Hostname: page-component-848d4c4894-8kt4b Total loading time: 0 Render date: 2024-06-21T05:00:14.185Z Has data issue: false hasContentIssue false

Metabolic signatures of immune cells in chronic kidney disease

Published online by Cambridge University Press:  21 October 2022

Jie Li
Affiliation:
Department of Nephrology, Tongji Hospital Affiliated to Tongji Medical College, Huazhong University of Science and Technology, Wuhan 430030, China
Yi Yang
Affiliation:
Department of Nephrology, Tongji Hospital Affiliated to Tongji Medical College, Huazhong University of Science and Technology, Wuhan 430030, China
Yanan Wang
Affiliation:
Department of Nephrology, Tongji Hospital Affiliated to Tongji Medical College, Huazhong University of Science and Technology, Wuhan 430030, China
Qing Li*
Affiliation:
Department of Nephrology, Tongji Hospital Affiliated to Tongji Medical College, Huazhong University of Science and Technology, Wuhan 430030, China
Fan He*
Affiliation:
Department of Nephrology, Tongji Hospital Affiliated to Tongji Medical College, Huazhong University of Science and Technology, Wuhan 430030, China
*
Author for correspondence: Fan He, E-mail: fhe@tjh.tjmu.edu.cn; Qing Li, E-mail: qing.li@tjh.tjmu.edu.cn
Author for correspondence: Fan He, E-mail: fhe@tjh.tjmu.edu.cn; Qing Li, E-mail: qing.li@tjh.tjmu.edu.cn
Rights & Permissions [Opens in a new window]

Abstract

Immune cells play a key role in maintaining renal dynamic balance and dealing with renal injury. The physiological and pathological functions of immune cells are intricately connected to their metabolic characteristics. However, immunometabolism in chronic kidney disease (CKD) is not fully understood. Pathophysiologically, disruption of kidney immune cells homeostasis causes inflammation and tissue damage via triggering metabolic reprogramming. The diverse metabolic characteristics of immune cells at different stages of CKD are strongly associated with their different pathological effect. In this work, we reviewed the metabolic characteristics of immune cells (macrophages, natural killer cells, T cells, natural killer T cells and B cells) and several non-immune cells, as well as potential treatments targeting immunometabolism in CKD. We attempt to elaborate on the metabolic signatures of immune cells and their intimate correlation with non-immune cells in CKD.

Type
Review
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
Copyright © The Author(s), 2022. Published by Cambridge University Press

Introduction

Chronic kidney disease (CKD) is tightly associated with high levels of morbidity and mortality, the global prevalence of CKD is estimated at about 14% (Ref. Reference Lv and Zhang1), which poses a significant threat to human health and economic burden especially when it progresses to uraemia (Ref. Reference Andrade-Oliveira2). The aetiology of CKD includes autoimmune disease, hypertension, diabetes mellitus, infections, polycystic kidney disease and hereditary diseases (Refs Reference Basso, Andrade-Oliveira and Camara3, Reference Ammirati4). It is worth mentioning that the immune system plays a crucial role in most forms of kidney injury, such as in diverse glomerular nephritis (GN) and IgA nephropathy (IgAN) (Ref. Reference Kurts5). In immune-mediated kidney diseases, immune cells are activated and respond to the stimuli. Sustained immune response based on inflammatory mediators including complement, reactive oxygen species and cytokines leads to persistent renal structural and functional damage, and ultimately causes irreversible renal tubulointerstitium fibrosis (Refs Reference Bohle6, Reference Markovic-Lipkovski7, Reference Risdon, Sloper and De Wardener8).

The kidney is frequently the target organ of autoimmune and systemic immune disorders. Immune responses mediated by resident immune cells or circulating recruited immune cells lead to renal damage and clinical symptoms. In kidney diseases, immune cells such as T cells, macrophages and dendritic cells (DCs) are persistently activated during the whole process. Importantly, increasing evidence indicates that the proliferation and activation of immune cells are regulated by the metabolic pathways (Ref. Reference Bonacina9). Under the condition of homeostasis, the six main metabolic pathways including glucose, fatty acid and amino acid metabolism cooperate with each other to maintain the stability of immunometabolism (Fig. 1) (Refs Reference Bonacina9, Reference Stewart10). However, pathophysiologically, metabolic reprogramming of immune cells was triggered.

Fig. 1. Six main metabolic pathways in cells. Glucose is used to produce ATP by glycolysis and OXPHOS. G-6-P is catabolised to R-5-P and NADPH through the PPP. TCA cycle, ETC and β-oxidation occur in the mitochondrial matrix, while glycolysis, FAO, FAS and glutamine metabolism occur in the cytoplasm. The intermediates of the TCA cycle are interconnected with glutamine metabolism, glycolysis and FAS. Abbreviations: R-5-P, ribose 5-phosphate; PPP, pentose phosphate pathway; FAS, fatty acid synthesis; FAO, fatty acid oxidation; G6PDH, glucose-6-phosphate dehydrogenase; HK2, hexokinase 2; PFK-1, phosphofructokinase 1; F-6-P, fructose 6 phosphate; F-1,6-2P, fructose 1,6 diphosphate; G-3-P, glyceraldehyde triphosphate; G-1,3-2P, 1,3-bisphosphoglycerate; PKM2, pyruvate kinase isozymes M2; PDH, pyruvate dehydrogenase; PEP, phosphoenolpyruvate; LADH, lactate dehydrogenase A; ACC1, acetyl-CoA carboxylase 1; CPT-1, carnitine palmitoyl transferase 1; CPS, citrate-pyruvate shuttle; α-KG, α-ketoglutarate; ETC, electron transport chain; I, II, III, IV; respiratory chain enzyme complexes I, II, III, IV; Cyt c, cytochrome c; Co Q, coenzyme Q. Red letters indicate rate-limiting enzyme.

Here we mainly introduce the immune response-mediated CKD and explore the changes in the metabolism of immune cells and non-immune cells during the progression of the disease, and expect to provide a novel direction for the remedy and retard the progression.

Immune cells and non-immune cells in CKD

The kidney immune cells include resident immune cells such as macrophages and DCs, and circulating recruited immune cells such as neutrophils, natural killer (NK) lymphocytes, natural killer T (NKT) lymphocytes, T lymphocytes and B lymphocytes (Refs Reference Kurts5, Reference Stewart10). The function of kidney immune cells has been well reviewed (Refs Reference Kurts, Ginhoux and Panzer11, Reference Tang, Nikolic-Paterson and Lan12, Reference Turner13, Reference Turner14, Reference Oleinika, Mauri and Salama15, Reference Brahler16). Kidney parenchymal cells include endothelial cells, tubular epithelial cells (TECs), mesangial cells and podocytes (Ref. Reference Kurts5). Those parenchymal cells intimately cooperate with immune cells to maintain renal homeostatic condition (Refs Reference Andrade-Oliveira2, Reference Basso, Andrade-Oliveira and Camara3). Under pathological conditions, kidney parenchymal cells express a subset of Toll-like receptors and induce an innate immune response. Immune cells mediate renal inflammation by presenting antigens (macrophages, DCs) to T cells, producing cytokines (tumour necrosis factor (TNF)-α, interleukin, interferon (IFN)-γ), chemokines, inducible nitric oxide synthase, hyaluronic acid, and so on (Ref. Reference Suarez-Fueyo17). Eventually, kidney damage and immune-related kidney diseases can occur.

