Next Article in Journal
Cognitive Decline in Early and Premature Menopause
Next Article in Special Issue
Preclinical Characterization of the 177Lu-Labeled Prostate Stem Cell Antigen (PSCA)-Specific Monoclonal Antibody 7F5
Previous Article in Journal
Cadmium Exposure Is Associated with Behavioral Deficits and Neuroimmune Dysfunction in BTBR T+ Itpr3tf/J Mice
Previous Article in Special Issue
Combining Radiation- with Immunotherapy in Prostate Cancer: Influence of Radiation on T Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advances in the Pathogenesis and Treatment of Systemic Lupus Erythematosus

1
Division of Clinical Immunology, Department of Clinical, Anesthesiologic and Cardiovascular Sciences, Sapienza University of Rome, 00185 Rome, Italy
2
Eye Clinic, Department of Sense Organs, Sapienza University of Rome, 00185 Rome, Italy
3
Post Graduate School of Public Health, University of Siena, 53100 Siena, Italy
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(7), 6578; https://doi.org/10.3390/ijms24076578
Submission received: 19 February 2023 / Revised: 24 March 2023 / Accepted: 28 March 2023 / Published: 31 March 2023

Abstract

:
Systemic lupus erythematosus (SLE) is a genetically predisposed, female-predominant disease, characterized by multiple organ damage, that in its most severe forms can be life-threatening. The pathogenesis of SLE is complex and involves cells of both innate and adaptive immunity. The distinguishing feature of SLE is the production of autoantibodies, with the formation of immune complexes that precipitate at the vascular level, causing organ damage. Although progress in understanding the pathogenesis of SLE has been slower than in other rheumatic diseases, new knowledge has recently led to the development of effective targeted therapies, that hold out hope for personalized therapy. However, the new drugs available to date are still an adjunct to conventional therapy, which is known to be toxic in the short and long term. The purpose of this review is to summarize recent advances in understanding the pathogenesis of the disease and discuss the results obtained from the use of new targeted drugs, with a look at future therapies that may be used in the absence of the current standard of care or may even cure this serious systemic autoimmune disease.

1. Introduction

Systemic lupus erythematosus (SLE) is an autoimmune disease, characterized by phases of flare-ups and remission, that can cause severe damage to many organs and tissues. The organs most affected by SLE are the kidneys, nervous system, joints, and skin. The hallmark of SLE is the production of circulating autoantibodies, with the formation of immune complexes that precipitate in vessels, with activation of potent inflammatory responses ultimately responsible for multi-organ damage [1,2]. Fortunately, in recent decades, the pathogenesis of SLE has been greatly elucidated, with the identification of dysregulation of cells of the innate and adaptive systems. It has also been shown that a predominant role is played by type-I interferon (IFN), which is responsible for the hyperactivation of genes coding for pro-inflammatory molecules by target cells, a phenomenon termed type-I IFN signature [3]. Although SLE therapy is still based on nonspecific immunomodulatory and immunosuppressive drugs [4], new treatments directed against specific targets of the immune system have recently been developed, and some have been approved by regulatory agencies [5]. Such new drugs, however, still need to be combined with conventional therapy to achieve acceptable disease control. In this review, we will discuss new advances in our understanding of the mechanisms underlying the pathogenesis of SLE, the therapeutic potential of available targeted drugs, and the future development of novel therapeutic strategies that will hopefully lead to safe personalized therapy, possibly avoiding combination with conventional therapy.

2. Epidemiology

In recent decades, the incidence and prevalence of SLE have increased in all countries. This increase in prevalence has been attributed to better diagnosis and the availability of data from international registries. It is currently estimated that the incidence of overall SLE ranges from 0.3 to 23.2 cases/100,000 person-years, depending on the geographical region considered [6]. The highest incidence has been reported in North America, while countries in sub-Saharan Africa, Europe, and Australia have a lower incidence. These differences have been attributed to different genetic predisposition [7], and socioeconomic [8] and environmental factors [9]. Women of reproductive age are more prone to SLE than males, with the incidence ratio between females and males varying between 8:1 and 15:1 [9]. African American populations have the highest SLE incidence and mortality, followed by Hispanic and Asian populations, while Caucasian populations have the lowest rates of the disease [10]. On the other hand, it has been observed that African populations are more susceptible to SLE and are more resistant to systemic treatment with corticosteroids and immunosuppressants [11]. In total, it has been calculated that the risk of mortality in SLE patients is increased about 2.6-fold compared with the general population [12]. Delayed diagnosis, renal involvement, high disease activity index, presence of infections, and major cardiovascular events were found to be the main predictors of mortality [11].

3. Diagnosis, Management, and Activity Criteria for SLE

At least one positive antinuclear antibody (ANA) test is required as an entry criterion for SLE, according to the most recently published classification from the European Alliance of Associations for Rheumatology (EULAR)/American College of Rheumatology (ACR), in 2019 [5]. However, the presence of ANAs is not exclusive to SLE. They can be found in healthy individuals and patients with other autoimmune and nonautoimmune diseases [13,14,15]. Moreover, about 30% of patients with a clinical diagnosis of SLE are ANA-negative [16]. The anti-dsDNA antibody is used for the diagnosis of SLE. This antibody is also strongly associated with disease activity [17]. Anti-extractable nuclear antigen (ENA) antibodies are more specific than ANA for the diagnosis of SLE. In particular, the anti-Sm antibody is a marker associated with SLE. Often, it is associated with the presence of anti-U1-ribonucleoprotein (U1-RNP) antibodies, because both bind small nuclear ribonucleoprotein (snRNP). The presence of such high titer antibodies is found in the serum of patients with mixed connective tissue disease [18]. The characteristic antibodies of Sjögren’s syndrome, anti-SSA and anti-SSB, are found in 24–60% of patients with SLE, and are associated with neonatal SLE [19]. Anti-histone antibodies are associated with drug-induced lupus, while anti-ribosomal antibodies are associated with lupus. Antiphospholipid antibodies, which include lupus anticoagulant, anti-cardiolipin antibodies, and anti-β2 glycoprotein 1 antibody, are important indicators of vascular inflammation and thromboembolic risk. Antiphospholipid antibodies can also be observed in primary antiphospholipid syndrome not associated with SLE, or other inflammatory rheumatic conditions. Such antibodies are associated with recurrent early abortions, thrombosis, and intrauterine livedoid vasculitis, as well as neuropsychiatric and cardiovascular complications [20,21]. In addition, the diagnosis of SLE requires at least ten additive points, accumulated from seven clinical and three immunological domains. The new criteria have been found to have a sensitivity of 96.1% and a specificity of 93.4% [22]. Periodic follow-up and evaluation are necessary for the long-term management of patients after the diagnosis of the disease. Given the complex multisystem features of SLE, accurate measurement of the disease activity remains a challenge [23]. The most widely used tool is the SLE Disease Activity Index-2K (SLEDAI-2K). Patients are therefore classified as having SLE, from mild to severe, depending on a score ranging from ≤ 6 to 12 [24]. The 2004 British Isles Lupus Activity Group (BILAG) index, in eight organ systems, has been found to provide a more comprehensive systems-based measure than SLEDAI-2K [25]. For organ damage assessment, the Systemic Lupus International Collaborating Clinics (SLICC)/ACR damage index (SDI) is an internationally recognized tool [26]. Another disease assessment index commonly used in clinical trials, is the SLE Responder Index (SRI), which integrates criteria from the Safety of Estrogens in Lupus Erythematosus National Assessment (SELENA)-SLEDAI, Physician Global Assessment (PGA), and BILAG 2004 [27]. Figure 1 summarizes the latest classification criteria for the diagnosis of SLE and how disease activity and organ damage are measured.

4. SLE Pathogenesis

4.1. The Role of Adaptive Immunity

4.1.1. B Cells and Autoantibodies in SLE

B lymphocytes are characterized by the expression on their membrane of the B-cell receptor (BCR). This receptor is physiologically devoted to the recognition of pathogens and the subsequent production of specific antibodies [28]. During the process of B-cell development, autoreactive B cells can also be generated. Although the development of these host-dangerous cells is controlled by immunological tolerance systems, such as clonal deletion or induction of peripheral anergy, these control mechanisms can sometimes fail. This allows the unwanted expansion and activation of such autoreactive B cells, with the possible onset of autoimmune diseases [29,30,31]. After development, B cells, including self-reactive B cells, require the intervention of soluble factors to ensure their survival and proliferation. Among these, the most important is the B-cell activating factor (BAFF), also known as B lymphocyte stimulator (BLys) [32,33]. The repertoire of autoantibodies produced by autoreactive B cells, targets mainly nuclear antigens. A key role in the generation of these autoantibodies is played by toll-like receptors (TLRs). Abnormal engagement of TLRs TLR7 and TLR9 subtypes in SLE, has been shown to effectively promote the production of autoantibodies against double-stranded DNA (dsDNA) and RNA-associated autoantigens, respectively [34,35,36,37]. Long-lived plasma cells (LLPCs) derived from the terminal differentiation of B cells, are an important source of autoantibody production in SLE. Short-lived plasmablasts, after interaction with CD4+ T cells in the germinal centers of the lymph nodes, have been shown to transform into high-affinity plasma cells that migrate to specific niches in the bone marrow, where they are protected from external events, being able to survive for a long time and continuing to produce autoantibodies [38]. Spontaneous formation of germinal centers, favoring the generation of LLPCs, is observed in both murine and human SLE, suggesting that this phenomenon is strictly involved in the genesis of autoantibody production [39]. Importantly, B lymphocytes may also play a role as antigen-presenting cells (APC) to autoreactive T lymphocytes in SLE, as demonstrated in mouse models [40,41]. A debated issue is the pathogenetic role of autoantibodies. Because the presence of autoantibodies can be detected in serum even years before the clinical signs of SLE, this has been considered an indication that these antibodies are a biomarker of the disease rather than a pathogenic factor. However, much evidence suggests their central role in the immunopathogenesis of SLE. Of particular importance, is the observation of the presence of immune complexes in lupus nephritis, at the glomerular level, formed by various autoantibodies, including anti-dsDNA antibodies, whose removal leads to amelioration of the disease [42,43,44]. Moreover, neonatal lupus erythematosus (NLE) develops as a result of the passive transfer of maternal autoantibodies across the placenta, which does not allow the passage of cells, including those of the immune system [45]. From these and other observations, it is possible to conclude that autoantibodies may contribute, at least to some extent, to the clinical manifestations of SLE.

4.1.2. T Cells in SLE Pathogenesis

Self-reactive T cells play a key role in the genesis of SLE. T-helper 1 (Th1) cells play a central role in the pathogenesis of SLE, by promoting oxidative stress related to IFNγ production [46]. In contrast, the number of IL-4-producing Th2 cells is reduced in the peripheral blood of SLE patients, suggesting their potential protective role, and that SLE activity may be associated with an increased IFNγ/IL-4 ratio [47]. T-helper 17 (Th17) cells are also involved in SLE pathogenesis. These cells are the main source of IL-17, a family of cytokines with potent inflammatory effects. In addition to their defensive action against pathogens, members of the IL-17 family can exacerbate tissue injury, because of their pro-inflammatory activity. IL-17 induces neutrophil recruitment, activation of the innate immune system, and enhancement of B-lymphocyte functions [48]. It has been reported that IL-17 levels correlate with SLEDAI in patients with LN [49,50]. Regulatory T cells (Tregs) are critical in maintaining peripheral tolerance to self-antigens. Although quantitative and qualitative differences in Tregs have been described in SLE, studies to date have shown conflicting results, and their role in SLE is still undefined. However, some studies have proposed that these cells, due to their ability to suppress effector T lymphocytes, could be considered in the basic cell therapy of SLE [51,52,53]. T-follicular helper (Tfh) cells are located in germinal centers and extrafollicular foci. These cells have been involved in the generation of autoreactive B-cell clones in murine and human SLE [54]. Tfh cells were found to aggregate in renal tissue with B cells, similar to what is observed in germinal centers in LN [55]. All these findings support the concept that interactions between CD4+ T cells and B cells are crucial in the development of autoimmunity, as they contribute decisively to the development and maintenance of autoreactive B cells and their differentiation into autoantibody-producing plasma cells. CD8+ T cells are also involved in the immunopathogenesis of SLE. Circulating CD8 T lymphocytes of SLE patients exhibit functional defects, including impaired cytolytic function, with reduced production of granzyme and perforin [56]. A depleted phenotype of circulating CD8 T lymphocytes in SLE patients has been associated with lower disease flare rates [57]. However, the qualitative abnormalities of CD8+ T lymphocytes are also related to the susceptibility of SLE patients to infections, which may be further exacerbated by the use of immunosuppressive drugs [58]. Finally, γδ-T lymphocytes were found in a higher percentage in SLE patients than in controls, suggesting their role in the autoimmune response [59,60].

4.2. The Role of Innate Immunity

4.2.1. Role of Neutrophils in SLE

It has been observed that in SLE, the neutrophil function is abnormal at several levels. First, neutrophils show reduced phagocytic capacity [61] and the inability to remove apoptotic cells, which are a known source of normally hidden self-antigens [62,63]. Variants of ITGAM, NCF1, and NCF2 genes, have been reported to be risk factors for SLE development, since they induce alteration of phagocytosis and dysregulate reactive oxide species production (ROS) [64,65,66]. It has also been reported that neutrophils can produce type-I IFN independently of TLRs stimulation and promote abnormal B-cell development in the bone marrow of SLE patients [67,68]. A subtype of neutrophils, called low-density granulocytes (LDG), is highly represented in the peripheral blood of SLE patients. These cells are associated with the presence of IFN signature and disease severity [69,70,71], and activate CD4+ T-cells to produce IFNγ and TNFα [69]. In SLE, LDGs are characterized by an increased ability to form neutrophil extracellular traps (NETs), released during their apoptosis (NETosis) [72]. NETs are rich in decondensed nucleic acids, and chromatin expelled outside the cells during the formation process can induce specific autoreactive immune responses against nucleic acid antigens [73]. Neutrophils are also characterized by the generation of ROS which, in normal conditions are responsible for cell killing, but in SLE contribute to endothelial damage [73]. Several genetic polymorphisms related to neutrophil dysregulation, that increase NET formation, have been described in SLE [74,75,76]. Moreover, neutrophils from SLE patients with mutations resulting in loss of STAT3 function, form NETs more spontaneously than healthy controls [77]. The increased formation of NETs and their reduced clearance may also lead to increased inflammasome activation in macrophages, amplifying the inflammatory response [78]. Taken together, these observations indicate that neutrophils, through NET formation, have very important immunostimulatory effects in SLE, contributing significantly to the immune dysregulation that leads to tissue damage.

