Next Article in Journal
Peripheral Inflammation Featuring Eosinophilia or Neutrophilia Is Associated with the Survival and Infiltration of Eosinophils within the Tumor among Various Histological Subgroups of Patients with NSCLC
Previous Article in Journal
A Possible Link between Cell Plasticity and Renin Expression in the Collecting Duct: A Narrative Review
Previous Article in Special Issue
AAV-Mediated Expression of miR-17 Enhances Neurite and Axon Regeneration In Vitro
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Identifying microRNAs Possibly Implicated in Myalgic Encephalomyelitis/Chronic Fatigue Syndrome and Fibromyalgia: A Review

by
Maria Tsamou
*,
Fabiënne A. C. Kremers
,
Keano A. Samaritakis
and
Erwin L. Roggen
ToxGenSolutions (TGS), 6229 EV Maastricht, The Netherlands
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(17), 9551; https://doi.org/10.3390/ijms25179551
Submission received: 9 July 2024 / Revised: 28 August 2024 / Accepted: 28 August 2024 / Published: 3 September 2024
(This article belongs to the Special Issue Non-coding RNA in Physiology and Pathophysiology)

Abstract

:
Myalgic encephalomyelitis/chronic fatigue syndrome (ME/CFS) and fibromyalgia (FM) are chronic syndromes of unknown etiology, accompanied by numerous symptoms affecting neurological and physical conditions. Despite frequent revisions of the diagnostic criteria, clinical practice guidelines are often outdated, leading to underdiagnosis and ineffective treatment. Our aim was to identify microRNA (miRNA) biomarkers implicated in pathological mechanisms underlying these diseases. A comprehensive literature review using publicly accessible databases was conducted. Interesting miRNAs were extracted from relevant publications on ME/CFS and/or FM, and were then linked to pathophysiological processes possibly manifesting these chronic diseases. Dysregulated miRNAs in ME/CFS and FM may serve as promising biomarkers for these diseases. Key identified miRNAs, such as miR-29c, miR-99b, miR-128, miR-374b, and miR-766, were frequently mentioned for their roles in immune response, mitochondrial dysfunction, oxidative stress, and central sensitization, while miR-23a, miR-103, miR-152, and miR-320 were implicated in multiple crucial pathological processes for FM and/or ME/CFS. In summary, both ME/CFS and FM seem to share many dysregulated biological or molecular processes, which may contribute to their commonly shared symptoms. This miRNA-based approach offers new angles for discovering molecular markers urgently needed for early diagnosis or therapeutics to tackle the pathology of these medically unexplained chronic diseases.

1. Introduction

Among chronic diseases, fibromyalgia (FM) and myalgic encephalomyelitis/chronic fatigue syndrome (ME/CFS) have been estimated to affect around 1.0–2.7% of the global population [1,2]. The prevalence of both FM and ME/CFS is higher in women (4.0% and 1.7%, respectively) than in men (2.4% and 0.9%, respectively) [1,3,4], especially in middle-aged individuals between 30 and 50 years old [5]. Due to the lack of medical education programs and the outdated guidance for practicing clinicians, these diseases are not accurately diagnosed [6,7,8], adversely affecting not only the quality of life of the patients, but also the associated costs incurred by healthcare systems.
There is a substantial clinical overlap between ME/CFS and FM. Both ME/CFS and FM are chronic syndromes of unknown etiology, characterized by neurological and physical conditions, which commonly share the symptoms of persistent debilitating fatigue and chronic pain that are medically unexplained for more than 3 months in the case of FM and more than 6 months for ME/CFS [9]. Multisystem symptoms have been described for these chronic diseases, including extreme fatigue, myalgia (muscle and joint pain), post-exertional malaise (PEM), brain fog, cognitive dysfunction, disturbed sleep patterns, anxiety, depression, mood disorders, and gastrointestinal disturbances [10,11]. Based on the Institute of Medicine (IOM) diagnostic criteria for ME/CFS (2015), three symptoms (extreme fatigue, PEM, and disturbed sleep) and at least one of two additional manifestations (cognitive impairment and orthostatic intolerance) are prerequisites for diagnosis [12]. As for FM diagnosis, according to the American College of Rheumatology (ACR) diagnostic criteria (2016), bilateral pain on the left and right sides of the body both above and below the waist and chronic generalized pain (on the cervical spine, anterior chest, thoracic spine, or low back) in at least 11 of the 18 tender points should be detected [13,14]. Several risk factors have been implicated in the onset of these chronic diseases [15]. These include genetic and family predisposition [16], environmental factors related to toxins [17], toxic metals [18], (agricultural) chemicals [19,20,21], and viral and bacterial infections [17,22,23,24,25]. Additionally, autoimmunity [26,27], stress conditions [28], and nutritional deficiencies [29,30] may also play significant roles.
A large body of evidence on the potential pathological mechanisms underlying either ME/CFS or FM supports the notion that the multiple system dysregulation of the central nervous system (CNS), immune system, metabolism, and ion transport may play a role in these chronic-fatigue-related diseases [10]. Importantly, these diseases are overrepresented by women, exhibiting an apparent gender/sex-specific pattern in disease susceptibility. Although many studies have invested substantial efforts into better understanding the pathophysiological basis of these chronic diseases and to improve their treatment, their etiology and even their diagnosis remain undetermined. There is an urgent need for biomarkers that can efficiently detect these diseases at an early stage and contribute to effective personalized treatments to alleviate pain and improve patients’ clinical conditions.
MiRNAs are short (18–21 nucleotides) non-coding RNA molecules known as master regulators of gene expression that control more than 60% of the human genome, binding to a specific sequence at the 3-untranslated region (3-UTR) of their target genes and inducing translational repression [31]. MiRNAs regulate a wide range of cellular and biological processes [32]. Several miRNAs have been suggested as promising biomarkers for many pathological conditions. Evidence has reported aberrant miRNA expression in chronic diseases [33,34]. Identifying candidate miRNA biomarkers implicated in the underlying mechanisms of these chronic fatigue syndromes is of great importance.
This review aims to strengthen the current understanding of the complex etiology and challenging diagnosis of FM and ME/CFS, with a particular focus on emerging miRNA biomarkers. We linked the identified miRNAs with dysregulated biological and cellular processes that have been implicated in the pathophysiology of FM and ME/CFS. This may help to develop more targeted and effective therapeutic strategies. Ultimately, by using this mechanistic-based approach, we aim to bridge the gaps in current research and suggest innovative directions for future studies. This, in turn, could improve both the diagnosis and management of FM and ME/CFS and improve the quality of life of affected individuals.

2. Methods

Literature Review for miRNAs in ME/CFS and FM

For the selection of miRNAs relevant to the chronic conditions of interest, we performed a literature search using public bibliographic databases (PubMed). Our research strategy included the keywords “miRNA”, “chronic fatigue syndrome (CFS)”, “myalgic encephalomyelitis (ME)”, and “fibromyalgia (FM)”. We applied specific search filters to refine the results, limiting the selection to studies that were open access, published after 2010, available in English, and focused on human studies. Additionally, “off-topic” and retracted articles were excluded from consideration. The final set of selected articles (n = 25) had publication dates prior to 10 January 2024. A visual representation of our literature review process is provided in Figure 1.

