrane potential (∆Ψm), an early checkpoint of T cell activation [122]. Intracellular GSH is depleted in lupus T cells, resulting in mitochondrial hyperpolarization (MHP) and adenosine triphosphate (ATP), which predisposes to pro-inflammatory cell death via necrosis [123,124,125]. MHP is linked to mTOR activation, which can be reversed by replenishing GSH by providing its rate-limited antioxidant precursor, N-acetylcysteine (NAC) [126]. Increased ROS production and GSH depletion has been attributed to the accumulation of mitochondria due to deficient recycling via mitophagy [127]. In turn, deficient mitophagy is caused by the Rab4A-mediated depletion of mitochondrial fission initiator dynamin-related protein 1 (Drp1) [127]. The targeting of Drp1 to lysosomal degradation is reversible by inhibiting the enzymatic activity of geranylgeranyl transferase II with 2-hydroxy-2-phosphono-3-(pyridin-3-yl) propanoic acid (3-PEHPC), which also provides preliminary evidence for therapeutic blockade of glomerulonephritis in lupus-prone MLR/lpr mice [127].

3.3. Transferrin Receptor, CD71, and Iron Metabolism

Iron plays an important role in hemopoiesis and mitochondrial function [128]. Excess iron can impair mitochondria, and it is one of the metabolic abnormalities found in SLE T cells [129]. Two iron molecules bind to transferrin, a soluble transporter protein. Once iron binds to transferrin, it binds to the transferrin receptor, CD71, on the cell surface [130]. Then, CD71 becomes internalized and fuses with the endosome. Ferrous ion leaves the endosome via an iron transporter, divalent metal transporter 1 (SLC11A2, DMT-1), into the cytoplasm [129] (Figure 2).
SLE patients’ CD4+ T cells contain higher intracellular iron levels than those of healthy controls [131,132]. Increased iron levels are correlated with epigenetic changes that promote Tfh differentiation [132]. Missense mutations of the CD71 gene, TFRC, causes immunodeficiency [133] and defective T cell proliferation [134], showing that CD71 expression and the resulting iron flux are important for T cell activation. A recent study showed that in SLE-prone mice and patient T cells, CD71 cell surface expression is increased, resulting from increased endosomal recycling to the plasma membrane. Upon T cell activation, the recycling of CD71 is increased by the fast endocytic sorting pathway controlled by Rab5 and Rab11a [135] (Figure 1). SLE patients with high CD71 expression on their Th17 cells have increased disease severity. In lupus mice, T cells that had increased CD71 expression level also had elevated intracellular iron levels [93]. Blocking CD71 inhibits Th17 differentiation by increasing IL-2 expression and reduces the recruitment of RORγt to the IL17a locus [136]. The deletion of Tfrc in mouse CD4+ T cells showed stronger fitness advantage to Tregs [93]. Interestingly, in Th cells, iron controls pathogenicity by promoting glucose metabolism [137].

3.4. NAD+ Synthesis and Metabolism

In the last few years, decreased cellular nicotinamide adenine dinucleotide (NAD+) levels have been associated with aging [138] and various diseases such as obesity [139], neurodegenerative diseases [140], cancer [141,142], and SLE [143]. NAD+ is a metabolite that plays an important role in cellular homeostasis. Cellular NAD+ levels are regulated through NAD+ metabolism, synthesis, and NAD+-dependent non-reduction–oxidation reactions. NAD+ is a cofactor in energy metabolism reduction–oxidation (redox) reactions, such as glycolysis, oxidative phosphorylation (OXPHOS), the tricarboxylic acid (TCA) cycle, and fatty acid oxidation [144,145]. NAD+ also participates in other non-redox processes, including post-translational modifications [146], mitochondrial metabolism [147], cell signaling [148], inflammatory responses [149], and apoptosis [150]. In non-redox reactions, NAD+ is a substrate to enzymes, such as sirtuins [151], poly-ADP-ribose polymerases (PARPs) [152], and the cyclic adenosine diphosphate-ribose (cADPR) family of ectoenzymes [153]. Sirtuin is a family of seven essential histone deacetylases that are involved in several cell signaling regulatory pathways, such as cell survival and longevity [138,151]. PARPs are a family of DNA repair enzymes that mediate a process where NAD+ acts as a donor of ADP-ribose moieties [154]. Since NAD+ is required for both redox and non-redox processes, cells synthesize NAD+ in different pathways.
NAD+ synthesis arises from different dietary precursors via (1) the de novo pathway from the amino acid precursor, tryptophan; (2) the Preiss–Handler pathway from nicotinic acid (NA); and (3) the nucleoside pathway from nicotinic acid riboside (NAR) or nicotinamide riboside (NR) [155]. However, the salvage pathway is preferred, where nicotinamide (NAM), the catabolic product of NAD+-consuming enzymes (sirtuins, PARPs, and NAD+ hydrolases), is recycled for NAD+ synthesis [156] (Figure 3).

3.4.1. Amino Acid Transporter, CD98

In T cells, kynurenine is transported by System L amino acid transporters, which are heterodimers with a heavy chain, SLC3A2 (CD98), and an amino acid transporting light chain, SLC7A5 (LAT1) [157] (Figure 3). In SLE, tryptophan is depleted, while its catabolite, kynurenine, is increased, which may affect the de novo pathway of NAD+ synthesis [158,159,160]. CD98 expression was increased in a lupus-prone mouse model [161], so the endosomal recycling pathway that may regulate its expression may play an important role in intracellular NAD+ synthesis. After endocytosis, CD98 is recognized and sorted on endosomes by Rab22a and Hook1, a cargo-tethering protein, routing the cargo to recycling endosomes [162] (Figure 1). As mentioned above, in CD4+ T cells, the trafficking of LAT1 between the IS and the TGN is controlled by Rab6 and Syntaxin-16 [88]. The accumulation of kynurenine has been attributed to the overgrowth of tryptophan-producing bacteria in the gut microbiota of lupus-prone mice [163,164]. Kynurenine is a most predictive metabolic biomarker in SLE, which triggers mechanistic target of rapamycin (mTOR) activation and is responsive to treatment with NAC [165]. Direct mTOR blockade with sirolimus also reverses inflammatory T cell activation in SLE [10,127,166,167,168].
In glutaminolysis, glutaminase converts glutamine to glutamate. In addition to glycolysis, glutaminolysis also plays an important role in energy production in T cells [111]. Alanine–serine–cysteine transporter 2 (ASCT2, SLC1A5) transports amino acids into the cell, including glutamine (Figure 2). The recycling of glutamine transporters is regulated by retromer, an effector of Rab7 [169], which mediates the transfer of cargo between the endosome and TGN [170]. The deletion of Slc1a5 impairs Th17 differentiation, confirming the importance of glutaminolysis in Th17 cells [171].

3.4.2. NAD+ Hydrolases, CD38, CD157, and SARM1

Genes encoding CD38 and its homologue, CD157, are both located on the human chromosome 4 [172,173]. CD38 and CD157 are both receptors and ectoenzymes that have ADP-ribosyl cyclase activity and can convert NAD+ into cADPR and hydrolyze cADPR to form ADP-ribose (ADPR), although CD157 is one hundred times less efficient than CD38 (Figure 3) [174,175,176,177]. CD38 was first observed on T cells as an activation marker [178], and now, it is considered ubiquitous in the immune system, with variable levels of expression [176]. CD157 is mainly expressed by myeloid cells, especially neutrophils and monocytes [176,179], and its expression is increased in rheumatoid arthritis patients [180].
CD38 expression levels on T cells have been studied in numerous diseases. CD38bright CD8+ T cells can predict acute graft-versus-host disease [181], phenotypic changes in CD38+ CD4+ T cells can predict the severity of inflammatory bowel disease [182], and CD38 expression in CD4+, CD8+, or CD25+ T cells is significantly higher in SLE patients than in healthy controls [183]. CD38 has a short cytoplasmic tail, and it controls both intra- [177] and extra-cellular [184] NAD+ levels. It takes 100 NAD+ molecules to generate one cADPR molecule, which indicates that CD38 is a significant NAD+ consumer [185,186]. In human CD8+ T cells and the Jurkat CD4+ T cell line, CD38 decreases NAD+ levels [143]. Since the level of CD38 expression on T cells impacts NAD+ levels, which is used in a variety of redox and non-redox pathways, determining their functions [187], it is important to determine how its expression is controlled by endosomal recycling. To date, what we know thus far is that in human lymphocytes and T cell leukemia lines, the downregulation of CD38 expression by endocytosis is not a key step in its intracellular signaling; rather, it is a negative feedback mechanism [188]. Upon TCR engagement, CD38 is actively recruited at the immune synapse from the plasma membrane and recycling endosomes [189].

3.4.3. CD73

CD73 is an ecto-5′-nucleotidase that is upregulated in cancer cells [190,191,192], while in SLE patients, it is selectively silenced in B cells and its expression decreased in T cells [193,194,195]. CD73 is suggested to have dual enzymatic functions: (1) cleaving NAD+ to NMN and adenosine monophosphate (AMP) and (2) hydrolyzing NMN to NR [196,197,198] (Figure 3). There is an ongoing debate regarding the role of CD73 in contributing to intracellular levels of NAD+ in human primary cells [199,200]. Interestingly, CD73+ Tregs inhibit CD4+ CD25 T cell proliferation via the Treg-derived exosomes from late endosomes [201,202] (Figure 1). The expression of CD73 in these CD73+ Treg-derived exosomes [202] suggests the need for further study to determine its endosomal pathway.
In the small intestine, SLC12A8, an NMN transporter, is highly expressed, which suggests that NMN may be an entry point into the nucleoside NAD+ synthesis pathway [203]. It is also overexpressed on bladder cancer cells and associated with tumor immune cell infiltration [204]. However, its expression and mechanisms of NMN uptake in immune cells remain unclear.

3.5. Role of Endosome Traffic in Toll-like Receptor-Mediated Signaling

Toll-like receptors (TLRs) are classified by the location of their expression: intracellular or extracellular. TLR2 and TLR4 are expressed on the surface of CD4+ T cells [205], and when they bind to their ligand, i.e., bacterial membrane components [206,207], they are endocytosed to the endosomes, a step required for the activation of nuclear factor-κB (NF-κB) and AP-1 [208,209]. In endosomes, Rab11a promotes TLR4 signaling [210], and Rab10 traffics TLR4 back to the cell surface [211]. Recently, it has been shown that Rab8 and Rab11 recruits adaptor protein-3 (AP-3) to TLR2, promoting the secretion of IL-6 by phagocytic cells [212]. TLR2 expression level is increased in SLE patients’ CD4+ T cells, CD8+ T cells, and B cells [213]. TLR2 activation in human Tregs reduces their suppressive functions, and it promotes Th17 differentiation in naïve CD4+ T cells [214]. On the other hand, intracellular TLR7 and TLR8 are confined to endosomes, where they sense viral single-stranded RNAs [215]. Genetic variants of TLR7 and TLR8 are risk factors for SLE [216,217], and their expression is upregulated in SLE patients [218]. TLR7 activation in CD4+ T cells induces calcium flux [219] and TLR8 signaling in human Tregs inhibits glucose uptake and glycolysis [220].

4. Proinflammatory Signaling Pathways Impacted by Trafficked Receptors in CD4+ T Cells

SLE results from multiple predisposing genetic traits and environmental stimuli [21]. Environment stimuli can induce alterations to membrane trafficking to increase or decrease surface protein expression [26]. The changes in surface receptor expression of a cell affects how the cell detects stimuli, such as cytokines from other immune cells, and the downstream effects of how it responds by producing other cytokines. A defect in the cell’s response, such as an uncontrolled cytokine production, can lead to immune activation and tissue damage in SLE, and aberrant signaling pathways contribute to this phenomenon [221]. Many studies have revealed that metabolism is important in T cell differentiation and function [111]. We will discuss how the above discussed changes in metabolites resulting from surface receptor recycling may lead to proinflammatory and anti-inflammatory imbalance in SLE (Table 1).

4.1. IL-2 and Tregs Are Decreased

Cytokine abnormalities are found in the pathogenesis of SLE [21]. IL-2 is produced by activated Th cells and plays an important role in expansion and homeostasis [21]. IL-2 production is decreased in the T cells of SLE patients and contributes to the Th17/Treg imbalance in SLE [222]. Naïve CD4+ T cells can be skewed to become Tregs when their TCR is stimulated in the presence of IL-2 and TGFβ [223]. The expansion of Tregs is greatly dependent on IL-2, which suggests that IL-2 depletion contributes to the reduced number and function of Tregs in SLE patients [224]. Tregs control the expansion of autoreactive T cells, and thus are important in inhibiting autoimmunity [13]. Even though it is well established that SLE patients have defective IL-2 production, the underlying mechanism is poorly understood.

4.1.1. CREM and CREB Control IL-2 Production

Several transcription factors control IL2 transcription, but the most well described in SLE patients is the imbalance between cyclic AMP responsive element-binding protein (CREB) and cyclic AMP element modulator (CREM). CREB is a positive regulator and CREM is a negative regulator, and they compete to bind to the IL2 promoter. In resting T cells, CREB is bound, and when the T cell is activated, CREB is phosphorylated, leading to IL2 transcription (Figure 2). However, when phosphorylated, CREB is replaced with phosphorylated CREM, and IL2 transcription is repressed [225].
In SLE patients’ T cells, CREB is decreased, while CREM is abnormally increased [225]. Furthermore, there is evidence that calcium/calmodulin-dependent protein kinase IV (CaMKIV), an enzyme that phosphorylates CREM, increases the binding of the repressor CREM to the IL2 promoter, suppressing IL2 transcription [13]. Circulating autoantibodies and autoantigens common in SLE patients, such as anti-TCR antibody, can also activate CaMKIV [226]. Moreover, SLE T cells demonstrate hypomethylation of protein phosphatase 2 (PP2A) promoter, increasing the levels of serine/threonine phosphatase, PP2A [227], which binds to CaMKIV and keeps it catalytically inactive in the cytoplasm [228]. Increased calcium flux in SLE T cells promotes the accumulation of Ca2+/calmodulin (CaM), which replaces PP2A from CaMKIV, activating it. This inhibition of CaMKIV/PP2A results in increased CaMKIV-mediated gene transcription [228]. PP2A is responsible for the dephosphorylation of CREB [229], and it also dephosphorylates and activates SP-1, which then promotes CREM transcription [230], ultimately decreasing IL-2 production (Figure 2). CREM transcription is increased by the abnormally increased amounts of activated transcription factor SP-1 shown in SLE [230].

4.1.2. NFAT and AP-1 Control IL-2 Production

In SLE T cells, there is an increase in calcium influx [231]. The increase in cytosolic calcium concentration leads to increased calcineurin activation. Activated calcineurin dephosphorylates the nuclear factor of activated T cells (NFAT) in the cytoplasm, which is abnormally high in SLE [232]. The dephosphorylated NFAT then translocates to the nucleus, where it would normally bind to the promoters of IL2 genes [233]. Calcium signaling and NFAT activity are reinforced by aerobic glycolysis-derived PEP [119] (Figure 2). However, in SLE, NFAT does not promote IL-2 production because the IL2 promoter requires the binding of a transcription factor AP-1 to adjacent sites [232].
The AP-1 transcription factor family is formed by heterodimers and homodimers of Fos and Jun proteins [234]. When TCR binds to an antigen, the Fos and Jun proteins are expressed, and AP-1 binds to the IL2 promoter (Figure 2). However, c-Fos expression is decreased in SLE T cells, which reduces AP-1 binding to the IL2 promoter [235]. c-Fos contains cyclic AMP responsive element (CRE) sites in its promoter and CREM downregulates c-Fos activity [236]. Since CREM is increased in SLE T cells, CREM binds to the c-Fos promoter and decreases c-Fos production [237]. Therefore, even though NFAT expression is increased in SLE, due to the reduced AP-1 activity, IL-2 production is still reduced [232].
The sirtuins remove acetyl groups from transcription factors and histones, inhibiting gene transcription [238]. In CD4+ T cells, sirtuin-2 deacetylates c-Jun and histones at the IL2 gene, decreasing IL-2 production [239] (Figure 3). In experimental autoimmune encephalomyelitis (EAE) mice and lupus-prone mice, a sirtuin-2 inhibitor, AK-7, ameliorated disease severity [239]. Since sirtuins are NAD+-dependent histone deacetylases, this suggests that NAD+-modulating cell surface molecules may play an important role.
In SLE T cells, blocking CD71 with an antibody normalizes T cell activation and IL-2 production [93,240]. However, in a Jurkat CD4+ T cell line, blocking CD71 with an antibody did not change the activation profile of NFAT, AP-1, and nuclear factor kappa-light-enhancer of activated B cells (NF-κB) [240]. Further study is needed to investigate the pathways through which CD71 affects IL-2 production.
When TCR is stimulated, CRACR2A transmits the signal to activate the Ca2+-NFAT and JNK-AP1 pathways [108]. As mentioned above, biallelic CRACR2A mutations cause autoimmune inflammatory complications, and STIM1, which CRACR2A traffics, is significantly increased in lupus mice kidneys and contributes to renal damage [108,241]. In a Jurkat CD4+ T cell line, Loureirin B, a constituent from a traditional Chinese medicine, modulates IL-2 secretion by inhibiting STIM1/ORAI1 channels and decreasing calcium influx [242]. In SLE patients’ T cells, forcing CD3ζ expression returns calcium fluxing and IL-2 production to normal, which indicates that these phenomena are downstream effects of altered calcium signaling [243]. This decrease in IL-2 production has effects on other cytokine and lineage development, such as hindering IL17a expression and Th17 differentiation. This may explain the increased IL-17 production in SLE [244,245].

4.2. IL-17 and Th17 Are Increased

SLE patients have increased serum levels of IL-17 [246] and an increased frequency of IL-17-producing cells in the peripheral blood [247,248,249]. IL-17 is pro-inflammatory because it induces IL-6, granulocyte monocyte-colony stimulating factor (GM-CSF), and granulocyte colony stimulating factor (G-CSF) production [250,251,252,253,254]. This recruits monocytes and neutrophils, leading to inflammation and tissue damage. Furthermore, in the presence of the B cell activating factor, IL-17 increases B cell activation, proliferation, and differentiation into immunoglobulin-secreting cells [255]. Th17 cells play an important role in the pathogenesis of SLE by amplifying inflammation [256].
IL-17 is produced rapidly and in large amounts by γδT cells, DN TCRαβ T cells, and Th17 cells [247,257,258,259]. DN T cells are increased and represent a major source of IL-17 in SLE patients [247]. TCRαβ DN T cells derive from CD8 T cells [260], and their effector functions play a key role in tissue damage seen in SLE. They also produce other pro-inflammatory mediators, namely interferon (IFN)-γ, IL-1β, CXCL2, and CXCL3, and infiltrate the kidneys, the site of lupus nephritis, one of the leading causes of death in SLE [247,260,261]. On the other hand, Th17 cells derive from naïve CD4+ T cells primed in the presence of TGFβ, IL-6, IL-21, and IL-1β [20,262,263]. IL-23, which is produced by antigen-presenting cells, can also induce the expansion of Th17 cells [19].