4.2.2. Role of Plasmacytoid Dendritic Cells

Lymphoid-origin plasmacytoid dendritic cells (pDCs) are characterized by the ability to produce high levels of type-I IFN, thus playing a key role in the pathogenesis of SLE [79,80,81]. Production of type-I IFN occurs primarily in response to circulating ssRNA and dsDNA, that are internalized by pDCs through Fc-gamma receptor IIA (FcγRIIa). Nucleic acids are then recognized by TLR7 and TLR9 in the cytoplasm [82]. These receptors can also be activated by endogenous nucleic acids present in NETs. Once activated, TLRs trigger signaling pathways mainly involving the myeloid differentiation response gene-88 (Myd88) and interleukin-1 receptor-associated kinase 4 (IRAK4), leading to the activation of interferon regulatory factors 3 (IRF3) and IRF7 for IFN-I production [83,84,85]. The pDC-dependent production of type-I IFN is also important in linking innate and adaptive immunity. This occurs through complex interactions involving monocytes, neutrophils, natural killer cells, and T and B cells [86,87]. In this regard, the production of IFN-I by pDCs, can promote the differentiation of extrafollicular B cells into short-lived plasmablasts, which produce anti-dsDNA antibodies, thus creating a positive feedback loop supporting the autoimmune response, as demonstrated in animal models [88]. Activation of pDCs and high levels of IFN-I production, also increase the number and recruitment of pro-inflammatory T cells. This occurs particularly in the arterial wall. This finding has been associated with the development of accelerated atherosclerosis, as commonly observed in the course of SLE [89]. It should be noted, however, that pDCs may have also a tolerogenic function, through the generation of regulatory T cells (Treg). As mentioned earlier in this review, these cells are known to inhibit the activation of effector T cells [90,91,92]. Moreover, pDCs also facilitate the differentiation of immature B cells into IL-10-producing regulatory B cells (Breg), which can limit IFN-I production by pDCs, thus forming a regulatory feedback, the dysregulation of which is one of the most important components of SLE pathogenesis [93]. PDCs are therefore another possible target of SLE cell therapies. Figure 2 shows the relationships between the innate and adaptive immune systems in the pathogenesis of SLE.

4.3. The Role of Mitochondria

Mitochondria are organelles that provide the energy necessary for cell metabolism and survival, being the major source of adenosine triphosphate (ATP) synthesis [94]. These ancestral structures can release high quantities of mitochondrial DNA (mtDNA) after destruction, following cell apoptosis. MtDNA is extremely unstable, and its easy degradation can generate antigenic fragments [95]. MtDNA has been found to induce specific autoreactive T lymphocytes in patients with SLE [96,97]. These in turn may induce B cells to produce anti-DNA antibodies. It has also been observed that sequences of mtDNA are analogous to those of bacteria and therefore able to activate TLRs. TLR recognition triggers powerful downstream inflammatory responses, including type-I IFN production [98]. This pro-inflammatory response, further contributes to the breakdown of tolerance. It has also been reported that several mitochondrial gene variants are linked to the risk of developing SLE, as demonstrated in mouse models [99,100,101]. Mitochondrial polymorphisms increase oxidative stress [102], as demonstrated by the accumulation of oxidized nucleic acids in the mitochondria of neutrophils of patients with SLE [103]. The accumulated oxidized mtDNA can be therefore extruded during NETosis, potentially triggering type-I IFN activation by plasmacytoid dendritic cells (pDCs) [63]. Finally, mitochondrial RNA (mtRNA) is another source of autoantigens, and titers of RNA autoantibodies against mtRNA are significantly higher in patients with SLE compared with controls [104].

4.4. The Role of Apoptosis

It has recently been shown that enzymes such as nucleases are critical for nucleic acid digestion and maintenance of tolerance [105]. A deficiency of nucleases is responsible for lupus-like manifestations in mouse models [106]. Recent studies have identified neutralizing antibodies against DNASE1L3 in some SLE patients, resulting in the accumulation of extracellular DNA and the formation of immune complexes [107]. Mice KO for DNaseII genes does not survive due to undigested DNA in phagocytes [108,109]. Mice KO for genes encoding TREX1 nuclease develops lupus-like symptoms, including skin lesions, vasculitis, nephritis, and sometimes systemic inflammation [110,111]. These observations lead to the conclusion that undigested nucleosomes are an initial inducer of the autoimmune responses, as observed in SLE patients. Apoptotic cell clearance can induce the breakdown of immune tolerance through several mechanisms. These include the activation of signals mediated by pattern recognition receptors (PRRS) [112,113,114]. The role of apoptosis defects in SLE has been confirmed in human studies, showing that patients with SLE have a defect in apoptotic cell clearance [115,116]. However, it should be noted that, in some mouse models with impaired apoptosis, autoimmunity does not occur, indicating that other events are required to induce the onset of SLE [117]. NETs are also involved in the apoptosis deficit in SLE. The molecular components of NETs can be complexed with DNA, making it resistant to enzymatic digestion by DNases, inducing type-I IFN production by plasmacytoid dendritic cells [118].

4.5. The Role of Interferons in SLE

A recent finding on the pathogenesis of SLE that opened new lines of research for innovative drug development, was the recognition of a high type-I IFN signature in SLE patients [119]. IFNs play a pivotal role in defense against pathogens. Three types of IFNs can be distinguished: type-I IFN, which is a family composed of IFNα (13 subtypes), IFNβ, IFNω, IFNκ, and IFNε; type-II IFN, also known as IFNγ; and type-III IFN (IFNλ) [3,120,121]. Type-I IFN, particularly the IFNα and IFNβ family members, is the one mainly involved in the pathogenesis of SLE. Type-I IFN is induced by the activation of pathogen recognition receptors (PRRs) such as toll-like receptors (TLRs), retinoic acid-inducible gene I (RIG-I), and melanoma differentiation-associated protein 5 (MDA5). All these molecules are activated by nucleic acids or by bacterial products such as lipopolysaccharides and peptidoglycan [122]. Although virtually any cell type can produce type-I IFN [123], very high levels of this cytokine are synthesized by pDCs, as already discussed [124,125,126]. Type-I IFNs recognize the IFNα receptor (IFNAR), a heterodimeric complex that in turn activates intracellular signaling through Janus kinase 1 (JAK1) and tyrosine kinase 2 (TYK2). These proteins phosphorylate transcriptional signal transducers and activators STAT 1 and STAT 2. These intracellular molecules bind IRF7 and IRF9, to form the ISGF3 complex. This complex translocates into the nucleus, where it induces transcription of genes named IFN-sensitive response elements (ISREs), encoding for proteins contributing to the inflammatory cascade [127,128]. Initial experimental animal model studies showed that the administration of type-I IFN was able to induce the production of autoantibodies and contribute to organ damage [129]. The earliest evidence suggesting that type one interferon could play a key role in the genesis of SLE in humans, came from the observation that patients treated with IFNα for hepatitis C [130] or neoplastic disease [131,132], could develop antinuclear antibody positivity and in some cases lupus-like syndromes. These clinical conditions regressed after discontinuation of IFNα treatment [133,134]. It has been shown that polymorphisms in genes along the type-I IFN signaling pathways represent important genetic risk factors for the occurrence of SLE-including IRF genes [135,136,137]. Polymorphisms have been also described in STAT4, STAT3, and TYK2 genes [77,138]. The IFN signature is also emerging as a possible biomarker for precision treatment with novel anti-IFN therapeutic agents, as discussed in more detail in the section on SLE therapy.

5. SLE Treatment

5.1. The EULAR/ACR Recommendations

Unlike other rheumatic diseases, the treatment of SLE has so far not taken decisive steps to replace traditional therapy [139]. According to recent recommendations of the European League Against Rheumatism (EULAR), the goal of treatment is to achieve remission, or at least a state of low disease activity [4,140,141]. The concept of "treat-to-target," initially formulated for the treatment of rheumatoid arthritis, has thus been extended to SLE therapies as well [4]. EULAR and the ACR periodically update their recommendations for the treatment of SLE. Regarding pharmacological treatment with conventional disease-modifying drugs (csDMARDs), the use of hydroxychloroquine (HCQ) was recommended for all patients, at a dose of no more than 5 mg/kg body weight, to minimize retinal complications. HQC, a drug that has been shown to be safe in the long term, reduces episodes of disease flare and the occurrence of cardiovascular events [142,143,144]. It also has a low cost, that makes it usable even in low-income geographic areas. Therefore, it is likely that this drug will be used for a long time to come in the treatment of SLE, in combination with innovative drugs. For chronic treatment, glucocorticoids should be reduced to less than 7.5 mg/d prednisone equivalent and, if possible, discontinued. In fact, steroids are burdened with side effects when administered for prolonged periods or at high doses, such as induction of osteoporosis, worsening of diabetes and hypertension, and susceptibility to infections [145]. Interruption of steroid administration is sometimes possible, through the use of so-called steroid-sparing agents, that include the immunosuppressants azathioprine, methotrexate, mycophenolate mofetil, and cyclophosphamide (CYC). However, side effects and toxicity limit the use of these drugs. Mycophenolate, generally more effective than azathioprine, is teratogenic and cannot be used during pregnancy [146]. CYC has been shown to be highly effective in organ-threatening manifestations, such as central nervous system involvement and LN, even at a lower dose than before [147]. CYC, however, is burdened by several important side effects, including cancer [148]. Therefore, for patients who do not respond, or respond poorly, to conventional immunosuppressants, to avoid severe side effects and allow safer pregnancy, new therapies are highly desirable. As pointed out earlier, the reduction in steroids is a particularly important recommendation, although they remain an important and often unavoidable therapeutic adjunct. In addition, the use of other conventional immunosuppressive therapies is still necessary. For all these considerations, additional treatment with the available innovative or biologic drugs is recommended only in addition to conventional therapy, when the latter is not sufficient to control the disease [149].

5.2. Rituximab

The use of rituximab, a chimeric anti-CD20 monoclonal antibody capable of causing B-lymphocyte depletion, has a great rationale for use in SLE therapy. However, two major clinical trials failed to achieve their primary endpoints, probably due to faulty study design. Notably, both clinical trials, EXPLORER for moderately-to-severely active SLE [150], and LUNAR for lupus nephritis [151], failed to meet their respective endpoints. However, it should be noted that some important favorable therapeutic effects of rituximab therapy have emerged from further data analysis. In this regard, patients treated with rituximab showed improvements in proteinuria in the LUNAR study [151], and a subgroup analysis of African Americans and Hispanics, suggested a clinical benefit in the EXPLORER study [150]. Another possibility, besides a flawed study design, that may explain the disappointing results obtained with the use of rituximab in SLE, is the inefficient depletion of tissue-resident CD201 B cells. Indeed, although circulating B cells are rapidly depleted, the persistence of B cells within inflamed target organs has been observed in patients with rheumatoid arthritis or Sjogren’s syndrome treated with rituximab [152,153]. In mouse models of lupus, complete ablation of B cells by treatment with chimeric anti-CD19 receptors, has been shown to produce lasting remission of the disease [154]. Thus, only complete ablation of peripheral B cells may correlate with an effective therapeutic response. This was highlighted in a post hoc analysis of data obtained from the LUNAR study [155]. For this reason, the use of novel anti-CD20 agents, designed for more efficient B-cell ablation, has been proposed [156]. Obinutuzumab, a type II monoclonal antibody showing more effective B-cell depletion than rituximab [156,157], is currently being tested in class III/IV lupus nephritis (NOBILITY study). The available data show that patients with active class III/IV nephritic lupus treated with obinutuzumab, have improved renal response, with 12.7% greater efficacy than a placebo. However, some infectious complications have emerged during therapy with this agent [158].

5.3. Belimumab

Anti-BAFF belimumab received Food and Drug Administration (FDA) approval in 2011, for the treatment of moderate to severe SLE, in patients older than 18 years. This approval was based on the efficacy results of the BLISS-52 and BLISS-76 trials [159,160]. In BLISS-52, enrolled subjects were randomized to receive standard therapy plus belimumab, administered intravenously, or placebo, every 4 weeks, with a loading dose administered at week 2. After 52 weeks, subjects receiving belimumab 10 mg/kg or 1 mg/kg, showed a statistically higher probability of achieving an SRI-4 response than those receiving placebo (57.6% and 51.4% vs. 43.6%, respectively). Similarly, in BLISS-72, subjects randomized to belimumab administered intravenously at the same dosages as BLISS-52, achieved an SRI-4 response in a percentage statistically more significant than those who received placebo (43.2% and 40.6% vs. 33.5%, respectively). It was also observed that, the patients who were most likely to respond to therapy, were those with high disease activity (SELENA-SLEDAI ≥ 10), hypocomplementemia, and anti-dsDNA positivity and/or prednisone use at baseline [161]. A reduction in organ damage was also demonstrated in subsequent studies [162,163,164]. The subcutaneous route of administration of belimumab was evaluated in the BLISS-SC study, in which treated subjects achieved an SRI-4 in 61% of cases, compared to 48% of those who received the placebo [165]. These results led to the approval of subcutaneous belimumab, in 2017, for active SLE. In 2019, belimumab was approved for use in pediatric patients with SLE, following an international multicenter study (PLUTO) in subjects aged 5 to 17 years. In this study, subjects who received belimumab achieved a significantly higher percentage of SRI-4 than subjects who received the placebo (52.8% vs. 43.6%) [166]. Belimumab, both intravenous and subcutaneous, then received FDA approval also for the treatment of active lupus nephritis, based on the results obtained in the BLISS-LN study [167]. In this phase III study, patients with class III, IV (with or without class V), or V LN were treated intravenously with belimumab or placebo, along with standard therapy. The patients who received belimumab were significantly more likely to achieve a primary efficacy of renal response (PERR) at week 104. PERR was defined by a urinary protein/creatinine ratio (UPCR) ≤ 0.7, an estimated glomerular filtration rate (eGFR) not less than 20% of basal value, or ≥60 mL per minute. A complete renal response (CRR) was defined as a urinary protein ratio < 0.5, an eGFR not ≥10% less than the basal value, or ≥90 mL per minute. The results showed that, patients treated with belimumab achieved a 43% response compared with 32% in the placebo group, and 30% compared with 20% in the placebo group, in PERR and CRR, respectively. The effect of belimumab in the preservation of kidney function was further confirmed in secondary data analysis [168].