3. Discussion

3.1. Dysregulated Biological and Cellular Processes Shared by ME/CFS and FM

Numerous studies have shown that abnormal changes in the immune system [35], chronic inflammatory pathways and oxidative stress pathways [36,37], mitochondrial dysfunction [38], reduced vascular function [39,40], tryptophan metabolism [41], transient receptor potential (TRP) ion channel disorder [42], and central sensitization [43] may be involved in both ME/CFS and FM. Additionally, alterations in the autonomic and neuroendocrine system involving the Hypothalamic–Pituitary–Adrenal (HPA) axis, the activation of cell signaling networks, and disturbances in metabolic pathways have also been implicated [44,45]. These factors contribute to the common symptoms of these conditions, such as chronic (widespread) pain, chronic fatigue, memory deficits, and depression [46,47].
In more detail, the dysregulation of the immune system is considered to contribute to the etiology of chronic fatigue diseases, inducing changes in cytokine profiles, immunoglobin levels, and the function of T and B cells and natural killer (NK) cells [30,48,49,50,51]. In particular, in ME/CFS, several immunological abnormalities including the impaired cytotoxic function of NK cells and CD8+ T cells, the presence of autoantibodies mainly targeting the central nervous system (CNS), and increased levels of proinflammatory cytokines have been described [52]. T cells have also been implicated in pain development and central sensitization, suggesting that numerous T cell subpopulations, including CD3+ T cells, CD4+ T helper cells, and CD8+ cytotoxic T cells, influence the pathophysiology of FM [48]. Increased levels of systemic and neuroinflammatory markers have been seen in both FM and ME/CFS [35,53]. Theoharides et al. [54] suggested that the role of mast cells as an “immune gate to the brain”, which can be stimulated by neuropeptides such as substance P, known to be increased in the cerebrospinal fluid (CSF) of FM patients [55], or sex hormones such as estrogens, possibly accounts for women’s overrepresentation in FM and CFS/ME [56,57,58].
In both ME/CFS and FM, mitochondrial dysfunction and reduced ATP levels in muscle and neural cells have been attributed to the widespread musculoskeletal pain seen in FM and sometimes in ME/CFS [38]. An imbalance between oxidants and antioxidants has been linked to these chronic diseases [36,37], supporting the role of oxidative stress in the development and progression of the diseases. Morris et al. [59] suggested that increased Nuclear Factor kappa B (NF-kB), causing the release of pro-inflammatory cytokines, and decreased tumor suppressor protein p53, causing aerobic mitochondrial dysfunction, are the key mechanisms in ME/CFS associated with the increased production of reactive oxygen species (ROS), mitochondrial exhaustion, and the additional need for ATP production.
A deficiency in the CoQ10 enzyme, which has been reported in both ME/CFS and FM [60,61], can lead to mitochondrial dysfunction because of the decreased activity of mitochondrial respiration and mitochondrial membrane potential, resulting in ROS production and, subsequently, in the selective degradation of mitochondria mediated by autophagy, known as mitophagy [62]. Autophagy is an important cellular process that involves the degradation of damaged cytosolic components by lysosomes and ensures balance in energy sources [48]. Changes in autophagic genes or proteins have been reported in FM [61] and ME/CFS [63].
Growing evidence supports the implication of reduced vascular function in both FM [39] and ME/CFS [40]. Endothelial dysfunction, which refers to the abnormal function of endothelial cells likely leading to decreased vasodilation along with reduced nitric oxide (NO) (endothelium-relaxing mediator) availability, has been associated with the severity of symptoms of these chronic diseases [64,65]. The decreased post-occlusive hyperemic response of branchial artery and forearm skin microcirculation were observed in ME/CFS patients compared to healthy controls [66]. The reduced brachial artery flow-mediated dilatation in FM patients in response to increased blood flow or NO has been attributed to the unavailability of NO to smooth muscle cells, to muscle cell dysfunction, or to the inability of the arteries to relax [65].
Also, TRP ion channel disorder has been linked to chronic fatigue. The TRP ion channels, activated by G-protein-coupled receptors (GPCRs), are non-selective cation channels that possess high permeability for calcium (Ca2+) implicated in store-operated calcium entry (SOCE) in the white matter of the CNS, and their dysfunction can lead to reduced intracellular Ca2+ mobilization [67,68]. TRP ion channels are susceptible to stressors such as viruses and have a key role in the regulation of Ca2+ signaling, which is important for cell functions, intracellular signaling pathways, and cellular homeostasis [42].
Chronic widespread pain has been described as a consequence of central sensitization, which is characterized by the increased and abnormal responsiveness of nociceptive neurons to a variety of stimuli, leading to disturbed pain processing in the CNS [43]. The activation of ionotropic N-methyl-D-Aspartate (NMDA), a-amino-3-hydroxy-5-methyl-4-isoxazole propionate (AMPA) receptors, metabotropic glutamate receptor subtypes (mGluR), brain-derived neurotrophic factor (BDNF), substance P and calcitonin gene-related peptide (CGRP), NO, and bradykinin have been implicated in initiating and maintaining central sensitization activity and pain hypersensitivity [69]. Also, the activation of Toll-like receptors (TLRs), such as increased TLR-4 signaling in glia and sensory neurons, has been implicated in the sustained proinflammatory condition within spinal cord and neuronal central sensitization, affecting the nociceptive pathways and thereby resulting in persistent pain [70,71].
Central HPA dysfunction has been implicated in stress-related disorders including ME/CFS and FM [44,45]. In ME/CFS and FM, several potential mechanisms have been described as underlying hypercortisolism or hypocortisolism, characterized by high or low levels of circulating cortisol (corticosteroids), respectively, which involve the release of neurohormones, corticotropin-releasing hormone (CRH), and arginine vasopressin (AVP) from the hypothalamus, stimulating their receptors and leading to the release of adrenocorticotropic hormone (ACTH) from the pituitary into systemic circulation, affecting the release of glucocorticoids from the adrenal cortex [45,72,73]. Cortisol acts via a negative feedback mechanism, causing a decrease in the secretion of CRH, AVP, and ACTH [74]. During chronic stress, the sustained activation of the HPA axis has been described [45]. Also, under chronic inflammatory conditions, a shift from CRH to AVP dominance in HPA activity has been described [75].
Tryptophan metabolism mediates the interactions among the blood, brain, and immune system [41]. Tryptophan breaks down via the methoxyindole pathway (1%), leading to the production of serotonin (5-hydroxytryptamine, 5-HT), or via the kynurenine (KYN) pathway (>95%), leading to the de novo synthesis of nicotinamide adenine dinucleotide (NAD+). In the KYN pathway, tryptophan 2,3-dioxygenase (TDO) is stimulated by cortisol, while indoleamine 2,3-dioxygenases (IDOs) are upregulated by lipopolysaccharides and proinflammatory cytokines and downregulated by antioxidants and anti-inflammatory cytokines [76]. Among the KYN metabolites, quinolinic acid (QA) is the most neurotoxic; it decreases NAD levels and acts as an agonist of NMDAR, causing neuronal and astrocytic damage by increasing the intracellular Ca2+. Kynurenic acid (KYNA) is neuroprotective; it is produced by astrocytes and acts as an antagonist of NMDAR [41]. Tryptophan dysmetabolism via the KYN pathway and its metabolites, mainly IDOs, 5-HT, KYNA, and QA, may contribute to fatigue, gastrointestinal disorders, depression, neuroinflammation, cognitive dysfunction, brain fog, sleep disorders, and immune system abnormalities [41], all of which have been described as shared symptoms by ME/CFS and FM.
Lastly, metabolic syndrome and obesity have been associated with symptom severity in chronic fatigue disorders [77]. Women with FM were found to be 5.5 times more likely than healthy women to acquire metabolic syndrome and have a larger waist circumference [78]. In FM, a high prevalence (17.3%) of deficiency in growth hormone (GH)/insulin-like growth factor 1 (IGF1) signaling in obesity has also been described [79,80,81]. In women with FM, higher levels of total and low-density cholesterol have been reported, as well as higher levels of glycosylated hemoglobin and triglycerides in serum and higher levels of norepinephrine (NE)/epinephrine and NE/cortisol ratios in urine, leading to higher systolic and diastolic blood pressure [78]. Finally, an altered fatty acid and lipid metabolism was revealed by metabolic profiling in the plasma of ME/CFS patients, indicating disturbances in taurine and glycerophospholipid metabolism among other fatty-acid-metabolism-related pathways, as well as in purines (ADP and ATP), pyrimidines, amino acids, glucose, and oxaloacetate metabolic pathways [82].

3.2. miRNAs Potentially Implicated in FM and/or ME/CFS

A compelling body of evidence has supported the implication of altered expression of miRNAs in ME/CFS and/or FM syndromes [83,84,85,86]. The identified miRNAs from relevant studies in the literature (Table S1 in Supplementary Materials) were further linked to the suggested dysregulated processes in these chronic diseases, accordingly. Figure 2 illustrates the most frequently cited miRNAs identified in at least two independent studies. Additionally, for each pathophysiological process, we highlight the miRNAs that were reported.
Considering the miRNAs that have been reported to be involved in immunological and inflammatory disturbances in FM and/or ME/CFS (n = 26), miR-17-5p [87,88,89], miR-21-5p [87,89,90], miR-127-3p [83,86,91,92,93], miR-146a-5p [87,88,89,94], miR-150-5p [83,86,88,93], and miR-320b [95,96,97,98] were found to have the greatest number of mentions (>3) in different publications. Concerning the most frequently mentioned miRNAs (>3 publications) regulating mitochondrial dysfunction and oxidative stress (n = 21) in these chronic conditions, miR-29a-3p [83,93], miR-34a-5p [90,99], miR-99b [87,100], miR-103a-3p [89,95], miR-146a [87,88,89,94], miR-183-5p [88,91], miR-328-3p [91,101], miR-374b-5p [83,93,95], and miR-486 [83,93] were indicated.
Among the autophagy-related miRNAs (n = 8), miR-103 [87,89,95] was most frequently reported as being dysregulated. Regarding the miRNAs implicated in vascular dysfunction (n = 26) in these conditions, miR-21-5p [85,87,89,90], miR-143-3p [86,91,92], and miR-320b [95,96,97,98] were the most frequently (>2) mentioned.
In addition, with mentions in more than three publications, the most studied miRNAs involved in the central sensitization process (n = 29) in these diseases include miR-17-5p [87,88,89], miR-21-5p [85,87,89,90], miR-127-3p [83,86,91,92,93], miR-140-5p [83,86,88,93], miR-142 [86,92,95], miR-143 [86,91,92], miR-150-5p [83,86,93], and miR-320b [95,96,97,98]. As for the dysregulated miRNAs in metabolic disorders (n = 37) in ME/CFS and/or FM, miR-127-3p [83,86,92,93], miR-374b-5p [33,83,93,95], miR-107 [95,102], miR-143 [86,91,92], miR-320a [95,96,97,98], and miR-374b-5p [83,93,95] were identified in at least studies. Only miR-183-5p [91] was found to be involved in TRP ion channel disorder and miR-320(a) [95,96,97,98] in the dysfunctional HPA axis in these pain conditions, while no miRNAs were identified as being involved in tryptophan dysmetabolism in ME/CFS and/or FM.
The overlap of miRNAs among the pathophysiological processes in ME/CFS and/or FM (Figure 3) revealed that miR-152 was found to regulate most of the pathological processes manifesting these chronic diseases, including immunological and inflammatory disturbances, mitochondrial dysfunction, oxidative stress, impaired autophagy, and vascular dysfunction. In more detail, the significantly decreased expression of miR-152 was reported in the NK cells of ME/CFS patients compared to non-fatigued controls [89]. In that study, it was suggested that miR-152 may be involved in the reduced cytotoxic activity often reported in patients with ME/CFS, via targeting inhibitory human leukocyte antigen G (HLA-G). In another study, it was demonstrated that miR-152(-3p) regulated chronic-pain-induced depression-like behaviors by targeting DNA methyltransferase 1 (DNMT1) [103].
In addition, miR-23a was shown to be involved in immunological and inflammatory disturbances, mitochondrial dysfunction/oxidative stress, metabolic disorders, and vascular dysfunction. The downregulation of miR-23a has been reported in FM patients [98,104] compared to controls. This miRNA is a member of the miR-23a-27a-24-2 cluster, which is implicated in many diseases, including muscle atrophy via targeting muscle-specific F-box protein (MAFx)/atrogin-1 [105]. Also, miR-23b (from the same cluster) has been shown to regulate the expression of μ-opioid receptors [106] and to be underexpressed in various autoimmune diseases via suppressing interleukin 17 (IL-17), tumor necrosis factor α (TNF- α), or IL-1β-induced NF-kB activation [107].
Moreover, miR-320 was implicated in the regulation of the HPA axis, mitochondrial function, oxidative stress, inflammation, and metabolic disorders. In patients with FM, miR-320a has been reported to be upregulated [95,108], while miR-320b was downregulated [98] compared to healthy controls. It has been shown that miR-320 targets genes involved in pain disorders, particularly by binding to FK506 binding protein 5 (FKBP5) [109], which plays an important role in the HPA axis and subsequently, in pain processing and the stress response. Additionally, miR-320a has been shown to regulate neuroinflammation by inhibiting its targets, transforming growth factor-beta receptor 2 (TGFBR2) and SMAD Family Member 2 (SMAD2), resulting in increased inflammatory cytokines, mitochondrial dysfunction, apoptosis, and oxidative stress [110,111]. Finally, miR-320(a) has been shown to target IGF-1 receptor and/or IGF-1 [111,112].
MiR-103(a) was involved in the immune inflammatory response, mitochondrial dysfunction and oxidative stress, impaired autophagy, central sensitization, and metabolic disorders. In FM patients, miR-103a-3p was underexpressed compared to healthy controls and was associated with pain and sleep quantity [104]. In patients with ME/CFS, the expression of miR-103 was lower in NK cells in comparison to the non-fatigued controls [89]. MiR-103 has been suggested as a therapeutic target, as its downregulation led to hypersensitivity to pain via the regulation of L-type calcium channel (in an animal model) [113].
Five miRNAs, namely, miR-29c-5p, miR-99b, miR-128-3p, miR-374b-5p, and miR-766, were found to be commonly implicated in metabolic disorders and mitochondrial dysfunction/oxidative stress. MiR-29c is a member of the miR-29 family, which has a key role in metabolism and metabolic disease [114], among numerous other human diseases [115]. MiR-29c targets the muscle atrophy genes muscle RING-finger protein-1 (MuRF1), Atrogin-1, and histone deacetylase 4 (HDAC4) and has been associated with skeletal muscle size and function [116]. Overexpressed miR-99b in blood and NK cells in patients with ME/CFS was previously reported [100], suggesting an mTOR signaling pathway and, subsequently, a reduction in NK cell cytotoxic and cytokine effector function [117,118]. The upregulation of miR-128-3p has been reported in patients with FM [91]. MiR-128-3p suppresses bone formation via sirtuin 6 (SIRT6) [119] and targets the cholinergic system in FM [91]. It has also been implicated in the neuronal oxidative stress response [120]. MiR-374b-5p was found to be underexpressed in both FM [104] and ME/CFS patients [83] and is implicated in induced pain perception. MiR-374b-5p targets vascular endothelial growth factor A (VEGFA), involved in vasculogenesis and angiogenesis, which regulate the capillary supply in skeletal muscle and thereby may be linked to the post-exertional malaise and fatigue seen in patients with ME/CFS and FM [121]. MiR-766-3p was underexpressed in blood samples collected from patients with FM [91]. In addition, miR-766 regulated multiple pathways including apoptosis, extracellular matrix, and inflammatory reactions in human intervertebral disc degeneration [122].
Figure 3. Overlapped miRNAs among the pathophysiological processes possibly manifesting fibromyalgia (FM) and/or myalgic encephalomyelitis/chronic fatigue syndrome (ME/CFS). Abbreviations: Metab Dis: metabolic disorder (n = 37 miRNAs), Centr Sens: central sensitization (n = 29 miRNAs), Immun & Inflam: immunological and inflammatory disturbances (n = 26 miRNAs), Vasc Dys: vascular dysfunction (n = 26 miRNAs), Mit & OS: mitochondrial dysfunction and oxidative stress (n = 21 miRNAs), Autophagy (n = 8 miRNAs). The InteractiVenn tool was used to generate the diagram [123].
Figure 3. Overlapped miRNAs among the pathophysiological processes possibly manifesting fibromyalgia (FM) and/or myalgic encephalomyelitis/chronic fatigue syndrome (ME/CFS). Abbreviations: Metab Dis: metabolic disorder (n = 37 miRNAs), Centr Sens: central sensitization (n = 29 miRNAs), Immun & Inflam: immunological and inflammatory disturbances (n = 26 miRNAs), Vasc Dys: vascular dysfunction (n = 26 miRNAs), Mit & OS: mitochondrial dysfunction and oxidative stress (n = 21 miRNAs), Autophagy (n = 8 miRNAs). The InteractiVenn tool was used to generate the diagram [123].
Ijms 25 09551 g003