4.2.1. JAK/STAT3 Pathway Regulates Th17 Differentiation

Calcium influx is increased in SLE T cells due to the loss of TCRζ [100]. The increase in cytosolic calcium concentration leads to increased calcineurin activation. Activated calcineurin dephosphorylates inactive NFAT, which is expressed at abnormally high levels in SLE T cells [232]. The dephosphorylated NFAT then translocates to the nucleus, where it binds and activates the promoters of CD40 ligand (CD40L) [233]. CD40-CD40L signaling plays an important role in the differentiation of Th17 cells [264].
CD40 is expressed by APCs and CD40L is on T cells. Various methylation-sensitive genes that functionally contribute to SLE pathogenesis are overexpressed in SLE patients’ CD4+ T cells, including CD11A, CD70, and CD40L [265,266]. Upon exposure to high antigen doses, T cells upregulate CD40L, providing efficient co-stimulation to APCs to produce IL-6 [264], which is elevated in SLE [18]. IL-6 is produced in many cell types, like monocytes and B cells [267]. IL-6 signaling induces Th17-related gene expression via the signal transducer and activator of transcription (STAT)3 [268].
Most cells respond to IL-6 trans-signaling, which is achieved through IL-6 binding to soluble IL-6 receptor (sIL-6R), rather than the classic signaling via IL-6R on the cell’s surface [269]. In human peripheral blood mononuclear cells, the cleavage of IL-6R to generate sIL-6R is induced by the activation of TLR2 [270]. When IL-6 binds to IL-6R, the IL-6R subunit-β (gp130; IL-6Rβ) initiates intracellular signaling and activates gp130-associated Janus kinases (JAKs) [271]. The activated JAKs phosphorylate gp130, initiating the JAK/STAT3 pathway [271]. The SRC homology domain 2 (SH2) domain of STAT3 binds to the gp130 phosphotyrosine docking sites. When STAT3 is within proximity of active JAKs, STAT3 is phosphorylated at Tyr705, resulting in the dimerization of the STAT3 protein and nuclear translocation [272]. STAT3 binds to the promoters of target Th17-related genes, such as IL17a, IL17f, and IL23r [273] (Figure 2). IL-23 is critical for Th17 maintenance [274]. Therefore, increased calcium influx due to Rab4A-mediated lysosomal degradation of TCRζ in SLE T cells leads to the activation of NFAT, which upregulates CD40L expression. CD40L stimulation leads APCs to produce IL-6 [264], which leads to the activation of JAK/STAT3 pathway in the T cell [271], causing an increase in IL-17 production and Th17 development [264].

4.2.2. mTOR Regulates IL-17 Production and Th17 Differentiation

CaMKIV also plays an important role in IL-17 production and Th17 development using two pathways: (1) increasing the binding of CREMα to IL17 genes and (2) via the AKT/mTOR pathway. CaMKIV activity is the highest under Th17 polarizing conditions. In mice, when CaMKIV is inhibited, Th17 and Treg differentiation is affected but not Th1 or Th2 [275].
CREM plays an important role in IL-2 production, but it also plays a role in the Th17 differentiation and IL-17 production through epigenetic remodeling [276]. The suppressor isoform, CREMα, is increased in SLE T cells and controls IL-17A expression by reducing DNA methylation of the IL17A locus [277]. DNA methylation is when methyl groups are added to DNA molecules, and in eukaryotes, cytosine is methylated at the 5-carbon [278]. DNA methylation occurs at CpG sites, the regions of DNA where cytosine is adjacent to a guanine nucleotide reading in the 5′ to 3′ direction. The addition of methyl groups represses gene expression [279]. Therefore, the reduction in CpG-DNA methylation by CREMα lifts the repression of gene expression, leading to increased IL17 transcription [277].
Furthermore, CaMKIV also promotes Th17 differentiation and IL-17 production via the PI3K/AKT/mTOR pathway [275] (Figure 2). The PI3K/AKT/mTOR pathway plays an important role in proliferation, growth, and survival [280]. mTOR is a component of two complexes, mTOR complex 1 (mTORC1) and mTOR complex 2 (mTORC2). A key function of mTORC1 is its ability to phosphorylate key regulators of mRNA translation [281], while a key function of mTORC2 is the phosphorylation and activation of serine/threonine protein kinase B (PKB; AKT) through phosphorylation at Ser473 [282]. A key effector downstream of AKT is mTORC1 [281]. The activation of mTORC1 enhances Th17 differentiation [283] and the disruption of mTORC1 impairs Th17 differentiation [284]. The TCR signal triggers mTORC1 activation by inducing System L amino acid transporters to uptake leucine [285]. Rapamycin inhibits mTORC1 but not mTORC2. Rapamycin has shown to normalize TCR-induced calcium influx [286] and Th17/Treg balance [287] in SLE patients.
In SLE, FcεRIγ replaces CD3ζ [96], resulting in the recruitment of Syk instead of ZAP-70 [59]. Syk has been shown to activate the PI3K/AKT/mTOR pathway. Active Syk increases the catalytic activity of PI3K, resulting in the conversion of PIP2 to phosphatidylinositol-3,4,5-triphosphate (PIP3) [288]. This then recruits phosphatidylinositol-dependent kinase 1 (PDK1) and AKT to the cell membrane [289]. Then, PDK1 is phosphorylated at Thr308 in the AKT kinase catalytic region [290], and mTORC2 phosphorylates AKT [276]. Activated AKT then phosphorylates tumor suppressor TSC2, leading to mTORC1 activation [291] (Figure 2). In SLE T cells, mTORC1 activity is increased, while mTORC2 is reduced [292].
mTORC1 phosphorylates the ribosomal S6 kinases (S6K) [293]. There are two homologs of S6K: S6K1 and S6K2. More research has been conducted on S6K1 than S6K2 due to the belief that their high degree of homology leads them to behave similarly [294] (Figure 2). However, S6K1 and S6K2 lead to different pathways that both positively regulate Th17 differentiation [283].
The PI3K/AKT/mTORC1/S6K1 axis promotes Th17 differentiation via growth factor-independent 1 transcriptional repressor (Gfi1) [283]. Gfi1 is a transcriptional repressor that recruits histone-modifying enzymes to the target gene promoters [295]. The downregulation of Gfi1 is a critical event for Th17 differentiation [296]. The PI3K/AKT/mTORC1/S6K1 axis induces early growth response 2 (EGR2), a transcription regulatory factor that contains zinc finger DNA-binding sites, and its increased gene expression is a risk factor for SLE [283,297]. EGR2 downregulates the expression of Gfi1 and increases Th17 differentiation [283] (Figure 2).
On the other hand, the PI3K/AKT/mTORC1/S6K2 axis promotes the nuclear translocation of RORγt [283], a critical transcription factor for the initiation of Th17 differentiation [298]. S6K2 interacts directly with RORγt and enhances its nuclear translocation [283]. The PI3K/AKT/mTORC1 pathway enhances the expression of S6K2, accelerating RORγt nuclear translocation during Th17 differentiation [283] (Figure 2).
Pyruvate kinase muscle isozyme 2 (PKM2) acts at the last step of glycolysis, and it is required for Th17 differentiation [299]. It translocates to the nucleus and activates with STAT3, increasing Th17 differentiation [300]. PKM2 also binds to CaMKIV, and CaMKIV enhances pyruvate kinase activity and glycolysis through the glucose transporter [301] (Figure 2).
CREM also enhances RORγt. Inducible cyclic AMP early repressor (ICER) is a transcriptional repressor isoform of CREM, and it enhances the accumulation of RORγt and binds to the IL17a promoter, leading to IL-17 production [302]. ICER also induces Th17 differentiation via the sirtuin-2/mTORC1/HIF-1α pathway [239]. Following TCR signaling and mTORC1, hypoxia-inducible factor 1-alpha (HIF-1α) is induced, which is normally overexpressed in Th17 cells [303]. ICER binds to the Sirt2 promoter, sirtuin-2 deacetylates p70S6K, activating the mTORC1/HIF-1α/RORγt pathway and inducing Th17 differentiation [239] (Figure 3). HIF-1α also upregulates GLUT1, promoting increased glucose uptake, and upregulates PDK1, reinforcing glycolysis [304] (Figure 2).
The STAT3 phosphorylation and activation is not limited to JAKs [271]. It was generally believed that JAK/STAT3 signaling pathway contributed to STAT3 activation through tyrosine and serine phosphorylation. However, studies have suggested that STAT3 can be activated through Ser727 phosphorylation in the absence of Tyr705 phosphorylation, the site of phosphorylation by JAKs [305]. Tyr705 phosphorylation by JAKs is involved in STAT3 dimerization and activation [306], while Ser727 phosphorylation modulates STAT3 activity [307,308]. mTOR phosphorylates STAT3 on Ser727 [309].
As discussed above, in SLE T cells, mTORC1 activity is increased while mTORC2 is reduced [292]. Rapamycin has shown to reduce STAT3 activation and the number of IL-17 producing cells in SLE patients [310]. Therefore, the increase in calcium flux in SLE T cells due to the lysosomal degradation of CD3ζ can lead to an increase in Th17 development through two pathways: S6K2/RORγt pathway and STAT3 pathway (Figure 2).

4.2.3. Histone Modification Regulates Th17 Differentiation and IL-17 Production

In mice, the loss of sirtuin-1 functions results in increased T cell activation and a lupus-like phenotype [311]. One of the targets for sirtuin-1 is STAT3. In addition to phosphorylation, acetylation on Lys685 regulates the transcriptional activity of STAT3, and this can be inhibited by sirtuin-1 [312,313]. Sirtuin-1 deacetylates STAT3, reducing its ability to translocate to the nucleus [314]. In cancer cell lines, decreased NAD+ activates STAT3 [315]. As mentioned above, sirtuin-2 deacetylates p70S6K, inducing Th17 differentiation via the mTORC1/HIF-1α/RORγt pathway [239] (Figure 3). Therefore, NAD+ levels may have an important role in STAT3 activity by targeting Th17-related genes, suggesting the importance of surface molecules that modulate intracellular NAD+ levels.
TLR2 activation in CD4+ T cells leads to pro-inflammatory skewing [214]. In vitro stimulation of TLR2 on SLE patient CD4+ T cells with a synthetic bacterial lipopeptide, Pam3-Cys-Ser-Lys4 (Pam3CSK4), significantly induces IL-17A and IL-17F production by upregulating H3K4 tri-methylation levels in the IL-17A promoter region and H4 acetylation levels in both IL-17A and IL-17F promoter regions [213].

4.3. Regulation of Tfh Development

DNA demethylation plays an important role in CD4+ T cell differentiation [316,317]. The methylome of SLE patients’ CD4+ T cells differ from those of healthy controls [318]. PP2A inhibits DNA methyltransferase 1 (DNMT1) expression [319]. Under the presence of Fe2+, the ten-eleven translocation (TET) enzymes oxidize 5-methylcytosine to 5-hydroxymethylcytosine in DNA, hypomethylating DNA and controlling gene transcription [320] (Figure 2). Therefore, genes involved in SLE pathogenesis may undergo hypomethylation, becoming upregulated [111].
SLE is about nine times more common in women than men in the United States [321]. However, SLE tends to be more severe in men with a higher prevalence of renal disease [322]. To achieve an SLE disease flare with a severity equal to women, men require a higher genetic risk and a greater degree of CD4+ T cell demethylation [323]. Compared to female SLE patients, male patients’ CD4+ T cells show significant hypomethylation of ELAVL1, UHRF1, and SMAD2, increasing their gene expressions [324]. Embryonic lethal vision-like protein 1 (ELAVL1) is an RNA-binding protein (RBP) that interacts with immune-specific transcripts, influencing the intensity of the immune response [325,326,327]. ELAVL1 stabilizes the mRNA of Ubiquitin-like with PHD and RING finger domains 1 (UHRF1), augmenting its effects [324]. UHRF1 regulates DNA methylation during the S phase of the cell cycle [328]; therefore, increased UHRF1 expression could lead to increased CD4+ T cell proliferation in male SLE patients [324]. UHRF1 also promotes Treg proliferation by hypermethylating p21 [329] and reduces follicular helper T cell (Tfh) differentiation [330]. Thus, UHRF1 may play a mixed role in SLE pathogenesis in male SLE patients [324].
Pyruvate, the product of glycolysis, can be broken down into acetyl-CoA by pyruvate dehydrogenase (PDH) (Figure 2). Histone acetylases transfer the acetyl group from acetyl-CoA to histones [331]. SLE patients’ CD4+ T cells have hypoacetylated histones [265,332], and treating lupus-prone mice with histone deacetylase inhibitors alleviated disease severity by downregulating proinflammatory cytokines [333].
In vitro, IL-6 and IL-21, which act through STAT3, drive Tfh differentiation [334]; however, in vivo, IL-6 and IL-21 are not absolutely required for Tfh development [335], suggesting the role of other cytokines in Tfh development. Type I IFN activates STAT1 to bind to B cell lymphoma 6 (Bcl6), a transcription factor that is required for Tfh development [336], and in human and SLE mouse model, type I IFN signaling activates STAT4 to produce IL-21 and IFNγ [337]. In HeLa cells, Rab7 is required for early endosomal sorting of the subunit 1 of the type I IFN receptor (IFNAR1) [338], and recently, anifrolumab, a monoclonal antibody targeting IFNAR1, has been FDA-approved for SLE [339]. Therefore, endosome traffic of IFNAR1 in CD4+ T cells may modulate the clinical efficacy of this new therapeutic intervention in SLE.
Table 1. Receptors that are recycled by Rab GTPase, downstream metabolites, and subsequent signaling/epigenetic regulation on Treg/Th17 differentiation. Expression level in SLE T cells is indicated in parentheses, if known. ↑ = increased in SLE; ↓ = decreased in SLE; ? = Unknown.
Table 1. Receptors that are recycled by Rab GTPase, downstream metabolites, and subsequent signaling/epigenetic regulation on Treg/Th17 differentiation. Expression level in SLE T cells is indicated in parentheses, if known. ↑ = increased in SLE; ↓ = decreased in SLE; ? = Unknown.
Recycled ReceptorsResponsible Rab GTPaseDownstream MetabolitesSubsequent Signaling/Epigenetic PathwaysEffects on T Cell Subset/CytokineReferences
CD3ζ (↓)Rab4A (↑)Ca2+ (↑)Ca2+/CaM/PP2A/de-pCREBDecrease IL-2 (↓)[10,229]
Ca2+/CaM/PP2A/SP-1/CREM (↑)[10,230,231]
Calcineurin/NFAT (↑) (no AP-1)[174,175,176,177,232]
CaMKIV/CREMαIncrease IL-17 (↑)
and Th17 (↑)
[10,277]
CaMKIV/PI3K/AKT/mTORC1 (S6K)[10,275]
Syk/PI3K/PIP3/PDK1/AKT/pTSC2/mTORC1 (S6K)[10,59,283,288,289,291]
CD38(↑)/
CD157
?CREM/ICER/RORγt[174,175,176,177,302]
ICER/Sirtuin-2/mTORC1/HIF-1α/RORγt[174,175,176,177,303]
CREM/ICER/Sirtuin-2/p70S6K/HIF-1α/RORγt[174,175,176,177,239,255]
NAD+ (↓)Sirtuin-1/STAT3Increase Th17[174,175,176,177,312,313]
Sirtuin-2/de-Ac c-JunDecrease IL-2[157,162,239]
CD98 (↑)Rab22aKynurenineNAD+ synthesis-[158,159,160]
CD4 (↓)Rab4A (↑)???[61]
CD71 (TfR) (↑)Rab4A, Rab5, Rab11a Fe2+TET/DNA hypomethylationTfh[61,93,135,320]
GLUT1 (↑)Rab25GlucosePEP/Ca2+/NFAT (no AP-1)Decrease IL-2 (↓)[117,119,232]
PMK2/STAT3Increase Th17 (↑)[299,300]

5. Conclusions

The factors that mediate SLE pathogenesis and organ damage span the whole immune system. Therefore, it is not surprising that the cytokine networks that orchestrate immune cell functions are dysregulated. The dysregulation of IL-2 is fundamental for SLE pathogenesis, as it reduces Treg differentiation, disturbing immunologic homeostasis [13]. The IL-17 signature plays an important role in tissue damage mediated by effector cells, as it recruits other pro-inflammatory cells and increases B cell activation and antibody production [255].
mTOR is a sensor of the mitochondrial transmembrane potential, which is increased in SLE T cells [10]. The activation of mTOR leads to the overexpression of the Rab5A and Rab4A small GTPases, which regulate the endocytic recycling of surface receptors, as described above [10]. Rab4A-dependent lysosomal degradation contributes to the loss of TCRζ and CD4 in SLE T cells [10], causing changes in the calcium signaling pathway compared to normal T cells. The dysregulated TCR signaling in SLE leads to a decrease in IL-2 production and increases in IL-6 and IL-17 production, causing a decrease in Treg differentiation and an increase in Th17 differentiation. GLUT1 also plays a role in activating the calcium signaling pathway through generating PEP during glycolysis [119].
Upon TCR engagement, CD38 is recycled to the immune synapse from the recycling endosomes [189]. CD38 and its homologue, CD157, convert NAD+ into cADPR, controlling both intra- [177] and extra-cellular [184] NAD+ levels. The intracellular NAD+ levels control the activity of sirtuins, histone deacetylases that inhibit gene transcription by removing acetyl groups from histones and transcription factors [238]. Sirtuin-1 deacetylates STAT3, reducing its nuclear translocation [314]. Sirtuin-2 deacetylates c-Jun and decreases IL-2 production while activating the mTORC1/HIF-1α/RORγt pathway and inducing Th17 differentiation [239]. mTOR regulates iron homeostasis by modulating CD71 stability [340]. The recycling of CD71 is increased upon T cell activation [135], and on SLE patients’ Th17 cells, elevated CD71 expression has been correlated with disease severity [93]. Furthermore, TET requires Fe2+ to hypomethylate DNA and control gene transcription, affecting Tfh cell differentiation [320].
This correlation between receptor recycling, mTOR activation, metabolic pathways, and T cell reactivity warrants further study on how Rab GTPases ultimately contribute to the pathogenesis of SLE.

