Next Article in Journal
Perception of Pathologists in Poland of Artificial Intelligence and Machine Learning in Medical Diagnosis—A Cross-Sectional Study
Next Article in Special Issue
Immune Portrayal of a New Therapy Targeting Microbiota in an Animal Model of Psoriasis
Previous Article in Journal
Prevalence of Sarcopenia in Older Patients in Rehabilitation Wards
Previous Article in Special Issue
Causes of Polyserositis: A Systematic Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Spectrum of Extraglandular Manifestations in Primary Sjögren’s Syndrome

by
Ancuta Mihai
1,2,
Constantin Caruntu
3,4,
Ciprian Jurcut
1,
Florin Cristian Blajut
5,6,
Mihnea Casian
7,8,
Daniela Opris-Belinski
9,10,
Ruxandra Ionescu
9 and
Ana Caruntu
11,12,*
1
Department of Internal Medicine, Carol Davila Central Military Emergency Hospital, 010825 Bucharest, Romania
2
Department of Rheumatology, Faculty of General Medicine, Titu Maiorescu University, 031593 Bucharest, Romania
3
Department of Physiology, Carol Davila University of Medicine and Pharmacy, 020021 Bucharest, Romania
4
Department of Dermatology, Prof. N.C. Paulescu National Institute of Diabetes, Nutrition and Metabolic Diseases, 011233 Bucharest, Romania
5
Department of General Surgery, Carol Davila Central Military Emergency Hospital, 010825 Bucharest, Romania
6
Department of Medical-Surgical Specialties, “Titu Maiorescu” University of Bucharest, 040441 Bucharest, Romania
7
Emergency Institute for Cardiovascular Diseases Prof. Dr. C.C. Iliescu, 022328 Bucharest, Romania
8
Department of Cardiology, Carol Davila University of Medicine and Pharmacy, 020021 Bucharest, Romania
9
Internal Medicine and Rheumatology Department, Sfanta Maria Clinical Hospital, 011172 Bucharest, Romania
10
Internal Medicine and Rheumatology Department, Carol Davila University of Medicine and Pharmacy, 020021 Bucharest, Romania
11
Department of Oral and Maxillofacial Surgery, Carol Davila Central Military Emergency Hospital, 010825 Bucharest, Romania
12
Department of Oral and Maxillofacial Surgery, Faculty of Dental Medicine, Titu Maiorescu University, 031593 Bucharest, Romania
*
Author to whom correspondence should be addressed.
J. Pers. Med. 2023, 13(6), 961; https://doi.org/10.3390/jpm13060961
Submission received: 30 April 2023 / Revised: 2 June 2023 / Accepted: 5 June 2023 / Published: 7 June 2023
(This article belongs to the Special Issue The Challenges and Therapeutic Prospects in Autoimmune Diseases)

Abstract

:
Extraglandular manifestations (EGMs) in primary Sjogren’s syndrome (pSS) represent the clinical expression of the systemic involvement in this disease. EGMs are characterized by a wide heterogeneity; virtually any organ or system can be affected, with various degrees of dysfunction. The existing gaps of knowledge in this complex domain of extraglandular extension in pSS need to be overcome in order to increase the diagnostic accuracy of EGMs in pSS. The timely identification of EGMs, as early as from subclinical stages, can be facilitated using highly specific biomarkers, thus preventing decompensated disease and severe complications. To date, there is no general consensus on the diagnostic criteria for the wide range of extraglandular involvement in pSS, which associates important underdiagnosing of EGMs, subsequent undertreatment and progression to severe organ dysfunction in these patients. This review article presents the most recent basic and clinical science research conducted to investigate pathogenic mechanisms leading to EGMs in pSS patients. In addition, it presents the current diagnostic and treatment recommendations and the trends for future therapeutic strategies based on personalized treatment, as well as the latest research in the field of diagnostic and prognostic biomarkers for extraglandular involvement in pSS.

1. Introduction

Primary Sjögren Syndrome (pSS) is a systemic chronic autoimmune rheumatic disorder of unknown etiology, characterized by lymphocytic infiltration with immune-mediated destruction of exocrine glands, primarily including salivary and lacrimal glands [1]. The dryness may affect other mucosal surfaces such as the airways, digestive tract, and vagina, leading to the clinical picture of the sicca syndrome [2]. In addition to these, pSS can involve any organ system, expressed in various and complex clinical extraglandular manifestations (EGMs). Extraglandular symptoms are divided into non-visceral, represented by musculoskeletal and cutaneous manifestations, and visceral symptoms, which include neurological, renal, hematological, pulmonary, gastrointestinal, and cardiovascular manifestations (Figure 1) [3]. The clinical expression of pSS is similar to secondary Sjögren Syndrome (sSS), which is characterized by simultaneous association with other autoimmune diseases, such as rheumatoid arthritis (RA), systemic sclerosis (SSc) or systemic lupus erythematosus (SLE), or with concomitant organ-specific autoimmune diseases, such as autoimmune thyroiditis, primary biliary cholangitis (PBC), and autoimmune hepatitis (AIH) [4]. Therefore, the distinction between pSS and sSS is reflected only by the overlap with other autoimmune diseases. The patients’ management is similar in both types of disease [1].
The classification criteria for pSS were published in 2016 by ACR/EULAR (American College of Rheumatology/European League against Rheumatism) [5]. These classification criteria apply to any patient with symptoms of ocular or oral dryness according to American European Consensus Group criteria (AECG) questions or to patients with the positivity of at least one of the domains of the EULAR Sjögren’s syndrome disease activity index (ESSDAI) questionnaire. The pSS diagnostic is considered in subjects that have a total score ≥ 4, obtained of the five items: anti-SSA/Ro antibody positivity, labial salivary gland biopsy with focal lymphocytic sialadenitis and focus score of ≥1 foci/4 mm2, abnormal Ocular Staining Score (OSS) of ≥5 (or van Bijsterveld score of ≥4), Schirmer’s test result of ≤5 mm/5 min and an unstimulated salivary flow rate of ≤0.1 mL/min [5].

2. Pathogenic Mechanisms of pSS

The interaction between genetic and environmental factors is thought to play a crucial role in susceptible individuals, leading to the dysregulation of the immune system and pSS development [6]. The function of specific cytokines and chemokines, and their expression by cells of the innate and adaptive immune systems are actively involved in pSS pathogenesis, including extraglandular involvement [1,7].

2.1. Viral/Infectious Factors

Different infectious agents, especially viruses, have been considered potential pSS pathogenetic triggers [8]. For example, Epstein–Barr virus (EBV) was identified in saliva samples and in salivary and lacrimal gland biopsies of pSS patients with EGMs, particularly with myopathies [9]. Furthermore, EBV has a well-established tropism for B cells, favoring the development of lymphoproliferative processes, the most severe extraglandular complication in pSS [10]. However, to date, no clear association with viral infections, such as human herpes virus-6 (HHV-6), hepatitis C or B viruses, human immunodeficiency virus (HIV), human T-lymphotropic virus type 1 (HTLV1), or Coxsackie A virus, have been reported in pSS [11].

2.2. Genetic and Epigenetic Factors

Genetic factors play an important role in pSS pathogenesis [12]. Thus, associations between Human Leukocyte Antigen (HLA) alleles, such as DRB1*03:01, DQA1*05:01, DQB1*02:01, and pSS susceptibility, were identified by genomic studies [13]. Moreover, six non-HLA regions were shown to be involved in pSS, including interferon regulatory factor 5 (IRF5), signal transducer and activator of transcription 4 (STAT4), BLK, interleukin (IL)-12A, TNFAIP3 interacting protein 1 (TNIP1), and C-X-C motif chemokine receptor 5 (CXCR5). The HLA-DQB1*0201 allele and the expression of IRF5 and STAT4 seem to have the strongest association with pSS [14,15].
Furthermore, based on genome-wide studies, key steps in pSS triggering were identified, such as aberrant activation of the innate immune response, through the IFN and NF-kB pathways, atypical recruitment to lymphoid sites, and T-cell activation with HLA susceptibility ascending [16].

2.3. Acquired Immunity

T cells significantly contribute to pSS pathogenesis. CD4+ T cells differentiate into the two subtypes, T helper (Th)1 and Th2 cells [1,17]. Th1 cells mainly produce pro-inflammatory cytokines such as IFN-γ and IL-2 [18], while Th2 cells produce anti-inflammatory cytokines such as IL-4, IL-5, IL-9, IL-10, IL-13, and IL-25 [18,19]. Abnormal Th1 activation was detected in the salivary gland infiltrate from pSS patients, accompanied by elevated levels of IFN-γ and Th1 cells in the blood [20]. Th2-related marker transcripts were identified in germinal centers from salivary gland biopsies of pSS patients, alongside an intense B cell infiltration [21]. Th17 cells play a fundamental role in maintaining mucosal barrier integrity by inducing the synthesis of tight junction proteins and playing a defensive role against infections in healthy individuals [22,23,24]. In autoimmune disease, Th17 cells produce IL-17 and other inflammatory cytokines such as TNF-α, IL-22, and IL-26, inducing and mediating pro-inflammatory responses [25]. The affected salivary glands of pSS patients represent a perfect environment for the recruitment of Th17 cells [26]. T-regulatory cells (Tregs), responsible for immune homeostasis, suppression of autoreactive lymphocytes and release of different cytokines, such as IL-10 and TGF-β [27], have been detected with increased values in blood samples from pSS patients with EGMs, while a reduction in Treg cells seems to prevent the emergence of EGMs in these patients [28,29]. Follicular regulatory T cells (Tfr), a subtype of Treg specialized in the regulation and suppression of T helper follicular and B cell activity [30], have been reported in higher numbers in blood and salivary glands analysis of pSS patients [31,32].
B lymphocytes represent one of the hallmarks of pSS, and their dysregulation also plays a key role in autoimmunity processes and extraglandular manifestations, particularly in lymphoma development [33,34]. In pSS patients, the properties of regulatory B cells have been attributed to their ability to secrete cytokines, such as IL-6 and IL-10 [35]. IL-6 is a primary cytokine which plays a pivotal role in promoting the synthesis of autoantibodies through secondary cytokine production by local B lymphocytes [36]. Moreover, IL-6 may synergize with IL-1β and transforming growth factor (TGF) β to modify the polarization of Th cells into Th17 [37]. In experimental studies, IL-6 deficiency may reduce autoantibody production and the subsequent inflammation in specific disease organs [38]. IL-10-producing regulatory B (Breg) cells play a critical role in maintaining immune tolerance in inflammatory reactions. IL-10, a pleiotropic and immunoregulatory cytokine, contributes to the delicate balance between inflammation and immunoregulation [39]. This anti-inflammatory cytokine may diminish the production of pro-inflammatory cytokines and chemokines, including IL-1β, IL-6, IL-8, TNF-α, and IL-12, controlling organ-specific inflammation [40]. Furthermore, in salivary glands, germinal centers have been identified, sites able to promote chronic activation of B lymphocytes followed by lymphoma development in pSS patients [41]. B cell receptor CXCR5 seems to be an important element in the formation of germinal centers in pSS with extraglandular involvement [42]. The association between B cell lymphomas and pSS will be further detailed in the hematologic manifestations section of the review.
B cell activating factor (BAFF, also known as BLyS) is produced by monocytes, macrophages, and dendritic cells (DCs), and is part of the TNF family, playing a vital role in B cell survival [43]. In pSS patients, salivary epithelial cells, T and B cells can produce BAFF [44]. Additionally, type I and II IFNs were shown to induce BAFF production [45]. In pSS patients, BAFF levels are significantly increased, not only in the blood, where they correlate with the levels of anti-Ro/SSA and anti-La/SSB antibodies and disease activity but also in the salivary glands [46]. Patients with high BAFF levels have a more pronounced B cell activation that leads to lymphoid proliferations in pSS patients with EGMs [47].

2.4. Innate Immunity

In pSS, DCs present an aberrant phenotype leading to their accumulation in the salivary glands. In saliva specimens from pSS patients, an upregulation of chemokines receptors and ligands was detected, with impact on the effective migration of DCs to inflamed tissues [1,48]. On the other hand, lower numbers of circulating DCs reported in pSS patients may be secondary to the aberrant regulation of apoptosis [48]. DCs can be activated by self-antigens, Toll-like receptor (TLR) type, leading to the production of type I interferon (IFN) [49]. Furthermore, IFN induces the production of the BAFF by monocyte circulating cells, a scenario in which the DCs contribute to the activation and differentiation of B cells into plasma cells, resulting in the secretion of antibodies [50]. These mechanisms could also influence the development of lymphoma, the most feared EGM in pSS [51].
Epithelial cells are considered major players in the pathogenesis of pSS, representing the target of the autoimmune process and also the triggers of immune activation [52]. They also regulate the processes of the expression of ribonucleoprotein complexes, Ro/SSA, and La/SSB, secondary to apoptotic mechanisms [53]. In addition to the fact that they modulate the production of cytokines, such as BAFF, epithelial cells also regulate the expression of chemokines, responsible for the recruitment of leukocytes [54,55]. Furthermore, through the expression of costimulatory proteins on their surface, epithelial cells control the interaction with T cell population [31]. In pSS, local inflammation and production of proinflammatory cytokines, such as IFN-gamma and tumor necrosis factor (TNF)-alpha, lead to the disruption of the tight junction between epithelial cells, which secondarily contributes to the emergence of both glandular and extraglandular dysfunctions in pSS patients [56,57].
Natural Killer (NK) cells have also been incriminated in pSS pathogenesis [58]. In the salivary glands of pSS patients, subsets of unconventional NK cells that may produce inflammatory cytokines such as IL-22 were identified [59]. In addition, NK cells can facilitate BAFF production [60]. Furthermore, in pSS patients with EGMs, a higher number of NK cells was detected, implying their role in clinically aggressive disease [61].
Transcriptional analyses yielded an overexpression of IFN-inducible genes, also known as type I and II IFN-signature, which were identified in peripheral blood mononuclear cells and salivary gland tissue specimens from pSS patients [62]. Both IFNs signatures demonstrated their association with the development of EGMs in pSS [63]. Particularly, the type II IFN signature seems to be associated with a higher risk for lymphoma development [64]. A bidirectional interaction between the IFN pathway and B lymphocyte activation was suggested [65]. Therefore, IFN induces B cell hyperresponsiveness, which in turn favor the production of autoantibodies [66]. The significant roles of BAFF and IFN signatures in pSS pathogenesis could represent an opportunity for novel therapeutic targets in pSS [43].

3. Extraglandular Manifestations

During disease progression, most pSS patients will develop EGMs [12]. The severity and type of symptoms can vary widely from one patient to another and can significantly impact their quality of life [67]. The effective management of EGMs implies an early diagnosis, if possible before their clinical expression by means of predictive biomarkers, efficient and accurate investigation tools and scores, as well as personalized treatments, aiming to prevent complications and improve the patient’s quality of life. In current clinical practice, the wide panel of extraglandular manifestations found in pSS patients is included in the ESSDAI scoring system [68].

3.1. Musculoskeletal Manifestations

Musculoskeletal involvement is present in the majority of pSS patients, consisting of arthralgias, arthritis, and myalgias [69]. The prevalence of arthralgia is up to 96% [70], whereas arthritis has been reported in 16.6% [71]. Arthralgias may appear concurrently with the sicca symptoms and correlate with anti-Ro/SSA anti-La/SSB antibodies positivity [72]. Arthritis was reported to be intermittent, predominantly polyarticular, symmetric, and non-destructive, occasionally involving the proximal interphalangeal and metacarpophalangeal joints and wrists [73,74]. Fewer cases of monoarthritis and axial involvement have been reported [75]. Over 70% of pSS patients may complain of myalgias [76], while active myositis was reported in 0.85–14% [77]. Within the broad myositis spectrum, the inclusion body myositis (IBM), a late complication of pSS, was present in 0.5% of pSS patients [77,78]. The coexistence of myositis-specific antibodies, such as anti cytosolic 5′-nucleotidase 1A (NT5c1A) antibodies, with anti-SSA/Ro antibodies, is reported in up to 12% of pSS patients [79].

3.1.1. Pathophysiology

Synovitis can be present in joint involvement, described as inflamed synovial tissue hyperplasia of the intimal lining layer, due to the accumulation of macrophages and proliferation of fibroblast-like synoviocytes [80]. Furthermore, neoangiogenesis with endothelial activation in the synovial tissue and accumulation of inflammatory cells, such as macrophages, DCs, lymphocytes, and mast cells, under the synovial lining have been reported [81]. These infiltrating cells are activated and produce a wide range of pro-inflammatory mediators that contribute to synovitis and lead, in some patients, to cartilage and bone destruction [82]. In muscular involvement, the deregulation of autophagy and anti-cN1A auto-immunity in HLA-DR3 genetic background were observed [77]. The main finding on muscle biopsy samples seems to be a perivascular lymphocytic inflammatory infiltrate, composed of CD 4+ T lymphocytes and B-cells [83]. Nonspecific myositis, vasculitis, and necrotic lesions were reported in isolated cases [76].

3.1.2. Diagnosis

In musculoskeletal manifestations, the diagnosis is mainly clinical, and can be completed with laboratory and imaging investigations. In addition to standard X-rays, the musculoskeletal ultrasound sonography with power Doppler can be useful, especially in active arthritis [84]. In complex cases, magnetic resonance imaging (MRI) is recommended. In muscular involvement, muscular weakness, high values in creatinine phosphokinase (CPK), and an abnormal electromyogram (EMG) can be observed, and further muscle biopsy with immunohistochemical staining can be performed, as the gold standard for the diagnosis [83].

3.1.3. Perspectives for Therapeutic Management of Musculoskeletal Manifestations in pSS

The recommended treatment in mild and intermittent inflammatory joint pain may consist of non-steroidal anti-inflammatory drugs (NSAIDs), while, in acute flares, intra-articular, intra-muscular, or oral steroids may be an option, allowing an immediate response [85]. For most patients with joint symptoms, hydroxychloroquine (HCQ) is the next treatment option [86]. Even though the studies using HCQ in pSS are inconclusive, this approach is recommended in the EULAR guidelines for the management of pSS [85]. Methotrexate (MTX) is strongly suggested as an alternative treatment for arthritis, either alone or in addition to HCQ [85,87]. In refractory cases to HCQ and/or MTX, alternative options, such as corticosteroids, leflunomide, sulfasalazine, azathioprine (AZA), cyclosporine A, or biologic drugs may be considered [88]. Anti-TNF-α and anti-CD-20 drugs have demonstrated a limited benefit in the control of systemic manifestations, while the CTLA fusion protein (CTLA4), abatacept, seems to bring hope in the alleviation of musculoskeletal manifestations of pSS patients [89,90,91]. Anti-BLyS/BAFF therapy, belimumab, seems to improve arthritis manifestations in pSS patients [92], while ianalumab (VAY736), another B cell-depleting BAFF-R blocker, provided promising results in recent studies [93]. However, to date, no biologic drugs have been approved by the regulatory agencies, the Food and Drug Administration (FDA) or the European Medicines Agency (EMA), for the treatment of pSS. Their use is experimental/investigational in ongoing clinical trials (Table 1).

3.2. Dermatological Manifestations

A large spectrum of skin manifestations may be present in patients with pSS, from common xeroderma to severe vasculitis, including other rare associated conditions [110]. The incidence of cutaneous manifestations in pSS has a female predominance, with a prevalence of up to 72%, making skin involvement one of the most common EGMs in pSS (Table 2) [111,112]. Xeroderma is the most typical cutaneous manifestation of pSS, with a prevalence of up to 72% [112]. Similar to xerosis, eyelid dermatitis has been reported in 42% of pSS patients as a lichenification or thickening of the skin, along with erythema, pigmentation, or papules [111,113]. Another cutaneous manifestation is Raynaud’s phenomenon (RP), with a prevalence between 16–35% in pSS patients [114].
Cutaneous vasculitis lesions (CVL) can be observed in pSS patients, manifested as palpable purpura and non-palpable purpura [115,116]. CVL is considered the most clinically and prognostically significant cutaneous complication of pSS patients [110]. The prevalence of CVL has been reported in 10–30% of pSS patients [111,117], while its clinical manifestation, palpable purpura, was found in 80–90% of CVL [118]. Therefore, CVL are frequently associated with other systemic manifestations, more severe disease, lymphoma, and poor prognosis, especially when serum cryoglobulins are present [115,119]. Patients with vasculitis have a higher prevalence of anti-Ro/SSA and/or anti-La/SSB antibodies, and about one-third of them have positive cryoglobulins [111]. Alongside the CVL, cutaneous ulcers, urticarial vasculitis, or skin nodules may appear in pSS patients [116,120].
Cutaneous amyloidosis is rarely found in pSS, and can manifest as a single nodule, or sometimes multiple nodules, mainly on the legs, arms, trunk, and face [115]. Localized amyloid light-chain (AL) amyloidosis can occur in the skin, lungs, eyes, and bladder of pSS patients [121,122], while cutaneous amyloid A (AA) amyloidosis, is uncommon in pSS and has been reported in association with celiac disease [123]. Annular erythema (AE), an erythematous non-scarring dermatosis, characterized by a wide elevated border and a central pallor area, is commonly found in Asian patients [124], compared with only 9% in non-Asian populations [125]. In subacute cutaneous lupus (scLE), an entity similar to AE, studies have reported a strong association with the positivity for anti-Ro/SS-A and/or anti-La/SS-B autoantibodies [8,111].
Table 2. The most frequent cutaneous manifestations in pSS.
Table 2. The most frequent cutaneous manifestations in pSS.
Cutaneous Manifestations% pSS Patients
Xeroderma72% [110]
Eyelid dermatitis42% [113]
Raynaud phenomenon16–35% [114]
Cutaneous vasculitis lesions10–30% [117]
Urticarial vasculitis0.8–21% [120]
Annular erythema9% [125]
Less common dermatologic manifestations noted in pSS include pruritus, vitiligo, alopecia, anetoderma, Sweet syndrome, lichen planus, granulomatous panniculitis, subcorneal pustular dermatosis, erythema elevatum diutinum, erythema multiforme-like, erythema perstans-like, erythema nodosum-like lesions, lymphomatoid papulosis, and cutaneous T-cell lymphoma [115].