Immune cells and non-immune cells and their metabolic characteristics in CKD

The role of dendritic cells and macrophages in CKD and their metabolic characteristics

Kidney resident antigen-presenting cells, macrophages and DCs are mostly enriched in the tubulointerstitium rather than the glomerulus (Refs Reference Basso, Andrade-Oliveira and Camara3, Reference Kelly and O'Neill18, Reference Krawczyk19, Reference Kruger20, Reference Kaissling and Le Hir21). Numerous studies have demonstrated that DCs are pathogenic in the mouse model of adriamycin glomerulopathy and IgAN, indicating that DCs play an important role in the progression of kidney diseases (Refs Reference Cao22, Reference Wardowska23, Reference Hochheiser, Tittel and Kurts24, Reference Zhang25). CD103 + DCs enhance the activity of CD8 + T cells and promote glomerular injury (Refs Reference Cao22, Reference Heymann26). Inhibition of CD103 + DCs attenuates kidney injury and fibrosis by reducing T helper 17 (Th17) cells and increasing regulatory T cells (Tregs) (Fig. 2d) (Refs Reference Liu27, Reference Wang28). As for macrophages, they are activated into M1 macrophages or M2 macrophages under different stimuli (Table 1) (Ref. Reference Murray29). Most forms of renal inflammation are characterised by M1 macrophage infiltration in the early phase but M2 macrophage infiltration in the chronic phase. The persistence of M2 macrophages is strongly associated with renal fibrosis and progressive CKD (Refs Reference Huen and Cantley30, Reference Eardley31, Reference Cao32, Reference Klessens33).

Fig. 2. The association between the pathogenic roles and metabolic signatures of macrophages, NK cells and DCs. (a) The metabolism signatures of macrophages and the metabolism signatures when they activate and differentiate into M1. (b) Metabolism characteristics of M1 and M2 under different antigen stimulation and corresponding kidney damage. (c) The activation and differentiation of immature NK cells and metabolism characteristics of immature NK cells and mature NK cells. (d) The metabolism characteristics of mature DCs and the interaction between DCs and T cells. M1, macrophage 1; M2, macrophage 2; mTORC1, mammalian target of rapamycin complex 1; HIF-1α, hypoxia-inducible factor 1α; OXPHOS, oxidative phosphorylation; FAS, fatty acid synthesis; PPP, pentose phosphate pathway; TLR, Toll-like receptor 4; LPS, lipopolysaccharides; NO, nitric oxide; cDC, conventional dendritic cells; NK cells, natural killer cells; iNOS, inducible nitric oxide synthase; AMPK, AMP-activated serine/threonine protein kinase; RAPA, rapamycin; MHC, major histocompatibility complex. Red lines indicate suppression.

Table 1. The metabolic characteristics of macrophages

M1, macrophage 1; M2, macrophage 2; OXPHOS, oxidative phosphorylation; iNOS, inducible nitric oxide synthase; mTORC1, mammalian target of rapamycin complex 1; HIF-1α, hypoxia-inducible factor 1 α; PPP, pentose phosphate pathway; mTORC2, mammalian target of rapamycin complex 2; LPS, lipopolysaccharides; IFN-γ, interferon-γ; IL, interleukins; MMIF, macrophage migration inhibitory factor; PAMP, pathogen-associated molecular patterns; DAMP, danger-associated molecular patterns; CRP, C-reactive protein.

Under the stimulation of pro-inflammatory signals, M1 macrophages (Table 1) and DCs are activated and undergo a metabolic switch from oxidative phosphorylation (OXPHOS) to glycolysis (Table 2) (Refs Reference Basso, Andrade-Oliveira and Camara3, Reference Kelly and O'Neill18). It is thought that cells rely on pyruvate to convert to lactate only when oxygen is insufficient. However, growing evidence suggests that tumour cells use glycolysis rather than OXPHOS pathways for energy generation even in the presence of adequate oxygen availability for mitochondrial OXPHOS (Refs Reference Pavlova and Thompson34, Reference Koppenol, Bounds and Dang35). Likewise, stimulated immune cells use glycolysis under oxygen-rich conditions. Increased aerobic glycolysis is the key metabolic phenotype of polycystic kidney disease and renal cell carcinoma (Refs Reference Rowe36, Reference Hoefflin37). Oxygen levels and changes in nutrient availability influence the metabolic reprogramming of cells. Hypoxia-inducible factor 1α (HIF-1α) is crucial for this course, which products sustained adenosine triphosphate (ATP) by inducing the expression of glucose transporter and glycolytic enzymes (Refs Reference Obach38, Reference Riddle39) (Fig. 2a). The phenotype switch of macrophages towards M1, there is an increase in glycolysis but a decrease in OXPHOS, which leads to renal injury in the early stage of kidney diseases. In contrast, M2 macrophages keep high OXPHOS and play a significant role in renal fibrosis in the chronic stage of kidney injury (Refs Reference Odegaard and Chawla40, Reference Galvan-Pena and O'Neill41) (Fig. 2b). Likewise, a recent study suggested that in lupus nephritis, macrophages metabolism in human and mouse undergoes a shift to glycolysis in answer to IgG immune complex-induced inflammation, this metabolic reprogramming was dependent on the mammalian target of rapamycin (mTOR) and HIF-1α, inhibition of glycolysis caused a decrease in the number of renal macrophages (Ref. Reference Jing42). Another research found that β-activated kinase 1-binding protein 1, a transforming growth factor, can regulate glycolysis and activate macrophages via activating nuclear factor κB (NF-κB)/HIF-1α signalling pathway in diabetic nephropathy (Ref. Reference Zeng43). Besides, treatment with glycolysis inhibitors in the mouse model of unilateral ureteric obstruction (UUO) significantly reduces macrophage infiltration and renal interstitial fibrosis (Ref. Reference Wei44), suggesting that regulating macrophage glycolytic metabolism is crucial for kidney disease progression.

Table 2. The metabolic characteristics of DCs and NK cells

DCs, dendritic cells; NK cells, natural killer cells; PAMPs, pathogen-associated molecular patterns; TCR, T cell receptor; NLR, NOD-like receptor; MHC, major histocompatibility complex; TBK, serine/threonine-protein kinase; NF-κB, nuclear factor κB; AKT, RAC-α serine/threonine protein kinase; mTORC1, mammalian target of rapamycin complex 1; HIF-1α, hypoxia-inducible factor 1α; FAO, fatty acid oxidation.

NK cell and its metabolism in CKD

NK cells are one of the specialised subpopulations of innate lymphoid cells (Ref. Reference Molofsky45). They play a key role in the immune response to anti-virus, killing tumours, aseptic inflammation, as well as kidney allograft rejection, and produce inflammatory cytokines (TNF-α, IFN-γ) and chemokines (MIP-1α, MIP-1β and RANTES) (Ref. Reference Fauriat46). Some evidence suggested that NK cells participate in acute kidney ischaemia injury and kidney allograft rejection (Refs Reference Turner13, Reference Angelo47, Reference Victorino48). Recently, Law et al. demonstrated that human tubulointerstitial CD56bright NK cells lead to the progression of CKD and kidney fibrosis by the production of proinflammatory cytokine IFN-γ (Ref. Reference Law49). IFN-γ can also induce proinflammation in kidney parenchymal cells and promotes M1 activation to exert pro-inflammation responses (Ref. Reference Turner13).

During the activation and differentiation of NK cells, metabolism changed. Immature NK cells upregulate glucose transporter and nutrient receptor expression. Whilst upon the use of interleukin-15 (IL-15) in vitro, immature NK cells differentiate and activate to be mature NK cells, which increases fatty acid oxidation (FAO) and OXPHOS (Table 2) (Refs Reference Marcais50, Reference Poznanski51). Upregulated OXPHOS pathway is critical for the production of IFN-γ for activation-induced NK cells. Furthermore, the metabolic regulator mTOR is critical for IL-15 signalling during the development and activation of NK cells and the production of IFN-γ. mTOR deficiency deeply impaired the activation of NK cells in the early stage (Ref. Reference Marcais50). mTOR is known to integrate cellular growth and metabolism via boosting cell proliferation and glycolysis (Ref. Reference Shaw52). mTORC1 is a complex composed of mTOR and other three subunits, which participate in glycolysis by facilitating the expression of HIF-1α and c-Myc. Its inhibitor rapamycin abrogates inflammation-induced priming of NK cells (Ref. Reference Marcais50). The energy sensor AMP-activated serine/threonine protein kinase (AMPK) also inhibits mTORC1 and thus inhibits the production of INF-γ (Ref. Reference Shaw52), therefore alleviating renal fibrosis (Ref. Reference Law49) (Fig. 2c).