Funding

This work was supported in part by grants AI072648, AI122176, and AR076092 from the National Institutes of Health, the Phillips Lupus and Autoimmunity Center of Excellence, and the Central New York Community Foundation.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

3-PEHPC2-hydroxy-2-phosphono-3-(pyridin-3-yl) propanoic acid
AKTserine/threonine protein kinase B
AMPadenosine monophosphate
AP-1activator protein 1
AP-3adaptor protein 3
APCantigen presenting cell
Bcl6B cell lymphoma 6
CaMcalmodulin
CaMKIVcalcium/calmodulin-dependent protein kinase IV
CCR6C-C-motif chemokine receptor 6
CDcluster of differentiation
CD40LCD40 ligand
cGVHDchronic graft versus host disease
CIEclathrin-independent endocytosis
CMEclathrin-mediated endocytosis
CpGcytosine and guanine separated by a phosphate
CREcyclic AMP responsive element
CREBcyclic AMP responsive element-binding protein
CREMcyclic AMP element modulator
CXCLchemokine (C-X-C motif) ligand
DN Tdouble negative T cell
Drp1dynamin-related protein 1
EAEexperimental autoimmune encephalomyelitis
EEA1early endosomal antigen 1
EGR2early growth response 2
ELAVL1embryonic lethal vision like protein 1
ERCendosomal recycling compartment
FcεRIγFc epsilon receptor I gamma chain
FoxP3forkhead box P3
GAPGTPase-activating protein
G-CSFgranulocyte colony stimulating factor
GDPguanosine diphosphate
GEFguanine nucleotide exchange factor
Gfi1growth factor independent 1 transcriptional repressor
GM-CSFgranulocyte monocyte-colony stimulating factor
GTPguanosine triphosphate
GTPaseguanosine triphosphatase
HIF-1αhypoxia-inducible factor 1α
HRES-1HTLV-1 related endogenous retroviral sequence 1
ICERinducible cyclic AMP early repressor
IFNinterferon
IFNAR1Type I IFN receptor
ILinterleukin
IL-6Rinterleukin 6 receptor
ISimmunological synapse
ITAMimmunoreceptor tyrosine-based activation motif
JAKJanus kinase
LATlinker for activation of T cells
Lcklymphocyte-specific proteins tyrosine kinase
MHCmajor histocompatibility complex
miRNAmicroRNA
mRNAmessenger RNA
MRL-lprMurphy Roths Large-lymphoproliferation
mTORmechanistic target of rapamycin
mTORCmechanistic target of rapamycin complex
NFATnuclear factor of activated T cells
NF-κB nuclear factor-κB
OXPHOSoxidative phosphorylation
Pam3CSK4Pam3-Cys-Ser-Lys4
PDK1phosphatidylinositol-dependent kinase 1
PI3Kphosphoinositol-3 kinase
PI3Pphosphoinositol-3-phosphate
PIP2phosphatidylinositol-4,5-biphosphate
PIP3phosphatidylinositol-3,4,5-triphosphate
PKBserine/threonine protein kinase B
PKCprotein kinase C
PLC-γphospholipase Cγ
PP2Aprotein phosphatase 2
RORγtretinoic-acid-receptor-related orphan nuclear receptor γ
S6Kribosomal S6 kinase
SH2SRC homology domain 2
sIL-6Rsoluble IL-6 receptor
SLEsystemic lupus erythematosus
SLP-76SH2-domain-containing leukocyte protein of 76 kDa
SMACsupramolecular activation complex
SNPssingle nucleotide polymorphisms
SNX27sorting nexin 27
SP-1specificity protein 1
STATsignal transducer and activator of transcription
Sykspleen tyrosine kinase
TCRT cell receptor
Tfhfollicular helper T cell
TGFβtransforming growth factor β
TGNtrans-Golgi network
Th17helper T 17 cell
TLRtoll-like receptor
Tregregulatory T cell
TSC2tuberous sclerosis complex 2
WASpWiscott–Aldrich syndrome protein
ZAP-70ζ-associated protein kinase 70