3.2.1. Pathophysiology

In xeroderma, an alteration in the protective function of the outer layer of the skin was described, with decreased sebaceous and sweat gland secretion, and the so called autoimmune epithelitis, defined as an increased infiltration with autoreactive T and B cells [126,127]. It also involves circulating immune complexes, complement activation, cytokine production, as well as endothelial cell damage, resulting in the loss of their fibrinolytic properties, with fibrin deposition and degeneration of affected vessels [128]. Eyelid dermatitis etiology is attributed to chronic mechanical trauma through the rubbing of the periorbital area. The histopathologic examination revealed interface dermatitis, multiple melanophages, and a dense lymphocytic infiltration around hair follicles [129]. In vasculitis, the most common finding on pathology specimens, aspects of leukocytoclastic vasculitis followed by cryoglobulinemic and urticarial vasculitis were described [115,130]. This leukocytoclastic vasculitis is characterized by the fibrinoid necrosis of the vessel walls, leukocytosis and extravasation of erythrocytes, and the presence of IgM, IgG, and C3 around the vessel [111,131]. In urticarial vasculitis, lesions arise due to the activation of mast cells, which release histamine, resulting in vasodilatation, increased vascular permeability, and dermal oedema [111]. Amyloidosis is characterized by the extracellular deposition and accumulation of amyloid fibrils, AL or AA proteins, that can be objectified in green fluorescence under polarized light microscopy or by Congo red staining of tissue samples [121]. For annular erythema/scLE, the histology exam showed perivascular and periadnexial lymphocytic infiltration with dermal mucin deposits [125].

3.2.2. Diagnosis

The diagnosis of cutaneous manifestations is mostly clinical (Figure 2), while skin biopsy is recommended in complex cases. Recent clinical studies have identified biological elements, such as monocytes to lymphocyte ratio (NLR), platelet to lymphocyte ratio (PLR), monocytes to lymphocyte ratio (MLR), or gammaglobulins as predictive parameters for cutaneous involvement in pSS patients [116,132].

3.2.3. Treatment of Cutaneous Manifestations in pSS

Treatment of cutaneous manifestations in pSS patients varies from local emollients in the xeroderma to systemic immunosuppression in CVL. In cutaneous vasculitis, the treatment choice depends on the extent and degree of the manifestation and requires glucocorticoids (GCs) with or without systemic immunosuppression, such as AZA or MTX [133]. In refractory cases, cyclophosphamide (CYC) is most commonly prescribed, while in cryoglobulinemic vasculitis, administration of rituximab, or belimumab, provided promising results [134,135]. Annular erythema has a slow response to topical therapy, while GCs, calcineurin inhibitors, and hydroxychloroquine (HCQ) showed a good perspective [124]. The therapy for localized cutaneous amyloidosis is challenging and involves cryotherapy, electrodissection and curettage, intralesional triamcinolone injections, or ablative laser therapy [136].

3.3. Neurologic Manifestations

Neurologic manifestations may involve the central nervous system (CNS), with a prevalence of around 5% [137], or the peripheral nervous system (PNS), with an incidence between 3.7% to 16% in pSS patients (Table 3) [138,139,140]. CNS involvement in pSS varies from mild cognitive dysfunction to transverse myelitis and paralysis [141]. CNS involvement can include demyelinating diseases, such as neuromyelitis optica, optic neuritis, multiple sclerosis-like disorders, transverse myelitis, lymphocytic meningitis, and possible cerebral vasculitis [142]. Demyelinating CNS lesions can occur in the white matter of the brain and spinal cord of patients with pSS, with an incidence of 3.6–68% [142]. It can mimic the primary progressive forms of multiple sclerosis, with various symptoms, including visual loss, paresis of limbs, ataxia, sphincter dysfunction, cognitive dysfunction, and sensory symptoms [141]. Cranial neuropathy may be present, usually as unilateral pure sensory trigeminal neuralgia that affects the maxillary branch of the trigeminal nerve [141]. Patients with pSS may also present facial nerve neuropathy, and cochlear nerve damage, with both hearing loss and vestibular symptoms [143].
PNS manifestations in pSS patients have various clinical aspects, from axonal sensory and sensorimotor polyneuropathies, small fiber sensory neuropathy, sensory ataxic neuronopathy, cranial nerve neuropathies, radiculoneuropathy, mononeuropathy multiplex, autonomic neuropathies, to chronic inflammatory demyelinating polyneuropathy [141,144]. The most common patterns are pure sensory polyneuropathies and sensorimotor neuropathies, with a prevalence between 40 to 49% and 28 to 56.45%, respectively [140,141]. Distal sensory polyneuropathy affects large nerve fibers and can present indolent and mild paresthesia of the extremities, while sensorimotor polyneuropathy occurs when there is weakness at the same time [139,140]. Mononeuritis multiplex is a painful condition in which damage to two or more nerves occurs in succession leading to sensory and motor deficits [145]. It can be associated with pSS cryoglobulinemic vasculitis and with active systemic disease [142].
Table 3. The prevalence of the main neurologic manifestations.
Table 3. The prevalence of the main neurologic manifestations.
Neurologic Manifestations% pSS Patients
Central nervous system involvement5% [137]
  Demyelinating lesions3.6–68% * [142]
  Cranial neuropathy16–20% * [141]
  Cognitive dysfunction53% * [141]
Peripheral nervous system involvement3.7–16% [138,140]
  Pure sensory neuropathy40–49% ** [138,140]
  Sensorimotor polyneuropathies28–56.45% ** [138,140]
Autonomic nervous system involvement3–50% [146]
* from patients with central nervous system involvement. ** from patients with peripheral nervous system involvement.
Autonomic nervous system dysfunction was also reported in pSS patients and can manifest as excessive postural tachycardia, orthostatic hypotension, bladder dysfunction, gastrointestinal dysmotility, tonic pupil, segmental hypohidrosis, and diminished sweating [146,147].

3.3.1. Pathophysiology

Different pathogenic mechanisms have been suggested based on the histological and serological findings in pSS patients with neurologic involvement. Vasculitis of the vasa nervorum was described, with lymphocytic, macrophage, and T cell infiltration, as well as necrotizing vasculitis and anti-neuronal antibodies, according to the type of nerve involved [139]. Moreover, perineurial infiltration was observed on nerve biopsies of patients with sensorimotor neuropathy [148]. While in mononeuritis multiplex, inflammation of epineural and perineural blood vessels that perfuse the involved nerves leads to infarction [149] in autonomic nervous system involvement, potential etiologies include cholinergic neurotransmission blockade by cytokines or autoantibodies, T-cell infiltration, and autonomic nerve destruction [141].

3.3.2. Diagnosis

Clinical neurologic symptoms and signs, electromyographic results, and nerve biopsy are the main elements in the diagnosis of peripheric neuropathy [150]. Pseudo blocks corresponding to the areas of nerve ischemia are the distinctive sign detected through electrophysiologic investigation [151]. Skin biopsies revealed a reduced density of epidermal nerve fibers in these patients [152]. In sensory ataxic neuronopathy, magnetic resonance imaging (MRI) may show increased hyperintensity of T2-weighted images in the posterior columns [144]. For demyelinating CNS lesions, MRI may describe a rare form of widespread inflammation of the spinal cord causing T2 hyperintensity, extending across three or more vertebral segments, an aspect suggestive of longitudinally extensive transverse myelitis [153]. The examination of cerebrospinal fluid is mandatory, especially in differential diagnosis with other pathologies [154]. In cognitive dysfunction, neuropsychological tests are the most helpful tools to characterize the nature and degree of impairment [155]. Autonomic testing includes the quantitative sudomotor tests of axonal reflexes and intestinal motility tests for the esophagus, stomach, and small and large bowel [156]. In contrast, the cardiovagal function may be assessed through the measurement of heart rate variability and blood pressure responses with the Valsalva maneuver or tilt-table testing [147]. Cryoglobulinemia, NLR, MLR, gammaglobulins, C4 or vitamin D were suggested in recent studies as predictive markers for neurological involvement in pSS patients [157,158].

3.3.3. Treatment Perspectives in Neurologic Involvement in pSS Patients

As first-line treatment for neuropathic pain, tricyclic antidepressants (TCAs), such as clomipramine and imipramine, and also the serotonin-norepinephrine reuptake inhibitors (SNRIs), such as duloxetine and venlafaxine, have demonstrated their effectiveness [159]. Antiepileptic drugs, such as gabapentin and pregabalin, helped in neuropathic pain alleviation [160,161]. In patients with progressive or refractory symptoms of axonal sensory and sensorimotor neuropathies, oral GCs, intravenous immunoglobulins (IVIg), or mycophenolate mofetil (MMF) are recommended [87,142]. Furthermore, pulse therapy with GCs should be initiated rapidly in vasculitis neuropathies, followed by the addition of CYC [162]. If patients develop CYC toxicity or as maintenance therapy, AZA, and MTX can be used [144,145]. Rituximab, a monoclonal antibody targeting CD20 on B lymphocytes, and long-term anti-BLyS/BAFF therapy, belimumab, revealed promising results in neuroimmune abnormalities in pSS [142,163] (Table 1).

3.4. Renal Manifestations

Clinical renal disease is unusual in pSS, being reported in 5% of patients, but it is probably underestimated [8,111,164]. Chronic tubulointerstitial nephritis is the predominant form of pSS-associated renal involvement, which clinically translates mostly into distal renal tubular acidosis (RTA). Furthermore, Type I distal RTA is characterized by a cortical collecting duct dysfunction leading to an impaired H+ elimination [164]. Secondary to tubular defects, patients develop systemic metabolic acidosis or the inability to acidify urine following an oral acid intake [164,165]. Weakness or paralysis due to hypokalemia, renal calculi, or osteomalacia may be present in pSS patients with distal RTA [166,167]. Other dysfunctions involving the cortical collecting duct, the proximal tubular loop of Henle, and the distal convoluted tubule have been reported [167]. Glomerulonephritis is also present in pSS patients, classified as membranoproliferative glomerulonephritis. Cryoglobulinemic vasculitis, characterized by the deposition of immune complexes, is one of the most severe renal manifestations in pSS [164]. Overall, among pSS with renal involvement, the loss of renal function was reported in only 5–10% and may progress to end-stage kidney disease [168].

3.4.1. Pathophysiology

Secondary to systemic inflammation, the infiltration of B and T lymphocytes and plasma cells into the renal interstitium was observed [169]. Interstitial fibrosis occurs as a result of tubulitis as well as the local production of autoantibodies against sodium chloride co-transporter and carbonic anhydrase in the collecting duct, which results in tubular dysfunction and renal tubular acidosis [170]. In glomerular involvement, an immune complex-mediated mesangioproliferative glomerulonephritis is present [171].

3.4.2. Diagnosis

The diagnosis is a compound of clinical symptoms, such as edema in nephrotic syndrome, plus routine analyses with low proteinuria, elevated serum creatinine, and metabolic acidosis [164,168].

3.4.3. Treatment

The prognosis of kidney disease associated with pSS is generally favorable. However, the treatment is very much dependent upon disease progression [164]. For the most common renal manifestations, especially in severe or active interstitial nephritis, treatment with systemic corticosteroids is recommended [172]. During disease relapse, a steroid-sparing agent is required, and AZA [172] or MMF [173] have been used with success. In type I distal RTA, alkaline products are prescribed for acidemia, while in persistent hypokalemia, supplemental potassium is required [171]. In membranoproliferative glomerulonephritis, and cryoglobulinemic vasculitis, specific treatment with immunosuppressants is recommended [174]. In severe cases, plasma exchange may be an option [171].

3.5. Hematologic Manifestations

In pSS, the hematologic manifestations can be divided into cellular and humoral components (Table 4). The cytopenias are the main cellular abnormalities, while hyper- and hypogammaglobulinemia, monoclonal gammopathy, cryoglobulinemia, and the presence of autoantibodies are the most frequent humoral manifestations [175]. Leukopenia was reported in 19% to 22% of pSS patients [176], while neutropenia and lymphopenia were noted in 14 to 27.3% [177] and in 14 to 23.9% of the pSS patients, respectively [116,178,179]. Secondary to chronic inflammation, anemia may occur in between 17.1 and 23.9% of patients [180], while thrombocytopenia was reported between 29.3 to 30.5% of pSS patients [181,182]. Hypergammaglobulinemia occur in 41.8% [183], while hypogammaglobulinemie is very rare [184]. Monoclonal gammopathy has been observed in few cases of pSS patients [185]. Low levels of C3 have been reported in 10–15% of patients [185], while low levels of C4 were detected in 5–20% [111,185].
Lymphoma is considered the most severe complications in pSS patients [184]. The most common type of lymphoma associated with pSS is the low-grade B cell non-Hodgkin lymphoma (NHL) of the marginal zone histologic type, especially that of mucosa-associated lymphoid tissue (MALT) [186].
Table 4. The prevalence of the main hematologic manifestations.
Table 4. The prevalence of the main hematologic manifestations.
Hematologic Manifestations% pSS Patients
Cellular manifestations
  Anemia17.1–23.9% [177,187]
  Leukopenia19–22% [176,179,187]
  Thrombocytopenia29.3–30.5% [118,182]
  Neutropenia14–27.3% [118,177]
  Lymphopenia14–23.9% [118,178]
Humoral manifestations
  Hypergammaglobulinemia41.8% [183]
  Hypocomplementemia C3/C410–15%/5–20% [185]
Lymphoma2.7–9.9% [184]

3.5.1. Pathophysiology

Various environmental factors, such as viral infections, induce lymphocyte activation and their migration to target tissues [188]. In the process of lymphocyte migration and infiltration, chemokine receptors and their ligands play an important role [189]. The CC-chemokine receptor 7 (CCR7) has been recognized as a fundamental regulator directing lymphocytes to inflammatory lesions [181,190]. BAFF has been involved in the pathogenesis of pSS due to its role in B-cell regulation and proliferation [43]. Mutations of the BAFF receptor (BAFF-R) confer a higher risk for lymphoproliferation through the activation of the Nuclear Factor-κB (NF-κB) signaling pathways and of PI3K signaling pathways, and by inhibiting additive apoptotic pathways in pSS patients [46,191]. These interactions can play an important role in lymphoma progression. Seronegative pSS patients have less hyperactive B cells and thus a low risk for lymphoma development [192].

3.5.2. Diagnostic

The diagnostic for hematologic manifestations is primarily biological. Additionally, in lymphoma, the clinical aspect is suggestive, with persistently swollen parotid glands secondary to autoimmune inflammatory sialadenitis, infection, and obstruction [193]. NHL manifests typically as a unilateral, persistent, and sometimes indurated nodule [184]. Ultrasound and MRI scans are helpful in the diagnosis and evaluation, while biopsy is mandatory for the final diagnosis [194]. The prognosis in NHL associated with pSS is usually good, especially with the MALT subtype [148]. The most frequently reported predictive factors in NHL include parotid enlargement, lymphadenopathy, palpable purpura, low C4 level, and cryoglobulinemia [35,176]. Elevated levels for rheumatoid factor, a focus score greater than three, and the detection of germinal centers in salivary biopsies are also highly predictive for lymphoma in pSS [195,196].

3.5.3. Treatment

The treatment for hematologic involvement includes GCs, immunosuppressants, and biologic therapy [197,198]. Given the key role of BAFF in B cell clonal expansion and lymphoma development [195], it seemed that targeting BAFF and CD20 simultaneously through belimumab/rituximab co-administration could represent a promising therapeutic approach in for MALT subtype of pSS [92]. For patients with disseminated MALT, a personalized treatment should be considered, applying a B cell depletion strategy, that can be associated with chemotherapy [6,199] (Table 1). Alkylating agents such as CYC, doxorubicin, vincristine plus prednisone (R-CHOP), chlorambucil [176], the purine analog cladribine (2-cdA) [200], or fludarabine, can also be associated with standard chemotherapy [176]. Anti-TNF alpha therapy has been associated with an increased risk of lymphoma development in pSS patients. The reported mean time between initiation of therapy and the onset of the first symptoms of lymphoma was 23 months [201].

3.6. Pulmonary Manifestations

Pulmonary manifestations occur in up to 20% of pSS patients [202,203]. The most frequently reported pulmonary manifestations are airway disease [204], interstitial lung disease (ILD), and xerotrachea [111]. Upper airway dryness can promote chronic non-productive cough, nasal crusting, epistaxis, rhinosinusitis, and hoarseness in pSS patients [205]. Persistent chronic dryness predisposes to atelectasis, bronchiectasis, bronchiolitis, and recurrent episodes of respiratory tract infections [206]. Therefore, chronic cough is reported in up to 60% of pSS patients with pulmonary involvement [206]. The most common respiratory complication of pulmonary involvement is ILD, with a prevalence between 6 to 70% [207]. ILD can manifest as non-specific interstitial pneumonia (NSIP), with a prevalence between 29 to 42%, followed by lymphocytic interstitial pneumonia (LIP) between 4 to 15% and organizing pneumonia (OP) [208], while 11 to 43% of ILD patients may develop the usual interstitial pneumonia (UIP) [203,209]. Lymphoma prevalence in pSS is estimated to be between 1–2%, represented by low-grade extranodal marginal B-cell lymphoma of the MALT type [204].
Other rare complications in pSS patients with pulmonary involvement are amyloidosis, thromboembolic disease, and pulmonary arterial hypertension (Table 5) [204].

3.6.1. Pathophysiology

A complex interaction of genetic environmental, and hormonal factors has been incriminated in pulmonary manifestations pathogenesis. Coughing symptoms are secondary to dryness from exocrine gland dysfunction and lymphocytic infiltration of the trachea, bronchi, and bronchioles [206,211]. In pSS patients, an increased formation of autoantibodies against the M3 R muscarinic receptor may lead to a compensatory increase in M3R expression, which finally leads to cholinergic hyperresponsiveness [212]. Human T lymphotropic virus type I (HTLV-1) seems to be one of the pathogens involved in the occurrence of pulmonary manifestations in pSS [213].

3.6.2. Diagnostic

All asymptomatic patients should perform chest X-rays and pulmonary function tests (PFTs) every 6 to 12 months, while in those with symptoms, a bronchoalveolar lavage (BAL) and high-resolution CT scan (HRCT) is recommended [214,215]. In ILD, a restrictive pattern is typically noted on PFTs with a diminished diffusing capacity of carbon monoxide [204,216]. On HRCT, the NSIP presents typically as a symmetrical involvement with reticular changes, traction bronchiectasis, and ground glass opacities [204,217]. In LIP, the HRCT pattern describes nodules, ground-glass opacities, thickening of the interlobular septa, and cysts [218]. Additionally, the histopathology of LIP consists of diffuse polyclonal lymphocytic interstitial infiltrate with lymphoid follicles and germinal centers [219]. OP is typically represented by multiple areas of consolidation in the periphery with ground glass opacities and centrilobular nodules [215]. On biopsy samples, chronic inflammation and polypoid intraluminal masses of fibroblasts, myofibroblasts, and collagen in the alveolar ducts and adjacent spaces are observed [220]. In UIP, the HRCT pattern is characterized by reticular changes, bronchiectasis, and honeycombing at the bases and periphery [221,222]. Histopathology reveals minimal interstitial inflammation and patches of interstitial fibrosis [223,224]. In pSS patients with consolidating nodules, mass-like opacities, and mediastinal adenopathy, a lung biopsy must be performed to exclude lymphoma or other malignancies [204,215,225]. A systemic screening of all patients with pulmonary involvement is recommended according to the latest consensus in pSS management [225].
Predictive factors for ILD development in pSS patients include dry cough, dyspnea, Raynaud’s phenomenon, and anti-Ro52 antibodies [218]. Lower levels of forced vital capacity (FVC) and higher levels of serum Krebs von den Lungen-6 (KL-6) are predictive factors for worse prognosis in these patients [226].