5.4. Anifrolumab

Anifrolumab is a fully human immunoglobulin G1 monoclonal antibody, that targets and inhibits type-I IFN receptor subunit 1 signaling. Anifrolumab was approved for the treatment of moderate to severe SLE in combination with standard of care, at a dose of 300 mg administered every four weeks intravenously. Approval by the regulatory agencies was based on efficacy data from the phase IIb MUSE study and the phase III TULIP-1 and -2 studies [169]. In the MUSE trial, a group of patients was randomized to receive anifrolumab 300 or 1000 mg intravenously, or placebo, every 4 weeks for 48 weeks. Patients also received glucocorticoids (GC), an antimalarial, azathioprine, mizoribine, mycophenolate mofetil or mycophenolic acid, or methotrexate [170]. The primary endpoint was the percentage of patients achieving the Systemic erythematosus lupus response index-4 (SRI-4) at week 24, with concomitant oral GC dose reduction at <4 weeks [170]. More patients treated with anifrolumab 300 mg (34.3%) and 1000 mg (28.8%), responded significantly to treatment compared to placebo (17.6%). Better responses were also observed for systemic lupus erythematosus responder index-4 (SRI-4) and BILAG-based composite lupus assessment (BICLA) at 52 weeks. Therefore, two phase III studies, Treatment of Uncontrolled Lupus via the Interferon Pathway (TULIP)-1 and TULIP-2, were conducted. TULIP-1 included an additional lower dose of anifrolumab of 150 mg. The placebo-controlled TULIP-1 study, involved the administration of anifrolumab every 4 weeks for 48 weeks [171]. The primary endpoint was the achievement of SRI-4 at week 52. There was no significant difference in the percentages of patients who reached the primary endpoint between the anifrolumab 300 mg group and the placebo group (36% vs. 40% of patients, respectively), out of a total of 364 subjects who completed the study. In addition, no significant response in the anifrolumab 150 mg group was observed. However, post hoc analysis considering the results, taking into account NSAID use, showed that all outcome measures improved, even though the primary endpoint had not yet been achieved. Of note, the BICLA responses at week 52 were in favor of anifrolumab 300 mg (46%) compared with placebo (30%). Pharmacodynamic evaluation of patients with a type-I IFN-high genetic signature, showed a neutralization of the type-I IFN 21 gene panel of 12.6% as early as week 12 and throughout the study period, for the anifrolumab 300 mg group, but not for the placebo group. Therefore, the experience of the TULIP-1 study led to a modification of the study protocol of the sister study, TULIP-2, before data unblinding [172]. Specifically, the primary endpoint was changed from an SRI-4 response to a BICLA response, at week 52. Disease relapse was defined as ≥1 new BILAG-2004 A-item or ≥2 new BILAG-2004 B-items, as compared with the baseline. After this protocol modification, TULIP-2 met the primary endpoint, as 47.8% of patients on anifrolumab, compared with 31.5% in the placebo group, responded according to the above BICLA criteria. Secondary outcomes were also achieved in the type-I IFN-high group (48% vs. 30.5% in the placebo group), with a reduction in GC use and ≥50% of cutaneous LE activity index (CLASI) at week 12 (49% vs. 25%). Data from the two TULIP studies were then pooled and analyzed [173,174]. More patients in the anifrolumab 300 mg group achieved a BICLA response than those in the placebo group (47.5%). Similarly, more patients in the anifrolumab 300 mg group achieved a response to SRI-4 (52.2% vs. 40.1%), sustained tapering of GC (50.5% vs. 31.8%), a ≥50% reduction in CLASI activity (46% vs. 24.9%), and ≥50% reduction in active joints (49.4% vs. 36.8%), compared with the placebo-treated group [173]. Moreover, the group treated with anifrolumab 300 mg, had a lower annualized flare rate (AFR) than the placebo (0.51% vs. 0.8%), a longer median time to first relapse (140 days vs. 119 days), and fewer patients with ≥1 relapse (33.6% vs. 42.9%). Importantly, among patients who achieved sustained reductions in GC, more patients remained flare-free with anifrolumab (40%) compared with placebo (17%). In the TULIP-LN study, the primary endpoint was the change in the 24-hour urinary protein–creatinine ratio (UPCR) at week 52, for the combined anifrolumab (BR and IR) groups compared with placebo [175]. The secondary endpoint was a CRR, defined as 24-hour UPCR ≤ 0.7 mg/mg, eGFR ≥ 60 ml/min/1.73 m2, or no decrease ≥ 20% from baseline. There was no difference in the primary endpoint between the anifrolumab-treated group and the placebo group. However, there were numerical improvements in the CRR achieved by the IR group compared with placebo (45.5% vs. 31.1%), and in the sustained reduction in GC dose (55.6% vs. 33.3%) in the same group. Overall, these data confirmed the efficacy of anifrolumab on many clinical outcomes, including BICLA and SRI-4, mucocutaneous and musculoskeletal manifestations, lower flare rates, relapse rates, and successful tapering from GC to ≤7.5 mg/day. All of these data led, as mentioned, to the approval of anifrolumab as an adjunctive treatment of active moderate to severe SLE, at the recommended dose of 300 mg, administered as a 30-minute intravenous infusion every four weeks, with possible discontinuation of treatment if no improvement in disease control is observed after six months. Table 1 summarizes the main clinical trials of new drugs currently used in SLE therapy.

5.5. Voclosporin

Voclosporin (VCS) is an oral calcineurin inhibitor, belonging to the same drug class as tacrolimus and cyclosporine. VCS was approved in January 2021, by the FDA, for the treatment of active lupus nephritis, in combination with conventional immunosuppressive therapy [176]. Two randomized controlled trial (RCT) studies demonstrated improved renal response rate and reduced proteinuria when VCS was added to mycophenolate mofetil (MMF) and steroids, compared with the group receiving MMF and steroids alone [177,178]. Preliminary interim data from a two year extension study, showed sustained reductions in proteinuria and stability of renal function after VCS therapy for up to 30 months [70].

6. Future Therapies

6.1. Targeting of Plasma Cells

LLPCs, which are responsible for antibody production in the bone marrow, do not express CD19 and CD20 molecules and do not require BAFF for long-term survival [179]. Therefore, they are resistant to available B-cell-targeted therapies. This observation suggested that plasma cell depletion might be a more effective target in SLE therapy. To this end, several strategies, such as inhibition of the proteasome, which is necessary for LLPCs’ survival, are being investigated. In animal models, proteasome inhibitor bortezomib reduced the number of plasma cells and prolonged survival in a mouse model of SLE [180]. However, in a small randomized trial in human SLE, a high rate of treatment discontinuation was found, due to severe side effects and minimal impact on dsDNA titers [181]. Possibly, compensatory augmentation of constitutive proteasome components may protect plasma cells from death, by significantly limiting therapeutic capping [182]. Other studies have exploited the experiences of onco-hematology in the treatment of multiple myeloma [183]. In this regard, pilot studies have shown that daratumumab, an anti-CD38 monoclonal antibody expressed on plasma cells, was able to induce significant clinical benefits in refractory SLE [184]. However, because CD38 is expressed by other cells of the immune system important for pathogen defense and immunoregulation, alternative strategies to target plasma cells in SLE should be pursued. In this regard, potential therapeutic agents appear to be elotuzumab, targeting SLAMF7, which is expressed in myeloma cells and nonmalignant plasma cells, as well as activated DN2 cells and circulating antibody-secreting cells (ASCs), in human SLE [185].

6.2. Cell-Based Therapies

6.2.1. Hematopoietic Stem Cell Transplantation

The hematopoietic stem cell transplantation (HSCT) approach, commonly used in hematology, was initially employed for life-threatening SLE. In recent times, thanks to improvements in biotechnology, this treatment option has become a possibility for patients who are less severe but refractory to standard therapy [186,187]. The rationale for this therapy is to eliminate self-reactive memory T and B lymphocytes from the recipient’s immune system, as well as plasma cells that, as described above, are refractory to standard B-cell depletion therapy and are not sensitive to anti-BAFF agents, and to replace them with normal cells. In contrast, plasma cells are susceptible to conditioning treatment in the presence of anti-thymocyte globulin, to avoid graft-versus-host disease (GVHD), followed by regeneration of the hematopoietic and immune system by stem cell transplantation [188,189,190]. To date, more than 300 patients have received autologous transplantation for SLE. Reported results have shown that 50%–66% of treated patients achieved remission at five years after discontinuation of immunosuppressive therapy [191]. Most recent studies have reported treatment-related mortality to be less than 5%, falling significantly from the initial studies. Patients who respond, are usually free of clinical symptoms, and can regain seronegativity for antinuclear antibodies, a condition very difficult to achieve with standard therapy [140]. Early use of HSCT has also been found to protect against organ failure and drug toxicity, and to improve quality of life [192]. Allogeneic HSCT (allo-HSCT) can also be used to restore a dysfunctional immune system, although its wide application has been limited by the risk of graft-versus-host disease (GvHD) and other procedure-related complications. A retrospective analysis of the EBMT registry, published in 2019, reported five SLE patients successfully treated with allo-HSCT, and three other SLE cases in the literature achieved complete remission of autoimmune manifestations [193,194,195,196]. This evidence provides the principle for the use of allo-HSCT as a potentially curative approach.

6.2.2. Chimeric Antigen Receptor T Cells

Cancer immunotherapy has experienced a new phase of development in recent years. One method is to artificially harness the immune system to fight cancer cells, after the fusion of T cells with chimeric molecules specific for tumor antigens. Such genetically engineered cells are called chimeric antigen receptor T cells (CAR-T). Chimeric antigen receptors (CARs) are composed of an extracellular domain and an intracellular portion. The extracellular domain is responsible for target recognition, being in most cases derived from the light chain and heavy chain of the variable portion of monoclonal antibodies bound together [197,198,199]. The intracellular portion mediates signal transduction in the T cell, and consists of one or more domains [200]. These two portions are linked together by a peptide linker or spacer [201,202]. CARs have been extensively studied in the treatment of cancer. Currently, several products, based on autologous CAR-T cells targeting the surface antigen CD19, have been approved for the treatment of B-cell malignancies [203]. Recently, the first data were published on the use of an anti-CD19 CAR-T cell-based strategy in a patient with refractory SLE. These data demonstrated rapid clinical remission, with no notable adverse effects, accompanied by sustained depletion of circulating B lymphocytes and rapid disappearance of native serum anti-DNA antibodies [204]. Subsequently, four additional patients with SLE, presenting with a form of disease refractory to standard therapy, were treated. Preliminary results on safety and efficacy were encouraging, but data on long-term follow-up are needed [205]. The toxicities of the therapy consist mainly of cytokine storm syndrome, that can be life-threatening for patients, although in most cases effective specific strategies can be used that can block released pro-inflammatory cytokines, including the use of monoclonal antibodies that can block IL-6 activity [206]. In SLE, B-cell depletion is a potentially curative approach, although a complete blockade of antibody production may increase susceptibility to infection. For these considerations, CAR-T cells represent an interesting and promising approach for SLE therapy, but some questions remain open, such as the duration of responses during B-cell repopulation and the identification of an appropriate target population.

6.3. Other Promising Therapies

Some attempts have been made to enhance Treg in SLE, due to their immunosuppressant properties, with promising results, by adoptive transfer of polyclonal Treg cells [207,208]. The possibility of targeting pDCs with BDCA2-specific antibody, because of the critical role of such cells in the pathogenesis of SLE, is an attractive therapeutic strategy [209]. The administration of allogeneic mesenchymal cells with immunoregulatory properties represents a further therapeutic opportunity [210]. In preclinical studies, inhibition of peptidyl arginine deiminase 4 (PAD4), an enzyme playing an important role in NET formation, has been found to exert a favorable effect in nephritis mouse models [211]. Inhibition of various receptor-associated kinases of several pro-inflammatory molecules, including JAK1 and TYK2, has shown promise, and RCT studies are underway to evaluate their efficacy [212]. Other studies are examining strategies to increase the immunosuppressive activity of regulatory T cells, through their activation by low-dose IL-2 administration [213]. Table 2 shows the innovative therapies that are still under investigation for possible future treatment of SLE.

7. Conclusions

SLE is a disease whose genesis is not yet well understood. The involvement of both the innate and adaptive immune systems underscore its great complexity. However, recent advances in research have made it possible to unravel at least some of the fundamental mechanisms underlying the disease. The identification of the key role played by type-I IFN has given further impetus to research. The results obtained so far have made it possible to develop new selective therapies, but these have proven effective only in combination with conventional therapy. In addition, other important issues need to be addressed. These include, assessing whether the side effects of the newly listed agents are acceptable, compared with steroids and conventional therapies. In addition, the high price of the new drugs may not be sustainable in the long term in most countries; finally, it has not yet been clarified for how long the new drugs should be administered. Further research is therefore needed to develop safe and effective therapies, that can be used without the need for steroids and conventional immunosuppressive drugs, which are burdened by many side effects, with a significant impact on patient morbidity and mortality