References

  1. Zerial, M.; McBride, H. Rab proteins as membrane organizers. Nat. Rev. Mol. Cell Biol. 2001, 2, 107–117. [Google Scholar] [CrossRef] [PubMed]
  2. Wheeler, D.B.; Zoncu, R.; Root, D.E.; Sabatini, D.M.; Sawyers, C.L. Identification of an oncogenic RAB protein. Science 2015, 350, 211–217. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Tzeng, H.-T.; Wang, Y.-C. Rab-mediated vesicle trafficking in cancer. J. Biomed. Sci. 2016, 23, 70. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Chua, C.E.L.; Tang, B.L. The role of the small GTPase Rab31 in cancer. J. Cell. Mol. Med. 2015, 19, 1–10. [Google Scholar] [CrossRef]
  5. Wang, S.; Hu, C.; Wu, F.; He, S. Rab25 GTPase: Functional roles in cancer. Oncotarget 2017, 8, 64591–64599. [Google Scholar] [CrossRef] [Green Version]
  6. Veleri, S.; Punnakkal, P.; Dunbar, G.L.; Maiti, P. Molecular Insights into the Roles of Rab Proteins in Intracellular Dynamics and Neurodegenerative Diseases. Neuromol. Med. 2018, 20, 18–36. [Google Scholar] [CrossRef]
  7. Kiral, F.R.; Kohrs, F.E.; Jin, E.J.; Hiesinger, P.R. Rab GTPases and Membrane Trafficking in Neurodegeneration. Curr. Biol. 2018, 28, R471–R486. [Google Scholar] [CrossRef] [Green Version]
  8. Tang, B.L. Rabs, Membrane Dynamics, and Parkinson’s Disease. J. Cell. Physiol. 2017, 232, 1626–1633. [Google Scholar] [CrossRef]
  9. Prashar, A.; Schnettger, L.; Bernard, E.M.; Gutierrez, M.G. Rab GTPases in Immunity and Inflammation. Front. Cell Infect. Microbiol. 2017, 7, 435. [Google Scholar] [CrossRef] [Green Version]
  10. Fernandez, D.R.; Telarico, T.; Bonilla, E.; Li, Q.; Banerjee, S.; Middleton, F.A.; Phillips, P.E.; Crow, M.K.; Oess, S.; Muller-Esterl, W.; et al. Activation of mTOR controls the loss of TCRζ in lupus T cells through HRES-1/Rab4-regulated lysosomal degradation. J. Immunol. 2009, 182, 2063–2073. [Google Scholar] [CrossRef] [Green Version]
  11. Evnouchidou, I.; Caillens, V.; Koumantou, D.; Saveanu, L. The role of endocytic trafficking in antigen T cell receptor activation. Biomed. J. 2022, 45, 310–320. [Google Scholar] [CrossRef] [PubMed]
  12. Perl, A. Oxidative stress and endosome recycling are complementary mechanisms reorganizing the T-cell receptor signaling complex in SLE. Clin. Immunol. 2012, 142, 219. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Lieberman, L.A.; Tsokos, G.C. The IL-2 Defect in Systemic Lupus Erythematosus Disease Has an Expansive Effect on Host Immunity. J. Biomed. Biotechnol. 2010, 2010, e740619. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Bonelli, M.; Savitskaya, A.; von Dalwigk, K.; Steiner, C.W.; Aletaha, D.; Smolen, J.S.; Scheinecker, C. Quantitative and qualitative deficiencies of regulatory T cells in patients with systemic lupus erythematosus (SLE). Int. Immunol. 2008, 20, 861–868. [Google Scholar] [CrossRef]
  15. Miyara, M.; Amoura, Z.; Parizot, C.; Badoual, C.; Dorgham, K.; Trad, S.; Nochy, D.; Debré, P.; Piette, J.-C.; Gorochov, G. Global natural regulatory T cell depletion in active systemic lupus erythematosus. J. Immunol. 2005, 175, 8392–8400. [Google Scholar] [CrossRef] [Green Version]
  16. Sakaguchi, S.; Ono, M.; Setoguchi, R.; Yagi, H.; Hori, S.; Fehervari, Z.; Shimizu, J.; Takahashi, T.; Nomura, T. Foxp3+ CD25+ CD4+ natural regulatory T cells in dominant self-tolerance and autoimmune disease. Immunol. Rev. 2006, 212, 8–27. [Google Scholar] [CrossRef]
  17. Josefowicz, S.Z.; Lu, L.-F.; Rudensky, A.Y. Regulatory T cells: Mechanisms of differentiation and function. Annu. Rev. Immunol. 2012, 30, 531–564. [Google Scholar] [CrossRef]
  18. Linker-Israeli, M.; Deans, R.J.; Wallace, D.J.; Prehn, J.; Ozeri-Chen, T.; Klinenberg, J.R. Elevated levels of endogenous IL-6 in systemic lupus erythematosus. A putative role in pathogenesis. J. Immunol. 1991, 147, 117–123. [Google Scholar] [CrossRef]
  19. Langrish, C.L.; Chen, Y.; Blumenschein, W.M.; Mattson, J.; Basham, B.; Sedgwick, J.D.; McClanahan, T.; Kastelein, R.A.; Cua, D.J. IL-23 drives a pathogenic T cell population that induces autoimmune inflammation. J. Exp. Med. 2005, 201, 233–240. [Google Scholar] [CrossRef] [Green Version]
  20. Veldhoen, M.; Hocking, R.J.; Atkins, C.J.; Locksley, R.M.; Stockinger, B. TGFβ in the Context of an Inflammatory Cytokine Milieu Supports De Novo Differentiation of IL-17-Producing T Cells. Immunity 2006, 24, 179–189. [Google Scholar] [CrossRef] [Green Version]
  21. Apostolidis, S.A.; Lieberman, L.A.; Kis-Toth, K.; Crispín, J.C.; Tsokos, G.C. The Dysregulation of Cytokine Networks in Systemic Lupus Erythematosus. J. Interferon Cytokine Res. 2011, 31, 769–779. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. King, C.; Tangye, S.G.; Mackay, C.R. T follicular helper (TFH) cells in normal and dysregulated immune responses. Annu. Rev. Immunol. 2008, 26, 741–766. [Google Scholar] [CrossRef] [PubMed]
  23. Mountz, J.D.; Hsu, H.-C.; Ballesteros-Tato, A. Dysregulation of T Follicular Helper Cells in Lupus. J. Immunol. 2019, 202, 1649–1658. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Li, D.; Guo, B.; Wu, H.; Tan, L.; Chang, C.; Lu, Q. Interleukin-17 in systemic lupus erythematosus: A comprehensive review. Autoimmunity 2015, 48, 353–361. [Google Scholar] [CrossRef]
  25. Linker-Israeli, M.; Bakke, A.C.; Kitridou, R.C.; Gendler, S.; Gillis, S.; Horwitz, D.A. Defective production of interleukin 1 and interleukin 2 in patients with systemic lupus erythematosus (SLE). J. Immunol. 1983, 130, 2651–2655. [Google Scholar] [CrossRef]
  26. Maxfield, F.R.; McGraw, T.E. Endocytic recycling. Nat. Rev. Mol. Cell Biol. 2004, 5, 121–132. [Google Scholar] [CrossRef]
  27. Grant, B.D.; Donaldson, J.G. Pathways and mechanisms of endocytic recycling. Nat. Rev. Mol. Cell Biol. 2009, 10, 597–608. [Google Scholar] [CrossRef] [Green Version]
  28. Bitsikas, V.; Corrêa, I.R., Jr.; Nichols, B.J. Clathrin-independent pathways do not contribute significantly to endocytic flux. eLife 2014, 3, e03970. [Google Scholar] [CrossRef]
  29. Huotari, J.; Helenius, A. Endosome maturation. EMBO J. 2011, 30, 3481–3500. [Google Scholar] [CrossRef]
  30. Novick, P.; Zerial, M. The diversity of Rab proteins in vesicle transport. Curr. Opin. Cell Biol. 1997, 9, 496–504. [Google Scholar] [CrossRef]
  31. Barr, F.; Lambright, D.G. Rab GEFs and GAPs. Curr. Opin. Cell Biol. 2010, 22, 461–470. [Google Scholar] [CrossRef] [PubMed]
  32. Steinman, R.M.; Mellman, I.S.; Muller, W.A.; Cohn, Z.A. Cohn Endocytosis and the recycling of plasma membrane. J. Cell Biol. 1983, 96, 1–27. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Mellman, I. Endocytosis and molecular sorting. Annu. Rev. Cell Dev. Biol. 1996, 12, 575–625. [Google Scholar] [CrossRef] [Green Version]
  34. Wang, J.; Fedoseienko, A.; Chen, B.; Burstein, E.; Jia, D.; Billadeau, D.D. Endosomal Receptor Trafficking: Retromer and Beyond. Traffic 2018, 19, 578–590. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Sönnichsen, B.; De Renzis, S.; Nielsen, E.; Rietdorf, J.; Zerial, M. Distinct membrane domains on endosomes in the recycling pathway visualized by multicolor imaging of Rab4, Rab5, and Rab11. J. Cell Biol. 2000, 149, 901–914. [Google Scholar] [CrossRef] [PubMed]
  36. Sheff, D.R.; Daro, E.A.; Hull, M.; Mellman, I. The receptor recycling pathway contains two distinct populations of early endosomes with different sorting functions. J. Cell Biol. 1999, 145, 123–139. [Google Scholar] [CrossRef] [Green Version]
  37. Riederer, M.A.; Soldati, T.; Shapiro, A.D.; Lin, J.; Pfeffer, S.R. Lysosome biogenesis requires Rab9 function and receptor recycling from endosomes to the trans-Golgi network. J. Cell Biol. 1994, 125, 573–582. [Google Scholar] [CrossRef] [Green Version]
  38. Poteryaev, D.; Datta, S.; Ackema, K.; Zerial, M.; Spang, A. Identification of the switch in early-to-late endosome transition. Cell 2010, 141, 497–508. [Google Scholar] [CrossRef] [Green Version]
  39. Bogan, J.S. Regulation of glucose transporter translocation in health and diabetes. Annu. Rev. Biochem. 2012, 81, 507–532. [Google Scholar] [CrossRef]
  40. Sorkin, A.; von Zastrow, M. Endocytosis and signalling: Intertwining molecular networks. Nat. Rev. Mol. Cell Biol. 2009, 10, 609–622. [Google Scholar] [CrossRef] [Green Version]
  41. Sorkin, A.; Von Zastrow, M. Signal transduction and endocytosis: Close encounters of many kinds. Nat. Rev. Mol. Cell Biol. 2002, 3, 600–614. [Google Scholar] [CrossRef]
  42. Jones, M.C.; Caswell, P.T.; Norman, J.C. Endocytic recycling pathways: Emerging regulators of cell migration. Curr. Opin. Cell Biol. 2006, 18, 549–557. [Google Scholar] [CrossRef] [PubMed]
  43. Gorvel, J.P.; Chavrier, P.; Zerial, M.; Gruenberg, J. rab5 controls early endosome fusion in vitro. Cell 1991, 64, 915–925. [Google Scholar] [CrossRef] [PubMed]
  44. Bucci, C.; Parton, R.G.; Mather, I.H.; Stunnenberg, H.; Simons, K.; Hoflack, B.; Zerial, M. The small GTPase rab5 functions as a regulatory factor in the early endocytic pathway. Cell 1992, 70, 715–728. [Google Scholar] [CrossRef] [Green Version]
  45. Horiuchi, H.; Giner, A.; Hoflack, B.; Zerial, M. A GDP/GTP Exchange-stimulatory Activity for the Rab5-RabGDI Complex on Clathrin-coated Vesicles from Bovine Brain. J. Biol. Chem. 1995, 270, 11257–11262. [Google Scholar] [CrossRef] [Green Version]
  46. Christoforidis, S.; Miaczynska, M.; Ashman, K.; Wilm, M.; Zhao, L.; Yip, S.C.; Waterfield, M.D.; Backer, J.M.; Zerial, M. Phosphatidylinositol-3-OH kinases are Rab5 effectors. Nat. Cell Biol. 1999, 1, 249–252. [Google Scholar] [CrossRef]
  47. Schu, P.V.; Takegawa, K.; Fry, M.J.; Stack, J.H.; Waterfield, M.D.; Emr, S.D. Phosphatidylinositol 3-kinase encoded by yeast VPS34 gene essential for protein sorting. Science 1993, 260, 88–91. [Google Scholar] [CrossRef]
  48. Patki, V.; Virbasius, J.; Lane, W.S.; Toh, B.H.; Shpetner, H.S.; Corvera, S. Identification of an early endosomal protein regulated by phosphatidylinositol 3-kinase. Proc. Natl. Acad. Sci. USA 1997, 94, 7326–7330. [Google Scholar] [CrossRef] [Green Version]
  49. Mills, I.G.; Jones, A.T.; Clague, M.J. Involvement of the endosomal autoantigen EEA1 in homotypic fusion of early endosomes. Curr. Biol. 1998, 8, 881–884. [Google Scholar] [CrossRef] [Green Version]
  50. Simonsen, A.; Lippé, R.; Christoforidis, S.; Gaullier, J.M.; Brech, A.; Callaghan, J.; Toh, B.H.; Murphy, C.; Zerial, M.; Stenmark, H. EEA1 links PI(3)K function to Rab5 regulation of endosome fusion. Nature 1998, 394, 494–498. [Google Scholar] [CrossRef] [PubMed]
  51. Christoforidis, S.; McBride, H.M.; Burgoyne, R.D.; Zerial, M. The Rab5 effector EEA1 is a core component of endosome docking. Nature 1999, 397, 621–625. [Google Scholar] [CrossRef]
  52. Dumas, J.J.; Merithew, E.; Sudharshan, E.; Rajamani, D.; Hayes, S.; Lawe, D.; Corvera, S.; Lambright, D.G. Multivalent endosome targeting by homodimeric EEA1. Mol. Cell 2001, 8, 947–958. [Google Scholar] [CrossRef]
  53. Lawe, D.C.; Chawla, A.; Merithew, E.; Dumas, J.; Carrington, W.; Fogarty, K.; Lifshitz, L.; Tuft, R.; Lambright, D.; Corvera, S. Sequential roles for phosphatidylinositol 3-phosphate and Rab5 in tethering and fusion of early endosomes via their interaction with EEA1. J. Biol. Chem. 2002, 277, 8611–8617. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. McBride, H.M.; Rybin, V.; Murphy, C.; Giner, A.; Teasdale, R.; Zerial, M. Oligomeric complexes link Rab5 effectors with NSF and drive membrane fusion via interactions between EEA1 and syntaxin 13. Cell 1999, 98, 377–386. [Google Scholar] [CrossRef] [Green Version]
  55. Kouranti, I.; Sachse, M.; Arouche, N.; Goud, B.; Echard, A. Rab35 Regulates an Endocytic Recycling Pathway Essential for the Terminal Steps of Cytokinesis. Curr. Biol. 2006, 16, 1719–1725. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Choudhury, A.; Sharma, D.K.; Marks, D.L.; Pagano, R.E. Elevated endosomal cholesterol levels in Niemann-Pick cells inhibit rab4 and perturb membrane recycling. Mol. Biol. Cell 2004, 15, 4500–4511. [Google Scholar] [CrossRef]
  57. Yudowski, G.A.; Puthenveedu, M.A.; Henry, A.G.; von Zastrow, M. Cargo-Mediated Regulation of a Rapid Rab4-Dependent Recycling Pathway. Mol. Biol. Cell 2009, 20, 2774–2784. [Google Scholar] [CrossRef] [Green Version]
  58. McCaffrey, M.W.; Bielli, A.; Cantalupo, G.; Mora, S.; Roberti, V.; Santillo, M.; Drummond, F.; Bucci, C. Rab4 affects both recycling and degradative endosomal trafficking. FEBS Lett. 2001, 495, 21–30. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Nag, S.; Rani, S.; Mahanty, S.; Bissig, C.; Arora, P.; Azevedo, C.; Saiardi, A.; van der Sluijs, P.; Delevoye, C.; van Niel, G.; et al. Rab4A organizes endosomal domains for sorting cargo to lysosome-related organelles. J. Cell Sci. 2018, 131, jcs216226. [Google Scholar] [CrossRef] [Green Version]
  60. van der Sluijs, P.; Hull, M.; Webster, P.; Mâle, P.; Goud, B.; Mellman, I. The small GTP-binding protein rab4 controls an early sorting event on the endocytic pathway. Cell 1992, 70, 729–740. [Google Scholar] [CrossRef]
  61. Nagy, G.; Ward, J.; Mosser, D.D.; Koncz, A.; Gergely, P.; Stancato, C.; Qian, Y.; Fernandez, D.; Niland, B.; Grossman, C.E.; et al. Regulation of CD4 Expression via Recycling by HRES-1/RAB4 Controls Susceptibility to HIV Infection. J. Biol. Chem. 2006, 281, 34574–34591. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Busch, R.; Rinderknecht, C.H.; Roh, S.; Lee, A.W.; Harding, J.J.; Burster, T.; Hornell, T.M.C.; Mellins, E.D. Achieving stability through editing and chaperoning: Regulation of MHC class II peptide binding and expression. Immunol. Rev. 2005, 207, 242–260. [Google Scholar] [CrossRef] [PubMed]
  63. Lagattuta, K.A.; Kang, J.B.; Nathan, A.; Pauken, K.E.; Jonsson, A.H.; Rao, D.A.; Sharpe, A.H.; Ishigaki, K.; Raychaudhuri, S. Repertoire analyses reveal T cell antigen receptor sequence features that influence T cell fate. Nat. Immunol. 2022, 23, 446–457. [Google Scholar] [CrossRef] [PubMed]
  64. Pérez-Montesinos, G.; López-Ortega, O.; Piedra-Reyes, J.; Bonifaz, L.C.; Moreno, J. Dynamic Changes in the Intracellular Association of Selected Rab Small GTPases with MHC Class II and DM during Dendritic Cell Maturation. Front. Immunol. 2017, 8, 340. [Google Scholar] [CrossRef] [Green Version]
  65. Théry, C.; Ostrowski, M.; Segura, E. Membrane vesicles as conveyors of immune responses. Nat. Rev. Immunol. 2009, 9, 581–593. [Google Scholar] [CrossRef]
  66. Lazzarino, D.A.; Blier, P.; Mellman, I. The Monomeric Guanosine Triphosphatase rab4 Controls an Essential Step on the Pathway of Receptor-mediated Antigen Processing in B Cells. J. Exp. Med. 1998, 188, 1769–1774. [Google Scholar] [CrossRef] [Green Version]
  67. Krawczyk, M.; Leimgruber, E.; Seguín-Estévez, Q.; Dunand-Sauthier, I.; Barras, E.; Reith, W. Expression of RAB4B, a protein governing endocytic recycling, is co-regulated with MHC class II genes. Nucleic Acids Res. 2007, 35, 595–605. [Google Scholar] [CrossRef] [Green Version]
  68. Schorey, J.S.; Bhatnagar, S. Exosome function: From tumor immunology to pathogen biology. Traffic 2008, 9, 871–881. [Google Scholar] [CrossRef] [Green Version]
  69. Ahmed, I.; Akram, Z.; Iqbal, H.M.N.; Munn, A.L. The regulation of Endosomal Sorting Complex Required for Transport and accessory proteins in multivesicular body sorting and enveloped viral budding—An overview. Int. J. Biol. Macromol. 2019, 127, 1–11. [Google Scholar] [CrossRef] [Green Version]
  70. Ostrowski, M.; Carmo, N.B.; Krumeich, S.; Fanget, I.; Raposo, G.; Savina, A.; Moita, C.F.; Schauer, K.; Hume, A.N.; Freitas, R.P.; et al. Rab27a and Rab27b control different steps of the exosome secretion pathway. Nat. Cell Biol. 2010, 12, 19–30. [Google Scholar] [CrossRef] [Green Version]
  71. Kalluri, R.; LeBleu, V.S. The biology, function, and biomedical applications of exosomes. Science 2020, 367, eaau6977. [Google Scholar] [CrossRef] [PubMed]
  72. Smith, V.L.; Cheng, Y.; Bryant, B.R.; Schorey, J.S. Exosomes function in antigen presentation during an in vivo Mycobacterium tuberculosis infection. Sci. Rep. 2017, 7, 43578. [Google Scholar] [CrossRef] [Green Version]
  73. Tan, L.; Zhao, M.; Wu, H.; Zhang, Y.; Tong, X.; Gao, L.; Zhou, L.; Lu, Q.; Zeng, J. Downregulated Serum Exosomal miR-451a Expression Correlates With Renal Damage and Its Intercellular Communication Role in Systemic Lupus Erythematosus. Front. Immunol. 2021, 12, 630112. [Google Scholar] [CrossRef] [PubMed]
  74. Monks, C.R.; Freiberg, B.A.; Kupfer, H.; Sciaky, N.; Kupfer, A. Three-dimensional segregation of supramolecular activation clusters in T cells. Nature 1998, 395, 82–86. [Google Scholar] [CrossRef]
  75. Anderson, G.; Moore, N.C.; Owen, J.J.T.; Jenkinson, E.J. Cellular Interactions in Thymocyte Development. Annu. Rev. Immunol. 1996, 14, 73–99. [Google Scholar] [CrossRef] [PubMed]
  76. Krissansen, G.W.; Owen, M.J.; Verbi, W.; Crumpton, M.J. Primary structure of the T3 gamma subunit of the T3/T cell antigen receptor complex deduced from cDNA sequences: Evolution of the T3 gamma and delta subunits. EMBO J. 1986, 5, 1799–1808. [Google Scholar] [CrossRef]
  77. Kane, L.P.; Lin, J.; Weiss, A. Signal transduction by the TCR for antigen. Curr. Opin. Immunol. 2000, 12, 242–249. [Google Scholar] [CrossRef]
  78. Freiberg, B.A.; Kupfer, H.; Maslanik, W.; Delli, J.; Kappler, J.; Zaller, D.M.; Kupfer, A. Staging and resetting T cell activation in SMACs. Nat. Immunol. 2002, 3, 911–917. [Google Scholar] [CrossRef] [PubMed]
  79. Batista, A.; Millán, J.; Mittelbrunn, M.; Sánchez-Madrid, F.; Alonso, M.A. Recruitment of transferrin receptor to immunological synapse in response to TCR engagement. J. Immunol. 2004, 172, 6709–6714. [Google Scholar] [CrossRef] [Green Version]
  80. Kao, H.; Lin, J.; Littman, D.R.; Shaw, A.S.; Allen, P.M. Regulated Movement of CD4 In and Out of the Immunological Synapse1. J. Immunol. 2008, 181, 8248–8257. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Bouchet, J.; Del Río-Iñiguez, I.; Vázquez-Chávez, E.; Lasserre, R.; Agüera-González, S.; Cuche, C.; McCaffrey, M.W.; Di Bartolo, V.; Alcover, A. Rab11-FIP3 Regulation of Lck Endosomal Traffic Controls TCR Signal Transduction. J. Immunol. 2017, 198, 2967–2978. [Google Scholar] [CrossRef] [Green Version]
  82. Thill, P.A.; Weiss, A.; Chakraborty, A.K. Phosphorylation of a Tyrosine Residue on Zap70 by Lck and Its Subsequent Binding via an SH2 Domain May Be a Key Gatekeeper of T Cell Receptor Signaling In Vivo. Mol. Cell. Biol. 2016, 36, 2396–2402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Samelson, L.E. Signal transduction mediated by the T cell antigen receptor: The role of adapter proteins. Annu. Rev. Immunol. 2002, 20, 371–394. [Google Scholar] [CrossRef] [Green Version]