T cells and their metabolism in CKD

T lymphocytes, the major effector cells in cellular immunity, produce cytokines in responses to mediate inflammation and coordinate other types of immune cells. In kidney diseases, T cells are well known for causing acute and CKDs, particularly in immune-mediated renal disease. Multiple evidence had shown that T cells play crucial roles in the initiation and progression of kidney diseases (Refs Reference Tapmeier53, Reference Hirooka54, Reference Mu55).

Recently, the correlation between the cellular metabolism of T cells and their function has been increasingly emphasised. In quiescent cells such as naïve T and memory T (Tm) cells, catabolic metabolism of glucose and amino acid facilitates its energy and biosynthesis requirement (Refs Reference Geltink, Kyle and Pearce56, Reference Chapman, Boothby and Chi57). After activation, T cells undergo drastic changes in function, rapid cellular growth, and the burst of cellular proliferation are engaged in (Ref. Reference Wang58). T cells reprogramme their metabolic pathways from mitochondrial OXPHOS, fatty acids β-oxidation to aerobic glycolysis, pentose-phosphate and glutaminolysis pathways. Activated T cells go from producing large amounts of ATP to producing sufficient ATP and a mass of intermediates for the generation of biomass (Ref. Reference Donnelly and Finlay59). The mTOR plays a key role in glycolytic metabolism (Refs Reference Wang58, Reference Pollizzi and Powell60). mTOR complex 1 (mTORC1) participates in the induction of T cells glycolysis and holds aerobic glycolysis of effector T cells (Refs Reference Donnelly and Finlay59, Reference Pollizzi and Powell60, Reference Finlay61). A mouse experiment found that treatment with rapamycin, an allosteric inhibitor of the mTOR, not only reduces the proportion of Th1, Th2 and Th17 cells but also boosts DCs to alleviate kidney damage in BALB/C mice with systemic lupus erythematosus (SLE) (Ref. Reference Song62). Furthermore, HIF-1α is critical for the activation of T cells (Ref. Reference Corcoran and O'Neill63). Under the stimuli of various cytokines and molecules, a certain T cell subset was activated, and the metabolism changed (Table 3). Under the induction of IL-12 or T-bet, naïve T cells differentiate into Th1 cells, which perform a high OXOHOS and glycolysis metabolism and prevent renal fibrosis (Refs Reference Geltink, Kyle and Pearce56, Reference Dong64). Activation of the Th2 and T follicular helper cells also predominately utilises OXOHOS and glycolysis metabolism to promote renal fibrosis under diverse stimuli (Refs Reference Geltink, Kyle and Pearce56, Reference Dong64, Reference MacIver, Michalek and Rathmell65, Reference Riedel66). Activated Th17 cells exploit glycolysis, OXOHOS and fatty acid synthesis (FAS) to exert pro-fibrosis function (Refs Reference Geltink, Kyle and Pearce56, Reference Dong64, Reference MacIver, Michalek and Rathmell65). Tm and naïve cells make full use of OXPHOS and FAO to sustain longevity (Refs Reference Basso, Andrade-Oliveira and Camara3, Reference Geltink, Kyle and Pearce56, Reference MacIver, Michalek and Rathmell65). Besides, it has been proposed that suppression of acetyl-CoA carboxylase 1 restrains the FAS, thus inhibiting the glycolytic-lipogenic pathway, which facilitates Tregs development and curbs Th17 cell differentiation (Ref. Reference Berod67).

Table 3. T cells subset and their metabolic characteristics

Th, T helper; Tfh, T follicular helper; Treg cells, regulatory T cells; Tn, naive T cell; TNF, tumour necrosis factor; Tm cell, memory T cells; IFN-γ, Interferon-γ; IL, iinterleukins; OXPHOS, oxidative phosphorylation; FAO, fatty acid oxidation; FAS, fatty acid synthesis; ROR, RAR-related orphan receptor; GATA-3, transcription factors GATA binding protein-3; Bcl6, B-cell CLL/lymphoma 6; TGF, tumour necrosis growth factor; STAT, selective transcription factor; IRF4, interferon regulatory factor 4; BATF, basic leucine transcription factor; c-Maf, transcription factor c-Maf; RUNX1, Runt-related transcription factor 1; SMAD, interacting transcription factor; IRI, ischaemia-reperfusion injury; Tox2, transcription factor Tox2; Blimp-1, transcription factor Blimp-1; CD40L, recombinant cluster of differentiation 40 ligand; CXCR, chemokine receptor.

NKT cells and their metabolism in CKD

NKT cells are CD1d-restricted, glycolipid antigens-reactive T cells, which express semi-invariant T cell antigen receptor (TCR) and support cell-mediated immunity against infection, autoimmunity, cancer, allergy, allograft rejection and graft-versus-host disease (Refs Reference Godfrey, Stankovic and Baxter68, Reference Bendelac, Savage and Teyton69). NKT cells are divided into invariant natural killer (iNKT) cells and type II NKT cells two populations by surface markers and their receptor TCR usage (Ref. Reference Turner14). In kidney ischaemia-reperfusion injury (IRI), iNKT cells and type II NKT cells play opposite roles. Inhibiting the production of IFN-γ by iNKT cells protects the kidney from IRI in murine model (Ref. Reference Zhang70). However, type II NKT cells protect the kidney by decreasing IFN-γ and IL-6 levels and enhancing IL-4 and IL-10 (Ref. Reference Yang71). In the murine model of experimental crescent glomerulonephritis and anti-glomerular basement membrane (anti-GBM) glomerulonephritis, a study found a reduction of iNKT cells aggravates disease progression, whereas activation of iNKT ameliorates GN (Refs Reference Riedel72, Reference Yang73, Reference Mesnard74). Similarly, diminished iNKT cell counts are observed in patients with CKD, especially in end-stage renal disease (ESRD), which revert to normal levels after renal transplantation (Refs Reference Dounousi75, Reference Peukert76). In an experimental model of lupus nephritis, repeated administration of α-GalCer to BWF1 mice ameliorates lupus disease activity and slows the development of lupus nephritis by suppressing Th2 immune responses in NKT cells (Ref. Reference Uchida77). However, Zeng et al. found administration of α-GalCer in BWF1 mice exacerbates lupus nephritis via inducing Th1-type immune responses (Ref. Reference Zeng78). These contradictory results were believed that influenced by the dose, dosing interval or the age of the mouse used, yet the molecular mechanisms behind these differences are not clear. Recently, NKT cells are reported to promote M2 macrophage-to-myofibroblast transition in kidney fibrosis by substantive IL-4 expression (Ref. Reference Liu79).