Author Contributions

Conceptualization, D.A., R.C. and M.P.; methodology, D.A. and M.P.; software, L.S. and M.P.; validation, D.A., R.C., M.P.P. and M.P.; formal analysis, D.A., R.C., M.P.P, C.G., B.L.N., L.S. and M.P.; investigation, D.A., R.C., and M.P.; resources, M.P.; data curation, D.A., R.C., M.P.P., C.G., B.L.N., L.S. and M.P.; writing—original draft preparation, M.P.; writing—review and editing, D.A., R.C. and M.P.; visualization, D.A., R.C. and M.P.; supervision, D.A. and M.P.; project administration, M.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fava, A.; Petri, M. Systemic lupus erythematosus: Diagnosis and clinical management. J. Autoimmun. 2019, 96, 1–13. [Google Scholar] [CrossRef] [PubMed]
  2. Justiz Vaillant, A.A.; Goyal, A.; Varacallo, M. Systemic Lupus Erythematosus. In StatPearls; StatPearls Publishing: Treasure Island, FL, USA, 2022. [Google Scholar]
  3. Barrat, F.J.; Crow, M.K.; Ivashkiv, L.B. Interferon target-gene expression and epigenomic signatures in health and disease. Nat. Immunol. 2019, 20, 1574–1583. [Google Scholar] [CrossRef] [PubMed]
  4. Fanouriakis, A.; Kostopoulou, M.; Alunno, A.; Aringer, M.; Bajema, I.; Boletis, J.N.; Cervera, R.; Doria, A.; Gordon, C.; Govoni, M.; et al. 2019 update of the EULAR recommendations for the management of systemic lupus erythematosus. Ann. Rheum. Dis. 2019, 78, 736–745. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Fanouriakis, A.; Tziolos, N.; Bertsias, G.; Boumpas, D.T. Update omicronn the diagnosis and management of systemic lupus erythematosus. Ann. Rheum. Dis. 2021, 80, 14–25. [Google Scholar] [CrossRef]
  6. Rees, F.; Doherty, M.; Grainge, M.J.; Lanyon, P.; Zhang, W. The worldwide incidence and prevalence of systemic lupus erythematosus: A systematic review of epidemiological studies. Rheumatology 2017, 56, 1945–1961. [Google Scholar] [CrossRef] [Green Version]
  7. Niewold, T.B. Advances in lupus genetics. Curr. Opin. Rheumatol. 2015, 27, 440–447. [Google Scholar] [CrossRef] [Green Version]
  8. Salgado-Galicia, N.A.; Hernandez-Dono, S.; Ruiz-Gomez, D.; Jakez-Ocampo, J.; Zuniga, J.; Vargas-Alarcon, G.; Acuna, V.; Hernandez, M.T.; Marquez-Garcia, J.E.; Garcia-Lechuga, M.; et al. The role of socioeconomic status in the susceptibility to develop systemic lupus erythematosus in Mexican patients. Clin. Rheumatol. 2020, 39, 2151–2161. [Google Scholar] [CrossRef]
  9. Danchenko, N.; Satia, J.A.; Anthony, M.S. Epidemiology of systemic lupus erythematosus: A comparison of worldwide disease burden. Lupus 2006, 15, 308–318. [Google Scholar] [CrossRef]
  10. Stojan, G.; Petri, M. Epidemiology of systemic lupus erythematosus: An update. Curr. Opin. Rheumatol. 2018, 30, 144–150. [Google Scholar] [CrossRef]
  11. Ocampo-Piraquive, V.; Nieto-Aristizabal, I.; Canas, C.A.; Tobon, G.J. Mortality in systemic lupus erythematosus: Causes, predictors and interventions. Expert Rev. Clin. Immunol. 2018, 14, 1043–1053. [Google Scholar] [CrossRef]
  12. Lee, Y.H.; Choi, S.J.; Ji, J.D.; Song, G.G. Overall and cause-specific mortality in systemic lupus erythematosus: An updated meta-analysis. Lupus 2016, 25, 727–734. [Google Scholar] [CrossRef] [PubMed]
  13. Gauderon, A.; Roux-Lombard, P.; Spoerl, D. Antinuclear Antibodies With a Homogeneous and Speckled Immunofluorescence Pattern Are Associated With Lack of Cancer While Those With a Nucleolar Pattern With the Presence of Cancer. Front. Med. 2020, 7, 165. [Google Scholar] [CrossRef] [PubMed]
  14. Applbaum, E.; Lichtbroun, A. Novel Sjogren’s autoantibodies found in fibromyalgia patients with sicca and/or xerostomia. Autoimmun. Rev. 2019, 18, 199–202. [Google Scholar] [CrossRef] [PubMed]
  15. Lawrence, J.M., 3rd; Moore, T.L.; Osborn, T.G.; Nesher, G.; Madson, K.L.; Kinsella, M.B. Autoantibody studies in juvenile rheumatoid arthritis. Semin. Arthritis Rheum. 1993, 22, 265–274. [Google Scholar] [CrossRef] [PubMed]
  16. Pisetsky, D.S.; Lipsky, P.E. New insights into the role of antinuclear antibodies in systemic lupus erythematosus. Nat. Rev. Rheumatol. 2020, 16, 565–579. [Google Scholar] [CrossRef]
  17. Petri, M.A.; van Vollenhoven, R.F.; Buyon, J.; Levy, R.A.; Navarra, S.V.; Cervera, R.; Zhong, Z.J.; Freimuth, W.W.; Bliss; Groups, B.-S. Baseline predictors of systemic lupus erythematosus flares: Data from the combined placebo groups in the phase III belimumab trials. Arthritis Rheum. 2013, 65, 2143–2153. [Google Scholar] [CrossRef]
  18. Negoro, N.; Kanayama, Y.; Takeda, T.; Amatsu, K.; Koda, S.; Inoue, Y.; Kim, T.; Okamura, M.; Inoue, T. Clinical significance of U1-RNP immune complexes in mixed connective tissue disease and systemic lupus erythematosus. Rheumatol. Int. 1987, 7, 7–11. [Google Scholar] [CrossRef]
  19. Novak, G.V.; Marques, M.; Balbi, V.; Gormezano, N.W.; Kozu, K.; Sakamoto, A.P.; Pereira, R.M.; Terreri, M.T.; Magalhaes, C.S.; Guariento, A.; et al. Anti-RO/SSA and anti-La/SSB antibodies: Association with mild lupus manifestations in 645 childhood-onset systemic lupus erythematosus. Autoimmun. Rev. 2017, 16, 132–135. [Google Scholar] [CrossRef] [Green Version]
  20. Ramirez, G.A.; Canti, V.; Del Rosso, S.; Erra, R.; Moiola, L.; Magnoni, M.; Bozzolo, E.P.; Manfredi, A.A.; Rovere-Querini, P. Diagnostic performance of aPS/PT antibodies in neuropsychiatric lupus and cardiovascular complications of systemic lupus erythematosus. Autoimmunity 2020, 53, 21–27. [Google Scholar] [CrossRef]
  21. Unlu, O.; Zuily, S.; Erkan, D. The clinical significance of antiphospholipid antibodies in systemic lupus erythematosus. Eur. J. Rheumatol. 2016, 3, 75–84. [Google Scholar] [CrossRef]
  22. Aringer, M.; Costenbader, K.; Daikh, D.; Brinks, R.; Mosca, M.; Ramsey-Goldman, R.; Smolen, J.S.; Wofsy, D.; Boumpas, D.T.; Kamen, D.L.; et al. 2019 European League Against Rheumatism/American College of Rheumatology classification criteria for systemic lupus erythematosus. Ann. Rheum. Dis. 2019, 78, 1151–1159. [Google Scholar] [CrossRef] [Green Version]
  23. Mikdashi, J.; Nived, O. Measuring disease activity in adults with systemic lupus erythematosus: The challenges of administrative burden and responsiveness to patient concerns in clinical research. Arthritis Res. Ther. 2015, 17, 183. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Gladman, D.D.; Ibanez, D.; Urowitz, M.B. Systemic lupus erythematosus disease activity index 2000. J. Rheumatol. 2002, 29, 288–291. [Google Scholar]
  25. Isenberg, D.A.; Rahman, A.; Allen, E.; Farewell, V.; Akil, M.; Bruce, I.N.; D’Cruz, D.; Griffiths, B.; Khamashta, M.; Maddison, P.; et al. BILAG 2004. Development and initial validation of an updated version of the British Isles Lupus Assessment Group’s disease activity index for patients with systemic lupus erythematosus. Rheumatology 2005, 44, 902–906. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Gladman, D.; Ginzler, E.; Goldsmith, C.; Fortin, P.; Liang, M.; Urowitz, M.; Bacon, P.; Bombardieri, S.; Hanly, J.; Hay, E.; et al. The development and initial validation of the Systemic Lupus International Collaborating Clinics/American College of Rheumatology damage index for systemic lupus erythematosus. Arthritis Rheum. 1996, 39, 363–369. [Google Scholar] [CrossRef]
  27. Luijten, K.M.; Tekstra, J.; Bijlsma, J.W.; Bijl, M. The Systemic Lupus Erythematosus Responder Index (SRI); a new SLE disease activity assessment. Autoimmun. Rev. 2012, 11, 326–329. [Google Scholar] [CrossRef] [PubMed]
  28. Robinson, R. Distinct B cell receptor functions are determined by phosphorylation. PLoS Biol. 2006, 4, e231. [Google Scholar] [CrossRef] [Green Version]
  29. Nemazee, D. Mechanisms of central tolerance for B cells. Nat. Rev. Immunol. 2017, 17, 281–294. [Google Scholar] [CrossRef]
  30. Zikherman, J.; Parameswaran, R.; Weiss, A. Endogenous antigen tunes the responsiveness of naive B cells but not T cells. Nature 2012, 489, 160–164. [Google Scholar] [CrossRef] [Green Version]
  31. Burnett, D.L.; Langley, D.B.; Schofield, P.; Hermes, J.R.; Chan, T.D.; Jackson, J.; Bourne, K.; Reed, J.H.; Patterson, K.; Porebski, B.T.; et al. Germinal center antibody mutation trajectories are determined by rapid self/foreign discrimination. Science 2018, 360, 223–226. [Google Scholar] [CrossRef] [Green Version]
  32. Mackay, F.; Schneider, P. Cracking the BAFF code. Nat. Rev. Immunol. 2009, 9, 491–502. [Google Scholar] [CrossRef] [Green Version]
  33. Jackson, S.W.; Davidson, A. BAFF inhibition in SLE-Is tolerance restored? Immunol. Rev. 2019, 292, 102–119. [Google Scholar] [CrossRef]
  34. Christensen, S.R.; Shupe, J.; Nickerson, K.; Kashgarian, M.; Flavell, R.A.; Shlomchik, M.J. Toll-like receptor 7 and TLR9 dictate autoantibody specificity and have opposing inflammatory and regulatory roles in a murine model of lupus. Immunity 2006, 25, 417–428. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Berland, R.; Fernandez, L.; Kari, E.; Han, J.H.; Lomakin, I.; Akira, S.; Wortis, H.H.; Kearney, J.F.; Ucci, A.A.; Imanishi-Kari, T. Toll-like receptor 7-dependent loss of B cell tolerance in pathogenic autoantibody knockin mice. Immunity 2006, 25, 429–440. [Google Scholar] [CrossRef] [Green Version]
  36. Christensen, S.R.; Kashgarian, M.; Alexopoulou, L.; Flavell, R.A.; Akira, S.; Shlomchik, M.J. Toll-like receptor 9 controls anti-DNA autoantibody production in murine lupus. J. Exp. Med. 2005, 202, 321–331. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Lartigue, A.; Courville, P.; Auquit, I.; Francois, A.; Arnoult, C.; Tron, F.; Gilbert, D.; Musette, P. Role of TLR9 in anti-nucleosome and anti-DNA antibody production in lpr mutation-induced murine lupus. J. Immunol. 2006, 177, 1349–1354. [Google Scholar] [CrossRef] [Green Version]
  38. Nutt, S.L.; Hodgkin, P.D.; Tarlinton, D.M.; Corcoran, L.M. The generation of antibody-secreting plasma cells. Nat. Rev. Immunol. 2015, 15, 160–171. [Google Scholar] [CrossRef]
  39. Vinuesa, C.G.; Sanz, I.; Cook, M.C. Dysregulation of germinal centres in autoimmune disease. Nat. Rev. Immunol. 2009, 9, 845–857. [Google Scholar] [CrossRef]
  40. Chan, O.; Shlomchik, M.J. A new role for B cells in systemic autoimmunity: B cells promote spontaneous T cell activation in MRL-lpr/lpr mice. J. Immunol. 1998, 160, 51–59. [Google Scholar] [CrossRef] [PubMed]
  41. Chan, O.T.; Hannum, L.G.; Haberman, A.M.; Madaio, M.P.; Shlomchik, M.J. A novel mouse with B cells but lacking serum antibody reveals an antibody-independent role for B cells in murine lupus. J. Exp. Med. 1999, 189, 1639–1648. [Google Scholar] [CrossRef]
  42. Koffler, D.; Schur, P.H.; Kunkel, H.G. Immunological studies concerning the nephritis of systemic lupus erythematosus. J. Exp. Med. 1967, 126, 607–624. [Google Scholar] [CrossRef] [PubMed]
  43. Ehrenstein, M.R.; Katz, D.R.; Griffiths, M.H.; Papadaki, L.; Winkler, T.H.; Kalden, J.R.; Isenberg, D.A. Human IgG anti-DNA antibodies deposit in kidneys and induce proteinuria in SCID mice. Kidney Int. 1995, 48, 705–711. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Mannik, M.; Merrill, C.E.; Stamps, L.D.; Wener, M.H. Multiple autoantibodies form the glomerular immune deposits in patients with systemic lupus erythematosus. J. Rheumatol. 2003, 30, 1495–1504. [Google Scholar] [PubMed]
  45. Izmirly, P.M.; Rivera, T.L.; Buyon, J.P. Neonatal lupus syndromes. Rheum. Dis. Clin. N. Am. 2007, 33, 267–285. [Google Scholar] [CrossRef] [PubMed]
  46. Shah, D.; Kiran, R.; Wanchu, A.; Bhatnagar, A. Oxidative stress in systemic lupus erythematosus: Relationship to Th1 cytokine and disease activity. Immunol. Lett. 2010, 129, 7–12. [Google Scholar] [CrossRef]
  47. Sugimoto, K.; Morimoto, S.; Kaneko, H.; Nozawa, K.; Tokano, Y.; Takasaki, Y.; Hashimoto, H. Decreased IL-4 producing CD4+ T cells in patients with active systemic lupus erythematosus-relation to IL-12R expression. Autoimmunity 2002, 35, 381–387. [Google Scholar] [CrossRef]
  48. Paroli, M.; Caccavale, R.; Fiorillo, M.T.; Spadea, L.; Gumina, S.; Candela, V.; Paroli, M.P. The Double Game Played by Th17 Cells in Infection: Host Defense and Immunopathology. Pathogens 2022, 11, 1547. [Google Scholar] [CrossRef]
  49. Lopez, P.; Rodriguez-Carrio, J.; Caminal-Montero, L.; Mozo, L.; Suarez, A. A pathogenic IFNalpha, BLyS and IL-17 axis in Systemic Lupus Erythematosus patients. Sci. Rep. 2016, 6, 20651. [Google Scholar] [CrossRef] [Green Version]
  50. Zickert, A.; Amoudruz, P.; Sundstrom, Y.; Ronnelid, J.; Malmstrom, V.; Gunnarsson, I. IL-17 and IL-23 in lupus nephritis—association to histopathology and response to treatment. BMC Immunol. 2015, 16, 7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. 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]
  52. Alexander, T.; Sattler, A.; Templin, L.; Kohler, S.; Gross, C.; Meisel, A.; Sawitzki, B.; Burmester, G.R.; Arnold, R.; Radbruch, A.; et al. Foxp3+ Helios+ regulatory T cells are expanded in active systemic lupus erythematosus. Ann. Rheum. Dis. 2013, 72, 1549–1558. [Google Scholar] [CrossRef] [PubMed]
  53. La Cava, A. Tregs in SLE: An Update. Curr. Rheumatol. Rep. 2018, 20, 6. [Google Scholar] [CrossRef]
  54. Chen, P.M.; Tsokos, G.C. T Cell Abnormalities in the Pathogenesis of Systemic Lupus Erythematosus: An Update. Curr. Rheumatol. Rep. 2021, 23, 12. [Google Scholar] [CrossRef] [PubMed]