  84. Krishna, S.; Zhong, X. Role of diacylglycerol kinases in T cell development and function. Crit. Rev. Immunol. 2013, 33, 97–118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Soares, H.; Lasserre, R.; Alcover, A. Orchestrating cytoskeleton and intracellular vesicle traffic to build functional immunological synapses. Immunol. Rev. 2013, 256, 118–132. [Google Scholar] [CrossRef] [PubMed]
  86. Finetti, F.; Patrussi, L.; Masi, G.; Onnis, A.; Galgano, D.; Lucherini, O.M.; Pazour, G.J.; Baldari, C.T. Specific recycling receptors are targeted to the immune synapse by the intraflagellar transport system. J. Cell Sci. 2014, 127, 1924–1937. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Degli Esposti, M.; Matarrese, P.; Tinari, A.; Longo, A.; Recalchi, S.; Khosravi-Far, R.; Malorni, W.; Misasi, R.; Garofalo, T.; Sorice, M. Changes in membrane lipids drive increased endocytosis following Fas ligation. Apoptosis 2017, 22, 681–695. [Google Scholar] [CrossRef] [Green Version]
  88. Carpier, J.-M.; Zucchetti, A.E.; Bataille, L.; Dogniaux, S.; Shafaq-Zadah, M.; Bardin, S.; Lucchino, M.; Maurin, M.; Joannas, L.D.; Magalhaes, J.G.; et al. Rab6-dependent retrograde traffic of LAT controls immune synapse formation and T cell activation. J. Exp. Med. 2018, 215, 1245–1265. [Google Scholar] [CrossRef]
  89. González-Mancha, N.; Rodríguez-Rodríguez, C.; Alcover, A.; Merida, I. Sorting Nexin 27 Enables MTOC and Secretory Machinery Translocation to the Immune Synapse. Front. Immunol. 2022, 12, 814570. [Google Scholar] [CrossRef]
  90. Temkin, P.; Lauffer, B.; Jäger, S.; Cimermancic, P.; Krogan, N.J.; von Zastrow, M. SNX27 mediates retromer tubule entry and endosome-to-plasma membrane trafficking of signalling receptors. Nat. Cell Biol. 2011, 13, 715–721. [Google Scholar] [CrossRef] [Green Version]
  91. Chen, P.-M.; Tsokos, G.C. The role of CD8+ T-cell systemic lupus erythematosus pathogenesis: An update. Curr. Opin. Rheumatol. 2021, 33, 586–591. [Google Scholar] [CrossRef]
  92. Yin, Z.; Bai, L.; Li, W.; Zeng, T.; Tian, H.; Cui, J. Targeting T cell metabolism in the tumor microenvironment: An anti-cancer therapeutic strategy. J. Exp. Clin. Cancer Res. 2019, 38, 403. [Google Scholar] [CrossRef] [PubMed]
  93. Voss, K.; Sewell, A.E.; Krystofiak, E.S.; Gibson-Corley, K.N.; Young, A.C.; Basham, J.H.; Sugiura, A.; Arner, E.N.; Beavers, W.N.; Kunkle, D.E.; et al. Elevated transferrin receptor impairs T cell metabolism and function in systemic lupus erythematosus. Sci. Immunol. 2023, 8, eabq0178. [Google Scholar] [CrossRef] [PubMed]
  94. Enyedy, E.J.; Nambiar, M.P.; Liossis, S.N.; Dennis, G.; Kammer, G.M.; Tsokos, G.C. Fc epsilon receptor type I gamma chain replaces the deficient T cell receptor zeta chain in T cells of patients with systemic lupus erythematosus. Arthritis Rheum. 2001, 44, 1114–1121. [Google Scholar] [CrossRef] [PubMed]
  95. Takeuchi, T.; Suzuki, K.; Kondo, T.; Yoshimoto, K.; Tsuzaka, K. CD3 ζ defects in systemic lupus erythematosus. Ann. Rheum. Dis. 2012, 71, i78–i81. [Google Scholar] [CrossRef]
  96. Takai, T. Fc receptors and their role in immune regulation and autoimmunity. J. Clin. Immunol. 2005, 25, 1–18. [Google Scholar] [CrossRef]
  97. Liu, C.P.; Lin, W.J.; Huang, M.; Kappler, J.W.; Marrack, P. Development and function of T cells in T cell antigen receptor/CD3 zeta knockout mice reconstituted with Fc epsilon RI gamma. Proc. Natl. Acad. Sci. USA 1997, 94, 616–621. [Google Scholar] [CrossRef] [Green Version]
  98. Liossis, S.N.; Ding, X.Z.; Dennis, G.J.; Tsokos, G.C. Altered pattern of TCR/CD3-mediated protein-tyrosyl phosphorylation in T cells from patients with systemic lupus erythematosus. Deficient expression of the T cell receptor zeta chain. J. Clin. Investig. 1998, 101, 1448–1457. [Google Scholar] [CrossRef] [Green Version]
  99. Krishnan, S.; Warke, V.G.; Nambiar, M.P.; Tsokos, G.C.; Farber, D.L. The FcR gamma subunit and Syk kinase replace the CD3 zeta-chain and ZAP-70 kinase in the TCR signaling complex of human effector CD4 T cells. J. Immunol. 2003, 170, 4189–4195. [Google Scholar] [CrossRef] [Green Version]
  100. Choi, O.H.; Kim, J.-H. Calcium mobilization via sphingosine kinase in signalling by the FctRI antigen receptor. Nature 1996, 380, 634–636. [Google Scholar] [CrossRef]
  101. Savina, A.; Furlán, M.; Vidal, M.; Colombo, M.I. Exosome release is regulated by a calcium-dependent mechanism in K562 cells. J. Biol. Chem. 2003, 278, 20083–20090. [Google Scholar] [CrossRef] [Green Version]
  102. Feske, S.; Wulff, H.; Skolnik, E.Y. Ion channels in innate and adaptive immunity. Annu. Rev. Immunol. 2015, 33, 291–353. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Feske, S.; Gwack, Y.; Prakriya, M.; Srikanth, S.; Puppel, S.-H.; Tanasa, B.; Hogan, P.G.; Lewis, R.S.; Daly, M.; Rao, A. A mutation in Orai1 causes immune deficiency by abrogating CRAC channel function. Nature 2006, 441, 179–185. [Google Scholar] [CrossRef]
  104. Zhou, Y.; Srinivasan, P.; Razavi, S.; Seymour, S.; Meraner, P.; Gudlur, A.; Stathopulos, P.B.; Ikura, M.; Rao, A.; Hogan, P.G. Initial activation of STIM1, the regulator of store-operated calcium entry. Nat. Struct. Mol. Biol. 2013, 20, 973–981. [Google Scholar] [CrossRef] [Green Version]
  105. de Souza, L.B.; Ong, H.L.; Liu, X.; Ambudkar, I.S. Fast endocytic recycling determines TRPC1-STIM1 clustering in ER-PM junctions and plasma membrane function of the channel. Biochim. Biophys. Acta 2015, 1853, 2709–2721. [Google Scholar] [CrossRef] [Green Version]
  106. Chen, G.; Panicker, S.; Lau, K.-Y.; Apparsundaram, S.; Patel, V.A.; Chen, S.-L.; Soto, R.; Jung, J.K.C.; Ravindran, P.; Okuhara, D.; et al. Characterization of a novel CRAC inhibitor that potently blocks human T cell activation and effector functions. Mol. Immunol. 2013, 54, 355–367. [Google Scholar] [CrossRef]
  107. Wang, Y.; Huynh, W.; Skokan, T.D.; Lu, W.; Weiss, A.; Vale, R.D. CRACR2a is a calcium-activated dynein adaptor protein that regulates endocytic traffic. J. Cell Biol. 2019, 218, 1619–1633. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Wu, B.; Rice, L.; Shrimpton, J.; Lawless, D.; Walker, K.; Carter, C.; McKeown, L.; Anwar, R.; Doody, G.M.; Srikanth, S.; et al. Biallelic mutations in calcium release activated channel regulator 2A (CRACR2A) cause a primary immunodeficiency disorder. Elife 2021, 10, e72559. [Google Scholar] [CrossRef]
  109. Pearce, E.L.; Pearce, E.J. Metabolic Pathways in Immune Cell Activation and Quiescence. Immunity 2013, 38, 633–643. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Teng, X.; Cornaby, C.; Li, W.; Morel, L. Metabolic regulation of pathogenic autoimmunity: Therapeutic targeting. Curr. Opin. Immunol. 2019, 61, 10–16. [Google Scholar] [CrossRef]
  111. Sharabi, A.; Tsokos, G.C. T cell metabolism: New insights in systemic lupus erythematosus pathogenesis and therapy. Nat. Rev. Rheumatol. 2020, 16, 100–112. [Google Scholar] [CrossRef]
  112. Wieman, H.L.; Wofford, J.A.; Rathmell, J.C. Cytokine Stimulation Promotes Glucose Uptake via Phosphatidylinositol-3 Kinase/Akt Regulation of Glut1 Activity and Trafficking. MBoC 2007, 18, 1437–1446. [Google Scholar] [CrossRef] [Green Version]
  113. Wofford, J.A.; Wieman, H.L.; Jacobs, S.R.; Zhao, Y.; Rathmell, J.C. IL-7 promotes Glut1 trafficking and glucose uptake via STAT5-mediated activation of Akt to support T-cell survival. Blood 2008, 111, 2101–2111. [Google Scholar] [CrossRef] [Green Version]
  114. Cheng, K.W.; Agarwal, R.; Mitra, S.; Lee, J.-S.; Carey, M.; Gray, J.W.; Mills, G.B. Rab25 increases cellular ATP and glycogen stores protecting cancer cells from bioenergetic stress. EMBO Mol. Med. 2012, 4, 125–141. [Google Scholar] [CrossRef]
  115. Macintyre, A.N.; Gerriets, V.A.; Nichols, A.G.; Michalek, R.D.; Rudolph, M.C.; Deoliveira, D.; Anderson, S.M.; Abel, E.D.; Chen, B.J.; Hale, L.P.; et al. The glucose transporter Glut1 is selectively essential for CD4 T cell activation and effector function. Cell Metab. 2014, 20, 61–72. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Michalek, R.D.; Gerriets, V.A.; Jacobs, S.R.; Macintyre, A.N.; MacIver, N.J.; Mason, E.F.; Sullivan, S.A.; Nichols, A.G.; Rathmell, J.C. Cutting edge: Distinct glycolytic and lipid oxidative metabolic programs are essential for effector and regulatory CD4+ T cell subsets. J. Immunol. 2011, 186, 3299–3303. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Koga, T.; Sato, T.; Furukawa, K.; Morimoto, S.; Endo, Y.; Umeda, M.; Sumiyoshi, R.; Fukui, S.; Kawashiri, S.; Iwamoto, N.; et al. Promotion of Calcium/Calmodulin-Dependent Protein Kinase 4 by GLUT1-Dependent Glycolysis in Systemic Lupus Erythematosus. Arthritis Rheumatol. 2019, 71, 766–772. [Google Scholar] [CrossRef]
  118. Li, W.; Qu, G.; Choi, S.-C.; Cornaby, C.; Titov, A.; Kanda, N.; Teng, X.; Wang, H.; Morel, L. Targeting T Cell Activation and Lupus Autoimmune Phenotypes by Inhibiting Glucose Transporters. Front. Immunol. 2019, 10, 833. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Ho, P.-C.; Bihuniak, J.D.; Macintyre, A.N.; Staron, M.; Liu, X.; Amezquita, R.; Tsui, Y.-C.; Cui, G.; Micevic, G.; Perales, J.C.; et al. Phosphoenolpyruvate Is a Metabolic Checkpoint of Anti-tumor T Cell Responses. Cell 2015, 162, 1217–1228. [Google Scholar] [CrossRef] [Green Version]
  120. Perl, A.; Hanczko, R.; Telarico, T.; Oaks, Z.; Landas, S. Oxidative stress, inflammation and carcinogenesis are controlled through the pentose phosphate pathway by transaldolase. Trends Mol. Med. 2011, 17, 395–403. [Google Scholar] [CrossRef] [Green Version]
  121. Banki, K.; Hutter, E.; Colombo, E.; Gonchoroff, N.J.; Perl, A. Glutathione levels and sensitivity to apoptosis are regulated by changes in transaldolase expression. J. Biol. Chem. 1996, 271, 32994–33001. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Banki, K.; Hutter, E.; Gonchoroff, N.J.; Perl, A. Elevation of Mitochondrial Transmembrane Potential and Reactive Oxygen Intermediate Levels Are Early Events and Occur Independently from Activation of Caspases in Fas Signaling1. J. Immunol. 1999, 162, 1466–1479. [Google Scholar] [CrossRef]
  123. Gergely, P.; Grossman, C.; Niland, B.; Puskas, F.; Neupane, H.; Allam, F.; Banki, K.; Phillips, P.E.; Perl, A. Mitochondrial hyperpolarization and ATP depletion in patients with systemic lupus erythematosus. Arthritis Rheum. 2002, 46, 175–190. [Google Scholar] [CrossRef] [PubMed]
  124. Gergely, P.; Niland, B.; Gonchoroff, N.; Pullmann, R.; Phillips, P.E.; Perl, A. Persistent Mitochondrial Hyperpolarization, Increased Reactive Oxygen Intermediate Production, and Cytoplasmic Alkalinization Characterize Altered IL-10 Signaling in Patients with Systemic Lupus Erythematosus. J. Immunol. 2002, 169, 1092–1101. [Google Scholar] [CrossRef] [Green Version]
  125. Perl, A.; Gergely, P.; Nagy, G.; Koncz, A.; Banki, K. Mitochondrial hyperpolarization: A checkpoint of T-cell life, death and autoimmunity. Trends Immunol. 2004, 25, 360–367. [Google Scholar] [CrossRef] [Green Version]
  126. Lai, Z.W.; Hanczko, R.; Bonilla, E.; Caza, T.N.; Clair, B.; Bartos, A.; Miklossy, G.; Jimah, J.; Doherty, E.; Tily, H.; et al. N-acetylcysteine reduces disease activity by blocking mTOR in T cells of lupus patients. Arthiritis Rhem. 2012, 64, 2937–2946. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Caza, T.N.; Fernandez, D.R.; Talaber, G.; Oaks, Z.; Haas, M.; Madaio, M.P.; Lai, Z.-W.; Miklossy, G.; Singh, R.R.; Chudakov, D.M.; et al. HRES-1/Rab4-mediated depletion of Drp1 impairs mitochondrial homeostasis and represents a target for treatment in SLE. Ann. Rheum. Dis. 2014, 73, 1888–1897. [Google Scholar] [CrossRef]
  128. Shaw, G.C.; Cope, J.J.; Li, L.; Corson, K.; Hersey, C.; Ackermann, G.E.; Gwynn, B.; Lambert, A.J.; Wingert, R.A.; Traver, D.; et al. Mitoferrin is essential for erythroid iron assimilation. Nature 2006, 440, 96–100. [Google Scholar] [CrossRef]
  129. Wincup, C.; Sawford, N.; Rahman, A. Pathological mechanisms of abnormal iron metabolism and mitochondrial dysfunction in systemic lupus erythematosus. Expert. Rev. Clin. Immunol. 2021, 17, 957–967. [Google Scholar] [CrossRef]
  130. Kleven, M.D.; Jue, S.; Enns, C.A. The Transferrin Receptors, TfR1 and TfR2 Bind Transferrin through Differing Mechanisms. Biochemistry 2018, 57, 1552–1559. [Google Scholar] [CrossRef]
  131. Zhao, M.; Li, M.-Y.; Gao, X.-F.; Jia, S.-J.; Gao, K.-Q.; Zhou, Y.; Zhang, H.-H.; Huang, Y.; Wang, J.; Wu, H.-J.; et al. Downregulation of BDH2 modulates iron homeostasis and promotes DNA demethylation in CD4+ T cells of systemic lupus erythematosus. Clin. Immunol. 2018, 187, 113–121. [Google Scholar] [CrossRef] [PubMed]
  132. Gao, X.; Song, Y.; Wu, J.; Lu, S.; Min, X.; Liu, L.; Hu, L.; Zheng, M.; Du, P.; Yu, Y.; et al. Iron-dependent epigenetic modulation promotes pathogenic T cell differentiation in lupus. J. Clin. Investig. 2022, 132, e152345. [Google Scholar] [CrossRef]
  133. Jabara, H.H.; Boyden, S.E.; Chou, J.; Ramesh, N.; Massaad, M.J.; Benson, H.; Bainter, W.; Fraulino, D.; Rahimov, F.; Sieff, C.; et al. A missense mutation in TFRC, encoding transferrin receptor 1, causes combined immunodeficiency. Nat. Genet. 2016, 48, 74–78. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Aljohani, A.H.; Al-Mousa, H.; Arnaout, R.; Al-Dhekri, H.; Mohammed, R.; Alsum, Z.; Nicolas-Jilwan, M.; Alrogi, F.; Al-Muhsen, S.; Alazami, A.M.; et al. Clinical and Immunological Characterization of Combined Immunodeficiency Due to TFRC Mutation in Eight Patients. J. Clin. Immunol. 2020, 40, 1103–1110. [Google Scholar] [CrossRef] [PubMed]
  135. Rossatti, P.; Redpath, G.M.I.; Ziegler, L.; Samson, G.P.B.; Clamagirand, C.D.; Legler, D.F.; Rossy, J. Rapid increase in transferrin receptor recycling promotes adhesion during T cell activation. BMC Biol. 2022, 20, 189. [Google Scholar] [CrossRef]
  136. Voulgarelis, M.; Kokori, S.I.; Ioannidis, J.P.; Tzioufas, A.G.; Kyriaki, D.; Moutsopoulos, H.M. Anaemia in systemic lupus erythematosus: Aetiological profile and the role of erythropoietin. Ann. Rheum. Dis. 2000, 59, 217–222. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Lai, Y.; Zhao, S.; Chen, B.; Huang, Y.; Guo, C.; Li, M.; Ye, B.; Wang, S.; Zhang, H.; Yang, N. Iron controls T helper cell pathogenicity by promoting glucose metabolism in autoimmune myopathy. Clin. Transl. Med. 2022, 12, e999. [Google Scholar] [CrossRef]
  138. Imai, S.; Guarente, L. NAD+ and sirtuins in aging and disease. Trends Cell Biol. 2014, 24, 464–471. [Google Scholar] [CrossRef]
  139. Jukarainen, S.; Heinonen, S.; Rämö, J.T.; Rinnankoski-Tuikka, R.; Rappou, E.; Tummers, M.; Muniandy, M.; Hakkarainen, A.; Lundbom, J.; Lundbom, N.; et al. Obesity Is Associated With Low NAD+/SIRT Pathway Expression in Adipose Tissue of BMI-Discordant Monozygotic Twins. J. Clin. Endocrinol. Metab. 2016, 101, 275–283. [Google Scholar] [CrossRef] [Green Version]
  140. Gong, B.; Pan, Y.; Vempati, P.; Zhao, W.; Knable, L.; Ho, L.; Wang, J.; Sastre, M.; Ono, K.; Sauve, A.A.; et al. Nicotinamide riboside restores cognition through an upregulation of proliferator-activated receptor-γ coactivator 1α regulated β-secretase 1 degradation and mitochondrial gene expression in Alzheimer’s mouse models. Neurobiol. Aging 2013, 34, 1581–1588. [Google Scholar] [CrossRef] [Green Version]
  141. Gujar, A.D.; Le, S.; Mao, D.D.; Dadey, D.Y.A.; Turski, A.; Sasaki, Y.; Aum, D.; Luo, J.; Dahiya, S.; Yuan, L.; et al. An NAD+-dependent transcriptional program governs self-renewal and radiation resistance in glioblastoma. Proc. Natl. Acad. Sci. USA 2016, 113, E8247–E8256. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Kennedy, B.E.; Sharif, T.; Martell, E.; Dai, C.; Kim, Y.; Lee, P.W.K.; Gujar, S.A. NAD+ salvage pathway in cancer metabolism and therapy. Pharmacol. Res. 2016, 114, 274–283. [Google Scholar] [CrossRef] [PubMed]
  143. Katsuyama, E.; Suarez-Fueyo, A.; Bradley, S.J.; Mizui, M.; Marin, A.V.; Mulki, L.; Krishfield, S.; Malavasi, F.; Yoon, J.; Ho Sui, S.J.; et al. The CD38/NAD/SIRTUIN1/EZH2 Axis Mitigates Cytotoxic CD8 T Cell Function and Identifies Patients with SLE Prone to Infections. Cell Rep. 2020, 30, 112–123. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Krebs, H.A.; Veech, R.L. Equilibrium relations between pyridine nucleotides and adenine nucleotides and their roles in the regulation of metabolic processes. Adv. Enzym. Regul. 1969, 7, 397–413. [Google Scholar] [CrossRef]
  145. Yang, Y.; Sauve, A.A. NAD+ metabolism: Bioenergetics, signaling and manipulation for therapy. Biochim. Biophys. Acta 2016, 1864, 1787–1800. [Google Scholar] [CrossRef] [Green Version]
  146. Anderson, K.A.; Madsen, A.S.; Olsen, C.A.; Hirschey, M.D. Metabolic control by sirtuins and other enzymes that sense NAD+, NADH, or their ratio. Biochim. Biophys. Acta 2017, 1858, 991–998. [Google Scholar] [CrossRef]
  147. Fletcher, R.S.; Lavery, G.G. The emergence of the nicotinamide riboside kinases in the regulation of NAD+ metabolism. J. Mol. Endocrinol. 2018, 61, R107–R121. [Google Scholar] [CrossRef] [Green Version]
  148. Chalkiadaki, A.; Guarente, L. The multifaceted functions of sirtuins in cancer. Nat. Rev. Cancer 2015, 15, 608–624. [Google Scholar] [CrossRef]
  149. Nacarelli, T.; Zhang, R. NAD+ metabolism controls inflammation during senescence. Mol. Cell Oncol. 2019, 6, 1605819. [Google Scholar] [CrossRef]
  150. Poljsak, B. NAD+ in Cancer Prevention and Treatment: Pros and Cons. J. Clin. Exp. Oncol. 2018, 2016, 2. [Google Scholar] [CrossRef]
  151. Imai, S.; Armstrong, C.M.; Kaeberlein, M.; Guarente, L. Transcriptional silencing and longevity protein Sir2 is an NAD-dependent histone deacetylase. Nature 2000, 403, 795–800. [Google Scholar] [CrossRef]
  152. Chambon, P.; Weill, J.D.; Mandel, P. Nicotinamide mononucleotide activation of new DNA-dependent polyadenylic acid synthesizing nuclear enzyme. Biochem. Biophys. Res. Commun. 1963, 11, 39–43. [Google Scholar] [CrossRef] [PubMed]
  153. De Flora, A.; Zocchi, E.; Guida, L.; Franco, L.; Bruzzone, S. Autocrine and paracrine calcium signaling by the CD38/NAD+/cyclic ADP-ribose system. Ann. N. Y. Acad. Sci. 2004, 1028, 176–191. [Google Scholar] [CrossRef]
  154. Hottiger, M.O.; Hassa, P.O.; Lüscher, B.; Schüler, H.; Koch-Nolte, F. Toward a unified nomenclature for mammalian ADP-ribosyltransferases. Trends Biochem. Sci. 2010, 35, 208–219. [Google Scholar] [CrossRef]