3.3. Sex- and Age-specific Patterns

There is around a three and six times higher prevalence of ME/CFS and FM in females, respectively [124]. Female sex is the most reproducible risk factor for both ME/CFS and FM [125,126], and the rates seem to increase with age [125,127,128]. The perception of pain has been shown to be different between biological men and women, with the latter exhibiting remarkably lower pain thresholds, which could be possibly and partially attributed to their different hormonal profiles [129,130]. Presumably, the differences in gender and/or sex that affect hormones and immune responses can also influence susceptibility in chronic pain conditions [131]. Moreover, hormonal fluctuations during the menstrual cycle and reproductive phase have also been reported to be involved in chronic pain disorders [132]. Among gonadal steroid hormones, estrogens have been shown to regulate nociceptive pathways and, subsequently, pain via targeting intracellular receptors such as GPCRs of the CNS and peripheral nervous system, and modulating serotonergic, noradrenergic, dopaminergic, and endogenous kappa (k) or μ-opioid pathways [129,133,134,135,136]. Estrogens modulate the synthesis and metabolism of serotonin, the dysregulation of which has been associated with pain disorders [137]. In contrast, androgens, which seem to inhibit the expression of estrogen receptors, act as a nociceptive protective agent against stresses [138]. Low levels of adrenal androgens, particularly serum concentrations of neurosteroids and dehydroepiandrosterone sulphate (DHEAS), have been observed in both ME/CFS and FM [139]. Women in a high-estrogen state showed decreased pain sensitivity, increased brain mu-opioid receptor binding, and higher levels of activated endogenous opioid neurotransmission during pain, whereas women in a low-estradiol state exhibited significantly decreased endogenous opioid neurotransmission associated with hyperalgesia [135].
Besides sex and/or gender, age is also a contributing factor for acquiring ME/CFS [140] and FM [141]. Two peaks in the incidence of ME/CFS, between 10 and 19 years and between 30 and 39 years, have been described [139]. In another study, older age was associated with the risk of severe fatigue syndrome, but not with the risk of developing ME/CFS [142]. Higher fatigue scores were observed in older patients with ME/CFS compared to younger ones [143]. The impact of age on the development of FM has been possibly attributed to the observed decrease in gray matter, which, in turn, has been associated with pain sensitivity [144]. It has been demonstrated that patients with FM experienced an age-related reduction in gray matter at a rate 3.3 times faster than that of healthy individuals, with each year of the FM equating to 9.5 times the normal aging loss in terms of gray matter [145]. In addition, increasing age was associated with an increase in symptom duration, but a decrease in FM symptoms [141]. In particular, decreased pain and depressive symptoms have been reported in older patients with chronic pain conditions compared to younger patients [146].

4. Conclusions

Chronic-fatigue-related syndromes, such as ME/CFS and FM, share an overlapping clinical picture, described by common symptoms. Although several risk factors have been described, the pathophysiology underlying these syndromes is not well defined, impeding their accurate diagnosis and effective treatment. Many dysregulated processes have been implicated in both diseases. A better understanding of the mechanisms involved in these diseases may lead to the identification of biomarkers, such as some of the miRNAs discussed in this review. Within this context, the identification of miRNAs implicated in these chronic diseases can be valuable for the timely diagnosis of FM and ME/CFS. Filling this gap can certainly benefit not only patients’ quality of life, but also future drug discovery for these chronic conditions.

5. Future Directions

In this review, several miRNAs, commonly dysregulated in pathophysiological processes, were identified using existing human data from the literature. These miRNAs appear to play potential roles in various biological processes such as immune response, inflammatory disturbances, mitochondrial dysfunction, oxidative stress, autophagy, vascular dysfunction, central sensitization, and metabolic disorders. This miRNA-based approach provides a clear direction for future research and potential clinical applications. These findings may also provide therapeutic targets. In most of our reviewed studies, a higher percentage of participants were women, as is true of most studies of ME/CFS and FM. Another significant strength of our study is its focus on miRNAs as molecular biomarkers, which are known to be extremely stable due to their short length, particularly in biofluids collected by minimally invasive methods. These miRNAs may serve as sensitive indicators of pathological processes, contributing to early diagnosis and targeted therapeutic interventions.
Nevertheless, this study has also its limitations. Our research relied on published studies, which may introduce publication bias and may overestimate the importance of certain miRNAs. In addition, these studies are often cross-sectional, making it difficult to draw conclusions about the causality of the relationship between miRNA dysregulations and the development of or progression of these conditions. In addition, confounding and lifestyle factors may also influence the interpretation of miRNA dysregulations. Lastly, while the study identifies potential miRNA biomarkers, further clinical validation is necessary to confirm their efficacy and reliability in diagnosing ME/CFS and FM.

Supplementary Materials

Author Contributions

Conceptualization, M.T., F.A.C.K. and E.L.R.; methodology, M.T. and F.A.C.K.; writing—original draft preparation, M.T.; writing—review and editing, M.T., F.A.C.K., K.A.S. and E.L.R.; visualization, M.T. and F.A.C.K.; supervision, E.L.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by ToxGenSolutions BV and 3Rs Management and Consulting ApS (https://3rsmc-aps.com (accessed on 1 August 2024)).

Conflicts of Interest

Authors were employed by the company ToxGenSolutions. The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Avellaneda Fernández, A.; Pérez Martín, Á.; Izquierdo Martínez, M.; Arruti Bustillo, M.; Barbado Hernández, F.J.; de la Cruz Labrado, J.; Díaz-Delgado Peñas, R.; Gutiérrez Rivas, E.; Palacín Delgado, C.; Rivera Redondo, J.; et al. Chronic fatigue syndrome: Aetiology, diagnosis and treatment. BMC Psychiatry 2009, 9, S1. [Google Scholar] [CrossRef] [PubMed]
  2. Queiroz, L.P. Worldwide Epidemiology of Fibromyalgia. Curr. Pain Headache Rep. 2013, 17, 356. [Google Scholar] [CrossRef]
  3. Zambolin, F.; Duro-Ocana, P.; Faisal, A.; Bagley, L.; Gregory, W.J.; Jones, A.W.; McPhee, J.S. Fibromyalgia and Chronic Fatigue Syndromes: A systematic review and meta-analysis of cardiorespiratory fitness and neuromuscular function compared with healthy individuals. PLoS ONE 2022, 17, e0276009. [Google Scholar] [CrossRef] [PubMed]
  4. Collado-Mateo, D.; Olivares, P.R.; Adsuar, J.C.; Gusi, N. Impact of fibromyalgia on sexual function in women. J. Back Musculoskelet. Rehabil. 2020, 33, 355–361. [Google Scholar] [CrossRef]
  5. Saral, I.; Sindel, D.; Esmaeilzadeh, S.; Sertel-Berk, H.O.; Oral, A. The effects of long- and short-term interdisciplinary treatment approaches in women with fibromyalgia: A randomized controlled trial. Rheumatol. Int. 2016, 36, 1379–1389. [Google Scholar] [CrossRef]
  6. Clayton, E.W. Beyond Myalgic Encephalomyelitis/Chronic Fatigue Syndrome: An IOM Report on Redefining an Illness. JAMA 2015, 313, 1101–1102. [Google Scholar] [CrossRef] [PubMed]
  7. Friedberg, F.; Sunnquist, M.; Nacul, L. Rethinking the Standard of Care for Myalgic Encephalomyelitis/Chronic Fatigue Syndrome. J. Gen. Intern. Med. 2020, 35, 906–909. [Google Scholar] [CrossRef]
  8. Brimmer, D.J.; Fridinger, F.; Lin, J.-M.S.; Reeves, W.C. U.S. healthcare providers’ knowledge, attitudes, beliefs, and perceptions concerning Chronic Fatigue Syndrome. BMC Fam. Pract. 2010, 11, 28. [Google Scholar] [CrossRef]
  9. Van Houdenhove, B.; Luyten, P. Customizing Treatment of Chronic Fatigue Syndrome and Fibromyalgia: The Role of Perpetuating Factors. Psychosomatics 2008, 49, 470–477. [Google Scholar] [CrossRef] [PubMed]
  10. Carruthers, B.M.; van de Sande, M.I.; De Meirleir, K.L.; Klimas, N.G.; Broderick, G.; Mitchell, T.; Staines, D.; Powles, A.C.P.; Speight, N.; Vallings, R.; et al. Myalgic encephalomyelitis: International Consensus Criteria. J. Intern. Med. 2011, 270, 327–338. [Google Scholar] [CrossRef]
  11. Arnold, L.M.; Clauw, D.J.; McCarberg, B.H. Improving the recognition and diagnosis of fibromyalgia. Mayo Clin. Proc. 2011, 86, 457–464. [Google Scholar] [CrossRef] [PubMed]
  12. Jason, L.A.; McManimen, S.; Sunnquist, M.; Brown, A.; Newton, J.L.; Strand, E.B. Examining the Institute of Medicine’s Recommendations Regarding Chronic Fatigue Syndrome: Clinical Versus Research Criteria. J. Neurol. Psychol. 2015, 2015 (Suppl. S2). Available online: http://www.avensonline.org/wp-content/uploads/JNP-2332-3469-S2-0002.pdf (accessed on 1 August 2024).
  13. Wolfe, F.; Clauw, D.J.; Fitzcharles, M.-A.; Goldenberg, D.L.; Häuser, W.; Katz, R.L.; Mease, P.J.; Russell, A.S.; Russell, I.J.; Walitt, B. 2016 Revisions to the 2010/2011 fibromyalgia diagnostic criteria. Semin. Arthritis Rheum. 2016, 46, 319–329. [Google Scholar] [CrossRef] [PubMed]
  14. Berwick, R.B.C.; Goebel, A. The diagnosis of fibromyalgia syndrome. Clin. Med. 2022, 22, 570–574. [Google Scholar]
  15. Bjørklund, G.; Dadar, M.; Pivina, L.; Doşa, M.D.; Semenova, Y.; Maes, M. Environmental, Neuro-immune, and Neuro-oxidative Stress Interactions in Chronic Fatigue Syndrome. Mol. Neurobiol. 2020, 57, 4598–4607. [Google Scholar] [CrossRef]
  16. Arnold, L.M.; Hudson, J.I.; Hess, E.V.; Ware, A.E.; Fritz, D.A.; Auchenbach, M.B.; Starck, L.O.; Keck, P.E., Jr. Family study of fibromyalgia. Arthritis Rheum. 2004, 50, 944–952. [Google Scholar] [CrossRef]
  17. Racciatti, D.; Vecchiet, J.; Ceccomancini, A.; Ricci, F.; Pizzigallo, E. Chronic fatigue syndrome following a toxic exposure. Sci. Total Environ. 2001, 270, 27–31. [Google Scholar] [CrossRef]
  18. Sterzl, I.; Procházková, J.; Hrdá, P.; Bártová, J.; Matucha, P.; Stejskal, V.D. Mercury and nickel allergy: Risk factors in fatigue and autoimmunity. Neuro Endocrinol. Lett. 1999, 20, 221–228. [Google Scholar]
  19. Tahmaz, N.; Soutar, A.; Cherrie, J.W. Chronic Fatigue and Organophosphate Pesticides in Sheep Farming: A Retrospective Study Amongst People Reporting to a UK Pharmacovigilance Scheme. Ann. Occup. Hyg. 2003, 47, 261–267. [Google Scholar] [CrossRef]
  20. Dunstan, R.H.; Donohoe, M.; Taylor, W.; Roberts, T.K.; Murdoch, R.N.; Watkins, J.A.; McGregor, N.R. A preliminary investigation of chlorinated hydrocarbons and chronic fatigue syndrome. Med. J. Aust. 1995, 163, 294–297. [Google Scholar] [CrossRef]
  21. Godfrey, R. Fibromyalgia as a manifestation of petroleum fume toxicity in a family of four. J. Clin. Rheumatol. Pract. Rep. Rheum. Musculoskelet. Dis. 1997, 3, 54–57. [Google Scholar] [CrossRef] [PubMed]
  22. Rasa-Dzelzkaleja, S.; Krumina, A.; Capenko, S.; Nora-Krukle, Z.; Gravelsina, S.; Vilmane, A.; Ievina, L.; Shoenfeld, Y.; Murovska, M. The persistent viral infections in the development and severity of myalgic encephalomyelitis/chronic fatigue syndrome. J. Transl. Med. 2023, 21, 33. [Google Scholar] [CrossRef] [PubMed]
  23. Shikova, E.; Reshkova, V.; Kumanova, A.; Raleva, S.; Alexandrova, D.; Capo, N.; Murovska, M. Cytomegalovirus, Epstein-Barr virus, and human herpesvirus-6 infections in patients with myalgic encephalomyelitis/chronic fatigue syndrome. J. Med. Virol. 2020, 92, 3682–3688. [Google Scholar] [CrossRef] [PubMed]
  24. Hunskar, G.S.; Rortveit, G.; Litleskare, S.; Eide, G.E.; Hanevik, K.; Langeland, N.; Wensaas, K.-A. Prevalence of fibromyalgia 10 years after infection with Giardia lamblia: A controlled prospective cohort study. Scand. J. Pain 2022, 22, 348–355. [Google Scholar] [CrossRef] [PubMed]
  25. Buskila, D.; Atzeni, F.; Sarzi-Puttini, P. Etiology of fibromyalgia: The possible role of infection and vaccination. Autoimmun. Rev. 2008, 8, 41–43. [Google Scholar] [CrossRef]
  26. Brenu, E.W.; van Driel, M.L.; Staines, D.R.; Ashton, K.J.; Ramos, S.B.; Keane, J.; Klimas, N.G.; Marshall-Gradisnik, S.M. Immunological abnormalities as potential biomarkers in Chronic Fatigue Syndrome/Myalgic Encephalomyelitis. J. Transl. Med. 2011, 9, 81. [Google Scholar] [CrossRef]
  27. Staud, R. Cytokine and immune system abnormalities in fibromyalgia and other central sensitivity syndromes. Curr. Rheumatol. Rev. 2015, 11, 109–115. [Google Scholar] [CrossRef]
  28. Van Houdenhove, B.; Luyten, P.; Kempke, S. Chronic fatigue syndrome/fibromyalgia: A “stress-adaptation” model. Fatigue Biomed. Health Behav. 2013, 1, 137–147. [Google Scholar] [CrossRef]
  29. Bjørklund, G.; Dadar, M.; Chirumbolo, S.; Aaseth, J. Fibromyalgia and nutrition: Therapeutic possibilities? Biomed. Pharmacother. 2018, 103, 531–538. [Google Scholar] [CrossRef]
  30. Bjørklund, G.; Dadar, M.; Pen, J.J.; Chirumbolo, S.; Aaseth, J. Chronic fatigue syndrome (CFS): Suggestions for a nutritional treatment in the therapeutic approach. Biomed. Pharmacother. 2019, 109, 1000–1007. [Google Scholar] [CrossRef]
  31. Zaporozhchenko, I.A.; Rykova, E.Y.; Laktionov, P.P. The Fundamentals of miRNA Biology: Structure, Biogenesis, and Regulatory Functions. Russ. J. Bioorganic Chem. 2020, 46, 1–13. [Google Scholar] [CrossRef]
  32. Saliminejad, K.; Khorram Khorshid, H.R.; Soleymani Fard, S.; Ghaffari, S.H. An overview of microRNAs: Biology, functions, therapeutics, and analysis methods. J. Cell. Physiol. 2019, 234, 5451–5465. [Google Scholar] [CrossRef]
  33. Kaczmarek, M.P. Heterogenous circulating miRNA changes in ME/CFS converge on a unified cluster of target genes: A computational analysis. PLoS ONE 2023, 18, e0296060. [Google Scholar] [CrossRef] [PubMed]
  34. Polli, A.; Godderis, L.; Ghosh, M.; Ickmans, K.; Nijs, J. Epigenetic and miRNA Expression Changes in People with Pain: A Systematic Review. J. Pain 2020, 21, 763–780. [Google Scholar] [CrossRef]
  35. Groven, N.; Fors, E.A.; Reitan, S.K. Patients with Fibromyalgia and Chronic Fatigue Syndrome show increased hsCRP compared to healthy controls. Brain Behav. Immun. 2019, 81, 172–177. [Google Scholar] [CrossRef]
  36. Atamer, Y.; Sarac, S.; Asık, H.K.; Sahbaz, T. Serum paraoxonase activities, nitric oxide, and malondialdehyde levels are altered in patients with primary fibromyalgia syndrome. Ir. J. Med. Sci. 2023, 192, 2541–2547. [Google Scholar] [CrossRef] [PubMed]
  37. Lee, J.S.; Kim, H.G.; Lee, D.S.; Son, C.G. Oxidative Stress is a Convincing Contributor to Idiopathic Chronic Fatigue. Sci. Rep. 2018, 8, 12890. [Google Scholar] [CrossRef]
  38. Meeus, M.; Nijs, J.; Hermans, L.; Goubert, D.; Calders, P. The role of mitochondrial dysfunctions due to oxidative and nitrosative stress in the chronic pain or chronic fatigue syndromes and fibromyalgia patients: Peripheral and central mechanisms as therapeutic targets? Expert Opin. Ther. Targets 2013, 17, 1081–1089. [Google Scholar] [CrossRef] [PubMed]
  39. Lee, J.H.; Cho, K.I.; Kim, S.M.; Lee, H.G.; Kim, T.I. Arterial stiffness in female patients with fibromyalgia and its relationship to chronic emotional and physical stress. Korean Circ. J. 2011, 41, 596–602. [Google Scholar] [CrossRef]
  40. Sandvik, M.K.; Sørland, K.; Leirgul, E.; Rekeland, I.G.; Stavland, C.S.; Mella, O.; Fluge, Ø. Endothelial dysfunction in ME/CFS patients. PLoS ONE 2023, 18, e0280942. [Google Scholar] [CrossRef]
  41. Kavyani, B.; Lidbury, B.A.; Schloeffel, R.; Fisher, P.R.; Missailidis, D.; Annesley, S.J.; Dehhaghi, M.; Heng, B.; Guillemin, G.J. Could the kynurenine pathway be the key missing piece of Myalgic Encephalomyelitis/Chronic Fatigue Syndrome (ME/CFS) complex puzzle? Cell. Mol. Life Sci. 2022, 79, 412. [Google Scholar] [CrossRef] [PubMed]
  42. Nguyen, T.; Staines, D.; Nilius, B.; Smith, P.; Marshall-Gradisnik, S. Novel identification and characterisation of Transient receptor potential melastatin 3 ion channels on Natural Killer cells and B lymphocytes: Effects on cell signalling in Chronic fatigue syndrome/Myalgic encephalomyelitis patients. Biol. Res. 2016, 49, 27. [Google Scholar] [CrossRef] [PubMed]
  43. Cagnie, B.; Coppieters, I.; Denecker, S.; Six, J.; Danneels, L.; Meeus, M. Central sensitization in fibromyalgia? A systematic review on structural and functional brain MRI. Semin. Arthritis Rheum. 2014, 44, 68–75. [Google Scholar] [CrossRef] [PubMed]
  44. Holtorf, K. Diagnosis and Treatment of Hypothalamic-Pituitary-Adrenal (HPA) Axis Dysfunction in Patients with Chronic Fatigue Syndrome (CFS) and Fibromyalgia (FM). J. Chronic Fatigue Syndr. 2007, 14, 59–88. [Google Scholar] [CrossRef]
  45. Crofford, L.J. The hypothalamic-pituitary-adrenal stress axis in fibromyalgia and chronic fatigue syndrome. Z. Für Rheumatol. 1998, 57, S67–S71. [Google Scholar] [CrossRef]
  46. Romano, G.F.; Tomassi, S.; Russell, A.; Mondelli, V.; Pariante, C.M. Fibromyalgia and chronic fatigue: The underlying biology and related theoretical issues. Adv. Psychosom. Med. 2015, 34, 61–77. [Google Scholar] [CrossRef]
  47. Maes, M. Inflammatory and Oxidative and Nitrosative Stress. Cascades as New Drug Targets in Myalgic Encephalomyelitis and Chronic Fatigue Syndrome. Mod. Trends Pharmacopsychiatry 2013, 28, 162–174. [Google Scholar] [CrossRef]
  48. Banfi, G.; Diani, M.; Pigatto, P.D.; Reali, E. T Cell Subpopulations in the Physiopathology of Fibromyalgia: Evidence and Perspectives. Int. J. Mol. Sci. 2020, 21, 1186. [Google Scholar] [CrossRef]
  49. Groven, N.; Fors, E.A.; Stunes, A.K.; Reitan, S.K. MCP-1 is increased in patients with CFS and FM, whilst several other immune markers are significantly lower than healthy controls. Brain Behav. Immun.-Health 2020, 4, 100067. [Google Scholar] [CrossRef]
  50. Brenu, E.W.; Hardcastle, S.L.; Atkinson, G.M.; van Driel, M.L.; Kreijkamp-Kaspers, S.; Ashton, K.J.; Staines, D.R.; Marshall-Gradisnik, S.M. Natural killer cells in patients with severe chronic fatigue syndrome. Autoimmun. Highlights 2013, 4, 69–80. [Google Scholar] [CrossRef]
  51. Mensah, F.K.F.; Bansal, A.S.; Ford, B.; Cambridge, G. Chronic fatigue syndrome and the immune system: Where are we now? Neurophysiol. Clin. 2017, 47, 131–138. [Google Scholar] [CrossRef]
  52. Komaroff, A.L. Inflammation correlates with symptoms in chronic fatigue syndrome. Proc. Natl. Acad. Sci. USA 2017, 114, 8914–8916. [Google Scholar] [CrossRef] [PubMed]
  53. Tate, W.; Walker, M.; Sweetman, E.; Helliwell, A.; Peppercorn, K.; Edgar, C.; Blair, A.; Chatterjee, A. Molecular Mechanisms of Neuroinflammation in ME/CFS and Long COVID to Sustain Disease and Promote Relapses. Front. Neurol. 2022, 13, 877772. [Google Scholar] [CrossRef] [PubMed]
  54. Theoharides, T.C.; Tsilioni, I.; Bawazeer, M. Mast Cells, Neuroinflammation and Pain in Fibromyalgia Syndrome. Front. Cell. Neurosci. 2019, 13, 353. [Google Scholar] [CrossRef] [PubMed]
  55. Russell, I.J.; Vaeroy, H.; Javors, M.; Nyberg, F. Cerebrospinal fluid biogenic amine metabolites in fibromyalgia/fibrositis syndrome and rheumatoid arthritis. Arthritis Rheum. 1992, 35, 550–556. [Google Scholar] [CrossRef] [PubMed]
  56. Schertzinger, M.; Wesson-Sides, K.; Parkitny, L.; Younger, J. Daily Fluctuations of Progesterone and Testosterone Are Associated with Fibromyalgia Pain Severity. J. Pain 2018, 19, 410–417. [Google Scholar] [CrossRef]
  57. de Kruijf, M.; Stolk, L.; Zillikens, M.C.; de Rijke, Y.B.; Bierma-Zeinstra, S.M.A.; Hofman, A.; Huygen, F.J.P.M.; Uitterlinden, A.G.; van Meurs, J.B.J. Lower sex hormone levels are associated with more chronic musculoskeletal pain in community-dwelling elderly women. Pain 2016, 157, 1425–1431. [Google Scholar] [CrossRef]
  58. Boneva, R.S.; Maloney, E.M.; Lin, J.-M.; Jones, J.F.; Wieser, F.; Nater, U.M.; Heim, C.M.; Reeves, W.C. Gynecological History in Chronic Fatigue Syndrome: A Population-Based Case-Control Study. J. Women’s Health 2010, 20, 21–28. [Google Scholar] [CrossRef]
  59. Morris, G.; Maes, M. Increased nuclear factor-κB and loss of p53 are key mechanisms in Myalgic Encephalomyelitis/chronic fatigue syndrome (ME/CFS). Med. Hypotheses 2012, 79, 607–613. [Google Scholar] [CrossRef]
  60. McArdle, A.; McArdle, F.; Jackson, M.J.; Page, S.F.; Fahal, I.; Edwards, R.H.T. Investigation by Polymerase Chain Reaction of Enteroviral Infection in Patients with Chronic Fatigue Syndrome. Clin. Sci. 1996, 90, 295–300. [Google Scholar] [CrossRef]
  61. Cordero, M.D.; De Miguel, M.; Moreno Fernández, A.M.; Carmona López, I.M.; Garrido Maraver, J.; Cotán, D.; Gómez Izquierdo, L.; Bonal, P.; Campa, F.; Bullon, P.; et al. Mitochondrial dysfunction and mitophagy activation in blood mononuclear cells of fibromyalgia patients: Implications in the pathogenesis of the disease. Arthritis Res. Ther. 2010, 12, R17. [Google Scholar] [CrossRef]
  62. Oezel, L.; Then, H.; Jung, A.L.; Jabari, S.; Bonaterra, G.A.; Wissniowski, T.T.; Önel, S.F.; Ocker, M.; Thieme, K.; Kinscherf, R.; et al. Fibromyalgia syndrome: Metabolic and autophagic processes in intermittent cold stress mice. Pharmacol. Res. Perspect. 2016, 4, e00248. [Google Scholar] [CrossRef] [PubMed]
  63. Gottschalk, G.; Peterson, D.; Knox, K.; Maynard, M.; Whelan, R.J.; Roy, A. Elevated ATG13 in serum of patients with ME/CFS stimulates oxidative stress response in microglial cells via activation of receptor for advanced glycation end products (RAGE). Mol. Cell. Neurosci. 2022, 120, 103731. [Google Scholar] [CrossRef] [PubMed]
  64. Scherbakov, N.; Szklarski, M.; Hartwig, J.; Sotzny, F.; Lorenz, S.; Meyer, A.; Grabowski, P.; Doehner, W.; Scheibenbogen, C. Peripheral endothelial dysfunction in myalgic encephalomyelitis/chronic fatigue syndrome. ESC Heart Fail. 2020, 7, 1064–1071. [Google Scholar] [CrossRef] [PubMed]
  65. Cho, K.I.; Lee, J.H.; Kim, S.M.; Lee, H.G.; Kim, T.I. Assessment of endothelial function in patients with fibromyalgia—Cardiac ultrasound study. Clin. Rheumatol. 2011, 30, 647–654. [Google Scholar] [CrossRef] [PubMed]
  66. Newton, D.J.; Kennedy, G.; Chan, K.K.F.; Lang, C.C.; Belch, J.J.F.; Khan, F. Large and small artery endothelial dysfunction in chronic fatigue syndrome. Int. J. Cardiol. 2012, 154, 335–336. [Google Scholar] [CrossRef]
  67. Cabanas, H.; Muraki, K.; Eaton, N.; Balinas, C.; Staines, D.; Marshall-Gradisnik, S. Loss of Transient Receptor Potential Melastatin 3 ion channel function in natural killer cells from Chronic Fatigue Syndrome/Myalgic Encephalomyelitis patients. Mol. Med. 2018, 24, 44. [Google Scholar] [CrossRef]
  68. Veldhuis, N.A.; Poole, D.P.; Grace, M.; McIntyre, P.; Bunnett, N.W. The G Protein–Coupled Receptor–Transient Receptor Potential Channel Axis: Molecular Insights for Targeting Disorders of Sensation and Inflammation. Pharmacol. Rev. 2015, 67, 36–73. [Google Scholar] [CrossRef]
  69. Latremoliere, A.; Woolf, C.J. Central sensitization: A generator of pain hypersensitivity by central neural plasticity. J. Pain 2009, 10, 895–926. [Google Scholar] [CrossRef]
  70. Nicotra, L.; Loram, L.C.; Watkins, L.R.; Hutchinson, M.R. Toll-like receptors in chronic pain. Exp. Neurol. 2012, 234, 316–329. [Google Scholar] [CrossRef]
  71. Lacagnina, M.J.; Watkins, L.R.; Grace, P.M. Toll-like receptors and their role in persistent pain. Pharmacol. Ther. 2018, 184, 145–158. [Google Scholar] [CrossRef] [PubMed]
  72. Van Den Eede, F.; Moorkens, G.; Van Houdenhove, B.; Cosyns, P.; Claes, S.J. Hypothalamic-Pituitary-Adrenal Axis Function in Chronic Fatigue Syndrome. Neuropsychobiology 2007, 55, 112–120. [Google Scholar] [CrossRef] [PubMed]
  73. Riva, R.; Mork, P.J.; Westgaard, R.H.; Rø, M.; Lundberg, U. Fibromyalgia Syndrome is Associated with Hypocortisolism. Int. J. Behav. Med. 2010, 17, 223–233. [Google Scholar] [CrossRef] [PubMed]
  74. Tomas, C.; Newton, J.; Watson, S. A Review of Hypothalamic-Pituitary-Adrenal Axis Function in Chronic Fatigue Syndrome. ISRN Neurosci. 2013, 2013, 784520. [Google Scholar] [CrossRef] [PubMed]
  75. Altemus, M.; Dale, J.K.; Michelson, D.; Demitrack, M.A.; Gold, P.W.; Straus, S.E. Abnormalities in response to vasopressin infusion in chronic fatigue syndrome. Psychoneuroendocrinology 2001, 26, 175–188. [Google Scholar] [CrossRef] [PubMed]
  76. Tanaka, M.; Vécsei, L. Monitoring the kynurenine system: Concentrations, ratios or what else? Adv. Clin. Exp. Med. 2021, 30, 775–778. [Google Scholar] [CrossRef] [PubMed]
  77. Maloney, E.M.; Boneva, R.S.; Lin, J.-M.S.; Reeves, W.C. Chronic fatigue syndrome is associated with metabolic syndrome: Results from a case-control study in Georgia. Metab. Clin. Exp. 2010, 59, 1351–1357. [Google Scholar] [CrossRef] [PubMed]
  78. Loevinger, B.L.; Muller, D.; Alonso, C.; Coe, C.L. Metabolic syndrome in women with chronic pain. Metabolism: Clinical and experimental. Metabolism 2007, 56, 87–93. [Google Scholar] [CrossRef]
  79. Bjersing, J.L.; Dehlin, M.; Erlandsson, M.; Bokarewa, M.I.; Mannerkorpi, K. Changes in pain and insulin-like growth factor 1 in fibromyalgia during exercise: The involvement of cerebrospinal inflammatory factors and neuropeptides. Arthritis Res. Ther. 2012, 14, R162. [Google Scholar] [CrossRef]
  80. Cuatrecasas, G.; Gonzalez, M.J.; Alegre, C.; Sesmilo, G.; Fernandez-Solà, J.; Casanueva, F.F.; Garcia-Fructuoso, F.; Poca-Dias, V.; Izquierdo, J.P.; Puig-Domingo, M. High prevalence of growth hormone deficiency in severe fibromyalgia syndromes. J. Clin. Endocrinol. Metab. 2010, 95, 4331–4337. [Google Scholar] [CrossRef]
  81. Bennett, R.M. Adult growth hormone deficiency in patients with fibromyalgia. Curr. Rheumatol. Rep. 2002, 4, 306–312. [Google Scholar] [CrossRef] [PubMed]
  82. Germain, A.; Ruppert, D.; Levine, S.M.; Hanson, M.R. Metabolic profiling of a myalgic encephalomyelitis/chronic fatigue syndrome discovery cohort reveals disturbances in fatty acid and lipid metabolism. Mol. Biosyst. 2017, 13, 371–379. [Google Scholar] [CrossRef]
  83. Nepotchatykh, E.; Caraus, I.; Elremaly, W.; Leveau, C.; Elbakry, M.; Godbout, C.; Rostami-Afshari, B.; Petre, D.; Khatami, N.; Franco, A.; et al. Circulating microRNA expression signatures accurately discriminate myalgic encephalomyelitis from fibromyalgia and comorbid conditions. Sci. Rep. 2023, 13, 1896. [Google Scholar] [CrossRef] [PubMed]
  84. Almenar-Pérez, E.; Sánchez-Fito, T.; Ovejero, T.; Nathanson, L.; Oltra, E. Impact of Polypharmacy on Candidate Biomarker miRNomes for the Diagnosis of Fibromyalgia and Myalgic Encephalomyelitis/Chronic Fatigue Syndrome: Striking Back on Treatments. Pharmaceutics 2019, 11, 126. [Google Scholar] [CrossRef]
  85. Cerdá-Olmedo, G.; Mena-Durán, A.V.; Monsalve, V.; Oltra, E. Identification of a MicroRNA Signature for the Diagnosis of Fibromyalgia. PLoS ONE 2015, 10, e0121903. [Google Scholar] [CrossRef] [PubMed]
  86. Soffritti, I.; Gravelsina, S.; D’Accolti, M.; Bini, F.; Mazziga, E.; Vilmane, A.; Rasa-Dzelzkaleja, S.; Nora-Krukle, Z.; Krumina, A.; Murovska, M.; et al. Circulating miRNAs Expression in Myalgic Encephalomyelitis/Chronic Fatigue Syndrome. Int. J. Mol. Sci. 2023, 24, 10582. [Google Scholar] [CrossRef]
  87. Yang, X.; Li, F.; Ma, J.; Liu, Y.; Wang, X.; Wang, R.; Zhang, Y.; Zhang, W.; He, Q.; Song, D.; et al. Study on the Relationship between the miRNA-centered ceRNA Regulatory Network and Fatigue. J. Mol. Neurosci. 2021, 71, 1967–1974. [Google Scholar] [CrossRef]
  88. Almenar-Pérez, E.; Sarría, L.; Nathanson, L.; Oltra, E. Assessing diagnostic value of microRNAs from peripheral blood mononuclear cells and extracellular vesicles in Myalgic Encephalomyelitis/Chronic Fatigue Syndrome. Sci. Rep. 2020, 10, 2064. [Google Scholar] [CrossRef]
  89. Brenu, E.W.; Ashton, K.J.; van Driel, M.; Staines, D.R.; Peterson, D.; Atkinson, G.M.; Marshall-Gradisnik, S.M. Cytotoxic lymphocyte microRNAs as prospective biomarkers for Chronic Fatigue Syndrome/Myalgic Encephalomyelitis. J. Affect. Disord. 2012, 141, 261–269. [Google Scholar] [CrossRef]
  90. Blauensteiner, J.; Bertinat, R.; León, L.E.; Riederer, M.; Sepúlveda, N.; Westermeier, F. Altered endothelial dysfunction-related miRs in plasma from ME/CFS patients. Sci. Rep. 2021, 11, 10604. [Google Scholar] [CrossRef]
  91. Erbacher, C.; Vaknine, S.; Moshitzky, G.; Lobentanzer, S.; Eisenberg, L.; Evdokimov, D.; Sommer, C.; Greenberg, D.S.; Soreq, H.; Üçeyler, N. Distinct CholinomiR Blood Cell Signature as a Potential Modulator of the Cholinergic System in Women with Fibromyalgia Syndrome. Cells 2022, 11, 1276. [Google Scholar] [CrossRef] [PubMed]
  92. Brenu, E.W.; Ashton, K.J.; Batovska, J.; Staines, D.R.; Marshall-Gradisnik, S.M. High-Throughput Sequencing of Plasma MicroRNA in Chronic Fatigue Syndrome/Myalgic Encephalomyelitis. PLoS ONE 2014, 9, e102783. [Google Scholar] [CrossRef] [PubMed]
  93. Nepotchatykh, E.; Elremaly, W.; Caraus, I.; Godbout, C.; Leveau, C.; Chalder, L.; Beaudin, C.; Kanamaru, E.; Kosovskaia, R.; Lauzon, S.; et al. Profile of circulating microRNAs in myalgic encephalomyelitis and their relation to symptom severity, and disease pathophysiology. Sci. Rep. 2020, 10, 19620. [Google Scholar] [CrossRef]
  94. Al-Rawaf, H.A.; Alghadir, A.H.; Gabr, S.A. MicroRNAs as Biomarkers of Pain Intensity in Patients with Chronic Fatigue Syndrome. Pain Pract. 2019, 19, 848–860. [Google Scholar] [CrossRef] [PubMed]
  95. Hussein, M.; Fathy, W.; Abdelaleem, E.A.; Nasser, M.; Yehia, A.; Elanwar, R. The Impact of Micro RNA-320a Serum Level on Severity of Symptoms and Cerebral Processing of Pain in Patients with Fibromyalgia. Pain Med. 2022, 23, 2061–2072. [Google Scholar] [CrossRef]
  96. Clos-Garcia, M.; Andrés-Marin, N.; Fernández-Eulate, G.; Abecia, L.; Lavín, J.L.; van Liempd, S.; Cabrerag, D.; Royoa, F.; Valeroh, A.; Errazquin, N.; et al. Gut microbiome and serum metabolome analyses identify molecular biomarkers and altered glutamate metabolism in fibromyalgia. eBioMedicine 2019, 46, 499–511. [Google Scholar] [CrossRef]
  97. Berg, F.; Moser, D.A.; Hagena, V.; Streit, F.; Mosch, B.; Kumsta, R.; Herpertz, S.; Diers, M. MicroRNA-Related Polymorphism and Their Association with Fibromyalgia. Genes 2023, 14, 1312. [Google Scholar] [CrossRef]
  98. Masotti, A.; Baldassarre, A.; Guzzo, M.P.; Iannuccelli, C.; Barbato, C.; Di Franco, M. Circulating microRNA Profiles as Liquid Biopsies for the Characterization and Diagnosis of Fibromyalgia Syndrome. Mol. Neurobiol. 2017, 54, 7129–7136. [Google Scholar] [CrossRef]
  99. Andersen, H.H.; Duroux, M.; Gazerani, P. Serum MicroRNA Signatures in Migraineurs During Attacks and in Pain-Free Periods. Mol. Neurobiol. 2016, 53, 1494–1500. [Google Scholar] [CrossRef]
  100. Petty, R.D.; McCarthy, N.E.; Le Dieu, R.; Kerr, J.R. MicroRNAs hsa-miR-99b, hsa-miR-330, hsa-miR-126 and hsa-miR-30c: Potential Diagnostic Biomarkers in Natural Killer (NK) Cells of Patients with Chronic Fatigue Syndrome (CFS)/Myalgic Encephalomyelitis (ME). PLoS ONE 2016, 11, e0150904. [Google Scholar] [CrossRef]
  101. Baraniuk, J.N.; Shivapurkar, N. Exercise–induced changes in cerebrospinal fluid miRNAs in Gulf War Illness, Chronic Fatigue Syndrome and sedentary control subjects. Sci. Rep. 2017, 7, 15338. [Google Scholar] [CrossRef] [PubMed]
  102. Braun, A.; Evdokimov, D.; Frank, J.; Sommer, C.; Üçeyler, N. MiR103a-3p and miR107 are related to adaptive coping in a cluster of fibromyalgia patients. PLoS ONE 2020, 15, e0239286. [Google Scholar] [CrossRef]
  103. Ding, X.; Lin, Y.; Yan, B.; Jiao, X.; Liu, Q.; Miao, H.; Wu, Y.; Zhou, C. LncRNA XR_351665 Contributes to Chronic Pain-Induced Depression by Upregulating DNMT1 via Sponging miR-152-3p. J. Pain 2023, 24, 449–462. [Google Scholar] [CrossRef] [PubMed]
  104. Bjersing, J.L.; Lundborg, C.; Bokarewa, M.I.; Mannerkorpi, K. Profile of Cerebrospinal microRNAs in Fibromyalgia. PLoS ONE 2013, 8, e78762. [Google Scholar] [CrossRef] [PubMed]
  105. Chhabra, R.; Dubey, R.; Saini, N. Cooperative and individualistic functions of the microRNAs in the miR-23a~27a~24-2 cluster and its implication in human diseases. Mol. Cancer 2010, 9, 232. [Google Scholar] [CrossRef]
  106. Rodriguez, R.E. Morphine and microRNA Activity: Is There a Relation with Addiction? Front. Genet. 2012, 3, 223. [Google Scholar] [CrossRef]
  107. Zhu, S.; Pan, W.; Song, X.; Liu, Y.; Shao, X.; Tang, Y.; Liang, D.; He, D.; Wang, H.; Liu, W.; et al. The microRNA miR-23b suppresses IL-17-associated autoimmune inflammation by targeting TAB2, TAB3 and IKK-α. Nat. Med. 2012, 18, 1077–1086. [Google Scholar] [CrossRef]
  108. Bjersing, J.L.; Bokarewa, M.I.; Mannerkorpi, K. Profile of circulating microRNAs in fibromyalgia and their relation to symptom severity: An exploratory study. Rheumatol. Int. 2015, 35, 635–642. [Google Scholar] [CrossRef]
  109. Linnstaedt, S.D.; Riker, K.D.; Walker, M.G.; Nyland, J.E.; Zimny, E.; Lewandowski, C.; Hendry, P.L.; Damiron, K.; Pearson, C.; Velilla, M.-A.; et al. MicroRNA 320a Predicts Chronic Axial and Widespread Pain Development Following Motor Vehicle Collision in a Stress-Dependent Manner. J. Orthop. Sports Phys. Ther. 2016, 46, 911–919. [Google Scholar] [CrossRef]
  110. Talebi, F.; Ghourbani, S.; Vojgani, M.; Noorbakhsh, F. miR-320 and Inflammation Regulation in Experimental Autoimmune Encephalomyelitis Through Interference with Tumor Growth Factor-β Signaling Pathway. Immunoregulation 2020, 2, 111–120. [Google Scholar] [CrossRef]
  111. Wang, X.; Qian, R.; Zhang, W.; Chen, S.; Jin, H.; Hu, R. MicroRNA-320 expression in myocardial microvascular endothelial cells and its relationship with insulin-like growth factor-1 in type 2 diabetic rats. Clin. Exp. Pharmacol. Physiol. 2009, 36, 181–188. [Google Scholar] [CrossRef] [PubMed]
  112. Li, Y.; Huang, J.; Hu, C.; Zhou, J.; Xu, D.; Hou, Y.; Wu, C.; Zhao, J.; Li, M.; Zeng, X.; et al. MicroRNA-320a: An important regulator in the fibrotic process in interstitial lung disease of systemic sclerosis. Arthritis Res. Ther. 2021, 11, 21. [Google Scholar] [CrossRef] [PubMed]
  113. Favereaux, A.; Thoumine, O.; Bouali-Benazzouz, R.; Roques, V.; Papon, M.A.; Salam, S.A.; Drutel, G.; Léger, C.; Calas, A.; Nagy, F.; et al. Bidirectional integrative regulation of Cav1.2 calcium channel by microRNA miR-103: Role in pain. EMBO J. 2011, 30, 3830–3841. [Google Scholar] [CrossRef] [PubMed]
  114. Dalgaard, L.T.; Sørensen, A.E.; Hardikar, A.A.; Joglekar, M.V. The microRNA-29 family: Role in metabolism and metabolic disease. Am. J. Physiol.-Cell Physiol. 2022, 323, C367–C377. [Google Scholar] [CrossRef]
  115. Horita, M.; Farquharson, C.; Stephen, L.A. The role of miR-29 family in disease. J. Cell Biochem. 2021, 122, 696–715. [Google Scholar] [CrossRef] [PubMed]
  116. Silva, W.J.; Graça, F.; Cruz, A.; Silvestre, J.G.; Labeit, S.; Miyabara, E.H.; Yan, C.Y.I.; Wang, D.Z.; Moriscot, A.S. miR-29c improves skeletal muscle mass and function throughout myocyte proliferation and differentiation and by repressing atrophy-related genes. Acta Physiol. 2019, 226, e13278. [Google Scholar] [CrossRef]
  117. Jin, Y.; Tymen, S.D.; Chen, D.; Fang, Z.J.; Zhao, Y.; Dragas, D.; Dai, Y.; Marucha, P.T.; Zhou, X. MicroRNA-99 Family Targets AKT/mTOR Signaling Pathway in Dermal Wound Healing. PLoS ONE 2013, 8, e64434. [Google Scholar] [CrossRef]
  118. Nandagopal, N.; Ali, A.K.; Komal, A.K.; Lee, S.-H. The Critical Role of IL-15–PI3K–mTOR Pathway in Natural Killer Cell Effector Functions. Front. Immunol. 2014, 5, 187. [Google Scholar] [CrossRef]
  119. Zhao, J.; Liu, S.; Zhang, W.; Ni, L.; Hu, Z.; Sheng, Z.; Yin, B. MiR-128 inhibits the osteogenic differentiation in osteoporosis by down-regulating SIRT6 expression. Biosci. Rep. 2019, 39, BSR20191405. [Google Scholar] [CrossRef] [PubMed]
  120. Bruno, I.G.; Karam, R.; Huang, L.; Bhardwaj, A.; Lou, C.H.; Shum, E.Y.; Song, H.-W.; Corbett, M.A.; Gifford, W.D.; Gecz, J.; et al. Identification of a MicroRNA that Activates Gene Expression by Repressing Nonsense-Mediated RNA Decay. Mol. Cell 2011, 42, 500–510. [Google Scholar] [CrossRef]
  121. Liao, Y.; Tsai, H.; Chou, P.; Wang, S.; Chen, H.; Lin, Y.; Chiang, I.; Chang, T.; Hsu, S.; Chou, M.; et al. CCL3 promotes angiogenesis by dysregulation of miR-374b/ VEGF-A axis in human osteosarcoma cells. Oncotarget 2016, 7, 4310–4325. [Google Scholar] [CrossRef] [PubMed]
  122. Cui, S.; Zhou, Z.; Liu, X.; Richards, R.G.; Alini, M.; Peng, S.; Liu, S.; Zou, X.; Li, Z.; Grad, S. Identification and Characterization of Serum microRNAs as Biomarkers for Human Disc Degeneration: An RNA Sequencing Analysis. Diagnostics 2020, 10, 1063. [Google Scholar] [CrossRef]
  123. Heberle, H.; Meirelles, G.V.; da Silva, F.R.; Telles, G.P.; Minghim, R. InteractiVenn: A web-based tool for the analysis of sets through Venn diagrams. BMC Bioinform. 2015, 16, 169. [Google Scholar] [CrossRef]
  124. Mogil, J.S. Sex differences in pain and pain inhibition: Multiple explanations of a controversial phenomenon. Nat. Rev. Neurosci. 2012, 13, 859–866. [Google Scholar] [CrossRef] [PubMed]
  125. Reyes, M.; Nisenbaum, R.; Hoaglin, D.C.; Unger, E.R.; Emmons, C.; Randall, B.; Stewart, J.A.; Abbey, S.; Jones, J.F.; Gantz, N.; et al. Prevalence and Incidence of Chronic Fatigue Syndrome in Wichita, Kansas. Arch. Intern. Med. 2003, 163, 1530–1536. [Google Scholar] [CrossRef] [PubMed]
  126. Lacerda, E.M.; Geraghty, K.; Kingdon, C.C.; Palla, L.; Nacul, L. A logistic regression analysis of risk factors in ME/CFS pathogenesis. BMC Neurol. 2019, 19, 275. [Google Scholar] [CrossRef]
  127. Wolfe, F.; Ross, K.; Anderson, J.; Russell, I.J.; Hebert, L. The prevalence and characteristics of fibromyalgia in the general population. Arthritis Rheum. 1995, 38, 19–28. [Google Scholar] [CrossRef]
  128. Haddad, H.W.; Mallepalli, N.R.; Scheinuk, J.E.; Bhargava, P.; Cornett, E.M.; Urits, I.; Kaye, A.D. The Role of Nutrient Supplementation in the Management of Chronic Pain in Fibromyalgia: A Narrative Review. Pain Ther. 2021, 10, 827–848. [Google Scholar] [CrossRef]
  129. Bartley, E.J.; Fillingim, R.B. Sex differences in pain: A brief review of clinical and experimental findings. Br. J. Anaesth. 2013, 111, 52–58. [Google Scholar] [CrossRef]
  130. Craft, R.M.; Mogil, J.S.; Aloisi, A.M. Sex differences in pain and analgesia: The role of gonadal hormones. Eur. J. Pain 2004, 8, 397–411. [Google Scholar] [CrossRef]
  131. Osborne, N.R.; Davis, K.D. Chapter Eight–Sex and gender differences in pain. In International Review of Neurobiology; Moro, E., Arabia, G., Tartaglia, M.C., Ferretti, M.T., Eds.; Academic Press: Cambridge, MA, USA, 2022; Volume 164, pp. 277–307. [Google Scholar] [CrossRef]
  132. Martin, V.T. Ovarian hormones and pain response: A review of clinical and basic science studies. Gend. Med. 2009, 6, 168–192. [Google Scholar] [CrossRef] [PubMed]
  133. Manolagas, S.C.; Kousteni, S. Perspective: Nonreproductive Sites of Action of Reproductive Hormones*. Endocrinology 2001, 142, 2200–2204. [Google Scholar] [CrossRef]
  134. Craft, R.M. Modulation of pain by estrogens. Pain 2007, 132, S3–S12. [Google Scholar] [CrossRef] [PubMed]
  135. Smith, Y.R.; Stohler, C.S.; Nichols, T.E.; Bueller, J.A.; Koeppe, R.A.; Zubieta, J.-K. Pronociceptive and Antinociceptive Effects of Estradiol through Endogenous Opioid Neurotransmission in Women. J. Neurosci. 2006, 26, 5777–5785. [Google Scholar] [CrossRef] [PubMed]
  136. Mogil, J.S.; Wilson, S.G.; Chesler, E.J.; Rankin, A.L.; Nemmani, K.V.S.; Lariviere, W.R.; Groce, M.K.; Wallace, M.R.; Kaplan, L.; Staud, R.; et al. The melanocortin-1 receptor gene mediates female-specific mechanisms of analgesia in mice and humans. Proc. Natl. Acad. Sci. USA 2003, 100, 4867–4872. [Google Scholar] [CrossRef]
  137. Paredes, S.; Cantillo, S.; Candido, K.D.; Knezevic, N.N. An Association of Serotonin with Pain Disorders and Its Modulation by Estrogens. Int. J. Mol. Sci. 2019, 20, 5729. [Google Scholar] [CrossRef] [PubMed]
  138. Athnaiel, O.; Cantillo, S.; Paredes, S.; Knezevic, N.N. The Role of Sex Hormones in Pain-Related Conditions. Int. J. Mol. Sci. 2023, 24, 1866. [Google Scholar] [CrossRef]
  139. Dessein, P.H.; Shipton, E.A. Hydrocortisone and chronic fatigue syndrome. Lancet 1999, 353, 1618. [Google Scholar] [CrossRef]
  140. Bakken, I.J.; Tveito, K.; Gunnes, N.; Ghaderi, S.; Stoltenberg, C.; Trogstad, L.; Håberg, S.E.; Magnus, P. Two age peaks in the incidence of chronic fatigue syndrome/myalgic encephalomyelitis: A population-based registry study from Norway 2008–2012. BMC Med. 2014, 12, 167. [Google Scholar] [CrossRef]
  141. Cronan, T.A.; Serber, E.R.; Walen, H.R.; Jaffe, M. The Influence of Age on Fibromyalgia Symptoms. J. Aging Health 2002, 14, 370–384. [Google Scholar] [CrossRef] [PubMed]
  142. Palacios, N.; Molsberry, S.; Fitzgerald, K.C.; Komaroff, A.L. Different risk factors distinguish myalgic encephalomyelitis/chronic fatigue syndrome from severe fatigue. Sci. Rep. 2023, 13, 2469. [Google Scholar] [CrossRef] [PubMed]
  143. Kidd, E.; Brown, A.; McManimen, S.; Jason, L.A.; Newton, J.L.; Strand, E.B. The Relationship between Age and Illness Duration in Chronic Fatigue Syndrome. Diagnostics 2016, 6, 16. [Google Scholar] [CrossRef] [PubMed]
  144. Ceko, M.; Bushnell, M.C.; Fitzcharles, M.-A.; Schweinhardt, P. Fibromyalgia interacts with age to change the brain. NeuroImage Clin. 2013, 3, 249–260. [Google Scholar] [CrossRef]
  145. Kuchinad, A.; Schweinhardt, P.; Seminowicz, D.A.; Wood, P.B.; Chizh, B.A.; Bushnell, M.C. Accelerated brain gray matter loss in fibromyalgia patients: Premature aging of the brain? J. Neurosci. Off. J. Soc. Neurosci. 2007, 27, 4004–4007. [Google Scholar] [CrossRef] [PubMed]
  146. Corran, T.M.; Farrell, M.J.; Helme, R.D.; Gibson, S.J. The Classification of Patients with Chronic Pain: Age as a Contributing Factor. Clin. J. Pain. 1997, 13, 207–214. [Google Scholar] [CrossRef] [PubMed]
  147. Moatar, A.I.; Chis, A.R.; Romanescu, M.; Ciordas, P.-D.; Nitusca, D.; Marian, C.; Oancea, C.; Sirbu, I.-O. Plasma miR-195-5p predicts the severity of Covid-19 in hospitalized patients. Sci. Rep. 2023, 13, 13806. [Google Scholar] [CrossRef]
  148. Kasimir, F.; Toomey, D.; Liu, Z.; Kaiping, A.C.; Ariza, M.E.; Prusty, B.K. Tissue specific signature of HHV-6 infection in ME/CFS. Front. Mol. Biosci. 2022, 9, 1044964. [Google Scholar] [CrossRef]
  149. Lv, Y.; Zhang, T.; Cai, J.; Huang, C.; Zhan, S.; Liu, J. Bioinformatics and systems biology approach to identify the pathogenetic link of Long COVID and Myalgic Encephalomyelitis/Chronic Fatigue Syndrome. Front. Immunol. 2022, 13, 952987. [Google Scholar] [CrossRef]