3.6.3. Treatment

Oral corticosteroids are the first-line therapy in symptomatic patients with progressive pulmonary impairment [204]. Furthermore, AZA or MMF, as steroid-sparing agents proved efficacy in ILD [224]. CYC combined with prednisone was successfully used in NSIP, OP, and combined patterns [208]. In patients with ILD following treatment with intravenous rituximab improvement was reported [227,228]. In a previous study on patients with LIP, a possible synergy of tacrolimus and a selective T-cell costimulatory inhibitor (CTLA4-Ig), abatacept, encouraging results of multitarget therapy were reported [88,229]. Low-dose IL-2 showed beneficial effects for patients with pSS and associated pulmonary lesions [222] (Table 1).

3.7. Gastrointestinal Manifestations

A spectrum of gastrointestinal (GI) manifestations has been described in pSS (Table 6). Esophageal involvement is represented by dysphagia in 65%, while gastroesophageal reflux (GER) is present in 13 to 60% of pSS patients with gastrointestinal manifestations [230,231]. Chronic diarrhea has been described in up to 9% of pSS patients and represents a diagnostic and therapeutic challenge [231,232]. Severe abdominal pain, GI bleeding, bowel infarction, or perforation are other manifestations reported in pSS patients [233]. Celiac disease associated with pSS was found with a prevalence between 4.5 to 15% [234,235,236]. Vasculitis involving the gastrointestinal tract in pSS is uncommon, and usually in association with cryoglobulinemia [119]. However, intestinal vasculitis should be considered in any pSS patient with severe abdominal pain, GI bleeding, infarcted bowel, or perforation [231]. Dysautonomia and chronic gastric inflammation with mucosal atrophy have also been described in pSS patients [3]. Therefore, autonomic dysfunction and gastroparesis were reported in 29 to 69% of pSS patients with gastric involvement [237,238]. In recent years, intestinal microbiota and its role in the pathogenesis of autoimmune diseases, including pSS, have been investigated [239,240].
Pancreatic involvement is typically asymptomatic and relates mostly to pancreatic exocrine insufficiency, with a prevalence of 36–63% [119,231]. Primary biliary cirrhosis (PBC), chronic active autoimmune hepatitis (AIH), and sclerosing cholangitis (SC) are liver manifestations in pSS [8,241]. PBC, the autoimmune disease of the bile ducts leading to bile duct destruction, cholestasis, and liver failure, was reported with a prevalence of 4% up to 9% [241], while AIH was confirmed in about 1 to 4% of pSS patients with liver involvement and was characterized by autoimmune destruction of hepatocytes and an increased serum level of autoantibodies and enzymes [242,243]. SC, characterized by progressive inflammation and fibrosis of the intra- and extra-hepatic bile ducts, is an exceptional manifestation in pSS patients [242].
Table 6. The prevalence of the main gastrointestinal manifestations.
Table 6. The prevalence of the main gastrointestinal manifestations.
Gastrointestinal Manifestations% pSS Patients
Esophagus
      Dysphagia65% * [231]
      Gastroesophageal Reflux13–60% * [231]
Stomach
      Gastritis36–65% ** [244]
      Gastrointestinal motility disfunction29–69% ** [237]
Pancreas36–63% [231]
Liver49% [235]
      Autoimmune hepatitis1–4% *** [242,243]
      Primary biliary cirrhosis4–9% *** [243]
Small Intestine
      Chronic diarrhea9% [231,232]
      Celiac disease4.5–15% [234,235]
* from patients with esophagus involvement. ** from patients with gastric involvement. *** from patients with hepatic involvement.

3.7.1. Pathophysiology

The acid-neutralizing capacity of saliva is diminished due to decreased volume and altered pH [245]. Additionally, gastric acid production is inhibited due to a decrease in epidermal growth factor secretion from the submandibular glands in pSS patients [246]. The dysmotility in pSS patients is attributed to autoantibody activity that may inhibit muscarinic receptor-mediated cholinergic neurotransmission [247].

3.7.2. Diagnostic

For the diagnosis of gastrointestinal involvement various paraclinical investigations might be used, such as endoscopy, barium swallow X-ray, or esophageal manometry [248,249]. The biopsy is indicated in atrophic gastritis, showing mononuclear cell infiltration of the mucosa and glandular atrophy with varying degrees of intestinal metaplasia [235]. Celiac disease also requires histopathologic confirmation [234]. Liver manifestations are diagnosed by laboratory parameters, ultrasonography, and in some cases, by biopsy. The prognosis of PBC is generally good, while the AIH can progress to cirrhosis and even hepatocellular carcinoma (HCC) [242,250].

3.7.3. Treatment

In mild GI involvement, simple treatment strategies with secretagogue medication can improve the symptoms [88]. In more severe conditions, such as celiac disease or PBC, specific treatment is required [251]. In patients with PBC associated with pSS, the early use of ursodeoxycholic acid could prevent progression to cirrhosis [252]. In autoimmune hepatitis, treatment with prednisone followed by AZA was reported to be efficient [252], whereas in SC endoscopic interventions even liver transplant can be considered [242,253].

3.8. Cardiovascular Manifestations

Cardiovascular manifestations are rarely reported in pSS and are the main elements of the organ-specific group of non-ESSDAI features [3]. Cardiovascular events in pSS patients can be classified according to the interconnection between the traditional risk factors, glandular involvement [234,254,255,256] and the association with extra-glandular disease activity and longer disease duration [254]. A higher risk for major cardiovascular events, cerebrovascular events, and coronary events have been reported in pSS patients [257,258,259]. Furthermore, a higher prevalence of valvular regurgitation, systolic dysfunction, pericardial effusion, and lower coronary reserve have been diagnosed in pSS patients [260,261]. Similarly, an increased prevalence of pulmonary arterial hypertension (PH), reaching 12.5%, has been reported in pSS patients [262,263]. A clinical study has found that pSS patients with positive Ro/SS-A and La/SS-B antibodies had a higher prevalence of cerebrovascular events [264]. Antiphospholipid antibodies are found more frequently in pSS patients than in the general population [265]. However, only one-third of these patients will develop the antiphospholipid syndrome [266]. In pregnant women with positive Ro/SS-A antibodies, especially the anti-Ro 52 antibodies, an increased risk of developing autoimmune congenital atrioventricular block (CAVB) in the fetus has been reported. The first-born child has a risk for CAVB between 2 and 5%, that may increase up to 12% if the woman had a previous child with CAVB [267].

3.8.1. Pathophysiology

The mechanisms for cardiovascular involvement remain unknown. Various studies have identified endothelial dysfunction, carotid intima-media thickness changes with resulting loss in vessel wall compliance in pSS patients [260,268]. Persistent endothelial dysfunction-related subclinical atherosclerosis may be found in pSS patients, who have increased values of circulating endothelial microparticles, endothelial progenitor cells [257], and angiotensin [269]. Endothelial dysfunction is considered one of the earliest changes that characterize atherosclerosis and seems to be more frequent in patients with higher disease activity scores [139]. Furthermore, BMPR2 mutation can cause the proliferation of pulmonary vascular smooth muscle cells, leading to PH [270]. The 5-HT4 serotoninergic receptor, the α1C and the α1D subunits of the L-type calcium channel, and the T-type calcium channel are in vitro fetus proteins that may be cross-reactive targets for anti-Ro 52 antibodies of women with pSS [271]. Thus, the arrhythmogenic effects of anti-Ro52 antibodies and their direct effect on fetus cardiocyte function were correlated with the inhibition of membrane calcium channels [272].

3.8.2. Diagnostic

Potential biomarkers evaluating cardiac involvement and disease severity, such as C-reactive protein (CRP), IL-6, calprotectin, and dickkopf-related protein 1 (DKK-1), can be assessed in pSS patients [273]. Creatine kinase-MB, cardiac troponin I and T, brain natriuretic peptide, and pro-brain natriuretic peptide may also be used for monitoring cardiovascular involvement in pSS patients [273,274]. For long-QT syndrome diagnosis, ECG should be performed, while echocardiography with Tissue Doppler imaging and speckle-tracking technique can detect subclinical myocardial alterations [275]. When PH is suspected, right heart catheterization is recommended, and the diagnosis is defined as a pulmonary artery pressure above 25 mmHg [276]. Cardiac CT is recommended when pericardial involvement is suspected [255], while cardiac magnetic resonance (CMR) imaging [277] with tissue characterization sequences and positron emission tomography (PET) [278] provides additional insight regarding the presence of non-ischemic inflammatory myocardial involvement [279]. In recent years, CMR feature tracking (CMR-FT) has been used for the assessment of myocardial deformation with ventricular strain [277]. Overall, a high risk of cardiovascular events in pSS patients has been reported with longer disease progression and higher disease activity, with active immunologic and clinical features [262].

3.8.3. Treatment

pSS patients can be considered at risk for cardiovascular diseases due to the systemic inflammation and immune dysregulation [258,280]. General preventive measures for cardiovascular manifestations represent the main treatment strategy in pSS patients as well. Having no specific recommendations for pSS patients with cardiovascular involvement, individualized treatment for each condition is recommended, according to the existing international guidelines [281].

3.9. Other Constitutional Symptoms

Fatigue is a challenge from both diagnostic and therapeutic perspectives in pSS patients. It is a widespread manifestation in pSS, alongside dry eye and dry mouth symptoms, being reported in up to 70% of patients [282]. Sleep disorders, dry mouth, anxiety, depression, fibromyalgia, anemia, or hypothyroidism have been incriminated in the pathogenesis of chronic fatigue [283,284]. Studies have shown that cytokines and other markers of inflammation are associated with fatigue in pSS [282,285]. Part of middle-aged women diagnosed with pSS may present a clinical triad characterized by dryness, pain, and fatigue [286]. Fatigue in pSS tends to persist over time, and regular physical exercise may prove beneficial [287]. HCQ has been mentioned in previous reports as a possible treatment option for fatigue, but there is currently no evidence of significant improvement [288,289]. Rituximab use for fatigue alleviation remains controversial [95,103]. In recent studies, low-dose IL-2 therapy has improved fatigue symptoms in pSS patients [222,290].

4. Future Perspectives for pSS Management

Genetic, proteomic and transcriptomic analyses are of interest in the characterization of the molecular and clinical variabilities among pSS patients. Analysis of the peripheral blood mononuclear cells (PBMCs) from pSS patients by single-cell RNA-sequencing (scRNA-seq) identified CD14+ monocytes (Mos), CD14+CD16+ Mos, CD16+ Mos, CD8+ CTLs, and CD56 CD16+FCER1G+ NK cells as the main players in pSS pathogenesis [291]. The weighted gene co-expression network analysis (WGCNA) indicated LINC00487 and SOX4 as key genes associated with the dysregulation of B cells in pSS patients [66]. Combining WGCNA and scRNA-seq, the ICOS gene up-regulation was found in salivary glands and PBMCs of pSS patients [292]. It was suggested that ICOS gene expression may be associated with lymphocytic infiltration in pSS patients and may become a useful biomarker for the detection of pSS and its complications [293]. Another study has detected an upregulation of Interferon induced with helicase c domain 1 (IFIH1) key gene in peripheral blood samples from mice and patients with pSS [294]. IFIH1 has been considered as a new diagnostic biomarker and potential therapeutic target in pSS patients. However, the clinical relevance of genetic testing in patients with pSS is still scarce, and the association of different gene variants with specific glandular and EGMs needs further investigation.
Proteomic studies have identified the upregulation of several salivary proteins, including neutrophil elastase, tripartite motif-containing protein 29 (TRIM29), calreticulin, clusterin, salivary NGAL, siglec-5, CA-VI, and vitronectin in pSS patients [295,296]. These unregulated proteins may help monitor the disease activity and predict the response to therapy in pSS patients. Furthermore, the TRIM29 protein might become an important marker due to its high diagnostic accuracy, particularly in pSS patients with anti-SSA/Ro antibodies [295].
Recent genome-wide transcriptome studies of salivary glands from mice observed that marginal zone B (MZB) cells are recruited during the early stage of the disease [297,298]. Blocking the lymphotoxin activity required for MZB cell ontogeny may prevent lymphomagenesis in pSS with EGMs [299]. Type 2 conventional dendritic cells (cDC2s) from patients with pSS are transcriptionally altered, inducing increased chemokine receptor CXCR5 expression and proliferation of tissue homing CD4+ T cells in pSS salivary glands [300]. Further investigation of cDC2s pathway in pSS may lead to future, more efficient therapies for patients [301].
The ongoing clinical trials investigating the benefits of biological therapies in pSS patients currently provide only preliminary but at the same time promising results that could lead to a radical shift in the overall management of this disease.

5. Conclusions

Diagnostic and therapeutic management in pSS, especially when extraglandular involvement is associated, is a major challenge for the clinician. It is very unlikely that a single therapy will provide satisfying results or long-term disease control in pSS patients with EGMs, given the heterogeneity of clinical and biological phenotypes. Therefore, a better knowledge of the pathogenesis and biological profile for each type of EGM associated with pSS is mandatory. Continued research in the pathogenic mechanisms and biomarkers field are necessary for the timely detection of EGMs in pSS patients, thus preventing serious complications. Biological technologies such as genetic, proteomic, or transcriptomic analysis could lead to updated, highly accurate diagnostic guidelines in pSS and could become the fundament for future personalized therapies in the management of this disease.