  155. Ijichi, H.; Ichiyama, A.; Hayaishi, O. Studies on the biosynthesis of nicotinamide adenine dinucleotide. 3. Comparative in vivo studies on nicotinic acid, nicotinamide, and quinolinic acid as precursors of nicotinamide adenine dinucleotide. J. Biol. Chem. 1966, 241, 3701–3707. [Google Scholar] [CrossRef]
  156. Navas, L.E.; Carnero, A. NAD+ metabolism, stemness, the immune response, and cancer. Sig Transduct. Target. 2021, 6, 2. [Google Scholar] [CrossRef]
  157. Sinclair, L.V.; Neyens, D.; Ramsay, G.; Taylor, P.M.; Cantrell, D.A. Single cell analysis of kynurenine and System L amino acid transport in T cells. Nat. Commun. 2018, 9, 1981. [Google Scholar] [CrossRef] [PubMed]
  158. Widner, B.; Sepp, N.; Kowald, E.; Kind, S.; Schmuth, M.; Fuchs, D. Degradation of tryptophan in patients with systemic lupus erythematosus. Adv. Exp. Med. Biol. 1999, 467, 571–577. [Google Scholar] [CrossRef] [PubMed]
  159. Widner, B.; Sepp, N.; Kowald, E.; Ortner, U.; Wirleitner, B.; Fritsch, P.; Baier-Bitterlich, G.; Fuchs, D. Enhanced tryptophan degradation in systemic lupus erythematosus. Immunobiology 2000, 201, 621–630. [Google Scholar] [CrossRef]
  160. Oaks, Z.; Perl, A. Metabolic control of the epigenome in systemic Lupus erythematosus. Autoimmunity 2014, 47, 256–264. [Google Scholar] [CrossRef] [Green Version]
  161. Yin, Y.; Choi, S.-C.; Xu, Z.; Perry, D.J.; Seay, H.; Croker, B.P.; Sobel, E.S.; Brusko, T.M.; Morel, L. Normalization of CD4+ T Cell Metabolism Reverses Lupus. Sci. Transl. Med. 2015, 7, 274ra18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Maldonado-Báez, L.; Cole, N.B.; Krämer, H.; Donaldson, J.G. Microtubule-dependent endosomal sorting of clathrin-independent cargo by Hook1. J. Cell Biol. 2013, 201, 233–247. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Choi, S.-C.; Brown, J.; Gong, M.; Ge, Y.; Zadeh, M.; Li, W.; Croker, B.P.; Michailidis, G.; Garrett, T.J.; Mohamadzadeh, M.; et al. Gut microbiota dysbiosis and altered tryptophan catabolism contribute to autoimmunity in lupus-susceptible mice. Sci. Transl. Med. 2020, 12, eaax2220. [Google Scholar] [CrossRef] [PubMed]
  164. Brown, J.; Abboud, G.; Ma, L.; Choi, S.-C.; Kanda, N.; Zeumer-Spataro, L.; Lee, J.; Peng, W.; Cagmat, J.; Faludi, T.; et al. Microbiota-mediated skewing of tryptophan catabolism modulates CD4+ T cells in lupus-prone mice. iScience 2022, 25, 104241. [Google Scholar] [CrossRef]
  165. Perl, A.; Hanczko, R.; Lai, Z.W.; Oaks, Z.; Kelly, R.; Borsuk, R.; Asara, J.M.; Phillips, P.E. Comprehensive metabolome analyses reveal N-acetylcysteine-responsive accumulation of kynurenine in systemic lupus erythematosus: Implications for activation of the mechanistic target of rapamycin. Metabolomics 2015, 11, 1157–1174. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Lai, Z.W.; Borsuk, R.; Shadakshari, A.; Yu, J.; Dawood, M.; Garcia, R.; Francis, L.; Tily, H.; Bartos, A.; Faraone, S.V.; et al. Mechanistic target of rapamycin activation triggers IL-4 production and necrotic death of double-negative T cells in patients with systemic lupus erythematosus. J. Immunol. 2013, 191, 2236–2246. [Google Scholar] [CrossRef] [Green Version]
  167. Lai, Z.-W.; Kelly, R.; Winans, T.; Marchena, I.; Shadakshari, A.; Yu, J.; Dawood, M.; Garcia, R.; Tily, H.; Francis, L.; et al. Sirolimus in patients with clinically active systemic lupus erythematosus resistant to, or intolerant of, conventional medications: A single-arm, open-label, phase 1/2 trial. Lancet 2018, 391, 1186–1196. [Google Scholar] [CrossRef]
  168. Kato, H.; Perl, A. The IL-21-mtor axis blocks treg differentiation and function by suppression of autophagy in patients with systemic lupus erythematosus. Arthritis Rheumatol. 2018, 70, 427. [Google Scholar] [CrossRef] [Green Version]
  169. Liu, T.-T.; Gomez, T.S.; Sackey, B.K.; Billadeau, D.D.; Burd, C.G. Rab GTPase regulation of retromer-mediated cargo export during endosome maturation. Mol. Biol. Cell 2012, 23, 2505–2515. [Google Scholar] [CrossRef]
  170. Curnock, R.; Calcagni, A.; Ballabio, A.; Cullen, P.J. TFEB controls retromer expression in response to nutrient availability. J. Cell Biol. 2019, 218, 3954–3966. [Google Scholar] [CrossRef]
  171. Nakaya, M.; Xiao, Y.; Zhou, X.; Chang, J.-H.; Chang, M.; Cheng, X.; Blonska, M.; Lin, X.; Sun, S.-C. Inflammatory T cell responses rely on amino acid transporter ASCT2 facilitation of glutamine uptake and mTORC1 kinase activation. Immunity 2014, 40, 692–705. [Google Scholar] [CrossRef] [Green Version]
  172. Katz, F.; Povey, S.; Parkar, M.; Schneider, C.; Sutherland, R.; Stanley, K.; Solomon, E.; Greaves, M. Chromosome assignment of monoclonal antibody-defined determinants on human leukemic cells. Eur. J. Immunol. 1983, 13, 1008–1013. [Google Scholar] [CrossRef] [PubMed]
  173. Muraoka, O.; Tanaka, H.; Itoh, M.; Ishihara, K.; Hirano, T. Genomic structure of human BST-1. Immunol. Lett. 1996, 54, 1–4. [Google Scholar] [CrossRef] [PubMed]
  174. Aarhus, R.; Graeff, R.M.; Dickey, D.M.; Walseth, T.F.; Hon, C.L. ADP-ribosyl Cyclase and CD38 Catalyze the Synthesis of a Calcium-mobilizing Metabolite from NADP+(∗). J. Biol. Chem. 1995, 270, 30327–30333. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Malavasi, F.; Deaglio, S.; Funaro, A.; Ferrero, E.; Horenstein, A.L.; Ortolan, E.; Vaisitti, T.; Aydin, S. Evolution and function of the ADP ribosyl cyclase/CD38 gene family in physiology and pathology. Physiol. Rev. 2008, 88, 841–886. [Google Scholar] [CrossRef] [Green Version]
  176. Quarona, V.; Zaccarello, G.; Chillemi, A.; Brunetti, E.; Singh, V.K.; Ferrero, E.; Funaro, A.; Horenstein, A.L.; Malavasi, F. CD38 and CD157: A long journey from activation markers to multifunctional molecules. Cytom. Part B Clin. Cytom. 2013, 84B, 207–217. [Google Scholar] [CrossRef]
  177. Aksoy, P.; White, T.A.; Thompson, M.; Chini, E.N. Regulation of intracellular levels of NAD: A novel role for CD38. Biochem. Biophys. Res. Commun. 2006, 345, 1386–1392. [Google Scholar] [CrossRef]
  178. Bhan, A.K.; Reinherz, E.L.; Poppema, S.; McCluskey, R.T.; Schlossman, S.F. Location of T cell and major histocompatibility complex antigens in the human thymus. J. Exp. Med. 1980, 152, 771–782. [Google Scholar] [CrossRef] [Green Version]
  179. Ortolan, E.; Augeri, S.; Fissolo, G.; Musso, I.; Funaro, A. CD157: From immunoregulatory protein to potential therapeutic target. Immunol. Lett. 2019, 205, 59–64. [Google Scholar] [CrossRef] [Green Version]
  180. Lee, B.O.; Ishihara, K.; Denno, K.; Kobune, Y.; Itoh, M.; Muraoka, O.; Kaisho, T.; Sasaki, T.; Ochi, T.; Hirano, T. Elevated levels of the soluble form of bone marrow stromal cell antigen 1 in the sera of patients with severe rheumatoid arthritis. Arthritis Rheum. 1996, 39, 629–637. [Google Scholar] [CrossRef]
  181. Khandelwal, P.; Lane, A.; Chaturvedi, V.; Owsley, E.; Davies, S.M.; Marmer, D.; Filipovich, A.H.; Jordan, M.B.; Marsh, R.A. Peripheral Blood CD38 Bright CD8+ Effector Memory T Cells Predict Acute Graft-versus-Host Disease. Biol. Blood Marrow Transplant. 2015, 21, 1215–1222. [Google Scholar] [CrossRef] [Green Version]
  182. Joosse, M.E.; Menckeberg, C.L.; de Ruiter, L.F.; Raatgeep, H.C.; van Berkel, L.A.; Simons-Oosterhuis, Y.; Tindemans, I.; Muskens, A.M.; Hendriks, R.W.; Hoogenboezem, R.M.; et al. Frequencies of circulating regulatory TIGIT+CD38+ effector T cells correlate with the course of inflammatory bowel disease. Mucosal Immunol. 2019, 12, 154–163. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Pavón, E.J.; Zumaquero, E.; Rosal-Vela, A.; Khoo, K.-M.; Cerezo-Wallis, D.; García-Rodríguez, S.; Carrascal, M.; Abian, J.; Graeff, R.; Callejas-Rubio, J.-L.; et al. Increased CD38 expression in T cells and circulating anti-CD38 IgG autoantibodies differentially correlate with distinct cytokine profiles and disease activity in systemic lupus erythematosus patients. Cytokine 2013, 62, 232–243. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Chini, E.N. CD38 as a regulator of cellular NAD: A novel potential pharmacological target for metabolic conditions. Curr. Pharm. Des. 2009, 15, 57–63. [Google Scholar] [CrossRef] [Green Version]
  185. Kim, H.; Jacobson, E.L.; Jacobson, M.K. Synthesis and degradation of cyclic ADP-ribose by NAD glycohydrolases. Science 1993, 261, 1330–1333. [Google Scholar] [CrossRef]
  186. Zielinska, W.; Barata, H.; Chini, E.N. Metabolism of cyclic ADP-ribose: Zinc is an endogenous modulator of the cyclase/NAD glycohydrolase ratio of a CD38-like enzyme from human seminal fluid. Life Sci. 2004, 74, 1781–1790. [Google Scholar] [CrossRef]
  187. Kar, A.; Mehrotra, S.; Chatterjee, S. CD38: T Cell Immuno-Metabolic Modulator. Cells 2020, 9, 1716. [Google Scholar] [CrossRef] [PubMed]
  188. Funaro, A.; Reiniš, M.; Trubiani, O.; Santi, S.; Di Primio, R.; Malavasi, F. CD38 Functions Are Regulated Through an Internalization Step1. J. Immunol. 1998, 160, 2238–2247. [Google Scholar] [CrossRef] [PubMed]
  189. Muñoz, P.; Mittelbrunn, M.; de la Fuente, H.; Pérez-Martínez, M.; García-Pérez, A.; Ariza-Veguillas, A.; Malavasi, F.; Zubiaur, M.; Sánchez-Madrid, F.; Sancho, J. Antigen-induced clustering of surface CD38 and recruitment of intracellular CD38 to the immunologic synapse. Blood 2008, 111, 3653–3664. [Google Scholar] [CrossRef] [Green Version]
  190. Allard, B.; Turcotte, M.; Stagg, J. Targeting CD73 and downstream adenosine receptor signaling in triple-negative breast cancer. Expert Opin. Ther. Targets 2014, 18, 863–881. [Google Scholar] [CrossRef]
  191. Koszałka, P.; Gołuńska, M.; Stanisławowski, M.; Urban, A.; Stasiłojć, G.; Majewski, M.; Wierzbicki, P.; Składanowski, A.C.; Bigda, J. CD73 on B16F10 melanoma cells in CD73-deficient mice promotes tumor growth, angiogenesis, neovascularization, macrophage infiltration and metastasis. Int. J. Biochem. Cell Biol. 2015, 69, 1–10. [Google Scholar] [CrossRef]
  192. Turcotte, M.; Spring, K.; Pommey, S.; Chouinard, G.; Cousineau, I.; George, J.; Chen, G.M.; Gendoo, D.M.A.; Haibe-Kains, B.; Karn, T.; et al. CD73 Is Associated with Poor Prognosis in High-Grade Serous Ovarian Cancer. Cancer Res. 2015, 75, 4494–4503. [Google Scholar] [CrossRef] [Green Version]
  193. Hesse, J.; Siekierka-Harreis, M.; Steckel, B.; Alter, C.; Schallehn, M.; Honke, N.; Schnieringer, M.-L.; Wippich, M.; Braband, R.; Schneider, M.; et al. Profound inhibition of CD73-dependent formation of anti-inflammatory adenosine in B cells of SLE patients. EBioMedicine 2021, 73, 103616. [Google Scholar] [CrossRef] [PubMed]
  194. Li, D.; Li, X.; Li, X.; Wang, G.; Ma, Y.; Zhao, S.; Zheng, S. Expression of FOXP3 in CD4+ CD39+ T cells of patients with systemic lupus erythematosus and dynamic observation of treatment with glucocorticoid. Zhonghua Yi Xue Za Zhi 2009, 89, 1636–1638. [Google Scholar]
  195. Li, D.; Li, X.; Zhang, J.; Hu, S.; Xiao, B.; Chen, W.; Zeng, X. The expression of CD73 in CD4+ regulatory T cells in patients with new-onset systemic lupus erythematosus. Zhonghua Nei Ke Za Zhi 2010, 49, 772–775. [Google Scholar] [PubMed]
  196. Grozio, A.; Sociali, G.; Sturla, L.; Caffa, I.; Soncini, D.; Salis, A.; Raffaelli, N.; De Flora, A.; Nencioni, A.; Bruzzone, S. CD73 protein as a source of extracellular precursors for sustained NAD+ biosynthesis in FK866-treated tumor cells. J. Biol. Chem. 2013, 288, 25938–25949. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Sociali, G.; Raffaghello, L.; Magnone, M.; Zamporlini, F.; Emionite, L.; Sturla, L.; Bianchi, G.; Vigliarolo, T.; Nahimana, A.; Nencioni, A.; et al. Antitumor effect of combined NAMPT and CD73 inhibition in an ovarian cancer model. Oncotarget 2016, 7, 2968–2984. [Google Scholar] [CrossRef] [Green Version]
  198. Garavaglia, S.; Bruzzone, S.; Cassani, C.; Canella, L.; Allegrone, G.; Sturla, L.; Mannino, E.; Millo, E.; De Flora, A.; Rizzi, M. The high-resolution crystal structure of periplasmic Haemophilus influenzae NAD nucleotidase reveals a novel enzymatic function of human CD73 related to NAD metabolism. Biochem. J. 2012, 441, 131–141. [Google Scholar] [CrossRef] [PubMed]
  199. Ratajczak, J.; Joffraud, M.; Trammell, S.A.J.; Ras, R.; Canela, N.; Boutant, M.; Kulkarni, S.S.; Rodrigues, M.; Redpath, P.; Migaud, M.E.; et al. NRK1 controls nicotinamide mononucleotide and nicotinamide riboside metabolism in mammalian cells. Nat. Commun. 2016, 7, 13103. [Google Scholar] [CrossRef] [Green Version]
  200. Wilk, A.; Hayat, F.; Cunningham, R.; Li, J.; Garavaglia, S.; Zamani, L.; Ferraris, D.M.; Sykora, P.; Andrews, J.; Clark, J.; et al. Extracellular NAD+ enhances PARP-dependent DNA repair capacity independently of CD73 activity. Sci. Rep. 2020, 10, 651. [Google Scholar] [CrossRef] [Green Version]
  201. Romio, M.; Reinbeck, B.; Bongardt, S.; Hüls, S.; Burghoff, S.; Schrader, J. Extracellular purine metabolism and signaling of CD73-derived adenosine in murine Treg and Teff cells. Am. J. Physiol. Cell Physiol. 2011, 301, C530–C539. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  202. Smyth, L.A.; Ratnasothy, K.; Tsang, J.Y.S.; Boardman, D.; Warley, A.; Lechler, R.; Lombardi, G. CD73 expression on extracellular vesicles derived from CD4+CD25+Foxp3+ T cells contributes to their regulatory function. Eur. J. Immunol. 2013, 43, 2430–2440. [Google Scholar] [CrossRef] [PubMed]
  203. Grozio, A.; Mills, K.F.; Yoshino, J.; Bruzzone, S.; Sociali, G.; Tokizane, K.; Lei, H.C.; Cunningham, R.; Sasaki, Y.; Migaud, M.E.; et al. Slc12a8 is a nicotinamide mononucleotide transporter. Nat. Metab. 2019, 1, 47–57. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Zhang, Q.; Liu, Y.; Chen, P.; Shi, X.; Liu, Y.; Shi, L.; Cong, P.; Mao, S.; Tong, C.; Du, C.; et al. Solute carrier family 12 member 8 (SLC12A8) is a potential biomarker and related to tumor immune cell infiltration in bladder cancer. Bioengineered 2021, 12, 4946–4961. [Google Scholar] [CrossRef]
  205. Castro, C.; Oyamada, H.A.A.; Cafasso, M.O.S.D.; Lopes, L.M.; Monteiro, C.; Sacramento, P.M.; Alves-Leon, S.V.; da Fontoura Galvão, G.; Hygino, J.; de Souza, J.P.B.M.; et al. Elevated proportion of TLR2- and TLR4-expressing Th17-like cells and activated memory B cells was associated with clinical activity of cerebral cavernous malformations. J. Neuroinflamm. 2022, 19, 28. [Google Scholar] [CrossRef]
  206. Poltorak, A.; He, X.; Smirnova, I.; Liu, M.-Y.; Huffel, C.V.; Du, X.; Birdwell, D.; Alejos, E.; Silva, M.; Galanos, C.; et al. Defective LPS Signaling in C3H/HeJ and C57BL/10ScCr Mice: Mutations in Tlr4 Gene. Science 1998, 282, 2085–2088. [Google Scholar] [CrossRef] [Green Version]
  207. Anders, H.-J. Toll-like receptors and danger signaling in kidney injury. J. Am. Soc. Nephrol. 2010, 21, 1270–1274. [Google Scholar] [CrossRef] [Green Version]
  208. Brandt, K.J.; Fickentscher, C.; Kruithof, E.K.; de Moerloose, P. TLR2 Ligands Induce NF-κB Activation from Endosomal Compartments of Human Monocytes. PLoS ONE 2013, 8, e80743. [Google Scholar] [CrossRef] [Green Version]
  209. Zanoni, I.; Ostuni, R.; Marek, L.R.; Barresi, S.; Barbalat, R.; Barton, G.M.; Granucci, F.; Kagan, J.C. CD14 controls the LPS-induced endocytosis of Toll-like receptor 4. Cell 2011, 147, 868–880. [Google Scholar] [CrossRef] [Green Version]
  210. Husebye, H.; Aune, M.H.; Stenvik, J.; Samstad, E.; Skjeldal, F.; Halaas, O.; Nilsen, N.J.; Stenmark, H.; Latz, E.; Lien, E.; et al. The Rab11a GTPase controls Toll-like receptor 4-induced activation of interferon regulatory factor-3 on phagosomes. Immunity 2010, 33, 583–596. [Google Scholar] [CrossRef] [Green Version]
  211. Wang, D.; Lou, J.; Ouyang, C.; Chen, W.; Liu, Y.; Liu, X.; Cao, X.; Wang, J.; Lu, L. Ras-related protein Rab10 facilitates TLR4 signaling by promoting replenishment of TLR4 onto the plasma membrane. Proc. Natl. Acad. Sci. USA 2010, 107, 13806–13811. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Petnicki-Ocwieja, T.; Sharma, B.; Powale, U.; Pathak, D.; Tan, S.; Hu, L.T. An AP-3-dependent pathway directs phagosome fusion with Rab8 and Rab11 vesicles involved in TLR2 signaling. Traffic 2022, 23, 558–567. [Google Scholar] [CrossRef]
  213. Liu, Y.; Liao, J.; Zhao, M.; Wu, H.; Yung, S.; Chan, T.M.; Yoshimura, A.; Lu, Q. Increased expression of TLR2 in CD4+ T cells from SLE patients enhances immune reactivity and promotes IL-17 expression through histone modifications. Eur. J. Immunol. 2015, 45, 2683–2693. [Google Scholar] [CrossRef] [PubMed]
  214. Nyirenda, M.H.; Sanvito, L.; Darlington, P.J.; O’Brien, K.; Zhang, G.-X.; Constantinescu, C.S.; Bar-Or, A.; Gran, B. TLR2 Stimulation Drives Human Naive and Effector Regulatory T Cells into a Th17-Like Phenotype with Reduced Suppressive Function. J. Immunol. 2011, 187, 2278–2290. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Heil, F.; Hemmi, H.; Hochrein, H.; Ampenberger, F.; Kirschning, C.; Akira, S.; Lipford, G.; Wagner, H.; Bauer, S. Species-specific recognition of single-stranded RNA via toll-like receptor 7 and 8. Science 2004, 303, 1526–1529. [Google Scholar] [CrossRef] [Green Version]
  216. Lee, Y.H.; Choi, S.J.; Ji, J.D.; Song, G.G. Association between toll-like receptor polymorphisms and systemic lupus erythematosus: A meta-analysis update. Lupus 2016, 25, 593–601. [Google Scholar] [CrossRef]
  217. Wang, C.-M.; Chang, S.-W.; Wu, Y.-J.J.; Lin, J.-C.; Ho, H.-H.; Chou, T.-C.; Yang, B.; Wu, J.; Chen, J.-Y. Genetic variations in Toll-like receptors (TLRs 3/7/8) are associated with systemic lupus erythematosus in a Taiwanese population. Sci. Rep. 2014, 4, 3792. [Google Scholar] [CrossRef] [Green Version]
  218. Guo, Y.; Chai, Q.; Zhao, Y.; Li, P.; Qiao, J.; Huang, J. Increased activation of toll-like receptors-7 and -8 of peripheral blood mononuclear cells and upregulated serum cytokines in patients with pediatric systemic lupus erythematosus. Int. J. Clin. Exp. Med. 2015, 8, 20472–20480. [Google Scholar]
  219. Dominguez-Villar, M.; Gautron, A.-S.; de Marcken, M.; Keller, M.J.; Hafler, D.A. TLR7 induces anergy in human CD4+ T cells. Nat. Immunol. 2015, 16, 118–128. [Google Scholar] [CrossRef]
  220. Li, L.; Liu, X.; Sanders, K.L.; Edwards, J.L.; Ye, J.; Si, F.; Gao, A.; Huang, L.; Hsueh, E.C.; Ford, D.A.; et al. TLR8-Mediated Metabolic Control of Human Treg Function: A Mechanistic Target for Cancer Immunotherapy. Cell Metab. 2019, 29, 103–123. [Google Scholar] [CrossRef] [Green Version]
  221. Hedrich, C.M. Epigenetics in SLE. Curr. Rheumatol. Rep. 2017, 19, 58. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. Alcocer-Varela, J.; Alarcón-Segovia, D. Decreased production of and response to interleukin-2 by cultured lymphocytes from patients with systemic lupus erythematosus. J. Clin. Investig. 1982, 69, 1388–1392. [Google Scholar] [CrossRef] [PubMed]