Metabolic dysregulation changes the fate and function of iNKT cells, which is tightly associated with the progression of immune-related renal disease. Normally, activated effector T cells undergo a metabolic switch from OXPHOS to glycolysis and glutaminolysis to generate high levels of ATP for rapid proliferation and fuelling substrate biogenesis and effector functions (Refs Reference van der Windt and Pearce80, Reference Pearce81). Tm cells and Tregs utilise FAO metabolism for their survivorship and durability in nutrient-poor environments (Refs Reference Salmond82, Reference Roy, Rizvi and Awasthi83), upon reactivation, Tm cells also switch to glycolytic metabolism and under polarisation to effector cells (Ref. Reference Gubser84). Nevertheless, metabolic pathways related to iNKT cells in immune-related diseases remain little known. Recently, a study demonstrated that after TCR stimulation, iNKT cells activate glycolysis, which in turn increases iNKT cell proliferation and IFN-γ production by promoting TCR recycling. Inhibiting glycolysis reduced the activity of protein kinase C and phosphoinositide 3-kinase–protein kinase B, which suppressed TCR accumulation, thus reducing the iNKT cells and immune injury (Ref. Reference Fu85). Similarly, Webb et al. revealed that pretreatment with 2-deoxyglucose, cobalt chloride and AICAR, which decrease glycolysis, upregulate HIF-1α and activate AMPK, respectively, significantly enhanced iNKT cell activation (Ref. Reference Webb86). Furthermore, Khurana et al. recently demonstrated that human iNKT cells utilise fatty acids but not glucose or glutamine as oxidative substrates exerting effector functions more than conventional T cells in the tumour microenvironment (Ref. Reference Khurana87). They possess a ‘memory-like’ metabolic phenotype but sustained high-level fatty acid metabolism upon stimulation. Nevertheless, how the metabolic programming alters the iNKT cells in CKD was not well elaborated.

B cells and their metabolism in CKD

B lymphocytes consist of several subsets, including B1 cells, B2 cells, Bregs and memory B cells. B cells play a central role in renal autoimmune diseases and renal transplantation by producing autoantibodies such as anti-neutrophil cytoplasmic antibodies (ANCA), antinuclear antibody, anti-GBM antibody and IgA (Refs Reference Wilson and Dixon88, Reference van der Woude89, Reference Tomana90), presenting antigens to T cells (Ref. Reference Rodriguez-Pinto91), and producing cytokines IL-10, transforming growth factor β (TGF-β) and IL-35 (Ref. Reference Long92). Regulatory B cells (Bregs) attenuate inflammation and conduce to the maintenance of immune tolerance. Bregs deficiency spoils renal function in the SLE, ANCA antibody-associated vasculitis, as well as in the renal transplant rejection and tolerance (Ref. Reference Oleinika, Mauri and Salama15).

The link between immunometabolism and autoimmunity has been extensively explored over the past decade. B cells, however, have received relatively little attention. Naïve B cells rely on FAO to generate ATP. During activation, OXPHOS, tricarboxylic acid (TCA) cycle and nucleotide biosynthesis increase, and FAO is downregulated (Refs Reference Raza and Clarke93, Reference Caro-Maldonado94). Plasma cells, the immunoglobulins antibody-secreting cells, take up glucose via the hexosamine pathway. They also rely on long-chain fatty acids and glutamine as substrates for oxidative metabolism to feed basal OXPHOS (Refs Reference Raza and Clarke93, Reference Price95). However, few studies have directly explored the correlation between metabolism reprogramming and B cell-related kidney diseases. More work is needed to investigate the characteristics and consequences of metabolic dysregulation in autoimmune B cell-related kidney diseases.

Metabolic characteristics in the progression of renal fibrosis

Fibrosis is characterised by aberrant extracellular matrix deposition, which leads to morbidity, dysfunction and even death. Upon activation of certain immune cells such as Th1 and CD8 + T cells restrains fibroblast-induced collagen synthesis and exerts an antifibrotic effect (Refs Reference Zhang and Zhang96, Reference Wynn97). On the contrary, activation of Th2 cell, Th17 cell and γδ T cells plays a key role in kidney fibrosis (Ref. Reference Zhang and Zhang96) (Fig. 3).

Fig. 3. The antifibrotic and profibrotic functions of various T cell subsets. PDGF, platelet-derived growth factor; Th, T helper; ECM, extracellular matrix; FGF, fibroblast growth factor; CTGF, connective tissue growth factor; TGF, transforming growth factor; α-SMA, α-smooth muscle actin; IFN-γ, Interferon-γ; IL, Interleukins; Tn, naive T cell; APCs, antigen-presenting cells; DCs, dendritic cells; EMT, epithelial-mesenchymal transition. Red lines indicate suppression.

Nowadays, preventing kidney fibrosis by regulating cellular metabolism brings novel therapeutic opportunities. Recently, a study demonstrates that orphan nuclear receptor COUP-TFII, which is a key regulator of glucose and lipid metabolism, enhances myofibroblast glycolysis leading to kidney fibrosis (Ref. Reference Li98). Moreover, it is revealed that tuberous sclerosis complex 1, a negative regulatory factor of the mTORC1, promotes the progression of kidney interstitial fibrosis by changing the level of glycolysis in proximal TECs in a mouse model of UUO (Ref. Reference Cao99). Lemos et al. found that IL-1β facilitates Myc-dependent metabolic switch from OXPHOS to glycolysis enzymes in kidney stromal cells, which accelerates the development of tubulointerstitial fibrosis (Ref. Reference Lemos100). It is reported that hypoxia plays a crucial role in the progression of renal fibrosis in ESRD (Refs Reference Nangaku101, Reference Tanaka102, Reference Fu, Colgan and Shelley103, Reference Tanaka and Nangaku104). Hypoxia-responsive signalling pathways including pathways mediated by HIF, TGF-β and NF-κB leading to renal fibrosis had been well reviewed (Ref. Reference Liu105), which also regulate cellular metabolism. For instance, HIF-1α is crucial for anaerobic glycolysis of immune cells, increasing glycolysis metabolism in NK cells and enhancing IFN-γ production, which in turn aggravates renal fibrosis (Ref. Reference Law49) (Fig. 2). TGF-β enhances glutamine metabolism, which causes lung fibroblasts by facilitating collagen protein synthesis (Ref. Reference Hewitson and Smith106). Besides, TGF-β is a strong activator of glycolysis in mesenchymal cells, it triggers the cellular metabolism shift towards glycolysis (Refs Reference Jiang107, Reference Kang108). The pentose phosphate pathway (PPP) is related to an increased requirement for R-5-P and NADPH, utilised in nucleotide synthesis and reductive biosynthesis, respectively. Increased functional activity PPP is associated with diabetic renal hypertrophy (Ref. Reference Steer, Sochor and McLean109) and renal cell carcinomas (Ref. Reference Massari110). Oxidative metabolism of glutamine produces citrate and acetyl-CoA for lipid synthesis. A key metabolic hallmark of renal cell carcinoma is increased glutamine utilisation compared to normal kidney tissues (Ref. Reference Wettersten111). However, how glutamine metabolism alters in CKD has not been well researched.

In addition to changes in glucose metabolism and glutamine metabolism, it has been confirmed that defective FAO in kidney TECs is tightly associated with renal fibrotic damage (Ref. Reference Kang112). FAO is a critical energy source for renal TECs and other high-energy demanding cells such as cardiac myocytes (Refs Reference Kang112, Reference Goldberg, Trent and Schulze113). However, excess accumulation of triglycerides induces cellular lipotoxicity. It was uncovered that renal fibrosis is associated with decreased FAO, whereas improving FAO in mouse models of renal fibrosis reduced fibrotic injury (Ref. Reference Kang112). Fierro-Fernandez et al. demonstrated that in the UUO mouse, miR-9-5p protects the kidney by mitigating the downregulation of genes associated with significant metabolic pathways, such as OXPHOS, glycolysis and FAO (Ref. Reference Fierro-Fernandez114). Furthermore, impairment of peroxisome proliferator-activated receptor α and the FAO pathway in the renal tubule epithelium aggravates age-associated renal fibrosis (Ref. Reference Chung115). In diabetic kidney disease, metabolism reprogrammes shift from FAO to glycolysis in TECs and HIF-1α plays a key role in this progress, dapagliflozin prevents the high glucose-induced metabolic shift in TECs by inhibiting HIF-1α (Ref. Reference Cai116). The amount of these studies suggested that lipotoxicity via impaired fatty acid oxidisation metabolism could trigger cell death, and inflammation and aggravate CKD (Ref. Reference Jang117).

Treatments targeting immune cells metabolic signature in CKD

Targeting immune or non-immune cell metabolism can promote or retard the progression of CKD (Table S1). For instance, tryptophan metabolism exerts immunosuppression effects by depleting tryptophan and increasing immunosuppressive metabolites in the kynurenine pathway (Ref. Reference Lee118). Overexpressing tryptophan-degrading enzyme IDO in DCs alleviates renal damage in the IgAN mouse model (Ref. Reference Liu27). Treatment with glycolysis inhibitors (dichloroacetate or shikonin) in M1 macrophages alleviates renal fibrosis in the UUO mouse model (Ref. Reference Wei44). Furthermore, pharmacological mTOR inhibitors inhibit glycolysis of NK cells both in vivo and in vitro, thus ameliorating mice renal fibrosis in the UUO model (Refs Reference Law49, Reference Marcais50). Besides, treatment with mTOR inhibitor rapamycin, which inhibits glycolysis, can reduce the proportion of Th1, Th2 and Th17 cells and boost the proportion of DCs in the kidney to alleviate kidney damage in BALB/C mice with SLE (Ref. Reference Song62). However, in addition to inhibiting metabolic pathways of immune cells that can alleviate renal damage or renal fibrosis, so do non-immune cells. Glycolysis inhibitors shikonin and 2-deoxyglucose mitigate renal interstitial fibroblasts activation and renal fibrosis in the UUO mouse model (Ref. Reference Ding119). Likewise, tsc1-associated mTORC1 signalling activation or treatment with glycolysis inhibitor 2-deoxyglucose attenuates TEC proliferation and kidney fibrosis (Ref. Reference Cao99). Kang et al. found that fenofibrate administration protected mice from the development of renal tubular epithelial fibrosis by strongly inducing transcriptional expression of FAO-related genes in a folate-induced injury model and a UUO model (Ref. Reference Kang112). Nevertheless, there are already numerous pieces of research in experimental mouse models of kidney diseases but few in clinical studies. In the near future, a large number of clinical studies of inhibitors or activators that directly act on metabolic pathways are called for.

Summary and outlook

In brief, immunometabolism as an emerging hot spot has a significant role in cancer as well as acute or chronic immune disease. It is significant and urgent to make clear the metabolic signature correlations between immune cells and parenchymal cells in the kidney. Yet, the intricate connection between immune or non-immune cells and their metabolic features has not been fully studied. In this review, we describe the metabolic reprogramming of immune cells and parenchymal cells during pathogenic processes in immune-related kidney diseases. The agents that directly act on some metabolites, metabolic pathway regulators or rate-limiting enzymes may become a novel direction for the treatment of CKD.

Supplementary material

The supplementary material for this article can be found at https://doi.org/10.1017/erm.2022.35.

Financial support

This work was supported by the National Natural Science Foundation of China (NSFC 81974089), 2019 Wuhan Huanghe talents (outstanding young talents) and the Frontier Application Basic Project of the Wuhan Science and Technology Bureau (Grant No. 2020020601012235).

Conflict of interest

None.

References

Lv, JC and Zhang, LX (2019) Prevalence and disease burden of chronic kidney disease. Advances in Experimental Medicine and Biology 1165, 315.CrossRefGoogle ScholarPubMed
Andrade-Oliveira, V et al. (2019) Inflammation in renal diseases: new and old players. Frontiers in Pharmacology 10, 1192.CrossRefGoogle ScholarPubMed
Basso, PJ, Andrade-Oliveira, V and Camara, NOS (2021) Targeting immune cell metabolism in kidney diseases. Nature Reviews. Nephrology 17, 465480.CrossRefGoogle ScholarPubMed
Ammirati, AL (2020) Chronic kidney disease. Revista da Associacao Medica Brasileira (1992) 66(suppl. 1), s03s09.CrossRefGoogle Scholar
Kurts, C et al. (2013) The immune system and kidney disease: basic concepts and clinical implications. Nature Reviews Immunology 13, 738753.CrossRefGoogle ScholarPubMed
Bohle, A et al. (1989) The pathogenesis of chronic renal failure. Pathology – Research and Practice 185, 421440.CrossRefGoogle ScholarPubMed
Markovic-Lipkovski, J et al. (1990) Association of glomerular and interstitial mononuclear leukocytes with different forms of glomerulonephritis. Nephrology Dialysis Transplantation 5, 1017.CrossRefGoogle ScholarPubMed
Risdon, RA, Sloper, JC and De Wardener, HE (1968) Relationship between renal function and histological changes found in renal-biopsy specimens from patients with persistent glomerular nephritis. The Lancet 292, 363366.CrossRefGoogle Scholar
Bonacina, F et al. (2019) The interconnection between immuno-metabolism, diabetes, and CKD. Current Diabetes Reports 19, 21.CrossRefGoogle Scholar
Stewart, BJ et al. (2019) Spatiotemporal immune zonation of the human kidney. Science 365, 14611466.CrossRefGoogle ScholarPubMed
Kurts, C, Ginhoux, F and Panzer, U (2020) Kidney dendritic cells: fundamental biology and functional roles in health and disease. Nature Reviews. Nephrology 16, 391407.CrossRefGoogle ScholarPubMed
Tang, PM, Nikolic-Paterson, DJ and Lan, HY (2019) Macrophages: versatile players in renal inflammation and fibrosis. Nature Reviews. Nephrology 15, 144158.CrossRefGoogle ScholarPubMed
Turner, JE et al. (2019) Natural killer cells in kidney health and disease. Frontiers in Immunology 10, 587.CrossRefGoogle ScholarPubMed
Turner, JE et al. (2018) Tissue-resident lymphocytes in the kidney. Journal of the American Society of Nephrology 29, 389399.CrossRefGoogle ScholarPubMed
Oleinika, K, Mauri, C and Salama, AD (2019) Effector and regulatory B cells in immune-mediated kidney disease. Nature Reviews. Nephrology 15, 1126.CrossRefGoogle ScholarPubMed
Brahler, S et al. (2018) Opposing roles of dendritic cell subsets in experimental GN. Journal of the American Society of Nephrology 29, 138154.CrossRefGoogle ScholarPubMed
Suarez-Fueyo, A et al. (2017) T cells and autoimmune kidney disease. Nature Reviews. Nephrology 13, 329343.CrossRefGoogle ScholarPubMed
Kelly, B and O'Neill, LA (2015) Metabolic reprogramming in macrophages and dendritic cells in innate immunity. Cell Research 25, 771784.CrossRefGoogle ScholarPubMed
Krawczyk, CM et al. (2010) Toll-like receptor-induced changes in glycolytic metabolism regulate dendritic cell activation. Blood 115, 47424749.CrossRefGoogle ScholarPubMed
Kruger, T et al. (2004) Identification and functional characterization of dendritic cells in the healthy murine kidney and in experimental glomerulonephritis. Journal of the American Society of Nephrology 15, 613621.CrossRefGoogle ScholarPubMed
Kaissling, B and Le Hir, M (1994) Characterization and distribution of interstitial cell types in the renal cortex of rats. Kidney International 45, 709720.CrossRefGoogle ScholarPubMed
Cao, Q et al. (2016) CD103 + dendritic cells elicit CD8 + T cell responses to accelerate kidney injury in adriamycin nephropathy. Journal of the American Society of Nephrology 27, 13441360.CrossRefGoogle ScholarPubMed
Wardowska, A et al. (2019) Transcriptomic and epigenetic alterations in dendritic cells correspond with chronic kidney disease in lupus nephritis. Frontiers in Immunology 10, 2026.CrossRefGoogle ScholarPubMed
Hochheiser, K, Tittel, A and Kurts, C (2011) Kidney dendritic cells in acute and chronic renal disease. International Journal of Experimental Pathology 92, 193201.CrossRefGoogle ScholarPubMed
Zhang, F et al. (2020) Mesenchymal stem cells alleviate rat diabetic nephropathy by suppressing CD103(+) DCs-mediated CD8(+) T cell responses. Journal of Cellular and Molecular Medicine 24, 58175831.CrossRefGoogle ScholarPubMed
Heymann, F et al. (2009) Kidney dendritic cell activation is required for progression of renal disease in a mouse model of glomerular injury. Journal of Clinical Investigation 119, 12861297.CrossRefGoogle Scholar
Liu, K et al. (2020) The therapeutic effect of dendritic cells expressing indoleamine 2,3-dioxygenase (IDO) on an IgA nephropathy mouse model. International Urology and Nephrology 52, 399407.CrossRefGoogle Scholar
Wang, R et al. (2019) Flt3 inhibition alleviates chronic kidney disease by suppressing CD103 + dendritic cell-mediated T cell activation. Nephrology Dialysis Transplantation 34, 18531863.CrossRefGoogle ScholarPubMed
Murray, PJ et al. (2014) Macrophage activation and polarization: nomenclature and experimental guidelines. Immunity 41, 1420.CrossRefGoogle ScholarPubMed
Huen, SC and Cantley, LG (2017) Macrophages in renal injury and repair. Annual Review of Physiology 79, 449469.CrossRefGoogle ScholarPubMed
Eardley, KS et al. (2006) The relationship between albuminuria, MCP-1/CCL2, and interstitial macrophages in chronic kidney disease. Kidney International 69, 11891197.CrossRefGoogle ScholarPubMed
Cao, Q et al. (2010) IL-10/TGF-beta-modified macrophages induce regulatory T cells and protect against adriamycin nephrosis. Journal of the American Society of Nephrology 21, 933942.CrossRefGoogle ScholarPubMed
Klessens, CQF et al. (2017) Macrophages in diabetic nephropathy in patients with type 2 diabetes. Nephrology Dialysis Transplantation 32, 13221329.Google ScholarPubMed
Pavlova, NN and Thompson, CB (2016) The emerging hallmarks of cancer metabolism. Cell Metabolism 23, 2747.CrossRefGoogle ScholarPubMed
Koppenol, WH, Bounds, PL and Dang, CV (2011) Otto Warburg's contributions to current concepts of cancer metabolism. Nature Reviews Cancer 11, 325337.CrossRefGoogle ScholarPubMed
Rowe, I et al. (2013) Defective glucose metabolism in polycystic kidney disease identifies a new therapeutic strategy. Nature Medicine 19, 488493.CrossRefGoogle ScholarPubMed
Hoefflin, R et al. (2020) HIF-1alpha and HIF-2alpha differently regulate tumour development and inflammation of clear cell renal cell carcinoma in mice. Nature Communications 11, 4111.CrossRefGoogle ScholarPubMed
Obach, M et al. (2004) 6-Phosphofructo-2-kinase (pfkfb3) gene promoter contains hypoxia-inducible factor-1 binding sites necessary for transactivation in response to hypoxia. Journal of Biological Chemistry 279, 5356253570.CrossRefGoogle ScholarPubMed
Riddle, SR et al. (2000) Hypoxia induces hexokinase II gene expression in human lung cell line A549. American Journal of Physiology. Lung Cellular and Molecular Physiology 278, L407L416.CrossRefGoogle ScholarPubMed
Odegaard, JI and Chawla, A (2011) Alternative macrophage activation and metabolism. Annual Review of Pathology 6, 275297.CrossRefGoogle ScholarPubMed
Galvan-Pena, S and O'Neill, LA (2014) Metabolic reprograming in macrophage polarization. Frontiers in Immunology 5, 420.Google ScholarPubMed
Jing, C et al. (2020) Macrophage metabolic reprogramming presents a therapeutic target in lupus nephritis. Proceedings of the National Academy of Sciences of the USA 117, 1516015171.CrossRefGoogle ScholarPubMed
Zeng, H et al. (2020) TAB1 regulates glycolysis and activation of macrophages in diabetic nephropathy. Inflammation Research 69, 12151234.CrossRefGoogle ScholarPubMed
Wei, Q et al. (2019) Glycolysis inhibitors suppress renal interstitial fibrosis via divergent effects on fibroblasts and tubular cells. American Journal of Physiology. Renal Physiology 316, F1162F1F72.CrossRefGoogle ScholarPubMed
Molofsky, AB et al. (2013) Innate lymphoid type 2 cells sustain visceral adipose tissue eosinophils and alternatively activated macrophages. Journal of Experimental Medicine 210, 535549.CrossRefGoogle ScholarPubMed
Fauriat, C et al. (2010) Regulation of human NK-cell cytokine and chemokine production by target cell recognition. Blood 115, 21672176.CrossRefGoogle ScholarPubMed
Angelo, LS et al. (2015) Practical NK cell phenotyping and variability in healthy adults. Immunologic Research 62, 341356.CrossRefGoogle ScholarPubMed
Victorino, F et al. (2015) Tissue-resident NK cells mediate ischemic kidney injury and are not depleted by anti-asialo-GM1 antibody. Journal of Immunology 195, 49734985.CrossRefGoogle Scholar
Law, BMP et al. (2017) Interferon-gamma production by tubulointerstitial human CD56(bright) natural killer cells contributes to renal fibrosis and chronic kidney disease progression. Kidney International 92, 7988.CrossRefGoogle ScholarPubMed
Marcais, A et al. (2014) The metabolic checkpoint kinase mTOR is essential for IL-15 signaling during the development and activation of NK cells. Nature Immunology 15, 749757.CrossRefGoogle ScholarPubMed
Poznanski, SM et al. (2018) Immunometabolism of T cells and NK cells: metabolic control of effector and regulatory function. Inflammation Research 67, 813828.CrossRefGoogle ScholarPubMed
Shaw, RJ (2009) LKB1 and AMP-activated protein kinase control of mTOR signalling and growth. Acta Physiologica 196, 6580.CrossRefGoogle ScholarPubMed
Tapmeier, TT et al. (2010) Pivotal role of CD4 + T cells in renal fibrosis following ureteric obstruction. Kidney International 78, 351362.CrossRefGoogle ScholarPubMed
Hirooka, Y et al. (2020) Foxp3-positive regulatory T cells contribute to antifibrotic effects in renal fibrosis via an interleukin-18 receptor signaling pathway. Frontiers in Medicine 7, 604656.CrossRefGoogle ScholarPubMed
Mu, Y et al. (2020) CD226 deficiency on regulatory T cells aggravates renal fibrosis via up-regulation of Th2 cytokines through miR-340. Journal of Leukocyte Biology 107, 573587.CrossRefGoogle ScholarPubMed
Geltink, RIK, Kyle, RL and Pearce, EL (2018) Unraveling the complex interplay between T cell metabolism and function. Annual Review of Immunology 36, 461488.CrossRefGoogle ScholarPubMed
Chapman, NM, Boothby, MR and Chi, H (2020) Metabolic coordination of T cell quiescence and activation. Nature Reviews Immunology 20, 5570.CrossRefGoogle ScholarPubMed
Wang, R et al. (2011) The transcription factor Myc controls metabolic reprogramming upon T lymphocyte activation. Immunity 35, 871882.CrossRefGoogle ScholarPubMed
Donnelly, RP and Finlay, DK (2015) Glucose, glycolysis and lymphocyte responses. Molecular Immunology 68, 513519.CrossRefGoogle ScholarPubMed
Pollizzi, KN and Powell, JD (2014) Integrating canonical and metabolic signalling programmes in the regulation of T cell responses. Nature Reviews Immunology 14, 435446.CrossRefGoogle ScholarPubMed
Finlay, DK et al. (2012) PDK1 regulation of mTOR and hypoxia-inducible factor 1 integrate metabolism and migration of CD8 + T cells. Journal of Experimental Medicine 209, 24412453.CrossRefGoogle ScholarPubMed
Song, X et al. (2021) Rapamycin alleviates renal damage in mice with systemic lupus erythematosus through improving immune response and function. Biomedicine & Pharmacotherapy 137, 111289.CrossRefGoogle ScholarPubMed
Corcoran, SE and O'Neill, LA (2016) HIF1alpha and metabolic reprogramming in inflammation. Journal of Clinical Investigation 126, 36993707.CrossRefGoogle ScholarPubMed
Dong, C (2021) Cytokine regulation and function in T cells. Annual Review of Immunology 39, 5176.CrossRefGoogle ScholarPubMed
MacIver, NJ, Michalek, RD and Rathmell, JC (2013) Metabolic regulation of T lymphocytes. Annual Review of Immunology 31, 259283.CrossRefGoogle ScholarPubMed
Riedel, JH et al. (2017) IL-33-mediated expansion of type 2 innate lymphoid cells protects from progressive glomerulosclerosis. Journal of the American Society of Nephrology 28, 20682080.CrossRefGoogle ScholarPubMed
Berod, L et al. (2014) De novo fatty acid synthesis controls the fate between regulatory T and T helper 17 cells. Nature Medicine 20, 13271333.CrossRefGoogle ScholarPubMed
Godfrey, DI, Stankovic, S and Baxter, AG (2010) Raising the NKT cell family. Nature Immunology 11, 197206.CrossRefGoogle ScholarPubMed
Bendelac, A, Savage, PB and Teyton, L (2007) The biology of NKT cells. Annual Review of Immunology 25, 297336.CrossRefGoogle ScholarPubMed
Zhang, J et al. (2016) Hypoxia-inducible factor-2alpha limits natural killer T cell cytotoxicity in renal ischemia/reperfusion injury. Journal of the American Society of Nephrology 27, 92106.CrossRefGoogle ScholarPubMed
Yang, SH et al. (2011) Sulfatide-reactive natural killer T cells abrogate ischemia-reperfusion injury. Journal of the American Society of Nephrology 22, 13051314.CrossRefGoogle ScholarPubMed
Riedel, JH et al. (2012) Immature renal dendritic cells recruit regulatory CXCR6(+) invariant natural killer T cells to attenuate crescentic GN. Journal of the American Society of Nephrology 23, 19872000.CrossRefGoogle ScholarPubMed
Yang, SH et al. (2008) NKT cells inhibit the development of experimental crescentic glomerulonephritis. Journal of the American Society of Nephrology 19, 16631671.CrossRefGoogle ScholarPubMed
Mesnard, L et al. (2009) Invariant natural killer T cells and TGF-beta attenuate anti-GBM glomerulonephritis. Journal of the American Society of Nephrology 20, 12821292.CrossRefGoogle ScholarPubMed
Dounousi, E et al. (2021) The innate immune system and cardiovascular disease in ESKD: monocytes and natural killer cells. Current Vascular Pharmacology 19, 6376.CrossRefGoogle ScholarPubMed
Peukert, K et al. (2014) Invariant natural killer T cells are depleted in renal impairment and recover after kidney transplantation. Nephrology Dialysis Transplantation 29, 10201028.CrossRefGoogle ScholarPubMed
Uchida, T et al. (2018) Repeated administration of alpha-galactosylceramide ameliorates experimental lupus nephritis in mice. Scientific Reports 8, 8225.CrossRefGoogle ScholarPubMed
Zeng, D et al. (2003) Activation of natural killer T cells in NZB/W mice induces Th1-type immune responses exacerbating lupus. Journal of Clinical Investigation 112, 12111222.CrossRefGoogle ScholarPubMed
Liu, B et al. (2021) Natural killer T cell/IL-4 signaling promotes bone marrow-derived fibroblast activation and M2 macrophage-to-myofibroblast transition in renal fibrosis. International Immunopharmacology 98, 107907.CrossRefGoogle ScholarPubMed
van der Windt, GJ and Pearce, EL (2012) Metabolic switching and fuel choice during T-cell differentiation and memory development. Immunological Reviews 249, 2742.CrossRefGoogle ScholarPubMed
Pearce, EL et al. (2013) Fueling immunity: insights into metabolism and lymphocyte function. Science 342, 1242454.CrossRefGoogle ScholarPubMed
Salmond, RJ (2018) mTOR regulation of glycolytic metabolism in T cells. Frontiers in Cell and Developmental Biology 6, 122.CrossRefGoogle ScholarPubMed
Roy, S, Rizvi, ZA and Awasthi, A (2018) Metabolic checkpoints in differentiation of helper T cells in tissue inflammation. Frontiers in Immunology 9, 3036.CrossRefGoogle ScholarPubMed
Gubser, PM et al. (2013) Rapid effector function of memory CD8 + T cells requires an immediate-early glycolytic switch. Nature Immunology 14, 10641072.CrossRefGoogle ScholarPubMed
Fu, S et al. (2019) Immunometabolism regulates TCR recycling and iNKT cell functions. Science Signaling 12, eaau1788.CrossRefGoogle ScholarPubMed
Webb, TJ et al. (2016) Alterations in cellular metabolism modulate CD1d-mediated NKT-cell responses. Pathogens and Disease 74, ftw055.CrossRefGoogle ScholarPubMed
Khurana, P et al. (2021) Distinct bioenergetic features of human invariant natural killer T cells enable retained functions in nutrient-deprived states. Frontiers in Immunology 12, 700374.CrossRefGoogle ScholarPubMed
Wilson, CB and Dixon, FJ (1973) Anti-glomerular basement membrane antibody-induced glomerulonephritis. Kidney International 3, 7489.CrossRefGoogle ScholarPubMed
van der Woude, FJ et al. (1985) Autoantibodies against neutrophils and monocytes: tool for diagnosis and marker of disease activity in Wegener's granulomatosis. Lancet 1, 425429.CrossRefGoogle ScholarPubMed
Tomana, M et al. (1999) Circulating immune complexes in IgA nephropathy consist of IgA1 with galactose-deficient hinge region and antiglycan antibodies. Journal of Clinical Investigation 104, 7381.CrossRefGoogle ScholarPubMed
Rodriguez-Pinto, D (2005) B cells as antigen presenting cells. Cellular Immunology 238, 6775.CrossRefGoogle ScholarPubMed
Long, W et al. (2021) The role of regulatory B cells in kidney diseases. Frontiers in Immunology 12, 683926.CrossRefGoogle ScholarPubMed
Raza, IGA and Clarke, AJ (2021) B cell metabolism and autophagy in autoimmunity. Frontiers in Immunology 12, 681105.CrossRefGoogle ScholarPubMed
Caro-Maldonado, A et al. (2014) Metabolic reprogramming is required for antibody production that is suppressed in anergic but exaggerated in chronically BAFF-exposed B cells. Journal of Immunology 192, 36263636.CrossRefGoogle ScholarPubMed
Price, MJ et al. (2018) Progressive upregulation of oxidative metabolism facilitates plasmablast differentiation to a T-independent antigen. Cell Reports 23, 31523159.CrossRefGoogle ScholarPubMed
Zhang, M and Zhang, S (2020) T cells in fibrosis and fibrotic diseases. Frontiers in Immunology 11, 1142.CrossRefGoogle ScholarPubMed
Wynn, TA (2004) Fibrotic disease and the T(H)1/T(H)2 paradigm. Nature Reviews Immunology 4, 583594.CrossRefGoogle Scholar
Li, L et al. (2021) Orphan nuclear receptor COUP-TFII enhances myofibroblast glycolysis leading to kidney fibrosis. EMBO Reports 22, e51169.CrossRefGoogle ScholarPubMed
Cao, H et al. (2020) Tuberous sclerosis 1 (Tsc1) mediated mTORC1 activation promotes glycolysis in tubular epithelial cells in kidney fibrosis. Kidney International 98, 686698.CrossRefGoogle ScholarPubMed
Lemos, DR et al. (2018) Interleukin-1beta activates a MYC-dependent metabolic switch in kidney stromal cells necessary for progressive tubulointerstitial fibrosis. Journal of the American Society of Nephrology 29, 16901705.CrossRefGoogle ScholarPubMed
Nangaku, M (2006) Chronic hypoxia and tubulointerstitial injury: a final common pathway to end-stage renal failure. Journal of the American Society of Nephrology 17, 1725.CrossRefGoogle ScholarPubMed
Tanaka, T (2016) Expanding roles of the hypoxia-response network in chronic kidney disease. Clinical and Experimental Nephrology 20, 835844.CrossRefGoogle ScholarPubMed
Fu, Q, Colgan, SP and Shelley, CS (2016) Hypoxia: the force that drives chronic kidney disease. Clinical Medicine & Research 14, 1539.CrossRefGoogle ScholarPubMed
Tanaka, T and Nangaku, M (2010) The role of hypoxia, increased oxygen consumption, and hypoxia-inducible factor-1 alpha in progression of chronic kidney disease. Current Opinion in Nephrology and Hypertension 19, 4350.CrossRefGoogle ScholarPubMed
Liu, M et al. (2017) Signalling pathways involved in hypoxia-induced renal fibrosis. Journal of Cellular and Molecular Medicine 21, 12481259.CrossRefGoogle ScholarPubMed
Hewitson, TD and Smith, ER (2021) A metabolic reprogramming of glycolysis and glutamine metabolism Is a requisite for renal fibrogenesis-why and how? Frontiers in Physiology 12, 645857.CrossRefGoogle ScholarPubMed
Jiang, L et al. (2015) Metabolic reprogramming during TGFbeta1-induced epithelial-to-mesenchymal transition. Oncogene 34, 39083916.CrossRefGoogle ScholarPubMed
Kang, YP et al. (2016) Metabolic profiling regarding pathogenesis of idiopathic pulmonary fibrosis. Journal of Proteome Research 15, 17171724.CrossRefGoogle ScholarPubMed
Steer, KA, Sochor, M and McLean, P (1985) Renal hypertrophy in experimental diabetes. Changes in pentose phosphate pathway activity. Diabetes 34, 485490.CrossRefGoogle ScholarPubMed
Massari, F et al. (2015) Metabolic alterations in renal cell carcinoma. Cancer Treatment Reviews 41, 767776.CrossRefGoogle ScholarPubMed
Wettersten, HI et al. (2017) Metabolic reprogramming in clear cell renal cell carcinoma. Nature Reviews. Nephrology 13, 410419.CrossRefGoogle ScholarPubMed
Kang, HM et al. (2015) Defective fatty acid oxidation in renal tubular epithelial cells has a key role in kidney fibrosis development. Nature Medicine 21, 3746.CrossRefGoogle Scholar
Goldberg, IJ, Trent, CM and Schulze, PC (2012) Lipid metabolism and toxicity in the heart. Cell Metabolism 15, 805812.CrossRefGoogle ScholarPubMed
Fierro-Fernandez, M et al. (2020) MiR-9-5p protects from kidney fibrosis by metabolic reprogramming. FASEB Journal 34, 410431.CrossRefGoogle ScholarPubMed
Chung, KW et al. (2018) Impairment of PPARalpha and the fatty acid oxidation pathway aggravates renal fibrosis during aging. Journal of the American Society of Nephrology 29, 12231237.CrossRefGoogle ScholarPubMed
Cai, T et al. (2020) Sodium-glucose cotransporter 2 inhibition suppresses HIF-1alpha-mediated metabolic switch from lipid oxidation to glycolysis in kidney tubule cells of diabetic mice. Cell death & disease 11, 390.CrossRefGoogle ScholarPubMed
Jang, HS et al. (2020) Defective mitochondrial fatty acid oxidation and lipotoxicity in kidney diseases. Frontiers in Medicine 7, 65.CrossRefGoogle ScholarPubMed
Lee, GK et al. (2002) Tryptophan deprivation sensitizes activated T cells to apoptosis prior to cell division. Immunology 107, 452460.CrossRefGoogle ScholarPubMed
Ding, H et al. (2017) Inhibiting aerobic glycolysis suppresses renal interstitial fibroblast activation and renal fibrosis. American Journal of Physiology. Renal Physiology 313, F561FF75.CrossRefGoogle ScholarPubMed
Figure 0

Fig. 1. Six main metabolic pathways in cells. Glucose is used to produce ATP by glycolysis and OXPHOS. G-6-P is catabolised to R-5-P and NADPH through the PPP. TCA cycle, ETC and β-oxidation occur in the mitochondrial matrix, while glycolysis, FAO, FAS and glutamine metabolism occur in the cytoplasm. The intermediates of the TCA cycle are interconnected with glutamine metabolism, glycolysis and FAS. Abbreviations: R-5-P, ribose 5-phosphate; PPP, pentose phosphate pathway; FAS, fatty acid synthesis; FAO, fatty acid oxidation; G6PDH, glucose-6-phosphate dehydrogenase; HK2, hexokinase 2; PFK-1, phosphofructokinase 1; F-6-P, fructose 6 phosphate; F-1,6-2P, fructose 1,6 diphosphate; G-3-P, glyceraldehyde triphosphate; G-1,3-2P, 1,3-bisphosphoglycerate; PKM2, pyruvate kinase isozymes M2; PDH, pyruvate dehydrogenase; PEP, phosphoenolpyruvate; LADH, lactate dehydrogenase A; ACC1, acetyl-CoA carboxylase 1; CPT-1, carnitine palmitoyl transferase 1; CPS, citrate-pyruvate shuttle; α-KG, α-ketoglutarate; ETC, electron transport chain; I, II, III, IV; respiratory chain enzyme complexes I, II, III, IV; Cyt c, cytochrome c; Co Q, coenzyme Q. Red letters indicate rate-limiting enzyme.

Figure 1

Fig. 2. The association between the pathogenic roles and metabolic signatures of macrophages, NK cells and DCs. (a) The metabolism signatures of macrophages and the metabolism signatures when they activate and differentiate into M1. (b) Metabolism characteristics of M1 and M2 under different antigen stimulation and corresponding kidney damage. (c) The activation and differentiation of immature NK cells and metabolism characteristics of immature NK cells and mature NK cells. (d) The metabolism characteristics of mature DCs and the interaction between DCs and T cells. M1, macrophage 1; M2, macrophage 2; mTORC1, mammalian target of rapamycin complex 1; HIF-1α, hypoxia-inducible factor 1α; OXPHOS, oxidative phosphorylation; FAS, fatty acid synthesis; PPP, pentose phosphate pathway; TLR, Toll-like receptor 4; LPS, lipopolysaccharides; NO, nitric oxide; cDC, conventional dendritic cells; NK cells, natural killer cells; iNOS, inducible nitric oxide synthase; AMPK, AMP-activated serine/threonine protein kinase; RAPA, rapamycin; MHC, major histocompatibility complex. Red lines indicate suppression.

Figure 2

Table 1. The metabolic characteristics of macrophages

Figure 3

Table 2. The metabolic characteristics of DCs and NK cells

Figure 4

Table 3. T cells subset and their metabolic characteristics

Figure 5

Fig. 3. The antifibrotic and profibrotic functions of various T cell subsets. PDGF, platelet-derived growth factor; Th, T helper; ECM, extracellular matrix; FGF, fibroblast growth factor; CTGF, connective tissue growth factor; TGF, transforming growth factor; α-SMA, α-smooth muscle actin; IFN-γ, Interferon-γ; IL, Interleukins; Tn, naive T cell; APCs, antigen-presenting cells; DCs, dendritic cells; EMT, epithelial-mesenchymal transition. Red lines indicate suppression.

Supplementary material: File

Li et al. supplementary material

Table S1

Download Li et al. supplementary material(File)
File 13.1 KB