Author Contributions

Conceptualization, A.M. and A.C.; methodology, A.M. and A.C.; writing—original draft preparation, A.M., C.C., C.J., M.C., F.C.B., D.O.-B., R.I. and A.C. writing—review and editing, A.M., C.C. and A.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chivasso, C.; Sarrand, J.; Perret, J.; Delporte, C.; Soyfoo, M.S. The Involvement of Innate and Adaptive Immunity in the Initiation and Perpetuation of Sjögren’s Syndrome. Int. J. Mol. Sci. 2021, 22, 658. [Google Scholar] [CrossRef] [PubMed]
  2. Both, T.; Dalm, V.A.; van Hagen, P.M.; van Daele, P.L. Reviewing primary Sjögren’s syndrome: Beyond the dryness—From pathophysiology to diagnosis and treatment. Int. J. Med. Sci. 2017, 14, 191–200. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Parisis, D.; Chivasso, C.; Perret, J.; Soyfoo, M.S.; Delporte, C. Current State of Knowledge on Primary Sjögren’s Syndrome, an Autoimmune Exocrinopathy. J. Clin. Med. 2020, 9, 2299. [Google Scholar] [CrossRef] [PubMed]
  4. André, F.; Böckle, B.C. Sjögren’s syndrome. J. Dtsch. Dermatol. Ges. J. Ger. Soc. Dermatol. JDDG 2022, 20, 980–1002. [Google Scholar] [CrossRef]
  5. Shiboski, C.H.; Shiboski, S.C.; Seror, R.; Criswell, L.A.; Labetoulle, M.; Lietman, T.M.; Rasmussen, A.; Scofield, H.; Vitali, C.; Bowman, S.J.; et al. 2016 American College of Rheumatology/European League Against Rheumatism Classification Criteria for Primary Sjögren’s Syndrome: A Consensus and Data-Driven Methodology Involving Three International Patient Cohorts. Arthritis Rheumatol. 2017, 69, 35–45. [Google Scholar] [CrossRef]
  6. Negrini, S.; Emmi, G.; Greco, M.; Borro, M.; Sardanelli, F.; Murdaca, G.; Indiveri, F.; Puppo, F. Sjögren’s syndrome: A systemic autoimmune disease. Clin. Exp. Med. 2022, 22, 9–25. [Google Scholar] [CrossRef]
  7. Shimizu, T.; Nakamura, H.; Kawakami, A. Role of the Innate Immunity Signaling Pathway in the Pathogenesis of Sjögren’s Syndrome. Int. J. Mol. Sci. 2021, 22, 3090. [Google Scholar] [CrossRef]
  8. Brito-Zerón, P.; Baldini, C.; Bootsma, H.; Bowman, S.J.; Jonsson, R.; Mariette, X.; Sivils, K.; Theander, E.; Tzioufas, A.; Ramos-Casals, M. Sjögren syndrome. Nat. Rev. Dis. Prim. 2016, 2, 16047. [Google Scholar] [CrossRef]
  9. Nakamura, H.; Shimizu, T.; Kawakami, A. Role of Viral Infections in the Pathogenesis of Sjögren’s Syndrome: Different Characteristics of Epstein-Barr Virus and HTLV-1. J. Clin. Med. 2020, 9, 1459. [Google Scholar] [CrossRef]
  10. Houen, G.; Trier, N.H. Epstein-Barr Virus and Systemic Autoimmune Diseases. Front. Immunol. 2020, 11, 587380. [Google Scholar] [CrossRef]
  11. Maslinska, M.; Kostyra-Grabczak, K. The role of virus infections in Sjögren’s syndrome. Front. Immunol. 2022, 13, 823659. [Google Scholar] [CrossRef] [PubMed]
  12. Bombardieri, M.; Argyropoulou, O.D.; Ferro, F.; Coleby, R.; Pontarini, E.; Governato, G.; Lucchesi, D.; Fulvio, G.; Tzioufas, A.G.; Baldini, C. One year in review 2020: Pathogenesis of primary Sjögren’s syndrome. Clin. Exp. Rheumatol. 2020, 38 (Suppl. S126), 3–9. [Google Scholar]
  13. Trutschel, D.; Bost, P.; Mariette, X.; Bondet, V.; Llibre, A.; Posseme, C.; Charbit, B.; Thorball, C.W.; Jonsson, R.; Lessard, C.J.; et al. Variability of Primary Sjögren’s Syndrome Is Driven by Interferon-α and Interferon-α Blood Levels Are Associated with the Class II HLA-DQ Locus. Arthritis Rheumatol. 2022, 74, 1991–2002. [Google Scholar] [CrossRef] [PubMed]
  14. Imgenberg-Kreuz, J.; Rasmussen, A.; Sivils, K.; Nordmark, G. Genetics and epigenetics in primary Sjögren’s syndrome. Rheumatology 2021, 60, 2085–2098. [Google Scholar] [CrossRef] [Green Version]
  15. Arvaniti, P.; Le Dantec, C.; Charras, A.; Arleevskaya, M.A.; Hedrich, C.M.; Zachou, K.; Dalekos, G.N.; Renaudineau, Y. Linking genetic variation with epigenetic profiles in Sjögren’s syndrome. Clin. Immunol. 2020, 210, 108314. [Google Scholar] [CrossRef] [PubMed]
  16. Khatri, B.; Tessneer, K.L.; Rasmussen, A.; Aghakhanian, F.; Reksten, T.R.; Adler, A.; Alevizos, I.; Anaya, J.M.; Aqrawi, L.A.; Baecklund, E.; et al. Genome-wide association study identifies Sjögren’s risk loci with functional implications in immune and glandular cells. Nat. Commun. 2022, 13, 4287. [Google Scholar] [CrossRef]
  17. Yao, Y.; Ma, J.F.; Chang, C.; Xu, T.; Gao, C.Y.; Gershwin, M.E.; Lian, Z.X. Immunobiology of T Cells in Sjögren’s Syndrome. Clin. Rev. Allergy Immunol. 2021, 60, 111–131. [Google Scholar] [CrossRef]
  18. Raphael, I.; Nalawade, S.; Eagar, T.N.; Forsthuber, T.G. T cell subsets and their signature cytokines in autoimmune and inflammatory diseases. Cytokine 2015, 74, 5–17. [Google Scholar] [CrossRef] [Green Version]
  19. Youinou, P.; Pers, J.O. Disturbance of cytokine networks in Sjögren’s syndrome. Arthritis Res. 2011, 13, 227. [Google Scholar] [CrossRef] [Green Version]
  20. Zamani, A.; Salehi, I.; Alahgholi-Hajibehzad, M. Moderate Exercise Enhances the Production of Interferon-γ and Interleukin-12 in Peripheral Blood Mononuclear Cells. Immune Netw. 2017, 17, 186–191. [Google Scholar] [CrossRef] [Green Version]
  21. Joachims, M.L.; Leehan, K.M.; Dozmorov, M.G.; Georgescu, C.; Pan, Z.; Lawrence, C.; Marlin, M.C.; Macwana, S.; Rasmussen, A.; Radfar, L.; et al. Sjögren’s Syndrome Minor Salivary Gland CD4(+) Memory T Cells Associate with Glandular Disease Features and have a Germinal Center T Follicular Helper Transcriptional Profile. J. Clin. Med. 2020, 9, 2164. [Google Scholar] [CrossRef] [PubMed]
  22. Abusleme, L.; Moutsopoulos, N.M. IL-17: Overview and role in oral immunity and microbiome. Oral Dis. 2017, 23, 854–865. [Google Scholar] [CrossRef] [PubMed]
  23. Xin, X.; Wang, Q.; Qing, J.; Song, W.; Gui, Y.; Li, X.; Li, Y. Th17 cells in primary Sjögren’s syndrome negatively correlate with increased Roseburia and Coprococcus. Front. Immunol. 2022, 13, 974648. [Google Scholar] [CrossRef] [PubMed]
  24. Wang, S.S.; Tang, Y.L.; Pang, X.; Zheng, M.; Tang, Y.J.; Liang, X.H. The maintenance of an oral epithelial barrier. Life Sci. 2019, 227, 129–136. [Google Scholar] [CrossRef]
  25. Yamagata, T.; Skepner, J.; Yang, J. Targeting Th17 Effector Cytokines for the Treatment of Autoimmune Diseases. Arch. Immunol. Exp. 2015, 63, 405–414. [Google Scholar] [CrossRef]
  26. Psianou, K.; Panagoulias, I.; Papanastasiou, A.D.; de Lastic, A.L.; Rodi, M.; Spantidea, P.I.; Degn, S.E.; Georgiou, P.; Mouzaki, A. Clinical and immunological parameters of Sjögren’s syndrome. Autoimmun. Rev. 2018, 17, 1053–1064. [Google Scholar] [CrossRef]
  27. Reale, M.; D’Angelo, C.; Costantini, E.; Laus, M.; Moretti, A.; Croce, A. MicroRNA in Sjögren’s Syndrome: Their Potential Roles in Pathogenesis and Diagnosis. J. Immunol. Res. 2018, 2018, 7510174. [Google Scholar] [CrossRef] [Green Version]
  28. Qi, J.; Zhang, Z.; Tang, X.; Li, W.; Chen, W.; Yao, G. IL-27 Regulated CD4(+)IL-10(+) T Cells in Experimental Sjögren Syndrome. Front. Immunol. 2020, 11, 1699. [Google Scholar] [CrossRef]
  29. Witas, R.; Gupta, S.; Nguyen, C.Q. Contributions of Major Cell Populations to Sjögren’s Syndrome. J. Clin. Med. 2020, 9, 3057. [Google Scholar] [CrossRef]
  30. Sage, P.T.; Sharpe, A.H. The multifaceted functions of follicular regulatory T cells. Curr. Opin. Immunol. 2020, 67, 68–74. [Google Scholar] [CrossRef]
  31. Verstappen, G.M.; Nakshbandi, U.; Mossel, E.; Haacke, E.A.; van der Vegt, B.; Vissink, A.; Bootsma, H.; Kroese, F.G.M. Is the T Follicular Regulatory:Follicular Helper T Cell Ratio in Blood a Biomarker for Ectopic Lymphoid Structure Formation in Sjögren’s Syndrome? Comment on the Article by Fonseca et al. Arthritis Rheumatol. 2018, 70, 1354–1355. [Google Scholar] [CrossRef] [Green Version]
  32. Fonseca, V.R.; Romão, V.C.; Agua-Doce, A.; Santos, M.; López-Presa, D.; Ferreira, A.C.; Fonseca, J.E.; Graca, L. The Ratio of Blood T Follicular Regulatory Cells to T Follicular Helper Cells Marks Ectopic Lymphoid Structure Formation While Activated Follicular Helper T Cells Indicate Disease Activity in Primary Sjögren’s Syndrome. Arthritis Rheumatol. 2018, 70, 774–784. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Kroese, F.G.; Abdulahad, W.H.; Haacke, E.; Bos, N.A.; Vissink, A.; Bootsma, H. B-cell hyperactivity in primary Sjögren’s syndrome. Expert Rev. Clin. Immunol. 2014, 10, 483–499. [Google Scholar] [CrossRef] [PubMed]
  34. Ferro, F.; Marcucci, E.; Orlandi, M.; Baldini, C.; Bartoloni-Bocci, E. One year in review 2017: Primary Sjögren’s syndrome. Clin. Exp. Rheumatol. 2017, 35, 179–191. [Google Scholar] [PubMed]
  35. Nocturne, G.; Mariette, X. B cells in the pathogenesis of primary Sjögren syndrome. Nat. Rev. Rheumatol. 2018, 14, 133–145. [Google Scholar] [CrossRef]
  36. Benchabane, S.; Boudjelida, A.; Toumi, R.; Belguendouz, H.; Youinou, P.; Touil-Boukoffa, C. A case for IL-6, IL-17A, and nitric oxide in the pathophysiology of Sjögren’s syndrome. Int. J. Immunopathol. Pharmacol. 2016, 29, 386–397. [Google Scholar] [CrossRef] [Green Version]
  37. Alunno, A.; Carubbi, F.; Bistoni, O.; Caterbi, S.; Bartoloni, E.; Mirabelli, G.; Cannarile, F.; Cipriani, P.; Giacomelli, R.; Gerli, R. T Regulatory and T Helper 17 Cells in Primary Sjögren’s Syndrome: Facts and Perspectives. Mediat. Inflamm. 2015, 2015, 243723. [Google Scholar] [CrossRef]
  38. Kaneko, Y.; Takeuchi, T. An update on the pathogenic role of IL-6 in rheumatic diseases. Cytokine 2021, 146, 155645. [Google Scholar] [CrossRef]
  39. Lin, X.; Wang, X.; Xiao, F.; Ma, K.; Liu, L.; Wang, X.; Xu, D.; Wang, F.; Shi, X.; Liu, D.; et al. IL-10-producing regulatory B cells restrain the T follicular helper cell response in primary Sjögren’s syndrome. Cell. Mol. Immunol. 2019, 16, 921–931. [Google Scholar] [CrossRef]
  40. Mielle, J.; Nutz, A.; Guillpain, P.; Audo, R.; Gaujoux-Viala, C.; Combe, B.; Morel, J.; Daien, C. IL-10-producing regulatory B cells are present and functional in primary Sjögren patients. Immunol. Res. 2021, 69, 107–113. [Google Scholar] [CrossRef]
  41. Szabo, K.; Papp, G.; Dezso, B.; Zeher, M. The histopathology of labial salivary glands in primary Sjögren’s syndrome: Focusing on follicular helper T cells in the inflammatory infiltrates. Mediat. Inflamm. 2014, 2014, 631787. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Bombardieri, M.; Lewis, M.; Pitzalis, C. Ectopic lymphoid neogenesis in rheumatic autoimmune diseases. Nat. Rev. Rheumatol. 2017, 13, 141–154. [Google Scholar] [CrossRef] [PubMed]
  43. Yoshimoto, K.; Suzuki, K.; Takei, E.; Ikeda, Y.; Takeuchi, T. Elevated expression of BAFF receptor, BR3, on monocytes correlates with B cell activation and clinical features of patients with primary Sjögren’s syndrome. Arthritis Res. 2020, 22, 157. [Google Scholar] [CrossRef]
  44. Ha, Y.J.; Choi, Y.S.; Kang, E.H.; Chung, J.H.; Cha, S.; Song, Y.W.; Lee, Y.J. Increased expression of interferon-λ in minor salivary glands of patients with primary Sjögren’s syndrome and its synergic effect with interferon-α on salivary gland epithelial cells. Clin. Exp. Rheumatol. 2018, 36 (Suppl. S112), 31–40. [Google Scholar]
  45. Lee, J.; Lee, J.; Kwok, S.K.; Baek, S.; Jang, S.G.; Hong, S.M.; Min, J.W.; Choi, S.S.; Lee, J.; Cho, M.L.; et al. JAK-1 Inhibition Suppresses Interferon-Induced BAFF Production in Human Salivary Gland: Potential Therapeutic Strategy for Primary Sjögren’s Syndrome. Arthritis Rheumatol. 2018, 70, 2057–2066. [Google Scholar] [CrossRef] [Green Version]
  46. Carrillo-Ballesteros, F.J.; Palafox-Sánchez, C.A.; Franco-Topete, R.A.; Muñoz-Valle, J.F.; Orozco-Barocio, G.; Martínez-Bonilla, G.E.; Gómez-López, C.E.; Marín-Rosales, M.; López-Villalobos, E.F.; Luquin, S.; et al. Expression of BAFF and BAFF receptors in primary Sjögren’s syndrome patients with ectopic germinal center-like structures. Clin. Exp. Med. 2020, 20, 615–626. [Google Scholar] [CrossRef]
  47. Cornec, D.; Costa, S.; Devauchelle-Pensec, V.; Jousse-Joulin, S.; Marcorelles, P.; Berthelot, J.M.; Chiche, L.; Hachulla, E.; Hatron, P.Y.; Goeb, V.; et al. Blood and salivary-gland BAFF-driven B-cell hyperactivity is associated to rituximab inefficacy in primary Sjögren’s syndrome. J. Autoimmun. 2016, 67, 102–110. [Google Scholar] [CrossRef]
  48. Hillen, M.R.; Pandit, A.; Blokland, S.L.M.; Hartgring, S.A.Y.; Bekker, C.P.J.; van der Heijden, E.H.M.; Servaas, N.H.; Rossato, M.; Kruize, A.A.; van Roon, J.A.G.; et al. Plasmacytoid DCs From Patients With Sjögren’s Syndrome Are Transcriptionally Primed for Enhanced Pro-inflammatory Cytokine Production. Front. Immunol. 2019, 10, 2096. [Google Scholar] [CrossRef] [Green Version]
  49. Ainola, M.; Porola, P.; Takakubo, Y.; Przybyla, B.; Kouri, V.P.; Tolvanen, T.A.; Hänninen, A.; Nordström, D.C. Activation of plasmacytoid dendritic cells by apoptotic particles—Mechanism for the loss of immunological tolerance in Sjögren’s syndrome. Clin. Exp. Immunol. 2018, 191, 301–310. [Google Scholar] [CrossRef] [Green Version]
  50. Swiecki, M.; Colonna, M. The multifaceted biology of plasmacytoid dendritic cells. Nat. Rev. Immunol. 2015, 15, 471–485. [Google Scholar] [CrossRef]
  51. Goules, A.V.; Kapsogeorgou, E.K.; Tzioufas, A.G. Insight into pathogenesis of Sjögren’s syndrome: Dissection on autoimmune infiltrates and epithelial cells. Clin. Immunol. 2017, 182, 30–40. [Google Scholar] [CrossRef] [PubMed]
  52. Tseng, Y.C.; Yang, H.Y.; Lin, W.T.; Chang, C.B.; Chien, H.C.; Wang, H.P.; Chen, C.M.; Wang, J.T.; Li, C.; Wu, S.F.; et al. Salivary dysbiosis in Sjögren’s syndrome and a commensal-mediated immunomodulatory effect of salivary gland epithelial cells. NPJ Biofilm. Microbiomes 2021, 7, 21. [Google Scholar] [CrossRef] [PubMed]
  53. Martín-Nares, E.; Hernández-Molina, G. Novel autoantibodies in Sjögren’s syndrome: A comprehensive review. Autoimmun. Rev. 2019, 18, 192–198. [Google Scholar] [CrossRef] [PubMed]
  54. Gong, Y.Z.; Nititham, J.; Taylor, K.; Miceli-Richard, C.; Sordet, C.; Wachsmann, D.; Bahram, S.; Georgel, P.; Criswell, L.A.; Sibilia, J.; et al. Differentiation of follicular helper T cells by salivary gland epithelial cells in primary Sjögren’s syndrome. J. Autoimmun. 2014, 51, 57–66. [Google Scholar] [CrossRef]
  55. Thompson, N.; Isenberg, D.A.; Jury, E.C.; Ciurtin, C. Exploring BAFF: Its expression, receptors and contribution to the immunopathogenesis of Sjögren’s syndrome. Rheumatology 2016, 55, 1548–1555. [Google Scholar] [CrossRef] [Green Version]
  56. Baker, O.J. Current trends in salivary gland tight junctions. Tissue Barriers 2016, 4, e1162348. [Google Scholar] [CrossRef] [Green Version]
  57. Cong, X.; Mao, X.D.; Wu, L.L.; Yu, G.Y. The role and mechanism of tight junctions in the regulation of salivary gland secretion. Oral Dis. 2023, 00, 1–20. [Google Scholar] [CrossRef]
  58. Ming, B.; Wu, T.; Cai, S.; Hu, P.; Tang, J.; Zheng, F.; Ye, C.; Dong, L. The Increased Ratio of Blood CD56(bright) NK to CD56(dim) NK Is a Distinguishing Feature of Primary Sjögren’s Syndrome. J. Immunol. Res. 2020, 2020, 7523914. [Google Scholar] [CrossRef]
  59. Liu, M.; Liang, S.; Zhang, C. NK Cells in Autoimmune Diseases: Protective or Pathogenic? Front. Immunol. 2021, 12, 624687. [Google Scholar] [CrossRef]
  60. Quah, P.S.; Sutton, V.; Whitlock, E.; Figgett, W.A.; Andrews, D.M.; Fairfax, K.A.; Mackay, F. The effects of B-cell-activating factor on the population size, maturation and function of murine natural killer cells. Immunol. Cell Biol. 2022, 100, 761–776. [Google Scholar] [CrossRef]
  61. Rizzo, C.; La Barbera, L.; Lo Pizzo, M.; Ciccia, F.; Sireci, G.; Guggino, G. Invariant NKT Cells and Rheumatic Disease: Focus on Primary Sjogren Syndrome. Int. J. Mol. Sci. 2019, 20, 5435. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Teruel, M.; Barturen, G.; Martínez-Bueno, M.; Castellini-Pérez, O.; Barroso-Gil, M.; Povedano, E.; Kerick, M.; Català-Moll, F.; Makowska, Z.; Buttgereit, A.; et al. Integrative epigenomics in Sjögren’s syndrome reveals novel pathways and a strong interaction between the HLA, autoantibodies and the interferon signature. Sci. Rep. 2021, 11, 23292. [Google Scholar] [CrossRef] [PubMed]
  63. Del Papa, N.; Minniti, A.; Lorini, M.; Carbonelli, V.; Maglione, W.; Pignataro, F.; Montano, N.; Caporali, R.; Vitali, C. The Role of Interferons in the Pathogenesis of Sjögren’s Syndrome and Future Therapeutic Perspectives. Biomolecules 2021, 11, 251. [Google Scholar] [CrossRef] [PubMed]
  64. Nezos, A.; Gravani, F.; Tassidou, A.; Kapsogeorgou, E.K.; Voulgarelis, M.; Koutsilieris, M.; Crow, M.K.; Mavragani, C.P. Type I and II interferon signatures in Sjogren’s syndrome pathogenesis: Contributions in distinct clinical phenotypes and Sjogren’s related lymphomagenesis. J. Autoimmun. 2015, 63, 47–58. [Google Scholar] [CrossRef] [Green Version]
  65. Pontarini, E.; Lucchesi, D.; Bombardieri, M. Current views on the pathogenesis of Sjögren’s syndrome. Curr. Opin. Rheumatol. 2018, 30, 215–221. [Google Scholar] [CrossRef]
  66. Inamo, J.; Suzuki, K.; Takeshita, M.; Kassai, Y.; Takiguchi, M.; Kurisu, R.; Okuzono, Y.; Tasaki, S.; Yoshimura, A.; Takeuchi, T. Identification of novel genes associated with dysregulation of B cells in patients with primary Sjögren’s syndrome. Arthritis Res. 2020, 22, 153. [Google Scholar] [CrossRef]