  223. Fu, S.; Zhang, N.; Yopp, A.C.; Chen, D.; Mao, M.; Chen, D.; Zhang, H.; Ding, Y.; Bromberg, J.S. TGF-beta induces Foxp3 + T-regulatory cells from CD4 + CD25—Precursors. Am. J. Transplant. 2004, 4, 1614–1627. [Google Scholar] [CrossRef]
  224. Malek, T.R.; Yu, A.; Vincek, V.; Scibelli, P.; Kong, L. CD4 regulatory T cells prevent lethal autoimmunity in IL-2Rbeta-deficient mice. Implications for the nonredundant function of IL-2. Immunity 2002, 17, 167–178. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Solomou, E.E.; Juang, Y.T.; Gourley, M.F.; Kammer, G.M.; Tsokos, G.C. Molecular basis of deficient IL-2 production in T cells from patients with systemic lupus erythematosus. J. Immunol. 2001, 166, 4216–4222. [Google Scholar] [CrossRef] [Green Version]
  226. Juang, Y.-T.; Wang, Y.; Solomou, E.E.; Li, Y.; Mawrin, C.; Tenbrock, K.; Kyttaris, V.C.; Tsokos, G.C. Systemic lupus erythematosus serum IgG increases CREM binding to the IL-2 promoter and suppresses IL-2 production through CaMKIV. J. Clin. Investig. 2005, 115, 996–1005. [Google Scholar] [CrossRef]
  227. Sunahori, K.; Juang, Y.-T.; Kyttaris, V.C.; Tsokos, G.C. Promoter Hypomethylation Results in Increased Expression of Protein Phosphatase 2A in T Cells from Patients with Systemic Lupus Erythematosus. J. Immunol. 2011, 186, 4508–4517. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Anderson, K.A.; Noeldner, P.K.; Reece, K.; Wadzinski, B.E.; Means, A.R. Regulation and function of the calcium/calmodulin-dependent protein kinase IV/protein serine/threonine phosphatase 2A signaling complex. J. Biol. Chem. 2004, 279, 31708–31716. [Google Scholar] [CrossRef] [Green Version]
  229. Katsiari, C.G.; Kyttaris, V.C.; Juang, Y.-T.; Tsokos, G.C. Protein phosphatase 2A is a negative regulator of IL-2 production in patients with systemic lupus erythematosus. J. Clin. Investig. 2005, 115, 3193–3204. [Google Scholar] [CrossRef] [Green Version]
  230. Juang, Y.-T.; Rauen, T.; Wang, Y.; Ichinose, K.; Benedyk, K.; Tenbrock, K.; Tsokos, G.C. Transcriptional Activation of the cAMP-responsive Modulator Promoter in Human T Cells Is Regulated by Protein Phosphatase 2A-mediated Dephosphorylation of SP-1 and Reflects Disease Activity in Patients with Systemic Lupus Erythematosus. J. Biol. Chem. 2011, 286, 1795–1801. [Google Scholar] [CrossRef] [Green Version]
  231. Vassilopoulos, D.; Kovacs, B.; Tsokos, G.C. TCR/CD3 complex-mediated signal transduction pathway in T cells and T cell lines from patients with systemic lupus erythematosus. J. Immunol. 1995, 155, 2269–2281. [Google Scholar] [CrossRef]
  232. Moulton, V.R.; Tsokos, G.C. Abnormalities of T cell signaling in systemic lupus erythematosus. Arthritis Res. 2011, 13, 207. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  233. Kyttaris, V.C.; Wang, Y.; Juang, Y.-T.; Weinstein, A.; Tsokos, G.C. Increased Levels of NF-ATc2 Differentially Regulate CD154 and IL-2 Genes in T Cells from Patients with Systemic Lupus Erythematosus. J. Immunol. 2007, 178, 1960–1966. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  234. Karin, M.; Liu, Z.; Zandi, E. AP-1 function and regulation. Curr. Opin. Cell Biol. 1997, 9, 240–246. [Google Scholar] [CrossRef]
  235. Kyttaris, V.C.; Juang, Y.-T.; Tenbrock, K.; Weinstein, A.; Tsokos, G.C. Cyclic adenosine 5′-monophosphate response element modulator is responsible for the decreased expression of c-fos and activator protein-1 binding in T cells from patients with systemic lupus erythematosus. J. Immunol. 2004, 173, 3557–3563. [Google Scholar] [CrossRef] [Green Version]
  236. Foulkes, N.S.; Laoide, B.M.; Schlotter, F.; Sassone-Corsi, P. Transcriptional antagonist cAMP-responsive element modulator (CREM) down-regulates c-fos cAMP-induced expression. Proc. Natl. Acad. Sci. USA 1991, 88, 5448–5452. [Google Scholar] [CrossRef] [Green Version]
  237. Tenbrock, K.; Tsokos, G.C. Transcriptional Regulation of Interlekin 2 in Sle T Cells. Int. Rev. Immunol. 2004, 23, 333–345. [Google Scholar] [CrossRef] [PubMed]
  238. Michishita, E.; Park, J.Y.; Burneskis, J.M.; Barrett, J.C.; Horikawa, I. Evolutionarily conserved and nonconserved cellular localizations and functions of human SIRT proteins. Mol. Biol. Cell 2005, 16, 4623–4635. [Google Scholar] [CrossRef] [Green Version]
  239. Hisada, R.; Yoshida, N.; Umeda, M.; Burbano, C.; Bhargava, R.; Scherlinger, M.; Kono, M.; Kyttaris, V.C.; Krishfield, S.; Tsokos, G.C. The deacetylase SIRT2 contributes to autoimmune disease pathogenesis by modulating IL-17A and IL-2 transcription. Cell Mol. Immunol. 2022, 19, 738–750. [Google Scholar] [CrossRef]
  240. Berg, V.; Modak, M.; Brell, J.; Puck, A.; Künig, S.; Jutz, S.; Steinberger, P.; Zlabinger, G.J.; Stöckl, J. Iron Deprivation in Human T Cells Induces Nonproliferating Accessory Helper Cells. ImmunoHorizons 2020, 4, 165–177. [Google Scholar] [CrossRef] [Green Version]
  241. Zhang, F.; Zhou, X.; Cui, D.; Zhang, W.; Lai, J.; Li, X.; Ruan, Y.; Xie, Y.; Shi, M.; Xiao, Y.; et al. The role of Stim1 in the progression of lupus nephritis in mice. Int. J. Clin. Exp. Pathol. 2020, 13, 3021–3032. [Google Scholar] [PubMed]
  242. Shi, S.; Zhao, Q.; Ke, C.; Long, S.; Zhang, F.; Zhang, X.; Li, Y.; Liu, X.; Hu, H.; Yin, S. Loureirin B Exerts its Immunosuppressive Effects by Inhibiting STIM1/Orai1 and KV1.3 Channels. Front. Pharmacol. 2021, 12, 685092. [Google Scholar] [CrossRef]
  243. Nambiar, M.P.; Fisher, C.U.; Warke, V.G.; Krishnan, S.; Mitchell, J.P.; Delaney, N.; Tsokos, G.C. Reconstitution of deficient T cell receptor zeta chain restores T cell signaling and augments T cell receptor/CD3-induced interleukin-2 production in patients with systemic lupus erythematosus. Arthritis Rheum. 2003, 48, 1948–1955. [Google Scholar] [CrossRef] [PubMed]
  244. Harada, T.; Kyttaris, V.; Li, Y.; Juang, Y.-T.; Wang, Y.; Tsokos, G.C. Increased expression of STAT3 in SLE T cells contributes to enhanced chemokine-mediated cell migration. Autoimmunity 2007, 40, 1–8. [Google Scholar] [CrossRef] [PubMed]
  245. Laurence, A.; Tato, C.M.; Davidson, T.S.; Kanno, Y.; Chen, Z.; Yao, Z.; Blank, R.B.; Meylan, F.; Siegel, R.; Hennighausen, L.; et al. Interleukin-2 signaling via STAT5 constrains T helper 17 cell generation. Immunity 2007, 26, 371–381. [Google Scholar] [CrossRef] [Green Version]
  246. Wong, C.K.; Ho, C.Y.; Li, E.K.; Lam, C.W. Elevation of proinflammatory cytokine (IL-18, IL-17, IL-12) and Th2 cytokine (IL-4) concentrations in patients with systemic lupus erythematosus. Lupus 2000, 9, 589–593. [Google Scholar] [CrossRef]
  247. Crispín, J.C.; Oukka, M.; Bayliss, G.; Cohen, R.A.; Van Beek, C.A.; Stillman, I.E.; Kyttaris, V.C.; Juang, Y.-T.; Tsokos, G.C. Expanded double negative T cells in patients with systemic lupus erythematosus produce IL-17 and infiltrate the kidneys. J. Immunol. 2008, 181, 8761–8766. [Google Scholar] [CrossRef] [Green Version]
  248. Shah, K.; Lee, W.-W.; Lee, S.-H.; Kim, S.H.; Kang, S.W.; Craft, J.; Kang, I. Dysregulated balance of Th17 and Th1 cells in systemic lupus erythematosus. Arthritis Res. 2010, 12, R53. [Google Scholar] [CrossRef] [Green Version]
  249. Yang, J.; Chu, Y.; Yang, X.; Gao, D.; Zhu, L.; Yang, X.; Wan, L.; Li, M. Th17 and natural Treg cell population dynamics in systemic lupus erythematosus. Arthritis Rheum. 2009, 60, 1472–1483. [Google Scholar] [CrossRef]
  250. Schwarzenberger, P.; La Russa, V.; Miller, A.; Ye, P.; Huang, W.; Zieske, A.; Nelson, S.; Bagby, G.J.; Stoltz, D.; Mynatt, R.L.; et al. IL-17 stimulates granulopoiesis in mice: Use of an alternate, novel gene therapy-derived method for in vivo evaluation of cytokines. J. Immunol. 1998, 161, 6383–6389. [Google Scholar] [CrossRef]
  251. Schwarzenberger, P.; Huang, W.; Ye, P.; Oliver, P.; Manuel, M.; Zhang, Z.; Bagby, G.; Nelson, S.; Kolls, J.K. Requirement of endogenous stem cell factor and granulocyte-colony-stimulating factor for IL-17-mediated granulopoiesis. J. Immunol. 2000, 164, 4783–4789. [Google Scholar] [CrossRef] [Green Version]
  252. Schwarzenberger, P.; Huang, W.; Oliver, P.; Byrne, P.; La Russa, V.; Zhang, Z.; Kolls, J.K. IL-17 Mobilizes Peripheral Blood Stem Cells with Short- and Long-Term Repopulating Ability in Mice. J. Immunol. 2001, 167, 2081–2086. [Google Scholar] [CrossRef] [Green Version]
  253. Tan, W.; Huang, W.; Zhong, Q.; Schwarzenberger, P. IL-17 receptor knockout mice have enhanced myelotoxicity and impaired hemopoietic recovery following gamma irradiation. J. Immunol. 2006, 176, 6186–6193. [Google Scholar] [CrossRef] [Green Version]
  254. Von Vietinghoff, S.; Ley, K. IL-17A Controls IL-17F Production and Maintains Blood Neutrophil Counts in Mice. J. Immunol. 2009, 183, 865–873. [Google Scholar] [CrossRef] [Green Version]
  255. Mitsdoerffer, M.; Lee, Y.; Jäger, A.; Kim, H.-J.; Korn, T.; Kolls, J.K.; Cantor, H.; Bettelli, E.; Kuchroo, V.K. Proinflammatory T helper type 17 cells are effective B-cell helpers. Proc. Natl. Acad. Sci. USA 2010, 107, 14292–14297. [Google Scholar] [CrossRef] [Green Version]
  256. Crispín, J.C.; Liossis, S.-N.C.; Kis-Toth, K.; Lieberman, L.A.; Kyttaris, V.C.; Juang, Y.-T.; Tsokos, G.C. Pathogenesis of human systemic lupus erythematosus: Recent advances. Trends Mol. Med. 2010, 16, 47–57. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  257. Riol-Blanco, L.; Lazarevic, V.; Awasthi, A.; Mitsdoerffer, M.; Wilson, B.S.; Croxford, A.; Waisman, A.; Kuchroo, V.K.; Glimcher, L.H.; Oukka, M. IL-23 Receptor Regulates Unconventional IL-17–Producing T Cells That Control Bacterial Infections. J. Immunol. 2010, 184, 1710–1720. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  258. Park, H.; Li, Z.; Yang, X.O.; Chang, S.H.; Nurieva, R.; Wang, Y.-H.; Wang, Y.; Hood, L.; Zhu, Z.; Tian, Q.; et al. A distinct lineage of CD4 T cells regulates tissue inflammation by producing interleukin 17. Nat. Immunol. 2005, 6, 1133–1141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  259. Harrington, L.E.; Hatton, R.D.; Mangan, P.R.; Turner, H.; Murphy, T.L.; Murphy, K.M.; Weaver, C.T. Interleukin 17–producing CD4 + effector T cells develop via a lineage distinct from the T helper type 1 and 2 lineages. Nat. Immunol. 2005, 6, 1123–1132. [Google Scholar] [CrossRef]
  260. Crispín, J.C.; Tsokos, G.C. Human TCR-αβ+ CD4− CD8− T Cells Can Derive from CD8+ T Cells and Display an Inflammatory Effector Phenotype. J. Immunol. 2009, 183, 4675–4681. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  261. Tsokos, G.C. Systemic Lupus Erythematosus. N. Engl. J. Med. 2011, 365, 2110–2121. [Google Scholar] [CrossRef] [Green Version]
  262. Mangan, P.R.; Harrington, L.E.; O’Quinn, D.B.; Helms, W.S.; Bullard, D.C.; Elson, C.O.; Hatton, R.D.; Wahl, S.M.; Schoeb, T.R.; Weaver, C.T. Transforming growth factor-β induces development of the TH17 lineage. Nature 2006, 441, 231–234. [Google Scholar] [CrossRef] [PubMed]
  263. Yang, L.; Anderson, D.E.; Baecher-Allan, C.; Hastings, W.D.; Bettelli, E.; Oukka, M.; Kuchroo, V.K.; Hafler, D.A. IL-21 and TGF-β are required for differentiation of human TH17 cells. Nature 2008, 454, 350–352. [Google Scholar] [CrossRef] [Green Version]
  264. Iezzi, G.; Sonderegger, I.; Ampenberger, F.; Schmitz, N.; Marsland, B.J.; Kopf, M. CD40-CD40L cross-talk integrates strong antigenic signals and microbial stimuli to induce development of IL-17-producing CD4+ T cells. Proc. Natl. Acad. Sci. USA 2009, 106, 876–881. [Google Scholar] [CrossRef] [Green Version]
  265. Javierre, B.M.; Richardson, B. A new epigenetic challenge: Systemic lupus erythematosus. Adv. Exp. Med. Biol. 2011, 711, 117–136. [Google Scholar] [CrossRef]
  266. Ghodke-Puranik, Y.; Niewold, T.B. Immunogenetics of systemic lupus erythematosus: A comprehensive review. J. Autoimmun. 2015, 64, 125–136. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  267. Hirano, T. Interleukin 6 and its Receptor: Ten Years Later. Int. Rev. Immunol. 1998, 16, 249–284. [Google Scholar] [CrossRef] [PubMed]
  268. Harbour, S.N.; DiToro, D.F.; Witte, S.J.; Zindl, C.L.; Gao, M.; Schoeb, T.R.; Jones, G.W.; Jones, S.A.; Hatton, R.D.; Weaver, C.T. TH17 cells require ongoing classic IL-6 receptor signaling to retain transcriptional and functional identity. Sci. Immunol. 2020, 5, eaaw2262. [Google Scholar] [CrossRef]
  269. Chalaris, A.; Garbers, C.; Rabe, B.; Rose-John, S.; Scheller, J. The soluble Interleukin 6 receptor: Generation and role in inflammation and cancer. Eur. J. Cell Biol. 2011, 90, 484–494. [Google Scholar] [CrossRef]
  270. La Belle Flynn, A.; Calhoun, B.C.; Sharma, A.; Chang, J.C.; Almasan, A.; Schiemann, W.P. Autophagy inhibition elicits emergence from metastatic dormancy by inducing and stabilizing Pfkfb3 expression. Nat. Commun. 2019, 10, 3668. [Google Scholar] [CrossRef] [Green Version]
  271. Johnson, D.E.; O’Keefe, R.A.; Grandis, J.R. Targeting the IL-6/JAK/STAT3 signalling axis in cancer. Nat. Rev. Clin. Oncol. 2018, 15, 234–248. [Google Scholar] [CrossRef] [PubMed]
  272. Cimica, V.; Chen, H.-C.; Iyer, J.K.; Reich, N.C. Dynamics of the STAT3 transcription factor: Nuclear import dependent on Ran and importin-β1. PLoS ONE 2011, 6, e20188. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  273. Zhou, L.; Ivanov, I.I.; Spolski, R.; Min, R.; Shenderov, K.; Egawa, T.; Levy, D.E.; Leonard, W.J.; Littman, D.R. IL-6 programs T(H)-17 cell differentiation by promoting sequential engagement of the IL-21 and IL-23 pathways. Nat. Immunol. 2007, 8, 967–974. [Google Scholar] [CrossRef]
  274. McGeachy, M.J.; Chen, Y.; Tato, C.M.; Laurence, A.; Joyce-Shaikh, B.; Blumenschein, W.M.; McClanahan, T.K.; O’Shea, J.J.; Cua, D.J. The interleukin 23 receptor is essential for the terminal differentiation of interleukin 17-producing effector T helper cells in vivo. Nat. Immunol. 2009, 10, 314–324. [Google Scholar] [CrossRef] [Green Version]
  275. Koga, T.; Hedrich, C.M.; Mizui, M.; Yoshida, N.; Otomo, K.; Lieberman, L.A.; Rauen, T.; Crispín, J.C.; Tsokos, G.C. CaMK4-dependent activation of AKT/mTOR and CREM-α underlies autoimmunity-associated Th17 imbalance. J. Clin. Investig. 2014, 124, 2234–2245. [Google Scholar] [CrossRef] [Green Version]
  276. Katsuyama, T.; Tsokos, G.C.; Moulton, V.R. Aberrant T Cell Signaling and Subsets in Systemic Lupus Erythematosus. Front. Immunol. 2018, 9, 1088. [Google Scholar] [CrossRef] [Green Version]
  277. Hedrich, C.M.; Crispin, J.C.; Rauen, T.; Ioannidis, C.; Apostolidis, S.A.; Lo, M.S.; Kyttaris, V.C.; Tsokos, G.C. cAMP response element modulator α controls IL2 and IL17A expression during CD4 lineage commitment and subset distribution in lupus. Proc. Natl. Acad. Sci. USA 2012, 109, 16606–16611. [Google Scholar] [CrossRef] [Green Version]
  278. Richardson, B. Primer: Epigenetics of autoimmunity. Nat. Clin. Pract. Rheumatol. 2007, 3, 521–527. [Google Scholar] [CrossRef] [PubMed]
  279. Ulrey, C.L.; Liu, L.; Andrews, L.G.; Tollefsbol, T.O. The impact of metabolism on DNA methylation. Hum. Mol. Genet. 2005, 14, R139–R147. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  280. Laplante, M.; Sabatini, D.M. mTOR signaling in growth control and disease. Cell 2012, 149, 274–293. [Google Scholar] [CrossRef] [Green Version]
  281. Zoncu, R.; Efeyan, A.; Sabatini, D.M. mTOR: From growth signal integration to cancer, diabetes and ageing. Nat. Rev. Mol. Cell Biol. 2011, 12, 21–35. [Google Scholar] [CrossRef] [Green Version]
  282. Sarbassov, D.D.; Guertin, D.A.; Ali, S.M.; Sabatini, D.M. Phosphorylation and regulation of Akt/PKB by the rictor-mTOR complex. Science 2005, 307, 1098–1101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  283. Kurebayashi, Y.; Nagai, S.; Ikejiri, A.; Ohtani, M.; Ichiyama, K.; Baba, Y.; Yamada, T.; Egami, S.; Hoshii, T.; Hirao, A.; et al. PI3K-Akt-mTORC1-S6K1/2 Axis Controls Th17 Differentiation by Regulating Gfi1 Expression and Nuclear Translocation of RORγ. Cell Rep. 2012, 1, 360–373. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  284. Delgoffe, G.M.; Pollizzi, K.N.; Waickman, A.T.; Heikamp, E.; Meyers, D.J.; Horton, M.R.; Xiao, B.; Worley, P.F.; Powell, J.D. The kinase mTOR regulates the differentiation of helper T cells through the selective activation of signaling by mTORC1 and mTORC2. Nat. Immunol. 2011, 12, 295–303. [Google Scholar] [CrossRef] [Green Version]
  285. Sinclair, L.V.; Rolf, J.; Emslie, E.; Shi, Y.-B.; Taylor, P.M.; Cantrell, D.A. Control of amino-acid transport by antigen receptors coordinates the metabolic reprogramming essential for T cell differentiation. Nat. Immunol. 2013, 14, 500–508. [Google Scholar] [CrossRef] [Green Version]
  286. Fernandez, D.; Bonilla, E.; Mirza, N.; Niland, B.; Perl, A. Rapamycin reduces disease activity and normalizes T cell activation-induced calcium fluxing in patients with systemic lupus erythematosus. Arthiritis Rhem. 2006, 54, 2983–2988. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  287. Chu, Y.; Zhao, C.; Zhang, B.; Wang, X.; Wang, Y.; An, J.; Chen, J. Restoring T-helper 17 cell/regulatory T-cell balance and decreasing disease activity by rapamycin and all-trans retinoic acid in patients with systemic lupus erythematosus. Lupus 2019, 28, 1397–1406. [Google Scholar] [CrossRef] [PubMed]
  288. Singh, R. Central role of PI3K–SYK interaction in fibrinogen-induced lamellipodia and filopodia formation in platelets. FEBS Open. Bio 2016, 6, 1285–1296. [Google Scholar] [CrossRef]
  289. Simioni, C.; Martelli, A.M.; Zauli, G.; Vitale, M.; McCubrey, J.A.; Capitani, S.; Neri, L.M. Targeting the phosphatidylinositol 3-kinase/Akt/mechanistic target of rapamycin signaling pathway in B-lineage acute lymphoblastic leukemia: An update. J. Cell. Physiol. 2018, 233, 6440–6454. [Google Scholar] [CrossRef] [Green Version]
  290. Szymonowicz, K.; Oeck, S.; Malewicz, N.M.; Jendrossek, V. New Insights into Protein Kinase B/Akt Signaling: Role of Localized Akt Activation and Compartment-Specific Target Proteins for the Cellular Radiation Response. Cancers 2018, 10, 78. [Google Scholar] [CrossRef] [Green Version]
  291. Henske, E.P.; Jóźwiak, S.; Kingswood, J.C.; Sampson, J.R.; Thiele, E.A. Tuberous sclerosis complex. Nat. Rev. Dis. Prim. 2016, 2, 16035. [Google Scholar] [CrossRef] [PubMed]