  150. Cheema, A.K.; Sarria, L.; Bekheit, M.; Collado, F.; Almenar-Pérez, E.; Martín-Martínez, E.; Alegre, J.; Castro-Marrero, J.; Fletcher, M.A.; Klimas, N.G.; et al. Unravelling myalgic encephalomyelitis/chronic fatigue syndrome (ME/CFS): Gender-specific changes in the microRNA expression profiling in ME/CFS. J. Cell. Mol. Med. 2020, 24, 5865–5877. [Google Scholar] [CrossRef]
  151. Takakura, S.; Oka, T.; Sudo, N. Changes in circulating microRNA after recumbent isometric yoga practice by patients with myalgic encephalomyelitis/chronic fatigue syndrome: An explorative pilot study. Biopsychosoc. Med. 2019, 13, 29. [Google Scholar] [CrossRef]
  152. Leinders, M.; Doppler, K.; Klein, T.; Deckart, M.; Rittner, H.; Sommer, C.; Üçeyler, N. Increased cutaneous miR-let-7d expression correlates with small nerve fiber pathology in patients with fibromyalgia syndrome. Pain 2016, 157, 2493–2503. [Google Scholar] [CrossRef]
  153. Frietze, S.; O'Geen, H.; Blahnik, K.R.; Jin, V.X.; Farnham, P.J. ZNF274 Recruits the Histone Methyltransferase SETDB1 to the 3′ Ends of ZNF Genes. PLoS ONE 2010, 5, e15082. [Google Scholar] [CrossRef] [PubMed]
  154. Shlopov, B.V.; Gumanovskaya, M.L.; Hasty, K.A. Autocrine regulation of collagenase 3 (matrix metalloproteinase 13) during osteoarthritis. Arthritis Rheum. 2000, 43, 195–205. [Google Scholar] [CrossRef]
  155. Fallowfield, J.A.; Mizuno, M.; Kendall, T.J.; Constandinou, C.M.; Benyon, R.C.; Duffield, J.S.; Iredale, J.P. Scar-Associated Macrophages Are a Major Source of Hepatic Matrix Metalloproteinase-13 and Facilitate the Resolution of Murine Hepatic Fibrosis. J. Immunol. 2007, 178, 5288–5295. [Google Scholar] [CrossRef] [PubMed]
  156. Heinrich, P.C.; Behrmann, I.; Müller-Newen, G.; Schaper, F.; Graeve, L. Interleukin-6-type cytokine signalling through the gp130/Jak/STAT pathway. Biochem. J. 1998, 334, 297–314. [Google Scholar] [CrossRef]
  157. Deshmane, S.L.; Kremlev, S.; Amini, S.; Sawaya, B.E. Monocyte Chemoattractant Protein-1 (MCP-1): An Overview. J. Interf. Cytokine Res. 2009, 29, 313–326. [Google Scholar] [CrossRef]
  158. Thomson, R.; Finkelstein, A. Human trypanolytic factor APOL1 forms pH-gated cation-selective channels in planar lipid bilayers: Relevance to trypanosome lysis. Proc. Natl. Acad. Sci. USA 2015, 112, 2894–2899. [Google Scholar] [CrossRef] [PubMed]
  159. Atmaramani, R.R.; Black, B.J.; de la Peña, J.B.; Campbell, Z.T.; Pancrazio, J.J. Conserved Expression of Nav1.7 and Nav1.8 Contribute to the Spontaneous and Thermally Evoked Excitability in IL-6 and NGF-Sensitized Adult Dorsal Root Ganglion Neurons In Vitro. Bioengineering 2020, 7, 44. [Google Scholar] [CrossRef]
  160. Kochumon, S.; Al Madhoun, A.; Al-Rashed, F.; Azim, R.; Al-Ozairi, E.; Al-Mulla, F.; Ahmad, R. Adipose tissue gene expression of CXCL10 and CXCL11 modulates inflammatory markers in obesity: Implications for metabolic inflammation and insulin resistance. Ther. Adv. Endocrinol. Metab. 2020, 11, 2042018820930902. [Google Scholar] [CrossRef] [PubMed]
  161. Tiwari, N.; Garbi, N.; Reinheckel, T.; Moldenhauer, G.; Hämmerling, G.J.; Momburg, F. A Transporter Associated with Antigen-Processing Independent Vacuolar Pathway for the MHC Class I-Mediated Presentation of Endogenous Transmembrane Proteins. J. Immunol. 2007, 178, 7932–7942. [Google Scholar] [CrossRef]
  162. Miao, L.; St Clair, D.K. Regulation of superoxide dismutase genes: Implications in disease. Free Radic. Biol. Med. 2009, 47, 344–356. [Google Scholar] [CrossRef]
  163. Hayden, M.S.; Ghosh, S. Shared Principles in NF-κB Signaling. Cell 2008, 132, 344–362. [Google Scholar] [CrossRef] [PubMed]
  164. Akira, S.; Takeda, K. Toll-like receptor signalling. Nat. Rev. Immunol. 2004, 4, 499–511. [Google Scholar] [CrossRef] [PubMed]
  165. Yu, H.; Pardoll, D.; Jove, R. STATs in cancer inflammation and immunity: A leading role for STAT3. Nat. Rev. Cancer 2009, 9, 798–809. [Google Scholar] [CrossRef] [PubMed]
  166. Cronk, J.C.; Derecki, N.C.; Ji, E.; Xu, Y.; Lampano, A.E.; Smirnov, I.; Baker, W.; Norris, G.T.; Marin, I.; Coddington, N.; et al. Methyl-CpG Binding Protein 2 Regulates Microglia and Macrophage Gene Expression in Response to Inflammatory Stimuli. Immunity 2015, 42, 679–691. [Google Scholar] [CrossRef]
  167. Zhang, X.; Zuo, X.; Yang, B.; Li, Z.; Xue, Y.; Zhou, Y.; Huang, J.; Zhao, X.; Zhou, J.; Yan, Y.; et al. MicroRNA Directly Enhances Mitochondrial Translation during Muscle Differentiation. Cell 2014, 158, 607–619. [Google Scholar] [CrossRef]
  168. Ying, S.-W.; Futter, M.; Rosenblum, K.; Webber, M.J.; Hunt, S.P.; Bliss, T.V.P.; Bramham, C.R. Brain-Derived Neurotrophic Factor Induces Long-Term Potentiation in Intact Adult Hippocampus: Requirement for ERK Activation Coupled to CREB and Upregulation of Arc Synthesis. J. Neurosci. 2002, 22, 1532–1540. [Google Scholar] [CrossRef]
  169. Bradley, J. TNF-mediated inflammatory disease. J. Pathol. 2007, 214, 149–160. [Google Scholar] [CrossRef]
  170. Forrester, S.J.; Booz, G.W.; Sigmund, C.D.; Coffman, T.M.; Kawai, T.; Rizzo, V.; Scalia, R.; Eguchi, S. Angiotensin II Signal Transduction: An Update on Mechanisms of Physiology and Pathophysiology. Physiol. Rev. 2018, 98, 1627–1738. [Google Scholar] [CrossRef]
  171. Goumans, M.-J.; Liu, Z.; Dijke, P.T. TGF-β signaling in vascular biology and dysfunction. Cell Res. 2008, 19, 116–127. [Google Scholar] [CrossRef]
  172. Sommer, C.; Kress, M. Recent findings on how proinflammatory cytokines cause pain: Peripheral mechanisms in inflammatory and neuropathic hyperalgesia. Neurosci. Lett. 2004, 361, 184–187. [Google Scholar] [CrossRef] [PubMed]
  173. Zhang, J.; Alcaide, P.; Liu, L.; Sun, J.; He, A.; Luscinskas, F.W.; Shi, G.-P. Regulation of Endothelial Cell Adhesion Molecule Expression by Mast Cells, Macrophages, and Neutrophils. PLoS ONE 2011, 6, e14525. [Google Scholar] [CrossRef] [PubMed]
  174. Tamagno, E.; Parola, M.; Bardini, P.; Piccini, A.; Borghi, R.; Guglielmotto, M.; Santoro, G.; Davit, A.; Danni, O.; Smith, M.A.; et al. β-Site APP cleaving enzyme up-regulation induced by 4-hydroxynonenal is mediated by stress-activated protein kinases pathways. J. Neurochem. 2005, 92, 628–636. [Google Scholar] [CrossRef] [PubMed]
  175. Hwang, S.-Y.; Kim, J.-Y.; Kim, K.-W.; Park, M.-K.; Moon, Y.; Kim, W.-U.; Kim, H.-Y. IL-17 induces production of IL-6 and IL-8 in rheumatoid arthritis synovial fibroblasts via NF-κB- and PI3-kinase/Akt-dependent pathways. Arthritis Res. Ther. 2004, 6, R120–R128. [Google Scholar] [CrossRef] [PubMed]
  176. Waugh, D.J.; Wilson, C. The Interleukin-8 Pathway in Cancer. Clin. Cancer Res. 2008, 14, 6735–6741. [Google Scholar] [CrossRef]
  177. Harty, J.T.; Badovinac, V.P. Shaping and reshaping CD8+ T-cell memory. Nat. Rev. Immunol. 2008, 8, 107–119. [Google Scholar] [CrossRef]
  178. Salvadores, N.; Gerónimo-Olvera, C.; Court, F.A. Axonal Degeneration in AD: The Contribution of Aβ and Tau. Front. Aging Neurosci. 2020, 12, 581767. [Google Scholar] [CrossRef]
  179. Ma, Q.; Chen, Z.; Barrantes, I.d.B.; de la Pompa, J.L.; Anderson, D.J. neurogenin1 Is Essential for the Determination of Neuronal Precursors for Proximal Cranial Sensory Ganglia. Neuron 1998, 20, 469–482. [Google Scholar] [CrossRef] [PubMed]
  180. Brindley, D.N.; Pilquil, C. Lipid phosphate phosphatases and signaling. J. Lipid Res. 2009, 50, S225–S230. [Google Scholar] [CrossRef]
  181. Venkatachalam, K.; Montell, C. TRP Channels. Annu. Rev. Biochem. 2007, 76, 387–417. [Google Scholar] [CrossRef]
  182. Kawai, T.; Akira, S. The role of pattern-recognition receptors in innate immunity: Update on Toll-like receptors. Nat. Immunol. 2010, 11, 373–384. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Flow diagram of the literature screening process for the studies included in our review.
Figure 1. Flow diagram of the literature screening process for the studies included in our review.
Ijms 25 09551 g001
Figure 2. Overview of miRNAs implicated in dysregulated processes manifesting myalgic encephalitis/chronic fatigue syndrome (ME/CFS) and/or fibromyalgia (FM) that were reported at least two times in different research studies.
Figure 2. Overview of miRNAs implicated in dysregulated processes manifesting myalgic encephalitis/chronic fatigue syndrome (ME/CFS) and/or fibromyalgia (FM) that were reported at least two times in different research studies.
Ijms 25 09551 g002
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

Tsamou, M.; Kremers, F.A.C.; Samaritakis, K.A.; Roggen, E.L. Identifying microRNAs Possibly Implicated in Myalgic Encephalomyelitis/Chronic Fatigue Syndrome and Fibromyalgia: A Review. Int. J. Mol. Sci. 2024, 25, 9551. https://doi.org/10.3390/ijms25179551

AMA Style

Tsamou M, Kremers FAC, Samaritakis KA, Roggen EL. Identifying microRNAs Possibly Implicated in Myalgic Encephalomyelitis/Chronic Fatigue Syndrome and Fibromyalgia: A Review. International Journal of Molecular Sciences. 2024; 25(17):9551. https://doi.org/10.3390/ijms25179551

Chicago/Turabian Style

Tsamou, Maria, Fabiënne A. C. Kremers, Keano A. Samaritakis, and Erwin L. Roggen. 2024. "Identifying microRNAs Possibly Implicated in Myalgic Encephalomyelitis/Chronic Fatigue Syndrome and Fibromyalgia: A Review" International Journal of Molecular Sciences 25, no. 17: 9551. https://doi.org/10.3390/ijms25179551

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