  67. Greenan, E.; Pilson, Q.; Gabhann-Dromgoole, J.N.; Murphy, C.C. Relationship between clinical parameters and quality of life in primary Sjögren’s Syndrome: A prospective study. Eye 2023, 1, 1–8. [Google Scholar] [CrossRef]
  68. Seror, R.; Bowman, S.J.; Brito-Zeron, P.; Theander, E.; Bootsma, H.; Tzioufas, A.; Gottenberg, J.E.; Ramos-Casals, M.; Dörner, T.; Ravaud, P.; et al. EULAR Sjögren’s syndrome disease activity index (ESSDAI): A user guide. RMD Open 2015, 1, e000022. [Google Scholar] [CrossRef] [Green Version]
  69. Leone, M.C.; Alunno, A.; Cafaro, G.; Valentini, V.; Marcucci, E.; Bartoloni, E.; Gerli, R. The clinical spectrum of primary Sjögren’s syndrome: Beyond exocrine glands. Reumatismo 2017, 69, 93–100. [Google Scholar] [CrossRef] [Green Version]
  70. Rozis, M.; Vlamis, J.; Vasiliadis, E.; Mavragani, C.; Pneumaticos, S.; Evangelopoulos, D.S. Musculoskeletal Manifestations in Sjogren’s Syndrome: An Orthopedic Point of View. J. Clin. Med. 2021, 10, 1574. [Google Scholar] [CrossRef]
  71. Gao, R.; Pu, J.; Wu, Z.; Tang, J.; Wang, X. Osteoarthritis or arthritis? Toward understanding of primary Sjögren’s syndrome patients with arthralgia. J. Orthop. Surg. Res. 2023, 18, 41. [Google Scholar] [CrossRef] [PubMed]
  72. Singh, J.A.; Cleveland, J.D. Sjogren’s syndrome is associated with higher rate of non-home discharge after primary hip arthroplasty and higher transfusion rates after primary hip or knee arthroplasty: A U.S. cohort study. BMC Musculoskelet. Disord. 2020, 21, 492. [Google Scholar] [CrossRef] [PubMed]
  73. Guedes, L.K.N.; Leon, E.P.; Bocate, T.S.; Bonfigliolli, K.R.; Lourenço, S.V.; Bonfa, E.; Pasoto, S.G. Characterizing hand and wrist ultrasound pattern in primary Sjögren’s syndrome: A case-control study. Clin. Rheumatol. 2020, 39, 1907–1918. [Google Scholar] [CrossRef] [PubMed]
  74. Jacques, T.; Sudoł-Szopińska, I.; Larkman, N.; O’Connor, P.; Cotten, A. Musculoskeletal Manifestations of Non-RA Connective Tissue Diseases: Scleroderma, Systemic Lupus Erythematosus, Still’s Disease, Dermatomyositis/Polymyositis, Sjögren’s Syndrome, and Mixed Connective Tissue Disease. Semin. Musculoskelet. Radiol. 2018, 22, 166–179. [Google Scholar] [CrossRef]
  75. ter Borg, E.J.; Kelder, J.C. Polyarthritis in primary Sjögren’s syndrome represents a distinct subset with less pronounced B cell proliferation a Dutch cohort with long-term follow-up. Clin. Rheumatol. 2016, 35, 649–655. [Google Scholar] [CrossRef] [PubMed]
  76. Espitia-Thibault, A.; Masseau, A.; Néel, A.; Espitia, O.; Toquet, C.; Mussini, J.M.; Hamidou, M. Sjögren’s syndrome-associated myositis with germinal centre-like structures. Autoimmun. Rev. 2017, 16, 154–158. [Google Scholar] [CrossRef]
  77. Felten, R.; Giannini, M.; Nespola, B.; Lannes, B.; Levy, D.; Seror, R.; Vittecoq, O.; Hachulla, E.; Perdriger, A.; Dieude, P.; et al. Refining myositis associated with primary Sjögren’s syndrome: Data from the prospective cohort ASSESS. Rheumatology 2021, 60, 675–681. [Google Scholar] [CrossRef]
  78. Giannini, M.; Felten, R.; Gottenberg, J.E.; Geny, B.; Meyer, A. Inclusion body myositis and Sjögren’s syndrome: The association works both ways. Acta Neuropathol. Commun. 2022, 10, 152. [Google Scholar] [CrossRef]
  79. Chung, S.H.; Bent, E.I.; Weiss, M.D.; Gardner, G.C. Sporadic inclusion body myositis and primary Sjogren’s syndrome: An overlooked diagnosis. Clin. Rheumatol. 2021, 40, 4089–4094. [Google Scholar] [CrossRef]
  80. Sanchez-Lopez, E.; Coras, R.; Torres, A.; Lane, N.E.; Guma, M. Synovial inflammation in osteoarthritis progression. Nat. Rev. Rheumatol. 2022, 18, 258–275. [Google Scholar] [CrossRef]
  81. MacDonald, I.J.; Liu, S.C.; Su, C.M.; Wang, Y.H.; Tsai, C.H.; Tang, C.H. Implications of Angiogenesis Involvement in Arthritis. Int. J. Mol. Sci. 2018, 19, 2012. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Bhattaram, P.; Chandrasekharan, U. The joint synovium: A critical determinant of articular cartilage fate in inflammatory joint diseases. Semin. Cell Dev. Biol. 2017, 62, 86–93. [Google Scholar] [CrossRef] [PubMed]
  83. Colafrancesco, S.; Priori, R.; Gattamelata, A.; Picarelli, G.; Minniti, A.; Brancatisano, F.; D’Amati, G.; Giordano, C.; Cerbelli, B.; Maset, M.; et al. Myositis in primary Sjögren’s syndrome: Data from a multicentre cohort. Clin. Exp. Rheumatol. 2015, 33, 457–464. [Google Scholar]
  84. Shabana, K.; Okamoto, N.; Sugita, Y.; Shindo, K.; Murata, T.; Tamai, H.; Fujiwara, K. The findings of musculoskeletal ultrasonography on primary Sjögren’s syndrome patients in childhood with articular manifestations and the impact of anti-cyclic citrullinated peptide antibody. Mod. Rheumatol. 2019, 29, 821–828. [Google Scholar] [CrossRef]
  85. Ramos-Casals, M.; Brito-Zerón, P.; Bombardieri, S.; Bootsma, H.; De Vita, S.; Dörner, T.; Fisher, B.A.; Gottenberg, J.E.; Hernandez-Molina, G.; Kocher, A.; et al. EULAR recommendations for the management of Sjögren’s syndrome with topical and systemic therapies. Ann. Rheum. Dis. 2020, 79, 3–18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Dima, A.; Jurcut, C.; Arnaud, L. Hydroxychloroquine in systemic and autoimmune diseases: Where are we now? Jt. Bone Spine 2021, 88, 105143. [Google Scholar] [CrossRef] [PubMed]
  87. Pereira, P.R.; Viala, K.; Maisonobe, T.; Haroche, J.; Mathian, A.; Hié, M.; Amoura, Z.; Cohen Aubart, F. Sjögren Sensory Neuronopathy (Sjögren Ganglionopathy): Long-Term Outcome and Treatment Response in a Series of 13 Cases. Medicine 2016, 95, e3632. [Google Scholar] [CrossRef]
  88. Fox, R.I.; Fox, C.M.; Gottenberg, J.E.; Dörner, T. Treatment of Sjögren’s syndrome: Current therapy and future directions. Rheumatology 2021, 60, 2066–2074. [Google Scholar] [CrossRef]
  89. Machado, A.C.; Dos Santos, L.C.; Fidelix, T.; Lekwitch, I.; Soares, S.B.; Gasparini, A.F.; Augusto, J.V.; Junior, N.C.; Trevisani, V.F.M. Effectiveness and safety of abatacept for the treatment of patients with primary Sjögren’s syndrome. Clin. Rheumatol. 2020, 39, 243–248. [Google Scholar] [CrossRef]
  90. Mavragani, C.P.; Moutsopoulos, H.M. Sjögren’s syndrome: Old and new therapeutic targets. J. Autoimmun. 2020, 110, 102364. [Google Scholar] [CrossRef]
  91. Carubbi, F.; Cipriani, P.; Marrelli, A.; Benedetto, P.; Ruscitti, P.; Berardicurti, O.; Pantano, I.; Liakouli, V.; Alvaro, S.; Alunno, A.; et al. Efficacy and safety of rituximab treatment in early primary Sjögren’s syndrome: A prospective, multi-center, follow-up study. Arthritis Res. 2013, 15, R172. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Álvarez-Rivas, N.; Sang-Park, H.; Díaz Del Campo, P.; Fernández-Castro, M.; Corominas, H.; Andreu, J.L.; Navarro-Compán, V. Efficacy of belimumab in Primary Sjögren’s syndrome: A systematic review. Reum. Clin. 2021, 17, 170–174. [Google Scholar] [CrossRef] [PubMed]
  93. van den Hoogen, L.L.; van Laar, J.M. Targeted therapies in systemic sclerosis, myositis, antiphospholipid syndrome, and Sjögren’s syndrome. Best Pract. Res. Clin. Rheumatol. 2020, 34, 101485. [Google Scholar] [CrossRef] [PubMed]
  94. Devauchelle-Pensec, V.; Mariette, X.; Jousse-Joulin, S.; Berthelot, J.M.; Perdriger, A.; Puéchal, X.; Le Guern, V.; Sibilia, J.; Gottenberg, J.E.; Chiche, L.; et al. Treatment of primary Sjögren syndrome with rituximab: A randomized trial. Ann. Intern. Med. 2014, 160, 233–242. [Google Scholar] [CrossRef] [PubMed]
  95. Bowman, S.J.; Everett, C.C.; O’Dwyer, J.L.; Emery, P.; Pitzalis, C.; Ng, W.F.; Pease, C.T.; Price, E.J.; Sutcliffe, N.; Gendi, N.S.T.; et al. Randomized Controlled Trial of Rituximab and Cost-Effectiveness Analysis in Treating Fatigue and Oral Dryness in Primary Sjögren’s Syndrome. Arthritis Rheumatol. 2017, 69, 1440–1450. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Fisher, B.A.; Everett, C.C.; Rout, J.; O’Dwyer, J.L.; Emery, P.; Pitzalis, C.; Ng, W.F.; Carr, A.; Pease, C.T.; Price, E.J.; et al. Effect of rituximab on a salivary gland ultrasound score in primary Sjögren’s syndrome: Results of the TRACTISS randomised double-blind multicentre substudy. Ann. Rheum. Dis. 2018, 77, 412–416. [Google Scholar] [CrossRef]
  97. Mariette, X.; Seror, R.; Quartuccio, L.; Baron, G.; Salvin, S.; Fabris, M.; Desmoulins, F.; Nocturne, G.; Ravaud, P.; De Vita, S. Efficacy and safety of belimumab in primary Sjögren’s syndrome: Results of the BELISS open-label phase II study. Ann. Rheum. Dis. 2015, 74, 526–531. [Google Scholar] [CrossRef]
  98. De Vita, S.; Quartuccio, L.; Seror, R.; Salvin, S.; Ravaud, P.; Fabris, M.; Nocturne, G.; Gandolfo, S.; Isola, M.; Mariette, X. Efficacy and safety of belimumab given for 12 months in primary Sjögren’s syndrome: The BELISS open-label phase II study. Rheumatology 2015, 54, 2249–2256. [Google Scholar] [CrossRef] [Green Version]
  99. Quartuccio, L.; Salvin, S.; Corazza, L.; Gandolfo, S.; Fabris, M.; De Vita, S. Efficacy of belimumab and targeting of rheumatoid factor-positive B-cell expansion in Sjögren’s syndrome: Follow-up after the end of the phase II open-label BELISS study. Clin. Exp. Rheumatol. 2016, 34, 311–314. [Google Scholar]
  100. Mariette, X.; Barone, F.; Baldini, C.; Bootsma, H.; Clark, K.L.; De Vita, S.; Gardner, D.H.; Henderson, R.B.; Herdman, M.; Lerang, K.; et al. A randomized, phase II study of sequential belimumab and rituximab in primary Sjögren’s syndrome. JCI Insight 2022, 7, e163030. [Google Scholar] [CrossRef]
  101. Adler, S.; Körner, M.; Förger, F.; Huscher, D.; Caversaccio, M.D.; Villiger, P.M. Evaluation of histologic, serologic, and clinical changes in response to abatacept treatment of primary Sjögren’s syndrome: A pilot study. Arthritis Care Res. 2013, 65, 1862–1868. [Google Scholar] [CrossRef] [PubMed]
  102. Haacke, E.A.; van der Vegt, B.; Meiners, P.M.; Vissink, A.; Spijkervet, F.K.; Bootsma, H.; Kroese, F.G. Abatacept treatment of patients with primary Sjögren’s syndrome results in a decrease of germinal centres in salivary gland tissue. Clin. Exp. Rheumatol. 2017, 35, 317–320. [Google Scholar] [PubMed]
  103. Verstappen, G.M.; van Nimwegen, J.F.; Vissink, A.; Kroese, F.G.M.; Bootsma, H. The value of rituximab treatment in primary Sjögren’s syndrome. Clin. Immunol. 2017, 182, 62–71. [Google Scholar] [CrossRef] [PubMed]
  104. Baer, A.N.; Gottenberg, J.E.; St Clair, E.W.; Sumida, T.; Takeuchi, T.; Seror, R.; Foulks, G.; Nys, M.; Mukherjee, S.; Wong, R.; et al. Efficacy and safety of abatacept in active primary Sjögren’s syndrome: Results of a phase III, randomised, placebo-controlled trial. Ann. Rheum. Dis. 2021, 80, 339–348. [Google Scholar] [CrossRef] [PubMed]
  105. de Wolff, L.; van Nimwegen, J.F.; Mossel, E.; van Zuiden, G.S.; Stel, A.J.; Majoor, K.I.; Olie, L.; Los, L.I.; Vissink, A.; Spijkervet, F.K.L.; et al. Long-term abatacept treatment for 48 weeks in patients with primary Sjögren’s syndrome: The open-label extension phase of the ASAP-III trial. Semin. Arthritis Rheum. 2022, 53, 151955. [Google Scholar] [CrossRef]
  106. Felten, R.; Devauchelle-Pensec, V.; Seror, R.; Duffau, P.; Saadoun, D.; Hachulla, E.; Pierre Yves, H.; Salliot, C.; Perdriger, A.; Morel, J.; et al. Interleukin 6 receptor inhibition in primary Sjögren syndrome: A multicentre double-blind randomised placebo-controlled trial. Ann. Rheum. Dis. 2021, 80, 329–338. [Google Scholar] [CrossRef]
  107. Bowman, S.J. Safety and efficacy of subcutaneous ianalumab (VAY736) in patients with primary Sjögren’s syndrome: A randomised, double-blind, placebo-controlled, phase 2b dose-finding trial. Lancet 2022, 399, 161–171. [Google Scholar] [CrossRef]
  108. Diekhoff, T.; Fischer, T.; Schefer, Q.; Posch, M.G.; Dörner, T.; Laurent, D.; Li, Y.; Wagner, F.D.; Oliver, S.J. Ianalumab (VAY736) in primary Sjögren’s syndrome: Assessing disease activity using multi-modal ultrasound. Clin. Exp. Rheumatol. 2020, 38 (Suppl. S126), 228–236. [Google Scholar]
  109. He, J.; Chen, J.; Miao, M.; Zhang, R.; Cheng, G.; Wang, Y.; Feng, R.; Huang, B.; Luan, H.; Jia, Y.; et al. Efficacy and Safety of Low-Dose Interleukin 2 for Primary Sjögren Syndrome: A Randomized Clinical Trial. JAMA Netw. Open 2022, 5, e2241451. [Google Scholar] [CrossRef]
  110. Orgeolet, L.; Foulquier, N.; Misery, L.; Redou, P.; Pers, J.O.; Devauchelle-Pensec, V.; Saraux, A. Can artificial intelligence replace manual search for systematic literature? Review on cutaneous manifestations in primary Sjögren’s syndrome. Rheumatology 2020, 59, 811–819. [Google Scholar] [CrossRef]
  111. Ramos-Casals, M.; Brito-Zerón, P.; Seror, R.; Bootsma, H.; Bowman, S.J.; Dörner, T.; Gottenberg, J.E.; Mariette, X.; Theander, E.; Bombardieri, S.; et al. Characterization of systemic disease in primary Sjögren’s syndrome: EULAR-SS Task Force recommendations for articular, cutaneous, pulmonary and renal involvements. Rheumatology 2015, 54, 2230–2238. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Torrente-Segarra, V.; Corominas, H.; Sánchez-Piedra, C.; Fernández-Castro, M.; Andreu, J.L.; Martínez-Taboada, V.M.; Olivé, A.; Rosas, J.; Sánchez-Alonso, F. Fibromyalgia prevalence and associated factors in primary Sjögren’s syndrome patients in a large cohort from the Spanish Society of Rheumatology registry (SJOGRENSER). Clin. Exp. Rheumatol. 2017, 35 (Suppl. S105), 28–34. [Google Scholar] [PubMed]
  113. Katayama, I. Abberant Sudomotor Functions in Sjögren’s Syndrome: Comparable Study with Atopic Dermatitis on Dry Skin Manifestation. Curr. Probl. Derm. 2016, 51, 62–74. [Google Scholar] [CrossRef]
  114. Rischmueller, M.; Tieu, J.; Lester, S. Primary Sjögren’s syndrome. Best Pract. Res. Clin. Rheumatol. 2016, 30, 189–220. [Google Scholar] [CrossRef] [PubMed]
  115. Jhorar, P.; Torre, K.; Lu, J. Cutaneous features and diagnosis of primary Sjögren syndrome: An update and review. J. Am. Acad. Dermatol. 2018, 79, 736–745. [Google Scholar] [CrossRef] [PubMed]
  116. Mihai, A.; Caruntu, A.; Opris-Belinski, D.; Jurcut, C.; Dima, A.; Caruntu, C.; Ionescu, R. The Predictive Role of Neutrophil-to-Lymphocyte Ratio (NLR), Platelet-to-Lymphocyte Ratio (PLR), Monocytes-to-Lymphocyte Ratio (MLR) and Gammaglobulins for the Development of Cutaneous Vasculitis Lesions in Primary Sjögren’s Syndrome. J. Clin. Med. 2022, 11, 5525. [Google Scholar] [CrossRef]
  117. Brito-Zerón, P.; Acar-Denizli, N.; Ng, W.F.; Horváth, I.F.; Rasmussen, A.; Seror, R.; Li, X.; Baldini, C.; Gottenberg, J.E.; Danda, D.; et al. Epidemiological profile and north-south gradient driving baseline systemic involvement of primary Sjögren’s syndrome. Rheumatology 2020, 59, 2350–2359. [Google Scholar] [CrossRef] [Green Version]
  118. Baldini, C.; Pepe, P.; Quartuccio, L.; Priori, R.; Bartoloni, E.; Alunno, A.; Gattamelata, A.; Maset, M.; Modesti, M.; Tavoni, A.; et al. Primary Sjogren’s syndrome as a multi-organ disease: Impact of the serological profile on the clinical presentation of the disease in a large cohort of Italian patients. Rheumatology 2014, 53, 839–844. [Google Scholar] [CrossRef] [Green Version]
  119. Retamozo, S.; Gheitasi, H.; Quartuccio, L.; Kostov, B.; Corazza, L.; Bové, A.; Sisó-Almirall, A.; Gandía, M.; Ramos-Casals, M.; De Vita, S.; et al. Cryoglobulinaemic vasculitis at diagnosis predicts mortality in primary Sjögren syndrome: Analysis of 515 patients. Rheumatology 2016, 55, 1443–1451. [Google Scholar] [CrossRef] [Green Version]
  120. Ramos-Casals, M.; Anaya, J.M.; García-Carrasco, M.; Rosas, J.; Bové, A.; Claver, G.; Diaz, L.A.; Herrero, C.; Font, J. Cutaneous vasculitis in primary Sjögren syndrome: Classification and clinical significance of 52 patients. Medicine 2004, 83, 96–106. [Google Scholar] [CrossRef]
  121. Ikeda, S.I.; Hineno, A.; Yoshinaga, T.; Matsuo, K.; Suga, T.; Shiina, T.; Otsuki, T.; Hoshii, Y. Sjögren syndrome-related plasma cell disorder and multifocal nodular AL amyloidosis: Clinical picture and pathological findings. Amyloid 2019, 26, 225–233. [Google Scholar] [CrossRef] [PubMed]
  122. Li, H.; Lu, Y. Pulmonary amyloidosis and cystic lung disease in primary Sjögren’s syndrome: A case report and literature review. Clin. Rheumatol. 2021, 40, 3345–3350. [Google Scholar] [CrossRef] [PubMed]
  123. Hernandez-Molina, G.; Faz-Munoz, D.; Astudillo-Angel, M.; Iturralde-Chavez, A.; Reyes, E. Coexistance of Amyloidosis and Primary Sjögren’s Syndrome: An Overview. Curr. Rheumatol. Rev. 2018, 14, 231–238. [Google Scholar] [CrossRef] [PubMed]
  124. Arakawa, H.; Tanese, K.; Tanaka, R.; Murakami, K.; Sujino, K.; Miyamoto, J.; Amagai, M.; Tanikawa, A. Efficacy of hydroxychloroquine for treating annular erythema associated with Sjögren’s syndrome. J. Dermatol. 2021, 48, 1526–1532. [Google Scholar] [CrossRef]
  125. Brito-Zerón, P.; Retamozo, S.; Akasbi, M.; Gandía, M.; Perez-De-Lis, M.; Soto-Cardenas, M.J.; Diaz-Lagares, C.; Kostov, B.; Bove, A.; Bosch, X.; et al. Annular erythema in primary Sjogren’s syndrome: Description of 43 non-Asian cases. Lupus 2014, 23, 166–175. [Google Scholar] [CrossRef]
  126. Olewicz-Gawlik, A.; Polańska, A.; Trzybulska, D.; Nowak-Gabryel, M.; Błochowiak, K.; Kocięcki, J.; Sokalski, J.; Żaba, R.; Adamski, Z.; Dańczak-Pazdrowska, A. Skin Barrier Function in Patients with Primary and Secondary Sjögren’s Syndrome. Acta Dermatovenerol. Croat. 2018, 26, 153–156. [Google Scholar]