  292. Kato, H.; Perl, A. Mechanistic Target of Rapamycin Complex 1 Expands Th17 and IL-4+ CD4 CD8 Double-Negative T Cells and Contracts Regulatory T Cells in Systemic Lupus Erythematosus. J. Immunol. 2014, 192, 4134–4144. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  293. Hay, N.; Sonenberg, N. Upstream and downstream of mTOR. Genes Dev. 2004, 18, 1926–1945. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  294. Sridharan, S.; Basu, A. Distinct Roles of mTOR Targets S6K1 and S6K2 in Breast Cancer. Int. J. Mol. Sci. 2020, 21, 1199. [Google Scholar] [CrossRef] [Green Version]
  295. GFI1 Growth Factor Independent 1 Transcriptional Repressor [Homo Sapiens (Human)]—Gene—NCBI. Available online: https://www.ncbi.nlm.nih.gov/gene/2672 (accessed on 1 June 2021).
  296. Ichiyama, K.; Hashimoto, M.; Sekiya, T.; Nakagawa, R.; Wakabayashi, Y.; Sugiyama, Y.; Komai, K.; Saba, I.; Moroy, T.; Yoshimura, A. Gfi1 negatively regulates Th17 differentiation by inhibiting ROR t activity. Int. Immunol. 2009, 21, 881–889. [Google Scholar] [CrossRef] [Green Version]
  297. EGR2 Early Growth Response 2 [Homo Sapiens (Human)]—Gene—NCBI. Available online: https://www.ncbi.nlm.nih.gov/gene/1959 (accessed on 1 June 2021).
  298. Ivanov, I.I.; McKenzie, B.S.; Zhou, L.; Tadokoro, C.E.; Lepelley, A.; Lafaille, J.J.; Cua, D.J.; Littman, D.R. The orphan nuclear receptor RORgammat directs the differentiation program of proinflammatory IL-17+ T helper cells. Cell 2006, 126, 1121–1133. [Google Scholar] [CrossRef] [Green Version]
  299. Kono, M. New insights into the metabolism of Th17 cells. Immunol. Med. 2022, 46, 15–24. [Google Scholar] [CrossRef] [PubMed]
  300. Damasceno, L.E.A.; Prado, D.S.; Veras, F.P.; Fonseca, M.M.; Toller-Kawahisa, J.E.; Rosa, M.H.; Públio, G.A.; Martins, T.V.; Ramalho, F.S.; Waisman, A.; et al. PKM2 promotes Th17 cell differentiation and autoimmune inflammation by fine-tuning STAT3 activation. J. Exp. Med. 2020, 217, e20190613. [Google Scholar] [CrossRef] [PubMed]
  301. Kono, M.; Maeda, K.; Stocton-Gavanescu, I.; Pan, W.; Umeda, M.; Katsuyama, E.; Burbano, C.; Orite, S.Y.K.; Vukelic, M.; Tsokos, M.G.; et al. Pyruvate kinase M2 is requisite for Th1 and Th17 differentiation. JCI Insight 2019, 4, e127395. [Google Scholar] [CrossRef] [Green Version]
  302. Yoshida, N.; Comte, D.; Mizui, M.; Otomo, K.; Rosetti, F.; Mayadas, T.N.; Crispín, J.C.; Bradley, S.J.; Koga, T.; Kono, M.; et al. ICER is requisite for Th17 differentiation. Nat. Commun. 2016, 7, 12993. [Google Scholar] [CrossRef] [Green Version]
  303. Dang, E.V.; Barbi, J.; Yang, H.-Y.; Jinasena, D.; Yu, H.; Zheng, Y.; Bordman, Z.; Fu, J.; Kim, Y.; Yen, H.-R.; et al. Control of T(H)17/T(reg) balance by hypoxia-inducible factor 1. Cell 2011, 146, 772–784. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  304. Kim, J.; Tchernyshyov, I.; Semenza, G.L.; Dang, C.V. HIF-1-mediated expression of pyruvate dehydrogenase kinase: A metabolic switch required for cellular adaptation to hypoxia. Cell Metab. 2006, 3, 177–185. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  305. Liu, H.; Ma, Y.; Cole, S.M.; Zander, C.; Chen, K.-H.; Karras, J.; Pope, R.M. Serine phosphorylation of STAT3 is essential for Mcl-1 expression and macrophage survival. Blood 2003, 102, 344–352. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  306. Calò, V.; Migliavacca, M.; Bazan, V.; Macaluso, M.; Buscemi, M.; Gebbia, N.; Russo, A. STAT proteins: From normal control of cellular events to tumorigenesis. J. Cell Physiol. 2003, 197, 157–168. [Google Scholar] [CrossRef]
  307. Wen, Z.; Zhong, Z.; Darnell, J.E. Maximal activation of transcription by Stat1 and Stat3 requires both tyrosine and serine phosphorylation. Cell 1995, 82, 241–250. [Google Scholar] [CrossRef] [Green Version]
  308. Decker, T.; Kovarik, P. Serine phosphorylation of STATs. Oncogene 2000, 19, 2628–2637. [Google Scholar] [CrossRef] [Green Version]
  309. Kim, J.-H.; Yoon, M.-S.; Chen, J. Signal transducer and activator of transcription 3 (STAT3) mediates amino acid inhibition of insulin signaling through serine 727 phosphorylation. J. Biol. Chem. 2009, 284, 35425–35432. [Google Scholar] [CrossRef] [Green Version]
  310. Kshirsagar, S.; Riedl, M.; Billing, H.; Tönshoff, B.; Thangavadivel, S.; Steuber, C.; Staude, H.; Wechselberger, G.; Edelbauer, M. Akt-dependent enhanced migratory capacity of Th17 cells from children with lupus nephritis. J. Immunol. 2014, 193, 4895–4903. [Google Scholar] [CrossRef] [Green Version]
  311. Sequeira, J.; Boily, G.; Bazinet, S.; Saliba, S.; He, X.; Jardine, K.; Kennedy, C.; Staines, W.; Rousseaux, C.; Mueller, R.; et al. sirt1-null mice develop an autoimmune-like condition. Exp. Cell Res. 2008, 314, 3069–3074. [Google Scholar] [CrossRef]
  312. Nie, Y.; Erion, D.M.; Yuan, Z.; Dietrich, M.; Shulman, G.I.; Horvath, T.L.; Gao, Q. STAT3 inhibition of gluconeogenesis is downregulated by SirT1. Nat. Cell Biol. 2009, 11, 492–500. [Google Scholar] [CrossRef] [Green Version]
  313. Yuan, Z.; Guan, Y.; Chatterjee, D.; Chin, Y.E. Stat3 Dimerization Regulated by Reversible Acetylation of a Single Lysine Residue. Science 2005, 307, 269–273. [Google Scholar] [CrossRef] [PubMed]
  314. Limagne, E.; Thibaudin, M.; Euvrard, R.; Berger, H.; Chalons, P.; Végan, F.; Humblin, E.; Boidot, R.; Rébé, C.; Derangère, V.; et al. Sirtuin-1 Activation Controls Tumor Growth by Impeding Th17 Differentiation via STAT3 Deacetylation. Cell Rep. 2017, 19, 746–759. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  315. Wang, W.; Hu, Y.; Yang, C.; Zhu, S.; Wang, X.; Zhang, Z.; Deng, H. Decreased NAD Activates STAT3 and Integrin Pathways to Drive Epithelial-Mesenchymal Transition. Mol. Cell. Proteom. 2018, 17, 2005–2017. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  316. Suarez-Alvarez, B.; Rodriguez, R.M.; Fraga, M.F.; López-Larrea, C. DNA methylation: A promising landscape for immune system-related diseases. Trends Genet. 2012, 28, 506–514. [Google Scholar] [CrossRef] [PubMed]
  317. Tsagaratou, A.; Lio, C.-W.J.; Yue, X.; Rao, A. TET Methylcytosine Oxidases in T Cell and B Cell Development and Function. Front. Immunol. 2017, 8, 220. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  318. Ballestar, E.; Sawalha, A.H.; Lu, Q. Clinical value of DNA methylation markers in autoimmune rheumatic diseases. Nat. Rev. Rheumatol. 2020, 16, 514–524. [Google Scholar] [CrossRef]
  319. Sunahori, K.; Nagpal, K.; Hedrich, C.M.; Mizui, M.; Fitzgerald, L.M.; Tsokos, G.C. The catalytic subunit of protein phosphatase 2A (PP2Ac) promotes DNA hypomethylation by suppressing the phosphorylated mitogen-activated protein kinase/extracellular signal-regulated kinase (ERK) kinase (MEK)/phosphorylated ERK/DNMT1 protein pathway in T-cells from controls and systemic lupus erythematosus patients. J. Biol. Chem. 2013, 288, 21936–21944. [Google Scholar] [CrossRef] [Green Version]
  320. Tamanaha, E.; Guan, S.; Marks, K.; Saleh, L. Distributive Processing by the Iron(II)/α-Ketoglutarate-Dependent Catalytic Domains of the TET Enzymes Is Consistent with Epigenetic Roles for Oxidized 5-Methylcytosine Bases. J. Am. Chem. Soc. 2016, 138, 9345–9348. [Google Scholar] [CrossRef] [Green Version]
  321. Izmirly, P.M.; Parton, H.; Wang, L.; McCune, W.J.; Lim, S.S.; Drenkard, C.; Ferucci, E.D.; Dall’Era, M.; Gordon, C.; Helmick, C.G.; et al. Prevalence of Systemic Lupus Erythematosus in the United States: Estimates From a Meta-Analysis of the Centers for Disease Control and Prevention National Lupus Registries. Arthritis Rheumatol. 2021, 73, 991–996. [Google Scholar] [CrossRef]
  322. Hughes, T.; Adler, A.; Merrill, J.T.; Kelly, J.A.; Kaufman, K.M.; Williams, A.; Langefeld, C.D.; Gilkeson, G.S.; Sanchez, E.; Martin, J.; et al. Analysis of autosomal genes reveals gene–sex interactions and higher total genetic risk in men with systemic lupus erythematosus. Ann. Rheum. Dis. 2012, 71, 694–699. [Google Scholar] [CrossRef]
  323. Sawalha, A.H.; Wang, L.; Nadig, A.; Somers, E.C.; McCune, W.J.; Michigan Lupus Cohort; Hughes, T.; Merrill, J.T.; Scofield, R.H.; Strickland, F.M.; et al. Sex-specific differences in the relationship between genetic susceptibility, T cell DNA demethylation and lupus flare severity. J. Autoimmun. 2012, 38, J216–J222. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  324. Ali, M.; Coit, P.; Sawalha, A.H. Sex-based comparison of CD4+ T cell DNA methylation in lupus reveals proinflammatory epigenetic changes in men. Clin. Immunol. 2022, 243, 109116. [Google Scholar] [CrossRef] [PubMed]
  325. Hao, S.; Baltimore, D. The stability of mRNA influences the temporal order of the induction of genes encoding inflammatory molecules. Nat. Immunol. 2009, 10, 281–288. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  326. Kafasla, P.; Skliris, A.; Kontoyiannis, D.L. Post-transcriptional coordination of immunological responses by RNA-binding proteins. Nat. Immunol. 2014, 15, 492–502. [Google Scholar] [CrossRef]
  327. Rothamel, K.; Arcos, S.; Kim, B.; Reasoner, C.; Lisy, S.; Mukherjee, N.; Ascano, M. ELAVL1 primarily couples mRNA stability with the 3′ UTRs of interferon-stimulated genes. Cell Rep. 2021, 35, 109178. [Google Scholar] [CrossRef]
  328. Mancini, M.; Magnani, E.; Macchi, F.; Bonapace, I.M. The multi-functionality of UHRF1: Epigenome maintenance and preservation of genome integrity. Nucleic Acids Res. 2021, 49, 6053–6068. [Google Scholar] [CrossRef]
  329. Obata, Y.; Furusawa, Y.; Endo, T.A.; Sharif, J.; Takahashi, D.; Atarashi, K.; Nakayama, M.; Onawa, S.; Fujimura, Y.; Takahashi, M.; et al. The epigenetic regulator Uhrf1 facilitates the proliferation and maturation of colonic regulatory T cells. Nat. Immunol. 2014, 15, 571–579. [Google Scholar] [CrossRef]
  330. Liu, L.; Hu, L.; Yang, L.; Jia, S.; Du, P.; Min, X.; Wu, J.; Wu, H.; Long, H.; Lu, Q.; et al. UHRF1 downregulation promotes T follicular helper cell differentiation by increasing BCL6 expression in SLE. Clin. Epigenetics 2021, 13, 31. [Google Scholar] [CrossRef]
  331. Lee, K.K.; Workman, J.L. Histone acetyltransferase complexes: One size doesn’t fit all. Nat. Rev. Mol. Cell Biol. 2007, 8, 284–295. [Google Scholar] [CrossRef]
  332. Hu, N.; Qiu, X.; Luo, Y.; Yuan, J.; Li, Y.; Lei, W.; Zhang, G.; Zhou, Y.; Su, Y.; Lu, Q. Abnormal histone modification patterns in lupus CD4+ T cells. J. Rheumatol. 2008, 35, 804–810. [Google Scholar]
  333. Mishra, N.; Reilly, C.M.; Brown, D.R.; Ruiz, P.; Gilkeson, G.S. Histone deacetylase inhibitors modulate renal disease in the MRL-lpr/lpr mouse. J. Clin. Investig. 2003, 111, 539–552. [Google Scholar] [CrossRef] [PubMed]
  334. Lu, K.T.; Kanno, Y.; Cannons, J.L.; Handon, R.; Bible, P.; Elkahloun, A.G.; Anderson, S.M.; Wei, L.; Sun, H.; O’Shea, J.J.; et al. Functional and epigenetic studies reveal multistep differentiation and plasticity of in vitro-generated and in vivo-derived follicular T helper cells. Immunity 2011, 35, 622–632. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  335. Poholek, A.C.; Hansen, K.; Hernandez, S.G.; Eto, D.; Chandele, A.; Weinstein, J.S.; Dong, X.; Odegard, J.M.; Kaech, S.M.; Dent, A.L.; et al. In vivo regulation of Bcl6 and T follicular helper cell development. J. Immunol. 2010, 185, 313–326. [Google Scholar] [CrossRef] [Green Version]
  336. Nakayamada, S.; Poholek, A.C.; Lu, K.T.; Takahashi, H.; Kato, M.; Iwata, S.; Hirahara, K.; Cannons, J.L.; Schwartzberg, P.L.; Vahedi, G.; et al. Type I Interferon induces binding of STAT1 to Bcl6: Divergent Roles of STAT-family transcription factors in the TFH cell genetic program. J. Immunol. 2014, 192, 2156–2166. [Google Scholar] [CrossRef] [Green Version]
  337. Dong, X.; Antao, O.Q.; Song, W.; Sanchez, G.M.; Zembrzuski, K.; Koumpouras, F.; Lemenze, A.; Craft, J.; Weinstein, J.S. Type 1 interferon activated STAT4 regulates T follicular helper (Tfh)-cell dependent cytokine and immunoglobulin production in lupus. Arthritis Rheumatol. 2021, 73, 478–489. [Google Scholar] [CrossRef] [PubMed]
  338. Girard, E.; Chmiest, D.; Fournier, N.; Johannes, L.; Paul, J.-L.; Vedie, B.; Lamaze, C. Rab7 Is Functionally Required for Selective Cargo Sorting at the Early Endosome. Traffic 2014, 15, 309–326. [Google Scholar] [CrossRef]
  339. Burki, T.K. FDA approval for anifrolumab in patients with lupus. Lancet Rheumatol. 2021, 3, e689. [Google Scholar] [CrossRef]
  340. Bayeva, M.; Khechaduri, A.; Puig, S.; Chang, H.-C.; Patial, S.; Blackshear, P.J.; Ardehali, H. mTOR Regulates Cellular Iron Homeostasis through Tristetraprolin. Cell Metab. 2012, 16, 645–657. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Endosomal trafficking and recycling pathways are illustrated. Cell membrane proteins are endocytosed, then the internalized cargoes converge into a common early endosome. There are two pathways of recycling back to the plasma membrane: (1) the rapid-recycling pathway, directly from the early endosome, and (2) the slow-recycling pathway, via the recycling endosome. The cargoes can be released out of the cell as exosomes through the late endosome or the multi-vesicular body. Some cargoes from endosomes are routed to the lysosome for degradation. Rab GTPases play important roles in receptor recycling, including endocytosis, intracellular vesicle trafficking, and exosome secretion. Key Rab family proteins are highlighted.
Figure 1. Endosomal trafficking and recycling pathways are illustrated. Cell membrane proteins are endocytosed, then the internalized cargoes converge into a common early endosome. There are two pathways of recycling back to the plasma membrane: (1) the rapid-recycling pathway, directly from the early endosome, and (2) the slow-recycling pathway, via the recycling endosome. The cargoes can be released out of the cell as exosomes through the late endosome or the multi-vesicular body. Some cargoes from endosomes are routed to the lysosome for degradation. Rab GTPases play important roles in receptor recycling, including endocytosis, intracellular vesicle trafficking, and exosome secretion. Key Rab family proteins are highlighted.
Ijms 24 10749 g001
Figure 2. Metabolic abnormalities in SLE T cells and how they contribute to the signaling pathways that control IL-2 and IL-17 production. CD71, which controls the iron flux, is increased in SLE T cells. In place of the CD3ζ protein, FcεRIγ is substituted. In SLE T cells, FcεRIγ recruits Syk instead of ZAP-70. FcεRIγ–Syk interaction is significantly stronger than CD3ζ–ZAP-70 interaction, resulting in higher calcium influx. CD38, which is increased in SLE T cells, converts NAD+ into cADPR, which mediates calcium-mobilizing activity. GLUT1, a glucose transporter, and ASCT2, a glutamine transporter, are increased in SLE T cells, and the resulting glycolysis, glutaminolysis, and the TCA cycle are shown. The effects of these metabolites on pathways that affect IL-2 and IL-17 production are shown. Red arrows, increased in SLE T cells. Blue arrows, decreased in SLE T cells.
Figure 2. Metabolic abnormalities in SLE T cells and how they contribute to the signaling pathways that control IL-2 and IL-17 production. CD71, which controls the iron flux, is increased in SLE T cells. In place of the CD3ζ protein, FcεRIγ is substituted. In SLE T cells, FcεRIγ recruits Syk instead of ZAP-70. FcεRIγ–Syk interaction is significantly stronger than CD3ζ–ZAP-70 interaction, resulting in higher calcium influx. CD38, which is increased in SLE T cells, converts NAD+ into cADPR, which mediates calcium-mobilizing activity. GLUT1, a glucose transporter, and ASCT2, a glutamine transporter, are increased in SLE T cells, and the resulting glycolysis, glutaminolysis, and the TCA cycle are shown. The effects of these metabolites on pathways that affect IL-2 and IL-17 production are shown. Red arrows, increased in SLE T cells. Blue arrows, decreased in SLE T cells.
Ijms 24 10749 g002
Figure 3. Receptors that contribute to NAD+ synthesis pathways in T cells and how NAD+ affects IL-2 and IL-17 production. NAD+ synthesis pathways shown are: (1) de novo pathway, (2) Preiss–Handler pathway, (3) nucleoside pathway, and (4) salvage pathway. The kynurenine transported by System L amino acid transporters, which are heterodimers of CD98 and LAT1 contribute to the de novo pathway. CD73 has dual enzymatic functions, (1) cleaving NAD+ to NMN and (2) hydrolyzing NMN to NR, which both contribute to the nucleoside pathway after being imported into the cell by the nicotinamide mononucleotide transporter, Slc12a8 and equilibrative nucleoside transporter (ENT), respectively. CD38 and CD157 are ADP-ribosyl cyclases, converting NAD+ into cADPR and hydrolyzing cADPR to form ADPR. The NAD-dependent histone deacetylases, sirtuins, and their effects on IL-2 and IL-17 production are shown. Red arrows, increased in SLE T cells. Blue arrows, decreased in SLE T cells.
Figure 3. Receptors that contribute to NAD+ synthesis pathways in T cells and how NAD+ affects IL-2 and IL-17 production. NAD+ synthesis pathways shown are: (1) de novo pathway, (2) Preiss–Handler pathway, (3) nucleoside pathway, and (4) salvage pathway. The kynurenine transported by System L amino acid transporters, which are heterodimers of CD98 and LAT1 contribute to the de novo pathway. CD73 has dual enzymatic functions, (1) cleaving NAD+ to NMN and (2) hydrolyzing NMN to NR, which both contribute to the nucleoside pathway after being imported into the cell by the nicotinamide mononucleotide transporter, Slc12a8 and equilibrative nucleoside transporter (ENT), respectively. CD38 and CD157 are ADP-ribosyl cyclases, converting NAD+ into cADPR and hydrolyzing cADPR to form ADPR. The NAD-dependent histone deacetylases, sirtuins, and their effects on IL-2 and IL-17 production are shown. Red arrows, increased in SLE T cells. Blue arrows, decreased in SLE T cells.
Ijms 24 10749 g003
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Park, J.S.; Perl, A. Endosome Traffic Modulates Pro-Inflammatory Signal Transduction in CD4+ T Cells—Implications for the Pathogenesis of Systemic Lupus Erythematosus. Int. J. Mol. Sci. 2023, 24, 10749. https://doi.org/10.3390/ijms241310749

AMA Style

Park JS, Perl A. Endosome Traffic Modulates Pro-Inflammatory Signal Transduction in CD4+ T Cells—Implications for the Pathogenesis of Systemic Lupus Erythematosus. International Journal of Molecular Sciences. 2023; 24(13):10749. https://doi.org/10.3390/ijms241310749

Chicago/Turabian Style

Park, Joy S., and Andras Perl. 2023. "Endosome Traffic Modulates Pro-Inflammatory Signal Transduction in CD4+ T Cells—Implications for the Pathogenesis of Systemic Lupus Erythematosus" International Journal of Molecular Sciences 24, no. 13: 10749. https://doi.org/10.3390/ijms241310749

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop