Summary
Rheumatoid arthritis is characterized by synovial proliferation, neovascularization and leucocyte extravasation leading to joint destruction and functional disability. The blood vessels in the inflamed synovium are highly dysregulated, resulting in poor delivery of oxygen; this, along with the increased metabolic demand of infiltrating immune cells and inflamed resident cells, results in the lack of key nutrients at the site of inflammation. In these adverse conditions synovial cells must adapt to generate sufficient energy to support their proliferation and activation status, and thus switch their cell metabolism from a resting regulatory state to a highly metabolically active state. This alters redox‐sensitive signalling pathways and also results in the accumulation of metabolic intermediates which, in turn, can act as signalling molecules that further exacerbate the inflammatory response. The RA synovium is a multi‐cellular tissue, and while many cell types interact to promote the inflammatory response, their metabolic requirements differ. Thus, understanding the complex interplay between hypoxia‐induced signalling pathways, metabolic pathways and the inflammatory response will provide better insight into the underlying mechanisms of disease pathogenesis.
Keywords: arthritis (including rheumatoid arthritis), autoimmunity, inflammation
Hypoxia and metabolism in RA
Rheumatoid arthritis (RA) is a chronic progressive autoimmune disease, and is characterized by proliferation of the synovial membrane leading to degradation of articular cartilage and bone, and thus functional disability 1. An increase in the number of blood vessels with an immature phenotype, displaying incomplete pericyte coverage, has been demonstrated in the RA synovium 2, 3, 4. These blood vessels are dysregulated, and are thought to remain in a ‘plastic state’, thus primed for endothelial cell activation and sprouting, further facilitating immune cell recruitment 2, 3, 5. The increased metabolic demand of the expanding synovial pannus leads to oxidative stress, altered cellular bioenergetics and a hypoxic microenvironment, which further promotes abnormal cell function and synovial invasiveness within the RA joint 6. The original hypothesis that the inflamed joint is hypoxic was based on studies measuring surrogate markers of hypoxia in RA synovial fluids, where an increase in glycolytic metabolites in the joint was observed 7, 8. Subsequent studies have shown that the inflamed joint is profoundly hypoxic 9, 10, 11, levels of which are inversely associated with increased synovitis, dysfunctional vascularity and microscopic inflammation 11, 12. Furthermore, hypoxia‐mediated effects within the RA inflamed joint are not only dependent on HIF‐1α but involve complex interactions between key proinflammatory signalling pathways, hypoxia‐inducible factor 1‐alpha (HIF‐1α), nuclear factor kappa light‐chain‐enhancer of activated B cells (NF‐κB), Notch‐1 intracellular domain (Notch‐1), Janus kinase/signal transducers and activators of transcription (JAK‐STAT) and phosphoinositide 3‐kinase‐protein kinase B (PI3K‐AKT) in synovial tissue and cells 13, 14, 15, 16.
Evidence of a key role for metabolism in the regulation of synovial inflammation has recently emerged, where proliferation and rapid activation of immune and stromal cells requires a switch in cell metabolism from a resting regulatory state to a highly metabolically active state, in order to maintain energy homeostasis 6, 17. Under normoxic conditions, one glucose molecule enters the cell and is oxidized by glycolysis, generating two molecules of pyruvate, and in the presence of oxygen, pyruvate is decarboxylated by pyruvate dehydrogenase (PDH) then enters the tricarboxylic acid (TCA) cycle to produce 36 molecules of adenosine triphosphate (ATP) via a process known as oxidative phosphorylation (OXPHOS) (Fig. 1). In the absence of oxygen, pyruvate is diverted away from the TCA cycle, and instead pyruvate is converted to lactate by lactate dehydrogenase (LDH) in the cytosol, generating two molecules of ATP (Fig. 1). However, these metabolic changes can also occur in the presence of oxygen, a process known as the ‘Warburg effect’. Although glycolysis is less efficient in terms of ATP production compared to the TCA cycle it produces ATP more rapidly, thus better meets the energy demands of activated synovial immune cells. In addition, many glycolytic‐derived intermediate metabolites diverge into parallel pathways to promote nucleotide synthesis through the pentose phosphate pathway (PPP) to support cell proliferation and survival 18 and de‐novo fatty acid synthesis required for the expansion of the endoplasmic and Golgi membranes for the synthesis, trafficking and secretion of proteins 19.
Figure 1.

Schematic illustration of the key metabolic pathways. Glucose enters the cell via glucose transporters and enters the glycolytic pathway. Hexokinase 2 (HK2) converts glucose into glucose 6‐phosphate dehydrogenase (G6PD). Glycolysis generates pyruvate from glucose with the help of pyruvate kinase M2 (PKM2). This process generates energy in the form of adenosine triphosphate (ATP). Pyruvate is then either converted to lactate and secreted out of the cell or decarboxylated by pyruvate dehydrogenase and converted to acetyl CoA, which enters the tricarboxylic acid (TCA) cycle. The TCA cycle generates nicotinamide adenine dinucleotide (NADH), flavin adenine dinucleotide (FADH2) to feed into the electron transport chain (ETC), which produces 36 molecules of ATP. Glycolysis also feeds into the pentose phosphate pathway (PPP) to produce ribose, NADPH and amino acids. Amino acid metabolism can also feed into the TCA cycle to drive ATP production by the ETC. TCA intermediate citrate drives fatty acid synthesis while fatty acid oxidation drives TCA cycle further by generating acetyl CoA.
Evidence for this metabolic switch in RA has been demonstrated by several studies showing an increase in mitochondrial dysfunction in the RA synovium 20, along with increased expression of glucose transporters (GLUTs) 21 and glycolytic enzymes, including hexokinase 2 (HK2), pyruvate kinase M2 (PKM2), glyceraldehyde 3‐phosphate dehydrogenase (GAPDH), glucose‐6‐phosphate isomerase (GPI) and LDH 7, 8, 21, 22, 23, 24, 25. Marked accumulation of metabolic intermediates including lactate, glutamine, succinate 26 and itaconate 27 have been demonstrated in the RA joint which, in turn, can activate synovial cells 21, 28, 29, further perpetuating disease. In animal models of arthritis and in ex‐vivo RA synovial tissue explants, several studies have shown that re‐programming of these pathways with specific inhibitors leads to resolution of inflammation 21, 30, 31, 32, 33, 34.
The RA synovium is a multi‐cellular tissue, and many cell types interact to promote the inflammatory response. While synovial cells including T cells, macrophages, dendritic cells and synovial fibroblasts co‐exist in this microenvironment, they utilize metabolites differently. Indeed, metabolites produced by one cell can have profound regulatory effects on the function of another cell within the RA synovium. This highlights the need to dissect metabolic pathways and utilization of metabolites in different cell types within the RA synovium, as any potential metabolic therapeutic treatment may be cell type‐specific. While there is a significant number of studies examining metabolic pathways in different immune cell subtypes, few studies have investigated these pathways in the context of the individual cells in the RA joint.
The role of T cell metabolism in RA
Growing understanding of the relationship between metabolic processes within T cells and their activation, differentiation and effector functions is emerging. It is now well established that whereas resting T cells primarily utilize oxidative phosphorylation, aerobic glycolysis is promoted within minutes of T cell receptor (TCR) activation independently of transcription or translation via activation of PDHK1 35, followed by a more sustained induction of the glycolytic machinery 34. This rapid metabolic reprogramming towards aerobic glycolysis provides the energy and substrates necessary for the proliferation and effector functions of activated T cells. In addition, the differentiation of T helper (Th) into Th17 versus regulatory T cells (Treg) cell is profoundly influenced by metabolism. Th17 cell differentiation requires glycolysis and HIF‐1α, whereas HIF‐1α suppresses Treg cell differentiation 31, 36. Despite these recent advances, little is known about how metabolism influences human T cell responses, in particular those of already differentiated memory T cells that are found at sites of inflammation such as the RA joint. There are a number of site‐specific factors that are likely to impact upon the metabolism of T cells within the inflamed RA joint, including hypoxia and activation of HIF‐1α, altered nutrient availability and the accumulation of metabolites.
The hypoxic environment of the RA joint might be expected to drive glycolysis in T cells via the induction of HIF‐1α. However, in the circulation there is recent evidence to suggest that this is not, in fact, the case. Rather, naive peripheral blood CD4 T cells from RA patients have been shown to have a defect in 6‐phosphofructo‐2‐kinase/fructose‐2,6‐biphosphatase 3 (PFKFB3), which is the rate‐limiting enzyme in the glycolytic pathway 37. This inhibits their ability to utilize glycolysis upon activation and instead diverts glucose‐6‐phosphate towards the PPP, which runs in parallel to glycolysis, PPP supporting mainly nucleotide synthesis 37. The PPP generates the 5‐carbon building blocks for nucleic acids in addition to significant reductive power in the form of nicotinamide adenine dinucleotide phosphate (NADPH) 37. NADPH is required for the process of fatty acid synthesis; indeed, subsequent work by the same group revealed that increased fatty acid synthesis and accumulation of lipid droplets within the cell was a consequence of the reduced glycolytic flux within RA T cells 34. Furthermore, these conditions resulted in the up‐regulation of tyrosine kinase substrate with five SH3 domains (TKS5) expression, which promoted tissue invasiveness of the RA T cells. RA T cell invasiveness could be reversed via inhibition of fatty acid synthesis, suggesting that the fatty acid synthesis pathway may represent a therapeutic target in RA 34. These studies have provided important insights into the metabolic status of CD4 T cells in RA. The role of CD8 T cells in RA has received less attention than that of CD4 T cells; however, future studies will be important to determine whether CD8 T cells in RA are subject to similar metabolic control.
The inflammatory environment of the RA joint leads to the accumulation of various metabolites 26, 28, and it has now become apparent that these metabolites are not simply by‐products of metabolism but also exert important immune modulating effects. Interestingly, lactate has been shown to modulate T cell function directly via specific cell surface lactate transporters 28. Sodium lactate inhibits the migration of CD4 T cells by inhibition of glycolysis via direct inhibition of HK and PFK 28. In addition, sodium lactate also promotes the expression of interleukin (IL)‐17 by CD4 T cells 28. Furthermore, expression of the lactate transporter SLC5a12 in synovial tissue was shown to correlate with the clinical T cell score, and in‐vivo blockade of lactate transporters resulted in the release of T cells from the inflammatory site 28. These data suggest that the elevated lactate concentrations observed in the RA joint inhibit the glycolysis and migration of effector T cells, resulting in their retention at the site of inflammation and also increasing their production of IL‐17. These findings provide a rationale for the therapeutic targeting of specific lactate transporters on T cells in autoimmune diseases such as RA (reviewed in 38). Consistent with this, expression levels of the lactate transporter MCT4 and glycolytic enzymes HK2, GPI, triosephosphate isomerase (TPI), enolase 1 (Eno 1), PKM2 and LDH are significantly reduced in Th17 cells obtained from HIF1α–/– compared to wild‐type (WT) mice 31, demonstrating the dependence of increased glycolytic enzymatic activity on the oxygen sensing pathway. Furthermore, in animal models of arthritis 3‐bromopyruvate (BrPA), a specific HK2 inhibitor, significantly decreased clinical arthritis scores in SKG mice, paralleled by an increase in the Treg/Th17 ratio 30. While the exact underlying mechanisms involved are unclear, it is hypothesized that BrPA may alter the Treg/Th17 ratio through differential regulation of their respective transcription factors forkhead box protein 3 (FoxP3) and retinoic acid receptor‐related orphan nuclear receptor gamma t (ROR‐γt). Finally, the TCA cycle metabolite succinate has been shown to exert proinflammatory effects on murine macrophages via induction of IL‐1β 39 which, in turn, might be expected to drive Th17 responses. Indeed, a recent study showed that deficiency of the succinate receptor GPR91 attenuates the severity of arthritis in Sucnr1–/– mice, reducing expansion of Th17 cells 40.
The role of monocyte and macrophage metabolism in RA
Activated macrophages promote a number of proinflammatory mechanisms in the RA synovium through abundant secretion of proinflammatory cytokines; in addition, macrophages induce nitric oxide synthase, present antigen to T and B cells and drive bone resorption 41. RA synovial CD68+ macrophages correlate strongly with mitochondrial dysfunction and oxidative stress and are inversely related to in‐vivo synovial pO2 levels 11, 12, 20. HIF‐1α is expressed abundantly by macrophages in the RA synovium, compared to osteoarthritis (OA) and healthy control synovial macrophages 42. Differential signalling mechanisms in monocytes and macrophages under hypoxic conditions have been observed where monocytes preferentially utilize NF‐κB1, while macrophages utilize HIF‐1α 43. Furthermore, in collagen‐induced arthritis (CIA) models, decreased infiltration of myeloid cells to the joint, reduced paw swelling and disease development was observed in animals with HIF‐1α‐deficient macrophages 44.
In the context of metabolic changes, studies have demonstrated comprehensively that classically activated M1 macrophages have an ardent appetite for glucose, indicating a reliance on glycolysis, in contrast to M2‐like macrophages which rely on OXPHOS 45, 46. Indeed, recent studies have shown that the TCA cycle is broken at two key steps in M1 macrophages – after citrate and after succinate 39, 46. The majority of research has, however, focused on in‐vitro monocyte‐derived M1 and M2 macrophage models, and not synovial macrophages, due to the difficulty of isolating these cells; however, it is now becoming apparent that the M1 versus M2 paradigm may be an over‐simplification, and that a spectrum of activation states exists between these two poles within the synovial joint. In the context of RA, studies have shown increased lactic acid, citrate and succinate in RA synovial fluids 26, consistent with studies showing increased glycolysis and a broken TCA cycle in M1 macrophages. Lactic acid enhances secretion of IL‐6 and IL‐23 from monocytes and macrophages 47. Accumulation of succinate in macrophages promotes HIF‐1α activation which, in turn, induces IL‐1β production 39 and in animal models of RA, mice lacking the succinate receptor GPR91 show reduced macrophage activation and secretion of IL‐1β 33. Itaconate, a metabolic inhibitor of succinate dehydrogenase which has been shown to regulate succinate levels and secretion of inflammatory cytokines in activated macrophages 48, is also increased in the RA joint, and is associated with disease activity and response to therapy in animal models of arthritis 27. RA macrophages express high amounts of the glycolytic enzyme α‐enolase, which through autoantibody recognition induces secretion of proinflammatory cytokines 49. High concentrations of glucose have also been shown to increase IL‐1β secretion from RA monocytes through an NLRP3‐dependent mechanism 50. Solute carrier family 7 member 5 (SLC7A5), a key amino acid transporter, is increased in RA monocytes and macrophages, silencing of which leads to a significant reduction of IL‐1β 51. Finally, activation of 5′ AMP‐activated protein kinase (AMPK) in macrophages inhibits IL‐6 production, differentiation of M2 macrophages from synovial fluid monocytes and macrophage expression of IL‐6, tumour necrosis factor (TNF)‐α and NF‐κB in animal models of arthritis 52, 53.
More recent data have shown that macrophages from both RA and coronary artery disease (CAD) patients share metabolic abnormalities to promote inflammation. Disease macrophages appear to be in a hypermetabolic state, addicted to glucose consumption and producing more ATP compared to healthy macrophages 54. In addition, it has been demonstrated that macrophages from CAD patients are capable of memorizing both the metabolic and inflammatory signatures of their precursor monocytes, indicating that there is a memory bias towards a hyperinflammatory and hypermetabolic phenotype in disease macrophages 55. There has been speculation that epigenetic regulation may be the cause of this memorized immune response in both monocytes and macrophages 56, 57, 58. Thus, epigenetic reprogramming is emerging as a key mechanism in macrophage immunometabolism.
The role of metabolism in DC activation in RA
Dendritic cells (DCs) are key players in immunity, and link the innate and adaptive immune response through antigen presentation and cytokine production. Resting DCs, which differ from activated DCs as they are less motile, secretory and immunogenic, rely predominantly on OXPHOS for their energetic needs 59, 60, 61. Activation of DCs, either by antigen exposure or stimulation by Toll‐like receptor (TLR) ligands, results in a two‐phased metabolic shift comprised of an initial glycolytic shift with maintained OXPHOS, followed by elevated glycolysis and the cessation of OXPHOS in long‐term activated DCs. While the early increase in glycolysis is associated with elevated GLUT1 expression 61, 62 and lactate production 59, 60, 61, 62, 63, 64, the increased glucose demand by activated DCs is associated with the requirement of glucose‐derived metabolic intermediates into the PPP and fatty acid synthesis to facilitate amino acid and protein synthesis for the secretion of proinflammatory mediators associated with DC activation. Inhibition of glycolysis via HK2 blockade or fatty acid synthase in lipopolysaccharide (LPS)‐stimulated DCs results in a marked reduction in DC activation and immunogenicity 60. Molecularly, TBK‐IKKε/AKT signalling pathways and downstream target mammalian target of rapamycin (mTOR) have been strongly implicated in early activation of the glycolytic shift, which can be prevented by adhesion‐related kinase (ARK) blockade 62. In the context of RA, while also modulating the Treg/Th17 ratio, BrPA‐mediated blockade of HK2 suppressed DC activation and cytokine expression 30, suggesting that the glycolysis pathway may represent a potential therapeutic target in RA.
During the later stages of DC activation, mTOR‐mediated induction of inducible nitric oxide synthase (iNOS) and stabilization of HIF‐1α is important for the complete commitment of activated DCs to glycolysis 59, 62, with iNOS‐derived nitric oxide (NO) blocking the electron transport chain (ETC) and sequestering OXPHOS 59 while elevating HIF1α‐dependent glycolytic gene expression 62, 63, 65. While these metabolic changes remain to be studied in the complex nutrient and/or oxygen‐deprived microenvironments, such as those found in the inflamed synovium, similar to macrophages, DC can also sense and respond to extracellular metabolites, such as succinate, butyrate and ATP 33, 66, 67, 68, 69 to potentially trigger a co‐ordinated cellular response reflective of the metabolic state of the surrounding microenvironment. A recent study demonstrated that the succinate receptor, GPR91, acts as a chemotactic that facilitates migration of DCs into the lymph nodes, which induces the expansion of Th17 cells and subsequent development of experimental antigen‐induced arthritis 40. This is consistent with previous studies that demonstrated that succinate can promote chemotaxis of DCs through activation of the succinate receptor 66, 67.
Role of metabolism in synovial fibroblast activation in RA
Synovial fibroblasts (FLS) are fundamental to disease progression and are active drivers of joint destruction in RA 70. FLS are characterized by increased proliferation, resistance to apoptosis and are potent producers of proinflammatory cytokines and matrix degradation enzymes, resulting in a highly invasive phenotype 71, 72. Previous studies have demonstrated increased mitochondrial dysfunction, coupled by changes in the ultrastructure of mitochondria in RA‐FLS in response to inflammatory stimuli, in addition to altered expression of mitochondrial genes associated with apoptosis, redox balance and mitochondrial protein transport 21, 23, 73. Metabolic profiling of FLS demonstrated increases in sugar metabolism (glycolysis and PPP) and amino acid metabolism (tyrosine and catecholamine biosynthesis and protein biosynthesis) 74. RA‐FLS also demonstrate a reliance on glutamine metabolism, with in‐vitro studies demonstrating inhibition of RA‐FLS proliferative and invasive functions under glutamine‐deprived conditions 29. Hypoxia, oxidative stress, TLR2 and TNF‐α activation all promote a switch in the metabolic profile of RA‐FLS, where an increase in the glycolysis : OXPHOS ratio is observed, paralleled by increases in surrogate markers of glycolysis; PFKFB3, PKM2 and GLUT1 and a more invasive phenotype 21, 23, 75, 76. Glycolytic inhibitors, including the PFKFB3 inhibitor 3‐(3‐pyridinyl)‐1‐(4‐pyridinyl)‐2‐propen‐1‐one (3PO) 21 and HK2 inhibitor, BrPA 76 significantly inhibit RA‐FLS invasion and migration capacity, secretion of proinflammatory mediators and activation of HIF‐1α, pSTAT‐3, NF‐κB and Notch‐1IC. Furthermore, accumulation of metabolic intermediates, including lactic acid, glutamine and succinate in the RA joint, have all been shown to further induce the RA‐FLS invasive phenotype 21, 76. Other key pathways involved in FLS activation include the PI3K/AKT1/mTOR, blockade of which resulted in repressed RA‐FLS function 77. Studies have shown that mTOR blockade of RA‐FLS invasion is mediated in part through regulation of focal adhesion kinase (FAK) signalling pathways and cytoskeletal rearrangement, a key mechanism involved in RA‐FLS movement 78. Furthermore, a recent study has shown that TNF‐α signalling co‐opts the mTOR pathway, directing specific signalling pathways that regulate the RA‐FLS response to inflammatory stimuli, an effect coupled with nutrient availability of specific amino acids 79. This is consistent with another study showing that IL‐17‐induced RA‐FLS migration is mediated through the amino acid transporter (LAT1) via mTOR 80.
Metabolomic profiling has shown changes in lipid metabolism of RA‐FLS compared to OA cells 76; however, the role of lipids in the regulation of FLS function has not been studied extensively. Interestingly, studies have shown that molecules in the choline pathway, which can interact with lipids, are highly activated in RA‐FLS, with the choline kinase (ChoKα) enzyme 81 and choline transporters 82 increased in RA‐FLS. Furthermore, a ChoKα inhibitor suppressed the migrative/invasive mechanisms of cultured RA‐FLS and in vivo ameliorated inflammation in the KxBN model 81. Thus, RA‐FLS are transformed from a quiescent state to an aggressive, invasive phenotype in this adverse environment through adaptation of metabolic pathways in order to meet their energy demands, which allows them to resist apoptosis and persist within the inflamed joint.
The above findings in FLS, however, are all performed in RA‐FLS which are maintained in culture in vitro, so while they provide some understanding of the metabolic pathways activated in RA‐FLS in response to specific stimuli, we still lack an understanding of the metabolic demands of RA‐FLS subtypes within this adverse inflammatory joint environment. Indeed, it has now become apparent that there are subsets of synovial fibroblasts within the inflamed joint that display pro‐ and anti‐inflammatory phenotypes, with effects also dependent upon positional memory 83, 84, 85. Therefore, further studies are required to examine if different metabolic profiles are observed depending on the RA‐FLS subtype and the anatomical location.
Metabolism of synovial endothelium
Endothelial cells (EC) rely heavily on glycolysis, with 85% of the ATP requirements coming from the conversion of glucose to lactate 86. The three EC subsets (tip, stalk and phalanx cells) differ in their metabolic requirements reflective of their individual functions (migration, proliferation and quiescence, respectively). When ECs are activated, metabolic changes dictate phenotypical differentiation, with tip and stalk cells showing increased glycolytic rates compared to phalanx EC 87. In the context of RA, few studies have extensively studied metabolic regulation of synovial vessels; however, they display similar dysfunctional morphology to that of the tumour vasculature 2. RA synovial blood vessel instability and oxidative damage correlate inversely with synovial pO2 levels 2, and display enhanced expression of the glucose transporter GLUT1 and the glycolytic enzymes GAPDH and PKM2 21. Exposure of ECs to hypoxic conditions and oxidative stress induces tube formation, migration and proinflammatory mediators, including vascular endothelial growth factor (VEGF), angiopoietins, monocyte chemoattractant protein 1 (MCP‐1), IL‐8 and matrix metalloproteinases (MMPs) 2, 75, 88. In RA‐FLS and EC, hypoxia induces activation of intracellular Notch‐1IC and its ligand DLL‐4, interaction of which is critical for EC tip cell selection, and lateral inhibition of the trailing stalk cell 13, 14. Blockade of the glycolytic enzyme, PFKFB3, inhibits angiogenic tube formation, secretion of proinflammatory/angiogenic mediators and key signalling pathways in RA‐FLS and EC 21. In addition, enriched expression of G6PI in synovial endothelial cells has recently been shown, with in‐vitro G6PI loss‐of‐function assays demonstrating the requirement of G6PI in mediating hypoxia‐induced angiogenesis in RA 89. Finally, in animal models of arthritis, succinate has been shown to induce synovial angiogenesis through VEGF‐dependent HIF‐1α pathways 90. Therefore, a deeper understanding of the metabolic perturbations in pathological conditions and their cross‐talk with immune and stromal cells might offer novel therapeutic opportunities, especially in early disease.
Targeting metabolism in rheumatoid arthritis
As highlighted above, studies examining metabolic pathways in RA have identified specific pathways, enzymes or metabolic intermediates that could potentially be targeted; however, many of our current treatment strategies are already known to alter metabolic pathways. For instance, glucocorticoids which are used routinely as first‐line treatment and regulate transcription of many metabolic genes 91 associated with glycolytic, autophagy and mTOR pathways 92, 93. Conventional disease‐modifying anti‐rheumatic drugs (DMARDs) for RA and psoriatic arthritis (PsA), including methotrexate, leflunomide and apremilast, are anti‐metabolic, where they target purine or pyrimidine nucleotide metabolism, the effects of which are known to inhibit both T cells and synovial fibroblast proliferation. Studies have shown that anti‐TNF‐α treatment decreases expression of GLUT1 and key glycolytic enzymes PKM2 and GAPDH in RA synovium in TNFi responders versus non‐responders 21. Tocilizumab, an anti‐IL‐6 receptor antibody, improved endothelial function and inhibited oxidative stress in RA leucocytes 94. Tofacitinib, a JAK1 and JAK3 inhibitor, reduces glycolysis in activated synovial fibroblasts, and in RA synovial explants inhibits key glycolytic enzymes, paralleled by reduced expression of key proinflammatory mediators and synovial fibroblast outgrowths 95. Indeed, interaction of STAT‐3 and PKM2 leads to activation of HIF1α, with subsequent induction of cellular invasive mechanisms creating a vicious PKM2/STAT‐3/HIF1 feedback loop 96. Furthermore, STAT‐3 blockade inhibits Notch signalling, which is involved in endothelial tip cell selection and fibroblast invasive mechanisms 1.
Directly targeting specific metabolic pathways has been demonstrated both in in‐vitro and in‐vivo models of arthritis where blockade of key glycolytic enzymes inhibits proinflammatory mechanisms 21, 30, 55, 97. Targeting metabolic intermediates including lactate acid, succinate, citrate, itaconate and lipids are also promising therapeutic avenues, where their cellular accumulation regulates synovial fibroblast invasiveness 21, T cell differentiation and migration 28, 31, 34, in addition to macrophage polarization 54. Furthermore, the anti‐cancer drug rapamycin, which targets mTOR, plays a critical role in directing T cell differentiation and function, inhibition of which promotes Treg cell generation 98. mTOR is also a major repressor of autophagy, and in animal models of arthritis, systemic administration of rapamycin was shown to induce autophagy, paralleled by decreased severity of synovitis and reduction in IL‐1β expression 99. Finally, metformin, the anti‐diabetic drug, which acts in part by indirectly activating the energy sensor 5' AMP‐activated protein kinase (AMPK), has been shown to attenuate disease in mouse models of arthritis 100, effects that are mediated by inhibition of mTOR activity, enhanced autophagic flux, suppression of NF‐κB signalling and inflammatory cytokine production 101.
Summary
Lack of nutrients and a poor oxygen supply, paralleled by the increased metabolic demand of the expanding synovial pannus, leads to a bioenergetic crisis. In this environment synovial cells show adaptive survival responses by switching their utilization of specific metabolic pathways (Fig. 2) in order to satisfy their energy demands. This, in turn, activates key transcriptional signalling pathways which further exacerbates inflammation. However, this environment is complex, with many different cell types co‐existing within the inflamed synovium that display different metabolic requirements. Furthermore, metabolites secreted from one cell type have the ability to regulate the pathogenic phenotype of another, thus amplifying the inflammatory response. Understanding the opposing metabolic requirements of the different cell types will provide significant insights into their relevant pathogenic contribution to disease. However, currently the metabolic regulation of specific cell‐types/subsets within the inflamed synovium are poorly understood. Metabolic regulation can also have opposing effects depending on the disease settings. For instance, while Haas et al. elegantly described that high lactate levels at the site of inflammation act to entrap CD4+ T cells in RA by inhibiting T cell motility 28, in the tumour microenvironment accumulation of lactate impairs T cell function and hinders their cytotoxicity 102. Thus, lactate serves to boost proinflammatory mechanisms in the inflammatory milieu, yet suppresses immunity in the cancer setting. Enolase has been shown to induce monocyte/macrophage activation in models of RA 49, but promotes Treg cell development in cancer models 103. Arginine metabolism drives a pro‐glycolytic phenotype in macrophages and DCs 60, 104; however, it promotes a metabolic switch to OXPHOS, exerting anti‐inflammatory effects in T cells 105, 106. Finally, the amino acid transporter SLC7A5, through leucine influx, induces proinflammatory cytokine secretion in RA monocytes and macrophages via mTORC1‐induced glycolytic reprogramming 107, and has also been demonstrated to play a key role in tumour survival and growth 108, 109. Thus, significant additional research is required if therapeutic strategies targeting metabolism are to be successful.
Figure 2.

Schematic illustration of the glycolytic switch in different cell types. The hypoxic conditions of the synovial joint drives hypoxia‐inducible factor 1‐alpha (HIF‐1a)‐induced glycolysis in some of the major cell types of the synovium. HIF‐1a induces expression of some of the key molecular switches to encode proinflammatory and pro‐glycolytic mechanisms. Specifically, 6‐phosphofructo‐2‐kinase/fructose‐2,6‐biphosphatase 3 (PFKFB3) is up‐regulated in endothelial cells and synovial fibroblasts in response to hypoxia. The mammalian target of rapamycin (mTOR) pathway is also involved in synovial fibroblast activation. Pyruvate dehydrogenase (PKM2) plays a central role in the metabolic switch observed in inflammatory macrophages and activated monocytes. DC metabolism can be directed by protein kinase B (AKT) signalling pathways in early activation while mTOR‐mediated induction of inducible nitric oxide synthase (iNOS) is important during later stages of DC activation, while glucose‐6‐phosphate dehydrogenase (G6PD) and PFKFB3 are key players in the metabolic switch observed in effector T cells.
Future research will require in‐depth characterization of immune and stromal cells at the site of inflammation. Isolation of single cell suspensions from the synovium for cell sorting, single cell transcriptomics and CyTOF analysis, paralleled by advanced imaging technology of synovial tissue sections, will improve our understanding with regard to which metabolic pathways are driving pathogenic phenotypes at the site of inflammation. Additional characterization of specific subsets, which have been recently described for synovial fibroblasts and macrophages, will further facilitate the development of new treatments that will specifically target pathogenic cell types as opposed to protective phenotypes. Furthermore, analysis of these pathways in pre‐RA (arthralgia), early disease RA, established RA and in responders versus non‐responders to current treatment strategies will also aid in our understanding of metabolism and its role in disease pathogenesis in RA.
Disclosure
None.
OTHER ARTICLES PUBLISHED IN THIS REVIEW SERIES
Translating immunometabolism: towards curing human diseases by targeting metabolic processes underpinning the immune response. Clinical and Experimental Immunology 2019, 197: 141–142.
T cell metabolism in chronic viral infection. Clinical and Experimental Immunology 2019, 197: 143–152.
Sculpting tumor microenvironment with immune system: from immunometabolism to immunoediting. Clinical and Experimental Immunology 2019, 197: 153–160.
Sensing between reactions – how the metabolic microenvironment shapes immunity. Clinical and Experimental Immunology 2019, 197: 161–169.
Immune cell metabolism in autoimmunity. Clinical and Experimental Immunology 2019, 197: 181–192.
Metabolism at the centre of the host‐microbe relationship. Clinical and Experimental Immunology 2019, 197: 193–204.
References
- 1. McInnes IB, Schett G. Pathogenetic insights from the treatment of rheumatoid arthritis. Lancet (Lond) 2017;389:2328–37. [DOI] [PubMed] [Google Scholar]
- 2. Kennedy A, Ng CT, Biniecka M et al Angiogenesis and blood vessel stability in inflammatory arthritis. Arthritis Rheum 2010;62:711–21. [DOI] [PubMed] [Google Scholar]
- 3. Izquierdo E, Cañete JD, Celis R et al Immature blood vessels in rheumatoid synovium are selectively depleted in response to anti‐TNF therapy. PLOS ONE 2009;4:e8131. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4. Tas SW, Maracle CX, Balogh E et al Targeting of proangiogenic signalling pathways in chronic inflammation. Nat Rev Rheumatol 2016;12:111–22. [DOI] [PubMed] [Google Scholar]
- 5. Gao W, Sweeney C, Walsh C et al Notch signalling pathways mediate synovial angiogenesis in response to vascular endothelial growth factor and angiopoietin 2. Ann Rheum Dis 2013;72:1080–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6. Fearon U, Canavan M, Biniecka M et al Hypoxia, mitochondrial dysfunction and synovial invasiveness in rheumatoid arthritis. Nat Rev Rheumatol 2016;12:385–97. [DOI] [PubMed] [Google Scholar]
- 7. Naughton DP, Haywood R, Blake DR et al A comparative evaluation of the metabolic profiles of normal and inflammatory knee‐joint synovial fluids by high resolution proton NMR spectroscopy. FEBS Lett 1993;332:221–5. [DOI] [PubMed] [Google Scholar]
- 8. Treuhaft PS, McCarty DJ. Synovial fluid pH, lactate, oxygen and carbon dioxide partial pressure in various joint diseases. Arthritis Rheum 1971;14:475–84. [DOI] [PubMed] [Google Scholar]
- 9. Etherington PJ, Winlove P, Taylor P et al VEGF release is associated with reduced oxygen tensions in experimental inflammatory arthritis. Clin Exp Rheumatol 2002;20:799–805. [PubMed] [Google Scholar]
- 10. Sivakumar B, Akhavani MA, Winlove CP et al Synovial hypoxia as a cause of tendon rupture in rheumatoid arthritis. J Hand Surg Am 2008;33:49–58. [DOI] [PubMed] [Google Scholar]
- 11. Ng CT, Biniecka M, Kennedy A et al Synovial tissue hypoxia and inflammation in vivo . Ann Rheum Dis 2010;69:1389–95. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12. Kennedy A, Ng CT, Chang TC et al Tumor necrosis factor blocking therapy alters joint inflammation and hypoxia. Arthritis Rheum 2011;63:923–32. [DOI] [PubMed] [Google Scholar]
- 13. Oliver KM, Garvey JF, Ng CT et al Hypoxia activates NF‐κB–dependent gene expression through the canonical signaling pathway. Antioxid Redox Signal 2009;11:2057–64. [DOI] [PubMed] [Google Scholar]
- 14. Gao W, McCormick J, Connolly M et al Hypoxia and STAT3 signalling interactions regulate pro‐inflammatory pathways in rheumatoid arthritis. Ann Rheum Dis 2015;74:1275–83. [DOI] [PubMed] [Google Scholar]
- 15. Gao W, Sweeney C, Connolly M et al Notch‐1 mediates hypoxia‐induced angiogenesis in rheumatoid arthritis. Arthritis Rheum 2012;64:2104–13. [DOI] [PubMed] [Google Scholar]
- 16. Li G‐Q, Zhang Y, Liu D et al PI3 kinase/Akt/HIF‐1α pathway is associated with hypoxia‐induced epithelial‐mesenchymal transition in fibroblast‐like synoviocytes of rheumatoid arthritis. Mol Cell Biochem 2013;372:221–31. [DOI] [PubMed] [Google Scholar]
- 17. Weyand CM, Goronzy JJ. Immunometabolism in early and late stages of rheumatoid arthritis. Nat Rev Rheumatol 2017;13:291–301. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18. Buchakjian MR, Kornbluth S. The engine driving the ship: metabolic steering of cell proliferation and death. Nat Rev Mol Cell Biol 2010;11:715–27. [DOI] [PubMed] [Google Scholar]
- 19. Röhrig F, Schulze A. The multifaceted roles of fatty acid synthesis in cancer. Nat Rev Cancer 2016;16:732–49. [DOI] [PubMed] [Google Scholar]
- 20. Harty LC, Biniecka M, O’Sullivan J et al Mitochondrial mutagenesis correlates with the local inflammatory environment in arthritis. Ann Rheum Dis 2012;71:582–8. [DOI] [PubMed] [Google Scholar]
- 21. Biniecka M, Canavan M, McGarry T et al Dysregulated bioenergetics: a key regulator of joint inflammation. Ann Rheum Dis 2016;75:2192–200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22. Morten KJ, Badder L, Knowles HJ. Differential regulation of HIF‐mediated pathways increases mitochondrial metabolism and ATP production in hypoxic osteoclasts. J Pathol 2013;229:755–64. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23. McGarry T, Biniecka M, Gao W et al Resolution of TLR2‐induced inflammation through manipulation of metabolic pathways in Rheumatoid Arthritis. Sci Rep 2017;7:43165. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24. Pejovic M, Stankovic A, Mitrovic DR. Lactate dehydrogenase activity and its isoenzymes in serum and synovial fluid of patients with rheumatoid arthritis and osteoarthritis. J Rheumatol 1992;19:529–33. [PubMed] [Google Scholar]
- 25. Lindy S, Uitto J, Turto H et al Lactate dehydrogenase in the synovial tissue in rheumatoid arthritis: total activity and isoenzyme composition. Clin Chim Acta 1971;31:19–23. [DOI] [PubMed] [Google Scholar]
- 26. Kim S, Hwang J, Xuan J et al Global metabolite profiling of synovial fluid for the specific diagnosis of rheumatoid arthritis from other inflammatory arthritis. PLOS ONE 2014;9:e97501. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27. Michopoulos F, Karagianni N, Whalley NM et al Targeted metabolic profiling of the Tg197 mouse model reveals itaconic acid as a marker of rheumatoid arthritis. J Proteome Res 2016;15:4579–90. [DOI] [PubMed] [Google Scholar]
- 28. Haas R, Smith J, Rocher‐Ros V et al Lactate regulates metabolic and pro‐inflammatory circuits in control of T cell migration and effector functions. PLOS Biol 2015;13:e1002202. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29. Takahashi S, Saegusa J, Sendo S et al Glutaminase 1 plays a key role in the cell growth of fibroblast‐like synoviocytes in rheumatoid arthritis. Arthritis Res Ther 2017;19:76. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30. Okano T, Saegusa J, Nishimura K et al 3‐bromopyruvate ameliorate autoimmune arthritis by modulating Th17/Treg cell differentiation and suppressing dendritic cell activation. Sci Rep 2017;7:42412. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31. Shi LZ, Wang R, Huang G et al HIF1alpha‐dependent glycolytic pathway orchestrates a metabolic checkpoint for the differentiation of TH17 and Treg cells. J Exp Med 2011;208:1367–76. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32. Schaller M, Burton DR, Ditzel HJ. Autoantibodies to GPI in rheumatoid arthritis: linkage between an animal model and human disease. Nat Immunol 2001;2:746–53. [DOI] [PubMed] [Google Scholar]
- 33. Littlewood‐Evans A, Sarret S, Apfel V et al GPR91 senses extracellular succinate released from inflammatory macrophages and exacerbates rheumatoid arthritis. J Exp Med 2016;213:1655–62. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34. Shen Y, Wen Z, Li Y et al Metabolic control of the scaffold protein TKS5 in tissue‐invasive, proinflammatory T cells. Nat Immunol 2017;18:1025–34. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35. Menk AV, Scharping NE, Moreci RS et al Early TCR signaling induces rapid aerobic glycolysis enabling distinct acute T cell effector functions. Cell Rep 2018;22:1509–21. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36. Bettelli E, Carrier Y, Gao W et al Reciprocal developmental pathways for the generation of pathogenic effector TH17 and regulatory T cells. Nature 2006;441:235–8. [DOI] [PubMed] [Google Scholar]
- 37. Yang Z, Fujii H, Mohan SV et al Phosphofructokinase deficiency impairs ATP generation, autophagy, and redox balance in rheumatoid arthritis T cells. J Exp Med 2013;210:2119–34. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38. Pucino V, Bombardieri M, Pitzalis C et al Lactate at the crossroads of metabolism, inflammation, and autoimmunity. Eur J Immunol 2017;47:14–21. [DOI] [PubMed] [Google Scholar]
- 39. Tannahill GM, Curtis AM, Adamik J et al Succinate is an inflammatory signal that induces IL‐1β through HIF‐1α. Nature 2013;496:238–42. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40. Saraiva AL, Veras FP, Peres RS et al Succinate receptor deficiency attenuates arthritis by reducing dendritic cell traffic and expansion of Th17 cells in the lymph nodes. FASEB J 2018;201800285. [DOI] [PubMed] [Google Scholar]
- 41. Kurowska‐Stolarska M, Alivernini S. Synovial tissue macrophages: friend or foe? RMD Open 2017;3:e000527. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42. Hollander AP, Corke KP, Freemont AJ et al Expression of hypoxia‐inducible factor 1alpha by macrophages in the rheumatoid synovium: implications for targeting of therapeutic genes to the inflamed joint. Arthritis Rheum 2001;44:1540–4. [DOI] [PubMed] [Google Scholar]
- 43. Fangradt M, Hahne M, Gaber T et al Human monocytes and macrophages differ in their mechanisms of adaptation to hypoxia. Arthritis Res Ther 2012;14:R181. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44. Cramer T, Yamanishi Y, Clausen BE et al HIF‐1alpha is essential for myeloid cell‐mediated inflammation. Cell 2003;112:645–57. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45. Ganesh K, Das A, Dickerson R et al Prostaglandin E₂ induces oncostatin M expression in human chronic wound macrophages through Axl receptor tyrosine kinase pathway. J Immunol 2012;189:2563–73. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46. O’Neill LAJ. A broken Krebs cycle in macrophages. Immunity 2015;42:393–4. [DOI] [PubMed] [Google Scholar]
- 47. Shime H, Yabu M, Akazawa T et al Tumor‐secreted lactic acid promotes IL‐23/IL‐17 proinflammatory pathway. J Immunol 2008;180:7175–83. [DOI] [PubMed] [Google Scholar]
- 48. Lampropoulou V, Sergushichev A, Bambouskova M et al Itaconate links inhibition of succinate dehydrogenase with macrophage metabolic remodeling and regulation of inflammation. Cell Metab 2016;24:158–66. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49. Bae S, Kim H, Lee N et al α‐Enolase expressed on the surfaces of monocytes and macrophages induces robust synovial inflammation in rheumatoid arthritis. J Immunol 2012;189:365–72. [DOI] [PubMed] [Google Scholar]
- 50. Ruscitti P, Cipriani P, Di Benedetto P et al Monocytes from patients with rheumatoid arthritis and type 2 diabetes mellitus display an increased production of interleukin (IL)‐1β via the nucleotide‐binding domain and leucine‐rich repeat containing family pyrin 3(NLRP3)‐inflammasome activation: a pos. Clin Exp Immunol 2015;182:35–44. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51. Yoon BR, Oh Y‐J, Kang SW et al Role of SLC7A5 in metabolic reprogramming of human monocyte/macrophage immune responses. Front Immunol 2018;9:53. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52. Guma M, Wang Y, Viollet B et al AMPK activation by A‐769662 controls IL‐6 expression in inflammatory arthritis. PLOS ONE 2015;10:e0140452. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53. Yan H, Zhou H‐F, Hu Y et al Suppression of experimental arthritis through AMP‐activated protein kinase activation and autophagy modulation. J Rheum Dis Treat 2015;1:5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54. Zeisbrich M, Yanes RE, Zhang H et al Hypermetabolic macrophages in rheumatoid arthritis and coronary artery disease due to glycogen synthase kinase 3b inactivation. Ann Rheum Dis 2018;77:1053–62. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55. Shirai T, Nazarewicz RR, Wallis BB et al The glycolytic enzyme PKM2 bridges metabolic and inflammatory dysfunction in coronary artery disease. J Exp Med 2016;213:337–54. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56. Cheng S‐C, Quintin J, Cramer RA et al mTOR‐ and HIF‐1α‐mediated aerobic glycolysis as metabolic basis for trained immunity. Science 2014;345:1250684. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57. Saeed S, Quintin J, Kerstens HHD et al Epigenetic programming of monocyte‐to‐macrophage differentiation and trained innate immunity. Science 2014;345:1251086. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58. Neele AE, Van den Bossche J, Hoeksema MA et al Epigenetic pathways in macrophages emerge as novel targets in atherosclerosis. Eur J Pharmacol 2015;763:79–89. [DOI] [PubMed] [Google Scholar]
- 59. Everts B, Amiel E, van der Windt GJW et al Commitment to glycolysis sustains survival of NO‐producing inflammatory dendritic cells. Blood 2012;120:1422–31. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60. Everts B, Amiel E, Huang SC‐C et al TLR‐driven early glycolytic reprogramming via the kinases TBK1‐IKKɛ supports the anabolic demands of dendritic cell activation. Nat Immunol 2014;15:323–32. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61. Krawczyk CM, Holowka T, Sun J et al Toll‐like receptor‐induced changes in glycolytic metabolism regulate dendritic cell activation. Blood 2010;115:4742–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62. Guak H, Al Habyan S, Ma EH et al Glycolytic metabolism is essential for CCR62 oligomerization and dendritic cell migration. Nat Commun 2018;9:2463. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63. Jantsch J, Chakravortty D, Turza N et al Hypoxia and hypoxia‐inducible factor‐1 alpha modulate lipopolysaccharide‐induced dendritic cell activation and function. J Immunol 2008;180:4697–705. [DOI] [PubMed] [Google Scholar]
- 64. Pantel A, Teixeira A, Haddad E et al Direct type I IFN but not MDA5/TLR3 activation of dendritic cells is required for maturation and metabolic shift to glycolysis after poly IC stimulation. PLOS Biol 2014;12:e1001759. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65. Lawless SJ, Kedia‐Mehta N, Walls JF et al Glucose represses dendritic cell‐induced T cell responses. Nat Commun 2017;8:15620. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66. He W, Miao FJ‐P, Lin DC‐H et al Citric acid cycle intermediates as ligands for orphan G‐protein‐coupled receptors. Nature 2004;429:188–93. [DOI] [PubMed] [Google Scholar]
- 67. Rubic T, Lametschwandtner G, Jost S et al Triggering the succinate receptor GPR91 on dendritic cells enhances immunity. Nat Immunol 2008;9:1261–9. [DOI] [PubMed] [Google Scholar]
- 68. Millard AL, Mertes PM, Ittelet D et al Butyrate affects differentiation, maturation and function of human monocyte‐derived dendritic cells and macrophages. Clin Exp Immunol 2002;130:245–55. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69. Li L, Huang L, Ye H et al Dendritic cells tolerized with adenosine A₂AR agonist attenuate acute kidney injury. J Clin Invest 2012;122:3931–42. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70. Ospelt C. Synovial fibroblasts in 2017. RMD Open 2017;3:e000471. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71. Huber LC, Distler O, Tarner I et al Synovial fibroblasts: key players in rheumatoid arthritis. Rheumatology 2006;45:669–75. [DOI] [PubMed] [Google Scholar]
- 72. Serratì S, Margheri F, Chillà A et al Reduction of in vitro invasion and in vivo cartilage degradation in a SCID mouse model by loss of function of the fibrinolytic system of rheumatoid arthritis synovial fibroblasts. Arthritis Rheum 2011;63:2584–94. [DOI] [PubMed] [Google Scholar]
- 73. Biniecka M, Fox E, Gao W et al Hypoxia induces mitochondrial mutagenesis and dysfunction in inflammatory arthritis. Arthritis Rheum 2011;63:2172–82. [DOI] [PubMed] [Google Scholar]
- 74. Ahn JK, Kim S, Hwang J et al GC/TOF‐MS‐based metabolomic profiling in cultured fibroblast‐like synoviocytes from rheumatoid arthritis. Jt Bone Spine 2016;83:707–13. [DOI] [PubMed] [Google Scholar]
- 75. Balogh E, Veale DJ, McGarry T et al Oxidative stress impairs energy metabolism in primary cells and synovial tissue of patients with rheumatoid arthritis. Arthritis Res Ther 2018;20:95. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76. Garcia‐Carbonell R, Divakaruni AS, Lodi A et al Critical role of glucose metabolism in rheumatoid arthritis fibroblast‐like synoviocytes. Arthritis Rheumatol 2016;68:1614–26. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77. Mitra A, Raychaudhuri SK, Raychaudhuri SP. IL22 induced cell proliferation is regulated by PI3K/Atk/mTOR signaling cascade. Cytokine 2012;60:38–42. [DOI] [PubMed] [Google Scholar]