  127. Katsiougiannis, S.; Tenta, R.; Skopouli, F.N. Autoimmune epithelitis (Sjögren’s syndrome); the impact of metabolic status of glandular epithelial cells on auto-immunogenicity. J. Autoimmun. 2019, 104, 102335. [Google Scholar] [CrossRef]
  128. Ogawa, Y.; Takeuchi, T.; Tsubota, K. Autoimmune Epithelitis and Chronic Inflammation in Sjögren’s Syndrome-Related Dry Eye Disease. Int. J. Mol. Sci. 2021, 22, 11820. [Google Scholar] [CrossRef]
  129. Takeo, N.; Sakai, T.; Saito-Shono, T.; Ishikawa, K.; Hatano, Y.; Katagiri, K.; Takahashi, Y.; Kawano, K.; Kimoto, K.; Kubota, T.; et al. Three cases of pigmented cosmetic dermatitis-like eruptions associated with primary Sjögren’s syndrome or anti-SSA antibody. J. Dermatol. 2016, 43, 947–950. [Google Scholar] [CrossRef]
  130. Jordán-González, P.; Gago-Piñero, R.; Varela-Rosario, N.; Pérez-Ríos, N.; Vilá, L.M. Characterization of a subset of patients with primary Sjögren’s syndrome initially presenting with C3 or C4 hypocomplementemia. Eur. J. Rheumatol. 2020, 7, 112–117. [Google Scholar] [CrossRef]
  131. Fraticelli, P.; Benfaremo, D.; Gabrielli, A. Diagnosis and management of leukocytoclastic vasculitis. Intern. Emerg. Med. 2021, 16, 831–841. [Google Scholar] [CrossRef] [PubMed]
  132. Mihai, A.; Mardale, D.; Opris-Belinski, D.; Ionescu, R.; Jurcut, C. Neutrophil to lymphocyte ratio independently predicts cutaneous manifestations in patients with Sjogren syndrome. Ann. Rheum. Dis. 2020, 79, 1513–1514. [Google Scholar] [CrossRef]
  133. Argyropoulou, O.D.; Tzioufas, A.G. Common and rare forms of vasculitis associated with Sjögren’s syndrome. Curr. Opin. Rheumatol. 2020, 32, 21–28. [Google Scholar] [CrossRef] [PubMed]
  134. Chevalier, K.; Belkhir, R.; Seror, R.; Mariette, X.; Nocturne, G. Efficacity of a sequential treatment by anti-CD 20 monoclonal antibody and belimumab in type II cryoglobulinaemia associated with primary Sjögren syndrome refractory to rituximab alone. Ann. Rheum. Dis. 2020, 79, 1257–1259. [Google Scholar] [CrossRef]
  135. Pouchelon, C.; Visentini, M.; Emmi, G.; le Guern, V.; Quartuccio, L.; Samson, M.; Venhoff, N.; Briantais, A.; Casato, M.; Chatelus, E.; et al. Management of nonviral mixed cryoglobulinemia vasculitis refractory to rituximab: Data from a European collaborative study and review of the literature. Autoimmun. Rev. 2022, 21, 103034. [Google Scholar] [CrossRef]
  136. Weidner, T.; Illing, T.; Elsner, P. Primary Localized Cutaneous Amyloidosis: A Systematic Treatment Review. Am. J. Clin. Dermatol. 2017, 18, 629–642. [Google Scholar] [CrossRef]
  137. Ye, W.; Chen, S.; Huang, X.; Qin, W.; Zhang, T.; Zhu, X.; Zhu, X.; Lin, C.; Wang, X. Clinical features and risk factors of neurological involvement in Sjögren’s syndrome. BMC Neurosci. 2018, 19, 26. [Google Scholar] [CrossRef] [Green Version]
  138. Cafaro, G.; Perricone, C.; Carubbi, F.; Baldini, C.; Quartuccio, L.; Priori, R.; Berardicurti, O.; Ferro, F.; Gandolfo, S.; Gattamelata, A.; et al. Peripheral Nervous System Involvement in Sjögren’s Syndrome: Analysis of a Cohort From the Italian Research Group on Sjögren’s Syndrome. Front. Immunol. 2021, 12, 615656. [Google Scholar] [CrossRef]
  139. Alunno, A.; Carubbi, F.; Bartoloni, E.; Cipriani, P.; Giacomelli, R.; Gerli, R. The kaleidoscope of neurological manifestations in primary Sjögren’s syndrome. Clin. Exp. Rheumatol. 2019, 37 (Suppl. S118), 192–198. [Google Scholar]
  140. Carvajal Alegria, G.; Guellec, D.; Mariette, X.; Gottenberg, J.E.; Dernis, E.; Dubost, J.J.; Trouvin, A.P.; Hachulla, E.; Larroche, C.; Le Guern, V.; et al. Epidemiology of neurological manifestations in Sjögren’s syndrome: Data from the French ASSESS Cohort. RMD Open 2016, 2, e000179. [Google Scholar] [CrossRef] [Green Version]
  141. Perzyńska-Mazan, J.; Maślińska, M.; Gasik, R. Neurological manifestations of primary Sjögren’s syndrome. Reumatologia 2018, 56, 99–105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Mekinian, A.; Tennenbaum, J.; Lahuna, C.; Dellal, A.; Belfeki, N.; Capron, J.; Januel, E.; Stankoff, B.; Alamowitch, S.; Fain, O. Primary Sjögren’s syndrome: Central and peripheral nervous system involvements. Clin. Exp. Rheumatol. 2020, 38 (Suppl. S126), 103–109. [Google Scholar] [PubMed]
  143. Alinezhad, N.; Habibagahi, Z.; Ostovan, V.R. Bilateral facial palsy as the first presentation of primary Sjögren’s syndrome: A case report and review of literature. Acta Neurol. Belg. 2020, 120, 999–1001. [Google Scholar] [CrossRef] [PubMed]
  144. McCoy, S.S.; Baer, A.N. Neurological Complications of Sjögren’s Syndrome: Diagnosis and Management. Curr. Treatm. Opt. Rheumatol. 2017, 3, 275–288. [Google Scholar] [CrossRef]
  145. Blaes, F. Diagnosis and therapeutic options for peripheral vasculitic neuropathy. Adv. Musculoskelet. Dis. 2015, 7, 45–55. [Google Scholar] [CrossRef] [Green Version]
  146. Davies, K.; Ng, W.F. Autonomic Nervous System Dysfunction in Primary Sjögren’s Syndrome. Front. Immunol. 2021, 12, 702505. [Google Scholar] [CrossRef]
  147. Goodman, B.P.; Crepeau, A.; Dhawan, P.S.; Khoury, J.A.; Harris, L.A. Spectrum of Autonomic Nervous System Impairment in Sjögren Syndrome. Neurologist 2017, 22, 127–130. [Google Scholar] [CrossRef]
  148. Baldini, C.; Zabotti, A.; Filipovic, N.; Vukicevic, A.; Luciano, N.; Ferro, F.; Lorenzon, M.; De Vita, S. Imaging in primary Sjögren’s syndrome: The ‘obsolete and the new’. Clin. Exp. Rheumatol. 2018, 36 (Suppl. S112), 215–221. [Google Scholar]
  149. Abdelhakim, S.; Klapholz, J.D.; Roy, B.; Weiss, S.A.; McGuone, D.; Corbin, Z.A. Mononeuritis multiplex as a rare and severe neurological complication of immune checkpoint inhibitors: A case report. J. Med. Case Rep. 2022, 16, 81. [Google Scholar] [CrossRef]
  150. Siao, P.; Kaku, M. A Clinician’s Approach to Peripheral Neuropathy. Semin. Neurol. 2019, 39, 519–530. [Google Scholar] [CrossRef]
  151. Sivadasan, A.; Muthusamy, K.; Patel, B.; Benjamin, R.N.; Prabhakar, A.T.; Mathew, V.; Aaron, S.; Alexander, M. Clinical Spectrum, Therapeutic Outcomes, and Prognostic Predictors in Sjogren’s Syndrome-associated Neuropathy. Ann. Indian Acad. Neurol. 2017, 20, 278–283. [Google Scholar] [CrossRef] [PubMed]
  152. Gwathmey, K.G.; Satkowiak, K. Peripheral nervous system manifestations of rheumatological diseases. J. Neurol. Sci. 2021, 424, 117421. [Google Scholar] [CrossRef] [PubMed]
  153. Pars, K.; Pul, R.; Schwenkenbecher, P.; Sühs, K.W.; Wurster, U.; Witte, T.; Bronzlik, P.; Stangel, M.; Skripuletz, T. Cerebrospinal Fluid Findings in Neurological Diseases Associated with Sjögren’s Syndrome. Eur. Neurol. 2017, 77, 91–102. [Google Scholar] [CrossRef] [PubMed]
  154. Butryn, M.; Neumann, J.; Rolfes, L.; Bartels, C.; Wattjes, M.P.; Mahmoudi, N.; Seeliger, T.; Konen, F.F.; Thiele, T.; Witte, T.; et al. Clinical, Radiological, and Laboratory Features of Spinal Cord Involvement in Primary Sjögren’s Syndrome. J. Clin. Med. 2020, 9, 1482. [Google Scholar] [CrossRef] [PubMed]
  155. Koçer, B.; Tezcan, M.E.; Batur, H.Z.; Haznedaroğlu, Ş.; Göker, B.; İrkeç, C.; Çetinkaya, R. Cognition, depression, fatigue, and quality of life in primary Sjögren’s syndrome: Correlations. Brain Behav. 2016, 6, e00586. [Google Scholar] [CrossRef]
  156. Goodman, B.P. Immunoresponsive Autonomic Neuropathy in Sjögren Syndrome-Case Series and Literature Review. Am. J. Ther. 2019, 26, e66–e71. [Google Scholar] [CrossRef]
  157. Jamilloux, Y.; Magy, L.; Hurtevent, J.F.; Gondran, G.; de Seze, J.; Launay, D.; Ly, K.H.; Lambert, M.; Hachulla, E.; Hatron, P.Y.; et al. Immunological profiles determine neurological involvement in Sjögren’s syndrome. Eur. J. Intern. Med. 2014, 25, 177–181. [Google Scholar] [CrossRef]
  158. Mihai, A.C.; Chitimus, D.M.; Jurcut, C.; Blajut, F.C.; Opris-Belinski, D.; Caruntu, C.; Ionescu, R.; Caruntu, A. Comparative Analysis of Hematological and Immunological Parameters in Patients with Primary Sjögren’s Syndrome and Peripheral Neuropathy. J. Clin. Med. 2022, 12, 3672. [Google Scholar] [CrossRef]
  159. Sène, D. Small fiber neuropathy: Diagnosis, causes, and treatment. Jt. Bone Spine 2018, 85, 553–559. [Google Scholar] [CrossRef]
  160. Trouvin, A.P.; Perrot, S.; Lloret-Linares, C. Efficacy of Venlafaxine in Neuropathic Pain: A Narrative Review of Optimized Treatment. Clin. Ther. 2017, 39, 1104–1122. [Google Scholar] [CrossRef] [Green Version]
  161. Omdal, R.; Mellgren, S.I.; Norheim, K.B. Pain and fatigue in primary Sjögren’s syndrome. Rheumatology 2021, 60, 3099–3106. [Google Scholar] [CrossRef]
  162. Liampas, A.; Parperis, K.; Erotocritou, M.F.; Nteveros, A.; Papadopoulou, M.; Moschovos, C.; Akil, M.; Coaccioli, S.; Hadjigeorgiou, G.M.; Hadjivassiliou, M.; et al. Primary Sjögren syndrome-related peripheral neuropathy: A systematic review and meta-analysis. Eur. J. Neurol. 2023, 30, 255–265. [Google Scholar] [CrossRef] [PubMed]
  163. Letaief, H.; Lukas, C.; Barnetche, T.; Gaujoux-Viala, C.; Combe, B.; Morel, J. Efficacy and safety of biological DMARDs modulating B cells in primary Sjögren’s syndrome: Systematic review and meta-analysis. Jt. Bone Spine 2018, 85, 15–22. [Google Scholar] [CrossRef] [PubMed]
  164. François, H.; Mariette, X. Renal involvement in Sjögren’s syndrome. Néphrologie Thérapeutique 2020, 16, 440–452. [Google Scholar] [CrossRef] [PubMed]
  165. Sandhya, P. Comprehensive analysis of clinical and laboratory features of 440 published cases of Sjögren’s syndrome and renal tubular acidosis. Int. J. Rheum. Dis. 2023, 26, 278–285. [Google Scholar] [CrossRef] [PubMed]
  166. Jung, S.W.; Park, E.J.; Kim, J.S.; Lee, T.W.; Ihm, C.G.; Lee, S.H.; Moon, J.Y.; Kim, Y.G.; Jeong, K.H. Renal Tubular Acidosis in Patients with Primary Sjögren’s Syndrome. Electrolyte Blood Press. 2017, 15, 17–22. [Google Scholar] [CrossRef] [Green Version]
  167. Aiyegbusi, O.; McGregor, L.; McGeoch, L.; Kipgen, D.; Geddes, C.C.; Stevens, K.I. Renal Disease in Primary Sjögren’s Syndrome. Rheumatol. Ther. 2021, 8, 63–80. [Google Scholar] [CrossRef]
  168. Evans, R.D.; Laing, C.M.; Ciurtin, C.; Walsh, S.B. Tubulointerstitial nephritis in primary Sjögren syndrome: Clinical manifestations and response to treatment. BMC Musculoskelet. Disord. 2016, 17, 2. [Google Scholar] [CrossRef] [Green Version]
  169. Jasiek, M.; Karras, A.; Le Guern, V.; Krastinova, E.; Mesbah, R.; Faguer, S.; Jourde-Chiche, N.; Fauchais, A.L.; Chiche, L.; Dernis, E.; et al. A multicentre study of 95 biopsy-proven cases of renal disease in primary Sjögren’s syndrome. Rheumatology 2017, 56, 362–370. [Google Scholar] [CrossRef] [Green Version]
  170. Chatterjee, R.; Balakrishnan, A.; Kharbanda, R.; Rathore, U.; Gupta, L.; Misra, D.P.; Agarwal, V.; Aggarwal, A.; Lawrence, A. Renal involvement in Sjőgren’s syndrome: Predictors and impact on patient outcomes. Rheumatol. Int. 2023, 43, 1297–1306. [Google Scholar] [CrossRef]
  171. François, H.; Mariette, X. Renal involvement in primary Sjögren syndrome. Nat. Rev. Nephrol. 2016, 12, 82–93. [Google Scholar] [CrossRef] [PubMed]
  172. Du, T.; Liu, X.; Ye, W.; Ye, W.; Li, C. Primary Sjögren syndrome-associated acute interstitial nephritis and type 3 renal tubular acidosis in a patient with thin basement membrane nephropathy: A case report. Medicine 2020, 99, e21644. [Google Scholar] [CrossRef] [PubMed]
  173. Koratala, A.; Reeves, W.H.; Segal, M.S. Tubulointerstitial Nephritis in Sjögren Syndrome Treated With Mycophenolate Mofetil. J. Clin. Rheumatol. 2017, 23, 402–403. [Google Scholar] [CrossRef] [PubMed]
  174. Chen, R.; Wang, J.; Xie, Q.; Xue, J.; Hao, C. Sjögren’s syndrome complicated with membranous nephropathy, a cause or coincidence? Int. J. Rheum. Dis. 2021, 24, 1086–1094. [Google Scholar] [CrossRef]
  175. García-Carrasco, M.; Mendoza-Pinto, C.; Jiménez-Hernández, C.; Jiménez-Hernández, M.; Nava-Zavala, A.; Riebeling, C. Serologic features of primary Sjögren’s syndrome: Clinical and prognostic correlation. Int. J. Clin. Rheumtol. 2012, 7, 651–659. [Google Scholar] [CrossRef] [Green Version]
  176. Stergiou, I.E.; Kapsogeorgou, E.E.; Tzioufas, A.G.; Voulgarelis, M.; Goules, A.V. Clinical Phenotype and Mechanisms of Leukopenia/Neutropenia in Patients with Primary Sjögren’s Syndrome. Mediterr. J. Rheumatol. 2022, 33, 99–101. [Google Scholar] [CrossRef]
  177. Koh, J.H.; Lee, J.; Chung, S.H.; Kwok, S.K.; Park, S.H. Relation of Autoimmune Cytopenia to Glandular and Systemic Manifestations in Primary Sjögren Syndrome: Analysis of 113 Korean Patients. J. Rheumatol. 2015, 42, 1817–1824. [Google Scholar] [CrossRef]
  178. Wei, L.; Zhifei, X.; Xiaoran, N.; Meilu, L.; Yang, L.; Yixuan, L.; Xiuying, R.; Yashuang, S.; Jingjing, C.; Shaoying, G.; et al. Patients with early-onset primary Sjögren’s syndrome have distinctive clinical manifestations and circulating lymphocyte profiles. Rheumatology 2022, 61, 597–605. [Google Scholar] [CrossRef]
  179. Malladi, A.S.; Sack, K.E.; Shiboski, S.C.; Shiboski, C.H.; Baer, A.N.; Banushree, R.; Dong, Y.; Helin, P.; Kirkham, B.W.; Li, M.; et al. Primary Sjögren’s syndrome as a systemic disease: A study of participants enrolled in an international Sjögren’s syndrome registry. Arthritis Care Res. 2012, 64, 911–918. [Google Scholar] [CrossRef] [Green Version]
  180. Flores-Chávez, A.; Kostov, B.; Solans, R.; Fraile, G.; Maure, B.; Feijoo-Massó, C.; Rascón, F.J.; Pérez-Alvarez, R.; Zamora-Pasadas, M.; García-Pérez, A.; et al. Severe, life-threatening phenotype of primary Sjögren’s syndrome: Clinical characterisation and outcomes in 1580 patients (GEAS-SS Registry). Clin. Exp. Rheumatol. 2018, 36 (Suppl. S112), 121–129. [Google Scholar]
  181. Wu, C.; Yang, P.; Liu, H.; Xiao, W.; Zhao, L. Increased frequency of CCR7(+)CD4(+) T cells from patients with primary Sjögren’s syndrome: An indicator of disease activity rather than of damage severity. Cytokine 2018, 110, 9–17. [Google Scholar] [CrossRef] [PubMed]
  182. Luo, J.; Song, W.J.; Chen, J.Q.; Yang, G.Y.; Yang, J.Y.; Yu, X.B.; Huang, Z.W.; Xu, Y.; Wang, J.M.; Tao, Q.W. Factors associated with secondary immune thrombocytopenia in patients with primary Sjögren’s syndrome: A retrospective study of 639 cases. Clin. Exp. Rheumatol. 2022, 40, 2245–2252. [Google Scholar] [CrossRef] [PubMed]
  183. Ter Borg, E.J.; Kelder, J.C. Is extra-glandular organ damage in primary Sjögren’s syndrome related to the presence of systemic auto-antibodies and/or hypergammaglobulinemia? A long-term cohort study with 110 patients from the Netherlands. Int. J. Rheum. Dis. 2017, 20, 875–881. [Google Scholar] [CrossRef] [PubMed]
  184. Goules, A.V.; Tzioufas, A.G. Lymphomagenesis in Sjögren’s syndrome: Predictive biomarkers towards precision medicine. Autoimmun. Rev. 2019, 18, 137–143. [Google Scholar] [CrossRef]
  185. Sebastian, A.; Madej, M.; Sebastian, M.; Butrym, A.; Woytala, P.; Hałoń, A.; Wiland, P. Prevalence and clinical presentation of lymphoproliferative disorder in patients with primary Sjögren’s syndrome. Rheumatol. Int. 2020, 40, 399–404. [Google Scholar] [CrossRef] [Green Version]
  186. Retamozo, S.; Brito-Zerón, P.; Ramos-Casals, M. Prognostic markers of lymphoma development in primary Sjögren syndrome. Lupus 2019, 28, 923–936. [Google Scholar] [CrossRef]
  187. Ramos-Casals, M.; Brito-Zerón, P.; Solans, R.; Camps, M.T.; Casanovas, A.; Sopeña, B.; Díaz-López, B.; Rascón, F.J.; Qanneta, R.; Fraile, G.; et al. Systemic involvement in primary Sjogren’s syndrome evaluated by the EULAR-SS disease activity index: Analysis of 921 Spanish patients (GEAS-SS Registry). Rheumatology 2014, 53, 321–331. [Google Scholar] [CrossRef] [Green Version]
  188. Ambrosi, A.; Wahren-Herlenius, M. Update on the immunobiology of Sjögren’s syndrome. Curr. Opin. Rheumatol. 2015, 27, 468–475. [Google Scholar] [CrossRef]
  189. Blokland, S.L.M.; Hillen, M.R.; Kruize, A.A.; Meller, S.; Homey, B.; Smithson, G.M.; Radstake, T.; van Roon, J.A.G. Increased CCL25 and T Helper Cells Expressing CCR9 in the Salivary Glands of Patients With Primary Sjögren’s Syndrome: Potential New Axis in Lymphoid Neogenesis. Arthritis Rheumatol. 2017, 69, 2038–2051. [Google Scholar] [CrossRef] [Green Version]
  190. Hauser, M.A.; Legler, D.F. Common and biased signaling pathways of the chemokine receptor CCR7 elicited by its ligands CCL19 and CCL21 in leukocytes. J. Leukoc. Biol. 2016, 99, 869–882. [Google Scholar] [CrossRef] [Green Version]
  191. Zhu, H.; Zheng, J.; Zhou, Y.; Wu, T.; Zhu, T. Knockdown of RSAD2 attenuates B cell hyperactivity in patients with primary Sjögren’s syndrome (pSS) via suppressing NF-κb signaling pathway. Mol. Cell. Biochem. 2021, 476, 2029–2037. [Google Scholar] [CrossRef] [PubMed]
  192. Chiu, Y.H.; Chung, C.H.; Lin, K.T.; Lin, C.S.; Chen, J.H.; Chen, H.C.; Huang, R.Y.; Wu, C.T.; Liu, F.C.; Chien, W.C. Predictable biomarkers of developing lymphoma in patients with Sjögren syndrome: A nationwide population-based cohort study. Oncotarget 2017, 8, 50098–50108. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Bowman, S.J. Primary Sjögren’s syndrome. Lupus 2018, 27, 32–35. [Google Scholar] [CrossRef]
  194. Manfrè, V.; Giovannini, I.; Zandonella Callegher, S.; Lorenzon, M.; Pegolo, E.; Tel, A.; Gandolfo, S.; Quartuccio, L.; De Vita, S.; Zabotti, A. Ultrasound and Bioptic Investigation of Patients with Primary Sjögren’s Syndrome. J. Clin. Med. 2021, 10, 1171. [Google Scholar] [CrossRef] [PubMed]