  55. Liarski, V.M.; Kaverina, N.; Chang, A.; Brandt, D.; Yanez, D.; Talasnik, L.; Carlesso, G.; Herbst, R.; Utset, T.O.; Labno, C.; et al. Cell distance mapping identifies functional T follicular helper cells in inflamed human renal tissue. Sci. Transl. Med. 2014, 6, 230ra246. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Comte, D.; Karampetsou, M.P.; Yoshida, N.; Kis-Toth, K.; Kyttaris, V.C.; Tsokos, G.C. Signaling Lymphocytic Activation Molecule Family Member 7 Engagement Restores Defective Effector CD8+ T Cell Function in Systemic Lupus Erythematosus. Arthritis Rheumatol. 2017, 69, 1035–1044. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. McKinney, E.F.; Lee, J.C.; Jayne, D.R.; Lyons, P.A.; Smith, K.G. T-cell exhaustion, co-stimulation and clinical outcome in autoimmunity and infection. Nature 2015, 523, 612–616. [Google Scholar] [CrossRef] [Green Version]
  58. Katsuyama, E.; Suarez-Fueyo, A.; Bradley, S.J.; Mizui, M.; Marin, A.V.; Mulki, L.; Krishfield, S.; Malavasi, F.; Yoon, J.; Sui, S.J.H.; 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.e114. [Google Scholar] [CrossRef] [Green Version]
  59. Li, X.; Kang, N.; Zhang, X.; Dong, X.; Wei, W.; Cui, L.; Ba, D.; He, W. Generation of human regulatory gammadelta T cells by TCRgammadelta stimulation in the presence of TGF-beta and their involvement in the pathogenesis of systemic lupus erythematosus. J. Immunol. 2011, 186, 6693–6700. [Google Scholar] [CrossRef] [Green Version]
  60. Robak, E.; Niewiadomska, H.; Robak, T.; Bartkowiak, J.; Blonski, J.Z.; Wozniacka, A.; Pomorski, L.; Sysa-Jedrezejowska, A. Lymphocyctes Tgammadelta in clinically normal skin and peripheral blood of patients with systemic lupus erythematosus and their correlation with disease activity. Mediat. Inflamm. 2001, 10, 179–189. [Google Scholar] [CrossRef] [Green Version]
  61. Brandt, L.; Hedberg, H. Impaired phagocytosis by peripheral blood granulocytes in systemic lupus erythematosus. Scand. J. Haematol. 1969, 6, 348–353. [Google Scholar] [CrossRef]
  62. Donnelly, S.; Roake, W.; Brown, S.; Young, P.; Naik, H.; Wordsworth, P.; Isenberg, D.A.; Reid, K.B.; Eggleton, P. Impaired recognition of apoptotic neutrophils by the C1q/calreticulin and CD91 pathway in systemic lupus erythematosus. Arthritis Rheum. 2006, 54, 1543–1556. [Google Scholar] [CrossRef]
  63. Lood, C.; Blanco, L.P.; Purmalek, M.M.; Carmona-Rivera, C.; De Ravin, S.S.; Smith, C.K.; Malech, H.L.; Ledbetter, J.A.; Elkon, K.B.; Kaplan, M.J. Neutrophil extracellular traps enriched in oxidized mitochondrial DNA are interferogenic and contribute to lupus-like disease. Nat. Med. 2016, 22, 146–153. [Google Scholar] [CrossRef] [Green Version]
  64. Hom, G.; Graham, R.R.; Modrek, B.; Taylor, K.E.; Ortmann, W.; Garnier, S.; Lee, A.T.; Chung, S.A.; Ferreira, R.C.; Pant, P.V.; et al. Association of systemic lupus erythematosus with C8orf13-BLK and ITGAM-ITGAX. N. Engl. J. Med. 2008, 358, 900–909. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Olsson, L.M.; Johansson, A.C.; Gullstrand, B.; Jonsen, A.; Saevarsdottir, S.; Ronnblom, L.; Leonard, D.; Wettero, J.; Sjowall, C.; Svenungsson, E.; et al. A single nucleotide polymorphism in the NCF1 gene leading to reduced oxidative burst is associated with systemic lupus erythematosus. Ann. Rheum. Dis. 2017, 76, 1607–1613. [Google Scholar] [CrossRef]
  66. Jacob, C.O.; Eisenstein, M.; Dinauer, M.C.; Ming, W.; Liu, Q.; John, S.; Quismorio, F.P., Jr.; Reiff, A.; Myones, B.L.; Kaufman, K.M.; et al. Lupus-associated causal mutation in neutrophil cytosolic factor 2 (NCF2) brings unique insights to the structure and function of NADPH oxidase. Proc. Natl. Acad. Sci. USA 2012, 109, E59–E67. [Google Scholar] [CrossRef] [Green Version]
  67. Lindau, D.; Mussard, J.; Rabsteyn, A.; Ribon, M.; Kotter, I.; Igney, A.; Adema, G.J.; Boissier, M.C.; Rammensee, H.G.; Decker, P. TLR9 independent interferon alpha production by neutrophils on NETosis in response to circulating chromatin, a key lupus autoantigen. Ann. Rheum. Dis. 2014, 73, 2199–2207. [Google Scholar] [CrossRef] [PubMed]
  68. Palanichamy, A.; Bauer, J.W.; Yalavarthi, S.; Meednu, N.; Barnard, J.; Owen, T.; Cistrone, C.; Bird, A.; Rabinovich, A.; Nevarez, S.; et al. Neutrophil-mediated IFN activation in the bone marrow alters B cell development in human and murine systemic lupus erythematosus. J. Immunol. 2014, 192, 906–918. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Rahman, S.; Sagar, D.; Hanna, R.N.; Lightfoot, Y.L.; Mistry, P.; Smith, C.K.; Manna, Z.; Hasni, S.; Siegel, R.M.; Sanjuan, M.A.; et al. Low-density granulocytes activate T cells and demonstrate a non-suppressive role in systemic lupus erythematosus. Ann. Rheum. Dis. 2019, 78, 957–966. [Google Scholar] [CrossRef] [Green Version]
  70. Midgley, A.; Beresford, M.W. Increased expression of low density granulocytes in juvenile-onset systemic lupus erythematosus patients correlates with disease activity. Lupus 2016, 25, 407–411. [Google Scholar] [CrossRef]
  71. Denny, M.F.; Yalavarthi, S.; Zhao, W.; Thacker, S.G.; Anderson, M.; Sandy, A.R.; McCune, W.J.; Kaplan, M.J. A distinct subset of proinflammatory neutrophils isolated from patients with systemic lupus erythematosus induces vascular damage and synthesizes type I IFNs. J. Immunol. 2010, 184, 3284–3297. [Google Scholar] [CrossRef] [Green Version]
  72. Boeltz, S.; Amini, P.; Anders, H.J.; Andrade, F.; Bilyy, R.; Chatfield, S.; Cichon, I.; Clancy, D.M.; Desai, J.; Dumych, T.; et al. To NET or not to NET:current opinions and state of the science regarding the formation of neutrophil extracellular traps. Cell Death Differ. 2019, 26, 395–408. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Apel, F.; Zychlinsky, A.; Kenny, E.F. The role of neutrophil extracellular traps in rheumatic diseases. Nat. Rev. Rheumatol. 2018, 14, 467–475. [Google Scholar] [CrossRef] [PubMed]
  74. Odqvist, L.; Jevnikar, Z.; Riise, R.; Oberg, L.; Rhedin, M.; Leonard, D.; Yrlid, L.; Jackson, S.; Mattsson, J.; Nanda, S.; et al. Genetic variations in A20 DUB domain provide a genetic link to citrullination and neutrophil extracellular traps in systemic lupus erythematosus. Ann. Rheum. Dis. 2019, 78, 1363–1370. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Chang, H.H.; Dwivedi, N.; Nicholas, A.P.; Ho, I.C. The W620 Polymorphism in PTPN22 Disrupts Its Interaction With Peptidylarginine Deiminase Type 4 and Enhances Citrullination and NETosis. Arthritis Rheumatol. 2015, 67, 2323–2334. [Google Scholar] [CrossRef] [PubMed]
  76. Li, D.; Matta, B.; Song, S.; Nelson, V.; Diggins, K.; Simpfendorfer, K.R.; Gregersen, P.K.; Linsley, P.; Barnes, B.J. IRF5 genetic risk variants drive myeloid-specific IRF5 hyperactivation and presymptomatic SLE. JCI Insight 2020, 5, e124020. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Goel, R.R.; Nakabo, S.; Dizon, B.L.P.; Urban, A.; Waldman, M.; Howard, L.; Darnell, D.; Buhaya, M.; Carmona-Rivera, C.; Hasni, S.; et al. Lupus-like autoimmunity and increased interferon response in patients with STAT3-deficient hyper-IgE syndrome. J. Allergy Clin. Immunol. 2021, 147, 746–749.e9. [Google Scholar] [CrossRef]
  78. Kahlenberg, J.M.; Carmona-Rivera, C.; Smith, C.K.; Kaplan, M.J. Neutrophil extracellular trap-associated protein activation of the NLRP3 inflammasome is enhanced in lupus macrophages. J. Immunol. 2013, 190, 1217–1226. [Google Scholar] [CrossRef] [Green Version]
  79. Sisirak, V.; Ganguly, D.; Lewis, K.L.; Couillault, C.; Tanaka, L.; Bolland, S.; D’Agati, V.; Elkon, K.B.; Reizis, B. Genetic evidence for the role of plasmacytoid dendritic cells in systemic lupus erythematosus. J. Exp. Med. 2014, 211, 1969–1976. [Google Scholar] [CrossRef] [Green Version]
  80. Fiore, N.; Castellano, G.; Blasi, A.; Capobianco, C.; Loverre, A.; Montinaro, V.; Netti, S.; Torres, D.; Manno, C.; Grandaliano, G.; et al. Immature myeloid and plasmacytoid dendritic cells infiltrate renal tubulointerstitium in patients with lupus nephritis. Mol. Immunol. 2008, 45, 259–265. [Google Scholar] [CrossRef]
  81. Sakata, K.; Nakayamada, S.; Miyazaki, Y.; Kubo, S.; Ishii, A.; Nakano, K.; Tanaka, Y. Up-Regulation of TLR7-Mediated IFN-alpha Production by Plasmacytoid Dendritic Cells in Patients With Systemic Lupus Erythematosus. Front. Immunol. 2018, 9, 1957. [Google Scholar] [CrossRef] [Green Version]
  82. Bave, U.; Magnusson, M.; Eloranta, M.L.; Perers, A.; Alm, G.V.; Ronnblom, L. Fc gamma RIIa is expressed on natural IFN-alpha-producing cells (plasmacytoid dendritic cells) and is required for the IFN-alpha production induced by apoptotic cells combined with lupus IgG. J. Immunol. 2003, 171, 3296–3302. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Garcia-Romo, G.S.; Caielli, S.; Vega, B.; Connolly, J.; Allantaz, F.; Xu, Z.; Punaro, M.; Baisch, J.; Guiducci, C.; Coffman, R.L.; et al. Netting neutrophils are major inducers of type I IFN production in pediatric systemic lupus erythematosus. Sci. Transl. Med. 2011, 3, 73ra20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Hjorton, K.; The DISSECT consortium; Hagberg, N.; Israelsson, E.; Jinton, L.; Berggren, O.; Sandling, J.K.; Thörn, K.; Mo, J.; Eloranta, M.-L.; et al. Cytokine production by activated plasmacytoid dendritic cells and natural killer cells is suppressed by an IRAK4 inhibitor. Arthritis Res. Ther. 2018, 20, 238. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Corzo, C.A.; Varfolomeev, E.; Setiadi, A.F.; Francis, R.; Klabunde, S.; Senger, K.; Sujatha-Bhaskar, S.; Drobnick, J.; Do, S.; Suto, E.; et al. The kinase IRAK4 promotes endosomal TLR and immune complex signaling in B cells and plasmacytoid dendritic cells. Sci. Signal. 2020, 13, eaaz1053. [Google Scholar] [CrossRef]
  86. Swiecki, M.; Colonna, M. The multifaceted biology of plasmacytoid dendritic cells. Nat. Rev. Immunol. 2015, 15, 471–485. [Google Scholar] [CrossRef]
  87. Jego, G.; Palucka, A.K.; Blanck, J.P.; Chalouni, C.; Pascual, V.; Banchereau, J. Plasmacytoid dendritic cells induce plasma cell differentiation through type I interferon and interleukin 6. Immunity 2003, 19, 225–234. [Google Scholar] [CrossRef] [Green Version]
  88. Soni, C.; Perez, O.A.; Voss, W.N.; Pucella, J.N.; Serpas, L.; Mehl, J.; Ching, K.L.; Goike, J.; Georgiou, G.; Ippolito, G.C.; et al. Plasmacytoid Dendritic Cells and Type I Interferon Promote Extrafollicular B Cell Responses to Extracellular Self-DNA. Immunity 2020, 52, 1022–1038.e1027. [Google Scholar] [CrossRef]
  89. Clement, M.; Charles, N.; Escoubet, B.; Guedj, K.; Chauveheid, M.P.; Caligiuri, G.; Nicoletti, A.; Papo, T.; Sacre, K. CD4+CXCR3+ T cells and plasmacytoid dendritic cells drive accelerated atherosclerosis associated with systemic lupus erythematosus. J. Autoimmun. 2015, 63, 59–67. [Google Scholar] [CrossRef]
  90. Ito, T.; Yang, M.; Wang, Y.H.; Lande, R.; Gregorio, J.; Perng, O.A.; Qin, X.F.; Liu, Y.J.; Gilliet, M. Plasmacytoid dendritic cells prime IL-10-producing T regulatory cells by inducible costimulator ligand. J. Exp. Med. 2007, 204, 105–115. [Google Scholar] [CrossRef]
  91. Jahrsdorfer, B.; Vollmer, A.; Blackwell, S.E.; Maier, J.; Sontheimer, K.; Beyer, T.; Mandel, B.; Lunov, O.; Tron, K.; Nienhaus, G.U.; et al. Granzyme B produced by human plasmacytoid dendritic cells suppresses T-cell expansion. Blood 2010, 115, 1156–1165. [Google Scholar] [CrossRef] [Green Version]
  92. Moseman, E.A.; Liang, X.; Dawson, A.J.; Panoskaltsis-Mortari, A.; Krieg, A.M.; Liu, Y.J.; Blazar, B.R.; Chen, W. Human plasmacytoid dendritic cells activated by CpG oligodeoxynucleotides induce the generation of CD4+CD25+ regulatory T cells. J. Immunol. 2004, 173, 4433–4442. [Google Scholar] [CrossRef] [Green Version]
  93. Menon, M.; Blair, P.A.; Isenberg, D.A.; Mauri, C. A Regulatory Feedback between Plasmacytoid Dendritic Cells and Regulatory B Cells Is Aberrant in Systemic Lupus Erythematosus. Immunity 2016, 44, 683–697. [Google Scholar] [CrossRef] [Green Version]
  94. Abrahams, J.P.; Leslie, A.G.; Lutter, R.; Walker, J.E. Structure at 2.8 A resolution of F1-ATPase from bovine heart mitochondria. Nature 1994, 370, 621–628. [Google Scholar] [CrossRef] [PubMed]
  95. Alam, K.; Moinuddin; Jabeen, S. Immunogenicity of mitochondrial DNA modified by hydroxyl radical. Cell. Immunol. 2007, 247, 12–17. [Google Scholar] [CrossRef] [PubMed]
  96. Chen, L.; Duvvuri, B.; Grigull, J.; Jamnik, R.; Wither, J.E.; Wu, G.E. Experimental evidence that mutated-self peptides derived from mitochondrial DNA somatic mutations have the potential to trigger autoimmunity. Hum. Immunol. 2014, 75, 873–879. [Google Scholar] [CrossRef]
  97. Lopez-Lopez, L.; Nieves-Plaza, M.; Castro Mdel, R.; Font, Y.M.; Torres-Ramos, C.A.; Vila, L.M.; Ayala-Pena, S. Mitochondrial DNA damage is associated with damage accrual and disease duration in patients with systemic lupus erythematosus. Lupus 2014, 23, 1133–1141. [Google Scholar] [CrossRef] [Green Version]
  98. Zhang, Q.; Raoof, M.; Chen, Y.; Sumi, Y.; Sursal, T.; Junger, W.; Brohi, K.; Itagaki, K.; Hauser, C.J. Circulating mitochondrial DAMPs cause inflammatory responses to injury. Nature 2010, 464, 104–107. [Google Scholar] [CrossRef] [Green Version]
  99. Vyshkina, T.; Sylvester, A.; Sadiq, S.; Bonilla, E.; Canter, J.A.; Perl, A.; Kalman, B. Association of common mitochondrial DNA variants with multiple sclerosis and systemic lupus erythematosus. Clin. Immunol. 2008, 129, 31–35. [Google Scholar] [CrossRef] [Green Version]
  100. Jonsen, A.; Yu, X.; Truedsson, L.; Nived, O.; Sturfelt, G.; Ibrahim, S.; Bengtsson, A. Mitochondrial DNA polymorphisms are associated with susceptibility and phenotype of systemic lupus erythematosus. Lupus 2009, 18, 309–312. [Google Scholar] [CrossRef]
  101. Tang, Y.; Wang, L.; Zhu, M.; Yang, M.; Zhong, K.; Du, Q.; Zhang, H.; Gui, M. Association of mtDNA M/N haplogroups with systemic lupus erythematosus: A case-control study of Han Chinese women. Sci. Rep. 2015, 5, 10817. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Yu, X.; Wester-Rosenlof, L.; Gimsa, U.; Holzhueter, S.A.; Marques, A.; Jonas, L.; Hagenow, K.; Kunz, M.; Nizze, H.; Tiedge, M.; et al. The mtDNA nt7778 G/T polymorphism affects autoimmune diseases and reproductive performance in the mouse. Hum. Mol. Genet. 2009, 18, 4689–4698. [Google Scholar] [CrossRef] [Green Version]
  103. Caielli, S.; Athale, S.; Domic, B.; Murat, E.; Chandra, M.; Banchereau, R.; Baisch, J.; Phelps, K.; Clayton, S.; Gong, M.; et al. Oxidized mitochondrial nucleoids released by neutrophils drive type I interferon production in human lupus. J. Exp. Med. 2016, 213, 697–713. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Becker, Y.; Marcoux, G.; Allaeys, I.; Julien, A.S.; Loignon, R.C.; Benk-Fortin, H.; Rollet-Labelle, E.; Rauch, J.; Fortin, P.R.; Boilard, E. Autoantibodies in Systemic Lupus Erythematosus Target Mitochondrial RNA. Front. Immunol. 2019, 10, 1026. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Sisirak, V.; Sally, B.; D’Agati, V.; Martinez-Ortiz, W.; Ozcakar, Z.B.; David, J.; Rashidfarrokhi, A.; Yeste, A.; Panea, C.; Chida, A.S.; et al. Digestion of Chromatin in Apoptotic Cell Microparticles Prevents Autoimmunity. Cell 2016, 166, 88–101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Al-Mayouf, S.M.; Sunker, A.; Abdwani, R.; Abrawi, S.A.; Almurshedi, F.; Alhashmi, N.; Al Sonbul, A.; Sewairi, W.; Qari, A.; Abdallah, E.; et al. Loss-of-function variant in DNASE1L3 causes a familial form of systemic lupus erythematosus. Nat. Genet. 2011, 43, 1186–1188. [Google Scholar] [CrossRef] [PubMed]
  107. Hartl, J.; Serpas, L.; Wang, Y.; Rashidfarrokhi, A.; Perez, O.A.; Sally, B.; Sisirak, V.; Soni, C.; Khodadadi-Jamayran, A.; Tsirigos, A.; et al. Autoantibody-mediated impairment of DNASE1L3 activity in sporadic systemic lupus erythematosus. J. Exp. Med. 2021, 218, e20201138. [Google Scholar] [CrossRef]
  108. Kawane, K.; Fukuyama, H.; Kondoh, G.; Takeda, J.; Ohsawa, Y.; Uchiyama, Y.; Nagata, S. Requirement of DNase II for definitive erythropoiesis in the mouse fetal liver. Science 2001, 292, 1546–1549. [Google Scholar] [CrossRef] [PubMed]
  109. Yoshida, H.; Okabe, Y.; Kawane, K.; Fukuyama, H.; Nagata, S. Lethal anemia caused by interferon-beta produced in mouse embryos carrying undigested DNA. Nat. Immunol. 2005, 6, 49–56. [Google Scholar] [CrossRef]
  110. Grieves, J.L.; Fye, J.M.; Harvey, S.; Grayson, J.M.; Hollis, T.; Perrino, F.W. Exonuclease TREX1 degrades double-stranded DNA to prevent spontaneous lupus-like inflammatory disease. Proc. Natl. Acad. Sci. USA 2015, 112, 5117–5122. [Google Scholar] [CrossRef] [Green Version]
  111. Crow, Y.J.; Hayward, B.E.; Parmar, R.; Robins, P.; Leitch, A.; Ali, M.; Black, D.N.; van Bokhoven, H.; Brunner, H.G.; Hamel, B.C.; et al. Mutations in the gene encoding the 3′-5′ DNA exonuclease TREX1 cause Aicardi-Goutieres syndrome at the AGS1 locus. Nat. Genet. 2006, 38, 917–920. [Google Scholar] [CrossRef]
  112. Shinde, R.; Hezaveh, K.; Halaby, M.J.; Kloetgen, A.; Chakravarthy, A.; da Silva Medina, T.; Deol, R.; Manion, K.P.; Baglaenko, Y.; Eldh, M.; et al. Apoptotic cell-induced AhR activity is required for immunological tolerance and suppression of systemic lupus erythematosus in mice and humans. Nat. Immunol. 2018, 19, 571–582. [Google Scholar] [CrossRef] [PubMed]
  113. Ravishankar, B.; Liu, H.; Shinde, R.; Chandler, P.; Baban, B.; Tanaka, M.; Munn, D.H.; Mellor, A.L.; Karlsson, M.C.; McGaha, T.L. Tolerance to apoptotic cells is regulated by indoleamine 2,3-dioxygenase. Proc. Natl. Acad. Sci. USA 2012, 109, 3909–3914. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Ravishankar, B.; Liu, H.; Shinde, R.; Chaudhary, K.; Xiao, W.; Bradley, J.; Koritzinsky, M.; Madaio, M.P.; McGaha, T.L. The amino acid sensor GCN2 inhibits inflammatory responses to apoptotic cells promoting tolerance and suppressing systemic autoimmunity. Proc. Natl. Acad. Sci. USA 2015, 112, 10774–10779. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Vaux, D.L.; Flavell, R.A. Apoptosis genes and autoimmunity. Curr. Opin. Immunol. 2000, 12, 719–724. [Google Scholar] [CrossRef]
  116. Henson, P.M.; Hume, D.A. Apoptotic cell removal in development and tissue homeostasis. Trends Immunol. 2006, 27, 244–250. [Google Scholar] [CrossRef]
  117. Devitt, A.; Parker, K.G.; Ogden, C.A.; Oldreive, C.; Clay, M.F.; Melville, L.A.; Bellamy, C.O.; Lacy-Hulbert, A.; Gangloff, S.C.; Goyert, S.M.; et al. Persistence of apoptotic cells without autoimmune disease or inflammation in CD14-/- mice. J. Cell Biol. 2004, 167, 1161–1170. [Google Scholar] [CrossRef]
  118. Lande, R.; Ganguly, D.; Facchinetti, V.; Frasca, L.; Conrad, C.; Gregorio, J.; Meller, S.; Chamilos, G.; Sebasigari, R.; Riccieri, V.; et al. Neutrophils activate plasmacytoid dendritic cells by releasing self-DNA-peptide complexes in systemic lupus erythematosus. Sci. Transl. Med. 2011, 3, 73ra19. [Google Scholar] [CrossRef] [Green Version]
  119. Crow, M.K.; Wohlgemuth, J. Microarray analysis of gene expression in lupus. Arthritis Res. Ther. 2003, 5, 279–287. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Ivashkiv, L.B. IFNgamma: Signalling, epigenetics and roles in immunity, metabolism, disease and cancer immunotherapy. Nat. Rev. Immunol. 2018, 18, 545–558. [Google Scholar] [CrossRef] [PubMed]
  121. Lee, S.; Baldridge, M.T. Interferon-Lambda: A Potent Regulator of Intestinal Viral Infections. Front. Immunol. 2017, 8, 749. [Google Scholar] [CrossRef] [Green Version]
  122. Ivashkiv, L.B.; Donlin, L.T. Regulation of type I interferon responses. Nat. Rev. Immunol. 2014, 14, 36–49. [Google Scholar] [CrossRef] [Green Version]
  123. Chyuan, I.T.; Tzeng, H.T.; Chen, J.Y. Signaling Pathways of Type I and Type III Interferons and Targeted Therapies in Systemic Lupus Erythematosus. Cells 2019, 8, 963. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Liu, Y.J. IPC: Professional type 1 interferon-producing cells and plasmacytoid dendritic cell precursors. Annu. Rev. Immunol. 2005, 23, 275–306. [Google Scholar] [CrossRef] [PubMed]
  125. Petro, T.M. IFN Regulatory Factor 3 in Health and Disease. J. Immunol. 2020, 205, 1981–1989. [Google Scholar] [CrossRef] [PubMed]
  126. Siegal, F.P.; Kadowaki, N.; Shodell, M.; Fitzgerald-Bocarsly, P.A.; Shah, K.; Ho, S.; Antonenko, S.; Liu, Y.J. The nature of the principal type 1 interferon-producing cells in human blood. Science 1999, 284, 1835–1837. [Google Scholar] [CrossRef] [Green Version]
  127. Schoggins, J.W. Interferon-Stimulated Genes: What Do They All Do? Annu. Rev. Virol. 2019, 6, 567–584. [Google Scholar] [CrossRef] [PubMed]
  128. Stark, G.R.; Darnell, J.E., Jr. The JAK-STAT pathway at twenty. Immunity 2012, 36, 503–514. [Google Scholar] [CrossRef] [Green Version]
  129. Steinberg, A.D.; Baron, S.; Talal, N. The pathogenesis of autoimmunity in New Zealand mice, I. Induction of antinucleic acid antibodies by polyinosinic-polycytidylic acid. Proc. Natl. Acad. Sci. USA 1969, 63, 1102–1107. [Google Scholar] [CrossRef] [Green Version]
  130. Okanoue, T.; Sakamoto, S.; Itoh, Y.; Minami, M.; Yasui, K.; Sakamoto, M.; Nishioji, K.; Katagishi, T.; Nakagawa, Y.; Tada, H.; et al. Side effects of high-dose interferon therapy for chronic hepatitis C. J. Hepatol. 1996, 25, 283–291. [Google Scholar] [CrossRef]
  131. Gota, C.; Calabrese, L. Induction of clinical autoimmune disease by therapeutic interferon-alpha. Autoimmunity 2003, 36, 511–518. [Google Scholar] [CrossRef]
  132. Niewold, T.B. Interferon alpha-induced lupus: Proof of principle. J. Clin. Rheumatol. 2008, 14, 131–132. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Niewold, T.B.; Swedler, W.I. Systemic lupus erythematosus arising during interferon-alpha therapy for cryoglobulinemic vasculitis associated with hepatitis C. Clin. Rheumatol. 2005, 24, 178–181. [Google Scholar] [CrossRef] [PubMed]
  134. Wilson, L.E.; Widman, D.; Dikman, S.H.; Gorevic, P.D. Autoimmune disease complicating antiviral therapy for hepatitis C virus infection. Semin. Arthritis Rheum. 2002, 32, 163–173. [Google Scholar] [CrossRef] [PubMed]
  135. Graham, R.R.; Kozyrev, S.V.; Baechler, E.C.; Reddy, M.V.; Plenge, R.M.; Bauer, J.W.; Ortmann, W.A.; Koeuth, T.; Gonzalez Escribano, M.F.; the Argentine and Spanish Collaborative Groups; et al. A common haplotype of interferon regulatory factor 5 (IRF5) regulates splicing and expression and is associated with increased risk of systemic lupus erythematosus. Nat. Genet. 2006, 38, 550–555. [Google Scholar] [CrossRef] [PubMed]
  136. International Consortium for Systemic Lupus Erythematosus Genetics (SLEGEN); Harley, J.B.; Alarcon-Riquelme, M.E.; Criswell, L.A.; Jacob, C.O.; Kimberly, R.P.; Moser, K.L.; Tsao, B.P.; Vyse, T.J.; Langefeld, C.D.; et al. Genome-wide association scan in women with systemic lupus erythematosus identifies susceptibility variants in ITGAM, PXK, KIAA1542 and other loci. Nat. Genet. 2008, 40, 204–210. [Google Scholar] [CrossRef] [PubMed]
  137. Lessard, C.J.; Adrianto, I.; Ice, J.A.; Wiley, G.B.; Kelly, J.A.; Glenn, S.B.; Adler, A.J.; Li, H.; Rasmussen, A.; Williams, A.H.; et al. Identification of IRF8, TMEM39A, and IKZF3-ZPBP2 as susceptibility loci for systemic lupus erythematosus in a large-scale multiracial replication study. Am. J. Hum. Genet. 2012, 90, 648–660. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Ghodke-Puranik, Y.; Niewold, T.B. Genetics of the type I interferon pathway in systemic lupus erythematosus. Int. J. Clin. Rheumtol. 2013, 8, 657–669. [Google Scholar] [CrossRef] [Green Version]
  139. van Vollenhoven, R.; Askanase, A.D.; Bomback, A.S.; Bruce, I.N.; Carroll, A.; Dall’Era, M.; Daniels, M.; Levy, R.A.; Schwarting, A.; Quasny, H.A.; et al. Conceptual framework for defining disease modification in systemic lupus erythematosus: A call for formal criteria. Lupus Sci. Med. 2022, 9, e000634. [Google Scholar] [CrossRef]
  140. van Vollenhoven, R.; Voskuyl, A.; Bertsias, G.; Aranow, C.; Aringer, M.; Arnaud, L.; Askanase, A.; Balazova, P.; Bonfa, E.; Bootsma, H.; et al. A framework for remission in SLE: Consensus findings from a large international task force on definitions of remission in SLE (DORIS). Ann. Rheum. Dis. 2017, 76, 554–561. [Google Scholar] [CrossRef]
  141. Franklyn, K.; Lau, C.S.; Navarra, S.V.; Louthrenoo, W.; Lateef, A.; Hamijoyo, L.; Wahono, C.S.; Chen, S.L.; Jin, O.; Morton, S.; et al. Definition and initial validation of a Lupus Low Disease Activity State (LLDAS). Ann. Rheum. Dis. 2016, 75, 1615–1621. [Google Scholar] [CrossRef]
  142. Almeida-Brasil, C.C.; Hanly, J.G.; Urowitz, M.; Clarke, A.E.; Ruiz-Irastorza, G.; Gordon, C.; Ramsey-Goldman, R.; Petri, M.; Ginzler, E.M.; Wallace, D.J.; et al. Flares after hydroxychloroquine reduction or discontinuation: Results from the Systemic Lupus International Collaborating Clinics (SLICC) inception cohort. Ann. Rheum. Dis. 2022, 81, 370–378. [Google Scholar] [CrossRef] [PubMed]
  143. Shinjo, S.K.; Bonfa, E.; Wojdyla, D.; Borba, E.F.; Ramirez, L.A.; Scherbarth, H.R.; Brenol, J.C.; Chacon-Diaz, R.; Neira, O.J.; Berbotto, G.A.; et al. Antimalarial treatment may have a time-dependent effect on lupus survival: Data from a multinational Latin American inception cohort. Arthritis Rheum. 2010, 62, 855–862. [Google Scholar] [CrossRef] [PubMed]
  144. Ruiz-Irastorza, G.; Ramos-Casals, M.; Brito-Zeron, P.; Khamashta, M.A. Clinical efficacy and side effects of antimalarials in systemic lupus erythematosus: A systematic review. Ann. Rheum. Dis. 2010, 69, 20–28. [Google Scholar] [CrossRef] [PubMed]
  145. Ugarte-Gil, M.F.; Mak, A.; Leong, J.; Dharmadhikari, B.; Kow, N.Y.; Reategui-Sokolova, C.; Elera-Fitzcarrald, C.; Aranow, C.; Arnaud, L.; Askanase, A.D.; et al. Impact of glucocorticoids on the incidence of lupus-related major organ damage: A systematic literature review and meta-regression analysis of longitudinal observational studies. Lupus Sci. Med. 2021, 8, e000590. [Google Scholar] [CrossRef] [PubMed]
  146. Anderka, M.T.; Lin, A.E.; Abuelo, D.N.; Mitchell, A.A.; Rasmussen, S.A. Reviewing the evidence for mycophenolate mofetil as a new teratogen: Case report and review of the literature. Am. J. Med. Genet. A 2009, 149A, 1241–1248. [Google Scholar] [CrossRef]
  147. Houssiau, F.A.; Vasconcelos, C.; D’Cruz, D.; Sebastiani, G.D.; Garrido, E.D.R.; Danieli, M.G.; Abramovicz, D.; Blockmans, D.; Mathieu, A.; Direskeneli, H.; et al. Immunosuppressive therapy in lupus nephritis: The Euro-Lupus Nephritis Trial, a randomized trial of low-dose versus high-dose intravenous cyclophosphamide. Arthritis Rheum. 2002, 46, 2121–2131. [Google Scholar] [CrossRef]
  148. Emadi, A.; Jones, R.J.; Brodsky, R.A. Cyclophosphamide and cancer: Golden anniversary. Nat. Rev. Clin. Oncol. 2009, 6, 638–647. [Google Scholar] [CrossRef]
  149. Dorner, T.; Furie, R. Novel paradigms in systemic lupus erythematosus. Lancet 2019, 393, 2344–2358. [Google Scholar] [CrossRef]
  150. Merrill, J.T.; Neuwelt, C.M.; Wallace, D.J.; Shanahan, J.C.; Latinis, K.M.; Oates, J.C.; Utset, T.O.; Gordon, C.; Isenberg, D.A.; Hsieh, H.J.; et al. Efficacy and safety of rituximab in moderately-to-severely active systemic lupus erythematosus: The randomized, double-blind, phase II/III systemic lupus erythematosus evaluation of rituximab trial. Arthritis Rheum. 2010, 62, 222–233. [Google Scholar] [CrossRef] [Green Version]
  151. Rovin, B.H.; Furie, R.; Latinis, K.; Looney, R.J.; Fervenza, F.C.; Sanchez-Guerrero, J.; Maciuca, R.; Zhang, D.; Garg, J.P.; Brunetta, P.; et al. Efficacy and safety of rituximab in patients with active proliferative lupus nephritis: The Lupus Nephritis Assessment with Rituximab study. Arthritis Rheum. 2012, 64, 1215–1226. [Google Scholar] [CrossRef]
  152. Teng, Y.K.; Levarht, E.W.; Toes, R.E.; Huizinga, T.W.; van Laar, J.M. Residual inflammation after rituximab treatment is associated with sustained synovial plasma cell infiltration and enhanced B cell repopulation. Ann. Rheum. Dis. 2009, 68, 1011–1016. [Google Scholar] [CrossRef] [PubMed]
  153. Pijpe, J.; Meijer, J.M.; Bootsma, H.; van der Wal, J.E.; Spijkervet, F.K.; Kallenberg, C.G.; Vissink, A.; Ihrler, S. Clinical and histologic evidence of salivary gland restoration supports the efficacy of rituximab treatment in Sjogren’s syndrome. Arthritis Rheum. 2009, 60, 3251–3256. [Google Scholar] [CrossRef]
  154. Kansal, R.; Richardson, N.; Neeli, I.; Khawaja, S.; Chamberlain, D.; Ghani, M.; Ghani, Q.U.; Balazs, L.; Beranova-Giorgianni, S.; Giorgianni, F.; et al. Sustained B cell depletion by CD19-targeted CAR T cells is a highly effective treatment for murine lupus. Sci. Transl. Med. 2019, 11, eaav1648. [Google Scholar] [CrossRef] [PubMed]
  155. Gomez Mendez, L.M.; Cascino, M.D.; Garg, J.; Katsumoto, T.R.; Brakeman, P.; Dall’Era, M.; Looney, R.J.; Rovin, B.; Dragone, L.; Brunetta, P. Peripheral Blood B Cell Depletion after Rituximab and Complete Response in Lupus Nephritis. Clin. J. Am. Soc. Nephrol. 2018, 13, 1502–1509. [Google Scholar] [CrossRef] [Green Version]
  156. Meyer, S.; Evers, M.; Jansen, J.H.M.; Buijs, J.; Broek, B.; Reitsma, S.E.; Moerer, P.; Amini, M.; Kretschmer, A.; Ten Broeke, T.; et al. New insights in Type I and II CD20 antibody mechanisms-of-action with a panel of novel CD20 antibodies. Br. J. Haematol. 2018, 180, 808–820. [Google Scholar] [CrossRef] [Green Version]
  157. Marinov, A.D.; Wang, H.; Bastacky, S.I.; van Puijenbroek, E.; Schindler, T.; Speziale, D.; Perro, M.; Klein, C.; Nickerson, K.M.; Shlomchik, M.J. The Type II Anti-CD20 Antibody Obinutuzumab (GA101) Is More Effective Than Rituximab at Depleting B Cells and Treating Disease in a Murine Lupus Model. Arthritis Rheumatol. 2021, 73, 826–836. [Google Scholar] [CrossRef] [PubMed]
  158. Mysler, E.F.; Spindler, A.J.; Guzman, R.; Bijl, M.; Jayne, D.; Furie, R.A.; Houssiau, F.A.; Drappa, J.; Close, D.; Maciuca, R.; et al. Efficacy and safety of ocrelizumab in active proliferative lupus nephritis: Results from a randomized, double-blind, phase III study. Arthritis Rheum. 2013, 65, 2368–2379. [Google Scholar] [CrossRef]
  159. Navarra, S.V.; Guzman, R.M.; Gallacher, A.E.; Hall, S.; Levy, R.A.; Jimenez, R.E.; Li, E.K.; Thomas, M.; Kim, H.Y.; Leon, M.G.; et al. Efficacy and safety of belimumab in patients with active systemic lupus erythematosus: A randomised, placebo-controlled, phase 3 trial. Lancet 2011, 377, 721–731. [Google Scholar] [CrossRef]
  160. Furie, R.; Petri, M.; Zamani, O.; Cervera, R.; Wallace, D.J.; Tegzova, D.; Sanchez-Guerrero, J.; Schwarting, A.; Merrill, J.T.; Chatham, W.W.; et al. A phase III, randomized, placebo-controlled study of belimumab, a monoclonal antibody that inhibits B lymphocyte stimulator, in patients with systemic lupus erythematosus. Arthritis Rheum. 2011, 63, 3918–3930. [Google Scholar] [CrossRef] [Green Version]
  161. van Vollenhoven, R.F.; Petri, M.A.; Cervera, R.; Roth, D.A.; Ji, B.N.; Kleoudis, C.S.; Zhong, Z.J.; Freimuth, W. Belimumab in the treatment of systemic lupus erythematosus: High disease activity predictors of response. Ann. Rheum. Dis. 2012, 71, 1343–1349. [Google Scholar] [CrossRef]
  162. Urowitz, M.B.; Ohsfeldt, R.L.; Wielage, R.C.; Kelton, K.A.; Asukai, Y.; Ramachandran, S. Organ damage in patients treated with belimumab versus standard of care: A propensity score-matched comparative analysis. Ann. Rheum. Dis. 2019, 78, 372–379. [Google Scholar] [CrossRef] [Green Version]
  163. Furie, R.A.; Wallace, D.J.; Aranow, C.; Fettiplace, J.; Wilson, B.; Mistry, P.; Roth, D.A.; Gordon, D. Long-Term Safety and Efficacy of Belimumab in Patients With Systemic Lupus Erythematosus: A Continuation of a Seventy-Six-Week Phase III Parent Study in the United States. Arthritis Rheumatol. 2018, 70, 868–877. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. van Vollenhoven, R.F.; Navarra, S.V.; Levy, R.A.; Thomas, M.; Heath, A.; Lustine, T.; Adamkovic, A.; Fettiplace, J.; Wang, M.L.; Ji, B.; et al. Long-term safety and limited organ damage in patients with systemic lupus erythematosus treated with belimumab: A Phase III study extension. Rheumatology 2020, 59, 281–291. [Google Scholar] [CrossRef] [Green Version]
  165. Stohl, W.; Schwarting, A.; Okada, M.; Scheinberg, M.; Doria, A.; Hammer, A.E.; Kleoudis, C.; Groark, J.; Bass, D.; Fox, N.L.; et al. Efficacy and Safety of Subcutaneous Belimumab in Systemic Lupus Erythematosus: A Fifty-Two-Week Randomized, Double-Blind, Placebo-Controlled Study. Arthritis Rheumatol. 2017, 69, 1016–1027. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Brunner, H.I.; Abud-Mendoza, C.; Viola, D.O.; Calvo Penades, I.; Levy, D.; Anton, J.; Calderon, J.E.; Chasnyk, V.G.; Ferrandiz, M.A.; Keltsev, V.; et al. Safety and efficacy of intravenous belimumab in children with systemic lupus erythematosus: Results from a randomised, placebo-controlled trial. Ann. Rheum. Dis. 2020, 79, 1340–1348. [Google Scholar] [CrossRef] [PubMed]
  167. Furie, R.; Rovin, B.H.; Houssiau, F.; Malvar, A.; Teng, Y.K.O.; Contreras, G.; Amoura, Z.; Yu, X.; Mok, C.C.; Santiago, M.B.; et al. Two-Year, Randomized, Controlled Trial of Belimumab in Lupus Nephritis. N. Engl. J. Med. 2020, 383, 1117–1128. [Google Scholar] [CrossRef]
  168. Rovin, B.H.; Furie, R.; Teng, Y.K.O.; Contreras, G.; Malvar, A.; Yu, X.; Ji, B.; Green, Y.; Gonzalez-Rivera, T.; Bass, D.; et al. A secondary analysis of the Belimumab International Study in Lupus Nephritis trial examined effects of belimumab on kidney outcomes and preservation of kidney function in patients with lupus nephritis. Kidney Int. 2022, 101, 403–413. [Google Scholar] [CrossRef]
  169. Tanaka, Y.; Tummala, R. Anifrolumab, a monoclonal antibody to the type I interferon receptor subunit 1, for the treatment of systemic lupus erythematosus: An overview from clinical trials. Mod. Rheumatol. 2021, 31, 1–12. [Google Scholar] [CrossRef]
  170. Furie, R.; Khamashta, M.; Merrill, J.T.; Werth, V.P.; Kalunian, K.; Brohawn, P.; Illei, G.G.; Drappa, J.; Wang, L.; Yoo, S.; et al. Anifrolumab, an Anti-Interferon-alpha Receptor Monoclonal Antibody, in Moderate-to-Severe Systemic Lupus Erythematosus. Arthritis Rheumatol. 2017, 69, 376–386. [Google Scholar] [CrossRef] [Green Version]
  171. Furie, R.A.; Morand, E.F.; Bruce, I.N.; Manzi, S.; Kalunian, K.C.; Vital, E.M.; Ford, T.L.; Gupta, R.; Hiepe, F.; Santiago, M. Type I interferon inhibitor anifrolumab in active systemic lupus erythematosus (TULIP-1): A randomised, controlled, phase 3 trial. Lancet Rheumatol. 2019, 1, e208–e219. [Google Scholar] [CrossRef]
  172. Morand, E.F.; Furie, R.; Tanaka, Y.; Bruce, I.N.; Askanase, A.D.; Richez, C.; Bae, S.C.; Brohawn, P.Z.; Pineda, L.; Berglind, A.; et al. Trial of Anifrolumab in Active Systemic Lupus Erythematosus. N. Engl. J. Med. 2020, 382, 211–221. [Google Scholar] [CrossRef] [PubMed]
  173. Furie, R.; Morand, E.F.; Askanase, A.D.; Vital, E.M.; Merrill, J.T.; Kalyani, R.N.; Abreu, G.; Pineda, L.; Tummala, R. Anifrolumab reduces flare rates in patients with moderate to severe systemic lupus erythematosus. Lupus 2021, 30, 1254–1263. [Google Scholar] [CrossRef]
  174. Vital, E.M.; Merrill, J.T.; Morand, E.F.; Furie, R.A.; Bruce, I.N.; Tanaka, Y.; Manzi, S.; Kalunian, K.C.; Kalyani, R.N.; Streicher, K.; et al. Anifrolumab efficacy and safety by type I interferon gene signature and clinical subgroups in patients with SLE: Post hoc analysis of pooled data from two phase III trials. Ann. Rheum. Dis. 2022, 81, 951–961. [Google Scholar] [CrossRef] [PubMed]
  175. Jayne, D.; Rovin, B.; Mysler, E.F.; Furie, R.A.; Houssiau, F.A.; Trasieva, T.; Knagenhjelm, J.; Schwetje, E.; Chia, Y.L.; Tummala, R.; et al. Phase II randomised trial of type I interferon inhibitor anifrolumab in patients with active lupus nephritis. Ann. Rheum. Dis. 2022, 81, 496–506. [Google Scholar] [CrossRef] [PubMed]
  176. Heo, Y.A. Voclosporin: First Approval. Drugs 2021, 81, 605–610. [Google Scholar] [CrossRef]
  177. Rovin, B.H.; Teng, Y.K.O.; Ginzler, E.M.; Arriens, C.; Caster, D.J.; Romero-Diaz, J.; Gibson, K.; Kaplan, J.; Lisk, L.; Navarra, S.; et al. Efficacy and safety of voclosporin versus placebo for lupus nephritis (AURORA 1): A double-blind, randomised, multicentre, placebo-controlled, phase 3 trial. Lancet 2021, 397, 2070–2080. [Google Scholar] [CrossRef] [PubMed]
  178. Rovin, B.H.; Solomons, N.; Pendergraft, W.F., 3rd; Dooley, M.A.; Tumlin, J.; Romero-Diaz, J.; Lysenko, L.; Navarra, S.V.; Huizinga, R.B.; Group, A.-L.S. A randomized, controlled double-blind study comparing the efficacy and safety of dose-ranging voclosporin with placebo in achieving remission in patients with active lupus nephritis. Kidney Int. 2019, 95, 219–231. [Google Scholar] [CrossRef]
  179. Scholz, J.L.; Crowley, J.E.; Tomayko, M.M.; Steinel, N.; O’Neill, P.J.; Quinn, W.J., 3rd; Goenka, R.; Miller, J.P.; Cho, Y.H.; Long, V.; et al. BLyS inhibition eliminates primary B cells but leaves natural and acquired humoral immunity intact. Proc. Natl. Acad. Sci. USA 2008, 105, 15517–15522. [Google Scholar] [CrossRef] [Green Version]
  180. Neubert, K.; Meister, S.; Moser, K.; Weisel, F.; Maseda, D.; Amann, K.; Wiethe, C.; Winkler, T.H.; Kalden, J.R.; Manz, R.A.; et al. The proteasome inhibitor bortezomib depletes plasma cells and protects mice with lupus-like disease from nephritis. Nat. Med. 2008, 14, 748–755. [Google Scholar] [CrossRef]
  181. Ishii, T.; Tanaka, Y.; Kawakami, A.; Saito, K.; Ichinose, K.; Fujii, H.; Shirota, Y.; Shirai, T.; Fujita, Y.; Watanabe, R.; et al. Multicenter double-blind randomized controlled trial to evaluate the effectiveness and safety of bortezomib as a treatment for refractory systemic lupus erythematosus. Mod. Rheumatol. 2018, 28, 986–992. [Google Scholar] [CrossRef]
  182. Ladi, E.; Everett, C.; Stivala, C.E.; Daniels, B.E.; Durk, M.R.; Harris, S.F.; Huestis, M.P.; Purkey, H.E.; Staben, S.T.; Augustin, M.; et al. Design and Evaluation of Highly Selective Human Immunoproteasome Inhibitors Reveal a Compensatory Process That Preserves Immune Cell Viability. J. Med. Chem. 2019, 62, 7032–7041. [Google Scholar] [CrossRef] [PubMed]
  183. Anderson, K.C. Progress and Paradigms in Multiple Myeloma. Clin. Cancer Res. 2016, 22, 5419–5427. [Google Scholar] [CrossRef] [Green Version]
  184. Ostendorf, L.; Burns, M.; Durek, P.; Heinz, G.A.; Heinrich, F.; Garantziotis, P.; Enghard, P.; Richter, U.; Biesen, R.; Schneider, U.; et al. Targeting CD38 with Daratumumab in Refractory Systemic Lupus Erythematosus. N. Engl. J. Med. 2020, 383, 1149–1155. [Google Scholar] [CrossRef] [PubMed]
  185. Jenks, S.A.; Cashman, K.S.; Zumaquero, E.; Marigorta, U.M.; Patel, A.V.; Wang, X.; Tomar, D.; Woodruff, M.C.; Simon, Z.; Bugrovsky, R.; et al. Distinct Effector B Cells Induced by Unregulated Toll-like Receptor 7 Contribute to Pathogenic Responses in Systemic Lupus Erythematosus. Immunity 2018, 49, 725–739.e726. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Duarte, R.F.; Labopin, M.; Bader, P.; Basak, G.W.; Bonini, C.; Chabannon, C.; Corbacioglu, S.; Dreger, P.; Dufour, C.; Gennery, A.R.; et al. Indications for haematopoietic stem cell transplantation for haematological diseases, solid tumours and immune disorders: Current practice in Europe, 2019. Bone Marrow Transpl. 2019, 54, 1525–1552. [Google Scholar] [CrossRef] [PubMed]
  187. Alexander, T.; Greco, R.; Snowden, J.A. Hematopoietic Stem Cell Transplantation for Autoimmune Disease. Annu. Rev. Med. 2021, 72, 215–228. [Google Scholar] [CrossRef]
  188. Zand, M.S.; Vo, T.; Pellegrin, T.; Felgar, R.; Liesveld, J.L.; Ifthikharuddin, J.J.; Abboud, C.N.; Sanz, I.; Huggins, J. Apoptosis and complement-mediated lysis of myeloma cells by polyclonal rabbit antithymocyte globulin. Blood 2006, 107, 2895–2903. [Google Scholar] [CrossRef]
  189. Alexander, T.; Arnold, R.; Hiepe, F.; Radbruch, A. Resetting the immune system with immunoablation and autologous haematopoietic stem cell transplantation in autoimmune diseases. Clin. Exp. Rheumatol. 2016, 34, 53–57. [Google Scholar]
  190. Swart, J.F.; Delemarre, E.M.; van Wijk, F.; Boelens, J.J.; Kuball, J.; van Laar, J.M.; Wulffraat, N.M. Haematopoietic stem cell transplantation for autoimmune diseases. Nat. Rev. Rheumatol. 2017, 13, 244–256. [Google Scholar] [CrossRef]
  191. Shifa, I.; Hazlewood, G.S.; Durand, C.; Barr, S.G.; Mydlarski, P.R.; Beck, P.L.; Burton, J.M.; Khan, F.M.; Jamani, K.; Osman, M.; et al. Efficacy of Allogeneic Hematopoietic Cell Transplantation for Autoimmune Diseases. Transpl. Cell. Ther. 2021, 27, 489.e1–489.e9. [Google Scholar] [CrossRef]
  192. Burt, R.K.; Han, X.; Gozdziak, P.; Yaung, K.; Morgan, A.; Clendenan, A.M.; Henry, J.; Calvario, M.A.; Datta, S.K.; Helenowski, I.; et al. Five year follow-up after autologous peripheral blood hematopoietic stem cell transplantation for refractory, chronic, corticosteroid-dependent systemic lupus erythematosus: Effect of conditioning regimen on outcome. Bone Marrow Transpl. 2018, 53, 692–700. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Khorshid, O.; Hosing, C.; Bibawi, S.; Ueno, N.; Reveille, J.; Mayes, M.D.; Champlin, R.E. Nonmyeloablative stem cell transplant in a patient with advanced systemic sclerosis and systemic lupus erythematosus. J. Rheumatol. 2004, 31, 2513–2516. [Google Scholar] [PubMed]
  194. Lu, Q.; Lu, L.; Niu, X.; Guo, Y.; Parino, G.R.; Liu, D. Non-myeloablative allogeneic stem cell transplant in a patient with refractory systemic lupus erythematosus. Bone Marrow Transpl. 2006, 37, 979–981. [Google Scholar] [CrossRef] [Green Version]
  195. Marmont, A.M.; Bacigalupo, A.; Gualandi, F.; Bregante, S.; van Lint, M.T.; Geroldi, S. Systemic lupus erythematosus complicated with thymoma and pure red cell aplasia (PCRA). CR of both complications following thymectomy and allogeneic haematopoietic SCT (HSCT), but persistence of antinuclear antibodies (ANA). Bone Marrow Transpl. 2014, 49, 982–983. [Google Scholar] [CrossRef] [Green Version]
  196. Greco, R.; Labopin, M.; Badoglio, M.; Veys, P.; Furtado Silva, J.M.; Abinun, M.; Gualandi, F.; Bornhauser, M.; Ciceri, F.; Saccardi, R.; et al. Allogeneic HSCT for Autoimmune Diseases: A Retrospective Study From the EBMT ADWP, IEWP, and PDWP Working Parties. Front. Immunol. 2019, 10, 1570. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Eshhar, Z.; Waks, T.; Gross, G.; Schindler, D.G. Specific activation and targeting of cytotoxic lymphocytes through chimeric single chains consisting of antibody-binding domains and the gamma or zeta subunits of the immunoglobulin and T-cell receptors. Proc. Natl. Acad. Sci. USA 1993, 90, 720–724. [Google Scholar] [CrossRef] [Green Version]
  198. Jiang, X.; Chen, X.; Carpenter, T.J.; Wang, J.; Zhou, R.; Davis, H.M.; Heald, D.L.; Wang, W. Development of a Target cell-Biologics-Effector cell (TBE) complex-based cell killing model to characterize target cell depletion by T cell redirecting bispecific agents. MAbs 2018, 10, 876–889. [Google Scholar] [CrossRef]
  199. Ellebrecht, C.T.; Bhoj, V.G.; Nace, A.; Choi, E.J.; Mao, X.; Cho, M.J.; Di Zenzo, G.; Lanzavecchia, A.; Seykora, J.T.; Cotsarelis, G.; et al. Reengineering chimeric antigen receptor T cells for targeted therapy of autoimmune disease. Science 2016, 353, 179–184. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  200. Sadelain, M.; Riviere, I.; Riddell, S. Therapeutic T cell engineering. Nature 2017, 545, 423–431. [Google Scholar] [CrossRef] [Green Version]
  201. Hudecek, M.; Sommermeyer, D.; Kosasih, P.L.; Silva-Benedict, A.; Liu, L.; Rader, C.; Jensen, M.C.; Riddell, S.R. The nonsignaling extracellular spacer domain of chimeric antigen receptors is decisive for in vivo antitumor activity. Cancer Immunol. Res. 2015, 3, 125–135. [Google Scholar] [CrossRef] [Green Version]
  202. Casucci, M.; Falcone, L.; Camisa, B.; Norelli, M.; Porcellini, S.; Stornaiuolo, A.; Ciceri, F.; Traversari, C.; Bordignon, C.; Bonini, C.; et al. Extracellular NGFR Spacers Allow Efficient Tracking and Enrichment of Fully Functional CAR-T Cells Co-Expressing a Suicide Gene. Front. Immunol. 2018, 9, 507. [Google Scholar] [CrossRef] [Green Version]
  203. Hayden, P.J.; Roddie, C.; Bader, P.; Basak, G.W.; Bonig, H.; Bonini, C.; Chabannon, C.; Ciceri, F.; Corbacioglu, S.; Ellard, R.; et al. Management of adults and children receiving CAR T-cell therapy: 2021 best practice recommendations of the European Society for Blood and Marrow Transplantation (EBMT) and the Joint Accreditation Committee of ISCT and EBMT (JACIE) and the European Haematology Association (EHA). Ann. Oncol. 2022, 33, 259–275. [Google Scholar] [CrossRef] [PubMed]
  204. Mougiakakos, D.; Kronke, G.; Volkl, S.; Kretschmann, S.; Aigner, M.; Kharboutli, S.; Boltz, S.; Manger, B.; Mackensen, A.; Schett, G. CD19-Targeted CAR T Cells in Refractory Systemic Lupus Erythematosus. N. Engl. J. Med. 2021, 385, 567–569. [Google Scholar] [CrossRef]
  205. Mackensen, A.; Muller, F.; Mougiakakos, D.; Boltz, S.; Wilhelm, A.; Aigner, M.; Volkl, S.; Simon, D.; Kleyer, A.; Munoz, L.; et al. Anti-CD19 CAR T cell therapy for refractory systemic lupus erythematosus. Nat. Med. 2022, 28, 2124–2132. [Google Scholar] [CrossRef]
  206. Jarczak, D.; Nierhaus, A. Cytokine Storm-Definition, Causes, and Implications. Int. J. Mol. Sci. 2022, 23, 11740. [Google Scholar] [CrossRef]
  207. Ferreira, L.M.R.; Muller, Y.D.; Bluestone, J.A.; Tang, Q. Next-generation regulatory T cell therapy. Nat. Rev. Drug Discov. 2019, 18, 749–769. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  208. Raffin, C.; Vo, L.T.; Bluestone, J.A. T(reg) cell-based therapies: Challenges and perspectives. Nat. Rev. Immunol. 2020, 20, 158–172. [Google Scholar] [CrossRef] [PubMed]
  209. Furie, R.; Werth, V.P.; Merola, J.F.; Stevenson, L.; Reynolds, T.L.; Naik, H.; Wang, W.; Christmann, R.; Gardet, A.; Pellerin, A.; et al. Monoclonal antibody targeting BDCA2 ameliorates skin lesions in systemic lupus erythematosus. J. Clin. Investig. 2019, 129, 1359–1371. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  210. Yuan, X.; Qin, X.; Wang, D.; Zhang, Z.; Tang, X.; Gao, X.; Chen, W.; Sun, L. Mesenchymal stem cell therapy induces FLT3L and CD1c(+) dendritic cells in systemic lupus erythematosus patients. Nat. Commun. 2019, 10, 2498. [Google Scholar] [CrossRef] [Green Version]
  211. Knight, J.S.; Subramanian, V.; O’Dell, A.A.; Yalavarthi, S.; Zhao, W.; Smith, C.K.; Hodgin, J.B.; Thompson, P.R.; Kaplan, M.J. Peptidylarginine deiminase inhibition disrupts NET formation and protects against kidney, skin and vascular disease in lupus-prone MRL/lpr mice. Ann. Rheum. Dis. 2015, 74, 2199–2206. [Google Scholar] [CrossRef] [Green Version]
  212. Mok, C.C. The Jakinibs in systemic lupus erythematosus: Progress and prospects. Expert Opin. Investig. Drugs 2019, 28, 85–92. [Google Scholar] [CrossRef] [PubMed]
  213. Robinson, S.; Thomas, R. Potential for Antigen-Specific Tolerizing Immunotherapy in Systematic Lupus Erythematosus. Front. Immunol. 2021, 12, 654701. [Google Scholar] [CrossRef] [PubMed]
Figure 1. SLE classification according to EULAR/ACR 2019 criteria and measurement of disease activity and organ damage. The entry criterion consists of antinuclear antibody (ANA) positivity on at least one occasion. Definitive diagnosis requires 10 points derived from clinical and immunologic criteria. Disease activity is assessed by the SLE Disease Activity Index-2000 (SLEDAI-2K), British Isles Lupus Activity Group 2004 (BILAG 2004), and Physician Global Assessment (PGA). Organ damage is assessed by the SLICC/ACR damage index (SDI).
Figure 1. SLE classification according to EULAR/ACR 2019 criteria and measurement of disease activity and organ damage. The entry criterion consists of antinuclear antibody (ANA) positivity on at least one occasion. Definitive diagnosis requires 10 points derived from clinical and immunologic criteria. Disease activity is assessed by the SLE Disease Activity Index-2000 (SLEDAI-2K), British Isles Lupus Activity Group 2004 (BILAG 2004), and Physician Global Assessment (PGA). Organ damage is assessed by the SLICC/ACR damage index (SDI).
Ijms 24 06578 g001
Figure 2. Both innate and adaptive immunity participate in the pathogenesis of SLE. In individuals with a genetic predisposition, and with the contribution of environmental factors, there is an accumulation of apoptotic cells and activation of NET production, by neutrophils. Cell nucleic acids are then exposed to the immune system and activate dendritic cells via toll-like receptors, to produce type-I IFN. This cytokine is responsible for activation of specific genes for pro-inflammatory factors by target cells (IFN signature). Dendritic cells also produce BAFF, which is necessary for B-cell activation and survival, and activate T cells, through the presentation of nuclear self-antigens. Self-reactive B cells are then activated to produce antibodies, and differentiate into long-lived plasma cells, localized in niches in the bone marrow, which are an additional source of autoantibodies. Autoantibodies form immunocomplexes with specific nuclear self-antigens that precipitate in tissues, contributing to organ damage.
Figure 2. Both innate and adaptive immunity participate in the pathogenesis of SLE. In individuals with a genetic predisposition, and with the contribution of environmental factors, there is an accumulation of apoptotic cells and activation of NET production, by neutrophils. Cell nucleic acids are then exposed to the immune system and activate dendritic cells via toll-like receptors, to produce type-I IFN. This cytokine is responsible for activation of specific genes for pro-inflammatory factors by target cells (IFN signature). Dendritic cells also produce BAFF, which is necessary for B-cell activation and survival, and activate T cells, through the presentation of nuclear self-antigens. Self-reactive B cells are then activated to produce antibodies, and differentiate into long-lived plasma cells, localized in niches in the bone marrow, which are an additional source of autoantibodies. Autoantibodies form immunocomplexes with specific nuclear self-antigens that precipitate in tissues, contributing to organ damage.
Ijms 24 06578 g002
Table 1. Major clinical trials of new drugs currently used for SLE therapy.
Table 1. Major clinical trials of new drugs currently used for SLE therapy.
DrugTargetMolecular StructureTrialDosing Primary Endpoint
(PE)
ResultFDA Approval
RituximabPan-B-cell marker CD20 Chimeric mAbEXPLORER1000 mg or placebo on days 1, 15, 168, and 182BILAG response versus placebo at week 52PE not metNo
Rituximab__LUNAR1000 mg or placebo on days 1, 15, 168, and 182Complete or partial response at week 52 in LN patients PE not metNo
BelimumabBAFFHuman mAb IgG-1 lambaBLISS-5210 mg/kg or 1 mg/kg or placebo every 4 weeksSRI-4 response versus placebo at week 52PE met_
Belimumab__BLISS-7610 mg/kg or 1 mg/kg i.v. or placebo every 4 weeksSRI-4 response versus placebo at week 52PE metYes (adults with ANA+, active SLE plus standard therapy)
Belimumab__BLISS-SC200 mg s.c. weeklySRI-4 response versus placebo at week 52PE metYes (adults with ANA+, active SLE plus standard therapy)
Belimumab__BLISS-LN10 mg/kg i.v. or placebo every 4 weeksPERR at week 104 in patients with active LNPE metYes (adults with active LN plus standard therapy)
Belimumab__PLUTO10 mg/kg i.v. or placebo every 4 weeksSRI-4 response versus placebo at week 52 in children aged 5 to 17 yearsPE metYes (children 5 years and older with SLE and LN plus standard therapy)
AnifrolumabType-I IFNR
subunit 1
Human mAb IgG-1 kappaTULIP-1300 mg or 150 mg or placebo
every 4 weeks
SRI-4 response of anifrolumab 300 mg versus placebo at week 52PE not met_
Anifrolumab__TULIP-2300 mg or placebo every 4 weeks BICLA response of anifrolumab 300 mg versus placebo at week 52PE metYes (adults with moderate to severe SLE plus standard therapy)
VoclosporinT-cell inhibition and kidney podocytes stabilizationCalcineurin inhibitionAURORAVoclosporin 23.7 mg twice daily + MM 1 g daily or MM 1 g daily CRR in voclosporin + MM versus MM alone at week 52PE metYes (adults with active LN plus standard therapy)
mAb = monoclonal antibody; BAFF = B-cell activating factor; IFNR = interferon receptor; PERR = primary efficacy renal response; CRR = complete renal response.
Table 2. Promising future therapies for SLE.
Table 2. Promising future therapies for SLE.
DrugMechanism of ActionReference
OcrelizumabCD20+ B-cell depletion [158]
BortezomibProteasome inhibition of LLPCs[181]
CAR T cellsCD19+ B-cell depletion[205]
BDCA2Anti-pDC antibody[209]
IL-2Treg enhancement[208]
JAK inhibitorsType-I and type-II IFN signaling inhibition[212]
LLPCs = long-lived plasma cells; pDC = plasmacytoid dendritic cells; Treg = T regulatory cells; JAK = Janus kinase.
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

Accapezzato, D.; Caccavale, R.; Paroli, M.P.; Gioia, C.; Nguyen, B.L.; Spadea, L.; Paroli, M. Advances in the Pathogenesis and Treatment of Systemic Lupus Erythematosus. Int. J. Mol. Sci. 2023, 24, 6578. https://doi.org/10.3390/ijms24076578

AMA Style

Accapezzato D, Caccavale R, Paroli MP, Gioia C, Nguyen BL, Spadea L, Paroli M. Advances in the Pathogenesis and Treatment of Systemic Lupus Erythematosus. International Journal of Molecular Sciences. 2023; 24(7):6578. https://doi.org/10.3390/ijms24076578

Chicago/Turabian Style

Accapezzato, Daniele, Rosalba Caccavale, Maria Pia Paroli, Chiara Gioia, Bich Lien Nguyen, Luca Spadea, and Marino Paroli. 2023. "Advances in the Pathogenesis and Treatment of Systemic Lupus Erythematosus" International Journal of Molecular Sciences 24, no. 7: 6578. https://doi.org/10.3390/ijms24076578

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