- 78. Laragione T, Gulko PS. mTOR regulates the invasive properties of synovial fibroblasts in rheumatoid arthritis. Mol Med 2010;16:352–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79. Karonitsch T, Kandasamy RK, Kartnig F et al mTOR senses environmental cues to shape the fibroblast‐like synoviocyte response to inflammation. Cell Rep 2018;23:2157–67. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80. Yu Z, Lin W, Rui Z et al Fibroblast‐like synoviocyte migration is enhanced by IL‐17‐mediated overexpression of L‐type amino acid transporter 1 (LAT1) via the mTOR/4E‐BP1 pathway. Amino Acids 2018;50:331–40. [DOI] [PubMed] [Google Scholar]
- 81. Guma M, Sanchez‐Lopez E, Lodi A et al Choline kinase inhibition in rheumatoid arthritis. Ann Rheum Dis 2015;74:1399–407. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82. Seki M, Kawai Y, Ishii C et al Functional analysis of choline transporters in rheumatoid arthritis synovial fibroblasts. Mod Rheumatol 2017;27:995–1003. [DOI] [PubMed] [Google Scholar]
- 83. Croft AP, Naylor AJ, Marshall JL et al Rheumatoid synovial fibroblasts differentiate into distinct subsets in the presence of cytokines and cartilage. Arthritis Res Ther 2016;18:270. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84. Mizoguchi F, Slowikowski K, Wei K et al Functionally distinct disease‐associated fibroblast subsets in rheumatoid arthritis. Nat Commun 2018;9:789. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85. Frank‐Bertoncelj M, Trenkmann M, Klein K et al Epigenetically‐driven anatomical diversity of synovial fibroblasts guides joint‐specific fibroblast functions. Nat Commun 2017;8:14852. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86. Eelen G, de Zeeuw P, Treps L, Harjes U, Wong BW, Carmeliet P. Endothelial Cell Metabolism. Physiol Rev 2018;98:3–58. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87. Cruys B, Wong BW, Kuchnio A et al Glycolytic regulation of cell rearrangement in angiogenesis. Nat Commun 2016;7:12240. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88. Akhavani MA, Madden L, Buysschaert I et al Hypoxia upregulates angiogenesis and synovial cell migration in rheumatoid arthritis. Arthritis Res Ther 2009;1:R64. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89. Lu Y, Yu SS, Zong M et al Glucose‐6‐phosphate isomerase (G6PI) mediates hypoxia‐induced angiogenesis in rheumatoid arthritis. Sci Rep 2017;7:40274. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90. Li Y, Liu Y, Wang C et al Succinate induces synovial angiogenesis in rheumatoid arthritis through metabolic remodeling and HIF1a/VEGF axis. Free Radic Biol Med 2018;26:1–14. [DOI] [PubMed] [Google Scholar]
- 91. Lu Y, Liu H, Bi Y et al Glucocorticoid receptor promotes the function of myeloid‐derived suppressor cells by suppressing HIF1a‐dependent glycolysis. Cell Mol Immunol 2018;15:618–629. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92. Moreno‐Aurioles VR, Sobrino F. Glucocorticoids inhibit fructose 2,6‐bisphosphate synthesis in rat thymocytes: opposite effect of cycloheximide. Biochim Biophys Acta 1991;1091:96–100. [DOI] [PubMed] [Google Scholar]
- 93. Swerdlow S, McColl K, Rong Y et al Apoptosis inhibition by Bcl‐2 gives way to autophagy in glucocorticoid‐treated lymphocytes. Autophagy 2008;4:612–20. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94. Ruiz‐Limón P, Ortega R, Arias de la Rosa I et al Tocilizumab improves the proatherothrombotic profile of rheumatoid arthritis patients modulating endothelial dysfunction, NETosis, and inflammation. Transl Res 2017;183:87–103. [DOI] [PubMed] [Google Scholar]
- 95. McGarry T, Orr C, Wade S et al JAK‐STAT blockade alters synovial bioenergetics, mitochondrial function and pro‐inflammatory mediators in Rheumatoid arthritis. Arthritis Rheumatol 2018. 10.1002/art.40569. [DOI] [PubMed] [Google Scholar]
- 96. Demaria M, Poli V. PKM2, STAT3 and HIF‐1alpha: the Warburg’s vicious circle. JAKSTAT 2012;1:194–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97. Bustamante MF, Oliveira PG, Garcia‐Carbonell R et al Hexokinase 2 as a novel selective metabolic target for rheumatoid arthritis. Ann Rheum Dis 2018;77:1636–1643. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98. Delgoffe GM, Powell JD. mTOR: taking cues from the immune microenvironment. Immunology 2009;127:459–65. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99. Carames B, Hasegawa A, Taniguchi N et al Autophagy activation by rapamycin reduces severity of experimental osteoarthritis. Ann Rheum Dis 2012;71:575–81. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100. Son HJ, Lee J, Lee SY et al Metformin attenuates experimental autoimmune arthritis through reciprocal regulation of Th17/Treg balance and osteoclastogenesis. Mediat Inflamm 2014;97398. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101. Yan H, Zhou HF, Hu Y, Pham CT. Suppression of experimental arthritis through AMP‐activated protein kinase activation and autophagy modulation. J Rheum Dis Treat 2015;1:5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102. Brand A, Singer K, Koehl GE et al LDHA‐associated lactic acid production blunts tumor immunosurveillance by T and NK cells. Cell Metab 2016;24:657–71. [DOI] [PubMed] [Google Scholar]
- 103. Amedei A, Niccolai E, Benagiano M et al Ex vivo analysis of pancreatic cancer‐infiltrating T lymphocytes reveals that ENO‐specific Tregs accumulate in tumor tissue and inhibit Th1/Th17 effector cell functions. Cancer Immunol Immunother 2013;62:1249–60. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104. Clementi E, Brown GC, Feelisch M, Moncada S. Persistent inhibition of cell respiration by nitric oxide: crucial role of S‐nitrosylation of mitochondrial complex I and protective action of glutathione. Proc Natl Acad Sci USA 1998;95:7631–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105. Geiger R, Rieckmann JC, Wolf T et al L‐arginine modulates T cell metabolism and enhances survival and anti‐tumor activity. Cell 2016;167:829–842. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106. Niedbala W, Cai B, Liew FY. Role of nitric oxide in the regulation of T cell functions. Ann Rheum Dis 2006;65(Suppl. 3):iii37–iii40. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107. Yoon BR, Oh Y‐J, Kang SW, Lee EB, Lee W‐W. Role of SLC7A5 in metabolic reprogramming of human monocyte/macrophage immune responses. Front Immunol 2018;9:53. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108. Oda K, Hosoda N, Endo H et al l‐Type amino acid transporter 1 inhibitors inhibit tumor cell growth. Cancer Sci 2010;101:173–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109. Wang Q, Holst J. L‐type amino acid transport and cancer: targeting the mTORC1 pathway to inhibit neoplasia. Am J Cancer Res 2015;5:1281–94. [PMC free article] [PubMed] [Google Scholar]