  195. Nocturne, G.; Pontarini, E.; Bombardieri, M.; Mariette, X. Lymphomas complicating primary Sjögren’s syndrome: From autoimmunity to lymphoma. Rheumatology 2021, 60, 3513–3521. [Google Scholar] [CrossRef] [PubMed]
  196. Bautista-Vargas, M.; Vivas, A.J.; Tobón, G.J. Minor salivary gland biopsy: Its role in the classification and prognosis of Sjögren’s syndrome. Autoimmun. Rev. 2020, 19, 102690. [Google Scholar] [CrossRef] [PubMed]
  197. Retamozo, S.; Sisó-Almirall, A.; Flores-Chávez, A.; Ramos-Casals, M.; Brito-Zerón, P. An update of targeted therapeutic options for primary Sjögren syndrome: Current status and future development. Expert Opin. Pharm. 2021, 22, 2359–2371. [Google Scholar] [CrossRef] [PubMed]
  198. Priori, R.; Mastromanno, L.; Izzo, R. What about glucocorticoids in primary Sjögren’s syndrome? Clin. Exp. Rheumatol. 2020, 38 (Suppl. S126), 237–244. [Google Scholar] [PubMed]
  199. Gandolfo, S.; De Vita, S. Double anti-B cell and anti-BAFF targeting for the treatment of primary Sjögren’s syndrome. Clin. Exp. Rheumatol. 2019, 37 (Suppl. S118), 199–208. [Google Scholar]
  200. Routsias, J.G.; Goules, J.D.; Charalampakis, G.; Tzima, S.; Papageorgiou, A.; Voulgarelis, M. Malignant lymphoma in primary Sjögren’s syndrome: An update on the pathogenesis and treatment. Semin. Arthritis Rheum. 2013, 43, 178–186. [Google Scholar] [CrossRef]
  201. Mariette, X.; Tubach, F.; Bagheri, H.; Bardet, M.; Berthelot, J.M.; Gaudin, P.; Heresbach, D.; Martin, A.; Schaeverbeke, T.; Salmon, D.; et al. Lymphoma in patients treated with anti-TNF: Results of the 3-year prospective French RATIO registry. Ann. Rheum. Dis. 2010, 69, 400–408. [Google Scholar] [CrossRef] [PubMed]
  202. Kampolis, C.F.; Fragkioudaki, S.; Mavragani, C.P.; Zormpala, A.; Samakovli, A.; Moutsopoulos, H.M. Prevalence and spectrum of symptomatic pulmonary involvement in primary Sjögren’s syndrome. Clin. Exp. Rheumatol. 2018, 36 (Suppl. S112), 94–101. [Google Scholar] [PubMed]
  203. Roca, F.; Dominique, S.; Schmidt, J.; Smail, A.; Duhaut, P.; Lévesque, H.; Marie, I. Interstitial lung disease in primary Sjögren’s syndrome. Autoimmun. Rev. 2017, 16, 48–54. [Google Scholar] [CrossRef] [PubMed]
  204. Luppi, F.; Sebastiani, M.; Sverzellati, N.; Cavazza, A.; Salvarani, C.; Manfredi, A. Lung complications of Sjogren syndrome. Eur. Respir. Rev. Off. J. Eur. Respir. Soc. 2020, 29, 200021. [Google Scholar] [CrossRef] [PubMed]
  205. Stojan, G.; Baer, A.N.; Danoff, S.K. Pulmonary manifestations of Sjögren’s syndrome. Curr. Allergy Asthma Rep. 2013, 13, 354–360. [Google Scholar] [CrossRef] [Green Version]
  206. Flament, T.; Bigot, A.; Chaigne, B.; Henique, H.; Diot, E.; Marchand-Adam, S. Pulmonary manifestations of Sjögren’s syndrome. Eur. Respir. Rev. Off. J. Eur. Respir. Soc. 2016, 25, 110–123. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Kreider, M.; Highland, K. Pulmonary involvement in Sjögren syndrome. Semin. Respir. Crit. Care Med. 2014, 35, 255–264. [Google Scholar] [CrossRef] [Green Version]
  208. Sogkas, G.; Hirsch, S.; Olsson, K.M.; Hinrichs, J.B.; Thiele, T.; Seeliger, T.; Skripuletz, T.; Schmidt, R.E.; Witte, T.; Jablonka, A.; et al. Lung Involvement in Primary Sjögren’s Syndrome-An Under-Diagnosed Entity. Front. Med. 2020, 7, 332. [Google Scholar] [CrossRef]
  209. Dong, X.; Zhou, J.; Guo, X.; Li, Y.; Xu, Y.; Fu, Q.; Lu, Y.; Zheng, Y. A retrospective analysis of distinguishing features of chest HRCT and clinical manifestation in primary Sjögren’s syndrome-related interstitial lung disease in a Chinese population. Clin. Rheumatol. 2018, 37, 2981–2988. [Google Scholar] [CrossRef]
  210. Mandl, T.; Diaz, S.; Ekberg, O.; Hesselstrand, R.; Piitulainen, E.; Wollmer, P.; Theander, E. Frequent development of chronic obstructive pulmonary disease in primary SS—Results of a longitudinal follow-up. Rheumatology 2012, 51, 941–946. [Google Scholar] [CrossRef] [Green Version]
  211. Chatzis, L.; Vlachoyiannopoulos, P.G.; Tzioufas, A.G.; Goules, A.V. New frontiers in precision medicine for Sjogren’s syndrome. Expert Rev. Clin. Immunol. 2021, 17, 127–141. [Google Scholar] [CrossRef] [PubMed]
  212. Gupta, S.; Ferrada, M.A.; Hasni, S.A. Pulmonary Manifestations of Primary Sjögren’s Syndrome: Underlying Immunological Mechanisms, Clinical Presentation, and Management. Front. Immunol. 2019, 10, 1327. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Kakugawa, T.; Sakamoto, N.; Ishimoto, H.; Shimizu, T.; Nakamura, H.; Nawata, A.; Ito, C.; Sato, S.; Hanaka, T.; Oda, K.; et al. Lymphocytic focus score is positively related to airway and interstitial lung diseases in primary Sjögren’s syndrome. Respir. Med. 2018, 137, 95–102. [Google Scholar] [CrossRef] [Green Version]
  214. Cereser, L.; Giovannini, I.; Caronia, G.; Zabotti, A.; De Vita, S.; Zuiani, C.; Quartuccio, L.; Girometti, R. Chest high-resolution computed tomography in primary Sjögren’s syndrome: An up-to-date primer for rheumatologists. Clin. Exp. Rheumatol. 2022, 40, 2450–2462. [Google Scholar] [CrossRef]
  215. Kim, Y.J.; Choe, J.; Kim, H.J.; Song, J.W. Long-term clinical course and outcome in patients with primary Sjögren syndrome-associated interstitial lung disease. Sci. Rep. 2021, 11, 12827. [Google Scholar] [CrossRef] [PubMed]
  216. Natalini, J.G.; Johr, C.; Kreider, M. Pulmonary Involvement in Sjögren Syndrome. Clin. Chest Med. 2019, 40, 531–544. [Google Scholar] [CrossRef]
  217. Gao, H.; Sun, Y.; Zhang, X.Y.; Xie, L.; Zhang, X.W.; Zhong, Y.C.; Zhang, J.; Hou, Y.K.; Li, Z.G. Characteristics and mortality in primary Sjögren syndrome-related interstitial lung disease. Medicine 2021, 100, e26777. [Google Scholar] [CrossRef]
  218. Lin, W.; Xin, Z.; Zhang, J.; Liu, N.; Ren, X.; Liu, M.; Su, Y.; Liu, Y.; Yang, L.; Guo, S.; et al. Interstitial lung disease in Primary Sjögren’s syndrome. BMC Pulm. Med. 2022, 22, 73. [Google Scholar] [CrossRef]
  219. Dong, X.; Gao, Y.L.; Lu, Y.; Zheng, Y. Characteristics of primary Sjögren’s syndrome related lymphocytic interstitial pneumonia. Clin. Rheumatol. 2021, 40, 601–612. [Google Scholar] [CrossRef]
  220. Wang, Y.; Zhao, S.; Du, G.; Ma, S.; Lin, Q.; Lin, J.; Zheng, K.; Zhang, G.; Matucci-Cerinic, M. Acute fibrinous and organizing pneumonia as initial presentation of primary Sjögren’s syndrome: A case report and literature review. Clin. Rheumatol. 2018, 37, 2001–2005. [Google Scholar] [CrossRef]
  221. Zhang, T.; Yuan, F.; Xu, L.; Sun, W.; Liu, L.; Xue, J. Characteristics of patients with primary Sjögren’s syndrome associated interstitial lung disease and relevant features of disease progression. Clin. Rheumatol. 2020, 39, 1561–1568. [Google Scholar] [CrossRef] [PubMed]
  222. He, S.H.; He, Y.J.; Guo, K.J.; Liang, X.; Li, S.S.; Li, T.F. Risk factors for progression of interstitial lung disease in Sjögren’s syndrome: A single-centered, retrospective study. Clin. Rheumatol. 2022, 41, 1153–1161. [Google Scholar] [CrossRef] [PubMed]
  223. Reina, D.; Roig Vilaseca, D.; Torrente-Segarra, V.; Cerdà, D.; Castellví, I.; Díaz Torné, C.; Moreno, M.; Narváez, J.; Ortiz, V.; Blavia, R.; et al. Sjögren’s syndrome-associated interstitial lung disease: A multicenter study. Reum. Clin. 2016, 12, 201–205. [Google Scholar] [CrossRef] [PubMed]
  224. Amlani, B.; Elsayed, G.; Barvalia, U.; Kanne, J.P.; Meyer, K.C.; Sandbo, N.; Li, Z.; McCoy, S.S. Treatment of primary sjögren’s syndrome-related interstitial lung disease: A retrospective cohort study. Sarcoidosis Vasc. Diffus. Lung Dis. 2020, 37, 136–147. [Google Scholar] [CrossRef]
  225. Lee, A.S.; Scofield, R.H.; Hammitt, K.M.; Gupta, N.; Thomas, D.E.; Moua, T.; Ussavarungsi, K.; St Clair, E.W.; Meehan, R.; Dunleavy, K.; et al. Consensus Guidelines for Evaluation and Management of Pulmonary Disease in Sjögren’s. Chest 2021, 159, 683–698. [Google Scholar] [CrossRef]
  226. Kamiya, Y.; Fujisawa, T.; Kono, M.; Nakamura, H.; Yokomura, K.; Koshimizu, N.; Toyoshima, M.; Imokawa, S.; Sumikawa, H.; Johkoh, T.; et al. Prognostic factors for primary Sjögren’s syndrome-associated interstitial lung diseases. Respir. Med. 2019, 159, 105811. [Google Scholar] [CrossRef]
  227. Robles-Perez, A.; Dorca, J.; Castellví, I.; Nolla, J.M.; Molina-Molina, M.; Narváez, J. Rituximab effect in severe progressive connective tissue disease-related lung disease: Preliminary data. Rheumatol. Int. 2020, 40, 719–726. [Google Scholar] [CrossRef]
  228. Klinowski, G.; Gozzi, F.; Trentacosti, F.; Andrisani, D.; Sebastiani, M.; Clini, E.M. Rituximab for the treatment of acute onset Interstitial Lung Disease in primary Sjogren’s syndrome. Pulmonology 2021, 27, 575–578. [Google Scholar] [CrossRef]
  229. Thompson, G.; McLean-Tooke, A.; Wrobel, J.; Lavender, M.; Lucas, M. Sjögren Syndrome With Associated Lymphocytic Interstitial Pneumonia Successfully Treated With Tacrolimus and Abatacept as an Alternative to Rituximab. Chest 2018, 153, e41–e43. [Google Scholar] [CrossRef] [Green Version]
  230. Crincoli, V.; Di Comite, M.; Guerrieri, M.; Rotolo, R.P.; Limongelli, L.; Tempesta, A.; Iannone, F.; Rinaldi, A.; Lapadula, G.; Favia, G. Orofacial Manifestations and Temporomandibular Disorders of Sjögren Syndrome: An Observational Study. Int. J. Med. Sci. 2018, 15, 475–483. [Google Scholar] [CrossRef] [Green Version]
  231. Vivino, F.B.; Bunya, V.Y.; Massaro-Giordano, G.; Johr, C.R.; Giattino, S.L.; Schorpion, A.; Shafer, B.; Peck, A.; Sivils, K.; Rasmussen, A.; et al. Sjogren’s syndrome: An update on disease pathogenesis, clinical manifestations and treatment. Clin. Immunol. 2019, 203, 81–121. [Google Scholar] [CrossRef] [PubMed]
  232. Xu, L.; Gui, M.; Sun, C.; Yau, V.; Sun, C.; Qi, J. Refractory diarrhea in a patient with Sjogren’s syndrome: A case report. Front. Nutr. 2023, 10, 1086967. [Google Scholar] [CrossRef] [PubMed]
  233. Grossman, S.; Tagliavini, L.B. Managing Sjogren’s Syndrome. Home Healthc. Now 2015, 33, 487–492. [Google Scholar] [CrossRef] [PubMed]
  234. Bartoloni, E.; Bistoni, O.; Alunno, A.; Cavagna, L.; Nalotto, L.; Baldini, C.; Priori, R.; Fischetti, C.; Fredi, M.; Quartuccio, L.; et al. Celiac Disease Prevalence is Increased in Primary Sjögren’s Syndrome and Diffuse Systemic Sclerosis: Lessons from a Large Multi-Center Study. J. Clin. Med. 2019, 8, 540. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Popov, Y.; Salomon-Escoto, K. Gastrointestinal and Hepatic Disease in Sjogren Syndrome. Rheum. Dis. Clin. N. Am. 2018, 44, 143–151. [Google Scholar] [CrossRef]
  236. Balaban, D.V.; Mihai, A.; Dima, A.; Popp, A.; Jinga, M.; Jurcut, C. Celiac disease and Sjögren’s syndrome: A case report and review of literature. World J. Clin. Cases 2020, 8, 4151–4161. [Google Scholar] [CrossRef]
  237. Parreau, S.; Jacques, J.; Dumonteil, S.; Palat, S.; Geyl, S.; Gondran, G.; Bezanahary, H.; Liozon, E.; Azaïs, J.; Colombie, S.; et al. Abdominal symptoms during Sjogren’s syndrome: A pilot study. Adv. Rheumatol. 2021, 61, 5. [Google Scholar] [CrossRef]
  238. Adamec, I.; Žarković, K.; Sentić, M.; Habek, M. Autonomic failure in Sjögren’s syndrome. Clin. Auton. Res. 2016, 26, 165–166. [Google Scholar] [CrossRef]
  239. Cano-Ortiz, A.; Laborda-Illanes, A.; Plaza-Andrades, I.; Membrillo Del Pozo, A.; Villarrubia Cuadrado, A.; Rodríguez Calvo de Mora, M.; Leiva-Gea, I.; Sanchez-Alcoholado, L.; Queipo-Ortuño, M.I. Connection between the Gut Microbiome, Systemic Inflammation, Gut Permeability and FOXP3 Expression in Patients with Primary Sjögren’s Syndrome. Int. J. Mol. Sci. 2020, 21, 8733. [Google Scholar] [CrossRef]
  240. Doaré, E.; Héry-Arnaud, G.; Devauchelle-Pensec, V.; Alegria, G.C. Healthy Patients Are Not the Best Controls for Microbiome-Based Clinical Studies: Example of Sjögren’s Syndrome in a Systematic Review. Front. Immunol. 2021, 12, 699011. [Google Scholar] [CrossRef]
  241. Deng, X.; Li, J.; Hou, S.; Ci, B.; Liu, B.; Xu, K. Prevalence and impact of Sjögren’s syndrome in primary biliary cholangitis: A systematic review and meta-analysis. Ann. Hepatol. 2022, 27, 100746. [Google Scholar] [CrossRef] [PubMed]
  242. Selmi, C.; Generali, E.; Gershwin, M.E. Rheumatic Manifestations in Autoimmune Liver Disease. Rheum. Dis. Clin. N. Am. 2018, 44, 65–87. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  243. Zeron, P.B.; Retamozo, S.; Bové, A.; Kostov, B.A.; Sisó, A.; Ramos-Casals, M. Diagnosis of Liver Involvement in Primary Sjögren Syndrome. J. Clin. Transl. Hepatol. 2013, 1, 94–102. [Google Scholar] [CrossRef] [Green Version]
  244. Hedström, A.; Kvarnström, M.; Lindberg, G.; Alsabeah, S.; Alsabeah, H.; Ndegwa, N.; Löhr, J.M.; Haas, S.L.; Vujasinovic, M. High prevalence of gastrointestinal symptoms in patients with primary Sjögren’s syndrome cannot be attributed to pancreatic exocrine insufficiency. Scand. J. Gastroenterol. 2022, 57, 1250–1256. [Google Scholar] [CrossRef] [PubMed]
  245. Assy, Z.; Bikker, F.J.; Mashhour, E.; Asadi, M.; Brand, H.S. Preferences of Sjögren’s syndrome patients regarding potential new saliva substitutes. Clin. Oral. Investig. 2022, 26, 6245–6252. [Google Scholar] [CrossRef]
  246. Azuma, N.; Katada, Y.; Sano, H. Deterioration in saliva quality in patients with Sjögren’s syndrome: Impact of decrease in salivary epidermal growth factor on the severity of intraoral manifestations. Inflamm. Regen. 2018, 38, 6. [Google Scholar] [CrossRef] [Green Version]
  247. Park, K.; Park, S.; Jackson, M.W. The inhibitory effects of antimuscarinic autoantibodies in the sera of primary Sjogren syndrome patients on the gastrointestinal motility. Mol. Immunol. 2013, 56, 583–587. [Google Scholar] [CrossRef]
  248. Liao, C.Y.; Chien, S.T.; Wang, C.C.; Chen, I.H.; Chiu, H.W.; Liu, M.Y.; Lin, C.H.; Ben, R.J.; Tsai, M.K. Sjögren’s syndrome associated with protein losing gastroenteropathy manifested by intestinal lymphangiectasia successfully treated with prednisolone and hydroxychloroquine. Lupus 2015, 24, 1552–1556. [Google Scholar] [CrossRef]
  249. Imrich, R.; Alevizos, I.; Bebris, L.; Goldstein, D.S.; Holmes, C.S.; Illei, G.G.; Nikolov, N.P. Predominant Glandular Cholinergic Dysautonomia in Patients With Primary Sjögren’s Syndrome. Arthritis Rheumatol. 2015, 67, 1345–1352. [Google Scholar] [CrossRef] [Green Version]
  250. Zhong, H.; Liu, S.; Wang, Y.; Xu, D.; Li, M.; Zhao, Y.; Zeng, X. Primary Sjögren’s syndrome is associated with increased risk of malignancies besides lymphoma: A systematic review and meta-analysis. Autoimmun. Rev. 2022, 21, 103084. [Google Scholar] [CrossRef]
  251. Price, E.J.; Baer, A.N. How to treat Sjögren’s syndrome. Rheumatology 2021, 60, 2574–2587. [Google Scholar] [CrossRef] [PubMed]
  252. Ni, P.; Men, R.; Shen, M.; Wang, T.; Huang, C.; Fan, X.; Yang, L. Concomitant Sjögren’s Syndrome Was Not Associated with a Poorer Response or Outcomes in Ursodeoxycholic Acid-Treated Patients with Primary Biliary Cholangitis. Can. J. Gastroenterol. Hepatol. 2019, 2019, 7396870. [Google Scholar] [CrossRef] [Green Version]
  253. Greverath, L.M.; Leicht, E.; Wald de Chamorro, N.; Wilde, A.B.; Steinhagen, L.M.; Lieb, C.; Schmelzle, M.; Chopra, S.; Shibolet, O.; Fischer, J.; et al. Evaluation of muscarinic acetylcholine receptor type 3 gene polymorphisms in patients with primary biliary cholangitis and primary sclerosing cholangitis. Hepatol. Res. 2020, 50, 321–329. [Google Scholar] [CrossRef]
  254. Manfrè, V.; Cafaro, G.; Riccucci, I.; Zabotti, A.; Perricone, C.; Bootsma, H.; De Vita, S.; Bartoloni, E. One year in review 2020: Comorbidities, diagnosis and treatment of primary Sjögren’s syndrome. Clin. Exp. Rheumatol. 2020, 38 (Suppl. S126), 10–22. [Google Scholar]
  255. Cai, X.; Luo, J.; Wei, T.; Qin, W.; Wang, X.; Li, X. Risk of Cardiovascular Involvement in Patients with Primary Sjögren’s Syndrome: A large-scale cross-sectional cohort study. Acta Reum. Port. 2019, 44, 71–77. [Google Scholar]
  256. Bartoloni, E.; Baldini, C.; Ferro, F.; Alunno, A.; Carubbi, F.; Cafaro, G.; Bombardieri, S.; Gerli, R.; Grossi, E. Application of artificial neural network analysis in the evaluation of cardiovascular risk in primary Sjögren’s syndrome: A novel pathogenetic scenario? Clin. Exp. Rheumatol. 2019, 37 (Suppl. S118), 133–139. [Google Scholar]
  257. Bartoloni, E.; Alunno, A.; Bistoni, O.; Caterbi, S.; Luccioli, F.; Santoboni, G.; Mirabelli, G.; Cannarile, F.; Gerli, R. Characterization of circulating endothelial microparticles and endothelial progenitor cells in primary Sjögren’s syndrome: New markers of chronic endothelial damage? Rheumatology 2015, 54, 536–544. [Google Scholar] [CrossRef] [Green Version]
  258. Beltai, A.; Barnetche, T.; Daien, C.; Lukas, C.; Gaujoux-Viala, C.; Combe, B.; Morel, J. Cardiovascular Morbidity and Mortality in Primary Sjögren’s Syndrome: A Systematic Review and Meta-Analysis. Arthritis Care Res. 2020, 72, 131–139. [Google Scholar] [CrossRef]
  259. Casian, M.; Jurcut, C.; Dima, A.; Mihai, A.; Stanciu, S.; Jurcut, R. Cardiovascular Disease in Primary Sjögren’s Syndrome: Raising Clinicians’ Awareness. Front. Immunol. 2022, 13, 865373. [Google Scholar] [CrossRef]
  260. Atzeni, F.; Sarzi-Puttini, P.; Signorello, M.C.; Gianturco, L.; Stella, D.; Boccassini, L.; Ricci, C.; Bodini, B.D.; Batticciotto, A.; De Gennaro-Colonna, V.; et al. New parameters for identifying subclinical atherosclerosis in patients with primary Sjögren’s syndrome: A pilot study. Clin. Exp. Rheumatol. 2014, 32, 361–368. [Google Scholar] [CrossRef]
  261. Ciçek, O.F.; Bayram, N.A.; Ayhan, H.; Erten, S.; Aslan, A.N.; Sarı, C.; Ozen, M.B.; Bİlen, E.; Bastuğ, S.; Durmaz, T.; et al. Assessment of the relationship between aortic stiffness and left ventricular functions with echocardiography in patients with Sjögren’s syndrome. Int. J. Rheum. Dis. 2014, 17, 658–663. [Google Scholar] [CrossRef] [PubMed]
  262. Goulabchand, R.; Roubille, C.; Montani, D.; Fesler, P.; Bourdin, A.; Malafaye, N.; Morel, J.; Arnaud, E.; Lattuca, B.; Barateau, L.; et al. Cardiovascular Events, Sleep Apnoea, and Pulmonary Hypertension in Primary Sjögren’s Syndrome: Data from the French Health Insurance Database. J. Clin. Med. 2021, 10, 5115. [Google Scholar] [CrossRef]
  263. Wang, J.; Li, M.; Wang, Q.; Zhang, X.; Qian, J.; Zhao, J.; Xu, D.; Tian, Z.; Wei, W.; Zuo, X.; et al. Pulmonary arterial hypertension associated with primary Sjögren’s syndrome: A multicentre cohort study from China. Eur. Respir. J. 2020, 56, 1902157. [Google Scholar] [CrossRef] [PubMed]
  264. Mofors, J.; Holmqvist, M.; Westermark, L.; Björk, A.; Kvarnström, M.; Forsblad-d’Elia, H.; Magnusson Bucher, S.; Eriksson, P.; Theander, E.; Mandl, T.; et al. Concomitant Ro/SSA and La/SSB antibodies are biomarkers for the risk of venous thromboembolism and cerebral infarction in primary Sjögren’s syndrome. J. Intern. Med. 2019, 286, 458–468. [Google Scholar] [CrossRef] [Green Version]
  265. Pasoto, S.G.; Chakkour, H.P.; Natalino, R.R.; Viana, V.S.; Bueno, C.; Lianza, A.C.; de Andrade, J.L.; Neto, M.L.; Fuller, R.; Bonfa, E. Lupus anticoagulant: A marker for stroke and venous thrombosis in primary Sjögren’s syndrome. Clin. Rheumatol. 2012, 31, 1331–1338. [Google Scholar] [CrossRef] [PubMed]
  266. Ramos-Casals, M.; Nardi, N.; Brito-Zerón, P.; Aguiló, S.; Gil, V.; Delgado, G.; Bové, A.; Font, J. Atypical autoantibodies in patients with primary Sjögren syndrome: Clinical characteristics and follow-up of 82 cases. Semin. Arthritis Rheum. 2006, 35, 312–321. [Google Scholar] [CrossRef]
  267. Li, J. The Role of Autoantibodies in Arrhythmogenesis. Curr. Cardiol. Rep. 2020, 23, 3. [Google Scholar] [CrossRef]
  268. Gültuna, S.; Can Sandıkçı, S.; Kaplanoğlu, H.; Aydın, F.N.; Özalp Ateş, F.S. Lipoprotein-associated phospholipase A2 and carotid intima-media thickness in primary Sjögren syndrome. Arch. Rheumatol. 2022, 37, 40–48. [Google Scholar] [CrossRef]
  269. Łuczak, A.; Małecki, R.; Kulus, M.; Madej, M.; Szahidewicz-Krupska, E.; Doroszko, A. Cardiovascular Risk and Endothelial Dysfunction in Primary Sjogren Syndrome Is Related to the Disease Activity. Nutrients 2021, 13, 2072. [Google Scholar] [CrossRef]
  270. Savale, L.; Guignabert, C.; Weatherald, J.; Humbert, M. Precision medicine and personalising therapy in pulmonary hypertension: Seeing the light from the dawn of a new era. Eur. Respir. Rev. Off. J. Eur. Respir. Soc. 2018, 27, 180004. [Google Scholar] [CrossRef] [Green Version]
  271. Popescu, M.R.; Dudu, A.; Jurcut, C.; Ciobanu, A.M.; Zagrean, A.M.; Panaitescu, A.M. A Broader Perspective on Anti-Ro Antibodies and Their Fetal Consequences-A Case Report and Literature Review. Diagnostics 2020, 10, 478. [Google Scholar] [CrossRef] [PubMed]
  272. Lazzerini, P.E.; Laghi-Pasini, F.; Boutjdir, M.; Capecchi, P.L. Anti-Ro/SSA Antibodies and the Autoimmune Long-QT Syndrome. Front. Med. 2021, 8, 730161. [Google Scholar] [CrossRef] [PubMed]
  273. Qin, L.; Zhang, Y.; Yang, X.; Luo, Q.; Wang, H. Cardiac involvement in primary Sjögren’s syndrome. Rheumatol. Int. 2022, 42, 179–189. [Google Scholar] [CrossRef] [PubMed]
  274. Yan, S.; Li, M.; Wang, H.; Yang, X.; Zhao, J.; Wang, Q.; Liu, Y.; Lai, J.; Tian, Z.; Song, H.; et al. Characteristics and risk factors of pulmonary arterial hypertension in patients with primary Sjögren’s syndrome. Int. J. Rheum. Dis. 2018, 21, 1068–1075. [Google Scholar] [CrossRef]
  275. Akaycan, J.; Hidayet, Ş.; Bayramoğlu, A.; Yolbaş, S.; Karaca, Y.; Yiğit, Y.; Ulutaş, Z. Subclinical left ventricular dysfunction in Sjögren’s syndrome assessed by four-dimensional speckle tracking echocardiography. Echocardiography 2020, 37, 1803–1808. [Google Scholar] [CrossRef]
  276. Guerreso, K.; Conner, E.A. Possible role of anti-SSA/Ro antibodies in the pathogenesis of pulmonary hypertension. Respir. Med. Case Rep. 2016, 17, 47–49. [Google Scholar] [CrossRef] [Green Version]
  277. Yokoe, I.; Kobayashi, H.; Nishiwaki, A.; Nagasawa, Y.; Kitamura, N.; Haraoka, M.; Kobayashi, Y.; Takei, M.; Nakamura, H. Asymptomatic myocardial dysfunction was revealed by feature tracking cardiac magnetic resonance imaging in patients with primary Sjögren’s syndrome. Int. J. Rheum. Dis. 2021, 24, 1482–1490. [Google Scholar] [CrossRef]
  278. Fujimoto, K.; Norikane, T.; Yamamoto, Y.; Takami, Y.; Murota, M.; Shimada, H.; Dobashi, H.; Nishiyama, Y. Cardiac Sarcoidosis Mimicking Lymphoma in a Patient With Sjogren’s Syndrome. Korean Circ. J. 2022, 52, 715–716. [Google Scholar] [CrossRef]
  279. Caforio, A.L.P.; Adler, Y.; Agostini, C.; Allanore, Y.; Anastasakis, A.; Arad, M.; Böhm, M.; Charron, P.; Elliott, P.M.; Eriksson, U.; et al. Diagnosis and management of myocardial involvement in systemic immune-mediated diseases: A position statement of the European Society of Cardiology Working Group on Myocardial and Pericardial Disease. Eur. Heart J. 2017, 38, 2649–2662. [Google Scholar] [CrossRef] [Green Version]
  280. Yong, W.C.; Sanguankeo, A.; Upala, S. Association between primary Sjogren’s syndrome, arterial stiffness, and subclinical atherosclerosis: A systematic review and meta-analysis. Clin. Rheumatol. 2019, 38, 447–455. [Google Scholar] [CrossRef]
  281. Alunno, A.; Carubbi, F.; Mariani, F.M.; Martini, C.; Campanozzi, E.; Ferri, C. The Interplay between Cardiovascular Risk, Cardiovascular Events, and Disease Activity in Primary Sjögren’s Syndrome: Is Uric Acid the Missing Link? Nutrients 2023, 15, 1563. [Google Scholar] [CrossRef] [PubMed]
  282. Bodewes, I.L.A.; van der Spek, P.J.; Leon, L.G.; Wijkhuijs, A.J.M.; van Helden-Meeuwsen, C.G.; Tas, L.; Schreurs, M.W.J.; van Daele, P.L.A.; Katsikis, P.D.; Versnel, M.A. Fatigue in Sjögren’s Syndrome: A Search for Biomarkers and Treatment Targets. Front. Immunol. 2019, 10, 312. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  283. Bucourt, E.; Martaillé, V.; Goupille, P.; Joncker-Vannier, I.; Huttenberger, B.; Réveillère, C.; Mulleman, D.; Courtois, A.R. A Comparative Study of Fibromyalgia, Rheumatoid Arthritis, Spondyloarthritis, and Sjögren’s Syndrome; Impact of the Disease on Quality of Life, Psychological Adjustment, and Use of Coping Strategies. Pain Med. 2021, 22, 372–381. [Google Scholar] [CrossRef] [PubMed]
  284. Loganathan, M.; Ladani, A.; Lippmann, S. Fibromyalgia, Sjogren’s & depression: Linked? Postgrad Med. 2020, 132, 575–580. [Google Scholar] [CrossRef] [PubMed]
  285. Davies, K.; Mirza, K.; Tarn, J.; Howard-Tripp, N.; Bowman, S.J.; Lendrem, D.; Ng, W.F. Fatigue in primary Sjögren’s syndrome (pSS) is associated with lower levels of proinflammatory cytokines: A validation study. Rheumatol. Int. 2019, 39, 1867–1873. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  286. Duret, P.M.; Meyer, N.; Saraux, A.; Devauchelle-Pensec, V.; Seror, R.; Le-Guern, V.; Larroche, C.; Perdriger, A.; Sibilia, J.; Guardiolle, V.; et al. Seasonal effect on fatigue, pain and dryness in primary Sjögren’s syndrome. Arthritis Res. 2020, 22, 39. [Google Scholar] [CrossRef] [Green Version]
  287. Miyamoto, S.T.; Valim, V.; Carletti, L.; Ng, W.F.; Perez, A.J.; Lendrem, D.W.; Trennel, M.; Giovelli, R.A.; Dias, L.H.; Serrano, É.V.; et al. Supervised walking improves cardiorespiratory fitness, exercise tolerance, and fatigue in women with primary Sjögren’s syndrome: A randomized-controlled trial. Rheumatol. Int. 2019, 39, 227–238. [Google Scholar] [CrossRef]
  288. Wang, X.; Zhang, T.; Guo, Z.; Pu, J.; Riaz, F.; Feng, R.; Fang, X.; Song, J.; Liang, Y.; Wu, Z.; et al. The Efficiency of Hydroxychloroquine for the Treatment of Primary Sjögren’s Syndrome: A Systematic Review and Meta-Analysis. Front. Pharm. 2021, 12, 693796. [Google Scholar] [CrossRef]
  289. Collins, A.; Lendrem, D.; Wason, J.; Tarn, J.; Howard-Tripp, N.; Bodewes, I.; Versnel, M.A.; Gottenberg, J.E.; Seror, R.; Mariette, X.; et al. Revisiting the JOQUER trial: Stratification of primary Sjögren’s syndrome and the clinical and interferon response to hydroxychloroquine. Rheumatol. Int. 2021, 41, 1593–1600. [Google Scholar] [CrossRef]
  290. Wen, J.; Zhu, F.; Yu, X.; Xie, H.; Li, C. Low-dose interleukin-2 can improve salivary secretion but not lymphocyte infiltration of salivary glands in a murine model of Sjögren’s syndrome. BMC Immunol. 2022, 23, 49. [Google Scholar] [CrossRef]
  291. Liu, J.; Gao, H.; Li, C.; Zhu, F.; Wang, M.; Xu, Y.; Wu, B. Expression and regulatory characteristics of peripheral blood immune cells in primary Sjögren’s syndrome patients using single-cell transcriptomic. iScience 2022, 25, 105509. [Google Scholar] [CrossRef] [PubMed]
  292. Luo, J.; Liao, X.; Zhang, L.; Xu, X.; Ying, S.; Yu, M.; Zhu, L.; Lin, S.; Wang, X. Transcriptome Sequencing Reveals Potential Roles of ICOS in Primary Sjögren’s Syndrome. Front. Cell Dev. Biol. 2020, 8, 592490. [Google Scholar] [CrossRef] [PubMed]
  293. García-Espinoza, J.A.; Muñoz-Valle, J.F.; García-Chagollán, M.; Hernández-Bello, J.; Palafox-Sánchez, C.A.; López-Villalobos, E.F.; Sánchez-Zuno, G.A.; Martínez-Bonilla, G.E.; Cerpa-Cruz, S.; Carrillo-Ballesteros, F.J.; et al. ICOS Gene Polymorphisms (IVS1 + 173 T/C and c. 1624 C/T) in Primary Sjögren’s Syndrome Patients: Analysis of ICOS Expression. Curr. Issues Mol. Biol. 2022, 44, 764–776. [Google Scholar] [CrossRef] [PubMed]
  294. Wu, D.; Chen, L.; Wang, D.; Wang, Y.; Yao, G.; Sun, L. IFIH1 was predicted as a key biomarker in primary Sjögren’s syndrome based on transcriptome analysis and experimental verification in patients and mice. Int. J. Rheum. Dis. 2023, 26, 895–906. [Google Scholar] [CrossRef] [PubMed]
  295. Sembler-Møller, M.L.; Belstrøm, D.; Locht, H.; Pedersen, A.M.L. Combined serum anti-SSA/Ro and salivary TRIM29 reveals promising high diagnostic accuracy in patients with primary Sjögren’s syndrome. PLoS ONE 2021, 16, e0258428. [Google Scholar] [CrossRef]
  296. Jung, J.Y.; Kim, J.W.; Kim, H.A.; Suh, C.H. Salivary Biomarkers in Patients with Sjögren’s Syndrome-A Systematic Review. Int. J. Mol. Sci. 2021, 22, 12903. [Google Scholar] [CrossRef]
  297. Peck, A.B.; Nguyen, C.Q.; Ambrus, J. Early Covert Appearance of Marginal Zone B Cells in Salivary Glands of Sjögren’s Syndrome-Susceptible Mice: Initiators of Subsequent Overt Clinical Disease. Int. J. Mol. Sci. 2021, 22, 1919. [Google Scholar] [CrossRef]
  298. Peck, A.B.; Nguyen, C.Q.; Ambrus, J.L. Upregulated Chemokine and Rho-GTPase Genes Define Immune Cell Emigration into Salivary Glands of Sjögren’s Syndrome-Susceptible C57BL/6.NOD-Aec1Aec2 Mice. Int. J. Mol. Sci. 2021, 22, 7176. [Google Scholar] [CrossRef]
  299. Peck, A.B.; Nguyen, C.Q.; Ambrus, J.L., Jr. A MZB Cell Activation Profile Present in the Lacrimal Glands of Sjögren’s Syndrome-Susceptible C57BL/6.NOD-Aec1Aec2 Mice Defined by Global RNA Transcriptomic Analyses. Int. J. Mol. Sci. 2022, 23, 6106. [Google Scholar] [CrossRef]
  300. Blokland, S.L.M.; van Vliet-Moret, F.M.; Hillen, M.R.; Pandit, A.; Goldschmeding, R.; Kruize, A.A.; Bouma, G.; van Maurik, A.; Olek, S.; Hoffmueller, U.; et al. Epigenetically quantified immune cells in salivary glands of Sjögren’s syndrome patients: A novel tool that detects robust correlations of T follicular helper cells with immunopathology. Rheumatology 2020, 59, 335–343. [Google Scholar] [CrossRef]
  301. Lopes, A.P.; Hillen, M.R.; Hinrichs, A.C.; Blokland, S.L.; Bekker, C.P.; Pandit, A.; Kruize, A.A.; Radstake, T.R.; van Roon, J.A. Deciphering the role of cDC2s in Sjögren’s syndrome: Transcriptomic profile links altered antigen processes with IFN signature and autoimmunity. Ann. Rheum. Dis. 2023, 82, 374–383. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Extraglandular manifestations in pSS. Abbreviation: RP, Raynaud phenomenon; CVL, cutaneous vasculitis lesions; CNS, central nervous system; PNS, peripheral nervous system; ANS, autonomic nervous system; RTA, renal tubular acidosis; GN, glomerulonephritis; ILD, interstitial lung disease; MR- MCNT, inhibit muscarinic receptor-mediated cholinergic neurotransmission; TIA, transient ischemic attacks; PH, pulmonary hypertension; CAVB, congenital atrioventricular block. ↓, low; ↑, high.
Figure 1. Extraglandular manifestations in pSS. Abbreviation: RP, Raynaud phenomenon; CVL, cutaneous vasculitis lesions; CNS, central nervous system; PNS, peripheral nervous system; ANS, autonomic nervous system; RTA, renal tubular acidosis; GN, glomerulonephritis; ILD, interstitial lung disease; MR- MCNT, inhibit muscarinic receptor-mediated cholinergic neurotransmission; TIA, transient ischemic attacks; PH, pulmonary hypertension; CAVB, congenital atrioventricular block. ↓, low; ↑, high.
Jpm 13 00961 g001
Figure 2. Cutaneous manifestations in pSS patients. (a) Palpable purpura; (b) Non-palpable purpura; (c) Cutaneous ulcers in vasculitis; (d) Erythema nodosum-like lesions; (e) Raynaud syndrome sign; (f) Chronic chilblains.
Figure 2. Cutaneous manifestations in pSS patients. (a) Palpable purpura; (b) Non-palpable purpura; (c) Cutaneous ulcers in vasculitis; (d) Erythema nodosum-like lesions; (e) Raynaud syndrome sign; (f) Chronic chilblains.
Jpm 13 00961 g002
Table 1. Clinical trials that investigated the effects of biological therapies in pSS.
Table 1. Clinical trials that investigated the effects of biological therapies in pSS.
TherapyReferences
Year
Study CharacteristicsOutcome
Type/PhaseCohort Size (n =)Follow-Up Period (w)
RituximabCarubbi et al., 2013 [91]RCT41120↓ ESSDAI activity, ↓ glandular infiltrate, ↓ ectopic GC
Devauchelle- Pensec et al., 2014 [94]RCT12024Fatigue alleviation
Cornec et al., 2016 [47]Open label4524Mild glandular B-cell depletion
Bowman et al., 2017 [95]RCT/III13326No clinical efficacy
Fisher et al., 2018 [96]RCT/III5248Improvement in ultrasound score
BelimumabMariette et al., 2013 [97]Open label/II3028Reduction in parotid swelling, Mild ↓ of B cell activation
de Vita et al., 2015 [98]Open label/II3052Improvement in ESSDAI score
Quartuccio et al., 2016 [99]Open label/II1352No significant changes in type II IFN scores
Belimumab/RituximabMariette et al., 2022 [100]RCT/II8668B cell depletion in salivary gland
AbataceptAdler et al., 2013 [101]Open label1124↓ glandular inflammation, ↑ saliva production
Haacke et al., 2017 [102]Pilot RCT1524Inhibition of local formation of memory B-cells
Verstappen et al., 2017 [103]Open label1548↓ cTfh-cells and expression levels of the activation marker ICOS on T-cells
Baer et al., 2020 [104]RCT/III18724No significant clinical efficacy
de Wolff et al., 2022 [105]RCT/III4048Improvement in ESSDAI activity and eyes dryness.
TocilizumabFelten et al., 2020 [106]RCT11044No significant clinical efficacy
Ianalumab (VAY736)Bowman et al., 2022 [107]RCT/IIb19024↓ ESSDAI activity
Diekhoff et al., 2022 [108]RCT2724Improvement in salivary gland ultrasound score
LD-IL-2He et al., 2022 [109]RCT/II6012Restore the balance of T and B cell subsets
Abbreviations: RCT, randomized control trial; ↓, low; ↑, high; ESSDAI, EULAR Sjögren’s syndrome disease activity index; GC, germinal centers; IFN, interferon; cTfh cells, circulating T follicular helper cells; ICOS, inducible costimulator; LD-IL-2, low-dose-interleukin-2.
Table 5. The prevalence of the main pulmonary manifestations.
Table 5. The prevalence of the main pulmonary manifestations.
Pulmonary Manifestations% pSS Patients
Airway disease
  Cough—Xerotrachea41–61% [206]
  Bronchiectasis7–54% [206,210]
Interstitial lung disease6–70% [207,210]
  Non- specific interstitial pneumonia (NSIP)29–42% * [208,209]
  Usual interstitial pneumonia (UIP)11–43% * [203,209]
  Lymphocytic interstitial pneumonia (LIP)4–15% * [111,208]
  Organizing pneumonia (OP)9.5% ** [208]
Others
  MALT lymphoma1–2% [204]
* from patients with airway disease. ** from interstitial lung disease.
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

Mihai, A.; Caruntu, C.; Jurcut, C.; Blajut, F.C.; Casian, M.; Opris-Belinski, D.; Ionescu, R.; Caruntu, A. The Spectrum of Extraglandular Manifestations in Primary Sjögren’s Syndrome. J. Pers. Med. 2023, 13, 961. https://doi.org/10.3390/jpm13060961

AMA Style

Mihai A, Caruntu C, Jurcut C, Blajut FC, Casian M, Opris-Belinski D, Ionescu R, Caruntu A. The Spectrum of Extraglandular Manifestations in Primary Sjögren’s Syndrome. Journal of Personalized Medicine. 2023; 13(6):961. https://doi.org/10.3390/jpm13060961

Chicago/Turabian Style

Mihai, Ancuta, Constantin Caruntu, Ciprian Jurcut, Florin Cristian Blajut, Mihnea Casian, Daniela Opris-Belinski, Ruxandra Ionescu, and Ana Caruntu. 2023. "The Spectrum of Extraglandular Manifestations in Primary Sjögren’s Syndrome" Journal of Personalized Medicine 13, no. 6: 961. https://doi.org/10.3390/jpm13060961

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